Abstract
We report all-atom molecular dynamics simulations of asymmetric bilayers containing phosphoinositides in the presence of monovalent and divalent cations. We have characterized the molecular mechanism by which these divalent cations interact with phosphoinositides. Ca2+ desolvates more readily, consistent with single-molecule calculations, and forms a network of ionic-like bonds that serve as a ‘molecular glue’ that allows a single ion to coordinate with up to three phosphatidylinositol-(4,5)-bisphosphate (PI(4, 5)P2) lipids. The phosphatidylinositol-(3,5)-bisphosphate isomer shows no such effect and neither does PI(4, 5)P2 in the presence of Mg2+. The resulting network of Ca2+-mediated lipid-lipid bonds grows to span the entire simulation space and therefore has implications for the lateral distribution of phosophoinositides in the bilayer. We observe context-specific differences in lipid diffusion rates, lipid surface densities and bilayer structure. The molecular-scale delineation of ion-lipid arrangements reported here provides insight into similar nanocluster formation induced by peripheral proteins to regulate the formation of functional signalling complexes on the membrane.
Keywords: molecular dynamics, asymmetric bilayer, phosphoinositides
1. Introduction
The interface between the intracellular and extracellular environment, defined by the boundary of the cell’s plasma membrane, is the location of bidirectional outside-in and inside-out signalling responses. Such signalling pathways often involve a class of rare phospholipids known as polyphosphoinositides (PPIs) which can act as the starting point for second messengers and serve directly in transmitting signals. PPIs are among the most high-charge-density lipids on the cell membrane, their negative charge density originating through the deprotonation of two phosphomonoester groups that have pKa values in the physiological range of 6–8 and the negatively charged phosphodiester linker [1,2]. PPIs with two phosphate groups connected to the inositol ring—we focus on phosphatidylinositol-(4,5)-bisphosphate (PI(4, 5)P2) and phosphatidylinositol-(3,5)-bisphosphate (PI(3, 5)P2) in this study—can range from − 3e to − 5e depending on pH and the counterions present, whereas most mammalian lipids are neutral, zwitterionic, or carry a charge of − 1e, see figure 1 [3].
Figure 1.
Chemical structures of the membrane constituents simulated in this work. The net charge is noted under each structure.
Despite their low abundance in terms of the overall numbers, numerous cellular processes, including cytoskeletal assembly, are controlled by the level and spatial localization of PPIs on the cell membrane [4]. Among the PPIs, PI(4, 5)P2, at about 1% mole per cent of all phospholipids, [5] is the most well studied. PI(4, 5)P2 has been shown to act as a signalling beacon and a platform for microscale membrane clusters containing proteins, phospholipids, and cholesterol in vivo [6,7]. Hundreds of cytosolic proteins encompassing at least 10 different domains have been shown to bind PI(4, 5)P2 either through non-specific electrostatic interactions or specific coordination of the phosphomonoester groups [8]. Furthermore, PPIs bind many cytoskeletal proteins and proteins that cause or sense membrane curvature. The shape of the cell membrane is controlled by changes in membrane-cytoskeletal linkages and forces generated by actin polymerization or cytoskeletal motors, all of which are sensitive to PPIs and their spatial organization [9–11]. Aside from their interactions with proteins, PPIs have been implicated in a number of diseases [12–14].
The positioning of phosphate groups on the inositol ring appears to bestow distinct biological roles to PI(4, 5)P2 and PI(3, 5)P2, despite their similar chemical structures. The significant charge of these molecules makes them not only sensitive to charged amino acids on protein domains, [8] but also to positive ions present in biological contexts, such as the monovalent ions Na+ or K+ and the divalent ions Ca2+ and Mg2+ [15–17]. The detection and visualization of PPIs in vivo has thus far relied on indirect perception through labelled proteins or chemical modification of PPIs with bulky dyes.
Owing to these limitations, one of the unresolved issues in understanding how phosphoinositides regulate cytoskeletal assembly and membrane curvature is how these scarce lipids affect the activity of many proteins [18]. Measurement of PI(4, 5)P2 diffusion shows that most of the plasma membrane PI(4, 5)P2 pool is bound or sequestered to some extent [19]. A major question is how PI(4, 5)P2 distributes laterally within the plasma membrane [20,21]. Several recent studies show the relevance of nano-scale PI(4, 5)P2 clusters to critical cellular functions [7,22–24]. PI(4, 5)P2-dependent clustering of syntaxin 1 in the presynaptic plasma membrane is reduced when cholesterol is depleted [25], but PI(4, 5)P2 clustering is not reduced when cortical actin assembly is disrupted with inhibitors, suggesting that PI(4, 5)P2 clusters precede and do not result from actin nucleation. Strikingly, even in the absence of proteins, μM Ca2+ induces PI(4, 5)P2 clusters with a similar size distribution to those found in vivo [26]. In recent work, Wen et al. [27,28] reported the formation of multivalent cation-bridged PI(4, 5)P2 clusters (approx. 8 nm) by using fluorescence self-quenching and Förster resonance energy transfer assays. Those experiments showed that concentrations as low as 0.05 mol% of PI(4, 5)P2 cluster in the presence of divalent cations, with Ca2+ being more efficient than Mg2+ (i.e. clusters form at a lower ion concentration). However, the authors noted that they do not yet have a molecular model to explain the mechanism by which multivalent metal ions promote clustering [27].
In this work, we define how PPIs behave in model asymmetric membranes (i.e.multi-component lipid bilayers with a distinct inner leaflet), characterize their interactions with cations, and describe their propensity to initiate nanoscale clusters using all-atom molecular dynamics simulations. We note that owing to the limitations of large system sizes and computational cost, we do not aim to model cluster growth using atomistic models, but rather aim to delineate how the molecular context of the lipid-ion interactions can lead to conditions that initiate or seed clusters of phosphoinositides. Many modelling studies of biological membranes have used symmetric leaflet compositions and predominantly feature short-length acyl chains [10,11,29–31]. We modelled multiple membrane compositions to independently test several variables. Specifically, we varied the membrane composition, the presence or absence of cholesterol, phosphoinositide net charge and counterion species. We also used membranes with symmetric leaflets to compare with the asymmetric membranes. The effects of Na+ and K+ were contrasted with those of Ca2+ and Mg2+. We likewise compared the self-association and ion-binding properties of PI(3, 5)P2 and PI(4, 5)P2 isomers. We controlled for non-specific electrostatic attraction by running identical simulations with the net charge of PI(4, 5)P2 of − 3e, − 4e and − 5e. Simulation compositions are provided in the electronic supplementary material, table S1.
Our study proposes a model for context-specific cluster initiation observed in recent experimental studies [27]. Molecular dynamics simulations are capable of resolving the driving forces intrinsic to the biological function of PPIs which are typically beyond the scope of experimental techniques because they require combined knowledge of the lipid-lipid, lipid-ion and ion-water configurations. While our focus in this work is on ion-mediated PPI interactions, extension to protein binding and signalling are logical next steps [10,11]. The implications beyond the results stated are relevant to protein recruitment and subsequent control of cellular processes such as cytoskeletal rearrangements [11].
2. Methods
We employ atomistic molecular dynamics simulations in order to resolve lipid-lipid, lipid-ion, and ion-water binding in atomic detail. The physical properties of simulated membranes are sensitive to the force field parameters. We use the CHARMM36 (C36) lipid force that was optimized to correctly model the surface tension and area per lipid of choline- and ethanolamine-containing head groups with both saturated and unsaturated chains and also includes cholesterol [30,32,33]. Notably, parameters for PPIs exist within C36 and have been independently validated against other condensed phase measurements, such as lipid tilt, membrane electrostatic potential, and electron density [31,34]. We also use the NBFIX correction for ion interactions [35]. Among a comparison of lipid-specific force fields, C36 was found to most accurately describe lipid dynamics, with a slight bias towards lipid order relative to experiment, and overall good computational efficiency [36]. Although there can be overly favourable anion-cation pair interactions in many additive force fields, including C36 [37,38], a full reparameterization of divalent cations was not attempted here; nevertheless, we note that our results are qualitatively consistent with previous quantum mechanics/molecular mechanics calculations [39]. We used GROMACS 4.6.3 to simulate our model bilayers [40], with net charge and protonation states for PI(4, 5)P2 and PI(3, 5)P2 based on quantum calculations [41]. Longer (500 ns) simulations for two key systems, Ca2+ versus Mg2+, were performed using GROMACS 2018.2. We note, based on earlier works reported in the literature, that the timescale for ion binding and equilibration is 50–100 ns [42], while timescales for lipid mixing and rearrangements occur in the timescale of 500–1000 ns [43]. Of course, we cannot entirely rule out a scenario which could show a rearrangement with respect to ion binding at a longer time frame. However, we note that the results reported here reflect reproducibility of the observations across our replicates. Therefore, we expect our results to be valid for ion binding and equilibration and initiation of cluster formation but not fully capture growth of clusters.
We have investigated fifteen distinct compositions. Each bilayer has 400 molecules (i.e. lipids and cholesterol) in each leaflet. The physiologically composed bilayers have an asymmetric composition with 75% 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC) and 25% cholesterol in the outer leaflet and 50% 1,2-dioleoyl-sn-glycero-3-phosphoethanolamine (DOPE), 25% cholesterol, 15% 1,2-dioleoyl-sn-glycero-3-phospho-l-serine (DOPS), and 10% PI, PI(4, 5)P2, or PI(3, 5)P2 in the inner leaflet. Our control systems included symmetric bilayers containing a 4 : 1 mixture of 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) and DOPS along with 10% PI(4, 5)P2 or PI(3, 5)P2 in both leaflets. A comprehensive list of simulation compositions can be found in the electronic supplementary material, table S1. We note that in preparing the asymmetric leaflets, we chose to focus on the compositions and balanced the number of lipids on each leaflet based on area per head group and the total leaflet area. These effects are in part mediated by our use of semi-isotropic pressure coupling to produce an approximately tensionless bilayer. In general, it is quite difficult to ensure the differential stress build-up across both the leaflets in an asymmetric simulation. Recent work [44] has indicated that area asymmetries below 5% have negligible effects on bilayer properties (e.g. thickness, diffusion, and order parameters). Our area mismatch is well below this 5% threshold.
A randomized grid of 400 molecular structures was assembled for each leaflet and composition and arranged with a regular 1 nm spacing. The lipids were fixed with mild position restraints of 500 kJ/(mol · nm2) in the normal direction and then gently packed into a bilayer using a vacuum equilibration procedure in order to ensure that no lipids flipped to the opposite leaflet. A number of similar bilayer-construction tools are available [45], however, our procedure is closely aligned with the tools described in recent studies [46]. The resulting bilayer configurations were solvated with water and then one of four different cations (Na+, K+, Mg2+ or Ca2+) and Cl− ions at an ionic strength of 150 mM. As per prescription of Klauda et al. [30], we employed CHARMM ‘special’ water (TIPS3P) which includes Lennard–Jones interactions on water hydrogen atoms in order to prevent artefacts in lipid motion.
The asymmetric lipid compositions were chosen because they reflect the physiological composition of mammalian cell membranes [47,48]. As for soluble ion compositions, they are different inside and outside the cell, but they are much more similar than the lipid asymmetry in the bilayer. As for the physiological ion compositions, we note that the ion compositions of the combined monovalent ions (Na+ plus K+ combined) are roughly equal in the cytosol and the extracellular medium. The Mg2+ compositions are also similar between the cytosol and the extracellular medium. Only the Ca2+ concentrations are very different on either side of the cell mammalian membrane. But because the Ca2+ concentration is higher on the extracellular side where the concentrations of highly charged polyanionic lipids like PI(4, 5)P2 are negligible, the ion does not bind the lipids at this interface significantly. Therefore, the single bilayer set-up with periodic boundaries considered here will not influence the main conclusions. In particular, the electrical double layer is formed and stable adjacent to the inner leaflet. The total ionic strength is very similar, and the largest relevant difference is the much higher concentration of calcium ions outside the cell, an effect that would only strengthen our interpretation. The double bilayer set-up is too prohibitive on the basis of computational cost, but is a worthwhile exercise for future studies of the systems we have considered here. Our systems were equilibrated for at least 20 ns followed by production runs lasting 80–500 ns. These systems were maintained at a temperature of 310 K using the velocity-rescaling thermostat of Bussi et al. [49] with a coupling frequency of 0.5 ps− 1 and atmospheric pressure via the Parrinello–Rahman barostat with a coupling constant of 2.0 ps− 1. The LINCS algorithm constrained hydrogen bond distances allowing for a 2 fs timestep. Van der Waals forces were switched off smoothly from 0.8 to 1.2 nm, and electrostatics were computed according to the particle-mesh Ewald summation with a Fourier spacing of 0.16 nm.
We report a number of bilayer properties using a set of analysis codes written in Python which make use of several libraries based on Python’s NumPy [50], namely MDAnalysis [51] for reading and selecting molecules and SciPy [52] for measuring lipid-ion-water distances under periodic boundary conditions using a k-dimensional tree [53]. The distances and angles were also used to identify associations including hydrogen bonds and salt bridges. Diffusion coefficients are computed from linear fits of the lateral mean-squared displacement after accounting for periodic boundary conditions and our bilayers have no net drift artefacts.
We quantify the lipid-lipid lateral distribution using two-dimensional meshes of the lipid centres of mass. We identify the observed numbers of pairs (two adjacent lipids) and triplets (three adjacent lipids) from a Delaunay triangulation generated using the quickhull algorithm [54] within SciPy with modifications to generate triangulations across periodic boundary conditions. We used the same algorithm to compute a Voronoi tesselation that allowed us to partition the leaflet area among each lipid species. This procedure contributed to lipid area calculations and a description of lipid association propensities discussed in the Results section. These methods are available in open-source format (via https://biophyscode.github.io/). Unless specified otherwise in this manuscript, PIP2 refers to either PI(4, 5)P2 or PI(3, 5)P2.
All simulations were performed in replicates on several supercomputing platforms provided by XSEDE as well as on a local cluster. Simulations contained up to 360 000 atoms including water and counterions (see the electronic supplementary material, table S1) and achieved simulation speeds of up to 16 ns d−1 on 6 × 16-processor platforms. The simulations ran for an aggregate production time over 2 µs. Convergence was assessed by comparing the radial distribution functions along the trajectories of different lengths, see the electronic supplementary material, figures S1 and S2.
3. Results
A typical simulation snapshot is provided in figure 2. All such snapshots, including the electronic supplementary material videos, have been rendered with visual molecular dynamics (VMD) [55]. An example simulation is depicted in the electronic supplementary material, video SV1.
Figure 2.
Simulation snapshot of a bilayer containing 800 lipids with PIP2 (purple) found in the inner leaflet (top) along with DOPE (light blue), DOPS (pink) and cholesterol (dark grey). The outer leaflet (bottom) contains POPC (white) and cholesterol. Ca2+ cations are shown in green and chloride ions are grey spheres.
3.1. Ion-lipid head group coordination drives ion-specific clustering of PIP2
To test the central hypothesis that lipid-lipid, lipid-water and lipid-counterion interactions are essential determinants of the distribution of PPIs within a membrane, we first consider the formation of ion-lipid complexes. We classify the multivalent lipid-cation interactions by the number of lipids bound within a specific cutoff distance of each ion. We call these structures N-bridges where N is the number of bound lipids. We select a cutoff distance that is equal to the radius of the first solvation shell in order to ensure that we only identify ions that are bound tightly enough to exclude water. These cut-offs were identified from the first local minima of the water-ion radial distribution functions (see the electronic supplementary material, figure S1), and are set to 3 Å for Na+ and K+, 2.3 Å for Mg2+ and 2.6 Å for Ca2+. Figure 3 shows the counts of 1-bridges, 2-bridges and 3-bridges for the 500 ns simulations.
Figure 3.
Counts of closely associated ions distinguished by the number of lipids which are also associated with these ions. We define bound ions according to the radius of the first solvation shell, which is 3.0 Å for Na+ and K+, 2.3 Å for Mg2+, and 2.6 Å for Ca2+. Colours correspond to one, two or three closely associated lipids, and are stacked in that order, i.e. from bottom to top as marked on the right. Solid, coloured bars show the number of ions associated with clusters with at least one PIP2. In the case of the lightest colours, the bound ions are therefore bound to exactly one PIP2 while the darkest colour includes ions associated with one or more PIP2 along with other lipid species. The grey, hatched bars represent ions which are associated with other lipids, in this case DOPS or DOPE. We have excluded the outer leaflet lipids (POPC) in this calculation in order to highlight the inner leaflet. Bar colours correspond to the chemical identity of the ion (blue for Ca2+, and red for Mg2+). While Ca2+ has a larger number of bound ions, they are also increasing over the duration of the simulation. Note, Ptdlns denotes phosphatidylinositol.
The counts in figure 3 capture a striking result: namely that Ca2+ forms tight, nearly ionic bonds with lipids—that is, those capable of excluding water—at a much higher rate and with a higher charge density than other ion species. The high − 4e charge on PIP2 ensures that most of the ions in the simulation are drawn to the bilayer, forming an electric double layer. We refer to this phenomenon as charging [56]. Ca2+ prefers to bind to PIP2 more than any other lipid, as evidenced by the large solid bars, denoting ions bound to the 10% PIP2 in the inner leaflet, compared to the hatched, grey bars which show bonds to the remaining 90% of the lipids in the bilayer. In contrast to Mg2+, Na+ and K+ (see the electronic supplementary material, figure S3), we find that Ca2+ forms many multivalent complexes that include at least one PIP2. The ion binding results for additional systems, including those containing PI(3, 5)P2, are shown in the electronic supplementary material, figure S3.
The distinction between Ca2+ and Mg2+ is also evident in the electronic supplementary material, video SV2 which shows that the lifetime of Mg2+ binding to phosphate groups is shorter than Ca2+, which rarely dissociates from the phosphate groups. This key difference is reflected in the number of Ca2+ bound to at least three lipids (i.e. 3-bridges) including PIP2, which grows progressively over the course of the simulation.
These rearrangements accompany a mild additional net increase in total bound Ca2+ which is not seen in the other ions. This demonstrates that the bilayer is charging on the relatively slow, 500 ns timescale of the simulation, facilitated by a reorganization of the lipid configurations around the ions.
Mg2+ is not able to generate 2-bridges or 3-bridges to any significant degree, nor does it accumulate on the membrane surface. It is interesting to note that Na+ can also facilitate a smaller number of 3-bridges (see the electronic supplementary material, figure S3), however, these clusters rarely include PIP2 and do not change appreciably with time. This suggests that Na+ occupies the positions within pre-existing lipid configurations rather than affecting a reorganization that maximizes lipid contact with the ions. In membranes with identical leaflet compositions, we find that the Na+ condensation is the highest when PIP2 carries the highest charge (for example, when deprotonated compared to doubly protonated; see the electronic supplementary material, figure S3). This calculation shows almost no measurable differences between the PI(4, 5)P2 and PI(3, 5)P2 isoforms. Na+ binds PI less than PIP2, indicating that many of the bound cations occupy the exposed phosphate groups. In the presence of PIP2, K+ forms significantly fewer bonds than Na+.
3.2. Lipid-ion-lipid bridging drives the growth of nanoclusters
To characterize the implications for the lateral distributions at length scales beyond the size of our simulations, we define a cluster as a set of lipids which are bound by at least one N-bridge and study the distribution of clusters over time. We derive the clusters from the N-bridges by reducing each lipid-cation-lipid bond to a lipid-lipid bond which forms an edge in the graph for that cluster. Figure 4 provides diagrams of both the N-bridges and the resulting clusters for systems containing either Mg2+ or Ca2+. We derive the clusters by collapsing each lipid-cation-lipid bond which contributes to an N-bridge into a single edge in a graph of associated lipids. The size of each graph (Nlipids) measures the cluster size, while the distribution of these clusters (see figure 4, right panel) quantifies the extent to which the lipid motion is hindered by the ‘molecular glue’ formed by the cations to form a larger collective lipid mass. The largest clusters of up to 60 lipids are only observed for the combination of PI(4, 5)P2 and Ca2+. Simulations containing an alternate isomer, for example PI(3, 5)P2 or PI, or a different ion, notably Mg2+, fail to generate these large and growing clusters.
Figure 4.
Diagram of observed N-bridges (left) and the resulting bond network (centre) alongside a quantification of the resulting clusters (right) over the duration of each simulation. Instantaneous snapshots (left and centre) were taken at the conclusion (t = 100 ns) of two replicates. The left panels show an overhead snapshot of the Voronoi tesselation of lipids on the inner leaflet distinguished by colour (PIP2 is red; DOPS is green; DOPE is purple). Black dots correspond to bound cations, with black lines indicating the tile (i.e. lipid) to which they are bound. Only the cation-lipid bonds for a single periodic image are shown. In the schematic in the centre panels, we reduce the lipid-cation-lipid bonds each to a single lipid-lipid edge drawn between tiles and repeated for periodic images. The right panel depicts histograms of the resulting lipid clusters according to the number of lipids in each cluster. Each symbol represents a cluster of lipids. The symbol area is proportional to the number of lipids in the cluster, which is also measured on the vertical axis. Only the unique combination of PI(4, 5)P2 and Ca2+ shows large and growing clusters.
The following sections characterize the atomic basis for the lipid-cation bonds and derive the effects of these bond networks on the collective lipid dynamics by studying their diffusion rates and surface densities. In contrast to these local, nano-scale measures of the molecular structure of the bilayer, however, the network of strong bonds quantified here is large enough to span the simulation box and therefore has implications for the lipid distributions at larger length scales.
In a recently published article, Han et al. [57] characterized PIP2 clusters in a somewhat similar fashion, albeit in monolayers composed entirely of PIP2. Our results build on their recommendation to extend this investigation to bilayers and bilayers with other lipid species. The Han et al. analysis proceeds from a different definition of an ‘edge’ in the lipid-association graph, namely the proximity between phosphate oxygen groups. Our definition of an ‘edge’ is a lipid-cation-lipid bond, that is, a desolvated cation shared between two lipids. Both definitions are valid ways to describe the concept of a cluster, however, we are unlikely to observe the edges defined by Han et al. in our simulations because we have a much lower PIP2 concentration and direct PIP2 interactions are statistically much less likely because there are many other lipids. This is the primary reason for the alternate definition of an ‘edge’ within a cluster employed in this work (see also the Discussion section).
3.3. Ion hydration mediates ion-lipid interactions to modulate ion-specific lipid clustering
The main result of transient Mg2+ binding versus sustained Ca2+ binding in figure 3 is primarily governed by the level of ion hydration. This is apparent in the electronic supplementary material, video SV2, where a single PIP2 in a bilayer is selectively depicted to highlight the interplay between the lipid, ions and water. The persistent binding of Ca2+ is consistent with Ca2+ interacting much more strongly with PIP2 by rearranging its hydration shell, while the short-lived binding of Mg2+ is consistent with the almost full hydration shell around the ion as it interacts with PIP2, thereby leading to a greater separation between the ion and PIP2. Before we quantify these observations to demonstrate that this qualitative picture is borne out, we note that the observations of a stronger dehydrated Ca2+ − PIP2 bond and a weaker water-mediated (i.e. non-dehydrated) Mg2+ − PIP2 bond agrees with single-molecule free energy calculations reported previously [41].
To examine the differences in ion-lipid-water complexes, we quantify the number of waters in the first hydration shell of these ions as they approach the lipids. We count the waters in the solvation shell according to equation (3.1):
3.1 |
In this equation, the indices c, l, w refer to cations, lipids, and water molecules, respectively; ε = 0.05 Å sets the observation windows for the cation-lipid distances; represents the minimum distance between any two atoms m, n in a pair i, j of molecules; t is time; and ds refers to the radius of the first solvation shell. We discretize minimum distance between atoms (rij) into bins of size 0.05 Å. Note that the bracket operators yield Boolean values used to count each cation-lipid and cation-water distance in the correct distance bin. Equation (3.1) thereby quantifies the cations bound to lipids (at a distance dc,l) and the waters bound to the cations (at a distance dc,w); see snapshots in figure 5.
Figure 5.
Snapshots of lipid-cation-lipid bonds. Each panel shows an instantaneous diagram of two PIP2 (black), their associated cations (Mg2+ in red and Ca2+ in blue), along with associated water molecules in the first solvation shell (red for oxygen and white for hydrogen, rendered as a tube). Hydrogen bonds between lipids are depicted with a purple tube. We find that bonds between PIP2 and Mg2+ (a–c) show more water in the solvation shell than bonds between PIP2 and Ca2+ (d–f).
Figure 6 depicts the result, namely the distribution for Nwaters(r,t) across a range of cation-lipid distances along with the number of ions (Nions) found at these distances. Although Mg2+ forms a bond with PIP2 at a closer distance than Ca2+, there are many more Ca2+ ions bound between 1.75− 2 Å. Ca2+ density in this zone is more than fourfold as high as Mg2+, a result which is also reflected in the charging curves found in figure 3. Mg2+ has a nearly equal preference for making either an ionic bond (approx. 2 Å) or occupying the region between 3 and 5 Å.
Figure 6.
Number of bound ions Nions (top) as a function of minimum lipid-ion distance (dc,l) along with the corresponding number of water molecules bound to each ion Nwaters (bottom). In the bottom panel, the distribution of the number of waters at each cation-lipid distance is reflected in the colour intensity in each (integer) bin on the vertical axis, while a black dot indicates the mean number of hydrated waters at that distance. The colour intensity is normalized for each bin, so even if there are comparatively few observations of ions Nions at that distance (for example, between 2.5 and 3.3 Å), the preferred number of waters in the first shell is still shown. This scheme highlights some rare events in which Ca2+ acquires up to seven waters at an intermediate distance (2.8 Å) from the lipids. Both Mg2+ (red) and Ca2+ are plotted together for comparison.
We find significant differences in hydration between Ca2+ and Mg2+. First, Ca2+ more easily loses water, retaining only 2–4 waters at 2 Å. The strength of the Ca2+ − lipid bond leaves comparatively few ions distant from any lipids, where they are coordinated by 4–6 waters, compared to 5–7 waters for Mg2+. These findings suggest that both ions lose 1–2 waters upon forming an ionic bond with phosphate oxygen atoms, beginning around 3 Å from the lipid. We find almost no difference between simulations containing PI(4, 5)P2 and PI(3, 5)P2 in the histogram of ion counts and the hydration of each ion species (see the contrast in the electronic supplementary material, figure S4).
Snapshots of the lipid-ion-water configurations for Mg2+ and Ca2+ are shown in figure 5 and call attention to the less-structured and less-hydrated nature of Ca2+ − PI(4, 5)P2 binding. One can see that Mg2+ ions are almost always buffered by water molecules while Ca2+ bonds directly with the lipids.
The ion-water radial distribution functions (see the electronic supplementary material, figure S1) show that water adopts a different structure around Mg2+ versus Ca2+ at both near (less than or equal to 2.2 Å) and far (greater than or equal to 4.6 Å) distances from the lipids. Waters in the first and second solvation shell are closer to Mg2+ than Ca2+ in good agreement with an earlier report [58]. In contrast to Mg2+, when Ca2+ is close to lipids (electronic supplementary material, figure S1, bold blue line), the density of water in its second hydration shell drops 25% relative to when it is far (light blue line).
The fact that Ca2+ forms tighter bonds with PIP2, and has fewer waters in its first and second solvation shells when it does so, compared to Mg2+, suggests that lipid-lipid interactions are context-specific, modulated by the ion species along with the particular structure of its hydration shell.
3.4. Inter-lipid hydrogen bonding
The charging curves in main text figure 3 quantify the number of closely bound cations, the hydration of those ions is characterized in figure 6 and the electronic supplementary material, figure S4, and their lipid-binding sites are characterized in figure 7. In this section, we analyse the effect of ion binding on the native hydrogen-bonding network formed between the lipids, irrespective of the mediating cations. These lipid-lipid hydrogen bonds represent an alternate mechanism by which lipids self-associate which could either supplement or compete with bonds mediated by cations. Electronic supplementary material, figure S5 depicts lipids directly engaged in hydrogen bonds along with the flanking ions. Hydrogen bonds are defined by a donor-acceptor distance of 3.4 Å with a donor-hydrogen-acceptor angle greater than 160°.
Figure 7.
Salt bridges between two PIP2 or (PI, upper left) for selected simulations. Each heat map shows the number of lipid-cation-lipid salt bridges (between heavy atoms within 3.4 Å) for each pair of valid atoms. Each axis uses the same list of atom names; atoms are greyed out if they are not members of that molecule. The heat maps are normalized to show the number of bridges per lipid. Each tile is normalized to a different maximum number of bridges indicated by the colour bar. Counts are provided in the electronic supplementary material: of atom-specific hydrogen bonds (electronic supplementary material, figure S8) and salt bridges (electronic supplementary material, figure S9). Atom names follow the CHARMM36 convention in which the OP4 prefix indicates the 4-phosphate oxygens. Our PIP2 are protonated at the OP52 position where there is a corresponding absence of bridges (lower intensity above). The prefix OP1 indicates the diester phosphate.
In the electronic supplementary material, figure S5, it is clear that PIP2 readily bonds with DOPE, but otherwise has mostly opportunistic bonding with other partners, including cholesterol. Overall, however, the hydrogen-bonding network is not strongly influenced by the identity of the cation. The first conclusion to draw is that the strong network of 2- and 3-bridges which we have described in the preceding sections does not compete with or interrupt the formation of hydrogen bonds. We find that PIP2 forms one hydrogen bond on average 100% of the time with DOPE, compared to roughly 30% on average with DOPS. Bonds between two PIP2 or bonds between a PIP2 and cholesterol are comparatively rare, representing 10% of the duration of the simulation time. Hence salt bridges, and not hydrogen bonds, appear to dominate interactions between PIP2, emphasizing the importance of their headgroups. This is further underscored by the comparison between PIP2 and PI; the latter forms a hydrogen bond only 60% of the time with DOPE.
Electronic supplementary material, figures S6 and S7 show a complete tabulation of normalized salt bridges and hydrogen bonds. It is important to note that the salt bridge calculation captures the N-bridges described in the main text because it includes mediating cations, while the hydrogen-bonding patterns are measured between lipids, independent of the cations. The electronic supplementary material, figures S8 and S9 describe the lipid atoms which participate in these bonds.
3.5. Which atoms facilitate lipid-ion-lipid bridging?
The quantification of ion hydration shells in the previous section shows that each ion interacts differently with water, and that Ca2+ is more likely to have waters in its first hydration shell replaced by phosphate oxygens when it is close to the lipids. In this section, we identify the specific parts of each lipid that form a lipid-cation-lipid salt bridge in order to isolate the isomer- and ion-specific differences in these interactions. Figure 7 shows a heat map of the atoms that participate in these salt bridges. Note that salt bridges are mediated by cations and hence form the basis for the N-bridges.
We find that K+ salt bridging occurs almost exclusively on the 4 −phosphate group (e.g. atoms OP42, OP43 and OP44 in figure 7, named according to the C36 [30] convention). PI forms comparatively few salt bridges with other PI, and the same is true for all lipids when Mg2+ is present (see the electronic supplementary material, figure S13). Interestingly, PI(4, 5)P2 is more likely to form a Ca2+-mediated salt bridge between the diester phosphate on one lipid and the 4– and 5-phosphate groups on a different lipid, forming an asymmetric bond between lipids. By contrast, PI(3, 5)P2 forms bonds between the 3- and 5-phosphate groups of different lipids, and also has slightly more overall salt bridges than Ca2+ and PI(4, 5)P2.
3.6. Effect of ion-lipid molecular association on macroscopic properties
Thus far, we have provided evidence that the strength and multivalency of Ca2+ − PIP2 bonds are qualitatively different from PIP2 bonds formed with other ions, even with the same charge, namely Mg2+. These bond networks form clusters which can span the simulation box. Here, we consider the possible implications of this large network on the behaviour of many lipids acting in concert. The features of the lipid-ion associations have a measurable effect on the collective material properties of the bilayer including the lipid diffusion rates and the lateral area of the leaflets containing PIP2 molecules. While these properties are averaged over measurements of individual lipids, they are emblems of the larger structure of the bilayer.
3.6.1. Diffusion coefficients
The strong lipid-lipid association owing to Ca2+ creates a larger effective mass for diffusing lipids connected by ions. In particular, we hypothesize that these strong bonds, and specifically 3-bridges, make the effective diffusive unit larger in proportion to the number of additional lipids that are associated with the PIP2 molecule.
Figure 8 shows the lipid diffusion rates for the asymmetric bilayers. We find that Ca2+ slows the diffusion of all the lipids, in qualitative agreement with experiments [59]. PIP2 diffuses the slowest of all lipids in each simulation, while PI diffuses about as fast as DOPS. Differences in lipid diffusion appear to be more sensitive to the ion species than the presence of cholesterol (see the electronic supplementary material, figure S10 for diffusion rates in the cholesterol-free, symmetric bilayers). We note that diffusion rates in finite simulations are subject to a number of measurement errors, most notably owing to the periodic nature of the simulation [60]. The diffusion rates reported here may be underestimating the true diffusion rates observed in a much larger simulation.
Figure 8.
Diffusion rates according to lipid and cation identity. Lipid diffusion rates for PPI in our simulations. Error bars describe the standard deviation of the diffusion constant over the 40 PIP2 molecules in each simulation. Simulations with K+ and Na+ have similar diffusion whereas Ca2+ appears to reduce the diffusion rate slightly. See the electronic supplementary material, figure S10 for diffusion rates for the remaining simulations.
3.6.2. Bilayer area
The area per lipid is an important biophysical quantity because it encodes information about both the density and rigidity of lipid bilayers. It can be measured by experiments [30,61] and is one of the crucial quantities of comparison between molecular models and experiments. We have computed the lipid areas in two ways: first, using the three-dimensional (3D), rugged surface of each leaflet and secondly, by projecting the lipid centres of mass onto a two-dimensional (2D) surface, which controls for the excess out-of-plane area contribution of the PIP2 headgroup. The 3D areas are depicted in figure 9 (see the electronic supplementary material, figure S11 for the 2D areas).
Figure 9.
3D leaflet areas for asymmetric bilayers. Leaflet areas are computed from the Delaunay mesh between lipid centres of mass under periodic boundary conditions. The top and bottom of each bar indicates the mean inner and outer leaflet area, respectively. We find that the PI simulation has statistically similar leaflet areas, while the PIP2 simulations, which contain the same total number of lipids and cholesterol molecules, have an additional inner leaflet area.
The projected lipid areas provide evidence that Ca2+ condenses the total 2D area of the bilayer (see the electronic supplementary material, figure S11). We observe a reduction of an average of ∼1.5 Å2 per lipid in the simulations with Ca2+ compared to the other conditions. Having constructed these bilayers with equal numbers of lipids in both leaflets, we find that the inner leaflet carries a larger 3D area across all systems which contain PIP2 (figure 9, right panel), although the differential between leaflets varies slightly by cation. This extra inner-leaflet area is notably absent when PI is present, suggesting that the extra area is created by the additional phosphate groups carried by PIP2. Even when these phosphate groups are absent, as is the case for PI, the Ca2+ causes the bilayer to contract compared to the other conditions. The additional leaflet area produced by PPI species is underscored by the area distributions across individual lipids shown in the electronic supplementary material, figure S12. We also find that the area compressibility is somewhat elevated when Ca2+ is present (see the electronic supplementary material, figure S13). The simultaneous 2D contraction along with extra 3D area suggests that PPIs under these conditions have a larger intrinsic curvature, which may affect their lateral distribution, and that their situation in the bilayer changes such that they ‘bulge’ more into solvent, providing greater access to the cations. We note that the observed effects on bilayer area condensation are consistent with experimental observations of area reduction in monolayer experiments reported by Levental et al. [62]. However, the magnitudes of such effects observed in our simulations (0–5% reduction in area) are smaller than those in the monolayer experiments (0–40%). While we cannot rule out force-field artefacts as cause for this deviation, it is reasonable that the deviations are attributable to monolayer versus bilayer and differences in lateral pressures between the simulations and the experiments.
3.7. Ion-lipid context mediates lipid association propensities
The simulations presented in this work span a timescale that is long enough to capture the motion of cations and corresponding changes in lipid structure which generate N-bridges and the bilayer-spanning bond networks depicted in figure 4. While this timescale is too short to observe long-range lipid diffusion, we can infer lipid association preferences by comparing the spatial arrangement of lipids between conditions.
We quantify the lipid-lipid association preferences by performing a Delaunay triangulation of the lipid centres of mass, including the effects of the periodic boundary. The diagrams in figure 4 illustrate the calculation by drawing the (Voronoi) mesh, in which each line separating lipid tiles represents an association or adjacency between two lipids and an edge in the corresponding Delaunay triangulation. To characterize the lipid-lipid association preferences, we counted the number of lipid-lipid associations of each type, for example an edge between PIP2 − DOPS, and compared these to the expected number for a random triangulation of a bilayer with an identical composition. The result provides a score relative to 1.0, which captures the extent to which that lipid association is favoured above (greater than 1.0) or below (less than 1.0) chance. We have also applied this calculation to groups of three lipids which we call a triplet, represented by a triangle on the Delaunay mesh.
We hypothesize that conditions which produce lipid pairs and triplets above chance will be more likely to create large and growing clusters in biological systems. The results are pictured in the electronic supplementary material, figure S14. We find that pure PPI triplets (for example, three adjacent PI(4, 5)P2) are disfavoured relatively to chance, suggesting that their bulky headgroups inhibit direct adjacency. A comparison of the two 500 ns trajectories containing divalent cations shows that the presence of Ca2+ correlates with heightened PIP2 pairs, but not triplets, compared to Mg2+, which occur below chance. Observations of adjacent Ca2+-mediated PIP2 pairs both above chance and in greater numbers than those observed with Ca2+ suggests that these structures are more energetically favourable.
4. Discussion
This study considers the chemical and physical characteristics of asymmetric membranes containing the rare, yet biologically crucial, phosphoinositide class of lipids. Although cation-induced cluster formation has been observed in vitro and in vivo [7,18–28], the physical basis and mechanism of clustering has remained unclear on an atomic level. What are the molecular driving forces and principles of clustering and how are they affected by the chemical context when clusters are forming? The answer to these questions can suggest mechanisms for determining the size and prevalence of plasma membrane clusters rich in PIP2, how they are regulated by, and how they regulate cell signalling on the membrane interface. We used a set of 15 simulations to independently vary the ionic conditions, net charge of the membrane, cholesterol content, leaflet composition asymmetry, and number and position of the phosphate groups on the inositol ring (PI, PI(3, 5)P2, or PI(4, 5)P2). We hypothesized that the close association driven by electrostatics and hydration status of divalent cations—particularly Ca2+—create a ‘molecular glue’ capable of driving the self-association of PPIs.
Over the course of the simulations, we observe condensation of monovalent and divalent counterions on the negatively charged surface of the asymmetric membranes (figure 3). This effect has been observed experimentally with PI(4, 5)P2-containing micelles [63], supported monolayers, [62,64] liposomes [27], large unilamellar vesicles and giant unilamellar vesicles [59], and cells [7]. Coarse-grained simulations of monolayers containing PI(4, 5)P2 [3] and atomistic simulations of DOPC/DOPS membranes [65] have also shown Ca2+ to lead to clustering. These results implicate the identity of the counterion in the formation of nano-clusters. Although Ca2+ and Mg2+ carry the same charge, they evoke distinctly different effects on charged bilayers. Hence, many of these differences must be attributed to their size and hydration free energy, which in turn generate different bonding patterns.
4.1. Charging
We find Ca2+ is most strongly associated with the membrane, followed by Na+, K+, and then Mg2+ (see the electronic supplementary material, figure S13). Ca2+ in particular, is able to simultaneously bind to one, two, or three lipids, and of those connections, roughly half are interactions with PIP2, whereas Mg2+ predominantly binds just one lipid at a time, with PIP2 in the minority. Monovalent ions fail to generate comparable charge density at close distances. K+ forms far fewer bonds than Na+, and both are typically limited to binding a single lipid. Ca2+ binds PI less than PIP2, and forms fewer 3-bridges.
The charge brought to the membrane surface by Ca2 + is +400e, representing significant overcharging above the net − 300e charge carried by the lipids. This overcharging leads to the formation of a double layer containing the co-ion Cl−. It is notable that the Cl− concentration in the double layer decays much more gradually when Mg2+ is present compared to Ca2+ (results not pictured here). When the divalent cations bind to the phosphate groups of PPIs, Mg2+ binds closer to PPI phosphate groups than Ca2+ but much more Ca2+ is present (figure 6). Simulations containing cholesterol have approximately 20% more bound Ca2+, however, these simulations also have an asymmetric composition in this study.
This charging effect, in tandem with the formation of cation-mediated 3-bridges, represents the driving force for much of the resulting redistribution of the lipids in this bilayer. The slow accumulation of Ca2+-driven 3-bridges, at the expense of 1- and 2-bridges, suggests the bilayer continues to undergo rearrangements as these nano-clusters develop. Coarse-grained models which account for both the condensed (Stern) and diffuse layers of cations at the lipid surface indicate that there is a cooperative effect in which cation binding is more favourable when lipids are clustered rather than dispersed [66]. These models suggest that the heightened N-bridges and resulting growth of lipid-lipid bond networks may proceed via feedback in which cations bind and enrich PIP2, which attracts further cations as it forms larger or denser clusters.
4.2. Hydration
The strength of ion-lipid interactions is determined by the relationship between hydration and coordination of the ion in bulk coupled with the general geometry constraints from the membrane and the specific partial desolvation of the ion by PIP2. Because Ca2+ and Mg2+ have different preferred coordination in bulk (with Mg2+ ranging 6− 8 and Ca2+ preferring octahedral 6 coordination) and corresponding hydration free energies, we expect this to influence their ability to incorporate a PIP2 phosphate group in their first solvation shell. We find Mg2+ retains more water, with more ordering in its first hydration shell (see the electronic supplementary material, figure S1). The high hydration free energy of Mg2+ underlies its short-lived binding (see the electronic supplementary material, video SV2).
4.3. Ions drive rearrangements
Despite condensing on the phosphate groups of PIP2, divalent cations do not significantly disrupt endogenous hydrogen bonds between lipids. It is not surprising that Ca2+ directly mediates a very large number of intermolecular salt bridges between headgroup atoms which combine to form N-bridges. The absence of large compensatory changes to lipid-lipid hydrogen bonds is evidence that Ca2+ acts as a ‘molecular glue’ by forming O − Ca2+ − O salt bridges at the level of the lipid headgroups. We attribute the differences in cluster formation, diffusion, and lipid area across conditions to ion-mediated lipid interactions. That the cation binding effect does not interfere with lipid-hydrogen bonding implies that these bonds are available for manipulation by other components, for example lysine-containing peptides, which may use them to stabilize further clustering when they are present [67].
