ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/xDissertation/mc.tex
Revision: 3383
Committed: Wed Apr 16 21:56:34 2008 UTC (16 years, 5 months ago) by xsun
Content type: application/x-tex
File size: 32364 byte(s)
Log Message:
formatting check

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 xsun 3383 series of Me\-trop\-o\-lis Monte Carlo simulations of moderately-sized
143     (34.3 $\sigma$ on a side) patches of membrane hosted on both
144     triangular ($\gamma = a/b = \sqrt{3}$) and distorted ($\gamma \neq
145     \sqrt{3}$) lattices. The linear extent of one edge of the monolayer
146     was $20 a$ and the system was kept roughly square. The average
147     distance that coplanar dipoles were positioned from their six nearest
148     neighbors was 1 $\sigma$ (on both triangular and distorted lattices).
149     Typical system sizes were 1360 dipoles for the triangular lattices and
150 xsun 3354 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 xsun 3383 parameters on the formation of ripple-like phases. The error bars in
163     our results are one $\sigma$ on each side of the average values, where
164     $\sigma$ is the standard deviation obtained from repeated observations
165     of many configurations.
166 xsun 3354
167     \section{Results and Analysis}
168     \label{mc:sec:results}
169    
170     \subsection{Dipolar Ordering and Coexistence Temperatures}
171     The principal method for observing the orientational ordering
172 xsun 3362 transition in dipolar or liquid crystalline systems is the $P_2$ order
173     parameter (defined as $1.5 \times \lambda_{max}$, where
174     $\lambda_{max}$ is the largest eigenvalue of the matrix,
175 xsun 3354 \begin{equation}
176     {\mathsf{S}} = \frac{1}{N} \sum_i \left(
177     \begin{array}{ccc}
178     u^{x}_i u^{x}_i-\frac{1}{3} & u^{x}_i u^{y}_i & u^{x}_i u^{z}_i \\
179     u^{y}_i u^{x}_i & u^{y}_i u^{y}_i -\frac{1}{3} & u^{y}_i u^{z}_i \\
180     u^{z}_i u^{x}_i & u^{z}_i u^{y}_i & u^{z}_i u^{z}_i -\frac{1}{3}
181     \end{array} \right).
182     \label{mceq:opmatrix}
183     \end{equation}
184     Here $u^{\alpha}_i$ is the $\alpha=x,y,z$ component of the unit vector
185     for dipole $i$. $P_2$ will be $1.0$ for a perfectly-ordered system
186     and near $0$ for a randomized system. Note that this order parameter
187     is {\em not} equal to the polarization of the system. For example,
188     the polarization of the perfect anti-ferroelectric system is $0$, but
189     $P_2$ for an anti-ferroelectric system is $1$. The eigenvector of
190     $\mathsf{S}$ corresponding to the largest eigenvalue is familiar as
191     the director axis, which can be used to determine a privileged dipolar
192     axis for dipole-ordered systems. The top panel in Fig. \ref{mcfig:phase}
193     shows the values of $P_2$ as a function of temperature for both
194     triangular ($\gamma = 1.732$) and distorted ($\gamma=1.875$)
195     lattices.
196    
197     \begin{figure}
198     \includegraphics[width=\linewidth]{./figures/mcPhase.pdf}
199 xsun 3383 \caption[ The $P_2$ dipolar order parameter as
200     a function of temperature and the phase diagram for the dipolar
201     membrane model]{\label{mcfig:phase} Top panel: The $P_2$ dipolar order
202     parameter as a function of temperature for both triangular ($\gamma =
203     1.732$) and distorted ($\gamma = 1.875$) lattices. Bottom Panel: The
204     phase diagram for the dipolar membrane model. The line denotes the
205     division between the dipolar ordered (anti-ferroelectric) and
206     disordered phases. An enlarged view near the triangular lattice is
207     shown inset.}
208 xsun 3354 \end{figure}
209    
210     There is a clear order-disorder transition in evidence from this data.
211     Both the triangular and distorted lattices have dipolar-ordered
212 xsun 3383 low-temperature phases, and ori\-en\-ta\-tion\-al\-ly-disordered high
213 xsun 3354 temperature phases. The coexistence temperature for the triangular
214     lattice is significantly lower than for the distorted lattices, and
215     the bulk polarization is approximately $0$ for both dipolar ordered
216     and disordered phases. This gives strong evidence that the dipolar
217     ordered phase is anti-ferroelectric. We have verified that this
218     dipolar ordering transition is not a function of system size by
219     performing identical calculations with systems twice as large. The
220     transition is equally smooth at all system sizes that were studied.
221     Additionally, we have repeated the Monte Carlo simulations over a wide
222     range of lattice ratios ($\gamma$) to generate a dipolar
223 xsun 3383 order/disorder phase diagram. The bottom panel in
224     Fig. \ref{mcfig:phase} shows that the triangular lattice is a
225     low-temperature cusp in the $T^{*}-\gamma$ phase diagram.
226 xsun 3354
227     This phase diagram is remarkable in that it shows an
228     anti-ferroelectric phase near $\gamma=1.732$ where one would expect
229     lattice frustration to result in disordered phases at all
230     temperatures. Observations of the configurations in this phase show
231     clearly that the system has accomplished dipolar ordering by forming
232     large ripple-like structures. We have observed anti-ferroelectric
233     ordering in all three of the equivalent directions on the triangular
234     lattice, and the dipoles have been observed to organize perpendicular
235     to the membrane normal (in the plane of the membrane). It is
236     particularly interesting to note that the ripple-like structures have
237     also been observed to propagate in the three equivalent directions on
238     the lattice, but the {\em direction of ripple propagation is always
239     perpendicular to the dipole director axis}. A snapshot of a typical
240     anti-ferroelectric rippled structure is shown in
241     Fig. \ref{mcfig:snapshot}.
242    
243     \begin{figure}
244     \includegraphics[width=\linewidth]{./figures/mcSnapshot.pdf}
245 xsun 3383 \caption[ Top and Side views of a representative
246 xsun 3354 configuration for the dipolar ordered phase supported on the
247 xsun 3383 triangular lattice]{\label{mcfig:snapshot} Top and Side views of a
248     representative configuration for the dipolar ordered phase supported
249     on the triangular lattice. Note the anti-ferroelectric ordering and
250     the long wavelength buckling of the membrane. Dipolar ordering has
251     been observed in all three equivalent directions on the triangular
252     lattice, and the ripple direction is always perpendicular to the
253     director axis for the dipoles.}
254 xsun 3354 \end{figure}
255    
256     Although the snapshot in Fig. \ref{mcfig:snapshot} gives the appearance
257     of three-row stair-like structures, these appear to be transient. On
258     average, the corrugation of the membrane is a relatively smooth,
259     long-wavelength phenomenon, with occasional steep drops between
260     adjacent lines of anti-aligned dipoles.
261    
262     The height-dipole correlation function ($C_{\textrm{hd}}(r, \cos
263     \theta)$) makes the connection between dipolar ordering and the wave
264     vector of the ripple even more explicit. $C_{\textrm{hd}}(r, \cos
265     \theta)$ is an angle-dependent pair distribution function. The angle
266     ($\theta$) is the angle between the intermolecular vector
267     $\vec{r}_{ij}$ and direction of dipole $i$,
268     \begin{equation}
269     C_{\textrm{hd}} = \frac{\langle \frac{1}{n(r)} \sum_{i}\sum_{j>i}
270     h_i \cdot h_j \delta(r - r_{ij}) \delta(\cos \theta_{ij} -
271     \cos \theta)\rangle} {\langle h^2 \rangle}
272     \end{equation}
273     where $\cos \theta_{ij} = \hat{\mu}_{i} \cdot \hat{r}_{ij}$ and
274     $\hat{r}_{ij} = \vec{r}_{ij} / r_{ij}$. $n(r)$ is the number of
275     dipoles found in a cylindrical shell between $r$ and $r+\delta r$ of
276     the central particle. Fig. \ref{mcfig:CrossCorrelation} shows contours
277     of this correlation function for both anti-ferroelectric, rippled
278     membranes as well as for the dipole-disordered portion of the phase
279     diagram.
280    
281     \begin{figure}
282     \includegraphics[width=\linewidth]{./figures/mcHdc.pdf}
283 xsun 3383 \caption[Contours of the height-dipole
284     correlation function]{\label{mcfig:CrossCorrelation} Contours of the
285     height-dipole correlation function as a function of the dot product
286     between the dipole ($\hat{\mu}$) and inter-dipole separation vector
287     ($\hat{r}$) and the distance ($r$) between the dipoles. Perfect
288     height correlation (contours approaching 1) are present in the ordered
289     phase when the two dipoles are in the same head-to-tail line.
290 xsun 3354 Anti-correlation (contours below 0) is only seen when the inter-dipole
291     vector is perpendicular to the dipoles. In the dipole-disordered
292     portion of the phase diagram, there is only weak correlation in the
293     dipole direction and this correlation decays rapidly to zero for
294     intermolecular vectors that are not dipole-aligned.}
295     \end{figure}
296    
297     The height-dipole correlation function gives a map of how the topology
298     of the membrane surface varies with angular deviation around a given
299     dipole. The upper panel of Fig. \ref{mcfig:CrossCorrelation} shows that
300     in the anti-ferroelectric phase, the dipole heights are strongly
301     correlated for dipoles in head-to-tail arrangements, and this
302     correlation persists for very long distances (up to 15 $\sigma$). For
303     portions of the membrane located perpendicular to a given dipole, the
304     membrane height becomes anti-correlated at distances of 10 $\sigma$.
305     The correlation function is relatively smooth; there are no steep
306     jumps or steps, so the stair-like structures in
307     Fig. \ref{mcfig:snapshot} are indeed transient and disappear when
308     averaged over many configurations. In the dipole-disordered phase,
309     the height-dipole correlation function is relatively flat (and hovers
310     near zero). The only significant height correlations are for axial
311     dipoles at very short distances ($r \approx
312     \sigma$).
313    
314     \subsection{Discriminating Ripples from Thermal Undulations}
315    
316     In order to be sure that the structures we have observed are actually
317     a rippled phase and not simply thermal undulations, we have computed
318     the undulation spectrum,
319     \begin{equation}
320     h(\vec{q}) = A^{-1/2} \int d\vec{r}
321     h(\vec{r}) e^{-i \vec{q}\cdot\vec{r}}
322     \end{equation}
323     where $h(\vec{r})$ is the height of the membrane at location $\vec{r}
324     = (x,y)$.~\cite{Safran94,Seifert97} In simple (and more complicated)
325     elastic continuum models, it can shown that in the $NVT$ ensemble, the
326     absolute value of the undulation spectrum can be written,
327     \begin{equation}
328     \langle | h(q) |^2 \rangle_{NVT} = \frac{k_B T}{k_c q^4 +
329     \gamma q^2},
330     \label{mceq:fit}
331     \end{equation}
332     where $k_c$ is the bending modulus for the membrane, and $\gamma$ is
333     the mechanical surface tension.~\cite{Safran94} The systems studied in
334     this paper have essentially zero bending moduli ($k_c$) and relatively
335     large mechanical surface tensions ($\gamma$), so a much simpler form
336     can be written,
337     \begin{equation}
338 xsun 3361 \langle | h(q) |^2 \rangle_{NVT} = \frac{k_B T}{\gamma q^2}.
339 xsun 3354 \label{mceq:fit2}
340     \end{equation}
341    
342     The undulation spectrum is computed by superimposing a rectangular
343     grid on top of the membrane, and by assigning height ($h(\vec{r})$)
344     values to the grid from the average of all dipoles that fall within a
345     given $\vec{r}+d\vec{r}$ grid area. Empty grid pixels are assigned
346     height values by interpolation from the nearest neighbor pixels. A
347     standard 2-d Fourier transform is then used to obtain $\langle |
348     h(q)|^2 \rangle$. Alternatively, since the dipoles sit on a Bravais
349     lattice, one could use the heights of the lattice points themselves as
350     the grid for the Fourier transform (without interpolating to a square
351     grid). However, if lateral translational freedom is added to this
352     model (a likely extension), an interpolated grid method for computing
353     undulation spectra will be required.
354    
355     As mentioned above, the best fits to our undulation spectra are
356     obtained by setting the value of $k_c$ to 0. In Fig. \ref{mcfig:fit} we
357     show typical undulation spectra for two different regions of the phase
358     diagram along with their fits from the Landau free energy approach
359     (Eq. \ref{mceq:fit2}). In the high-temperature disordered phase, the
360     Landau fits can be nearly perfect, and from these fits we can estimate
361     the tension in the surface. In reduced units, typical values of
362     $\gamma^{*} = \gamma / \epsilon = 2500$ are obtained for the
363     disordered phase ($\gamma^{*} = 2551.7$ in the top panel of
364     Fig. \ref{mcfig:fit}).
365    
366     Typical values of $\gamma^{*}$ in the dipolar-ordered phase are much
367     higher than in the dipolar-disordered phase ($\gamma^{*} = 73,538$ in
368     the lower panel of Fig. \ref{mcfig:fit}). For the dipolar-ordered
369     triangular lattice near the coexistence temperature, we also observe
370     long wavelength undulations that are far outliers to the fits. That
371     is, the Landau free energy fits are well within error bars for most of
372     the other points, but can be off by {\em orders of magnitude} for a
373     few low frequency components.
374    
375     We interpret these outliers as evidence that these low frequency modes
376     are {\em non-thermal undulations}. We take this as evidence that we
377     are actually seeing a rippled phase developing in this model system.
378    
379     \begin{figure}
380     \includegraphics[width=\linewidth]{./figures/mcLogFit.pdf}
381 xsun 3383 \caption[Evidence that the observed ripples are {\em not} thermal
382     undulations]{\label{mcfig:fit} Evidence that the observed ripples are
383     {\em not} thermal undulations is obtained from the 2-d Fourier
384     transform $\langle |h^{*}(\vec{q})|^2 \rangle$ of the height profile
385     ($\langle h^{*}(x,y) \rangle$). Rippled samples show low-wavelength
386     peaks that are outliers on the Landau free energy fits by an order of
387     magnitude. Samples exhibiting only thermal undulations fit
388     Eq. \ref{mceq:fit} remarkably well.}
389 xsun 3354 \end{figure}
390    
391     \subsection{Effects of Potential Parameters on Amplitude and Wavelength}
392    
393     We have used two different methods to estimate the amplitude and
394     periodicity of the ripples. The first method requires projection of
395     the ripples onto a one dimensional rippling axis. Since the rippling
396     is always perpendicular to the dipole director axis, we can define a
397     ripple vector as follows. The largest eigenvector, $s_1$, of the
398     $\mathsf{S}$ matrix in Eq. \ref{mceq:opmatrix} is projected onto a
399     planar director axis,
400     \begin{equation}
401     \vec{d} = \left(\begin{array}{c}
402     \vec{s}_1 \cdot \hat{i} \\
403     \vec{s}_1 \cdot \hat{j} \\
404     0
405     \end{array} \right).
406     \end{equation}
407     ($\hat{i}$, $\hat{j}$ and $\hat{k}$ are unit vectors along the $x$,
408     $y$, and $z$ axes, respectively.) The rippling axis is in the plane of
409     the membrane and is perpendicular to the planar director axis,
410     \begin{equation}
411     \vec{q}_{\mathrm{rip}} = \vec{d} \times \hat{k}
412     \end{equation}
413     We can then find the height profile of the membrane along the ripple
414     axis by projecting heights of the dipoles to obtain a one-dimensional
415     height profile, $h(q_{\mathrm{rip}})$. Ripple wavelengths can be
416     estimated from the largest non-thermal low-frequency component in the
417     Fourier transform of $h(q_{\mathrm{rip}})$. Amplitudes can be
418     estimated by measuring peak-to-trough distances in
419     $h(q_{\mathrm{rip}})$ itself.
420    
421     A second, more accurate, and simpler method for estimating ripple
422     shape is to extract the wavelength and height information directly
423     from the largest non-thermal peak in the undulation spectrum. For
424     large-amplitude ripples, the two methods give similar results. The
425     one-dimensional projection method is more prone to noise (particularly
426     in the amplitude estimates for the distorted lattices). We report
427     amplitudes and wavelengths taken directly from the undulation spectrum
428     below.
429    
430     In the triangular lattice ($\gamma = \sqrt{3}$), the rippling is
431     observed for temperatures ($T^{*}$) from $61-122$. The wavelength of
432     the ripples is remarkably stable at 21.4~$\sigma$ for all but the
433     temperatures closest to the order-disorder transition. At $T^{*} =
434     122$, the wavelength drops to 17.1~$\sigma$.
435    
436     The dependence of the amplitude on temperature is shown in the top
437     panel of Fig. \ref{mcfig:Amplitude}. The rippled structures shrink
438     smoothly as the temperature rises towards the order-disorder
439     transition. The wavelengths and amplitudes we observe are
440     surprisingly close to the $\Lambda / 2$ phase observed by Kaasgaard
441     {\it et al.} in their work on PC-based lipids.\cite{Kaasgaard03}
442     However, this is coincidental agreement based on a choice of 7~\AA~as
443     the mean spacing between lipids.
