ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/xDissertation/md.tex
(Generate patch)

Comparing trunk/xDissertation/md.tex (file contents):
Revision 3336 by xsun, Wed Jan 30 16:01:02 2008 UTC vs.
Revision 3358 by xsun, Wed Mar 5 02:18:08 2008 UTC

# Line 1 | Line 1
1 < \chapter{\label{chap:md}MOLECULAR DYNAMICS}
1 > \chapter{\label{chap:md}Dipolar ordering in the ripple phases of
2 > molecular-scale models of lipid membranes}
3 >
4 > \section{Introduction}
5 > \label{mdsec:Int}
6 > Fully hydrated lipids will aggregate spontaneously to form bilayers
7 > which exhibit a variety of phases depending on their temperatures and
8 > compositions. Among these phases, a periodic rippled phase
9 > ($P_{\beta'}$) appears as an intermediate phase between the gel
10 > ($L_\beta$) and fluid ($L_{\alpha}$) phases for relatively pure
11 > phosphatidylcholine (PC) bilayers.  The ripple phase has attracted
12 > substantial experimental interest over the past 30 years. Most
13 > structural information of the ripple phase has been obtained by the
14 > X-ray diffraction~\cite{Sun96,Katsaras00} and freeze-fracture electron
15 > microscopy (FFEM).~\cite{Copeland80,Meyer96} Recently, Kaasgaard {\it
16 > et al.} used atomic force microscopy (AFM) to observe ripple phase
17 > morphology in bilayers supported on mica.~\cite{Kaasgaard03} The
18 > experimental results provide strong support for a 2-dimensional
19 > hexagonal packing lattice of the lipid molecules within the ripple
20 > phase.  This is a notable change from the observed lipid packing
21 > within the gel phase,~\cite{Cevc87} although Tenchov {\it et al.} have
22 > recently observed near-hexagonal packing in some phosphatidylcholine
23 > (PC) gel phases.\cite{Tenchov2001} The X-ray diffraction work by
24 > Katsaras {\it et al.} showed that a rich phase diagram exhibiting both
25 > {\it asymmetric} and {\it symmetric} ripples is possible for lecithin
26 > bilayers.\cite{Katsaras00}
27 >
28 > A number of theoretical models have been presented to explain the
29 > formation of the ripple phase. Marder {\it et al.} used a
30 > curvature-dependent Landau-de~Gennes free-energy functional to predict
31 > a rippled phase.~\cite{Marder84} This model and other related
32 > continuum models predict higher fluidity in convex regions and that
33 > concave portions of the membrane correspond to more solid-like
34 > regions.  Carlson and Sethna used a packing-competition model (in
35 > which head groups and chains have competing packing energetics) to
36 > predict the formation of a ripple-like phase.  Their model predicted
37 > that the high-curvature portions have lower-chain packing and
38 > correspond to more fluid-like regions.  Goldstein and Leibler used a
39 > mean-field approach with a planar model for {\em inter-lamellar}
40 > interactions to predict rippling in multilamellar
41 > phases.~\cite{Goldstein88} McCullough and Scott proposed that the {\em
42 > anisotropy of the nearest-neighbor interactions} coupled to
43 > hydrophobic constraining forces which restrict height differences
44 > between nearest neighbors is the origin of the ripple
45 > phase.~\cite{McCullough90} Lubensky and MacKintosh introduced a Landau
46 > theory for tilt order and curvature of a single membrane and concluded
47 > that {\em coupling of molecular tilt to membrane curvature} is
48 > responsible for the production of ripples.~\cite{Lubensky93} Misbah,
49 > Duplat and Houchmandzadeh proposed that {\em inter-layer dipolar
50 > interactions} can lead to ripple instabilities.~\cite{Misbah98}
51 > Heimburg presented a {\em coexistence model} for ripple formation in
52 > which he postulates that fluid-phase line defects cause sharp
53 > curvature between relatively flat gel-phase regions.~\cite{Heimburg00}
54 > Kubica has suggested that a lattice model of polar head groups could
55 > be valuable in trying to understand bilayer phase
56 > formation.~\cite{Kubica02} Bannerjee used Monte Carlo simulations of
57 > lamellar stacks of hexagonal lattices to show that large head groups
58 > and molecular tilt with respect to the membrane normal vector can
59 > cause bulk rippling.~\cite{Bannerjee02} Recently, Kranenburg and Smit
60 > described the formation of symmetric ripple-like structures using a
61 > coarse grained solvent-head-tail bead model.\cite{Kranenburg2005}
62 > Their lipids consisted of a short chain of head beads tied to the two
63 > longer ``chains''.
64 >
65 > In contrast, few large-scale molecular modeling studies have been
66 > done due to the large size of the resulting structures and the time
67 > required for the phases of interest to develop.  With all-atom (and
68 > even unified-atom) simulations, only one period of the ripple can be
69 > observed and only for time scales in the range of 10-100 ns.  One of
70 > the most interesting molecular simulations was carried out by de~Vries
71 > {\it et al.}~\cite{deVries05}. According to their simulation results,
72 > the ripple consists of two domains, one resembling the gel bilayer,
73 > while in the other, the two leaves of the bilayer are fully
74 > interdigitated.  The mechanism for the formation of the ripple phase
75 > suggested by their work is a packing competition between the head
76 > groups and the tails of the lipid molecules.\cite{Carlson87} Recently,
77 > the ripple phase has also been studied by Lenz and Schmid using Monte
78 > Carlo simulations.\cite{Lenz07} Their structures are similar to the De
79 > Vries {\it et al.} structures except that the connection between the
80 > two leaves of the bilayer is a narrow interdigitated line instead of
81 > the fully interdigitated domain.  The symmetric ripple phase was also
82 > observed by Lenz {\it et al.}, and their work supports other claims
83 > that the mismatch between the size of the head group and tail of the
84 > lipid molecules is the driving force for the formation of the ripple
85 > phase. Ayton and Voth have found significant undulations in
86 > zero-surface-tension states of membranes simulated via dissipative
87 > particle dynamics, but their results are consistent with purely
88 > thermal undulations.~\cite{Ayton02}
89 >
90 > Although the organization of the tails of lipid molecules are
91 > addressed by these molecular simulations and the packing competition
92 > between head groups and tails is strongly implicated as the primary
93 > driving force for ripple formation, questions about the ordering of
94 > the head groups in ripple phase have not been settled.
95 >
96 > In a recent paper, we presented a simple ``web of dipoles'' spin
97 > lattice model which provides some physical insight into relationship
98 > between dipolar ordering and membrane buckling.\cite{Sun2007} We found
99 > that dipolar elastic membranes can spontaneously buckle, forming
100 > ripple-like topologies.  The driving force for the buckling of dipolar
101 > elastic membranes is the anti-ferroelectric ordering of the dipoles.
102 > This was evident in the ordering of the dipole director axis
103 > perpendicular to the wave vector of the surface ripples.  A similar
104 > phenomenon has also been observed by Tsonchev {\it et al.} in their
105 > work on the spontaneous formation of dipolar peptide chains into
106 > curved nano-structures.\cite{Tsonchev04,Tsonchev04II}
107 >
108 > In this paper, we construct a somewhat more realistic molecular-scale
109 > lipid model than our previous ``web of dipoles'' and use molecular
110 > dynamics simulations to elucidate the role of the head group dipoles
111 > in the formation and morphology of the ripple phase.  We describe our
112 > model and computational methodology in section \ref{mdsec:method}.
113 > Details on the simulations are presented in section
114 > \ref{mdsec:experiment}, with results following in section
115 > \ref{mdsec:results}.  A final discussion of the role of dipolar heads in
116 > the ripple formation can be found in section
117 > \ref{mdsec:discussion}.
118 >
119 > \section{Computational Model}
120 > \label{mdsec:method}
121 >
122 > \begin{figure}
123 > \centering
124 > \includegraphics[width=\linewidth]{./figures/mdLipidModels.pdf}
125 > \caption{Three different representations of DPPC lipid molecules,
126 > including the chemical structure, an atomistic model, and the
127 > head-body ellipsoidal coarse-grained model used in this
128 > work.\label{mdfig:lipidModels}}
129 > \end{figure}
130 >
131 > Our simple molecular-scale lipid model for studying the ripple phase
132 > is based on two facts: one is that the most essential feature of lipid
133 > molecules is their amphiphilic structure with polar head groups and
134 > non-polar tails. Another fact is that the majority of lipid molecules
135 > in the ripple phase are relatively rigid (i.e. gel-like) which makes
136 > some fraction of the details of the chain dynamics negligible.  Figure
137 > \ref{mdfig:lipidModels} shows the molecular structure of a DPPC
138 > molecule, as well as atomistic and molecular-scale representations of
139 > a DPPC molecule.  The hydrophilic character of the head group is
140 > largely due to the separation of charge between the nitrogen and
141 > phosphate groups.  The zwitterionic nature of the PC headgroups leads
142 > to abnormally large dipole moments (as high as 20.6 D), and this
143 > strongly polar head group interacts strongly with the solvating water
144 > layers immediately surrounding the membrane.  The hydrophobic tail
145 > consists of fatty acid chains.  In our molecular scale model, lipid
146 > molecules have been reduced to these essential features; the fatty
147 > acid chains are represented by an ellipsoid with a dipolar ball
148 > perched on one end to represent the effects of the charge-separated
149 > head group.  In real PC lipids, the direction of the dipole is
150 > nearly perpendicular to the tail, so we have fixed the direction of
151 > the point dipole rigidly in this orientation.  
152 >
153 > The ellipsoidal portions of the model interact via the Gay-Berne
154 > potential which has seen widespread use in the liquid crystal
155 > community.  Ayton and Voth have also used Gay-Berne ellipsoids for
156 > modeling large length-scale properties of lipid
157 > bilayers.\cite{Ayton01} In its original form, the Gay-Berne potential
158 > was a single site model for the interactions of rigid ellipsoidal
159 > molecules.\cite{Gay81} It can be thought of as a modification of the
160 > Gaussian overlap model originally described by Berne and
161 > Pechukas.\cite{Berne72} The potential is constructed in the familiar
162 > form of the Lennard-Jones function using orientation-dependent
163 > $\sigma$ and $\epsilon$ parameters,
164 > \begin{equation*}
165 > V_{ij}({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j}, {\mathbf{\hat
166 > r}_{ij}}) = 4\epsilon ({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j},
167 > {\mathbf{\hat r}_{ij}})\left[\left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u}_i},
168 > {\mathbf{\hat u}_j}, {\mathbf{\hat r}_{ij}})+\sigma_0}\right)^{12}
169 > -\left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j},
170 > {\mathbf{\hat r}_{ij}})+\sigma_0}\right)^6\right]
171 > \label{mdeq:gb}
172 > \end{equation*}
173 >
174 > The range $(\sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
175 > \hat{r}}_{ij}))$, and strength $(\epsilon({\bf \hat{u}}_{i},{\bf
176 > \hat{u}}_{j},{\bf \hat{r}}_{ij}))$ parameters
177 > are dependent on the relative orientations of the two molecules (${\bf
178 > \hat{u}}_{i},{\bf \hat{u}}_{j}$) as well as the direction of the
179 > intermolecular separation (${\bf \hat{r}}_{ij}$).  $\sigma$ and
180 > $\sigma_0$ are also governed by shape mixing and anisotropy variables,
181 > \begin {eqnarray*}
182 > \sigma_0 & = & \sqrt{d_i^2 + d_j^2} \\ \\
183 > \chi & = & \left[ \frac{\left( l_i^2 - d_i^2 \right) \left(l_j^2 -
184 > d_j^2 \right)}{\left( l_j^2 + d_i^2 \right) \left(l_i^2 +
185 > d_j^2 \right)}\right]^{1/2} \\ \\
186 > \alpha^2 & = & \left[ \frac{\left( l_i^2 - d_i^2 \right) \left(l_j^2 +
187 > d_i^2 \right)}{\left( l_j^2 - d_j^2 \right) \left(l_i^2 +
188 > d_j^2 \right)}\right]^{1/2},
189 > \end{eqnarray*}
190 > where $l$ and $d$ describe the length and width of each uniaxial
191 > ellipsoid.  These shape anisotropy parameters can then be used to
192 > calculate the range function,
193 > \begin{equation*}
194 > \sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf \hat{r}}_{ij}) = \sigma_{0}
195 > \left[ 1- \left\{ \frac{ \chi \alpha^2 ({\bf \hat{u}}_i \cdot {\bf
196 > \hat{r}}_{ij} ) + \chi \alpha^{-2} ({\bf \hat{u}}_j \cdot {\bf
197 > \hat{r}}_{ij} ) - 2 \chi^2 ({\bf \hat{u}}_i \cdot {\bf
198 > \hat{r}}_{ij} )({\bf \hat{u}}_j \cdot {\bf
199 > \hat{r}}_{ij} ) ({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j)}{1 - \chi^2
200 > \left({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j\right)^2} \right\}
201 > \right]^{-1/2}
202 > \end{equation*}
203 >
204 > Gay-Berne ellipsoids also have an energy scaling parameter,
205 > $\epsilon^s$, which describes the well depth for two identical
206 > ellipsoids in a {\it side-by-side} configuration.  Additionally, a well
207 > depth aspect ratio, $\epsilon^r = \epsilon^e / \epsilon^s$, describes
208 > the ratio between the well depths in the {\it end-to-end} and
209 > side-by-side configurations.  As in the range parameter, a set of
210 > mixing and anisotropy variables can be used to describe the well
211 > depths for dissimilar particles,
212 > \begin {eqnarray*}
213 > \epsilon_0 & = & \sqrt{\epsilon^s_i  * \epsilon^s_j} \\ \\
214 > \epsilon^r & = & \sqrt{\epsilon^r_i  * \epsilon^r_j} \\ \\
215 > \chi' & = & \frac{1 - (\epsilon^r)^{1/\mu}}{1 + (\epsilon^r)^{1/\mu}}
216 > \\ \\
217 > \alpha'^2 & = & \left[1 + (\epsilon^r)^{1/\mu}\right]^{-1}
218 > \end{eqnarray*}
219 > The form of the strength function is somewhat complicated,
220 > \begin {eqnarray*}
221 > \epsilon({\bf \hat{u}}_{i}, {\bf \hat{u}}_{j},{\bf \hat{r}}_{ij}) & = &
222 > \epsilon_{0}  \epsilon_{1}^{\nu}({\bf \hat{u}}_{i}.{\bf \hat{u}}_{j})
223 > \epsilon_{2}^{\mu}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
224 > \hat{r}}_{ij}) \\ \\
225 > \epsilon_{1}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j}) & = &
226 > \left[1-\chi^{2}({\bf \hat{u}}_{i}.{\bf
227 > \hat{u}}_{j})^{2}\right]^{-1/2} \\ \\
228 > \epsilon_{2}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf \hat{r}}_{ij}) &
229 > = &
230 > 1 - \left\{ \frac{ \chi' \alpha'^2 ({\bf \hat{u}}_i \cdot {\bf
231 > \hat{r}}_{ij} ) + \chi' \alpha'^{-2} ({\bf \hat{u}}_j \cdot {\bf
232 > \hat{r}}_{ij} ) - 2 \chi'^2 ({\bf \hat{u}}_i \cdot {\bf
233 > \hat{r}}_{ij} )({\bf \hat{u}}_j \cdot {\bf
234 > \hat{r}}_{ij} ) ({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j)}{1 - \chi'^2
235 > \left({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j\right)^2} \right\},
236 > \end {eqnarray*}
237 > although many of the quantities and derivatives are identical with
238 > those obtained for the range parameter. Ref. \citen{Luckhurst90}
239 > has a particularly good explanation of the choice of the Gay-Berne
240 > parameters $\mu$ and $\nu$ for modeling liquid crystal molecules. An
241 > excellent overview of the computational methods that can be used to
242 > efficiently compute forces and torques for this potential can be found
243 > in Ref. \citen{Golubkov06}
244 >
245 > The choices of parameters we have used in this study correspond to a
246 > shape anisotropy of 3 for the chain portion of the molecule.  In
247 > principle, this could be varied to allow for modeling of longer or
248 > shorter chain lipid molecules. For these prolate ellipsoids, we have:
249 > \begin{equation}
250 > \begin{array}{rcl}
251 > d & < & l \\
252 > \epsilon^{r} & < & 1
253 > \end{array}
254 > \end{equation}
255 > A sketch of the various structural elements of our molecular-scale
256 > lipid / solvent model is shown in figure \ref{mdfig:lipidModel}.  The
257 > actual parameters used in our simulations are given in table
258 > \ref{mdtab:parameters}.
259 >
260 > \begin{figure}
261 > \centering
262 > \includegraphics[width=\linewidth]{./figures/md2LipidModel.pdf}
263 > \caption{The parameters defining the behavior of the lipid
264 > models. $\sigma_h / d$ is the ratio of the head group to body diameter.
265 > Molecular bodies had a fixed aspect ratio of 3.0.  The solvent model
266 > was a simplified 4-water bead ($\sigma_w \approx d$) that has been
267 > used in other coarse-grained simulations.  The dipolar strength
268 > (and the temperature and pressure) were the only other parameters that
269 > were varied systematically.\label{mdfig:lipidModel}}
270 > \end{figure}
271 >
272 > To take into account the permanent dipolar interactions of the
273 > zwitterionic head groups, we have placed fixed dipole moments $\mu_{i}$ at
274 > one end of the Gay-Berne particles.  The dipoles are oriented at an
275 > angle $\theta = \pi / 2$ relative to the major axis.  These dipoles
276 > are protected by a head ``bead'' with a range parameter ($\sigma_h$) which we have
277 > varied between $1.20 d$ and $1.41 d$.  The head groups interact with
278 > each other using a combination of Lennard-Jones,
279 > \begin{equation}
280 > V_{ij}(r_{ij}) = 4\epsilon_h \left[\left(\frac{\sigma_h}{r_{ij}}\right)^{12} -
281 > \left(\frac{\sigma_h}{r_{ij}}\right)^6\right],
282 > \end{equation}
283 > and dipole-dipole,
284 > \begin{equation}
285 > V_{ij}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
286 > \hat{r}}_{ij})) = \frac{|\mu|^2}{4 \pi \epsilon_0 r_{ij}^3}
287 > \left[ \hat{\bf u}_i \cdot \hat{\bf u}_j - 3 (\hat{\bf u}_i \cdot
288 > \hat{\bf r}_{ij})(\hat{\bf u}_j \cdot \hat{\bf r}_{ij}) \right]
289 > \end{equation}
290 > potentials.  
291 > In these potentials, $\mathbf{\hat u}_i$ is the unit vector pointing
292 > along dipole $i$ and $\mathbf{\hat r}_{ij}$ is the unit vector
293 > pointing along the inter-dipole vector $\mathbf{r}_{ij}$.  
294 >
295 > Since the charge separation distance is so large in zwitterionic head
296 > groups (like the PC head groups), it would also be possible to use
297 > either point charges or a ``split dipole'' approximation,
298 > \begin{equation}
299 > V_{ij}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
300 > \hat{r}}_{ij})) = \frac{1}{{4\pi \epsilon_0 }}\left[ {\frac{{\mu _i  \cdot \mu _j }}{{R_{ij}^3 }} -
301 > \frac{{3\left( {\mu _i  \cdot r_{ij} } \right)\left( {\mu _i  \cdot
302 > r_{ij} } \right)}}{{R_{ij}^5 }}} \right]
303 > \end{equation}
304 > where $\mu _i$ and $\mu _j$ are the dipole moments of sites $i$ and
305 > $j$, $r_{ij}$ is vector between the two sites, and $R_{ij}$ is given
306 > by,
307 > \begin{equation}
308 > R_{ij}  = \sqrt {r_{ij}^2  + \frac{{d_i^2 }}{4} + \frac{{d_j^2
309 > }}{4}}.
310 > \end{equation}
311 > Here, $d_i$ and $d_j$ are charge separation distances associated with
312 > each of the two dipolar sites. This approximation to the multipole
313 > expansion maintains the fast fall-off of the multipole potentials but
314 > lacks the normal divergences when two polar groups get close to one
315 > another.
316 >
317 > For the interaction between nonequivalent uniaxial ellipsoids (in this
318 > case, between spheres and ellipsoids), the spheres are treated as
319 > ellipsoids with an aspect ratio of 1 ($d = l$) and with an well depth
320 > ratio ($\epsilon^r$) of 1 ($\epsilon^e = \epsilon^s$).  The form of
321 > the Gay-Berne potential we are using was generalized by Cleaver {\it
322 > et al.} and is appropriate for dissimilar uniaxial
323 > ellipsoids.\cite{Cleaver96}
324 >
325 > The solvent model in our simulations is similar to the one used by
326 > Marrink {\it et al.}  in their coarse grained simulations of lipid
327 > bilayers.\cite{Marrink04} The solvent bead is a single site that
328 > represents four water molecules (m = 72 amu) and has comparable
329 > density and diffusive behavior to liquid water.  However, since there
330 > are no electrostatic sites on these beads, this solvent model cannot
331 > replicate the dielectric properties of water.  Note that although we
332 > are using larger cutoff and switching radii than Marrink {\it et al.},
333 > our solvent density at 300 K remains at 0.944 g cm$^{-3}$, and the
334 > solvent diffuses at 0.43 $\AA^2 ps^{-1}$ (only twice as fast as liquid
335 > water).
336 >
337 > \begin{table*}
338 > \begin{minipage}{\linewidth}
339 > \begin{center}
340 > \caption{Potential parameters used for molecular-scale coarse-grained
341 > lipid simulations}
342 > \begin{tabular}{llccc}
343 > \hline
344 >  & &  Head & Chain & Solvent \\
345 > \hline
346 > $d$ (\AA) & & varied & 4.6  & 4.7 \\
347 > $l$ (\AA) & & $= d$ & 13.8 & 4.7 \\
348 > $\epsilon^s$ (kcal/mol) & & 0.185 & 1.0 & 0.8 \\
349 > $\epsilon_r$ (well-depth aspect ratio)& & 1 & 0.2 &  1 \\
350 > $m$ (amu) & & 196 & 760 & 72.06 \\
351 > $\overleftrightarrow{\mathsf I}$ (amu \AA$^2$) & & & & \\
352 > \multicolumn{2}c{$I_{xx}$} & 1125 & 45000 & N/A \\
353 > \multicolumn{2}c{$I_{yy}$} & 1125 & 45000 & N/A \\
354 > \multicolumn{2}c{$I_{zz}$} &  0 &    9000 & N/A \\
355 > $\mu$ (Debye) & & varied & 0 & 0 \\
356 > \end{tabular}
357 > \label{mdtab:parameters}
358 > \end{center}
359 > \end{minipage}
360 > \end{table*}
361 >
362 > \section{Experimental Methodology}
363 > \label{mdsec:experiment}
364 >
365 > The parameters that were systematically varied in this study were the
366 > size of the head group ($\sigma_h$), the strength of the dipole moment
367 > ($\mu$), and the temperature of the system.  Values for $\sigma_h$
368 > ranged from 5.5 \AA\ to 6.5 \AA.  If the width of the tails is taken
369 > to be the unit of length, these head groups correspond to a range from
370 > $1.2 d$ to $1.41 d$.  Since the solvent beads are nearly identical in
371 > diameter to the tail ellipsoids, all distances that follow will be
372 > measured relative to this unit of distance.  Because the solvent we
373 > are using is non-polar and has a dielectric constant of 1, values for
374 > $\mu$ are sampled from a range that is somewhat smaller than the 20.6
375 > Debye dipole moment of the PC head groups.
376 >
377 > To create unbiased bilayers, all simulations were started from two
378 > perfectly flat monolayers separated by a 26 \AA\ gap between the
379 > molecular bodies of the upper and lower leaves.  The separated
380 > monolayers were evolved in a vacuum with $x-y$ anisotropic pressure
381 > coupling. The length of $z$ axis of the simulations was fixed and a
382 > constant surface tension was applied to enable real fluctuations of
383 > the bilayer. Periodic boundary conditions were used, and $480-720$
384 > lipid molecules were present in the simulations, depending on the size
385 > of the head beads.  In all cases, the two monolayers spontaneously
386 > collapsed into bilayer structures within 100 ps. Following this
387 > collapse, all systems were equilibrated for $100$ ns at $300$ K.
388 >
389 > The resulting bilayer structures were then solvated at a ratio of $6$
390 > solvent beads (24 water molecules) per lipid. These configurations
391 > were then equilibrated for another $30$ ns. All simulations utilizing
392 > the solvent were carried out at constant pressure ($P=1$ atm) with
393 > $3$D anisotropic coupling, and small constant surface tension
394 > ($\gamma=0.015$ N/m). Given the absence of fast degrees of freedom in
395 > this model, a time step of $50$ fs was utilized with excellent energy
396 > conservation.  Data collection for structural properties of the
397 > bilayers was carried out during a final 5 ns run following the solvent
398 > equilibration.  Orientational correlation functions and diffusion
399 > constants were computed from 30 ns simulations in the microcanonical
400 > (NVE) ensemble using the average volume from the end of the constant
401 > pressure and surface tension runs.  The timestep on these final
402 > molecular dynamics runs was 25 fs.  No appreciable changes in phase
403 > structure were noticed upon switching to a microcanonical ensemble.
404 > All simulations were performed using the {\sc oopse} molecular
405 > modeling program.\cite{Meineke05}
406 >
407 > A switching function was applied to all potentials to smoothly turn
408 > off the interactions between a range of $22$ and $25$ \AA.  The
409 > switching function was the standard (cubic) function,
410 > \begin{equation}
411 > s(r) =
412 >        \begin{cases}
413 >        1 & \text{if $r \le r_{\text{sw}}$},\\
414 >        \frac{(r_{\text{cut}} + 2r - 3r_{\text{sw}})(r_{\text{cut}} - r)^2}
415 >        {(r_{\text{cut}} - r_{\text{sw}})^3}
416 >        & \text{if $r_{\text{sw}} < r \le r_{\text{cut}}$}, \\
417 >        0 & \text{if $r > r_{\text{cut}}$.}
418 >        \end{cases}
419 > \label{mdeq:dipoleSwitching}
420 > \end{equation}
421 >
422 > \section{Results}
423 > \label{mdsec:results}
424 >
425 > The membranes in our simulations exhibit a number of interesting
426 > bilayer phases.  The surface topology of these phases depends most
427 > sensitively on the ratio of the size of the head groups to the width
428 > of the molecular bodies.  With heads only slightly larger than the
429 > bodies ($\sigma_h=1.20 d$) the membrane exhibits a flat bilayer.
430 >
431 > Increasing the head / body size ratio increases the local membrane
432 > curvature around each of the lipids.  With $\sigma_h=1.28 d$, the
433 > surface is still essentially flat, but the bilayer starts to exhibit
434 > signs of instability.  We have observed occasional defects where a
435 > line of lipid molecules on one leaf of the bilayer will dip down to
436 > interdigitate with the other leaf.  This gives each of the two bilayer
437 > leaves some local convexity near the line defect.  These structures,
438 > once developed in a simulation, are very stable and are spaced
439 > approximately 100 \AA\ away from each other.
440 >
441 > With larger heads ($\sigma_h = 1.35 d$) the membrane curvature
442 > resolves into a ``symmetric'' ripple phase.  Each leaf of the bilayer
443 > is broken into several convex, hemicylinderical sections, and opposite
444 > leaves are fitted together much like roof tiles.  There is no
445 > interdigitation between the upper and lower leaves of the bilayer.
446 >
447 > For the largest head / body ratios studied ($\sigma_h = 1.41 d$) the
448 > local curvature is substantially larger, and the resulting bilayer
449 > structure resolves into an asymmetric ripple phase.  This structure is
450 > very similar to the structures observed by both de~Vries {\it et al.}
451 > and Lenz {\it et al.}.  For a given ripple wave vector, there are two
452 > possible asymmetric ripples, which is not the case for the symmetric
453 > phase observed when $\sigma_h = 1.35 d$.
454 >
455 > \begin{figure}
456 > \centering
457 > \includegraphics[width=\linewidth]{./figures/mdPhaseCartoon.pdf}
458 > \caption{The role of the ratio between the head group size and the
459 > width of the molecular bodies is to increase the local membrane
460 > curvature.  With strong attractive interactions between the head
461 > groups, this local curvature can be maintained in bilayer structures
462 > through surface corrugation.  Shown above are three phases observed in
463 > these simulations.  With $\sigma_h = 1.20 d$, the bilayer maintains a
464 > flat topology.  For larger heads ($\sigma_h = 1.35 d$) the local
465 > curvature resolves into a symmetrically rippled phase with little or
466 > no interdigitation between the upper and lower leaves of the membrane.
467 > The largest heads studied ($\sigma_h = 1.41 d$) resolve into an
468 > asymmetric rippled phases with interdigitation between the two
469 > leaves.\label{mdfig:phaseCartoon}}
470 > \end{figure}
471 >
472 > Sample structures for the flat ($\sigma_h = 1.20 d$), symmetric
473 > ($\sigma_h = 1.35 d$, and asymmetric ($\sigma_h = 1.41 d$) ripple
474 > phases are shown in Figure \ref{mdfig:phaseCartoon}.  
475 >
476 > It is reasonable to ask how well the parameters we used can produce
477 > bilayer properties that match experimentally known values for real
478 > lipid bilayers.  Using a value of $l = 13.8$ \AA for the ellipsoidal
479 > tails and the fixed ellipsoidal aspect ratio of 3, our values for the
480 > area per lipid ($A$) and inter-layer spacing ($D_{HH}$) depend
481 > entirely on the size of the head bead relative to the molecular body.
482 > These values are tabulated in table \ref{mdtab:property}.  Kucera {\it
483 > et al.}  have measured values for the head group spacings for a number
484 > of PC lipid bilayers that range from 30.8 \AA\ (DLPC) to 37.8 \AA\ (DPPC).
485 > They have also measured values for the area per lipid that range from
486 > 60.6
487 > \AA$^2$ (DMPC) to 64.2 \AA$^2$
488 > (DPPC).\cite{NorbertKucerka04012005,NorbertKucerka06012006} Only the
489 > largest of the head groups we modeled ($\sigma_h = 1.41 d$) produces
490 > bilayers (specifically the area per lipid) that resemble real PC
491 > bilayers.  The smaller head beads we used are perhaps better models
492 > for PE head groups.
493 >
494 > \begin{table*}
495 > \begin{minipage}{\linewidth}
496 > \begin{center}
497 > \caption{Phase, bilayer spacing, area per lipid, ripple wavelength
498 > and amplitude observed as a function of the ratio between the head
499 > beads and the diameters of the tails.  Ripple wavelengths and
500 > amplitudes are normalized to the diameter of the tail ellipsoids.}
501 > \begin{tabular}{lccccc}
502 > \hline
503 > $\sigma_h / d$ & type of phase & bilayer spacing (\AA) & area per
504 > lipid (\AA$^2$) & $\lambda / d$ & $A / d$\\
505 > \hline
506 > 1.20 & flat & 33.4 & 49.6 & N/A & N/A \\
507 > 1.28 & flat & 33.7 & 54.7 & N/A & N/A \\
508 > 1.35 & symmetric ripple & 42.9 & 51.7 & 17.2 & 2.2 \\
509 > 1.41 & asymmetric ripple & 37.1 & 63.1 & 15.4 & 1.5 \\
510 > \end{tabular}
511 > \label{mdtab:property}
512 > \end{center}
513 > \end{minipage}
514 > \end{table*}
515 >
516 > The membrane structures and the reduced wavelength $\lambda / d$,
517 > reduced amplitude $A / d$ of the ripples are summarized in Table
518 > \ref{mdtab:property}. The wavelength range is $15 - 17$ molecular bodies
519 > and the amplitude is $1.5$ molecular bodies for asymmetric ripple and
520 > $2.2$ for symmetric ripple. These values are reasonably consistent
521 > with experimental measurements.\cite{Sun96,Katsaras00,Kaasgaard03}
522 > Note, that given the lack of structural freedom in the tails of our
523 > model lipids, the amplitudes observed from these simulations are
524 > likely to underestimate of the true amplitudes.
525 >
526 > \begin{figure}
527 > \centering
528 > \includegraphics[width=\linewidth]{./figures/mdTopDown.pdf}
529 > \caption{Top views of the flat (upper), symmetric ripple (middle),
530 > and asymmetric ripple (lower) phases.  Note that the head-group
531 > dipoles have formed head-to-tail chains in all three of these phases,
532 > but in the two rippled phases, the dipolar chains are all aligned {\it
533 > perpendicular} to the direction of the ripple.  Note that the flat
534 > membrane has multiple vortex defects in the dipolar ordering, and the
535 > ordering on the lower leaf of the bilayer can be in an entirely
536 > different direction from the upper leaf.\label{mdfig:topView}}
537 > \end{figure}
538 >
539 > The principal method for observing orientational ordering in dipolar
540 > or liquid crystalline systems is the $P_2$ order parameter (defined
541 > as $1.5 \times \lambda_{max}$, where $\lambda_{max}$ is the largest
542 > eigenvalue of the matrix,
543 > \begin{equation}
544 > {\mathsf{S}} = \frac{1}{N} \sum_i \left(
545 > \begin{array}{ccc}
546 >        u^{x}_i u^{x}_i-\frac{1}{3} & u^{x}_i u^{y}_i & u^{x}_i u^{z}_i \\
547 >        u^{y}_i u^{x}_i  & u^{y}_i u^{y}_i -\frac{1}{3} & u^{y}_i u^{z}_i \\
548 >        u^{z}_i u^{x}_i & u^{z}_i u^{y}_i  & u^{z}_i u^{z}_i -\frac{1}{3}
549 > \end{array} \right).
550 > \label{mdeq:opmatrix}
551 > \end{equation}
552 > Here $u^{\alpha}_i$ is the $\alpha=x,y,z$ component of the unit vector
553 > for molecule $i$.  (Here, $\hat{\bf u}_i$ can refer either to the
554 > principal axis of the molecular body or to the dipole on the head
555 > group of the molecule.)  $P_2$ will be $1.0$ for a perfectly-ordered
556 > system and near $0$ for a randomized system.  Note that this order
557 > parameter is {\em not} equal to the polarization of the system.  For
558 > example, the polarization of a perfect anti-ferroelectric arrangement
559 > of point dipoles is $0$, but $P_2$ for the same system is $1$.  The
560 > eigenvector of $\mathsf{S}$ corresponding to the largest eigenvalue is
561 > familiar as the director axis, which can be used to determine a
562 > privileged axis for an orientationally-ordered system.  Since the
563 > molecular bodies are perpendicular to the head group dipoles, it is
564 > possible for the director axes for the molecular bodies and the head
565 > groups to be completely decoupled from each other.
566 >
567 > Figure \ref{mdfig:topView} shows snapshots of bird's-eye views of the
568 > flat ($\sigma_h = 1.20 d$) and rippled ($\sigma_h = 1.35, 1.41 d$)
569 > bilayers.  The directions of the dipoles on the head groups are
570 > represented with two colored half spheres: blue (phosphate) and yellow
571 > (amino).  For flat bilayers, the system exhibits signs of
572 > orientational frustration; some disorder in the dipolar head-to-tail
573 > chains is evident with kinks visible at the edges between differently
574 > ordered domains.  The lipids can also move independently of lipids in
575 > the opposing leaf, so the ordering of the dipoles on one leaf is not
576 > necessarily consistent with the ordering on the other.  These two
577 > factors keep the total dipolar order parameter relatively low for the
578 > flat phases.
579 >
580 > With increasing head group size, the surface becomes corrugated, and
581 > the dipoles cannot move as freely on the surface. Therefore, the
582 > translational freedom of lipids in one layer is dependent upon the
583 > position of the lipids in the other layer.  As a result, the ordering of
584 > the dipoles on head groups in one leaf is correlated with the ordering
585 > in the other leaf.  Furthermore, as the membrane deforms due to the
586 > corrugation, the symmetry of the allowed dipolar ordering on each leaf
587 > is broken. The dipoles then self-assemble in a head-to-tail
588 > configuration, and the dipolar order parameter increases dramatically.
589 > However, the total polarization of the system is still close to zero.
590 > This is strong evidence that the corrugated structure is an
591 > anti-ferroelectric state.  It is also notable that the head-to-tail
592 > arrangement of the dipoles is always observed in a direction
593 > perpendicular to the wave vector for the surface corrugation.  This is
594 > a similar finding to what we observed in our earlier work on the
595 > elastic dipolar membranes.\cite{Sun2007}
596 >
597 > The $P_2$ order parameters (for both the molecular bodies and the head
598 > group dipoles) have been calculated to quantify the ordering in these
599 > phases.  Figure \ref{mdfig:rP2} shows that the $P_2$ order parameter for
600 > the head-group dipoles increases with increasing head group size. When
601 > the heads of the lipid molecules are small, the membrane is nearly
602 > flat. Since the in-plane packing is essentially a close packing of the
603 > head groups, the head dipoles exhibit frustration in their
604 > orientational ordering.
605 >
606 > The ordering trends for the tails are essentially opposite to the
607 > ordering of the head group dipoles. The tail $P_2$ order parameter
608 > {\it decreases} with increasing head size. This indicates that the
609 > surface is more curved with larger head / tail size ratios. When the
610 > surface is flat, all tails are pointing in the same direction (normal
611 > to the bilayer surface).  This simplified model appears to be
612 > exhibiting a smectic A fluid phase, similar to the real $L_{\beta}$
613 > phase.  We have not observed a smectic C gel phase ($L_{\beta'}$) for
614 > this model system.  Increasing the size of the heads results in
615 > rapidly decreasing $P_2$ ordering for the molecular bodies.
616 >
617 > \begin{figure}
618 > \centering
619 > \includegraphics[width=\linewidth]{./figures/mdRP2.pdf}
620 > \caption{The $P_2$ order parameters for head groups (circles) and
621 > molecular bodies (squares) as a function of the ratio of head group
622 > size ($\sigma_h$) to the width of the molecular bodies ($d$). \label{mdfig:rP2}}
623 > \end{figure}
624 >
625 > In addition to varying the size of the head groups, we studied the
626 > effects of the interactions between head groups on the structure of
627 > lipid bilayer by changing the strength of the dipoles.  Figure
628 > \ref{mdfig:sP2} shows how the $P_2$ order parameter changes with
629 > increasing strength of the dipole.  Generally, the dipoles on the head
630 > groups become more ordered as the strength of the interaction between
631 > heads is increased and become more disordered by decreasing the
632 > interaction strength.  When the interaction between the heads becomes
633 > too weak, the bilayer structure does not persist; all lipid molecules
634 > become dispersed in the solvent (which is non-polar in this
635 > molecular-scale model).  The critical value of the strength of the
636 > dipole depends on the size of the head groups.  The perfectly flat
637 > surface becomes unstable below $5$ Debye, while the  rippled
638 > surfaces melt at $8$ Debye (asymmetric) or $10$ Debye (symmetric).
639 >
640 > The ordering of the tails mirrors the ordering of the dipoles {\it
641 > except for the flat phase}. Since the surface is nearly flat in this
642 > phase, the order parameters are only weakly dependent on dipolar
643 > strength until it reaches $15$ Debye.  Once it reaches this value, the
644 > head group interactions are strong enough to pull the head groups
645 > close to each other and distort the bilayer structure. For a flat
646 > surface, a substantial amount of free volume between the head groups
647 > is normally available.  When the head groups are brought closer by
648 > dipolar interactions, the tails are forced to splay outward, first forming
649 > curved bilayers, and then inverted micelles.
650 >
651 > When $\sigma_h=1.28 d$, the $P_2$ order parameter decreases slightly
652 > when the strength of the dipole is increased above $16$ Debye. For
653 > rippled bilayers, there is less free volume available between the head
654 > groups. Therefore increasing dipolar strength weakly influences the
655 > structure of the membrane.  However, the increase in the body $P_2$
656 > order parameters implies that the membranes are being slightly
657 > flattened due to the effects of increasing head-group attraction.
658 >
659 > A very interesting behavior takes place when the head groups are very
660 > large relative to the molecular bodies ($\sigma_h = 1.41 d$) and the
661 > dipolar strength is relatively weak ($\mu < 9$ Debye). In this case,
662 > the two leaves of the bilayer become totally interdigitated with each
663 > other in large patches of the membrane.   With higher dipolar
664 > strength, the interdigitation is limited to single lines that run
665 > through the bilayer in a direction perpendicular to the ripple wave
666 > vector.
667 >
668 > \begin{figure}
669 > \centering
670 > \includegraphics[width=\linewidth]{./figures/mdSP2.pdf}
671 > \caption{The $P_2$ order parameters for head group dipoles (a) and
672 > molecular bodies (b) as a function of the strength of the dipoles.
673 > These order parameters are shown for four values of the head group /
674 > molecular width ratio ($\sigma_h / d$). \label{mdfig:sP2}}
675 > \end{figure}
676 >
677 > Figure \ref{mdfig:tP2} shows the dependence of the order parameters on
678 > temperature.  As expected, systems are more ordered at low
679 > temperatures, and more disordered at high temperatures.  All of the
680 > bilayers we studied can become unstable if the temperature becomes
681 > high enough.  The only interesting feature of the temperature
682 > dependence is in the flat surfaces ($\sigma_h=1.20 d$ and
683 > $\sigma_h=1.28 d$).  Here, when the temperature is increased above
684 > $310$K, there is enough jostling of the head groups to allow the
685 > dipolar frustration to resolve into more ordered states.  This results
686 > in a slight increase in the $P_2$ order parameter above this
687 > temperature.
688 >
689 > For the rippled surfaces ($\sigma_h=1.35 d$ and $\sigma_h=1.41 d$),
690 > there is a slightly increased orientational ordering in the molecular
691 > bodies above $290$K.  Since our model lacks the detailed information
692 > about the behavior of the lipid tails, this is the closest the model
693 > can come to depicting the ripple ($P_{\beta'}$) to fluid
694 > ($L_{\alpha}$) phase transition.  What we are observing is a
695 > flattening of the rippled structures made possible by thermal
696 > expansion of the tightly-packed head groups.  The lack of detailed
697 > chain configurations also makes it impossible for this model to depict
698 > the ripple to gel ($L_{\beta'}$) phase transition.
699 >
700 > \begin{figure}
701 > \centering
702 > \includegraphics[width=\linewidth]{./figures/mdTP2.pdf}
703 > \caption{The $P_2$ order parameters for head group dipoles (a) and
704 > molecular bodies (b) as a function of temperature.
705 > These order parameters are shown for four values of the head group /
706 > molecular width ratio ($\sigma_h / d$).\label{mdfig:tP2}}
707 > \end{figure}
708 >
709 > Fig. \ref{mdfig:phaseDiagram} shows a phase diagram for the model as a
710 > function of the head group / molecular width ratio ($\sigma_h / d$)
711 > and the strength of the head group dipole moment ($\mu$).  Note that
712 > the specific form of the bilayer phase is governed almost entirely by
713 > the head group / molecular width ratio, while the strength of the
714 > dipolar interactions between the head groups governs the stability of
715 > the bilayer phase.  Weaker dipoles result in unstable bilayer phases,
716 > while extremely strong dipoles can shift the equilibrium to an
717 > inverted micelle phase when the head groups are small.   Temperature
718 > has little effect on the actual bilayer phase observed, although higher
719 > temperatures can cause the unstable region to grow into the higher
720 > dipole region of this diagram.
721 >
722 > \begin{figure}
723 > \centering
724 > \includegraphics[width=\linewidth]{./figures/mdPhaseDiagram.pdf}
725 > \caption{Phase diagram for the simple molecular model as a function
726 > of the head group / molecular width ratio ($\sigma_h / d$) and the
727 > strength of the head group dipole moment
728 > ($\mu$).\label{mdfig:phaseDiagram}}
729 > \end{figure}
730 >
731 > We have computed translational diffusion constants for lipid molecules
732 > from the mean-square displacement,
733 > \begin{equation}
734 > D = \lim_{t \rightarrow \infty} \frac{1}{6 t} \langle {|\left({\bf r}_{i}(t) - {\bf r}_{i}(0) \right)|}^2 \rangle,
735 > \end{equation}
736 > of the lipid bodies. Translational diffusion constants for the
737 > different head-to-tail size ratios (all at 300 K) are shown in table
738 > \ref{mdtab:relaxation}.  We have also computed orientational correlation
739 > times for the head groups from fits of the second-order Legendre
740 > polynomial correlation function,
741 > \begin{equation}
742 > C_{\ell}(t)  =  \langle P_{\ell}\left({\bf \mu}_{i}(t) \cdot {\bf
743 > \mu}_{i}(0) \right) \rangle
744 > \end{equation}
745 > of the head group dipoles.  The orientational correlation functions
746 > appear to have multiple components in their decay: a fast ($12 \pm 2$
747 > ps) decay due to librational motion of the head groups, as well as
748 > moderate ($\tau^h_{\rm mid}$) and slow ($\tau^h_{\rm slow}$)
749 > components.  The fit values for the moderate and slow correlation
750 > times are listed in table \ref{mdtab:relaxation}.  Standard deviations
751 > in the fit time constants are quite large (on the order of the values
752 > themselves).
753 >
754 > Sparrman and Westlund used a multi-relaxation model for NMR lineshapes
755 > observed in gel, fluid, and ripple phases of DPPC and obtained
756 > estimates of a correlation time for water translational diffusion
757 > ($\tau_c$) of 20 ns.\cite{Sparrman2003} This correlation time
758 > corresponds to water bound to small regions of the lipid membrane.
759 > They further assume that the lipids can explore only a single period
760 > of the ripple (essentially moving in a nearly one-dimensional path to
761 > do so), and the correlation time can therefore be used to estimate a
762 > value for the translational diffusion constant of $2.25 \times
763 > 10^{-11} m^2 s^{-1}$.  The translational diffusion constants we obtain
764 > are in reasonable agreement with this experimentally determined
765 > value. However, the $T_2$ relaxation times obtained by Sparrman and
766 > Westlund are consistent with P-N vector reorientation timescales of
767 > 2-5 ms.  This is substantially slower than even the slowest component
768 > we observe in the decay of our orientational correlation functions.
769 > Other than the dipole-dipole interactions, our head groups have no
770 > shape anisotropy which would force them to move as a unit with
771 > neighboring molecules.  This would naturally lead to P-N reorientation
772 > times that are too fast when compared with experimental measurements.
773 >
774 > \begin{table*}
775 > \begin{minipage}{\linewidth}
776 > \begin{center}
777 > \caption{Fit values for the rotational correlation times for the head
778 > groups ($\tau^h$) and molecular bodies ($\tau^b$) as well as the
779 > translational diffusion constants for the molecule as a function of
780 > the head-to-body width ratio.  All correlation functions and transport
781 > coefficients were computed from microcanonical simulations with an
782 > average temperture of 300 K.  In all of the phases, the head group
783 > correlation functions decay with an fast librational contribution ($12
784 > \pm 1$ ps).  There are additional moderate ($\tau^h_{\rm mid}$) and
785 > slow $\tau^h_{\rm slow}$ contributions to orientational decay that
786 > depend strongly on the phase exhibited by the lipids.  The symmetric
787 > ripple phase ($\sigma_h / d = 1.35$) appears to exhibit the slowest
788 > molecular reorientation.}
789 > \begin{tabular}{lcccc}
790 > \hline
791 > $\sigma_h / d$ & $\tau^h_{\rm mid} (ns)$ & $\tau^h_{\rm
792 > slow} (\mu s)$ & $\tau^b (\mu s)$ & $D (\times 10^{-11} m^2 s^{-1})$ \\
793 > \hline
794 > 1.20 & $0.4$ &  $9.6$ & $9.5$ & $0.43(1)$ \\
795 > 1.28 & $2.0$ & $13.5$ & $3.0$ & $5.91(3)$ \\
796 > 1.35 & $3.2$ &  $4.0$ & $0.9$ & $3.42(1)$ \\
797 > 1.41 & $0.3$ & $23.8$ & $6.9$ & $7.16(1)$ \\
798 > \end{tabular}
799 > \label{mdtab:relaxation}
800 > \end{center}
801 > \end{minipage}
802 > \end{table*}
803 >
804 > \section{Discussion}
805 > \label{mdsec:discussion}
806 >
807 > Symmetric and asymmetric ripple phases have been observed to form in
808 > our molecular dynamics simulations of a simple molecular-scale lipid
809 > model. The lipid model consists of an dipolar head group and an
810 > ellipsoidal tail.  Within the limits of this model, an explanation for
811 > generalized membrane curvature is a simple mismatch in the size of the
812 > heads with the width of the molecular bodies.  With heads
813 > substantially larger than the bodies of the molecule, this curvature
814 > should be convex nearly everywhere, a requirement which could be
815 > resolved either with micellar or cylindrical phases.
816 >
817 > The persistence of a {\it bilayer} structure therefore requires either
818 > strong attractive forces between the head groups or exclusionary
819 > forces from the solvent phase.  To have a persistent bilayer structure
820 > with the added requirement of convex membrane curvature appears to
821 > result in corrugated structures like the ones pictured in
822 > Fig. \ref{mdfig:phaseCartoon}.  In each of the sections of these
823 > corrugated phases, the local curvature near a most of the head groups
824 > is convex.  These structures are held together by the extremely strong
825 > and directional interactions between the head groups.
826 >
827 > The attractive forces holding the bilayer together could either be
828 > non-directional (as in the work of Kranenburg and
829 > Smit),\cite{Kranenburg2005} or directional (as we have utilized in
830 > these simulations).  The dipolar head groups are key for the
831 > maintaining the bilayer structures exhibited by this particular model;
832 > reducing the strength of the dipole has the tendency to make the
833 > rippled phase disappear.  The dipoles are likely to form attractive
834 > head-to-tail configurations even in flat configurations, but the
835 > temperatures are high enough that vortex defects become prevalent in
836 > the flat phase.  The flat phase we observed therefore appears to be
837 > substantially above the Kosterlitz-Thouless transition temperature for
838 > a planar system of dipoles with this set of parameters.  For this
839 > reason, it would be interesting to observe the thermal behavior of the
840 > flat phase at substantially lower temperatures.
841 >
842 > One feature of this model is that an energetically favorable
843 > orientational ordering of the dipoles can be achieved by forming
844 > ripples.  The corrugation of the surface breaks the symmetry of the
845 > plane, making vortex defects somewhat more expensive, and stabilizing
846 > the long range orientational ordering for the dipoles in the head
847 > groups.  Most of the rows of the head-to-tail dipoles are parallel to
848 > each other and the system adopts a bulk anti-ferroelectric state.  We
849 > believe that this is the first time the organization of the head
850 > groups in ripple phases has been addressed.
851 >
852 > Although the size-mismatch between the heads and molecular bodies
853 > appears to be the primary driving force for surface convexity, the
854 > persistence of the bilayer through the use of rippled structures is a
855 > function of the strong, attractive interactions between the heads.
856 > One important prediction we can make using the results from this
857 > simple model is that if the dipole-dipole interaction is the leading
858 > contributor to the head group attractions, the wave vectors for the
859 > ripples should always be found {\it perpendicular} to the dipole
860 > director axis.  This echoes the prediction we made earlier for simple
861 > elastic dipolar membranes, and may suggest experimental designs which
862 > will test whether this is really the case in the phosphatidylcholine
863 > $P_{\beta'}$ phases.  The dipole director axis should also be easily
864 > computable for the all-atom and coarse-grained simulations that have
865 > been published in the literature.\cite{deVries05}
866 >
867 > Experimental verification of our predictions of dipolar orientation
868 > correlating with the ripple direction would require knowing both the
869 > local orientation of a rippled region of the membrane (available via
870 > AFM studies of supported bilayers) as well as the local ordering of
871 > the membrane dipoles. Obtaining information about the local
872 > orientations of the membrane dipoles may be available from
873 > fluorescence detected linear dichroism (LD).  Benninger {\it et al.}
874 > have recently used axially-specific chromophores
875 > 2-(4,4-difluoro-5,7-dimethyl-4-bora-3a,4a-diaza-s-indacene-3-pentanoyl)-1-hexadecanoyl-sn-glycero-3-phospocholine
876 > ($\beta$-BODIPY FL C5-HPC or BODIPY-PC) and 3,3'
877 > dioctadecyloxacarbocyanine perchlorate (DiO) in their
878 > fluorescence-detected linear dichroism (LD) studies of plasma
879 > membranes of living cells.\cite{Benninger:2005qy} The DiO dye aligns
880 > its transition moment perpendicular to the membrane normal, while the
881 > BODIPY-PC transition dipole is parallel with the membrane normal.
882 > Without a doubt, using fluorescence detection of linear dichroism in
883 > concert with AFM surface scanning would be difficult experiments to
884 > carry out.  However, there is some hope of performing experiments to
885 > either verify or falsify the predictions of our simulations.
886 >
887 > Although our model is simple, it exhibits some rich and unexpected
888 > behaviors.  It would clearly be a closer approximation to reality if
889 > we allowed bending motions between the dipoles and the molecular
890 > bodies, and if we replaced the rigid ellipsoids with ball-and-chain
891 > tails.  However, the advantages of this simple model (large system
892 > sizes, 50 fs time steps) allow us to rapidly explore the phase diagram
893 > for a wide range of parameters.  Our explanation of this rippling
894 > phenomenon will help us design more accurate molecular models for
895 > corrugated membranes and experiments to test whether or not
896 > dipole-dipole interactions exert an influence on membrane rippling.

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines