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 3270 by xsun, Fri Oct 26 22:19:44 2007 UTC

# Line 24 | Line 24
24  
25   \bibliographystyle{achemso}
26  
27 < \title{Dipolar Ordering in Molecular-scale Models of the Ripple Phase
28 < in Lipid Membranes}
27 > \title{Dipolar ordering in the ripple phases of molecular-scale models
28 > of lipid membranes}
29   \author{Xiuquan Sun and J. Daniel Gezelter \\
30   Department of Chemistry and Biochemistry,\\
31   University of Notre Dame, \\
# Line 38 | Line 38 | The ripple phase in phosphatidylcholine (PC) bilayers
38   \maketitle
39  
40   \begin{abstract}
41 < The ripple phase in phosphatidylcholine (PC) bilayers has never been
42 < completely explained.
41 > Symmetric and asymmetric ripple phases have been observed to form in
42 > molecular dynamics simulations of a simple molecular-scale lipid
43 > model. The lipid model consists of an dipolar head group and an
44 > ellipsoidal tail.  Within the limits of this model, an explanation for
45 > generalized membrane curvature is a simple mismatch in the size of the
46 > heads with the width of the molecular bodies.  The persistence of a
47 > {\it bilayer} structure requires strong attractive forces between the
48 > head groups.  One feature of this model is that an energetically
49 > favorable orientational ordering of the dipoles can be achieved by
50 > out-of-plane membrane corrugation.  The corrugation of the surface
51 > stabilizes the long range orientational ordering for the dipoles in the
52 > head groups which then adopt a bulk anti-ferroelectric state. We
53 > observe a common feature of the corrugated dipolar membranes: the wave
54 > vectors for the surface ripples are always found to be perpendicular
55 > to the dipole director axis.  
56   \end{abstract}
57  
58   %\maketitle
59 + \newpage
60  
61   \section{Introduction}
62   \label{sec:Int}
# Line 61 | Line 75 | within the gel phase.~\cite{Cevc87}
75   experimental results provide strong support for a 2-dimensional
76   hexagonal packing lattice of the lipid molecules within the ripple
77   phase.  This is a notable change from the observed lipid packing
78 < within the gel phase.~\cite{Cevc87}
78 > within the gel phase,~\cite{Cevc87} although Tenchov {\it et al.} have
79 > recently observed near-hexagonal packing in some phosphatidylcholine
80 > (PC) gel phases.\cite{Tenchov2001} The X-ray diffraction work by
81 > Katsaras {\it et al.} showed that a rich phase diagram exhibiting both
82 > {\it asymmetric} and {\it symmetric} ripples is possible for lecithin
83 > bilayers.\cite{Katsaras00}
84  
85   A number of theoretical models have been presented to explain the
86   formation of the ripple phase. Marder {\it et al.} used a
87 < curvature-dependent Landau-de Gennes free-energy functional to predict
88 < a rippled phase.~\cite{Marder84} This model and other related continuum
89 < models predict higher fluidity in convex regions and that concave
90 < portions of the membrane correspond to more solid-like regions.
91 < Carlson and Sethna used a packing-competition model (in which head
92 < groups and chains have competing packing energetics) to predict the
93 < formation of a ripple-like phase.  Their model predicted that the
94 < high-curvature portions have lower-chain packing and correspond to
95 < more fluid-like regions.  Goldstein and Leibler used a mean-field
96 < approach with a planar model for {\em inter-lamellar} interactions to
97 < predict rippling in multilamellar phases.~\cite{Goldstein88} McCullough
98 < and Scott proposed that the {\em anisotropy of the nearest-neighbor
99 < interactions} coupled to hydrophobic constraining forces which
100 < restrict height differences between nearest neighbors is the origin of
101 < the ripple phase.~\cite{McCullough90} Lubensky and MacKintosh
102 < introduced a Landau theory for tilt order and curvature of a single
103 < membrane and concluded that {\em coupling of molecular tilt to membrane
104 < curvature} is responsible for the production of
105 < ripples.~\cite{Lubensky93} Misbah, Duplat and Houchmandzadeh proposed
106 < that {\em inter-layer dipolar interactions} can lead to ripple
107 < instabilities.~\cite{Misbah98} Heimburg presented a {\em coexistence
108 < model} for ripple formation in which he postulates that fluid-phase
109 < line defects cause sharp curvature between relatively flat gel-phase
110 < regions.~\cite{Heimburg00} Kubica has suggested that a lattice model of
111 < polar head groups could be valuable in trying to understand bilayer
112 < phase formation.~\cite{Kubica02} Bannerjee used Monte Carlo simulations
113 < of lamellar stacks of hexagonal lattices to show that large headgroups
114 < and molecular tilt with respect to the membrane normal vector can
115 < cause bulk rippling.~\cite{Bannerjee02}
87 > curvature-dependent Landau-de~Gennes free-energy functional to predict
88 > a rippled phase.~\cite{Marder84} This model and other related
89 > continuum models predict higher fluidity in convex regions and that
90 > concave portions of the membrane correspond to more solid-like
91 > regions.  Carlson and Sethna used a packing-competition model (in
92 > which head groups and chains have competing packing energetics) to
93 > predict the formation of a ripple-like phase.  Their model predicted
94 > that the high-curvature portions have lower-chain packing and
95 > correspond to more fluid-like regions.  Goldstein and Leibler used a
96 > mean-field approach with a planar model for {\em inter-lamellar}
97 > interactions to predict rippling in multilamellar
98 > phases.~\cite{Goldstein88} McCullough and Scott proposed that the {\em
99 > anisotropy of the nearest-neighbor interactions} coupled to
100 > hydrophobic constraining forces which restrict height differences
101 > between nearest neighbors is the origin of the ripple
102 > phase.~\cite{McCullough90} Lubensky and MacKintosh introduced a Landau
103 > theory for tilt order and curvature of a single membrane and concluded
104 > that {\em coupling of molecular tilt to membrane curvature} is
105 > responsible for the production of ripples.~\cite{Lubensky93} Misbah,
106 > Duplat and Houchmandzadeh proposed that {\em inter-layer dipolar
107 > interactions} can lead to ripple instabilities.~\cite{Misbah98}
108 > Heimburg presented a {\em coexistence model} for ripple formation in
109 > which he postulates that fluid-phase line defects cause sharp
110 > curvature between relatively flat gel-phase regions.~\cite{Heimburg00}
111 > Kubica has suggested that a lattice model of polar head groups could
112 > be valuable in trying to understand bilayer phase
113 > formation.~\cite{Kubica02} Bannerjee used Monte Carlo simulations of
114 > lamellar stacks of hexagonal lattices to show that large head groups
115 > and molecular tilt with respect to the membrane normal vector can
116 > cause bulk rippling.~\cite{Bannerjee02} Recently, Kranenburg and Smit
117 > described the formation of symmetric ripple-like structures using a
118 > coarse grained solvent-head-tail bead model.\cite{Kranenburg2005}
119 > Their lipids consisted of a short chain of head beads tied to the two
120 > longer ``chains''.
121  
122 < In contrast, few large-scale molecular modelling studies have been
122 > In contrast, few large-scale molecular modeling studies have been
123   done due to the large size of the resulting structures and the time
124   required for the phases of interest to develop.  With all-atom (and
125   even unified-atom) simulations, only one period of the ripple can be
126 < observed and only for timescales in the range of 10-100 ns.  One of
127 < the most interesting molecular simulations was carried out by De Vries
126 > observed and only for time scales in the range of 10-100 ns.  One of
127 > the most interesting molecular simulations was carried out by de~Vries
128   {\it et al.}~\cite{deVries05}. According to their simulation results,
129   the ripple consists of two domains, one resembling the gel bilayer,
130   while in the other, the two leaves of the bilayer are fully
# Line 122 | Line 146 | between headgroups and tails is strongly implicated as
146  
147   Although the organization of the tails of lipid molecules are
148   addressed by these molecular simulations and the packing competition
149 < between headgroups and tails is strongly implicated as the primary
149 > between head groups and tails is strongly implicated as the primary
150   driving force for ripple formation, questions about the ordering of
151 < the head groups in ripple phase has not been settled.
151 > the head groups in ripple phase have not been settled.
152  
153   In a recent paper, we presented a simple ``web of dipoles'' spin
154   lattice model which provides some physical insight into relationship
155   between dipolar ordering and membrane buckling.\cite{Sun2007} We found
156   that dipolar elastic membranes can spontaneously buckle, forming
157 < ripple-like topologies.  The driving force for the buckling in dipolar
158 < elastic membranes the antiferroelectric ordering of the dipoles, and
159 < this was evident in the ordering of the dipole director axis
160 < perpendicular to the wave vector of the surface ripples.  A similiar
157 > ripple-like topologies.  The driving force for the buckling of dipolar
158 > elastic membranes is the anti-ferroelectric ordering of the dipoles.
159 > This was evident in the ordering of the dipole director axis
160 > perpendicular to the wave vector of the surface ripples.  A similar
161   phenomenon has also been observed by Tsonchev {\it et al.} in their
162   work on the spontaneous formation of dipolar peptide chains into
163   curved nano-structures.\cite{Tsonchev04,Tsonchev04II}
# Line 167 | Line 191 | some fraction of the details of the chain dynamics neg
191   non-polar tails. Another fact is that the majority of lipid molecules
192   in the ripple phase are relatively rigid (i.e. gel-like) which makes
193   some fraction of the details of the chain dynamics negligible.  Figure
194 < \ref{fig:lipidModels} shows the molecular strucure of a DPPC
194 > \ref{fig:lipidModels} shows the molecular structure of a DPPC
195   molecule, as well as atomistic and molecular-scale representations of
196   a DPPC molecule.  The hydrophilic character of the head group is
197   largely due to the separation of charge between the nitrogen and
# Line 186 | Line 210 | modelling large length-scale properties of lipid
210   The ellipsoidal portions of the model interact via the Gay-Berne
211   potential which has seen widespread use in the liquid crystal
212   community.  Ayton and Voth have also used Gay-Berne ellipsoids for
213 < modelling large length-scale properties of lipid
213 > modeling large length-scale properties of lipid
214   bilayers.\cite{Ayton01} In its original form, the Gay-Berne potential
215   was a single site model for the interactions of rigid ellipsoidal
216   molecules.\cite{Gay81} It can be thought of as a modification of the
# Line 236 | Line 260 | ellipsoids in a {\it side-by-side} configuration.  Add
260  
261   Gay-Berne ellipsoids also have an energy scaling parameter,
262   $\epsilon^s$, which describes the well depth for two identical
263 < ellipsoids in a {\it side-by-side} configuration.  Additionaly, a well
263 > ellipsoids in a {\it side-by-side} configuration.  Additionally, a well
264   depth aspect ratio, $\epsilon^r = \epsilon^e / \epsilon^s$, describes
265   the ratio between the well depths in the {\it end-to-end} and
266   side-by-side configurations.  As in the range parameter, a set of
# Line 303 | Line 327 | zwitterionic head groups, we place fixed dipole moment
327   \end{figure}
328  
329   To take into account the permanent dipolar interactions of the
330 < zwitterionic head groups, we place fixed dipole moments $\mu_{i}$ at
330 > zwitterionic head groups, we have placed fixed dipole moments $\mu_{i}$ at
331   one end of the Gay-Berne particles.  The dipoles are oriented at an
332   angle $\theta = \pi / 2$ relative to the major axis.  These dipoles
333 < are protected by a head ``bead'' with a range parameter which we have
333 > are protected by a head ``bead'' with a range parameter ($\sigma_h$) which we have
334   varied between $1.20 d$ and $1.41 d$.  The head groups interact with
335   each other using a combination of Lennard-Jones,
336   \begin{equation}
# Line 325 | Line 349 | For the interaction between nonequivalent uniaxial ell
349   along dipole $i$ and $\mathbf{\hat r}_{ij}$ is the unit vector
350   pointing along the inter-dipole vector $\mathbf{r}_{ij}$.  
351  
352 + Since the charge separation distance is so large in zwitterionic head
353 + groups (like the PC head groups), it would also be possible to use
354 + either point charges or a ``split dipole'' approximation,
355 + \begin{equation}
356 + V_{ij}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf
357 + \hat{r}}_{ij})) = \frac{1}{{4\pi \epsilon_0 }}\left[ {\frac{{\mu _i  \cdot \mu _j }}{{R_{ij}^3 }} -
358 + \frac{{3\left( {\mu _i  \cdot r_{ij} } \right)\left( {\mu _i  \cdot
359 + r_{ij} } \right)}}{{R_{ij}^5 }}} \right]
360 + \end{equation}
361 + where $\mu _i$ and $\mu _j$ are the dipole moments of sites $i$ and
362 + $j$, $r_{ij}$ is vector between the two sites, and $R_{ij}$ is given
363 + by,
364 + \begin{equation}
365 + R_{ij}  = \sqrt {r_{ij}^2  + \frac{{d_i^2 }}{4} + \frac{{d_j^2
366 + }}{4}}.
367 + \end{equation}
368 + Here, $d_i$ and $d_j$ are effect charge separation distances
369 + associated with each of the two dipolar sites. This approximation to
370 + the multipole expansion maintains the fast fall-off of the multipole
371 + potentials but lacks the normal divergences when two polar groups get
372 + close to one another.
373 +
374   For the interaction between nonequivalent uniaxial ellipsoids (in this
375   case, between spheres and ellipsoids), the spheres are treated as
376   ellipsoids with an aspect ratio of 1 ($d = l$) and with an well depth
# Line 333 | Line 379 | The solvent model in our simulations is identical to o
379   et al.} and is appropriate for dissimilar uniaxial
380   ellipsoids.\cite{Cleaver96}
381  
382 < The solvent model in our simulations is identical to one used by
383 < Marrink {\it et al.}  in their dissipative particle dynamics (DPD)
384 < simulation of lipid bilayers.\cite{Marrink04} This solvent bead is a single
385 < site that represents four water molecules (m = 72 amu) and has
386 < comparable density and diffusive behavior to liquid water.  However,
387 < since there are no electrostatic sites on these beads, this solvent
388 < model cannot replicate the dielectric properties of water.
382 > The solvent model in our simulations is similar to the one used by
383 > Marrink {\it et al.}  in their coarse grained simulations of lipid
384 > bilayers.\cite{Marrink04} The solvent bead is a single site that
385 > represents four water molecules (m = 72 amu) and has comparable
386 > density and diffusive behavior to liquid water.  However, since there
387 > are no electrostatic sites on these beads, this solvent model cannot
388 > replicate the dielectric properties of water.  Note that although we
389 > are using larger cutoff and switching radii than Marrink {\it et al.},
390 > our solvent density at 300 K remains at 0.944 g cm$^{-3}$, and the
391 > solvent diffuses at 0.43 $\AA^2 ps^{-1}$ (roughly twice as fast as
392 > liquid water).
393 >
394   \begin{table*}
395   \begin{minipage}{\linewidth}
396   \begin{center}
# Line 365 | Line 416 | A switching function has been applied to all potential
416   \end{minipage}
417   \end{table*}
418  
419 < A switching function has been applied to all potentials to smoothly
420 < turn off the interactions between a range of $22$ and $25$ \AA.
419 > \section{Experimental Methodology}
420 > \label{sec:experiment}
421  
422   The parameters that were systematically varied in this study were the
423   size of the head group ($\sigma_h$), the strength of the dipole moment
424   ($\mu$), and the temperature of the system.  Values for $\sigma_h$
425 < ranged from 5.5 \AA\  to 6.5 \AA\ .  If the width of the tails is
426 < taken to be the unit of length, these head groups correspond to a
427 < range from $1.2 d$ to $1.41 d$.  Since the solvent beads are nearly
428 < identical in diameter to the tail ellipsoids, all distances that
429 < follow will be measured relative to this unit of distance.
425 > ranged from 5.5 \AA\ to 6.5 \AA\ .  If the width of the tails is taken
426 > to be the unit of length, these head groups correspond to a range from
427 > $1.2 d$ to $1.41 d$.  Since the solvent beads are nearly identical in
428 > diameter to the tail ellipsoids, all distances that follow will be
429 > measured relative to this unit of distance.  Because the solvent we
430 > are using is non-polar and has a dielectric constant of 1, values for
431 > $\mu$ are sampled from a range that is somewhat smaller than the 20.6
432 > Debye dipole moment of the PC head groups.
433  
380 \section{Experimental Methodology}
381 \label{sec:experiment}
382
434   To create unbiased bilayers, all simulations were started from two
435   perfectly flat monolayers separated by a 26 \AA\ gap between the
436   molecular bodies of the upper and lower leaves.  The separated
437 < monolayers were evolved in a vaccum with $x-y$ anisotropic pressure
437 > monolayers were evolved in a vacuum with $x-y$ anisotropic pressure
438   coupling. The length of $z$ axis of the simulations was fixed and a
439   constant surface tension was applied to enable real fluctuations of
440   the bilayer. Periodic boundary conditions were used, and $480-720$
441   lipid molecules were present in the simulations, depending on the size
442   of the head beads.  In all cases, the two monolayers spontaneously
443   collapsed into bilayer structures within 100 ps. Following this
444 < collapse, all systems were equlibrated for $100$ ns at $300$ K.
444 > collapse, all systems were equilibrated for $100$ ns at $300$ K.
445  
446   The resulting bilayer structures were then solvated at a ratio of $6$
447   solvent beads (24 water molecules) per lipid. These configurations
448   were then equilibrated for another $30$ ns. All simulations utilizing
449   the solvent were carried out at constant pressure ($P=1$ atm) with
450 < $3$D anisotropic coupling, and constant surface tension
451 < ($\gamma=0.015$ UNIT). Given the absence of fast degrees of freedom in
452 < this model, a timestep of $50$ fs was utilized with excellent energy
450 > $3$D anisotropic coupling, and small constant surface tension
451 > ($\gamma=0.015$ N/m). Given the absence of fast degrees of freedom in
452 > this model, a time step of $50$ fs was utilized with excellent energy
453   conservation.  Data collection for structural properties of the
454   bilayers was carried out during a final 5 ns run following the solvent
455 < equilibration.  All simulations were performed using the OOPSE
456 < molecular modeling program.\cite{Meineke05}
455 > equilibration.  Orientational correlation functions and diffusion
456 > constants were computed from 30 ns simulations in the microcanonical
457 > (NVE) ensemble using the average volume from the end of the constant
458 > pressure and surface tension runs.  The timestep on these final
459 > molecular dynamics runs was 25 fs.  No appreciable changes in phase
460 > structure were noticed upon switching to a microcanonical ensemble.
461 > All simulations were performed using the {\sc oopse} molecular
462 > modeling program.\cite{Meineke05}
463  
464 + A switching function was applied to all potentials to smoothly turn
465 + off the interactions between a range of $22$ and $25$ \AA.  The
466 + switching function was the standard (cubic) function,
467 + \begin{equation}
468 + s(r) =
469 +        \begin{cases}
470 +        1 & \text{if $r \le r_{\text{sw}}$},\\
471 +        \frac{(r_{\text{cut}} + 2r - 3r_{\text{sw}})(r_{\text{cut}} - r)^2}
472 +        {(r_{\text{cut}} - r_{\text{sw}})^3}
473 +        & \text{if $r_{\text{sw}} < r \le r_{\text{cut}}$}, \\
474 +        0 & \text{if $r > r_{\text{cut}}$.}
475 +        \end{cases}
476 + \label{eq:dipoleSwitching}
477 + \end{equation}
478 +
479   \section{Results}
480   \label{sec:results}
481  
482 < Snapshots in Figure \ref{fig:phaseCartoon} show that the membrane is
483 < more corrugated with increasing size of the head groups. The surface
484 < is nearly flat when $\sigma_h=1.20 d$. With $\sigma_h=1.28 d$,
485 < although the surface is still flat, the bilayer starts to splay
486 < inward; the upper leaf of the bilayer is connected to the lower leaf
487 < with an interdigitated line defect. Two periodicities with $100$ \AA\
488 < wavelengths were observed in the simulation. This structure is very
489 < similiar to the structure observed by de Vries and Lenz {\it et
490 < al.}. The same basic structure is also observed when $\sigma_h=1.41
491 < d$, but the wavelength of the surface corrugations depends sensitively
492 < on the size of the ``head'' beads. From the undulation spectrum, the
493 < corrugation is clearly non-thermal.
482 > The membranes in our simulations exhibit a number of interesting
483 > bilayer phases.  The surface topology of these phases depends most
484 > sensitively on the ratio of the size of the head groups to the width
485 > of the molecular bodies.  With heads only slightly larger than the
486 > bodies ($\sigma_h=1.20 d$) the membrane exhibits a flat bilayer.
487 >
488 > Increasing the head / body size ratio increases the local membrane
489 > curvature around each of the lipids.  With $\sigma_h=1.28 d$, the
490 > surface is still essentially flat, but the bilayer starts to exhibit
491 > signs of instability.  We have observed occasional defects where a
492 > line of lipid molecules on one leaf of the bilayer will dip down to
493 > interdigitate with the other leaf.  This gives each of the two bilayer
494 > leaves some local convexity near the line defect.  These structures,
495 > once developed in a simulation, are very stable and are spaced
496 > approximately 100 \AA\ away from each other.
497 >
498 > With larger heads ($\sigma_h = 1.35 d$) the membrane curvature
499 > resolves into a ``symmetric'' ripple phase.  Each leaf of the bilayer
500 > is broken into several convex, hemicylinderical sections, and opposite
501 > leaves are fitted together much like roof tiles.  There is no
502 > interdigitation between the upper and lower leaves of the bilayer.
503 >
504 > For the largest head / body ratios studied ($\sigma_h = 1.41 d$) the
505 > local curvature is substantially larger, and the resulting bilayer
506 > structure resolves into an asymmetric ripple phase.  This structure is
507 > very similar to the structures observed by both de~Vries {\it et al.}
508 > and Lenz {\it et al.}.  For a given ripple wave vector, there are two
509 > possible asymmetric ripples, which is not the case for the symmetric
510 > phase observed when $\sigma_h = 1.35 d$.
511 >
512   \begin{figure}[htb]
513   \centering
514   \includegraphics[width=4in]{phaseCartoon}
515 < \caption{A sketch to discribe the structure of the phases observed in
516 < our simulations.\label{fig:phaseCartoon}}
515 > \caption{The role of the ratio between the head group size and the
516 > width of the molecular bodies is to increase the local membrane
517 > curvature.  With strong attractive interactions between the head
518 > groups, this local curvature can be maintained in bilayer structures
519 > through surface corrugation.  Shown above are three phases observed in
520 > these simulations.  With $\sigma_h = 1.20 d$, the bilayer maintains a
521 > flat topology.  For larger heads ($\sigma_h = 1.35 d$) the local
522 > curvature resolves into a symmetrically rippled phase with little or
523 > no interdigitation between the upper and lower leaves of the membrane.
524 > The largest heads studied ($\sigma_h = 1.41 d$) resolve into an
525 > asymmetric rippled phases with interdigitation between the two
526 > leaves.\label{fig:phaseCartoon}}
527   \end{figure}
528  
529 < When $\sigma_h=1.35 d$, we observed another corrugated surface
530 < morphology.  This structure is different from the asymmetric rippled
531 < surface; there is no interdigitation between the upper and lower
532 < leaves of the bilayer. Each leaf of the bilayer is broken into several
533 < hemicylinderical sections, and opposite leaves are fitted together
534 < much like roof tiles. Unlike the surface in which the upper
535 < hemicylinder is always interdigitated on the leading or trailing edge
536 < of lower hemicylinder, this ``symmetric'' ripple has no prefered
537 < direction.  The corresponding structures are shown in Figure
538 < \ref{fig:phaseCartoon} for elucidation of the detailed structures of
539 < different phases.  The top panel in figure \ref{fig:phaseCartoon} is
540 < the flat phase, the middle panel shows the asymmetric ripple phase
541 < corresponding to $\sigma_h = 1.41 d$ and the lower panel shows the
542 < symmetric ripple phase observed when $\sigma_h=1.35 d$. In the
543 < symmetric ripple, the bilayer is continuous over the whole membrane,
544 < however, in asymmetric ripple phase, the bilayer domains are connected
545 < by thin interdigitated monolayers that share molecules between the
546 < upper and lower leaves.
529 > Sample structures for the flat ($\sigma_h = 1.20 d$), symmetric
530 > ($\sigma_h = 1.35 d$, and asymmetric ($\sigma_h = 1.41 d$) ripple
531 > phases are shown in Figure \ref{fig:phaseCartoon}.  
532 >
533 > It is reasonable to ask how well the parameters we used can produce
534 > bilayer properties that match experimentally known values for real
535 > lipid bilayers.  Using a value of $l = 13.8$ \AA for the ellipsoidal
536 > tails and the fixed ellipsoidal aspect ratio of 3, our values for the
537 > area per lipid ($A$) and inter-layer spacing ($D_{HH}$) depend
538 > entirely on the size of the head bead relative to the molecular body.
539 > These values are tabulated in table \ref{tab:property}.  Kucera {\it
540 > et al.}  have measured values for the head group spacings for a number
541 > of PC lipid bilayers that range from 30.8 \AA (DLPC) to 37.8 (DPPC).
542 > They have also measured values for the area per lipid that range from
543 > 60.6
544 > \AA$^2$ (DMPC) to 64.2 \AA$^2$
545 > (DPPC).\cite{NorbertKucerka04012005,NorbertKucerka06012006} Only the
546 > largest of the head groups we modeled ($\sigma_h = 1.41 d$) produces
547 > bilayers (specifically the area per lipid) that resemble real PC
548 > bilayers.  The smaller head beads we used are perhaps better models
549 > for PE head groups.
550 >
551   \begin{table*}
552   \begin{minipage}{\linewidth}
553   \begin{center}
554 < \caption{Phases, ripple wavelengths and amplitudes observed as a
555 < function of the ratio between the head beads and the diameters of the
556 < tails.  All lengths are normalized to the diameter of the tail
557 < ellipsoids.}
558 < \begin{tabular}{lccc}
554 > \caption{Phase, bilayer spacing, area per lipid, ripple wavelength
555 > and amplitude observed as a function of the ratio between the head
556 > beads and the diameters of the tails.  Ripple wavelengths and
557 > amplitudes are normalized to the diameter of the tail ellipsoids.}
558 > \begin{tabular}{lccccc}
559   \hline
560 < $\sigma_h / d$ & type of phase & $\lambda / d$ & $A / d$\\
560 > $\sigma_h / d$ & type of phase & bilayer spacing (\AA) & area per
561 > lipid (\AA$^2$) & $\lambda / d$ & $A / d$\\
562   \hline
563 < 1.20 & flat & N/A & N/A \\
564 < 1.28 & asymmetric ripple or flat & 21.7 & N/A \\
565 < 1.35 & symmetric ripple & 17.2 & 2.2 \\
566 < 1.41 & asymmetric ripple & 15.4 & 1.5 \\
563 > 1.20 & flat & 33.4 & 49.6 & N/A & N/A \\
564 > 1.28 & flat & 33.7 & 54.7 & N/A & N/A \\
565 > 1.35 & symmetric ripple & 42.9 & 51.7 & 17.2 & 2.2 \\
566 > 1.41 & asymmetric ripple & 37.1 & 63.1 & 15.4 & 1.5 \\
567   \end{tabular}
568   \label{tab:property}
569   \end{center}
# Line 467 | Line 572 | reduced amplitude $A / d$ of the ripples are summarize
572  
573   The membrane structures and the reduced wavelength $\lambda / d$,
574   reduced amplitude $A / d$ of the ripples are summarized in Table
575 < \ref{tab:property}. The wavelength range is $15~21$ molecular bodies
575 > \ref{tab:property}. The wavelength range is $15 - 17$ molecular bodies
576   and the amplitude is $1.5$ molecular bodies for asymmetric ripple and
577 < $2.2$ for symmetric ripple. These values are consistent to the
578 < experimental results.  Note, that given the lack of structural freedom
579 < in the tails of our model lipids, the amplitudes observed from these
580 < simulations are likely to underestimate of the true amplitudes.
577 > $2.2$ for symmetric ripple. These values are reasonably consistent
578 > with experimental measurements.\cite{Sun96,Katsaras00,Kaasgaard03}
579 > Note, that given the lack of structural freedom in the tails of our
580 > model lipids, the amplitudes observed from these simulations are
581 > likely to underestimate of the true amplitudes.
582  
583   \begin{figure}[htb]
584   \centering
585   \includegraphics[width=4in]{topDown}
586 < \caption{Top views of the flat (upper), asymmetric ripple (middle),
587 < and symmetric ripple (lower) phases.  Note that the head-group dipoles
588 < have formed head-to-tail chains in all three of these phases, but in
589 < the two rippled phases, the dipolar chains are all aligned
590 < {\it perpendicular} to the direction of the ripple.  The flat membrane
591 < has multiple point defects in the dipolar orientational ordering, and
592 < the dipolar ordering on the lower leaf of the bilayer can be in a
593 < different direction from the upper leaf.\label{fig:topView}}
586 > \caption{Top views of the flat (upper), symmetric ripple (middle),
587 > and asymmetric ripple (lower) phases.  Note that the head-group
588 > dipoles have formed head-to-tail chains in all three of these phases,
589 > but in the two rippled phases, the dipolar chains are all aligned {\it
590 > perpendicular} to the direction of the ripple.  Note that the flat
591 > membrane has multiple vortex defects in the dipolar ordering, and the
592 > ordering on the lower leaf of the bilayer can be in an entirely
593 > different direction from the upper leaf.\label{fig:topView}}
594   \end{figure}
595  
596   The principal method for observing orientational ordering in dipolar
# Line 516 | Line 622 | flat ($\sigma_h = 1.20 d$) and rippled ($\sigma_h = 1.
622   groups to be completely decoupled from each other.
623  
624   Figure \ref{fig:topView} shows snapshots of bird's-eye views of the
625 < flat ($\sigma_h = 1.20 d$) and rippled ($\sigma_h = 1.41, 1.35 d$)
625 > flat ($\sigma_h = 1.20 d$) and rippled ($\sigma_h = 1.35, 1.41 d$)
626   bilayers.  The directions of the dipoles on the head groups are
627   represented with two colored half spheres: blue (phosphate) and yellow
628   (amino).  For flat bilayers, the system exhibits signs of
# Line 539 | Line 645 | antiferroelectric state.  It is also notable that the
645   configuration, and the dipolar order parameter increases dramatically.
646   However, the total polarization of the system is still close to zero.
647   This is strong evidence that the corrugated structure is an
648 < antiferroelectric state.  It is also notable that the head-to-tail
648 > anti-ferroelectric state.  It is also notable that the head-to-tail
649   arrangement of the dipoles is always observed in a direction
650   perpendicular to the wave vector for the surface corrugation.  This is
651   a similar finding to what we observed in our earlier work on the
# Line 580 | Line 686 | interaction stength.  When the interaction between the
686   increasing strength of the dipole.  Generally, the dipoles on the head
687   groups become more ordered as the strength of the interaction between
688   heads is increased and become more disordered by decreasing the
689 < interaction stength.  When the interaction between the heads becomes
689 > interaction strength.  When the interaction between the heads becomes
690   too weak, the bilayer structure does not persist; all lipid molecules
691   become dispersed in the solvent (which is non-polar in this
692 < molecular-scale model).  The critial value of the strength of the
692 > molecular-scale model).  The critical value of the strength of the
693   dipole depends on the size of the head groups.  The perfectly flat
694   surface becomes unstable below $5$ Debye, while the  rippled
695   surfaces melt at $8$ Debye (asymmetric) or $10$ Debye (symmetric).
# Line 596 | Line 702 | dipolar interactions, the tails are forced to splay ou
702   close to each other and distort the bilayer structure. For a flat
703   surface, a substantial amount of free volume between the head groups
704   is normally available.  When the head groups are brought closer by
705 < dipolar interactions, the tails are forced to splay outward, forming
706 < first curved bilayers, and then inverted micelles.
705 > dipolar interactions, the tails are forced to splay outward, first forming
706 > curved bilayers, and then inverted micelles.
707  
708   When $\sigma_h=1.28 d$, the $P_2$ order parameter decreases slightly
709 < when the strength of the dipole is increased above $16$ debye. For
709 > when the strength of the dipole is increased above $16$ Debye. For
710   rippled bilayers, there is less free volume available between the head
711   groups. Therefore increasing dipolar strength weakly influences the
712   structure of the membrane.  However, the increase in the body $P_2$
# Line 657 | Line 763 | molecular width ratio ($\sigma_h / d$).\label{fig:tP2}
763   molecular width ratio ($\sigma_h / d$).\label{fig:tP2}}
764   \end{figure}
765  
766 < \section{Discussion}
767 < \label{sec:discussion}
766 > Fig. \ref{fig:phaseDiagram} shows a phase diagram for the model as a
767 > function of the head group / molecular width ratio ($\sigma_h / d$)
768 > and the strength of the head group dipole moment ($\mu$).  Note that
769 > the specific form of the bilayer phase is governed almost entirely by
770 > the head group / molecular width ratio, while the strength of the
771 > dipolar interactions between the head groups governs the stability of
772 > the bilayer phase.  Weaker dipoles result in unstable bilayer phases,
773 > while extremely strong dipoles can shift the equilibrium to an
774 > inverted micelle phase when the head groups are small.   Temperature
775 > has little effect on the actual bilayer phase observed, although higher
776 > temperatures can cause the unstable region to grow into the higher
777 > dipole region of this diagram.
778  
779 < The ripple phases have been observed in our molecular dynamic
780 < simulations using a simple molecular lipid model. The lipid model
781 < consists of an anisotropic interacting dipolar head group and an
782 < ellipsoid shape tail. According to our simulations, the explanation of
783 < the formation for the ripples are originated in the size mismatch
784 < between the head groups and the tails. The ripple phases are only
785 < observed in the studies using larger head group lipid models. However,
786 < 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.
779 > \begin{figure}[htb]
780 > \centering
781 > \includegraphics[width=\linewidth]{phaseDiagram}
782 > \caption{Phase diagram for the simple molecular model as a function
783 > of the head group / molecular width ratio ($\sigma_h / d$) and the
784 > strength of the head group dipole moment
785 > ($\mu$).\label{fig:phaseDiagram}}
786 > \end{figure}
787  
788 < Dipolar head groups are the key elements for the maintaining of the
789 < bilayer structure. The lipids are solvated in water when lowering the
790 < the strength of the dipole on the head groups. The long range
791 < orientational ordering of the dipoles can be achieved by forming the
792 < ripples, although the dipoles are likely to form head-to-tail
793 < configurations even in flat surface, the frustration prevents the
794 < formation of the long range orientational ordering for dipoles. The
795 < corrugation of the surface breaks the frustration and stablizes the
796 < long range oreintational ordering for the dipoles in the head groups
797 < of the lipid molecules. Many rows of the head-to-tail dipoles are
798 < parallel to each other and adopt the antiferroelectric state as a
799 < whole. This is the first time the organization of the head groups in
800 < ripple phases of the lipid bilayer has been addressed.
788 > We have computed translational diffusion constants for lipid molecules
789 > from the mean-square displacement,
790 > \begin{equation}
791 > D = \lim_{t\rightarrow \infty} \frac{1}{6 t} \langle {|\left({\bf r}_{i}(t) - {\bf r}_{i}(0) \right)|}^2 \rangle,
792 > \end{equation}
793 > of the lipid bodies. Translational diffusion constants for the
794 > different head-to-tail size ratios (all at 300 K) are shown in table
795 > \ref{tab:relaxation}.  We have also computed orientational correlation
796 > times for the head groups from fits of the second-order Legendre
797 > polynomial correlation function,
798 > \begin{equation}
799 > C_{\ell}(t)  =  \langle P_{\ell}\left({\bf \mu}_{i}(t) \cdot {\bf
800 > \mu}_{i}(0) \right)
801 > \end{equation}
802 > of the head group dipoles.  The orientational correlation functions
803 > appear to have multiple components in their decay: a fast ($12 \pm 2$
804 > ps) decay due to librational motion of the head groups, as well as
805 > moderate ($\tau^h_{\rm mid}$) and slow ($\tau^h_{\rm slow}$)
806 > components.  The fit values for the moderate and slow correlation
807 > times are listed in table \ref{tab:relaxation}.  Standard deviations
808 > in the fit time constants are quite large (on the order of the values
809 > themselves).
810  
811 < The most important prediction we can make using the results from this
812 < simple model is that if dipolar ordering is driving the surface
813 < corrugation, the wave vectors for the ripples should always found to
814 < be {\it perpendicular} to the dipole director axis.  This prediction
815 < should suggest experimental designs which test whether this is really
816 < true in the phosphatidylcholine $P_{\beta'}$ phases.  The dipole
817 < director axis should also be easily computable for the all-atom and
818 < coarse-grained simulations that have been published in the literature.
811 > Sparrman and Westlund used a multi-relaxation model for NMR lineshapes
812 > observed in gel, fluid, and ripple phases of DPPC and obtained
813 > estimates of a correlation time for water translational diffusion
814 > ($\tau_c$) of 20 ns.\cite{Sparrman2003} This correlation time
815 > corresponds to water bound to small regions of the lipid membrane.
816 > They further assume that the lipids can explore only a single period
817 > of the ripple (essentially moving in a nearly one-dimensional path to
818 > do so), and the correlation time can therefore be used to estimate a
819 > value for the translational diffusion constant of $2.25 \times
820 > 10^{-11} m^2 s^{-1}$.  The translational diffusion constants we obtain
821 > are in reasonable agreement with this experimentally determined
822 > value. However, the $T_2$ relaxation times obtained by Sparrman and
823 > Westlund are consistent with P-N vector reorientation timescales of
824 > 2-5 ms.  This is substantially slower than even the slowest component
825 > we observe in the decay of our orientational correlation functions.
826 > Other than the dipole-dipole interactions, our head groups have no
827 > shape anisotropy which would force them to move as a unit with
828 > neighboring molecules.  This would naturally lead to P-N reorientation
829 > times that are too fast when compared with experimental measurements.
830  
831 + \begin{table*}
832 + \begin{minipage}{\linewidth}
833 + \begin{center}
834 + \caption{Fit values for the rotational correlation times for the head
835 + groups ($\tau^h$) and molecular bodies ($\tau^b$) as well as the
836 + translational diffusion constants for the molecule as a function of
837 + the head-to-body width ratio (all at 300 K).  In all of the phases,
838 + the head group correlation functions decay with an fast librational
839 + contribution ($12 \pm 1$ ps).  There are additional moderate
840 + ($\tau^h_{\rm mid}$) and slow $\tau^h_{\rm slow}$ contributions to
841 + orientational decay that depend strongly on the phase exhibited by the
842 + lipids.  The symmetric ripple phase ($\sigma_h / d = 1.35$) appears to
843 + exhibit the slowest molecular reorientation.}
844 + \begin{tabular}{lcccc}
845 + \hline
846 + $\sigma_h / d$ & $\tau^h_{\rm mid} (ns)$ & $\tau^h_{\rm
847 + slow} (\mu s)$ & $\tau_b (\mu s)$ & $D (\times 10^{-11} m^2 s^{-1})$ \\
848 + \hline
849 + 1.20 & $0.4$ &  $9.6$ & $9.5$ & $0.43(1)$ \\
850 + 1.28 & $2.0$ & $13.5$ & $3.0$ & $5.91(3)$ \\
851 + 1.35 & $3.2$ &  $4.0$ & $0.9$ & $3.42(1)$ \\
852 + 1.41 & $0.3$ & $23.8$ & $6.9$ & $7.16(1)$ \\
853 + \end{tabular}
854 + \label{tab:relaxation}
855 + \end{center}
856 + \end{minipage}
857 + \end{table*}
858 +
859 + \section{Discussion}
860 + \label{sec:discussion}
861 +
862 + Symmetric and asymmetric ripple phases have been observed to form in
863 + our molecular dynamics simulations of a simple molecular-scale lipid
864 + model. The lipid model consists of an dipolar head group and an
865 + ellipsoidal tail.  Within the limits of this model, an explanation for
866 + generalized membrane curvature is a simple mismatch in the size of the
867 + heads with the width of the molecular bodies.  With heads
868 + substantially larger than the bodies of the molecule, this curvature
869 + should be convex nearly everywhere, a requirement which could be
870 + resolved either with micellar or cylindrical phases.
871 +
872 + The persistence of a {\it bilayer} structure therefore requires either
873 + strong attractive forces between the head groups or exclusionary
874 + forces from the solvent phase.  To have a persistent bilayer structure
875 + with the added requirement of convex membrane curvature appears to
876 + result in corrugated structures like the ones pictured in
877 + Fig. \ref{fig:phaseCartoon}.  In each of the sections of these
878 + corrugated phases, the local curvature near a most of the head groups
879 + is convex.  These structures are held together by the extremely strong
880 + and directional interactions between the head groups.
881 +
882 + The attractive forces holding the bilayer together could either be
883 + non-directional (as in the work of Kranenburg and
884 + Smit),\cite{Kranenburg2005} or directional (as we have utilized in
885 + these simulations).  The dipolar head groups are key for the
886 + maintaining the bilayer structures exhibited by this particular model;
887 + reducing the strength of the dipole has the tendency to make the
888 + rippled phase disappear.  The dipoles are likely to form attractive
889 + head-to-tail configurations even in flat configurations, but the
890 + temperatures are high enough that vortex defects become prevalent in
891 + the flat phase.  The flat phase we observed therefore appears to be
892 + substantially above the Kosterlitz-Thouless transition temperature for
893 + a planar system of dipoles with this set of parameters.  For this
894 + reason, it would be interesting to observe the thermal behavior of the
895 + flat phase at substantially lower temperatures.
896 +
897 + One feature of this model is that an energetically favorable
898 + orientational ordering of the dipoles can be achieved by forming
899 + ripples.  The corrugation of the surface breaks the symmetry of the
900 + plane, making vortex defects somewhat more expensive, and stabilizing
901 + the long range orientational ordering for the dipoles in the head
902 + groups.  Most of the rows of the head-to-tail dipoles are parallel to
903 + each other and the system adopts a bulk anti-ferroelectric state.  We
904 + believe that this is the first time the organization of the head
905 + groups in ripple phases has been addressed.
906 +
907 + Although the size-mismatch between the heads and molecular bodies
908 + appears to be the primary driving force for surface convexity, the
909 + persistence of the bilayer through the use of rippled structures is a
910 + function of the strong, attractive interactions between the heads.
911 + One important prediction we can make using the results from this
912 + simple model is that if the dipole-dipole interaction is the leading
913 + contributor to the head group attractions, the wave vectors for the
914 + ripples should always be found {\it perpendicular} to the dipole
915 + director axis.  This echoes the prediction we made earlier for simple
916 + elastic dipolar membranes, and may suggest experimental designs which
917 + will test whether this is really the case in the phosphatidylcholine
918 + $P_{\beta'}$ phases.  The dipole director axis should also be easily
919 + computable for the all-atom and coarse-grained simulations that have
920 + been published in the literature.\cite{deVries05}
921 +
922   Although our model is simple, it exhibits some rich and unexpected
923 < behaviors.  It would clearly be a closer approximation to the reality
924 < if we allowed greater translational freedom to the dipoles and
925 < replaced the somewhat artificial lattice packing and the harmonic
926 < elastic tension with more realistic molecular modeling potentials.
927 < What we have done is to present a simple model which exhibits bulk
928 < non-thermal corrugation, and our explanation of this rippling
923 > behaviors.  It would clearly be a closer approximation to reality if
924 > we allowed bending motions between the dipoles and the molecular
925 > bodies, and if we replaced the rigid ellipsoids with ball-and-chain
926 > tails.  However, the advantages of this simple model (large system
927 > sizes, 50 fs time steps) allow us to rapidly explore the phase diagram
928 > for a wide range of parameters.  Our explanation of this rippling
929   phenomenon will help us design more accurate molecular models for
930 < corrugated membranes and experiments to test whether rippling is
931 < dipole-driven or not.
720 <
930 > corrugated membranes and experiments to test whether or not
931 > dipole-dipole interactions exert an influence on membrane rippling.
932   \newpage
933   \bibliography{mdripple}
934   \end{document}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines