ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/xDissertation/md.tex
Revision: 3370
Committed: Fri Mar 14 21:38:07 2008 UTC (16 years, 5 months ago) by xsun
Content type: application/x-tex
File size: 45245 byte(s)
Log Message:
writing.

File Contents

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