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 3266 by xsun, Fri Oct 19 21:33:05 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} The X-ray diffraction work by
79 > Katsaras {\it et al.} showed that a rich phase diagram exhibiting both
80 > {\it asymmetric} and {\it symmetric} ripples is possible for lecithin
81 > bilayers.\cite{Katsaras00}
82  
83   A number of theoretical models have been presented to explain the
84   formation of the ripple phase. Marder {\it et al.} used a
85 < curvature-dependent Landau-de Gennes free-energy functional to predict
85 > curvature-dependent Landau-de~Gennes free-energy functional to predict
86   a rippled phase.~\cite{Marder84} This model and other related continuum
87   models predict higher fluidity in convex regions and that concave
88   portions of the membrane correspond to more solid-like regions.
# Line 91 | Line 108 | of lamellar stacks of hexagonal lattices to show that
108   regions.~\cite{Heimburg00} Kubica has suggested that a lattice model of
109   polar head groups could be valuable in trying to understand bilayer
110   phase formation.~\cite{Kubica02} Bannerjee used Monte Carlo simulations
111 < of lamellar stacks of hexagonal lattices to show that large headgroups
111 > of lamellar stacks of hexagonal lattices to show that large head groups
112   and molecular tilt with respect to the membrane normal vector can
113   cause bulk rippling.~\cite{Bannerjee02}
114  
115 < In contrast, few large-scale molecular modelling studies have been
115 > In contrast, few large-scale molecular modeling studies have been
116   done due to the large size of the resulting structures and the time
117   required for the phases of interest to develop.  With all-atom (and
118   even unified-atom) simulations, only one period of the ripple can be
119 < observed and only for timescales in the range of 10-100 ns.  One of
120 < the most interesting molecular simulations was carried out by De Vries
119 > observed and only for time scales in the range of 10-100 ns.  One of
120 > the most interesting molecular simulations was carried out by de~Vries
121   {\it et al.}~\cite{deVries05}. According to their simulation results,
122   the ripple consists of two domains, one resembling the gel bilayer,
123   while in the other, the two leaves of the bilayer are fully
# Line 122 | Line 139 | between headgroups and tails is strongly implicated as
139  
140   Although the organization of the tails of lipid molecules are
141   addressed by these molecular simulations and the packing competition
142 < between headgroups and tails is strongly implicated as the primary
142 > between head groups and tails is strongly implicated as the primary
143   driving force for ripple formation, questions about the ordering of
144 < the head groups in ripple phase has not been settled.
144 > the head groups in ripple phase have not been settled.
145  
146   In a recent paper, we presented a simple ``web of dipoles'' spin
147   lattice model which provides some physical insight into relationship
148   between dipolar ordering and membrane buckling.\cite{Sun2007} We found
149   that dipolar elastic membranes can spontaneously buckle, forming
150 < ripple-like topologies.  The driving force for the buckling in dipolar
151 < elastic membranes the antiferroelectric ordering of the dipoles, and
152 < this was evident in the ordering of the dipole director axis
153 < perpendicular to the wave vector of the surface ripples.  A similiar
150 > ripple-like topologies.  The driving force for the buckling of dipolar
151 > elastic membranes is the anti-ferroelectric ordering of the dipoles.