4.4. Lipid-lipid interactions
Intermolecular bonds do not convey the full scope of the attraction and repulsion between two lipids. Understanding those forces is necessary to predict the lateral distribution of lipids. One possible proxy for those attractive forces is the statistical likelihood that two lipids will occupy adjacent spaces in the bilayer. These likelihoods may reflect lipid-lipid association preferences if they are well-sampled either by lengthy simulations or sampling many lipids. We have studied these associations via Delaunay triangulation described in the Results section and identified some notable patterns in the results.
If we restrict our attention to the longest-running simulations, we find that almost all pair association probabilities show decreased preferences for adjacent PIP2. This is consistent with steric repulsion owing to the bulky headgroup (see the electronic supplementary material, figure S14). By contrast, DOPE has a smaller headgroup and shows heightened associations with PIP2, and extra opportunities for hydrogen bonding, reflected in the electronic supplementary material, figure S6 may also facilitate these interactions. While these pair association probabilities have an intuitive explanation, we surprisingly find that the specific combination of PI(4, 5)P2 and Ca2+ shows PIP2 associations above simulations with Mg2+. These probabilities also exceed the random probabilities of observing adjacent PIP2. To demonstrate a causal relation between the N-bridges or lipid-lipid network and the propensity for PIP2 to form enriched clusters, these heightened associations must be confirmed with free energy calculations or extremely long trajectories. They nevertheless provide a plausible connection to experiments.
Analysis of these likelihoods does not include additional interactions that could drive clustering. Snapshots in the electronic supplementary material, figure S5 indicate that Ca2+ ions can enable PIP2 to ‘reach over’ nearby lipids to form associations with next-nearest neighbours. That this binding is less prevalent in simulations with Mg2+ suggests yet another specific feature of Ca2+ binding that might influence the lateral distribution of the PIP2. A more complete characterization of PIP2 would include these next-nearest-neighbour interactions, explain the heightened association between otherwise bulky lipids, and account for the timescale by which these clusters form in vivo. We note that while cholesterol was included to mimic the physiological composition, the systematic effect of cholesterol was not studied in this work; we refer to a recent related study that investigates the effect of cholesterol on protein binding to PIP2-containing bilayers [68].
4.5. Clusters
The implications of PIP2 cluster growth depend on the precise conditions under which they form. Han et al. [57] have studied the structure of cation-stabilized cluster formation in a slightly different context, namely monolayers composed exclusively of PIP2. They report a similar increase in area compressibility in the presence of Ca2+ and find mixtures of ions may further stabilize clusters. They characterize clusters by direct lipid-lipid bonds, providing a more natural definition for a PIP2-rich microenvironment and revealing linear, string-like clusters. Our results reported here for ion-mediated PIP2 clusters are in qualitative agreement with Han et al. despite the difference in lipid microenvironment (their monolayers versus our bilayers, our inclusion of other lipids, and our significantly lower PIP2 concentration). In either interpretation, the nanoclusters are viewed as assemblies generated by connected graphs (i.e. lipids act as nodes linked together either directly or through ions), rather than 2D domains associated with microphase separation.
4.6. Mechanical/structural properties
In order to relate highly-local molecular rearrangements to larger-length scale changes in the structure and dynamics of the bilayer, we computed representative macroscopic properties, namely diffusion coefficients and areas per lipid. Ca2+ slows the diffusion of all lipids in our simulations (see the electronic supplementary material, figure S10) despite the fact that it coordinates the most with PIP2 which represent only 10% of the inner leaflet. The area per lipid, however, is a more illuminating descriptor of the lipid packing. It reflects the denominator of the local charge density at a fixed lipid composition and also serves as an important order parameter for describing phase changes in lipid bilayers because it couples to the bilayer thickness and surface pressure.
In the presence of Ca2+, we find a small but significant contraction of roughly 3% in both the projected bilayer areas and in the leaflet areas measured on a 3D mesh. In the small simulations produced for this study, this area contraction is strong enough to condense the entire bilayer even when the PIP2 only populates the inner leaflet. In larger bilayers, this contraction might be less frustrated by periodic boundaries, and hence more likely to create spontaneous curvature. The area contraction owing to Ca2+ is roughly equal and opposite to the area asymmetry generated by PIP2 with phosphorylation at positions 4, 5 and 3, 5 compared to PI, which lacks these extra groups.
These results, taken together with the observation that PIP2 has the slowest diffusion coefficient of all the membrane constituents we tested, even more so in the presence of Ca2+, and that Ca2+ can bridge multiple PIP2 molecules, provide a potential mechanism to explain experimental Ca2+-induced PIP2 clusters. Specifically, enrichment of PIP2 in a particular region that grows from 2- and 3-bridges to the nanometre scale, can lead to an area and fluidity mismatch between the two leaflets that is dissipated through the formation of curvature. This is consistent with bulges found on supported lipid monolayers containing PI(4, 5)P2 in the presence of micromolar Ca2+ imaged with atomic force microscopy [59] and shape instabilities on giant unilamellar vesicles containing PI(4, 5)P2 exposed to excess Ca2+ on the lipid exterior [69].
It is notable that the additional area owing to poly-phosphorylated inositol rings is consistent across ionic conditions in our simulations. This means that it is the solvent-facing phosphate groups, and not the Ca2+, that provide the excess leaflet area. Hence any bulging area owing to Ca2+ in experiments could be owing to its ability to locally enrich PPIs to a higher concentration compared to other cations. Second, our simulations might likewise predict that Ca2+ bound to a leaflet containing a negative charge can generate an area contraction that creates invaginations, as long as the pre-existing PIP2 concentration is high enough. We estimate that at 10% PIP2, an area contraction of 3% across leaflets with centres of mass separated by 3 nm would correspond to the spontaneous curvature of a 235 nm vesicle (positive if the area contraction is on the outside). The modulation of lipid area by Ca2+ might contribute to PIP2 phase separation directly, or facilitate variation in spontaneous curvature, which causes lipid sorting. In either case, the specific features of headgroup-ion interactions identified in these simulations can plausibly play a role in sequestering PIP2 into highly enriched microdomains.
5. Conclusion
In conclusion, we have used advances in molecular dynamics to investigate the molecular mechanisms driving lateral organization of phosphoinositides in asymmetric bilayers. A wealth of experimental data is available showing interactions of phosphoinositdes with cytoskeletal-binding proteins, and a natural extension of this work will be to elucidate how the lipid bilayer couples to the protein framework immediately beneath it in the cell membrane. Together, the molecular interactions at the membrane-cytoskeletal interface can produce an integrated model for how these fascinating lipids perform their many cellular functions.
Supplementary Material
Supplementary Material
Supplementary Material
Data accessibility
The protocols are available as part of the electronic supplementary material. The codes and scripts along with complete documentation are available as part of the biophyscode software developed by the authors. The distribution of biophyscode is through an open-source license through GitHub and can be accessed using the link: https://biophyscode.github.io.
Authors' contributions
R.P.B. and D.S. performed the molecular simulations, R.B.P., D.S., P.A.G., and R.R., performed all the analysis. All authors contributed to the discussion and the writing of the manuscript. R.P.B. and D.S. contributed to the work equally.
Competing interests
There are no conflicts to declare.
Funding
This work was supported in part by grant nos. NIH GM111942 and GM136259 (P.A.G. and R.R.), NIH CA193417 (R.R.) and NSF DMR1120901 (R.R.). Computational resources were provided in part by the grant no. MCB060006 from XSEDE, by Penn MRSEC, and through Penn Institute for Computational Science (PICS).
References
- 1.Kooijman EE, King KE, Gangoda M, Gericke A. 2009. Ionization properties of phosphatidylinositol polyphosphates in mixed model membranes. Biochemistry 48, 9360–9371. ( 10.1021/bi9008616) [DOI] [PubMed] [Google Scholar]
- 2.Loew S, Kooijman EE, May S. 2013. Increased pH-sensitivity of protein binding to lipid membranes through the electrostatic-hydrogen bond switch. Chem. Phys. Lipids 169, 9–18. ( 10.1016/j.chemphyslip.2013.01.008) [DOI] [PubMed] [Google Scholar]
- 3.Ellenbroek WG, Wang Y-H, Christian DA, Discher DE, Janmey PA, Liu AJ. 2011. Divalent cation-dependent formation of electrostatic PIP2 clusters in lipid monolayers. Biophys. J. 101, 2178–2184. ( 10.1016/j.bpj.2011.09.039) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Gamper N, Shapiro MS. 2007. Target-specific PIP2 signalling: how might it work? J. Physiol. 582, 967–975. ( 10.1113/jphysiol.2007.132787) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Cockcroft S. 2000. Biology of phosphoinositides. Oxford, UK: Oxford University Press. [Google Scholar]
- 6.McLaughlin S, Wang J, Gambhir A, Murray D. 2002. PIP2 and proteins: interactions, organization, and information flow. Annu. Rev. Biophys. Biomol. Struct. 31, 151–175. ( 10.1146/annurev.biophys.31.082901.134259) [DOI] [PubMed] [Google Scholar]
- 7.van den Bogaart G. et al. 2011. Membrane protein sequestering by ionic protein–lipid interactions. Nature 479, 552–555. ( 10.1038/nature10545) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Lemmon MA. 2008. Membrane recognition by phospholipid-binding domains. Nat. Rev. Mol. Cell Biol. 9, 99–111. ( 10.1038/nrm2328) [DOI] [PubMed] [Google Scholar]
- 9.Kapus A, Janmey P. 2013. Plasma membrane-cortical cytoskeleton interactions: a cell biology approach with biophysical considerations. Hoboken, NJ: John Wiley & Sons, Inc. [DOI] [PubMed] [Google Scholar]
- 10.Slochower DR, Wang Y-H, Radhakrishnan R, Janmey PA. 2019. Handbook of lipid membranes, molecular and materials aspects. Boca Raton, FL: CRC Press. [Google Scholar]
- 11.Janmey PA, Bucki R, Radhakrishnan R. 2018. Regulation of actin assembly by PI(4,5)P2 and other inositol phospholipids: an update on possible mechanisms. Biochem. Biophys. Res. Trans. 506, 307–314. ( 10.1016/j.bbrc.2018.07.155) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Volpicelli-Daley L, Camilli PD. 2007. Phosphoinositides’ link to neurodegeneration. Nat. Med. 13, 784–786. ( 10.1038/nm0707-784) [DOI] [PubMed] [Google Scholar]
- 13.Vicinanza M, D’Angelo G, Di Campli A, De Matteis MA. 2008. Function and dysfunction of the PI system in membrane trafficking. EMBO J. 27, 2457–2470. ( 10.1038/emboj.2008.169) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Vicinanza M, D’Angelo G, Di Campli A, De Matteis MA. 2008. Membrane traffic in the secretory pathway. Cell. Mol. Life Sci. 65, 2833–2841. ( 10.1007/s00018-008-8353-2) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Balla T. 2013. Phosphoinositides: tiny lipids with giant impact on cell regulation. Physiol. Rev. 93, 1019–1137. ( 10.1152/physrev.00028.2012) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Schink KO, Tan K-W, Stenmark H. 2016. Phosphoinositides in control of membrane dynamics. Annu. Rev. Cell Dev. Biol. 32, 143–171. ( 10.1146/annurev-cellbio-111315-125349) [DOI] [PubMed] [Google Scholar]
- 17.Brown DA. 2015. PIP2 clustering: from model membranes to cells. Chem. Phys. Lipids 192, 33–40. ( 10.1016/j.chemphyslip.2015.07.021) [DOI] [PubMed] [Google Scholar]
- 18.Catimel B, Schieber C, Condron M, Patsiouras H, Connolly L, Catimel J, Nice EC, Burgess AW, Holmes AB. 2008. The PI(3,5)P2 and PI(4,5)P2 interactomes. J. Proteome Res. 7, 5295–5313. ( 10.1021/pr800540h) [DOI] [PubMed] [Google Scholar]
- 19.Golebiewska U, Nyako M, Woturski W, Zaitseva I, McLaughlin S. 2008. Diffusion coefficient of fluorescent phosphatidylinositol 4,5-bisphosphate in the plasma membrane of cells. Mol. Biol. Cell 19, 1663–1669. ( 10.1091/mbc.e07-12-1208) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Slochower DR, Wang Y-H, Tourdot RW, Radhakrishnan R, Janmey PA. 2014. Counterion-mediated pattern formation in membranes containing anionic lipids. Adv. Colloid Interface Sci. 208, 177–188. ( 10.1016/j.cis.2014.01.016) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Wang Y-H, Slochower DR, Janmey PA. 2014. Counterion-mediated cluster formation by polyphosphoinositides. Chem. Phys. Lipids 182, 38–51. ( 10.1016/j.chemphyslip.2014.01.001) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Chierico L, Joseph AS, Lewis AL, Battaglia G. 2014. Live cell imaging of membrane/cytoskeleton interactions and membrane topology. Sci. Rep. 4, 6056 ( 10.1038/srep06056) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Honigmann A. et al. 2013. Phosphatidylinositol 4,5-bisphosphate clusters act as molecular beacons for vesicle recruitment. Nat. Struct. Mol. Biol. 20, 679–686. ( 10.1038/nsmb.2570) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Wang J, Richards DA. 2012. Segregation of PIP2 and PIP3 into distinct nanoscale regions within the plasma membrane. Biol. Open 1, 857–862. ( 10.1242/bio.20122071) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Dason JS, Smith AJ, Marin L, Charlton MP. 2014. Cholesterol and F-actin are required for clustering of recycling synaptic vesicle proteins in the presynaptic plasma membrane. J. Physiol. 592, 621–633. ( 10.1113/jphysiol.2013.265447) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Wang Y-H, Collins A, Guo L, Smith-Dupont KB, Gai F, Svitkina T, Janmey PA. 2012. Divalent cation-induced cluster formation by polyphosphoinositides in model membranes. J. Am. Chem. Soc. 134, 3387–3395. ( 10.1021/ja208640t) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Wen Y, Vogt VM, Feigenson GW. 2018. Multivalent cation-bridged PI(4,5)P2 clusters form at very low concentrations. Biophys. J. 114, 2630–2639. ( 10.1016/j.bpj.2018.04.048) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Gericke A. 2018. Is calcium fine-tuning phosphoinositide-mediated signaling events through clustering? Biophys. J. 114, 2483–2484. ( 10.1016/j.bpj.2018.05.004) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Yu H, Schulten K. 2013. Membrane sculpting by F-BAR domains studied by molecular dynamics simulations. PLoS Comput. Biol. 9, e1002892–15 ( 10.1371/journal.pcbi.1002892) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Klauda JB, Venable RM, Freites JA, O’Connor JW, Tobias DJ, Mondragon-Ramirez C, Vorobyov I, MacKerell AD Jr, Pastor RW. 2010. Update of the CHARMM all-atom additive force field for lipids: validation on six lipid types. J. Phys. Chem. B 114, 7830–7843. ( 10.1021/jp101759q) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Li Z, Venable RM, Rogers LA, Murray D, Pastor RW. 2009. Molecular dynamics simulations of PIP2 and PIP3 in lipid bilayers: determination of ring orientation, and the effects of surface roughness on a poisson-Boltzmann description. Biophys. J. 97, 155–163. ( 10.1016/j.bpj.2009.04.037) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Feller SE, MacKerell AD. 2000. An improved empirical potential energy function for molecular simulations of phospholipids. J. Phys. Chem. B 104, 7510–7515. ( 10.1021/jp0007843) [DOI] [Google Scholar]
- 33.Lim JB, Rogaski B, Klauda JB. 2011. Update of the cholesterol force field parameters in CHARMM. J. Phys. Chem. B 116, 203–210. ( 10.1021/jp207925m) [DOI] [PubMed] [Google Scholar]
- 34.Hatcher ER, Guvench O, MacKerell AD Jr. 2009. CHARMM Additive all-atom force field for acyclic polyalcohols, acyclic carbohydrates, and inositol. J. Chem. Theory Comput. 5, 1315–1327. ( 10.1021/ct9000608) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Venable RM, Luo Y, Gawrisch K, Roux B, Pastor RW. 2013. Simulations of anionic lipid membranes: development of interaction-specific ion parameters and validation using NMR data. J. Phys. Chem. B 117, 10183–10192. ( 10.1021/jp401512z) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Pluhackova K, Kirsch SA, Han J, Sun L, Jiang Z, Unruh T, Böckmann RA. 2016. A critical comparison of biomembrane force fields: structure and dynamics of model DMPC, POPC, and POPE bilayers. J. Phys. Chem. B 120, 3888–3903. ( 10.1021/acs.jpcb.6b01870) [DOI] [PubMed] [Google Scholar]
- 37.Yoo J, Wilson J, Aksimentiev A. 2016. Improved model of hydrated calcium ion for molecular dynamics simulations using classical biomolecular force fields. Biopolymers 105, 752–763. ( 10.1002/bip.22868) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Yoo J, Aksimentiev A. 2018. New tricks for old dogs: improving the accuracy of biomolecular force fields by pair-specific corrections to non-bonded interactions. Phys. Chem. Chem. Phys. 20, 8432–8449. ( 10.1039/C7CP08185E) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Slochower DR, Huwe PJ, Radhakrishnan R, Janmey PA. 2013. Quantum and all-atom molecular dynamics simulations of protonation and divalent ion binding to phosphatidylinositol 4,5-bisphosphate (PIP2). J. Phys. Chem. B 117, 8322–8329. ( 10.1021/jp401414y) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Van Der Spoel D, Lindahl E, Hess B, Groenhof G, Mark AE, Berendsen HJC. 2005. GROMACS: fast, flexible, and free. J. Comput. Chem. 26, 1701–1718. ( 10.1002/jcc.20291) [DOI] [PubMed] [Google Scholar]
- 41.Slochower DR, Wang Y-H, Radhakrishnan R, Janmey PA. 2015. Physical chemistry and membrane properties of two phosphatidylinositol bisphosphate isomers. Phys. Chem. Chem. Phys. 17, 12 608–12 615. ( 10.1039/C5CP00862J) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Böckmann RA, Grubmuller H. 2004. Multistep binding of divalent cations to phospholipid bilayers: a molecular dynamics study. Angew. Chem. Int. Ed. 43, 1021–1024. ( 10.1002/anie.200352784) [DOI] [PubMed] [Google Scholar]
- 43.Hong C, Tieleman DP, Wang Y. 2014. Microsecond molecular dynamics simulations of lipid mixing. Langmuir 30, 11 993–12 001. ( 10.1021/la502363b) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Park S, Beaven AH, Klauda JB, Im W. 2015. How tolerant are membrane simulations with mismatch in area per lipid between leaflets? J. Chem. Theory Comput. 11, 3466–3477. ( 10.1021/acs.jctc.5b00232) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Sommer B. 2013. Membrane packing problems: a short review on computational membrane modeling methods and tools. Comput. Struct. Biotechnol. J. 5, e201302014 ( 10.5936/csbj.201302014) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Radhakrishnan R, Jordan EJ, Bradley RP. 2018. See https://biophyscode.github.io/.
- 47.Milo R, Phillips R. 2015. Cell biology by the numbers. New York, NY: Garland Science. [Google Scholar]
- 48.Lodish H, Berk A, Zipursky L, Matsudaira P, Baltimore D, Darnell J. 2000. Molecular cell biology, 4th edn. New York, NY: W. H. Freeman. [Google Scholar]
- 49.Bussi G, Bussi G, Donadio D, Donadio D, Parrinello M, Parrinello M. 2007. Canonical sampling through velocity rescaling. J. Chem. Phys. 126, 014101 ( 10.1063/1.2408420) [DOI] [PubMed] [Google Scholar]
- 50.van der Walt S, Colbert SC, Varoquaux G. 2011. The NumPy array: a structure for efficient numerical computation. arXiv: 1102.1523. ( 10.1109/MCSE.2011.37) [DOI]
- 51.Michaud-Agrawal N, Denning EJ, Woolf TB, Beckstein O. 2011. MDAnalysis: a toolkit for the analysis of molecular dynamics simulations. J. Comput. Chem. 32, 2319–2327. ( 10.1002/jcc.21787) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Jones E, Oliphant T, Pet P. 2001. SciPy: Open source scientific tools for Python. See http://www.scipy.org/.
- 53.Maneewongvatana S, Mount DM. 1999 Analysis of approximate nearest neighbor searching with clustered point sets. arXiv: cs/9901013.
- 54.Barber CB, Dobkin DP, Huhdanpaa H. 1996. The quickhull algorithm for convex hulls. ACM Trans. Math. Software (TOMS) 22, 469–483. ( 10.1145/235815.235821) [DOI] [Google Scholar]
- 55.Humphrey W, Dalke A, Schulten K. 1996. VMD: visual molecular dynamics. J. Mol. Graphics 14, 33–38. ( 10.1016/0263-7855(96)00018-5) [DOI] [PubMed] [Google Scholar]
- 56.Henderson D, Boda D. 2009. Insights from theory and simulation on the electrical double layer. Phys. Chem. Chem. Phys. 11, 3822–3830. ( 10.1039/b815946g) [DOI] [PubMed] [Google Scholar]
- 57.Han K, Gericke A, Pastor RW. 2020. Characterization of specific ion effects on PI(4,5)P2 clustering: molecular dynamics simulations and graph-theoretic analysis. J. Phys. Chem. B 124, 1183–1196. ( 10.1021/acs.jpcb.9b10951) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Pavlov M, Siegbahn PEM, Sandström M. 1998. Hydration of beryllium, magnesium, calcium, and zinc ions using density functional theory. J. Phys. Chem. A 102, 219–228. ( 10.1021/jp972072r) [DOI] [Google Scholar]
- 59.Wang Y-H, Collins A, Guo L, Smith-Dupont KB, Gai F, Svitkina T, Janmey PA. 2012. Divalent cation-induced cluster formation by polyphosphoinositides in model membranes. J. Am. Chem. Soc. 134, 3387–3395. ( 10.1021/ja208640t) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Venable RM, Ingólfsson HI, Lerner MG, Perrin BS Jr, Camley BA, Marrink SJ, Brown FLH, Pastor RW. 2017. Lipid and peptide diffusion in bilayers: the Saffman–Delbrück model and periodic boundary conditions. J. Phys. Chem. B 121, 3443–3457. ( 10.1021/acs.jpcb.6b09111) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Kučerka N, Nagle JF, Sachs JN, Feller SE, Pencer J, Jackson A, Katsaras J. 2008. Lipid bilayer structure determined by the simultaneous analysis of neutron and X-ray scattering data. Biophys. J. 95, 2356–2367. ( 10.1529/biophysj.108.132662) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Levental I, Christian DA, Wang Y-H, Madara JJ, Discher DE, Janmey PA. 2009. Calcium-dependent lateral organization in phosphatidylinositol 4,5-bisphosphate (PIP2)- and cholesterol-containing monolayers. Biochemistry 48, 8241–8248. ( 10.1021/bi9007879) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Flanagan LA, Cunningham CC, Chen J, Prestwich GD, Kosik KS, Janmey PA. 1997. The structure of divalent cation-induced aggregates of PIP2 and their alteration by gelsolin and tau. Biophys. J. 73, 1440–1447. ( 10.1016/S0006-3495(97)78176-1) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Levental I, Cebers A, Janmey PA. 2008. Combined electrostatics and hydrogen bonding determine intermolecular interactions between polyphosphoinositides. J. Am. Chem. Soc. 130, 9025–9030. ( 10.1021/ja800948c) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Melcrová A, Pokorna S, Pullanchery S, Kohagen M, Jurkiewicz P, Hof M, Jungwirth P, Cremer PS, Cwiklik L. 2016. The complex nature of calcium cation interactions with phospholipid bilayers. Sci. Rep. 6, 38035 ( 10.1038/srep38035) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Vangaveti S, Travesset A. 2014. Separation of the Stern and diffuse layer in coarse-grained models: the cases of phosphatidyl serine, phosphatidic acid, and PIP2 monolayers. J. Chem. Phys. 141, 245102 ( 10.1063/1.4904885) [DOI] [PubMed] [Google Scholar]
- 67.Lorenz CD, Faraudo J, Travesset A. 2008. Hydrogen bonding and binding of polybasic residues with negatively charged mixed lipid monolayers. Langmuir 24, 1654–1658. ( 10.1021/la703550t) [DOI] [PubMed] [Google Scholar]
- 68.Bucki R, Wang Y-H, Yang C, Kandy SK, Fatunmbi O, Bradley R, Pogoda K, Svitkina T, Radhakrishnan R, Janmey PA. 2019. Lateral distribution of phosphatidylinositol 4,5-bisphosphate in membranes regulates formin- and ARP2/3-mediated actin nucleation. J. Biol. Chem. 294, 4704–4722. ( 10.1074/jbc.RA118.005552) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Graber ZT, Shi Z, Baumgart T. 2017. Cations induce shape remodeling of negatively charged phospholipid membranes. Phys. Chem. Chem. Phys. 19, 15 285–15 295. ( 10.1039/C7CP00718C) [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
The protocols are available as part of the electronic supplementary material. The codes and scripts along with complete documentation are available as part of the biophyscode software developed by the authors. The distribution of biophyscode is through an open-source license through GitHub and can be accessed using the link: https://biophyscode.github.io.