1 |
\chapter{\label{chap:mc}SPONTANEOUS CORRUGATION OF DIPOLAR MEMBRANES} |
2 |
|
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} The X-ray diffraction |
33 |
work by Katsaras {\it et al.} showed that a rich phase diagram |
34 |
exhibiting both {\it asymmetric} and {\it symmetric} ripples is |
35 |
possible for lecithin bilayers.\cite{Katsaras00} Recently, Kaasgaard |
36 |
{\it et al.} used atomic force microscopy (AFM) to observe ripple |
37 |
phase morphology in bilayers supported on mica.~\cite{Kaasgaard03} The |
38 |
experimental results provide strong support for a 2-dimensional |
39 |
triangular packing lattice of the lipid molecules within the ripple |
40 |
phase. This is a notable change from the observed lipid packing |
41 |
within the gel phase,~\cite{Cevc87} although Tenchov {\it et al.} have |
42 |
recently observed near-hexagonal packing in some phosphatidylcholine |
43 |
(PC) gel phases.~\cite{Tenchov2001} There have been a number of |
44 |
theoretical |
45 |
approaches~\cite{Marder84,Carlson87,Goldstein88,McCullough90,Lubensky93,Misbah98,Heimburg00,Kubica02,Bannerjee02} |
46 |
(and some heroic |
47 |
simulations~\cite{Ayton02,Jiang04,Brannigan04a,deVries05,deJoannis06}) |
48 |
undertaken to try to explain this phase, but to date, none have looked |
49 |
specifically at the contribution of the dipolar character of the lipid |
50 |
head groups towards this corrugation. Lipid chain interdigitation |
51 |
certainly plays a major role, and the structures of the ripple phase |
52 |
are highly ordered. The model we investigate here lacks chain |
53 |
interdigitation (as well as the chains themselves!) and will not be |
54 |
detailed enough to rule in favor of (or against) any of these |
55 |
explanations for the $P_{\beta'}$ phase. |
56 |
|
57 |
Membranes containing electrostatic dipoles can also exhibit the |
58 |
flexoelectric effect,\cite{Todorova2004,Harden2006,Petrov2006} which |
59 |
is the ability of mechanical deformations to result in electrostatic |
60 |
organization of the membrane. This phenomenon is a curvature-induced |
61 |
membrane polarization which can lead to potential differences across a |
62 |
membrane. Reverse flexoelectric behavior (in which applied currents |
63 |
effect membrane curvature) has also been observed. Explanations of |
64 |
the details of these effects have typically utilized membrane |
65 |
polarization perpendicular to the face of the |
66 |
membrane,\cite{Petrov2006} and the effect has been observed in both |
67 |
biological,\cite{Raphael2000} bent-core liquid |
68 |
crystalline,\cite{Harden2006} and polymer-dispersed liquid crystalline |
69 |
membranes.\cite{Todorova2004} |
70 |
|
71 |
The problem with using atomistic and even coarse-grained approaches to |
72 |
study membrane buckling phenomena is that only a relatively small |
73 |
number of periods of the corrugation (i.e. one or two) can be |
74 |
realistically simulated given current technology. Also, simulations |
75 |
of lipid bilayers are traditionally carried out with periodic boundary |
76 |
conditions in two or three dimensions and these have the potential to |
77 |
enhance the periodicity of the system at that wavelength. To avoid |
78 |
this pitfall, we are using a model which allows us to have |
79 |
sufficiently large systems so that we are not causing artificial |
80 |
corrugation through the use of periodic boundary conditions. |
81 |
|
82 |
The simplest dipolar membrane is one in which the dipoles are located |
83 |
on fixed lattice sites. Ferroelectric states (with long-range dipolar |
84 |
order) can be observed in dipolar systems with non-triangular |
85 |
packings. However, {\em triangularly}-packed 2-D dipolar systems are |
86 |
inherently frustrated and one would expect a dipolar-disordered phase |
87 |
to be the lowest free energy |
88 |
configuration.\cite{Toulouse1977,Marland1979} Dipolar lattices already |
89 |
have rich phase behavior, but in order to allow the membrane to |
90 |
buckle, a single degree of freedom (translation normal to the membrane |
91 |
face) must be added to each of the dipoles. It would also be possible |
92 |
to allow complete translational freedom. This approach |
93 |
is similar in character to a number of elastic Ising models that have |
94 |
been developed to explain interesting mechanical properties in |
95 |
magnetic alloys.\cite{Renard1966,Zhu2005,Zhu2006,Jiang2006} |
96 |
|
97 |
What we present here is an attempt to find the simplest dipolar model |
98 |
which will exhibit buckling behavior. We are using a modified XYZ |
99 |
lattice model; details of the model can be found in section |
100 |
\ref{mc:sec:model}, results of Monte Carlo simulations using this model |
101 |
are presented in section |
102 |
\ref{mc:sec:results}, and section \ref{mc:sec:discussion} contains our conclusions. |
103 |
|
104 |
\section{2-D Dipolar Membrane} |
105 |
\label{mc:sec:model} |
106 |
|
107 |
The point of developing this model was to arrive at the simplest |
108 |
possible theoretical model which could exhibit spontaneous corrugation |
109 |
of a two-dimensional dipolar medium. Since molecules in polymerized |
110 |
membranes and in the $P_{\beta'}$ ripple phase have limited |
111 |
translational freedom, we have chosen a lattice to support the dipoles |
112 |
in the x-y plane. The lattice may be either triangular (lattice |
113 |
constants $a/b = |
114 |
\sqrt{3}$) or distorted. However, each dipole has 3 degrees of |
115 |
freedom. They may move freely {\em out} of the x-y plane (along the |
116 |
$z$ axis), and they have complete orientational freedom ($0 <= \theta |
117 |
<= \pi$, $0 <= \phi < 2 |
118 |
\pi$). This is essentially a modified X-Y-Z model with translational |
119 |
freedom along the z-axis. |
120 |
|
121 |
The potential energy of the system, |
122 |
\begin{equation} |
123 |
\begin{split} |
124 |
V = \sum_i &\left( \sum_{j>i} \frac{|\mu|^2}{4\pi \epsilon_0 r_{ij}^3} \left[ |
125 |
{\mathbf{\hat u}_i} \cdot {\mathbf{\hat u}_j} - |
126 |
3({\mathbf{\hat u}_i} \cdot {\mathbf{\hat |
127 |
r}_{ij}})({\mathbf{\hat u}_j} \cdot {\mathbf{\hat r}_{ij}})\right] \right. \\ |
128 |
& \left. + \sum_{j \in NN_i}^6 \frac{k_r}{2}\left( |
129 |
r_{ij}-\sigma \right)^2 \right) |
130 |
\end{split} |
131 |
\label{mceq:pot} |
132 |
\end{equation} |
133 |
|
134 |
In this potential, $\mathbf{\hat u}_i$ is the unit vector pointing |
135 |
along dipole $i$ and $\mathbf{\hat r}_{ij}$ is the unit vector |
136 |
pointing along the inter-dipole vector $\mathbf{r}_{ij}$. The entire |
137 |
potential is governed by three parameters, the dipolar strength |
138 |
($\mu$), the harmonic spring constant ($k_r$) and the preferred |
139 |
intermolecular spacing ($\sigma$). In practice, we set the value of |
140 |
$\sigma$ to the average inter-molecular spacing from the planar |
141 |
lattice, yielding a potential model that has only two parameters for a |
142 |
particular choice of lattice constants $a$ (along the $x$-axis) and |
143 |
$b$ (along the $y$-axis). We also define a set of reduced parameters |
144 |
based on the length scale ($\sigma$) and the energy of the harmonic |
145 |
potential at a deformation of 2 $\sigma$ ($\epsilon = k_r \sigma^2 / |
146 |
2$). Using these two constants, we perform our calculations using |
147 |
reduced distances, ($r^{*} = r / \sigma$), temperatures ($T^{*} = 2 |
148 |
k_B T / k_r \sigma^2$), densities ($\rho^{*} = N \sigma^2 / L_x L_y$), |
149 |
and dipole moments ($\mu^{*} = \mu / \sqrt{4 \pi \epsilon_0 \sigma^5 |
150 |
k_r / 2}$). It should be noted that the density ($\rho^{*}$) depends |
151 |
only on the mean particle spacing in the $x-y$ plane; the lattice is |
152 |
fully populated. |
153 |
|
154 |
To investigate the phase behavior of this model, we have performed a |
155 |
series of Metropolis Monte Carlo simulations of moderately-sized (34.3 |
156 |
$\sigma$ on a side) patches of membrane hosted on both triangular |
157 |
($\gamma = a/b = \sqrt{3}$) and distorted ($\gamma \neq \sqrt{3}$) |
158 |
lattices. The linear extent of one edge of the monolayer was $20 a$ |
159 |
and the system was kept roughly square. The average distance that |
160 |
coplanar dipoles were positioned from their six nearest neighbors was |
161 |
1 $\sigma$ (on both triangular and distorted lattices). Typical |
162 |
system sizes were 1360 dipoles for the triangular lattices and |
163 |
840-2800 dipoles for the distorted lattices. Two-dimensional periodic |
164 |
boundary conditions were used, and the cutoff for the dipole-dipole |
165 |
interaction was set to 4.3 $\sigma$. This cutoff is roughly 2.5 times |
166 |
the typical real-space electrostatic cutoff for molecular systems. |
167 |
Since dipole-dipole interactions decay rapidly with distance, and |
168 |
since the intrinsic three-dimensional periodicity of the Ewald sum can |
169 |
give artifacts in 2-d systems, we have chosen not to use it in these |
170 |
calculations. Although the Ewald sum has been reformulated to handle |
171 |
2-D systems,\cite{Parry75,Parry76,Heyes77,deLeeuw79,Rhee89} these |
172 |
methods are computationally expensive,\cite{Spohr97,Yeh99} and are not |
173 |
necessary in this case. All parameters ($T^{*}$, $\mu^{*}$, and |
174 |
$\gamma$) were varied systematically to study the effects of these |
175 |
parameters on the formation of ripple-like phases. |
176 |
|
177 |
\section{Results and Analysis} |
178 |
\label{mc:sec:results} |
179 |
|
180 |
\subsection{Dipolar Ordering and Coexistence Temperatures} |
181 |
The principal method for observing the orientational ordering |
182 |
transition in dipolar or liquid crystalline systems is the $P_2$ order |
183 |
parameter (defined as $1.5 \times \lambda_{max}$, where |
184 |
$\lambda_{max}$ is the largest eigenvalue of the matrix, |
185 |
\begin{equation} |
186 |
{\mathsf{S}} = \frac{1}{N} \sum_i \left( |
187 |
\begin{array}{ccc} |
188 |
u^{x}_i u^{x}_i-\frac{1}{3} & u^{x}_i u^{y}_i & u^{x}_i u^{z}_i \\ |
189 |
u^{y}_i u^{x}_i & u^{y}_i u^{y}_i -\frac{1}{3} & u^{y}_i u^{z}_i \\ |
190 |
u^{z}_i u^{x}_i & u^{z}_i u^{y}_i & u^{z}_i u^{z}_i -\frac{1}{3} |
191 |
\end{array} \right). |
192 |
\label{mceq:opmatrix} |
193 |
\end{equation} |
194 |
Here $u^{\alpha}_i$ is the $\alpha=x,y,z$ component of the unit vector |
195 |
for dipole $i$. $P_2$ will be $1.0$ for a perfectly-ordered system |
196 |
and near $0$ for a randomized system. Note that this order parameter |
197 |
is {\em not} equal to the polarization of the system. For example, |
198 |
the polarization of the perfect anti-ferroelectric system is $0$, but |
199 |
$P_2$ for an anti-ferroelectric system is $1$. The eigenvector of |
200 |
$\mathsf{S}$ corresponding to the largest eigenvalue is familiar as |
201 |
the director axis, which can be used to determine a privileged dipolar |
202 |
axis for dipole-ordered systems. The top panel in Fig. \ref{mcfig:phase} |
203 |
shows the values of $P_2$ as a function of temperature for both |
204 |
triangular ($\gamma = 1.732$) and distorted ($\gamma=1.875$) |
205 |
lattices. |
206 |
|
207 |
\begin{figure} |
208 |
\includegraphics[width=\linewidth]{./figures/mcPhase.pdf} |
209 |
\caption{\label{mcfig:phase} Top panel: The $P_2$ dipolar order parameter as |
210 |
a function of temperature for both triangular ($\gamma = 1.732$) and |
211 |
distorted ($\gamma = 1.875$) lattices. Bottom Panel: The phase |
212 |
diagram for the dipolar membrane model. The line denotes the division |
213 |
between the dipolar ordered (anti-ferroelectric) and disordered phases. |
214 |
An enlarged view near the triangular lattice is shown inset.} |
215 |
\end{figure} |
216 |
|
217 |
There is a clear order-disorder transition in evidence from this data. |
218 |
Both the triangular and distorted lattices have dipolar-ordered |
219 |
low-temperature phases, and orientationally-disordered high |
220 |
temperature phases. The coexistence temperature for the triangular |
221 |
lattice is significantly lower than for the distorted lattices, and |
222 |
the bulk polarization is approximately $0$ for both dipolar ordered |
223 |
and disordered phases. This gives strong evidence that the dipolar |
224 |
ordered phase is anti-ferroelectric. We have verified that this |
225 |
dipolar ordering transition is not a function of system size by |
226 |
performing identical calculations with systems twice as large. The |
227 |
transition is equally smooth at all system sizes that were studied. |
228 |
Additionally, we have repeated the Monte Carlo simulations over a wide |
229 |
range of lattice ratios ($\gamma$) to generate a dipolar |
230 |
order/disorder phase diagram. The bottom panel in Fig. \ref{mcfig:phase} |
231 |
shows that the triangular lattice is a low-temperature cusp in the |
232 |
$T^{*}-\gamma$ phase diagram. |
233 |
|
234 |
This phase diagram is remarkable in that it shows an |
235 |
anti-ferroelectric phase near $\gamma=1.732$ where one would expect |
236 |
lattice frustration to result in disordered phases at all |
237 |
temperatures. Observations of the configurations in this phase show |
238 |
clearly that the system has accomplished dipolar ordering by forming |
239 |
large ripple-like structures. We have observed anti-ferroelectric |
240 |
ordering in all three of the equivalent directions on the triangular |
241 |
lattice, and the dipoles have been observed to organize perpendicular |
242 |
to the membrane normal (in the plane of the membrane). It is |
243 |
particularly interesting to note that the ripple-like structures have |
244 |
also been observed to propagate in the three equivalent directions on |
245 |
the lattice, but the {\em direction of ripple propagation is always |
246 |
perpendicular to the dipole director axis}. A snapshot of a typical |
247 |
anti-ferroelectric rippled structure is shown in |
248 |
Fig. \ref{mcfig:snapshot}. |
249 |
|
250 |
\begin{figure} |
251 |
\includegraphics[width=\linewidth]{./figures/mcSnapshot.pdf} |
252 |
\caption{\label{mcfig:snapshot} Top and Side views of a representative |
253 |
configuration for the dipolar ordered phase supported on the |
254 |
triangular lattice. Note the anti-ferroelectric ordering and the long |
255 |
wavelength buckling of the membrane. Dipolar ordering has been |
256 |
observed in all three equivalent directions on the triangular lattice, |
257 |
and the ripple direction is always perpendicular to the director axis |
258 |
for the dipoles.} |
259 |
\end{figure} |
260 |
|
261 |
Although the snapshot in Fig. \ref{mcfig:snapshot} gives the appearance |
262 |
of three-row stair-like structures, these appear to be transient. On |
263 |
average, the corrugation of the membrane is a relatively smooth, |
264 |
long-wavelength phenomenon, with occasional steep drops between |
265 |
adjacent lines of anti-aligned dipoles. |
266 |
|
267 |
The height-dipole correlation function ($C_{\textrm{hd}}(r, \cos |
268 |
\theta)$) makes the connection between dipolar ordering and the wave |
269 |
vector of the ripple even more explicit. $C_{\textrm{hd}}(r, \cos |
270 |
\theta)$ is an angle-dependent pair distribution function. The angle |
271 |
($\theta$) is the angle between the intermolecular vector |
272 |
$\vec{r}_{ij}$ and direction of dipole $i$, |
273 |
\begin{equation} |
274 |
C_{\textrm{hd}} = \frac{\langle \frac{1}{n(r)} \sum_{i}\sum_{j>i} |
275 |
h_i \cdot h_j \delta(r - r_{ij}) \delta(\cos \theta_{ij} - |
276 |
\cos \theta)\rangle} {\langle h^2 \rangle} |
277 |
\end{equation} |
278 |
where $\cos \theta_{ij} = \hat{\mu}_{i} \cdot \hat{r}_{ij}$ and |
279 |
$\hat{r}_{ij} = \vec{r}_{ij} / r_{ij}$. $n(r)$ is the number of |
280 |
dipoles found in a cylindrical shell between $r$ and $r+\delta r$ of |
281 |
the central particle. Fig. \ref{mcfig:CrossCorrelation} shows contours |
282 |
of this correlation function for both anti-ferroelectric, rippled |
283 |
membranes as well as for the dipole-disordered portion of the phase |
284 |
diagram. |
285 |
|
286 |
\begin{figure} |
287 |
\includegraphics[width=\linewidth]{./figures/mcHdc.pdf} |
288 |
\caption{\label{mcfig:CrossCorrelation} Contours of the height-dipole |
289 |
correlation function as a function of the dot product between the |
290 |
dipole ($\hat{\mu}$) and inter-dipole separation vector ($\hat{r}$) |
291 |
and the distance ($r$) between the dipoles. Perfect height |
292 |
correlation (contours approaching 1) are present in the ordered phase |
293 |
when the two dipoles are in the same head-to-tail line. |
294 |
Anti-correlation (contours below 0) is only seen when the inter-dipole |
295 |
vector is perpendicular to the dipoles. In the dipole-disordered |
296 |
portion of the phase diagram, there is only weak correlation in the |
297 |
dipole direction and this correlation decays rapidly to zero for |
298 |
intermolecular vectors that are not dipole-aligned.} |
299 |
\end{figure} |
300 |
|
301 |
The height-dipole correlation function gives a map of how the topology |
302 |
of the membrane surface varies with angular deviation around a given |
303 |
dipole. The upper panel of Fig. \ref{mcfig:CrossCorrelation} shows that |
304 |
in the anti-ferroelectric phase, the dipole heights are strongly |
305 |
correlated for dipoles in head-to-tail arrangements, and this |
306 |
correlation persists for very long distances (up to 15 $\sigma$). For |
307 |
portions of the membrane located perpendicular to a given dipole, the |
308 |
membrane height becomes anti-correlated at distances of 10 $\sigma$. |
309 |
The correlation function is relatively smooth; there are no steep |
310 |
jumps or steps, so the stair-like structures in |
311 |
Fig. \ref{mcfig:snapshot} are indeed transient and disappear when |
312 |
averaged over many configurations. In the dipole-disordered phase, |
313 |
the height-dipole correlation function is relatively flat (and hovers |
314 |
near zero). The only significant height correlations are for axial |
315 |
dipoles at very short distances ($r \approx |
316 |
\sigma$). |
317 |
|
318 |
\subsection{Discriminating Ripples from Thermal Undulations} |
319 |
|
320 |
In order to be sure that the structures we have observed are actually |
321 |
a rippled phase and not simply thermal undulations, we have computed |
322 |
the undulation spectrum, |
323 |
\begin{equation} |
324 |
h(\vec{q}) = A^{-1/2} \int d\vec{r} |
325 |
h(\vec{r}) e^{-i \vec{q}\cdot\vec{r}} |
326 |
\end{equation} |
327 |
where $h(\vec{r})$ is the height of the membrane at location $\vec{r} |
328 |
= (x,y)$.~\cite{Safran94,Seifert97} In simple (and more complicated) |
329 |
elastic continuum models, it can shown that in the $NVT$ ensemble, the |
330 |
absolute value of the undulation spectrum can be written, |
331 |
\begin{equation} |
332 |
\langle | h(q) |^2 \rangle_{NVT} = \frac{k_B T}{k_c q^4 + |
333 |
\gamma q^2}, |
334 |
\label{mceq:fit} |
335 |
\end{equation} |
336 |
where $k_c$ is the bending modulus for the membrane, and $\gamma$ is |
337 |
the mechanical surface tension.~\cite{Safran94} The systems studied in |
338 |
this paper have essentially zero bending moduli ($k_c$) and relatively |
339 |
large mechanical surface tensions ($\gamma$), so a much simpler form |
340 |
can be written, |
341 |
\begin{equation} |
342 |
\langle | h(q) |^2 \rangle_{NVT} = \frac{k_B T}{\gamma q^2}. |
343 |
\label{mceq:fit2} |
344 |
\end{equation} |
345 |
|
346 |
The undulation spectrum is computed by superimposing a rectangular |
347 |
grid on top of the membrane, and by assigning height ($h(\vec{r})$) |
348 |
values to the grid from the average of all dipoles that fall within a |
349 |
given $\vec{r}+d\vec{r}$ grid area. Empty grid pixels are assigned |
350 |
height values by interpolation from the nearest neighbor pixels. A |
351 |
standard 2-d Fourier transform is then used to obtain $\langle | |
352 |
h(q)|^2 \rangle$. Alternatively, since the dipoles sit on a Bravais |
353 |
lattice, one could use the heights of the lattice points themselves as |
354 |
the grid for the Fourier transform (without interpolating to a square |
355 |
grid). However, if lateral translational freedom is added to this |
356 |
model (a likely extension), an interpolated grid method for computing |
357 |
undulation spectra will be required. |
358 |
|
359 |
As mentioned above, the best fits to our undulation spectra are |
360 |
obtained by setting the value of $k_c$ to 0. In Fig. \ref{mcfig:fit} we |
361 |
show typical undulation spectra for two different regions of the phase |
362 |
diagram along with their fits from the Landau free energy approach |
363 |
(Eq. \ref{mceq:fit2}). In the high-temperature disordered phase, the |
364 |
Landau fits can be nearly perfect, and from these fits we can estimate |
365 |
the tension in the surface. In reduced units, typical values of |
366 |
$\gamma^{*} = \gamma / \epsilon = 2500$ are obtained for the |
367 |
disordered phase ($\gamma^{*} = 2551.7$ in the top panel of |
368 |
Fig. \ref{mcfig:fit}). |
369 |
|
370 |
Typical values of $\gamma^{*}$ in the dipolar-ordered phase are much |
371 |
higher than in the dipolar-disordered phase ($\gamma^{*} = 73,538$ in |
372 |
the lower panel of Fig. \ref{mcfig:fit}). For the dipolar-ordered |
373 |
triangular lattice near the coexistence temperature, we also observe |
374 |
long wavelength undulations that are far outliers to the fits. That |
375 |
is, the Landau free energy fits are well within error bars for most of |
376 |
the other points, but can be off by {\em orders of magnitude} for a |
377 |
few low frequency components. |
378 |
|
379 |
We interpret these outliers as evidence that these low frequency modes |
380 |
are {\em non-thermal undulations}. We take this as evidence that we |
381 |
are actually seeing a rippled phase developing in this model system. |
382 |
|
383 |
\begin{figure} |
384 |
\includegraphics[width=\linewidth]{./figures/mcLogFit.pdf} |
385 |
\caption{\label{mcfig:fit} Evidence that the observed ripples are {\em |
386 |
not} thermal undulations is obtained from the 2-d Fourier transform |
387 |
$\langle |h^{*}(\vec{q})|^2 \rangle$ of the height profile ($\langle |
388 |
h^{*}(x,y) \rangle$). Rippled samples show low-wavelength peaks that |
389 |
are outliers on the Landau free energy fits by an order of magnitude. |
390 |
Samples exhibiting only thermal undulations fit Eq. \ref{mceq:fit} |
391 |
remarkably well.} |
392 |
\end{figure} |
393 |
|
394 |
\subsection{Effects of Potential Parameters on Amplitude and Wavelength} |
395 |
|
396 |
We have used two different methods to estimate the amplitude and |
397 |
periodicity of the ripples. The first method requires projection of |
398 |
the ripples onto a one dimensional rippling axis. Since the rippling |
399 |
is always perpendicular to the dipole director axis, we can define a |
400 |
ripple vector as follows. The largest eigenvector, $s_1$, of the |
401 |
$\mathsf{S}$ matrix in Eq. \ref{mceq:opmatrix} is projected onto a |
402 |
planar director axis, |
403 |
\begin{equation} |
404 |
\vec{d} = \left(\begin{array}{c} |
405 |
\vec{s}_1 \cdot \hat{i} \\ |
406 |
\vec{s}_1 \cdot \hat{j} \\ |
407 |
0 |
408 |
\end{array} \right). |
409 |
\end{equation} |
410 |
($\hat{i}$, $\hat{j}$ and $\hat{k}$ are unit vectors along the $x$, |
411 |
$y$, and $z$ axes, respectively.) The rippling axis is in the plane of |
412 |
the membrane and is perpendicular to the planar director axis, |
413 |
\begin{equation} |
414 |
\vec{q}_{\mathrm{rip}} = \vec{d} \times \hat{k} |
415 |
\end{equation} |
416 |
We can then find the height profile of the membrane along the ripple |
417 |
axis by projecting heights of the dipoles to obtain a one-dimensional |
418 |
height profile, $h(q_{\mathrm{rip}})$. Ripple wavelengths can be |
419 |
estimated from the largest non-thermal low-frequency component in the |
420 |
Fourier transform of $h(q_{\mathrm{rip}})$. Amplitudes can be |
421 |
estimated by measuring peak-to-trough distances in |
422 |
$h(q_{\mathrm{rip}})$ itself. |
423 |
|
424 |
A second, more accurate, and simpler method for estimating ripple |
425 |
shape is to extract the wavelength and height information directly |
426 |
from the largest non-thermal peak in the undulation spectrum. For |
427 |
large-amplitude ripples, the two methods give similar results. The |
428 |
one-dimensional projection method is more prone to noise (particularly |
429 |
in the amplitude estimates for the distorted lattices). We report |
430 |
amplitudes and wavelengths taken directly from the undulation spectrum |
431 |
below. |
432 |
|
433 |
In the triangular lattice ($\gamma = \sqrt{3}$), the rippling is |
434 |
observed for temperatures ($T^{*}$) from $61-122$. The wavelength of |
435 |
the ripples is remarkably stable at 21.4~$\sigma$ for all but the |
436 |
temperatures closest to the order-disorder transition. At $T^{*} = |
437 |
122$, the wavelength drops to 17.1~$\sigma$. |
438 |
|
439 |
The dependence of the amplitude on temperature is shown in the top |
440 |
panel of Fig. \ref{mcfig:Amplitude}. The rippled structures shrink |
441 |
smoothly as the temperature rises towards the order-disorder |
442 |
transition. The wavelengths and amplitudes we observe are |
443 |
surprisingly close to the $\Lambda / 2$ phase observed by Kaasgaard |
444 |
{\it et al.} in their work on PC-based lipids.\cite{Kaasgaard03} |
445 |
However, this is coincidental agreement based on a choice of 7~\AA~as |
446 |
the mean spacing between lipids. |
447 |
|
448 |
\begin{figure} |
449 |
\includegraphics[width=\linewidth]{./figures/mcProperties_sq.pdf} |
450 |
\caption{\label{mcfig:Amplitude} a) The amplitude $A^{*}$ of the ripples |
451 |
vs. temperature for a triangular lattice. b) The amplitude $A^{*}$ of |
452 |
the ripples vs. dipole strength ($\mu^{*}$) for both the triangular |
453 |
lattice (circles) and distorted lattice (squares). The reduced |
454 |
temperatures were kept fixed at $T^{*} = 94$ for the triangular |
455 |
lattice and $T^{*} = 106$ for the distorted lattice (approximately 2/3 |
456 |
of the order-disorder transition temperature for each lattice).} |
457 |
\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 |
\caption{\label{mcfig:systemsize} The ripple wavelength (top) and |
510 |
amplitude (bottom) as a function of system size for a triangular |
511 |
lattice ($\gamma=1.732$) at $T^{*} = 122$.} |
512 |
\end{figure} |
513 |
|
514 |
There is substantial dependence on system size for small (less than |
515 |
29~$\sigma$) periodic boxes. Notably, there are resonances apparent |
516 |
in the ripple amplitudes at box lengths of 17.3 and 29.5 $\sigma$. |
517 |
For larger systems, the behavior of the ripples appears to have |
518 |
stabilized and is on a trend to slightly smaller amplitudes (and |
519 |
slightly longer wavelengths) than were observed from the 34.3 $\sigma$ |
520 |
box sizes that were used for most of the calculations. |
521 |
|
522 |
It is interesting to note that system sizes which are multiples of the |
523 |
default ripple wavelength can enhance the amplitude of the observed |
524 |
ripples, but appears to have only a minor effect on the observed |
525 |
wavelength. It would, of course, be better to use system sizes that |
526 |
were many multiples of the ripple wavelength to be sure that the |
527 |
periodic box is not driving the phenomenon, but at the largest system |
528 |
size studied (70 $\sigma$ $\times$ 70 $\sigma$), the number of dipoles |
529 |
(5440) made long Monte Carlo simulations prohibitively expensive. |
530 |
|
531 |
\section{Discussion} |
532 |
\label{mc:sec:discussion} |
533 |
|
534 |
We have been able to show that a simple dipolar lattice model which |
535 |
contains only molecular packing (from the lattice), anisotropy (in the |
536 |
form of electrostatic dipoles) and a weak elastic tension (in the form |
537 |
of a nearest-neighbor harmonic potential) is capable of exhibiting |
538 |
stable long-wavelength non-thermal surface corrugations. The best |
539 |
explanation for this behavior is that the ability of the dipoles to |
540 |
translate out of the plane of the membrane is enough to break the |
541 |
symmetry of the triangular lattice and allow the energetic benefit |
542 |
from the formation of a bulk anti-ferroelectric phase. Were the weak |
543 |
elastic tension absent from our model, it would be possible for the |
544 |
entire lattice to ``tilt'' using $z$-translation. Tilting the lattice |
545 |
in this way would yield an effectively non-triangular lattice which |
546 |
would avoid dipolar frustration altogether. With the elastic tension |
547 |
in place, bulk tilt causes a large strain, and the least costly way to |
548 |
release this strain is between two rows of anti-aligned dipoles. |
549 |
These ``breaks'' will result in rippled or sawtooth patterns in the |
550 |
membrane, and allow small stripes of membrane to form |
551 |
anti-ferroelectric regions that are tilted relative to the averaged |
552 |
membrane normal. |
553 |
|
554 |
Although the dipole-dipole interaction is the major driving force for |
555 |
the long range orientational ordered state, the formation of the |
556 |
stable, smooth ripples is a result of the competition between the |
557 |
elastic tension and the dipole-dipole interactions. This statement is |
558 |
supported by the variation in $\mu^{*}$. Substantially weaker dipoles |
559 |
relative to the surface tension can cause the corrugated phase to |
560 |
disappear. |
561 |
|
562 |
The packing of the dipoles into a nearly-triangular lattice is clearly |
563 |
an important piece of the puzzle. The dipolar head groups of lipid |
564 |
molecules are sterically (as well as electrostatically) anisotropic, |
565 |
and would not pack in triangular arrangements without the steric |
566 |
interference of adjacent molecular bodies. Since we only see rippled |
567 |
phases in the neighborhood of $\gamma=\sqrt{3}$, this implies that |
568 |
even if this dipolar mechanism is the correct explanation for the |
569 |
ripple phase in realistic bilayers, there would still be a role played |
570 |
by the lipid chains in the in-plane organization of the triangularly |
571 |
ordered phases which could support ripples. The present model is |
572 |
certainly not detailed enough to answer exactly what drives the |
573 |
formation of the $P_{\beta'}$ phase in real lipids, but suggests some |
574 |
avenues for further experiments. |
575 |
|
576 |
The most important prediction we can make using the results from this |
577 |
simple model is that if dipolar ordering is driving the surface |
578 |
corrugation, the wave vectors for the ripples should always found to |
579 |
be {\it perpendicular} to the dipole director axis. This prediction |
580 |
should suggest experimental designs which test whether this is really |
581 |
true in the phosphatidylcholine $P_{\beta'}$ phases. The dipole |
582 |
director axis should also be easily computable for the all-atom and |
583 |
coarse-grained simulations that have been published in the literature. |
584 |
|
585 |
Our other observation about the ripple and dipolar directionality is |
586 |
that the dipole director axis can be found to be parallel to any of |
587 |
the three equivalent lattice vectors in the triangular lattice. |
588 |
Defects in the ordering of the dipoles can cause the dipole director |
589 |
(and consequently the surface corrugation) of small regions to be |
590 |
rotated relative to each other by 120$^{\circ}$. This is a similar |
591 |
behavior to the domain rotation seen in the AFM studies of Kaasgaard |
592 |
{\it et al.}\cite{Kaasgaard03} |
593 |
|
594 |
Although our model is simple, it exhibits some rich and unexpected |
595 |
behaviors. It would clearly be a closer approximation to the reality |
596 |
if we allowed greater translational freedom to the dipoles and |
597 |
replaced the somewhat artificial lattice packing and the harmonic |
598 |
elastic tension with more realistic molecular modeling potentials. |
599 |
What we have done is to present a simple model which exhibits bulk |
600 |
non-thermal corrugation, and our explanation of this rippling |
601 |
phenomenon will help us design more accurate molecular models for |
602 |
corrugated membranes and experiments to test whether rippling is |
603 |
dipole-driven or not. |