ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/xDissertation/mc.tex
Revision: 3374
Committed: Thu Mar 20 22:21:43 2008 UTC (16 years, 5 months ago) by xsun
Content type: application/x-tex
File size: 31649 byte(s)
Log Message:
writing.

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 xsun 3372 substantial experimental interest over the past 30
30 xsun 3374 years,~\cite{Sun96,Katsaras00,Copeland80,Meyer96,Kaasgaard03} and
31 xsun 3372 there have been a number of theoretical
32 xsun 3361 approaches~\cite{Marder84,Carlson87,Goldstein88,McCullough90,Lubensky93,Misbah98,Heimburg00,Kubica02,Bannerjee02}
33 xsun 3354 (and some heroic
34     simulations~\cite{Ayton02,Jiang04,Brannigan04a,deVries05,deJoannis06})
35     undertaken to try to explain this phase, but to date, none have looked
36     specifically at the contribution of the dipolar character of the lipid
37     head groups towards this corrugation. Lipid chain interdigitation
38     certainly plays a major role, and the structures of the ripple phase
39     are highly ordered. The model we investigate here lacks chain
40     interdigitation (as well as the chains themselves!) and will not be
41     detailed enough to rule in favor of (or against) any of these
42     explanations for the $P_{\beta'}$ phase.
43    
44     Membranes containing electrostatic dipoles can also exhibit the
45     flexoelectric effect,\cite{Todorova2004,Harden2006,Petrov2006} which
46     is the ability of mechanical deformations to result in electrostatic
47     organization of the membrane. This phenomenon is a curvature-induced
48     membrane polarization which can lead to potential differences across a
49     membrane. Reverse flexoelectric behavior (in which applied currents
50     effect membrane curvature) has also been observed. Explanations of
51     the details of these effects have typically utilized membrane
52     polarization perpendicular to the face of the
53     membrane,\cite{Petrov2006} and the effect has been observed in both
54     biological,\cite{Raphael2000} bent-core liquid
55     crystalline,\cite{Harden2006} and polymer-dispersed liquid crystalline
56     membranes.\cite{Todorova2004}
57    
58     The problem with using atomistic and even coarse-grained approaches to
59     study membrane buckling phenomena is that only a relatively small
60     number of periods of the corrugation (i.e. one or two) can be
61     realistically simulated given current technology. Also, simulations
62     of lipid bilayers are traditionally carried out with periodic boundary
63     conditions in two or three dimensions and these have the potential to
64     enhance the periodicity of the system at that wavelength. To avoid
65     this pitfall, we are using a model which allows us to have
66     sufficiently large systems so that we are not causing artificial
67     corrugation through the use of periodic boundary conditions.
68    
69     The simplest dipolar membrane is one in which the dipoles are located
70     on fixed lattice sites. Ferroelectric states (with long-range dipolar
71     order) can be observed in dipolar systems with non-triangular
72     packings. However, {\em triangularly}-packed 2-D dipolar systems are
73     inherently frustrated and one would expect a dipolar-disordered phase
74     to be the lowest free energy
75     configuration.\cite{Toulouse1977,Marland1979} Dipolar lattices already
76     have rich phase behavior, but in order to allow the membrane to
77     buckle, a single degree of freedom (translation normal to the membrane
78     face) must be added to each of the dipoles. It would also be possible
79     to allow complete translational freedom. This approach
80     is similar in character to a number of elastic Ising models that have
81     been developed to explain interesting mechanical properties in
82     magnetic alloys.\cite{Renard1966,Zhu2005,Zhu2006,Jiang2006}
83    
84     What we present here is an attempt to find the simplest dipolar model
85     which will exhibit buckling behavior. We are using a modified XYZ
86     lattice model; details of the model can be found in section
87     \ref{mc:sec:model}, results of Monte Carlo simulations using this model
88     are presented in section
89     \ref{mc:sec:results}, and section \ref{mc:sec:discussion} contains our conclusions.
