ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/xDissertation/md.tex
Revision: 3361
Committed: Thu Mar 6 23:22:15 2008 UTC (16 years, 6 months ago) by xsun
Content type: application/x-tex
File size: 46359 byte(s)
Log Message:
writing the dissertation.

File Contents

# User Rev Content
1 xsun 3360 \chapter{\label{chap:md}DIPOLAR ORDERING IN THE RIPPLE PHASES OF
2     MOLECULAR-SCALE MODELS OF LIPID MEMBRANES}
3 xsun 3358
4     \section{Introduction}
5     \label{mdsec:Int}
6    
7     A number of theoretical models have been presented to explain the
8     formation of the ripple phase. Marder {\it et al.} used a
9     curvature-dependent Landau-de~Gennes free-energy functional to predict
10     a rippled phase.~\cite{Marder84} This model and other related
11     continuum models predict higher fluidity in convex regions and that
12     concave portions of the membrane correspond to more solid-like
13     regions. Carlson and Sethna used a packing-competition model (in
14     which head groups and chains have competing packing energetics) to
15 xsun 3361 predict the formation of a ripple-like phase~\cite{Carlson87}. Their
16     model predicted that the high-curvature portions have lower-chain
17     packing and correspond to more fluid-like regions. Goldstein and
18     Leibler used a mean-field approach with a planar model for {\em
19     inter-lamellar} interactions to predict rippling in multilamellar
20 xsun 3358 phases.~\cite{Goldstein88} McCullough and Scott proposed that the {\em
21     anisotropy of the nearest-neighbor interactions} coupled to
22     hydrophobic constraining forces which restrict height differences
23     between nearest neighbors is the origin of the ripple
24     phase.~\cite{McCullough90} Lubensky and MacKintosh introduced a Landau
25     theory for tilt order and curvature of a single membrane and concluded
26     that {\em coupling of molecular tilt to membrane curvature} is
27     responsible for the production of ripples.~\cite{Lubensky93} Misbah,
28     Duplat and Houchmandzadeh proposed that {\em inter-layer dipolar
29     interactions} can lead to ripple instabilities.~\cite{Misbah98}
30     Heimburg presented a {\em coexistence model} for ripple formation in
31     which he postulates that fluid-phase line defects cause sharp
32     curvature between relatively flat gel-phase regions.~\cite{Heimburg00}
33     Kubica has suggested that a lattice model of polar head groups could
34     be valuable in trying to understand bilayer phase
35     formation.~\cite{Kubica02} Bannerjee used Monte Carlo simulations of
36     lamellar stacks of hexagonal lattices to show that large head groups
37     and molecular tilt with respect to the membrane normal vector can
38     cause bulk rippling.~\cite{Bannerjee02} Recently, Kranenburg and Smit
39     described the formation of symmetric ripple-like structures using a
40     coarse grained solvent-head-tail bead model.\cite{Kranenburg2005}
41     Their lipids consisted of a short chain of head beads tied to the two
42 xsun 3361 longer ``chains''.
43 xsun 3358
44     In contrast, few large-scale molecular modeling studies have been
45     done due to the large size of the resulting structures and the time
46     required for the phases of interest to develop. With all-atom (and
47     even unified-atom) simulations, only one period of the ripple can be
48     observed and only for time scales in the range of 10-100 ns. One of
49     the most interesting molecular simulations was carried out by de~Vries
50     {\it et al.}~\cite{deVries05}. According to their simulation results,
51     the ripple consists of two domains, one resembling the gel bilayer,
52     while in the other, the two leaves of the bilayer are fully
53     interdigitated. The mechanism for the formation of the ripple phase
54     suggested by their work is a packing competition between the head
55     groups and the tails of the lipid molecules.\cite{Carlson87} Recently,
56     the ripple phase has also been studied by Lenz and Schmid using Monte
57     Carlo simulations.\cite{Lenz07} Their structures are similar to the De
58     Vries {\it et al.} structures except that the connection between the
59     two leaves of the bilayer is a narrow interdigitated line instead of
60     the fully interdigitated domain. The symmetric ripple phase was also
61     observed by Lenz {\it et al.}, and their work supports other claims
62     that the mismatch between the size of the head group and tail of the
63     lipid molecules is the driving force for the formation of the ripple
64     phase. Ayton and Voth have found significant undulations in
65     zero-surface-tension states of membranes simulated via dissipative
66     particle dynamics, but their results are consistent with purely
67     thermal undulations.~\cite{Ayton02}
68    
69     Although the organization of the tails of lipid molecules are
70     addressed by these molecular simulations and the packing competition
71     between head groups and tails is strongly implicated as the primary
72     driving force for ripple formation, questions about the ordering of
73     the head groups in ripple phase have not been settled.
74    
75 xsun 3361 In Ch.~\ref{chap:mc}, we presented a simple ``web of dipoles'' spin
76 xsun 3358 lattice model which provides some physical insight into relationship
77 xsun 3359 between dipolar ordering and membrane buckling.\cite{sun:031602} We
78     found that dipolar elastic membranes can spontaneously buckle, forming
79 xsun 3358 ripple-like topologies. The driving force for the buckling of dipolar
80     elastic membranes is the anti-ferroelectric ordering of the dipoles.
81     This was evident in the ordering of the dipole director axis
82     perpendicular to the wave vector of the surface ripples. A similar
83     phenomenon has also been observed by Tsonchev {\it et al.} in their
84     work on the spontaneous formation of dipolar peptide chains into
85     curved nano-structures.\cite{Tsonchev04,Tsonchev04II}
86    
87 xsun 3361 In this chapter, we construct a somewhat more realistic molecular-scale
88 xsun 3358 lipid model than our previous ``web of dipoles'' and use molecular
89     dynamics simulations to elucidate the role of the head group dipoles
90     in the formation and morphology of the ripple phase. We describe our
91     model and computational methodology in section \ref{mdsec:method}.
92     Details on the simulations are presented in section
93     \ref{mdsec:experiment}, with results following in section
94     \ref{mdsec:results}. A final discussion of the role of dipolar heads in
95     the ripple formation can be found in section
96     \ref{mdsec:discussion}.