444    
445     \begin{figure}
446     \includegraphics[width=\linewidth]{./figures/mcProperties_sq.pdf}
447 xsun 3383 \caption[ The amplitude $A^{*}$ of the ripples
448     vs. temperature and dipole strength
449     ($\mu^{*}$)]{\label{mcfig:Amplitude} a) The amplitude $A^{*}$ of the
450     ripples vs. temperature for a triangular lattice. b) The amplitude
451     $A^{*}$ of the ripples vs. dipole strength ($\mu^{*}$) for both the
452     triangular lattice (circles) and distorted lattice (squares). The
453     reduced temperatures were kept fixed at $T^{*} = 94$ for the
454     triangular lattice and $T^{*} = 106$ for the distorted lattice
455     (approximately 2/3 of the order-disorder transition temperature for
456     each lattice).}
457 xsun 3354 \end{figure}
458    
459     The ripples can be made to disappear by increasing the internal
460     elastic tension (i.e. by increasing $k_r$ or equivalently, reducing
461     the dipole moment). The amplitude of the ripples depends critically
462     on the strength of the dipole moments ($\mu^{*}$) in Eq. \ref{mceq:pot}.
463     If the dipoles are weakened substantially (below $\mu^{*}$ = 20) at a
464     fixed temperature of 94, the membrane loses dipolar ordering
465     and the ripple structures. The ripples reach a peak amplitude of
466     3.7~$\sigma$ at a dipolar strength of 25. We show the dependence
467     of ripple amplitude on the dipolar strength in
468     Fig. \ref{mcfig:Amplitude}.
469    
470     \subsection{Distorted lattices}
471    
472     We have also investigated the effect of the lattice geometry by
473     changing the ratio of lattice constants ($\gamma$) while keeping the
474     average nearest-neighbor spacing constant. The anti-ferroelectric state
475     is accessible for all $\gamma$ values we have used, although the
476     distorted triangular lattices prefer a particular director axis due to
477     the anisotropy of the lattice.
478    
479     Our observation of rippling behavior was not limited to the triangular
480     lattices. In distorted lattices the anti-ferroelectric phase can
481     develop nearly instantaneously in the Monte Carlo simulations, and
482     these dipolar-ordered phases tend to be remarkably flat. Whenever
483     rippling has been observed in these distorted lattices
484     (e.g. $\gamma = 1.875$), we see relatively short ripple wavelengths
485     (14 $\sigma$) and amplitudes of 2.4~$\sigma$. These ripples are
486     weakly dependent on dipolar strength (see Fig. \ref{mcfig:Amplitude}),
487     although below a dipolar strength of $\mu^{*} = 20$, the membrane
488     loses dipolar ordering and displays only thermal undulations.
489    
490     The ripple phase does {\em not} appear at all values of $\gamma$. We
491     have only observed non-thermal undulations in the range $1.625 <
492     \gamma < 1.875$. Outside this range, the order-disorder transition in
493     the dipoles remains, but the ordered dipolar phase has only thermal
494     undulations. This is one of our strongest pieces of evidence that
495     rippling is a symmetry-breaking phenomenon for triangular and
496     nearly-triangular lattices.
497    
498     \subsection{Effects of System Size}
499     To evaluate the effect of finite system size, we have performed a
500     series of simulations on the triangular lattice at a reduced
501     temperature of 122, which is just below the order-disorder transition
502     temperature ($T^{*} = 139$). These conditions are in the
503     dipole-ordered and rippled portion of the phase diagram. These are
504     also the conditions that should be most susceptible to system size
505     effects.
506    
507     \begin{figure}
508     \includegraphics[width=\linewidth]{./figures/mcSystemSize.pdf}
509 xsun 3383 \caption[The ripple wavelength and amplitude as a function of system
510     size]{\label{mcfig:systemsize} The ripple wavelength (top) and
511 xsun 3354 amplitude (bottom) as a function of system size for a triangular
512     lattice ($\gamma=1.732$) at $T^{*} = 122$.}
513     \end{figure}
514    
515     There is substantial dependence on system size for small (less than
516     29~$\sigma$) periodic boxes. Notably, there are resonances apparent
517     in the ripple amplitudes at box lengths of 17.3 and 29.5 $\sigma$.
518     For larger systems, the behavior of the ripples appears to have
519     stabilized and is on a trend to slightly smaller amplitudes (and
520     slightly longer wavelengths) than were observed from the 34.3 $\sigma$
521     box sizes that were used for most of the calculations.
522    
523     It is interesting to note that system sizes which are multiples of the
524     default ripple wavelength can enhance the amplitude of the observed
525     ripples, but appears to have only a minor effect on the observed
526     wavelength. It would, of course, be better to use system sizes that
527     were many multiples of the ripple wavelength to be sure that the
528     periodic box is not driving the phenomenon, but at the largest system
529     size studied (70 $\sigma$ $\times$ 70 $\sigma$), the number of dipoles
530     (5440) made long Monte Carlo simulations prohibitively expensive.
531    
532     \section{Discussion}
533     \label{mc:sec:discussion}
534    
535     We have been able to show that a simple dipolar lattice model which
536     contains only molecular packing (from the lattice), anisotropy (in the
537     form of electrostatic dipoles) and a weak elastic tension (in the form
538     of a nearest-neighbor harmonic potential) is capable of exhibiting
539     stable long-wavelength non-thermal surface corrugations. The best
540     explanation for this behavior is that the ability of the dipoles to
541     translate out of the plane of the membrane is enough to break the
542     symmetry of the triangular lattice and allow the energetic benefit
543     from the formation of a bulk anti-ferroelectric phase. Were the weak
544     elastic tension absent from our model, it would be possible for the
545     entire lattice to ``tilt'' using $z$-translation. Tilting the lattice
546     in this way would yield an effectively non-triangular lattice which
547     would avoid dipolar frustration altogether. With the elastic tension
548     in place, bulk tilt causes a large strain, and the least costly way to
549     release this strain is between two rows of anti-aligned dipoles.
550     These ``breaks'' will result in rippled or sawtooth patterns in the
551     membrane, and allow small stripes of membrane to form
552     anti-ferroelectric regions that are tilted relative to the averaged
553     membrane normal.
554    
555     Although the dipole-dipole interaction is the major driving force for
556     the long range orientational ordered state, the formation of the
557     stable, smooth ripples is a result of the competition between the
558     elastic tension and the dipole-dipole interactions. This statement is
559     supported by the variation in $\mu^{*}$. Substantially weaker dipoles
560     relative to the surface tension can cause the corrugated phase to
561     disappear.
562    
563     The packing of the dipoles into a nearly-triangular lattice is clearly
564     an important piece of the puzzle. The dipolar head groups of lipid
565     molecules are sterically (as well as electrostatically) anisotropic,
566     and would not pack in triangular arrangements without the steric
567     interference of adjacent molecular bodies. Since we only see rippled
568     phases in the neighborhood of $\gamma=\sqrt{3}$, this implies that
569     even if this dipolar mechanism is the correct explanation for the
570     ripple phase in realistic bilayers, there would still be a role played
571     by the lipid chains in the in-plane organization of the triangularly
572     ordered phases which could support ripples. The present model is
573     certainly not detailed enough to answer exactly what drives the
574     formation of the $P_{\beta'}$ phase in real lipids, but suggests some
575     avenues for further experiments.
576    
577     The most important prediction we can make using the results from this
578     simple model is that if dipolar ordering is driving the surface
579     corrugation, the wave vectors for the ripples should always found to
580     be {\it perpendicular} to the dipole director axis. This prediction
581     should suggest experimental designs which test whether this is really
582     true in the phosphatidylcholine $P_{\beta'}$ phases. The dipole
583     director axis should also be easily computable for the all-atom and
584     coarse-grained simulations that have been published in the literature.
585    
586     Our other observation about the ripple and dipolar directionality is
587     that the dipole director axis can be found to be parallel to any of
588     the three equivalent lattice vectors in the triangular lattice.
589     Defects in the ordering of the dipoles can cause the dipole director
590     (and consequently the surface corrugation) of small regions to be
591     rotated relative to each other by 120$^{\circ}$. This is a similar
592     behavior to the domain rotation seen in the AFM studies of Kaasgaard
593     {\it et al.}\cite{Kaasgaard03}
594    
595     Although our model is simple, it exhibits some rich and unexpected
596     behaviors. It would clearly be a closer approximation to the reality
597     if we allowed greater translational freedom to the dipoles and
598     replaced the somewhat artificial lattice packing and the harmonic
599     elastic tension with more realistic molecular modeling potentials.
600     What we have done is to present a simple model which exhibits bulk
601     non-thermal corrugation, and our explanation of this rippling
602     phenomenon will help us design more accurate molecular models for
603     corrugated membranes and experiments to test whether rippling is
604     dipole-driven or not.