90    
91     \section{2-D Dipolar Membrane}
92     \label{mc:sec:model}
93    
94     The point of developing this model was to arrive at the simplest
95     possible theoretical model which could exhibit spontaneous corrugation
96     of a two-dimensional dipolar medium. Since molecules in polymerized
97     membranes and in the $P_{\beta'}$ ripple phase have limited
98     translational freedom, we have chosen a lattice to support the dipoles
99     in the x-y plane. The lattice may be either triangular (lattice
100     constants $a/b =
101     \sqrt{3}$) or distorted. However, each dipole has 3 degrees of
102     freedom. They may move freely {\em out} of the x-y plane (along the
103     $z$ axis), and they have complete orientational freedom ($0 <= \theta
104     <= \pi$, $0 <= \phi < 2
105     \pi$). This is essentially a modified X-Y-Z model with translational
106     freedom along the z-axis.
107    
108     The potential energy of the system,
109 xsun 3361 \begin{equation}
110     \begin{split}
111     V = \sum_i &\left( \sum_{j>i} \frac{|\mu|^2}{4\pi \epsilon_0 r_{ij}^3} \left[
112 xsun 3354 {\mathbf{\hat u}_i} \cdot {\mathbf{\hat u}_j} -
113     3({\mathbf{\hat u}_i} \cdot {\mathbf{\hat
114 xsun 3361 r}_{ij}})({\mathbf{\hat u}_j} \cdot {\mathbf{\hat r}_{ij}})\right] \right. \\
115     & \left. + \sum_{j \in NN_i}^6 \frac{k_r}{2}\left(
116 xsun 3354 r_{ij}-\sigma \right)^2 \right)
117 xsun 3361 \end{split}
118 xsun 3354 \label{mceq:pot}
119 xsun 3361 \end{equation}
120 xsun 3354
121     In this potential, $\mathbf{\hat u}_i$ is the unit vector pointing
122     along dipole $i$ and $\mathbf{\hat r}_{ij}$ is the unit vector
123     pointing along the inter-dipole vector $\mathbf{r}_{ij}$. The entire
124     potential is governed by three parameters, the dipolar strength
125     ($\mu$), the harmonic spring constant ($k_r$) and the preferred
126     intermolecular spacing ($\sigma$). In practice, we set the value of
127     $\sigma$ to the average inter-molecular spacing from the planar
128     lattice, yielding a potential model that has only two parameters for a
129     particular choice of lattice constants $a$ (along the $x$-axis) and
130     $b$ (along the $y$-axis). We also define a set of reduced parameters
131     based on the length scale ($\sigma$) and the energy of the harmonic
132     potential at a deformation of 2 $\sigma$ ($\epsilon = k_r \sigma^2 /
133     2$). Using these two constants, we perform our calculations using
134     reduced distances, ($r^{*} = r / \sigma$), temperatures ($T^{*} = 2
135     k_B T / k_r \sigma^2$), densities ($\rho^{*} = N \sigma^2 / L_x L_y$),
136     and dipole moments ($\mu^{*} = \mu / \sqrt{4 \pi \epsilon_0 \sigma^5
137     k_r / 2}$). It should be noted that the density ($\rho^{*}$) depends
138     only on the mean particle spacing in the $x-y$ plane; the lattice is
139     fully populated.
140    
141     To investigate the phase behavior of this model, we have performed a
142     series of Metropolis Monte Carlo simulations of moderately-sized (34.3
143     $\sigma$ on a side) patches of membrane hosted on both triangular
144     ($\gamma = a/b = \sqrt{3}$) and distorted ($\gamma \neq \sqrt{3}$)
145     lattices. The linear extent of one edge of the monolayer was $20 a$
146     and the system was kept roughly square. The average distance that
147     coplanar dipoles were positioned from their six nearest neighbors was
148     1 $\sigma$ (on both triangular and distorted lattices). Typical
149     system sizes were 1360 dipoles for the triangular lattices and
150     840-2800 dipoles for the distorted lattices. Two-dimensional periodic
151     boundary conditions were used, and the cutoff for the dipole-dipole
152     interaction was set to 4.3 $\sigma$. This cutoff is roughly 2.5 times
153     the typical real-space electrostatic cutoff for molecular systems.