97    
98     \section{Computational Model}
99     \label{mdsec:method}
100    
101     \begin{figure}
102     \centering
103     \includegraphics[width=\linewidth]{./figures/mdLipidModels.pdf}
104     \caption{Three different representations of DPPC lipid molecules,
105     including the chemical structure, an atomistic model, and the
106     head-body ellipsoidal coarse-grained model used in this
107     work.\label{mdfig:lipidModels}}
108     \end{figure}
109    
110     Our simple molecular-scale lipid model for studying the ripple phase
111     is based on two facts: one is that the most essential feature of lipid
112     molecules is their amphiphilic structure with polar head groups and
113     non-polar tails. Another fact is that the majority of lipid molecules
114     in the ripple phase are relatively rigid (i.e. gel-like) which makes
115     some fraction of the details of the chain dynamics negligible. Figure
116     \ref{mdfig:lipidModels} shows the molecular structure of a DPPC
117     molecule, as well as atomistic and molecular-scale representations of
118     a DPPC molecule. The hydrophilic character of the head group is
119     largely due to the separation of charge between the nitrogen and
120     phosphate groups. The zwitterionic nature of the PC headgroups leads
121     to abnormally large dipole moments (as high as 20.6 D), and this
122     strongly polar head group interacts strongly with the solvating water
123     layers immediately surrounding the membrane. The hydrophobic tail
124     consists of fatty acid chains. In our molecular scale model, lipid
125     molecules have been reduced to these essential features; the fatty
126     acid chains are represented by an ellipsoid with a dipolar ball
127     perched on one end to represent the effects of the charge-separated
128     head group. In real PC lipids, the direction of the dipole is
129     nearly perpendicular to the tail, so we have fixed the direction of
130     the point dipole rigidly in this orientation.
131    
132     The ellipsoidal portions of the model interact via the Gay-Berne
133     potential which has seen widespread use in the liquid crystal
134     community. Ayton and Voth have also used Gay-Berne ellipsoids for
135     modeling large length-scale properties of lipid
136     bilayers.\cite{Ayton01} In its original form, the Gay-Berne potential
137     was a single site model for the interactions of rigid ellipsoidal
138 xsun 3359 molecules.\cite{Gay1981} It can be thought of as a modification of the
139 xsun 3358 Gaussian overlap model originally described by Berne and
140     Pechukas.\cite{Berne72} The potential is constructed in the familiar
141     form of the Lennard-Jones function using orientation-dependent
142     $\sigma$ and $\epsilon$ parameters,
143 xsun 3361 \begin{equation}
144     \begin{split}
145 xsun 3358 V_{ij}({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j}, {\mathbf{\hat
146 xsun 3361 r}_{ij}}) = & 4\epsilon ({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j},
147     {\mathbf{\hat r}_{ij}})\left[ \left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u}_i},
148     {\mathbf{\hat u}_j}, {\mathbf{\hat r}_{ij}})+\sigma_0}\right)^{12} \right.\\
149     &\left. -\left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u}_i},
150     {\mathbf{\hat u}_j}, {\mathbf{\hat
151     r}_{ij}})+\sigma_0}\right)^6\right]
152     \end{split}
153 xsun 3358 \label{mdeq:gb}
154 xsun 3361 \end{equation}
155 xsun 3358
156     The range $(\sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
157     \hat{r}}_{ij}))$, and strength $(\epsilon({\bf \hat{u}}_{i},{\bf
158     \hat{u}}_{j},{\bf \hat{r}}_{ij}))$ parameters
159     are dependent on the relative orientations of the two molecules (${\bf
160     \hat{u}}_{i},{\bf \hat{u}}_{j}$) as well as the direction of the
161     intermolecular separation (${\bf \hat{r}}_{ij}$). $\sigma$ and
162     $\sigma_0$ are also governed by shape mixing and anisotropy variables,
163     \begin {eqnarray*}
164     \sigma_0 & = & \sqrt{d_i^2 + d_j^2} \\ \\
165     \chi & = & \left[ \frac{\left( l_i^2 - d_i^2 \right) \left(l_j^2 -
166     d_j^2 \right)}{\left( l_j^2 + d_i^2 \right) \left(l_i^2 +
167     d_j^2 \right)}\right]^{1/2} \\ \\
168     \alpha^2 & = & \left[ \frac{\left( l_i^2 - d_i^2 \right) \left(l_j^2 +
169     d_i^2 \right)}{\left( l_j^2 - d_j^2 \right) \left(l_i^2 +
170     d_j^2 \right)}\right]^{1/2},
171     \end{eqnarray*}
172     where $l$ and $d$ describe the length and width of each uniaxial
173     ellipsoid. These shape anisotropy parameters can then be used to
174     calculate the range function,
175 xsun 3361 \begin{equation}
176     \begin{split}
177     & \sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf \hat{r}}_{ij}) =
178     \sigma_{0} \times \\
179     & \left[ 1- \left\{ \frac{ \chi \alpha^2 ({\bf \hat{u}}_i \cdot {\bf
180 xsun 3358 \hat{r}}_{ij} ) + \chi \alpha^{-2} ({\bf \hat{u}}_j \cdot {\bf
181     \hat{r}}_{ij} ) - 2 \chi^2 ({\bf \hat{u}}_i \cdot {\bf
182     \hat{r}}_{ij} )({\bf \hat{u}}_j \cdot {\bf
183     \hat{r}}_{ij} ) ({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j)}{1 - \chi^2
184     \left({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j\right)^2} \right\}
185 xsun 3361 \right]^{-1/2}
186     \end{split}
187     \end{equation}
188 xsun 3358
189     Gay-Berne ellipsoids also have an energy scaling parameter,
190     $\epsilon^s$, which describes the well depth for two identical
191     ellipsoids in a {\it side-by-side} configuration. Additionally, a well
192     depth aspect ratio, $\epsilon^r = \epsilon^e / \epsilon^s$, describes
193     the ratio between the well depths in the {\it end-to-end} and
194     side-by-side configurations. As in the range parameter, a set of
195     mixing and anisotropy variables can be used to describe the well
196     depths for dissimilar particles,
197     \begin {eqnarray*}
198     \epsilon_0 & = & \sqrt{\epsilon^s_i * \epsilon^s_j} \\ \\
199     \epsilon^r & = & \sqrt{\epsilon^r_i * \epsilon^r_j} \\ \\
200     \chi' & = & \frac{1 - (\epsilon^r)^{1/\mu}}{1 + (\epsilon^r)^{1/\mu}}
201     \\ \\
202     \alpha'^2 & = & \left[1 + (\epsilon^r)^{1/\mu}\right]^{-1}
203     \end{eqnarray*}
204     The form of the strength function is somewhat complicated,
205 xsun 3361 \begin{eqnarray*}
206 xsun 3358 \epsilon({\bf \hat{u}}_{i}, {\bf \hat{u}}_{j},{\bf \hat{r}}_{ij}) & = &
207     \epsilon_{0} \epsilon_{1}^{\nu}({\bf \hat{u}}_{i}.{\bf \hat{u}}_{j})
208     \epsilon_{2}^{\mu}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
209     \hat{r}}_{ij}) \\ \\
210     \epsilon_{1}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j}) & = &
211     \left[1-\chi^{2}({\bf \hat{u}}_{i}.{\bf
212 xsun 3361 \hat{u}}_{j})^{2}\right]^{-1/2}
213     \end{eqnarray*}
214     \begin{equation*}
215     \begin{split}
216     & \epsilon_{2}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf \hat{r}}_{ij})
217     = 1 - \\
218     & \left\{ \frac{ \chi' \alpha'^2 ({\bf \hat{u}}_i \cdot {\bf
219 xsun 3358 \hat{r}}_{ij} ) + \chi' \alpha'^{-2} ({\bf \hat{u}}_j \cdot {\bf
220     \hat{r}}_{ij} ) - 2 \chi'^2 ({\bf \hat{u}}_i \cdot {\bf
221     \hat{r}}_{ij} )({\bf \hat{u}}_j \cdot {\bf
222     \hat{r}}_{ij} ) ({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j)}{1 - \chi'^2
223     \left({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j\right)^2} \right\},
224 xsun 3361 \end{split}
225     \end{equation*}
226 xsun 3358 although many of the quantities and derivatives are identical with
227     those obtained for the range parameter. Ref. \citen{Luckhurst90}
228     has a particularly good explanation of the choice of the Gay-Berne
229     parameters $\mu$ and $\nu$ for modeling liquid crystal molecules. An
230     excellent overview of the computational methods that can be used to
231     efficiently compute forces and torques for this potential can be found
232     in Ref. \citen{Golubkov06}
233    
234     The choices of parameters we have used in this study correspond to a
235     shape anisotropy of 3 for the chain portion of the molecule. In
236     principle, this could be varied to allow for modeling of longer or
237     shorter chain lipid molecules. For these prolate ellipsoids, we have:
238     \begin{equation}
239     \begin{array}{rcl}
240     d & < & l \\
241     \epsilon^{r} & < & 1
242     \end{array}
243     \end{equation}
244     A sketch of the various structural elements of our molecular-scale
245     lipid / solvent model is shown in figure \ref{mdfig:lipidModel}. The
246     actual parameters used in our simulations are given in table
247     \ref{mdtab:parameters}.
248    
249     \begin{figure}
250     \centering
251     \includegraphics[width=\linewidth]{./figures/md2LipidModel.pdf}
252     \caption{The parameters defining the behavior of the lipid
253     models. $\sigma_h / d$ is the ratio of the head group to body diameter.
254     Molecular bodies had a fixed aspect ratio of 3.0. The solvent model
255     was a simplified 4-water bead ($\sigma_w \approx d$) that has been
256     used in other coarse-grained simulations. The dipolar strength
257     (and the temperature and pressure) were the only other parameters that
258     were varied systematically.\label{mdfig:lipidModel}}
259     \end{figure}
260    
261     To take into account the permanent dipolar interactions of the
262 xsun 3359 zwitterionic head groups, we have placed fixed dipole moments
263     $\mu_{i}$ at one end of the Gay-Berne particles. The dipoles are
264     oriented at an angle $\theta = \pi / 2$ relative to the major axis.
265     These dipoles are protected by a head ``bead'' with a range parameter
266     ($\sigma_h$) which we have varied between $1.20 d$ and $1.41 d$. The
267     head groups interact with each other using a combination of
268     Lennard-Jones,
269 xsun 3358 \begin{equation}
270     V_{ij}(r_{ij}) = 4\epsilon_h \left[\left(\frac{\sigma_h}{r_{ij}}\right)^{12} -
271     \left(\frac{\sigma_h}{r_{ij}}\right)^6\right],
272     \end{equation}
273     and dipole-dipole,
274     \begin{equation}
275     V_{ij}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
276     \hat{r}}_{ij})) = \frac{|\mu|^2}{4 \pi \epsilon_0 r_{ij}^3}
277     \left[ \hat{\bf u}_i \cdot \hat{\bf u}_j - 3 (\hat{\bf u}_i \cdot
278     \hat{\bf r}_{ij})(\hat{\bf u}_j \cdot \hat{\bf r}_{ij}) \right]
279     \end{equation}
280     potentials.
281     In these potentials, $\mathbf{\hat u}_i$ is the unit vector pointing
282     along dipole $i$ and $\mathbf{\hat r}_{ij}$ is the unit vector
283     pointing along the inter-dipole vector $\mathbf{r}_{ij}$.
284    
285     Since the charge separation distance is so large in zwitterionic head
286     groups (like the PC head groups), it would also be possible to use
287     either point charges or a ``split dipole'' approximation,
288     \begin{equation}
289     V_{ij}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
290     \hat{r}}_{ij})) = \frac{1}{{4\pi \epsilon_0 }}\left[ {\frac{{\mu _i \cdot \mu _j }}{{R_{ij}^3 }} -
291     \frac{{3\left( {\mu _i \cdot r_{ij} } \right)\left( {\mu _i \cdot
292     r_{ij} } \right)}}{{R_{ij}^5 }}} \right]
293     \end{equation}
294     where $\mu _i$ and $\mu _j$ are the dipole moments of sites $i$ and
295     $j$, $r_{ij}$ is vector between the two sites, and $R_{ij}$ is given
296     by,
297     \begin{equation}
298     R_{ij} = \sqrt {r_{ij}^2 + \frac{{d_i^2 }}{4} + \frac{{d_j^2
299     }}{4}}.
300     \end{equation}
301     Here, $d_i$ and $d_j$ are charge separation distances associated with
302     each of the two dipolar sites. This approximation to the multipole
303     expansion maintains the fast fall-off of the multipole potentials but
304     lacks the normal divergences when two polar groups get close to one
305     another.
306    
307     For the interaction between nonequivalent uniaxial ellipsoids (in this
308     case, between spheres and ellipsoids), the spheres are treated as
309     ellipsoids with an aspect ratio of 1 ($d = l$) and with an well depth
310     ratio ($\epsilon^r$) of 1 ($\epsilon^e = \epsilon^s$). The form of
311     the Gay-Berne potential we are using was generalized by Cleaver {\it
312     et al.} and is appropriate for dissimilar uniaxial
313     ellipsoids.\cite{Cleaver96}
314    
315     The solvent model in our simulations is similar to the one used by
316     Marrink {\it et al.} in their coarse grained simulations of lipid
317 xsun 3359 bilayers.\cite{Marrink2004} The solvent bead is a single site that
318 xsun 3358 represents four water molecules (m = 72 amu) and has comparable
319     density and diffusive behavior to liquid water. However, since there
320     are no electrostatic sites on these beads, this solvent model cannot
321     replicate the dielectric properties of water. Note that although we
322     are using larger cutoff and switching radii than Marrink {\it et al.},
323     our solvent density at 300 K remains at 0.944 g cm$^{-3}$, and the
324 xsun 3359 solvent diffuses at 0.43 \AA$^2 ps^{-1}$ (only twice as fast as liquid
325 xsun 3358 water).
326    
327     \begin{table*}
328     \begin{minipage}{\linewidth}
329     \begin{center}
330     \caption{Potential parameters used for molecular-scale coarse-grained
331     lipid simulations}
332     \begin{tabular}{llccc}
333     \hline
334     & & Head & Chain & Solvent \\
335     \hline
336     $d$ (\AA) & & varied & 4.6 & 4.7 \\
337     $l$ (\AA) & & $= d$ & 13.8 & 4.7 \\
338     $\epsilon^s$ (kcal/mol) & & 0.185 & 1.0 & 0.8 \\
339     $\epsilon_r$ (well-depth aspect ratio)& & 1 & 0.2 & 1 \\
340     $m$ (amu) & & 196 & 760 & 72.06 \\
341     $\overleftrightarrow{\mathsf I}$ (amu \AA$^2$) & & & & \\
342     \multicolumn{2}c{$I_{xx}$} & 1125 & 45000 & N/A \\
343     \multicolumn{2}c{$I_{yy}$} & 1125 & 45000 & N/A \\
344     \multicolumn{2}c{$I_{zz}$} & 0 & 9000 & N/A \\
345     $\mu$ (Debye) & & varied & 0 & 0 \\
346     \end{tabular}
347     \label{mdtab:parameters}
348     \end{center}
349     \end{minipage}
350     \end{table*}
351    
352     \section{Experimental Methodology}
353     \label{mdsec:experiment}
354    
355     The parameters that were systematically varied in this study were the
356     size of the head group ($\sigma_h$), the strength of the dipole moment
357     ($\mu$), and the temperature of the system. Values for $\sigma_h$
358     ranged from 5.5 \AA\ to 6.5 \AA. If the width of the tails is taken
359     to be the unit of length, these head groups correspond to a range from
360     $1.2 d$ to $1.41 d$. Since the solvent beads are nearly identical in
361     diameter to the tail ellipsoids, all distances that follow will be
362     measured relative to this unit of distance. Because the solvent we
363     are using is non-polar and has a dielectric constant of 1, values for
364     $\mu$ are sampled from a range that is somewhat smaller than the 20.6
365     Debye dipole moment of the PC head groups.
366    
367     To create unbiased bilayers, all simulations were started from two
368     perfectly flat monolayers separated by a 26 \AA\ gap between the
369     molecular bodies of the upper and lower leaves. The separated
370     monolayers were evolved in a vacuum with $x-y$ anisotropic pressure
371     coupling. The length of $z$ axis of the simulations was fixed and a
372     constant surface tension was applied to enable real fluctuations of
373     the bilayer. Periodic boundary conditions were used, and $480-720$
374     lipid molecules were present in the simulations, depending on the size
375     of the head beads. In all cases, the two monolayers spontaneously
376     collapsed into bilayer structures within 100 ps. Following this
377     collapse, all systems were equilibrated for $100$ ns at $300$ K.
378    
379     The resulting bilayer structures were then solvated at a ratio of $6$
380     solvent beads (24 water molecules) per lipid. These configurations
381     were then equilibrated for another $30$ ns. All simulations utilizing
382     the solvent were carried out at constant pressure ($P=1$ atm) with
383     $3$D anisotropic coupling, and small constant surface tension
384     ($\gamma=0.015$ N/m). Given the absence of fast degrees of freedom in
385     this model, a time step of $50$ fs was utilized with excellent energy
386     conservation. Data collection for structural properties of the
387     bilayers was carried out during a final 5 ns run following the solvent
388     equilibration. Orientational correlation functions and diffusion
389     constants were computed from 30 ns simulations in the microcanonical
390     (NVE) ensemble using the average volume from the end of the constant
391     pressure and surface tension runs. The timestep on these final
392     molecular dynamics runs was 25 fs. No appreciable changes in phase
393     structure were noticed upon switching to a microcanonical ensemble.
394     All simulations were performed using the {\sc oopse} molecular
395 xsun 3359 modeling program.\cite{Meineke2005}
396 xsun 3358
397     A switching function was applied to all potentials to smoothly turn
398     off the interactions between a range of $22$ and $25$ \AA. The
399     switching function was the standard (cubic) function,
400     \begin{equation}
401     s(r) =
402     \begin{cases}
403     1 & \text{if $r \le r_{\text{sw}}$},\\
404     \frac{(r_{\text{cut}} + 2r - 3r_{\text{sw}})(r_{\text{cut}} - r)^2}
405     {(r_{\text{cut}} - r_{\text{sw}})^3}
406     & \text{if $r_{\text{sw}} < r \le r_{\text{cut}}$}, \\
407     0 & \text{if $r > r_{\text{cut}}$.}
408     \end{cases}
409     \label{mdeq:dipoleSwitching}
410     \end{equation}
411    
412     \section{Results}
413     \label{mdsec:results}
414    
415     The membranes in our simulations exhibit a number of interesting
416     bilayer phases. The surface topology of these phases depends most
417     sensitively on the ratio of the size of the head groups to the width
418     of the molecular bodies. With heads only slightly larger than the
419     bodies ($\sigma_h=1.20 d$) the membrane exhibits a flat bilayer.
420    
421     Increasing the head / body size ratio increases the local membrane
422     curvature around each of the lipids. With $\sigma_h=1.28 d$, the
423     surface is still essentially flat, but the bilayer starts to exhibit
424     signs of instability. We have observed occasional defects where a
425     line of lipid molecules on one leaf of the bilayer will dip down to
426     interdigitate with the other leaf. This gives each of the two bilayer
427     leaves some local convexity near the line defect. These structures,
428     once developed in a simulation, are very stable and are spaced
429     approximately 100 \AA\ away from each other.
430    
431     With larger heads ($\sigma_h = 1.35 d$) the membrane curvature
432     resolves into a ``symmetric'' ripple phase. Each leaf of the bilayer
433     is broken into several convex, hemicylinderical sections, and opposite
434     leaves are fitted together much like roof tiles. There is no
435     interdigitation between the upper and lower leaves of the bilayer.
436    
437     For the largest head / body ratios studied ($\sigma_h = 1.41 d$) the
438     local curvature is substantially larger, and the resulting bilayer
439     structure resolves into an asymmetric ripple phase. This structure is
440     very similar to the structures observed by both de~Vries {\it et al.}
441     and Lenz {\it et al.}. For a given ripple wave vector, there are two
442     possible asymmetric ripples, which is not the case for the symmetric
443     phase observed when $\sigma_h = 1.35 d$.
444    
445     \begin{figure}
446     \centering
447     \includegraphics[width=\linewidth]{./figures/mdPhaseCartoon.pdf}
448     \caption{The role of the ratio between the head group size and the
449     width of the molecular bodies is to increase the local membrane
450     curvature. With strong attractive interactions between the head
451     groups, this local curvature can be maintained in bilayer structures
452     through surface corrugation. Shown above are three phases observed in
453     these simulations. With $\sigma_h = 1.20 d$, the bilayer maintains a
454     flat topology. For larger heads ($\sigma_h = 1.35 d$) the local
455     curvature resolves into a symmetrically rippled phase with little or
456     no interdigitation between the upper and lower leaves of the membrane.
457     The largest heads studied ($\sigma_h = 1.41 d$) resolve into an
458     asymmetric rippled phases with interdigitation between the two
459     leaves.\label{mdfig:phaseCartoon}}
460     \end{figure}
461    
462     Sample structures for the flat ($\sigma_h = 1.20 d$), symmetric
463     ($\sigma_h = 1.35 d$, and asymmetric ($\sigma_h = 1.41 d$) ripple
464     phases are shown in Figure \ref{mdfig:phaseCartoon}.
465    
466     It is reasonable to ask how well the parameters we used can produce
467     bilayer properties that match experimentally known values for real
468     lipid bilayers. Using a value of $l = 13.8$ \AA for the ellipsoidal
469     tails and the fixed ellipsoidal aspect ratio of 3, our values for the
470     area per lipid ($A$) and inter-layer spacing ($D_{HH}$) depend
471     entirely on the size of the head bead relative to the molecular body.
472     These values are tabulated in table \ref{mdtab:property}. Kucera {\it
473     et al.} have measured values for the head group spacings for a number
474     of PC lipid bilayers that range from 30.8 \AA\ (DLPC) to 37.8 \AA\ (DPPC).
475     They have also measured values for the area per lipid that range from
476     60.6
477     \AA$^2$ (DMPC) to 64.2 \AA$^2$
478     (DPPC).\cite{NorbertKucerka04012005,NorbertKucerka06012006} Only the
479     largest of the head groups we modeled ($\sigma_h = 1.41 d$) produces
480     bilayers (specifically the area per lipid) that resemble real PC
481     bilayers. The smaller head beads we used are perhaps better models
482     for PE head groups.
483    
484     \begin{table*}
485     \begin{minipage}{\linewidth}
486     \begin{center}
487     \caption{Phase, bilayer spacing, area per lipid, ripple wavelength
488     and amplitude observed as a function of the ratio between the head
489     beads and the diameters of the tails. Ripple wavelengths and
490     amplitudes are normalized to the diameter of the tail ellipsoids.}
491     \begin{tabular}{lccccc}
492     \hline
493     $\sigma_h / d$ & type of phase & bilayer spacing (\AA) & area per
494     lipid (\AA$^2$) & $\lambda / d$ & $A / d$\\
495     \hline
496     1.20 & flat & 33.4 & 49.6 & N/A & N/A \\
497     1.28 & flat & 33.7 & 54.7 & N/A & N/A \\
498     1.35 & symmetric ripple & 42.9 & 51.7 & 17.2 & 2.2 \\
499     1.41 & asymmetric ripple & 37.1 & 63.1 & 15.4 & 1.5 \\
500     \end{tabular}
501     \label{mdtab:property}
502     \end{center}
503     \end{minipage}
504     \end{table*}
505    
506     The membrane structures and the reduced wavelength $\lambda / d$,
507     reduced amplitude $A / d$ of the ripples are summarized in Table
508     \ref{mdtab:property}. The wavelength range is $15 - 17$ molecular bodies
509     and the amplitude is $1.5$ molecular bodies for asymmetric ripple and
510     $2.2$ for symmetric ripple. These values are reasonably consistent
511     with experimental measurements.\cite{Sun96,Katsaras00,Kaasgaard03}
512     Note, that given the lack of structural freedom in the tails of our
513     model lipids, the amplitudes observed from these simulations are
514     likely to underestimate of the true amplitudes.
515    
516     \begin{figure}
517     \centering
518     \includegraphics[width=\linewidth]{./figures/mdTopDown.pdf}
519     \caption{Top views of the flat (upper), symmetric ripple (middle),
520     and asymmetric ripple (lower) phases. Note that the head-group
521     dipoles have formed head-to-tail chains in all three of these phases,
522     but in the two rippled phases, the dipolar chains are all aligned {\it
523     perpendicular} to the direction of the ripple. Note that the flat
524     membrane has multiple vortex defects in the dipolar ordering, and the
525     ordering on the lower leaf of the bilayer can be in an entirely
526     different direction from the upper leaf.\label{mdfig:topView}}
527     \end{figure}
528    
529     The principal method for observing orientational ordering in dipolar
530     or liquid crystalline systems is the $P_2$ order parameter (defined
531     as $1.5 \times \lambda_{max}$, where $\lambda_{max}$ is the largest
532     eigenvalue of the matrix,
533     \begin{equation}
534     {\mathsf{S}} = \frac{1}{N} \sum_i \left(
535     \begin{array}{ccc}
536     u^{x}_i u^{x}_i-\frac{1}{3} & u^{x}_i u^{y}_i & u^{x}_i u^{z}_i \\
537     u^{y}_i u^{x}_i & u^{y}_i u^{y}_i -\frac{1}{3} & u^{y}_i u^{z}_i \\
538     u^{z}_i u^{x}_i & u^{z}_i u^{y}_i & u^{z}_i u^{z}_i -\frac{1}{3}
539     \end{array} \right).
540     \label{mdeq:opmatrix}
541     \end{equation}
542     Here $u^{\alpha}_i$ is the $\alpha=x,y,z$ component of the unit vector
543     for molecule $i$. (Here, $\hat{\bf u}_i$ can refer either to the
544     principal axis of the molecular body or to the dipole on the head
545     group of the molecule.) $P_2$ will be $1.0$ for a perfectly-ordered
546     system and near $0$ for a randomized system. Note that this order
547     parameter is {\em not} equal to the polarization of the system. For
548     example, the polarization of a perfect anti-ferroelectric arrangement
549     of point dipoles is $0$, but $P_2$ for the same system is $1$. The
550     eigenvector of $\mathsf{S}$ corresponding to the largest eigenvalue is
551     familiar as the director axis, which can be used to determine a
552     privileged axis for an orientationally-ordered system. Since the
553     molecular bodies are perpendicular to the head group dipoles, it is
554     possible for the director axes for the molecular bodies and the head
555     groups to be completely decoupled from each other.
556    
557     Figure \ref{mdfig:topView} shows snapshots of bird's-eye views of the
558     flat ($\sigma_h = 1.20 d$) and rippled ($\sigma_h = 1.35, 1.41 d$)
559     bilayers. The directions of the dipoles on the head groups are
560     represented with two colored half spheres: blue (phosphate) and yellow
561     (amino). For flat bilayers, the system exhibits signs of
562     orientational frustration; some disorder in the dipolar head-to-tail
563     chains is evident with kinks visible at the edges between differently
564     ordered domains. The lipids can also move independently of lipids in
565     the opposing leaf, so the ordering of the dipoles on one leaf is not
566     necessarily consistent with the ordering on the other. These two
567     factors keep the total dipolar order parameter relatively low for the
568     flat phases.
569    
570     With increasing head group size, the surface becomes corrugated, and
571     the dipoles cannot move as freely on the surface. Therefore, the
572     translational freedom of lipids in one layer is dependent upon the
573     position of the lipids in the other layer. As a result, the ordering of
574     the dipoles on head groups in one leaf is correlated with the ordering
575     in the other leaf. Furthermore, as the membrane deforms due to the
576     corrugation, the symmetry of the allowed dipolar ordering on each leaf
577     is broken. The dipoles then self-assemble in a head-to-tail
578     configuration, and the dipolar order parameter increases dramatically.
579     However, the total polarization of the system is still close to zero.
580     This is strong evidence that the corrugated structure is an
581     anti-ferroelectric state. It is also notable that the head-to-tail
582     arrangement of the dipoles is always observed in a direction
583     perpendicular to the wave vector for the surface corrugation. This is
584     a similar finding to what we observed in our earlier work on the
585 xsun 3359 elastic dipolar membranes.\cite{sun:031602}
586 xsun 3358
587     The $P_2$ order parameters (for both the molecular bodies and the head
588     group dipoles) have been calculated to quantify the ordering in these
589     phases. Figure \ref{mdfig:rP2} shows that the $P_2$ order parameter for
590     the head-group dipoles increases with increasing head group size. When
591     the heads of the lipid molecules are small, the membrane is nearly
592     flat. Since the in-plane packing is essentially a close packing of the
593     head groups, the head dipoles exhibit frustration in their
594     orientational ordering.
595    
596     The ordering trends for the tails are essentially opposite to the
597     ordering of the head group dipoles. The tail $P_2$ order parameter
598     {\it decreases} with increasing head size. This indicates that the
599     surface is more curved with larger head / tail size ratios. When the
600     surface is flat, all tails are pointing in the same direction (normal
601     to the bilayer surface). This simplified model appears to be
602     exhibiting a smectic A fluid phase, similar to the real $L_{\beta}$
603     phase. We have not observed a smectic C gel phase ($L_{\beta'}$) for
604     this model system. Increasing the size of the heads results in
605     rapidly decreasing $P_2$ ordering for the molecular bodies.
606    
607     \begin{figure}
608     \centering
609     \includegraphics[width=\linewidth]{./figures/mdRP2.pdf}
610     \caption{The $P_2$ order parameters for head groups (circles) and
611     molecular bodies (squares) as a function of the ratio of head group
612     size ($\sigma_h$) to the width of the molecular bodies ($d$). \label{mdfig:rP2}}
613     \end{figure}
614    
615     In addition to varying the size of the head groups, we studied the
616     effects of the interactions between head groups on the structure of
617     lipid bilayer by changing the strength of the dipoles. Figure
618     \ref{mdfig:sP2} shows how the $P_2$ order parameter changes with
619     increasing strength of the dipole. Generally, the dipoles on the head
620     groups become more ordered as the strength of the interaction between
621     heads is increased and become more disordered by decreasing the
622     interaction strength. When the interaction between the heads becomes
623     too weak, the bilayer structure does not persist; all lipid molecules
624     become dispersed in the solvent (which is non-polar in this
625     molecular-scale model). The critical value of the strength of the
626     dipole depends on the size of the head groups. The perfectly flat
627     surface becomes unstable below $5$ Debye, while the rippled
628     surfaces melt at $8$ Debye (asymmetric) or $10$ Debye (symmetric).
629    
630     The ordering of the tails mirrors the ordering of the dipoles {\it
631     except for the flat phase}. Since the surface is nearly flat in this
632     phase, the order parameters are only weakly dependent on dipolar
633     strength until it reaches $15$ Debye. Once it reaches this value, the
634     head group interactions are strong enough to pull the head groups
635     close to each other and distort the bilayer structure. For a flat
636     surface, a substantial amount of free volume between the head groups
637     is normally available. When the head groups are brought closer by
638     dipolar interactions, the tails are forced to splay outward, first forming
639     curved bilayers, and then inverted micelles.
640    
641     When $\sigma_h=1.28 d$, the $P_2$ order parameter decreases slightly
642     when the strength of the dipole is increased above $16$ Debye. For
643     rippled bilayers, there is less free volume available between the head
644     groups. Therefore increasing dipolar strength weakly influences the
645     structure of the membrane. However, the increase in the body $P_2$
646     order parameters implies that the membranes are being slightly
647     flattened due to the effects of increasing head-group attraction.
648    
649     A very interesting behavior takes place when the head groups are very
650     large relative to the molecular bodies ($\sigma_h = 1.41 d$) and the
651     dipolar strength is relatively weak ($\mu < 9$ Debye). In this case,
652     the two leaves of the bilayer become totally interdigitated with each
653     other in large patches of the membrane. With higher dipolar
654     strength, the interdigitation is limited to single lines that run
655     through the bilayer in a direction perpendicular to the ripple wave
656     vector.
657    
658     \begin{figure}
659     \centering
660     \includegraphics[width=\linewidth]{./figures/mdSP2.pdf}
661     \caption{The $P_2$ order parameters for head group dipoles (a) and
662     molecular bodies (b) as a function of the strength of the dipoles.
663     These order parameters are shown for four values of the head group /
664     molecular width ratio ($\sigma_h / d$). \label{mdfig:sP2}}
665     \end{figure}
666    
667     Figure \ref{mdfig:tP2} shows the dependence of the order parameters on
668     temperature. As expected, systems are more ordered at low
669     temperatures, and more disordered at high temperatures. All of the
670     bilayers we studied can become unstable if the temperature becomes
671     high enough. The only interesting feature of the temperature
672     dependence is in the flat surfaces ($\sigma_h=1.20 d$ and
673     $\sigma_h=1.28 d$). Here, when the temperature is increased above
674     $310$K, there is enough jostling of the head groups to allow the
675     dipolar frustration to resolve into more ordered states. This results
676     in a slight increase in the $P_2$ order parameter above this
677     temperature.
678    
679     For the rippled surfaces ($\sigma_h=1.35 d$ and $\sigma_h=1.41 d$),
680     there is a slightly increased orientational ordering in the molecular
681     bodies above $290$K. Since our model lacks the detailed information
682     about the behavior of the lipid tails, this is the closest the model
683     can come to depicting the ripple ($P_{\beta'}$) to fluid
684     ($L_{\alpha}$) phase transition. What we are observing is a
685     flattening of the rippled structures made possible by thermal
686     expansion of the tightly-packed head groups. The lack of detailed
687     chain configurations also makes it impossible for this model to depict
688     the ripple to gel ($L_{\beta'}$) phase transition.
689    
690     \begin{figure}
691     \centering
692     \includegraphics[width=\linewidth]{./figures/mdTP2.pdf}
693     \caption{The $P_2$ order parameters for head group dipoles (a) and
694     molecular bodies (b) as a function of temperature.
695     These order parameters are shown for four values of the head group /
696     molecular width ratio ($\sigma_h / d$).\label{mdfig:tP2}}
697     \end{figure}
698    
699     Fig. \ref{mdfig:phaseDiagram} shows a phase diagram for the model as a
700     function of the head group / molecular width ratio ($\sigma_h / d$)
701     and the strength of the head group dipole moment ($\mu$). Note that
702     the specific form of the bilayer phase is governed almost entirely by
703     the head group / molecular width ratio, while the strength of the
704     dipolar interactions between the head groups governs the stability of
705     the bilayer phase. Weaker dipoles result in unstable bilayer phases,
706     while extremely strong dipoles can shift the equilibrium to an
707     inverted micelle phase when the head groups are small. Temperature
708     has little effect on the actual bilayer phase observed, although higher
709     temperatures can cause the unstable region to grow into the higher
710     dipole region of this diagram.
711    
712     \begin{figure}
713     \centering
714     \includegraphics[width=\linewidth]{./figures/mdPhaseDiagram.pdf}
715     \caption{Phase diagram for the simple molecular model as a function
716     of the head group / molecular width ratio ($\sigma_h / d$) and the
717     strength of the head group dipole moment
718     ($\mu$).\label{mdfig:phaseDiagram}}
719     \end{figure}
720    
721     We have computed translational diffusion constants for lipid molecules
722     from the mean-square displacement,
723     \begin{equation}
724     D = \lim_{t \rightarrow \infty} \frac{1}{6 t} \langle {|\left({\bf r}_{i}(t) - {\bf r}_{i}(0) \right)|}^2 \rangle,
725     \end{equation}
726     of the lipid bodies. Translational diffusion constants for the
727     different head-to-tail size ratios (all at 300 K) are shown in table
728     \ref{mdtab:relaxation}. We have also computed orientational correlation
729     times for the head groups from fits of the second-order Legendre
730     polynomial correlation function,
731     \begin{equation}
732     C_{\ell}(t) = \langle P_{\ell}\left({\bf \mu}_{i}(t) \cdot {\bf
733     \mu}_{i}(0) \right) \rangle
734     \end{equation}
735     of the head group dipoles. The orientational correlation functions
736     appear to have multiple components in their decay: a fast ($12 \pm 2$
737     ps) decay due to librational motion of the head groups, as well as
738     moderate ($\tau^h_{\rm mid}$) and slow ($\tau^h_{\rm slow}$)
739     components. The fit values for the moderate and slow correlation
740     times are listed in table \ref{mdtab:relaxation}. Standard deviations
741     in the fit time constants are quite large (on the order of the values
742     themselves).
743    
744     Sparrman and Westlund used a multi-relaxation model for NMR lineshapes
745     observed in gel, fluid, and ripple phases of DPPC and obtained
746     estimates of a correlation time for water translational diffusion
747     ($\tau_c$) of 20 ns.\cite{Sparrman2003} This correlation time
748     corresponds to water bound to small regions of the lipid membrane.
749     They further assume that the lipids can explore only a single period
750     of the ripple (essentially moving in a nearly one-dimensional path to
751     do so), and the correlation time can therefore be used to estimate a
752     value for the translational diffusion constant of $2.25 \times
753     10^{-11} m^2 s^{-1}$. The translational diffusion constants we obtain
754     are in reasonable agreement with this experimentally determined
755     value. However, the $T_2$ relaxation times obtained by Sparrman and
756     Westlund are consistent with P-N vector reorientation timescales of
757     2-5 ms. This is substantially slower than even the slowest component
758     we observe in the decay of our orientational correlation functions.
759     Other than the dipole-dipole interactions, our head groups have no
760     shape anisotropy which would force them to move as a unit with
761     neighboring molecules. This would naturally lead to P-N reorientation
762     times that are too fast when compared with experimental measurements.
763    
764     \begin{table*}
765     \begin{minipage}{\linewidth}
766     \begin{center}
767     \caption{Fit values for the rotational correlation times for the head
768     groups ($\tau^h$) and molecular bodies ($\tau^b$) as well as the
769     translational diffusion constants for the molecule as a function of
770     the head-to-body width ratio. All correlation functions and transport
771     coefficients were computed from microcanonical simulations with an
772     average temperture of 300 K. In all of the phases, the head group
773     correlation functions decay with an fast librational contribution ($12
774     \pm 1$ ps). There are additional moderate ($\tau^h_{\rm mid}$) and
775     slow $\tau^h_{\rm slow}$ contributions to orientational decay that
776     depend strongly on the phase exhibited by the lipids. The symmetric
777     ripple phase ($\sigma_h / d = 1.35$) appears to exhibit the slowest
778     molecular reorientation.}
779     \begin{tabular}{lcccc}
780     \hline
781     $\sigma_h / d$ & $\tau^h_{\rm mid} (ns)$ & $\tau^h_{\rm
782     slow} (\mu s)$ & $\tau^b (\mu s)$ & $D (\times 10^{-11} m^2 s^{-1})$ \\
783     \hline
784     1.20 & $0.4$ & $9.6$ & $9.5$ & $0.43(1)$ \\
785     1.28 & $2.0$ & $13.5$ & $3.0$ & $5.91(3)$ \\
786     1.35 & $3.2$ & $4.0$ & $0.9$ & $3.42(1)$ \\
787     1.41 & $0.3$ & $23.8$ & $6.9$ & $7.16(1)$ \\
788     \end{tabular}
789     \label{mdtab:relaxation}
790     \end{center}
791     \end{minipage}
792     \end{table*}
793    
794     \section{Discussion}
795     \label{mdsec:discussion}
796    
797     Symmetric and asymmetric ripple phases have been observed to form in
798     our molecular dynamics simulations of a simple molecular-scale lipid
799     model. The lipid model consists of an dipolar head group and an
800     ellipsoidal tail. Within the limits of this model, an explanation for
801     generalized membrane curvature is a simple mismatch in the size of the
802     heads with the width of the molecular bodies. With heads
803     substantially larger than the bodies of the molecule, this curvature
804     should be convex nearly everywhere, a requirement which could be
805     resolved either with micellar or cylindrical phases.
806    
807     The persistence of a {\it bilayer} structure therefore requires either
808     strong attractive forces between the head groups or exclusionary
809     forces from the solvent phase. To have a persistent bilayer structure
810     with the added requirement of convex membrane curvature appears to
811     result in corrugated structures like the ones pictured in
812     Fig. \ref{mdfig:phaseCartoon}. In each of the sections of these
813     corrugated phases, the local curvature near a most of the head groups
814     is convex. These structures are held together by the extremely strong
815     and directional interactions between the head groups.
816    
817     The attractive forces holding the bilayer together could either be
818     non-directional (as in the work of Kranenburg and
819     Smit),\cite{Kranenburg2005} or directional (as we have utilized in
820     these simulations). The dipolar head groups are key for the
821     maintaining the bilayer structures exhibited by this particular model;
822     reducing the strength of the dipole has the tendency to make the
823     rippled phase disappear. The dipoles are likely to form attractive
824     head-to-tail configurations even in flat configurations, but the
825     temperatures are high enough that vortex defects become prevalent in
826     the flat phase. The flat phase we observed therefore appears to be
827     substantially above the Kosterlitz-Thouless transition temperature for
828     a planar system of dipoles with this set of parameters. For this
829     reason, it would be interesting to observe the thermal behavior of the
830     flat phase at substantially lower temperatures.
831    
832     One feature of this model is that an energetically favorable
833     orientational ordering of the dipoles can be achieved by forming
834     ripples. The corrugation of the surface breaks the symmetry of the
835     plane, making vortex defects somewhat more expensive, and stabilizing
836     the long range orientational ordering for the dipoles in the head
837     groups. Most of the rows of the head-to-tail dipoles are parallel to
838     each other and the system adopts a bulk anti-ferroelectric state. We
839     believe that this is the first time the organization of the head
840     groups in ripple phases has been addressed.
841    
842     Although the size-mismatch between the heads and molecular bodies
843     appears to be the primary driving force for surface convexity, the
844     persistence of the bilayer through the use of rippled structures is a
845     function of the strong, attractive interactions between the heads.
846     One important prediction we can make using the results from this
847     simple model is that if the dipole-dipole interaction is the leading
848     contributor to the head group attractions, the wave vectors for the
849     ripples should always be found {\it perpendicular} to the dipole
850     director axis. This echoes the prediction we made earlier for simple
851     elastic dipolar membranes, and may suggest experimental designs which
852     will test whether this is really the case in the phosphatidylcholine
853     $P_{\beta'}$ phases. The dipole director axis should also be easily
854     computable for the all-atom and coarse-grained simulations that have
855     been published in the literature.\cite{deVries05}
856    
857     Experimental verification of our predictions of dipolar orientation
858     correlating with the ripple direction would require knowing both the
859     local orientation of a rippled region of the membrane (available via
860     AFM studies of supported bilayers) as well as the local ordering of
861     the membrane dipoles. Obtaining information about the local
862     orientations of the membrane dipoles may be available from
863     fluorescence detected linear dichroism (LD). Benninger {\it et al.}
864     have recently used axially-specific chromophores
865     2-(4,4-difluoro-5,7-dimethyl-4-bora-3a,4a-diaza-s-indacene-3-pentanoyl)-1-hexadecanoyl-sn-glycero-3-phospocholine
866     ($\beta$-BODIPY FL C5-HPC or BODIPY-PC) and 3,3'
867     dioctadecyloxacarbocyanine perchlorate (DiO) in their
868     fluorescence-detected linear dichroism (LD) studies of plasma
869     membranes of living cells.\cite{Benninger:2005qy} The DiO dye aligns
870     its transition moment perpendicular to the membrane normal, while the
871     BODIPY-PC transition dipole is parallel with the membrane normal.
872     Without a doubt, using fluorescence detection of linear dichroism in
873     concert with AFM surface scanning would be difficult experiments to
874     carry out. However, there is some hope of performing experiments to
875     either verify or falsify the predictions of our simulations.
876    
877     Although our model is simple, it exhibits some rich and unexpected
878     behaviors. It would clearly be a closer approximation to reality if
879     we allowed bending motions between the dipoles and the molecular
880     bodies, and if we replaced the rigid ellipsoids with ball-and-chain
881     tails. However, the advantages of this simple model (large system
882     sizes, 50 fs time steps) allow us to rapidly explore the phase diagram
883     for a wide range of parameters. Our explanation of this rippling
884     phenomenon will help us design more accurate molecular models for
885     corrugated membranes and experiments to test whether or not
886     dipole-dipole interactions exert an influence on membrane rippling.