ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/xDissertation/md.tex
Revision: 3383
Committed: Wed Apr 16 21:56:34 2008 UTC (16 years, 2 months ago) by xsun
Content type: application/x-tex
File size: 45819 byte(s)
Log Message:
formatting check

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 xsun 3383 \caption[Three different representations of DPPC lipid
105     molecules]{Three different representations of DPPC lipid molecules,
106 xsun 3358 including the chemical structure, an atomistic model, and the
107     head-body ellipsoidal coarse-grained model used in this
108     work.\label{mdfig:lipidModels}}
109     \end{figure}
110    
111     Our simple molecular-scale lipid model for studying the ripple phase
112     is based on two facts: one is that the most essential feature of lipid
113     molecules is their amphiphilic structure with polar head groups and
114     non-polar tails. Another fact is that the majority of lipid molecules
115     in the ripple phase are relatively rigid (i.e. gel-like) which makes
116     some fraction of the details of the chain dynamics negligible. Figure
117     \ref{mdfig:lipidModels} shows the molecular structure of a DPPC
118     molecule, as well as atomistic and molecular-scale representations of
119     a DPPC molecule. The hydrophilic character of the head group is
120     largely due to the separation of charge between the nitrogen and
121     phosphate groups. The zwitterionic nature of the PC headgroups leads
122     to abnormally large dipole moments (as high as 20.6 D), and this
123     strongly polar head group interacts strongly with the solvating water
124     layers immediately surrounding the membrane. The hydrophobic tail
125     consists of fatty acid chains. In our molecular scale model, lipid
126     molecules have been reduced to these essential features; the fatty
127     acid chains are represented by an ellipsoid with a dipolar ball
128     perched on one end to represent the effects of the charge-separated
129     head group. In real PC lipids, the direction of the dipole is
130     nearly perpendicular to the tail, so we have fixed the direction of
131     the point dipole rigidly in this orientation.
132    
133     The ellipsoidal portions of the model interact via the Gay-Berne
134     potential which has seen widespread use in the liquid crystal
135     community. Ayton and Voth have also used Gay-Berne ellipsoids for
136     modeling large length-scale properties of lipid
137     bilayers.\cite{Ayton01} In its original form, the Gay-Berne potential
138     was a single site model for the interactions of rigid ellipsoidal
139 xsun 3359 molecules.\cite{Gay1981} It can be thought of as a modification of the
140 xsun 3358 Gaussian overlap model originally described by Berne and
141     Pechukas.\cite{Berne72} The potential is constructed in the familiar
142     form of the Lennard-Jones function using orientation-dependent
143     $\sigma$ and $\epsilon$ parameters,
144 xsun 3370 \begin{multline}
145 xsun 3358 V_{ij}({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j}, {\mathbf{\hat
146 xsun 3370 r}_{ij}}) = 4\epsilon ({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j},
147 xsun 3361 {\mathbf{\hat r}_{ij}})\left[ \left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u}_i},
148 xsun 3370 {\mathbf{\hat u}_j}, {\mathbf{\hat r}_{ij}})+\sigma_0}\right)^{12}
149     \right. \\
150     \left. - \left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u}_i},
151 xsun 3361 {\mathbf{\hat u}_j}, {\mathbf{\hat
152     r}_{ij}})+\sigma_0}\right)^6\right]
153 xsun 3358 \label{mdeq:gb}
154 xsun 3370 \end{multline}
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 3370 \begin{multline}
176     \sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf \hat{r}}_{ij}) = \\
177     \sigma_0 \left[ 1 - \left\{ \frac{ \chi \alpha^2 ({\bf \hat{u}}_i \cdot {\bf
178 xsun 3358 \hat{r}}_{ij} ) + \chi \alpha^{-2} ({\bf \hat{u}}_j \cdot {\bf
179     \hat{r}}_{ij} ) - 2 \chi^2 ({\bf \hat{u}}_i \cdot {\bf
180     \hat{r}}_{ij} )({\bf \hat{u}}_j \cdot {\bf
181     \hat{r}}_{ij} ) ({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j)}{1 - \chi^2
182     \left({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j\right)^2} \right\}
183 xsun 3370 \right]^{-1/2}
184     \end{multline}
185 xsun 3358
186     Gay-Berne ellipsoids also have an energy scaling parameter,
187     $\epsilon^s$, which describes the well depth for two identical
188     ellipsoids in a {\it side-by-side} configuration. Additionally, a well
189     depth aspect ratio, $\epsilon^r = \epsilon^e / \epsilon^s$, describes
190     the ratio between the well depths in the {\it end-to-end} and
191     side-by-side configurations. As in the range parameter, a set of
192     mixing and anisotropy variables can be used to describe the well
193     depths for dissimilar particles,
194     \begin {eqnarray*}
195     \epsilon_0 & = & \sqrt{\epsilon^s_i * \epsilon^s_j} \\ \\
196     \epsilon^r & = & \sqrt{\epsilon^r_i * \epsilon^r_j} \\ \\
197     \chi' & = & \frac{1 - (\epsilon^r)^{1/\mu}}{1 + (\epsilon^r)^{1/\mu}}
198     \\ \\
199     \alpha'^2 & = & \left[1 + (\epsilon^r)^{1/\mu}\right]^{-1}
200     \end{eqnarray*}
201     The form of the strength function is somewhat complicated,
202 xsun 3361 \begin{eqnarray*}
203 xsun 3358 \epsilon({\bf \hat{u}}_{i}, {\bf \hat{u}}_{j},{\bf \hat{r}}_{ij}) & = &
204     \epsilon_{0} \epsilon_{1}^{\nu}({\bf \hat{u}}_{i}.{\bf \hat{u}}_{j})
205     \epsilon_{2}^{\mu}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
206     \hat{r}}_{ij}) \\ \\
207     \epsilon_{1}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j}) & = &
208     \left[1-\chi^{2}({\bf \hat{u}}_{i}.{\bf
209 xsun 3361 \hat{u}}_{j})^{2}\right]^{-1/2}
210     \end{eqnarray*}
211 xsun 3370 \begin{multline*}
212     \epsilon_{2}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf \hat{r}}_{ij})
213     = \\
214     1 - \left\{ \frac{ \chi' \alpha'^2 ({\bf \hat{u}}_i \cdot {\bf
215 xsun 3358 \hat{r}}_{ij} ) + \chi' \alpha'^{-2} ({\bf \hat{u}}_j \cdot {\bf
216     \hat{r}}_{ij} ) - 2 \chi'^2 ({\bf \hat{u}}_i \cdot {\bf
217     \hat{r}}_{ij} )({\bf \hat{u}}_j \cdot {\bf
218     \hat{r}}_{ij} ) ({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j)}{1 - \chi'^2
219     \left({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j\right)^2} \right\},
220 xsun 3370 \end{multline*}
221 xsun 3358 although many of the quantities and derivatives are identical with
222     those obtained for the range parameter. Ref. \citen{Luckhurst90}
223     has a particularly good explanation of the choice of the Gay-Berne
224     parameters $\mu$ and $\nu$ for modeling liquid crystal molecules. An
225     excellent overview of the computational methods that can be used to
226     efficiently compute forces and torques for this potential can be found
227     in Ref. \citen{Golubkov06}
228    
229     The choices of parameters we have used in this study correspond to a
230     shape anisotropy of 3 for the chain portion of the molecule. In
231     principle, this could be varied to allow for modeling of longer or
232     shorter chain lipid molecules. For these prolate ellipsoids, we have:
233     \begin{equation}
234     \begin{array}{rcl}
235     d & < & l \\
236     \epsilon^{r} & < & 1
237     \end{array}
238     \end{equation}
239     A sketch of the various structural elements of our molecular-scale
240     lipid / solvent model is shown in figure \ref{mdfig:lipidModel}. The
241     actual parameters used in our simulations are given in table
242     \ref{mdtab:parameters}.
243    
244     \begin{figure}
245     \centering
246     \includegraphics[width=\linewidth]{./figures/md2LipidModel.pdf}
247 xsun 3383 \caption[The parameters defining the behavior of the lipid
248     models]{The parameters defining the behavior of the lipid
249     models. $\sigma_h / d$ is the ratio of the head group to body
250     diameter. Molecular bodies had a fixed aspect ratio of 3.0. The
251     solvent model was a simplified 4-water bead ($\sigma_w \approx d$)
252     that has been used in other coarse-grained simulations. The dipolar
253     strength (and the temperature and pressure) were the only other
254     parameters that were varied systematically.\label{mdfig:lipidModel}}
255 xsun 3358 \end{figure}
256    
257     To take into account the permanent dipolar interactions of the
258 xsun 3359 zwitterionic head groups, we have placed fixed dipole moments
259     $\mu_{i}$ at one end of the Gay-Berne particles. The dipoles are
260     oriented at an angle $\theta = \pi / 2$ relative to the major axis.
261     These dipoles are protected by a head ``bead'' with a range parameter
262     ($\sigma_h$) which we have varied between $1.20 d$ and $1.41 d$. The
263     head groups interact with each other using a combination of
264     Lennard-Jones,
265 xsun 3358 \begin{equation}
266     V_{ij}(r_{ij}) = 4\epsilon_h \left[\left(\frac{\sigma_h}{r_{ij}}\right)^{12} -
267     \left(\frac{\sigma_h}{r_{ij}}\right)^6\right],
268     \end{equation}
269     and dipole-dipole,
270     \begin{equation}
271     V_{ij}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
272     \hat{r}}_{ij})) = \frac{|\mu|^2}{4 \pi \epsilon_0 r_{ij}^3}
273     \left[ \hat{\bf u}_i \cdot \hat{\bf u}_j - 3 (\hat{\bf u}_i \cdot
274     \hat{\bf r}_{ij})(\hat{\bf u}_j \cdot \hat{\bf r}_{ij}) \right]
275     \end{equation}
276     potentials.
277     In these potentials, $\mathbf{\hat u}_i$ is the unit vector pointing
278     along dipole $i$ and $\mathbf{\hat r}_{ij}$ is the unit vector
279     pointing along the inter-dipole vector $\mathbf{r}_{ij}$.
280    
281     Since the charge separation distance is so large in zwitterionic head
282     groups (like the PC head groups), it would also be possible to use
283     either point charges or a ``split dipole'' approximation,
284     \begin{equation}
285     V_{ij}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
286     \hat{r}}_{ij})) = \frac{1}{{4\pi \epsilon_0 }}\left[ {\frac{{\mu _i \cdot \mu _j }}{{R_{ij}^3 }} -
287     \frac{{3\left( {\mu _i \cdot r_{ij} } \right)\left( {\mu _i \cdot
288     r_{ij} } \right)}}{{R_{ij}^5 }}} \right]
289     \end{equation}
290     where $\mu _i$ and $\mu _j$ are the dipole moments of sites $i$ and
291     $j$, $r_{ij}$ is vector between the two sites, and $R_{ij}$ is given
292     by,
293     \begin{equation}
294     R_{ij} = \sqrt {r_{ij}^2 + \frac{{d_i^2 }}{4} + \frac{{d_j^2
295     }}{4}}.
296     \end{equation}
297     Here, $d_i$ and $d_j$ are charge separation distances associated with
298     each of the two dipolar sites. This approximation to the multipole
299     expansion maintains the fast fall-off of the multipole potentials but
300     lacks the normal divergences when two polar groups get close to one
301     another.
302    
303     For the interaction between nonequivalent uniaxial ellipsoids (in this
304     case, between spheres and ellipsoids), the spheres are treated as
305     ellipsoids with an aspect ratio of 1 ($d = l$) and with an well depth
306     ratio ($\epsilon^r$) of 1 ($\epsilon^e = \epsilon^s$). The form of
307     the Gay-Berne potential we are using was generalized by Cleaver {\it
308     et al.} and is appropriate for dissimilar uniaxial
309     ellipsoids.\cite{Cleaver96}
310    
311     The solvent model in our simulations is similar to the one used by
312     Marrink {\it et al.} in their coarse grained simulations of lipid
313 xsun 3359 bilayers.\cite{Marrink2004} The solvent bead is a single site that
314 xsun 3358 represents four water molecules (m = 72 amu) and has comparable
315     density and diffusive behavior to liquid water. However, since there
316     are no electrostatic sites on these beads, this solvent model cannot
317     replicate the dielectric properties of water. Note that although we
318     are using larger cutoff and switching radii than Marrink {\it et al.},
319     our solvent density at 300 K remains at 0.944 g cm$^{-3}$, and the
320 xsun 3359 solvent diffuses at 0.43 \AA$^2 ps^{-1}$ (only twice as fast as liquid
321 xsun 3358 water).
322    
323     \begin{table*}
324     \begin{minipage}{\linewidth}
325     \begin{center}
326 xsun 3383 \caption{POTENTIAL PARAMETERS USED FOR MOLECULAR SCALE COARSE-GRAINED
327     LIPID SIMULATIONS}
328 xsun 3358 \begin{tabular}{llccc}
329     \hline
330     & & Head & Chain & Solvent \\
331     \hline
332     $d$ (\AA) & & varied & 4.6 & 4.7 \\
333     $l$ (\AA) & & $= d$ & 13.8 & 4.7 \\
334     $\epsilon^s$ (kcal/mol) & & 0.185 & 1.0 & 0.8 \\
335     $\epsilon_r$ (well-depth aspect ratio)& & 1 & 0.2 & 1 \\
336     $m$ (amu) & & 196 & 760 & 72.06 \\
337     $\overleftrightarrow{\mathsf I}$ (amu \AA$^2$) & & & & \\
338     \multicolumn{2}c{$I_{xx}$} & 1125 & 45000 & N/A \\
339     \multicolumn{2}c{$I_{yy}$} & 1125 & 45000 & N/A \\
340     \multicolumn{2}c{$I_{zz}$} & 0 & 9000 & N/A \\
341     $\mu$ (Debye) & & varied & 0 & 0 \\
342     \end{tabular}
343     \label{mdtab:parameters}
344     \end{center}
345     \end{minipage}
346     \end{table*}
347    
348 xsun 3383 \section{Simulation Methodology}
349     \label{mdsec:simulation}
350 xsun 3358
351     The parameters that were systematically varied in this study were the
352     size of the head group ($\sigma_h$), the strength of the dipole moment
353     ($\mu$), and the temperature of the system. Values for $\sigma_h$
354     ranged from 5.5 \AA\ to 6.5 \AA. If the width of the tails is taken
355     to be the unit of length, these head groups correspond to a range from
356     $1.2 d$ to $1.41 d$. Since the solvent beads are nearly identical in
357     diameter to the tail ellipsoids, all distances that follow will be
358     measured relative to this unit of distance. Because the solvent we
359     are using is non-polar and has a dielectric constant of 1, values for
360     $\mu$ are sampled from a range that is somewhat smaller than the 20.6
361     Debye dipole moment of the PC head groups.
362    
363     To create unbiased bilayers, all simulations were started from two
364     perfectly flat monolayers separated by a 26 \AA\ gap between the
365     molecular bodies of the upper and lower leaves. The separated
366     monolayers were evolved in a vacuum with $x-y$ anisotropic pressure
367     coupling. The length of $z$ axis of the simulations was fixed and a
368     constant surface tension was applied to enable real fluctuations of
369     the bilayer. Periodic boundary conditions were used, and $480-720$
370     lipid molecules were present in the simulations, depending on the size
371     of the head beads. In all cases, the two monolayers spontaneously
372     collapsed into bilayer structures within 100 ps. Following this
373     collapse, all systems were equilibrated for $100$ ns at $300$ K.
374    
375     The resulting bilayer structures were then solvated at a ratio of $6$
376     solvent beads (24 water molecules) per lipid. These configurations
377     were then equilibrated for another $30$ ns. All simulations utilizing
378     the solvent were carried out at constant pressure ($P=1$ atm) with
379     $3$D anisotropic coupling, and small constant surface tension
380     ($\gamma=0.015$ N/m). Given the absence of fast degrees of freedom in
381     this model, a time step of $50$ fs was utilized with excellent energy
382     conservation. Data collection for structural properties of the
383     bilayers was carried out during a final 5 ns run following the solvent
384     equilibration. Orientational correlation functions and diffusion
385     constants were computed from 30 ns simulations in the microcanonical
386     (NVE) ensemble using the average volume from the end of the constant
387     pressure and surface tension runs. The timestep on these final
388     molecular dynamics runs was 25 fs. No appreciable changes in phase
389     structure were noticed upon switching to a microcanonical ensemble.
390     All simulations were performed using the {\sc oopse} molecular
391 xsun 3359 modeling program.\cite{Meineke2005}
392 xsun 3358
393     A switching function was applied to all potentials to smoothly turn
394     off the interactions between a range of $22$ and $25$ \AA. The
395     switching function was the standard (cubic) function,
396     \begin{equation}
397     s(r) =
398     \begin{cases}
399     1 & \text{if $r \le r_{\text{sw}}$},\\
400     \frac{(r_{\text{cut}} + 2r - 3r_{\text{sw}})(r_{\text{cut}} - r)^2}
401     {(r_{\text{cut}} - r_{\text{sw}})^3}
402     & \text{if $r_{\text{sw}} < r \le r_{\text{cut}}$}, \\
403     0 & \text{if $r > r_{\text{cut}}$.}
404     \end{cases}
405     \label{mdeq:dipoleSwitching}
406     \end{equation}
407    
408     \section{Results}
409     \label{mdsec:results}
410    
411     The membranes in our simulations exhibit a number of interesting
412     bilayer phases. The surface topology of these phases depends most
413     sensitively on the ratio of the size of the head groups to the width
414     of the molecular bodies. With heads only slightly larger than the
415     bodies ($\sigma_h=1.20 d$) the membrane exhibits a flat bilayer.
416    
417     Increasing the head / body size ratio increases the local membrane
418     curvature around each of the lipids. With $\sigma_h=1.28 d$, the
419     surface is still essentially flat, but the bilayer starts to exhibit
420     signs of instability. We have observed occasional defects where a
421     line of lipid molecules on one leaf of the bilayer will dip down to
422     interdigitate with the other leaf. This gives each of the two bilayer
423     leaves some local convexity near the line defect. These structures,
424     once developed in a simulation, are very stable and are spaced
425     approximately 100 \AA\ away from each other.
426    
427     With larger heads ($\sigma_h = 1.35 d$) the membrane curvature
428     resolves into a ``symmetric'' ripple phase. Each leaf of the bilayer
429     is broken into several convex, hemicylinderical sections, and opposite
430     leaves are fitted together much like roof tiles. There is no
431     interdigitation between the upper and lower leaves of the bilayer.
432    
433     For the largest head / body ratios studied ($\sigma_h = 1.41 d$) the
434     local curvature is substantially larger, and the resulting bilayer
435     structure resolves into an asymmetric ripple phase. This structure is
436     very similar to the structures observed by both de~Vries {\it et al.}
437     and Lenz {\it et al.}. For a given ripple wave vector, there are two
438     possible asymmetric ripples, which is not the case for the symmetric
439     phase observed when $\sigma_h = 1.35 d$.
440    
441     \begin{figure}
442     \centering
443     \includegraphics[width=\linewidth]{./figures/mdPhaseCartoon.pdf}
444 xsun 3383 \caption[ three phases observed in the simulations]{The role of the
445     ratio between the head group size and the width of the molecular
446     bodies is to increase the local membrane curvature. With strong
447     attractive interactions between the head groups, this local curvature
448     can be maintained in bilayer structures through surface corrugation.
449     Shown above are three phases observed in these simulations. With
450     $\sigma_h = 1.20 d$, the bilayer maintains a flat topology. For
451     larger heads ($\sigma_h = 1.35 d$) the local curvature resolves into a
452     symmetrically rippled phase with little or no interdigitation between
453     the upper and lower leaves of the membrane. The largest heads studied
454     ($\sigma_h = 1.41 d$) resolve into an asymmetric rippled phases with
455     interdigitation between the two leaves.\label{mdfig:phaseCartoon}}
456 xsun 3358 \end{figure}
457    
458     Sample structures for the flat ($\sigma_h = 1.20 d$), symmetric
459     ($\sigma_h = 1.35 d$, and asymmetric ($\sigma_h = 1.41 d$) ripple
460     phases are shown in Figure \ref{mdfig:phaseCartoon}.
461    
462     It is reasonable to ask how well the parameters we used can produce
463     bilayer properties that match experimentally known values for real
464 xsun 3362 lipid bilayers. Using a value of $l = 13.8$ \AA~for the ellipsoidal
465 xsun 3358 tails and the fixed ellipsoidal aspect ratio of 3, our values for the
466     area per lipid ($A$) and inter-layer spacing ($D_{HH}$) depend
467     entirely on the size of the head bead relative to the molecular body.
468     These values are tabulated in table \ref{mdtab:property}. Kucera {\it
469     et al.} have measured values for the head group spacings for a number
470     of PC lipid bilayers that range from 30.8 \AA\ (DLPC) to 37.8 \AA\ (DPPC).
471     They have also measured values for the area per lipid that range from
472     60.6
473     \AA$^2$ (DMPC) to 64.2 \AA$^2$
474     (DPPC).\cite{NorbertKucerka04012005,NorbertKucerka06012006} Only the
475     largest of the head groups we modeled ($\sigma_h = 1.41 d$) produces
476     bilayers (specifically the area per lipid) that resemble real PC
477     bilayers. The smaller head beads we used are perhaps better models
478     for PE head groups.
479    
480     \begin{table*}
481     \begin{center}
482 xsun 3383 \caption{PHASE, BILAYER SPACING, AREA PER LIPID, RIPPLE WAVELENGTH AND
483     AMPLITUDE OBSERVED AS A FUNCTION OF THE RATIO BETWEEN THE HEAD BEADS
484     AND THE DIAMETERS OF THE TAILS}
485 xsun 3358 \begin{tabular}{lccccc}
486     \hline
487     $\sigma_h / d$ & type of phase & bilayer spacing (\AA) & area per
488     lipid (\AA$^2$) & $\lambda / d$ & $A / d$\\
489     \hline
490     1.20 & flat & 33.4 & 49.6 & N/A & N/A \\
491     1.28 & flat & 33.7 & 54.7 & N/A & N/A \\
492     1.35 & symmetric ripple & 42.9 & 51.7 & 17.2 & 2.2 \\
493     1.41 & asymmetric ripple & 37.1 & 63.1 & 15.4 & 1.5 \\
494     \end{tabular}
495 xsun 3383 \begin{minipage}{\linewidth}
496     %\centering
497     \vspace{2mm}
498     Ripple wavelengths and amplitudes are normalized to the diameter of
499     the tail ellipsoids.
500 xsun 3358 \label{mdtab:property}
501 xsun 3383 \end{minipage}
502 xsun 3358 \end{center}
503     \end{table*}
504    
505     The membrane structures and the reduced wavelength $\lambda / d$,
506     reduced amplitude $A / d$ of the ripples are summarized in Table
507     \ref{mdtab:property}. The wavelength range is $15 - 17$ molecular bodies
508     and the amplitude is $1.5$ molecular bodies for asymmetric ripple and
509     $2.2$ for symmetric ripple. These values are reasonably consistent
510     with experimental measurements.\cite{Sun96,Katsaras00,Kaasgaard03}
511     Note, that given the lack of structural freedom in the tails of our
512     model lipids, the amplitudes observed from these simulations are
513     likely to underestimate of the true amplitudes.
514    
515     \begin{figure}
516     \centering
517     \includegraphics[width=\linewidth]{./figures/mdTopDown.pdf}
518 xsun 3383 \caption[Top views of the flat, symmetric ripple, and asymmetric
519     ripple phases]{Top views of the flat (upper), symmetric ripple
520     (middle), and asymmetric ripple (lower) phases. Note that the
521     head-group dipoles have formed head-to-tail chains in all three of
522     these phases, but in the two rippled phases, the dipolar chains are
523     all aligned {\it perpendicular} to the direction of the ripple. Note
524     that the flat membrane has multiple vortex defects in the dipolar
525     ordering, and the ordering on the lower leaf of the bilayer can be in
526     an entirely different direction from the upper
527     leaf.\label{mdfig:topView}}
528 xsun 3358 \end{figure}
529    
530 xsun 3362 The orientational ordering in the system is observed by $P_2$ order
531     parameter, which is calculated from Eq.~\ref{mceq:opmatrix}
532     in Ch.~\ref{chap:mc}. Here, $\hat{\bf u}_i$ can refer either to the
533 xsun 3358 principal axis of the molecular body or to the dipole on the head
534 xsun 3362 group of the molecule. Since the molecular bodies are perpendicular to
535     the head group dipoles, it is possible for the director axes for the
536     molecular bodies and the head groups to be completely decoupled from
537     each other.
538 xsun 3358
539     Figure \ref{mdfig:topView} shows snapshots of bird's-eye views of the
540     flat ($\sigma_h = 1.20 d$) and rippled ($\sigma_h = 1.35, 1.41 d$)
541     bilayers. The directions of the dipoles on the head groups are
542     represented with two colored half spheres: blue (phosphate) and yellow
543     (amino). For flat bilayers, the system exhibits signs of
544     orientational frustration; some disorder in the dipolar head-to-tail
545     chains is evident with kinks visible at the edges between differently
546     ordered domains. The lipids can also move independently of lipids in
547     the opposing leaf, so the ordering of the dipoles on one leaf is not
548     necessarily consistent with the ordering on the other. These two
549     factors keep the total dipolar order parameter relatively low for the
550     flat phases.
551    
552     With increasing head group size, the surface becomes corrugated, and
553     the dipoles cannot move as freely on the surface. Therefore, the
554     translational freedom of lipids in one layer is dependent upon the
555     position of the lipids in the other layer. As a result, the ordering of
556     the dipoles on head groups in one leaf is correlated with the ordering
557     in the other leaf. Furthermore, as the membrane deforms due to the
558     corrugation, the symmetry of the allowed dipolar ordering on each leaf
559     is broken. The dipoles then self-assemble in a head-to-tail
560     configuration, and the dipolar order parameter increases dramatically.
561     However, the total polarization of the system is still close to zero.
562     This is strong evidence that the corrugated structure is an
563     anti-ferroelectric state. It is also notable that the head-to-tail
564     arrangement of the dipoles is always observed in a direction
565     perpendicular to the wave vector for the surface corrugation. This is
566     a similar finding to what we observed in our earlier work on the
567 xsun 3359 elastic dipolar membranes.\cite{sun:031602}
568 xsun 3358
569     The $P_2$ order parameters (for both the molecular bodies and the head
570     group dipoles) have been calculated to quantify the ordering in these
571     phases. Figure \ref{mdfig:rP2} shows that the $P_2$ order parameter for
572     the head-group dipoles increases with increasing head group size. When
573     the heads of the lipid molecules are small, the membrane is nearly
574     flat. Since the in-plane packing is essentially a close packing of the
575     head groups, the head dipoles exhibit frustration in their
576     orientational ordering.
577    
578     The ordering trends for the tails are essentially opposite to the
579     ordering of the head group dipoles. The tail $P_2$ order parameter
580     {\it decreases} with increasing head size. This indicates that the
581     surface is more curved with larger head / tail size ratios. When the
582     surface is flat, all tails are pointing in the same direction (normal
583     to the bilayer surface). This simplified model appears to be
584     exhibiting a smectic A fluid phase, similar to the real $L_{\beta}$
585     phase. We have not observed a smectic C gel phase ($L_{\beta'}$) for
586     this model system. Increasing the size of the heads results in
587     rapidly decreasing $P_2$ ordering for the molecular bodies.
588    
589     \begin{figure}
590     \centering
591     \includegraphics[width=\linewidth]{./figures/mdRP2.pdf}
592 xsun 3383 \caption[The $P_2$ order parameters as a function of the ratio of head group
593     size to the width of the molecular bodies]{The $P_2$ order parameters
594     for head groups (circles) and molecular bodies (squares) as a function
595     of the ratio of head group size ($\sigma_h$) to the width of the
596     molecular bodies ($d$). \label{mdfig:rP2}}
597 xsun 3358 \end{figure}
598    
599     In addition to varying the size of the head groups, we studied the
600     effects of the interactions between head groups on the structure of
601     lipid bilayer by changing the strength of the dipoles. Figure
602     \ref{mdfig:sP2} shows how the $P_2$ order parameter changes with
603     increasing strength of the dipole. Generally, the dipoles on the head
604     groups become more ordered as the strength of the interaction between
605     heads is increased and become more disordered by decreasing the
606     interaction strength. When the interaction between the heads becomes
607     too weak, the bilayer structure does not persist; all lipid molecules
608     become dispersed in the solvent (which is non-polar in this
609     molecular-scale model). The critical value of the strength of the
610     dipole depends on the size of the head groups. The perfectly flat
611     surface becomes unstable below $5$ Debye, while the rippled
612     surfaces melt at $8$ Debye (asymmetric) or $10$ Debye (symmetric).
613    
614     The ordering of the tails mirrors the ordering of the dipoles {\it
615     except for the flat phase}. Since the surface is nearly flat in this
616     phase, the order parameters are only weakly dependent on dipolar
617     strength until it reaches $15$ Debye. Once it reaches this value, the
618     head group interactions are strong enough to pull the head groups
619     close to each other and distort the bilayer structure. For a flat
620     surface, a substantial amount of free volume between the head groups
621     is normally available. When the head groups are brought closer by
622     dipolar interactions, the tails are forced to splay outward, first forming
623     curved bilayers, and then inverted micelles.
624    
625     When $\sigma_h=1.28 d$, the $P_2$ order parameter decreases slightly
626     when the strength of the dipole is increased above $16$ Debye. For
627     rippled bilayers, there is less free volume available between the head
628     groups. Therefore increasing dipolar strength weakly influences the
629     structure of the membrane. However, the increase in the body $P_2$
630     order parameters implies that the membranes are being slightly
631     flattened due to the effects of increasing head-group attraction.
632    
633     A very interesting behavior takes place when the head groups are very
634     large relative to the molecular bodies ($\sigma_h = 1.41 d$) and the
635     dipolar strength is relatively weak ($\mu < 9$ Debye). In this case,
636     the two leaves of the bilayer become totally interdigitated with each
637     other in large patches of the membrane. With higher dipolar
638     strength, the interdigitation is limited to single lines that run
639     through the bilayer in a direction perpendicular to the ripple wave
640     vector.
641    
642     \begin{figure}
643     \centering
644     \includegraphics[width=\linewidth]{./figures/mdSP2.pdf}
645 xsun 3383 \caption[The $P_2$ order parameters as a function of the strength of
646     the dipoles.]{The $P_2$ order parameters for head group dipoles (a)
647     and molecular bodies (b) as a function of the strength of the dipoles.
648 xsun 3358 These order parameters are shown for four values of the head group /
649     molecular width ratio ($\sigma_h / d$). \label{mdfig:sP2}}
650     \end{figure}
651    
652     Figure \ref{mdfig:tP2} shows the dependence of the order parameters on
653     temperature. As expected, systems are more ordered at low
654     temperatures, and more disordered at high temperatures. All of the
655     bilayers we studied can become unstable if the temperature becomes
656     high enough. The only interesting feature of the temperature
657     dependence is in the flat surfaces ($\sigma_h=1.20 d$ and
658     $\sigma_h=1.28 d$). Here, when the temperature is increased above
659     $310$K, there is enough jostling of the head groups to allow the
660     dipolar frustration to resolve into more ordered states. This results
661     in a slight increase in the $P_2$ order parameter above this
662     temperature.
663    
664     For the rippled surfaces ($\sigma_h=1.35 d$ and $\sigma_h=1.41 d$),
665     there is a slightly increased orientational ordering in the molecular
666     bodies above $290$K. Since our model lacks the detailed information
667     about the behavior of the lipid tails, this is the closest the model
668     can come to depicting the ripple ($P_{\beta'}$) to fluid
669     ($L_{\alpha}$) phase transition. What we are observing is a
670     flattening of the rippled structures made possible by thermal
671     expansion of the tightly-packed head groups. The lack of detailed
672     chain configurations also makes it impossible for this model to depict
673     the ripple to gel ($L_{\beta'}$) phase transition.
674    
675     \begin{figure}
676     \centering
677     \includegraphics[width=\linewidth]{./figures/mdTP2.pdf}
678 xsun 3383 \caption[The $P_2$ order parameters as a function of temperature]{The
679     $P_2$ order parameters for head group dipoles (a) and molecular bodies
680     (b) as a function of temperature. These order parameters are shown
681     for four values of the head group / molecular width ratio ($\sigma_h /
682     d$).\label{mdfig:tP2}}
683 xsun 3358 \end{figure}
684    
685     Fig. \ref{mdfig:phaseDiagram} shows a phase diagram for the model as a
686     function of the head group / molecular width ratio ($\sigma_h / d$)
687     and the strength of the head group dipole moment ($\mu$). Note that
688     the specific form of the bilayer phase is governed almost entirely by
689     the head group / molecular width ratio, while the strength of the
690     dipolar interactions between the head groups governs the stability of
691     the bilayer phase. Weaker dipoles result in unstable bilayer phases,
692     while extremely strong dipoles can shift the equilibrium to an
693     inverted micelle phase when the head groups are small. Temperature
694     has little effect on the actual bilayer phase observed, although higher
695     temperatures can cause the unstable region to grow into the higher
696     dipole region of this diagram.
697    
698     \begin{figure}
699     \centering
700     \includegraphics[width=\linewidth]{./figures/mdPhaseDiagram.pdf}
701 xsun 3383 \caption[Phase diagram for the simple molecular model]{Phase diagram
702     for the simple molecular model as a function of the head group /
703     molecular width ratio ($\sigma_h / d$) and the strength of the head
704     group dipole moment ($\mu$).\label{mdfig:phaseDiagram}}
705 xsun 3358 \end{figure}
706    
707     We have computed translational diffusion constants for lipid molecules
708     from the mean-square displacement,
709     \begin{equation}
710 xsun 3370 D = \lim_{t \rightarrow \infty} \frac{1}{6 t} \langle {|\left({\bf
711     r}_{i}(t) - {\bf r}_{i}(0) \right)|}^2 \rangle,
712     \label{mdeq:msdisplacement}
713 xsun 3358 \end{equation}
714     of the lipid bodies. Translational diffusion constants for the
715     different head-to-tail size ratios (all at 300 K) are shown in table
716     \ref{mdtab:relaxation}. We have also computed orientational correlation
717     times for the head groups from fits of the second-order Legendre
718     polynomial correlation function,
719     \begin{equation}
720     C_{\ell}(t) = \langle P_{\ell}\left({\bf \mu}_{i}(t) \cdot {\bf
721     \mu}_{i}(0) \right) \rangle
722     \end{equation}
723     of the head group dipoles. The orientational correlation functions
724     appear to have multiple components in their decay: a fast ($12 \pm 2$
725     ps) decay due to librational motion of the head groups, as well as
726     moderate ($\tau^h_{\rm mid}$) and slow ($\tau^h_{\rm slow}$)
727     components. The fit values for the moderate and slow correlation
728     times are listed in table \ref{mdtab:relaxation}. Standard deviations
729     in the fit time constants are quite large (on the order of the values
730     themselves).
731    
732     Sparrman and Westlund used a multi-relaxation model for NMR lineshapes
733     observed in gel, fluid, and ripple phases of DPPC and obtained
734     estimates of a correlation time for water translational diffusion
735     ($\tau_c$) of 20 ns.\cite{Sparrman2003} This correlation time
736     corresponds to water bound to small regions of the lipid membrane.
737     They further assume that the lipids can explore only a single period
738     of the ripple (essentially moving in a nearly one-dimensional path to
739     do so), and the correlation time can therefore be used to estimate a
740     value for the translational diffusion constant of $2.25 \times
741     10^{-11} m^2 s^{-1}$. The translational diffusion constants we obtain
742     are in reasonable agreement with this experimentally determined
743     value. However, the $T_2$ relaxation times obtained by Sparrman and
744     Westlund are consistent with P-N vector reorientation timescales of
745     2-5 ms. This is substantially slower than even the slowest component
746     we observe in the decay of our orientational correlation functions.
747     Other than the dipole-dipole interactions, our head groups have no
748     shape anisotropy which would force them to move as a unit with
749     neighboring molecules. This would naturally lead to P-N reorientation
750     times that are too fast when compared with experimental measurements.
751    
752     \begin{table*}
753     \begin{center}
754 xsun 3383 \caption{FIT VALUES FOR THE ROTATIONAL CORRELATION TIMES FOR THE HEAD
755     GROUPS ($\tau^h$) AND MOLECULAR BODIES ($\tau^b$) AS WELL AS THE
756     TRANSLATIONAL DIFFUSION CONSTANTS FOR THE MOL\-E\-CULE AS A FUNCTION
757     OF THE HEAD-TO-BODY WIDTH RATIO}
758 xsun 3358 \begin{tabular}{lcccc}
759     \hline
760     $\sigma_h / d$ & $\tau^h_{\rm mid} (ns)$ & $\tau^h_{\rm
761     slow} (\mu s)$ & $\tau^b (\mu s)$ & $D (\times 10^{-11} m^2 s^{-1})$ \\
762     \hline
763     1.20 & $0.4$ & $9.6$ & $9.5$ & $0.43(1)$ \\
764     1.28 & $2.0$ & $13.5$ & $3.0$ & $5.91(3)$ \\
765     1.35 & $3.2$ & $4.0$ & $0.9$ & $3.42(1)$ \\
766     1.41 & $0.3$ & $23.8$ & $6.9$ & $7.16(1)$ \\
767     \end{tabular}
768 xsun 3383 \begin{minipage}{\linewidth}
769     %\centering
770     \vspace{2mm}
771     All correlation functions and transport coefficients were computed
772     from microcanonical simulations with an average temperture of 300 K.
773     In all of the phases, the head group correlation functions decay with
774     an fast librational contribution ($12 \pm 1$ ps). There are
775     additional moderate ($\tau^h_{\rm mid}$) and slow $\tau^h_{\rm slow}$
776     contributions to orientational decay that depend strongly on the phase
777     exhibited by the lipids. The symmetric ripple phase ($\sigma_h / d =
778     1.35$) appears to exhibit the slowest molecular reorientation.
779 xsun 3358 \label{mdtab:relaxation}
780 xsun 3383 \end{minipage}
781 xsun 3358 \end{center}
782     \end{table*}
783    
784     \section{Discussion}
785     \label{mdsec:discussion}
786    
787     Symmetric and asymmetric ripple phases have been observed to form in
788     our molecular dynamics simulations of a simple molecular-scale lipid
789     model. The lipid model consists of an dipolar head group and an
790     ellipsoidal tail. Within the limits of this model, an explanation for
791     generalized membrane curvature is a simple mismatch in the size of the
792     heads with the width of the molecular bodies. With heads
793     substantially larger than the bodies of the molecule, this curvature
794     should be convex nearly everywhere, a requirement which could be
795     resolved either with micellar or cylindrical phases.
796    
797     The persistence of a {\it bilayer} structure therefore requires either
798     strong attractive forces between the head groups or exclusionary
799     forces from the solvent phase. To have a persistent bilayer structure
800     with the added requirement of convex membrane curvature appears to
801     result in corrugated structures like the ones pictured in
802     Fig. \ref{mdfig:phaseCartoon}. In each of the sections of these
803     corrugated phases, the local curvature near a most of the head groups
804     is convex. These structures are held together by the extremely strong
805     and directional interactions between the head groups.
806    
807     The attractive forces holding the bilayer together could either be
808     non-directional (as in the work of Kranenburg and
809     Smit),\cite{Kranenburg2005} or directional (as we have utilized in
810     these simulations). The dipolar head groups are key for the
811     maintaining the bilayer structures exhibited by this particular model;
812     reducing the strength of the dipole has the tendency to make the
813     rippled phase disappear. The dipoles are likely to form attractive
814     head-to-tail configurations even in flat configurations, but the
815     temperatures are high enough that vortex defects become prevalent in
816     the flat phase. The flat phase we observed therefore appears to be
817     substantially above the Kosterlitz-Thouless transition temperature for
818     a planar system of dipoles with this set of parameters. For this
819     reason, it would be interesting to observe the thermal behavior of the
820     flat phase at substantially lower temperatures.
821    
822     One feature of this model is that an energetically favorable
823     orientational ordering of the dipoles can be achieved by forming
824     ripples. The corrugation of the surface breaks the symmetry of the
825     plane, making vortex defects somewhat more expensive, and stabilizing
826     the long range orientational ordering for the dipoles in the head
827     groups. Most of the rows of the head-to-tail dipoles are parallel to
828     each other and the system adopts a bulk anti-ferroelectric state. We
829     believe that this is the first time the organization of the head
830     groups in ripple phases has been addressed.
831    
832     Although the size-mismatch between the heads and molecular bodies
833     appears to be the primary driving force for surface convexity, the
834     persistence of the bilayer through the use of rippled structures is a
835     function of the strong, attractive interactions between the heads.
836     One important prediction we can make using the results from this
837     simple model is that if the dipole-dipole interaction is the leading
838     contributor to the head group attractions, the wave vectors for the
839     ripples should always be found {\it perpendicular} to the dipole
840     director axis. This echoes the prediction we made earlier for simple
841     elastic dipolar membranes, and may suggest experimental designs which
842     will test whether this is really the case in the phosphatidylcholine
843     $P_{\beta'}$ phases. The dipole director axis should also be easily
844     computable for the all-atom and coarse-grained simulations that have
845     been published in the literature.\cite{deVries05}
846    
847     Experimental verification of our predictions of dipolar orientation
848     correlating with the ripple direction would require knowing both the
849     local orientation of a rippled region of the membrane (available via
850     AFM studies of supported bilayers) as well as the local ordering of
851     the membrane dipoles. Obtaining information about the local
852     orientations of the membrane dipoles may be available from
853     fluorescence detected linear dichroism (LD). Benninger {\it et al.}
854     have recently used axially-specific chromophores
855 xsun 3376 2-(4,4-difluoro-5,7-dimethyl-4-bora-3a,4a-diaza-s-indacene-3-pentanoyl)-1-hexadecanoyl-sn-glycero-3-\\
856     phospocholine ($\beta$-BODIPY FL C5-HPC or BODIPY-PC) and 3,3'
857 xsun 3358 dioctadecyloxacarbocyanine perchlorate (DiO) in their
858     fluorescence-detected linear dichroism (LD) studies of plasma
859     membranes of living cells.\cite{Benninger:2005qy} The DiO dye aligns
860     its transition moment perpendicular to the membrane normal, while the
861     BODIPY-PC transition dipole is parallel with the membrane normal.
862     Without a doubt, using fluorescence detection of linear dichroism in
863     concert with AFM surface scanning would be difficult experiments to
864     carry out. However, there is some hope of performing experiments to
865     either verify or falsify the predictions of our simulations.
866    
867     Although our model is simple, it exhibits some rich and unexpected
868     behaviors. It would clearly be a closer approximation to reality if
869     we allowed bending motions between the dipoles and the molecular
870     bodies, and if we replaced the rigid ellipsoids with ball-and-chain
871     tails. However, the advantages of this simple model (large system
872     sizes, 50 fs time steps) allow us to rapidly explore the phase diagram
873     for a wide range of parameters. Our explanation of this rippling
874     phenomenon will help us design more accurate molecular models for
875     corrugated membranes and experiments to test whether or not
876     dipole-dipole interactions exert an influence on membrane rippling.