154     Since dipole-dipole interactions decay rapidly with distance, and
155     since the intrinsic three-dimensional periodicity of the Ewald sum can
156     give artifacts in 2-d systems, we have chosen not to use it in these
157     calculations. Although the Ewald sum has been reformulated to handle
158     2-D systems,\cite{Parry75,Parry76,Heyes77,deLeeuw79,Rhee89} these
159     methods are computationally expensive,\cite{Spohr97,Yeh99} and are not
160     necessary in this case. All parameters ($T^{*}$, $\mu^{*}$, and
161     $\gamma$) were varied systematically to study the effects of these
162     parameters on the formation of ripple-like phases.
163    
164     \section{Results and Analysis}
165     \label{mc:sec:results}
166    
167     \subsection{Dipolar Ordering and Coexistence Temperatures}
168     The principal method for observing the orientational ordering
169 xsun 3362 transition in dipolar or liquid crystalline systems is the $P_2$ order
170     parameter (defined as $1.5 \times \lambda_{max}$, where
171     $\lambda_{max}$ is the largest eigenvalue of the matrix,
172 xsun 3354 \begin{equation}
173     {\mathsf{S}} = \frac{1}{N} \sum_i \left(
174     \begin{array}{ccc}
175     u^{x}_i u^{x}_i-\frac{1}{3} & u^{x}_i u^{y}_i & u^{x}_i u^{z}_i \\
176     u^{y}_i u^{x}_i & u^{y}_i u^{y}_i -\frac{1}{3} & u^{y}_i u^{z}_i \\
177     u^{z}_i u^{x}_i & u^{z}_i u^{y}_i & u^{z}_i u^{z}_i -\frac{1}{3}
178     \end{array} \right).
179     \label{mceq:opmatrix}
180     \end{equation}
181     Here $u^{\alpha}_i$ is the $\alpha=x,y,z$ component of the unit vector
182     for dipole $i$. $P_2$ will be $1.0$ for a perfectly-ordered system
183     and near $0$ for a randomized system. Note that this order parameter
184     is {\em not} equal to the polarization of the system. For example,
185     the polarization of the perfect anti-ferroelectric system is $0$, but
186     $P_2$ for an anti-ferroelectric system is $1$. The eigenvector of
187     $\mathsf{S}$ corresponding to the largest eigenvalue is familiar as
188     the director axis, which can be used to determine a privileged dipolar
189     axis for dipole-ordered systems. The top panel in Fig. \ref{mcfig:phase}
190     shows the values of $P_2$ as a function of temperature for both
191     triangular ($\gamma = 1.732$) and distorted ($\gamma=1.875$)
192     lattices.
193    
194     \begin{figure}
195     \includegraphics[width=\linewidth]{./figures/mcPhase.pdf}
196     \caption{\label{mcfig:phase} Top panel: The $P_2$ dipolar order parameter as
197     a function of temperature for both triangular ($\gamma = 1.732$) and
198     distorted ($\gamma = 1.875$) lattices. Bottom Panel: The phase
199     diagram for the dipolar membrane model. The line denotes the division
200     between the dipolar ordered (anti-ferroelectric) and disordered phases.
201     An enlarged view near the triangular lattice is shown inset.}
202     \end{figure}
203    
204     There is a clear order-disorder transition in evidence from this data.
205     Both the triangular and distorted lattices have dipolar-ordered
206     low-temperature phases, and orientationally-disordered high
207     temperature phases. The coexistence temperature for the triangular
208     lattice is significantly lower than for the distorted lattices, and
209     the bulk polarization is approximately $0$ for both dipolar ordered
210     and disordered phases. This gives strong evidence that the dipolar
211     ordered phase is anti-ferroelectric. We have verified that this
212     dipolar ordering transition is not a function of system size by
213     performing identical calculations with systems twice as large. The
214     transition is equally smooth at all system sizes that were studied.
215     Additionally, we have repeated the Monte Carlo simulations over a wide
216     range of lattice ratios ($\gamma$) to generate a dipolar
217     order/disorder phase diagram. The bottom panel in Fig. \ref{mcfig:phase}
218     shows that the triangular lattice is a low-temperature cusp in the
219     $T^{*}-\gamma$ phase diagram.
220    
221     This phase diagram is remarkable in that it shows an
222     anti-ferroelectric phase near $\gamma=1.732$ where one would expect
223     lattice frustration to result in disordered phases at all
224     temperatures. Observations of the configurations in this phase show
225     clearly that the system has accomplished dipolar ordering by forming
226     large ripple-like structures. We have observed anti-ferroelectric
227     ordering in all three of the equivalent directions on the triangular
228     lattice, and the dipoles have been observed to organize perpendicular
229     to the membrane normal (in the plane of the membrane). It is
230     particularly interesting to note that the ripple-like structures have
231     also been observed to propagate in the three equivalent directions on
232     the lattice, but the {\em direction of ripple propagation is always
233     perpendicular to the dipole director axis}. A snapshot of a typical
234     anti-ferroelectric rippled structure is shown in
235     Fig. \ref{mcfig:snapshot}.
236    
237     \begin{figure}
238     \includegraphics[width=\linewidth]{./figures/mcSnapshot.pdf}
239     \caption{\label{mcfig:snapshot} Top and Side views of a representative
240     configuration for the dipolar ordered phase supported on the
241     triangular lattice. Note the anti-ferroelectric ordering and the long
242     wavelength buckling of the membrane. Dipolar ordering has been
243     observed in all three equivalent directions on the triangular lattice,
244     and the ripple direction is always perpendicular to the director axis
245     for the dipoles.}
246     \end{figure}
247    
248     Although the snapshot in Fig. \ref{mcfig:snapshot} gives the appearance
249     of three-row stair-like structures, these appear to be transient. On
250     average, the corrugation of the membrane is a relatively smooth,
251     long-wavelength phenomenon, with occasional steep drops between
252     adjacent lines of anti-aligned dipoles.
253    
254     The height-dipole correlation function ($C_{\textrm{hd}}(r, \cos
255     \theta)$) makes the connection between dipolar ordering and the wave
256     vector of the ripple even more explicit. $C_{\textrm{hd}}(r, \cos
257     \theta)$ is an angle-dependent pair distribution function. The angle
258     ($\theta$) is the angle between the intermolecular vector
259     $\vec{r}_{ij}$ and direction of dipole $i$,
260     \begin{equation}
261     C_{\textrm{hd}} = \frac{\langle \frac{1}{n(r)} \sum_{i}\sum_{j>i}
262     h_i \cdot h_j \delta(r - r_{ij}) \delta(\cos \theta_{ij} -
263     \cos \theta)\rangle} {\langle h^2 \rangle}
264     \end{equation}
265     where $\cos \theta_{ij} = \hat{\mu}_{i} \cdot \hat{r}_{ij}$ and
266     $\hat{r}_{ij} = \vec{r}_{ij} / r_{ij}$. $n(r)$ is the number of
267     dipoles found in a cylindrical shell between $r$ and $r+\delta r$ of
268     the central particle. Fig. \ref{mcfig:CrossCorrelation} shows contours
269     of this correlation function for both anti-ferroelectric, rippled
270     membranes as well as for the dipole-disordered portion of the phase
271     diagram.
272    
273     \begin{figure}
274     \includegraphics[width=\linewidth]{./figures/mcHdc.pdf}
275     \caption{\label{mcfig:CrossCorrelation} Contours of the height-dipole
276     correlation function as a function of the dot product between the
277     dipole ($\hat{\mu}$) and inter-dipole separation vector ($\hat{r}$)
278     and the distance ($r$) between the dipoles. Perfect height
279     correlation (contours approaching 1) are present in the ordered phase
280     when the two dipoles are in the same head-to-tail line.
281     Anti-correlation (contours below 0) is only seen when the inter-dipole
282     vector is perpendicular to the dipoles. In the dipole-disordered
283     portion of the phase diagram, there is only weak correlation in the
284     dipole direction and this correlation decays rapidly to zero for
285     intermolecular vectors that are not dipole-aligned.}
286     \end{figure}
287    
288     The height-dipole correlation function gives a map of how the topology
289     of the membrane surface varies with angular deviation around a given
290     dipole. The upper panel of Fig. \ref{mcfig:CrossCorrelation} shows that
291     in the anti-ferroelectric phase, the dipole heights are strongly
292     correlated for dipoles in head-to-tail arrangements, and this
293     correlation persists for very long distances (up to 15 $\sigma$). For
294     portions of the membrane located perpendicular to a given dipole, the
295     membrane height becomes anti-correlated at distances of 10 $\sigma$.
296     The correlation function is relatively smooth; there are no steep
297     jumps or steps, so the stair-like structures in
298     Fig. \ref{mcfig:snapshot} are indeed transient and disappear when
299     averaged over many configurations. In the dipole-disordered phase,
300     the height-dipole correlation function is relatively flat (and hovers
301     near zero). The only significant height correlations are for axial
302     dipoles at very short distances ($r \approx
303     \sigma$).
304    
305     \subsection{Discriminating Ripples from Thermal Undulations}
306    
307     In order to be sure that the structures we have observed are actually
308     a rippled phase and not simply thermal undulations, we have computed
309     the undulation spectrum,
310     \begin{equation}
311     h(\vec{q}) = A^{-1/2} \int d\vec{r}
312     h(\vec{r}) e^{-i \vec{q}\cdot\vec{r}}
313     \end{equation}
314     where $h(\vec{r})$ is the height of the membrane at location $\vec{r}
315     = (x,y)$.~\cite{Safran94,Seifert97} In simple (and more complicated)
316     elastic continuum models, it can shown that in the $NVT$ ensemble, the
317     absolute value of the undulation spectrum can be written,
318     \begin{equation}
319     \langle | h(q) |^2 \rangle_{NVT} = \frac{k_B T}{k_c q^4 +
320     \gamma q^2},
321     \label{mceq:fit}
322     \end{equation}
323     where $k_c$ is the bending modulus for the membrane, and $\gamma$ is
324     the mechanical surface tension.~\cite{Safran94} The systems studied in
325     this paper have essentially zero bending moduli ($k_c$) and relatively
326     large mechanical surface tensions ($\gamma$), so a much simpler form
327     can be written,
328     \begin{equation}
329 xsun 3361 \langle | h(q) |^2 \rangle_{NVT} = \frac{k_B T}{\gamma q^2}.
330 xsun 3354 \label{mceq:fit2}
331     \end{equation}
332    
333     The undulation spectrum is computed by superimposing a rectangular
334     grid on top of the membrane, and by assigning height ($h(\vec{r})$)
335     values to the grid from the average of all dipoles that fall within a
336     given $\vec{r}+d\vec{r}$ grid area. Empty grid pixels are assigned
337     height values by interpolation from the nearest neighbor pixels. A
338     standard 2-d Fourier transform is then used to obtain $\langle |
339     h(q)|^2 \rangle$. Alternatively, since the dipoles sit on a Bravais
340     lattice, one could use the heights of the lattice points themselves as
341     the grid for the Fourier transform (without interpolating to a square
342     grid). However, if lateral translational freedom is added to this
343     model (a likely extension), an interpolated grid method for computing
344     undulation spectra will be required.
345    
346     As mentioned above, the best fits to our undulation spectra are
347     obtained by setting the value of $k_c$ to 0. In Fig. \ref{mcfig:fit} we
348     show typical undulation spectra for two different regions of the phase
349     diagram along with their fits from the Landau free energy approach
350     (Eq. \ref{mceq:fit2}). In the high-temperature disordered phase, the
351     Landau fits can be nearly perfect, and from these fits we can estimate
352     the tension in the surface. In reduced units, typical values of
353     $\gamma^{*} = \gamma / \epsilon = 2500$ are obtained for the
354     disordered phase ($\gamma^{*} = 2551.7$ in the top panel of
355     Fig. \ref{mcfig:fit}).
356    
357     Typical values of $\gamma^{*}$ in the dipolar-ordered phase are much
358     higher than in the dipolar-disordered phase ($\gamma^{*} = 73,538$ in
359     the lower panel of Fig. \ref{mcfig:fit}). For the dipolar-ordered
360     triangular lattice near the coexistence temperature, we also observe
361     long wavelength undulations that are far outliers to the fits. That
362     is, the Landau free energy fits are well within error bars for most of
363     the other points, but can be off by {\em orders of magnitude} for a
364     few low frequency components.
365    
366     We interpret these outliers as evidence that these low frequency modes
367     are {\em non-thermal undulations}. We take this as evidence that we
368     are actually seeing a rippled phase developing in this model system.
369    
370     \begin{figure}
371     \includegraphics[width=\linewidth]{./figures/mcLogFit.pdf}
372     \caption{\label{mcfig:fit} Evidence that the observed ripples are {\em
373     not} thermal undulations is obtained from the 2-d Fourier transform
374     $\langle |h^{*}(\vec{q})|^2 \rangle$ of the height profile ($\langle
375     h^{*}(x,y) \rangle$). Rippled samples show low-wavelength peaks that
376     are outliers on the Landau free energy fits by an order of magnitude.
377     Samples exhibiting only thermal undulations fit Eq. \ref{mceq:fit}
378     remarkably well.}
379     \end{figure}
380    
381     \subsection{Effects of Potential Parameters on Amplitude and Wavelength}
382    
383     We have used two different methods to estimate the amplitude and
384     periodicity of the ripples. The first method requires projection of
385     the ripples onto a one dimensional rippling axis. Since the rippling
386     is always perpendicular to the dipole director axis, we can define a
387     ripple vector as follows. The largest eigenvector, $s_1$, of the
388     $\mathsf{S}$ matrix in Eq. \ref{mceq:opmatrix} is projected onto a
389     planar director axis,
390     \begin{equation}
391     \vec{d} = \left(\begin{array}{c}
392     \vec{s}_1 \cdot \hat{i} \\
393     \vec{s}_1 \cdot \hat{j} \\
394     0
395     \end{array} \right).
396     \end{equation}
397     ($\hat{i}$, $\hat{j}$ and $\hat{k}$ are unit vectors along the $x$,
398     $y$, and $z$ axes, respectively.) The rippling axis is in the plane of
399     the membrane and is perpendicular to the planar director axis,
400     \begin{equation}
401     \vec{q}_{\mathrm{rip}} = \vec{d} \times \hat{k}
402     \end{equation}
403     We can then find the height profile of the membrane along the ripple
404     axis by projecting heights of the dipoles to obtain a one-dimensional
405     height profile, $h(q_{\mathrm{rip}})$. Ripple wavelengths can be
406     estimated from the largest non-thermal low-frequency component in the
407     Fourier transform of $h(q_{\mathrm{rip}})$. Amplitudes can be
408     estimated by measuring peak-to-trough distances in
409     $h(q_{\mathrm{rip}})$ itself.
410    
411     A second, more accurate, and simpler method for estimating ripple
412     shape is to extract the wavelength and height information directly
413     from the largest non-thermal peak in the undulation spectrum. For
414     large-amplitude ripples, the two methods give similar results. The
415     one-dimensional projection method is more prone to noise (particularly
416     in the amplitude estimates for the distorted lattices). We report
417     amplitudes and wavelengths taken directly from the undulation spectrum
418     below.
419    
420     In the triangular lattice ($\gamma = \sqrt{3}$), the rippling is
421     observed for temperatures ($T^{*}$) from $61-122$. The wavelength of
422     the ripples is remarkably stable at 21.4~$\sigma$ for all but the
423     temperatures closest to the order-disorder transition. At $T^{*} =
424     122$, the wavelength drops to 17.1~$\sigma$.
425    
426     The dependence of the amplitude on temperature is shown in the top
427     panel of Fig. \ref{mcfig:Amplitude}. The rippled structures shrink
428     smoothly as the temperature rises towards the order-disorder
429     transition. The wavelengths and amplitudes we observe are
430     surprisingly close to the $\Lambda / 2$ phase observed by Kaasgaard
431     {\it et al.} in their work on PC-based lipids.\cite{Kaasgaard03}
432     However, this is coincidental agreement based on a choice of 7~\AA~as
433     the mean spacing between lipids.
434    
435     \begin{figure}
436     \includegraphics[width=\linewidth]{./figures/mcProperties_sq.pdf}
437     \caption{\label{mcfig:Amplitude} a) The amplitude $A^{*}$ of the ripples
438     vs. temperature for a triangular lattice. b) The amplitude $A^{*}$ of
439     the ripples vs. dipole strength ($\mu^{*}$) for both the triangular
440     lattice (circles) and distorted lattice (squares). The reduced
441     temperatures were kept fixed at $T^{*} = 94$ for the triangular
442     lattice and $T^{*} = 106$ for the distorted lattice (approximately 2/3
443     of the order-disorder transition temperature for each lattice).}
444     \end{figure}
445    
446     The ripples can be made to disappear by increasing the internal
447     elastic tension (i.e. by increasing $k_r$ or equivalently, reducing
448     the dipole moment). The amplitude of the ripples depends critically
449     on the strength of the dipole moments ($\mu^{*}$) in Eq. \ref{mceq:pot}.
450     If the dipoles are weakened substantially (below $\mu^{*}$ = 20) at a
451     fixed temperature of 94, the membrane loses dipolar ordering
452     and the ripple structures. The ripples reach a peak amplitude of
453     3.7~$\sigma$ at a dipolar strength of 25. We show the dependence
454     of ripple amplitude on the dipolar strength in
455     Fig. \ref{mcfig:Amplitude}.
456    
457     \subsection{Distorted lattices}
458    
459     We have also investigated the effect of the lattice geometry by
460     changing the ratio of lattice constants ($\gamma$) while keeping the
461     average nearest-neighbor spacing constant. The anti-ferroelectric state
462     is accessible for all $\gamma$ values we have used, although the
463     distorted triangular lattices prefer a particular director axis due to
464     the anisotropy of the lattice.
465    
466     Our observation of rippling behavior was not limited to the triangular
467     lattices. In distorted lattices the anti-ferroelectric phase can
468     develop nearly instantaneously in the Monte Carlo simulations, and
469     these dipolar-ordered phases tend to be remarkably flat. Whenever
470     rippling has been observed in these distorted lattices
471     (e.g. $\gamma = 1.875$), we see relatively short ripple wavelengths
472     (14 $\sigma$) and amplitudes of 2.4~$\sigma$. These ripples are
473     weakly dependent on dipolar strength (see Fig. \ref{mcfig:Amplitude}),
474     although below a dipolar strength of $\mu^{*} = 20$, the membrane
475     loses dipolar ordering and displays only thermal undulations.
476    
477     The ripple phase does {\em not} appear at all values of $\gamma$. We
478     have only observed non-thermal undulations in the range $1.625 <
479     \gamma < 1.875$. Outside this range, the order-disorder transition in
480     the dipoles remains, but the ordered dipolar phase has only thermal
481     undulations. This is one of our strongest pieces of evidence that
482     rippling is a symmetry-breaking phenomenon for triangular and
483     nearly-triangular lattices.
484    
485     \subsection{Effects of System Size}
486     To evaluate the effect of finite system size, we have performed a
487     series of simulations on the triangular lattice at a reduced
488     temperature of 122, which is just below the order-disorder transition
489     temperature ($T^{*} = 139$). These conditions are in the
490     dipole-ordered and rippled portion of the phase diagram. These are
491     also the conditions that should be most susceptible to system size
492     effects.
493    
494     \begin{figure}
495     \includegraphics[width=\linewidth]{./figures/mcSystemSize.pdf}
496     \caption{\label{mcfig:systemsize} The ripple wavelength (top) and
497     amplitude (bottom) as a function of system size for a triangular
498     lattice ($\gamma=1.732$) at $T^{*} = 122$.}
499     \end{figure}
500    
501     There is substantial dependence on system size for small (less than
502     29~$\sigma$) periodic boxes. Notably, there are resonances apparent
503     in the ripple amplitudes at box lengths of 17.3 and 29.5 $\sigma$.
504     For larger systems, the behavior of the ripples appears to have
505     stabilized and is on a trend to slightly smaller amplitudes (and
506     slightly longer wavelengths) than were observed from the 34.3 $\sigma$
507     box sizes that were used for most of the calculations.
508    
509     It is interesting to note that system sizes which are multiples of the
510     default ripple wavelength can enhance the amplitude of the observed
511     ripples, but appears to have only a minor effect on the observed
512     wavelength. It would, of course, be better to use system sizes that
513     were many multiples of the ripple wavelength to be sure that the
514     periodic box is not driving the phenomenon, but at the largest system
515     size studied (70 $\sigma$ $\times$ 70 $\sigma$), the number of dipoles
516     (5440) made long Monte Carlo simulations prohibitively expensive.
517    
518     \section{Discussion}
519     \label{mc:sec:discussion}
520    
521     We have been able to show that a simple dipolar lattice model which
522     contains only molecular packing (from the lattice), anisotropy (in the
523     form of electrostatic dipoles) and a weak elastic tension (in the form
524     of a nearest-neighbor harmonic potential) is capable of exhibiting
525     stable long-wavelength non-thermal surface corrugations. The best
526     explanation for this behavior is that the ability of the dipoles to
527     translate out of the plane of the membrane is enough to break the
528     symmetry of the triangular lattice and allow the energetic benefit
529     from the formation of a bulk anti-ferroelectric phase. Were the weak
530     elastic tension absent from our model, it would be possible for the
531     entire lattice to ``tilt'' using $z$-translation. Tilting the lattice
532     in this way would yield an effectively non-triangular lattice which
533     would avoid dipolar frustration altogether. With the elastic tension
534     in place, bulk tilt causes a large strain, and the least costly way to
535     release this strain is between two rows of anti-aligned dipoles.
536     These ``breaks'' will result in rippled or sawtooth patterns in the
537     membrane, and allow small stripes of membrane to form
538     anti-ferroelectric regions that are tilted relative to the averaged
539     membrane normal.
540    
541     Although the dipole-dipole interaction is the major driving force for
542     the long range orientational ordered state, the formation of the
543     stable, smooth ripples is a result of the competition between the
544     elastic tension and the dipole-dipole interactions. This statement is
545     supported by the variation in $\mu^{*}$. Substantially weaker dipoles
546     relative to the surface tension can cause the corrugated phase to
547     disappear.
548    
549     The packing of the dipoles into a nearly-triangular lattice is clearly
550     an important piece of the puzzle. The dipolar head groups of lipid
551     molecules are sterically (as well as electrostatically) anisotropic,
552     and would not pack in triangular arrangements without the steric
553     interference of adjacent molecular bodies. Since we only see rippled
554     phases in the neighborhood of $\gamma=\sqrt{3}$, this implies that
555     even if this dipolar mechanism is the correct explanation for the
556     ripple phase in realistic bilayers, there would still be a role played
557     by the lipid chains in the in-plane organization of the triangularly
558     ordered phases which could support ripples. The present model is
559     certainly not detailed enough to answer exactly what drives the
560     formation of the $P_{\beta'}$ phase in real lipids, but suggests some
561     avenues for further experiments.
562    
563     The most important prediction we can make using the results from this
564     simple model is that if dipolar ordering is driving the surface
565     corrugation, the wave vectors for the ripples should always found to
566     be {\it perpendicular} to the dipole director axis. This prediction
567     should suggest experimental designs which test whether this is really
568     true in the phosphatidylcholine $P_{\beta'}$ phases. The dipole
569     director axis should also be easily computable for the all-atom and
570     coarse-grained simulations that have been published in the literature.
571    
572     Our other observation about the ripple and dipolar directionality is
573     that the dipole director axis can be found to be parallel to any of
574     the three equivalent lattice vectors in the triangular lattice.
575     Defects in the ordering of the dipoles can cause the dipole director
576     (and consequently the surface corrugation) of small regions to be
577     rotated relative to each other by 120$^{\circ}$. This is a similar
578     behavior to the domain rotation seen in the AFM studies of Kaasgaard
579     {\it et al.}\cite{Kaasgaard03}
580    
581     Although our model is simple, it exhibits some rich and unexpected
582     behaviors. It would clearly be a closer approximation to the reality
583     if we allowed greater translational freedom to the dipoles and
584     replaced the somewhat artificial lattice packing and the harmonic
585     elastic tension with more realistic molecular modeling potentials.
586     What we have done is to present a simple model which exhibits bulk
587     non-thermal corrugation, and our explanation of this rippling
588     phenomenon will help us design more accurate molecular models for
589     corrugated membranes and experiments to test whether rippling is
590     dipole-driven or not.