152 > This was evident in the ordering of the dipole director axis
153 > perpendicular to the wave vector of the surface ripples.  A similar
154   phenomenon has also been observed by Tsonchev {\it et al.} in their
155   work on the spontaneous formation of dipolar peptide chains into
156   curved nano-structures.\cite{Tsonchev04,Tsonchev04II}
# Line 167 | Line 184 | some fraction of the details of the chain dynamics neg
184   non-polar tails. Another fact is that the majority of lipid molecules
185   in the ripple phase are relatively rigid (i.e. gel-like) which makes
186   some fraction of the details of the chain dynamics negligible.  Figure
187 < \ref{fig:lipidModels} shows the molecular strucure of a DPPC
187 > \ref{fig:lipidModels} shows the molecular structure of a DPPC
188   molecule, as well as atomistic and molecular-scale representations of
189   a DPPC molecule.  The hydrophilic character of the head group is
190   largely due to the separation of charge between the nitrogen and
# Line 186 | Line 203 | modelling large length-scale properties of lipid
203   The ellipsoidal portions of the model interact via the Gay-Berne
204   potential which has seen widespread use in the liquid crystal
205   community.  Ayton and Voth have also used Gay-Berne ellipsoids for
206 < modelling large length-scale properties of lipid
206 > modeling large length-scale properties of lipid
207   bilayers.\cite{Ayton01} In its original form, the Gay-Berne potential
208   was a single site model for the interactions of rigid ellipsoidal
209   molecules.\cite{Gay81} It can be thought of as a modification of the
# Line 236 | Line 253 | ellipsoids in a {\it side-by-side} configuration.  Add
253  
254   Gay-Berne ellipsoids also have an energy scaling parameter,
255   $\epsilon^s$, which describes the well depth for two identical
256 < ellipsoids in a {\it side-by-side} configuration.  Additionaly, a well
256 > ellipsoids in a {\it side-by-side} configuration.  Additionally, a well
257   depth aspect ratio, $\epsilon^r = \epsilon^e / \epsilon^s$, describes
258   the ratio between the well depths in the {\it end-to-end} and
259   side-by-side configurations.  As in the range parameter, a set of
# Line 303 | Line 320 | zwitterionic head groups, we place fixed dipole moment
320   \end{figure}
321  
322   To take into account the permanent dipolar interactions of the
323 < zwitterionic head groups, we place fixed dipole moments $\mu_{i}$ at
323 > zwitterionic head groups, we have placed fixed dipole moments $\mu_{i}$ at
324   one end of the Gay-Berne particles.  The dipoles are oriented at an
325   angle $\theta = \pi / 2$ relative to the major axis.  These dipoles
326 < are protected by a head ``bead'' with a range parameter which we have
326 > are protected by a head ``bead'' with a range parameter ($\sigma_h$) which we have
327   varied between $1.20 d$ and $1.41 d$.  The head groups interact with
328   each other using a combination of Lennard-Jones,
329   \begin{equation}
# Line 335 | Line 352 | simulation of lipid bilayers.\cite{Marrink04} This sol
352  
353   The solvent model in our simulations is identical to one used by
354   Marrink {\it et al.}  in their dissipative particle dynamics (DPD)
355 < simulation of lipid bilayers.\cite{Marrink04} This solvent bead is a single
356 < site that represents four water molecules (m = 72 amu) and has
355 > simulation of lipid bilayers.\cite{Marrink04} This solvent bead is a
356 > single site that represents four water molecules (m = 72 amu) and has
357   comparable density and diffusive behavior to liquid water.  However,
358   since there are no electrostatic sites on these beads, this solvent
359 < model cannot replicate the dielectric properties of water.
359 > model cannot replicate the dielectric properties of water.
360 >
361   \begin{table*}
362   \begin{minipage}{\linewidth}
363   \begin{center}
# Line 365 | Line 383 | A switching function has been applied to all potential
383   \end{minipage}
384   \end{table*}
385  
386 < A switching function has been applied to all potentials to smoothly
387 < turn off the interactions between a range of $22$ and $25$ \AA.
386 > \section{Experimental Methodology}
387 > \label{sec:experiment}
388  
389   The parameters that were systematically varied in this study were the
390   size of the head group ($\sigma_h$), the strength of the dipole moment
391   ($\mu$), and the temperature of the system.  Values for $\sigma_h$
392 < ranged from 5.5 \AA\  to 6.5 \AA\ .  If the width of the tails is
393 < taken to be the unit of length, these head groups correspond to a
394 < range from $1.2 d$ to $1.41 d$.  Since the solvent beads are nearly
395 < identical in diameter to the tail ellipsoids, all distances that
396 < follow will be measured relative to this unit of distance.
392 > ranged from 5.5 \AA\ to 6.5 \AA\ .  If the width of the tails is taken
393 > to be the unit of length, these head groups correspond to a range from
394 > $1.2 d$ to $1.41 d$.  Since the solvent beads are nearly identical in
395 > diameter to the tail ellipsoids, all distances that follow will be
396 > measured relative to this unit of distance.  Because the solvent we
397 > are using is non-polar and has a dielectric constant of 1, values for
398 > $\mu$ are sampled from a range that is somewhat smaller than the 20.6
399 > Debye dipole moment of the PC head groups.
400  
380 \section{Experimental Methodology}
381 \label{sec:experiment}
382
401   To create unbiased bilayers, all simulations were started from two
402   perfectly flat monolayers separated by a 26 \AA\ gap between the
403   molecular bodies of the upper and lower leaves.  The separated
404 < monolayers were evolved in a vaccum with $x-y$ anisotropic pressure
404 > monolayers were evolved in a vacuum with $x-y$ anisotropic pressure
405   coupling. The length of $z$ axis of the simulations was fixed and a
406   constant surface tension was applied to enable real fluctuations of
407   the bilayer. Periodic boundary conditions were used, and $480-720$
408   lipid molecules were present in the simulations, depending on the size
409   of the head beads.  In all cases, the two monolayers spontaneously
410   collapsed into bilayer structures within 100 ps. Following this
411 < collapse, all systems were equlibrated for $100$ ns at $300$ K.
411 > collapse, all systems were equilibrated for $100$ ns at $300$ K.
412  
413   The resulting bilayer structures were then solvated at a ratio of $6$
414   solvent beads (24 water molecules) per lipid. These configurations
415   were then equilibrated for another $30$ ns. All simulations utilizing
416   the solvent were carried out at constant pressure ($P=1$ atm) with
417   $3$D anisotropic coupling, and constant surface tension
418 < ($\gamma=0.015$ UNIT). Given the absence of fast degrees of freedom in
419 < this model, a timestep of $50$ fs was utilized with excellent energy
418 > ($\gamma=0.015$ N/m). Given the absence of fast degrees of freedom in
419 > this model, a time step of $50$ fs was utilized with excellent energy
420   conservation.  Data collection for structural properties of the
421   bilayers was carried out during a final 5 ns run following the solvent
422   equilibration.  All simulations were performed using the OOPSE
423   molecular modeling program.\cite{Meineke05}
424  
425 + A switching function was applied to all potentials to smoothly turn
426 + off the interactions between a range of $22$ and $25$ \AA.
427 +
428   \section{Results}
429   \label{sec:results}
430  
431 < Snapshots in Figure \ref{fig:phaseCartoon} show that the membrane is
432 < more corrugated with increasing size of the head groups. The surface
433 < is nearly flat when $\sigma_h=1.20 d$. With $\sigma_h=1.28 d$,
434 < although the surface is still flat, the bilayer starts to splay
435 < inward; the upper leaf of the bilayer is connected to the lower leaf
436 < with an interdigitated line defect. Two periodicities with $100$ \AA\
437 < wavelengths were observed in the simulation. This structure is very
438 < similiar to the structure observed by de Vries and Lenz {\it et
439 < al.}. The same basic structure is also observed when $\sigma_h=1.41
440 < d$, but the wavelength of the surface corrugations depends sensitively
441 < on the size of the ``head'' beads. From the undulation spectrum, the
442 < corrugation is clearly non-thermal.
431 > The membranes in our simulations exhibit a number of interesting
432 > bilayer phases.  The surface topology of these phases depends most
433 > sensitively on the ratio of the size of the head groups to the width
434 > of the molecular bodies.  With heads only slightly larger than the
435 > bodies ($\sigma_h=1.20 d$) the membrane exhibits a flat bilayer.
436 >
437 > Increasing the head / body size ratio increases the local membrane
438 > curvature around each of the lipids.  With $\sigma_h=1.28 d$, the
439 > surface is still essentially flat, but the bilayer starts to exhibit
440 > signs of instability.  We have observed occasional defects where a
441 > line of lipid molecules on one leaf of the bilayer will dip down to
442 > interdigitate with the other leaf.  This gives each of the two bilayer
443 > leaves some local convexity near the line defect.  These structures,
444 > once developed in a simulation, are very stable and are spaced
445 > approximately 100 \AA\ away from each other.
446 >
447 > With larger heads ($\sigma_h = 1.35 d$) the membrane curvature
448 > resolves into a ``symmetric'' ripple phase.  Each leaf of the bilayer
449 > is broken into several convex, hemicylinderical sections, and opposite
450 > leaves are fitted together much like roof tiles.  There is no
451 > interdigitation between the upper and lower leaves of the bilayer.
452 >
453 > For the largest head / body ratios studied ($\sigma_h = 1.41 d$) the
454 > local curvature is substantially larger, and the resulting bilayer
455 > structure resolves into an asymmetric ripple phase.  This structure is
456 > very similar to the structures observed by both de~Vries {\it et al.}
457 > and Lenz {\it et al.}.  For a given ripple wave vector, there are two
458 > possible asymmetric ripples, which is not the case for the symmetric
459 > phase observed when $\sigma_h = 1.35 d$.
460 >
461   \begin{figure}[htb]
462   \centering
463   \includegraphics[width=4in]{phaseCartoon}
464 < \caption{A sketch to discribe the structure of the phases observed in
465 < our simulations.\label{fig:phaseCartoon}}
464 > \caption{The role of the ratio between the head group size and the
465 > width of the molecular bodies is to increase the local membrane
466 > curvature.  With strong attractive interactions between the head
467 > groups, this local curvature can be maintained in bilayer structures
468 > through surface corrugation.  Shown above are three phases observed in
469 > these simulations.  With $\sigma_h = 1.20 d$, the bilayer maintains a
470 > flat topology.  For larger heads ($\sigma_h = 1.35 d$) the local
471 > curvature resolves into a symmetrically rippled phase with little or
472 > no interdigitation between the upper and lower leaves of the membrane.
473 > The largest heads studied ($\sigma_h = 1.41 d$) resolve into an
474 > asymmetric rippled phases with interdigitation between the two
475 > leaves.\label{fig:phaseCartoon}}
476   \end{figure}
477  
478 < When $\sigma_h=1.35 d$, we observed another corrugated surface
479 < morphology.  This structure is different from the asymmetric rippled
480 < surface; there is no interdigitation between the upper and lower
481 < leaves of the bilayer. Each leaf of the bilayer is broken into several
482 < hemicylinderical sections, and opposite leaves are fitted together
483 < much like roof tiles. Unlike the surface in which the upper
484 < hemicylinder is always interdigitated on the leading or trailing edge
485 < of lower hemicylinder, this ``symmetric'' ripple has no prefered
486 < direction.  The corresponding structures are shown in Figure
487 < \ref{fig:phaseCartoon} for elucidation of the detailed structures of
488 < different phases.  The top panel in figure \ref{fig:phaseCartoon} is
489 < the flat phase, the middle panel shows the asymmetric ripple phase
490 < corresponding to $\sigma_h = 1.41 d$ and the lower panel shows the
491 < symmetric ripple phase observed when $\sigma_h=1.35 d$. In the
492 < symmetric ripple, the bilayer is continuous over the whole membrane,
493 < however, in asymmetric ripple phase, the bilayer domains are connected
494 < by thin interdigitated monolayers that share molecules between the
495 < upper and lower leaves.
478 > Sample structures for the flat ($\sigma_h = 1.20 d$), symmetric
479 > ($\sigma_h = 1.35 d$, and asymmetric ($\sigma_h = 1.41 d$) ripple
480 > phases are shown in Figure \ref{fig:phaseCartoon}.  
481 >
482 > It is reasonable to ask how well the parameters we used can produce
483 > bilayer properties that match experimentally known values for real
484 > lipid bilayers.  Using a value of $l = 13.8$ \AA for the ellipsoidal
485 > tails and the fixed ellipsoidal aspect ratio of 3, our values for the
486 > area per lipid ($A$) and inter-layer spacing ($D_{HH}$) depend
487 > entirely on the size of the head bead relative to the molecular body.
488 > These values are tabulated in table \ref{tab:property}.  Kucera {\it
489 > et al.}  have measured values for the head group spacings for a number
490 > of PC lipid bilayers that range from 30.8 \AA (DLPC) to 37.8 (DPPC).
491 > They have also measured values for the area per lipid that range from
492 > 60.6
493 > \AA$^2$ (DMPC) to 64.2 \AA$^2$
494 > (DPPC).\cite{NorbertKucerka04012005,NorbertKucerka06012006} Only the
495 > largest of the head groups we modeled ($\sigma_h = 1.41 d$) produces
496 > bilayers (specifically the area per lipid) that resemble real PC
497 > bilayers.  The smaller head beads we used are perhaps better models
498 > for PE head groups.
499 >
500   \begin{table*}
501   \begin{minipage}{\linewidth}
502   \begin{center}
503 < \caption{Phases, ripple wavelengths and amplitudes observed as a
504 < function of the ratio between the head beads and the diameters of the
505 < tails.  All lengths are normalized to the diameter of the tail
506 < ellipsoids.}
507 < \begin{tabular}{lccc}
503 > \caption{Phase, bilayer spacing, area per lipid, ripple wavelength
504 > and amplitude observed as a function of the ratio between the head
505 > beads and the diameters of the tails.  Ripple wavelengths and
506 > amplitudes are normalized to the diameter of the tail ellipsoids.}
507 > \begin{tabular}{lccccc}
508   \hline
509 < $\sigma_h / d$ & type of phase & $\lambda / d$ & $A / d$\\
509 > $\sigma_h / d$ & type of phase & bilayer spacing (\AA) & area per
510 > lipid (\AA$^2$) & $\lambda / d$ & $A / d$\\
511   \hline
512 < 1.20 & flat & N/A & N/A \\
513 < 1.28 & asymmetric ripple or flat & 21.7 & N/A \\
514 < 1.35 & symmetric ripple & 17.2 & 2.2 \\
515 < 1.41 & asymmetric ripple & 15.4 & 1.5 \\
512 > 1.20 & flat & 33.4 & 49.6 & N/A & N/A \\
513 > 1.28 & flat & 33.7 & 54.7 & N/A & N/A \\
514 > 1.35 & symmetric ripple & 42.9 & 51.7 & 17.2 & 2.2 \\
515 > 1.41 & asymmetric ripple & 37.1 & 63.1 & 15.4 & 1.5 \\
516   \end{tabular}
517   \label{tab:property}
518   \end{center}
# Line 467 | Line 521 | reduced amplitude $A / d$ of the ripples are summarize
521  
522   The membrane structures and the reduced wavelength $\lambda / d$,
523   reduced amplitude $A / d$ of the ripples are summarized in Table
524 < \ref{tab:property}. The wavelength range is $15~21$ molecular bodies
524 > \ref{tab:property}. The wavelength range is $15 - 17$ molecular bodies
525   and the amplitude is $1.5$ molecular bodies for asymmetric ripple and
526 < $2.2$ for symmetric ripple. These values are consistent to the
527 < experimental results.  Note, that given the lack of structural freedom
528 < in the tails of our model lipids, the amplitudes observed from these
529 < simulations are likely to underestimate of the true amplitudes.
526 > $2.2$ for symmetric ripple. These values are reasonably consistent
527 > with experimental measurements.\cite{Sun96,Katsaras00,Kaasgaard03}
528 > Note, that given the lack of structural freedom in the tails of our
529 > model lipids, the amplitudes observed from these simulations are
530 > likely to underestimate of the true amplitudes.
531  
532   \begin{figure}[htb]
533   \centering
534   \includegraphics[width=4in]{topDown}
535 < \caption{Top views of the flat (upper), asymmetric ripple (middle),
536 < and symmetric ripple (lower) phases.  Note that the head-group dipoles
537 < have formed head-to-tail chains in all three of these phases, but in
538 < the two rippled phases, the dipolar chains are all aligned
539 < {\it perpendicular} to the direction of the ripple.  The flat membrane
540 < has multiple point defects in the dipolar orientational ordering, and
541 < the dipolar ordering on the lower leaf of the bilayer can be in a
542 < different direction from the upper leaf.\label{fig:topView}}
535 > \caption{Top views of the flat (upper), symmetric ripple (middle),
536 > and asymmetric ripple (lower) phases.  Note that the head-group
537 > dipoles have formed head-to-tail chains in all three of these phases,
538 > but in the two rippled phases, the dipolar chains are all aligned {\it
539 > perpendicular} to the direction of the ripple.  Note that the flat
540 > membrane has multiple vortex defects in the dipolar ordering, and the
541 > ordering on the lower leaf of the bilayer can be in an entirely
542 > different direction from the upper leaf.\label{fig:topView}}
543   \end{figure}
544  
545   The principal method for observing orientational ordering in dipolar
# Line 516 | Line 571 | flat ($\sigma_h = 1.20 d$) and rippled ($\sigma_h = 1.
571   groups to be completely decoupled from each other.
572  
573   Figure \ref{fig:topView} shows snapshots of bird's-eye views of the
574 < flat ($\sigma_h = 1.20 d$) and rippled ($\sigma_h = 1.41, 1.35 d$)
574 > flat ($\sigma_h = 1.20 d$) and rippled ($\sigma_h = 1.35, 1.41 d$)
575   bilayers.  The directions of the dipoles on the head groups are
576   represented with two colored half spheres: blue (phosphate) and yellow
577   (amino).  For flat bilayers, the system exhibits signs of
# Line 539 | Line 594 | antiferroelectric state.  It is also notable that the
594   configuration, and the dipolar order parameter increases dramatically.
595   However, the total polarization of the system is still close to zero.
596   This is strong evidence that the corrugated structure is an
597 < antiferroelectric state.  It is also notable that the head-to-tail
597 > anti-ferroelectric state.  It is also notable that the head-to-tail
598   arrangement of the dipoles is always observed in a direction
599   perpendicular to the wave vector for the surface corrugation.  This is
600   a similar finding to what we observed in our earlier work on the
# Line 580 | Line 635 | interaction stength.  When the interaction between the
635   increasing strength of the dipole.  Generally, the dipoles on the head
636   groups become more ordered as the strength of the interaction between
637   heads is increased and become more disordered by decreasing the
638 < interaction stength.  When the interaction between the heads becomes
638 > interaction strength.  When the interaction between the heads becomes
639   too weak, the bilayer structure does not persist; all lipid molecules
640   become dispersed in the solvent (which is non-polar in this
641 < molecular-scale model).  The critial value of the strength of the
641 > molecular-scale model).  The critical value of the strength of the
642   dipole depends on the size of the head groups.  The perfectly flat
643   surface becomes unstable below $5$ Debye, while the  rippled
644   surfaces melt at $8$ Debye (asymmetric) or $10$ Debye (symmetric).
# Line 596 | Line 651 | dipolar interactions, the tails are forced to splay ou
651   close to each other and distort the bilayer structure. For a flat
652   surface, a substantial amount of free volume between the head groups
653   is normally available.  When the head groups are brought closer by
654 < dipolar interactions, the tails are forced to splay outward, forming
655 < first curved bilayers, and then inverted micelles.
654 > dipolar interactions, the tails are forced to splay outward, first forming
655 > curved bilayers, and then inverted micelles.
656  
657   When $\sigma_h=1.28 d$, the $P_2$ order parameter decreases slightly
658 < when the strength of the dipole is increased above $16$ debye. For
658 > when the strength of the dipole is increased above $16$ Debye. For
659   rippled bilayers, there is less free volume available between the head
660   groups. Therefore increasing dipolar strength weakly influences the
661   structure of the membrane.  However, the increase in the body $P_2$
# Line 657 | Line 712 | molecular width ratio ($\sigma_h / d$).\label{fig:tP2}
712   molecular width ratio ($\sigma_h / d$).\label{fig:tP2}}
713   \end{figure}
714  
715 < \section{Discussion}
716 < \label{sec:discussion}
715 > Fig. \ref{fig:phaseDiagram} shows a phase diagram for the model as a
716 > function of the head group / molecular width ratio ($\sigma_h / d$)
717 > and the strength of the head group dipole moment ($\mu$).  Note that
718 > the specific form of the bilayer phase is governed almost entirely by
719 > the head group / molecular width ratio, while the strength of the
720 > dipolar interactions between the head groups governs the stability of
721 > the bilayer phase.  Weaker dipoles result in unstable bilayer phases,
722 > while extremely strong dipoles can shift the equilibrium to an
723 > inverted micelle phase when the head groups are small.   Temperature
724 > has little effect on the actual bilayer phase observed, although higher
725 > temperatures can cause the unstable region to grow into the higher
726 > dipole region of this diagram.
727  
728 < The ripple phases have been observed in our molecular dynamic
729 < simulations using a simple molecular lipid model. The lipid model
730 < consists of an anisotropic interacting dipolar head group and an
731 < ellipsoid shape tail. According to our simulations, the explanation of
732 < the formation for the ripples are originated in the size mismatch
733 < between the head groups and the tails. The ripple phases are only
734 < observed in the studies using larger head group lipid models. However,
735 < 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.
728 > \begin{figure}[htb]
729 > \centering
730 > \includegraphics[width=\linewidth]{phaseDiagram}
731 > \caption{Phase diagram for the simple molecular model as a function
732 > of the head group / molecular width ratio ($\sigma_h / d$) and the
733 > strength of the head group dipole moment
734 > ($\mu$).\label{fig:phaseDiagram}}
735 > \end{figure}
736  
737 < Dipolar head groups are the key elements for the maintaining of the
738 < bilayer structure. The lipids are solvated in water when lowering the
739 < the strength of the dipole on the head groups. The long range
740 < orientational ordering of the dipoles can be achieved by forming the
741 < ripples, although the dipoles are likely to form head-to-tail
742 < configurations even in flat surface, the frustration prevents the
743 < formation of the long range orientational ordering for dipoles. The
744 < corrugation of the surface breaks the frustration and stablizes the
745 < long range oreintational ordering for the dipoles in the head groups
696 < of the lipid molecules. Many rows of the head-to-tail dipoles are
697 < parallel to each other and adopt the antiferroelectric state as a
698 < whole. This is the first time the organization of the head groups in
699 < ripple phases of the lipid bilayer has been addressed.
737 > We have computed translational diffusion coefficients for lipid
738 > molecules from the mean square displacement,
739 > \begin{eqnarray}
740 > \langle {|\left({\bf r}_{i}(t) - {\bf r}_{i}(0) \right)|}^2 \rangle \\ \\
741 > & = & 6Dt
742 > \end{eqnarray}
743 > of the lipid bodies. The values of the translational diffusion
744 > coefficient for different head-to-tail size ratio are shown in table
745 > \ref{tab:relaxation}.
746  
747 < The most important prediction we can make using the results from this
748 < simple model is that if dipolar ordering is driving the surface
749 < corrugation, the wave vectors for the ripples should always found to
750 < be {\it perpendicular} to the dipole director axis.  This prediction
751 < should suggest experimental designs which test whether this is really
752 < true in the phosphatidylcholine $P_{\beta'}$ phases.  The dipole
753 < director axis should also be easily computable for the all-atom and
754 < coarse-grained simulations that have been published in the literature.
747 > We have also computed orientational diffusion constants for the head
748 > groups from the relaxation of the second-order Legendre polynomial
749 > correlation function,
750 > \begin{eqnarray}
751 > C_{\ell}(t) & = & \langle P_{\ell}\left({\bf \mu}_{i}(t) \cdot {\bf
752 > \mu}_{i}(0) \right) \rangle  \\ \\
753 > & \approx & e^{-\ell(\ell + 1) \theta t},
754 > \end{eqnarray}
755 > of the head group dipoles.  In this last line, we have used a simple
756 > ``Debye''-like model for the relaxation of the correlation function,
757 > specifically in the case when $\ell = 2$.   The computed orientational
758 > diffusion constants are given in table \ref{tab:relaxation}.  The
759 > notable feature we observe is that the orientational diffusion
760 > constant for the head group exhibits an order of magnitude decrease
761 > upon entering the rippled phase.  Our orientational correlation times
762 > are substantially in excess of those provided by...
763  
764 +
765 + \begin{table*}
766 + \begin{minipage}{\linewidth}
767 + \begin{center}
768 + \caption{Rotational diffusion constants for the head groups
769 + ($\theta_h$) and molecular bodies ($\theta_b$) as well as the
770 + translational diffusion coefficients for the molecule as a function of
771 + the head-to-body width ratio.  The orientational mobility of the head
772 + groups experiences an {\it order of magnitude decrease} upon entering
773 + the rippled phase, which suggests that the rippling is tied to a
774 + freezing out of head group orientational freedom.  Uncertainties in
775 + the last digit are indicated by the values in parentheses.}
776 + \begin{tabular}{lccc}
777 + \hline
778 + $\sigma_h / d$ & $\theta_h (\mu s^{-1})$ & $\theta_b (1/fs)$ & $D (
779 + \times 10^{-11} m^2 s^{-1}) \\
780 + \hline
781 + 1.20 & $0.206(1) $ & $0.0175(5) $ & $0.43(1)$ \\
782 + 1.28 & $0.179(2) $ & $0.055(2)  $ & $5.91(3)$ \\
783 + 1.35 & $0.025(1) $ & $0.195(3)  $ & $3.42(1)$ \\
784 + 1.41 & $0.023(1) $ & $0.024(3)  $ & $7.16(1)$ \\
785 + \end{tabular}
786 + \label{tab:relaxation}
787 + \end{center}
788 + \end{minipage}
789 + \end{table*}
790 +
791 + \section{Discussion}
792 + \label{sec:discussion}
793 +
794 + Symmetric and asymmetric ripple phases have been observed to form in
795 + our molecular dynamics simulations of a simple molecular-scale lipid
796 + model. The lipid model consists of an dipolar head group and an
797 + ellipsoidal tail.  Within the limits of this model, an explanation for
798 + generalized membrane curvature is a simple mismatch in the size of the
799 + heads with the width of the molecular bodies.  With heads
800 + substantially larger than the bodies of the molecule, this curvature
801 + should be convex nearly everywhere, a requirement which could be
802 + resolved either with micellar or cylindrical phases.
803 +
804 + The persistence of a {\it bilayer} structure therefore requires either
805 + strong attractive forces between the head groups or exclusionary
806 + forces from the solvent phase.  To have a persistent bilayer structure
807 + with the added requirement of convex membrane curvature appears to
808 + result in corrugated structures like the ones pictured in
809 + Fig. \ref{fig:phaseCartoon}.  In each of the sections of these
810 + corrugated phases, the local curvature near a most of the head groups
811 + is convex.  These structures are held together by the extremely strong
812 + and directional interactions between the head groups.
813 +
814 + Dipolar head groups are key for the maintaining the bilayer structures
815 + exhibited by this model.  The dipoles are likely to form head-to-tail
816 + configurations even in flat configurations, but the temperatures are
817 + high enough that vortex defects become prevalent in the flat phase.
818 + The flat phase we observed therefore appears to be substantially above
819 + the Kosterlitz-Thouless transition temperature for a planar system of
820 + dipoles with this set of parameters.  For this reason, it would be
821 + interesting to observe the thermal behavior of the flat phase at
822 + substantially lower temperatures.
823 +
824 + One feature of this model is that an energetically favorable
825 + orientational ordering of the dipoles can be achieved by forming
826 + ripples.  The corrugation of the surface breaks the symmetry of the
827 + plane, making vortex defects somewhat more expensive, and stabilizing
828 + the long range orientational ordering for the dipoles in the head
829 + groups.  Most of the rows of the head-to-tail dipoles are parallel to
830 + each other and the system adopts a bulk anti-ferroelectric state.  We
831 + believe that this is the first time the organization of the head
832 + groups in ripple phases has been addressed.
833 +
834 + Although the size-mismatch between the heads and molecular bodies
835 + appears to be the primary driving force for surface convexity, the
836 + persistence of the bilayer through the use of rippled structures is a
837 + function of the strong, attractive interactions between the heads.
838 + One important prediction we can make using the results from this
839 + simple model is that if the dipole-dipole interaction is the leading
840 + contributor to the head group attractions, the wave vectors for the
841 + ripples should always be found {\it perpendicular} to the dipole
842 + director axis.  This echoes the prediction we made earlier for simple
843 + elastic dipolar membranes, and may suggest experimental designs which
844 + will test whether this is really the case in the phosphatidylcholine
845 + $P_{\beta'}$ phases.  The dipole director axis should also be easily
846 + computable for the all-atom and coarse-grained simulations that have
847 + been published in the literature.\cite{deVries05}
848 +
849   Although our model is simple, it exhibits some rich and unexpected
850 < behaviors.  It would clearly be a closer approximation to the reality
851 < if we allowed greater translational freedom to the dipoles and
852 < replaced the somewhat artificial lattice packing and the harmonic
853 < elastic tension with more realistic molecular modeling potentials.
854 < What we have done is to present a simple model which exhibits bulk
855 < non-thermal corrugation, and our explanation of this rippling
850 > behaviors.  It would clearly be a closer approximation to reality if
851 > we allowed bending motions between the dipoles and the molecular
852 > bodies, and if we replaced the rigid ellipsoids with ball-and-chain
853 > tails.  However, the advantages of this simple model (large system
854 > sizes, 50 fs time steps) allow us to rapidly explore the phase diagram
855 > for a wide range of parameters.  Our explanation of this rippling
856   phenomenon will help us design more accurate molecular models for
857 < corrugated membranes and experiments to test whether rippling is
858 < dipole-driven or not.
720 <
857 > corrugated membranes and experiments to test whether or not
858 > dipole-dipole interactions exert an influence on membrane rippling.
859   \newpage
860   \bibliography{mdripple}
861   \end{document}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines