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

Comparing trunk/mdRipple/mdripple.tex (file contents):
Revision 3202 by gezelter, Wed Aug 1 16:07:12 2007 UTC vs.
Revision 3351 by gezelter, Fri Feb 29 22:02:20 2008 UTC

# Line 1 | Line 1
1 %\documentclass[aps,pre,twocolumn,amssymb,showpacs,floatfix]{revtex4}
2 %\documentclass[aps,pre,preprint,amssymb]{revtex4}
1   \documentclass[12pt]{article}
2 + \usepackage{graphicx}
3   \usepackage{times}
4   \usepackage{mathptm}
5   \usepackage{tabularx}
6   \usepackage{setspace}
7   \usepackage{amsmath}
8   \usepackage{amssymb}
10 \usepackage{graphicx}
9   \usepackage[ref]{overcite}
10   \pagestyle{plain}
11   \pagenumbering{arabic}
# Line 18 | Line 16
16   \renewcommand{\baselinestretch}{1.2}
17   \renewcommand\citemid{\ } % no comma in optional reference note
18  
19 +
20   \begin{document}
21   %\renewcommand{\thefootnote}{\fnsymbol{footnote}}
22   %\renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
23  
24   \bibliographystyle{achemso}
25  
26 < \title{Dipolar Ordering in Molecular-scale Models of the Ripple Phase
27 < in Lipid Membranes}
26 > \title{Dipolar ordering in the ripple phases of molecular-scale models
27 > of lipid membranes}
28   \author{Xiuquan Sun and J. Daniel Gezelter \\
29   Department of Chemistry and Biochemistry,\\
30   University of Notre Dame, \\
# Line 38 | Line 37 | The ripple phase in phosphatidylcholine (PC) bilayers
37   \maketitle
38  
39   \begin{abstract}
40 < The ripple phase in phosphatidylcholine (PC) bilayers has never been
41 < completely explained.
40 > Symmetric and asymmetric ripple phases have been observed to form in
41 > molecular dynamics simulations of a simple molecular-scale lipid
42 > model. The lipid model consists of an dipolar head group and an
43 > ellipsoidal tail.  Within the limits of this model, an explanation for
44 > generalized membrane curvature is a simple mismatch in the size of the
45 > heads with the width of the molecular bodies.  The persistence of a
46 > {\it bilayer} structure requires strong attractive forces between the
47 > head groups.  One feature of this model is that an energetically
48 > favorable orientational ordering of the dipoles can be achieved by
49 > out-of-plane membrane corrugation.  The corrugation of the surface
50 > stabilizes the long range orientational ordering for the dipoles in the
51 > head groups which then adopt a bulk anti-ferroelectric state. We
52 > observe a common feature of the corrugated dipolar membranes: the wave
53 > vectors for the surface ripples are always found to be perpendicular
54 > to the dipole director axis.  
55   \end{abstract}
56  
57   %\maketitle
58 + \newpage
59  
60   \section{Introduction}
61   \label{sec:Int}
# Line 61 | Line 74 | within the gel phase.~\cite{Cevc87}
74   experimental results provide strong support for a 2-dimensional
75   hexagonal packing lattice of the lipid molecules within the ripple
76   phase.  This is a notable change from the observed lipid packing
77 < within the gel phase.~\cite{Cevc87}
77 > within the gel phase,~\cite{Cevc87} although Tenchov {\it et al.} have
78 > recently observed near-hexagonal packing in some phosphatidylcholine
79 > (PC) gel phases.\cite{Tenchov2001} The X-ray diffraction work by
80 > Katsaras {\it et al.} showed that a rich phase diagram exhibiting both
81 > {\it asymmetric} and {\it symmetric} ripples is possible for lecithin
82 > bilayers.\cite{Katsaras00}
83  
84   A number of theoretical models have been presented to explain the
85   formation of the ripple phase. Marder {\it et al.} used a
86 < curvature-dependent Landau-de Gennes free-energy functional to predict
87 < a rippled phase.~\cite{Marder84} This model and other related continuum
88 < models predict higher fluidity in convex regions and that concave
89 < portions of the membrane correspond to more solid-like regions.
90 < Carlson and Sethna used a packing-competition model (in which head
91 < groups and chains have competing packing energetics) to predict the
92 < formation of a ripple-like phase.  Their model predicted that the
93 < high-curvature portions have lower-chain packing and correspond to
94 < more fluid-like regions.  Goldstein and Leibler used a mean-field
95 < approach with a planar model for {\em inter-lamellar} interactions to
96 < predict rippling in multilamellar phases.~\cite{Goldstein88} McCullough
97 < and Scott proposed that the {\em anisotropy of the nearest-neighbor
98 < interactions} coupled to hydrophobic constraining forces which
99 < restrict height differences between nearest neighbors is the origin of
100 < the ripple phase.~\cite{McCullough90} Lubensky and MacKintosh
101 < introduced a Landau theory for tilt order and curvature of a single
102 < membrane and concluded that {\em coupling of molecular tilt to membrane
103 < curvature} is responsible for the production of
104 < ripples.~\cite{Lubensky93} Misbah, Duplat and Houchmandzadeh proposed
105 < that {\em inter-layer dipolar interactions} can lead to ripple
106 < instabilities.~\cite{Misbah98} Heimburg presented a {\em coexistence
107 < model} for ripple formation in which he postulates that fluid-phase
108 < line defects cause sharp curvature between relatively flat gel-phase
109 < regions.~\cite{Heimburg00} Kubica has suggested that a lattice model of
110 < polar head groups could be valuable in trying to understand bilayer
111 < phase formation.~\cite{Kubica02} Bannerjee used Monte Carlo simulations
112 < of lamellar stacks of hexagonal lattices to show that large headgroups
113 < and molecular tilt with respect to the membrane normal vector can
114 < cause bulk rippling.~\cite{Bannerjee02}
86 > curvature-dependent Landau-de~Gennes free-energy functional to predict
87 > a rippled phase.~\cite{Marder84} This model and other related
88 > continuum models predict higher fluidity in convex regions and that
89 > concave portions of the membrane correspond to more solid-like
90 > regions.  Carlson and Sethna used a packing-competition model (in
91 > which head groups and chains have competing packing energetics) to
92 > predict the formation of a ripple-like phase.  Their model predicted
93 > that the high-curvature portions have lower-chain packing and
94 > correspond to more fluid-like regions.  Goldstein and Leibler used a
95 > mean-field approach with a planar model for {\em inter-lamellar}
96 > interactions to predict rippling in multilamellar
97 > phases.~\cite{Goldstein88} McCullough and Scott proposed that the {\em
98 > anisotropy of the nearest-neighbor interactions} coupled to
99 > hydrophobic constraining forces which restrict height differences
100 > between nearest neighbors is the origin of the ripple
101 > phase.~\cite{McCullough90} Lubensky and MacKintosh introduced a Landau
102 > theory for tilt order and curvature of a single membrane and concluded
103 > that {\em coupling of molecular tilt to membrane curvature} is
104 > responsible for the production of ripples.~\cite{Lubensky93} Misbah,
105 > Duplat and Houchmandzadeh proposed that {\em inter-layer dipolar
106 > interactions} can lead to ripple instabilities.~\cite{Misbah98}
107 > Heimburg presented a {\em coexistence model} for ripple formation in
108 > which he postulates that fluid-phase line defects cause sharp
109 > curvature between relatively flat gel-phase regions.~\cite{Heimburg00}
110 > Kubica has suggested that a lattice model of polar head groups could
111 > be valuable in trying to understand bilayer phase
112 > formation.~\cite{Kubica02} Bannerjee used Monte Carlo simulations of
113 > lamellar stacks of hexagonal lattices to show that large head groups
114 > and molecular tilt with respect to the membrane normal vector can
115 > cause bulk rippling.~\cite{Bannerjee02} Recently, Kranenburg and Smit
116 > described the formation of symmetric ripple-like structures using a
117 > coarse grained solvent-head-tail bead model.\cite{Kranenburg2005}
118 > Their lipids consisted of a short chain of head beads tied to the two
119 > longer ``chains''.
120  
121 < In contrast, few large-scale molecular modelling studies have been
121 > In contrast, few large-scale molecular modeling studies have been
122   done due to the large size of the resulting structures and the time
123   required for the phases of interest to develop.  With all-atom (and
124   even unified-atom) simulations, only one period of the ripple can be
125 < observed and only for timescales in the range of 10-100 ns.  One of
126 < the most interesting molecular simulations was carried out by De Vries
125 > observed and only for time scales in the range of 10-100 ns.  One of
126 > the most interesting molecular simulations was carried out by de~Vries
127   {\it et al.}~\cite{deVries05}. According to their simulation results,
128   the ripple consists of two domains, one resembling the gel bilayer,
129   while in the other, the two leaves of the bilayer are fully
# Line 122 | Line 145 | between headgroups and tails is strongly implicated as
145  
146   Although the organization of the tails of lipid molecules are
147   addressed by these molecular simulations and the packing competition
148 < between headgroups and tails is strongly implicated as the primary
148 > between head groups and tails is strongly implicated as the primary
149   driving force for ripple formation, questions about the ordering of
150 < the head groups in ripple phase has not been settled.
150 > the head groups in ripple phase have not been settled.
151  
152   In a recent paper, we presented a simple ``web of dipoles'' spin
153   lattice model which provides some physical insight into relationship
154   between dipolar ordering and membrane buckling.\cite{Sun2007} We found
155   that dipolar elastic membranes can spontaneously buckle, forming
156 < ripple-like topologies.  The driving force for the buckling in dipolar
157 < elastic membranes the antiferroelectric ordering of the dipoles, and
158 < this was evident in the ordering of the dipole director axis
159 < perpendicular to the wave vector of the surface ripples.  A similiar
156 > ripple-like topologies.  The driving force for the buckling of dipolar
157 > elastic membranes is the anti-ferroelectric ordering of the dipoles.
158 > This was evident in the ordering of the dipole director axis
159 > perpendicular to the wave vector of the surface ripples.  A similar
160   phenomenon has also been observed by Tsonchev {\it et al.} in their
161   work on the spontaneous formation of dipolar peptide chains into
162   curved nano-structures.\cite{Tsonchev04,Tsonchev04II}
# Line 167 | Line 190 | some fraction of the details of the chain dynamics neg
190   non-polar tails. Another fact is that the majority of lipid molecules
191   in the ripple phase are relatively rigid (i.e. gel-like) which makes
192   some fraction of the details of the chain dynamics negligible.  Figure
193 < \ref{fig:lipidModels} shows the molecular strucure of a DPPC
193 > \ref{fig:lipidModels} shows the molecular structure of a DPPC
194   molecule, as well as atomistic and molecular-scale representations of
195   a DPPC molecule.  The hydrophilic character of the head group is
196   largely due to the separation of charge between the nitrogen and
# Line 186 | Line 209 | modelling large length-scale properties of lipid
209   The ellipsoidal portions of the model interact via the Gay-Berne
210   potential which has seen widespread use in the liquid crystal
211   community.  Ayton and Voth have also used Gay-Berne ellipsoids for
212 < modelling large length-scale properties of lipid
212 > modeling large length-scale properties of lipid
213   bilayers.\cite{Ayton01} In its original form, the Gay-Berne potential
214   was a single site model for the interactions of rigid ellipsoidal
215   molecules.\cite{Gay81} It can be thought of as a modification of the
# Line 236 | Line 259 | ellipsoids in a {\it side-by-side} configuration.  Add
259  
260   Gay-Berne ellipsoids also have an energy scaling parameter,
261   $\epsilon^s$, which describes the well depth for two identical
262 < ellipsoids in a {\it side-by-side} configuration.  Additionaly, a well
262 > ellipsoids in a {\it side-by-side} configuration.  Additionally, a well
263   depth aspect ratio, $\epsilon^r = \epsilon^e / \epsilon^s$, describes
264   the ratio between the well depths in the {\it end-to-end} and
265   side-by-side configurations.  As in the range parameter, a set of
# Line 294 | Line 317 | models. $l / d$ is the ratio of the head group to body
317   \centering
318   \includegraphics[width=4in]{2lipidModel}
319   \caption{The parameters defining the behavior of the lipid
320 < models. $l / d$ is the ratio of the head group to body diameter.
320 > models. $\sigma_h / d$ is the ratio of the head group to body diameter.
321   Molecular bodies had a fixed aspect ratio of 3.0.  The solvent model
322   was a simplified 4-water bead ($\sigma_w \approx d$) that has been
323 < used in other coarse-grained (DPD) simulations.  The dipolar strength
323 > used in other coarse-grained simulations.  The dipolar strength
324   (and the temperature and pressure) were the only other parameters that
325   were varied systematically.\label{fig:lipidModel}}
326   \end{figure}
327  
328   To take into account the permanent dipolar interactions of the
329 < zwitterionic head groups, we place fixed dipole moments $\mu_{i}$ at
329 > zwitterionic head groups, we have placed fixed dipole moments $\mu_{i}$ at
330   one end of the Gay-Berne particles.  The dipoles are oriented at an
331   angle $\theta = \pi / 2$ relative to the major axis.  These dipoles
332 < are protected by a head ``bead'' with a range parameter which we have
332 > are protected by a head ``bead'' with a range parameter ($\sigma_h$) which we have
333   varied between $1.20 d$ and $1.41 d$.  The head groups interact with
334   each other using a combination of Lennard-Jones,
335   \begin{equation}
# Line 325 | Line 348 | For the interaction between nonequivalent uniaxial ell
348   along dipole $i$ and $\mathbf{\hat r}_{ij}$ is the unit vector
349   pointing along the inter-dipole vector $\mathbf{r}_{ij}$.  
350  
351 + Since the charge separation distance is so large in zwitterionic head
352 + groups (like the PC head groups), it would also be possible to use
353 + either point charges or a ``split dipole'' approximation,
354 + \begin{equation}
355 + V_{ij}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
356 + \hat{r}}_{ij})) = \frac{1}{{4\pi \epsilon_0 }}\left[ {\frac{{\mu _i  \cdot \mu _j }}{{R_{ij}^3 }} -
357 + \frac{{3\left( {\mu _i  \cdot r_{ij} } \right)\left( {\mu _i  \cdot
358 + r_{ij} } \right)}}{{R_{ij}^5 }}} \right]
359 + \end{equation}
360 + where $\mu _i$ and $\mu _j$ are the dipole moments of sites $i$ and
361 + $j$, $r_{ij}$ is vector between the two sites, and $R_{ij}$ is given
362 + by,
363 + \begin{equation}
364 + R_{ij}  = \sqrt {r_{ij}^2  + \frac{{d_i^2 }}{4} + \frac{{d_j^2
365 + }}{4}}.
366 + \end{equation}
367 + Here, $d_i$ and $d_j$ are charge separation distances associated with
368 + each of the two dipolar sites. This approximation to the multipole
369 + expansion maintains the fast fall-off of the multipole potentials but
370 + lacks the normal divergences when two polar groups get close to one
371 + another.
372 +
373   For the interaction between nonequivalent uniaxial ellipsoids (in this
374   case, between spheres and ellipsoids), the spheres are treated as
375   ellipsoids with an aspect ratio of 1 ($d = l$) and with an well depth
# Line 333 | Line 378 | The solvent model in our simulations is identical to o
378   et al.} and is appropriate for dissimilar uniaxial
379   ellipsoids.\cite{Cleaver96}
380  
381 < The solvent model in our simulations is identical to one used by
382 < Marrink {\it et al.}  in their dissipative particle dynamics (DPD)
383 < simulation of lipid bilayers.\cite{Marrink04} This solvent bead is a single
384 < site that represents four water molecules (m = 72 amu) and has
385 < comparable density and diffusive behavior to liquid water.  However,
386 < since there are no electrostatic sites on these beads, this solvent
387 < model cannot replicate the dielectric properties of water.
381 > The solvent model in our simulations is similar to the one used by
382 > Marrink {\it et al.}  in their coarse grained simulations of lipid
383 > bilayers.\cite{Marrink04} The solvent bead is a single site that
384 > represents four water molecules (m = 72 amu) and has comparable
385 > density and diffusive behavior to liquid water.  However, since there
386 > are no electrostatic sites on these beads, this solvent model cannot
387 > replicate the dielectric properties of water.  Note that although we
388 > are using larger cutoff and switching radii than Marrink {\it et al.},
389 > our solvent density at 300 K remains at 0.944 g cm$^{-3}$, and the
390 > solvent diffuses at 0.43 $\AA^2 ps^{-1}$ (only twice as fast as liquid
391 > water).
392 >
393   \begin{table*}
394   \begin{minipage}{\linewidth}
395   \begin{center}
# Line 365 | Line 415 | A switching function has been applied to all potential
415   \end{minipage}
416   \end{table*}
417  
418 < A switching function has been applied to all potentials to smoothly
419 < turn off the interactions between a range of $22$ and $25$ \AA.
418 > \section{Experimental Methodology}
419 > \label{sec:experiment}
420  
421   The parameters that were systematically varied in this study were the
422   size of the head group ($\sigma_h$), the strength of the dipole moment
423   ($\mu$), and the temperature of the system.  Values for $\sigma_h$
424 < ranged from 5.5 \AA\  to 6.5 \AA\ .  If the width of the tails is
425 < taken to be the unit of length, these head groups correspond to a
426 < range from $1.2 d$ to $1.41 d$.  Since the solvent beads are nearly
427 < identical in diameter to the tail ellipsoids, all distances that
428 < follow will be measured relative to this unit of distance.
424 > ranged from 5.5 \AA\ to 6.5 \AA.  If the width of the tails is taken
425 > to be the unit of length, these head groups correspond to a range from
426 > $1.2 d$ to $1.41 d$.  Since the solvent beads are nearly identical in
427 > diameter to the tail ellipsoids, all distances that follow will be
428 > measured relative to this unit of distance.  Because the solvent we
429 > are using is non-polar and has a dielectric constant of 1, values for
430 > $\mu$ are sampled from a range that is somewhat smaller than the 20.6
431 > Debye dipole moment of the PC head groups.
432  
380 \section{Experimental Methodology}
381 \label{sec:experiment}
382
433   To create unbiased bilayers, all simulations were started from two
434   perfectly flat monolayers separated by a 26 \AA\ gap between the
435   molecular bodies of the upper and lower leaves.  The separated
436 < monolayers were evolved in a vaccum with $x-y$ anisotropic pressure
436 > monolayers were evolved in a vacuum with $x-y$ anisotropic pressure
437   coupling. The length of $z$ axis of the simulations was fixed and a
438   constant surface tension was applied to enable real fluctuations of
439   the bilayer. Periodic boundary conditions were used, and $480-720$
440   lipid molecules were present in the simulations, depending on the size
441   of the head beads.  In all cases, the two monolayers spontaneously
442   collapsed into bilayer structures within 100 ps. Following this
443 < collapse, all systems were equlibrated for $100$ ns at $300$ K.
443 > collapse, all systems were equilibrated for $100$ ns at $300$ K.
444  
445   The resulting bilayer structures were then solvated at a ratio of $6$
446   solvent beads (24 water molecules) per lipid. These configurations
447   were then equilibrated for another $30$ ns. All simulations utilizing
448   the solvent were carried out at constant pressure ($P=1$ atm) with
449 < $3$D anisotropic coupling, and constant surface tension
450 < ($\gamma=0.015$ UNIT). Given the absence of fast degrees of freedom in
451 < this model, a timestep of $50$ fs was utilized with excellent energy
449 > $3$D anisotropic coupling, and small constant surface tension
450 > ($\gamma=0.015$ N/m). Given the absence of fast degrees of freedom in
451 > this model, a time step of $50$ fs was utilized with excellent energy
452   conservation.  Data collection for structural properties of the
453   bilayers was carried out during a final 5 ns run following the solvent
454 < equilibration.  All simulations were performed using the OOPSE
455 < molecular modeling program.\cite{Meineke05}
454 > equilibration.  Orientational correlation functions and diffusion
455 > constants were computed from 30 ns simulations in the microcanonical
456 > (NVE) ensemble using the average volume from the end of the constant
457 > pressure and surface tension runs.  The timestep on these final
458 > molecular dynamics runs was 25 fs.  No appreciable changes in phase
459 > structure were noticed upon switching to a microcanonical ensemble.
460 > All simulations were performed using the {\sc oopse} molecular
461 > modeling program.\cite{Meineke05}
462  
463 + A switching function was applied to all potentials to smoothly turn
464 + off the interactions between a range of $22$ and $25$ \AA.  The
465 + switching function was the standard (cubic) function,
466 + \begin{equation}
467 + s(r) =
468 +        \begin{cases}
469 +        1 & \text{if $r \le r_{\text{sw}}$},\\
470 +        \frac{(r_{\text{cut}} + 2r - 3r_{\text{sw}})(r_{\text{cut}} - r)^2}
471 +        {(r_{\text{cut}} - r_{\text{sw}})^3}
472 +        & \text{if $r_{\text{sw}} < r \le r_{\text{cut}}$}, \\
473 +        0 & \text{if $r > r_{\text{cut}}$.}
474 +        \end{cases}
475 + \label{eq:dipoleSwitching}
476 + \end{equation}
477 +
478   \section{Results}
479   \label{sec:results}
480  
481 < Snapshots in Figure \ref{fig:phaseCartoon} show that the membrane is
482 < more corrugated with increasing size of the head groups. The surface
483 < is nearly flat when $\sigma_h=1.20 d$. With $\sigma_h=1.28 d$,
484 < although the surface is still flat, the bilayer starts to splay
485 < inward; the upper leaf of the bilayer is connected to the lower leaf
486 < with an interdigitated line defect. Two periodicities with $100$ \AA\
487 < wavelengths were observed in the simulation. This structure is very
488 < similiar to the structure observed by de Vries and Lenz {\it et
489 < al.}. The same basic structure is also observed when $\sigma_h=1.41
490 < d$, but the wavelength of the surface corrugations depends sensitively
491 < on the size of the ``head'' beads. From the undulation spectrum, the
492 < corrugation is clearly non-thermal.
481 > The membranes in our simulations exhibit a number of interesting
482 > bilayer phases.  The surface topology of these phases depends most
483 > sensitively on the ratio of the size of the head groups to the width
484 > of the molecular bodies.  With heads only slightly larger than the
485 > bodies ($\sigma_h=1.20 d$) the membrane exhibits a flat bilayer.
486 >
487 > Increasing the head / body size ratio increases the local membrane
488 > curvature around each of the lipids.  With $\sigma_h=1.28 d$, the
489 > surface is still essentially flat, but the bilayer starts to exhibit
490 > signs of instability.  We have observed occasional defects where a
491 > line of lipid molecules on one leaf of the bilayer will dip down to
492 > interdigitate with the other leaf.  This gives each of the two bilayer
493 > leaves some local convexity near the line defect.  These structures,
494 > once developed in a simulation, are very stable and are spaced
495 > approximately 100 \AA\ away from each other.
496 >
497 > With larger heads ($\sigma_h = 1.35 d$) the membrane curvature
498 > resolves into a ``symmetric'' ripple phase.  Each leaf of the bilayer
499 > is broken into several convex, hemicylinderical sections, and opposite
500 > leaves are fitted together much like roof tiles.  There is no
501 > interdigitation between the upper and lower leaves of the bilayer.
502 >
503 > For the largest head / body ratios studied ($\sigma_h = 1.41 d$) the
504 > local curvature is substantially larger, and the resulting bilayer
505 > structure resolves into an asymmetric ripple phase.  This structure is
506 > very similar to the structures observed by both de~Vries {\it et al.}
507 > and Lenz {\it et al.}.  For a given ripple wave vector, there are two
508 > possible asymmetric ripples, which is not the case for the symmetric
509 > phase observed when $\sigma_h = 1.35 d$.
510 >
511   \begin{figure}[htb]
512   \centering
513   \includegraphics[width=4in]{phaseCartoon}
514 < \caption{A sketch to discribe the structure of the phases observed in
515 < our simulations.\label{fig:phaseCartoon}}
514 > \caption{The role of the ratio between the head group size and the
515 > width of the molecular bodies is to increase the local membrane
516 > curvature.  With strong attractive interactions between the head
517 > groups, this local curvature can be maintained in bilayer structures
518 > through surface corrugation.  Shown above are three phases observed in
519 > these simulations.  With $\sigma_h = 1.20 d$, the bilayer maintains a
520 > flat topology.  For larger heads ($\sigma_h = 1.35 d$) the local
521 > curvature resolves into a symmetrically rippled phase with little or
522 > no interdigitation between the upper and lower leaves of the membrane.
523 > The largest heads studied ($\sigma_h = 1.41 d$) resolve into an
524 > asymmetric rippled phases with interdigitation between the two
525 > leaves.\label{fig:phaseCartoon}}
526   \end{figure}
527  
528 < When $\sigma_h=1.35 d$, we observed another corrugated surface
529 < morphology.  This structure is different from the asymmetric rippled
530 < surface; there is no interdigitation between the upper and lower
531 < leaves of the bilayer. Each leaf of the bilayer is broken into several
532 < hemicylinderical sections, and opposite leaves are fitted together
533 < much like roof tiles. Unlike the surface in which the upper
534 < hemicylinder is always interdigitated on the leading or trailing edge
535 < of lower hemicylinder, this ``symmetric'' ripple has no prefered
536 < direction.  The corresponding structures are shown in Figure
537 < \ref{fig:phaseCartoon} for elucidation of the detailed structures of
538 < different phases.  The top panel in figure \ref{fig:phaseCartoon} is
539 < the flat phase, the middle panel shows the asymmetric ripple phase
540 < corresponding to $\sigma_h = 1.41 d$ and the lower panel shows the
541 < symmetric ripple phase observed when $\sigma_h=1.35 d$. In the
542 < symmetric ripple, the bilayer is continuous over the whole membrane,
543 < however, in asymmetric ripple phase, the bilayer domains are connected
544 < by thin interdigitated monolayers that share molecules between the
545 < upper and lower leaves.
528 > Sample structures for the flat ($\sigma_h = 1.20 d$), symmetric
529 > ($\sigma_h = 1.35 d$, and asymmetric ($\sigma_h = 1.41 d$) ripple
530 > phases are shown in Figure \ref{fig:phaseCartoon}.  
531 >
532 > It is reasonable to ask how well the parameters we used can produce
533 > bilayer properties that match experimentally known values for real
534 > lipid bilayers.  Using a value of $l = 13.8$ \AA for the ellipsoidal
535 > tails and the fixed ellipsoidal aspect ratio of 3, our values for the
536 > area per lipid ($A$) and inter-layer spacing ($D_{HH}$) depend
537 > entirely on the size of the head bead relative to the molecular body.
538 > These values are tabulated in table \ref{tab:property}.  Kucera {\it
539 > et al.}  have measured values for the head group spacings for a number
540 > of PC lipid bilayers that range from 30.8 \AA\ (DLPC) to 37.8 \AA\ (DPPC).
541 > They have also measured values for the area per lipid that range from
542 > 60.6
543 > \AA$^2$ (DMPC) to 64.2 \AA$^2$
544 > (DPPC).\cite{NorbertKucerka04012005,NorbertKucerka06012006} Only the
545 > largest of the head groups we modeled ($\sigma_h = 1.41 d$) produces
546 > bilayers (specifically the area per lipid) that resemble real PC
547 > bilayers.  The smaller head beads we used are perhaps better models
548 > for PE head groups.
549 >
550   \begin{table*}
551   \begin{minipage}{\linewidth}
552   \begin{center}
553 < \caption{Phases, ripple wavelengths and amplitudes observed as a
554 < function of the ratio between the head beads and the diameters of the
555 < tails.  All lengths are normalized to the diameter of the tail
556 < ellipsoids.}
557 < \begin{tabular}{lccc}
553 > \caption{Phase, bilayer spacing, area per lipid, ripple wavelength
554 > and amplitude observed as a function of the ratio between the head
555 > beads and the diameters of the tails.  Ripple wavelengths and
556 > amplitudes are normalized to the diameter of the tail ellipsoids.}
557 > \begin{tabular}{lccccc}
558   \hline
559 < $\sigma_h / d$ & type of phase & $\lambda / d$ & $A / d$\\
559 > $\sigma_h / d$ & type of phase & bilayer spacing (\AA) & area per
560 > lipid (\AA$^2$) & $\lambda / d$ & $A / d$\\
561   \hline
562 < 1.20 & flat & N/A & N/A \\
563 < 1.28 & asymmetric ripple or flat & 21.7 & N/A \\
564 < 1.35 & symmetric ripple & 17.2 & 2.2 \\
565 < 1.41 & asymmetric ripple & 15.4 & 1.5 \\
562 > 1.20 & flat & 33.4 & 49.6 & N/A & N/A \\
563 > 1.28 & flat & 33.7 & 54.7 & N/A & N/A \\
564 > 1.35 & symmetric ripple & 42.9 & 51.7 & 17.2 & 2.2 \\
565 > 1.41 & asymmetric ripple & 37.1 & 63.1 & 15.4 & 1.5 \\
566   \end{tabular}
567   \label{tab:property}
568   \end{center}
# Line 467 | Line 571 | reduced amplitude $A / d$ of the ripples are summarize
571  
572   The membrane structures and the reduced wavelength $\lambda / d$,
573   reduced amplitude $A / d$ of the ripples are summarized in Table
574 < \ref{tab:property}. The wavelength range is $15~21$ molecular bodies
574 > \ref{tab:property}. The wavelength range is $15 - 17$ molecular bodies
575   and the amplitude is $1.5$ molecular bodies for asymmetric ripple and
576 < $2.2$ for symmetric ripple. These values are consistent to the
577 < experimental results.  Note, that given the lack of structural freedom
578 < in the tails of our model lipids, the amplitudes observed from these
579 < simulations are likely to underestimate of the true amplitudes.
576 > $2.2$ for symmetric ripple. These values are reasonably consistent
577 > with experimental measurements.\cite{Sun96,Katsaras00,Kaasgaard03}
578 > Note, that given the lack of structural freedom in the tails of our
579 > model lipids, the amplitudes observed from these simulations are
580 > likely to underestimate of the true amplitudes.
581  
582   \begin{figure}[htb]
583   \centering
584   \includegraphics[width=4in]{topDown}
585 < \caption{Top views of the flat (upper), asymmetric ripple (middle),
586 < and symmetric ripple (lower) phases.  Note that the head-group dipoles
587 < have formed head-to-tail chains in all three of these phases, but in
588 < the two rippled phases, the dipolar chains are all aligned
589 < {\it perpendicular} to the direction of the ripple.  The flat membrane
590 < has multiple point defects in the dipolar orientational ordering, and
591 < the dipolar ordering on the lower leaf of the bilayer can be in a
592 < different direction from the upper leaf.\label{fig:topView}}
585 > \caption{Top views of the flat (upper), symmetric ripple (middle),
586 > and asymmetric ripple (lower) phases.  Note that the head-group
587 > dipoles have formed head-to-tail chains in all three of these phases,
588 > but in the two rippled phases, the dipolar chains are all aligned {\it
589 > perpendicular} to the direction of the ripple.  Note that the flat
590 > membrane has multiple vortex defects in the dipolar ordering, and the
591 > ordering on the lower leaf of the bilayer can be in an entirely
592 > different direction from the upper leaf.\label{fig:topView}}
593   \end{figure}
594  
595   The principal method for observing orientational ordering in dipolar
# Line 516 | Line 621 | flat ($\sigma_h = 1.20 d$) and rippled ($\sigma_h = 1.
621   groups to be completely decoupled from each other.
622  
623   Figure \ref{fig:topView} shows snapshots of bird's-eye views of the
624 < flat ($\sigma_h = 1.20 d$) and rippled ($\sigma_h = 1.41, 1.35 d$)
624 > flat ($\sigma_h = 1.20 d$) and rippled ($\sigma_h = 1.35, 1.41 d$)
625   bilayers.  The directions of the dipoles on the head groups are
626   represented with two colored half spheres: blue (phosphate) and yellow
627   (amino).  For flat bilayers, the system exhibits signs of
# Line 539 | Line 644 | antiferroelectric state.  It is also notable that the
644   configuration, and the dipolar order parameter increases dramatically.
645   However, the total polarization of the system is still close to zero.
646   This is strong evidence that the corrugated structure is an
647 < antiferroelectric state.  It is also notable that the head-to-tail
647 > anti-ferroelectric state.  It is also notable that the head-to-tail
648   arrangement of the dipoles is always observed in a direction
649   perpendicular to the wave vector for the surface corrugation.  This is
650   a similar finding to what we observed in our earlier work on the
# Line 580 | Line 685 | interaction stength.  When the interaction between the
685   increasing strength of the dipole.  Generally, the dipoles on the head
686   groups become more ordered as the strength of the interaction between
687   heads is increased and become more disordered by decreasing the
688 < interaction stength.  When the interaction between the heads becomes
688 > interaction strength.  When the interaction between the heads becomes
689   too weak, the bilayer structure does not persist; all lipid molecules
690   become dispersed in the solvent (which is non-polar in this
691 < molecular-scale model).  The critial value of the strength of the
691 > molecular-scale model).  The critical value of the strength of the
692   dipole depends on the size of the head groups.  The perfectly flat
693   surface becomes unstable below $5$ Debye, while the  rippled
694   surfaces melt at $8$ Debye (asymmetric) or $10$ Debye (symmetric).
# Line 596 | Line 701 | dipolar interactions, the tails are forced to splay ou
701   close to each other and distort the bilayer structure. For a flat
702   surface, a substantial amount of free volume between the head groups
703   is normally available.  When the head groups are brought closer by
704 < dipolar interactions, the tails are forced to splay outward, forming
705 < first curved bilayers, and then inverted micelles.
704 > dipolar interactions, the tails are forced to splay outward, first forming
705 > curved bilayers, and then inverted micelles.
706  
707   When $\sigma_h=1.28 d$, the $P_2$ order parameter decreases slightly
708 < when the strength of the dipole is increased above $16$ debye. For
708 > when the strength of the dipole is increased above $16$ Debye. For
709   rippled bilayers, there is less free volume available between the head
710   groups. Therefore increasing dipolar strength weakly influences the
711   structure of the membrane.  However, the increase in the body $P_2$
# Line 657 | Line 762 | molecular width ratio ($\sigma_h / d$).\label{fig:tP2}
762   molecular width ratio ($\sigma_h / d$).\label{fig:tP2}}
763   \end{figure}
764  
765 < \section{Discussion}
766 < \label{sec:discussion}
765 > Fig. \ref{fig:phaseDiagram} shows a phase diagram for the model as a
766 > function of the head group / molecular width ratio ($\sigma_h / d$)
767 > and the strength of the head group dipole moment ($\mu$).  Note that
768 > the specific form of the bilayer phase is governed almost entirely by
769 > the head group / molecular width ratio, while the strength of the
770 > dipolar interactions between the head groups governs the stability of
771 > the bilayer phase.  Weaker dipoles result in unstable bilayer phases,
772 > while extremely strong dipoles can shift the equilibrium to an
773 > inverted micelle phase when the head groups are small.   Temperature
774 > has little effect on the actual bilayer phase observed, although higher
775 > temperatures can cause the unstable region to grow into the higher
776 > dipole region of this diagram.
777  
778 < The ripple phases have been observed in our molecular dynamic
779 < simulations using a simple molecular lipid model. The lipid model
780 < consists of an anisotropic interacting dipolar head group and an
781 < ellipsoid shape tail. According to our simulations, the explanation of
782 < the formation for the ripples are originated in the size mismatch
783 < between the head groups and the tails. The ripple phases are only
784 < observed in the studies using larger head group lipid models. However,
785 < there is a mismatch betweent the size of the head groups and the size
671 < of the tails in the simulations of the flat surface. This indicates
672 < the competition between the anisotropic dipolar interaction and the
673 < packing of the tails also plays a major role for formation of the
674 < ripple phase. The larger head groups provide more free volume for the
675 < tails, while these hydrophobic ellipsoids trying to be close to each
676 < other, this gives the origin of the spontanous curvature of the
677 < surface, which is believed as the beginning of the ripple phases. The
678 < lager head groups cause the spontanous curvature inward for both of
679 < leaves of the bilayer. This results in a steric strain when the tails
680 < of two leaves too close to each other. The membrane has to be broken
681 < to release this strain. There are two ways to arrange these broken
682 < curvatures: symmetric and asymmetric ripples. Both of the ripple
683 < phases have been observed in our studies. The difference between these
684 < two ripples is that the bilayer is continuum in the symmetric ripple
685 < phase and is disrupt in the asymmetric ripple phase.
778 > \begin{figure}[htb]
779 > \centering
780 > \includegraphics[width=\linewidth]{phaseDiagram}
781 > \caption{Phase diagram for the simple molecular model as a function
782 > of the head group / molecular width ratio ($\sigma_h / d$) and the
783 > strength of the head group dipole moment
784 > ($\mu$).\label{fig:phaseDiagram}}
785 > \end{figure}
786  
787 < Dipolar head groups are the key elements for the maintaining of the
788 < bilayer structure. The lipids are solvated in water when lowering the
789 < the strength of the dipole on the head groups. The long range
790 < orientational ordering of the dipoles can be achieved by forming the
791 < ripples, although the dipoles are likely to form head-to-tail
792 < configurations even in flat surface, the frustration prevents the
793 < formation of the long range orientational ordering for dipoles. The
794 < corrugation of the surface breaks the frustration and stablizes the
795 < long range oreintational ordering for the dipoles in the head groups
796 < of the lipid molecules. Many rows of the head-to-tail dipoles are
797 < parallel to each other and adopt the antiferroelectric state as a
798 < whole. This is the first time the organization of the head groups in
799 < ripple phases of the lipid bilayer has been addressed.
787 > We have computed translational diffusion constants for lipid molecules
788 > from the mean-square displacement,
789 > \begin{equation}
790 > D = \lim_{t \rightarrow \infty} \frac{1}{6 t} \langle {|\left({\bf r}_{i}(t) - {\bf r}_{i}(0) \right)|}^2 \rangle,
791 > \end{equation}
792 > of the lipid bodies. Translational diffusion constants for the
793 > different head-to-tail size ratios (all at 300 K) are shown in table
794 > \ref{tab:relaxation}.  We have also computed orientational correlation
795 > times for the head groups from fits of the second-order Legendre
796 > polynomial correlation function,
797 > \begin{equation}
798 > C_{\ell}(t)  =  \langle P_{\ell}\left({\bf \mu}_{i}(t) \cdot {\bf
799 > \mu}_{i}(0) \right) \rangle
800 > \end{equation}
801 > of the head group dipoles.  The orientational correlation functions
802 > appear to have multiple components in their decay: a fast ($12 \pm 2$
803 > ps) decay due to librational motion of the head groups, as well as
804 > moderate ($\tau^h_{\rm mid}$) and slow ($\tau^h_{\rm slow}$)
805 > components.  The fit values for the moderate and slow correlation
806 > times are listed in table \ref{tab:relaxation}.  Standard deviations
807 > in the fit time constants are quite large (on the order of the values
808 > themselves).
809  
810 < The most important prediction we can make using the results from this
811 < simple model is that if dipolar ordering is driving the surface
812 < corrugation, the wave vectors for the ripples should always found to
813 < be {\it perpendicular} to the dipole director axis.  This prediction
814 < should suggest experimental designs which test whether this is really
815 < true in the phosphatidylcholine $P_{\beta'}$ phases.  The dipole
816 < director axis should also be easily computable for the all-atom and
817 < coarse-grained simulations that have been published in the literature.
810 > Sparrman and Westlund used a multi-relaxation model for NMR lineshapes
811 > observed in gel, fluid, and ripple phases of DPPC and obtained
812 > estimates of a correlation time for water translational diffusion
813 > ($\tau_c$) of 20 ns.\cite{Sparrman2003} This correlation time
814 > corresponds to water bound to small regions of the lipid membrane.
815 > They further assume that the lipids can explore only a single period
816 > of the ripple (essentially moving in a nearly one-dimensional path to
817 > do so), and the correlation time can therefore be used to estimate a
818 > value for the translational diffusion constant of $2.25 \times
819 > 10^{-11} m^2 s^{-1}$.  The translational diffusion constants we obtain
820 > are in reasonable agreement with this experimentally determined
821 > value. However, the $T_2$ relaxation times obtained by Sparrman and
822 > Westlund are consistent with P-N vector reorientation timescales of
823 > 2-5 ms.  This is substantially slower than even the slowest component
824 > we observe in the decay of our orientational correlation functions.
825 > Other than the dipole-dipole interactions, our head groups have no
826 > shape anisotropy which would force them to move as a unit with
827 > neighboring molecules.  This would naturally lead to P-N reorientation
828 > times that are too fast when compared with experimental measurements.
829  
830 + \begin{table*}
831 + \begin{minipage}{\linewidth}
832 + \begin{center}
833 + \caption{Fit values for the rotational correlation times for the head
834 + groups ($\tau^h$) and molecular bodies ($\tau^b$) as well as the
835 + translational diffusion constants for the molecule as a function of
836 + the head-to-body width ratio.  All correlation functions and transport
837 + coefficients were computed from microcanonical simulations with an
838 + average temperture of 300 K.  In all of the phases, the head group
839 + correlation functions decay with an fast librational contribution ($12
840 + \pm 1$ ps).  There are additional moderate ($\tau^h_{\rm mid}$) and
841 + slow $\tau^h_{\rm slow}$ contributions to orientational decay that
842 + depend strongly on the phase exhibited by the lipids.  The symmetric
843 + ripple phase ($\sigma_h / d = 1.35$) appears to exhibit the slowest
844 + molecular reorientation.}
845 + \begin{tabular}{lcccc}
846 + \hline
847 + $\sigma_h / d$ & $\tau^h_{\rm mid} (ns)$ & $\tau^h_{\rm
848 + slow} (\mu s)$ & $\tau^b (\mu s)$ & $D (\times 10^{-11} m^2 s^{-1})$ \\
849 + \hline
850 + 1.20 & $0.4$ &  $9.6$ & $9.5$ & $0.43(1)$ \\
851 + 1.28 & $2.0$ & $13.5$ & $3.0$ & $5.91(3)$ \\
852 + 1.35 & $3.2$ &  $4.0$ & $0.9$ & $3.42(1)$ \\
853 + 1.41 & $0.3$ & $23.8$ & $6.9$ & $7.16(1)$ \\
854 + \end{tabular}
855 + \label{tab:relaxation}
856 + \end{center}
857 + \end{minipage}
858 + \end{table*}
859 +
860 + \section{Discussion}
861 + \label{sec:discussion}
862 +
863 + Symmetric and asymmetric ripple phases have been observed to form in
864 + our molecular dynamics simulations of a simple molecular-scale lipid
865 + model. The lipid model consists of an dipolar head group and an
866 + ellipsoidal tail.  Within the limits of this model, an explanation for
867 + generalized membrane curvature is a simple mismatch in the size of the
868 + heads with the width of the molecular bodies.  With heads
869 + substantially larger than the bodies of the molecule, this curvature
870 + should be convex nearly everywhere, a requirement which could be
871 + resolved either with micellar or cylindrical phases.
872 +
873 + The persistence of a {\it bilayer} structure therefore requires either
874 + strong attractive forces between the head groups or exclusionary
875 + forces from the solvent phase.  To have a persistent bilayer structure
876 + with the added requirement of convex membrane curvature appears to
877 + result in corrugated structures like the ones pictured in
878 + Fig. \ref{fig:phaseCartoon}.  In each of the sections of these
879 + corrugated phases, the local curvature near a most of the head groups
880 + is convex.  These structures are held together by the extremely strong
881 + and directional interactions between the head groups.
882 +
883 + The attractive forces holding the bilayer together could either be
884 + non-directional (as in the work of Kranenburg and
885 + Smit),\cite{Kranenburg2005} or directional (as we have utilized in
886 + these simulations).  The dipolar head groups are key for the
887 + maintaining the bilayer structures exhibited by this particular model;
888 + reducing the strength of the dipole has the tendency to make the
889 + rippled phase disappear.  The dipoles are likely to form attractive
890 + head-to-tail configurations even in flat configurations, but the
891 + temperatures are high enough that vortex defects become prevalent in
892 + the flat phase.  The flat phase we observed therefore appears to be
893 + substantially above the Kosterlitz-Thouless transition temperature for
894 + a planar system of dipoles with this set of parameters.  For this
895 + reason, it would be interesting to observe the thermal behavior of the
896 + flat phase at substantially lower temperatures.
897 +
898 + One feature of this model is that an energetically favorable
899 + orientational ordering of the dipoles can be achieved by forming
900 + ripples.  The corrugation of the surface breaks the symmetry of the
901 + plane, making vortex defects somewhat more expensive, and stabilizing
902 + the long range orientational ordering for the dipoles in the head
903 + groups.  Most of the rows of the head-to-tail dipoles are parallel to
904 + each other and the system adopts a bulk anti-ferroelectric state.  We
905 + believe that this is the first time the organization of the head
906 + groups in ripple phases has been addressed.
907 +
908 + Although the size-mismatch between the heads and molecular bodies
909 + appears to be the primary driving force for surface convexity, the
910 + persistence of the bilayer through the use of rippled structures is a
911 + function of the strong, attractive interactions between the heads.
912 + One important prediction we can make using the results from this
913 + simple model is that if the dipole-dipole interaction is the leading
914 + contributor to the head group attractions, the wave vectors for the
915 + ripples should always be found {\it perpendicular} to the dipole
916 + director axis.  This echoes the prediction we made earlier for simple
917 + elastic dipolar membranes, and may suggest experimental designs which
918 + will test whether this is really the case in the phosphatidylcholine
919 + $P_{\beta'}$ phases.  The dipole director axis should also be easily
920 + computable for the all-atom and coarse-grained simulations that have
921 + been published in the literature.\cite{deVries05}
922 +
923 + Experimental verification of our predictions of dipolar orientation
924 + correlating with the ripple direction would require knowing both the
925 + local orientation of a rippled region of the membrane (available via
926 + AFM studies of supported bilayers) as well as the local ordering of
927 + the membrane dipoles. Obtaining information about the local
928 + orientations of the membrane dipoles may be available from
929 + fluorescence detected linear dichroism (LD).  Benninger {\it et al.}
930 + have recently used axially-specific chromophores
931 + 2-(4,4-difluoro-5,7-dimethyl-4-bora-3a,4a-diaza-s-indacene-3-pentanoyl)-1-hexadecanoyl-sn-glycero-3-phospocholine
932 + ($\beta$-BODIPY FL C5-HPC or BODIPY-PC) and 3,3'
933 + dioctadecyloxacarbocyanine perchlorate (DiO) in their
934 + fluorescence-detected linear dichroism (LD) studies of plasma
935 + membranes of living cells.\cite{Benninger:2005qy} The DiO dye aligns
936 + its transition moment perpendicular to the membrane normal, while the
937 + BODIPY-PC transition dipole is parallel with the membrane normal.
938 + Without a doubt, using fluorescence detection of linear dichroism in
939 + concert with AFM surface scanning would be difficult experiments to
940 + carry out.  However, there is some hope of performing experiments to
941 + either verify or falsify the predictions of our simulations.
942 +
943   Although our model is simple, it exhibits some rich and unexpected
944 < behaviors.  It would clearly be a closer approximation to the reality
945 < if we allowed greater translational freedom to the dipoles and
946 < replaced the somewhat artificial lattice packing and the harmonic
947 < elastic tension with more realistic molecular modeling potentials.
948 < What we have done is to present a simple model which exhibits bulk
949 < non-thermal corrugation, and our explanation of this rippling
944 > behaviors.  It would clearly be a closer approximation to reality if
945 > we allowed bending motions between the dipoles and the molecular
946 > bodies, and if we replaced the rigid ellipsoids with ball-and-chain
947 > tails.  However, the advantages of this simple model (large system
948 > sizes, 50 fs time steps) allow us to rapidly explore the phase diagram
949 > for a wide range of parameters.  Our explanation of this rippling
950   phenomenon will help us design more accurate molecular models for
951 < corrugated membranes and experiments to test whether rippling is
952 < dipole-driven or not.
720 <
951 > corrugated membranes and experiments to test whether or not
952 > dipole-dipole interactions exert an influence on membrane rippling.
953   \newpage
954   \bibliography{mdripple}
955   \end{document}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines