--- trunk/xDissertation/md.tex 2008/01/30 16:01:02 3336 +++ trunk/xDissertation/md.tex 2008/03/05 23:34:04 3360 @@ -1 +1,897 @@ -\chapter{\label{chap:md}MOLECULAR DYNAMICS} +\chapter{\label{chap:md}DIPOLAR ORDERING IN THE RIPPLE PHASES OF +MOLECULAR-SCALE MODELS OF LIPID MEMBRANES} + +\section{Introduction} +\label{mdsec:Int} +Fully hydrated lipids will aggregate spontaneously to form bilayers +which exhibit a variety of phases depending on their temperatures and +compositions. Among these phases, a periodic rippled phase +($P_{\beta'}$) appears as an intermediate phase between the gel +($L_\beta$) and fluid ($L_{\alpha}$) phases for relatively pure +phosphatidylcholine (PC) bilayers. The ripple phase has attracted +substantial experimental interest over the past 30 years. Most +structural information of the ripple phase has been obtained by the +X-ray diffraction~\cite{Sun96,Katsaras00} and freeze-fracture electron +microscopy (FFEM).~\cite{Copeland80,Meyer96} Recently, Kaasgaard {\it +et al.} used atomic force microscopy (AFM) to observe ripple phase +morphology in bilayers supported on mica.~\cite{Kaasgaard03} The +experimental results provide strong support for a 2-dimensional +hexagonal packing lattice of the lipid molecules within the ripple +phase. This is a notable change from the observed lipid packing +within the gel phase,~\cite{Cevc87} although Tenchov {\it et al.} have +recently observed near-hexagonal packing in some phosphatidylcholine +(PC) gel phases.\cite{Tenchov2001} The X-ray diffraction work by +Katsaras {\it et al.} showed that a rich phase diagram exhibiting both +{\it asymmetric} and {\it symmetric} ripples is possible for lecithin +bilayers.\cite{Katsaras00} + +A number of theoretical models have been presented to explain the +formation of the ripple phase. Marder {\it et al.} used a +curvature-dependent Landau-de~Gennes free-energy functional to predict +a rippled phase.~\cite{Marder84} This model and other related +continuum models predict higher fluidity in convex regions and that +concave portions of the membrane correspond to more solid-like +regions. Carlson and Sethna used a packing-competition model (in +which head groups and chains have competing packing energetics) to +predict the formation of a ripple-like phase. Their model predicted +that the high-curvature portions have lower-chain packing and +correspond to more fluid-like regions. Goldstein and Leibler used a +mean-field approach with a planar model for {\em inter-lamellar} +interactions to predict rippling in multilamellar +phases.~\cite{Goldstein88} McCullough and Scott proposed that the {\em +anisotropy of the nearest-neighbor interactions} coupled to +hydrophobic constraining forces which restrict height differences +between nearest neighbors is the origin of the ripple +phase.~\cite{McCullough90} Lubensky and MacKintosh introduced a Landau +theory for tilt order and curvature of a single membrane and concluded +that {\em coupling of molecular tilt to membrane curvature} is +responsible for the production of ripples.~\cite{Lubensky93} Misbah, +Duplat and Houchmandzadeh proposed that {\em inter-layer dipolar +interactions} can lead to ripple instabilities.~\cite{Misbah98} +Heimburg presented a {\em coexistence model} for ripple formation in +which he postulates that fluid-phase line defects cause sharp +curvature between relatively flat gel-phase regions.~\cite{Heimburg00} +Kubica has suggested that a lattice model of polar head groups could +be valuable in trying to understand bilayer phase +formation.~\cite{Kubica02} Bannerjee used Monte Carlo simulations of +lamellar stacks of hexagonal lattices to show that large head groups +and molecular tilt with respect to the membrane normal vector can +cause bulk rippling.~\cite{Bannerjee02} Recently, Kranenburg and Smit +described the formation of symmetric ripple-like structures using a +coarse grained solvent-head-tail bead model.\cite{Kranenburg2005} +Their lipids consisted of a short chain of head beads tied to the two +longer ``chains''. + +In contrast, few large-scale molecular modeling studies have been +done due to the large size of the resulting structures and the time +required for the phases of interest to develop. With all-atom (and +even unified-atom) simulations, only one period of the ripple can be +observed and only for time scales in the range of 10-100 ns. One of +the most interesting molecular simulations was carried out by de~Vries +{\it et al.}~\cite{deVries05}. According to their simulation results, +the ripple consists of two domains, one resembling the gel bilayer, +while in the other, the two leaves of the bilayer are fully +interdigitated. The mechanism for the formation of the ripple phase +suggested by their work is a packing competition between the head +groups and the tails of the lipid molecules.\cite{Carlson87} Recently, +the ripple phase has also been studied by Lenz and Schmid using Monte +Carlo simulations.\cite{Lenz07} Their structures are similar to the De +Vries {\it et al.} structures except that the connection between the +two leaves of the bilayer is a narrow interdigitated line instead of +the fully interdigitated domain. The symmetric ripple phase was also +observed by Lenz {\it et al.}, and their work supports other claims +that the mismatch between the size of the head group and tail of the +lipid molecules is the driving force for the formation of the ripple +phase. Ayton and Voth have found significant undulations in +zero-surface-tension states of membranes simulated via dissipative +particle dynamics, but their results are consistent with purely +thermal undulations.~\cite{Ayton02} + +Although the organization of the tails of lipid molecules are +addressed by these molecular simulations and the packing competition +between head groups and tails is strongly implicated as the primary +driving force for ripple formation, questions about the ordering of +the head groups in ripple phase have not been settled. + +In a recent paper, we presented a simple ``web of dipoles'' spin +lattice model which provides some physical insight into relationship +between dipolar ordering and membrane buckling.\cite{sun:031602} We +found that dipolar elastic membranes can spontaneously buckle, forming +ripple-like topologies. The driving force for the buckling of dipolar +elastic membranes is the anti-ferroelectric ordering of the dipoles. +This was evident in the ordering of the dipole director axis +perpendicular to the wave vector of the surface ripples. A similar +phenomenon has also been observed by Tsonchev {\it et al.} in their +work on the spontaneous formation of dipolar peptide chains into +curved nano-structures.\cite{Tsonchev04,Tsonchev04II} + +In this paper, we construct a somewhat more realistic molecular-scale +lipid model than our previous ``web of dipoles'' and use molecular +dynamics simulations to elucidate the role of the head group dipoles +in the formation and morphology of the ripple phase. We describe our +model and computational methodology in section \ref{mdsec:method}. +Details on the simulations are presented in section +\ref{mdsec:experiment}, with results following in section +\ref{mdsec:results}. A final discussion of the role of dipolar heads in +the ripple formation can be found in section +\ref{mdsec:discussion}. + +\section{Computational Model} +\label{mdsec:method} + +\begin{figure} +\centering +\includegraphics[width=\linewidth]{./figures/mdLipidModels.pdf} +\caption{Three different representations of DPPC lipid molecules, +including the chemical structure, an atomistic model, and the +head-body ellipsoidal coarse-grained model used in this +work.\label{mdfig:lipidModels}} +\end{figure} + +Our simple molecular-scale lipid model for studying the ripple phase +is based on two facts: one is that the most essential feature of lipid +molecules is their amphiphilic structure with polar head groups and +non-polar tails. Another fact is that the majority of lipid molecules +in the ripple phase are relatively rigid (i.e. gel-like) which makes +some fraction of the details of the chain dynamics negligible. Figure +\ref{mdfig:lipidModels} shows the molecular structure of a DPPC +molecule, as well as atomistic and molecular-scale representations of +a DPPC molecule. The hydrophilic character of the head group is +largely due to the separation of charge between the nitrogen and +phosphate groups. The zwitterionic nature of the PC headgroups leads +to abnormally large dipole moments (as high as 20.6 D), and this +strongly polar head group interacts strongly with the solvating water +layers immediately surrounding the membrane. The hydrophobic tail +consists of fatty acid chains. In our molecular scale model, lipid +molecules have been reduced to these essential features; the fatty +acid chains are represented by an ellipsoid with a dipolar ball +perched on one end to represent the effects of the charge-separated +head group. In real PC lipids, the direction of the dipole is +nearly perpendicular to the tail, so we have fixed the direction of +the point dipole rigidly in this orientation. + +The ellipsoidal portions of the model interact via the Gay-Berne +potential which has seen widespread use in the liquid crystal +community. Ayton and Voth have also used Gay-Berne ellipsoids for +modeling large length-scale properties of lipid +bilayers.\cite{Ayton01} In its original form, the Gay-Berne potential +was a single site model for the interactions of rigid ellipsoidal +molecules.\cite{Gay1981} It can be thought of as a modification of the +Gaussian overlap model originally described by Berne and +Pechukas.\cite{Berne72} The potential is constructed in the familiar +form of the Lennard-Jones function using orientation-dependent +$\sigma$ and $\epsilon$ parameters, +\begin{equation*} +V_{ij}({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j}, {\mathbf{\hat +r}_{ij}}) = 4\epsilon ({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j}, +{\mathbf{\hat r}_{ij}})\left[\left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u}_i}, +{\mathbf{\hat u}_j}, {\mathbf{\hat r}_{ij}})+\sigma_0}\right)^{12} +-\left(\frac{\sigma_0}{r_{ij}-\sigma({\mathbf{\hat u}_i}, {\mathbf{\hat u}_j}, +{\mathbf{\hat r}_{ij}})+\sigma_0}\right)^6\right] +\label{mdeq:gb} +\end{equation*} + +The range $(\sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf +\hat{r}}_{ij}))$, and strength $(\epsilon({\bf \hat{u}}_{i},{\bf +\hat{u}}_{j},{\bf \hat{r}}_{ij}))$ parameters +are dependent on the relative orientations of the two molecules (${\bf +\hat{u}}_{i},{\bf \hat{u}}_{j}$) as well as the direction of the +intermolecular separation (${\bf \hat{r}}_{ij}$). $\sigma$ and +$\sigma_0$ are also governed by shape mixing and anisotropy variables, +\begin {eqnarray*} +\sigma_0 & = & \sqrt{d_i^2 + d_j^2} \\ \\ +\chi & = & \left[ \frac{\left( l_i^2 - d_i^2 \right) \left(l_j^2 - +d_j^2 \right)}{\left( l_j^2 + d_i^2 \right) \left(l_i^2 + +d_j^2 \right)}\right]^{1/2} \\ \\ +\alpha^2 & = & \left[ \frac{\left( l_i^2 - d_i^2 \right) \left(l_j^2 + +d_i^2 \right)}{\left( l_j^2 - d_j^2 \right) \left(l_i^2 + +d_j^2 \right)}\right]^{1/2}, +\end{eqnarray*} +where $l$ and $d$ describe the length and width of each uniaxial +ellipsoid. These shape anisotropy parameters can then be used to +calculate the range function, +\begin{equation*} +\sigma({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf \hat{r}}_{ij}) = \sigma_{0} + \left[ 1- \left\{ \frac{ \chi \alpha^2 ({\bf \hat{u}}_i \cdot {\bf +\hat{r}}_{ij} ) + \chi \alpha^{-2} ({\bf \hat{u}}_j \cdot {\bf +\hat{r}}_{ij} ) - 2 \chi^2 ({\bf \hat{u}}_i \cdot {\bf +\hat{r}}_{ij} )({\bf \hat{u}}_j \cdot {\bf +\hat{r}}_{ij} ) ({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j)}{1 - \chi^2 +\left({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j\right)^2} \right\} +\right]^{-1/2} +\end{equation*} + +Gay-Berne ellipsoids also have an energy scaling parameter, +$\epsilon^s$, which describes the well depth for two identical +ellipsoids in a {\it side-by-side} configuration. Additionally, a well +depth aspect ratio, $\epsilon^r = \epsilon^e / \epsilon^s$, describes +the ratio between the well depths in the {\it end-to-end} and +side-by-side configurations. As in the range parameter, a set of +mixing and anisotropy variables can be used to describe the well +depths for dissimilar particles, +\begin {eqnarray*} +\epsilon_0 & = & \sqrt{\epsilon^s_i * \epsilon^s_j} \\ \\ +\epsilon^r & = & \sqrt{\epsilon^r_i * \epsilon^r_j} \\ \\ +\chi' & = & \frac{1 - (\epsilon^r)^{1/\mu}}{1 + (\epsilon^r)^{1/\mu}} +\\ \\ +\alpha'^2 & = & \left[1 + (\epsilon^r)^{1/\mu}\right]^{-1} +\end{eqnarray*} +The form of the strength function is somewhat complicated, +\begin {eqnarray*} +\epsilon({\bf \hat{u}}_{i}, {\bf \hat{u}}_{j},{\bf \hat{r}}_{ij}) & = & +\epsilon_{0} \epsilon_{1}^{\nu}({\bf \hat{u}}_{i}.{\bf \hat{u}}_{j}) + \epsilon_{2}^{\mu}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf +\hat{r}}_{ij}) \\ \\ +\epsilon_{1}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j}) & = & +\left[1-\chi^{2}({\bf \hat{u}}_{i}.{\bf +\hat{u}}_{j})^{2}\right]^{-1/2} \\ \\ +\epsilon_{2}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf \hat{r}}_{ij}) & += & + 1 - \left\{ \frac{ \chi' \alpha'^2 ({\bf \hat{u}}_i \cdot {\bf +\hat{r}}_{ij} ) + \chi' \alpha'^{-2} ({\bf \hat{u}}_j \cdot {\bf +\hat{r}}_{ij} ) - 2 \chi'^2 ({\bf \hat{u}}_i \cdot {\bf +\hat{r}}_{ij} )({\bf \hat{u}}_j \cdot {\bf +\hat{r}}_{ij} ) ({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j)}{1 - \chi'^2 +\left({\bf \hat{u}}_i \cdot {\bf \hat{u}}_j\right)^2} \right\}, +\end {eqnarray*} +although many of the quantities and derivatives are identical with +those obtained for the range parameter. Ref. \citen{Luckhurst90} +has a particularly good explanation of the choice of the Gay-Berne +parameters $\mu$ and $\nu$ for modeling liquid crystal molecules. An +excellent overview of the computational methods that can be used to +efficiently compute forces and torques for this potential can be found +in Ref. \citen{Golubkov06} + +The choices of parameters we have used in this study correspond to a +shape anisotropy of 3 for the chain portion of the molecule. In +principle, this could be varied to allow for modeling of longer or +shorter chain lipid molecules. For these prolate ellipsoids, we have: +\begin{equation} +\begin{array}{rcl} +d & < & l \\ +\epsilon^{r} & < & 1 +\end{array} +\end{equation} +A sketch of the various structural elements of our molecular-scale +lipid / solvent model is shown in figure \ref{mdfig:lipidModel}. The +actual parameters used in our simulations are given in table +\ref{mdtab:parameters}. + +\begin{figure} +\centering +\includegraphics[width=\linewidth]{./figures/md2LipidModel.pdf} +\caption{The parameters defining the behavior of the lipid +models. $\sigma_h / d$ is the ratio of the head group to body diameter. +Molecular bodies had a fixed aspect ratio of 3.0. The solvent model +was a simplified 4-water bead ($\sigma_w \approx d$) that has been +used in other coarse-grained simulations. The dipolar strength +(and the temperature and pressure) were the only other parameters that +were varied systematically.\label{mdfig:lipidModel}} +\end{figure} + +To take into account the permanent dipolar interactions of the +zwitterionic head groups, we have placed fixed dipole moments +$\mu_{i}$ at one end of the Gay-Berne particles. The dipoles are +oriented at an angle $\theta = \pi / 2$ relative to the major axis. +These dipoles are protected by a head ``bead'' with a range parameter +($\sigma_h$) which we have varied between $1.20 d$ and $1.41 d$. The +head groups interact with each other using a combination of +Lennard-Jones, +\begin{equation} +V_{ij}(r_{ij}) = 4\epsilon_h \left[\left(\frac{\sigma_h}{r_{ij}}\right)^{12} - +\left(\frac{\sigma_h}{r_{ij}}\right)^6\right], +\end{equation} +and dipole-dipole, +\begin{equation} +V_{ij}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf +\hat{r}}_{ij})) = \frac{|\mu|^2}{4 \pi \epsilon_0 r_{ij}^3} +\left[ \hat{\bf u}_i \cdot \hat{\bf u}_j - 3 (\hat{\bf u}_i \cdot +\hat{\bf r}_{ij})(\hat{\bf u}_j \cdot \hat{\bf r}_{ij}) \right] +\end{equation} +potentials. +In these potentials, $\mathbf{\hat u}_i$ is the unit vector pointing +along dipole $i$ and $\mathbf{\hat r}_{ij}$ is the unit vector +pointing along the inter-dipole vector $\mathbf{r}_{ij}$. + +Since the charge separation distance is so large in zwitterionic head +groups (like the PC head groups), it would also be possible to use +either point charges or a ``split dipole'' approximation, +\begin{equation} +V_{ij}({\bf \hat{u}}_{i},{\bf \hat{u}}_{j},{\bf +\hat{r}}_{ij})) = \frac{1}{{4\pi \epsilon_0 }}\left[ {\frac{{\mu _i \cdot \mu _j }}{{R_{ij}^3 }} - +\frac{{3\left( {\mu _i \cdot r_{ij} } \right)\left( {\mu _i \cdot +r_{ij} } \right)}}{{R_{ij}^5 }}} \right] +\end{equation} +where $\mu _i$ and $\mu _j$ are the dipole moments of sites $i$ and +$j$, $r_{ij}$ is vector between the two sites, and $R_{ij}$ is given +by, +\begin{equation} +R_{ij} = \sqrt {r_{ij}^2 + \frac{{d_i^2 }}{4} + \frac{{d_j^2 +}}{4}}. +\end{equation} +Here, $d_i$ and $d_j$ are charge separation distances associated with +each of the two dipolar sites. This approximation to the multipole +expansion maintains the fast fall-off of the multipole potentials but +lacks the normal divergences when two polar groups get close to one +another. + +For the interaction between nonequivalent uniaxial ellipsoids (in this +case, between spheres and ellipsoids), the spheres are treated as +ellipsoids with an aspect ratio of 1 ($d = l$) and with an well depth +ratio ($\epsilon^r$) of 1 ($\epsilon^e = \epsilon^s$). The form of +the Gay-Berne potential we are using was generalized by Cleaver {\it +et al.} and is appropriate for dissimilar uniaxial +ellipsoids.\cite{Cleaver96} + +The solvent model in our simulations is similar to the one used by +Marrink {\it et al.} in their coarse grained simulations of lipid +bilayers.\cite{Marrink2004} The solvent bead is a single site that +represents four water molecules (m = 72 amu) and has comparable +density and diffusive behavior to liquid water. However, since there +are no electrostatic sites on these beads, this solvent model cannot +replicate the dielectric properties of water. Note that although we +are using larger cutoff and switching radii than Marrink {\it et al.}, +our solvent density at 300 K remains at 0.944 g cm$^{-3}$, and the +solvent diffuses at 0.43 \AA$^2 ps^{-1}$ (only twice as fast as liquid +water). + +\begin{table*} +\begin{minipage}{\linewidth} +\begin{center} +\caption{Potential parameters used for molecular-scale coarse-grained +lipid simulations} +\begin{tabular}{llccc} +\hline + & & Head & Chain & Solvent \\ +\hline +$d$ (\AA) & & varied & 4.6 & 4.7 \\ +$l$ (\AA) & & $= d$ & 13.8 & 4.7 \\ +$\epsilon^s$ (kcal/mol) & & 0.185 & 1.0 & 0.8 \\ +$\epsilon_r$ (well-depth aspect ratio)& & 1 & 0.2 & 1 \\ +$m$ (amu) & & 196 & 760 & 72.06 \\ +$\overleftrightarrow{\mathsf I}$ (amu \AA$^2$) & & & & \\ +\multicolumn{2}c{$I_{xx}$} & 1125 & 45000 & N/A \\ +\multicolumn{2}c{$I_{yy}$} & 1125 & 45000 & N/A \\ +\multicolumn{2}c{$I_{zz}$} & 0 & 9000 & N/A \\ +$\mu$ (Debye) & & varied & 0 & 0 \\ +\end{tabular} +\label{mdtab:parameters} +\end{center} +\end{minipage} +\end{table*} + +\section{Experimental Methodology} +\label{mdsec:experiment} + +The parameters that were systematically varied in this study were the +size of the head group ($\sigma_h$), the strength of the dipole moment +($\mu$), and the temperature of the system. Values for $\sigma_h$ +ranged from 5.5 \AA\ to 6.5 \AA. If the width of the tails is taken +to be the unit of length, these head groups correspond to a range from +$1.2 d$ to $1.41 d$. Since the solvent beads are nearly identical in +diameter to the tail ellipsoids, all distances that follow will be +measured relative to this unit of distance. Because the solvent we +are using is non-polar and has a dielectric constant of 1, values for +$\mu$ are sampled from a range that is somewhat smaller than the 20.6 +Debye dipole moment of the PC head groups. + +To create unbiased bilayers, all simulations were started from two +perfectly flat monolayers separated by a 26 \AA\ gap between the +molecular bodies of the upper and lower leaves. The separated +monolayers were evolved in a vacuum with $x-y$ anisotropic pressure +coupling. The length of $z$ axis of the simulations was fixed and a +constant surface tension was applied to enable real fluctuations of +the bilayer. Periodic boundary conditions were used, and $480-720$ +lipid molecules were present in the simulations, depending on the size +of the head beads. In all cases, the two monolayers spontaneously +collapsed into bilayer structures within 100 ps. Following this +collapse, all systems were equilibrated for $100$ ns at $300$ K. + +The resulting bilayer structures were then solvated at a ratio of $6$ +solvent beads (24 water molecules) per lipid. These configurations +were then equilibrated for another $30$ ns. All simulations utilizing +the solvent were carried out at constant pressure ($P=1$ atm) with +$3$D anisotropic coupling, and small constant surface tension +($\gamma=0.015$ N/m). Given the absence of fast degrees of freedom in +this model, a time step of $50$ fs was utilized with excellent energy +conservation. Data collection for structural properties of the +bilayers was carried out during a final 5 ns run following the solvent +equilibration. Orientational correlation functions and diffusion +constants were computed from 30 ns simulations in the microcanonical +(NVE) ensemble using the average volume from the end of the constant +pressure and surface tension runs. The timestep on these final +molecular dynamics runs was 25 fs. No appreciable changes in phase +structure were noticed upon switching to a microcanonical ensemble. +All simulations were performed using the {\sc oopse} molecular +modeling program.\cite{Meineke2005} + +A switching function was applied to all potentials to smoothly turn +off the interactions between a range of $22$ and $25$ \AA. The +switching function was the standard (cubic) function, +\begin{equation} +s(r) = + \begin{cases} + 1 & \text{if $r \le r_{\text{sw}}$},\\ + \frac{(r_{\text{cut}} + 2r - 3r_{\text{sw}})(r_{\text{cut}} - r)^2} + {(r_{\text{cut}} - r_{\text{sw}})^3} + & \text{if $r_{\text{sw}} < r \le r_{\text{cut}}$}, \\ + 0 & \text{if $r > r_{\text{cut}}$.} + \end{cases} +\label{mdeq:dipoleSwitching} +\end{equation} + +\section{Results} +\label{mdsec:results} + +The membranes in our simulations exhibit a number of interesting +bilayer phases. The surface topology of these phases depends most +sensitively on the ratio of the size of the head groups to the width +of the molecular bodies. With heads only slightly larger than the +bodies ($\sigma_h=1.20 d$) the membrane exhibits a flat bilayer. + +Increasing the head / body size ratio increases the local membrane +curvature around each of the lipids. With $\sigma_h=1.28 d$, the +surface is still essentially flat, but the bilayer starts to exhibit +signs of instability. We have observed occasional defects where a +line of lipid molecules on one leaf of the bilayer will dip down to +interdigitate with the other leaf. This gives each of the two bilayer +leaves some local convexity near the line defect. These structures, +once developed in a simulation, are very stable and are spaced +approximately 100 \AA\ away from each other. + +With larger heads ($\sigma_h = 1.35 d$) the membrane curvature +resolves into a ``symmetric'' ripple phase. Each leaf of the bilayer +is broken into several convex, hemicylinderical sections, and opposite +leaves are fitted together much like roof tiles. There is no +interdigitation between the upper and lower leaves of the bilayer. + +For the largest head / body ratios studied ($\sigma_h = 1.41 d$) the +local curvature is substantially larger, and the resulting bilayer +structure resolves into an asymmetric ripple phase. This structure is +very similar to the structures observed by both de~Vries {\it et al.} +and Lenz {\it et al.}. For a given ripple wave vector, there are two +possible asymmetric ripples, which is not the case for the symmetric +phase observed when $\sigma_h = 1.35 d$. + +\begin{figure} +\centering +\includegraphics[width=\linewidth]{./figures/mdPhaseCartoon.pdf} +\caption{The role of the ratio between the head group size and the +width of the molecular bodies is to increase the local membrane +curvature. With strong attractive interactions between the head +groups, this local curvature can be maintained in bilayer structures +through surface corrugation. Shown above are three phases observed in +these simulations. With $\sigma_h = 1.20 d$, the bilayer maintains a +flat topology. For larger heads ($\sigma_h = 1.35 d$) the local +curvature resolves into a symmetrically rippled phase with little or +no interdigitation between the upper and lower leaves of the membrane. +The largest heads studied ($\sigma_h = 1.41 d$) resolve into an +asymmetric rippled phases with interdigitation between the two +leaves.\label{mdfig:phaseCartoon}} +\end{figure} + +Sample structures for the flat ($\sigma_h = 1.20 d$), symmetric +($\sigma_h = 1.35 d$, and asymmetric ($\sigma_h = 1.41 d$) ripple +phases are shown in Figure \ref{mdfig:phaseCartoon}. + +It is reasonable to ask how well the parameters we used can produce +bilayer properties that match experimentally known values for real +lipid bilayers. Using a value of $l = 13.8$ \AA for the ellipsoidal +tails and the fixed ellipsoidal aspect ratio of 3, our values for the +area per lipid ($A$) and inter-layer spacing ($D_{HH}$) depend +entirely on the size of the head bead relative to the molecular body. +These values are tabulated in table \ref{mdtab:property}. Kucera {\it +et al.} have measured values for the head group spacings for a number +of PC lipid bilayers that range from 30.8 \AA\ (DLPC) to 37.8 \AA\ (DPPC). +They have also measured values for the area per lipid that range from +60.6 +\AA$^2$ (DMPC) to 64.2 \AA$^2$ +(DPPC).\cite{NorbertKucerka04012005,NorbertKucerka06012006} Only the +largest of the head groups we modeled ($\sigma_h = 1.41 d$) produces +bilayers (specifically the area per lipid) that resemble real PC +bilayers. The smaller head beads we used are perhaps better models +for PE head groups. + +\begin{table*} +\begin{minipage}{\linewidth} +\begin{center} +\caption{Phase, bilayer spacing, area per lipid, ripple wavelength +and amplitude observed as a function of the ratio between the head +beads and the diameters of the tails. Ripple wavelengths and +amplitudes are normalized to the diameter of the tail ellipsoids.} +\begin{tabular}{lccccc} +\hline +$\sigma_h / d$ & type of phase & bilayer spacing (\AA) & area per +lipid (\AA$^2$) & $\lambda / d$ & $A / d$\\ +\hline +1.20 & flat & 33.4 & 49.6 & N/A & N/A \\ +1.28 & flat & 33.7 & 54.7 & N/A & N/A \\ +1.35 & symmetric ripple & 42.9 & 51.7 & 17.2 & 2.2 \\ +1.41 & asymmetric ripple & 37.1 & 63.1 & 15.4 & 1.5 \\ +\end{tabular} +\label{mdtab:property} +\end{center} +\end{minipage} +\end{table*} + +The membrane structures and the reduced wavelength $\lambda / d$, +reduced amplitude $A / d$ of the ripples are summarized in Table +\ref{mdtab:property}. The wavelength range is $15 - 17$ molecular bodies +and the amplitude is $1.5$ molecular bodies for asymmetric ripple and +$2.2$ for symmetric ripple. These values are reasonably consistent +with experimental measurements.\cite{Sun96,Katsaras00,Kaasgaard03} +Note, that given the lack of structural freedom in the tails of our +model lipids, the amplitudes observed from these simulations are +likely to underestimate of the true amplitudes. + +\begin{figure} +\centering +\includegraphics[width=\linewidth]{./figures/mdTopDown.pdf} +\caption{Top views of the flat (upper), symmetric ripple (middle), +and asymmetric ripple (lower) phases. Note that the head-group +dipoles have formed head-to-tail chains in all three of these phases, +but in the two rippled phases, the dipolar chains are all aligned {\it +perpendicular} to the direction of the ripple. Note that the flat +membrane has multiple vortex defects in the dipolar ordering, and the +ordering on the lower leaf of the bilayer can be in an entirely +different direction from the upper leaf.\label{mdfig:topView}} +\end{figure} + +The principal method for observing orientational ordering in dipolar +or liquid crystalline systems is the $P_2$ order parameter (defined +as $1.5 \times \lambda_{max}$, where $\lambda_{max}$ is the largest +eigenvalue of the matrix, +\begin{equation} +{\mathsf{S}} = \frac{1}{N} \sum_i \left( +\begin{array}{ccc} + u^{x}_i u^{x}_i-\frac{1}{3} & u^{x}_i u^{y}_i & u^{x}_i u^{z}_i \\ + u^{y}_i u^{x}_i & u^{y}_i u^{y}_i -\frac{1}{3} & u^{y}_i u^{z}_i \\ + u^{z}_i u^{x}_i & u^{z}_i u^{y}_i & u^{z}_i u^{z}_i -\frac{1}{3} +\end{array} \right). +\label{mdeq:opmatrix} +\end{equation} +Here $u^{\alpha}_i$ is the $\alpha=x,y,z$ component of the unit vector +for molecule $i$. (Here, $\hat{\bf u}_i$ can refer either to the +principal axis of the molecular body or to the dipole on the head +group of the molecule.) $P_2$ will be $1.0$ for a perfectly-ordered +system and near $0$ for a randomized system. Note that this order +parameter is {\em not} equal to the polarization of the system. For +example, the polarization of a perfect anti-ferroelectric arrangement +of point dipoles is $0$, but $P_2$ for the same system is $1$. The +eigenvector of $\mathsf{S}$ corresponding to the largest eigenvalue is +familiar as the director axis, which can be used to determine a +privileged axis for an orientationally-ordered system. Since the +molecular bodies are perpendicular to the head group dipoles, it is +possible for the director axes for the molecular bodies and the head +groups to be completely decoupled from each other. + +Figure \ref{mdfig:topView} shows snapshots of bird's-eye views of the +flat ($\sigma_h = 1.20 d$) and rippled ($\sigma_h = 1.35, 1.41 d$) +bilayers. The directions of the dipoles on the head groups are +represented with two colored half spheres: blue (phosphate) and yellow +(amino). For flat bilayers, the system exhibits signs of +orientational frustration; some disorder in the dipolar head-to-tail +chains is evident with kinks visible at the edges between differently +ordered domains. The lipids can also move independently of lipids in +the opposing leaf, so the ordering of the dipoles on one leaf is not +necessarily consistent with the ordering on the other. These two +factors keep the total dipolar order parameter relatively low for the +flat phases. + +With increasing head group size, the surface becomes corrugated, and +the dipoles cannot move as freely on the surface. Therefore, the +translational freedom of lipids in one layer is dependent upon the +position of the lipids in the other layer. As a result, the ordering of +the dipoles on head groups in one leaf is correlated with the ordering +in the other leaf. Furthermore, as the membrane deforms due to the +corrugation, the symmetry of the allowed dipolar ordering on each leaf +is broken. The dipoles then self-assemble in a head-to-tail +configuration, and the dipolar order parameter increases dramatically. +However, the total polarization of the system is still close to zero. +This is strong evidence that the corrugated structure is an +anti-ferroelectric state. It is also notable that the head-to-tail +arrangement of the dipoles is always observed in a direction +perpendicular to the wave vector for the surface corrugation. This is +a similar finding to what we observed in our earlier work on the +elastic dipolar membranes.\cite{sun:031602} + +The $P_2$ order parameters (for both the molecular bodies and the head +group dipoles) have been calculated to quantify the ordering in these +phases. Figure \ref{mdfig:rP2} shows that the $P_2$ order parameter for +the head-group dipoles increases with increasing head group size. When +the heads of the lipid molecules are small, the membrane is nearly +flat. Since the in-plane packing is essentially a close packing of the +head groups, the head dipoles exhibit frustration in their +orientational ordering. + +The ordering trends for the tails are essentially opposite to the +ordering of the head group dipoles. The tail $P_2$ order parameter +{\it decreases} with increasing head size. This indicates that the +surface is more curved with larger head / tail size ratios. When the +surface is flat, all tails are pointing in the same direction (normal +to the bilayer surface). This simplified model appears to be +exhibiting a smectic A fluid phase, similar to the real $L_{\beta}$ +phase. We have not observed a smectic C gel phase ($L_{\beta'}$) for +this model system. Increasing the size of the heads results in +rapidly decreasing $P_2$ ordering for the molecular bodies. + +\begin{figure} +\centering +\includegraphics[width=\linewidth]{./figures/mdRP2.pdf} +\caption{The $P_2$ order parameters for head groups (circles) and +molecular bodies (squares) as a function of the ratio of head group +size ($\sigma_h$) to the width of the molecular bodies ($d$). \label{mdfig:rP2}} +\end{figure} + +In addition to varying the size of the head groups, we studied the +effects of the interactions between head groups on the structure of +lipid bilayer by changing the strength of the dipoles. Figure +\ref{mdfig:sP2} shows how the $P_2$ order parameter changes with +increasing strength of the dipole. Generally, the dipoles on the head +groups become more ordered as the strength of the interaction between +heads is increased and become more disordered by decreasing the +interaction strength. When the interaction between the heads becomes +too weak, the bilayer structure does not persist; all lipid molecules +become dispersed in the solvent (which is non-polar in this +molecular-scale model). The critical value of the strength of the +dipole depends on the size of the head groups. The perfectly flat +surface becomes unstable below $5$ Debye, while the rippled +surfaces melt at $8$ Debye (asymmetric) or $10$ Debye (symmetric). + +The ordering of the tails mirrors the ordering of the dipoles {\it +except for the flat phase}. Since the surface is nearly flat in this +phase, the order parameters are only weakly dependent on dipolar +strength until it reaches $15$ Debye. Once it reaches this value, the +head group interactions are strong enough to pull the head groups +close to each other and distort the bilayer structure. For a flat +surface, a substantial amount of free volume between the head groups +is normally available. When the head groups are brought closer by +dipolar interactions, the tails are forced to splay outward, first forming +curved bilayers, and then inverted micelles. + +When $\sigma_h=1.28 d$, the $P_2$ order parameter decreases slightly +when the strength of the dipole is increased above $16$ Debye. For +rippled bilayers, there is less free volume available between the head +groups. Therefore increasing dipolar strength weakly influences the +structure of the membrane. However, the increase in the body $P_2$ +order parameters implies that the membranes are being slightly +flattened due to the effects of increasing head-group attraction. + +A very interesting behavior takes place when the head groups are very +large relative to the molecular bodies ($\sigma_h = 1.41 d$) and the +dipolar strength is relatively weak ($\mu < 9$ Debye). In this case, +the two leaves of the bilayer become totally interdigitated with each +other in large patches of the membrane. With higher dipolar +strength, the interdigitation is limited to single lines that run +through the bilayer in a direction perpendicular to the ripple wave +vector. + +\begin{figure} +\centering +\includegraphics[width=\linewidth]{./figures/mdSP2.pdf} +\caption{The $P_2$ order parameters for head group dipoles (a) and +molecular bodies (b) as a function of the strength of the dipoles. +These order parameters are shown for four values of the head group / +molecular width ratio ($\sigma_h / d$). \label{mdfig:sP2}} +\end{figure} + +Figure \ref{mdfig:tP2} shows the dependence of the order parameters on +temperature. As expected, systems are more ordered at low +temperatures, and more disordered at high temperatures. All of the +bilayers we studied can become unstable if the temperature becomes +high enough. The only interesting feature of the temperature +dependence is in the flat surfaces ($\sigma_h=1.20 d$ and +$\sigma_h=1.28 d$). Here, when the temperature is increased above +$310$K, there is enough jostling of the head groups to allow the +dipolar frustration to resolve into more ordered states. This results +in a slight increase in the $P_2$ order parameter above this +temperature. + +For the rippled surfaces ($\sigma_h=1.35 d$ and $\sigma_h=1.41 d$), +there is a slightly increased orientational ordering in the molecular +bodies above $290$K. Since our model lacks the detailed information +about the behavior of the lipid tails, this is the closest the model +can come to depicting the ripple ($P_{\beta'}$) to fluid +($L_{\alpha}$) phase transition. What we are observing is a +flattening of the rippled structures made possible by thermal +expansion of the tightly-packed head groups. The lack of detailed +chain configurations also makes it impossible for this model to depict +the ripple to gel ($L_{\beta'}$) phase transition. + +\begin{figure} +\centering +\includegraphics[width=\linewidth]{./figures/mdTP2.pdf} +\caption{The $P_2$ order parameters for head group dipoles (a) and +molecular bodies (b) as a function of temperature. +These order parameters are shown for four values of the head group / +molecular width ratio ($\sigma_h / d$).\label{mdfig:tP2}} +\end{figure} + +Fig. \ref{mdfig:phaseDiagram} shows a phase diagram for the model as a +function of the head group / molecular width ratio ($\sigma_h / d$) +and the strength of the head group dipole moment ($\mu$). Note that +the specific form of the bilayer phase is governed almost entirely by +the head group / molecular width ratio, while the strength of the +dipolar interactions between the head groups governs the stability of +the bilayer phase. Weaker dipoles result in unstable bilayer phases, +while extremely strong dipoles can shift the equilibrium to an +inverted micelle phase when the head groups are small. Temperature +has little effect on the actual bilayer phase observed, although higher +temperatures can cause the unstable region to grow into the higher +dipole region of this diagram. + +\begin{figure} +\centering +\includegraphics[width=\linewidth]{./figures/mdPhaseDiagram.pdf} +\caption{Phase diagram for the simple molecular model as a function +of the head group / molecular width ratio ($\sigma_h / d$) and the +strength of the head group dipole moment +($\mu$).\label{mdfig:phaseDiagram}} +\end{figure} + +We have computed translational diffusion constants for lipid molecules +from the mean-square displacement, +\begin{equation} +D = \lim_{t \rightarrow \infty} \frac{1}{6 t} \langle {|\left({\bf r}_{i}(t) - {\bf r}_{i}(0) \right)|}^2 \rangle, +\end{equation} +of the lipid bodies. Translational diffusion constants for the +different head-to-tail size ratios (all at 300 K) are shown in table +\ref{mdtab:relaxation}. We have also computed orientational correlation +times for the head groups from fits of the second-order Legendre +polynomial correlation function, +\begin{equation} +C_{\ell}(t) = \langle P_{\ell}\left({\bf \mu}_{i}(t) \cdot {\bf +\mu}_{i}(0) \right) \rangle +\end{equation} +of the head group dipoles. The orientational correlation functions +appear to have multiple components in their decay: a fast ($12 \pm 2$ +ps) decay due to librational motion of the head groups, as well as +moderate ($\tau^h_{\rm mid}$) and slow ($\tau^h_{\rm slow}$) +components. The fit values for the moderate and slow correlation +times are listed in table \ref{mdtab:relaxation}. Standard deviations +in the fit time constants are quite large (on the order of the values +themselves). + +Sparrman and Westlund used a multi-relaxation model for NMR lineshapes +observed in gel, fluid, and ripple phases of DPPC and obtained +estimates of a correlation time for water translational diffusion +($\tau_c$) of 20 ns.\cite{Sparrman2003} This correlation time +corresponds to water bound to small regions of the lipid membrane. +They further assume that the lipids can explore only a single period +of the ripple (essentially moving in a nearly one-dimensional path to +do so), and the correlation time can therefore be used to estimate a +value for the translational diffusion constant of $2.25 \times +10^{-11} m^2 s^{-1}$. The translational diffusion constants we obtain +are in reasonable agreement with this experimentally determined +value. However, the $T_2$ relaxation times obtained by Sparrman and +Westlund are consistent with P-N vector reorientation timescales of +2-5 ms. This is substantially slower than even the slowest component +we observe in the decay of our orientational correlation functions. +Other than the dipole-dipole interactions, our head groups have no +shape anisotropy which would force them to move as a unit with +neighboring molecules. This would naturally lead to P-N reorientation +times that are too fast when compared with experimental measurements. + +\begin{table*} +\begin{minipage}{\linewidth} +\begin{center} +\caption{Fit values for the rotational correlation times for the head +groups ($\tau^h$) and molecular bodies ($\tau^b$) as well as the +translational diffusion constants for the molecule as a function of +the head-to-body width ratio. All correlation functions and transport +coefficients were computed from microcanonical simulations with an +average temperture of 300 K. In all of the phases, the head group +correlation functions decay with an fast librational contribution ($12 +\pm 1$ ps). There are additional moderate ($\tau^h_{\rm mid}$) and +slow $\tau^h_{\rm slow}$ contributions to orientational decay that +depend strongly on the phase exhibited by the lipids. The symmetric +ripple phase ($\sigma_h / d = 1.35$) appears to exhibit the slowest +molecular reorientation.} +\begin{tabular}{lcccc} +\hline +$\sigma_h / d$ & $\tau^h_{\rm mid} (ns)$ & $\tau^h_{\rm +slow} (\mu s)$ & $\tau^b (\mu s)$ & $D (\times 10^{-11} m^2 s^{-1})$ \\ +\hline +1.20 & $0.4$ & $9.6$ & $9.5$ & $0.43(1)$ \\ +1.28 & $2.0$ & $13.5$ & $3.0$ & $5.91(3)$ \\ +1.35 & $3.2$ & $4.0$ & $0.9$ & $3.42(1)$ \\ +1.41 & $0.3$ & $23.8$ & $6.9$ & $7.16(1)$ \\ +\end{tabular} +\label{mdtab:relaxation} +\end{center} +\end{minipage} +\end{table*} + +\section{Discussion} +\label{mdsec:discussion} + +Symmetric and asymmetric ripple phases have been observed to form in +our molecular dynamics simulations of a simple molecular-scale lipid +model. The lipid model consists of an dipolar head group and an +ellipsoidal tail. Within the limits of this model, an explanation for +generalized membrane curvature is a simple mismatch in the size of the +heads with the width of the molecular bodies. With heads +substantially larger than the bodies of the molecule, this curvature +should be convex nearly everywhere, a requirement which could be +resolved either with micellar or cylindrical phases. + +The persistence of a {\it bilayer} structure therefore requires either +strong attractive forces between the head groups or exclusionary +forces from the solvent phase. To have a persistent bilayer structure +with the added requirement of convex membrane curvature appears to +result in corrugated structures like the ones pictured in +Fig. \ref{mdfig:phaseCartoon}. In each of the sections of these +corrugated phases, the local curvature near a most of the head groups +is convex. These structures are held together by the extremely strong +and directional interactions between the head groups. + +The attractive forces holding the bilayer together could either be +non-directional (as in the work of Kranenburg and +Smit),\cite{Kranenburg2005} or directional (as we have utilized in +these simulations). The dipolar head groups are key for the +maintaining the bilayer structures exhibited by this particular model; +reducing the strength of the dipole has the tendency to make the +rippled phase disappear. The dipoles are likely to form attractive +head-to-tail configurations even in flat configurations, but the +temperatures are high enough that vortex defects become prevalent in +the flat phase. The flat phase we observed therefore appears to be +substantially above the Kosterlitz-Thouless transition temperature for +a planar system of dipoles with this set of parameters. For this +reason, it would be interesting to observe the thermal behavior of the +flat phase at substantially lower temperatures. + +One feature of this model is that an energetically favorable +orientational ordering of the dipoles can be achieved by forming +ripples. The corrugation of the surface breaks the symmetry of the +plane, making vortex defects somewhat more expensive, and stabilizing +the long range orientational ordering for the dipoles in the head +groups. Most of the rows of the head-to-tail dipoles are parallel to +each other and the system adopts a bulk anti-ferroelectric state. We +believe that this is the first time the organization of the head +groups in ripple phases has been addressed. + +Although the size-mismatch between the heads and molecular bodies +appears to be the primary driving force for surface convexity, the +persistence of the bilayer through the use of rippled structures is a +function of the strong, attractive interactions between the heads. +One important prediction we can make using the results from this +simple model is that if the dipole-dipole interaction is the leading +contributor to the head group attractions, the wave vectors for the +ripples should always be found {\it perpendicular} to the dipole +director axis. This echoes the prediction we made earlier for simple +elastic dipolar membranes, and may suggest experimental designs which +will test whether this is really the case in the phosphatidylcholine +$P_{\beta'}$ phases. The dipole director axis should also be easily +computable for the all-atom and coarse-grained simulations that have +been published in the literature.\cite{deVries05} + +Experimental verification of our predictions of dipolar orientation +correlating with the ripple direction would require knowing both the +local orientation of a rippled region of the membrane (available via +AFM studies of supported bilayers) as well as the local ordering of +the membrane dipoles. Obtaining information about the local +orientations of the membrane dipoles may be available from +fluorescence detected linear dichroism (LD). Benninger {\it et al.} +have recently used axially-specific chromophores +2-(4,4-difluoro-5,7-dimethyl-4-bora-3a,4a-diaza-s-indacene-3-pentanoyl)-1-hexadecanoyl-sn-glycero-3-phospocholine +($\beta$-BODIPY FL C5-HPC or BODIPY-PC) and 3,3' +dioctadecyloxacarbocyanine perchlorate (DiO) in their +fluorescence-detected linear dichroism (LD) studies of plasma +membranes of living cells.\cite{Benninger:2005qy} The DiO dye aligns +its transition moment perpendicular to the membrane normal, while the +BODIPY-PC transition dipole is parallel with the membrane normal. +Without a doubt, using fluorescence detection of linear dichroism in +concert with AFM surface scanning would be difficult experiments to +carry out. However, there is some hope of performing experiments to +either verify or falsify the predictions of our simulations. + +Although our model is simple, it exhibits some rich and unexpected +behaviors. It would clearly be a closer approximation to reality if +we allowed bending motions between the dipoles and the molecular +bodies, and if we replaced the rigid ellipsoids with ball-and-chain +tails. However, the advantages of this simple model (large system +sizes, 50 fs time steps) allow us to rapidly explore the phase diagram +for a wide range of parameters. Our explanation of this rippling +phenomenon will help us design more accurate molecular models for +corrugated membranes and experiments to test whether or not +dipole-dipole interactions exert an influence on membrane rippling.