Graphical abstract
Keywords: K+ channels, phospholipids, phosphatidylinositol-bisphosphate, protein-lipid interactions, molecular dynamics simulations, electrophysiology, flux assays, mass spectrometry, Super-resolution microscopy, X-ray crystallography, Cryo-EM microscopy
Highlights
-
•
Interactions of lipids with K+ channels.
-
•
Structural basis of PIP2 binding.
-
•
Experimental methods for quantifying PIP2-binding to K+ channels.
-
•
Molecular simulation methods for studying PIP2 binding.
Abstract
Potassium (K+) ion channels are crucial in numerous cellular processes as they hyperpolarise a cell through K+ conductance, returning a cell to its resting potential. K+ channel mutations result in multiple clinical complications such as arrhythmia, neonatal diabetes and migraines. Since 1995, the regulation of K+ channels by phospholipids has been heavily studied using a range of interdisciplinary methods such as cellular electrophysiology, structural biology and computational modelling. As a result, K+ channels are model proteins for the analysis of protein-lipid interactions. In this review, we will focus on the roles of lipids in the regulation of K+ channels, and how atomic-level structures, along with experimental techniques and molecular simulations, have helped guide our understanding of the importance of phospholipid interactions.
Introduction
In the mid 1990s, Rebecca Ball and Don Hilgemann identified that a lipid involved in cell signalling, phosphatidylinositol-4,5-bisphosphate (PIP2), was also able to directly interact with ion channels.1 Since then, multiple methodologies and concepts have been developed to understand the roles for lipids as regulators of ion channels.2 The determination of potassium (K+) channel structures, and the development of computational tools to study them, have also greatly contributed to our perspectives on the molecular basis of ion channel interactions with lipids.3, 4, 5
K+ channels are integral membrane proteins which facilitate the passive movement of K+ ions across the membrane. The selective permeation of ions induce ionic gradients that are crucial in cellular communications, metabolism and homeostasis.6 Mutations of K+ channels cause many clinical symptoms such as cardiac arrhythmia, neonatal diabetes and migraines.7, 8, 9 As K+ channels are transmembrane (TM) proteins, one of their key modulators is their lipid environment, with membrane lipids binding both specifically and/or indiscriminately to the proteins.10 Mechanistic understanding of protein-lipid interactions is essential to inform research on novel therapeutic agents that may modulate ion channel activity in diseases.11
In this review, we will discuss our current understanding of the diversity of K+ channels and how these channels are modulated by phospholipids. We discuss how the interactions with lipids impact on channel gating and/or stabilise distinct conformations of the proteins, and review how complementary techniques may be applied, detailing how experimental and computational approaches may be applied to calculate values for protein-lipid binding. Finally, we discuss how knowledge of ion channel gating has shed light on therapeutic approaches to clinical disease.
K+channel diversity
K+ ions are transported through a central pore domain that is common to all members of the K+ channel family (Figure 1). This domain comprises two TM helices linked by a pore (P) half-helix and a highly-conserved ion selectivity filter.12 Four copies of the pore topology must be present to form a functional pore, and therefore most K+ channels are tetramers. An exception is the 2-pore (K2P) family where each monomer has two repeated pore topologies, and therefore requires only a dimer of proteins (Figure 1(C)). K+ channels are further functionalised though additional domains that regulate gating of the channel.13
Figure 1.
Topological representation and atomic structures of K+ channels. (A) Voltage-gated potassium (Kv) channel topology. (B) A structural example for of a Kv channel: hKCNQ1 (PDB ID: 6V01).56(C) Two-pore potassium (K2P) channel topology. (D) Example structure of the K2P channel, mTREK-1 (PDB ID: 6V37). (E) Ca2+ gated potassium (KCa) channel topology. (F) Structure of the KCa channel hSlo1 (PDB ID: 6V38).143(G) Inwardly rectified potassium (Kir) channel topology. (H) Structure of the cKir2.2 channel (PDB ID: 3SPI).144 Topology of potassium channels where transmembrane helices are represented as cylinders. The structures are shown in cartoon. The voltage sensor domain (VSD) is coloured green. The pore domains are all represented in blue. The cytoplasmic domain of Kv, KCa and Kir are shown in yellow, red and lilac respectively. The cap domain of K2P channel is shown in orange. N and C represent each termini of the protein. The PIP2 cartoon represents a putative binding site based on experimental and/or structural data.
Voltage-gated K+ (Kv) channels gate in response to changes in the potential across the membrane. In addition to the canonical pore domain, the topology of a Kv channel contains four further TM helices, known as the voltage-sensor domain (VSD)14 (Figure 1(A),(B)). The fourth TM helix (S4) contains multiple basic residues that respond to a change in membrane potential, moving towards the extracellular side upon membrane depolarization.15, 16 This motion of the S4 helix is coupled to the channel pore through an amphipathic helix17; as the S4 helix translates, the linker is pulled away from the inner helices of the pore, therefore permitting channel opening.18 Conversely, hyperpolarization-activated cyclic nucleotide-gated (HCN) channels, which share a similar TM domain topology, open at hyperpolarized voltages, with the inward motion of the S4 helix resulting in channel opening.19, 20
K2P channels are denoted by the fact that each monomer has two pore domain sequences (Figure 1(C),(D)). The regulation of K2P channels differs between their members, and can involve responses to changes in pH (TASK21) and/or membrane pressure (TREK1, TREK2 and TRAAK).22, 23 The diversity and mechanisms of these proteins is discussed in more detail as part of this series of ion channel reviews.4
Calcium regulated K+ (KCa) channels are expressed ubiquitously throughout the human body and their single-channel conductance is the highest of all K+ channels.24 KCa channel opening is induced by an increase in intracellular concentration of Ca2+; with the divalent ion binding to the C-terminal domain of the K+ channel, as the intracellular ionic concentration rises from the nM to μM range. These channels also have a VSD, akin to Kv channel, and are therefore also voltage sensitive (Figure 1(E),(F)).25 This makes the channels activatable by both Ca2+-binding and membrane depolarization. In addition to the tetrameric pore, KCa can have auxiliary subunits which are tissue specific and can modify the Ca2+ sensitivity of the channel as well as gating properties.26
Inwardly rectifying K+ (Kir) channels conduct current at a negative potential, and cease to do so as the membrane become more depolarised, being inhibited by direct channel block by polyamines and divalent cations.27, 28 There are several publications that review the structures, regulation and function of Kir channels.29, 30 Kir channels generally have a short, 50–90 amino acid, N-terminal domain, and a larger C-terminal domain comprised of up to ~ 250 residues (Figure 1(G),(H)). These channels associate with other proteins to form larger complexes and are regulated by several factors.31 For example, Kir3.x channels are activated by the binding of Gβ and Gγ subunits from the Gi/o proteins.32 On the other hand, Kir6.2 associates with 4 regulatory sulfonylurea receptor (SUR) subunits33, 34, 35 to form the ATP-sensitive K (KATP) channel, and is inhibited by the binding of ATP to the pore subunits of the channel.36
Identifying a functional role for PIP2 binding to Kir channels
Many K+ channels are functionally modulated by specific interactions with lipids from the surrounding membrane. Lipids, in particular PIP2, were first identified as channel activators of the ATP-sensitive K+ (KATP) channel.1, 37, 38 Here, an increase in levels of PIP2 – by either activation of phospholipase C (PLC) or direct application of PIP2 – was shown to reduce channel inhibition by ATP.1, 37, 38 This could either be due direct binding site competition between ATP and PIP2 or allosteric modulation of the ligand-induced opening and closing events.39, 40
Application of PIP2 to the Kir1, Kir2, Kir3 and Kir6 channels has been shown to increase the size of the current of an excised patch and also elevates their open probability (Popen).41, 42, 43 To evaluate the response of the channel to PIP2 and to characterise the residues involved in coordinating PIP2, the size of the current after patch excision has been used as a readout.45 Mutations which decrease PIP2 binding to the Kir6.2 channel also reduce channel burst time duration, and result in a faster run-down; where the current flowing through the channel in an excised patch declines over time.46 The molecular mechanism of the run-down has been proposed to be associated with PIP2 disassociating from its binding site in the excised patch, or a depletion in the local concentration of PIP2, which ultimately causes the channel to shut.47 One method to reduce the level of run-down is to apply a fluoride vanadate and pyrophosphate (FVPP) solution, which inhibits the degradation of PIP2 by phosphatases.48 Together, this increases the local PIP2 concentration in the patch and enhances PIP2-channel interactions in the excised patch.49 A second method used to qualitatively assess channel activation by PIP2 is to introduce non-native voltage sensitive phosphatases (VSPs) to the cell of interest.50 This has been shown to be more effective than activation of PLC by Ca2+ due to a greater selectivity and specificity to the PIP2 in the membrane. These studies suggest that PIP2 is involved in stabilising the channel in its open conformation, as well as initially promoting channel opening. Interestingly, PIP2 binding to the Kir channel homologue (KirBac1.1) closes the channel, despite the absence of PIP2 in bacterial membranes.44
Regulation of other K+ channels by PIP2
Voltage gated K+ (Kv) channels, including KCNQ1 (Kv7.1),5, 51 hERG (Kv11.1),52 Shaker (Kv1.2)53 and Kv2.2,54 are primarily gated by voltage but can also be modulated by PIP2. Similar to Kir channels, application of PIP2 to the KCNQ1 and hERG channels slows channel run-down.58, 59 For KCNQ1, three binding sites have been proposed; one on the VSD, one on the linker between the VSD and pore domain, and one directly adjacent to the pore.55, 56, 57 The multiple binding sites therefore means that modulation of KCNQ1 by PIP2 is complex.
PIP2 is absolutely necessary to bridge the intramolecular interactions between the S4 helix of the VSD domain,60 this therefore enhances the VSD coupling to the pore.53, 61 PIP2 also appears to induce a large scale conformational change of the intracellular domain of the channel, which is also important for channel gating.56 In addition, PIP2 also has a dynamic interplay with other KCNQ1 regulator such as calmodulin (CaM).62
Application of PIP2 directly to K2P channel family members, including TREK-1, TASK1, TASK3 and TRAAK channels, increases the size of the current influx and modulates voltage dependency after tension is applied to the bilayer.63, 64 The binding of the PIP2 headgroup to TREK1 exhibits bimodal activity, with channel activation observed at low concentrations and inhibition at higher concentrations.65
KCa channels are also sensitive to PIP2, with structural studies revealing a KDRDD-loop to be essential for the interaction.66 While PIP2 may activate KCa channels alone, the primary effect of PIP2 binding has been shown to be enhancement of Ca2+ induced gating of the channel.67 PIP2-KCa channel interactions also depend on lipid tail length, subunit stoichiometry and tissue type.67, 68
Roles for other anionic phospholipids and cholesterol in modulating ion channel activity
Anionic phospholipid such as phosphatidylglycerol (PG) have been shown to provide a secondary mode of regulation for ion channels such as Kir2.1 channel.69, 70, 71 This regulation includes tuning the channel activation by PIP2, e.g. in Kir2.2 channels where PIP2 activation is enhanced by approximately 100-fold when PG, PS, PA, and PI are also present.69, 71 Based on the negative electrostatic charge of the headgroup, it could be proposed that the anionic lipids, e.g. PS and PG, compete with PIP2 binding for the primary binding site.72 For the Kir2.2 channel, PS has been shown to have a second binding site, adjacent to the primary PIP2 binding site. The affinity of PS for Kir2.2 increases when PIP2 is bound.73 Other inositol lipids, including PI4P and PI, have also been shown to activate Kir channels, but to a lesser extent.74, 75 Conversely, for PIP3, the additional phosphate group appears to reduce binding to the TREK1 channel.72 Therefore, channel activation by PIP lipids is not driven solely by charge.
Modulation of K+ channel activity by anionic phospholipids has been demonstrated directly through the application of the lipids to a patch,76 or by enzymatic activation such as phospholipase D2 (PLD2) and therefore generating a local intracellular anionic phospholipids.77 An example of a direct activator includes intracellular lysophosphatidic acid, which has been previously shown to directly activate TREK1, TREK2 and TRAAK channels.76 Phosphatidic acid (PA) has also been shown to activate a chimeric Kv channel (voltage sensor domain of Kv2.1 fused with a Kv1.2 pore domain).78 On the other hand, indirect activation mediated through PLD2 produces multiple products such as phosphatidic acid (PA), phosphatidylglycerol (PG) and phosphatidylethanol (PEtOH). These lipids have been previously suggested to activate multiple K+ channels such as Kir, TREK1 and TRAAK.79, 80, 81
In addition to PIP and other anionic lipids, cholesterol has also been shown to modulate a subset of K+ channels. Several reviews have been dedicated to the regulation of K+ channels by cholesterol.82, 83, 84 Cholesterol has been shown to inhibit Kir1.1, Kir2.2 and Kir6.2 channels by binding to the CD loop between the transmembrane domain and the cytoplasmic domain84 and closing the channel.85 Curiously, closure of Kir channels by cholesterol is independent of channel activation by PIP2.85 Overall, this suggests that cholesterol stabilises the closed state of the channel as it binds to the CD loop, rather than triggering a competitive gating mechanism, or competing for the PIP2 binding site.
Crystal structures highlight lipid binding sites and conformational change
In 2011, X-ray structures of Kir2.2 and GIRK (Kir3.2) channels revealed well-defined densities for PIP2.86, 87 These structures highlight the key amino acid residues coordinating the bound PIP2 lipid, as well as showing how PIP2 induces conformational changes and drives opening of the channel gate. The PIP2 binding site on a Kir channel is predominantly coordinated by basic residues (lysine and arginine), which engage with the phosphate groups (Figure 2(A)). These residues are located on both N-terminal and C-terminal domains of the protein, with the binding site being formed between adjacent subunits. At the N-terminal end of TM1 of Kir2.2, residue W79 engages with the inositol sugar ring and lipid tails, while R78 and R80 form hydrogen bonds with the 1′ phosphate. In the C-terminal domain, K183, R186 and K188, of the C-linker, interact with the 5′ phosphate and K189 coordinates both 4′ and 5′ phosphates.87
Figure 2.
PIP2 binding site on cKir2.2 and hKCNQ1 channel. (A) Atomic structure of the cKir2.2 channel in Apo (PDB ID: 3JYC)145 and PIP2-bound state (PDB ID: 3SPI).87 The pore domain is shown in blue and the cytoplasmic domain is shown in pink. (B) Atomic structure model of KCNQ1 channel in KCNE3-CaM bound (PDB ID: 6V00) and KCNE3-CaM-PIP2-bound state (PDB ID: 6V01).56 The pore domain is shown in blue, voltage sensor domain is coloured in green, S6 helices in the cytoplasm are shown in yellow, calmodulin proteins are shown in pink and KCNE3 subunits are shown in orange. In both panels, carbon atoms on PIP2 are represented in purple and others are coloured based on CPK convention. All PIP2 bound residues are shown as sticks.
Conformational changes of the Kir channels were proposed by comparing the bound and apo states of the structures. The binding of PIP2 to a Kir channel introduces a 6 Å upward conformational change and rotation of the C-terminal domain (Figure 2(B)). Similar structural changes have also been observed in the bacterial Kir channel, KirBac.88 This twists the TM helices of the pore and widens the distance at the I177 constriction point from 4.9 Å to 6.3 Å. However, this alone is insufficient to hydrate the lower gate of the pore, in part due to its hydrophobicity.
Capturing ion channels in different states has been a major technical challenge. Despite the wealth of information gleaned from X-ray crystallography, structures are generally captured at stable energy minima and thus transient intermediates are difficult to trap. This has been a major a bottleneck for studying the subtleties of lipid-dependent conformational change.
The Cryo-EM revolution and protein-lipid interactions
A major breakthrough for protein-lipid interaction studies has been the improvement to cryo-electron microscopy (cryo-EM) over recent years. This has yielded numerous atomic and near-atomic resolution structures of membrane proteins solubilised in detergents and nanodiscs. As a result, both novel structures and states have been observed, with many cryo-EM structures of membrane proteins revealing lipids, including phosphatidylglycerol (PG), cholesterol, PIP2 and phosphatidylinositol (PI), bound to the structures.10
Recent cryo-EM structures of the KCNQ1 and Kir3.2 channels have captured binding sites for PIP2. The quality of the electron microscopy density map allows precise identification of interactions between the lipid phosphate groups and the basic side chains. For the structures of the KCNQ1 channel, the PIP2 binding site is shown to be located on the VSD, adjacent to the S2-S3 linker (Figure 2(B)).56 In comparison to the apo-state structure, the PIP2 binding appears to induce a large conformational change in the S6 helix, as the helix bends at the RQKH motif. This causes the S6 helix and C-terminus of the protein to splay outwards and widens the cavity of the pore (Figure 2(B)).
Multiple cryo-EM structures of the Kir3.2 channel illustrate how PIP2 mediates an interaction between Kir3.2 and the α-subunit of a G-protein.89 Using cryo-EM, the study titrated in increasing concentrations of PIP2 to trap different states of Kir3.2. Similar to the chicken Kir2.2 structure, an upward motion of the C-terminal domain by approximately 6 Å is observed when PIP2 is bound to the channel. This conformational change shifts the position of the βL-βM loop upwards, allowing Gβγ to bind to the channel and triggers an opening event. In contrast to other models of regulation, the binding of Gβγ to the channel may only happen after the PIP2 binding event.89
Membrane embedding ion channels and identifying lipid binding sites with MD simulations.
Molecular dynamics (MD) simulations enable the unbiased prediction of lipid binding sites to ion channels within a membrane environment.90 Due to limits in simulation time and slow diffusion of lipids, coarse-grained (CG) MD simulations are regularly applied to study lipid binding. These simulations usually commence with lipids unbiasedly placed around an ion channel, either as a preformed bilayer or in random orientations to allow membrane self-assembly. As part of the simulation the lipids and protein are allowed to freely diffuse, with specific lipid-protein interactions recorded as the simulation proceeds.73, 91 From these simulations the nature of the binding site can be quantified by several factors, such as residue-lipid contacts, the relative residence time for the lipid within the binding site and membrane deformation around the protein.73 These methods have been shown to be have both excellent agreement with structurally determined sites73 and also for proposing uncharacterised binding sites.91 Therefore, these methods have the potential to identify bounds lipids based on unassigned electron density.92 In the future, this could be integrated into structural processing software to aid identification of a lipid at its site.
In addition to the PIP2 binding sites observed in the cryo-EM structure of the KCNQ1 channel, electrophysiology studies have proposed that PIP2 binds to multiple intracellular loops, the VSD and the pore.55, 93, 94, 95, 96 MD simulations with PIP2 bound at these sites have shown that the lipid bridges the VSD to the pore, and thereby enhances their coupling.60 MD simulations may therefore also be used to highlight potential gating mechanisms of an ion channel, induced by bound lipids. Understanding the residues involved in the conformational changes aids characterisation of the clinical phenotypes at the molecular level. This could potentially be used to probe clinical symptoms and outline potential therapies.
Computational tools to investigate the strength of protein-lipid interactions
With improvements to both computational hardware and the software for molecular simulation, longer timescale and/or larger simulations can be performed to identify lipid binding sites.73 Once a lipid-binding site is identified, the natural next step is to calculate binding free energy. However, all-atom computational binding free calculations can be highly costly due to convergence and sampling issues of lipids within a membrane environment. Therefore, CG approaches have been applied to overcome the limitations of, e.g., diffusion at an atomistic-level.97, 98 To date, a number of computational approaches have been used to identify lipid binding affinity, for a recent review on this topic see.98
One method is to use Potential of Mean Force (PMF) calculations to measure the free energy of lipid interactions from bound to unbound, along what is referred to as a collective variable (CV) (Figure 3(A)). By using umbrella sampling, lipids may be simulated at fixed, incremental distances from the protein binding site.99 This yields a calculated lipid binding free energy profile from bound to unbound, whilst also providing a total free energy difference between the bound and unbound states. The first attempt of PMF calculations to calculate protein-lipid affinity was for cardiolipin binding sites on the cytochrome c oxidase and complex IV.100, 101 Since this initial approach, the method has been applied to study lipid binding affinity to receptors, transporters and ion channels.92, 97, 99, 102, 103 The method has also shown to be useful when comparing differences in the affinities for a range of lipid types and provides details of their 1D energy landscape. There are other CV-based enhanced sampling methods that can also be used to calculate binding free energies across multiple dimensions, e.g. well-tempered metadynamics.99
Figure 3.
Computational methods associated with K+ Channel lipid interactions. (A) Potential of mean force (PMF) calculation. PIP2 lipids (acyl tail - green, glycerol - yellow, phosphates - orange, inositol ring - cyan) are sampled with a harmonic potential along the collective variable (CV) parallel to the bilayer. Each PIP2 represents a sampling position. The protein is displayed only in coarse-grained backbone where the transmembrane region is shown in blue and cytoplasmic region is shown in pink. PIP2 binding free energy to the protein is then calculated using weighted histogram analysis method (WHAM) based on the energies it samples along the CV (right). (B) The thermodynamic scheme underpinning free energy perturbation (FEP) calculation. Transformation of PIP2 to PC in both bound and free configurations are calculated in ΔG1 and ΔG2 (blue). This allows ΔG3 and ΔG4 to be obtained (green) using the equation shown (bottom).
A second approach for quantifying lipid binding affinity is free energy perturbation (FEP). FEP is often referred to as “alchemical transformation”, where molecules or subgroups of molecules are computationally mutated from one to another. From this transformation a relative binding free energy between the bound and free states can calculated104 (Figure 3(B)). Recent development of CG-FEP approaches illustrates strong agreement with the calculations using PMF approaches, but with a reduced computational cost.99, 105 The method can also be applied to amino acids to investigate the effects of site-directed mutagenesis on the binding free energy105, with the potential to develop this into a high-throughput screen for e.g. ion channel mutations that cause disease, which would be tightly coupled to experimentally determined measurements.
Experimental affinity, gating and efficacy of lipids
There are several approaches that have been used to quantify lipid binding affinities and the extent of channels activation by lipids both in vitro and in vivo.2, 106 In this section, we will discuss several experimental methods which could be used to measure the extent of channel activation/binding by PIP2.
Ion flux across the ion channels may be measured in reconstituted systems or in intact cells. Full details for configuring reconstituted system have been reviewed previously.107, 108, 109 Ion flux assays may be performed using K+ channel permeant ions such as Rb+ or Tl+ and/or fluorescence dyes.109 For example, K+ channels may be reconstituted into liposomes containing a Tl+-sensitive fluorescence dye and K+ ions. When the liposome is placed in an extracellular Tl+ solution, the Tl+ enters the liposomes through the K+ channels, and K+ diffuses out; with both ions permeating down their concentration gradients. The increased internal concentration of Tl+ quenches the fluorescence and thus provides a signal for K+ efflux.110 A second example is to establish a proton gradient, by applying a protonophore, such as Carbonyl cyanide m-chlorophenyl hydrazone (CCCP). In this instance the movement of protons into the liposome drives the efflux of K+ ions. The increase in proton flux may be measure by the quenching of a proton-sensitive fluorescence dye (Figure 4(A)).72 In both cases ligand may be added to investigate the impact on ion channel activity.
Figure 4.
Experimental methods associated with K+ channel lipid interactions. (A) Reconstituted flux Assay. Schematic diagram showing the fluorescence properties of the liposome based on.72 The liposome contains K+ and fluorescence sensitive dye (green). Upon an addition of CCCP - uncoupler (grey circle) and the opening of the channel (blue), K+ ions (pink) move out of the liposome, counteracting with the movement of proton (red) into the liposome. The protons then quench the fluorescence signal (green to grey). The degree of quenching is depending upon the concentration of the channel activator. (B) PIP2 sensitivity. The FVPP labels its application to prevent rundown. The channel is then allowed to rundown almost to 0 through the wash-out of FVPP. PIP2 at different concentrations (low - dark blue, high - yellow) are applied to the patch to observe an increase in channel current. The y-axis is labelled as the fraction of the maximal current. Neomycin sensitivity. (C) Neomycin at different concentrations (low - dark blue, high - yellow) are applied to the patch to observe decrease in channel current as PIP2 is quenched from the membrane. (D) Fluorescence binding assay. Decrease in fluorescence emission in an in vitro binding assay caused by FRET as different lipids (shown in different colour) compete with a PIP2 analogue at the binding site. Relative change in fluorescence recorded by the assay. (E) Native mass spectrometry. The traces are in different colours to show the influence of PIP2 concentration (low - dark blue, high - yellow), when applied to the protein in a native membrane or as a purified protein. Relative peak height can be used to calculated PIP2-bound saturation curve. (F) Super-resolution microscopy. Two states of the cells (A and B) display different sizes of lipid clustering observed under fluorescence microscope. Size of the cluster can be calculated to infer degrees of protein-lipid interaction. (G) Patch-clamp fluorometry. Fluorescence emission detected at specific wavelength. Different TNP-ATP concentrations (low - dark blue, high - yellow) are applied to a patch, resulting in a shift in the fluorescence emission peak caused by FRET. Relative fluorescence measured from the panel on the left are used to calculate TNP-ATP saturation curve.
Single-channel patch clamp electrophysiology remains a principal approach for studying ionic flux (Figure 4(B),(C)). Ligands, such as PIP2, may be applied to an excised patch, with their influence on the single-channel properties recorded. However, application of lipids directly to the channel faces solubility and concentration issues. One of the more straightforward means to overcome the solubility issues of PIP2 is to use diC8-PIP2, which has shorter 8-carbons acyl tails, rather than the native 1-stearoyl-2-arachidonyl acyl chains. Di-C8-PIP2 has been used to calculate the EC50 value PIP2 binding to Kir3.2.106, 111 This short-tail lipid was also crystallised with the Kir2.2 and Kir3.2 channels detailed earlier in this review.86, 87
An alternative approach is to add molecules that compete for the phospholipids in the membrane. Previous studies have applied ATP as a PIP2 competitor to the channel, and therefore used this as a metric to measure binding.1, 37, 38, 112 Alternatively, the cationic neomycin may be applied to directly bind to PIP2 and therefore inhibit lipid-protein interactions through PIP2 sequestration.113 This thereby reduces the free concentration of PIP2 available to bind and activate an ion channel (Figure 4(C)).114, 115 This method provides an advantage over di-C8 in terms of the relative solubility of neomycin, and has been used as a tool to screen for the PIP2 binding site on Kir6.2 channels.42 Later work has used neomycin to remove the background signal from PIP2 activation on TASK2 K2P channels in single channel recordings,116 or just simply to deplete the free PIP2 within the bilayer.117
Protein-lipid interactions have also been studied in vitro with spectroscopy methods, using a fluorescence PIP2 lipid analogue. Here, the fluorescent PIP2 analogue was used to measure competition at the lipid binding site. As the labelled PIP2 binds to a protein fused to GFP, FRET occurs between the lipid and protein fluorophores. By titrating in other lipids, to induce competition, the affinity of the lipid for the channel can be measured relative to the labelled PIP272 (Figure 4(D)). Direct measurement at the lipid binding site is also currently possible using isotopic labelling of residues. This allows binding kinetics of weaker binders (mM range) to be quantified using NMR spectroscopy.118
Native mass spectrometry (nMS) has emerged as a technique to characterise the interactions of membrane proteins and lipids.119, 120 By recording ligand-binding curves of lipids titrated into a purified membrane protein sample, binding affinities can be assessed, as well as quantification of the number of putative lipid binding sites121, 122, 123 (Figure 4(E)). In addition, nMS experiments are employed to investigate the effect of lipids on protein oligomeric state and their stability.121, 122 Early nMS work on ion channel lipid interactions focused on the effect of lipids (and other effects) on the stoichiometry of the mechanosensitive channel of large conductance (MscL) in different bacteria.121 Furthermore, nMS has been applied to characterise the selectivity of lipid binding to the isoforms of the TRAAK channel80 and to Kir3.2 where PIP2 was shown to bind selectively and single mutations influence the affinities and specificity.124 A more comprehensive review on characterising protein-lipid interactions through nMS can be found in.120 Recent advances in sample preparation have enabled the recording of mass spectra of membrane vesicles, allowing interactions to be investigated in their potentially most native state.125, 126
The use of fluorescence has extended to investigate protein-lipid interaction and clustering under super-resolution fluorescence microscopy such as direct stochastic optical reconstruction microscopy (dSTORM).127 By using a lipid specific dye, the size and the intensity of the lipid specific fluorescence can be quantified to infer PA clustering (Figure 4(F))128. Through the labelling of specific proteins and lipids, it is possible to quantify colocalization of proteins and lipid, allowing the extent of their interactions to be investigated.129, 130
To couple channel activation to the binding process, patch-clamp fluorometry has been introduced as a method to record both current (channel activation/inhibition) simultaneously with the binding kinetics (Figure 4(G)). An extensive review on the method is noted in.131 Initially, the method was implemented to investigate K+ channel conformational change, by measuring the change in the fluorescence signal of the labelled protein.132, 133 The method is then applied to investigate protein–ligand interactions where both protein and ligand are labelled. By modifying one of the amino acids near the ligand binding site to 6-propionyl-2-(N,N-dimethyl)aminonaphthalene (ANAP), this allows ATP analogue such as trinitrophenol-ATP (TNP-ATP) to quench to fluorescence signal on the ANAP.134, 135 As the binding of PIP2 influences the binding kinetics of TNP-ATP, this method indirectly assesses the affinity of PIP2 to its binding site.40 This simultaneously yields the ligand binding constant and the kinetics of ion channel gating.
Lipid binding and disease
Amino acid mutations within a lipid binding site have the potential to cause physiological defects to an ion channel and lead to clinical symptoms.6, 8 Mutations near the PIP2 binding site of Kir6.2 are associated with problems in insulin secretion.8 In the pancreatic β-cell, insulin secretion is triggered when the KATP channel is closed.6 Thus, mutations which alter channel opening cause defects in insulin secretion, for example mutations which enhanced PIP2 binding (such as K39R, E179A, E179K) are associated with neonatal diabetes (ND), whereas mutations which destabilise PIP2 binding (such as K67N) to the channel results in congenital hyperinsulinism (CHI) – an opposite phenotype.8, 42, 136, 137 Mutations to the PIP2 binding site of Kir2.1 (such as R218Q/W) are shown to be associated with Andersen-Tawil syndrome (ATS) which displays cardia arrythmia, periodic paralysis and dysmorphic features. Similar to Kir2.1, the Kir1.1 mutation at R312Q/W displays Bartter syndrome which results in polyhydramnios, premature delivery, hypokalemic alkalosis, and hypercalciuria.41
Mutations of the KCNQ channel are associated with QT syndrome, causing a longer QT hyperpolarised wave in the ECG signal.7, 138 Some of these mutations, e.g. R555H, are shown to be associated with PIP2 binding site. By understanding the consequences of PIP2 binding on the conformational changes of the protein, one can better appreciate how mutations are linked to clinical symptoms. In some cases this may also aid in the development of novel therapeutics, with a number of small molecule drugs known to bind to ion channel lipid binding sites or mediate allosteric coupling between PIP2 and the pore, e.g. CP1 which binds to the S4-S5 linker of the KCNQ1,57 or retigabine which binds to the pore domain and promotes channel opening through PIP2 mediated interactions.139
Future perspectives and conclusions
As further lipid-bound ion channel structures become available, from X-ray crystallography and the multiple conformational states captured in the class-averages from cryo-EM, it is important to understand how lipids induce conformational change and to then relate this to function. A direct linear motion between two structural states of a protein is unlikely to reflect the true nature of its conformational transition, which might bypass certain short-lived intermediate states. Therefore, molecular simulations on a longer timescale and/or enhanced (accelerated) sampling may be used to allow transition events to be mapped between these states. In addition to striving for longer molecular simulation timescales, it is also important to be able to simulate larger assemblies to reach the dimensions of the current resolution limits of super-resolution microscopy. Of note, recent studies with super-resolution microscopy have revealed that clustering of lipids in the bilayer is driven by membrane protein localization,140 with comparable clustering observed using MD simulations.141, 142
A further bottleneck in studying protein-lipid interaction is the determination of their binding affinity. MD simulations provide relative affinity between protein and lipids based on the structure available, but the challenge is to compare those to experimental data. Novel approaches in mass spectrometry and the use of fluorescence-based analogues provide values for comparison with the computationally calculated binding free energies. Comparative studies between methods would highlight the significance of lipid binding, modulation of ion channel activity and separation between lipid binding and signal transduction.
One of the first papers directly relating the binding of PIP2 to the activity of K+ ion channels was published in 1996.1 25 years later, we now have atomic level details of binding sites of PIP2, how PIP2 induces conformational change, the impact on single-channel kinetics and details of its potential roles in disease. Together this provides an interdisciplinary framework for understanding lipid modulation of K+ channels, which may be extended to membrane proteins in general.
CRediT authorship contribution statement
Tanadet Pipatpolkai: Conceptualization, Investigation, Methodology, Visualization, Writing - original draft, review & editing. Daniel Quetschlich: Conceptualization, Investigation, Methodology, Visualization, Writing - original draft, review & editing. Phillip J. Stansfeld: Conceptualization, Investigation, Methodology, Visualization, Supervision, Funding Acquistion, Writing - original draft, review & editing.
Edited by Lucie Delemotte
References
- 1.D.W. Hilgemann, R. Ball, Regulation of Cardiac Na+/Ca2+ Exchange and KATP Potassium Channels by PIP2 Science (80). 273 (1996) 956 LP – 959. https://doi.org/10.1126/science.273.5277.956. [DOI] [PubMed]
- 2.Saliba A.-E., Vonkova I., Gavin A.-C. The systematic analysis of protein–lipid interactions comes of age. Nat. Rev. Mol. Cell Biol. 2015;16:753–761. doi: 10.1038/nrm4080. [DOI] [PubMed] [Google Scholar]
- 3.Gu R.-X., de Groot B.L. Lipid-protein interactions modulate the conformational equilibrium of a potassium channel. Nat. Commun. 2020;11:2162. doi: 10.1038/s41467-020-15741-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.A.M. Natale, P.E. Deal, D.L. Minor Jr., Structural insights into the mechanisms and pharmacology of K2P potassium channels, J. Mol. Biol. (2021) 166995. https://doi.org/https://doi.org/10.1016/j.jmb.2021.166995. [DOI] [PMC free article] [PubMed]
- 5.Zaydman M.A., Cui J. PIP2 regulation of KCNQ channels: biophysical and molecular mechanisms for lipid modulation of voltage-dependent gating. Front. Physiol. 2014;5:195. doi: 10.3389/fphys.2014.00195. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Ashcroft F.M. From molecule to malady. Nature. 2006;440:440–447. doi: 10.1038/nature04707. [DOI] [PubMed] [Google Scholar]
- 7.Wu J., Ding W.-G., Horie M. Molecular pathogenesis of long QT syndrome type 1. J. Arrhythmia. 2016;32:381–388. doi: 10.1016/j.joa.2015.12.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Pipatpolkai T., Usher S., Stansfeld P.J., Ashcroft F.M. New insights into KATP channel gene mutations and neonatal diabetes mellitus. Nat. Rev. Endocrinol. 2020 doi: 10.1038/s41574-020-0351-y. [DOI] [PubMed] [Google Scholar]
- 9.C. Verkest, S. Häfner, P. Ávalos Prado, A. Baron, G. Sandoz, Migraine and Two-Pore-Domain Potassium Channels, Neurosci. (2020) 1073858420940949. https://doi.org/10.1177/1073858420940949. [DOI] [PubMed]
- 10.Duncan A.L., Song W., Sansom M.S.P. Lipid-dependent regulation of ion channels and G protein-coupled receptors: insights from structures and simulations. Annu. Rev. Pharmacol. Toxicol. 2019 doi: 10.1146/annurev-pharmtox-010919-023411. [DOI] [PubMed] [Google Scholar]
- 11.Thompson M.J., Baenziger J.E. Ion channels as lipid sensors: from structures to mechanisms. Nat. Chem. Biol. 2020;16:1331–1342. doi: 10.1038/s41589-020-00693-3. [DOI] [PubMed] [Google Scholar]
- 12.D.A. Doyle, J.M. Cabral, R.A. Pfuetzner, A. Kuo, J.M. Gulbis, S.L. Cohen, B.T. Chait, R. MacKinnon, The Structure of the Potassium Channel: Molecular Basis of K+ Conduction and Selectivity, Science (80), 280 (1998) 69 LP – 77. [DOI] [PubMed]
- 13.Anderson P.A., Greenberg R.M. Phylogeny of ion channels: clues to structure and function. Comp. Biochem. Physiol. B. Biochem. Mol. Biol. 2001;129:17–28. doi: 10.1016/s1096-4959(01)00376-1. [DOI] [PubMed] [Google Scholar]
- 14.Shih T.M., Goldin A.L. Topology of the Shaker potassium channel probed with hydrophilic epitope insertions. J. Cell Biol. 1997;136:1037–1045. doi: 10.1083/jcb.136.5.1037. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.S.K. Aggarwal, R. MacKinnon, Contribution of the S4 Segment to Gating Charge in the Shaker K+ Channel, Neuron 16 (1996) 1169–1177. https://doi.org/https://doi.org/10.1016/S0896-6273(00)80143-9. [DOI] [PubMed]
- 16.S.-A. Seoh, D. Sigg, D.M. Papazian, F. Bezanilla, Voltage-sensing residues in the S2 and S4 segments of the shaker K+ channel, Neuron 16 (1996) 1159–1167. https://doi.org/https://doi.org/10.1016/S0896-6273(00)80142-7. [DOI] [PubMed]
- 17.Mannuzzu L.M., Moronne M.M., Isacoff E.Y. Direct physical measure of conformational rearrangement underlying potassium channel gating. Science. 1996;271:213–216. doi: 10.1126/science.271.5246.213. [DOI] [PubMed] [Google Scholar]
- 18.Tao X., MacKinnon R. Cryo-EM structure of the KvAP channel reveals a non-domain-swapped voltage sensor topology. Elife. 2019;8 doi: 10.7554/eLife.52164. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Lee C.-H., MacKinnon R. Structures of the human HCN1 hyperpolarization-activated channel. Cell. 2017;168:111–120.e11. doi: 10.1016/j.cell.2016.12.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.C.-H. Lee, R. MacKinnon, Voltage sensor movements during hyperpolarization in the HCN channel, Cell 179 (2019) 1582-1589.e7. https://doi.org/https://doi.org/10.1016/j.cell.2019.11.006. [DOI] [PMC free article] [PubMed]
- 21.Duprat F., Lesage F., Fink M., Reyes R., Heurteaux C., Lazdunski M. TASK, a human background K+ channel to sense external pH variations near physiological pH. EMBO J. 1997;16:5464–5471. doi: 10.1093/emboj/16.17.5464. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.S.G. Brohawn, Z. Su, R. MacKinnon, Mechanosensitivity is mediated directly by the lipid membrane in TRAAK and TREK1 K+ channels, Proc. Natl. Acad. Sci. 111 (2014) 3614 LP – 3619. https://doi.org/10.1073/pnas.1320768111. [DOI] [PMC free article] [PubMed]
- 23.Aryal P., Jarerattanachat V., Clausen M.V., Schewe M., McClenaghan C., Argent L., Conrad L.J., Dong Y.Y., Pike A.C.W., Carpenter E.P., Baukrowitz T., Sansom M.S.P., Tucker S.J. Bilayer-mediated structural transitions control mechanosensitivity of the TREK-2 K2P channel. Structure. 2017;25:708–718.e2. doi: 10.1016/j.str.2017.03.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Schrecker M., Wunnicke D., Hänelt I. How RCK domains regulate gating of K+ channels. Biol. Chem. 2019;400:1303–1322. doi: 10.1515/hsz-2019-0153. [DOI] [PubMed] [Google Scholar]
- 25.Yang H., Zhang G., Cui J. BK channels: multiple sensors, one activation gate. Front. Physiol. 2015;6:29. doi: 10.3389/fphys.2015.00029. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Contreras G.F., Castillo K., Enrique N., Carrasquel-Ursulaez W., Castillo J.P., Milesi V., Neely A., Alvarez O., Ferreira G., González C., Latorre R. A BK (Slo1) channel journey from molecule to physiology. Channels. 2013;7:442–458. doi: 10.4161/chan.26242. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Matsuda H., Saigusa A., Irisawa H. Ohmic conductance through the inwardly rectifying K channel and blocking by internal Mg2+ Nature. 1987;325:156–159. doi: 10.1038/325156a0. [DOI] [PubMed] [Google Scholar]
- 28.Lopatin A.N., Makhina E.N., Nichols C.G. Potassium channel block by cytoplasmic polyamines as the mechanism of intrinsic rectification. Nature. 1994;372:366–369. doi: 10.1038/372366a0. [DOI] [PubMed] [Google Scholar]
- 29.Hibino H., Inanobe A., Furutani K., Murakami S., Findlay I., Kurachi Y. Inwardly rectifying potassium channels: their structure, function, and physiological roles. Physiol. Rev. 2010;90:291–366. doi: 10.1152/physrev.00021.2009. [DOI] [PubMed] [Google Scholar]
- 30.Nichols C.G., Lopatin A.N. Inward rectifier potassium channels. Annu. Rev. Physiol. 1997;59:171–191. doi: 10.1146/annurev.physiol.59.1.171. [DOI] [PubMed] [Google Scholar]
- 31.Sakura H., Wat N., Horton V., Millns H., Turner R.C., Ashcroft F.M. Sequence variations in the human Kir6.2 gene, a subunit of the beta-cell ATP-sensitive K-channel: no association with NIDDM in white caucasian subjects or evidence of abnormal function when expressed in vitro. Diabetologia. 1996;39:1233–1236. doi: 10.1007/BF02658512. [DOI] [PubMed] [Google Scholar]
- 32.Logothetis D.E., Kurachi Y., Galper J., Neer E.J., Clapham D.E. The beta gamma subunits of GTP-binding proteins activate the muscarinic K+ channel in heart. Nature. 1987;325:321–326. doi: 10.1038/325321a0. [DOI] [PubMed] [Google Scholar]
- 33.Sakura H., Ammälä C., Smith P.A., Gribble F.M., Ashcroft F.M. Cloning and functional expression of the cDNA encoding a novel ATP-sensitive potassium channel subunit expressed in pancreatic beta-cells, brain, heart and skeletal muscle. FEBS Lett. 1995 doi: 10.1016/0014-5793(95)01369-5. [DOI] [PubMed] [Google Scholar]
- 34.Inagaki N., Gonoi T., Clement J.P., IV, Namba N., Inazawa J., Gonzalez G., Aguilar-Bryan L., Seino S., Bryan J. Reconstitution of IKATP: An inward rectifier subunit plus the sulfonylurea receptor. Science (80-.) 1995 doi: 10.1126/science.270.5239.1166. [DOI] [PubMed] [Google Scholar]
- 35.L. Aguilar-Bryan, C.G. Nichols, S.W. Wechsler, J.P. Clement, A.E. Boyd, G. Gonzalez, H. Herrera-Sosa, K. Nguy, J. Bryan, D.A. Nelson, K. Kunjilwar, G. Gonzalez, M. Schwanstecher, U. Panten, L. Aguilar-Bryan, J. Bryan, S. Trapp, S. Haider, P. Jones, M.S.P. Sansom, F.M. Ashcroft, M.V. Mikhailov, J.D. Campbell, H. de Wet, K. Shimomura, B. Zadek, R.F. Collins, M.S.P. Sansom, R.C. Ford, F.M. Ashcroft, J. Aittoniemi, C. Fotinou, T.J. Craig, H. de Wet, P. Proks, F.M. Ashcroft, Cloning of the beta cell high-affinity sulfonylurea receptor: a regulator of insulin secretion, Neuron 18 (1995) 423 LP – 426. https://doi.org/https://doi.org/10.1016/S0896-6273(00)80321-9. [DOI] [PubMed]
- 36.Tucker S.J., Gribble F.M., Zhao C., Trapp S., Ashcroft F.M. Truncation of Kir6.2 produces ATP-sensitive K+ channels in the absence of the sulphonylurea receptor. Nature. 1997;387:179–183. doi: 10.1038/387179a0. [DOI] [PubMed] [Google Scholar]
- 37.S.L. Shyng, C.G. Nichols, Membrane phospholipid control of nucleotide sensitivity of KATP channels, Science (80). 282 (1998) 1138–1141. https://doi.org/10.1126/science.282.5391.1138. [DOI] [PubMed]
- 38.T. Baukrowitz, U. Schulte, D. Oliver, S. Herlitze, T. Krauter, S.J. Tucker, J.P. Ruppersberg, B. Fakler, PIP2 and PIP as determinants for ATP inhibition of KATP channels., Science (80). 282 (1998) 1141–1144. https://doi.org/10.1126/science.282.5391.1141. [DOI] [PubMed]
- 39.G.G. MacGregor, K. Dong, C.G. Vanoye, L. Tang, G. Giebisch, S.C. Hebert, Nucleotides and phospholipids compete for binding to the C terminus of KATP channels, Proc. Natl. Acad. Sci. 99 (2002) 2726 LP – 2731. https://doi.org/10.1073/pnas.042688899. [DOI] [PMC free article] [PubMed]
- 40.T. Pipatpolkai, S.G. Usher, N. Vedovato, F.M. Ashcroft, P.J. Stansfeld, The dynamic interplay of PIP2 and ATP in the regulation of the KATP channel, BioRxiv. (2021) 2021.05.06.442933. https://doi.org/10.1101/2021.05.06.442933. [DOI] [PMC free article] [PubMed]
- 41.C.M.B. Lopes, H. Zhang, T. Rohacs, T. Jin, J. Yang, D.E. Logothetis, alterations in conserved Kir channel-PIP2 interactions underlie channelopathies, Neuron 34 (2002) 933–944. https://doi.org/https://doi.org/10.1016/S0896-6273(02)00725-0. [DOI] [PubMed]
- 42.Haider S., Tarasov A.I., Craig T.J., Sansom M.S.P., Ashcroft F.M. Identification of the PIP2-binding site on Kir6.2 by molecular modelling and functional analysis. EMBO J. 2007;26:3749–3759. doi: 10.1038/sj.emboj.7601809. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Lacin E., Aryal P., Glaaser I.W., Bodhinathan K., Tsai E., Marsh N., Tucker S.J., Sansom M.S.P., Slesinger P.A. 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]
- 44.Enkvetchakul D., Jeliazkova I., Nichols C.G. Direct modulation of Kir channel gating by membrane phosphatidylinositol 4,5-bisphosphate. J. Biol. Chem. 2005;280:35785–35788. doi: 10.1074/jbc.C500355200. [DOI] [PubMed] [Google Scholar]
- 45.S.-L. Shyng, C.A. Cukras, J. Harwood, C.G. Nichols, Structural determinants of PIP2 regulation of inward rectifier KATP channels, J. Gen. Physiol. 116 (2000) 599 LP – 608. [DOI] [PMC free article] [PubMed]
- 46.Ribalet B., John S.A., Xie L.-H., Weiss J.N. ATP-sensitive K+ channels: regulation of bursting by the sulphonylurea receptor, PIP2 and regions of Kir6.2. J. Physiol. 2006;571:303–317. doi: 10.1113/jphysiol.2005.100719. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Suh B.-C., Hille B. PIP2 is a necessary cofactor for ion channel function: how and why? Annu. Rev. Biophys. 2008;37:175–195. doi: 10.1146/annurev.biophys.37.032807.125859. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Friedman Z.Y. Tamoxifen and vanadate synergize in causing accumulation of polyphosphoinositides in GH4C1 membranes. J. Pharmacol. Exp. Ther. 1993;267:617–623. [PubMed] [Google Scholar]
- 49.Huang C.-L., Feng S., Hilgemann D.W. 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]
- 50.Rjasanow A., Leitner M.G., Thallmair V., Halaszovich C.R., Oliver D. Ion channel regulation by phosphoinositides analyzed with VSPs-PI(4,5)P2 affinity, phosphoinositide selectivity, and PI(4,5)P2 pool accessibility. Front. Pharmacol. 2015;6:127. doi: 10.3389/fphar.2015.00127. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Zhang H., Craciun L.C., Mirshahi T., Rohács T., Lopes C.M.B., Jin T., Logothetis D.E. PIP(2) activates KCNQ channels, and its hydrolysis underlies receptor-mediated inhibition of M currents. Neuron. 2003;37:963–975. doi: 10.1016/s0896-6273(03)00125-9. [DOI] [PubMed] [Google Scholar]
- 52.Bian J.-S., Kagan A., McDonald T.V. Molecular analysis of PIP2 regulation of HERG and IKr. Am. J. Physiol. Heart Circ. Physiol. 2004;287:H2154–H2163. doi: 10.1152/ajpheart.00120.2004. [DOI] [PubMed] [Google Scholar]
- 53.Abderemane-Ali F., Es-Salah-Lamoureux Z., Delemotte L., Kasimova M.A., Labro A.J., Snyders D.J., Fedida D., Tarek M., Baró I., Loussouarn G. Dual effect of phosphatidylinositol (4,5)-bisphosphate PIP(2) on Shaker K(+) corrected channels. J. Biol. Chem. 2012;287:36158–36167. doi: 10.1074/jbc.M112.382085. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.M. Delgado-Ramírez, J.J. De Jesús-Pérez, I.A. Aréchiga-Figueroa, J. Arreola, S.K. Adney, C.A. Villalba-Galea, D.E. Logothetis, A.A. Rodríguez-Menchaca, Regulation of Kv2.1 channel inactivation by phosphatidylinositol 4,5-bisphosphate, Sci. Rep. 8 (2018) 1769. https://doi.org/10.1038/s41598-018-20280-w. [DOI] [PMC free article] [PubMed]
- 55.M.A. Zaydman, J.R. Silva, K. Delaloye, Y. Li, H. Liang, H.P. Larsson, J. Shi, J. Cui, Kv7.1 ion channels require a lipid to couple voltage sensing to pore opening, Proc. Natl. Acad. Sci. 110 (2013) 13180 LP – 13185. https://doi.org/10.1073/pnas.1305167110. [DOI] [PMC free article] [PubMed]
- 56.J. Sun, R. MacKinnon, Structural Basis of Human KCNQ1 Modulation and Gating, Cell. 180 (2020) 340-347.e9. https://doi.org/https://doi.org/10.1016/j.cell.2019.12.003. [DOI] [PMC free article] [PubMed]
- 57.Liu Y., Xu X., Gao J., Naffaa M.M., Liang H., Shi J., Wang H.Z., Yang N.-D., Hou P., Zhao W., White K.M., Kong W., Dou A., Cui A., Zhang G., Cohen I.S., Zou X., Cui J. A PIP2 substitute mediates voltage sensor-pore coupling in KCNQ activation. Commun. Biol. 2020;3:385. doi: 10.1038/s42003-020-1104-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Loussouarn G., Park K.-H., Bellocq C., Baró I., Charpentier F., Escande D. Phosphatidylinositol-4,5-bisphosphate, PIP2, controls KCNQ1/KCNE1 voltage-gated potassium channels: a functional homology between voltage-gated and inward rectifier K+ channels. EMBO J. 2003;22:5412–5421. doi: 10.1093/emboj/cdg526. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.B. Jinsong, C. Jie, M.T. V., HERG K+ channel activity is regulated by changes in phosphatidyl inositol 4,5-bisphosphate, Circ. Res. 89 (2001) 1168–1176. https://doi.org/10.1161/hh2401.101375. [DOI] [PubMed]
- 60.Kasimova M.A., Zaydman M.A., 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]
- 61.Rodriguez-Menchaca A.A., Adney S.K., Tang Q.-Y., Meng X.-Y., Rosenhouse-Dantsker A., Cui M., Logothetis D.E. PIP2 controls voltage-sensor movement and pore opening of Kv channels through the S4–S5 linker. Proc. Natl. Acad. Sci. U. S. A. 2012;109:E2399–E2408. doi: 10.1073/pnas.1207901109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Kang P.W., Westerlund A.M., Shi J., White K.M.F., Dou A.K., Cui A.H., Silva J.R., Delemotte L., Cui J. Calmodulin acts as a state-dependent switch to control a cardiac potassium channel opening. Sci. Adv. 2020;6:eabd6798. doi: 10.1126/sciadv.abd6798. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Chemin J., Patel A.J., Duprat F., Lauritzen I., Lazdunski M., Honoré E. A phospholipid sensor controls mechanogating of the K+ channel TREK-1. EMBO J. 2005;24:44–53. doi: 10.1038/sj.emboj.7600494. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Lopes C.M.B., Rohács T., Czirják G., Balla T., Enyedi P., Logothetis D.E. PIP2 hydrolysis underlies agonist-induced inhibition and regulates voltage gating of two-pore domain K+ channels. J. Physiol. 2005;564:117–129. doi: 10.1113/jphysiol.2004.081935. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Chemin J., Patel A.J., Duprat F., Sachs F., Lazdunski M., Honore E. Up- and down-regulation of the mechano-gated K2P channel TREK-1 by PIP2 and other membrane phospholipids. Pflügers Arch. - Eur. J. Physiol. 2007;455:97–103. doi: 10.1007/s00424-007-0250-2. [DOI] [PubMed] [Google Scholar]
- 66.Tang Q.-Y., Zhang Z., Meng X.-Y., Cui M., Logothetis D.E. Structural determinants of phosphatidylinositol 4,5-bisphosphate (PIP2) regulation of BK channel activity through the RCK1 Ca2+ coordination site. J. Biol. Chem. 2014;289:18860–18872. doi: 10.1074/jbc.M113.538033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Vaithianathan T., Bukiya A., Liu J., Liu P., Asuncion-Chin M., Fan Z., Dopico A. Direct regulation of BK channels by phosphatidylinositol 4,5-bisphosphate as a novel signaling pathway. J. Gen. Physiol. 2008;132:13–28. doi: 10.1085/jgp.200709913. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Tian Y., Ullrich F., Xu R., Heinemann S.H., Hou S., Hoshi T. Two distinct effects of PIP2 underlie auxiliary subunit-dependent modulation of Slo1 BK channels. J. Gen. Physiol. 2015;145:331–343. doi: 10.1085/jgp.201511363. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Lee S.-J., Wang S., Borschel W., Heyman S., Gyore J., Nichols C.G. Secondary anionic phospholipid binding site and gating mechanism in Kir2.1 inward rectifier channels. Nat. Commun. 2013;4:2786. doi: 10.1038/ncomms3786. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Lee S.-J., Ren F., Zangerl-Plessl E.-M., Heyman S., Stary-Weinzinger A., Yuan P., Nichols C.G. 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]
- 71.Cheng W.W.L., D’Avanzo N., Doyle D.A., Nichols C.G. 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]
- 72.Cabanos C., Wang M., Han X., Hansen S.B. A soluble fluorescent binding assay reveals PIP2 antagonism of TREK-1 channels. Cell Rep. 2017;20:1287–1294. doi: 10.1016/j.celrep.2017.07.034. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.A.L. Duncan, R.A. Corey, M.S.P. Sansom, Defining how multiple lipid species interact with inward rectifier potassium (Kir2) channels, Proc. Natl. Acad. Sci. 117 (2020) 7803 LP – 7813. https://doi.org/10.1073/pnas.1918387117. [DOI] [PMC free article] [PubMed]
- 74.Fan Z., Makielski J.C. Anionic phospholipids activate ATP-sensitive potassium channels. J. Biol. Chem. 1997;272:5388–5395. doi: 10.1074/jbc.272.9.5388. [DOI] [PubMed] [Google Scholar]
- 75.T. Rohács, C.M.B. Lopes, T. Jin, P.P. Ramdya, Z. Molnár, D.E. Logothetis, Specificity of activation by phosphoinositides determines lipid regulation of Kir channels, Proc. Natl. Acad. Sci. 100 (2003) 745 LP – 750. https://doi.org/10.1073/pnas.0236364100. [DOI] [PMC free article] [PubMed]
- 76.Chemin J., Patel A., Duprat F., Zanzouri M., Lazdunski M., Honoré E. Lysophosphatidic acid-operated K+ channels. J. Biol. Chem. 2005;280:4415–4421. doi: 10.1074/jbc.M408246200. [DOI] [PubMed] [Google Scholar]
- 77.Comoglio Y., Levitz J., Kienzler M.A., Lesage F., Isacoff E.Y., Sandoz G. Phospholipase D2 specifically regulates TREK potassium channels via direct interaction and local production of phosphatidic acid. Proc. Natl. Acad. Sci. U. S. A. 2014;111:13547–13552. doi: 10.1073/pnas.1407160111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Hite R.K., Butterwick J.A., MacKinnon R. Phosphatidic acid modulation of Kv channel voltage sensor function. Elife. 2014;3 doi: 10.7554/eLife.04366. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Petersen E.N., Chung H.-W., Nayebosadri A., Hansen S.B. Kinetic disruption of lipid rafts is a mechanosensor for phospholipase D. Nat. Commun. 2016;7:13873. doi: 10.1038/ncomms13873. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Schrecke S., Zhu Y., McCabe J.W., Bartz M., Packianathan C., Zhao M., Zhou M., Russell D., Laganowsky A. Selective regulation of human TRAAK channels by biologically active phospholipids. Nat. Chem. Biol. 2021;17:89–95. doi: 10.1038/s41589-020-00659-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Chung H.-W., Petersen E.N., Cabanos C., Murphy K.R., Pavel M.A., Hansen A.S., Ja W.W., Hansen S.B. A molecular target for an alcohol chain-length cutoff. J. Mol. Biol. 2019;431:196–209. doi: 10.1016/j.jmb.2018.11.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.F. Zakany, T. Kovacs, G. Panyi, Z. Varga, Direct and indirect cholesterol effects on membrane proteins with special focus on potassium channels, Biochim. Biophys. Acta - Mol. Cell Biol. Lipids. 1865 (2020) 158706. https://doi.org/https://doi.org/10.1016/j.bbalip.2020.158706. [DOI] [PubMed]
- 83.Levitan I., Fang Y., Rosenhouse-Dantsker A., Romanenko V. Cholesterol and ion channels. Subcell. Biochem. 2010;51:509–549. doi: 10.1007/978-90-481-8622-8_19. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Levitan I. Cholesterol and Kir channels. IUBMB Life. 2009;61:781–790. doi: 10.1002/iub.192. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Y. Epshtein, A.P. Chopra, A. Rosenhouse-Dantsker, G.B. Kowalsky, D.E. Logothetis, I. Levitan, Identification of a C-terminus domain critical for the sensitivity of Kir2.1 to cholesterol, Proc. Natl. Acad. Sci. 106 (2009) 8055 LP – 8060. https://doi.org/10.1073/pnas.0809847106. [DOI] [PMC free article] [PubMed]
- 86.Whorton M.R., MacKinnon R. Crystal structure of the mammalian GIRK2 K+ channel and gating regulation by G proteins, PIP2, and sodium. Cell. 2011;477:199–208. doi: 10.1016/j.cell.2011.07.046. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Hansen S.B., 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]
- 88.Clarke O.B., Caputo A.T., Hill A.P., Vandenberg J.I., Smith B.J., Gulbis J.M. Domain reorientation and rotation of an intracellular assembly regulate conduction in kir potassium channels. Cell. 2010;141:1018–1029. doi: 10.1016/j.cell.2010.05.003. [DOI] [PubMed] [Google Scholar]
- 89.Niu Y., Tao X., Touhara K.K., MacKinnon R. Cryo-EM analysis of PIP2 regulation in mammalian GIRK channels. Elife. 2020;9 doi: 10.7554/eLife.60552. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Chavent M., Duncan A.L., Sansom M.S. 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]
- 91.Stansfeld P.J., Hopkinson R., Ashcroft F.M., Sansom M.S.P. 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]
- 92.Wang Q., Corey R.A., Hedger G., Aryal P., Grieben M., Nasrallah C., Baronina A., Pike A.C.W., Shi J., Carpenter E.P., Sansom M.S.P. Lipid interactions of a ciliary membrane TRP channel: simulation and structural studies of polycystin-2. Structure. 2020;28:169–184.e5. doi: 10.1016/j.str.2019.11.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Telezhkin V., Reilly J.M., Thomas A.M., Tinker A., Brown D.A. Structural requirements of membrane phospholipids for M-type potassium channel activation and binding. J. Biol. Chem. 2012;287:10001–10012. doi: 10.1074/jbc.M111.322552. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Thomas A.M., Harmer S.C., Khambra T., Tinker A. Characterization of a binding site for anionic phospholipids on KCNQ1. J. Biol. Chem. 2011;286:2088–2100. doi: 10.1074/jbc.M110.153551. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Kyu-Ho P., Julien P., Shehrazade D., Jean M., Isabelle B., Denis E., Gildas L. Impaired KCNQ1–KCNE1 and phosphatidylinositol-4,5-bisphosphate interaction underlies the long QT syndrome. Circ. Res. 2005;96:730–739. doi: 10.1161/01.RES.0000161451.04649.a8. [DOI] [PubMed] [Google Scholar]
- 96.Q. Zhang, P. Zhou, Z. Chen, M. Li, H. Jiang, Z. Gao, H. Yang, Dynamic PIP2 interactions with voltage sensor elements contribute to KCNQ2 channel gating, Proc. Natl. Acad. Sci. 110 (2013) 20093 LP – 20098. https://doi.org/10.1073/pnas.1312483110. [DOI] [PMC free article] [PubMed]
- 97.Domański J., Hedger G., Best R.B., Stansfeld P.J., Sansom M.S.P. 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]
- 98.Corey R.A., Stansfeld P.J., Sansom M.S.P. The energetics of protein–lipid interactions as viewed by molecular simulations. Biochem. Soc. Trans. 2019;48:25–37. doi: 10.1042/BST20190149. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Corey R.A., Vickery O.N., Sansom M.S.P., Stansfeld P.J. 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]
- 100.Arnarez C., Marrink S.J., Periole X. Identification of cardiolipin binding sites on cytochrome c oxidase at the entrance of proton channels. Sci. Rep. 2013;3:1263. doi: 10.1038/srep01263. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Arnarez C., Mazat J.-P., Elezgaray J., Marrink S.-J., Periole X. Evidence for cardiolipin binding sites on the membrane-exposed surface of the cytochrome Bc1. J. Am. Chem. Soc. 2013;135:3112. doi: 10.1021/ja310577u. [DOI] [PubMed] [Google Scholar]
- 102.Hedger G., Rouse S.L., Domański J., Chavent M., Koldsø H., Sansom M.S.P. Lipid-loving ANTs: molecular simulations of cardiolipin interactions and the organization of the adenine nucleotide translocase in model mitochondrial membranes. Biochemistry. 2016;55:6238. doi: 10.1021/acs.biochem.6b00751. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Hedger G., Shorthouse D., Koldsø H., Sansom M.S.P. 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. doi: 10.1021/acs.jpcb.6b01387. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Aldeghi M., Bluck J.P., Biggin P.C. Absolute alchemical free energy calculations for ligand binding: a beginner’s guide. Methods Mol. Biol. 2018;1762:199. doi: 10.1007/978-1-4939-7756-7_11. [DOI] [PubMed] [Google Scholar]
- 105.Pipatpolkai T., Corey R.A., Proks P., Ashcroft F.M., Stansfeld P.J. 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]
- 106.Rohács T., Lopes C., Mirshahi T., Jin T., Zhang H., Logothetis D.E. Assaying phosphatidylinositol bisphosphate regulation of potassium channels. Methods Enzymol. 2002;345:71–92. doi: 10.1016/s0076-6879(02)45008-2. [DOI] [PubMed] [Google Scholar]
- 107.C. Miller, Ion channel reconstitution why bother? BT - ionic channels in cells and model systems, in: R. Latorre (Ed.), Springer US, Boston, MA, 1986: pp. 257–271. https://doi.org/10.1007/978-1-4684-5077-4_17.
- 108.J. Gaburjakova, M. Gaburjakova, Chapter five - reconstitution of ion channels in planar lipid bilayers: new approaches, in: A. Iglič, M. Rappolt, A.J.B.T.-A. in B. and L.S.-A. García-Sáez (Eds.), Academic Press, 2018: pp. 147–185. https://doi.org/https://doi.org/10.1016/bs.abl.2017.12.005.
- 109.Yu H., Li M., Wang W., Wang X. High throughput screening technologies for ion channels. Acta Pharmacol. Sin. 2016;37:34–43. doi: 10.1038/aps.2015.108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Posson D.J., Rusinova R., Andersen O.S., Nimigean C.M. Stopped-flow fluorometric ion flux assay for ligand-gated ion channel studies. Methods Mol. Biol. 2018;1684:223–235. doi: 10.1007/978-1-4939-7362-0_17. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Rohács T., Chen J., Prestwich G.D., Logothetis D.E. Distinct specificities of inwardly rectifying K(+) channels for phosphoinositides. J. Biol. Chem. 1999;274:36065–36072. doi: 10.1074/jbc.274.51.36065. [DOI] [PubMed] [Google Scholar]
- 112.Fan Z., Makielski J.C. Phosphoinositides decrease ATP sensitivity of the cardiac ATP-sensitive K(+) channel. A molecular probe for the mechanism of ATP-sensitive inhibition. J. Gen. Physiol. 1999;114:251–269. doi: 10.1085/jgp.114.2.251. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.Gabev E., Kasianowicz J., Abbott T., McLaughlin S. Binding of neomycin to phosphatidylinositol 4,5-bisphosphate (PIP2) Biochim. Biophys. Acta. 1989;979:105–112. doi: 10.1016/0005-2736(89)90529-4. [DOI] [PubMed] [Google Scholar]
- 114.Krauter T., Ruppersberg J.P., Baukrowitz T. Phospholipids as modulators of K(ATP) channels: distinct mechanisms for control of sensitivity to sulphonylureas, K(+) channel openers, and ATP. Mol. Pharmacol. 2001;59:1086–1093. [PubMed] [Google Scholar]
- 115.D. Schulze, T. Krauter, H. Fritzenschaft, M. Soom, T. Baukrowitz, Phosphatidylinositol 4,5-bisphosphate (PIP2) modulation of ATP and pH sensitivity in Kir channels. A tale of an active and a silent PIP2 site in the N terminus, J. Biol. Chem. 278 (2003) 10500–10505. https://doi.org/10.1074/jbc.M208413200. [DOI] [PubMed]
- 116.Niemeyer M.I., Cid L.P., Paulais M., Teulon J., Sepúlveda F.V. Phosphatidylinositol (4,5)-bisphosphate dynamically regulates the K2P background K+ channel TASK-2. Sci. Rep. 2017;7:45407. doi: 10.1038/srep45407. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.M.G. Leitner, C.R. Halaszovich, D. Oliver, Aminoglycosides inhibit KCNQ4 channels in cochlear outer hair cells via depletion of phosphatidylinositol(4,5)bisphosphate, Mol. Pharmacol. 79 (2011) 51 LP – 60. https://doi.org/10.1124/mol.110.068130. [DOI] [PubMed]
- 118.Zhang D., Howarth G.S., Parkin L.A., McDermott A.E. NMR studies of lipid regulation of the K+ channel KcsA. Biochim. Biophys. Acta - Biomembr. 2021;1863 doi: 10.1016/j.bbamem.2020.183491. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.Robinson C.V., Rohacs T., Hansen S.B. Tools for understanding nanoscale lipid regulation of ion channels. Trends Biochem. Sci. 2019;44:795–806. doi: 10.1016/j.tibs.2019.04.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120.Bolla J.R., Agasid M.T., Mehmood S., Robinson C.V. Membrane protein-lipid interactions probed using mass spectrometry. Annu. Rev. Biochem. 2019;88:85–111. doi: 10.1146/annurev-biochem-013118-111508. [DOI] [PubMed] [Google Scholar]
- 121.Reading E., Walton T.A., Liko I., Marty M.T., Laganowsky A., Rees D.C., Robinson C.V. The effect of detergent, temperature, and lipid on the oligomeric state of MscL constructs: insights from mass spectrometry. Chem. Biol. 2015;22:593–603. doi: 10.1016/j.chembiol.2015.04.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122.Laganowsky A., Reading E., Allison T.M., Ulmschneider M.B., Degiacomi M.T., Baldwin A.J., Robinson C.V. 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]
- 123.Gupta K., Li J., Liko I., Gault J., Bechara C., Wu D., Hopper J.T.S., Giles K., Benesch J.L.P., Robinson C.V. Identifying key membrane protein lipid interactions using mass spectrometry. Nat. Protoc. 2018;13:1106. doi: 10.1038/nprot.2018.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.P. Qiao, Y. Liu, T. Zhang, A. Benavides, A. Laganowsky, Insight into the Selectivity of Kir3.2 toward Phosphatidylinositides, Biochemistry 59 (2020) 2089–2099. https://doi.org/10.1021/acs.biochem.0c00163. [DOI] [PubMed]
- 125.D.S. Chorev, L.A. Baker, D. Wu, V. Beilsten-Edmands, S.L. Rouse, T. Zeev-Ben-Mordehai, C. Jiko, F. Samsudin, C. Gerle, S. Khalid, A.G. Stewart, S.J. Matthews, K. Grünewald, C. V Robinson, Protein assemblies ejected directly from native membranes yield complexes for mass spectrometry, Science (80). 362 (2018) 829 LP – 834. https://doi.org/10.1126/science.aau0976. [DOI] [PMC free article] [PubMed]
- 126.Marty M.T., Zhang H., Cui W., Blankenship R.E., Gross M.L., Sligar S.G. Native mass spectrometry characterization of intact nanodisc lipoprotein complexes. Anal. Chem. 2012;84:8957–8960. doi: 10.1021/ac302663f. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127.Moon S., Yan R., Kenny S.J., Shyu Y., Xiang L., Li W., Xu K. Spectrally resolved, functional super-resolution microscopy reveals nanoscale compositional heterogeneity in live-cell membranes. J. Am. Chem. Soc. 2017;139:10944–10947. doi: 10.1021/jacs.7b03846. [DOI] [PubMed] [Google Scholar]
- 128.Petersen E.N., Pavel M.A., Wang H., Hansen S.B. Disruption of palmitate-mediated localization; a shared pathway of force and anesthetic activation of TREK-1 channels. Biochim. Biophys. Acta - Biomembr. 2020;1862 doi: 10.1016/j.bbamem.2019.183091. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129.Sengupta P., Jovanovic-Talisman T., Skoko D., Renz M., Veatch S.L., Lippincott-Schwartz J. Probing protein heterogeneity in the plasma membrane using PALM and pair correlation analysis. Nat. Methods. 2011;8:969–975. doi: 10.1038/nmeth.1704. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130.Petersen E.N., Gudheti M., Pavel M.A., Murphy K.R., Ja W.W., Jorgensen E.M., Hansen S.B. Phospholipase D transduces force to TREK-1 channels in a biological membrane. BioRxiv. 2019 doi: 10.1101/758896. [DOI] [Google Scholar]
- 131.Kusch J., Zifarelli G. Patch-clamp fluorometry: electrophysiology meets fluorescence. Biophys. J. 2014;106:1250–1257. doi: 10.1016/j.bpj.2014.02.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132.Zheng J., Zagotta W.N. Gating rearrangements in cyclic nucleotide-gated channels revealed by patch-clamp fluorometry. Neuron. 2000;28:369–374. doi: 10.1016/s0896-6273(00)00117-3. [DOI] [PubMed] [Google Scholar]
- 133.Zheng J., Zagotta W.N. Patch-clamp fluorometry recording of conformational rearrangements of ion channels. Sci. STKE. 2003;2003:PL7. doi: 10.1126/stke.2003.176.pl7. [DOI] [PubMed] [Google Scholar]
- 134.Puljung M., Vedovato N., Usher S., Ashcroft F. Activation mechanism of ATP-sensitive K+ channels explored with real-time nucleotide binding. Elife. 2019;8 doi: 10.7554/eLife.41103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 135.Usher S.G., Ashcroft F.M., Puljung M.C. Nucleotide inhibition of the pancreatic ATP-sensitive K+ channel explored with patch-clamp fluorometry. Elife. 2020;9 doi: 10.7554/eLife.52775. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 136.Reimann F., Huopio H., Dabrowski M., Proks P., Gribble F.M., Laakso M., Otonkoski T., Ashcroft F.M. Characterisation of new KATP-channel mutations associated with congenital hyperinsulinism in the Finnish population. Diabetologia. 2003;46:241–249. doi: 10.1007/s00125-002-1014-3. [DOI] [PubMed] [Google Scholar]
- 137.Flanagan S.E., Patch A.-M., Mackay D.J.G., Edghill E.L., Gloyn A.L., Robinson D., Shield J.P.H., Temple K., Ellard S., Hattersley A.T. Mutations in ATP-sensitive K+ channel genes cause transient neonatal diabetes and permanent diabetes in childhood or adulthood. Diabetes. 2007;56:1930–1937. doi: 10.2337/db07-0043. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 138.Hedley P.L., Jørgensen P., Schlamowitz S., Moolman-Smook J., Kanters J.K., Corfield V.A., Christiansen M. The genetic basis of Brugada syndrome: A mutation update. Hum. Mutat. 2009;30:1256–1266. doi: 10.1002/humu.21066. [DOI] [PubMed] [Google Scholar]
- 139.R.Y. Kim, S.A. Pless, H.T. Kurata, PIP2 mediates functional coupling and pharmacology of neuronal KCNQ channels, Proc. Natl. Acad. Sci. 114 (2017) E9702 LP-E9711. https://doi.org/10.1073/pnas.1705802114. [DOI] [PMC free article] [PubMed]
- 140.van den Bogaart G., Meyenberg K., Risselada H.J., Amin H., Willig K.I., Hubrich B.E., Dier M., Hell S.W., Grubmüller H., 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]
- 141.Marrink S.J., Corradi V., Souza P.C.T., Ingólfsson H.I., Tieleman D.P., Sansom M.S.P. Computational modeling of realistic cell membranes. Chem. Rev. 2019;119:6184. doi: 10.1021/acs.chemrev.8b00460. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142.Ingólfsson H.I., Melo M.N., van Eerden F.J., Arnarez C., Lopez C.A., Wassenaar T.A., Periole X., de Vries A.H., Tieleman D.P., Marrink S.J. Lipid organization of the plasma membrane. J. Am. Chem. Soc. 2014;136:14554–14559. doi: 10.1021/ja507832e. [DOI] [PubMed] [Google Scholar]
- 143.Tao X., MacKinnon R. Molecular structures of the human Slo1 K(+) channel in complex with β4. Elife. 2019;8 doi: 10.7554/eLife.51409. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 144.Pope L., Lolicato M., Minor D.L.J. Polynuclear ruthenium amines inhibit K(2P) channels via a “finger in the dam” mechanism. Cell Chem. Biol. 2020;27:511–524.e4. doi: 10.1016/j.chembiol.2020.01.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145.Tao X., Avalos J.L., Chen J., MacKinnon R. Crystal structure of the eukaryotic strong inward-rectifier K+ channel Kir2.2 at 3.1 A resolution. Science. 2009;326:1668–1674. doi: 10.1126/science.1180310. [DOI] [PMC free article] [PubMed] [Google Scholar]