Skip to main content
UKPMC Funders Author Manuscripts logoLink to UKPMC Funders Author Manuscripts
. Author manuscript; available in PMC: 2025 Feb 12.
Published in final edited form as: J Mol Biol. 2025 Jan 9;437(4):168937. doi: 10.1016/j.jmb.2025.168937

The Influence of Phosphoinositide Lipids in the Molecular Biology of Membrane Proteins: Recent Insights from Simulations

George Hedger 1,*, Hsin-Yung Yen 2,3
PMCID: PMC7617384  EMSID: EMS202886  PMID: 39793883

Abstract

The phosphoinositide family of membrane lipids play diverse and critical roles in eukaryotic molecular biology. Much of this biological activity derives from interactions of phosphoinositide lipids with integral and peripheral membrane proteins, leading to modulation of protein structure, function, and cellular distribution. Since the discovery of phosphoinositides in the 1940s, combined molecular biology, biophysical, and structural approaches have made enormous progress in untangling this vast and diverse cellular network of interactions. More recently, in silico approaches such as molecular dynamics simulations have proven to be an asset in prospectively identifying, characterising, explaining the structural basis of these interactions, and in the best cases providing atomic level testable hypotheses on how such interactions control the function of a given membrane protein. This review details a number of recent seminal discoveries in phosphoinositide biology, enabled by advanced biomolecular simulation, and its integration with molecular biology, biophysical, and structural biology approaches. The results of the simulation studies agree well with experimental work, and in a number of notable cases have arrived at the key conclusion several years in advance of the experimental structures.

Keywords: phosphoinositide, PIP2, lipid-protein interaction, MD simulation, membrane protein

Introduction

Phosphoinositides are a family of membrane lipids with far reaching influences on cellular function.1 The study of phosphoinositides has a rich history dating back many decades.27 This work has revealed roles in cell signalling,8 trafficking,9 cytoskeletal organisation,10 and regulation of ion passage across membranes.11 Pathogenic variance in the kinases and phosphatases which control the relative cellular abundance of phosphoinositides are associated with a number of human diseases including growth disorders, cancers, and neurological conditions.1216 Underlying these cellular functions are molecular level interactions of phosphoinositides with integral and peripheral membrane proteins. Experimental methods such as X-ray crystallography have made significant early contributions to revealing the detailed architecture of membrane proteins in complex with phosphoinositides.17 The advent of technologies including atomic resolution cryo-EM18 and native mass spectrometry19 have accelerated that understanding.2022 Such structural studies have been coupled to intricate biochemical work to show how specific phosphoinositides control the function of certain membrane proteins. Despite progress, resolving bound phosphoinositides in structures23 and relating this information to function and dynamics at the molecular-level using experimental techniques remains non-trivial from static structures alone. In silico approaches such as molecular dynamics (MD) simulations have experienced rapid technological development in the area of membranes and membrane proteins.24,25 This has led to a surge in simulation studies, and a number of notable contributions to uncovering new aspects of phosphoinositide biology. Many of these simulations have been performed in an integrated fashion with experiments. The simulations serve to computationally reunite experimentally determined membrane protein structures with their complex lipid bilayer environment, and then capture the atomic level dynamics and structural aspects of how the membrane protein interacts with phosphoinositides. In a number of cases, simulations have prospectively predicted phosphoinositide binding sites on integral membrane proteins several years in advance of experimental structures being determined. In the very best examples the simulations have provided a detailed mechanistic view into how the structure and dynamics of the interaction then lead to the phosphoinositide altering the function of a given membrane protein.

Within this review we examine how simulations and complementary experiments have revealed new and exciting molecular-level aspects of phosphoinositide biology. We begin with a brief survey of cell membranes, phosphoinositides, and biomolecular simulation approaches.

Cell membranes

Cells and subcellular compartments are defined by membranes. Membranes are composed of a lipid bilayer and embedded membrane proteins. This bilayer is compositionally complex, containing hundreds of lipid species asymmetrically distributed between leaflets.26 Advances in structural biology of membrane proteins18 and in lipidomics27,28 have led to enormous progress in uncovering the compositional complexity and molecular level architecture of membranes. The structures of ca. 1700 unique membrane proteins have been determined (https://blanco.biomol.uci.edu/mpstruc/) alongside ca. 49,000 lipid structures (https://www.lipidmaps.org/databases/lmsd/overview). The composition of a membrane and its organisation varies substantially by membrane type, stage in the life cycle, metabolic state, signalling state, and health and disease.29,30 Lipids interact with membrane proteins and alter their structure, function, localisation, and oligomeric states.31 These interactions may take a variety of forms, including binding to specific well-defined sites on the membrane-exposed surface of a protein with complementary geometry, binding to sites deep within a protein, or more dynamic interactions with disordered patches of complementary charge or grease. They may also occur indirectly by influences on the local biophysical properties of the membrane. Lipids may even represent the native ligand for a membrane protein, as has been proposed for activation of Smoothened by cholesterol.32,33 Examples of regulation of membrane protein function by phosphoinositides include receptor tyrosine kinases (RTKs),34 inward rectifying potassium (Kir) channels,35 G-protein-coupled receptors (GPCRs),21 neurotransmitter transporters36,37 and syntaxins.38 In the case of phosphoinositides, another major type of lipid-protein interaction is the recruitment and orientation of soluble proteins, so called ‘peripheral membrane proteins’ (PMPs) to the membrane surface by direct association of the PMP with the phosphoinositide headgroup.39,40

The phosphoinositide family

Phosphoinositides are phosphorylated derivates of phosphatidylinositol (PI) (Figure 1). Differential decoration of the myo-inositol headgroup with varying numbers and positions of phosphate groups give rise to seven derivates in humans. The inositol group adopts a chair conformation with inorganic ions complexed to the negative phosphoryl groups.41 The acyl tails show unusual uniformity compared to other phospholipids, and are usually of the 1-stearoyl-2-arachidonoyl variety.42 The overall abundance of phosphoinositides is estimated at less than 1% of cellular phospholipids, with species such as PI(3,4,5)P3 being even rarer.43,44 This seeming dichotomy led to the famous adage ‘Tiny lipids with giant impact on cell regulation’.1 Precise measurements of abundance remain challenging, not least as it is a moving target. The relative abundance of different phosphoinositide species is under constant dynamic spatiotemporal control, and varies by cell type, signalling state, and myriad other cellular conditions. The distribution of phosphoinositide species also varies by subcellular location (Figure 1C),14,44,45 which confers unique biochemical properties to the host membranes. Localisation to specific membranes, as well as localisation to particular regions within the same membrane,38 can lead to substantially higher local concentrations than the total cellular fraction would imply. Phosphoinositides also exhibit leaflet asymmetry, with e.g. PI(4,5)P2 being found within the inner leaflet of the plasma membrane, with reports of it constituting up to 5% of total lipids.8,46 While the outer leaflet of human plasma membranes is abundant in enormous lipids with complex glycosylation patterns,47,48 within the inner leaflet it is phosphoinositides which bear one of the largest headgroups relative to other phospholipids, which protrude beyond the membrane surface into solution. They are also the most negatively charged of the major plasma membrane inner leaflet lipids. These two properties in particular set them apart and confer a suite of biochemical interactions not available to other phospholipids. During apoptosis this plasma membrane asymmetry is disrupted, and phosphoinositides have been reported to be present in elevated levels within the outer leaflet where they may act as ‘eat-me’ signals in cell death.49 Although not canonically referred to as phosphoinositides, membrane lipids containing the core phosphatidylinositol structure have been seen in other domains of life, for example in mycobacteria.50,51

Figure 1. The phosphoinositide family.

Figure 1

(A) Structure of the phosphatidylinositol headgroup depicting the chair conformation and numbering scheme of the myo-inositol moiety. (B) The predominant fatty acid tails. (C) Headgroup structures of the seven eukaryotic phosphoinositides, derived from phosphatidylinositol. The average cellular distributions are indicated. Note that this varies with cell type and myriad other variables. See Refs. 14,44 for further detailed discussion of cellular distribution.

Computational approaches for studying phosphoinositides

A variety of computational approaches are available to learn something about the interactions and dynamics of biomolecules. These include docking,52,53 artificial intelligence (AI) methods,5457 quantum mechanics/molecular mechanics (QM/MM) approaches,58,59 and molecular dynamics (MD) simulations.60,61 Amongst these, MD simulation has shown particular utility in the study of lipid–protein interactions.6265 At its core, MD simulation is a physics-based method to predict how a group of particles (representing the biomolecule(s) of interest) propagate through time and space, using classical mechanics and an underlying potential energy function which describes particle interactions. The functional form of this potential energy function and the set of parameters associated with it are termed the ‘force field’. Parameters for a given biomolecule are obtained from experiments and quantum mechanical calculations. The successful application of the simulation relies on having a well-parameterised model and sufficient sampling of conformational space, coupled to appropriate simulation design and analysis. Common force field choices for membrane protein systems include CHARMM,66,67 AMBER,6870 and MARTINI.24,71 Simulations may be performed at all-atom resolution, or at ‘coarse-grained’ (CG) resolution, in which groups of atoms are represented as single particles. This is the case for the popular MARTINI force field,24 which has shown particular utility in simulation of membrane systems and lipid-protein interactions.25,64,72,73 CG resolutions decrease the computational load of the calculations and allow greater sampling of time and space. The compromise on atomic detail may be tempered by conversion74,75 of the end-point of a CG simulation back to all-atom detail followed by subsequent all-atom simulation, i.e. multiscale simulation.76 Other approaches to addressing the sampling challenge include the use of special purpose supercomputers such as ANTON,77 and a group of techniques collectively termed ‘enhanced sampling’,78 which are essentially statistical mechanical and algorithmic tricks to enforce sampling of certain regions of space or ‘collective variables’. This review will share examples of all three of these strategies as it pertains to simulations of phosphoinositides.

Amongst phospholipids, phosphoinositides represent one of the more challenging lipid groups for MD simulation due to their highly charged headgroups. For example, while the net headgroup charge of PI(4,5)P2 lipids has been estimated as ca. 4 by 31P NMR and QM methods in simplified systems,41,79 the magnitude and charge distribution may vary substantially in vivo depending on physiological microenvironment.80,81 This last part is key, as a PI(4,5)P2 lipid in the vicinity of a cluster of basic residues may be induced toward the −5 state, while in other cellular scenarios may adopt forms with reduced net charge. The vast majority of simulation approaches employ a fixed charge model in which charge states are set at the beginning of the simulation and remain thus. The choice of charge state is therefore non-trivial. While this is a general challenge in MD simulations,82 it is perhaps more acute for certain phosphoinositides. Though -4 variants of PI(4,5)P2 are most common,83 -5 variants have been seen.84 What is most appropriate depends on the system in question and testing of multiple variants may be necessary. Despite these challenges, significant effort has been put into the continued improvement of phosphoinositide parameters, and upgraded versions continue to be released.83,85,86 This even extends beyond the core group of mammalian phosphoinositides to more exotic forms with additional modifications, such as those found in Mycobacterium tuberculosis.87

A theme which will become apparent in the course of this review is that amongst the various phosphoinositide species, PI(4,5)P2 has received particularly high levels of attention in simulation and structural studies. This may in part correlate with its inherent biological importance and its relative abundance under a range of cellular states. It may also reflect its localisation to (the inner leaflet of) the plasma membrane, where much membrane protein research has historically tended to focus, especially from a pharmacological perspective. Pragmatic factors such as the early availability of parameters and seminal simulation works on PI(4,5)P2 may also have contributed.38,84,88,89 We note that simulation parameters are now available for all phosphoinositides species in both all-atom and CG resolutions, and continue to be improved.83,85,90,91

More generally, parameters now exist for many thousands of different membrane lipid species.86,92,93 This explosion in parameter space has been coupled to knowledge from experimental lipidomics27,29 to allow simulations in complex asymmetric membranes which mimic e.g. cell-type specific plasma membranes94,95 and other subcellular membranes.25,9698 The mutual lipid-lipid interplay between species within complex biologically relevant membranes can be critical to the biological activity of phosphoinositides99 and the potential to capture this in simulations is a powerful ability. In addition to improvements in parameter quality, new tools for the setup and analysis of simulations continue to be developed.93,100105 In particular, specialised tools for the analysis of lipid-protein interaction have emerged, including PyLipID106 and ProLint,107 as well as refinement of more advanced simulation methodologies such as calculating Δ Δ Gs for lipid binding.108

These specific advances have occurred to the continual background march of hardware improvement109 and general force field quality.24 Notably, the widespread use of GPU computing and the corresponding updates to simulation code have dramatically increased the timescales accessible to research groups.110 CG simulations of small membrane proteins such as GPCRs may now reach hundreds of microseconds without breaking too much of a sweat, while length scales of hundreds of nanometres have also been seen.111116 On commodity compute devices atomistic simulations are comfortably of the order of several microseconds for most groups and system sizes, while special purpose devices may routinely reach hundreds of microseconds as a matter of course.117 The ‘dawn’ of exascale computing118 will further increase accessible time and length scales.119

Coupled with new experimental capabilities in structural biology18 and biophysics,120 these improvements in simulation approaches have been applied alongside experiments by hundreds of scientists to unlock the molecular basis for many exciting new facets of phosphoinositide biology. In the following section we highlight a range of case studies. We then draw it together to summarise an emerging picture of simulation-based analysis of phosphoinositide influence on membrane protein structure, function, and dynamics. We end by taking a critical look at challenges and imminent advances.

Recent Case Studies

G-protein-coupled receptors

GPCRs constitute the largest family of human membrane proteins121 and play key roles in signal transduction at the plasma membrane,122 and at subcellular membranes.123125 The core unit of these receptors consists of a seven transmembrane helical bundle, which may be differentially decorated with extracellular and intracellular domains.126 GPCRs are activated by a diverse array of stimuli including binding of small molecule odorants,127 peptide hormones,128 neurotransmitters,129 and by photons of light.130 This leads to a conformational change in the helical bundle, coupling to intracellular interaction partners such as G-proteins and arrestins, and signal propagation to downstream cascades ultimately leading to a cellular outcome.131

Membrane lipids bind and control GPCR function via diverse modes. The role of cholesterol has been the subject of particularly extensive study by experiment and simulation.32,132134 Phospholipids have been known to bind and modulate GPCR structure and function for some time.135138 However, it is only relatively recently that phosphoinositide modulation of GPCRs has moved into the spotlight.22

An integrated simulation and native mass spectrometry study in 2018 discovered that PI(4,5)P2 lipids were able to bind directly to Class A GPCRs and enhance G-protein coupling.21 In this work, the interactions of PI(4,5)P2 with the Neurotensin receptor 1 (NTSR1), the Adenosine A2A receptor (A2AR), and the β1 adrenergic receptor (β1AR) were first detected by structural native mass spectrometry. CG simulations of up to 100 μs duration were then employed to predict the locations of these interactions on the membrane exposed surface of the GPCR. Key residues seen to bind PI(4,5)P2 in the simulations were used to inform a mass spectrometry mutagenesis strategy which identified TM4 and TM1 as the major PI(4,5)P2 binding region on NTSR1, alongside several lower affinity sites (Figure 2A). Interestingly, in both simulations and mass spectrometry experiments, PI(4,5)P2 outcompeted all other phosphoinositide sub-types, including the more negatively charged PI(3,4,5)P3. This is indicative of specific complementary geometries and not only simple correlation with net headgroup charge. High-throughput CG simulations then expanded this analysis to nine other Class A GPCRs, demonstrating conserved PI(4,5)P2 interaction, albeit with variability in the precise interaction pattern. The functional consequences of this interaction were demonstrated by employing potential of mean force (PMF) simulations139 in which a mini-G-protein was pulled away from the A2AR under ±PI(4,5)P2 conditions. This showed a

Figure 2. PI(4,5)P2 binding sites on the NTSR1 GPCR identified by simulation and experiment.

Figure 2

(A) View from the cytosol onto the base on the NTSR1 showing simulation derived densities for PI(4,5)P2 around the TM helices.21 Densities are shown for an ensemble of short simulations (magenta), and one long simulation (blue) to demonstrate convergence of the simulation over these timescales. The site which showed the highest affinity for PI(4,5)P2 in the simulations and nMS is labelled. Helices with red patches indicate residues with high levels of direct phosphoinositide interaction. (B) Side view of a Cryo-EM structure of NTSR1 in complex with β-arrestin 1 showing short-chain dioctyl PI(4,5)P2 resolved at the same high affinity site (PDB ID: 6UP7).140

ca. 50% enhancement in the stability of GPCR-G-protein complex in the presence of PI(4,5)P2, which acted as a molecular glue bridging between the GPCR and G-protein. This agreed well with complementary mass spectrometry G-protein coupling experiments which showed selective enhancement of complex formation for mini-Gαs over mini-Gαi and mini-Gα12, and withfluorescence-based GTPase functional assays in the presence and absence of PI(4,5)P2.21 Intriguingly, two years after this discovery, a cryo-EM structure was published of NTSR1 bound to β-arrestin 1.140 This structure showed a short-chain dioctyl PI(4,5)P2 lipid bound at the same high affinity TM4 site previously identified in the simulations and mass spectrometry (Figure 2B). These combined works uncovered the structural and functional basis for a new biological role for PI(4,5)P2 in GPCR signalling, with broad consequences for cell biology.8,141 Of particular note, while most known roles of phosphoinositides have their origins in the culture plates and test tubes of cell biology and molecular biology approaches, the present story began with simulations and chemistry at the molecular level within computers and mass spectrometers.

Structural coverage of GPCRs has increased exponentially in recent years.142 The number of experimentally determined structures in the PDB and GPCRdb is now well over one thousand.143 The advent of predictive AI models such as the AlphaFold suite,54 Chai-1,57 and RoseTTAFold56 provide another means of obtaining initial structural coordinates for simulation,144146 alongside traditional homology modelling.147 This has presented new opportunities for molecular simulations. A number of studies have leveraged the increased availability of structures to perform simulations of GPCRs in both active and inactive states to assess the conformational dependence of phosphoinositide interactions.148154 Song et al., identified differences in the free energies of interaction for PI(4,5)P2 with binding sites on the A2AR between inactive, active, and active-miniG states.148 Ma et al., subsequently showed the stiffening effects of PI(4,5)P2 on both the inactive and active states of A2AR in limiting the accessible conformational landscape.151 Later experiments employing 19F NMR in nanodiscs suggested that in addition to the role of PI(4,5)P2 in enhancing G-protein coupling,21 it may also increase population of the A2AR active state in the absence of G-protein.155 CG simulations combined with experimental measurements have reported similar PI(4,5)P2 induced stabilisation of the active state of the Ghrelin receptor,154 as well as an influence on its coupling to β-arrestin 1.156 Most recently PI(4,5)P2 interaction with ICL3 has been suggested to stabilise the active state of β2AR.153

Schiøtt and colleagues performed detailed work on the Class B Glucagon receptor (GCGR), finding that in contrast to Class A GPCRs, PI(4,5)P2 preferentially bound the inactive state of the receptor.150 Interestingly, the authors also found that binding was modulated by acyl tail saturation pattern, supported by native mass spectrometry measurements. This is a novel observation as phosphoinositides in vivo have particularly uniform acyl tail composition of the stearoryl/arachidonoyl form,42 and most interactions seen in simulations and structures tend to be dominated by headgroup electrostatics, with tails being dynamic and/or unresolved due to inherentflexibility and the use of short-chain derivatives.17 An influence of acyl tail saturation pattern on PI(4,5)P2 clustering propensity has also been reported in simulations in the absence of protein.157

Amongst Class F GPCRs, CG simulations of the hedgehog signal transducer Smoothened (SMO) predicted PI(4,5)P2 encounter complexes.158 This is intriguing as SMO signalling is intricately linked to primary cilia, a region which contains special zones of phosphoinositide enrichment.159 Joubert’s syndrome, a human ciliopathy characterized by impaired Hh signaling and human birth defects, can be caused from mutations in a 5-positon phosphatase (Inpp5e) which lead to alterations in the distribution of ciliary PI(4,5)P2.12,159161

As well as focussed simulations of different conformational states of particular GPCRs, simulations may also assess trends across larger groups. Sejdiu and Tieleman performed simulations of 23 unique GPCRs and characterised their ‘lipid fingerprint’, that is, the unique lipid interaction profile of each GPCR.149 While phosphoinositide species interacted with all 23 GPCRs, the location and strength of interaction showed substantial variation. Most recently, CG simulations of 15 human Class B1 GPCRs were performed in both active and inactive states.162 A conserved state-dependent pattern of PI(4,5)P2 interaction was seen across the subfamily at the TM6/7 – H8 interface, in agreement with earlier CG simulations of GCGR.150,152 This observation is particularly intriguing as two antagonists of GCGR have been structurally resolved bound at this same interaction site (PDB ids: 5EE7 and 5XEZ).163,164 A positive allosteric modulator (PAM) used to stabilise structures of glucose-dependent insulinotropic polypeptide receptor (GIPR) also binds adjacent to this site, and has been used to stabilise the receptor in complex with the antidiabetic therapeutic tirzepatide (Mounjaro) (PDB id: 7RBT).165 This raises the intriguing prospect of possible direct phosphoinositide / drug competition and synergistic binding effects at this region.

