Abstract
Regulation of pancreatic KATP channels involves orchestrated interactions of their subunits, Kir6.2 and SUR1, and ligands. Previously we reported KATP channel cryo-EM structures in the presence and absence of pharmacological inhibitors and ATP, focusing on the mechanisms by which inhibitors act as pharmacological chaperones of KATP channels (Martin et al., 2019). Here we analyzed the same cryo-EM datasets with a focus on channel conformational dynamics to elucidate structural correlates pertinent to ligand interactions and channel gating. We found pharmacological inhibitors and ATP enrich a channel conformation in which the Kir6.2 cytoplasmic domain is closely associated with the transmembrane domain, while depleting one where the Kir6.2 cytoplasmic domain is extended away into the cytoplasm. This conformational change remodels a network of intra- and inter-subunit interactions as well as the ATP and PIP2 binding pockets. The structures resolved key contacts between the distal N-terminus of Kir6.2 and SUR1’s ABC module involving residues implicated in channel function and showed a SUR1 residue, K134, participates in PIP2 binding. Molecular dynamics simulations revealed two Kir6.2 residues, K39 and R54, that mediate both ATP and PIP2 binding, suggesting a mechanism for competitive gating by ATP and PIP2.
Graphical Abstract
Introduction
Pancreatic ATP-sensitive potassium (KATP) channels functionally couple glucose metabolism to insulin release and are crucial for glucose homeostasis (Ashcroft, 2005; Nichols, 2006). Structurally, the pancreatic KATP channel is an octameric complex composed of two distinct integral membrane proteins (Clement et al., 1997; Inagaki et al., 1997; Lee et al., 2017; Li et al., 2017; Martin et al., 2017b; Shyng and Nichols, 1997). A tetrameric core of Kir6.2 subunits form the central transmembrane pore of the channel. A coronal array of four sulfonylurea receptor 1 (SUR1) subunits surrounds the channel core, each SUR1 companioned with one Kir6.2 subunit. Genetic mutations of these subunits that dysregulate KATP channel activity are causes of neonatal diabetes (gain of function) and congenital hyperinsulinism (loss of function) (Ashcroft, 2005). KATP channels harbor multiple distinct and antagonistic binding sites for their primary physiological regulators, intracellular ATP and ADP, which close the ion channel through a binding site in Kir6.2, but open the channel through Mg-dependent binding on SUR1 (Nichols et al., 1996; Puljung, 2018). In addition, channel activity is operationally governed by binding sites for specific membrane phospholipids, particularly PIP2, which directly promote opening as well as antagonize the ATP inhibition at the Kir6.2 binding sites (Nichols et al., 1996). Finally, the pancreatic KATP channel is the drug binding target for sulfonylurea and glinide anti-diabetic medications, which inhibit channel activity and thus stimulate insulin secretion (Gribble and Reimann, 2003). The long held principal objective of KATP channel research has been to understand the protein dynamics by which these several ligand interactions, separately and in concert, ultimately determine levels of KATP channel activity, and hence control insulin release.
CryoEM structures of KATP channels have provided direct insights into the structural mechanisms of ligand recognition and gating regulation. In a previous study, we reported comparative cryoEM structures for pancreatic KATP channels in the absence of ligands (apo); in the presence of ATP; and in the combined presence of ATP with alternative pharmacological inhibitors: glibenclamide, repaglinide, or carbamazepine (Martin et al., 2019). The study found all pharmacological inhibitors occupy a common binding pocket located within SUR1 and that this binding pocket lies adjacent to the deep binding site for the Kir6.2 N-terminal tail, which courses through the prominent cleft between the two halves of the ABC (ATP Binding Cassette) module of SUR1. The findings offered mechanistic insight into how distinct pharmacological inhibitors inhibit channel activity and also facilitate channel assembly by stabilizing the interaction between Kir6.2 N-terminus and SUR1. However, it was noted during image analysis that each dataset in the study possessed considerable conformational heterogeneity, suggesting classification analyses within datasets may further illuminate channel structural dynamics relevant to ligand binding and gating.
Here, we show results from reprocessing of cyroEM datasets previously reported, focusing on conformational analysis and augmented by molecular dynamics (MD) simulations. Most notably, we found that the cytoplasmic domain (CTD) of Kir6.2 adopted two distinct conformations. In one, the CTD is tethered close to the membrane (Kir6.2-CTD-up). In the other, the CTD is counterclockwise corkscrewed away from the membrane, towards the cytoplasm (Kir6.2-CTD-down). Across structure datasets, the ratio of CTD-up versus CTD-down conformations strongly correlated with the occupation of inhibitory ligand binding sites. Importantly, drug binding and CTD conformation were associated with significant structural reorganization at the ATP and PIP2 binding sites, and at domain interfaces within and between subunits, suggesting ligands act as molecular glues to stabilize the channel in the Kir6.2-CTD-up conformation. Of further importance, improved cryoEM maps and functional analysis revealed that binding of the activating ligand PIP2 involves a direct interaction with SUR1 lysine-134 in TMD0, implicating a mechanism by which SUR1 enhances Kir6.2 PIP2 sensitivity. Moreover, MD simulations uncovered Kir6.2 residues that participate in both ATP and PIP2 binding, providing an explanation for how ATP and PIP2 compete to control KATP channel gating. Together, our findings provide a framework for understanding how ligands shift channel conformation to regulate channel activity and how mutations, now observed to affect key protein-protein and protein-ligand interfaces, disrupt channel function and cause disease.
Results
KATP channel conformation analysis
Focused 3D classification of the Kir6.2 tetramer core plus one SUR1 subunit (denoted K4S hereinafter) following symmetry expansion and signal subtraction (Scheres, 2016) was performed on our previously published five datasets: apo, ATP only, carbamazepine and ATP (CBZ/ATP), glibenclamide and ATP (GBC/ATP), repaglinide and ATP (RPG/ATP) (Martin et al., 2019) (see Methods). This strategy was employed to circumvent alignment difficulty due to flexible SUR1 (Fig.S1, S2). The analysis revealed two major K4S conformations: Kir6.2-CTD-up and Kir6.2-CTD-down, wherein the Kir6.2-CTD was alternatively located closer to, or further from, the Kir6.2 membrane spanning channel domains, respectively. Particularly, translation of the CTD between up and down conformations further involved a rotation, together comprising a corkscrew movement wherein the CTD (from an extracellular point of view) was rotated clockwise in Kir6.2-CTD-up conformation relative to Kir6.2-CTD-down (Fig.1). The CTD-up and CTD-down conformations appear qualitatively similar to the T(tense)-state and R(relaxed)-state previously reported by others using a fusion SUR1-Kir6.2 protein under three different ligand conditions, GBC+ATPγS, ATPγS, or MgADP, wherein the T-state exhibits 10.6–12.5° CW rotation viewed from the extracellular side and 3–4.2 Å translation towards the membrane relative to the R-state (Wu et al., 2018). Within both the CTD-up and CTD-down conformations, rocking and rotation of the Kir6.2-CTD was discernable using Relion 3 multibody refinement principal component analysis (Nakane et al., 2018). The heterogeneity was greater in the CTD-down than the CTD-up population of particles (Fig.S3), consistent with an increase in CTD mobility when detached from the membrane. Similar Kir6.2-CTD dynamics were observed using cryoSPARC 3D variability analysis (Punjani and Fleet, 2021).
Figure 1. Two distinct conformations of the Kir6.2-CTD in RPG/ATP bound KATP channels.
(A) CryoEM maps for the Kir6.2 tetramer core plus one SUR1 subunit are shown (semi-transparent grey, 1.0 σ contour), with the Kir6.2 subunit including KNtp in the Kir6.2-CTD-up (blue, 0.7 σ contour) and the Kir6.2-CTD-down (magenta, 0.7 σ contour) conformations. Compared to the CTD-up conformation, the Kir6.2-CTD in the CTD-down conformation is translocated from near the lipid bilayer towards the cytoplasm by ~4 Å (distance measured from the center of mass of G295- Cα in the G-loop gate of all four Kir6.2 subunits to the center of mass of F168-Cα in the helix bundle crossing gate of all four Kir6.2 subunits), and rotated counterclockwise by 12° (viewed from the extracellular side), measured by aligning structures onto the TM domain (residues 55–175) of the RPG CTD-up reference model, and calculating dihedral angles between K338-Cα of Chain A of Kir6.2 and the centers of mass of G245-Cα. (B) Fraction of particles in Kir6.2-CTD-up (blue) and Kir6.2-CTD-down (magenta) in KATP channels bound to different ligands or in apo state. (C) Variations in the extent of CTD translation and rotation observed for Kir6.2-CTD-up (blue dots) and Kir6.2-CTD-down (magenta dots) are shown for all datasets using RPG/ATP-Kir6.2-CTD-up as reference (circled blue dot). CTD translation away from the membrane is shown in negative value. There is a linear correlation between Kir6.2-CTD translation and Kir6.2-CTD-rotation (R2 = 0.9682, y = −0.3031x − 0.1755). Translation and rotation of the CTD from a human open KATP structure (Zhao and MacKinnon 2021, PDB: 7S5T) relative to the RPG/ATP Kir6.2-CTD-up reference structure is included for comparison (green dot).
Both the Kir6.2-CTD-up and Kir6.2-CTD-down conformations were observed in the GBC/ATP, RPG/ATP, CBZ/ATP and ATP-only datasets; however, relative abundance of the two conformations varied in the different liganded states (Fig. 1B, Table S1). The GBC/ATP and RPG/ATP datasets had the highest percentages of particles in the CTD-up conformation, with 92.5% and 71.2%, respectively. The CBZ/ATP dataset, which only had CBZ density in SUR1 but no clear ATP density in Kir6.2 likely due to lower concentrations of ATP used during sample preparation (see Methods) had a significantly lower percentage (22%) of particles in the CTD-up conformation, comparable to the ATP only dataset of 17.8%. In the absence of added ligands, i.e. the apo state, only Kir6.2-CTD-down conformation was observed. These findings show Kir6.2-CTD exists largely in two discrete conformations. That the distribution of the two states correlated with the binding of inhibitory ligands implicates the switching between these conformations as a crucial mechanistic event controlling channel opening and closing. The RPG/ATP dataset gave the highest resolution maps for both the Kir6.2-CTD-up and CTD-down conformations (3.4 Å and 3.6 Å, respectively; Fig.S1, S2, Table S1). The improved map quality compared to our previously published structure (Martin et al., 2019) allowed us to reevaluate ligand and protein densities that were previously ambiguous. We have therefore focused on the RPG/ATP dataset for structural analyses hereinafter.
In the RPG/ATP state, the predominant SUR1 conformation is arranged like a propeller when symmetrized, as described in our previous study (Martin et al., 2019). In addition, we identified a minor conformation (~27% of all particles; Fig.S1, Table S1) showing a large clockwise rotation of SUR1 towards the Kir6.2 tetramer (viewed from the extracellular side). This conformation is qualitatively similar to the quatrefoil conformation previously reported in the MgATP/MgADP-bound, NBDs-dimerized Kir6.2-SUR1 fusion channel structure (Lee et al., 2017), and likewise our recently reported quatrefoil-like Kir6.1-SUR2B vascular KATP channel structure bound to GBC and ATP with separate NBDs (Sung et al., 2021). The overall map resolution of this minor class is ~7 Å (Fig.S1), which precluded detailed structural analysis. Nonetheless, it reveals that even in the presence of RPG and ATP, a large rotation of SUR1 resembling that seen in NBDs dimerized SUR1 quatrefoil conformation occurs, albeit much less frequently. Heterogeneity of SUR1 within the dominant propeller conformation with more subtle rotations of SUR1’s ABC core around the Kir6.2 tetramer was also observed regardless of whether the Kir6.2-CTD is up or down. We explored the details of SUR1 dynamics using the Kir6.2-CTD-up class of particles, classifying the dynamic motion as discrete eignevectors using Relion 3 multibody refinement (see Methods). Particles with amplitudes between 5 and 20, and between −5 and −20, along eigenvector 1 were refined separately to generate two maps, referred to as SUR1-in and SUR1-out conformations at 3.9 Å and 3.8 Å (Fig.S4), respectively, for model building (Table S2) and structural analysis.
Comparison of different KATP conformations
The changes in conformation of Kir6.2-CTD and SUR1 were accompanied by remodeling of subunit and domain interfaces as well as protein-ligand interactions pertinent to gating. Comparing CTD-down to the CTD-up conformation, the Kir6.2-CTD is translated down into the cytoplasm by ~4 Å along an axis perpendicular to the embedding membrane, and simultaneously counterclockwise (CCW) rotated by 12° about that axis viewed from the extracellular side (Fig.1, movie 1). The descended location of Kir6.2-CTD reconfigured the interfacial (IF) helix (also called the slide helix, herein taken to include G53-D65) in the Kir6.2 N-terminus, and also the C-linker (herein taken to include H175-L181) by which inner helix M2 interacts with the CTD (Fig.2). Specifically, in the CTD-up conformation, the IF helix adopted a 310 helix characteristic (Vieira-Pires and Morais-Cabral, 2010) wherein a directional kink at D58 demarcated the helix into N-terminal and C-terminal halves, with the N-terminal half pivoted towards SUR1 instead of along the adjacent Kir6.2 subunit. In contrast, the IF helix in the CTD-down conformation formed a continuous helix that extended towards the neighboring Kir6.2 subunit. Respecting the C-linker, in the CTD-up conformation, the C-linker formed a helical structure that participated in membrane PIP2 binding; while in the CTD-down conformation, the C-linker was fully unraveled into a loop, in which a key PIP2 interacting residue R176 (Baukrowitz et al., 1998; Shyng and Nichols, 1998) was distant from the membrane and incapable of direct PIP2 interaction. In comparisons between the SUR1 propeller in and out structures (Fig.S1, S4), the ABC module in the SUR1-in structure is rotated clockwise closer to a neighboring Kir6.2 (viewed from the extracellular side). In this rotated position, the SUR1-L0 loop was pulled away from its interaction with SUR1’s direct Kir6.2 subunit partner. Specific molecular changes at the subunit and domain interfaces and ligand binding sites in different conformations and their relevance to gating are described below.
Figure 2. Structural comparison between RPG/ATP CTD-up and CTD-down conformations.
(A) Superposition of the Kir6.2-CTD-up (blue) and CTD-down (magenta) structures. The red boxed region shows significant secondary structural difference at the IF helix and the C-linker. (B) Close-up view with structural differences highlighted (green boxes) at the IF helix and the C-linker in the two conformations. (C) Close-up view of the inter-subunit H-bond interactions between D65 and T293, and an inter-subunit salt-bridge between D58 and R206, as well as an intra-subunit salt-bridge between R177 and D204, and an inter-subunit ATP-binding interaction from K39 in the Kir6.2-CTD-up conformation. (D) Same close-up view as in (C) but of the Kir6.2-CTD-down conformation showing disruption of those interactions seen in (C). Structural elements and residues from chain a or chain b are labeled with a or b at the end.
Intra-Kir6.2 and inter Kir6.2-Kir6.2 interactions--
Inspection of the Kir6.2 CTD-up and CTD-down structures revealed changes in intra- and inter-Kir6.2 interactions involving structural elements which occupy the transitional space between the membrane spanning and cytoplasmic domains of Kir6.2. These include the IF helix, the C-linker, the DE loop (the loop that connects βD and βE), the G-loop (the lower cytoplasmic gate) as well as the N-terminus and ATP. Specifically, in the CTD-up conformation, D204 at the start of the DE loop forms an intrasubunit salt bridge with R177 in the C-linker, which connects to Kir6.2 TM helix 2; and R206, also in the DE loop forms a salt bridge with D58 in the IF helix of the adjacent Kir6.2 subunit (Fig.2C). Two hydrogen bonds: one between D65 in the IF helix and T293 in the G-loop from adjacent subunits, and one between K39 at the N-terminus and ATP at a neighboring Kir6.2 subunit were also observed (Fig.2C). The salt bridges and hydrogen bonds bind the different structural elements together and hold the CTD in the up conformation. In contrast, the aforementioned salt bridges and hydrogen bonds were eliminated in the CTD-down structures (Fig.2D). In particular, separation of D204 from R177 in CTD-down was accompanied by uncoiling and extension at the end of the C-linker helix including R177, while separation of R206 from D58 was accompanied by a significant straightening of the kink in the IF helix and reorientation of the continuing chain, which connects to the base of the KNtp (Fig.2).
Consistent with the notion that these labile salt bridges have critical roles in channel gating, previously published functional studies have implicated the participating Kir6.2 residues in channel regulation. D58 was previously suggested to be involved in anchoring Kir6.2 CTD to the TM domain through results of targeted mutagenesis (Li et al., 2013). Our findings resolve the salt bridge partnership with R206 in the neighboring Kir6.2 polypeptide, and further reveal such tethering is dynamically incorporated into the conformational changes of ion channel activation. Consolidating this view, R206 has been separately implicated in channel activation by PIP2, through scanning mutagenesis investigations of positive residues involved in effecting bound PIP2, wherein mutation R206A was found to abolish PIP2 response and thus diminish channel activity (Li et al., 2013; Lin et al., 2005; Shyng et al., 2000). Similar to R206A, mutation R177A also abolishes or greatly attenuates channel activity by diminishing PIP2 response (Li et al., 2013; Shyng et al., 2000). Thus, in corroborating earlier functional studies, our structural findings here elucidate key molecular interactions that place the Kir6.2-CTD close to the membrane in position to interact with membrane-bound phospholipids for channel opening. These insights are directly relevant to human health. Mutations of each of the four residues in the above salt bridges have been identified in congenital hyperinsulinism, a disease caused by loss of function of KATP channels. These include D58V (De Franco et al., 2020), R177W (Arya et al., 2014), D204E (Pinney et al., 2008), and R206H (Boodhansingh et al., 2019). Our structures here provide a mechanistic illustration of how perturbation of residues involved in conformational dynamics cause loss of channel function and hyperinsulinism.
Kir6.2 and SUR1 interactions--
Our previous study of pancreatic KATP channel structure suggested that a key regulatory interface through which SUR1 controls Kir6.2 channel activity is formed by the extended N-terminus of Kir6.2 (residues 1–30; referred to as KNtp) that is inserted within its SUR1 subunit partner, wherein the KNtp is located within the SUR1 ABC transporter module (Martin et al., 2019). We showed in particular that the KNtp is located between the two transmembrane helix bundles (TMBs) of SUR1 ABC module, and adjacent to the drug binding pocket of GBC, RPG, and CBZ. More detailed structural analysis was hindered by insufficient resolution for the density of KNtp in those published cryoEM maps. The additional analysis methods applied in our current study yielded clearer, contiguous densities and produced significantly improved maps revealing specific interactions between residues in KNtp and SUR1 (Fig.3). In the distal part of KNtp, which lies deep in the SUR1 ABC core cavity, Kir6.2-R4 is in bonding position with SUR1-T1139 and N1301. In the middle section of KNtp, which is near the entrance to the SUR1 ABC core cavity, Kir6.2-L17 interacts with R826 of NBD1 and G1119 and N1123 of TMB2. In the proximal end of KNtp, cryoEM density corresponding to residues P24, Y26 and R27 comes into close contact with cryoEM density correponding to SUR1’s NBD1-TMD2 linker around residue S988. In the Kir6.2-CTD down structure, interactions at the distal and middle segments of KNtp with SUR1 remain largely unchanged; however, the proximal section of KNtp is significantly further away from the NBD1-TMD2 of SUR1 (Fig.3D). In a recent Kir6.1-SUR2B cryoEM structure we showed that the NBD1-TMD2 linker has a role in regulating MgADP-dependent interactions between SUR2B-NBD2 and Kir6.1-CTD (Sung et al., 2021). Our structures presented here reveal an additional contact between NBD1-TMD2 linker and KNtp. Whether this contact and changes at this interface in different conformations have functional significance warrants future investigation. As reported previously (Martin et al., 2019), the cryoEM density of KNtp is the strongest in the GBC/ATP, RPG/ATP, and CBZ/ATP datasets, followed by ATP only, and is the weakest in the apo state, indicating inhibitory ligands stabilize the KNtp-SUR1 interface. Deletion of KNtp is known to increase channel open probability (Babenko et al., 1999; Koster et al., 1999; Reimann et al., 1999), while immobilizing KNtp in the SUR1 ABC core cavity via engineered crosslinking between Kir6.2-L2C and SUR1-C1142 inhibits channel activity (Martin et al., 2019). Worth noting, mutations L2P (Alkorta-Aranburu et al., 2014), R4C/H, L17P, R24C, R27C/H in Kir6.2 (De Franco et al., 2020) as well as R826W (de Wet et al., 2008) and N1123D (Suzuki et al., 2007) of SUR1 have all been reported in neonatal diabetes or congenital hyperinsulinism, which underscores the importance of the KNtp-SUR1 interface in channel gating.
Figure 3. KNtp and SUR1 interface.
(A) KNtp (Kir6.2 aa 1–30) cryoEM density (purple mesh, 1.0 σ contour) in RPG-Kir6.2-CTD-up structure, with key residues that interact with SUR1 shown as orange spheres within red circles. The distal portion of KNtp is located deep in the cavity of the SUR1-ABC core module. The middle section of KNtp lies near the entrance of the cavity. The proximal part (i.e. C-terminal part) of KNtp including P24, Y26 and R27 contacts the SUR1 N1-T2 linker density (shown as green mesh; 1.0 σ contour) near S988. (B) Left: Close view of distal KNtp cryoEM density (purple mesh, 1.0 σ contour) showing interaction of R4 with SUR1 T1139 and N1301 in TMD2 and proximity to bound RPG. Right: Close-up view of middle KNtp in cryoEM density (purple mesh, 1.0 σ contour) showing interaction of L17 with SUR1 R826, G1119 and N1123. (C) KNtp viewed without SUR1, showing its interconnectivity with two ATP binding sites and the LM-loop in the CTD of a neighboring Kir6.2. (D) Superposition of Kir6.2-CTD-up (blue) and CTD-down (pink) structures viewed from the extracellular side showing divergent positions of proximal KNtp.
In addition to KNtp forming interactions with the SUR1-ABC module, regions C-terminal to KNtp in the Kir6.2-N terminal domain were also observed to be intimately involved in protein-protein and protein-ligand interactions. First, in the CTD-up structure, Kir6.2 R31-R34 is close to the short loop that connects βL and βM (LM loop) (Martin et al., 2017b) of the neighboring subunit (Fig.3C). A mutation D323K in the LM loop has been shown to disrupt ATP inhibition (Brennan et al., 2020). Second, further downstream K39 has its sidechain oriented towards ATP bound to the neighboring Kir6.2 on the other side. Thus, we found the Kir6.2 N-terminus is connected simultaneously to two ATP binding pockets. Dynamic movement of the KNtp between the CTD-up and CTD-down conformations may therefore impact the interactions of downstream Kir6.2-N terminal domain with neighboring subunits on both sides and with ATP. Finally, the refined structures showed that a loop (K47-Q52) N-terminal to the IF helix of Kir6.2 has close interaction with SUR1’s TMD0-intracellular loop 1 (ICL1), ICL2 and ICL3 (i.e. L0). Here, a compact network of interactions stabilizes the Kir6.2-CTD close to the membrane and also reinforces ATP binding. In the CTD-down conformation, the Kir6.2 pre-IF helix loop becomes more distant from the SUR1-ICLs such that the Kir6.2-CTD is no longer tethered close to the membrane, which also impacts the ATP binding pocket (see below).
ATP binding pocket--
Rapid and reversible closure upon non-hydrolytic binding of ATP at Kir6.2 is a cardinal feature of KATP channels (Nichols et al., 1996). The improved map quality in the current study allowed us to refine interpretation of the ATP cryoEM density and the interaction network that coordinates ATP binding and follow how the ATP binding pocket becomes reconfigured in different conformations.
In our improved maps, the ATP density could be modeled with ATP in two alternative poses. In the first, the γ-phosphate is oriented upward towards R50 of Kir6.2, which is consistent with functional studies indicating that R50 interacts with the γ-phosphate of ATP (Trapp et al., 2003). This pose was used to model ATP density in our previously published structure bound to GBC and ATP (PDB: 6BAA) (Martin et al., 2017a) and also to model ATPγS bound to a rodent SUR1–39aa-Kir6.2 fusion KATP channel (Wu et al., 2018). In the second pose, the ATP’s γ-phosphate is oriented downward facing N335. This alternative orientation is also supported by functional data showing that N335Q decreases ATP sensitivity (Drain et al., 1998) and used to model ATP density bound to Kir6.2 in cryoEM structures of a human SUR1–6aa-Kir6.2 fusion KATP channel (PDB: 6C3O and 6C3P) (Lee et al., 2017). Of note, we also observed an unassigned protruding density in ATP in our initial GBC/ATP map (EMD-7073), which we speculated may be contaminating Mg2+ (Martin et al., 2017a) but which can be well modeled by the alternative pose of the γ-phosphate. The simplest interpretation is that the cryoEM density of ATP we observed is likely an ensemble of the two possible γ-phosphate poses.
The improved map also showed clear cryoEM density for the side chain of K205 in the L0 of SUR1. We have previously proposed that K205 participates in ATP binding (Martin et al., 2017b) based on an early finding that K205E reduces ATP inhibition (Pratt et al., 2012). However, our previously published cryoEM map (EMD-7073) does not resolve the side chain density of K205 sufficiently to permit definitive conclusion. In our current map, K205 side chain was clearly oriented to the bound ATP (Fig.4B), stabilizing interactions with the β- and γ-phosphates of ATP. Similar observations have been reported by Ding et al. (Ding et al., 2019). The role of K205 in ATP binding is further substantiated by Usher et al. (Usher et al., 2020) in which binding affinity between a fluorescent ATP analogue and the channel was assessed by FRET measurements between the ATP analogue and a fluorescent unnatural amino acid ANAP (3-(6-acetylnaphthalen-2-ylamino)–2-aminopropanoic acid) engineered at Kir6.2 amino acid position 311. The study found that SUR1-K205A and K205E mutations reduce ATP binding affinity by ~5 and 10-fold.
Figure 4. Comparison of ATP binding pocket and SUR1-TMD0/Kir6.2-CTD interface between Kir6.2-CTD-up and CTD-down conformations.
(A) Surface representations of the SUR1-TMD0/Kir6.2-CTD interface and the ATP binding pocket in Kir6.2-CTD-up (left) and the CTD-down (right) conformations. SUR1 is shown in deep blue or pink hues, Kir6.2-CTD in pale blue or pink hues, and ATP as stick model. The arrows point to contacts between SUR1 and Kir6.2 in CTD-up panel which are lost (dashed arrows) in the CTD-down panel. Loss of the tight interaction between the SUR1-TMD0 and Kir6.2-CTD domains renders the ATP binding pocket less compact. (B) Detailed views of residues surrounding ATP in the Kir6.2 CTD-up conformation (blue, left) which become more distant from ATP and/or interacting partners in the CTD-down conformation (pink, right), including K205 of SUR1, Q52, R54 and K39 of Kir6.2. ATP cryoEM density (3.3 σ contour) is represented by a mesh.
The conformational dynamics we observed in Kir6.2-CTD and SUR1 had significant impact on the structure of the ATP binding site. In the Kir6.2-CTD-up conformation, the Kir6.2-CTD was packed tightly against SUR1’s ICL1, ICL2, and the initial segment of L0 (Fig.4A; contact surface area ~144 Å2, calculated using PDBePISA). In this conformation, ATP was fitted snugly into the pocket formed by the N- and C-terminal domains of adjacent Kir6.2 subunits and L0 of SUR1, and had an ATP interface area ~430 Å2. In the CTD-down conformation, the Kir6.2-CTD became disengaged from the SUR1-ICLs (Fig.4A; contact surface area ~9 Å2), which disrupted several interactions that had stabilized ATP binding, and reduced the ATP interface area to ~380 Å2. Specifically, R54, which was oriented towards the ATP in the CTD-up state became distant from the ATP binding pocket in the CTD-down structure. K39 in the N-terminus of neighboring Kir6.2, which also coordinated ATP binding in the CTD-up structure, was reoriented away from the ATP in the CTD-down structure (Fig.2C, D). Moreover, the distance between K205 in the L0 of SUR1 and ATP increased in the CTD-down conformation. These changes together weakened the ATP binding pocket and exposed ATP to solvent. In addition to impacting Kir6.2-CTD dynamics, SUR1 rotation also affected ATP binding. When SUR1-ABC module rotated toward the Kir6.2 tetramer core (SUR1-in), L0 pulled away from the ATP binding pocket. As a result, SUR1-K205 lost interaction with ATP, destabilizing binding (Fig.4B).
PIP2 binding site--
At the binding pocket where PIP2 is predicted to bind based on homology with Kir2 and Kir3 channels for which PIP2 bound structures are available (Hansen et al., 2011; Lee et al., 2016b; Niu et al., 2021), a lipid cryoEM density is seen in both Kir6.2-CTD-up and CTD-down conformations. Interestingly, the lipid density in the CTD-up conformation is significantly larger than that in the CTD-down conformation (Fig.5A). This was consistent for all datasets that include both conformations. We were able to fit, and tentatively model, the lipid density in the CTD-up structure with two phosphatidylserine (PS) molecules and that in the CTD-down structure with one PS molecule (Fig.S5). Since no PIP2 was added to our samples prior to imaging, we reasoned that the more abundant PS may have entered the binding pocket. In a recent study by Zhao and MacKinnon (Zhao and MacKinnon, 2021), it was shown that PIP2 is not required for KATP channel activity, suggesting other phospholipids that occupy the PIP2 binding site could potentially support channel activity. Whether the density in our structure represents PS, co-purified endogenous PIP2, or other phospholipids requires further investigation.
Figure 5. Comparison of the PIP2 binding pocket in Kir6.2-CTD-up and CTD-down conformations.
(A, B) PIP2 binding pocket of Kir6.2-CTD up (A, blue) and Kir6.2-CTD down (B, pink) conformations viewed from the side. Lipid densities are shown as a mesh (1.0 σ contour). In (A), in addition to Kir6.2 residues previously implicated in phospholipid binding, SUR1-K134 side chain is pointed directly at the lipid density. SUR1 is shown in deep blue or pink, Kir6.2 in pale hues. (C, D) Same as (A, B) viewed from the extracellular side with the helical bundle crossing shown (F168). Note the PIP2 binding pocket is more compressed in Kir6.2-CTD-down than CTD-up conformation due to secondary structural change at the IF helix that brings Q57 to interact closely with F60 and W68 (red dashed circle). (E, F) Inside-out patch-clamp recording (examples in E) show greater fold current increase in response to PIP2 of the SUR1-K134A mutant channel than WT channel (left), with statistically significant difference (*p<0.05, student’s t-test).
In the Kir6.2-CTD-up structure, in addition to the set of Kir6.2 residues K67 and W68 in the outer helix, and R176 in the C-linker previously implicated in PIP2 binding (Brundl et al., 2021; Cukras et al., 2002; Shyng and Nichols, 1998), we found K134 in TMD0 of SUR1 was in close contact with the density corresponding to lipid headgroups (Fig.5A). To test whether SUR1-K134 has a role in the PIP2 sensitivity of KATP channel opening, we functionally characterized a mutant KATP channel in which this residue is mutated to alanine (SUR1-K134A), using inside-out patch-clamp recording (see Methods). Compared to WT channels, the SUR1-K134A mutants exhibited substantially smaller initial currents in ATP-free solution. Upon PIP2 addition, however, the mutant channel currents increased by 5.93±2.17-fold, which is significantly higher than the 1.27±0.23-fold current increase seen in WT channels (Fig.5E, F), indicating the SUR1-K134A mutation reduced intrinsic Po and PIP2 interactions. It is well documented that channel Po, determined by channel interaction with PIP2, is primarily conferred by SUR1 association with Kir6.2. The Kir6.2 channel itself has low Po, but co-expression with SUR1 or just the TMD0 domain of SUR1 increases channel Po by more than 10-fold (Babenko and Bryan, 2003; Chan et al., 2003; Enkvetchakul et al., 2000; Pratt et al., 2011). Our results show SUR1-TMD0 participates in PIP2 interaction, at least in part through K134, which strengthens PIP2 interactions with Kir6.2.
In the CTD-down structure, the IF helix was closer to W68 near the cytoplasmic end of the outer helix of the neighboring Kir6.2 than in the CTD-up structure, causing compression of the PIP2 binding pocket (Fig.5A–D). This provides an explanation for why the lipid cryoEM density in the CTD-down structure was significantly weaker than that in the CTD-up structure and could be tentatively fit by only one PS molecule (Fig.S5). Moreover, the simultaneous unwinding of the C-linker in the CTD-down structure withdrew the key PIP2-interacting residue R176 to a position too distant for interaction. In this conformation, Kir6.2 is expected to be inactive.
Elucidating the relationship between ATP and PIP2 binding by MD simulations
ATP and PIP2 compete with each other to close and open the channel, respectively (Baukrowitz et al., 1998; Shyng and Nichols, 1998). However, the structural mechanism underlying this functional competition has remained unresolved. Previous mutation-function correlation studies led to a proposal that ATP and PIP2 have overlapping but non-identical binding residues (Cukras et al., 2002; Shyng et al., 2000; Tucker et al., 1998). To test this hypothesis, we conducted MD simulation studies using as a starting point the Kir6.2 (32–352) plus SUR1-TMD0 (1–193) tetramer part the RPG/ATP Kir6.2-CTD-up structure. Previous studies have shown that Kir6.2 and TMD0 of SUR1 form “mini KATP channels” (Babenko and Bryan, 2003; Chan et al., 2003), which like WT channels exhibit functional antagonism between ATP and PIP2 (Babenko and Bryan, 2003; Pratt et al., 2011). The mini KATP channel system is therefore suitable for simulating residues which may participate in binding of both ligands. Three independent 1 μs simulations were carried out for each of two ligand conditions, either without ATP or PIP2 (apo), or with both ATP and PIP2 in their respective binding pockets (ATP+PIP2) (Fig.S6A, B; for details see Methods).
Comparing the two different conditions, there was an overall increase in dynamics of the Kir6.2-CTD in the apo simulations versus the ATP+PIP2 simulations (Fig.6, Fig.S6, movie 2–5). First, significant secondary structural changes at the IF helix and the C-linker were observed in the apo simulations, resembling the changes from the CTD-up to the CTD-down conformation we observed in cryoEM structures. Second, the entire CTD relaxed towards the cytoplasm in the apo simulations. This was quantified by measuring the distance between the helix bundle crossing (HBC) gate at F168 and the G-loop gate at G295 (Fig.6A). In apo simulations, this distance increased over time in all three runs, whereas it remained relatively unchanged for the ATP+PIP2 simulations (Fig. 6B), except in run 2 during which the distance increased when ATP became partially dissociated at around 500 ns (Fig.6B). These findings show that in the absence of ligands, the Kir6.2-CTD has a tendency to relax toward the CTD-down conformation.
Figure 6. MD simulations of Apo versus ATP+PIP2 state.
(A) Positioning of the helix bundle crossing (HBC) and G-loop gate in a representative apo-simulation and ATP-PIP2-simulation. The time-varying position of the geometric center (plotted in red) of G295 (G-loop gate) Cα carbons of all four chains is overlaid on beginning and end snapshots of G295 Cα carbons, after aligning trajectories to the Kir6.2 TM domain. (B) Plots of the distance between the geometric centers of all four Cα atoms of F168 and of G295 for the entire simulations. Greater distances between the two gates in the apo state compared to the ATP+PIP2 state indicate relaxation of the CTD in the absence of ligands. In one apo simulation (light blue), ATP became partially dissociated at around 500 ns (red arrow). (C) Snapshots of simulations in the presence of ATP and PIP2 showing interactions of R54 and K39 with either ATP or PIP2. Kir6.2-K185 and SUR1-K134 which only interact with ATP or PIP2 respectively are also shown. (D) Heatmaps showing fraction of time residues spend at increasing distances from ATP (horizontal axis) or PIP2 (vertical axis). A representative trajectory corresponding to chain 1 in run 1 and colored to show time evolution is shown for each residue. K39 and R54 (top row) exhibit switching behavior and occasional simultaneous contact, while essentially exclusive ATP binding residues are shown in the middle row, and PIP2 binding residues in the bottom row.
Analysis of the minimum distance between each of the ligands and their surrounding residues within 4 Å over the entire simulation revealed that K39 and R54 of Kir6.2 engaged in both ATP and PIP2 binding. Fig.6D shows the fraction of time over the entire simulation each residue in each subunit and each run came into contact with ATP or PIP2. R54 and K39 each showed partial ATP and PIP2 occupancy (Fig.6C,D, Fig.S6C,D, movie 4, 5), which was in contrast to well established ATP binding residues, including R50 and K185, and PIP2 binding residues K67 and R176, which showed nearly 100% ATP or PIP2 occupancy. Of the two dual occupancy residues, R54 showed greater interactions compared to K39 with both ATP and especially with PIP2. K39, while showing interaction with ATP in all three runs, only showed significant interaction with PIP2 in one of the three runs (Fig.S6D). The analysis also identified residues that had specific, although less stable, interactions with either ATP or PIP2 as defined by distance between residue and ligand < 4 Å. In particular, Kir6.2-Q52 specifically interacted with ATP, and SUR1-K134 with PIP2, contrasting with the dual ligand binding mode of R54 and K39. A previous mutagenesis study has shown that mutation of either K39 or R54 to alanine reduces channel open probability as well as sensitivity to ATP inhibition (Cukras et al., 2002), which early implicated a role for these residues in channel gating by PIP2 and ATP. Confirming a role in physiological regulation, mutations R54C and R54H are linked to congenital hyperinsulinism (De Franco et al., 2020) and mutation K39R to transient neonatal diabetes (Zhang et al., 2015). Our MD simulation results suggest both residues participate directly in ATP and PIP2 binding, providing mechanistic insight into how mutation of these residues affect PIP2 and ATP sensitivities and cause disease.
Discussion
Cryo-preserved purified protein samples may contain multiple protein structures that represent distinct functional or transitional states and can provide mechanistic insight (Nogales and Scheres, 2015). In this study, analysis of five KATP channel cyroEM datasets collected in different ligand conditions revealed conformational heterogeneity of the Kir6.2-CTD and SUR1 ABC module. We observed the Kir6.2-CTD in either an “up” position tethered close to the plasma membrane, or a “down” position corkscrewed away from the membrane towards the cytoplasm. The ratio of the two conformations correlated with occupancy of inhibitory ligands at the SUR1 and Kir6.2 binding sites (Fig.1), suggesting inhibitory ligands help stabilize the Kir6.2-CTD close to the membrane. Furthermore, in both Kir6.2-CTD conformations the SUR1 ABC module was observed oriented with a range of rotation around the Kir6.2 tetramer central axis (Fig.S4). We observed a restructuring of protein-protein and protein-ligand interfaces in different conformations that sheds light on how ligands shift channel conformational dynamics to regulate gating.
Correlation between Kir6.2-CTD conformation and channel function
The structures analyzed in this study all represent closed channels. Recently, an open human KATP channel structure containing Kir6.2 C166S and G334D mutations was reported (Zhao and MacKinnon, 2021), which showed a Kir6.2-CTD that is further CW rotated (extracellular view) and slightly upward translated compared to our Kir6.2-CTD-up conformation (Fig.1C). Rearrangement of the molecular interactions between the IF helix, the C-linker, and TM residues as well as increased distance between the pre-IF helix loop and SUR1 L0 compared to the ATP-bound closed WT channel structure (equivalent to our Kir6.2-CTD-up structure) were observed (Fig.7). The restructuring widens the ATP-binding pocket, explaining the absence of ATP cryoEM density despite high concentrations of ATP in the sample, and stabilizes HBC gate opening via side chain interactions between F60 in the IF helix and the HBC gate residue F168. Another pre-open KATP channel structure using a rat SUR1–39aa-Kir6.2 H175K fusion construct published while this manuscript was under review (Wang et al., 2022) showed similar structural characteristics as the open channel structure. Taken together, a picture emerges wherein rotational and translational position of the Kir6.2-CTD determines the functional state of the channel. When the CTD is in the down position, the phospholipid binding pocket is compressed due to a change in secondary structures of the IF helix (Fig.5) and the C-linker is unwound by the increased separation of the CTD from the membrane domain. We propose this conformation corresponds to an “inactivated” state in which the CTD is unable to engage with membrane phospholipids and thus open the channel. When the CTD is in the clockwise up-screwed position with ATP and/or drug bound, the channel is primed for opening but is arrested in an inhibited state due to an interaction network between SUR1-L0 and the Kir6.2-N terminal domain, an interaction stabilized by ATP that prevents further rotation of the CTD needed to open the HBC gate. Upon ATP dissociation, the CTD is released for further CW rotation, enabling the gate to widen and open the channel.
Figure 7. Comparison of an inhibitor-bound CTD-up closed structure and a mutant open structure.
(A) Overlay of RPG+ATP CTD-up structure (blue) with a Kir6.2 double mutant (C166S, G334D) structure (PDB:7S61; yellow). Only two Kir6.2 subunits are colored. A small upward translation of the mutant open structure relative to the inhibitor-bound closed structure is indicated by the small red arrow next to the red box, which is shown in a 90° rotated enlarged view in B. (B) Overlay of the Kir6.2 tetramer with part of the SUR1-L0 viewed from the extracellular side. A CW rotation (5°) of the CTD from the inhibitor-bound closed structure to the mutant open structure is noted. Structural elements and residues from chain b are labeled with b at the end to distinguish from those from chain a. Residues which show significant differences in the two structures are labeled in red. (C) Same as B but with the two structures shown separately and residues in the C-linker.b visible. The red dashed circles illustrate the enlargement of the potassium ion path at the helix bundle crossing (F168) in the open structure.
To explain our structural data and a wealth of electrophysiological data we propose the Kir6.2-CTD undergoes transitions between four principal conformation states in dynamic equilibrium: CTD-down inactivated, CTD-up unliganded and closed, CTD-up bound to inhibitory ligands, and CTD-up open, with the probability to occupy a given conformation driven by ligands (Fig.8), similar to the model previously proposed by Borschel et al. (Borschel et al., 2017). In the absence of ligands the CTD-down conformation dominates, as seen in our apo state dataset, and the channel is inactivated. While not observed in our structural studies here, inactivated channels can transition into a short-lived unliganded CTD-up conformation at low probability. Binding of physiological inhibitor ATP at Kir6.2, and/or a pharmacological inhibitor at SUR1, shifts Kir6.2 towards a stable CTD-up but closed conformation. Binding of phospholipids such as PIP2, when coupled with unbinding of inhibitory ligands, shifts the equilibrium towards the Kir6.2-CTD-up open position. Under physiological conditions with high intracellular ATP concentrations and ambient phospholipids, we expect the Kir6.2-CTD to be mostly in an ATP inhibited CTD-up conformation, with a small fraction in a CTD-up state having the phospholipids bound and a further-rotated-open conformation, and rarely in an unliganded CTD-up conformation; the CTD-down inactivated conformation would also be rare. However, in pathological conditions the CTD-down inactivated state could be prevalent. We and others have previously reported several mutations including congenital hyperinsulinism-causing mutations at the Kir6.2 subunit-subunit interface (such as R192A, E229A, R314A, R301A/C/H) that promote channel inactivation (Borschel et al., 2017; Lin et al., 2008; Lin et al., 2003; Shyng et al., 2000). Channels containing such mutations briefly open upon patch excision into ATP-free solution but then quickly inactivate. Interestingly, inactivation can be overcome by transiently exposing channels to high concentrations of ATP, followed by washout of ATP. We propose that these mutations increase an energy barrier for Kir6.2-CTD to transition from the CTD-down conformation to CTD-up conformation, thus trapping the CTD in the down inactivated state. ATP thus effectively acts as a molecular glue at its Kir6.2 domain interfaces. Exposure to ATP shifts the Kir6.2-CTD back to the CTD-up position such that channels can open again when ATP is washed out. The model similarly explains the ability of PIP2 to prevent and reverse inactivation mutants from inactivation (Borschel et al., 2017; Lin et al., 2008; Lin et al., 2003; Shyng et al., 2000) by stabilizing Kir6.2-CTD in the up and further rotated open position.
Figure 8. Correlating Kir6.2-CTD structures with functional states of pancreatic KATP channels.
Cartoon representation of the structural and functional states of KATP channels. In the presence of high concentrations of intracellular ATP and ambient PIP2, channels are mostly in ATP-bound closed state in which the Kir6.2-CTD is in the up-conformation and docked near the membrane and rotated CW from an extracellular perspective. Removal of ATP results in a transient unliganded closed state with the CTD in the up position conducive to binding PIP2. Binding of PIP2 opens the channel in which the Kir6.2-CTD is further CW rotated and moved up towards the membrane. Inactivation occurs when the CTD in the unliganded state transitions into the down conformation, a process that is enhanced by inactivation mutations including those known to cause hyperinsulinism. ATP facilitates channel recovery from inactivation by shifting the equilibrium towards the ATP-bound closed state in which the Kir6.2-CTD is in the up-conformation to allow the channel to transition to the unliganded CTD-up closed state upon subsequent removal of ATP, primed for PIP2 binding and channel opening. Addition of PIP2 also counters inactivation by shifting the equilibrium via the unliganded CTD-up closed state towards PIP2 bound open state. Note the unliganded, CTD-up closed state shown to account for functional and kinetic modeling data in the literature is likely short-lived and its structure is yet to be captured by cryoEM.
The Kir6.2-CTD-up and CTD-down conformations observed in our structures are similar to the T and R states observed in a SUR1–39aa-Kir6.2 fusion channel alternatively bound to GBC+ ATPγS, ATPγS alone, or MgADP by Wu et al. (Wu et al., 2018). However, the percentage of particles in the T-state (corresponding to our Kir6.2-CTD-up state) in their ATPγS+GBC or ATPγS datasets is ~40% and 43% respectively, which differ significantly from the ~93% and 22% in our GBC+ATP and ATP alone datasets. Compared to our CTD-up state in the ATP condition, the higher percentage of T-state particles in the fusion channel’s ATPγS condition could potentially derive from the 10-fold higher concentrations of ATPγS used to generate the Wu et al. structure. An explanation for the markedly higher percentage of Kir6.2-CTD-up state particles in our GBC+ATP condition, over T-state particles in their GBC+ATPγS condition is not obvious. In principle, however, the extra 39aa linker between SUR1 C-terminus and Kir6.2 N-terminus in the fusion construct could uncouple drug binding from Kir6.2-CTD conformation. Consistent with this, GBC was shown to be ineffective in inhibiting the fusion channel in contrast to the WT channel formed by separate SUR1 and Kir6.2 proteins (Wu et al., 2018).
Previous KATP channel cryoEM studies have provided evidence that KNtp interacts with the central cavity of the SUR ABC core (Ding et al., 2019; Martin et al., 2019; Sung et al., 2021; Wu et al., 2018). The structures presented here refines our view of the molecular interactions between the KNtp and different parts of SUR1. We have previously shown that engineered crosslinking between Kir6.2-L2C and SUR1-C1142 reduces channel activity (Martin et al, 2019). A likely scenario is that stapling KNtp along SUR1 via the contact sites we observe stabilizes the inhibited Kir6.2-CTD-up conformation and prevents further rotation of the CTD needed to open the channel. This interpretation can explain why deletion of KNtp increases channel open probability (Babenko et al., 1999; Koster et al., 1999; Reimann et al., 1999), while drugs such as GBC, RPG and CBZ, which stabilize KNtp in the transmembrane cavity of the SUR1 ABC module, mimic the physiological inhibitor ATP and block channel activity (Devaraneni et al., 2015).
Comparison to other Kir channels
Differential proximity of the CTD to the membrane has also been reported in Kir2 (Hansen et al., 2011; Lee et al., 2016b; Zangerl-Plessl et al., 2020) and Kir3 (Niu et al., 2020), suggesting there may be a common theme in Kir channel conformation and gating transitions. Supporting this, mutations which disrupt cytoplasmic domain subunit interface in Kir2.1 also reduce channel activity, akin to the inactivation mutations reported in Kir6.2 (Borschel et al., 2017). Distinctively however, unlike Kir2 and Kir3 channels, KATP channels have an additional ATP-bound Kir6.2-CTD-up conformation stop between the CTD-down inactivated and CTD-up open conformations, which allows for a rapid and reversible inhibition of the channel in response to metabolic signals.
Another unique feature of Kir6.2 channels is the requirement of SUR1 co-assembly to achieve the high ATP sensitivity and open probability of native KATP channels (Inagaki et al., 1995; Tucker et al., 1997). Several studies have now provided functional, biochemical and structural evidence that SUR1 directly participates in ATP binding via K205 in the L0 linker (Ding et al., 2019; Pratt et al., 2012; Usher et al., 2020). SUR1 increases the open probability of Kir6.2 by more than 10-fold, an effect that is largely mediated by TMD0 (Babenko and Bryan, 2003; Chan et al., 2003). We show in our structure that K134 in SUR1-ICL2 is oriented towards the lipid headgroup density in the PIP2 binding pocket, suggesting SUR1-TMD0 increases channel open probability by directly contributing to binding of PIP2 or other phospholipids. Supporting this, MD simulations show K134-PIP2 interactions (Fig.6D, Fig.S6D) and functional experiments show that mutation of SUR1 K134 to alanine reduces channel Po (Fig.5E, F).
Structural basis of ATP and PIP2 antagonism
ATP and PIP2 functionally compete to inhibit and activate KATP channels, respectively (Baukrowitz et al., 1998; Shyng and Nichols, 1998). Molecular dynamics (MD) simulations reveal that two Kir6.2 residues, K39 and R54, interact with both ATP and PIP2, providing evidence that competition for binding residues between the two ligands underlies, at least in part, functional competition between ATP and PIP2. Interestingly, in one of the ATP+PIP2 simulation runs, ATP dissociates from its binding pocket (Fig.6B). This dissociation event is likely captured because SUR1-L0, which contains the ATP stabilizing residue K205 is not included in the simulation structure. It offers a view of how ATP may dissociate such that binding residues shared between ATP and PIP2 have greater freedom to interact with PIP2, favoring channel opening. The rotational movement of SUR1 towards the Kir6.2-tetramer observed in our multibody refinement analysis (Fig.S4) increases the distance between SUR1-K205 and bound ATP, which may initiate ATP dissociation by weakening ATP binding and thus provide a pathway for channel transition from ATP bound inhibited state to PIP2 bound open state.
In summary, the structural analysis, MD simulations, and functional studies presented here together with the recent open KATP channel structure reported by others (Zhao and MacKinnon, 2021) offer insight into several longstanding questions on KATP channel gating mechanisms. A principal unresolved question regards the full extent of the conformational dynamics of the SUR1 subunit and how it may relate to channel function. The large rotation of the SUR ABC module that leads to a quatrefoil channel conformation has previously been reported in a human SUR1–6aa-Kir6.2 fusion protein channel in which the NBDs are bound to MgATP/MgADP and dimerized (Lee et al., 2017). Recently, a similar large rotation is reported in a SUR2B/Kir6.1 vascular KATP channel bound to GBC and ATP (Sung et al., 2021). However, the quatrefoil conformation is not observed in the most recent human SUR1 NBDs dimerized-Kir6.2 mutant open channel (Zhao and MacKinnon, 2021). Whether the variable findings stem from protein preparation methods or involve differences in data processing will be important to resolve in order to fully understand KATP channel structure and function relationship.
Methods
Image processing and particle classification
CryoEM images of pancreatic KATP channels (co-assembled from hamster SUR1 and rat Kir6.2) collected in different liganded conditions in our previous publication (Martin et al., 2019) were reprocessed from the initial 2D classification step that we described previously using RELION-3.1 (Zivanov et al., 2018). Classes displaying fully and partially assembled complexes with high signal/noise were selected. The particles were re-extracted at 1.045 Å/pix for RPG/ATP, 1.399 Å/pix for GBC/ATP, 1.72 Å/pix for CBZ/ATP and ATP only, and 1.826 Å/pix for apo state datasets, and then used as input for 3D classification in RELION-3.1. Fig.S1 shows the data processing workflow for the RPG/ATP dataset. Channel particles refined in the final C4 reconstruction (150,707 total particles) were subjected to C4 symmetry expansion and yielded 4 fold more copies. Further refinement was performed without symmetry restraints or masking such that possible heterogeneous particles can be aligned without any restraints. 2D class averages from all data sets showed significant heterogeneity of the SUR1-ABC module, indicating dynamic SUR1 motions were captured during vitrification of the cryoEM samples. To probe potential novel conformations due to dynamics of SUR1-ABC module relative to the Kir6.2 tetramer, or for novel conformations that arise due to dynamic motions of other domains of the KATP channel, a soft mask that includes the Kir6.2 tetramer and one SUR1 in a propeller form was created in Chimera using our previously published model (PDB:6BAA) (Martin et al., 2017a), and extensive focused 3D classification was performed without particle alignment. This revealed two major classes with different Kir6.2-CTD conformations that are either anchored up towards the plasma membrane (CTD-up) or extended down further towards the cytoplasm (CTD-down).
Focused refinement of SUR1 was carried out after partial signal subtraction that removed signals outside the masked region, followed by further 3D classification without alignment at higher regularization T values (ranged from 6 to 20) and local refinement of signal subtracted particles (Scheres, 2016). Extensive 3D classification sorted out remaining minor groups of particles that did not align well with the propeller conformation but no other conformations emerged. The dominant class then underwent three iterations of CTF refinement and 3D refinement. To test whether some of the SUR1 particles adopt quatrefoil-like conformation reported previously, a mask that includes the Kir6.2 tetramer and one SUR1 was also created using a quatrefoil-like model of our previously published Kir6.1/SUR2B structure (PDB:7MJO) (Sung et al., 2021), which was used for classification following the same scheme described for the propeller form mask. A minor quatrefoil form at 7.1 Å overall resolution was identified from the RPG/ATP dataset (Fig.S1). Final maps were subjected to Map-modification implemented in Phenix with two independent half maps and corresponding mask and model as input. They were then sharpened with model-based auto sharpening with the corresponding model using Phenix, a step that was iterated during model building.
The same workflow was used to process the other four datasets, GBC/ATP, CBZ/ATP, ATP only and apo states. With the exception of the apo dataset which yielded only the CTD-down conformation, all other datasets showed both Kir6.2-CTD classes similar to those identified in the RPG/ATP dataset but with varying ratios of the two conformations. Particle distributions and final map resolutions for all datasets are summarized in Table S1. Upon carrying out this further analysis, we noted the CBZ/ATP dataset previously reported to be collected in the presence of 10μM CBZ and 1mM ATP did not yield a map with clear ATP density at the Kir6.2 ATP binding site. Upon inspection of the ATP used it was discovered that the concentration had been mistakenly reported as 1mM rather than 0.1mM, which likely explained the lack of clear ATP cryoEM density.
Multibody refinement
For Kir6.2 tetramer multibody refinement, particles pooled from both the RPG/ATP CTD-up and CTD-down classes and also each class separately were used (Nakane et al., 2018). To interrogate the Kir6.2 CTD movements relative to its TM domain, we masked out the SUR1 density and assigned the Kir6.2 TM domain (58–173) and CTD (174–352) as two separate rigid bodies (Fig.S3A). Principal component analysis showed that a dominant eigenvector accounted for 25.3% of the overall variance (Fig.S3B). Histograms of the amplitudes along this eigenvector shows a bimodal distribution with two peaks, indicating two conformationally distinct populations differing in the distance between the CTD and the TM domain and rotation of the CTD as expected. Rotation and rocking motions of the CTD were also observed within the Kir6.2-CTD-up and Kir6.2-CTD-down class particles. Although similar in degrees of motion, greater heterogeneity was seen in the CTD-down population of particles than in the CTD-up (Fig.S3C, D), consistent with increased mobility of the CTD when extended away from the membrane.
For (Kir6.2)4-SUR1 multibody refinement, the map was divided into three bodies: body 1, Kir6.2-tetramer (E30-D352); body 2, ABC-core (Q211-V1578) of SUR1 plus KNtp (M1-E19) of Kir6.2; and body 3, TMD0 (M1-L210) of SUR1 (Fig.S4A). Multibody refinement was repeated with varying standard deviations for the degrees of rotation, and pixels of translation, to rule out artifacts. Principal component analysis in the relion_flex_analyse program revealed that approximately 17.5% of the variance is explained by the dominant eigenvector 1 (Fig.S4B,C) corresponding to horizontal swinging motion of SUR1 (Nakane et al., 2018). We then used the program to generate two separate STAR files, each containing ~35,000 particles with eigenvalues less than −5 or greater than +5 along eigenvector 1 (Fig.S4D). These two sets of particles were further refined using a soft mask, yielding two maps which we refer to as SUR1-out and SUR1-in with an overall resolution of 3.8 and 3.9 Å, respectively.
Model building and refinement
The RPG/ATP dataset yielded the highest resolution maps and were used for modeling (Table S2). Initial models for the Kir6.24-SUR1 channel were obtained by docking Kir6.2-TMD (32–171) and Kir6.2-CTD (172–352) from our previously published model (PDB:6BAA) (Martin et al., 2017b), and TMD0/L0 (1–284), TMD1 (285–614), NBD1 (615–928), NBD1-TMD2-linker (992–999), TMD2 (1000–1319) and NBD2 (1320–1582) of SUR1 (PDB:6PZA) (Martin et al., 2019), into either the RPG-CTD-up or the RPG-CTD-down cryoEM density map by rigid-body fitting using Chimera’s ‘Fit in’ tool (Pettersen et al., 2004). Then different domains were combined using Chimera to form a composite model for further refinement. We used Coot to manually build and edit residues 32–78 and residues 79–361 at the interface of two Kir6.2 subunits (Emsley et al., 2010), we then copied those changes to each of the other four Kir6.2 subunits. Further edits and refinements were done independently to each chain without strict NCS restraints. The models were then iteratively built and refined in Coot (Emsley et al., 2010) and Phenix (Afonine et al., 2018), with Ramachandran restraints, secondary structure restraints, and side-chain rotamer restraints. The N-terminus of Kir6.2 (residues 1–31, KNtp) had sufficient continuous density and the density was sufficiently clear to allow modeling of several key interactions with SUR1, although accurate modeling of side chains was not possible for the entire KNtp. Similarly, the NBDs of SUR1 and particularly NBD2 had weaker density than most of the reconstructed map, imposing reliance on restraints and prior models. In addition to modeling the protein, ATP was modeled at the inhibitory ATP-binding site on each of the four Kir6.2 subunits, and at the nucleotide binding site in NBD1 of SUR1 which had sufficient cryoEM density. Phosphatidylserine, phosphatidylcholine, and phosphatidylethanolamine were modeled liberally into plausible lipid density; 17 lipids were modeled for the RPG-CTD-up model and 13 lipids modeled for the RPG-CTD-down model.
MD simulations and analysis
All MD simulations were performed at all-atom resolution using AMBER 16 (Case et al., 2016) with GPU acceleration. Initial coordinates were developed from the RPG/ATP-CTD-up model including four Kir6.2 (32–352) and four SUR1-TMD0 (1–193) without the SUR1-ABC core. ATP in the cryoEM structure was removed for simulations in the apo condition. For simulations in the presence of ATP and PIP2, the ATP from the cryoEM structure was kept, and a PIP2 molecule (DMPI24, di-myristoyl-inositol-(4,5)-bisphosphate) was docked in the PIP2 binding pocket using Kir3.2-PIP2 structure (PDB ID: 6M84) as a template.
The simulation starting structures were protonated by the H++ webserver (http://biophysics.cs.vt.edu/H++) at pH 7 and inserted in a bilayer membrane composed of 1-palmitoyl-2-oleoyl-phosphatidylcholine (POPC) lipids and surrounded by an aqueous solution of 0.15 M KCl. The optimal protein orientations in the membrane were obtained from the OPM database (Lomize et al., 2012). All systems contain 650–680 POPC lipids and ~87,000 water molecules, resulting in a total of ~385,000 atoms. They were assembled using the CHARMM-GUI webserver (Jo et al., 2008; Lee et al., 2016a; Wu et al., 2014), which also generated all simulation input files.
The CHARMM36m protein (Huang et al., 2017) and CHARMM36 lipid (Klauda et al., 2010; Pastor and Mackerell, 2011) force field parameters were used with the TIP3P water model (Jorgensen, 1983). Langevin dynamics (Pastor, 1988) were applied to control the temperature at 300K with a damping coefficient of 1/ps. Van der Waals (vdW) interactions were truncated via a force-based switching function with a switching distance of 10 Å and a cutoff distance of 12 Å. Short-range Coulomb interactions were cut off at 12 Å, long-range electrostatic interactions were calculated by the Particle-Mesh Ewald summation (Darden et al., 1993; Essmann et al., 1995). Bonds to hydrogen atoms were constrained using the SHAKE algorithm (Jean-Paul Ryckaert, 1977).
The atomic coordinates were first minimized for 5000 steps using the steepest-descent and conjugated gradient algorithms, followed by a ~2 ns equilibration simulation phase, during which dihedral restraints on lipid and protein heavy atoms were gradually removed from 250 to 0 kcal/mol/Å2, the simulation time step was increased from 1 fs to 2 fs, and the simulation ensemble was switched from NVT to NPT. To keep the pressure at 1 bar, a semi-isotropic pressure coupling was applied that allows the z-axis to expand and contract independently from the x-y plane (Martyna, 1994). The simulations were then run for over 1 μs with a time step of 4 fs enabled by hydrogen mass repartitioning (Balusek et al., 2019; Hopkins et al., 2015).
Analysis of ATP/PIP2 occupancy
ATP and PIP2 residue occupancies (Fig. 6D) were computed using the histogram of minimum hydrogen bond lengths between each residue and PIP2/ATP, to show the amount of time spent at different distances. Summarized residue occupancies (Fig. S6D) were calculated as the fraction of time each residue spent in contact with ATP/PIP2, where contact is defined as a minimum hydrogen bond length of below 4 Å. For both ligands, minimum bond lengths were used regardless of which pairs formed the bond. A 10 ns window average was used to smooth the minimum bond length time series data.
Functional studies
Point mutation SUR1-K134A was introduced into hamster SUR1 cDNA in pECE using the QuikChange site-directed mutagenesis kit (Stratagene). Mutation was confirmed by DNA sequencing. For electrophysiology, wild-type or mutant SUR1 cDNA and rat Kir6.2 in pcDNA1 along with cDNA for green fluorescent protein GFP (to facilitate identification of transfected cells) were co-transfected into COS cells using FuGENE®6, and plated onto glass coverslips 24 hours after transfection for recording, as described previously (Martin et al., 2019). Recording pipettes were pulled from non-heparinized Kimble glass (Fisher Scientific) on a horizontal puller (Sutter Instrument, Co., Novato, CA, USA). Electrode resistance was typically 1–2 MΩ when filled with K-INT solution containing 140 mM KCl, 10 mM K-HEPES, 1 mM K-EGTA, pH 7.3. ATP was added as the potassium salt. PI4,5P2 (Avanti Polar Lipids) was reconstituted in K-INT solution at 5 μM and bath sonicated in ice water for 20 min before use. Inside-out patches of cells bathed in K-INT were voltage-clamped with an Axopatch 1D amplifier (Axon Inc., Foster City, CA). Exposure of membrane patches to ATP- or PIP2-containing K-INT bath solution was as specified in Fig.5 legend. All currents were measured at room temperature at a membrane potential of −50 mV (pipette voltage = +50 mV) and inward currents shown as upward deflections. Data were analyzed using pCLAMP10 software (Axon Instrument). Off-line analysis was performed using Microsoft Excel programs. Data were presented as mean ± standard error of the mean (S.E.M) and statistical analysis was performed using two-tailed student’s t-test, with p<0.05 considered statistically significant.
Accession numbers
Coordinates and cryoEM density maps for KATP channel models of the Kir6.2 tetramer core plus one SUR1 subunit have been deposited to the Protein Data Bank and the Electron Microscopy Data Bank with accession numbers as follows: RPG/ATP Kir6.2-CTD-up (PDB code 7TYS, EMDB code EMD-26193); RPG/ATP Kir6.2-CTD-down (PDB code 7TYT, EMDB code EMD-26194); RPG/ATP Kir6.2-CTD-up SUR1-in (PDB code 7U1Q, EMDB code EMD-26303); RPG/ATP Kir6.2-CTD-up SUR1-out (PDB code 7U1S, EMDB code EMD-26304); GBC/ATP Kir6.2-CTD-up (PDB code 7U24, EMDB code EMD-26307); GBC/ATP Kir6.2-CTD-down (PDB code 7U6Y, EMDB code EMD-26308); CBZ/ATP Kir6.2-CTD-up (PDB code 7U7M, EMDB code EMD-26309); CBZ/ATP Kir6.2-CTD-down (PDB code 7U2X, EMDB code EMD-26321); ATP-only Kir6.2-CTD-up (PDB code 7UAA, EMDB EMD-26312); ATP-only Kir6.2-CTD-down (PDB 7U1E, EMDB code EMD-26299); Apo Kir6.2-CTD-down (PDB code 7UQR, EMDB code EMD-26320).
Supplementary Material
Highlights.
The activity of KATP channel is governed by the interplay between its ligands and its subunits, Kir6.2 and SUR1.
Inhibitory ligands bias Kir6.2-cytoplasmic domain conformation towards one close to the membrane instead of one extended away into the cytoplasm via a corkscrew motion.
SUR1 cooperates with Kir6.2 to stabilize inhibitor ATP and activator PIP2 binding thereby enhancing channel sensitivity to both.
ATP and PIP2 compete with each other to close or open the channel, respectively, by having overlapping but non-identical binding residues.
Ligand-dependent conformational dynamics of the Kir6.2-cytoplasmic domain offer insights into KATP channel gating mechanisms.
Acknowledgements
We thank Zhongying Yang for technical support and Assmaa ElSheikh for helpful discussions. The originial cryoEM datasets used for analaysis in the current manuscript were collected at the Multi-Scale Microscopy Core at the Oregon Health and Science University. The project was supported by the National Institutes of Health grant R01DK066485 (to S.-L. S.) and the National Science Foundation grant MCB 2119837 (to D.M.Z.).
Footnotes
Competing interests
The authors declare that they have no competing financial or non-financial interests with the contents of this article.
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
References
- Afonine PV, Poon BK, Read RJ, Sobolev OV, Terwilliger TC, Urzhumtsev A, and Adams PD. 2018. Real-space refinement in PHENIX for cryo-EM and crystallography. Acta Crystallogr D Struct Biol. 74:531–544. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Alkorta-Aranburu G, Carmody D, Cheng YW, Nelakuditi V, Ma L, Dickens JT, Das S, Greeley SAW, and Del Gaudio D. 2014. Phenotypic heterogeneity in monogenic diabetes: the clinical and diagnostic utility of a gene panel-based next-generation sequencing approach. Molecular genetics and metabolism. 113:315–320. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Arya VB, Guemes M, Nessa A, Alam S, Shah P, Gilbert C, Senniappan S, Flanagan SE, Ellard S, and Hussain K. 2014. Clinical and histological heterogeneity of congenital hyperinsulinism due to paternally inherited heterozygous ABCC8/KCNJ11 mutations. Eur J Endocrinol. 171:685–695. [DOI] [PubMed] [Google Scholar]
- Ashcroft FM 2005. ATP-sensitive potassium channelopathies: focus on insulin secretion. The Journal of clinical investigation. 115:2047–2058. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Babenko AP, and Bryan J. 2003. Sur domains that associate with and gate KATP pores define a novel gatekeeper. The Journal of biological chemistry. 278:41577–41580. [DOI] [PubMed] [Google Scholar]
- Babenko AP, Gonzalez G, and Bryan J. 1999. The N-terminus of KIR6.2 limits spontaneous bursting and modulates the ATP-inhibition of KATP channels. Biochemical and biophysical research communications. 255:231–238. [DOI] [PubMed] [Google Scholar]
- Balusek C, Hwang H, Lau CH, Lundquist K, Hazel A, Pavlova A, Lynch DL, Reggio PH, Wang Y, and Gumbart JC. 2019. Accelerating Membrane Simulations with Hydrogen Mass Repartitioning. Journal of chemical theory and computation. 15:4673–4686. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Baukrowitz T, Schulte U, Oliver D, Herlitze S, Krauter T, Tucker SJ, Ruppersberg JP, and Fakler B. 1998. PIP2 and PIP as determinants for ATP inhibition of KATP channels. Science. 282:1141–1144. [DOI] [PubMed] [Google Scholar]
- Boodhansingh KE, Kandasamy B, Mitteer L, Givler S, De Leon DD, Shyng SL, Ganguly A, and Stanley CA. 2019. Novel dominant KATP channel mutations in infants with congenital hyperinsulinism: Validation by in vitro expression studies and in vivo carrier phenotyping. Am J Med Genet A. 179:2214–2227. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Borschel WF, Wang S, Lee S, and Nichols CG. 2017. Control of Kir channel gating by cytoplasmic domain interface interactions. The Journal of general physiology. 149:561–576. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brennan S, Rubaiy HN, Imanzadeh S, Reid R, Lodwick D, Norman RI, and Rainbow RD. 2020. Kir6.2-D323 and SUR2A-Q1336: an intersubunit interaction pairing for allosteric information transfer in the KATP channel complex. The Biochemical journal. 477:671–689. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brundl M, Pellikan S, and Stary-Weinzinger A. 2021. Simulating PIP2-Induced Gating Transitions in Kir6.2 Channels. Front Mol Biosci. 8:711975. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Case DA, Betz RM, Cerutti DS, Cheatham TE III, Darden TA, Duke RE, Giese TJ, Gohlke H,, Goetz AW, Homeyer N, Izadi S, Janowski P, Kaus J, Kovalenko A, Lee TS, LeGrand S, Li P, , Lin C, Luchko T, Luo R, Madej B, Mermelstein D, Merz KM, Monard G, Nguyen H, Nguyen HT,, Omelyan I, Onufriev A, Roe DR, Roitberg A, Sagui C, Simmerling CL, Botello-Smith WM, Swails J, , and Walker RC, Wang J, Wolf RM, Wu X, Xiao L, and Kollman PA. 2016. AMBER 2016. [Google Scholar]
- Chan KW, Zhang H, and Logothetis DE. 2003. N-terminal transmembrane domain of the SUR controls trafficking and gating of Kir6 channel subunits. The EMBO journal. 22:3833–3843. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Clement J.P.t., Kunjilwar K, Gonzalez G, Schwanstecher M, Panten U, Aguilar-Bryan L, and Bryan J. 1997. Association and stoichiometry of K(ATP) channel subunits. Neuron. 18:827–838. [DOI] [PubMed] [Google Scholar]
- Cukras CA, Jeliazkova I, and Nichols CG. 2002. The role of NH2-terminal positive charges in the activity of inward rectifier KATP channels. The Journal of general physiology. 120:437–446. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Darden T, York D, and Pedersen L. 1993. Particle mesh Ewald: An N·log(N) method for Ewald sums in large systems. The Journal of chemical physics. 98:10089–10092. [Google Scholar]
- De Franco E, Saint-Martin C, Brusgaard K, Knight Johnson AE, Aguilar-Bryan L, Bowman P, Arnoux JB, Larsen AR, Sanyoura M, Greeley SAW, Calzada-Leon R, Harman B, Houghton JAL, Nishimura-Meguro E, Laver TW, Ellard S, Del Gaudio D, Christesen HT, Bellanne-Chantelot C, and Flanagan SE. 2020. Update of variants identified in the pancreatic beta-cell KATP channel genes KCNJ11 and ABCC8 in individuals with congenital hyperinsulinism and diabetes. Human mutation. 41:884–905. [DOI] [PMC free article] [PubMed] [Google Scholar]
- de Wet H, Proks P, Lafond M, Aittoniemi J, Sansom MS, Flanagan SE, Pearson ER, Hattersley AT, and Ashcroft FM. 2008. A mutation (R826W) in nucleotide-binding domain 1 of ABCC8 reduces ATPase activity and causes transient neonatal diabetes. EMBO reports. 9:648–654. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Devaraneni PK, Martin GM, Olson EM, Zhou Q, and Shyng SL. 2015. Structurally distinct ligands rescue biogenesis defects of the KATP channel complex via a converging mechanism. The Journal of biological chemistry. 290:7980–7991. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ding D, Wang M, Wu JX, Kang Y, and Chen L. 2019. The Structural Basis for the Binding of Repaglinide to the Pancreatic KATP Channel. Cell reports. 27:1848–1857 e1844. [DOI] [PubMed] [Google Scholar]
- Drain P, Li L, and Wang J. 1998. KATP channel inhibition by ATP requires distinct functional domains of the cytoplasmic C terminus of the pore-forming subunit. Proceedings of the National Academy of Sciences of the United States of America. 95:13953–13958. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Emsley P, Lohkamp B, Scott WG, and Cowtan K. 2010. Features and development of Coot. Acta crystallographica. Section D, Biological crystallography. 66:486–501. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Enkvetchakul D, Loussouarn G, Makhina E, Shyng SL, and Nichols CG. 2000. The kinetic and physical basis of K(ATP) channel gating: toward a unified molecular understanding. Biophysical journal. 78:2334–2348. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Essmann U, Perera L, Berkowitz ML, Darden T, Lee H, and Pedersen LG. 1995. A smooth particle mesh Ewald method. The Journal of chemical physics. 103:8577–8593. [Google Scholar]
- Gribble FM, and Reimann F. 2003. Sulphonylurea action revisited: the post-cloning era. Diabetologia. 46:875–891. [DOI] [PubMed] [Google Scholar]
- Hansen SB, Tao X, and MacKinnon R. 2011. Structural basis of PIP2 activation of the classical inward rectifier K+ channel Kir2.2. Nature. 477:495–498. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hopkins CW, Le Grand S, Walker RC, and Roitberg AE. 2015. Long-Time-Step Molecular Dynamics through Hydrogen Mass Repartitioning. Journal of chemical theory and computation. 11:1864–1874. [DOI] [PubMed] [Google Scholar]
- Huang J, Rauscher S, Nawrocki G, Ran T, Feig M, de Groot BL, Grubmuller H, and MacKerell AD Jr. 2017. CHARMM36m: an improved force field for folded and intrinsically disordered proteins. Nature methods. 14:71–73. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Inagaki N, Gonoi T, Clement J.P.t., Namba N, Inazawa J, Gonzalez G, Aguilar-Bryan L, Seino S, and Bryan J. 1995. Reconstitution of IKATP: an inward rectifier subunit plus the sulfonylurea receptor. Science. 270:1166–1170. [DOI] [PubMed] [Google Scholar]
- Inagaki N, Gonoi T, and Seino S. 1997. Subunit stoichiometry of the pancreatic beta-cell ATP-sensitive K+ channel. FEBS letters. 409:232–236. [DOI] [PubMed] [Google Scholar]
- Jean-Paul Ryckaert GC, Berendsen Herman J.C. 1977. Numerical integration of the cartesian equations of motion of a system with constraints: molecular dynamics of n-alkanes. Journal of Computational Physics. 23:327–341. [Google Scholar]
- Jo S, Kim T, Iyer VG, and Im W. 2008. CHARMM-GUI: a web-based graphical user interface for CHARMM. Journal of computational chemistry. 29:1859–1865. [DOI] [PubMed] [Google Scholar]
- Jorgensen WL, Chandrasekhar J, and Madura JD 1983. Comparison of simple potential functions for simulating liquid water. The Journal of Chemical Physics. 79:926–935. [Google Scholar]
- Klauda JB, Venable RM, Freites JA, O’Connor JW, Tobias DJ, Mondragon-Ramirez C, Vorobyov I, MacKerell AD Jr., and Pastor RW. 2010. Update of the CHARMM all-atom additive force field for lipids: validation on six lipid types. The journal of physical chemistry. B. 114:7830–7843. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Koster JC, Sha Q, Shyng S, and Nichols CG. 1999. ATP inhibition of KATP channels: control of nucleotide sensitivity by the N-terminal domain of the Kir6.2 subunit. The Journal of physiology. 515 ( Pt 1):19–30. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee J, Cheng X, Swails JM, Yeom MS, Eastman PK, Lemkul JA, Wei S, Buckner J, Jeong JC, Qi Y, Jo S, Pande VS, Case DA, Brooks CL 3rd, MacKerell AD Jr., Klauda JB, and Im W. 2016a. CHARMM-GUI Input Generator for NAMD, GROMACS, AMBER, OpenMM, and CHARMM/OpenMM Simulations Using the CHARMM36 Additive Force Field. Journal of chemical theory and computation. 12:405–413. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee KPK, Chen J, and MacKinnon R. 2017. Molecular structure of human KATP in complex with ATP and ADP. eLife. 6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee SJ, Ren F, Zangerl-Plessl EM, Heyman S, Stary-Weinzinger A, Yuan P, and Nichols CG. 2016b. Structural basis of control of inward rectifier Kir2 channel gating by bulk anionic phospholipids. The Journal of general physiology. 148:227–237. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li JB, Huang X, Zhang RS, Kim RY, Yang R, and Kurata HT. 2013. Decomposition of slide helix contributions to ATP-dependent inhibition of Kir6.2 channels. The Journal of biological chemistry. 288:23038–23049. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li N, Wu JX, Ding D, Cheng J, Gao N, and Chen L. 2017. Structure of a Pancreatic ATP-Sensitive Potassium Channel. Cell. 168:101–110 e110. [DOI] [PubMed] [Google Scholar]
- Lin CW, Yan F, Shimamura S, Barg S, and Shyng SL. 2005. Membrane phosphoinositides control insulin secretion through their effects on ATP-sensitive K+ channel activity. Diabetes. 54:2852–2858. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lin YW, Bushman JD, Yan FF, Haidar S, MacMullen C, Ganguly A, Stanley CA, and Shyng SL. 2008. Destabilization of ATP-sensitive potassium channel activity by novel KCNJ11 mutations identified in congenital hyperinsulinism. The Journal of biological chemistry. 283:9146–9156. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lin YW, Jia T, Weinsoft AM, and Shyng SL. 2003. Stabilization of the activity of ATP-sensitive potassium channels by ion pairs formed between adjacent Kir6.2 subunits. The Journal of general physiology. 122:225–237. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lomize MA, Pogozheva ID, Joo H, Mosberg HI, and Lomize AL. 2012. OPM database and PPM web server: resources for positioning of proteins in membranes. Nucleic acids research. 40:D370–376. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Martin GM, Kandasamy B, DiMaio F, Yoshioka C, and Shyng SL. 2017a. Anti-diabetic drug binding site in a mammalian KATP channel revealed by Cryo-EM. eLife. 6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Martin GM, Sung MW, Yang Z, Innes LM, Kandasamy B, David LL, Yoshioka C, and Shyng SL. 2019. Mechanism of pharmacochaperoning in a mammalian KATP channel revealed by cryo-EM. eLife. 8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Martin GM, Yoshioka C, Rex EA, Fay JF, Xie Q, Whorton MR, Chen JZ, and Shyng SL. 2017b. Cryo-EM structure of the ATP-sensitive potassium channel illuminates mechanisms of assembly and gating. eLife. 6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Martyna GJ 1994. Constant pressure molecular dynamics algorithms. The Journal of Chemical Physics. 101:4177–4189. [Google Scholar]
- Nakane T, Kimanius D, Lindahl E, and Scheres SH. 2018. Characterisation of molecular motions in cryo-EM single-particle data by multi-body refinement in RELION. eLife. 7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nichols CG 2006. KATP channels as molecular sensors of cellular metabolism. Nature. 440:470–476. [DOI] [PubMed] [Google Scholar]
- Nichols CG, Shyng SL, Nestorowicz A, Glaser B, Clement J.P.t., Gonzalez G, Aguilar-Bryan L, Permutt MA, and Bryan J. 1996. Adenosine diphosphate as an intracellular regulator of insulin secretion. Science. 272:1785–1787. [DOI] [PubMed] [Google Scholar]
- Niu Y, Tao X, Touhara KK, and MacKinnon R. 2020. Cryo-EM analysis of PIP2 regulation in mammalian GIRK channels. eLife. 9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Niu Y, Tao X, Vaisey G, Olinares PDB, Alwaseem H, Chait BT, and MacKinnon R. 2021. Analysis of the mechanosensor channel functionality of TACAN. Elife. 10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nogales E, and Scheres SH. 2015. Cryo-EM: A Unique Tool for the Visualization of Macromolecular Complexity. Mol Cell. 58:677–689. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pastor RW, Brooks BR, Szabo A 1988. An analysis of the accuracy of Langevin and molecular dynamics algorithms. Molecular Physics. 65:1409–1419. [Google Scholar]
- Pastor RW, and Mackerell AD Jr. 2011. Development of the CHARMM Force Field for Lipids. J Phys Chem Lett. 2:1526–1532. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pettersen EF, Goddard TD, Huang CC, Couch GS, Greenblatt DM, Meng EC, and Ferrin TE. 2004. UCSF Chimera--a visualization system for exploratory research and analysis. Journal of computational chemistry. 25:1605–1612. [DOI] [PubMed] [Google Scholar]
- Pinney SE, MacMullen C, Becker S, Lin YW, Hanna C, Thornton P, Ganguly A, Shyng SL, and Stanley CA. 2008. Clinical characteristics and biochemical mechanisms of congenital hyperinsulinism associated with dominant KATP channel mutations. The Journal of clinical investigation. 118:2877–2886. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pratt EB, Tewson P, Bruederle CE, Skach WR, and Shyng SL. 2011. N-terminal transmembrane domain of SUR1 controls gating of Kir6.2 by modulating channel sensitivity to PIP2. The Journal of general physiology. 137:299–314. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pratt EB, Zhou Q, Gay JW, and Shyng SL. 2012. Engineered interaction between SUR1 and Kir6.2 that enhances ATP sensitivity in KATP channels. The Journal of general physiology. 140:175–187. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Puljung MC 2018. Cryo-electron microscopy structures and progress toward a dynamic understanding of KATP channels. The Journal of general physiology. 150:653–669. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Punjani A, and Fleet DJ. 2021. 3D variability analysis: Resolving continuous flexibility and discrete heterogeneity from single particle cryo-EM. Journal of structural biology. 213:107702. [DOI] [PubMed] [Google Scholar]
- Reimann F, Tucker SJ, Proks P, and Ashcroft FM. 1999. Involvement of the n-terminus of Kir6.2 in coupling to the sulphonylurea receptor. The Journal of physiology. 518 ( Pt 2):325–336. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Scheres SH 2016. Processing of Structurally Heterogeneous Cryo-EM Data in RELION. Methods in enzymology. 579:125–157. [DOI] [PubMed] [Google Scholar]
- Shyng S, and Nichols CG. 1997. Octameric stoichiometry of the KATP channel complex. The Journal of general physiology. 110:655–664. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shyng SL, Cukras CA, Harwood J, and Nichols CG. 2000. Structural determinants of PIP(2) regulation of inward rectifier K(ATP) channels. The Journal of general physiology. 116:599–608. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shyng SL, and Nichols CG. 1998. Membrane phospholipid control of nucleotide sensitivity of KATP channels. Science. 282:1138–1141. [DOI] [PubMed] [Google Scholar]
- Sung MW, Yang Z, Driggers CM, Patton BL, Mostofian B, Russo JD, Zuckerman DM, and Shyng SL. 2021. Vascular KATP channel structural dynamics reveal regulatory mechanism by Mg-nucleotides. Proceedings of the National Academy of Sciences of the United States of America. 118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Suzuki S, Makita Y, Mukai T, Matsuo K, Ueda O, and Fujieda K. 2007. Molecular basis of neonatal diabetes in Japanese patients. The Journal of clinical endocrinology and metabolism. 92:3979–3985. [DOI] [PubMed] [Google Scholar]
- Trapp S, Haider S, Jones P, Sansom MS, and Ashcroft FM. 2003. Identification of residues contributing to the ATP binding site of Kir6.2. EMBO J. 22:2903–2912. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tucker SJ, Gribble FM, Proks P, Trapp S, Ryder TJ, Haug T, Reimann F, and Ashcroft FM. 1998. Molecular determinants of KATP channel inhibition by ATP. The EMBO journal. 17:3290–3296. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tucker SJ, Gribble FM, Zhao C, Trapp S, and Ashcroft FM. 1997. Truncation of Kir6.2 produces ATP-sensitive K+ channels in the absence of the sulphonylurea receptor. Nature. 387:179–183. [DOI] [PubMed] [Google Scholar]
- Usher SG, Ashcroft FM, and Puljung MC. 2020. Nucleotide inhibition of the pancreatic ATP-sensitive K+ channel explored with patch-clamp fluorometry. eLife. 9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vieira-Pires RS, and Morais-Cabral JH. 2010. 3(10) helices in channels and other membrane proteins. The Journal of general physiology. 136:585–592. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang M, Wu JX, Ding D, and Chen L. 2022. Structural insights into the mechanism of pancreatic KATP channel regulation by nucleotides. Nature communications. 13:2770. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu EL, Cheng X, Jo S, Rui H, Song KC, Davila-Contreras EM, Qi Y, Lee J, Monje-Galvan V, Venable RM, Klauda JB, and Im W. 2014. CHARMM-GUI Membrane Builder toward realistic biological membrane simulations. Journal of computational chemistry. 35:1997–2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu JX, Ding D, Wang M, Kang Y, Zeng X, and Chen L. 2018. Ligand binding and conformational changes of SUR1 subunit in pancreatic ATP-sensitive potassium channels. Protein & cell. 9:553–567. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zangerl-Plessl EM, Lee SJ, Maksaev G, Bernsteiner H, Ren F, Yuan P, Stary-Weinzinger A, and Nichols CG. 2020. Atomistic basis of opening and conduction in mammalian inward rectifier potassium (Kir2.2) channels. The Journal of general physiology. 152. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang M, Chen X, Shen S, Li T, Chen L, Hu M, Cao L, Cheng R, Zhao Z, and Luo F. 2015. Sulfonylurea in the treatment of neonatal diabetes mellitus children with heterogeneous genetic backgrounds. Journal of pediatric endocrinology & metabolism : JPEM. 28:877–884. [DOI] [PubMed] [Google Scholar]
- Zhao C, and MacKinnon R. 2021. Molecular structure of an open human KATP channel. Proceedings of the National Academy of Sciences of the United States of America. 118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zivanov J, Nakane T, Forsberg BO, Kimanius D, Hagen WJ, Lindahl E, and Scheres SH. 2018. New tools for automated high-resolution cryo-EM structure determination in RELION-3. eLife. 7. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.