ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/xDissertation/mc.tex
Revision: 3360
Committed: Wed Mar 5 23:34:04 2008 UTC (16 years, 4 months ago) by xsun
Content type: application/x-tex
File size: 32124 byte(s)
Log Message:
writing up the dissertation.

File Contents

# User Rev Content
1 xsun 3360 \chapter{\label{chap:mc}SPONTANEOUS CORRUGATION OF DIPOLAR MEMBRANES}
2 xsun 3354
3     \section{Introduction}
4     \label{mc:sec:Int}
5    
6     The properties of polymeric membranes are known to depend sensitively
7     on the details of the internal interactions between the constituent
8     monomers. A flexible membrane will always have a competition between
9     the energy of curvature and the in-plane stretching energy and will be
10     able to buckle in certain limits of surface tension and
11     temperature.\cite{Safran94} The buckling can be non-specific and
12     centered at dislocation~\cite{Seung1988} or grain-boundary
13     defects,\cite{Carraro1993} or it can be directional and cause long
14     ``roof-tile'' or tube-like structures to appear in
15     partially-polymerized phospholipid vesicles.\cite{Mutz1991}
16    
17     One would expect that anisotropic local interactions could lead to
18     interesting properties of the buckled membrane. We report here on the
19     buckling behavior of a membrane composed of harmonically-bound, but
20     freely-rotating electrostatic dipoles. The dipoles have strongly
21     anisotropic local interactions and the membrane exhibits coupling
22     between the buckling and the long-range ordering of the dipoles.
23    
24     Buckling behavior in liquid crystalline and biological membranes is a
25     well-known phenomenon. Relatively pure phosphatidylcholine (PC)
26     bilayers form a corrugated or ``rippled'' phase ($P_{\beta'}$) which
27     appears as an intermediate phase between the gel ($L_\beta$) and fluid
28     ($L_{\alpha}$) phases. The $P_{\beta'}$ phase has attracted
29     substantial experimental interest over the past 30 years. Most
30     structural information of the ripple phase has been obtained by the
31     X-ray diffraction~\cite{Sun96,Katsaras00} and freeze-fracture electron
32     microscopy (FFEM).~\cite{Copeland80,Meyer96} Recently, Kaasgaard {\it
33     et al.} used atomic force microscopy (AFM) to observe ripple phase
34     morphology in bilayers supported on mica.~\cite{Kaasgaard03} The
35     experimental results provide strong support for a 2-dimensional
36     triangular packing lattice of the lipid molecules within the ripple
37     phase. This is a notable change from the observed lipid packing
38     within the gel phase.~\cite{Cevc87} There have been a number of
39     theoretical
40     approaches~\cite{Marder84,Goldstein88,McCullough90,Lubensky93,Misbah98,Heimburg00,Kubica02,Bannerjee02}
41     (and some heroic
42     simulations~\cite{Ayton02,Jiang04,Brannigan04a,deVries05,deJoannis06})
43     undertaken to try to explain this phase, but to date, none have looked
44     specifically at the contribution of the dipolar character of the lipid
45     head groups towards this corrugation. Lipid chain interdigitation
46     certainly plays a major role, and the structures of the ripple phase
47     are highly ordered. The model we investigate here lacks chain
48     interdigitation (as well as the chains themselves!) and will not be
49     detailed enough to rule in favor of (or against) any of these
50     explanations for the $P_{\beta'}$ phase.
51    
52     Membranes containing electrostatic dipoles can also exhibit the
53     flexoelectric effect,\cite{Todorova2004,Harden2006,Petrov2006} which
54     is the ability of mechanical deformations to result in electrostatic
55     organization of the membrane. This phenomenon is a curvature-induced
56     membrane polarization which can lead to potential differences across a
57     membrane. Reverse flexoelectric behavior (in which applied currents
58     effect membrane curvature) has also been observed. Explanations of
59     the details of these effects have typically utilized membrane
60     polarization perpendicular to the face of the
61     membrane,\cite{Petrov2006} and the effect has been observed in both
62     biological,\cite{Raphael2000} bent-core liquid
63     crystalline,\cite{Harden2006} and polymer-dispersed liquid crystalline
64     membranes.\cite{Todorova2004}
65    
66     The problem with using atomistic and even coarse-grained approaches to
67     study membrane buckling phenomena is that only a relatively small
68     number of periods of the corrugation (i.e. one or two) can be
69     realistically simulated given current technology. Also, simulations
70     of lipid bilayers are traditionally carried out with periodic boundary
71     conditions in two or three dimensions and these have the potential to
72     enhance the periodicity of the system at that wavelength. To avoid
73     this pitfall, we are using a model which allows us to have
74     sufficiently large systems so that we are not causing artificial
75     corrugation through the use of periodic boundary conditions.
76    
77     The simplest dipolar membrane is one in which the dipoles are located
78     on fixed lattice sites. Ferroelectric states (with long-range dipolar
79     order) can be observed in dipolar systems with non-triangular
80     packings. However, {\em triangularly}-packed 2-D dipolar systems are
81     inherently frustrated and one would expect a dipolar-disordered phase
82     to be the lowest free energy
83     configuration.\cite{Toulouse1977,Marland1979} Dipolar lattices already
84     have rich phase behavior, but in order to allow the membrane to
85     buckle, a single degree of freedom (translation normal to the membrane
86     face) must be added to each of the dipoles. It would also be possible
87     to allow complete translational freedom. This approach
88     is similar in character to a number of elastic Ising models that have
89     been developed to explain interesting mechanical properties in
90     magnetic alloys.\cite{Renard1966,Zhu2005,Zhu2006,Jiang2006}
91    
92     What we present here is an attempt to find the simplest dipolar model
93     which will exhibit buckling behavior. We are using a modified XYZ
94     lattice model; details of the model can be found in section
95     \ref{mc:sec:model}, results of Monte Carlo simulations using this model
96     are presented in section
97     \ref{mc:sec:results}, and section \ref{mc:sec:discussion} contains our conclusions.
98    
99     \section{2-D Dipolar Membrane}
100     \label{mc:sec:model}
101    
102     The point of developing this model was to arrive at the simplest
103     possible theoretical model which could exhibit spontaneous corrugation
104     of a two-dimensional dipolar medium. Since molecules in polymerized
105     membranes and in the $P_{\beta'}$ ripple phase have limited
106     translational freedom, we have chosen a lattice to support the dipoles
107     in the x-y plane. The lattice may be either triangular (lattice
108     constants $a/b =
109     \sqrt{3}$) or distorted. However, each dipole has 3 degrees of
110     freedom. They may move freely {\em out} of the x-y plane (along the
111     $z$ axis), and they have complete orientational freedom ($0 <= \theta
112     <= \pi$, $0 <= \phi < 2
113     \pi$). This is essentially a modified X-Y-Z model with translational
114     freedom along the z-axis.
115    
116     The potential energy of the system,
117     \begin{eqnarray}
118     V = \sum_i & & \left( \sum_{j>i} \frac{|\mu|^2}{4\pi \epsilon_0 r_{ij}^3} \left[
119     {\mathbf{\hat u}_i} \cdot {\mathbf{\hat u}_j} -
120     3({\mathbf{\hat u}_i} \cdot {\mathbf{\hat
121     r}_{ij}})({\mathbf{\hat u}_j} \cdot {\mathbf{\hat r}_{ij}})\right]
122     \right. \nonumber \\
123     & & \left. + \sum_{j \in NN_i}^6 \frac{k_r}{2}\left(
124     r_{ij}-\sigma \right)^2 \right)
125     \label{mceq:pot}
126     \end{eqnarray}
127    
128     In this potential, $\mathbf{\hat u}_i$ is the unit vector pointing
129     along dipole $i$ and $\mathbf{\hat r}_{ij}$ is the unit vector
130     pointing along the inter-dipole vector $\mathbf{r}_{ij}$. The entire
131     potential is governed by three parameters, the dipolar strength
132     ($\mu$), the harmonic spring constant ($k_r$) and the preferred
133     intermolecular spacing ($\sigma$). In practice, we set the value of
134     $\sigma$ to the average inter-molecular spacing from the planar
135     lattice, yielding a potential model that has only two parameters for a
136     particular choice of lattice constants $a$ (along the $x$-axis) and
137     $b$ (along the $y$-axis). We also define a set of reduced parameters
138     based on the length scale ($\sigma$) and the energy of the harmonic
139     potential at a deformation of 2 $\sigma$ ($\epsilon = k_r \sigma^2 /
140     2$). Using these two constants, we perform our calculations using
141     reduced distances, ($r^{*} = r / \sigma$), temperatures ($T^{*} = 2
142     k_B T / k_r \sigma^2$), densities ($\rho^{*} = N \sigma^2 / L_x L_y$),
143     and dipole moments ($\mu^{*} = \mu / \sqrt{4 \pi \epsilon_0 \sigma^5
144     k_r / 2}$). It should be noted that the density ($\rho^{*}$) depends
145     only on the mean particle spacing in the $x-y$ plane; the lattice is
146     fully populated.
147    
148     To investigate the phase behavior of this model, we have performed a
149     series of Metropolis Monte Carlo simulations of moderately-sized (34.3
150     $\sigma$ on a side) patches of membrane hosted on both triangular
151     ($\gamma = a/b = \sqrt{3}$) and distorted ($\gamma \neq \sqrt{3}$)
152     lattices. The linear extent of one edge of the monolayer was $20 a$
153     and the system was kept roughly square. The average distance that
154     coplanar dipoles were positioned from their six nearest neighbors was
155     1 $\sigma$ (on both triangular and distorted lattices). Typical
156     system sizes were 1360 dipoles for the triangular lattices and
157     840-2800 dipoles for the distorted lattices. Two-dimensional periodic
158     boundary conditions were used, and the cutoff for the dipole-dipole
159     interaction was set to 4.3 $\sigma$. This cutoff is roughly 2.5 times
160     the typical real-space electrostatic cutoff for molecular systems.
161     Since dipole-dipole interactions decay rapidly with distance, and
162     since the intrinsic three-dimensional periodicity of the Ewald sum can
163     give artifacts in 2-d systems, we have chosen not to use it in these
164     calculations. Although the Ewald sum has been reformulated to handle
165     2-D systems,\cite{Parry75,Parry76,Heyes77,deLeeuw79,Rhee89} these
166     methods are computationally expensive,\cite{Spohr97,Yeh99} and are not
167     necessary in this case. All parameters ($T^{*}$, $\mu^{*}$, and
168     $\gamma$) were varied systematically to study the effects of these
169     parameters on the formation of ripple-like phases.
170    
171     \section{Results and Analysis}
172     \label{mc:sec:results}
173    
174     \subsection{Dipolar Ordering and Coexistence Temperatures}
175     The principal method for observing the orientational ordering
176     transition in dipolar systems is the $P_2$ order parameter (defined as
177     $1.5 \times \lambda_{max}$, where $\lambda_{max}$ is the largest
178     eigenvalue of the matrix,
179     \begin{equation}
180     {\mathsf{S}} = \frac{1}{N} \sum_i \left(
181     \begin{array}{ccc}
182     u^{x}_i u^{x}_i-\frac{1}{3} & u^{x}_i u^{y}_i & u^{x}_i u^{z}_i \\
183     u^{y}_i u^{x}_i & u^{y}_i u^{y}_i -\frac{1}{3} & u^{y}_i u^{z}_i \\
184     u^{z}_i u^{x}_i & u^{z}_i u^{y}_i & u^{z}_i u^{z}_i -\frac{1}{3}
185     \end{array} \right).
186     \label{mceq:opmatrix}
187     \end{equation}
188     Here $u^{\alpha}_i$ is the $\alpha=x,y,z$ component of the unit vector
189     for dipole $i$. $P_2$ will be $1.0$ for a perfectly-ordered system
190     and near $0$ for a randomized system. Note that this order parameter
191     is {\em not} equal to the polarization of the system. For example,
192     the polarization of the perfect anti-ferroelectric system is $0$, but
193     $P_2$ for an anti-ferroelectric system is $1$. The eigenvector of
194     $\mathsf{S}$ corresponding to the largest eigenvalue is familiar as
195     the director axis, which can be used to determine a privileged dipolar
196     axis for dipole-ordered systems. The top panel in Fig. \ref{mcfig:phase}
197     shows the values of $P_2$ as a function of temperature for both
198     triangular ($\gamma = 1.732$) and distorted ($\gamma=1.875$)
199     lattices.
200    
201     \begin{figure}
202     \includegraphics[width=\linewidth]{./figures/mcPhase.pdf}
203     \caption{\label{mcfig:phase} Top panel: The $P_2$ dipolar order parameter as
204     a function of temperature for both triangular ($\gamma = 1.732$) and
205     distorted ($\gamma = 1.875$) lattices. Bottom Panel: The phase
206     diagram for the dipolar membrane model. The line denotes the division
207     between the dipolar ordered (anti-ferroelectric) and disordered phases.
208     An enlarged view near the triangular lattice is shown inset.}
209     \end{figure}
210    
211     There is a clear order-disorder transition in evidence from this data.
212     Both the triangular and distorted lattices have dipolar-ordered
213     low-temperature phases, and orientationally-disordered high
214     temperature phases. The coexistence temperature for the triangular
215     lattice is significantly lower than for the distorted lattices, and
216     the bulk polarization is approximately $0$ for both dipolar ordered
217     and disordered phases. This gives strong evidence that the dipolar
218     ordered phase is anti-ferroelectric. We have verified that this
219     dipolar ordering transition is not a function of system size by
220     performing identical calculations with systems twice as large. The
221     transition is equally smooth at all system sizes that were studied.
222     Additionally, we have repeated the Monte Carlo simulations over a wide
223     range of lattice ratios ($\gamma$) to generate a dipolar
224     order/disorder phase diagram. The bottom panel in Fig. \ref{mcfig:phase}
225     shows that the triangular lattice is a low-temperature cusp in the
226     $T^{*}-\gamma$ phase diagram.
227    
228     This phase diagram is remarkable in that it shows an
229     anti-ferroelectric phase near $\gamma=1.732$ where one would expect
230     lattice frustration to result in disordered phases at all
231     temperatures. Observations of the configurations in this phase show
232     clearly that the system has accomplished dipolar ordering by forming
233     large ripple-like structures. We have observed anti-ferroelectric
234     ordering in all three of the equivalent directions on the triangular
235     lattice, and the dipoles have been observed to organize perpendicular
236     to the membrane normal (in the plane of the membrane). It is
237     particularly interesting to note that the ripple-like structures have
238     also been observed to propagate in the three equivalent directions on
239     the lattice, but the {\em direction of ripple propagation is always
240     perpendicular to the dipole director axis}. A snapshot of a typical
241     anti-ferroelectric rippled structure is shown in
242     Fig. \ref{mcfig:snapshot}.
243    
244     \begin{figure}
245     \includegraphics[width=\linewidth]{./figures/mcSnapshot.pdf}
246     \caption{\label{mcfig:snapshot} Top and Side views of a representative
247     configuration for the dipolar ordered phase supported on the
248     triangular lattice. Note the anti-ferroelectric ordering and the long
249     wavelength buckling of the membrane. Dipolar ordering has been
250     observed in all three equivalent directions on the triangular lattice,
251     and the ripple direction is always perpendicular to the director axis
252     for the dipoles.}
253     \end{figure}
254    
255     Although the snapshot in Fig. \ref{mcfig:snapshot} gives the appearance
256     of three-row stair-like structures, these appear to be transient. On
257     average, the corrugation of the membrane is a relatively smooth,
258     long-wavelength phenomenon, with occasional steep drops between
259     adjacent lines of anti-aligned dipoles.
260    
261     The height-dipole correlation function ($C_{\textrm{hd}}(r, \cos
262     \theta)$) makes the connection between dipolar ordering and the wave
263     vector of the ripple even more explicit. $C_{\textrm{hd}}(r, \cos
264     \theta)$ is an angle-dependent pair distribution function. The angle
265     ($\theta$) is the angle between the intermolecular vector
266     $\vec{r}_{ij}$ and direction of dipole $i$,
267     \begin{equation}
268     C_{\textrm{hd}} = \frac{\langle \frac{1}{n(r)} \sum_{i}\sum_{j>i}
269     h_i \cdot h_j \delta(r - r_{ij}) \delta(\cos \theta_{ij} -
270     \cos \theta)\rangle} {\langle h^2 \rangle}
271     \end{equation}
272     where $\cos \theta_{ij} = \hat{\mu}_{i} \cdot \hat{r}_{ij}$ and
273     $\hat{r}_{ij} = \vec{r}_{ij} / r_{ij}$. $n(r)$ is the number of
274     dipoles found in a cylindrical shell between $r$ and $r+\delta r$ of
275     the central particle. Fig. \ref{mcfig:CrossCorrelation} shows contours
276     of this correlation function for both anti-ferroelectric, rippled
277     membranes as well as for the dipole-disordered portion of the phase
278     diagram.
279    
280     \begin{figure}
281     \includegraphics[width=\linewidth]{./figures/mcHdc.pdf}
282     \caption{\label{mcfig:CrossCorrelation} Contours of the height-dipole
283     correlation function as a function of the dot product between the
284     dipole ($\hat{\mu}$) and inter-dipole separation vector ($\hat{r}$)
285     and the distance ($r$) between the dipoles. Perfect height
286     correlation (contours approaching 1) are present in the ordered phase
287     when the two dipoles are in the same head-to-tail line.
288     Anti-correlation (contours below 0) is only seen when the inter-dipole
289     vector is perpendicular to the dipoles. In the dipole-disordered
290     portion of the phase diagram, there is only weak correlation in the
291     dipole direction and this correlation decays rapidly to zero for
292     intermolecular vectors that are not dipole-aligned.}
293     \end{figure}
294    
295     The height-dipole correlation function gives a map of how the topology
296     of the membrane surface varies with angular deviation around a given
297     dipole. The upper panel of Fig. \ref{mcfig:CrossCorrelation} shows that
298     in the anti-ferroelectric phase, the dipole heights are strongly
299     correlated for dipoles in head-to-tail arrangements, and this
300     correlation persists for very long distances (up to 15 $\sigma$). For
301     portions of the membrane located perpendicular to a given dipole, the
302     membrane height becomes anti-correlated at distances of 10 $\sigma$.
303     The correlation function is relatively smooth; there are no steep
304     jumps or steps, so the stair-like structures in
305     Fig. \ref{mcfig:snapshot} are indeed transient and disappear when
306     averaged over many configurations. In the dipole-disordered phase,
307     the height-dipole correlation function is relatively flat (and hovers
308     near zero). The only significant height correlations are for axial
309     dipoles at very short distances ($r \approx
310     \sigma$).
311    
312     \subsection{Discriminating Ripples from Thermal Undulations}
313    
314     In order to be sure that the structures we have observed are actually
315     a rippled phase and not simply thermal undulations, we have computed
316     the undulation spectrum,
317     \begin{equation}
318     h(\vec{q}) = A^{-1/2} \int d\vec{r}
319     h(\vec{r}) e^{-i \vec{q}\cdot\vec{r}}
320     \end{equation}
321     where $h(\vec{r})$ is the height of the membrane at location $\vec{r}
322     = (x,y)$.~\cite{Safran94,Seifert97} In simple (and more complicated)
323     elastic continuum models, it can shown that in the $NVT$ ensemble, the
324     absolute value of the undulation spectrum can be written,
325     \begin{equation}
326     \langle | h(q) |^2 \rangle_{NVT} = \frac{k_B T}{k_c q^4 +
327     \gamma q^2},
328     \label{mceq:fit}
329     \end{equation}
330     where $k_c$ is the bending modulus for the membrane, and $\gamma$ is
331     the mechanical surface tension.~\cite{Safran94} The systems studied in
332     this paper have essentially zero bending moduli ($k_c$) and relatively
333     large mechanical surface tensions ($\gamma$), so a much simpler form
334     can be written,
335     \begin{equation}
336     \langle | h(q) |^2 \rangle_{NVT} = \frac{k_B T}{\gamma q^2},
337     \label{mceq:fit2}
338     \end{equation}
339    
340     The undulation spectrum is computed by superimposing a rectangular
341     grid on top of the membrane, and by assigning height ($h(\vec{r})$)
342     values to the grid from the average of all dipoles that fall within a
343     given $\vec{r}+d\vec{r}$ grid area. Empty grid pixels are assigned
344     height values by interpolation from the nearest neighbor pixels. A
345     standard 2-d Fourier transform is then used to obtain $\langle |
346     h(q)|^2 \rangle$. Alternatively, since the dipoles sit on a Bravais
347     lattice, one could use the heights of the lattice points themselves as
348     the grid for the Fourier transform (without interpolating to a square
349     grid). However, if lateral translational freedom is added to this
350     model (a likely extension), an interpolated grid method for computing
351     undulation spectra will be required.
352    
353     As mentioned above, the best fits to our undulation spectra are
354     obtained by setting the value of $k_c$ to 0. In Fig. \ref{mcfig:fit} we
355     show typical undulation spectra for two different regions of the phase
356     diagram along with their fits from the Landau free energy approach
357     (Eq. \ref{mceq:fit2}). In the high-temperature disordered phase, the
358     Landau fits can be nearly perfect, and from these fits we can estimate
359     the tension in the surface. In reduced units, typical values of
360     $\gamma^{*} = \gamma / \epsilon = 2500$ are obtained for the
361     disordered phase ($\gamma^{*} = 2551.7$ in the top panel of
362     Fig. \ref{mcfig:fit}).
363    
364     Typical values of $\gamma^{*}$ in the dipolar-ordered phase are much
365     higher than in the dipolar-disordered phase ($\gamma^{*} = 73,538$ in
366     the lower panel of Fig. \ref{mcfig:fit}). For the dipolar-ordered
367     triangular lattice near the coexistence temperature, we also observe
368     long wavelength undulations that are far outliers to the fits. That
369     is, the Landau free energy fits are well within error bars for most of
370     the other points, but can be off by {\em orders of magnitude} for a
371     few low frequency components.
372    
373     We interpret these outliers as evidence that these low frequency modes
374     are {\em non-thermal undulations}. We take this as evidence that we
375     are actually seeing a rippled phase developing in this model system.
376    
377     \begin{figure}
378     \includegraphics[width=\linewidth]{./figures/mcLogFit.pdf}
379     \caption{\label{mcfig:fit} Evidence that the observed ripples are {\em
380     not} thermal undulations is obtained from the 2-d Fourier transform
381     $\langle |h^{*}(\vec{q})|^2 \rangle$ of the height profile ($\langle
382     h^{*}(x,y) \rangle$). Rippled samples show low-wavelength peaks that
383     are outliers on the Landau free energy fits by an order of magnitude.
384     Samples exhibiting only thermal undulations fit Eq. \ref{mceq:fit}
385     remarkably well.}
386     \end{figure}
387    
388     \subsection{Effects of Potential Parameters on Amplitude and Wavelength}
389    
390     We have used two different methods to estimate the amplitude and
391     periodicity of the ripples. The first method requires projection of
392     the ripples onto a one dimensional rippling axis. Since the rippling
393     is always perpendicular to the dipole director axis, we can define a
394     ripple vector as follows. The largest eigenvector, $s_1$, of the
395     $\mathsf{S}$ matrix in Eq. \ref{mceq:opmatrix} is projected onto a
396     planar director axis,
397     \begin{equation}
398     \vec{d} = \left(\begin{array}{c}
399     \vec{s}_1 \cdot \hat{i} \\
400     \vec{s}_1 \cdot \hat{j} \\
401     0
402     \end{array} \right).
403     \end{equation}
404     ($\hat{i}$, $\hat{j}$ and $\hat{k}$ are unit vectors along the $x$,
405     $y$, and $z$ axes, respectively.) The rippling axis is in the plane of
406     the membrane and is perpendicular to the planar director axis,
407     \begin{equation}
408     \vec{q}_{\mathrm{rip}} = \vec{d} \times \hat{k}
409     \end{equation}
410     We can then find the height profile of the membrane along the ripple
411     axis by projecting heights of the dipoles to obtain a one-dimensional
412     height profile, $h(q_{\mathrm{rip}})$. Ripple wavelengths can be
413     estimated from the largest non-thermal low-frequency component in the
414     Fourier transform of $h(q_{\mathrm{rip}})$. Amplitudes can be
415     estimated by measuring peak-to-trough distances in
416     $h(q_{\mathrm{rip}})$ itself.
417    
418     A second, more accurate, and simpler method for estimating ripple
419     shape is to extract the wavelength and height information directly
420     from the largest non-thermal peak in the undulation spectrum. For
421     large-amplitude ripples, the two methods give similar results. The
422     one-dimensional projection method is more prone to noise (particularly
423     in the amplitude estimates for the distorted lattices). We report
424     amplitudes and wavelengths taken directly from the undulation spectrum
425     below.
426    
427     In the triangular lattice ($\gamma = \sqrt{3}$), the rippling is
428     observed for temperatures ($T^{*}$) from $61-122$. The wavelength of
429     the ripples is remarkably stable at 21.4~$\sigma$ for all but the
430     temperatures closest to the order-disorder transition. At $T^{*} =
431     122$, the wavelength drops to 17.1~$\sigma$.
432    
433     The dependence of the amplitude on temperature is shown in the top
434     panel of Fig. \ref{mcfig:Amplitude}. The rippled structures shrink
435     smoothly as the temperature rises towards the order-disorder
436     transition. The wavelengths and amplitudes we observe are
437     surprisingly close to the $\Lambda / 2$ phase observed by Kaasgaard
438     {\it et al.} in their work on PC-based lipids.\cite{Kaasgaard03}
439     However, this is coincidental agreement based on a choice of 7~\AA~as
440     the mean spacing between lipids.
441    
442     \begin{figure}
443     \includegraphics[width=\linewidth]{./figures/mcProperties_sq.pdf}
444     \caption{\label{mcfig:Amplitude} a) The amplitude $A^{*}$ of the ripples
445     vs. temperature for a triangular lattice. b) The amplitude $A^{*}$ of
446     the ripples vs. dipole strength ($\mu^{*}$) for both the triangular
447     lattice (circles) and distorted lattice (squares). The reduced
448     temperatures were kept fixed at $T^{*} = 94$ for the triangular
449     lattice and $T^{*} = 106$ for the distorted lattice (approximately 2/3
450     of the order-disorder transition temperature for each lattice).}
451     \end{figure}
452    
453     The ripples can be made to disappear by increasing the internal
454     elastic tension (i.e. by increasing $k_r$ or equivalently, reducing
455     the dipole moment). The amplitude of the ripples depends critically
456     on the strength of the dipole moments ($\mu^{*}$) in Eq. \ref{mceq:pot}.
457     If the dipoles are weakened substantially (below $\mu^{*}$ = 20) at a
458     fixed temperature of 94, the membrane loses dipolar ordering
459     and the ripple structures. The ripples reach a peak amplitude of
460     3.7~$\sigma$ at a dipolar strength of 25. We show the dependence
461     of ripple amplitude on the dipolar strength in
462     Fig. \ref{mcfig:Amplitude}.
463    
464     \subsection{Distorted lattices}
465    
466     We have also investigated the effect of the lattice geometry by
467     changing the ratio of lattice constants ($\gamma$) while keeping the
468     average nearest-neighbor spacing constant. The anti-ferroelectric state
469     is accessible for all $\gamma$ values we have used, although the
470     distorted triangular lattices prefer a particular director axis due to
471     the anisotropy of the lattice.
472    
473     Our observation of rippling behavior was not limited to the triangular
474     lattices. In distorted lattices the anti-ferroelectric phase can
475     develop nearly instantaneously in the Monte Carlo simulations, and
476     these dipolar-ordered phases tend to be remarkably flat. Whenever
477     rippling has been observed in these distorted lattices
478     (e.g. $\gamma = 1.875$), we see relatively short ripple wavelengths
479     (14 $\sigma$) and amplitudes of 2.4~$\sigma$. These ripples are
480     weakly dependent on dipolar strength (see Fig. \ref{mcfig:Amplitude}),
481     although below a dipolar strength of $\mu^{*} = 20$, the membrane
482     loses dipolar ordering and displays only thermal undulations.
483    
484     The ripple phase does {\em not} appear at all values of $\gamma$. We
485     have only observed non-thermal undulations in the range $1.625 <
486     \gamma < 1.875$. Outside this range, the order-disorder transition in
487     the dipoles remains, but the ordered dipolar phase has only thermal
488     undulations. This is one of our strongest pieces of evidence that
489     rippling is a symmetry-breaking phenomenon for triangular and
490     nearly-triangular lattices.
491    
492     \subsection{Effects of System Size}
493     To evaluate the effect of finite system size, we have performed a
494     series of simulations on the triangular lattice at a reduced
495     temperature of 122, which is just below the order-disorder transition
496     temperature ($T^{*} = 139$). These conditions are in the
497     dipole-ordered and rippled portion of the phase diagram. These are
498     also the conditions that should be most susceptible to system size
499     effects.
500    
501     \begin{figure}
502     \includegraphics[width=\linewidth]{./figures/mcSystemSize.pdf}
503     \caption{\label{mcfig:systemsize} The ripple wavelength (top) and
504     amplitude (bottom) as a function of system size for a triangular
505     lattice ($\gamma=1.732$) at $T^{*} = 122$.}
506     \end{figure}
507    
508     There is substantial dependence on system size for small (less than
509     29~$\sigma$) periodic boxes. Notably, there are resonances apparent
510     in the ripple amplitudes at box lengths of 17.3 and 29.5 $\sigma$.
511     For larger systems, the behavior of the ripples appears to have
512     stabilized and is on a trend to slightly smaller amplitudes (and
513     slightly longer wavelengths) than were observed from the 34.3 $\sigma$
514     box sizes that were used for most of the calculations.
515    
516     It is interesting to note that system sizes which are multiples of the
517     default ripple wavelength can enhance the amplitude of the observed
518     ripples, but appears to have only a minor effect on the observed
519     wavelength. It would, of course, be better to use system sizes that
520     were many multiples of the ripple wavelength to be sure that the
521     periodic box is not driving the phenomenon, but at the largest system
522     size studied (70 $\sigma$ $\times$ 70 $\sigma$), the number of dipoles
523     (5440) made long Monte Carlo simulations prohibitively expensive.
524    
525     \section{Discussion}
526     \label{mc:sec:discussion}
527    
528     We have been able to show that a simple dipolar lattice model which
529     contains only molecular packing (from the lattice), anisotropy (in the
530     form of electrostatic dipoles) and a weak elastic tension (in the form
531     of a nearest-neighbor harmonic potential) is capable of exhibiting
532     stable long-wavelength non-thermal surface corrugations. The best
533     explanation for this behavior is that the ability of the dipoles to
534     translate out of the plane of the membrane is enough to break the
535     symmetry of the triangular lattice and allow the energetic benefit
536     from the formation of a bulk anti-ferroelectric phase. Were the weak
537     elastic tension absent from our model, it would be possible for the
538     entire lattice to ``tilt'' using $z$-translation. Tilting the lattice
539     in this way would yield an effectively non-triangular lattice which
540     would avoid dipolar frustration altogether. With the elastic tension
541     in place, bulk tilt causes a large strain, and the least costly way to
542     release this strain is between two rows of anti-aligned dipoles.
543     These ``breaks'' will result in rippled or sawtooth patterns in the
544     membrane, and allow small stripes of membrane to form
545     anti-ferroelectric regions that are tilted relative to the averaged
546     membrane normal.
547    
548     Although the dipole-dipole interaction is the major driving force for
549     the long range orientational ordered state, the formation of the
550     stable, smooth ripples is a result of the competition between the
551     elastic tension and the dipole-dipole interactions. This statement is
552     supported by the variation in $\mu^{*}$. Substantially weaker dipoles
553     relative to the surface tension can cause the corrugated phase to
554     disappear.
555    
556     The packing of the dipoles into a nearly-triangular lattice is clearly
557     an important piece of the puzzle. The dipolar head groups of lipid
558     molecules are sterically (as well as electrostatically) anisotropic,
559     and would not pack in triangular arrangements without the steric
560     interference of adjacent molecular bodies. Since we only see rippled
561     phases in the neighborhood of $\gamma=\sqrt{3}$, this implies that
562     even if this dipolar mechanism is the correct explanation for the
563     ripple phase in realistic bilayers, there would still be a role played
564     by the lipid chains in the in-plane organization of the triangularly
565     ordered phases which could support ripples. The present model is
566     certainly not detailed enough to answer exactly what drives the
567     formation of the $P_{\beta'}$ phase in real lipids, but suggests some
568     avenues for further experiments.
569    
570     The most important prediction we can make using the results from this
571     simple model is that if dipolar ordering is driving the surface
572     corrugation, the wave vectors for the ripples should always found to
573     be {\it perpendicular} to the dipole director axis. This prediction
574     should suggest experimental designs which test whether this is really
575     true in the phosphatidylcholine $P_{\beta'}$ phases. The dipole
576     director axis should also be easily computable for the all-atom and
577     coarse-grained simulations that have been published in the literature.
578    
579     Our other observation about the ripple and dipolar directionality is
580     that the dipole director axis can be found to be parallel to any of
581     the three equivalent lattice vectors in the triangular lattice.
582     Defects in the ordering of the dipoles can cause the dipole director
583     (and consequently the surface corrugation) of small regions to be
584     rotated relative to each other by 120$^{\circ}$. This is a similar
585     behavior to the domain rotation seen in the AFM studies of Kaasgaard
586     {\it et al.}\cite{Kaasgaard03}
587    
588     Although our model is simple, it exhibits some rich and unexpected
589     behaviors. It would clearly be a closer approximation to the reality
590     if we allowed greater translational freedom to the dipoles and
591     replaced the somewhat artificial lattice packing and the harmonic
592     elastic tension with more realistic molecular modeling potentials.
593     What we have done is to present a simple model which exhibits bulk
594     non-thermal corrugation, and our explanation of this rippling
595     phenomenon will help us design more accurate molecular models for
596     corrugated membranes and experiments to test whether rippling is
597     dipole-driven or not.