Phosphoinositide interactions with GPCRs have also been modelled in large membrane patches measuring up to ca. 135 nm across and containing 144 GPCR proteins embedded in plasma membrane like lipid compositions166168 (Figure 3A). As well as providing excellent statistics on phosphoinositide behaviour, these simulations allow assessment of in vivo like factors including e.g. competition effects between different phosphoinositide species such PI(4,5)P2 and PI(3,4,5)P3, the involvement of membrane curvature, and of lateral protein–protein interactions. In this spirit, clustering of PI(4,5)P2 around the S1P receptor 1 (S1PR1)166 and NTSR1167 have been observed. Similarly, work on the A2AR embedded in complex membranes employed simulations coupled to Markov state modelling to predict PI(4,5)P2 mediated enhancement of A2AR oligomerisation168 (Figure 3B).

Figure 3. Phosphoinositides can mediate lateral GPCR interactions.

Figure 3

(A) 144 S1PR1 GPCRs (pink) in a membrane with complex asymmetric lipid distribution mimicking that of a mammalian plasma membrane. The image shown is from the end of a 10 μs CG simulation. (B) An A2AR dimer with interfacial PI(4,5)P2 lipids mediating the interaction. We thank Dr. Heidi Koldsø and Dr. Wanling Song for the figure. Figure (adapted) reprinted with permission from 166 and 168. Copyright (2015) American Chemical Society.

A common thread in all simulation studies of GPCRs in multicomponent membranes, is that despite comprising a small percentage of the simulated lipid composition, phosphoinositides routinely outcompete all other inner leaflet phospholipids for interaction with the GPCR. This is especially pertinent for lipid species which, like phosphoinositides, have net anionic headgroup charges such as phosphatidylserine (PS) and phosphatidylglycerol (PG) and which have previously been reported to modulate GPCRs.135,169171 This highlights the importance of including the biologically relevant phosphoinositide species in experiments and simulations investigating interactions of lipids with GPCRs. It also raises the question of how many anionic phospholipid interaction sites and functional effects previously reported from in vitro and in silico studies of GPCRs in simple membranes without phosphoinositides may be abrogated or enhanced by phosphoinositides in vivo.

Ion channels

Ion channels have a particularly close relationship with phosphoinositides.35 This relationship has been subject to extensive structural, functional, and computational study.172174 We focus on transient receptor potential (TRP) channels, for which phosphoinositide modulation is diverse and emerging, and on inward-rectifying potassium (Kir) channels, which require PI(4,5)P2 for activation.

Transient receptor potential channels

Transient receptor potential (TRP) channels are a major class of tetrameric ion channel,175 and their dysfunction has been implicated in a range of diseases.176178 TRP channels have been shown to be functionally modulated by lipids including sterols and phospholipids.179 The binding of phosphoinositides to TRP channels has been explored in a range of structural and simulation studies.20,180186 In 2020, a combined simulation, biochemical, and structural study explored the binding of a number of phosphoinositide species to polycystin-2181 (PC2; also known as TRPP2 or PKD2). Cryo-EM structures of PC2 in the presence of PI(3,5)P2 (to 3.4 Å), and of PI(4,5)P2 (to 3 Å) revealed non-protein density between the transmembrane helices S3, S4 and S5 (Figure 4). As is often the case in structural studies, lipid identity could not be unambiguously assigned from the experimental electron density alone. The authors addressed this challenge by performing CG simulations of PC2 in lipid bilayers, in which they observed spontaneous binding of PI(4,5)P2 lipids to the same S3, S4, S5 region seen in the PC2 cryo-EM structures (Figure 4A,B). This region had also previously been seen to bind PI, PI(4,5)P2, and vanilloids (e.g. capsaicin) in TRPV1/5 cryo-EM structures and MD simulations of TRPV6.20,180 Gao et al., noted in their phosphoinositide bound structure the possibility of the pocket being suited to bind a range of phosphoinositide species.20 Simulations designed to calculate free energies of lipid association with this site on PC2 from the PMF139 well-depth obtained values of ca. −9 kJ/mol (PS), −20 kJ/mol (PI), −28 kJ/mol (PI(3,4,5)P3), and −37 kJ/mol (PI(4,5)P2). While this suggested a specific interaction of PIP2 over PS, PI, and PI(3,4,5)P3, the authors noted the limitations of the CG simulation model in distinguishing between the PI(4,5)P2 and PI(3,5)P2 isoforms. A biochemical assay in the form of ‘PIP strips’ provides one route to experimental interrogation of lipid selectivity.187,188 This assay consists of a nitrocellulose membrane dotted with different lipid species. Binding of the protein of interest to these lipid spots can then be detected using antibodies. Wang et al., employed this approach for PC2 in detergent micelles to demonstrate selective binding of phosphoinositide species over other lipids (PA, PS, PE, S1P), in agreement with the simulations.181 However high variation between biological repeats circumscribed further dissection of selectivity between phosphoinositide species.

Figure 4. A phosphoinositide binding site on PC2.

Figure 4

(A) CG simulation-based identification of a PI(4,5)P2 binding site between helices S3/4/5 of PC2.181 (B) Lipid density seen at the same site in a cryo-EM structure of PC2 solved in the presence of PI(3,5P)2 (PDB ID: 6T9O, EMDB: 10419).181 (C) Phosphatidylinositol bound to the same site in a cryo-EM structure of TRPV1 (PDB ID: 5IRZ).20 Figure modified from 181.

Delling and colleagues recently applied detailed electrophysiology and simulation work to demonstrate binding of a ciliary-enriched oxysterol (7β,27-DHC) to PC2 at a site which included the S4-5 linker, and functional modulation of channel current.182 This oxysterol site overlapped with the putative phosphoinositide binding pocket previously identified from the structures and simulations.20,181 Remarkably, inclusion of either PI(3,5)P2 or PI(4,5)P2 decreased the modulatory effects of the oxysterol on PC1/2 current density, while PI(4)P completely abolished channel activation.182 Taken together, the high affinity competition indicated both the veracity of the previously identified phosphoinositide binding pocket, and suggests PI(4)P may be the key biologically relevant phosphoinositide species in controlling PC2 channel function at this site. This finding aligns with the cellular localisation of PC2 channels to cilia membranes, a region enriched in PI(4)P lipids.159

These studies highlight both the technical challenges of interrogating phosphoinositide interactions with membrane proteins, and how these challenges may be met by an integrated approach combining simulation and experiment. The paradigm of using CG simulation to assign molecular identity to putative lipid density in cryo-EM structures was recently developed into the LipIDens tool.23

In addition to phosphoinositide interaction sites with defined tertiary structure, interactions with intrinsically disordered regions (IDRs) of membrane proteins have been observed189,190. In the context of ion channels a fascinating example is provided by the ca. 150 residue N-terminal IDR of TRPV4.191 Goretzki et al., integrated a suite of experimental approaches and multiscale simulations to map the molecular level detail of PI(4,5)P2 interaction with the IDR ensemble, and its relationship to channel activity. CG simulations combined with NMR experiments delineated a mechanism whereby interaction of PI(4,5)P2 with a particular region of the IDR led to a mechanical pull-force on the ankyrin repeat domain (ARD), which ultimately led to force transduction to the TRPV4 core. The extent and structural basis for phosphoinositide selectivity when it comes to membrane-proximal IDR ensembles is intriguing to consider, and simulations combined with machine learning technologies are increasingly well positioned to further address this.192,193,66

Inward rectifying potassium channels

Potassium channels facilitate the selectiveflow of K+ ions across cellular membranes. These tetrameric ion channels are one of the best characterised and most ubiquitously expressed integral membrane proteins with wide ranging roles in physiology and disease.194 The three main groups of potassium channels are inward rectifying (Kir), voltage-gated (Kv), and two-pore (K2P) potassium channels. A variety of structural17,195 and functional196,197 studies have established central roles for lipids in potassium channel biology. A particularly intimate relationship is seen between Kir channels and PI(4,5)P2 lipids, which act as channel activators.198 This section highlights some of the key ways in which simulations have enriched our understanding of various aspects of phosphoinositide biology as it relates to Kir channels, as well as remaining challenges.

At the macroscale, it is now possible to simulate large patches of membrane containing hundreds of proteins embedded in complex membrane compositions.94,111 Thus Duncan et al., performed CG simulations of membrane patches containing >100 Kir2.2 channels in a plasma membrane like lipid composition112 (Figure 5A). Simulations under a range of defined lipid regimes and protein concentrations revealed a role for PI(4,5)P2 in influencing channel crowding and bilayerfluctuations. Follow up work assessed the nanoscale interactions of PI(4,5)P2 with Kir2.2199 by leveraging the large complex membranes to gather sufficient statistics to assess competition effects between different lipids. PI(4,5)P2 was seen to bind at the experimentally determined primary site on Kir2.217 (Figure 5). Interactions at a secondary200,201 anionic lipid site were also observed. Though PI(4,5)P2 interaction was most favourable over other lipid species at both sites, anionic PS lipids (net headgroup charge -1) were able to bind if the PI(4,5)P2 concentration was low enough. The molecular basis of PI(4,5)P2 selectivity over other phospholipids resulted from protrusion of the comparatively larger inositol headgroups into solution, with the 4′ and 5′ phosphoryl moieties able to interact with key basic residues of the cytoplasmic domains which were beyond the reach of phospholipids with smaller headgroups. Thus in this case selectivity resulted both from structure and from charge. Free energy perturbation (FEP) simulations were applied to estimate the ΔΔ G for transforming PI(4,5)P2 to PS at both lipid binding sites. The calculations showed the selectivity for PI(4,5)P2 over PS was significantly greater at the primary interaction site (ΔΔG = 43 k J/mol) than at the secondary interaction site (ΔΔG = 29 kJ/mol). Notably, the magnitude of the unfavourable ΔΔG for the secondary site was reduced from 29 kJ/mol to 17 kJ/mol if a PI(4,5)P2 lipid was also present within the primary site. That is, PS interaction at the secondary site became more favourable if a PI(4,5)P2 lipid was also present within the primary site. This delicate interplay between the two sites may provide the cell with an additional means of differentially modulating channel function by controlling the relative local abundance of competing lipid species.199

Figure 5. Primary PI(4,5)P2 binding site on Kir2.2 from simulation, experiment, and AI.

Figure 5

(A) View of the inner leaflet surface from a simulation containing multiple Kir2.2 channels embedded in a complex and crowded membrane environment.199 PI(4,5)P2 (magenta) is clustered around the channels (beige). (B) The primary PI(4,5)P2 sites identified on homotetrameric Kir2.2 by CG MD simulation.84,199 (C) the same site identified by X-ray crystallography (PDB id: 3SPI),17 and (D) by AI. The AI coordinates shown were predicted using Chai-157 (https://lab.chaidiscovery.com/), the monomeric FASTA sequence from PDB id: 3SPI, and the PubChem202 SMILES string for PI(4,5)P2. The red spheres denote the position of the phosphate group bridging the glycerol and inositol moieties of PI(4,5)P2. The perspective shown is from the extracellular side of the bilayer. (E–G) Lateral view of the same structures, with the PI(4,5)P2 lipids shown in magenta. Dotted lines denote the approximate position of the bilayer – solvent interface. We thank Dr. Anna Duncan for providing simulation coordinates used to produce the figure.

The location of the primary PI(4,5)P2 interaction site characterised by Duncan et al., was first predicted in simulations of KirBac1.1, a Kir3.1-KirBac1.3 chimera, and a homology model of Kir6.2 performed in 2009.84 This early work employed CG simulations of PI(4,5)P2 in simple membranes containing phosphatidylcholine (PC) and PI(4,5)P2. Simulations of 5 μs duration were sufficient to observe spontaneous and sustained binding of PI(4,5)P2. Key clusters of basic residues which formed the binding site for the inositol head-group had previously been suggested to be involved in Kir2.1 interactions with PI(4,5)P2.203205 A subsequently solved Kir2.2 crystal structure showed short-chain (dioctanoyl) PI(4,5)P2 density at the same site (Figure 5). Interestingly, in this case lipid binding seemed robust both to the acyl tail composition/absence, and the use of non-eukaryotic structures and homology models. Further simulation studies of Kir2.2 have continued to support the veracity of the PI(4,5)P2 site.206208

Interestingly, this site shows excellent agreement with an AI model we predicted for this review using the Chai-1 web interface57 (https://lab.chaidiscovery.com/), and only the primary sequence of the Kir2.2 monomer and the SMILES string for PI(4,5)P2 as input (Figure 5D, G). The authors of this model report similar capabilities to AlphaFold3, and the web tool places no limitations on small molecule types which may be assessed. The emerging ability of AI models to predict protein structures in complex with small molecules, including lipids, is likely to be a major accelerator for phosphoinositide research, providing initial coordinate sets from which to launch dynamical physics-based simulation and experimental investigations.

The functional consequences of PI(4,5)P2 binding for gating and Kir channel conductance have been explored in a range of simulations works.209213 Tour de force simulations from the D. E. Shaw Research group recently further defined the mechanism of PI(4,5)P2 activation of Kir2.2 in atomic detail.214 Starting from a closed-state structure,17 Jogini et al., introduced a previously characterised mutation at the bundle crossing to obtain an open conducting state, capture key conformational changes at the main activation gate and cytoplasmic domain, and demonstrate the dynamical basis for PI(4,5)P2 stabilisation of the open-state in all-atom detail (Figure 6). A particularly notable feature of this work is the degree of sampling achieved, with timescales of up to 200 μs reached for an individual simulation, and on aggregate over a millisecond across all simulations. Despite this immense sampling, the authors note that the presence of PI(4,5)P2 and bulk PS failed to evoke pore-opening of the WT channel, as would be expected in the event of a sufficiently accurate force field coupled to sufficient sampling. This suggests that while substantial progress continues to be made, there remains scope for development of all-atom MD in capturing certain mechanistic aspects of phosphoinositide biology occurring on long timescales.

Figure 6. Atomic architecture of a PI(4,5)P2 binding site on Kir2.2.214.

Figure 6

(A) Cross-section of an ANTON simulation snapshot showing a PI(4,5)P2 lipid (magenta) binding site on tetrameric Kir2.2 (beige) embedded in a lipid bilayer (blue surface). An adjacent PS lipid is shown in green. (B) Zoomed in view on the PI(4,5)P2 site showing how the headgroup is cradled by basic residues during the simulation. K+ ions within the selectivity filter are shown in the background as blue, yellow, and orange spheres. We thank Dr. Vishwanath Jogini, Dr. Morten Ø. Jensen, and D. E. Shaw Research for providing coordinates used to produce the figure.

Further studies have leveraged MD simulations to assess the molecular detail of phosphoinositide interaction in the context of Kv channels,215221 and K2P channels.222,223 Other examples of phosphoinositides controlling non-K+ ion channel structure and function include the calcium-dependent chloride ion channel TMEM16A/ANO1,224226 mechanosensitive channels such as Piezo1,227 and voltage-gated sodium channels.228 A particularly notable example is the inositol trisphosphate receptor (IP3R), where PI(4,5)P2 primes the receptor by partially occupying the orthosteric IP3-binding site.229

Peripheral membrane proteins

Peripheral membrane proteins (PMPs) exist largely in aqueous solution and interact with the hydrophilic surface of membranes. This is a particularly important group of proteins in relation for phosphoinositide biology, as it encompasses the kinases230 and phosphatases231 which synthesise and degrade the various phosphoinositide species and thereby regulate their relative abundance and localisation in membranes.232 The interactions of PMPs with phosphoinositides is also intimately involved in intracellular signalling cascades, for instance in the enzymatic cleavage of PI(4,5)P2 into the second messengers diacylglycerol and inositol-trisphosphate (IP3).233

Interactions of PMPs with membranes may take a variety of forms. These include binding of a specific lipid headgroup to a specific site on the PMP, more dynamic interactions between patches of complementary charge on the protein surface and the net anionic charge of the inner leaflet, and by insertions of hydrophobic loops or lipidic post translational modifications, e.g. myristoylation.234 Structure determination of PMPs in solution or indeed bound to individual lipids is now relatively routine. However, the main biological function of PMPs takes place on the surface of membranes. Stabilising PMPs on membranes to the level required for atomic level structure determination remains challenging.235 Simulations have played a critical role in reuniting experimentally determined structures of PMPs with phosphoinositide-containing lipid membranes in silico to determine their lipid interactions, binding orientation and dynamics. This is particularly important as the biologically relevant membrane context can influence the conformation of the PMP, and reveal new lipid binding sites.236

Extensive all-atom simulations have revealed the molecular basis for how PI(3,4,5)P3 drives binding of Bruton’s tyrosine kinase (Btk) to membranes and allosterically stabilises the experimentally determined dimeric form of the protein.237 Starting from two copies of the monomeric form of Btk in solution, this study was able to simulate spontaneous binding to PI(3,4,5)P3 membranes and Btk dimerization. This impressive feat was achieved through use of the ANTON supercomputer77 coupled to aflavour of enhanced sampling simulation termed ‘tempered binding’.238

Simulations may also be employed alongside complementary biophysical data to pre-orientate a given PMP on the membrane. Thus in early works in this area Lai et al., used electron paramagnetic resonance (EPR) spectroscopy data to obtain approximate geometries for the C2 domain of Protein kinase Cα relative to PI(4,5)P2 containing membranes.239 These geometries were used as starting points for simulations, thereby decreasing the level of sampling required compared to de novo binding from solution. This approach enabled refinement of C2 domain orientation, determination of lipid:protein stoichiometry, and a molecular level view of key PI(4,5)P2 interactions. This same approach was subsequently applied to reveal the molecular basis of stereospecific binding of the GRP1 PH domain to PI(3,4,5)P3.240

All-atom simulations of PI(4,5)P2 and PI(3,4,5)P3 binding to PMPs have also been performed241244 using the highly mobile membrane mimetic (HMMM) model.92 This approach replaces the hydrophobic core of the lipid bilayer with organic solvent, atop which phosphoinositide headgroups with shortened acyl tails diffuse. The net effect is to increase the rates of lipid diffusion and therefore enhance the rate at which they explore possible interaction modes with the PMP.245 In a recent example of the HMMM approach, Pant and colleagues addressed the orientational dynamics of Arf1 GTPase on PI(4,5)P2 membranes241 (Figure 7). This particular PMP is localised to the membrane via a myristoylated amphipathic N-terminal helix. The increased diffusion rates provided by the HMMM model were leveraged to identify three principle orientational states of myristoylated Arf1 on membranes with different lipid compositions, which were subsequently refined with microsecond all-atom simulations with full lipid representation. Interactions with PI(4,5)P2 were key for driving full membrane association and population of these three states, in qualitative agreement with neutron reflectometry (NR) and NMR experiments.241

Figure 7. Arf1 membrane interactions.241.

Figure 7

(A) Structure of myristoylated Arf1 and short-chain PC/PI(4,5)P2. (B) Simulation snapshot showing a cross-section of Arf1 bound to a PC/PI(4,5)P2 highly mobile membrane mimetic (HMMM) containing a DCLE organic solvent core. Interactions of Arf1 with PI(4,5)P2. We thank Dr. Shashank Pant and Prof. Emad Tajkhorshid for providing coordinates used to render the figure.

High-throughput CG simulations have been used to assess patterns of phosphoinositide interaction across whole families of pH domain-containing PMPs.236,246 Yamamoto and colleagues examined the spontaneous binding of the apo forms of 13 different PH domains to PI(3,4,5)P3 and PI(4,5)P2 membranes.236 Overlay of the simulation derived phosphoinositide interactions showed good agreement with available experimental X-ray crystallographic and NMR structures of the phosphoinositide bound forms of these proteins. In addition to the experimentally known interactions, the simulations also revealed a number of new non-canonical phosphoinositide binding sites for select PH domains. Mutation of key residues within these sites allowed delineation of the main contributors to phosphoinositide binding. For all PH domains, a degree of less specific phosphoinositide clustering around the protein was also indicated to contribute to membrane association. This was consistent with observations from large scale biochemical assessment of 91 PH domains,247 and was later supported by free energy calculations for the GRP1 PH domain interacting with PI(3,4,5)P3 clusters of varying size,248 and by a herculean body of work to extend analysis to 100 different mammalian PH domains.249

Further simulation works delineating phosphoinositide interaction with PMPs include on Vinculin,250 the Dok7 PH domain,251 the K-Ras G domain,252 KRas4b,244 KRas4b and RAF1,253 PLD2,254 focal adhesion kinase,255,256 PTEN and SHIP2 C2 domains,257 PTEN and voltage sensitive phosphatase,258 Arf/GEF,259 PLCγ1,260 Gelsolin,261 ORP/Osh proteins,262264 GRP1 PH domain,265 murine kindlin-3 PH domain,266 BIN1/M-Amphiphysin2,267 lymphocyte specific kinase (LCK),268 Smurf1 C2 domain,269 StarD4,270,271 SNX11,272 WD repeat domain phosphoinositide-interacting protein 2 (WIPI2),273 tubby domains,274 and phosphatidylinositol-transfer proteins.275 The number of different research groups turning to simulations to interrogate the molecular basis of how their favourite PMP interacts with phosphoinositides is indicative of both the need in this area, and the power of molecular simulation in meeting that need. Notably, while most works to date have dealt with mammalian PMP interaction with PI(4,5)P2 and PI(3,4,5)P3 lipids, simulations have also shown utility in exploring other phosphoinositide species in non-mammalian contexts, such as PI(4)P interactions with Plasmodium vivax Perforin-Like Proteins (PLPs).276 Future CG simulations of PMPs are likely to benefit from the GoMartini model of protein structure, which has recently been reported to further enhance the accuracy of conformational modelling of PLCδ1 interacting with PI(4,5)P2 bilayers.277 For a recent excellent review on the generalized application of molecular dynamics simulations to PMPs broadly.278

Receptor tyrosine kinases

Many fundamental cellular signals are transduced across the plasma membrane by human receptor tyrosine kinases (RTKs).279 This family of 58 cell surface receptors share a common molecular architecture consisting of an extracellular ligand binding ectodomain, a single-pass transmembrane (TM) helix, aflexible juxtamembrane (JM) region, and an intracellular protein kinase domain. A variety of additional soluble domains may decorate different types of RTK leading to substantial variety.280 Binding of ligands to the ectodomain canonically leads to receptor dimerization, activation of protein kinase activity, and sets in motion a variety of intracellular signalling cascades.281 Non-canonical forms of signalling by heterodimerisation of RTKs are also possible.282 Phosphoinositides have been shown experimentally to directly bind a range of RTKs,283 modulate their function,34 as well as control their localisation into high order multimers at the cell surface.284286

As major therapeutic targets287,288 (e.g. Herceptin against the Erbb2 RTK in the treatment of breast cancer) RTKs have been paid particular attention by structural biologists and atomic resolution structure availability is relatively good for select RTKs289291. However intrinsically disordered regions such as the JM region remain more mysterious, despite their known functional importance.292 This is an area where computational approaches are increasingly making impact.193 Simulations of the TM-JM constructs of all 58 human RTKs have predicted nanoclustering of PI(4,5)P2 around the disordered JM189 (Figure 8A, B). This behaviour was driven via direct interaction with a conserved polybasic sequence motif found in the region of the disordered JM immediately adjacent to the TM helix. More detailed work focussing on single RTKs including the epidermal growth factor receptor (EGFR),293 the tropomyosin receptor kinase A (TrkA),294 and the ephrin A2 receptor (EphA2R)295 identified similar JM driven PI(4,5)P2 interactions. Most recently a combined simulation and NMR study observed this phenomenon for the prolactin receptor JM.190 Such nanoclustering of PI(4,5)P2 lipids around individual receptors is especially pertinent when it is considered that PI(4,5)P2 is a key second messenger substrate in many of the downstream signal cascades which are activated by RTKs. Fluorescence microscopy approaches have indicated that a variety of RTKs are localised into higher order arrays at the cell surface.284,285,296,297 If each member of this array has its own local pool of PI(4,5)P2 lipids around it, as predicted in the simulations, this would overall add up to a high local concentration of phosphoinositide lipids, with potential implications for signal transduction efficiency and PMP recruitment. CG simulations of large membranes containing tens of full-length RTKs are now feasible, and are expected to shed further light on the molecular architecture of these clusters (Figure 8C, D).

Figure 8. Multiscale simulations of the human RTK superfamily.

Figure 8

(A) TM-JM models of all 58 known human RTKs, built from NMR data and secondary structure prediction programs and simulated in 189. (B) A nanocluster of PI(4,5)P2 lipids around the TM-JM monomer of EGFR/Erbb1. Image is the endpoint of a 1 μs CG simulation followed by 50 ns of all-atom simulation. (C) Full-length model of the EGFR/Erbb1 dimer built from NMR, X-ray and modelling data. (D) 49 full-length EGFR/Erbb1 dimers embedded in a patch of membrane with a lipid composition approximating that of a mammalian plasma membrane, including inner leaflet PI(4,5)P2 and PI(3,4,5)P3. The image shown is after 8 μs of simulation. Hedger et al., unpublished.

Simulations have also been used to compute the change in free energy for lipid binding to the EGFR TM-JM dimer within PC bilayers.298 As expected, PI(4,5)P2 binding is significantly more favourable than for other plasma membrane phospholipid species. Notably, the ΔΔG values are of a magnitude similar to those obtained for TM helix dimerization for a range of RTKs.299,300 This would seem to lend credence to the idea that one mechanism of PI(4,5)P2 modulation of RTK activity may be via direct competition with helices and/or modulation of TM dimer crossing angle.301,302 Barrera and colleagues subsequently tested this concept experimentally for the EphA2R and identified a marked effect of PI(4,5)P2 in promoting dimerization of a particular crossing angle mode.303 This is an exciting result which positions helix dimerization as yet another potential microswitch304 in phosphoinositide biology. Factors which will be important to address experimentally going forward include the influence of soluble domains, competition between different phosphoinositide species, and the universality or otherwise amongst different RTKs and indeed other single-pass membrane proteins.

Other ways in which phosphoinositides influence RTKs include in controlling how the intracellular kinase domains interact with the inner leaflet surface, and recruitment of RTK growth factor ligands to membranes. Chavent et al., used CG simulations to capture the spontaneous binding of the kinase domain to the membrane surface.295 The conceptual framework bears resemblance to the PMP – phosphoinositide simulations described in Section ‘Peripheral membrane proteins’, albeit in this case with the kinase domain tethered to the membrane by theflexible JM and TM helix (Figure 9). By coupling the simulations to liposome pull-down assays and prior modelling of EphA2R, a key role for PI(4,5)P2 was identified in orientating the kinase domains for trans-autophosphorylation within EphA2R clusters. Most recently, all-atom simulations have been used to examine heterodimeric EphA2R-EGFR kinase domain interactions with PI(4,5)P2.285 This study identified intriguing differences between EphA2R and EGFR, in the context of wider experimental work which demonstrated PI(4,5)P2 promoted both hetero and homomultimerization of EphA2R and EGFR.

Figure 9. Simulations identify orientations of the EphA2R kinase domain on PI(4,5)P2 containing membranes.

Figure 9

(A) Schematic of start and endpoint of simulations. (B) Representative snapshot of interface 1 identified in 295. The kinase domain (yellow) is shown interacting with a cluster of PI(4,5)P2 lipids (magenta). The JM is located within the cluster. We thank Dr. Matthieu Chavent for the coordinates used to produce the figure.

Phosphoinositide species also have key roles in the biology of certain RTK ligands, such as the fibroblast growth factor 2 (FGF2). FGF2 comes into contact with PI(4,5)P2 lipids in the inner leaflet during its secretion by the type I ‘unconventional protein secretion (UPS)’ pathway.305 PI(4,5)P2 interaction induces a fascinating piece of biology in which FGF2 undergoes oligomerisation to form a membrane pore, before exiting the other side of the membrane to arrive in the extracellular space.306 Simulations have been applied to capture PI(4,5)P2 interactions with FGF2 during this process.99,307 Most recently Lolicato et al. combined microsecond atomistic simulations with in vitro and in cellulo assays to address lipid synergy during the initial membrane recruitment of FGF2.99 The simulations identified the atomic basis of an intricate lipid interplay in which cholesterol interacted with PI(4,5)P2 lipids to increase the inositol headgroup visibility to FGF2 in a manner which resulted in faster binding kinetics and more stable FGF2 – PI(4,5)P2 complexes. This is an important finding for phosphoinositide biology generally, as it highlights the consequences of lipid-lipid interplay and complexity for certain phosphoinositide molecular interactions.

Other membrane proteins

The phosphoinositide interactions of a variety of other membrane proteins have been addressed by simulations. Examples of transporters include PI(4,5)P2 modulation of the structure and function of the dopamine transporter,37,308 solute carrier spinster homolog 2 (SPNS2),309 and SLC4 family transporters.310312 PI(3,5)P2 mediated oligomerisation of the endosomal sodium-proton exchanger NHE9,313 and phosphatidylinositol stabilisation of purine symporter UapA dimers314 have also been addressed with simulations. The influence of phosphoinositides on membrane proteins involved in viral biology,315321 cytoskeleton biology,322324 vesicle trafficking,325 autophagy,273 and pore formation during programmed cell death,326 have also been studied.

Phosphoinositide clustering

Beyond specific membrane proteins, simulation studies have also addressed the ability of certain phosphoinositides to self-interact and cluster into areas of enrichment in the absence of membrane proteins.85,327330 A key role for divalent calcium has been identified in bridging interactions between the phosphate groups of adjacent PI(4,5)P2 lipids.331 This effect did not extend to other common inorganic cations nor to PI(3,5)P2 lipids,332 in qualitative agreement with trends seen influorescence-based spectroscopy and microscopy experiments.333 This is particularly interesting in a cell biology context given the role of calcium signalling downstream of many of the signalling cascades in which PI(4,5)P2 lipids act as key PMP recruitment factors and second messenger substrates.283 Indeed, direct competition between calcium and PMPs for interaction with PI(4,5)P2 has been reported334. Dynamic modulation of local calcium levels could represent another means of controlling how phosphoinositide species interact with membrane proteins. Acyl tail saturation has also been seen to play a key role in modulating nanocluster formation in CG simulations.157

Nanocluster sizes may be increased substantially by the presence of proteins. In a landmark study, van den Bogaart and colleagues addressed the architecture of syntaxin-1A – PI(4,5)P2 clusters using super-resolution microscopy and CG simulations.38 Microdomains of up to 73 nm in size were experimentally identified in which PI(4,5)P2 was the dominant lipid species. Use of phosphatases to knockdown cellular PI(4,5)P2 levels is a common trick in phosphoinositide biology, and was applied in this case to demonstrate co-clustering of syntaxin-1A was dependent on PI(4,5)P2. Simulations were run in which 64 copies of the transmembrane helix of syntaxin-1A were embedded into a PI(4,5)P2:PC:PS membrane. The helices and PI(4,5)P2 lipids were seen to spontaneously separate into microdomains knitted together by ionic interactions between negative PI(4,5)P2 phosphate groups and basic residues located at the tips of the syntaxin-1A helices, similar to later observations made for RTKs189 and the prolactin receptor.190 These microdomains remained stable on simulation timescales of up to 6 μs, and could be recreated in vitro within giant unilamellar vesicle membranes. Notably, in this case the effects were robust to changes in lipid composition including of cholesterol, which has been seen to influence PI(4,5)P2 interactions in other contexts.99 Simulations have also played a key role in delineating how phosphoinositide – membrane protein interactions can lead to local remodelling of membranes. BAR domains in particular have been the subject of attention in this area.267,335338 Phosphoinositide interactions with a variety of other remodelling proteins including epsin N-terminal homology domain (ENTH),339 cavin1,340 and AP180 N-terminal homology domain (ANTH)341 have also been studied.

Summary and Concluding Remarks

Simulations have been applied in a variety of ways to address the molecular underpinnings of how phosphoinositides influence membrane proteins (Figure 10). By comparison to experiment it is evident they are capable of accurately identifying phosphoinositide binding sites. This was the case for Class A GPCRs21 and Kir channels,84 as confirmed by later X-ray17 and cryo-EM140 structures. Simulations may also be used synergistically to interpret ambiguous experimental lipid density in structures.23 Wang et al., thus used CG-simulations to identify a general phosphoinositide binding site on the PC2 ion channel, in agreement with their cryo-EM structures.181 This general phosphoinositide site was later refined to a possible PI(4)P site in a subsequent biochemical and cell-based functional study.182 Characterisation of the energetics of phosphoinositide interaction may be achieved using advanced simulation techniques such as FEP-derived ΔΔGs for lipid binding,199 and PMF-derived free energy landscapes for PMP association with membranes from solution.248 Analysis of the interaction patterns in simulations of different conformational states can show in atomic detail how functional effects are conferred by structure, for instance in PI(4,5)P2 stabilisation of the open state of Kir2.2.199,214 Function may also be prospectively predicted from the structure and dynamics seen in a simulation, for instance in simulation-based prediction that PI(4,5)P2 lipids act as a molecular glue to bridge A2AR – miniG interactions to enhance G-protein coupling, in agreement with nMS andfluorescence-based functional assays.21

Figure 10. Cellular roles of phosphoinositide lipids studied by simulations.

Figure 10

(A) Combined CG MD simulation and nMS based discovery of GPCR binding and functional modulation of G-protein coupling by PI(4,5)P2.21 (B) Multiscale simulation based identification and characterisation of PI(4,5)P2 binding sites on Kir channels.84,199,206,214 (C) CG MD simulation identification of a phosphoinositide binding site on PC2 with functional consequence.181,182 (D) Multiscale simulation of PI(4,5)P2 control of RTK kinase domain orientation, and TM-JM dynamics.189,293,295,303 (E) CG, all-atom, and HMMM simulation of PI(3)P, PI(4)P, PI(4,5)P2, and PI(3,4,5)P3, interaction with many PMPs: see section 2.3. (F) All-atom simulations of PI(4,5)P2 in FGF2 secretion.99,307 (G) All-atom simulations of cholesterol – PI(4,5)P2 interplay.99 (H) Multiscale simulations of phosphoinositide clustering.331,332,85,99,334 (I) Multiscale simulation of PI(4,5)P2 interaction with IDRs within receptors189 and channels.191 (J) Combined CG MD simulation and microscopy of phosphoinositide microdomain formation.38,189,284 (K) Multiscale and ultra-CG simulations of PI(4,5)P2 interaction with remodelling proteins.267,335,336,338,339

Increases in computational power110 and structure availability18,143 increasingly allow assessment of phosphoinositide interactions at scale across large groups of membrane proteins, for instance RTKs,189 GPCRs,149 and PH domains,249 facilitating assessments of conservation. Expansion of parameter space for phosphoinositides,83 and membrane lipids generally69,86,95,342 now allow routine simulation in complex lipid mixtures, which have revealed new aspects of phosphoinositide biology deriving from lipid-lipid interplay.99,199 The vast majority of simulation studies to date have focussed on PI(4,5)P2, which likely derives from its localisation in the plasma membrane, where membrane protein research has tended to focus. It will be especially welcome to see phosphoinositide research expand further into subcellular membranes containing other phosphoinositide species.

Existing limitations of phosphoinositide simulations, such as fixed charge force fields, and challenges in distinguishing closely related isomers e.g. PI(3,5)P2 and PI(4,5)P2 in CG models181 have also been seen. Continued developments in multiscale modelling,74 and constant pH simulations343,344 will be especially valuable in this area. The CG MARTINI model has shown exceptional utility in de novo prediction of phosphoinositide binding sites,64,72 (converged) free energy calculations,108,199,345 and modelling of large membrane phenomena.38,166 At the same time, we would caution the budding explorer of phosphoinositide biology against the risk of overinterpretation of CG models. For instance, in capturing the detailed pattern of phosphoinositide interactions and ensemble of poses within a given binding site, certain ion-induced effects,85 and differentiation between closely related isomers. Moreover, high-quality (ideally not detergent or chemical probe based) experimental measures of specific lipid interaction energies with membrane proteins are scarce, which makes direct comparison to CG interaction energies challenging.108 All-atom simulations (includingflavours such as HMMM92,245) have tended to excel in areas such as refinement of detailed interactions with known binding sites,206,237 specific ion interactions,85,329,332 and phosphoinositide-induced membrane protein conformational change.214,242 Ultra-CG models346 have also shown utility in modelling of large-scale phenomena such as membrane protrusion formation by I-BAR domains.335

In considering the present body of phosphoinositide simulations it is evident that they interact with a huge range of membrane proteins. Unsurprisingly these interactions nearly always feature basic residues on the protein. In the case of integral membrane proteins, these residues tend to be positioned slightly beyond the membrane surface, and are thus optimally positioned to interact with the 3′, 4′, and 5′ phosphoryl groups of phosphoinositides. Such positions are often inaccessible to other phospholipids with smaller headgroups and this is one way specificity is conferred. Given the body of simulation work on phosphoinositides, it should now be possible to derive more detailed heuristics on what characterises a phosphoinositide binding site. Can a suite of pharmacophore and QSAR models for phosphoinositides be developed? Can machine learning technologies be trained on the abundance of simulation and structural data on phosphoinositide interactions with membrane proteins? Leveraging simulation data in this way is a key challenge for computational biochemists. At present simulation data is generally stored individually within the filesystems of research groups, which severely circumscribes large scale meta-analysis of this nature. Upcoming initiatives such as the Molecular Dynamics Data Bank (MDDB)347 (see https://mddbr.eu/) and MDverse104 will be welcome advancements and should enable new insights into phosphoinositide biology. Such analysis would be especially welcome from a proof-of-concept therapeutic perspective, as a principal challenge to leveraging lipid binding sites on membrane proteins for structure-based drug design348,349 is likely to be selectivity.

The advent of AI models for biomolecular structure prediction is expected to further expand the range of questions which can be addressed by physics-based simulations. The widespread availability of increasingly accurate structure predictions of proteins and their conformations will provide new entry points from which to initiate dynamical simulations. This effect will be felt particularly for simulations of membrane proteins, where experimental structure determination remains challenging compared to soluble proteins.350 Recent developments in AI models which predict small molecules in complex with target proteins54,56,57 are anticipated to be especially valuable in rapid prediction of phosphoinositide binding sites on membrane proteins, and will serve to provide additional points from which to launch detailed simulation and experimental studies. We also note the evolving ability of cryo-EM to capture time-resolved conformational heterogeneity,351,352 offering the intriguing possibility for further simulation – experimental synergy in a temporal dimension.

Simulations have grown to occupy a unique place in the toolkit available to the interested connoisseur of phosphoinositide biology. Their synergistic application hand in glove with rigorous structural and functional experiment is likely to continue to reveal new facets of the molecular underpinnings of the cellular effects of phosphoinositides.

Summary.

Hedger and Yen review developments in simulations of phosphoinositides and membrane proteins.

Acknowledgements

G. H. is funded by the Wellcome Trust (301619/Z/23/Z). Research in H-Y. Y.’s group is funded by Academia Sinica (AS-GCP-113-L01). Dr. Anna Duncan, Dr. Shashank Pant, Prof. Emad Tajkhorshid, Dr. Matthieu Chavent, Dr. Vishwanath Jogini, Dr. Morten Ø. Jensen, and D. E. Shaw Research are thanked for providing simulation structures used to render figures. Prof. Mark S. P. Sansom is thanked for valuable discussions.

Abbreviations

MD

molecular dynamics

CG

coarse-grained

cryo-EM

cryogenic electron microscopy

nMS

native mass spectrometry

AI

artificial intelligence

ML

machine learning

NMR

nuclear magnetic resonance

ESR

electron spin resonance

EM

electron microscopy

PC

phosphatidylcholine

PG

phosphatidylglycerol

PS

phosphatidylserine

PE

phosphatidylethanolamine

PI

phosphatidylinositol

PI(4)P

phosphatidylinositol-4-phos phate

PI(4,5)P2

phosphatidylinositol-4,5-bisphosphate

PI(3,4,5)P3

phosphatidylinositol-3,4,5-trisphosphate

FEP

free energy perturbation

PMF

potential of mean force

GPCR

G-protein-coupled receptor

NTSR1

neurotensin receptor 1

A2AR

Adenosine2A receptor

RTK

receptor tyrosine kinase

JM

juxtamembrane

TRP

transient receptor potential

PC2

polycystin-2

Kir

inward rectifying potassium ion channel

PMP

peripheral membrane protein

Footnotes

CRediT authorship contribution statement

George Hedger: Writing – review & editing, Writing – original draft, Visualization, Validation, Supervision, Software, Resources, Project administration, Investigation, Formal analysis, Conceptualization. Hsin-Yung Yen: Writing – review & editing, Writing – original draft, Formal analysis.

Declaration of Competing Interest

The authors declare the following financial interests/personal relationships which may be considered as potential competing interests: “H-Y. Y. was previously an employee and is a founder of OMass Therapeutics.”.

References

  • 1.Balla T. Phosphoinositides: tiny lipids with giant impact on cell regulation. Physiol Rev. 2013;93:1019–1137. doi: 10.1152/physrev.00028.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Agranoff BW. Turtles All the Way: Reflections on myo-Inositol. J Biol Chem. 2009;284:21121–21126. doi: 10.1074/jbc.X109.004747. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Irvine RF. 20 years of Ins(1,4,5)P3, and 40 years before. Nature Rev Mol Cell Biol. 2003;4:586–590. doi: 10.1038/nrm1152. [DOI] [PubMed] [Google Scholar]
  • 4.Michell B. Early steps along the road to inositol-lipid-based signalling. Trends Biochem Sci. 1995;20:326–329. doi: 10.1016/s0968-0004(00)89061-1. [DOI] [PubMed] [Google Scholar]
  • 5.Hokin LE. The road to the phosphoinositide-generated second messengers. Trends Pharmacol Sci. 1987;8:53–56. doi: 10.1016/0165-6147(87)90009-5. [DOI] [Google Scholar]
  • 6.Jensen JB, Falkenburger BH, Dickson EJ, de la Cruz L, Dai G, Myeong J, Jung S-R, Kruse M, Vivas O, Suh B-C, Hille B. Biophysical physiology of phosphoinositide rapid dynamics and regulation in living cells. J Gen Physiol. 2022;154:e202113074. doi: 10.1085/jgp.202113074. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Folch J, Woolley DW. Inositol, a constituent of a brain phosphatide. J Biol Chem. 1942;142:963–964. doi: 10.1016/S0021-9258(18)45096-X. [DOI] [Google Scholar]
  • 8.Wills RC, Hammond GRV. PI(4,5)P2: signaling the plasma membrane. Biochem J. 2022;479:2311–2325. doi: 10.1042/BCJ20220445. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Marat AL, Haucke V. Phosphatidylinositol 3-phosphates—at the interface between cell signalling and membrane traffic. EMBO J. 2016;35:561–579. doi: 10.15252/embj.201593564. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Yin HL, Janmey PA. Phosphoinositide regulation of the actin cytoskeleton. Annu Rev Physiol. 2003;65:761–789. doi: 10.1146/annurev.physiol.65.092101.142517. [DOI] [PubMed] [Google Scholar]
  • 11.Suh B-C, Hille B. Regulation of ion channels by phosphatidylinositol 4,5-bisphosphate. Curr Opin Neurobiol. 2005;15:370–378. doi: 10.1016/j.conb.2005.05.005. [DOI] [PubMed] [Google Scholar]
  • 12.Bielas SL, Silhavy JL, Brancati F, Kisseleva MV, Al-Gazali L, Sztriha L, Bayoumi RA, Zaki MS, Abdel-Aleem A, Rosti RO, Kayserili H, et al. Mutations in INPP5E, encoding inositol polyphosphate-5-phosphatase E, link phosphatidyl inositol signaling to the ciliopathies. Nature Genet. 2009;41:1032–1036. doi: 10.1038/ng.423. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Samuels Y, Wang Z, Bardelli A, Silliman N, Ptak J, Szabo S, Yan H, Gazdar A, Powell SM, Riggins GJ, Willson JKV, et al. High frequency of mutations of the PIK3CA gene in human cancers. Science. 2004;304:554. doi: 10.1126/science.1096502. [DOI] [PubMed] [Google Scholar]
  • 14.Posor Y, Jang W, Haucke V. Phosphoinositides as membrane organizers. Nature Rev Mol Cell Biol. 2022;23:797–816. doi: 10.1038/s41580-022-00490-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Hammond GRV, Burke JE. Novel roles of phosphoinositides in signaling, lipid transport, and disease. Curr Opin Cell Biol. 2020;63:57–67. doi: 10.1016/j.ceb.2019.12.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Burke JE, Triscott J, Emerling BM, Hammond GRV. Beyond PI3Ks: targeting phosphoinositide kinases in disease. Nature Rev Drug Discov. 2023;22:357–386. doi: 10.1038/s41573-022-00582-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Hansen SB, Tao X, MacKinnon R. Structural basis of PIP2 activation of the classical inward rectifier K+ channel Kir2.2. Nature. 2011;477:495–498. doi: 10.1038/nature10370. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Cheng Y. Membrane protein structural biology in the era of single particle cryo-EM. Curr Opin Struct Biol. 2018;52:58–63. doi: 10.1016/j.sbi.2018.08.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Laganowsky A, Reading E, Allison TM, Ulmschneider MB, Degiacomi MT, Baldwin AJ, Robinson CV. Membrane proteins bind lipids selectively to modulate their structure and function. Nature. 2014;510:172–175. doi: 10.1038/nature13419. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Gao Y, Cao E, Julius D, Cheng Y. TRPV1 structures in nanodiscs reveal mechanisms of ligand and lipid action. Nature. 2016;534:347–351. doi: 10.1038/nature17964. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Yen H-Y, Hoi KK, Liko I, Hedger G, Horrell MR, Song W, Wu D, Heine P, Warne T, Lee Y, Carpenter B, et al. PtdIns(4,5)P2 stabilizes active states of GPCRs and enhances selectivity of G-protein coupling. Nature. 2018;559:423–427. doi: 10.1038/s41586-018-0325-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Xu P, Huang S, Zhang H, Mao C, Zhou XE, Cheng X, Simon IA, Shen D-D, Yen H-Y, Robinson CV, Harpsøe K, et al. Structural insights into the lipid and ligand regulation of serotonin receptors. Nature. 2021;592:469–473. doi: 10.1038/s41586-021-03376-8. [DOI] [PubMed] [Google Scholar]
  • 23.Ansell TB, Song W, Coupland CE, Carrique L, Corey RA, Duncan AL, Cassidy CK, Geurts MMG, Rasmussen T, Ward AB, Siebold C, et al. LipIDens: simulation assisted interpretation of lipid densities in cryo-EM structures of membrane proteins. Nature Commun. 2023;14:7774. doi: 10.1038/s41467-023-43392-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Souza PCT, Alessandri R, Barnoud J, Thallmair S, Faustino I, Grünewald F, Patmanidis I, Abdizadeh H, Bruininks BMH, Wassenaar TA, Kroon PC, et al. Martini 3: a general purpose force field for coarse-grained molecular dynamics. Nature Methods. 2021;18:382–388. doi: 10.1038/s41592-021-01098-3. [DOI] [PubMed] [Google Scholar]
  • 25.Marrink SJ, Corradi V, Souza PCT, Ingólfsson HI, Tieleman DP, Sansom MSP. Computational modeling of realistic cell membranes. Chem Rev. 2019;119:6184–6226. doi: 10.1021/acs.chemrev.8b00460. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Harayama T, Riezman H. Understanding the diversity of membrane lipid composition. Nature Rev Mol Cell Biol. 2018;19:281–296. doi: 10.1038/nrm.2017.138. [DOI] [PubMed] [Google Scholar]
  • 27.Han X, Gross RW. The foundations and development of lipidomics. J Lipid Res. 2022;63 doi: 10.1016/j.jlr.2021.100164. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Shevchenko A, Simons K. Lipidomics: coming to grips with lipid diversity. Nature Rev Mol Cell Biol. 2010;11:593–598. doi: 10.1038/nrm2934. [DOI] [PubMed] [Google Scholar]
  • 29.van Meer G, Voelker DR, Feigenson GW. Membrane lipids: where they are and how they behave. Nature Rev Mol Cell Biol. 2008;9:112–124. doi: 10.1038/nrm2330. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Casares D, Escribá PV, Rosselló CA. Membrane lipid composition: effect on membrane and organelle structure, function and compartmentalization and therapeutic avenues. Int J Mol Sci. 2019 doi: 10.3390/ijms20092167. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Levental I, Lyman E. Regulation of membrane protein structure and function by their lipid nano-environment. Nature Rev Mol Cell Biol. 2023;24:107–122. doi: 10.1038/s41580-022-00524-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Byrne EFX, Sircar R, Miller PS, Hedger G, Luchetti G, Nachtergaele S, Tully MD, Mydock-McGrane L, Covey DF, Rambo RP, Sansom MSP, et al. Structural basis of Smoothened regulation by its extracellular domains. Nature. 2016;535:517–522. doi: 10.1038/nature18934. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Deshpande I, Liang J, Hedeen D, Roberts KJ, Zhang Y, Ha B, Latorraca NR, Faust B, Dror RO, Beachy PA, Myers BR, et al. Smoothened stimulation by membrane sterols drives Hedgehog pathway activity. Nature. 2019;571:284–288. doi: 10.1038/s41586-019-1355-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Michailidis IE, Rusinova R, Georgakopoulos A, Chen Y, Iyengar R, Robakis NK, Logothetis DE, Baki L. Phosphatidylinositol-4,5-bisphosphate regulates epidermal growth factor receptor activation. Pflügers Archiv – Eur J Physiol. 2011;461:387–397. doi: 10.1007/s00424-010-0904-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Tucker SJ, Baukrowitz T. How highly charged anionic lipids bind and regulate ion channels. J Gen Physiol. 2008;131:431–438. doi: 10.1085/jgp.200709936. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Anderluh A, Hofmaier T, Klotzsch E, Kudlacek O, Stockner T, Sitte HH, Schütz GJ. Direct PIP2 binding mediates stable oligomer formation of the serotonin transporter. Nature Commun. 2017;8:14089. doi: 10.1038/ncomms14089. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Hamilton PJ, Belovich AN, Khelashvili G, Saunders C, Erreger K, Javitch JA, Sitte HH, Weinstein H, Matthies HJG, Galli A. PIP2 regulates psychostimulant behaviors through its interaction with a membrane protein. Nature Chem Biol. 2014;10:582–589. doi: 10.1038/nchembio.1545. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.van den Bogaart G, Meyenberg K, Risselada HJ, Amin H, Willig KI, Hubrich BE, Dier M, Hell SW, Grubmüller Diederichsen U, Jahn R. Membrane protein sequestering by ionic protein–lipid interactions. Nature. 2011;479:552–555. doi: 10.1038/nature10545. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Cho W, Stahelin RV. Membrane-protein interactions in cell signaling and membrane trafficking. Annu Rev Biophys. 2005;34:119–151. doi: 10.1146/annurev.biophys.33.110502.133337. [DOI] [PubMed] [Google Scholar]
  • 40.Lemmon MA. Membrane recognition by phospholipid-binding domains. Nature Rev Mol Cell Biol. 2008;9:99–111. doi: 10.1038/nrm2328. [DOI] [PubMed] [Google Scholar]
  • 41.Slochower DR, Huwe PJ, Radhakrishnan R, Janmey PA. Quantum and all-atom molecular dynamics simulations of protonation and divalent ion binding to phosphatidylinositol 4,5-bisphosphate (PIP2) J Phys Chem B. 2013;117:8322–8329. doi: 10.1021/jp401414y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Barneda D, Cosulich S, Stephens L, Hawkins P. How is the acyl chain composition of phosphoinositides created and does it matter? Biochem Soc Trans. 2019;47:1291–1305. doi: 10.1042/BST20190205. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Hernandez CC, Falkenburger B, Shapiro MS. Affinity for phosphatidylinositol 4,5-bisphosphate determines muscarinic agonist sensitivity of Kv7 K+ channels. J Gen Physiol. 2009;134:437–448. doi: 10.1085/jgp.200910313. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Dickson EJ, Hille B. Understanding phosphoinositides: rare, dynamic, and essential membrane phospholipids. Biochem J. 2019;476:1–23. doi: 10.1042/BCJ20180022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Viaud J, Boal F, Tronchère H, Gaits-Iacovoni F, Payrastre B. Phosphatidylinositol 5-phosphate: a nuclear stress lipid and a tuner of membranes and cytoskeleton dynamics. Bioessays. 2014;36:260–272. doi: 10.1002/bies.201300132. [DOI] [PubMed] [Google Scholar]
  • 46.Zewe JP, Miller AM, Sangappa S, Wills RC, Goulden BD, Hammond GRV. Probing the subcellular distribution of phosphatidylinositol reveals a surprising lack at the plasma membrane. J Cell Biol. 2020;219 doi: 10.1083/jcb.201906127. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Daleke DL. Regulation of transbilayer plasma membrane phospholipid asymmetry. J Lipid Res. 2003;44:233–242. doi: 10.1194/jlr.R200019-JLR200. [DOI] [PubMed] [Google Scholar]
  • 48.van Meer G, de Kroon AIPM. Lipid map of the mammalian cell. J Cell Sci. 2011;124:5–8. doi: 10.1242/jcs.071233. [DOI] [PubMed] [Google Scholar]
  • 49.Kim O-H, Kang G-H, Hur J, Lee J, Jung Y, Hong I-S, Lee H, Seo S-Y, Lee DH, Lee CS, Lee I-K, et al. Externalized phosphatidylinositides on apoptotic cells are eat-me signals recognized by CD14. Cell Death Differ. 2022;29:1423–1432. doi: 10.1038/s41418-022-00931-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Jackson M, Crick DC, Brennan PJ. Phosphatidylinositol is an essential phospholipid of mycobacteria*. J Biol Chem. 2000;275:30092–30099. doi: 10.1074/jbc.M004658200. [DOI] [PubMed] [Google Scholar]
  • 51.Ballou CE, Lee YC. The structure of a myoinositol mannoside from Mycobacterium tuberculosis glycolipid. Biochemistry. 1964;3:682–685. doi: 10.1021/bi00893a014. [DOI] [PubMed] [Google Scholar]
  • 52.Friesner RA, Banks JL, Murphy RB, Halgren TA, Klicic JJ, Mainz DT, Repasky MP, Knoll EH, Shelley M, Perry JK, Shaw DE, et al. Glide: a new approach for rapid, accurate docking and scoring. 1. Method and assessment of docking accuracy. J Med Chem. 2004;47:1739–1749. doi: 10.1021/jm0306430. [DOI] [PubMed] [Google Scholar]
  • 53.Brannigan G, Hénin J, Law R, Eckenhoff R, Klein ML. Embedded cholesterol in the nicotinic acetylcholine receptor. Proc Natl Acad Sci. 2008;105:14418–14423. doi: 10.1073/pnas.0803029105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Abramson J, Adler J, Dunger J, Evans R, Green T, Pritzel A, Ronneberger O, Willmore L, Ballard AJ, Bambrick J, Bodenstein SW, et al. Accurate structure prediction of biomolecular interactions with AlphaFold 3. Nature. 2024;630:493–500. doi: 10.1038/s41586-024-07487-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Baek M, DiMaio F, Anishchenko I, Dauparas J, Ovchinnikov S, Lee GR, Wang J, Cong Q, Kinch LN, Schaeffer RD, Millán C, et al. Accurate prediction of protein structures and interactions using a three-track neural network. Science. 2021;373:871–876. doi: 10.1126/science.abj8754. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Krishna R, Wang J, Ahern W, Sturmfels P, Venkatesh P, Kalvet I, Lee GR, Morey-Burrows FS, Anishchenko I, Humphreys IR, McHugh R, et al. Generalized biomolecular modeling and design with RoseTTAFold All-Atom. Science. 2024;384:eadl2528. doi: 10.1126/science.adl2528. [DOI] [PubMed] [Google Scholar]
  • 57.Boitreaud J, Dent J, McPartlon M, Meier J, Reis V, Rogozhonikov A, Wu K. Chai-1: decoding the molecular interactions of life. BioRxiv. 2024 doi: 10.1101/2024.10.10.615955. [DOI] [Google Scholar]
  • 58.Bannwarth C, Caldeweyher E, Ehlert S, Hansen A, Pracht P, Seibert J, Spicher S, Grimme S. Extended tight-binding quantum chemistry methods. WIREs Comput Mol Sci. 2021;11:e1493. doi: 10.1002/wcms.1493. [DOI] [Google Scholar]
  • 59.van der Kamp MW, Mulholland AJ. Combined quantum mechanics/molecular mechanics (QM/MM) methods in computational enzymology. Biochemistry. 2013;52:2708–2728. doi: 10.1021/bi400215w. [DOI] [PubMed] [Google Scholar]
  • 60.Hollingsworth SA, Dror RO. Molecular dynamics simulation for all. Neuron. 2018;99:1129–1143. doi: 10.1016/j.neuron.2018.08.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Karplus M, McCammon JA. Molecular dynamics simulations of biomolecules. Nature Struct Biol. 2002;9:646–652. doi: 10.1038/nsb0902-646. [DOI] [PubMed] [Google Scholar]
  • 62.Muller MP, Jiang T, Sun C, Lihan M, Pant S, Mahinthichaichan P, Trifan A, Tajkhorshid E. Characterization of lipid–protein interactions and lipid-mediated modulation of membrane protein function through molecular simulation. Chem Rev. 2019;119:6086–6161. doi: 10.1021/acs.chemrev.8b00608. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Enkavi G, Javanainen M, Kulig W, Róg T, Vattulainen I. Multiscale simulations of biological membranes: the challenge to understand biological phenomena in a living substance. Chem Rev. 2019;119:5607–5774. doi: 10.1021/acs.chemrev.8b00538. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Hedger G, Sansom MSP. Lipid interaction sites on channels, transporters and receptors: Recent insights from molecular dynamics simulations. Biochimica et Biophysica Acta (BBA) – B Biomembranes. 2016;1858:2390–2400. doi: 10.1016/j.bbamem.2016.02.037. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Simons K. Cell membranes: a subjective perspective. Biochimica et Biophysica Acta (BBA) – B Biomembranes. 2016;1858:2569–2572. doi: 10.1016/j.bbamem.2016.01.023. [DOI] [PubMed] [Google Scholar]
  • 66.Huang J, Rauscher S, Nawrocki G, Ran T, Feig M, de Groot BL, Grubmüller H, MacKerell AD. CHARMM36m: an improved force field for folded and intrinsically disordered proteins. Nature Methods. 2017;14:71–73. doi: 10.1038/nmeth.4067. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Klauda JB, Venable RM, Freites JA, O'Connor JW, Tobias DJ, Mondragon-Ramirez C, Vorobyov I, MacKerell AD, Jr, Pastor RW. Update of the CHARMM all-atom additive force field for lipids: validation on six lipid types. J Phys Chem B. 2010;114:7830–7843. doi: 10.1021/jp101759q. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Dickson CJ, Madej BD, Skjevik ÅA, Betz RM, Teigen K, Gould IR, Walker RC. Lipid14: the amber lipid force field. J Chem Theory Comput. 2014;10:865–879. doi: 10.1021/ct4010307. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Dickson CJ, Walker RC, Gould IR. Lipid21: complex lipid membrane simulations with AMBER. J Chem Theory Comput. 2022;18:1726–1736. doi: 10.1021/acs.jctc.1c01217. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Tian C, Kasavajhala K, Belfon KAA, Raguette L, Huang H, Migues AN, Bickel J, Wang Y, Pincay J, Wu Q, Simmerling C. ff19SB: amino-acid-specific protein backbone parameters trained against quantum mechanics energy surfaces in solution. J Chem Theory Comput. 2020;16:528–552. doi: 10.1021/acs.jctc.9b00591. [DOI] [PubMed] [Google Scholar]
  • 71.Marrink SJ, Risselada HJ, Yefimov S, Tieleman DP, de Vries AH. The MARTINI force field: coarse grained model for biomolecular simulations. J Phys Chem B. 2007;111:7812–7824. doi: 10.1021/jp071097f. [DOI] [PubMed] [Google Scholar]
  • 72.Corradi V, Sejdiu BI, Mesa-Galloso H, Abdizadeh H, Noskov SY, Marrink SJ, Tieleman DP. Emerging diversity in lipid–protein interactions. Chem Rev. 2019;119:5775–5848. doi: 10.1021/acs.chemrev.8b00451. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Chavent M, Duncan AL, Sansom MSP. Molecular dynamics simulations of membrane proteins and their interactions: from nanoscale to mesoscale. Curr Opin Struct Biol. 2016;40:8–16. doi: 10.1016/j.sbi.2016.06.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Bennett WFD, Bernardi A, Ozturk TN, Ingólfsson H, Fox SJ, Sun D, Maupin CM. ezAlign: a tool I. for converting coarse-grained molecular dynamics structures to atomistic resolution for multiscale modeling. Molecules. 2024;29 doi: 10.3390/molecules29153557. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Wassenaar TA, Pluhackova K, Böckmann RA, Marrink SJ, Tieleman DP. Going backward: a flexible geometric approach to reverse transformation from coarse grained to atomistic models. J Chem Theory Comput. 2014;10:676–690. doi: 10.1021/ct400617g. [DOI] [PubMed] [Google Scholar]
  • 76.Stansfeld PJ, Sansom MSP. Molecular simulation approaches to membrane proteins. Structure. 2011;19:1562–1572. doi: 10.1016/j.str.2011.10.002. [DOI] [PubMed] [Google Scholar]
  • 77.Shaw DE, Deneroff MM, Dror RO, Kuskin JS, Larson RH, Salmon JK, Young C, Batson B, Bowers KJ, Chao JC, Eastwood MP, et al. Anton, a special-purpose machine for molecular dynamics simulation. Commun ACM. 2008;51:91–97. doi: 10.1145/1364782.1364802. [DOI] [Google Scholar]
  • 78.Hénin J, Lelièvre T, Shirts MR, Valsson O, Delemotte L. Enhanced sampling methods for molecular dynamics simulations. arXiv preprint. 2022:arXiv:2202.04164. doi: 10.48550/arXiv.2202.04164. [DOI] [Google Scholar]
  • 79.Kooijman EE, King KE, Gangoda M, Gericke A. Ionization properties of phosphatidylinositol polyphosphates in mixed model membranes. Biochemistry. 2009;48:9360–9371. doi: 10.1021/bi9008616. [DOI] [PubMed] [Google Scholar]
  • 80.McLaughlin S, Wang J, Gambhir A, Murray D. PIP2 and proteins: interactions, organization, and information flow. Annu Rev Biophys. 2002;31:151–175. doi: 10.1146/annurev.biophys.31.082901.134259. [DOI] [PubMed] [Google Scholar]
  • 81.Borges-Araújo L, Fernandes F. Structure and lateral organization of phosphatidylinositol 4,5-bisphosphate. Molecules. 2020 doi: 10.3390/molecules25173885. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Buslaev P, Aho N, Jansen A, Bauer P, Hess B, Groenhof G. Best practices in constant pH MD simulations: accuracy and sampling. J Chem Theory Comput. 2022;18:6134–6147. doi: 10.1021/acs.jctc.2c00517. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Borges-Araújo L, Souza PCT, Fernandes F, Melo MN. Improved parameterization of phosphatidylinositide lipid headgroups for the Martini 3 coarse-grain force field. J Chem Theory Comput. 2022;18:357–373. doi: 10.1021/acs.jctc.1c00615. [DOI] [PubMed] [Google Scholar]
  • 84.Stansfeld PJ, Hopkinson R, Ashcroft FM, Sansom MSP. PIP2-binding site in kir channels: definition by multiscale biomolecular simulations. Biochemistry. 2009;48:10926–10933. doi: 10.1021/bi9013193. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Jarin Z, Venable RM, Han K, Pastor RW. Ion-induced PIP2 clustering with Martini3: modification of phosphate–ion interactions and comparison with CHARMM36. J Phys Chem B. 2024;128:2134–2143. doi: 10.1021/acs.jpcb.3c06523. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Pedersen KB, Ingólfsson HI, R-E DP, Borges-Araujo L, Andreasen MD, Empereur-mot C, Melcr J, Ozturk TN, Bennett WFD, Kjølbye LR, Brasnett C, Corradi V, et al. The Martini 3 lipidome: expanded and refined parameters improve lipid phase behavior. ChemRxiv. 2024 doi: 10.26434/chemrxiv-2024-8bjrr. [DOI] [Google Scholar]
  • 87.Brown CM, Corey RA, Grélard A, Gao Y, Choi YK, Luna E, Gilleron M, Destainville N, Nigou J, Loquet A, Fullam E, et al. Supramolecular organization and dynamics of mannosylated phosphatidylinositol lipids in the mycobacterial plasma membrane. Proc Natl Acad Sci. 2023;120:e2212755120. doi: 10.1073/pnas.2212755120. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Lupyan D, Mezei M, Logothetis DE, Osman R. A molecular dynamics investigation of lipid bilayer perturbation by PIP2. Biophys J. 2010;98:240–247. doi: 10.1016/j.bpj.2009.09.063. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Mallajosyula SS, Guvench O, Hatcher E, MacKerell AD., Jr CHARMM additive all-atom force field for phosphate and sulfate linked to carbohydrates. J Chem Theory Comput. 2012;8:759–776. doi: 10.1021/ct200792v. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Wu EL, Qi Y, Song KC, Klauda JB, Im W. Preferred orientations of phosphoinositides in bilayers and their implications in protein recognition mechanisms. J Phys Chem B. 2014;118:4315–4325. doi: 10.1021/jp500610t. [DOI] [PubMed] [Google Scholar]
  • 91.Lee J, Patel DS, Ståhle J, Park S-J, Kern NR, Kim S, Lee J, Cheng X, Valvano MA, Holst O, Knirel YA, et al. CHARMM-GUI membrane builder for complex biological membrane simulations with glycolipids and lipoglycans. J Chem Theory Comput. 2019;15:775–786. doi: 10.1021/acs.jctc.8b01066. [DOI] [PubMed] [Google Scholar]
  • 92.Qi Y, Cheng X, Lee J, Vermaas JV, Pogorelov TV, Tajkhorshid E, Park S, Klauda JB, Im W. CHARMM-GUI HMMM builder for membrane simulations with the highly mobile membrane-mimetic model. Biophys J. 2015;109:2012–2022. doi: 10.1016/j.bpj.2015.10.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Wu EL, Cheng X, Jo S, Rui H, Song KC, Dávila-Contreras EM, Qi Y, Lee J, Monje-Galvan V, Venable RM, Klauda JB, et al. CHARMM-GUI membrane builder toward realistic biological membrane simulations. J Comput Chem. 2014;35:1997–2004. doi: 10.1002/jcc.23702. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Ingólfsson HI, Melo MN, van Eerden FJ, Arnarez C, Lopez CA, Wassenaar TA, Periole X, de Vries AH, Tieleman DP, Marrink SJ. Lipid organization of the plasma membrane. J Am Chem Soc. 2014;136:14554–14559. doi: 10.1021/ja507832e. [DOI] [PubMed] [Google Scholar]
  • 95.Ingólfsson HI, Carpenter TS, Bhatia H, Bremer P-T, Marrink SJ, Lightstone FC. Computational lipidomics of the neuronal plasma membrane. Biophys J. 2017;113:2271–2280. doi: 10.1016/j.bpj.2017.10.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Duncan AL, Ruprecht JJ, Kunji ERS, Robinson AJ. Cardiolipin dynamics and binding to conserved residues in the mitochondrial ADP/ATP carrier. Biochimica et Biophysica Acta (BBA) – Biomembranes. 2018;1860:1035–1045. doi: 10.1016/j.bbamem.2018.01.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Hedger G, Rouse SL, Domański J, Chavent M, Koldsø H, Sansom MSP. Lipid-loving ANTs: molecular simulations of cardiolipin interactions and the organization of the adenine nucleotide translocase in model mitochondrial membranes. Biochemistry. 2016;55:6238–6249. doi: 10.1021/acs.biochem.6b00751. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Brown CM, Westendorp MSS, Zarmiento-Garcia R, Stevens JA, Rouse SL, Marrink SJ, Wassenaar TA. An integrative modelling approach to the mitochondrial cristae. BioRxiv. 2024 doi: 10.1101/2024.09.23.613389. [DOI] [Google Scholar]
  • 99.Lolicato F, Saleppico R, Griffo A, Meyer A, Scollo F, Pokrandt B, Müller H-M, Ewers H, Hähl H, Fleury J-B, Seemann R, et al. Cholesterol promotes clustering of PI(4,5)P2 driving unconventional secretion of FGF2. J Cell Biol. 2022;221 doi: 10.1083/jcb.202106123. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Lee J, Cheng X, Jo S, MacKerell AD, Jr, Klauda JB, Im W. CHARMM-GUI input generator for NAMD, gromacs, amber, openmm, and CHARMM/OpenMM simulations using the CHARMM36 additive force field. Biophys J. 2016;110:641a. doi: 10.1021/acs.jctc.5b00935. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Wassenaar TA, Ingólfsson HI, Böckmann RA, Tieleman DP, Marrink SJ. Computational lipidomics with insane: a versatile tool for generating custom membranes for molecular simulations. J Chem Theory Comput. 2015;11:2144–2155. doi: 10.1021/acs.jctc.5b00209. [DOI] [PubMed] [Google Scholar]
  • 102.Kroon PC, Grunewald F, Barnoud J, van Tilburg M, Souza PCT, Wassenaar TA, Marrink SJ. Martinize2 and vermouth: unified framework for topology generation. 2024 doi: 10.7554/elife.90627.2. [DOI] [Google Scholar]
  • 103.Andreasen MD, Souza PCT, Schiøtt B, Zuzic L. Creating coarse-grained systems with COBY: towards higher accuracy in membrane complexity. BioRxiv. 2024 doi: 10.1101/2024.07.23.604601. [DOI] [Google Scholar]
  • 104.Tiemann JKS, Szczuka M, Bouarroudj L, Oussaren M, Garcia S, Howard RJ, Delemotte L, Lindahl E, Baaden M, Lindorff-Larsen K, Chavent M, et al. Mdverse, shedding light on the dark matter of molecular dynamics simulations. eLife. 2024;12:RP90061. doi: 10.7554/eLife.90061. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Tiemann JKS, Guixà-González R, Hildebrand PW, Rose AS. MDsrv: viewing and sharing molecular dynamics simulations on the web. Nature Methods. 2017;14:1123–1124. doi: 10.1038/nmeth.4497. [DOI] [PubMed] [Google Scholar]
  • 106.Song W, Corey RA, Ansell TB, Cassidy CK, Horrell MR, Duncan AL, Stansfeld PJ, Sansom MSP. PyLipID: a python package for analysis of protein–lipid interactions from molecular dynamics simulations. J Chem Theory Comput. 2022;18:1188–1201. doi: 10.1021/acs.jctc.1c00708. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Sejdiu BI, Tieleman DP. ProLint: a web-based framework for the automated data analysis and visualization of lipid–protein interactions. Nucleic Acids Res. 2021;49:W544–W550. doi: 10.1093/nar/gkab409. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Corey RA, Vickery ON, Sansom MSP, Stansfeld PJ. Insights into membrane protein–lipid interactions from free energy calculations. J Chem Theory Comput. 2019;15:5727–5736. doi: 10.1021/acs.jctc.9b00548. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Leiserson CE, Thompson NC, Emer JS, Kuszmaul BC, Lampson BW, Sanchez D, Schardl TB. There’s plenty of room at the top: what will drive computer performance after Moore’s law? Science. 2020;368:eaam9744. doi: 10.1126/science.aam9744. [DOI] [PubMed] [Google Scholar]
  • 110.Páll S, Zhmurov A, Bauer P, Abraham M, Lundborg M, Gray A, Hess B, Lindahl E. Heterogeneous parallelization and acceleration of molecular dynamics simulations in GROMACS. J Chem Phys. 2020;153:134110. doi: 10.1063/5.0018516. [DOI] [PubMed] [Google Scholar]
  • 111.Vögele M, Köfinger J, Hummer G. Hydrodynamics of diffusion in lipid membrane simulations. Phys Rev Lett. 2018;120:268104. doi: 10.1103/PhysRevLett.120.268104. [DOI] [PubMed] [Google Scholar]
  • 112.Duncan AL, Reddy T, Koldsø H, Hélie J, Fowler PW, Chavent M, Sansom MSP. Protein crowding and lipid complexity influence the nanoscale dynamic organization of ion channels in cell membranes. Sci Rep. 2017;7:16647. doi: 10.1038/s41598-017-16865-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Fowler PW, Hélie J, Duncan A, Chavent M, Koldsø H, Sansom MSP. Membrane stiffness is modified by integral membrane proteins. Soft Matter. 2016;12:7792–7803. doi: 10.1039/c6sm01186a. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Hadden JA, Perilla JR. All-atom virus simulations. Curr Opin Virol. 2018;31:82–91. doi: 10.1016/j.coviro.2018.08.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Gupta C, Sarkar D, Tieleman DP, Singharoy A. The ugly, bad, and good stories of large-scale biomolecular simulations. Curr Opin Struct Biol. 2022;73:102338. doi: 10.1016/j.sbi.2022.102338. [DOI] [PubMed] [Google Scholar]
  • 116.Khalid S, Brandner AF, Juraschko N, Newman KE, Pedebos C, Prakaash D, Smith IPS, Waller C, Weerakoon D. Computational microbiology of bacteria: advancements in molecular dynamics simulations. Structure. 2023;31:1320–1327. doi: 10.1016/j.str.2023.09.012. [DOI] [PubMed] [Google Scholar]
  • 117.Shaw DE, Adams PJ, Azaria A, Bank JA, Batson B, Bell A, Bergdorf M, Bhatt J, Butts JA, Correia T, Dirks RM, et al. Anton 3: twenty microseconds of molecular dynamics simulation before lunch; Proceedings of the International Conference for High Performance Computing, Networking, Storage and Analysis; 2021. [DOI] [Google Scholar]
  • 118.Naddaf M. The science events to watch for in 2024. Nature. 2024;625:221–223. doi: 10.1038/d41586-023-04044-9. [DOI] [PubMed] [Google Scholar]
  • 119.Gapsys V, Kopec W, Matthes D, de Groot BL. Biomolecular simulations at the exascale: from drug design to organelles and beyond. Curr Opin Struct Biol. 2024;88:102887. doi: 10.1016/j.sbi.2024.102887. [DOI] [PubMed] [Google Scholar]
  • 120.Gault J, Liko I, Landreh M, Shutin D, Bolla JR, Jefferies D, Agasid M, Yen H-Y, Ladds MJGW, Lane DP, Khalid S, et al. Combining native and ‘omics’ mass spectrometry to identify endogenous ligands bound to membrane proteins. Nature Methods. 2020;17:505–508. doi: 10.1038/s41592-020-0821-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Hauser AS, Attwood MM, Rask-Andersen M, Schiöth HB, Gloriam DE. Trends in GPCR drug discovery: new agents, targets and indications. Nature Rev Drug Discov. 2017;16:829–842. doi: 10.1038/nrd.2017.178. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Hilger D, Masureel M, Kobilka BK. Structure and dynamics of GPCR signaling complexes. Nature Struct Mol Biol. 2018;25:4–12. doi: 10.1038/s41594-017-0011-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Fasciani I, Carli M, Petragnano F, Colaianni F, Aloisi G, Maggio R, Scarselli M, Rossi M. GPCRs in intracellular compartments: new targets for drug discovery. Biomolecules. 2022;12:1343. doi: 10.3390/biom12101343. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Vargas MV, Dunlap LE, Dong C, Carter SJ, Tombari RJ, Jami SA, Cameron LP, Patel SD, Hennessey JJ, Saeger HN, McCorvy JD, et al. Psychedelics promote neuroplasticity through the activation of intracellular 5-HT2A receptors. Science. 2023;379:700–706. doi: 10.1126/science.adf0435. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Austin G, ElEid L, Oqua AI, Manchanda Y, Poliakova Y, Bouzakri K, Montoya A, Withers DJ, Jones B, Millership SJ, Prokopenko I, et al. An inter-organelle contact between endosomal GLP-1R, ER VAP-B, and the mitochondrial AKAP SPHKAP triggers PKA-dependent MIC19 phosphorylation and β-cell mitochondrial remodelling. BioRxiv. 2024 doi: 10.1101/2024.04.28.591531. [DOI] [Google Scholar]
  • 126.Wootten D, Christopoulos A, Marti-Solano M, Babu MM, Sexton PM. Mechanisms of signalling and biased agonism in G protein-coupled receptors. Nature Rev Mol Cell Biol. 2018;19:638–653. doi: 10.1038/s41580-018-0049-3. [DOI] [PubMed] [Google Scholar]
  • 127.Billesbølle CB, de March CA, van der Velden WJC, Ma N, Tewari J, del Torrent CL, Li L, Faust B, Vaidehi N, Matsunami H, Manglik A. Structural basis of odorant recognition by a human odorant receptor. Nature. 2023;615:742–749. doi: 10.1038/s41586-023-05798-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Donnelly D. The structure and function of the glucagon-like peptide-1 receptor and its ligands. Br J Pharmacol. 2012;166:27–41. doi: 10.1111/j.1476-5381.2011.01687.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Niswender CM, Conn PJ. Metabotropic glutamate receptors: physiology, pharmacology, and disease. Annu Rev Pharmacol Toxicol. 2010;50:295–322. doi: 10.1146/annurev.pharmtox.011008.145533. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Palczewski K. G Protein–coupled receptor rhodopsin. Annu Rev Biochem. 2006;75:743–767. doi: 10.1146/annurev.biochem.75.103004.142743. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Liu S, Anderson PJ, Rajagopal S, Lefkowitz RJ, Rockman HA. G protein-coupled receptors: a century of research and discovery. Circ Res. 2024;135:174–197. doi: 10.1161/CIRCRESAHA.124.323067. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Sengupta D, Chattopadhyay A. Molecular dynamics simulations of GPCR–cholesterol interaction: an emerging paradigm. Biochim Biophys Acta Biomembr. 2015;1848:1775–1782. doi: 10.1016/j.bbamem.2015.03.018. [DOI] [PubMed] [Google Scholar]
  • 133.Hanson MA, Cherezov V, Griffith MT, Roth CB, Jaakola V-P, Chien EYT, Velasquez J, Kuhn P, Stevens RC. A specific cholesterol binding site is established by the 2.8 A structure of the human b2-adrenergic receptor. Structure. 2008;16:897–905. doi: 10.1016/j.str.2008.05.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Oqua AI, Chao K, Eid LE, Casteller L, Miguéns A, Barg S, Jones B, de la Serna JB, Rouse SL, Tomas A. Molecular mapping and functional validation of GLP-1R cholesterol binding sites in pancreatic beta cells. BioRxiv. 2024 doi: 10.1101/2024.06.22.600087. [DOI] [Google Scholar]
  • 135.Dawaliby R, Trubbia C, Delporte C, Masureel M, Van Antwerpen P, Kobilka BK, Govaerts C. Allosteric regulation of G protein–coupled receptor activity by phospholipids. Nature Chem Biol. 2016;12:35–39. doi: 10.1038/nchembio.1960. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Dijkman PM, Watts A. Lipid modulation of early G protein-coupled receptor signalling events. Biochim Biophys Acta Biomembr. 2015;1848:2889–2897. doi: 10.1016/j.bbamem.2015.08.004. [DOI] [PubMed] [Google Scholar]
  • 137.Inagaki S, Ghirlando R, White JF, Gvozdenovic-Jeremic J, Northup JK, Grisshammer R. Modulation of the interaction between neurotensin receptor NTS1 and Gq protein by lipid. J Mol Biol. 2012;417:95–111. doi: 10.1016/j.jmb.2012.01.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Velgy N, Hedger G, Biggin PC. In: Computational Methods for GPCR Drug Discovery. Heifetz A, editor. New York, NY: Springer, New York; 2018. GPCRs: what can we learn from molecular dynamics simulations? pp. 133–158. [DOI] [PubMed] [Google Scholar]
  • 139.Roux B. The calculation of the potential of mean force using computer simulations. Comput Phys Commun. 1995;91:275–282. doi: 10.1016/0010-4655(95)00053-I. [DOI] [Google Scholar]
  • 140.Huang W, Masureel M, Qu Q, Janetzko J, Inoue A, Kato HE, Robertson MJ, Nguyen KC, Glenn JS, Skiniotis G, Kobilka BK. Structure of the neurotensin receptor 1 in complex with β-arrestin 1. Nature. 2020;579:303–308. doi: 10.1038/s41586-020-1953-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Janetzko J, Kise R, Barsi-Rhyne B, Siepe DH, Heydenreich FM, Kawakami K, Masureel M, Maeda S, Garcia KC, von Zastrow M, Inoue A, et al. Membrane phosphoinositides regulate GPCR-beta-arrestin complex assembly and dynamics. Cell. 2022;185:4560–4573.:e4519. doi: 10.1016/j.cell.2022.10.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Zhang M, Chen T, Lu X, Lan X, Chen Z, Lu S. G protein-coupled receptors (GPCRs): advances in structures, mechanisms, and drug discovery. Signal Transduct Target Ther. 2024;9:88. doi: 10.1038/s41392-024-01803-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Pándy-Szekeres G, Munk C, Tsonkov TM, Mordalski S, Harpsøe K, Hauser AS, Bojarski AJ, Gloriam DE. GPCRdb in 2018: adding GPCR structure models and ligands. Nucleic Acids Res. 2017;46:D440–D446. doi: 10.1093/nar/gkx1109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Coskun D, Lihan M, Rodrigues JPGLM, Vass M, Robinson D, Friesner RA, Miller EB. Using AlphaFold and experimental structures for the prediction of the structure and binding affinities of GPCR complexes via induced fit docking and free energy perturbation. J Chem Theory Comput. 2024;20:477–489. doi: 10.1021/acs.jctc.3c00839. [DOI] [PubMed] [Google Scholar]
  • 145.Heo L, Feig M. Multi-state modeling of G-protein coupled receptors at experimental accuracy. Proteins: Struct Funct Bioinf. 2022;90:1873–1885. doi: 10.1002/prot.26382. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Miller EB, Hwang H, Shelley M, Placzek A, Rodrigues JPGLM, Suto RK, Wang L, Akinsanya K, Abel R. Enabling structure-based drug discovery utilizing predicted models. Cell. 2024;187:521–525. doi: 10.1016/j.cell.2023.12.034. [DOI] [PubMed] [Google Scholar]
  • 147.Waterhouse A, Bertoni M, Bienert S, Studer G, Tauriello G, Gumienny R, Heer FT, de Beer TAP, Rempfer C, Bordoli L, Lepore R, et al. SWISS-MODEL: homology modelling of protein structures and complexes. Nucleic Acids Res. 2018;46:W296–W303. doi: 10.1093/nar/gky427. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Song W, Yen H-Y, Robinson CV, Sansom MSP. State-dependent lipid interactions with the A2a receptor revealed by MD simulations using in vivo-mimetic membranes. Structure. 2019;27:392–403.:e393. doi: 10.1016/j.str.2018.10.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Sejdiu BI, Tieleman DP. Lipid-protein interactions are a unique property and defining feature of G protein-coupled receptors. Biophys J. 2020;118:1887–1900. doi: 10.1016/j.bpj.2020.03.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Kjølbye LR, Sørensen L, Yan J, Berglund NA, Ferkinghoff-Borg J, Robinson CV, Schiøtt B. Lipid modulation of a Class B GPCR: elucidating the modulatory role of PI(4,5)P2 lipids. J Chem Inf Model. 2022;62:6788–6802. doi: 10.1021/acs.jcim.2c00635. [DOI] [PubMed] [Google Scholar]
  • 151.Ma N, Lee S, Vaidehi N. Activation microswitches in adenosine receptor A2A function as rheostats in the cell membrane. Biochemistry. 2020;59:4059–4071. doi: 10.1021/acs.biochem.0c00626. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Ansell TB, Song W, Sansom MSP. The glycosphingolipid GM3 modulates conformational dynamics of the glucagon receptor. Biophys J. 2020;119:300–313. doi: 10.1016/j.bpj.2020.06.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Mukhaleva E, Yang T, Sadler F, Sivaramakrishnan S, Ma N, Vaidehi N. Cellular lipids regulate the conformational ensembles of the disordered intracellular loop 3 in beta2-adrenergic receptor. iScience. 2024;27 doi: 10.1016/j.isci.2024.110086. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Damian M, Louet M, Gomes AAS, M’Kadmi C, Denoyelle S, Cantel S, Mary S, Bisch PM, Fehrentz JA, Catoire LJ, Floquet N, et al. Allosteric modulation of ghrelin receptor signaling by lipids. Nature Commun. 2021;12:3938. doi: 10.1038/s41467-021-23756-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Thakur N, Ray AP, Sharp L, Jin B, Duong A, Pour NG, Obeng S, Wijesekara AV, Gao Z-G, McCurdy CR, Jacobson KA, et al. Anionic phospholipids control mechanisms of GPCR-G protein recognition. Nature Commun. 2023;14:794. doi: 10.1038/s41467-023-36425-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Gomes A, Di Michele M, Roessner RA, Damian M, Bisch PM, Sibille N, Louet M, Baneres J-L, Floquet N. Lipids modulate the dynamics of GPCR:β-Arrestin interaction. BioRxiv. 2024 doi: 10.1101/2024.03.16.585329. [DOI] [Google Scholar]
  • 157.Borges-Araújo L, Domingues MM, Fedorov A, Santos NC, Melo MN, Fernandes F. Acyl-chain saturation regulates the order of phosphatidylinositol 4,5-bisphosphate nanodomains. Commun Chem. 2021;4:164. doi: 10.1038/s42004-021-00603-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Hedger G, Koldsø H, Chavent M, Siebold C, Rohatgi R, Sansom MSP. Cholesterol interaction sites on the transmembrane domain of the hedgehog signal transducer and class F G protein-coupled receptor smoothened. Structure. 2019;27:549–559.:e542. doi: 10.1016/j.str.2018.11.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Garcia-Gonzalo FR, Phua SC, Roberson EC, Garcia G, III, Abedin M, Schurmans S, Inoue T, Reiter JF. Phosphoinositides regulate ciliary protein trafficking to modulate hedgehog signaling. Dev Cell. 2015;34:400–409. doi: 10.1016/j.devcel.2015.08.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Chávez M, Ena S, Van Sande J, de Kerchove d’Exaerde A, Schurmans S, Schiffmann SN. Modulation of ciliary phosphoinositide content regulates trafficking and sonic hedgehog signaling output. Dev Cell. 2015;34:338–350. doi: 10.1016/j.devcel.2015.06.016. [DOI] [PubMed] [Google Scholar]
  • 161.Nakatsu F. A phosphoinositide code for primary cilia. Dev Cell. 2015;34:379–380. doi: 10.1016/j.devcel.2015.08.008. [DOI] [PubMed] [Google Scholar]
  • 162.Chao KW, Wong L, Oqua A, Kalayan J, Manchanda Y, Gebbie-Rayet J, Hedger G, Tomas A, Rouse SL. Human class B1 GPCR modulation by plasma membrane lipids. BioRxiv. 2024 doi: 10.1101/2024.12.19.629375. [DOI] [Google Scholar]
  • 163.Zhang H, Qiao A, Yang D, Yang L, Dai A, de Graaf C, Reedtz-Runge S, Dharmarajan V, Zhang H, Han GW, Grant TD, et al. Structure of the full-length glucagon class B G-protein-coupled receptor. Nature. 2017;546:259–264. doi: 10.1038/nature22363. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Jazayeri A, Doré AS, Lamb D, Krishnamurthy H, Southall SM, Baig AH, Bortolato A, Koglin M, Robertson NJ, Errey JC, Andrews SP, et al. Extra-helical binding site of a glucagon receptor antagonist. Nature. 2016;533:274–277. doi: 10.1038/nature17414. [DOI] [PubMed] [Google Scholar]
  • 165.Sun B, Willard FS, Feng D, Alsina-Fernandez J, Chen Q, Vieth M, Ho JD, Showalter AD, Stutsman C, Ding L, Suter TM, et al. Structural determinants of dual incretin receptor agonism by tirzepatide. Proc Natl Acad Sci. 2022;119:e2116506119. doi: 10.1073/pnas.2116506119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Koldsø H, Sansom MSP. Organization and dynamics of receptor proteins in a plasma membrane. J Am Chem Soc. 2015;137:14694–14704. doi: 10.1021/jacs.5b08048. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Hedger G. Characterisation of lipid-protein interactions through computational modelling. PhD Thesis University of Oxford; 2017. [Google Scholar]
  • 168.Song W, Duncan AL, Sansom MSP. Modulation of adenosine A2a receptor oligomerization by receptor activation and PIP2 interactions. Structure. 2021;29:1312–1325.:e1313. doi: 10.1016/j.str.2021.06.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Bolivar JH, Muñoz-García JC, Castro-Dopico T, Dijkman PM, Stansfeld PJ, Watts A. Interaction of lipids with the neurotensin receptor 1. Biochimica et Biophysica Acta (BBA) – Biomembranes. 2016;1858:1278–1287. doi: 10.1016/j.bbamem.2016.02.032. [DOI] [PubMed] [Google Scholar]
  • 170.Eberle SA, Gustavsson M. Bilayer lipids modulate ligand binding to atypical chemokine receptor 3. Structure. 2024;32:1174–1183.:e1175. doi: 10.1016/j.str.2024.04.018. [DOI] [PubMed] [Google Scholar]
  • 171.Pluhackova K, Wilhelm FM, Müller DJ. Lipids and phosphorylation conjointly modulate complex formation of β2-adrenergic receptor and β-arrestin2. Front Cell Dev Biol. 2021;9 doi: 10.3389/fcell.2021.807913. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Hille B, Dickson EJ, Kruse M, Vivas O, Suh B-C. Phosphoinositides regulate ion channels. Biochimica et Biophysica Acta (BBA) – Mol Cell Biol Lipids. 2015;1851:844–856. doi: 10.1016/j.bbalip.2014.09.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Rohacs T. Phosphoinositide regulation of TRP channels: a functional overview in the structural era. Annu Rev Physiol. 2024;86:329–355. doi: 10.1146/annurev-physiol-042022-013956. [DOI] [PubMed] [Google Scholar]
  • 174.Thon O, Wang Z, Schmidpeter PAM, Nimigean CM. PIP2 inhibits pore opening of the cyclic nucleotide-gated channel SthK. Nature Commun. 2024;15:8230. doi: 10.1038/s41467-024-52469-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175.Ramsey IS, Delling M, Clapham DE. An introduction to trp channels. Annu Rev Physiol. 2006;68:619–647. doi: 10.1146/annurev.physiol.68.040204.100431. [DOI] [PubMed] [Google Scholar]
  • 176.Nilius B, Voets T, Peters J. TRP channels in disease. Science’s STKE. 2005;2005:re88. doi: 10.1126/stke.2952005re8. [DOI] [PubMed] [Google Scholar]
  • 177.Kaneko Y, Szallasi A. Transient receptor potential (TRP) channels: a clinical perspective. Br J Pharmacol. 2014;171:2474–2507. doi: 10.1111/bph.12414. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 178.Bergmann C, Guay-Woodford LM, Harris PC, Horie S, Peters DJM, Torres VE. Polycystic kidney disease. Nature Rev Dis Primers. 2018;4:50. doi: 10.1038/s41572-018-0047-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Taberner FJ, Fernández-Ballester G, Fernández-Carvajal A, Ferrer-Montiel A. TRP channels interaction with lipids and its implications in disease. Biochimica et Biophysica Acta (BBA) – Biomembranes. 2015;1848:1818–1827. doi: 10.1016/j.bbamem.2015.03.022. [DOI] [PubMed] [Google Scholar]
  • 180.Hughes TET, Pumroy RA, Yazici AT, Kasimova MA, Fluck EC, Huynh KW, Samanta A, Molugu SK, Zhou ZH, Carnevale V, Rohacs T, et al. Structural insights on TRPV5 gating by endogenous modulators. Nature Commun. 2018;9:4198. doi: 10.1038/s41467-018-06753-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.Wang Q, Corey RA, Hedger G, Aryal P, Grieben M, Nasrallah C, Baronina A, Pike ACW, Shi J, Carpenter EP, Sansom MSP. Lipid interactions of a ciliary membrane TRP channel: simulation and structural studies of Polycystin-2. Structure. 2020;28:169–184.:e165. doi: 10.1016/j.str.2019.11.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182.Ha K, Mundt-Machado N, Bisignano P, Pinedo A, Raleigh DR, Loeb G, Reiter JF, Cao E, Delling M. Cilia-enriched oxysterol 7β,27-DHC is required for polycystin ion channel activation. Nature Commun. 2024;15:6468. doi: 10.1038/s41467-024-50318-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Ciardo MG, Ferrer-Montiel A. Lipids as central modulators of sensory TRP channels. Biochim Biophys Acta Biomembr. 2017;1859:1615–1628. doi: 10.1016/j.bbamem.2017.04.012. [DOI] [PubMed] [Google Scholar]
  • 184.Clarke A, Skerjanz J, Gsell MAF, Wiedner P, Erkan-Candag H, Groschner K, Stockner T, Tiapko O. PIP2 modulates TRPC3 activity via TRP helix and S4-S5 linker. Nature Commun. 2024;15:5220. doi: 10.1038/s41467-024-49396-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Wang L, Cai R, Chen X-Z, Peng J-B. Molecular insights into the structural and dynamical changes of calcium channel TRPV6 induced by its interaction with phosphatidylinositol 4,5-bisphosphate. J Biomol Struct Dyn. 2023;41:6559–6568. doi: 10.1080/07391102.2022.2109752. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Fathizadeh A, Senning E, Elber R. Impact of the protonation state of phosphatidylinositol 4,5-bisphosphate (PIP2) on the binding kinetics and thermodynamics to transient receptor potential vanilloid (TRPV5): a milestoning study. J Phys Chem B. 2021;125:9547–9556. doi: 10.1021/acs.jpcb.1c04052. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187.Shirey CM, Scott JL, Stahelin RV. Notes and tips for improving quality of lipid-protein overlay assays. Anal Biochem. 2017;516:9–12. doi: 10.1016/j.ab.2016.10.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Seebohm G, Wrobel E, Pusch M, Dicks M, Terhag J, Matschke V, Rothenberg I, Ursu ON, Hertel F, Pott L, Lang F, et al. Structural basis of PI(4,5) P2-dependent regulation of GluA1 by phosphatidylinositol-5-phosphate 4-kinase, type II, alpha (PIP5K2A. Pflügers Archiv – Eur J Physiol. 2014;466:1885–1897. doi: 10.1007/s00424-013-1424-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 189.Hedger G, Sansom MSP, Koldsø H. The juxtamembrane regions of human receptor tyrosine kinases exhibit conserved interaction sites with anionic lipids. Sci Rep. 2015;5:9198. doi: 10.1038/srep09198. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190.Araya-Secchi R, Bugge K, Seiffert P, Petry A, Haxholm GW, Lindorff-Larsen K, Pedersen SF, Arleth L, Kragelund BB. The prolactin receptor scaffolds janus kinase 2 via co-structure formation with phosphoinositide-4,5-bisphosphate. eLife. 2023;12:e84645. doi: 10.7554/eLife.84645. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Goretzki B, Wiedemann C, McCray BA, Schäfer SL, Jansen J, Tebbe F, Mitrovic S-A, Nöth J, Cabezudo AC, Donohue JK, Jeffries CM, et al. Crosstalk between regulatory elements in disordered TRPV4 N-terminus modulates lipid-dependent channel activity. Nature Commun. 2023;14:4165. doi: 10.1038/s41467-023-39808-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.Thomasen FE, Skaalum T, Kumar A, Srinivasan S, Vanni S, Lindorff-Larsen K. Rescaling protein-protein interactions improves Martini 3 for flexible proteins in solution. Nature Commun. 2024;15:6645. doi: 10.1038/s41467-024-50647-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 193.Tesei G, Trolle AI, Jonsson N, Betz J, Knudsen FE, Pesce F, Johansson KE, Lindorff-Larsen K. Conformational ensembles of the human intrinsically disordered proteome. Nature. 2024;626:897–904. doi: 10.1038/s41586-023-07004-5. [DOI] [PubMed] [Google Scholar]
  • 194.Shieh C-C, Coghlan M, Sullivan JP, Gopalakrishnan M. Potassium channels: molecular defects, diseases, and therapeutic opportunities. Pharmacol Rev. 2000;52:557. [PubMed] [Google Scholar]
  • 195.Long SB, Tao X, Campbell EB, MacKinnon R. Atomic structure of a voltage-dependent K+ channel in a lipid membrane-like environment. Nature. 2007;450:376–382. doi: 10.1038/nature06265. [DOI] [PubMed] [Google Scholar]
  • 196.Valiyaveetil FI, Zhou Y, MacKinnon R. Lipids in the structure, folding, and function of the KcsA K+ channel. Biochemistry. 2002;41:10771–10777. doi: 10.1021/bi026215y. [DOI] [PubMed] [Google Scholar]
  • 197.Xu Y, Ramu Y, Lu Z. Removal of phospho-head groups of membrane lipids immobilizes voltage sensors of K+ channels. Nature. 2008;451:826–829. doi: 10.1038/nature06618. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 198.Hansen SB. Lipid agonism: The PIP2 paradigm of ligand-gated ion channels. Biochimica et Biophysica Acta (BBA) – Mol Cell Biol Lipids. 2015;1851:620–628. doi: 10.1016/j.bbalip.2015.01.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 199.Duncan AL, Corey RA, Sansom MSP. Defining how multiple lipid species interact with inward rectifier potassium (Kir2) channels. Proc Natl Acad Sci. 2020;117:7803–7813. doi: 10.1073/pnas.1918387117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 200.Cheng WWL, D'Avanzo N, Doyle DA, Nichols CG. Dual-mode phospholipid regulation of human inward rectifying potassium channels. Biophys J. 2011;100:620–628. doi: 10.1016/j.bpj.2010.12.3724. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201.Lee S-J, Wang S, Borschel W, Heyman S, Gyore J, Nichols CG. Secondary anionic phospholipid binding site and gating mechanism in Kir2.1 inward rectifier channels. Nature Commun. 2013;4:2786. doi: 10.1038/ncomms3786. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 202.Kim S, Chen J, Cheng T, Gindulyte A, He J, He S, Li Q, Shoemaker BA, Thiessen PA, Yu B, Zaslavsky L, et al. PubChem 2023 update. Nucleic Acids Res. 2023;51:D1373–D1380. doi: 10.1093/nar/gkac956. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 203.Lopes CMB, Zhang H, Rohacs T, Jin T, Yang J, Logothetis DE. Alterations in conserved kir channel-PIP2 interactions underlie channelopathies. Neuron. 2002;34:933–944. doi: 10.1016/s0896-6273(02)00725-0. [DOI] [PubMed] [Google Scholar]
  • 204.Zhang H, He C, Yan X, Mirshahi T, Logothetis DE. Activation of inwardly rectifying K+ channels by distinct PtdIns(4,5)P2 interactions. Nature Cell Biol. 1999;1:183–188. doi: 10.1038/11103. [DOI] [PubMed] [Google Scholar]
  • 205.Huang C-L, Feng S, Hilgemann DW. Direct activation of inward rectifier potassium channels by PIP2 and its stabilization by Gβγ. Nature. 1998;391:803–806. doi: 10.1038/35882. [DOI] [PubMed] [Google Scholar]
  • 206.Schmidt MR, Stansfeld PJ, Tucker SJ, Sansom MSP. Simulation-based prediction of phosphatidylinositol 4,5-bisphosphate binding to an ion channel. Biochemistry. 2013;52:279–281. doi: 10.1021/bi301350s. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 207.Pipatpolkai T, Corey RA, Proks P, Ashcroft FM, Stansfeld PJ. Evaluating inositol phospholipid interactions with inward rectifier potassium channels and characterising their role in disease. Commun Chem. 2020;3:147. doi: 10.1038/s42004-020-00391-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 208.Pipatpolkai T, Usher SG, Vedovato N, Ashcroft FM, Stansfeld PJ. The dynamic interplay of PIP2 and ATP in the regulation of the KATP channel. J Physiol. 2022;600:4503–4519. doi: 10.1113/JP283345. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 209.Bernsteiner H, Zangerl-Plessl E-M, Chen X, Stary-Weinzinger A. Conduction through a narrow inward-rectifier K+ channel pore. J Gen Physiol. 2019;151:1231–1246. doi: 10.1085/jgp.201912359. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 210.Chen X, Bründl M, Friesacher T, Stary-Weinzinger A. Computational insights into voltage dependence of polyamine block in a strong inwardly rectifying K+ channel. Front Pharmacol. 2020;11 doi: 10.3389/fphar.2020.00721. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 211.Bründl M, Pellikan S, Stary-Weinzinger A. Simulating PIP2-induced gating transitions in Kir6.2 channels. Front Mol Biosci. 2021;8 doi: 10.3389/fmolb.2021.711975. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 212.Lee S-J, Ren F, Zangerl-Plessl E-M, Heyman S, Stary-Weinzinger A, Yuan P, Nichols CG. Structural basis of control of inward rectifier Kir2 channel gating by bulk anionic phospholipids. J Gen Physiol. 2016;148:227–237. doi: 10.1085/jgp.201611616. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 213.Lacin E, Aryal P, Glaaser IW, Bodhinathan K, Tsai E, Marsh N, Tucker SJ, Sansom MSP, Slesinger PA. Dynamic role of the tether helix in PIP2-dependent gating of a G protein–gated potassium channel. J Gen Physiol. 2017;149:799–811. doi: 10.1085/jgp.201711801. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 214.Jogini V, Jensen MØ, Shaw DE. Gating and modulation of an inward-rectifier potassium channel. J Gen Physiol. 2022;155 doi: 10.1085/jgp.202213085. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215.Rodriguez-Menchaca AA, Adney SK, Tang Q-Y, Meng X-Y, Rosenhouse-Dantsker A, Cui M, Logothetis DE. PIP2 controls voltage-sensor movement and pore opening of Kv channels through the S4-S5 linker. Proc Natl Acad Sci. 2012;109:E2399–E2408. doi: 10.1073/pnas.1207901109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216.Kasimova MA, Zaydman MA, Cui J, Tarek M. PIP2-dependent coupling is prominent in Kv7.1 due to weakened interactions between S4-S5 and S6. Sci Rep. 2015;5:7474. doi: 10.1038/srep07474. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 217.Kasimova MA, Tarek M, Shaytan AK, Shaitan KV, Delemotte L. Voltage-gated ion channel modulation by lipids: Insights from molecular dynamics simulations. Biochimica et Biophysica Acta (BBA) – B Biomembranes. 2014;1838:1322–1331. doi: 10.1016/j.bbamem.2014.01.024. [DOI] [PubMed] [Google Scholar]
  • 218.Chen L, Zhang Q, Qiu Y, Li Z, Chen Z, Jiang H, Li Y, Yang H. Migration of PIP2 lipids on voltage-gated potassium channel surface influences channel deactivation. Sci Rep. 2015;5:15079. doi: 10.1038/srep15079. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 219.Zhang Q, Zhou P, Chen Z, Li M, Jiang H, Gao Z, Yang H. Dynamic PIP2 interactions with voltage sensor elements contribute to KCNQ2 channel gating. Proc Natl Acad Sci. 2013;110:20093–20098. doi: 10.1073/pnas.1312483110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 220.Abderemane-Ali F, Es-Salah-Lamoureux Z, Delemotte L, Kasimova MA, Labro AJ, Snyders DJ, Fedida D, Tarek M, Baró I, Loussouarn G. Dual effect of phosphatidyl (4,5)-bisphosphate PIP2 on Shaker K+ channels. J Biol Chem. 2012;287:36158–36167. doi: 10.1074/jbc.M112.382085. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 221.Pant S, Zhang J, Kim EC, Lam K, Chung HJ, Tajkhorshid E. PIP2-dependent coupling of voltage sensor and pore domains in Kv7.2 channel. Commun Biol. 2021;4:1189. doi: 10.1038/s42003-021-02729-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 222.Kirsch SA, Kugemann A, Carpaneto A, Böckmann RA, Dietrich P. Phosphatidylinositol-3,5-bisphosphate lipid-binding-induced activation of the human two-pore channel 2. Cell Mol Life Sci. 2018;75:3803–3815. doi: 10.1007/s00018-018-2829-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 223.Panasawatwong A, Pipatpolkai T, Tucker SJ. Transition between conformational states of the TREK-1 K2P channel promoted by interaction with PIP2. Biophys J. 2022;121:2380–2388. doi: 10.1016/j.bpj.2022.05.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 224.Yu K, Jiang T, Cui Y, Tajkhorshid E, Hartzell HC. A network of phosphatidylinositol 4,5-bisphosphate binding sites regulates gating of the Ca2+ activated Cl-channel ANO1 (TMEM16A) Proc Natl Acad Sci. 2019;116:19952–19962. doi: 10.1073/pnas.1904012116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 225.Jia Z, Chen J. Specific PIP2 binding promotes calcium activation of TMEM16A chloride channels. Commun Biol. 2021;4:259. doi: 10.1038/s42003-021-01782-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 226.Le SC, Jia Z, Chen J, Yang H. Molecular basis of PIP2-dependent regulation of the Ca2+-activated chloride channel TMEM16A. Nature Commun. 2019;10:3769. doi: 10.1038/s41467-019-11784-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 227.Buyan A, Cox CD, Barnoud J, Li J, Chan HSM, Martinac B, Marrink SJ, Corry B. Piezo1 forms specific, functionally important interactions with phosphoinositides and cholesterol. Biophys J. 2020;119:1683–1697. doi: 10.1016/j.bpj.2020.07.043. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 228.Lin Y, Tao E, Champion JP, Corry B. A binding site for phosphoinositides described by multiscale simulations explains their modulation of voltage-gated sodium channels. eLife. 2024;12:RP91218. doi: 10.7554/eLife.91218. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 229.Ivanova A, Atakpa-Adaji P, Rao S, Marti-Solano M, Taylor CW. Dual regulation of IP3 receptors by IP3 and PIP2 controls the transition from local to global Ca2+ signals. Mol Cell. 2024;84:3997–4015.:e3997. doi: 10.1016/j.molcel.2024.09.009. [DOI] [PubMed] [Google Scholar]
  • 230.Fruman DA, Meyers RE, Cantley LC. Phosphoinositide kinases. Annu Rev Biochem. 1998;67:481–507. doi: 10.1146/annurev.biochem.67.1.481. [DOI] [PubMed] [Google Scholar]
  • 231.Hsu F, Mao Y. The structure of phosphoinositide phosphatases: insights into substrate specificity and catalysis. Biochimica et Biophysica Acta (BBA) – Mol Cell Biol Lipids. 2015;1851:698–710. doi: 10.1016/j.bbalip.2014.09.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 232.Burke JE. Structural basis for regulation of phosphoinositide kinases and their involvement in human disease. Mol Cell. 2018;71:653–673. doi: 10.1016/j.molcel.2018.08.005. [DOI] [PubMed] [Google Scholar]
  • 233.Bunney TD, Katan M. Phosphoinositide signalling in cancer: beyond PI3K and PTEN. Nature Rev Cancer. 2010;10:342–352. doi: 10.1038/nrc2842. [DOI] [PubMed] [Google Scholar]
  • 234.Whited AM, Johs A. The interactions of peripheral membrane proteins with biological membranes. Chem Phys Lipids. 2015;192:51–59. doi: 10.1016/j.chemphyslip.2015.07.015. [DOI] [PubMed] [Google Scholar]
  • 235.Moravcevic K, Oxley CL, Lemmon MA. Conditional peripheral membrane proteins: facing up to limited specificity. Structure. 2012;20:15–27. doi: 10.1016/j.str.2011.11.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 236.Yamamoto E, Kalli AC, Yasuoka K, Sansom MSP. Interactions of pleckstrin homology domains with membranes: adding back the bilayer via high-throughput molecular dynamics. Structure. 2016;24:1421–1431. doi: 10.1016/j.str.2016.06.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 237.Wang Q, Pechersky Y, Sagawa S, Pan AC, Shaw DE. Structural mechanism for Bruton’s tyrosine kinase activation at the cell membrane. Proc Natl Acad Sci. 2019;116:9390–9399. doi: 10.1073/pnas.1819301116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 238.Pan AC, Jacobson D, Yatsenko K, Sritharan D, Weinreich TM, Shaw DE. Atomic-level characterization of protein–protein association. Proc Natl Acad Sci. 2019;116:4244–4249. doi: 10.1073/pnas.1815431116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 239.Lai C-L, Landgraf KE, Voth GA, Falke JJ. Membrane docking geometry and target lipid stoichiometry of membrane-bound PKCα C2 domain: a combined molecular dynamics and experimental study. J Mol Biol. 2010;402:301–310. doi: 10.1016/j.jmb.2010.07.037. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 240.Lai C-L, Srivastava A, Pilling C, Chase AR, Falke JJ, Voth GA. Molecular mechanism of membrane binding of the GRP1 PH domain. J Mol Biol. 2013;425:3073–3090. doi: 10.1016/j.jmb.2013.05.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 241.Zhang Y, Soubias O, Pant S, Heinrich F, Vogel A, Li J, Li Y, Clifton LA, Daum S, Bacia K, Huster D, et al. Myr-Arf1 conformational flexibility at the membrane surface sheds light on the interactions with ArfGAP ASAP1. Nature Commun. 2023;14:7570. doi: 10.1038/s41467-023-43008-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 242.Soubias O, Pant S, Heinrich F, Zhang Y, Roy NS, Li J, Jian X, Yohe ME, Randazzo PA, Lösche M, Tajkhorshid E, et al. Membrane surface recognition by the ASAP1 PH domain and consequences for interactions with the small GTPase Arf1. Sci Adv. 2020;6:eabd1882. doi: 10.1126/sciadv.abd1882. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 243.Pant S, Tajkhorshid E. Microscopic characterization of GRP1 PH domain interaction with anionic membranes. J Comput Chem. 2020;41:489–499. doi: 10.1002/jcc.26109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 244.Gregory MC, McLean MA, Sligar SG. Interaction of KRas4b with anionic membranes: a special role for PIP2. Biochem Biophys Res Commun. 2017;487:351–355. doi: 10.1016/j.bbrc.2017.04.063. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 245.Ohkubo YZ, Pogorelov TV, Arcario MJ, Christensen GA, Tajkhorshid E. Accelerating membrane insertion of peripheral proteins with a novel membrane mimetic model. Biophys J. 2012;102:2130–2139. doi: 10.1016/j.bpj.2012.03.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 246.Naughton FB, Kalli AC, Sansom MSP. Modes of interaction of pleckstrin homology domains with membranes: toward a computational biochemistry of membrane recognition. J Mol Biol. 2018;430:372–388. doi: 10.1016/j.jmb.2017.12.011. [DOI] [PubMed] [Google Scholar]
  • 247.Vonkova I, Saliba A-E, Deghou S, Anand K, Ceschia S, Doerks T, Galih A, Kugler KG, Maeda K, Rybin V, van Noort V, et al. Lipid cooperativity as a general membrane-recruitment principle for PH domains. Cell Rep. 2015;12:1519–1530. doi: 10.1016/j.celrep.2015.07.054. [DOI] [PubMed] [Google Scholar]
  • 248.Yamamoto E, Domański J, Naughton FB, Best RB, Kalli AC, Stansfeld PJ, Sansom MSP. Multiple lipid binding sites determine the affinity of PH domains for phosphoinositide-containing membranes. Sci Adv. 2020;6:eaay5736. doi: 10.1126/sciadv.aay5736. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 249.Le Huray KIP, Wang H, Sobott F, Kalli AC. Systematic simulation of the interactions of pleckstrin homology domains with membranes. Sci Adv. 2022;8:eabn6992. doi: 10.1126/sciadv.abn6992. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 250.Braun L, Schoen I, Vogel V. PIP2-induced membrane binding of the vinculin tail competes with its other binding partners. Biophys J. 2021;120:4608–4622. doi: 10.1016/j.bpj.2021.08.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 251.Buyan A, Kalli AC, Sansom MSP. Multiscale simulations suggest a mechanism for the association of the Dok7 PH domain with PIP-containing membranes. PLoS Comput Biol. 2016;12:e1005028. doi: 10.1371/journal.pcbi.1005028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 252.Cao S, Chung S, Kim S, Li Z, Manor D, Buck M. K-Ras G-domain binding with signaling lipid phosphatidylinositol (4,5)-phosphate (PIP2): membrane association, protein orientation, and function. J Biol Chem. 2019;294:7068–7084. doi: 10.1074/jbc.RA118.004021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 253.Shrestha R, Carpenter TS, Van QN, Agamasu C, Tonelli M, Aydin F, Chen D, Gulten G, Glosli JN, López CA, Oppelstrup T, et al. Membrane lipids drive formation of KRAS4b-RAF1 RBDCRD nanoclusters on the membrane. Commun Biol. 2024;7:242. doi: 10.1038/s42003-024-05916-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 254.Han K, Pastor RW, Fenollar-Ferrer C. PLD2–PI(4,5)P2 interactions in fluid phase membranes: structural modeling and molecular dynamics simulations. PLoS One. 2020;15:e0236201. doi: 10.1371/journal.pone.0236201. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 255.Herzog FA, Braun L, Schoen I, Vogel V. Structural insights how PIP2 imposes preferred binding orientations of FAK at lipid membranes. J Phys Chem B. 2017;121:3523–3535. doi: 10.1021/acs.jpcb.6b09349. [DOI] [PubMed] [Google Scholar]
  • 256.Zhou J, Bronowska A, Le Coq J, Lietha D, Gräter F. Allosteric regulation of focal adhesion kinase by PIP2 and ATP. Biophys J. 2015;108:698–705. doi: 10.1016/j.bpj.2014.11.3454. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 257.John LH, Naughton FB, Sansom MSP, Larsen AH. The role of C2 domains in two different phosphatases: PTEN and SHIP2. Membranes. 2023;13:408. doi: 10.3390/membranes13040408. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 258.Kalli AC, Devaney I, Sansom MSP. Interactions of phosphatase and tensin homologue (PTEN) proteins with phosphatidylinositol phosphates: insights from molecular dynamics simulations of PTEN and voltage sensitive phosphatase. Biochemistry. 2014;53:1724–1732. doi: 10.1021/bi5000299. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 259.Karandur D, Nawrotek A, Kuriyan J, Cherfils J. Multiple interactions between an Arf/GEF complex and charged lipids determine activation kinetics on the membrane. Proc Natl Acad Sci. 2017;114:11416–11421. doi: 10.1073/pnas.1707970114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 260.Le Huray KIP, Bunney TD, Pinotsis N, Kalli AC, Katan M. Characterization of the membrane interactions of phospholipase Cγ reveals key features of the active enzyme. Sci Adv. 2022;8:eabp9688. doi: 10.1126/sciadv.abp9688. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 261.Liepiną I, Czaplewski C, Janmey P, Liwo A. Molecular dynamics study of a gelsolin-derived peptide binding to a lipid bilayer containing phosphatidylinositol 4,5-bisphosphate. Pept Sci. 2003;71:49–70. doi: 10.1002/bip.10375. [DOI] [PubMed] [Google Scholar]
  • 262.Moser von Filseck J, Čopič A, Delfosse V, Vanni S, Jackson CL, Bourguet W, Drin G. Phosphatidylserine transport by ORP/Osh proteins is driven by phosphatidylinositol 4-phosphate. Science. 2015;349:432–436. doi: 10.1126/science.aab1346. [DOI] [PubMed] [Google Scholar]
  • 263.von Filseck JM, Vanni S, Mesmin B, Antonny B, Drin G. A phosphatidylinositol-4-phosphate powered exchange mechanism to create a lipid gradient between membranes. Nature Commun. 2015;6:6671. doi: 10.1038/ncomms7671. [DOI] [PubMed] [Google Scholar]
  • 264.Allsopp RJ, Klauda JB. Impact of PIP2 lipids, force field parameters, and mutational analysis on the binding of the Osh4’s α6–α7 domain. J Phys Chem B. 2021;125:5296–5308. doi: 10.1021/acs.jpcb.0c10393. [DOI] [PubMed] [Google Scholar]
  • 265.Naughton FB, Kalli AC, Sansom MSP. Association of peripheral membrane proteins with membranes: free energy of binding of GRP1 PH domain with phosphatidylinositol phosphate-containing model bilayers. J Phys Chem Lett. 2016;7:1219–1224. doi: 10.1021/acs.jpclett.6b00153. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 266.Ni T, Kalli AC, Naughton FB, Yates LA, Naneh O, Kozorog M, Anderluh G, Sansom MSP, Gilbert RJC. Structure and lipid-binding properties of the kindlin-3 pleckstrin homology domain. Biochem J. 2017;474:539–556. doi: 10.1042/BCJ20160791. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 267.Picas L, Viaud J, Schauer K, Vanni S, Hnia K, Fraisier V, Roux A, Bassereau P, Gaits-Iacovoni F, Payrastre B, Laporte J, et al. BIN1/M-Amphiphysin2 induces clustering of phosphoinositides to recruit its downstream partner dynamin. Nature Commun. 2014;5:5647. doi: 10.1038/ncomms6647. [DOI] [PubMed] [Google Scholar]
  • 268.Prakaash D, Fagnen C, Cook GP, Acuto O, Kalli AC. Molecular dynamics simulations reveal membrane lipid interactions of the full-length lymphocyte specific kinase (Lck) Sci Rep. 2022;12:21121. doi: 10.1038/s41598-022-25603-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 269.Scott JL, Frick CT, Johnson KA, Liu H, Yong SS, Varney AG, Wiest O, Stahelin RV. Molecular analysis of membrane targeting by the C2 domain of the E3 ubiquitin ligase Smurf1. Biomolecules. 2020;10:229. doi: 10.3390/biom10020229. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 270.Xie H, Weinstein H. The selective cargo loading and release functions of the cellular sterol transporter StarD4 are allosterically controlled by the recognition of specific PIP2 subtypes in the targeted membranes. BioRxiv. 2024 doi: 10.1101/2024.03.26.586881. [DOI] [Google Scholar]
  • 271.Talandashti R, van Ek L, Gehin C, Xue D, Moqadam M, Gavin A-C, Reuter N. Membrane specificity of the human cholesterol transfer protein STARD4. J Mol Biol. 2024;436:168572. doi: 10.1016/j.jmb.2024.168572. [DOI] [PubMed] [Google Scholar]
  • 272.Xu T, Gan Q, Wu B, Yin M, Xu J, Shu X, Liu J. Molecular basis for PI(3,5)P2 recognition by SNX11, a protein involved in lysosomal degradation and endosome homeostasis regulation. J Mol Biol. 2020;432:4750–4761. doi: 10.1016/j.jmb.2020.06.010. [DOI] [PubMed] [Google Scholar]
  • 273.Rao S, Skulsuppaisarn M, Strong LM, Ren X, Lazarou M, Hurley JH, Hummer G. Three-step docking by WIPI2, ATG16L1, and ATG3 delivers LC3 to the phagophore. Sci Adv. 2024;10:eadj8027. doi: 10.1126/sciadv.adj8027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 274.Thallmair V, Schultz L, Zhao W, Marrink SJ, Oliver D, Thallmair S. Two cooperative binding sites sensitize PI(4,5)P2 recognition by the tubby domain. Sci Adv. 2022;8:eabp9471. doi: 10.1126/sciadv.abp9471. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 275.Grabon A, Orłowski A, Tripathi A, Vuorio J, Javanainen M, Róg T, Lönnfors M, McDermott MI, Siebert G, Somerharju P, Vattulainen I, et al. Dynamics and energetics of the mammalian phosphatidylinositol transfer protein phospholipid exchange cycle. J Biol Chem. 2017;292:14438–14455. doi: 10.1074/jbc.M117.791467. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 276.Williams SI, Yu X, Ni T, Gilbert RJC, Stansfeld PJ. Structural, functional and computational studies of membrane recognition by plasmodium perforin-like proteins 1 and 2. J Mol Biol. 2022;434:167642. doi: 10.1016/j.jmb.2022.167642. [DOI] [PubMed] [Google Scholar]
  • 277.Souza PCT, Borges-Araújo L, Brasnett C, Moreira RA, Grünewald F, Park P, Wang L, Razmazma H, Borges-Araújo AC, Cofas-Vargas LF, Monticelli L, et al. GoMartini 3: From large conformational changes in proteins to environmental bias corrections. BioRxiv. 2024 doi: 10.1101/2024.04.15.589479. [DOI] [Google Scholar]
  • 278.Larsen AH, John LH, Sansom MSP, Corey RA. Specific interactions of peripheral membrane proteins with lipids: what can molecular simulations show us? Biosci Rep. 2022;42 doi: 10.1042/BSR20211406. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 279.Schlessinger J. Cell signaling by receptor tyrosine kinases. Cell. 2000;103:211–225. doi: 10.1016/s0092-8674(00)00114-8. [DOI] [PubMed] [Google Scholar]
  • 280.Lemmon MA, Schlessinger J. Cell signaling by receptor tyrosine kinases. Cell. 2010;141:1117–1134. doi: 10.1016/j.cell.2010.06.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 281.Karl K, Hristova K. Pondering the mechanism of receptor tyrosine kinase activation: the case for ligand-specific dimer microstate ensembles. Curr Opin Struct Biol. 2021;71:193–199. doi: 10.1016/j.sbi.2021.07.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 282.Paul MD, Hristova K. The RTK interactome: overview and perspective on RTK heterointeractions. Chem Rev. 2019;119:5881–5921. doi: 10.1021/acs.chemrev.8b00467. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 283.McLaughlin S, Smith SO, Hayman MJ, Murray D. An electrostatic engine model for autoinhibition and activation of the epidermal growth factor receptor (EGFR/ErbB) family. J Gen Physiol. 2005;126:41–53. doi: 10.1085/jgp.200509274. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 284.Wang Y, Gao J, Guo X, Tong T, Shi X, Li L, Qi M, Wang Y, Cai M, Jiang J, Xu C, et al. Regulation of EGFR nanocluster formation by ionic protein-lipid interaction. Cell Res. 2014;24:959–976. doi: 10.1038/cr.2014.89. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 285.Singh PK, Rybak JA, Schuck RJ, Sahoo AR, Buck M, Barrera FN, Smith AW. Phosphatidylinositol 4,5-bisphosphate drives the formation of EGFR and EphA2 complexes. Sci Adv. 2024;10:eadl0649. doi: 10.1126/sciadv.adl0649. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 286.Gonzalez-Magaldi M, Gullapalli A, Papoulas O, Liu C, Leung AYH, Guo L, Brilot A, Marcotte EM, Ke Z, Leahy DJ. Structure and organization of full-length epidermal growth factor receptor in extracellular vesicles by cryo-electron tomography. BioRxiv. 2024 doi: 10.1101/2024.11.25.625301. [DOI] [Google Scholar]
  • 287.Kumar R, Goel H, Solanki R, Rawat L, Tabasum S, Tanwar P, Pal S, Sabarwal A. Recent developments in receptor tyrosine kinase inhibitors: a promising mainstay in targeted cancer therapy. Med Drug Discovery. 2024;23:100195. doi: 10.1016/j.medidd.2024.100195. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 288.Gschwind A, Fischer OM, Ullrich A. The discovery of receptor tyrosine kinases: targets for cancer therapy. Nature Rev Cancer. 2004;4:361–370. doi: 10.1038/nrc1360. [DOI] [PubMed] [Google Scholar]
  • 289.Kovacs E, Zorn JA, Huang Y, Barros T, Kuriyan J. A structural perspective on the regulation of the epidermal growth factor receptor. Annu Rev Biochem. 2015;84:739–764. doi: 10.1146/annurev-biochem-060614-034402. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 290.Uchikawa E, Choi E, Shang G, Yu H, Bai X-C. Activation mechanism of the insulin receptor revealed by Cryo-EM structure of the fully liganded receptor–ligand complex. eLife. 2019;8:e48630. doi: 10.7554/eLife.48630. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 291.Schlessinger J. Receptor tyrosine kinases: legacy of the first two decades. Cold Spring Harb Perspect Biol. 2014;6 doi: 10.1101/cshperspect.a008912. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 292.Thiel KW, Carpenter G. Epidermal growth factor receptor juxtamembrane region regulates allosteric tyrosine kinase activation. Proc Natl Acad Sci. 2007;104:19238–19243. doi: 10.1073/pnas.0703854104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 293.Abd Halim KB, Koldsø H, Sansom MSP. Interactions of the EGFR juxtamembrane domain with PIP2-containing lipid bilayers: Insights from multiscale molecular dynamics simulations. Biochimica et Biophysica Acta (BBA) – Gen Subj. 2015;1850:1017–1025. doi: 10.1016/j.bbagen.2014.09.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 294.Wang Z, Fan H, Hu X, Khamo J, Diao J, Zhang K, Pogorelov TV. Coaction of electrostatic and hydrophobic interactions: dynamic constraints on disordered TrkA juxtamembrane domain. J Phys Chem B. 2019;123:10709–10717. doi: 10.1021/acs.jpcb.9b09352. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 295.Chavent M, Karia D, Kalli AC, Domański J, Duncan AL, Hedger G, Stansfeld PJ, Seiradake E, Jones EY, Sansom MSP. Interactions of the EphA2 kinase domain with PIPs in membranes: implications for receptor function. Structure. 2018;26:1025–1034.:e1022. doi: 10.1016/j.str.2018.05.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 296.Singh PK, Rybak JA, Schuck RJ, Barrera FN, Smith AW. Phosphatidylinositol (4,5)-bisphosphate drives the formation of EGFR and EphA2 complexes. BioRxiv. 2024 doi: 10.1101/2024.05.03.592400. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 297.Abe M, Yanagawa M, Hiroshima M, Kobayashi T, Sako Y. Bilateral regulation of EGFR activity and local PI(4,5)P2 dynamics in mammalian cells observed with superresolution microscopy. eLife. 2024;13:e101652. doi: 10.7554/eLife.101652. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 298.Hedger G, Shorthouse D, Koldsø H, Sansom MSP. Free energy landscape of lipid interactions with regulatory binding sites on the transmembrane domain of the EGF receptor. J Phys Chem B. 2016;120:8154–8163. doi: 10.1021/acs.jpcb.6b01387. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 299.Chavent M, Chetwynd AP, Stansfeld PJ, Sansom MSP. Dimerization of the EphA1 receptor tyrosine kinase transmembrane domain: insights into the mechanism of receptor activation. Biochemistry. 2014;53:6641–6652. doi: 10.1021/bi500800x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 300.Lelimousin M, Limongelli V, Sansom MSP. Conformational changes in the epidermal growth factor receptor: role of the transmembrane domain investigated by coarse-grained metadynamics free energy calculations. J Am Chem Soc. 2016;138:10611–10622. doi: 10.1021/jacs.6b05602. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 301.Bocharov EV, Mineev KS, Pavlov KV, Akimov SA, Kuznetsov AS, Efremov RG, Arseniev AS. Helix-helix interactions in membrane domains of bitopic proteins: Specificity and role of lipid environment. Biochimica et Biophysica Acta (BBA) – B Biomembranes. 2017;1859:561–576. doi: 10.1016/j.bbamem.2016.10.024. [DOI] [PubMed] [Google Scholar]
  • 302.Stangl M, Schneider D. Functional competition within a membrane: Lipid recognition vs. transmembrane helix oligomerization. Biochimica et Biophysica Acta (BBA) – Biomembranes. 2015;1848:1886–1896. doi: 10.1016/j.bbamem.2015.03.011. [DOI] [PubMed] [Google Scholar]
  • 303.Stefanski KM, Russell CM, Westerfield JM, Lamichhane R, Barrera FN. PIP2 promotes conformation-specific dimerization of the EphA2 membrane region. J Biol Chem. 2021;296 doi: 10.1074/jbc.RA120.016423. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 304.Lolicato F, Nickel W, Haucke V, Ebner M. Phosphoinositide switches in cell physiology – from molecular mechanisms to disease. J Biol Chem. 2024;300 doi: 10.1016/j.jbc.2024.105757. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 305.Dimou E, Nickel W. Unconventional mechanisms of eukaryotic protein secretion. Curr Biol. 2018;28:R406–R410. doi: 10.1016/j.cub.2017.11.074. [DOI] [PubMed] [Google Scholar]
  • 306.Pallotta MT, Nickel W. FGF2 and IL-1b – explorers of unconventional secretory pathways at a glance. J Cell Sci. 2020;133 doi: 10.1242/jcs.250449. [DOI] [PubMed] [Google Scholar]
  • 307.Steringer JP, Lange S, Čujová S, Šachl R, Poojari C, Lolicato F, Beutel O, Müller HM, Unger S, Coskun Ü, Honigmann A, et al. Key steps in unconventional secretion of fibroblast growth factor 2 reconstituted with purified components. eLife. 2017;6:e28985. doi: 10.7554/eLife.28985. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 308.Khelashvili G, Stanley N, Sahai MA, Medina J, LeVine MV, Shi L, De Fabritiis G, Weinstein H. Spontaneous inward opening of the dopamine transporter is triggered by PIP2-regulated dynamics of the N-terminus. ACS Chem Nerosci. 2015;6:1825–1837. doi: 10.1021/acschemneuro.5b00179. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 309.Tang H, Li H, Prakaash D, Pedebos C, Qiu X, Sauer DB, Khalid S, Duerr K, Robinson CV. The solute carrier SPNS2 recruits PI(4,5)P2 to synergistically regulate transport of sphingosine-1-phosphate. Mol Cell. 2023;83:2739–2752.:e2735. doi: 10.1016/j.molcel.2023.06.033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 310.Zhekova HR, Ramirez Echemendía DP, Sejdiu BI, Pushkin A, Tieleman DP, Kurtz I. Molecular dynamics simulations of lipid-protein interactions in SLC4 proteins. Biophys J. 2024;123:1705–1721. doi: 10.1016/j.bpj.2024.05.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 311.Qin X, Tieleman DP, Liang Q. Effects of cholesterol and PIP2 on interactions between glycophorin A and Band 3 in lipid bilayers. Biophys J. 2022;121:2069–2077. doi: 10.1016/j.bpj.2022.05.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 312.Chen T, Vallese F, Gil-Iturbe E, Kim K, Quick M, Clarke O, Tajkhorshid E. Impact of anionic lipids on the energy landscape of conformational transition in anion exchanger 1 (AE1) BioRxiv. 2024 doi: 10.1101/2024.10.15.616105. [DOI] [Google Scholar]
  • 313.Kokane S, Meier PF, Gulati A, Matsuoka R, Pipatpolkai T, Albano G, Delemotte L, Fuster D, Drew D. PI-(3,5)P2-mediated oligomerization of the endosomal sodium/proton exchanger NHE9. BioRxiv. 2024 doi: 10.1101/2023.09.10.557050. [DOI] [Google Scholar]
  • 314.Pyle E, Kalli AC, Amillis S, Hall Z, Lau AM, Hanyaloglu AC, Diallinas G, Byrne B, Politis A. Structural lipids enable the formation of functional oligomers of the eukaryotic purine symporter UapA. Cell Chem Biol. 2018;25:840–848.:e844. doi: 10.1016/j.chembiol.2018.03.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 315.Norris MJ, Husby ML, Kiosses WB, Yin J, Saxena R, Rennick LJ, Heiner A, Harkins SS, Pokhrel R, Schendel SL, Hastie KM, et al. Measles and Nipah virus assembly: specific lipid binding drives matrix polymerization. Sci Adv. 2022;8:eabn1440. doi: 10.1126/sciadv.abn1440. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 316.Gc JB, Gerstman BS, Chapagain PP. Membrane association and localization dynamics of the Ebola virus matrix protein VP40. Biochimica et Biophysica Acta (BBA) – Biomembranes. 2017;1859:2012–2020. doi: 10.1016/j.bbamem.2017.07.007. [DOI] [PubMed] [Google Scholar]
  • 317.Winski D, Parent MT, Wallace JN, Shivanna K, Aho BM, Waters H, Zimmerberg J, Sodt AJ, Hess ST. The cytoplasmic tail domain of influenza hemagglutinin interacts with phosphatidylinositol (4,5)-bisphosphate (PIP2) Biophys J. 2024;123:236a–237a. doi: 10.1016/j.bpj.2023.11.1498. [DOI] [Google Scholar]
  • 318.Cioffi MD, Sharma T, Motsa BB, Bhattarai N, Gerstman BS, Stahelin RV, Chapagain PP. Ebola virus matrix protein VP40 single mutations G198R and G201R significantly enhance plasma membrane localization. J Phys Chem B. 2024;128:11335–11344. doi: 10.1021/acs.jpcb.4c02700. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 319.Hudait A, Hurley JH, Voth GA. Dynamics of upstream ESCRT organization at the HIV-1 budding site. Biophys J. 2023;122:2655–2674. doi: 10.1016/j.bpj.2023.05.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 320.Banerjee P, Monje-Galvan V, Voth GA. Cooperative membrane binding of HIV-1 matrix proteins. J Phys Chem B. 2024;128:2595–2606. doi: 10.1021/acs.jpcb.3c06222. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 321.Monje-Galvan V, Voth GA. Binding mechanism of the matrix domain of HIV-1 gag on lipid membranes. eLife. 2020;9:e58621. doi: 10.7554/eLife.58621. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 322.Senju Y, Kalimeri M, Koskela EV, Somerharju P, Zhao H, Vattulainen I, Lappalainen P. Mechanistic principles underlying regulation of the actin cytoskeleton by phosphoinositides. Proc Natl Acad Sci. 2017;114:E8977–E8986. doi: 10.1073/pnas.1705032114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 323.Prakash S, Krishna A, Sengupta D. Cofilin-membrane interactions: electrostatic effects in phosphoinositide lipid binding. ChemPhysChem. 2023;24:e202200509. doi: 10.1002/cphc.202200509. [DOI] [PubMed] [Google Scholar]
  • 324.Fatunmbi O, Bradley RP, Kandy SK, Bucki R, Janmey PA, Radhakrishnan R. A multiscale biophysical model for the recruitment of actin nucleating proteins at the membrane interface. Soft Matter. 2020;16:4941–4954. doi: 10.1039/d0sm00267d. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 325.Honigmann A, van den Bogaart G, Iraheta E, Risselada HJ, Milovanovic D, Mueller V, Müllar S, Diederichsen U, Fasshauer D, Grubmüller H, Hell SW, et al. Phosphatidylinositol 4,5-bisphosphate clusters act as molecular beacons for vesicle recruitment. Nature Struct Mol Biol. 2013;20:679–686. doi: 10.1038/nsmb.2570. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 326.Schaefer SL, Hummer G. Sublytic Gasdermin-D pores captured in atomistic molecular simulations. eLife. 2022;11:e81432. doi: 10.7554/eLife.81432. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 327.Koldsø H, Shorthouse D, Hélie J, Sansom MSP. Lipid clustering correlates with membrane curvature as revealed by molecular simulations of complex lipid bilayers. PLoS Comput Biol. 2014;10:e1003911. doi: 10.1371/journal.pcbi.1003911. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 328.Lee KI, Im W, Pastor RW. Langevin dynamics simulations of charged model phosphatidylinositol lipids in the presence of diffusion barriers: toward an atomic level understanding of corralling of PIP2 by protein fences in biological membranes. BMC Biophys. 2014;7:13. doi: 10.1186/s13628-014-0013-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 329.Han K, Gericke A, Pastor RW. Characterization of specific ion effects on PI(4,5)P2 clustering: molecular dynamics simulations and graph-theoretic analysis. J Phys Chem B. 2020;124:1183–1196. doi: 10.1021/acs.jpcb.9b10951. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 330.Wang Y-H, Slochower DR, Janmey PA. Counterion-mediated cluster formation by polyphosphoinositides. Chem Phys Lipids. 2014;182:38–51. doi: 10.1016/j.chemphyslip.2014.01.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 331.Han K, Kim SH, Venable RM, Pastor RW. Design principles of PI(4,5)P2 clustering under protein-free conditions: specific cation effects and calcium-potassium synergy. Proc Natl Acad Sci. 2022;119:e2202647119. doi: 10.1073/pnas.2202647119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 332.Bradley RP, Slochower DR, Janmey PA, Radhakrishnan R. Divalent cations bind to phosphoinositides to induce ion and isomer specific propensities for nano-cluster initiation in bilayer membranes. R Soc Open Sci. 2020;7:192208. doi: 10.1098/rsos.192208. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 333.Sarmento MJ, Coutinho A, Fedorov A, Prieto M, Fernandes F. Membrane order is a key regulator of divalent cation-induced clustering of PI(3,5)P2 and PI(4,5)P2. Langmuir. 2017;33:12463–12477. doi: 10.1021/acs.langmuir.7b00666. [DOI] [PubMed] [Google Scholar]
  • 334.Bilkova E, Pleskot R, Rissanen S, Sun S, Czogalla A, Cwiklik L, Rog T, Vattulainen U, CCramer PS, Jungwirth P, Coskun U. Calcium Directly Regulates Phosphatidylinositol 4,5-Bisphosphate Headgroup Conformation and Recognition. J Am Chem Soc. 2017;139:4019–4024. doi: 10.1021/jacs.6b11760. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 335.Jarin Z, Pak AJ, Bassereau P, Voth GA. Lipid-composition-mediated forces can stabilize tubular assemblies of I-BAR proteins. Biophys J. 2021;120:46–54. doi: 10.1016/j.bpj.2020.11.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 336.Blood PD, Swenson RD, Voth GA. Factors influencing local membrane curvature induction by N-BAR domains as revealed by molecular dynamics simulations. Biophys J. 2008;95:1866–1876. doi: 10.1529/biophysj.107.121160. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 337.Tsai F-C, Henderson JM, Jarin Z, Kremneva E, Senju Y, Pernier J, Mikhajlov O, Manzi J, Kogan K, Le Clainche C, Voth GA, et al. Activated I-BAR IRSp53 clustering controls the formation of VASP-actin–based membrane protrusions. Sci Adv. 2022;8:eabp8677. doi: 10.1126/sciadv.abp8677. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 338.Kefauver JM, Hakala M, Zou L, Alba J, Espadas J, Tettamanti MG, Gajić J, Gabus C, Campomanes P, Estrozi LF, Sen NE, et al. Cryo-EM architecture of a near-native stretch-sensitive membrane microdomain. Nature. 2024;632:664–671. doi: 10.1038/s41586-024-07720-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 339.Belessiotis-Richards A, Higgins SG, Sansom MSP, Alexander-Katz A, Stevens MM. Coarse-grained simulations suggest the epsin N-terminal homology domain can sense membrane curvature without its terminal amphipathic helix. ACS Nano. 2020;14:16919–16928. doi: 10.1021/acsnano.0c05960. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 340.Liu K-C, Pace H, Larsson E, Hossain S, Kabedev A, Shukla A, Jerschabek V, Mohan J, Bergström CAS, Bally M, Schwieger C, et al. Membrane insertion mechanism of the caveola coat protein Cavin1. Proc Natl Acad Sci. 2022;119:e2202295119. doi: 10.1073/pnas.2202295119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 341.Belessiotis-Richards A, Larsen AH, Higgins SG, Stevens MM, Alexander-Katz A. Coarse-grained simulations suggest potential competing roles of phosphoinositides and amphipathic helix structures in membrane curvature sensing of the AP180 N-terminal homology domain. J Phys Chem B. 2022;126:2789–2797. doi: 10.1021/acs.jpcb.2c00239. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 342.Brown CM, Marrink SJ. Modeling membranes in situ. Curr Opin Struct Biol. 2024;87:102837. doi: 10.1016/j.sbi.2024.102837. [DOI] [PubMed] [Google Scholar]
  • 343.Grünewald F, Souza PCT, Abdizadeh H, Barnoud J, de Vries AH, Marrink SJ. Titratable Martini model for constant pH simulations. J Chem Phys. 2020;153:024118. doi: 10.1063/5.0014258. [DOI] [PubMed] [Google Scholar]
  • 344.Bignucolo O, Chipot C, Kellenberger S, Roux B. Galvani offset potential and constant-pH simulations of membrane proteins. J Phys Chem B. 2022;126:6868–6877. doi: 10.1021/acs.jpcb.2c04593. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 345.Domanski J, Hedger G, Best RB, Stansfeld PJ, Sansom MSP. Convergence and Sampling in Determining Free Energy Landscapes for Membrane Protein Association. J Phys Chem B. 2017;121:3364–3375. doi: 10.1021/acs.jpcb.6b08445. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 346.Grime J, Madsen JJ. Efficient simulation of tunable lipid assemblies across scales and resolutions. arXiv preprint. 2019:arXiv:1910.05362. doi: 10.48550/arXiv.1910.05362. [DOI] [Google Scholar]
  • 347.Amaro R, Åqvist J, Bahar I, Battistini F, Bellaiche A, Beltran D, Biggin PC, Bonomi M, Bowman GR, Bryce R. The need to implement FAIR principles in biomolecular simulations. arXiv preprint. 2024:arXiv:2407.16584. doi: 10.48550/arXiv.2407.16584. [DOI] [Google Scholar]
  • 348.Payandeh J, Volgraf M. Ligand binding at the protein–lipid interface: strategic considerations for drug design. Nature Rev Drug Discov. 2021;20:710–722. doi: 10.1038/s41573-021-00240-2. [DOI] [PubMed] [Google Scholar]
  • 349.Conflitti P, Lyman E, Sansom MSP, Hildebrand PW, Gutierrez-de-Teran H, Carloni P, Ansell TB, Yuan S, Barth P, Robinson AS, Tate CG, et al. Functional dynamics of G protein-couplied receptors reveal new routes for drug discovery. Nature Rev Drug Discov. 2025 doi: 10.1038/s41573-024-01083-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 350.Bai X-C, Gonen T, Gronenborn AM, Perrakis A, Thorn A, Yang J. Challenges and opportunities in macromolecular structure determination. Nature Rev Mol Cell Biol. 2024;25:7–12. doi: 10.1038/s41580-023-00659-y. [DOI] [PubMed] [Google Scholar]
  • 351.Papasergi-Scott MM, Pérez-Hernández G, Batebi H, Gao Y, Eskici G, Seven AB, Panova O, Hilger D, Casiraghi M, He F, Maul L, et al. Time-resolved cryo-EM of G-protein activation by a GPCR. Nature. 2024;629:1182–1191. doi: 10.1038/s41586-024-07153-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 352.Amann SJ, Keihsler D, Bodrug T, Brown NG, Haselbach D. Frozen in time: analyzing molecular dynamics with time-resolved cryo-EM. Structure. 2023;31:4–19. doi: 10.1016/j.str.2022.11.014. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES