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 3362 by xsun, Fri Mar 7 01:53:18 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 >
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 > 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 > 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 > longer ``chains''.
43 >
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 > In Ch.~\ref{chap:mc}, we presented a simple ``web of dipoles'' spin
76 > lattice model which provides some physical insight into relationship
77 > between dipolar ordering and membrane buckling.\cite{sun:031602} We
78 > found that dipolar elastic membranes can spontaneously buckle, forming
79 > 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 > In this chapter, we construct a somewhat more realistic molecular-scale
88 > lipid model than our previous ``web of dipoles'' and use molecular
89 > dynamics simulations to elucidate the role of the head group dipoles
90 > in the formation and morphology of the ripple phase.  We describe our
91 > model and computational methodology in section \ref{mdsec:method}.
92 > Details on the simulations are presented in section
93 > \ref{mdsec:experiment}, with results following in section
94 > \ref{mdsec:results}.  A final discussion of the role of dipolar heads in
95 > the ripple formation can be found in section
96 > \ref{mdsec:discussion}.
97 >
98 > \section{Computational Model}
99 > \label{mdsec:method}
100 >
101 > \begin{figure}
102 > \centering
103 > \includegraphics[width=\linewidth]{./figures/mdLipidModels.pdf}
104 > \caption{Three different representations of DPPC lipid molecules,
105 > including the chemical structure, an atomistic model, and the
106 > head-body ellipsoidal coarse-grained model used in this
107 > work.\label{mdfig:lipidModels}}
108 > \end{figure}
109 >
110 > Our simple molecular-scale lipid model for studying the ripple phase
111 > is based on two facts: one is that the most essential feature of lipid
112 > molecules is their amphiphilic structure with polar head groups and
113 > non-polar tails. Another fact is that the majority of lipid molecules
114 > in the ripple phase are relatively rigid (i.e. gel-like) which makes
115 > some fraction of the details of the chain dynamics negligible.  Figure
116 > \ref{mdfig:lipidModels} shows the molecular structure of a DPPC
117 > molecule, as well as atomistic and molecular-scale representations of
118 > a DPPC molecule.  The hydrophilic character of the head group is
119 > largely due to the separation of charge between the nitrogen and
120 > phosphate groups.  The zwitterionic nature of the PC headgroups leads
121 > to abnormally large dipole moments (as high as 20.6 D), and this
122 > strongly polar head group interacts strongly with the solvating water
123 > layers immediately surrounding the membrane.  The hydrophobic tail
124 > consists of fatty acid chains.  In our molecular scale model, lipid
125 > molecules have been reduced to these essential features; the fatty
126 > acid chains are represented by an ellipsoid with a dipolar ball
127 > perched on one end to represent the effects of the charge-separated
128 > head group.  In real PC lipids, the direction of the dipole is
129 > nearly perpendicular to the tail, so we have fixed the direction of
130 > the point dipole rigidly in this orientation.  
131 >
132 > The ellipsoidal portions of the model interact via the Gay-Berne
133 > potential which has seen widespread use in the liquid crystal
134 > community.  Ayton and Voth have also used Gay-Berne ellipsoids for
135 > modeling large length-scale properties of lipid
136 > bilayers.\cite{Ayton01} In its original form, the Gay-Berne potential
137 > was a single site model for the interactions of rigid ellipsoidal
138 > molecules.\cite{Gay1981} It can be thought of as a modification of the
139 > Gaussian overlap model originally described by Berne and
140 > Pechukas.\cite{Berne72} The potential is constructed in the familiar
141 > form of the Lennard-Jones function using orientation-dependent
142 > $\sigma$ and $\epsilon$ parameters,
143 > \begin{equation}
144 > \begin{split}
145 > V_{ij}({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j}, {\mathbf{\hat
146 > r}_{ij}}) = & 4\epsilon ({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j},
147 > {\mathbf{\hat r}_{ij}})\left[ \left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u}_i},
148 > {\mathbf{\hat u}_j}, {\mathbf{\hat r}_{ij}})+\sigma_0}\right)^{12} \right.\\
149 > &\left. -\left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u}_i},
150 > {\mathbf{\hat u}_j}, {\mathbf{\hat
151 > r}_{ij}})+\sigma_0}\right)^6\right]
152 > \end{split}
153 > \label{mdeq:gb}
154 > \end{equation}
155 >
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 > \begin{equation}
176 > \begin{split}
177 > & \sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf \hat{r}}_{ij}) =
178 > \sigma_{0} \times  \\
179 > & \left[ 1- \left\{ \frac{ \chi \alpha^2 ({\bf \hat{u}}_i \cdot {\bf
180 > \hat{r}}_{ij} ) + \chi \alpha^{-2} ({\bf \hat{u}}_j \cdot {\bf
181 > \hat{r}}_{ij} ) - 2 \chi^2 ({\bf \hat{u}}_i \cdot {\bf
182 > \hat{r}}_{ij} )({\bf \hat{u}}_j \cdot {\bf
183 > \hat{r}}_{ij} ) ({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j)}{1 - \chi^2
184 > \left({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j\right)^2} \right\}
185 > \right]^{-1/2}
186 > \end{split}
187 > \end{equation}
188 >
189 > Gay-Berne ellipsoids also have an energy scaling parameter,
190 > $\epsilon^s$, which describes the well depth for two identical
191 > ellipsoids in a {\it side-by-side} configuration.  Additionally, a well
192 > depth aspect ratio, $\epsilon^r = \epsilon^e / \epsilon^s$, describes
193 > the ratio between the well depths in the {\it end-to-end} and
194 > side-by-side configurations.  As in the range parameter, a set of
195 > mixing and anisotropy variables can be used to describe the well
196 > depths for dissimilar particles,
197 > \begin {eqnarray*}
198 > \epsilon_0 & = & \sqrt{\epsilon^s_i  * \epsilon^s_j} \\ \\
199 > \epsilon^r & = & \sqrt{\epsilon^r_i  * \epsilon^r_j} \\ \\
200 > \chi' & = & \frac{1 - (\epsilon^r)^{1/\mu}}{1 + (\epsilon^r)^{1/\mu}}
201 > \\ \\
202 > \alpha'^2 & = & \left[1 + (\epsilon^r)^{1/\mu}\right]^{-1}
203 > \end{eqnarray*}
204 > The form of the strength function is somewhat complicated,
205 > \begin{eqnarray*}
206 > \epsilon({\bf \hat{u}}_{i}, {\bf \hat{u}}_{j},{\bf \hat{r}}_{ij}) & = &
207 > \epsilon_{0}  \epsilon_{1}^{\nu}({\bf \hat{u}}_{i}.{\bf \hat{u}}_{j})
208 > \epsilon_{2}^{\mu}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
209 > \hat{r}}_{ij}) \\ \\
210 > \epsilon_{1}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j}) & = &
211 > \left[1-\chi^{2}({\bf \hat{u}}_{i}.{\bf
212 > \hat{u}}_{j})^{2}\right]^{-1/2}
213 > \end{eqnarray*}
214 > \begin{equation*}
215 > \begin{split}
216 > & \epsilon_{2}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf \hat{r}}_{ij})
217 > = 1 - \\
218 > & \left\{ \frac{ \chi' \alpha'^2 ({\bf \hat{u}}_i \cdot {\bf
219 > \hat{r}}_{ij} ) + \chi' \alpha'^{-2} ({\bf \hat{u}}_j \cdot {\bf
220 > \hat{r}}_{ij} ) - 2 \chi'^2 ({\bf \hat{u}}_i \cdot {\bf
221 > \hat{r}}_{ij} )({\bf \hat{u}}_j \cdot {\bf
222 > \hat{r}}_{ij} ) ({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j)}{1 - \chi'^2
223 > \left({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j\right)^2} \right\},
224 > \end{split}
225 > \end{equation*}
226 > although many of the quantities and derivatives are identical with
227 > those obtained for the range parameter. Ref. \citen{Luckhurst90}
228 > has a particularly good explanation of the choice of the Gay-Berne
229 > parameters $\mu$ and $\nu$ for modeling liquid crystal molecules. An
230 > excellent overview of the computational methods that can be used to
231 > efficiently compute forces and torques for this potential can be found
232 > in Ref. \citen{Golubkov06}
233 >
234 > The choices of parameters we have used in this study correspond to a
235 > shape anisotropy of 3 for the chain portion of the molecule.  In
236 > principle, this could be varied to allow for modeling of longer or
237 > shorter chain lipid molecules. For these prolate ellipsoids, we have:
238 > \begin{equation}
239 > \begin{array}{rcl}
240 > d & < & l \\
241 > \epsilon^{r} & < & 1
242 > \end{array}
243 > \end{equation}
244 > A sketch of the various structural elements of our molecular-scale
245 > lipid / solvent model is shown in figure \ref{mdfig:lipidModel}.  The
246 > actual parameters used in our simulations are given in table
247 > \ref{mdtab:parameters}.
248 >
249 > \begin{figure}
250 > \centering
251 > \includegraphics[width=\linewidth]{./figures/md2LipidModel.pdf}
252 > \caption{The parameters defining the behavior of the lipid
253 > models. $\sigma_h / d$ is the ratio of the head group to body diameter.
254 > Molecular bodies had a fixed aspect ratio of 3.0.  The solvent model
255 > was a simplified 4-water bead ($\sigma_w \approx d$) that has been
256 > used in other coarse-grained simulations.  The dipolar strength
257 > (and the temperature and pressure) were the only other parameters that
258 > were varied systematically.\label{mdfig:lipidModel}}
259 > \end{figure}
260 >
261 > To take into account the permanent dipolar interactions of the
262 > zwitterionic head groups, we have placed fixed dipole moments
263 > $\mu_{i}$ at one end of the Gay-Berne particles.  The dipoles are
264 > oriented at an angle $\theta = \pi / 2$ relative to the major axis.
265 > These dipoles are protected by a head ``bead'' with a range parameter
266 > ($\sigma_h$) which we have varied between $1.20 d$ and $1.41 d$.  The
267 > head groups interact with each other using a combination of
268 > Lennard-Jones,
269 > \begin{equation}
270 > V_{ij}(r_{ij}) = 4\epsilon_h \left[\left(\frac{\sigma_h}{r_{ij}}\right)^{12} -
271 > \left(\frac{\sigma_h}{r_{ij}}\right)^6\right],
272 > \end{equation}
273 > and dipole-dipole,
274 > \begin{equation}
275 > V_{ij}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
276 > \hat{r}}_{ij})) = \frac{|\mu|^2}{4 \pi \epsilon_0 r_{ij}^3}
277 > \left[ \hat{\bf u}_i \cdot \hat{\bf u}_j - 3 (\hat{\bf u}_i \cdot
278 > \hat{\bf r}_{ij})(\hat{\bf u}_j \cdot \hat{\bf r}_{ij}) \right]
279 > \end{equation}
280 > potentials.  
281 > In these potentials, $\mathbf{\hat u}_i$ is the unit vector pointing
282 > along dipole $i$ and $\mathbf{\hat r}_{ij}$ is the unit vector
283 > pointing along the inter-dipole vector $\mathbf{r}_{ij}$.  
284 >
285 > Since the charge separation distance is so large in zwitterionic head
286 > groups (like the PC head groups), it would also be possible to use
287 > either point charges or a ``split dipole'' approximation,
288 > \begin{equation}
289 > V_{ij}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
290 > \hat{r}}_{ij})) = \frac{1}{{4\pi \epsilon_0 }}\left[ {\frac{{\mu _i  \cdot \mu _j }}{{R_{ij}^3 }} -
291 > \frac{{3\left( {\mu _i  \cdot r_{ij} } \right)\left( {\mu _i  \cdot
292 > r_{ij} } \right)}}{{R_{ij}^5 }}} \right]
293 > \end{equation}
294 > where $\mu _i$ and $\mu _j$ are the dipole moments of sites $i$ and
295 > $j$, $r_{ij}$ is vector between the two sites, and $R_{ij}$ is given
296 > by,
297 > \begin{equation}
298 > R_{ij}  = \sqrt {r_{ij}^2  + \frac{{d_i^2 }}{4} + \frac{{d_j^2
299 > }}{4}}.
300 > \end{equation}
301 > Here, $d_i$ and $d_j$ are charge separation distances associated with
302 > each of the two dipolar sites. This approximation to the multipole
303 > expansion maintains the fast fall-off of the multipole potentials but
304 > lacks the normal divergences when two polar groups get close to one
305 > another.
306 >
307 > For the interaction between nonequivalent uniaxial ellipsoids (in this
308 > case, between spheres and ellipsoids), the spheres are treated as
309 > ellipsoids with an aspect ratio of 1 ($d = l$) and with an well depth
310 > ratio ($\epsilon^r$) of 1 ($\epsilon^e = \epsilon^s$).  The form of
311 > the Gay-Berne potential we are using was generalized by Cleaver {\it
312 > et al.} and is appropriate for dissimilar uniaxial
313 > ellipsoids.\cite{Cleaver96}
314 >
315 > The solvent model in our simulations is similar to the one used by
316 > Marrink {\it et al.}  in their coarse grained simulations of lipid
317 > bilayers.\cite{Marrink2004} The solvent bead is a single site that
318 > represents four water molecules (m = 72 amu) and has comparable
319 > density and diffusive behavior to liquid water.  However, since there
320 > are no electrostatic sites on these beads, this solvent model cannot
321 > replicate the dielectric properties of water.  Note that although we
322 > are using larger cutoff and switching radii than Marrink {\it et al.},
323 > our solvent density at 300 K remains at 0.944 g cm$^{-3}$, and the
324 > solvent diffuses at 0.43 \AA$^2 ps^{-1}$ (only twice as fast as liquid
325 > water).
326 >
327 > \begin{table*}
328 > \begin{minipage}{\linewidth}
329 > \begin{center}
330 > \caption{Potential parameters used for molecular-scale coarse-grained
331 > lipid simulations}
332 > \begin{tabular}{llccc}
333 > \hline
334 >  & &  Head & Chain & Solvent \\
335 > \hline
336 > $d$ (\AA) & & varied & 4.6  & 4.7 \\
337 > $l$ (\AA) & & $= d$ & 13.8 & 4.7 \\
338 > $\epsilon^s$ (kcal/mol) & & 0.185 & 1.0 & 0.8 \\
339 > $\epsilon_r$ (well-depth aspect ratio)& & 1 & 0.2 &  1 \\
340 > $m$ (amu) & & 196 & 760 & 72.06 \\
341 > $\overleftrightarrow{\mathsf I}$ (amu \AA$^2$) & & & & \\
342 > \multicolumn{2}c{$I_{xx}$} & 1125 & 45000 & N/A \\
343 > \multicolumn{2}c{$I_{yy}$} & 1125 & 45000 & N/A \\
344 > \multicolumn{2}c{$I_{zz}$} &  0 &    9000 & N/A \\
345 > $\mu$ (Debye) & & varied & 0 & 0 \\
346 > \end{tabular}
347 > \label{mdtab:parameters}
348 > \end{center}
349 > \end{minipage}
350 > \end{table*}
351 >
352 > \section{Experimental Methodology}
353 > \label{mdsec:experiment}
354 >
355 > The parameters that were systematically varied in this study were the
356 > size of the head group ($\sigma_h$), the strength of the dipole moment
357 > ($\mu$), and the temperature of the system.  Values for $\sigma_h$
358 > ranged from 5.5 \AA\ to 6.5 \AA.  If the width of the tails is taken
359 > to be the unit of length, these head groups correspond to a range from
360 > $1.2 d$ to $1.41 d$.  Since the solvent beads are nearly identical in
361 > diameter to the tail ellipsoids, all distances that follow will be
362 > measured relative to this unit of distance.  Because the solvent we
363 > are using is non-polar and has a dielectric constant of 1, values for
364 > $\mu$ are sampled from a range that is somewhat smaller than the 20.6
365 > Debye dipole moment of the PC head groups.
366 >
367 > To create unbiased bilayers, all simulations were started from two
368 > perfectly flat monolayers separated by a 26 \AA\ gap between the
369 > molecular bodies of the upper and lower leaves.  The separated
370 > monolayers were evolved in a vacuum with $x-y$ anisotropic pressure
371 > coupling. The length of $z$ axis of the simulations was fixed and a
372 > constant surface tension was applied to enable real fluctuations of
373 > the bilayer. Periodic boundary conditions were used, and $480-720$
374 > lipid molecules were present in the simulations, depending on the size
375 > of the head beads.  In all cases, the two monolayers spontaneously
376 > collapsed into bilayer structures within 100 ps. Following this
377 > collapse, all systems were equilibrated for $100$ ns at $300$ K.
378 >
379 > The resulting bilayer structures were then solvated at a ratio of $6$
380 > solvent beads (24 water molecules) per lipid. These configurations
381 > were then equilibrated for another $30$ ns. All simulations utilizing
382 > the solvent were carried out at constant pressure ($P=1$ atm) with
383 > $3$D anisotropic coupling, and small constant surface tension
384 > ($\gamma=0.015$ N/m). Given the absence of fast degrees of freedom in
385 > this model, a time step of $50$ fs was utilized with excellent energy
386 > conservation.  Data collection for structural properties of the
387 > bilayers was carried out during a final 5 ns run following the solvent
388 > equilibration.  Orientational correlation functions and diffusion
389 > constants were computed from 30 ns simulations in the microcanonical
390 > (NVE) ensemble using the average volume from the end of the constant
391 > pressure and surface tension runs.  The timestep on these final
392 > molecular dynamics runs was 25 fs.  No appreciable changes in phase
393 > structure were noticed upon switching to a microcanonical ensemble.
394 > All simulations were performed using the {\sc oopse} molecular
395 > modeling program.\cite{Meineke2005}
396 >
397 > A switching function was applied to all potentials to smoothly turn
398 > off the interactions between a range of $22$ and $25$ \AA.  The
399 > switching function was the standard (cubic) function,
400 > \begin{equation}
401 > s(r) =
402 >        \begin{cases}
403 >        1 & \text{if $r \le r_{\text{sw}}$},\\
404 >        \frac{(r_{\text{cut}} + 2r - 3r_{\text{sw}})(r_{\text{cut}} - r)^2}
405 >        {(r_{\text{cut}} - r_{\text{sw}})^3}
406 >        & \text{if $r_{\text{sw}} < r \le r_{\text{cut}}$}, \\
407 >        0 & \text{if $r > r_{\text{cut}}$.}
408 >        \end{cases}
409 > \label{mdeq:dipoleSwitching}
410 > \end{equation}
411 >
412 > \section{Results}
413 > \label{mdsec:results}
414 >
415 > The membranes in our simulations exhibit a number of interesting
416 > bilayer phases.  The surface topology of these phases depends most
417 > sensitively on the ratio of the size of the head groups to the width
418 > of the molecular bodies.  With heads only slightly larger than the
419 > bodies ($\sigma_h=1.20 d$) the membrane exhibits a flat bilayer.
420 >
421 > Increasing the head / body size ratio increases the local membrane
422 > curvature around each of the lipids.  With $\sigma_h=1.28 d$, the
423 > surface is still essentially flat, but the bilayer starts to exhibit
424 > signs of instability.  We have observed occasional defects where a
425 > line of lipid molecules on one leaf of the bilayer will dip down to
426 > interdigitate with the other leaf.  This gives each of the two bilayer
427 > leaves some local convexity near the line defect.  These structures,
428 > once developed in a simulation, are very stable and are spaced
429 > approximately 100 \AA\ away from each other.
430 >
431 > With larger heads ($\sigma_h = 1.35 d$) the membrane curvature
432 > resolves into a ``symmetric'' ripple phase.  Each leaf of the bilayer
433 > is broken into several convex, hemicylinderical sections, and opposite
434 > leaves are fitted together much like roof tiles.  There is no
435 > interdigitation between the upper and lower leaves of the bilayer.
436 >
437 > For the largest head / body ratios studied ($\sigma_h = 1.41 d$) the
438 > local curvature is substantially larger, and the resulting bilayer
439 > structure resolves into an asymmetric ripple phase.  This structure is
440 > very similar to the structures observed by both de~Vries {\it et al.}
441 > and Lenz {\it et al.}.  For a given ripple wave vector, there are two
442 > possible asymmetric ripples, which is not the case for the symmetric
443 > phase observed when $\sigma_h = 1.35 d$.
444 >
445 > \begin{figure}
446 > \centering
447 > \includegraphics[width=\linewidth]{./figures/mdPhaseCartoon.pdf}
448 > \caption{The role of the ratio between the head group size and the
449 > width of the molecular bodies is to increase the local membrane
450 > curvature.  With strong attractive interactions between the head
451 > groups, this local curvature can be maintained in bilayer structures
452 > through surface corrugation.  Shown above are three phases observed in
453 > these simulations.  With $\sigma_h = 1.20 d$, the bilayer maintains a
454 > flat topology.  For larger heads ($\sigma_h = 1.35 d$) the local
455 > curvature resolves into a symmetrically rippled phase with little or
456 > no interdigitation between the upper and lower leaves of the membrane.
457 > The largest heads studied ($\sigma_h = 1.41 d$) resolve into an
458 > asymmetric rippled phases with interdigitation between the two
459 > leaves.\label{mdfig:phaseCartoon}}
460 > \end{figure}
461 >
462 > Sample structures for the flat ($\sigma_h = 1.20 d$), symmetric
463 > ($\sigma_h = 1.35 d$, and asymmetric ($\sigma_h = 1.41 d$) ripple
464 > phases are shown in Figure \ref{mdfig:phaseCartoon}.  
465 >
466 > It is reasonable to ask how well the parameters we used can produce
467 > bilayer properties that match experimentally known values for real
468 > lipid bilayers.  Using a value of $l = 13.8$ \AA~for the ellipsoidal
469 > tails and the fixed ellipsoidal aspect ratio of 3, our values for the
470 > area per lipid ($A$) and inter-layer spacing ($D_{HH}$) depend
471 > entirely on the size of the head bead relative to the molecular body.
472 > These values are tabulated in table \ref{mdtab:property}.  Kucera {\it
473 > et al.}  have measured values for the head group spacings for a number
474 > of PC lipid bilayers that range from 30.8 \AA\ (DLPC) to 37.8 \AA\ (DPPC).
475 > They have also measured values for the area per lipid that range from
476 > 60.6
477 > \AA$^2$ (DMPC) to 64.2 \AA$^2$
478 > (DPPC).\cite{NorbertKucerka04012005,NorbertKucerka06012006} Only the
479 > largest of the head groups we modeled ($\sigma_h = 1.41 d$) produces
480 > bilayers (specifically the area per lipid) that resemble real PC
481 > bilayers.  The smaller head beads we used are perhaps better models
482 > for PE head groups.
483 >
484 > \begin{table*}
485 > \begin{minipage}{\linewidth}
486 > \begin{center}
487 > \caption{Phase, bilayer spacing, area per lipid, ripple wavelength
488 > and amplitude observed as a function of the ratio between the head
489 > beads and the diameters of the tails.  Ripple wavelengths and
490 > amplitudes are normalized to the diameter of the tail ellipsoids.}
491 > \begin{tabular}{lccccc}
492 > \hline
493 > $\sigma_h / d$ & type of phase & bilayer spacing (\AA) & area per
494 > lipid (\AA$^2$) & $\lambda / d$ & $A / d$\\
495 > \hline
496 > 1.20 & flat & 33.4 & 49.6 & N/A & N/A \\
497 > 1.28 & flat & 33.7 & 54.7 & N/A & N/A \\
498 > 1.35 & symmetric ripple & 42.9 & 51.7 & 17.2 & 2.2 \\
499 > 1.41 & asymmetric ripple & 37.1 & 63.1 & 15.4 & 1.5 \\
500 > \end{tabular}
501 > \label{mdtab:property}
502 > \end{center}
503 > \end{minipage}
504 > \end{table*}
505 >
506 > The membrane structures and the reduced wavelength $\lambda / d$,
507 > reduced amplitude $A / d$ of the ripples are summarized in Table
508 > \ref{mdtab:property}. The wavelength range is $15 - 17$ molecular bodies
509 > and the amplitude is $1.5$ molecular bodies for asymmetric ripple and
510 > $2.2$ for symmetric ripple. These values are reasonably consistent
511 > with experimental measurements.\cite{Sun96,Katsaras00,Kaasgaard03}
512 > Note, that given the lack of structural freedom in the tails of our
513 > model lipids, the amplitudes observed from these simulations are
514 > likely to underestimate of the true amplitudes.
515 >
516 > \begin{figure}
517 > \centering
518 > \includegraphics[width=\linewidth]{./figures/mdTopDown.pdf}
519 > \caption{Top views of the flat (upper), symmetric ripple (middle),
520 > and asymmetric ripple (lower) phases.  Note that the head-group
521 > dipoles have formed head-to-tail chains in all three of these phases,
522 > but in the two rippled phases, the dipolar chains are all aligned {\it
523 > perpendicular} to the direction of the ripple.  Note that the flat
524 > membrane has multiple vortex defects in the dipolar ordering, and the
525 > ordering on the lower leaf of the bilayer can be in an entirely
526 > different direction from the upper leaf.\label{mdfig:topView}}
527 > \end{figure}
528 >
529 > The orientational ordering in the system is observed by $P_2$ order
530 > parameter, which is calculated from Eq.~\ref{mceq:opmatrix}
531 > in Ch.~\ref{chap:mc}. Here, $\hat{\bf u}_i$ can refer either to the
532 > principal axis of the molecular body or to the dipole on the head
533 > group of the molecule. Since the molecular bodies are perpendicular to
534 > the head group dipoles, it is possible for the director axes for the
535 > molecular bodies and the head groups to be completely decoupled from
536 > each other.
537 >
538 > Figure \ref{mdfig:topView} shows snapshots of bird's-eye views of the
539 > flat ($\sigma_h = 1.20 d$) and rippled ($\sigma_h = 1.35, 1.41 d$)
540 > bilayers.  The directions of the dipoles on the head groups are
541 > represented with two colored half spheres: blue (phosphate) and yellow
542 > (amino).  For flat bilayers, the system exhibits signs of
543 > orientational frustration; some disorder in the dipolar head-to-tail
544 > chains is evident with kinks visible at the edges between differently
545 > ordered domains.  The lipids can also move independently of lipids in
546 > the opposing leaf, so the ordering of the dipoles on one leaf is not
547 > necessarily consistent with the ordering on the other.  These two
548 > factors keep the total dipolar order parameter relatively low for the
549 > flat phases.
550 >
551 > With increasing head group size, the surface becomes corrugated, and
552 > the dipoles cannot move as freely on the surface. Therefore, the
553 > translational freedom of lipids in one layer is dependent upon the
554 > position of the lipids in the other layer.  As a result, the ordering of
555 > the dipoles on head groups in one leaf is correlated with the ordering
556 > in the other leaf.  Furthermore, as the membrane deforms due to the
557 > corrugation, the symmetry of the allowed dipolar ordering on each leaf
558 > is broken. The dipoles then self-assemble in a head-to-tail
559 > configuration, and the dipolar order parameter increases dramatically.
560 > However, the total polarization of the system is still close to zero.
561 > This is strong evidence that the corrugated structure is an
562 > anti-ferroelectric state.  It is also notable that the head-to-tail
563 > arrangement of the dipoles is always observed in a direction
564 > perpendicular to the wave vector for the surface corrugation.  This is
565 > a similar finding to what we observed in our earlier work on the
566 > elastic dipolar membranes.\cite{sun:031602}
567 >
568 > The $P_2$ order parameters (for both the molecular bodies and the head
569 > group dipoles) have been calculated to quantify the ordering in these
570 > phases.  Figure \ref{mdfig:rP2} shows that the $P_2$ order parameter for
571 > the head-group dipoles increases with increasing head group size. When
572 > the heads of the lipid molecules are small, the membrane is nearly
573 > flat. Since the in-plane packing is essentially a close packing of the
574 > head groups, the head dipoles exhibit frustration in their
575 > orientational ordering.
576 >
577 > The ordering trends for the tails are essentially opposite to the
578 > ordering of the head group dipoles. The tail $P_2$ order parameter
579 > {\it decreases} with increasing head size. This indicates that the
580 > surface is more curved with larger head / tail size ratios. When the
581 > surface is flat, all tails are pointing in the same direction (normal
582 > to the bilayer surface).  This simplified model appears to be
583 > exhibiting a smectic A fluid phase, similar to the real $L_{\beta}$
584 > phase.  We have not observed a smectic C gel phase ($L_{\beta'}$) for
585 > this model system.  Increasing the size of the heads results in
586 > rapidly decreasing $P_2$ ordering for the molecular bodies.
587 >
588 > \begin{figure}
589 > \centering
590 > \includegraphics[width=\linewidth]{./figures/mdRP2.pdf}
591 > \caption{The $P_2$ order parameters for head groups (circles) and
592 > molecular bodies (squares) as a function of the ratio of head group
593 > size ($\sigma_h$) to the width of the molecular bodies ($d$). \label{mdfig:rP2}}
594 > \end{figure}
595 >
596 > In addition to varying the size of the head groups, we studied the
597 > effects of the interactions between head groups on the structure of
598 > lipid bilayer by changing the strength of the dipoles.  Figure
599 > \ref{mdfig:sP2} shows how the $P_2$ order parameter changes with
600 > increasing strength of the dipole.  Generally, the dipoles on the head
601 > groups become more ordered as the strength of the interaction between
602 > heads is increased and become more disordered by decreasing the
603 > interaction strength.  When the interaction between the heads becomes
604 > too weak, the bilayer structure does not persist; all lipid molecules
605 > become dispersed in the solvent (which is non-polar in this
606 > molecular-scale model).  The critical value of the strength of the
607 > dipole depends on the size of the head groups.  The perfectly flat
608 > surface becomes unstable below $5$ Debye, while the  rippled
609 > surfaces melt at $8$ Debye (asymmetric) or $10$ Debye (symmetric).
610 >
611 > The ordering of the tails mirrors the ordering of the dipoles {\it
612 > except for the flat phase}. Since the surface is nearly flat in this
613 > phase, the order parameters are only weakly dependent on dipolar
614 > strength until it reaches $15$ Debye.  Once it reaches this value, the
615 > head group interactions are strong enough to pull the head groups
616 > close to each other and distort the bilayer structure. For a flat
617 > surface, a substantial amount of free volume between the head groups
618 > is normally available.  When the head groups are brought closer by
619 > dipolar interactions, the tails are forced to splay outward, first forming
620 > curved bilayers, and then inverted micelles.
621 >
622 > When $\sigma_h=1.28 d$, the $P_2$ order parameter decreases slightly
623 > when the strength of the dipole is increased above $16$ Debye. For
624 > rippled bilayers, there is less free volume available between the head
625 > groups. Therefore increasing dipolar strength weakly influences the
626 > structure of the membrane.  However, the increase in the body $P_2$
627 > order parameters implies that the membranes are being slightly
628 > flattened due to the effects of increasing head-group attraction.
629 >
630 > A very interesting behavior takes place when the head groups are very
631 > large relative to the molecular bodies ($\sigma_h = 1.41 d$) and the
632 > dipolar strength is relatively weak ($\mu < 9$ Debye). In this case,
633 > the two leaves of the bilayer become totally interdigitated with each
634 > other in large patches of the membrane.   With higher dipolar
635 > strength, the interdigitation is limited to single lines that run
636 > through the bilayer in a direction perpendicular to the ripple wave
637 > vector.
638 >
639 > \begin{figure}
640 > \centering
641 > \includegraphics[width=\linewidth]{./figures/mdSP2.pdf}
642 > \caption{The $P_2$ order parameters for head group dipoles (a) and
643 > molecular bodies (b) as a function of the strength of the dipoles.
644 > These order parameters are shown for four values of the head group /
645 > molecular width ratio ($\sigma_h / d$). \label{mdfig:sP2}}
646 > \end{figure}
647 >
648 > Figure \ref{mdfig:tP2} shows the dependence of the order parameters on
649 > temperature.  As expected, systems are more ordered at low
650 > temperatures, and more disordered at high temperatures.  All of the
651 > bilayers we studied can become unstable if the temperature becomes
652 > high enough.  The only interesting feature of the temperature
653 > dependence is in the flat surfaces ($\sigma_h=1.20 d$ and
654 > $\sigma_h=1.28 d$).  Here, when the temperature is increased above
655 > $310$K, there is enough jostling of the head groups to allow the
656 > dipolar frustration to resolve into more ordered states.  This results
657 > in a slight increase in the $P_2$ order parameter above this
658 > temperature.
659 >
660 > For the rippled surfaces ($\sigma_h=1.35 d$ and $\sigma_h=1.41 d$),
661 > there is a slightly increased orientational ordering in the molecular
662 > bodies above $290$K.  Since our model lacks the detailed information
663 > about the behavior of the lipid tails, this is the closest the model
664 > can come to depicting the ripple ($P_{\beta'}$) to fluid
665 > ($L_{\alpha}$) phase transition.  What we are observing is a
666 > flattening of the rippled structures made possible by thermal
667 > expansion of the tightly-packed head groups.  The lack of detailed
668 > chain configurations also makes it impossible for this model to depict
669 > the ripple to gel ($L_{\beta'}$) phase transition.
670 >
671 > \begin{figure}
672 > \centering
673 > \includegraphics[width=\linewidth]{./figures/mdTP2.pdf}
674 > \caption{The $P_2$ order parameters for head group dipoles (a) and
675 > molecular bodies (b) as a function of temperature.
676 > These order parameters are shown for four values of the head group /
677 > molecular width ratio ($\sigma_h / d$).\label{mdfig:tP2}}
678 > \end{figure}
679 >
680 > Fig. \ref{mdfig:phaseDiagram} shows a phase diagram for the model as a
681 > function of the head group / molecular width ratio ($\sigma_h / d$)
682 > and the strength of the head group dipole moment ($\mu$).  Note that
683 > the specific form of the bilayer phase is governed almost entirely by
684 > the head group / molecular width ratio, while the strength of the
685 > dipolar interactions between the head groups governs the stability of
686 > the bilayer phase.  Weaker dipoles result in unstable bilayer phases,
687 > while extremely strong dipoles can shift the equilibrium to an
688 > inverted micelle phase when the head groups are small.   Temperature
689 > has little effect on the actual bilayer phase observed, although higher
690 > temperatures can cause the unstable region to grow into the higher
691 > dipole region of this diagram.
692 >
693 > \begin{figure}
694 > \centering
695 > \includegraphics[width=\linewidth]{./figures/mdPhaseDiagram.pdf}
696 > \caption{Phase diagram for the simple molecular model as a function
697 > of the head group / molecular width ratio ($\sigma_h / d$) and the
698 > strength of the head group dipole moment
699 > ($\mu$).\label{mdfig:phaseDiagram}}
700 > \end{figure}
701 >
702 > We have computed translational diffusion constants for lipid molecules
703 > from the mean-square displacement,
704 > \begin{equation}
705 > D = \lim_{t \rightarrow \infty} \frac{1}{6 t} \langle {|\left({\bf r}_{i}(t) - {\bf r}_{i}(0) \right)|}^2 \rangle,
706 > \end{equation}
707 > of the lipid bodies. Translational diffusion constants for the
708 > different head-to-tail size ratios (all at 300 K) are shown in table
709 > \ref{mdtab:relaxation}.  We have also computed orientational correlation
710 > times for the head groups from fits of the second-order Legendre
711 > polynomial correlation function,
712 > \begin{equation}
713 > C_{\ell}(t)  =  \langle P_{\ell}\left({\bf \mu}_{i}(t) \cdot {\bf
714 > \mu}_{i}(0) \right) \rangle
715 > \end{equation}
716 > of the head group dipoles.  The orientational correlation functions
717 > appear to have multiple components in their decay: a fast ($12 \pm 2$
718 > ps) decay due to librational motion of the head groups, as well as
719 > moderate ($\tau^h_{\rm mid}$) and slow ($\tau^h_{\rm slow}$)
720 > components.  The fit values for the moderate and slow correlation
721 > times are listed in table \ref{mdtab:relaxation}.  Standard deviations
722 > in the fit time constants are quite large (on the order of the values
723 > themselves).
724 >
725 > Sparrman and Westlund used a multi-relaxation model for NMR lineshapes
726 > observed in gel, fluid, and ripple phases of DPPC and obtained
727 > estimates of a correlation time for water translational diffusion
728 > ($\tau_c$) of 20 ns.\cite{Sparrman2003} This correlation time
729 > corresponds to water bound to small regions of the lipid membrane.
730 > They further assume that the lipids can explore only a single period
731 > of the ripple (essentially moving in a nearly one-dimensional path to
732 > do so), and the correlation time can therefore be used to estimate a
733 > value for the translational diffusion constant of $2.25 \times
734 > 10^{-11} m^2 s^{-1}$.  The translational diffusion constants we obtain
735 > are in reasonable agreement with this experimentally determined
736 > value. However, the $T_2$ relaxation times obtained by Sparrman and
737 > Westlund are consistent with P-N vector reorientation timescales of
738 > 2-5 ms.  This is substantially slower than even the slowest component
739 > we observe in the decay of our orientational correlation functions.
740 > Other than the dipole-dipole interactions, our head groups have no
741 > shape anisotropy which would force them to move as a unit with
742 > neighboring molecules.  This would naturally lead to P-N reorientation
743 > times that are too fast when compared with experimental measurements.
744 >
745 > \begin{table*}
746 > \begin{minipage}{\linewidth}
747 > \begin{center}
748 > \caption{Fit values for the rotational correlation times for the head
749 > groups ($\tau^h$) and molecular bodies ($\tau^b$) as well as the
750 > translational diffusion constants for the molecule as a function of
751 > the head-to-body width ratio.  All correlation functions and transport
752 > coefficients were computed from microcanonical simulations with an
753 > average temperture of 300 K.  In all of the phases, the head group
754 > correlation functions decay with an fast librational contribution ($12
755 > \pm 1$ ps).  There are additional moderate ($\tau^h_{\rm mid}$) and
756 > slow $\tau^h_{\rm slow}$ contributions to orientational decay that
757 > depend strongly on the phase exhibited by the lipids.  The symmetric
758 > ripple phase ($\sigma_h / d = 1.35$) appears to exhibit the slowest
759 > molecular reorientation.}
760 > \begin{tabular}{lcccc}
761 > \hline
762 > $\sigma_h / d$ & $\tau^h_{\rm mid} (ns)$ & $\tau^h_{\rm
763 > slow} (\mu s)$ & $\tau^b (\mu s)$ & $D (\times 10^{-11} m^2 s^{-1})$ \\
764 > \hline
765 > 1.20 & $0.4$ &  $9.6$ & $9.5$ & $0.43(1)$ \\
766 > 1.28 & $2.0$ & $13.5$ & $3.0$ & $5.91(3)$ \\
767 > 1.35 & $3.2$ &  $4.0$ & $0.9$ & $3.42(1)$ \\
768 > 1.41 & $0.3$ & $23.8$ & $6.9$ & $7.16(1)$ \\
769 > \end{tabular}
770 > \label{mdtab:relaxation}
771 > \end{center}
772 > \end{minipage}
773 > \end{table*}
774 >
775 > \section{Discussion}
776 > \label{mdsec:discussion}
777 >
778 > Symmetric and asymmetric ripple phases have been observed to form in
779 > our molecular dynamics simulations of a simple molecular-scale lipid
780 > model. The lipid model consists of an dipolar head group and an
781 > ellipsoidal tail.  Within the limits of this model, an explanation for
782 > generalized membrane curvature is a simple mismatch in the size of the
783 > heads with the width of the molecular bodies.  With heads
784 > substantially larger than the bodies of the molecule, this curvature
785 > should be convex nearly everywhere, a requirement which could be
786 > resolved either with micellar or cylindrical phases.
787 >
788 > The persistence of a {\it bilayer} structure therefore requires either
789 > strong attractive forces between the head groups or exclusionary
790 > forces from the solvent phase.  To have a persistent bilayer structure
791 > with the added requirement of convex membrane curvature appears to
792 > result in corrugated structures like the ones pictured in
793 > Fig. \ref{mdfig:phaseCartoon}.  In each of the sections of these
794 > corrugated phases, the local curvature near a most of the head groups
795 > is convex.  These structures are held together by the extremely strong
796 > and directional interactions between the head groups.
797 >
798 > The attractive forces holding the bilayer together could either be
799 > non-directional (as in the work of Kranenburg and
800 > Smit),\cite{Kranenburg2005} or directional (as we have utilized in
801 > these simulations).  The dipolar head groups are key for the
802 > maintaining the bilayer structures exhibited by this particular model;
803 > reducing the strength of the dipole has the tendency to make the
804 > rippled phase disappear.  The dipoles are likely to form attractive
805 > head-to-tail configurations even in flat configurations, but the
806 > temperatures are high enough that vortex defects become prevalent in
807 > the flat phase.  The flat phase we observed therefore appears to be
808 > substantially above the Kosterlitz-Thouless transition temperature for
809 > a planar system of dipoles with this set of parameters.  For this
810 > reason, it would be interesting to observe the thermal behavior of the
811 > flat phase at substantially lower temperatures.
812 >
813 > One feature of this model is that an energetically favorable
814 > orientational ordering of the dipoles can be achieved by forming
815 > ripples.  The corrugation of the surface breaks the symmetry of the
816 > plane, making vortex defects somewhat more expensive, and stabilizing
817 > the long range orientational ordering for the dipoles in the head
818 > groups.  Most of the rows of the head-to-tail dipoles are parallel to
819 > each other and the system adopts a bulk anti-ferroelectric state.  We
820 > believe that this is the first time the organization of the head
821 > groups in ripple phases has been addressed.
822 >
823 > Although the size-mismatch between the heads and molecular bodies
824 > appears to be the primary driving force for surface convexity, the
825 > persistence of the bilayer through the use of rippled structures is a
826 > function of the strong, attractive interactions between the heads.
827 > One important prediction we can make using the results from this
828 > simple model is that if the dipole-dipole interaction is the leading
829 > contributor to the head group attractions, the wave vectors for the
830 > ripples should always be found {\it perpendicular} to the dipole
831 > director axis.  This echoes the prediction we made earlier for simple
832 > elastic dipolar membranes, and may suggest experimental designs which
833 > will test whether this is really the case in the phosphatidylcholine
834 > $P_{\beta'}$ phases.  The dipole director axis should also be easily
835 > computable for the all-atom and coarse-grained simulations that have
836 > been published in the literature.\cite{deVries05}
837 >
838 > Experimental verification of our predictions of dipolar orientation
839 > correlating with the ripple direction would require knowing both the
840 > local orientation of a rippled region of the membrane (available via
841 > AFM studies of supported bilayers) as well as the local ordering of
842 > the membrane dipoles. Obtaining information about the local
843 > orientations of the membrane dipoles may be available from
844 > fluorescence detected linear dichroism (LD).  Benninger {\it et al.}
845 > have recently used axially-specific chromophores
846 > 2-(4,4-difluoro-5,7-dimethyl-4-bora-3a,4a-diaza-s-indacene-3-pentanoyl)-1-hexadecanoyl-sn-glycero-3-phospocholine
847 > ($\beta$-BODIPY FL C5-HPC or BODIPY-PC) and 3,3'
848 > dioctadecyloxacarbocyanine perchlorate (DiO) in their
849 > fluorescence-detected linear dichroism (LD) studies of plasma
850 > membranes of living cells.\cite{Benninger:2005qy} The DiO dye aligns
851 > its transition moment perpendicular to the membrane normal, while the
852 > BODIPY-PC transition dipole is parallel with the membrane normal.
853 > Without a doubt, using fluorescence detection of linear dichroism in
854 > concert with AFM surface scanning would be difficult experiments to
855 > carry out.  However, there is some hope of performing experiments to
856 > either verify or falsify the predictions of our simulations.
857 >
858 > Although our model is simple, it exhibits some rich and unexpected
859 > behaviors.  It would clearly be a closer approximation to reality if
860 > we allowed bending motions between the dipoles and the molecular
861 > bodies, and if we replaced the rigid ellipsoids with ball-and-chain
862 > tails.  However, the advantages of this simple model (large system
863 > sizes, 50 fs time steps) allow us to rapidly explore the phase diagram
864 > for a wide range of parameters.  Our explanation of this rippling
865 > phenomenon will help us design more accurate molecular models for
866 > corrugated membranes and experiments to test whether or not
867 > dipole-dipole interactions exert an influence on membrane rippling.

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines