ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/ripplePaper/ripple.tex
Revision: 3006
Committed: Tue Sep 19 14:45:15 2006 UTC (17 years, 11 months ago) by gezelter
Content type: application/x-tex
File size: 29570 byte(s)
Log Message:
latest version

File Contents

# User Rev Content
1 gezelter 3006 \documentclass[aps,endfloats*,preprint,amssymb]{revtex4}
2     \usepackage{epsfig}
3 gezelter 2143 \usepackage{times}
4 gezelter 3006 \usepackage{mathptm}
5 xsun 2138
6     \begin{document}
7 gezelter 3006 \renewcommand{\thefootnote}{\fnsymbol{footnote}}
8     \renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
9 xsun 2138
10 gezelter 3006 \bibliographystyle{pccp}
11    
12     \title{Symmetry breaking and the $P_{\beta'}$ Ripple phase}
13     \author{Xiuquan Sun and J. Daniel Gezelter}
14     \email[]{E-mail: gezelter@nd.edu}
15     \affiliation{Department of Chemistry and Biochemistry,\\
16     University of Notre Dame, \\
17 xsun 2138 Notre Dame, Indiana 46556}
18    
19 gezelter 2143 \date{\today}
20    
21 xsun 2138 \begin{abstract}
22 xsun 2254 The ripple phase in phosphatidylcholine (PC) bilayers has never been
23 gezelter 3006 completely explained. We present a simple (XYZ) spin-lattice model
24     that allows spins to vary their elevation as well as their
25 xsun 2254 orientation. The extra degree of freedom allows hexagonal lattices of
26     spins to find states that break out of the normally frustrated
27 gezelter 3006 randomized states and are stabilized by long-range antiferroelectric
28 xsun 2254 ordering. To break out of the frustrated states, the spins must form
29 gezelter 3006 ``rippled'' phases that make the lattices effectively
30     non-hexagonal. Our XYZ model contains a hydrophobic interaction and
31     dipole-dipole interactions to describe the interaction potential for
32     model lipid molecules. We find non-thermal ripple phases and note
33     that the wave vectors for the ripples are always perpendicular to the
34     director axis for the dipoles. Non-hexagonal lattices of dipoles are
35     not inherently frustrated, and are therefore less likely to form
36     ripple phases because they can easily form low-energy
37     antiferroelectric states. We see that the dipolar order-disorder
38     transition is substantially lower for hexagonal lattices and that the
39     ordered phase for this lattice is clearly rippled.
40 xsun 2138 \end{abstract}
41    
42 gezelter 3006 \maketitle
43    
44    
45 xsun 2138 \section{Introduction}
46     \label{Int}
47     Fully hydrated lipids will aggregate spontaneously to form bilayers
48 gezelter 3006 which exhibit a variety of phases depending on their temperatures and
49     compositions. Among these phases, a periodic rippled phase
50     ($P_{\beta'}$) appears as an intermediate phase between the gel
51     ($L_\beta$) and fluid ($L_{\alpha}$) phases for relatively pure
52     phosphatidylcholine (PC) bilayers. The ripple phase has attracted
53     substantial experimental interest over the past 30 years. Most
54     structural information of the ripple phase has been obtained by the
55     X-ray diffraction~\cite{Sun96,Katsaras00} and freeze-fracture electron
56     microscopy (FFEM).~\cite{Copeland80,Meyer96} Recently, Kaasgaard {\it
57     et al.} used atomic force microscopy (AFM) to observe ripple phase
58     morphology in bilayers supported on mica.~\cite{Kaasgaard03} The
59     experimental results provide strong support for a 2-dimensional
60     hexagonal packing lattice of the lipid molecules within the ripple
61     phase. This is a notable change from the observed lipid packing
62     within the gel phase.~\cite{Cevc87}
63 xsun 2138
64 gezelter 3006 A number of theoretical models have been presented to explain the
65     formation of the ripple phase. Marder {\it et al.} used a
66     curvature-dependent Landau-de Gennes free-energy functional to predict
67     a rippled phase.~\cite{Marder84} This model and other related continuum
68     models predict higher fluidity in convex regions and that concave
69     portions of the membrane correspond to more solid-like regions.
70     Carlson and Sethna used a packing-competition model (in which head
71     groups and chains have competing packing energetics) to predict the
72     formation of a ripple-like phase. Their model predicted that the
73     high-curvature portions have lower-chain packing and correspond to
74     more fluid-like regions. Goldstein and Leibler used a mean-field
75     approach with a planar model for {\em inter-lamellar} interactions to
76     predict rippling in multilamellar phases.~\cite{Goldstein88} McCullough
77     and Scott proposed that the {\em anisotropy of the nearest-neighbor
78     interactions} coupled to hydrophobic constraining forces which
79     restrict height differences between nearest neighbors is the origin of
80     the ripple phase.~\cite{McCullough90} Lubensky and MacKintosh
81     introduced a Landau theory for tilt order and curvature of a single
82     membrane and concluded that {\em coupling of molecular tilt to membrane
83     curvature} is responsible for the production of
84     ripples.~\cite{Lubensky93} Misbah, Duplat and Houchmandzadeh proposed
85     that {\em inter-layer dipolar interactions} can lead to ripple
86     instabilities.~\cite{Misbah98} Heimburg presented a {\em coexistence
87     model} for ripple formation in which he postulates that fluid-phase
88     line defects cause sharp curvature between relatively flat gel-phase
89     regions.~\cite{Heimburg00} Kubica has suggested that a lattice model of
90     polar head groups could be valuable in trying to understand bilayer
91     phase formation.~\cite{Kubica02} Bannerjee used Monte Carlo simulations
92     of lamellar stacks of hexagonal lattices to show that large headgroups
93     and molecular tilt with respect to the membrane normal vector can
94     cause bulk rippling.~\cite{Bannerjee02}
95 xsun 2138
96 gezelter 3006 Large-scale molecular dynamics simulations have also been performed on
97     rippled phases using united atom as well as molecular scale
98     models. De~Vries {\it et al.} studied the structure of lecithin ripple
99     phases via molecular dynamics and their simulations seem to support
100     the coexistence models (i.e. fluid-like chain dynamics was observed in
101     the kink regions).~\cite{deVries05} Ayton and Voth have found
102     significant undulations in zero-surface-tension states of membranes
103     simulated via dissipative particle dynamics, but their results are
104     consistent with purely thermal undulations.~\cite{Ayton02} Brannigan,
105     Tamboli and Brown have used a molecular scale model to elucidate the
106     role of molecular shape on membrane phase behavior and
107     elasticity.~\cite{Brannigan04b} They have also observed a buckled hexatic
108     phase with strong tail and moderate alignment attractions.~\cite{Brannigan04a}
109 gezelter 2143
110 gezelter 3006 Ferroelectric states (with long-range dipolar order) can be observed
111     in dipolar systems with non-hexagonal packings. However, {\em
112     hexagonally}-packed 2-D dipolar systems are inherently frustrated and
113     one would expect a dipolar-disordered phase to be the lowest free
114     energy configuration. Concomitantly, it would seem unlikely that a
115     frustrated lattice in a dipolar-disordered state could exhibit the
116     long-range periodicity in the range of 100-600 \AA (as exhibited in
117     the ripple phases studied by Kaasgard {\it et al.}).~\cite{Kaasgaard03}
118    
119     The various theoretical models have attributed membrane rippling to
120     various causes which appear contradictory. We are left with a number
121     of open questions: 1) Are inter-layer interactions required to explain
122     the ripple, or can a single bilayer (or even a single leaf) exhibit
123     the rippling? 2) To what degree is the dipolar anisotropy of the head
124     group important in determining the rippling? 3) Is chain fluidity
125     required? (i.e. are the coexistence models necessary to explain the
126     ripple phenomenon?) 4) How could a state with long-range order be
127     formed using a substrate consisting of 2-D hexagonally-packed dipolar
128     molecules? What we present here is an attempt to find the simplest
129     model which will exhibit this phenomenon. We are using a very simple
130     modified XYZ lattice model; details of the model can be found in
131     section \ref{sec:model}, results of Monte Carlo simulations using this
132     model are presented in section
133     \ref{sec:results}, and section \ref{sec:discussion} contains our conclusions.
134    
135     \section{The Web-of-Dipoles Model}
136     \label{sec:model}
137     The model used in our simulations is shown schematically in
138     Figs. \ref{fmod1} and \ref{fmod2}.
139    
140     \begin{figure}[ht]
141 xsun 2138 \centering
142 gezelter 3006 \caption{The modified X-Y-Z model used in our simulations. Point dipoles are
143     represented as arrows. Dipoles are locked to the lattice points in x-y
144     plane and connect to their nearest neighbors with harmonic
145     potentials. The lattice parameters $a$ and $b$ are indicated above.}
146     \includegraphics[width=\linewidth]{picture/WebOfDipoles.eps}
147 xsun 2138 \label{fmod1}
148     \end{figure}
149 gezelter 2143
150 gezelter 3006 \begin{figure}[ht]
151 gezelter 2143 \centering
152 gezelter 3006 \caption{The 6 coordinates describing the state of a 2-dipole system
153     in our extended X-Y-Z model. $z_i$ is the height of dipole $i$ from
154     an arbitrary x-y plane, $\theta_i$ is the angle that the dipole makes
155     with the laboratory z-axis and $\phi_i$ is the angle between
156     the projection of the dipole on x-y plane with the x axis.}
157 xsun 2138 \includegraphics[width=\linewidth]{picture/xyz.eps}
158     \label{fmod2}
159     \end{figure}
160 gezelter 2143
161 gezelter 3006 In this model, lipid molecules are represented by point-dipoles (which
162     is a reasonable approximation to the zwitterionic head groups of the
163     phosphatidylcholine head groups). The dipoles are locked in place to
164     their original lattice sites on the x-y plane. The original lattice
165     may be either hexagonal ($a/b = \sqrt{3}$) or non-hexagonal. However,
166     each dipole has 3 degrees of freedom. They may move freely {\em out} of the
167     x-y plane (along the $z$ axis), and they have complete orientational
168     freedom ($0 <= \theta <= \pi$, $0 <= \phi < 2 \pi$). This is a
169     modified X-Y-Z model with translational freedom along the z-axis.
170    
171     The potential energy of the system,
172 xsun 2138 \begin{equation}
173 gezelter 3006 V = \sum_i \left[ \sum_{j>i} V^{\mathrm{dd}}_{ij} + \frac{1}{2}\sum_{j
174     \in NN_i}^6 V^{\mathrm{harm}}_{ij} \right]
175 xsun 2138 \end{equation}
176 gezelter 3006 The dipolar head groups interact via a traditional point-dipolar
177     electrostatic potential,
178 xsun 2138 \begin{equation}
179 gezelter 3006 V^{\mathrm{dd}}_{ij} = \frac{|\mu|^2}{4\pi \epsilon_0 r_{ij}^3} \left[
180     {\mathbf{\hat u}_i} \cdot {\mathbf{\hat u}_j} -
181     3({\mathbf{\hat u}_i} \cdot {\mathbf{\hat
182     r}_{ij}})({\mathbf{\hat u}_j} \cdot {\mathbf{\hat r}_{ij}})
183     \right],
184     \label{eq:vdd}
185 xsun 2138 \end{equation}
186 gezelter 3006 and the hydrophobic interactions are approximated with a nearest
187     neighbor sum of harmonic interactions,
188     \begin{equation}
189     V^{\mathrm{harm}}_{ij} = \frac{k_r}{2} \left(r_{ij}-r_0\right)^2
190     \end{equation}
191     In these potentials, $\mathbf{\hat u}_i$ is the unit vector pointing
192     along dipole $i$ and $\mathbf{\hat r}_{ij}$ is the unit vector
193     pointing along the inter-dipole vector $\mathbf{r}_{ij}$. The entire
194     potential is governed by three parameters, the dipolar strength
195     ($\mu$), the harmonic spring constant ($k_r$) and the preferred
196     intermolecular spacing ($r_0$). In practice, we set the value of
197     $r_0$ to the average inter-molecular spacing from the planar lattice,
198     yielding a potential model that has only two parameters for a
199     particular choice of lattice constants $a$ (along the $x$-axis) and $b$
200     (along the $y$-axis).
201 xsun 2138
202 gezelter 3006 To investigate the phase behavior of this model, we have performed a
203     series of Metropolis Monte Carlo simulations of moderately-sized (24
204     nm on a side) patches of membrane hosted on both hexagonal ($\gamma =
205     a/b = \sqrt{3}$) and non-hexagonal ($\gamma \neq \sqrt{3}$) lattices.
206     The linear extent of one edge of the monolayer was $20 a$ and the
207     system was kept roughly square. The average distance that coplanar
208     dipoles were positioned from their six nearest neighbors was $7$ \AA.
209     Typical system sizes were 1360 lipids for the hexagonal lattices and
210     840-2800 lipids for the non-hexagonal lattices. Periodic boundary
211     conditions were used, and the cutoff for the dipole-dipole interaction
212     was set to 30 \AA. All parameters ($T$, $\mu$, $k_r$, $\gamma$) were
213     varied systematically to study the effects of these parameters on the
214     formation of ripple-like phases.
215 xsun 2138
216 gezelter 3006 \section{Results and Analysis}
217     \label{sec:results}
218    
219     \subsection{Dipolar Ordering and Coexistence Temperatures}
220     The principal method for observing the orientational ordering
221     transition in dipolar systems is the $P_2$ order parameter (defined as
222     $1.5 \times \lambda_{max}$, where $\lambda_{max}$ is the largest
223     eigenvalue of the matrix,
224     \begin{equation}
225     {\mathsf{S}} = \frac{1}{N} \sum_i \left(
226     \begin{array}{ccc}
227     u^{x}_i u^{x}_i-\frac{1}{3} & u^{x}_i u^{y}_i & u^{x}_i u^{z}_i \\
228     u^{y}_i u^{x}_i & u^{y}_i u^{y}_i -\frac{1}{3} & u^{y}_i u^{z}_i \\
229     u^{z}_i u^{x}_i & u^{z}_i u^{y}_i & u^{z}_i u^{z}_i -\frac{1}{3}
230     \end{array} \right).
231     \label{eq:opmatrix}
232     \end{equation}
233     Here $u^{\alpha}_i$ is the $\alpha=x,y,z$ component of the unit vector
234     for dipole $i$. $P_2$ will be $1.0$ for a perfectly-ordered system
235     and near $0$ for a randomized system. Note that this order parameter
236     is {\em not} equal to the polarization of the system. For example,
237     the polarization of the perfect antiferroelectric system is $0$, but
238     $P_2$ for an antiferroelectric system is $1$. The eigenvector of
239     $\mathsf{S}$ corresponding to the largest eigenvalue is familiar as
240     the director axis, which can be used to determine a priveleged dipolar
241     axis for dipole-ordered systems. Fig. \ref{t-op} shows the values of
242     $P_2$ as a function of temperature for both hexagonal ($\gamma =
243     1.732$) and non-hexagonal ($\gamma=1.875$) lattices.
244 gezelter 2143
245 gezelter 3006 \begin{figure}[ht]
246 gezelter 2143 \centering
247 gezelter 3006 \caption{The $P_2$ dipolar order parameter as a function of
248     temperature for both hexagonal ($\gamma = 1.732$) and non-hexagonal
249     ($\gamma = 1.875$) lattices}
250     \includegraphics[width=\linewidth]{picture/t-orderpara.eps}
251 gezelter 2143 \label{t-op}
252 xsun 2138 \end{figure}
253 gezelter 2143
254 gezelter 3006 There is a clear order-disorder transition in evidence from this data.
255     Both the hexagonal and non-hexagonal lattices have dipolar-ordered
256     low-temperature phases, and orientationally-disordered high
257     temperature phases. The coexistence temperature for the hexagonal
258     lattice is significantly lower than for the non-hexagonal lattices,
259     and the bulk polarization is approximately $0$ for both dipolar
260     ordered and disordered phases. This gives strong evidence that the
261     dipolar ordered phase is antiferroelectric. We have repeated the
262     Monte Carlo simulations over a wide range of lattice ratios ($\gamma$)
263     to generate a dipolar order/disorder phase diagram. Fig. \ref{phase}
264     shows that the hexagonal lattice is a low-temperature cusp in the
265     $T-\gamma$ phase diagram.
266 gezelter 2143
267 gezelter 3006 \begin{figure}[ht]
268 gezelter 2143 \centering
269 gezelter 3006 \caption{The phase diagram for the web-of-dipoles model. The line
270     denotes the division between the dipolar ordered (antiferroelectric)
271     and disordered phases. An enlarged view near the hexagonal lattice is
272     shown inset.}
273     \includegraphics[width=\linewidth]{picture/phase.eps}
274     \label{phase}
275 xsun 2138 \end{figure}
276 gezelter 2143
277 gezelter 3006 This phase diagram is remarkable in that it shows an antiferroelectric
278     phase near $\gamma=1.732$ where one would expect lattice frustration
279     to result in disordered phases at all temperatures. Observations of
280     the configurations in this phase show clearly that the system has
281     accomplished dipolar orderering by forming large ripple-like
282     structures. We have observed antiferroelectric ordering in all three
283     of the equivalent directions on the hexagonal lattice, and the dipoles
284     have been observed to organize perpendicular to the membrane normal
285     (in the plane of the membrane). It is particularly interesting to
286     note that the ripple-like structures have also been observed to
287     propagate in the three equivalent directions on the lattice, but the
288     {\em direction of ripple propagation is always perpendicular to the
289     dipole director axis}. A snapshot of a typical antiferroelectric
290     rippled structure is shown in Fig. \ref{fig:snapshot}.
291 xsun 2138
292 gezelter 3006 \begin{figure}[ht]
293     \centering
294     \caption{Top and Side views of a representative configuration for the
295     dipolar ordered phase supported on the hexagonal lattice. Note the
296     antiferroelectric ordering and the long wavelength buckling of the
297     membrane. Dipolar ordering has been observed in all three equivalent
298     directions on the hexagonal lattice, and the ripple direction is
299     always perpendicular to the director axis for the dipoles.}
300     \includegraphics[width=\linewidth]{picture/snapshot.eps}
301     \label{fig:snapshot}
302     \end{figure}
303 gezelter 2143
304 gezelter 3006 \subsection{Discriminating Ripples from Thermal Undulations}
305    
306     In order to be sure that the structures we have observed are actually
307     a rippled phase and not merely thermal undulations, we have computed
308     the undulation spectrum,
309     \begin{equation}
310     h(\vec{q}) = A^{-1/2} \int d\vec{r}
311     h(\vec{r}) e^{-i \vec{q}\cdot\vec{r}}
312     \end{equation}
313     where $h(\vec{r})$ is the height of the membrane at location $\vec{r}
314     = (x,y)$.~\cite{Safran94} In simple (and more complicated) elastic
315     continuum models, Brannigan {\it et al.} have shown that in the $NVT$
316     ensemble, the absolute value of the undulation spectrum can be
317     written,
318     \begin{equation}
319     \langle | h(q)|^2 \rangle_{NVT} = \frac{k_B T}{k_c |\vec{q}|^4 +
320     \tilde{\gamma}|\vec{q}|^2},
321     \label{eq:fit}
322     \end{equation}
323     where $k_c$ is the bending modulus for the membrane, and
324     $\tilde{\gamma}$ is the mechanical surface
325     tension.~\cite{Brannigan04b}
326    
327     The undulation spectrum is computed by superimposing a rectangular
328     grid on top of the membrane, and by assigning height ($h(\vec{r})$)
329     values to the grid from the average of all dipoles that fall within a
330     given $\vec{r}+d\vec{r}$ grid area. Empty grid pixels are assigned
331     height values by interpolation from the nearest neighbor pixels. A
332     standard 2-d Fourier transform is then used to obtain $\langle |
333     h(q)|^2 \rangle$.
334    
335     The systems studied in this paper have relatively small bending moduli
336     ($k_c$) and relatively large mechanical surface tensions
337     ($\tilde{\gamma}$). In practice, the best fits to our undulation
338     spectra are obtained by approximating the value of $k_c$ to 0. In
339     Fig. \ref{fig:fit} we show typical undulation spectra for two
340     different regions of the phase diagram along with their fits from the
341     Landau free energy approach (Eq. \ref{eq:fit}). In the
342     high-temperature disordered phase, the Landau fits can be nearly
343     perfect, and from these fits we can estimate the bending modulus and
344     the mechanical surface tension.
345    
346     For the dipolar-ordered hexagonal lattice near the coexistence
347     temperature, however, we observe long wavelength undulations that are
348     far outliers to the fits. That is, the Landau free energy fits are
349     well within error bars for all other points, but can be off by {\em
350     orders of magnitude} for a few (but not all) low frequency
351     components.
352    
353     We interpret these outliers as evidence that these low frequency modes
354     are {\em non-thermal undulations} which is clear evidence that we are
355     actually seeing a rippled phase developing in this model system.
356    
357     \begin{figure}[ht]
358 gezelter 2143 \centering
359 gezelter 3006 \caption{Evidence that the observed ripples are {\em not} thermal
360     undulations is obtained from the 2-d fourier transform $\langle
361     |h(\vec{q})|^2 \rangle$ of the height profile ($\langle h(x,y)
362     \rangle$). Rippled samples show low-wavelength peaks that are
363     outliers on the Landau free energy fits. Samples exhibiting only
364     thermal undulations fit Eq. \ref{eq:fit} remarkably well.}
365     \includegraphics[width=\linewidth]{picture/fit.eps}
366     \label{fig:fit}
367 xsun 2138 \end{figure}
368 gezelter 2143
369 gezelter 3006 \subsection{Effects of Parameters on Ripple Amplitude and Wavelength}
370 gezelter 2143
371 gezelter 3006 We have used two different methods to estimate the amplitude and
372     periodicity of the ripples. The first method requires projection of
373     the ripples onto a one dimensional rippling axis. Since the rippling
374     is always perpendicular to the dipole director axis, we can define a
375     ripple vector as follows. The largest eigenvector, $s_1$, of the
376     $\mathsf{S}$ matrix in Eq. \ref{eq:opmatrix} is projected onto a
377     planar director axis,
378     \begin{equation}
379     \vec{d} = \left(\begin{array}{c}
380     \vec{s}_1 \cdot \hat{i} \\
381     \vec{s}_1 \cdot \hat{j} \\
382     0
383     \end{array} \right).
384     \end{equation}
385     ($\hat{i}$, $\hat{j}$ and $\hat{k}$ are unit vectors along the $x$,
386     $y$, and $z$ axes, respectively.) The rippling axis is in the plane of
387     the membrane and is perpendicular to the planar director axis,
388     \begin{equation}
389     \vec{q}_{\mathrm{rip}} = \vec{d} \times \hat{k}
390     \end{equation}
391     We can then find the height profile of the membrane along the ripple
392     axis by projecting heights of the dipoles to obtain a one-dimensional
393     height profile, $h(q_{\mathrm{rip}})$. Ripple wavelengths can be
394     estimated from the largest non-thermal low-frequency component in the
395     fourier transform of $h(q_{\mathrm{rip}})$. Amplitudes can be
396     estimated by measuring peak-to-trough distances in
397     $h(q_{\mathrm{rip}})$ itself.
398    
399     A second, more accurate, and simpler method for estimating ripple
400     shape is to extract the wavelength and height information directly
401     from the largest non-thermal peak in the undulation spectrum. For
402     large-amplitude ripples, the two methods give similar results. The
403     one-dimensional projection method is more prone to noise (particularly
404     in the amplitude estimates for the non-hexagonal lattices). We report
405     amplitudes and wavelengths taken directly from the undulation spectrum
406     below.
407    
408     In the hexagonal lattice ($\gamma = \sqrt{3}$), the rippling is
409     observed from $150-300$ K. The wavelength of the ripples is
410     remarkably stable at 150~\AA~for all but the temperatures closest to
411     the order-disorder transition. At 300 K, the wavelength drops to 120
412     \AA.
413    
414     The dependence of the amplitude on temperature is shown in
415     Fig. \ref{fig:t-a}. The rippled structures shrink smoothly as the
416     temperature rises towards the order-disorder transition. The
417     wavelengths and amplitudes we observe are surprisingly close to the
418     $\Lambda / 2$ phase observed by Kaasgaard {\it et al.} in their work
419     on PC-based lipids,\cite{Kaasgaard03} although this may be
420     coincidental agreement given our choice of parameters.
421    
422     \begin{figure}[ht]
423 gezelter 2143 \centering
424 gezelter 3006 \caption{ The amplitude $A$ of the ripples vs. temperature for a
425     hexagonal lattice.}
426     \includegraphics[width=\linewidth]{picture/t-a-error.eps}
427     \label{fig:t-a}
428 xsun 2138 \end{figure}
429 gezelter 2143
430 gezelter 3006 The ripples can be made to disappear by increasing the internal
431     surface tension (i.e. by increasing $k_r$). In Fig. \ref{fig:kr-a}
432     we show the ripple amplitude as a function of the internal spring
433     constant for non-dipolar part of the lipid interaction potential.
434     Weaker ``hydrophobic'' interactions allow the lipid structure to be
435     dominated by the dipoles, and stronger ``hydrophobic'' interactions
436     result in much flatter membranes. Section \ref{sec:discussion}
437     contains further discussion of this effect.
438 xsun 2138
439 gezelter 3006 \begin{figure}[ht]
440     \centering
441     \caption{The amplitude $A$ of the ripples vs. the harmonic binding
442     constant $k_r$ for both the hexagonal lattice (circles) and
443     non-hexagonal lattice (squares). In both simulations the dipole
444     strength ($\mu$) was 7 Debye and the temperature was 260K.}
445     \includegraphics[width=\linewidth]{picture/k-a-error.eps}
446     \label{fig:kr-a}
447     \end{figure}
448 xsun 2138
449 gezelter 3006 The amplitude of the ripples depends critically on the strength of the
450     dipole moments ($\mu$) in Eq. \ref{eq:vdd}. If the dipoles are
451     weakened substantially (below $\mu$ = 5 Debye) at a fixed temperature
452     of 230 K, the membrane loses dipolar ordering and the ripple
453     structures. The ripples reach a peak amplitude
454     of 26~\AA~at a dipolar strength of 9 Debye. We show the dependence of
455     ripple amplitude on the dipolar strength in Fig. \ref{fig:s-a}.
456    
457     \begin{figure}[ht]
458     \centering
459     \caption{The amplitude $A$ of the ripples vs. dipole strength ($\mu$)
460     for both the hexagonal lattice (circles) and non-hexagonal lattice
461     (squares). In both simulations the dipole
462     strength ($k_r$) was kept constant at a value of $1.987 \times
463     10^{-4}$ kcal mol$^{-1}$ \AA$^{-2}$. The temperatures were also kept
464     fixed at 230K for the hexagonal lattice and 260K for the non-hexagonal
465     lattice (approximately 2/3 of the order-disorder transition
466     temperature for each lattice).}
467     \includegraphics[width=\linewidth]{picture/A-s.eps}
468     \label{fig:s-a}
469     \end{figure}
470    
471     \subsection{Non-hexagonal lattices}
472    
473     We have also investigated the effect of the lattice geometry by
474     changing the ratio of lattice constants ($\gamma$) while keeping the
475     average nearest-neighbor spacing constant. The antiferroelectric state
476     is accessible for all $\gamma$ values we have used, although the
477     distorted hexagonal lattices prefer a particular director axis due to
478     the anisotropy of the lattice.
479    
480     Our observation of rippling behavior was not limited to the hexagonal
481     lattices. In non-hexagonal lattices the antiferroelectric phase can
482     develop nearly instantaneously in the Monte Carlo simulations, and
483     these dipolar-ordered phases tend to be remarkably flat. Whenever
484     rippling has been observed in these non-hexagonal lattices
485     (e.g. $\gamma = 1.875$), we see relatively short ripple wavelengths
486     (98 \AA) and amplitudes of 17 \AA. These ripples are weakly dependent
487     on dipolar strength (see Fig. \ref{fig:s-a}), although below a dipolar
488     strength of 5.5 Debye, the membrane loses dipolar ordering and
489     displays only thermal undulations.
490    
491     The rippling in non-hexagonal lattices also shows a strong dependence
492     on the internal surface tension ($k_r$). It is possible to make the
493     ripples disappear by increasing the internal tension. The low-tension
494     limit appears to result in somewhat smaller ripples than in the
495     hexagonal lattice (see Fig. \ref{fig:kr-a}).
496    
497     The ripple phase does {\em not} appear at all values of $\gamma$. We
498     have only observed non-thermal undulations in the range $1.625 <
499     \gamma < 1.875$. Outside this range, the order-disorder transition in
500     the dipoles remains, but the ordered dipolar phase has only thermal
501     undulations. This is one of our strongest pieces of evidence that
502     rippling is a symmetry-breaking phenomenon for hexagonal and
503     nearly-hexagonal lattices.
504    
505     \subsection{Effects of System Size}
506     To evaluate the effect of finite system size, we have performed a
507     series of simulations on the hexagonal lattice at a temperature of 300K,
508     which is just below the order-disorder transition temperature (340K).
509     These conditions are in the dipole-ordered and rippled portion of the phase
510     diagram. These are also the conditions that should be most susceptible to
511     system size effects. The wavelength and amplitude of the observed
512     ripples as a function of system size are shown in Fig. \ref{fig:systemsize}.
513    
514     \begin{figure}[ht]
515     \centering
516     \caption{The ripple wavelength (top) and amplitude (bottom) as a function of
517     system size for a hexagonal lattice ($\gamma=1.732$) at 300K.}
518     \includegraphics[width=\linewidth]{picture/SystemSize.eps}
519     \label{fig:systemsize}
520     \end{figure}
521    
522     There is substantial dependence on system size for small (less than 200 \AA)
523     periodic boxes. Notably, there are resonances apparent in the ripple
524     amplitudes at box lengths of 121 \AA and 206 \AA. For larger systems,
525     the behavior of the ripples appears to have stabilized and is on a trend to
526     slightly smaller amplitudes (and slightly longer wavelenghts) than were
527     observed from the 240 \AA box sizes that were used for most of the calculations.
528    
529     It is interesting to note that system sizes which are multiples of the
530     default ripple wavelength can enhance the amplitude of the observed ripples,
531     but appears to have only a minor effect on the observed wavelength. It would,
532     of course, be better to use system sizes that were many multiples of the ripple
533     wavelength to be sure that the periodic box is not driving the phenomenon, but at
534     the largest system size studied (485 \AA $\times$ 485 \AA), the number of
535     molecules (5440) made long Monte Carlo simulations prohibitively expensive.
536     We recognize this as a possible flaw of our model for bilayer rippling, but
537     it is a flaw that will plague any molecular-scale computational model for
538     this phenomenon.
539    
540     \section{Discussion}
541     \label{sec:discussion}
542    
543     We have been able to show that a simple lattice model for membranes
544     which contains only molecular packing (from the lattice), head-group
545     anisotropy (in the form of electrostatic dipoles) and ``hydrophobic''
546     interactions (in the form of a nearest-neighbor harmonic potential) is
547     capable of exhibiting stable long-wavelength non-thermal ripple
548     structures. The best explanation for this behavior is that the
549     ability of the molecules to translate out of the plane of the membrane
550     is enough to break the symmetry of the hexagonal lattice and allow the
551     enormous energetic benefit from the formation of a bulk
552     antiferroelectric phase. Were the hydrophobic interactions absent
553     from our model, it would be possible for the entire lattice to
554     ``tilt'' using $z$-translation. Tilting the lattice in this way would
555     yield an effectively non-hexagonal lattice which would avoid dipolar
556     frustration altogether. With the hydrophobic interactions, bulk tilt
557     would cause a large strain, and the simplest way to release this
558     strain is along line defects. Line defects will result in rippled or
559     sawtooth patterns in the membrane, and allow small ``stripes'' of
560     membrane to form antiferroelectric regions that are tilted relative to
561     the averaged membrane normal.
562    
563     Although the dipole-dipole interaction is the major driving force for
564     the long range orientational ordered state, the formation of the
565     stable, smooth ripples is a result of the competition between the
566     hydrophobic and dipole-dipole interactions. This statement is
567     supported by the variations in both $\mu$ and $k_r$. Substantially
568     weaker dipoles or stronger hydrophobic forces can both cause the
569     ripple phase to disappear.
570    
571     Molecular packing also plays a role in the formation of the ripple
572     phase. It would be surprising if strongly anisotropic head groups
573     would be able to pack in hexagonal lattices without the underlying
574     steric interactions between the rest of the molecular bodies. Since
575     we only see rippled phases in the neighborhood of $\gamma=\sqrt{3}$,
576     this implies that there is a role played by the lipid chains in the
577     organization of the hexagonally ordered phases which support ripples.
578    
579     Our simple model would clearly be a closer approximation to reality if
580     we allowed greater translational freedom to the dipoles and replaced
581     the somewhat artificial lattice packing and the harmonic mimic of the
582     hydrophobic interaction with more realistic molecular modelling
583     potentials. What we have done is to present an extremely simple model
584     which exhibits bulk non-thermal rippling, and our explanation of the
585     rippling phenomenon will help us design more accurate molecular models
586     for the rippling phenomenon.
587    
588     \clearpage
589    
590     \bibliography{ripple}
591    
592     \clearpage
593    
594 xsun 2138 \end{document}