Abstract
Endothelin type B receptor (ETBR) plays a crucial role in regulating blood pressure and humoral homeostasis, making it an important therapeutic target for related diseases. ETBR activation by the endogenous peptide hormones endothelin (ET)−1–3 stimulates several signaling pathways, including Gs, Gi/o, Gq/11, G12/13, and β-arrestin. Although the conserved NPxxY motif in transmembrane helix 7 (TM7) is important during GPCR activation, ETBR possesses the lesser known NPxxL motif. In this study, we present the cryo-EM structure of the ETBR–Gi complex, complemented by MD simulations and functional studies. These investigations reveal an unusual movement of TM7 to the intracellular side during ETBR activation and the essential roles of the diverse NPxxL motif in stabilizing the active conformation of ETBR and organizing the assembly of the binding pocket for the α5 helix of Gi protein. These findings enhance our understanding of the interactions between GPCRs and G proteins, thereby advancing the development of therapeutic strategies.
Subject terms: Cryoelectron microscopy, Structural biology
Tani et al. present the cryo-EM structure of the endothelin type B receptor (ETBR) in complex with the Gi protein, emphasizing the movements of transmembrane helix 7 during ETBR activation. The study underscores the crucial role of the NPxxL motif in stabilizing the receptor's active state and organizing the Gi protein binding site, enhancing our understanding of GPCR activation and aiding therapeutic development.
Introduction
The endothelin (ET) family comprises three endogenous isoforms (ET-1–3), each of which contains 21 amino acid residues and two intramolecular disulfide bonds. ET-1, the primary isoform in the human cardiovascular system, is one of the most abundant, potent, and long-lasting constrictors of blood vessels. ET-1 plays a significant role in physiological processes, such as modulation of basal vascular tone, regulation of sodium balance, development of neural crest cells, and cell proliferation, and development of pathophysiological conditions, such as cardiovascular disease, neurological disorders, renal disease, and cancer1–5. The ET family exerts its effects through ET receptors, specifically subtypes ETA and ETB (ETAR and ETBR, respectively), which belong to the β-subfamily of class-A G-protein-coupled receptors (GPCRs). The ET-bound receptors transmit signals through heterotrimeric G proteins with promiscuous coupling properties and also interact with β-arrestins2,6–8.
GPCRs mediate cellular responses to various extracellular molecules, including lipids, nucleosides, neurotransmitters, hormones, and proteins. Ligand binding triggers structural changes in GPCRs, initiating signal transmission. Agonist-mediated GPCR activation is well understood, with specific conserved sequence regions, including C6.47W6.48xP6.50, P5.50I3.40F6.44, N7.49P7.50xxY7.53, and D3.49R3.50Y3.51 motifs (using Ballesteros–Weinstein numbering9 for class-A GPCRs), playing successive roles10–16. Furthermore, three highly conserved residues: R3.50 in DRY, Y5.58, and Y7.53 in NPxxY, play key roles in activating class-A GPCRs11–15. Y7.53 in NPxxY acts as a switch for water rearrangement, in addition to the inward movement of the cytoplasmic end of TM7 during activation17. During this process, N7.49 from NPxxY interacts directly with the conserved D2.50 and Y7.53 interacts with the highly conserved Y5.58 in TM5, either directly or through a bridging water molecule known as the “water lock” in the active state18,19. Because Y5.58 in TM5 undergoes rotation during activation and then stabilizes the orientation of R3.50 through a hydrogen bond, Y7.53 in NPxxY indirectly stabilizes the orientation of R3.50 in DRY. Thus, R3.50, Y5.58, and Y7.53 change their interactions during activation and structurally cooperate to generate the active state of class-A GPCRs. However, some class-A GPCRs have unique motifs such as the NPxxL found in ETBR; how these conserved or divergent motifs contribute to the formation of binding pockets for heterotrimeric G proteins is unclear.
We determined the crystal structures of thermostabilized ETBR in three forms: ET-1-bound, ligand-free, and antagonist bosentan-bound20–22. Although the ET-1-bound ETBR structure detailed the binding of ET-1 to the receptor, it did not explain the activation mechanism, because the intracellular side was fixed in an inactive state by the insertion of T4 lysozyme into ICL3. To better understand ETBR activation by ET-1 and its coupling with G proteins, we report the structure of the ET-1-bound ETBR–Gi complex, determined using cryo-electron microscopy (cryo-EM), and further evaluated with MD simulations and mutagenesis studies. We identified a unique feature—the downward motion of TM7 during activation through a diverse NPxxL motif. This motion stabilized the active conformation of ETBR, leading to the formation of a hydrophobic binding pocket for the C-terminal α5 helix of Gαi.
Results
Overall structure of the ET-1-bound ETBR–Gi complex
To facilitate complex formation, ETBR and Gi1 heterotrimer were expressed separately in Sf9 insect cells and combined after purification in lauryl maltose-neopentyl glycol (LMNG) and cholesteryl hemisuccinate (CHS). ETBR was stabilized by introducing the R124Y1.55 thermostabilizing mutation, which does not reduce ET-1 binding affinity and G-protein coupling ability21. Stabilization of the ETBR–Gi complex was achieved by introducing four dominant negative mutations into the Gαi1 subunit23. In addition, scFv1624 was used to stabilize interactions between the αi1 and β subunits (Supplementary Figs. 1, 2). First, the structure of the ET-1-bound ETBR–wild-type Gi1–scFv16 complex was analyzed by single particle cryo-EM at a global resolution of 4.6 Å (Table 1, Supplementary Figs. 1, 3). To improve resolution, the structure of the complex, including the dominant negative Gαi1 subunit (DNGαi1), was determined with a global resolution of 3.2 Å (Fig. 1, Table 1, Supplementary Figs. 2, 4). Furthermore, we performed focused 3D refinement to obtain receptor densities at a resolution of 3.6 Å. Receptor density was assessed in the ETBR–DNGi1 complex after adjusting the alignment center to the receptor (Table 1, Supplementary Figs. 2, 5). Both ETBR–Gi complex models are nearly identical—their Cα atoms have an RMSD of 0.662 Å (Supplementary Fig. 6a). Compared with the ET-1 bound ETBR model in ETBR–DNGi1, the small RMSD values of the Cα atoms and the similar residue conformations in the other two models indicate they are nearly identical (0.391 Å for ETBR–wild-type Gi1 and 0.364 Å for the focused 3D refinement of ETBR) (Supplementary Fig. 6b, c). The Gi1-bound ETBR structure displayed a typical outward movement of the cytoplasmic side of TM6 to a moderate extent (approximately 7 Å), similar to other class-A Gi-bound GPCRs (Supplementary Fig. 7). We used the ET-1-bound ETBR–DNGi1–scFv16 complex as the ETBR–Gi1 complex and analyzed structural changes in detail.
Table 1.
ET-1 bound ETBR-DNGi-scFv16 (EMDB-38740, PDB-8XWP) | ET-1 bound ETBR-DNGi-scFv16 (focused ETBR) (EMDB-60404, PDB-8ZRT) | ET-1 bound ETBR-wild type Gi-scFv16 (EMDB-38741, PDB-8XWQ) | ||
---|---|---|---|---|
Data collection and processing | ||||
Microscope | JEOL CRYO-ARM300 | TF Talos Arctica | ||
Camera | K3 | Falcon III | ||
Magnification | 60,000 | 100,000 | 92,000 | |
Voltage (kV) | 300 | 300 | 200 | |
Electron exposure (e–/Å2) | 53.3 | 49.2 | 40 | |
Defocus range (μm) | –0.5 to –2.7 | –0.6 to –3.5 | –0.7 to –2.9 | |
Calibrated pixel size (Å) | 0.816 | 0.507 | 1.094 | |
Detector physical pixel size (μm) | 5 | 14 | ||
Symmetry imposed | C1 | C1 | ||
Initial particle images (no.) | 1,193,302 | 1,757,339 | 1,038,215 | 954,972 |
Final particle images (no.) | 556,576 | 481,639 | 401,671 | 278,209 |
Map resolution (Å) | 3.2 | 3.6 | 4.6 | |
FSC threshold | 0.143 | 0.143 | 0.143 | |
Map resolution range (Å) | 5.4–3.1 | 8.7–3.6 | 6.9–4.4 | |
Refinement | ||||
Initial model used (PDB code) | 5GLH, 6OS9 | 8XWP | 5GLH, 6OS9 | |
Model resolution (Å) | 3.2 | 4.1 | 4.6 | |
FSC threshold | 0.5 | 0.5 | 0.5 | |
Model resolution range (Å) | 120–3.2 | 120–3.6 | 126–4.6 | |
Map sharpening B factor (Å2) | –100 | -100 | –263 | |
Model composition | ||||
Non-hydrogen atoms | 9076 | 2322 | 9076 | |
Protein residues | 1163 | 303 | 1163 | |
B factors (Å2) | ||||
Protein | 80.7 | 48.9 | 191.8 | |
R.m.s. deviations | ||||
Bond lengths (Å) | 0.002 | 0.003 | 0.003 | |
Bond angles (°) | 0.473 | 0.693 | 0.546 | |
Validation | ||||
MolProbity score | 2.31 | 1.82 | 2.59 | |
Clashscore | 8.97 | 13.56 | 14.24 | |
Poor rotamers (%) | 5.56 | 0.81 | 7.98 | |
Ramachandran plot | ||||
Favored (%) | 96.06 | 96.93 | 96.33 | |
Allowed (%) | 3.94 | 3.07 | 3.67 | |
Disallowed (%) | 0.00 | 0.00 | 0.00 |
Structure of Gi-stabilized active ETBR
The mode of ET-1 binding in the ETBR–Gi complex closely resembled that of the crystal structure of ET-1-bound ETBR. Y13ET-1 and F14ET-1 in the helical region of ET-1 played a pivotal role in the compact assembly of the N-terminal tail and the extracellular side of TM7, initiating helical rearrangements of ETBR (Fig. 2a). This assembly is essential for full G-protein activation25. The C-terminal region of ET-1 (L17ET-1–W21ET-1) fits into the transmembrane orthosteric pocket of the receptor, interacting with many hydrophobic (I1572.60, L2775.42, L3396.51, etc.) and hydrophilic (including K1823.33, K2735.38, R3436.55, D3687.35, etc.) residues20. The C-terminal side chain of W21ET-1 directly interacted with W3366.48 in the CWxP motif (Fig. 2b). Interactions between ET-1 and ETBR, both in the transmembrane region surrounding the C-terminal region of ET-1 and close to the extracellular side, played a role in ETBR activation. These ligand–receptor interactions influenced the helical rearrangement of ETBR through the conserved V1893.40P2855.50F3326.44 motif, resulting in an inward rotation of R1993.50 and Y2935.58, an outward movement of the cytoplasmic side of TM6 (Fig. 2b, c), and crevice formation on the cytoplasmic side of the receptor to accommodate Gαi.
As observed in other class-A GPCRs, the outward shift of TM6 disrupted the salt bridge between D1983.49 and R1993.50 of DRY, seen in the ET-1-bound inactive ETBR. R1993.50 extended toward TM7, stabilized by Y2935.58 through hydrogen bonding (Fig. 2d, Supplementary Fig. 8a). A simultaneous downward displacement (~ 1.5 Å) at the NPxxL motif (N3827.49, P3837.50, and L3867.53 in ETBR instead of Y7.53) was observed in the ET-1–ETBR–Gi complex (Fig. 2e). N3827.49 extended toward R1993.50, and L3867.53 formed a hydrophobic interaction with I1402.43 to stabilize the helical contacts between TM2 and TM7 (Figs. 2e and 3a). Because the residue at 3867.53 was leucine, and not the conserved tyrosine, the hallmark water-mediated hydrogen bonding network, including Y7.53 and R3.50, which is characteristic of class-A GPCR activation (Supplementary Fig. 8b), was not formed.
Structural comparisons across class-A GPCRs indicated a conserved rearrangement of residue contacts at positions 3.46 and 7.53 upon activation10. In many class-A GPCRs with the conserved Y7.53 sequence in the active state, distances between residues 3.46 and 7.53 are typically ≤4.5 Å, allowing for hydrophobic or van der Waals interactions. However, for ETBR, the distance between L1953.46 and L3867.53 was approximately 7.3 Å without direct contact, because the absence of Y7.53 and downward shift of TM7 created a space between them (Fig. 3a). The side chains of the rotated R1993.50 and N3827.49 extended toward this space, where possible water molecules were detected (Fig. 3a, Supplementary Fig. 8a). In a later section, we validated the presence of water molecules using MD simulations.
The downward shift of TM7 was stabilized by a hydrophobic interaction between L3867.53 and I1402.43, which simultaneously interacted with L1953.46 (Fig. 3a). Despite the considerable distances between residues 3.46 and 7.53, precluding direct contacts, this conformation could be maintained. Furthermore, V3897.56, located one turn downward from L3867.53 in TM7 of ETBR, contacted T3246.36 in TM6, as seen in other class-A GPCRs with the conserved Y7.53 (Fig. 3b). Remarkably, residues S3908.47 and V3256.37, which are adjacent and play crucial roles as binding sites for the α5 helix of Gαi (described in the next section), were appropriately arranged in the active conformation of ETBR through the downward motion of TM7. Hence, although the unique 382N7.49Pxx386L7.53 motif creates an unusual space between L1953.46, R1993.50, N3827.49, and L3867.53 (Fig. 3a, Supplementary Fig. 8a), a binding pocket for the α5 helix of Gαi was established in the active structure of ETBR (described ahead). This unusual space can be observed in the area between V1263.46 and Y3057.53 of NK1R, comprising the NPxxY motif26,27. The surrounding area demonstrates an active conformation like ETBR, characterized by a downward shift of the cytoplasmic end of TM7, contrasting with the other GPCRs with the NPxxY motif (Fig. 3c, Supplementary Fig. 8; see the Discussion section).
The biological importance of these interactions in the active conformation of ETBR was confirmed through the dissociation of heterotrimeric G proteins associated with its activation7,28 (Fig. 2f, g, Supplementary Fig. 9, Table 1). Mutations R1993.50A, Y2935.58F, and N3827.49A, resulted in nearly complete impairment in the Gi-protein dissociation assay. Additionally, although hydrophobic mutations of L3867.53 to Ile and Val reduced dissociation activities by approximately 50% when considering their expression levels (Supplementary Table 1), mutations of L3867.53 to hydrophilic or small residues, such as Tyr, Ala, or Asn, resulted in severely impaired activities. The importance of these residues in forming the active conformation was confirmed through the GloSensor cAMP accumulation assay (Promega) through Gs coupling. We observed severe impairment due to mutations, which is consistent with the findings of the Gi dissociation assay (Supplementary Fig. 10a, b, Supplementary Table 2). Thus, interactions between R1993.50, Y2935.58, and N3827.49 are biologically essential for Gi-protein activation, and the bulky hydrophobic residue leucine at 3867.53 is important for the active conformation of ETBR.
ETBR–Gi interface
The structure of the ETBR–Gi complex (Fig. 1a, Supplementary Fig. 7) revealed a mode of interaction similar to that in other Gi-bound receptors. However, the interactions between ETBR and Gi were exclusively mediated through the α5 helix of Gαi. This helix binds ETBR in a more vertical orientation in ETBR–Gi than in other GPCR–Gs or Gq structures (Supplementary Fig. 7). Consequently, the C-terminus of the α5 helix of Gαi dominantly bound ETBR, which confined the ETBR–Gi interface within a relatively narrow area.
A significant interface between ETBR and Gαi1 was formed by TM3, TM5, TM6, TM7, ICL1, and ICL2 of the receptor, in addition to the last 15 residues of the C-terminal α5 helix (residues 340–351) and the following three-residue wavy hook (352-GLF-354) of Gαi (Fig. 4a). In detail, as observed in many Gi-bound GPCR complexes, the apex of the α5-helix engaged with the end of TM7 and helix 8. At this interface, the backbone carbonyl of G353H5.24 (superscripts refer to the CGN numbering system)29 and C-terminal carboxylate of F354H5.26 formed hydrogen bonds with the side chain of S3908.47 and the backbone carbonyl of V3897.56 of ETBR. Gαi residues, including D341H5.13, N347H5.19, and D350H5.22, established four hydrogen bonds with ETBR residues N134AICL1, R208ICL2, K210 ICL2, and R3186.30 (Supplementary Table 3).
However, the amino acid residues located between the cytoplasmic cleft of ETBR and the α5 helix of Gαi predominantly formed van der Waals interactions for the pairs C351H5.23–R1993.50 and N347H5.19–A2023.53 (Fig. 4, Supplementary Table 3). Notably, the large side chains of L348H5.20 and L353H5.25 nestled deeply into the hydrophobic pocket formed by V2033.54, M2965.61, M3005.65, V3216.33, V3256.37, and V3897.56 in TM3, TM5, TM6, and TM7 of ETBR (Fig. 3b). The hydrophobic residues I344H5.16 and I343H5.15 formed interactions with W206ICL2 and I209ICL2, respectively. Although residues at the C-terminal side of the α5 helix of Gαi interacted with residues within the ETBR hydrophobic pocket, residues in the middle part of the α5 helix, such as T340H5.12, D341 H5.13, and I343 H5.15, interacted with residues in ICL2, such as W206ICL2 and I209ICL2, or close to ICL3, such as H3146.26 and R3186.30. Thus, the α5 helix of Gαi binding to ETBR showed a relatively vertical orientation (Supplementary Fig. 7). This resulted in a shorter cytoplasmic side of TM5 compared with other class-A GPCRs, and the ICL2 of ETBR formed a flexible loop.
ETBR-Gi dissociation assay
These structural observations were validated using an ETBR-stimulated Gi-protein dissociation assay to examine the recognition determinants. The each ETBR mutant receptor retained the affinity for ET-1 comparable to that of the wild-type (Supplementary Table 4). Among ETBR mutations, S3908.47A, M2965.61A, M3005.65A, and V3256.37A substantially reduced the coupling between the receptor and Gαi by approximately 50%, whereas N134ICL1A, H3146.26A, R3186.30A, V3897.56A, and K3918.48A mutations retained comparable or slightly reduced activities compared with wild-type, considering the expression of mutant receptors (Fig. 5a–c, Supplementary Table 1c–e). By contrast, among Gαi mutations, replacing L353H5.25 with alanine severely impaired coupling with ETBR, whereas G352AH5.24 and K345AH5.17 mutations decreased coupling by 50%. C351AH5.23 and F354AH5.26 mutations showed a slight reduction, whereas D341AH5.13 and D350AH5.22 mutations did not exhibit marked defects (Fig. 5d, Supplementary Table 1f). These findings are consistent with extensive mutagenesis studies of Gαi1 on the stability and formation of the rhodopsin–Gi complex, where L353AH5.25, G352AH5.24, and L348AH5.20 substitutions severely impaired coupling, and C351AH5.23, K345AH5.17, and I344AH5.16 substitutions reduced complex formation efficiencies to approximately 60%30. Therefore, coupling efficacies affected by mutations in ETBR and the α5 helix of Gαi corresponded well with each other, reflecting their interactions at the observed interface of the complex. Notably, interactions at the end of TM7 and helix 8 of ETBR with the C-terminus of the Gαi α5 helix, as well as the hydrophobic pocket composed of V2033.54, M2965.61, M3005.65, and V3256.37 with the C-terminal L348H5.20 and L353H5.25 of Gαi α5 helix, play crucial roles in ETBR–Gi coupling.
Most residues of the C-terminal α5 helix (T340–F354) interacted with ETBR in the complex, except K345H5.17, which interacted with F354H5.26 through a cation–π interaction, and with D341H5.13 and E318h4s6.12 through salt bridges within Gαi1 (Supplementary Fig. 11). In the GDP-bound form, K345H5.17 did not interact with D341H5.13 or E318h4s6.12, which was originally located at the end of the β6 sheet. The translation and twist of the α5 helix during coupling with ETBR led to K345H5.17 interacting with D341H5.13 and E318h4s6.12. This interaction stabilized the twisted α5 helix and the conformation of the shortened β6 sheet as well as the GDP-released β6-α5 loop. The K345AH5.17 mutation led to an approximately 50% reduction in the Gi dissociation assay (Fig. 5d) and the rhodopsin–Gi complex formation assay30, indicating that K345H5.17 plays a fundamental role in Gi activation. This role includes modulating the location of C-terminal F354H5.26 and stabilizing the ETBR–Gi complex.
ETBR coupled through the C-terminus of Gα
The α5 helix comprises conserved and variable residues across Gα proteins and could serve as a common mode of interaction with various types of GPCRs or as a selective mode of interaction based on receptor specificity29. The structural insights provided by the ETBR–Gi structure, in addition to the results of biological validation, suggest that conserved hydrophobic residues, particularly L348H5.20 and L353H5.25, play pivotal roles in coupling (Figs. 4b and 5d). These residues form numerous contacts with specific residues in the hydrophobic binding pocket of ETBR. When these residues are substituted with others, coupling is significantly impaired (Fig. 5b, Supplementary Fig. 10c). Additionally, subtype-specific residues involved in Gα selectivity, such as C351H5.23 and G352H5.24, occupied crucial positions in the complex and established contacts with the central residues of ETBR, including R1993.50 and L3867.53 (Figs. 2f, g, 5a, d, Supplementary Fig. 10, Tables 1, 2). Notably, the primary interactions of ETBR with the α5 helix of Gαi are limited to the transmembrane area. This is because the binding of the C-terminal α5 helix to ETBR occurs in a relatively vertical orientation, and ICL2 of ETBR is a flexible loop. Consequently, in the coupling of ETBR with other subfamilies, such as Gs, Gq, and G12, it is likely that the conserved L348H5.20 and L353H5.25 continue to play central roles as binding partners through a common mode of interactions (Fig. 3b, d). ETBR may further adapt to selectively accommodate subtype-specific residues, such as H5.23 and H5.24, based on the requirements of the G-protein subfamily29. These distinctive features would enable ETBR to exhibit promiscuity in coupling with G-protein subfamilies2,6,7.
ETBR–Gi interactions in molecular dynamics simulations
We performed molecular dynamics (MD) simulations of the ET-1–ETBR–Gi complex to evaluate the key interactions for ETBR–Gi activation. The simulations, each lasting 500 ns, were repeated three times with different initial velocities. The time evolutions of the Cα RMSDs of ETBR, Gαi, Gβ, and Gγ from the initial structures are shown in Supplementary Fig. 12a. The structures of ETBR, Gβ, and Gγ remained stable during MD simulations with consistent RMSD values of <3 Å. However, Gαi underwent substantial conformational changes due to the large flexibility of its activated form. The Cα RMSDs of ET-1 and the C-terminal α5 helix of Gαi (residues 335–354) were calculated after superposing the Cα atoms of ETBR on those of the initial structure (Supplementary Fig. 12b). No significant change occurred in either run, indicating stable binding of ET-1 and Gαi to ETBR. We calculated the probabilities of hydrogen-bond formation for pairs D341H5.13–R3186.30, N347H5.19–R208ICL2, D350H5.22–N134ICL1, and F354H5.26–S3908.47 to analyze the stability of intermolecular interactions (Fig. 6a, Supplementary Table 5). Hydrogen bonds for pairs D341H5.13–R3186.30 and F354H5.26–S3908.47 were stably formed with probabilities of approximately 0.7. Although the hydrogen bond between D350H5.22 and N134ICL1 was broken after 130 ns of run 3, it was formed in runs 1 and 2 with probabilities of approximately 0.7 and 0.4, respectively, indicating the formation of a weak bond. By contrast, N347H5.19 and R208ICL2 rarely formed a hydrogen bond, because R208ICL2 exhibited high structural flexibility. Next, we analyzed intermolecular hydrogen bonds within ETBR for pairs D1472.50–N3827.49, D1472.50–S3797.46, and R1993.50–Y2935.58. Hydrogen bond D1472.50–S3797.46 was stable in all runs. Hydrogen bonds for pairs D1472.50–N3827.49 and R1993.50–Y2935.58 were weak because they formed only in runs 1 and 2. Additionally, we calculated the average water occupancy in the intracellular cavity of ETBR using the 500 ns trajectory of run 1 to analyze water-mediated interactions (Fig. 6b). Water densities exceeding 2-fold bulk density were observed in the cavity surrounded by L1953.46, R1993.50, N3827.49, and L3867.53. Minimum distances for the pairs R1993.50–N3827.49 and L1953.46–L3867.53 settled at approximately 7 and 6 Å, respectively (Fig. 6c, d). Thus, MD simulations revealed a water-mediated hydrogen-bond network connecting the area of Y2935.58–R1993.50–water molecules–N3827.49–S3797.46–D1472.50 residing at the center of ETBR. Accordingly, a relatively bulky density at the tip of R1993.50 observed in the cryo-EM map can be attributed to water, contributing to the network (an arrowhead in Fig. 3a). This network was sealed by hydrophobic interactions through L1953.46, I1402.43, and L3867.53, and ultimately completed by the binding of the α5 helix of Gαi to the receptor.
Discussion
The diversity in residue L3867.53 within NPxxL in TM7 is crucial for the active conformation of ETBR. Surprisingly, L386Y, as well as L386N/A, mutant receptors severely impaired G-protein activation (Fig. 2g, Supplementary Fig. 10b). Only the hydrophobic mutant receptor L386I/V retained approximately 50% of the activity. The mutant receptors indicate that a bulky hydrophobic residue at position 7.53 is indispensable for the active conformation of ETBR. In the common rearrangement that occurs upon activation, direct contacts occur between residues at positions 7.53 and 3.4610. However, L3867.53 was distant from L1953.46 in ETBR and linked with it through I1402.43 through hydrophobic interactions, presumably to maintain hydrophilic interactions and form stable contacts in the active conformation of ETBR (Fig. 3a). In addition, downward-shifted L3867.53 positions V3897.56 one turn below in TM7 adequately to create the binding site for Gαi. Both V3897.56 and the adjacent S3908.47, located at the transition of TM7 to helix 8, interact with the C-terminal region of Gαi, specifically the backbone carbonyl of G352H5.24 and the C-terminal carboxylate of F354H5.26 (Figs. 3b and 4a). These interactions play crucial roles in coupling (Fig. 5a, d, Supplementary Fig. 10). V3897.56 interacts closely with T3246.36 in TM6, adjacent to V3256.37, which interacts with M2965.61 in TM5, under which M3005.65 is positioned one turn below, and which in turn is close to V2033.54. Altogether, V3256.37, M2965.61, M3005.65, and V2033.54 align to form a hydrophobic core to bind the C-terminal L353H5.25 and L348H5.20 of α5 helix of Gαi. These interactions constitute one of the primary binding determinants (Fig. 5b, d, Supplementary Fig. 10c). Coordinating with V3897.56, the diverse N7.49PxxL7.53 motif plays a structural role in the active conformation of ETBR through a downward shift, similar to NPxxY. In class-A GPCRs, approximately 4% of the receptors possess the N7.49P7.50xxX7.53 sequence (X is leucine, phenylalanine, threonine, histidine, and so on (GPCRdb, http://www.gpcrdb.org) on the cytoplasmic side of TM7, such as ETAR31 and GRPR/BB232. In these receptors, L7.53 may contribute to the organization of a binding pocket for Gα, similar to that observed in ETBR. Alternatively, NK1R (with NPxxY) shows an unusual downward shift of the cytoplasmic end of TM7 upon activation. Because N3017.49 of NK1R forms direct hydrogen bonds with E782.50 and Y3057.53 in the active state, and the cytoplasmic side of TM7 does not shift inward, but to the intracellular side upon activation, due to the longer side chain of E782.50 at position D2.50 (Fig. 3c)26,27. Consequently, the downward-shifted L3087.56 one turn below Y3057.53 plays an essential role as a structural pivot in the active conformation, as well as a member of the hydrophobic binding site for L353H5.20 and L358H5.25 of the C-terminal α5 helix of Gαq, in addition to M2496.36, V2466.33, L2235.65, I1343.54, and R1303.50 (Fig. 3d).
Ji et al. reported cryo-EM models of ET-1-bound ETAR and ETBR coupled to miniGs/q, as well as a selective peptide IRL1620-bound ETBR coupled to Gi, providing valuable structural insights into these complexes31. Their findings suggest that interface regions between ETRs and Gi/q in the structures of ETAR and ETBR bound to ET-1 resemble the interface observed in our ETBR–Gi complex structure. This implies that both ET-1-bound ETAR and ETBR engage Gi and Gq in a manner similar to the hydrophobic binding pocket of L348H5.20, L353H5.25, and S373/S3908.47, interacting with the C-terminal end of the α5 helix. However, the deposited structures (PDB code 8HCQ, 8HCX, 8HBD) show some ambiguities. Discrepancies in the extracellular region, such as lack of disulfide bonds C158/C239 and C69/C341 in ETAR, C174/C255 in ETBR, and C3/C11 in ET-1–ETBR, could affect structural interpretation. Furthermore, Sano et al. presented the cryo-EM structure of the ET-1–ETB–Gi complex33. They observed a downward shift of the cytoplasmic side of TM7, consistent with our results. Although they used different constructs for Gi protein, including the linker between ETBR and β subunit of Gi, their findings were consistent with the overall structure of the ET-1–ETBR–DNGi1–scFv16 complex. Notably, they described binding of the C-terminal α5 helix of Gαi to ETBR as “shallow;” however, we have highlighted that the C-terminal wavy hook of Gαi is in a relatively deeper position than that in other Gi-coupled GPCRs, indicating a more vertical orientation in binding. The nearly identical structures with significant differences collectively contribute to a deeper understanding of the structural basis of ETAR and ETBR activation, their interactions with various G proteins, and the details of the ligand binding interface.
Materials and methods
Expression and purification of ETBR
We used a previously described human ETBR construct with cleavable N- and C-terminal tags. The N-terminus was modified to include the hemagglutinin signal peptide followed by a Flag tag. Rhinovirus 3C protease recognition site (LEVLFQGP) was introduced between G57 and L66. The C-terminus was truncated at S407; three cysteine residues were mutated to alanine (C396A, C400A, and C405A), as described20; and fused with an EGFP-HiS9 tag22, following rhinovirus 3C protease recognition site. The R1241.55Y mutation was introduced to increase thermostability21. The resulting construct was introduced into the pFastBac vector. Recombinant baculovirus was prepared using the Bac-to-Bac baculovirus expression system (Invitrogen). Spodoptera frugiperda Sf9 insect cells (Invitrogen) were infected with the virus at a cell density of 3.0–4.0 ×106 cells/mL in Sf900 II medium and cultured for 48 h at 27 °C. To purify ETBR, harvested cells were lysed with hypotonic lysis buffer (20 mM HEPES [pH 7.5], 0.1 μM ET-1, and protease inhibitors) and centrifuged at 30,000 ×g for 20 min. The pellet was homogenized with a Dounce homogenizer in a solubilization buffer (1% lauryl maltose neopentyl glycol (LMNG, Anatrace), 0.1% cholesteryl hemisuccinate (CHS, Sigma-Aldrich), 20 mM HEPES [pH 7.5], 200 mM NaCl, 20% glycerol, 0.2 μM ET-1, and protease inhibitors) and solubilized for 1 h at 4 °C. The insoluble cell debris was removed by centrifugation (30,000 ×g, 20 min), and the supernatant was mixed with TALON cobalt resin (Clontech) for 2 h at 4 °C. The resin was collected in an open glass column, washed with 10 column volumes of wash buffer I (0.01% LMNG, 0.001% CHS, 20 mM HEPES [pH 7.5], 500 mM NaCl, 20% glycerol, and 10 mM imidazole), washed with 5 column volumes of wash buffer II (0.01% LMNG, 0.001% CHS, 20 mM HEPES [pH 7.5], 100 mM NaCl, 10% glycerol, and 10 mM imidazole), and eluted in wash buffer II supplemented with 250 mM imidazole. The eluate was concentrated, mixed with ET-1 to 1 μM, and dialyzed against a buffer containing 0.01% LMNG, 0.001% CHS, 20 mM HEPES (pH 7.5), 100 mM NaCl, 10% glycerol, 0.1 mM TECP, and His-tagged rhinovirus 3 C protease (made in-house) overnight at 4 °C. Following the cleavage of the N-terminus and EGFP–His10 by His-tagged 3 C protease, the sample was mixed with TALON resin for 1 h at 4 °C to remove cleaved EGFP–His10. The ETBR-containing flow-through was concentrated and purified by size-exclusion chromatography on a Superdex 200 Increase 10/300 GL gel-filtration column (Cytiva) in a final buffer (100 mM NaCl, 20 mM HEPES [pH 7.5], 5% glycerol, 0.01% LMNG, 0.001% CHS and 0.1 μM ET1). Peak fractions were pooled and concentrated to 4–5 mg/mL.
Expression and purification of heterotrimeric wild-type Gi1 and DNGi1
Wild-type Gi1 and DNGi1 heterotrimers were expressed in Sf9 or Trichoplusia ni Hi5 insect cells (Expression Systems) and purified as described34. In brief, insect cells were coinfected with two recombinant viruses: one encoding wild-type human Gαi1 or DNGαi1 containing four mutations (S47N, G203A, E245A, A326A) and another encoding wild-type human Gβ1 and Gγ2 subunits with a hexa-histidine tag inserted at the amino terminus of the Gβ1 subunit. Cultures were collected 48 h after infection. Cells were lysed in hypotonic buffer, and lipid-modified heterotrimeric Gi1 or DNGi1 was extracted in buffer containing 0.7% sodium cholate, 0.01% LMNG–0.001% CHS, 20 mM HEPES (pH 7.5), 100 mM NaCl, 5 mM MgCl2, 1 mM TCEP, 50 μM GDP, and protease inhibitors. The soluble fraction was purified using TALON cobalt resin, and the detergent was exchanged from sodium cholate to 0.01% LMNG–0.001% CHS on a column. After elution was complete, the concentrated protein was dialyzed against a buffer containing 20 mM HEPES (pH 7.5), 100 mM NaCl, 1 mM MgCl2, 0.1 mM TCEP, 10 μM GDP, 0.01% LMNG, 0.001% CHS, and His-tagged rhinovirus 3 C protease overnight at 4 °C to cleave the N-terminal His-tag. Then, the sample was mixed with TALON resin for 1 h at 4 °C to remove the cleaved His-tag. The flow-through fraction, containing wild-type Gi1 or DNGi1 heterotrimers, was concentrated and purified by size-exclusion chromatography on a Superdex 200 Increase 10/300 GL gel-filtration column in a final buffer (20 mM HEPES [pH 7.5], 100 mM NaCl, 1 mM MgCl2, 0.1 mM TCEP, 10 μM GDP, 0.01% LMNG and 0.001% CHS). Peak fractions were pooled and concentrated to approximately 20 mg/mL.
Expression and purification of scFv16
Single-chain Fab16 (scFv16) was expressed and purified as described18,24. In brief, scFv16 tagged with hexa-histidine at the C-terminus was expressed with a signal peptide in Hi5 insect cells using the Bac-to-Bac baculovirus expression system. The scFv16 secreted into the culture medium was purified by Ni-NTA (Qiagen) chromatography, following addition of Tris (pH 8.0) to the culture supernatant. The Ni-NTA eluent was dialyzed against a buffer consisting of 20 mM HEPES (pH 7.5), 100 mM NaCl, 0.1 mM TCEP, and rhinovirus 3 C protease overnight at 4 °C. The sample was mixed with TALON resin for 1 h at 4 °C to remove the cleaved His-tag. The flow-through fraction containing scFv16 was concentrated and purified by gel-filtration chromatography in a final buffer (100 mM NaCl and 20 mM HEPES [pH 7.5]). Peak fractions were pooled and concentrated to approximately 60 mg/mL.
Purification of the ETBR–Gi1–scFv16 complex
The ETBR–Gi1–scFv16 complex was prepared as described18,28. Purified ETBR was mixed with a 1.2 molar excess of wild-type or dominant negative Gi1 heterotrimer. The coupling reaction proceeded at 20–24 °C for 2 h, followed by incubation for 1 h at 4 °C with apyrase and λ-phosphatase (New England Biolabs) together with 1 mM MnCl2 and 5 mM MgCl2 for the hydrolysis of unbound GDP and dephosphorylation of proteins, respectively. Furthermore, 1.2 molar excess of scFv16 was added to the mixture and incubated for 2 h at 4 °C. The coupling mixture was incubated with 2A5 anti-ETBR immunoaffinity resin overnight at 4 °C35. Complex-bound resin was first washed in a buffer containing 0.1% LMNG, 0.01% CHS, and 0.0003% glyco-diosgenin (GDN), then washed in gradually decreasing concentrations of LMNG and increasing concentrations of GDN. The complex was eluted in 20 mM HEPES (pH 7.5), 150 mM NaCl, 0.1 mM TCEP, 2 mM EDTA, 5% glycerol, 0.00375% LMNG, 0.000375% CHS, 0.00125% GDN, 0.1 μM ET-1, and 300 μg/mL 2A5 peptide (VPKGDRTAGSPPRTI) at room temperature. Finally, the ETBR–Gi1–scFv16 complex was purified by size-exclusion chromatography on a Superdex 200 Increase 10/300 GL in 20 mM HEPES (pH 7.5), 100 mM NaCl, 0.1 mM TCEP, 0.1 μM ET-1, 0.00075% LMNG, 0.000075% CHS, and 0.00025% GDN. Peak fractions were concentrated to approximately 30 mg/mL for electron microscopy studies.
Collection of Cryo-EM Data
Proteins for cryo-EM were prepared to ~6 and 4 mg/mL for ET-1-bound ETBR–wild-type Gi1–scFv16 (ETBR–WTGi) and ET-1-bound ETBR–DNGi1–scFv16 (ETBR–DNGi), respectively. Protein solution (3 μL) was applied to glow-discharged holey carbon grids (200 mesh Quantifoil R2/2 molybdenum and 200 mesh Quantifoil R1.2/1.3 copper for ETBR–WTGi and ETBR–DNGi, respectively), blotted, and plunged into liquid ethane at −182 °C using an EM GP2 plunger (Leica, Microsystems, Vienna, Austria) and Vitrobot Mark IV (Thermo Fisher Scientific) for ETBR–WTGi and ETBR–DNGi, respectively. Data were collected at OIST on a Talos Arctica (Thermo Fisher Scientific, Hillsboro, USA) electron microscope at 200 kV, equipped with a Falcon 3 camera (Thermo Fisher Scientific) and at SPring-8 on a CRYO-ARM300 electron microscope (JEOL) at 300 kV, equipped with a K3 camera (Gatan) (Supplementary Figs. 1, 2). An in-column energy filter with a slit width of 20 eV was inserted to acquire movie frames using CRYO-ARM300. Movies were recorded using EPU software (Thermo Fisher Scientific) on a Talos Arctica at a nominal magnification of 92,000× in counting mode and a pixel size of 1.094 Å at the specimen level, with a dose rate of 0.93 e- per physical pixel per second. Exposure time was 51.3 s, resulting in an accumulated dose of 40 e- per Å2. Each movie included 40 fractioned frames. The movies were recorded using SerialEM36 and JAFIS Tool version 1 (JEOL) on a CRYO-ARM300 at nominal magnifications of 60,000× and 100,000× in counting mode. The AI detection of each center hole position was performed using yoneoLocr, which prevented any stage alignment failures37. The pixel sizes at the specimen level were 0.816 and 0.507 Å for magnifications of 60,000× and 100,000×, with dose rates of 8.3 and 3.4 e- per physical pixel per second, resulting in an accumulated dose of ~76 and ~65 e- per Å2 for 6.1 s and 4.9 s exposures, respectively. Each movie included 61 fractioned frames.
Image processing
All stacked frames were motion corrected with MotionCor238. Defocus was estimated using CTFFIND439. All the particles picked using crYOLO40 were analyzed with RELION 3.141 and selected by 2D classification (Table 1, Supplementary Figs. 1, 2). The initial 3D model was generated in RELION, and the particles were divided into four classes by 3D classification, resulting in only one good class. The 3D auto-refinement produced a map, after contrast transfer function refinement, Bayesian polishing, masking, and postprocessing. Particle projections were subjected to subtraction of the detergent micelle density followed by 3D auto-refinement, yielding a final map with resolutions of 4.61, 3.21, and 3.62 Å for ETBR–WTGi, ETBR–DNGi, and ETBR after focused 3D classification42, respectively, according to the gold-standard Fourier shell correlation using a criterion of 0.143 (Supplementary Figs. 3–5 for ETBR–WTGi, ETBR–DNGi, and ETBR, respectively)36. Local resolution maps were calculated using RELION.
Model building and refinement of the ETBR–Gi1 complex
The atomic models of ET-1 bound ETBR (PDB ID: 5GLH) and Gi–scFv (PDB ID: 6OS9) were fitted to cryo-EM maps of ETBR–WTGi and ETBR–DNGi, respectively, using Chimera43. Atomic model building was performed using COOT44. The manually modified model was refined in real space on PHENIX45, and the COOT/PHENIX refinement was iterated until the refinements converged. Finally, statistics calculated using MolProbity46 were checked. Figures were drawn using the Pymol Molecular Graphic System (Schrödinger)47, UCSF Chimera43, and UCSF ChimeraX48.
MD simulations
The intracellular loop between TM5 and TM6 (residues 302–311) of ETBR and α-helical domain of Gαi (residues 56–181, 234–240), which are missing in the cryo-EM structure, were modeled using modeller 9.2449. The X-ray crystallographic structures of the D2 dopamine receptor–Gi complex (PDB ID: 6VMS) and rhodopsin–Gi complex (PDB ID: 6CMO) were used as templates for modeling the intracellular loop of ETBR (Supplementary Fig. 12) and the α-helical domain of Gαi, respectively. The structure of ET-1-bound ETBR–Gi was embedded in a solvated 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC) bilayer using the CHARMM-GUI server50. The protein structure was protonated using the default settings of the CHARMM-GUI server. The system was composed of 453 POPC molecules, 64,293 water molecules, and 0.15 M K+/Cl− ions adjusted to neutralize the net charge of the entire system (Supplementary Table 6). The CHARMM36m force field51,52 was used for proteins, ions, and POPC molecules53. The TIP3P model54 was used for water. Energy minimization and equilibration were performed using the CHARMM-GUI protocol with additional distance restraints between the hydrogen-bond donor and acceptor atoms found in the cryo-EM structure. The parameters for the distance restraints were r0 = 0 nm, r1 = 0.3 nm, r2 = 0.4 nm, and k = 4000 kJ mol−1 nm−2. Then, additional three-step equilibrations were performed with decreasing force constant. Simulations of 50-, 30-, and 20-ns were performed with k = 4000, 1000, and 200 kJ mol−1 nm−2, respectively. After equilibrium simulations, a production run was performed in the constant-NPT ensemble for 500 ns. The temperature was maintained at 303.15 K using the Nose–Hoover thermostat55,56 with a coupling constant of 1.0 ps. The pressure was maintained at 1.0 bar using a Parrinello–Rahman barostat57 with a coupling constant of 5.0 ps. Electrostatic interactions were calculated using the particle mesh Ewald method58 with a real space cutoff of 1.2 nm. Van der Waals interactions were calculated with a modified Lennard–Jones potential, where the force was smoothly switched to zero between 1.0 and 1.2 nm. The lengths of the bonds involving hydrogen atoms were constrained using the LINCS algorithm59,60 to allow for the use of a time step of 2 fs. The simulations were repeated three times with different initial velocities. All simulations were performed using GROMACS 2022.461.
Probabilities of hydrogen-bond formation in the MD simulations were calculated using the “gmx hbond” tool with default settings. To calculate the density of water molecules, each snapshot of the trajectories was translated and rotated to superpose Cα atoms of ETBR on the corresponding atoms of the initial structure. A cubic grid with a spacing of 0.4 Å was then created. Water density (ρi) at grid point i was calculated as follows:
where T is the number of snapshots in the trajectories, N is the number of water molecules in the system, Vr is the volume of a sphere with radius r (r = 1 Å), xj,t represents the coordinates of the oxygen atom of the j-th water molecule of the t-th snapshot, ci is the coordinate of the grid point i, and H(x) is the Heaviside step function.
NanoBiT G-protein dissociation assay
Gi activation was measured using a NanoBiT G-protein dissociation assay7, in which heterotrimeric G-protein dissociation catalyzed by GPCR was monitored using a NanoBiT system (Promega). A large fragment (LgBiT) of NanoBiT luciferase was inserted into Gαi1, and a small fragment (SmBiT) was N-terminally fused to a C68S-mutated Gγ2. The amino acid sequences of the NanoBiT G-protein constructs used in this study are identical to those in Inoue et al. 7. The genes coding for the NanoBiT G-protein constructs, untagged Gβ1 construct, and Flag-tagged ETBR were synthesized and cloned into pCAG vectors (provided by Dr. Jun-ichi Miyazaki at Osaka University, Japan) or pcDNA3.1 expression plasmid by GenScript. Mixtures of plasmids prepared for transfection of HEK293A cells (Thermo Fisher Scientific) were prepared as described7. HEK293A cells were seeded in a 6-well culture plate at a concentration of 2 ×105 cells/mL (2 mL per well) one day before transfection. Transfection solution was prepared by combining 4 μL (per well hereafter) of polyethylenimine solution (Polysciences; 1 mg/mL) and a plasmid mixture consisting of 100 ng LgBiT-inserted Gα subunit (Gαi1), 500 ng Gβ1, 500 ng C68S-mutant SmBiT-fused Gγ2, and 200 ng wild-type or mutant ETBR in 200 μL of Opti-MEM (Thermo Fisher Scientific). After 1-day incubation, transfected cells were collected with 0.5 mM EDTA-containing Dulbecco’s PBS (D-PBS), centrifuged, and suspended in 2 mL of Hank’s Balanced Salt Solution containing 0.01% bovine serum albumin (BSA; fatty-acid-free grade; SERVA) and 5 mM HEPES (pH 7.4) (assay buffer). The cell suspension was dispensed into a white 96-well plate (Greiner Bio-one) at a volume of 80 μL per well and loaded with 20 μL of 50 μM coelenterazine (Carbosynth) diluted in the assay buffer. After 2-h incubation at room temperature in the dark, baseline luminescence was measured (GloMax Navigator, Promega). A range of ET-1 solutions (20 μL of 0–6 × 10−6 M) were added and incubated for 3–5 min at room temperature before the second measurement. Luminescence counts were normalized to the initial count, and fold-change signals over vehicle treatment were used to evaluate the G-protein dissociation response. The G-protein activation signals were fitted to a 3-parametric concentration–response curve (GraphPad Prism 9.4), and pEC50 values and span values (“Top”–“Bottom”) as Emax were obtained.
GloSensor cAMP assay
Gs activation was measured by the GloSensor cAMP accumulation assay, in which ETBR-induced cAMP accumulation was assayed in cells transiently expressing a biosensor variant, with a cAMP binding domain fused to a luciferase mutant, according to the manufacturer’s instructions (Promega). HEK293A cells were seeded in a 6-well culture plate at a density of 2.5 ×105 cells/mL (2 mL per well) one day before transfection. The cells were transfected with a mixture of pGloSensor cAMP 22 F plasmid (1.5 μg per well) and pCAG expression plasmid encoding ETBR or mutant receptors (0.5 μg per well) using 6 μL of FuGENE HD transfection reagent (Promega) in 200 μL of Opti-MEM I reduced serum medium (Thermo Fisher Scientific). After 24 h of incubation, the transfected cells were harvested with 0.5 mM EDTA-containing D-PBS, centrifuged, and suspended in 2 mL of CO2-independent medium containing 10% FBS (Invitrogen). The cell suspension was dispensed into a white 96-well plate at a volume of 80 μL per well and loaded with 20 μL of 5 mM D-luciferin in CO2-independent medium containing 10% FBS. After 2 h incubation at room temperature in the dark, baseline luminescence was measured (GloMax Navigator, Promega). Varying concentrations of ET-1 solution (20 μL of 0–6 ×10−6 M) were added and incubated for 5 min at room temperature before the second measurement. Luminescence counts were normalized to the initial count. To evaluate the Gs-activated response, fold-change signals over vehicle treatment were represented as percentage of wild-type Emax. The activation signals were fitted to a three-parametric concentration–response curve (GraphPad Prism 9.4), and pEC50 and relative Emax values were obtained. Although a slight decrease was observed in the baseline without ETBR expression plasmid (vehicle only) and increasing ET-1 concentration, we did not use phosphodiesterase inhibitors and pertussis toxins, because cAMP signals produced by ETBR expression were sufficiently high in HEK293 cells (Supplementary Fig. 10e).
[125I]ET-1 binding assay
In the NanoBiT G-protein dissociation assay and the cAMP accumulation assay, one quarter of the transfected cells were separately frozen in liquid nitrogen and stored at −80 °C. The numbers of wild-type or mutant ETBRs expressed in the transfected cells were monitored by residual [125I]ET-1 binding activity, reflecting correctly folded ETBRs. A single-point binding assay using hydroxyapatite resin was performed as described21. Briefly, transfected cells were suspended in 50–100 μL of binding buffer containing 50 mM sodium phosphate buffer (pH 7.5), 2 mM MgCl2, 0.1% BSA, and 0.1% digitonin. Then, 0.5–2 μL of samples (1.5–6 μg total protein) were incubated with approximately 150 pM [125I]ET-1 (PerkinElmer) in 50 μL binding buffer at room temperature for 30 min. Hydroxyapatite resin (30 μL, BioRad) in 15% slurry was added to absorb receptor proteins, and the mixtures were centrifuged at 2000 rpm for 2 min to remove unbound [125I]ET-1. Pelleted resin was washed with 0.3 mL of 50 mM sodium phosphate buffer (pH 7.5), 2 mM MgCl2, and 0.1% digitonin and measured using a γ-counter. The count of [125I]ET-1 bound in the presence of 100 nM ET-1 was subtracted as a background, which was approximately 10% or less of total binding. Each assay was performed in duplicate three times. Relative expression was represented as wild type 100%.
The apparent dissociation constants (Kd) of ET-1 for wild-type and mutant ETB receptors expressed in HEK293A cell membrane were measured using saturation binding assays with [125I]ET-1. The cell membranes containing ETB receptors were incubated with eight different concentrations of [125I]ET-1 ranging from 2.0 to 200 pM in 50 μl of 50 mM HEPES-NaOH, pH 7.5, 10 mM MgCl2 (Mg-HEPES) buffer containing 0.1% BSA at 37 °C for 2 h. Binding reactions were terminated by dilution with cold Mg-HEPES, then were filtered onto glass fiber filters in 96-well plates (multiscreen HTS FB, Merck Millipore) pretreated with 0.1% BSA in Mg-HEPES, to separate the unbound [125I]ET-1. After three washes with cold Mg-HEPES, the radioactivity captured by the filters was counted using a γ-counter. The non-specific binding of [125I]ET-1 in each reaction was assessed by including 100 nM ET-1 in the same reaction. Assays were performed in duplicate three times and analyzed by fitting to a one-site binding equation (total and nonspecific) using GraphPad Prism 9.4.
Statistics and reproducibility
NanoBiT G-protein dissociation assay and GloSensor cAMP assay were analyzed using GraphPad Prism 9.4 (GraphPad) and are presented as mean ± standard error of the mean (SEM) from three to five independent experiments conducted in duplicate or triplicate. Statistical analyses were performed using Prism 9.4 (GraphPad) with one-way analysis of variance followed by Dunnett’s multiple comparison of means test or Student’s t test. Significance levels in statistical differences are indicated as (****p < 0.0001, ***p < 0.001, **p < 0.01, *p < 0.05 vs. WT).
Reporting summary
Further information on research design is available in the Nature Portfolio Reporting Summary linked to this article.
Supplementary information
Acknowledgements
This research was partially supported by the Platform Project for Supporting Drug Discovery and Life Science Research (Basis for Supporting Innovative Drug Discovery and Life Science Research) from AMED under Grant Numbers JP21am0101118, JP21am0101116, JP22ama121006, JP23ama121004, and JP23ama121027, JST-Mirai Program Grant Number JPMJMI23G2. R.K., M.H., and B.M.H. acknowledge the generous support of the Okinawa Institute of Science and Technology (OIST), Scientific Computing & Data Analysis Section, and Scientific Imaging Section at OIST and the Japanese Cabinet Office. R.K. acknowledges the support of Prof. Tsumoru Shintake. This work was supported by JSPS KAKENHI Grant Number 20H03210 (T.D.), ISHIZUE 2019 from the Kyoto University Research Development Program, and Center for Quantum and Information Life Sciences, University of Tsukuba. T.D. acknowledges the generous support of the laboratory members of Prof. Tochio at Kyoto University.
Author contributions
K.T. and T.D. designed the research, K.T., S.M.-Y, R.K., T.H., and T.D. performed the research, K.T., M.H., A.M., B.M.H., K.Y., and T.D. analyzed the data, T.N, and T.T. performed MD simulation, K.T., T.N., T.T., and T.D. wrote the manuscript, and all authors made editorial contribution.
Peer review
Peer review information
Communications biology thanks Jagannath Maharana, Sadashiva Karnik and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Primary Handling Editors: Janesh Kumar and Laura Rodríguez Pérez. A peer review file is available.
Data availability
The map and model generated in this study have been deposited in the EMDB and PDB with accession codes: EMD-38741 and PDB-8XWQ for the ET-1-bound ETBR-wild-type Gi1-scFv16 complex, EMD-38740 and PDB-8XWP for the ET-1-bound ETBR–DNGi1-scFv16 complex and EMD-60404 and PDB-8ZRT for the focused 3D refinement of ETBR in the ET-1-bound ETBR–DNGi1-scFv16 complex. All other study data including uncropped gel images are included in the article and/or Supplementary Data.
Competing interests
The authors declare no competing interests.
Footnotes
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
These authors contributed equally: Kazutoshi Tani, Saori Maki-Yonekura.
These authors jointly supervised this work: Kazutoshi Tani, Tomoko Doi.
Contributor Information
Kazutoshi Tani, Email: ktani@ccs.tsukuba.ac.jp.
Tomoko Doi, Email: doi.tomoko.8n@kyoto-u.jp.
Supplementary information
The online version contains supplementary material available at 10.1038/s42003-024-06905-z.
References
- 1.Kedzierski, R. M. & Yanagisawa, M. Endothelin system: The double-edged swords in health and disease. Annu. Rev. Pharmacol. Toxicol.41, 851–876 (2001). [DOI] [PubMed] [Google Scholar]
- 2.Davenport, A. P. et al. Endothelin. Phamacol. Rev.68, 357–418 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Barton, M. & Yanagisawa, M. Endothelin: 30 years from discovery to therapy. Hypertension74, 1232–1265 (2019). [DOI] [PubMed] [Google Scholar]
- 4.Enevoldsen, F. C. et al. Endothelin receptor antagonists: status quo and future perspectives for targeted therapy. J. Clin. Med.9, 824 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Ranjan, A. K. & Gulati, A. Sovateltide mediated endothelin B receptors agonism and curbing neurological disorders. Int. J. Mol. Sci.23, 3146 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Doi, T., Sugimoto, H., Arimoto, I., Hiroaki, Y. & Fijiyoshi, Y. Interactions of endothelin receptor subtypes A and B with Gi, Go, and Gq in reconstituted phospholipid vesicles. Biochemistry38, 3090–3099 (1999). [DOI] [PubMed] [Google Scholar]
- 7.Inoue, A. et al. Illuminating G-Protein-Coupling Selectivity of GPCRs. Cell177, 1933–1947 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Avet, C. et al. Effector membrane translocation biosensors reveal G protein and βarrestin coupling profiles of 100 therapeutically relevant GPCRs. eLife11, e74101 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Ballesteros, J. A. & Weinstein, H. Integrated methods for the construction of three dimensional models and computational probing of structure-function relations in G protein-coupled receptors. Methods Neurosci.25, 366–428 (1995). [Google Scholar]
- 10.Venkatakrishnan, J. et al. Diverse activation pathways in class A GPCRs converge near the G-protein-coupling region. Nature536, 484–487 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Weis, W. I. & Kobilka, B. K. The molecular basis of G protein-coupled receptor activation. Annu. Rev. Biochem.87, 897–919 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Thal, D. M., Glukhova, A., Sexton, P. M. & Christopoulos, A. Structural insights into G-protein-coupled receptor allostery. Nature559, 45–53 (2018). [DOI] [PubMed] [Google Scholar]
- 13.Carpenter, B. & Tate, C. G. Active state structures of G protein-coupled receptors highlight the similarities and differences in the G protein and arrestin coupling interfaces. Curr. Opin. Struct. Biol.45, 124–132 (2017). [DOI] [PubMed] [Google Scholar]
- 14.Manglik, A. & Kruse, A. C. Structural Basis for G Protein-Coupled Receptor Activation. Biochemistry56, 5628–5634 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Glukhova, A. et al. Rules of Engagement: GPCRs and G Proteins. ACS Pharmacol. Transl. Sci.1, 73–83 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Hilger, D. The role of structural dynamics in GPCR-mediated signaling. FEBS J.288, 2461–2489 (2021). [DOI] [PubMed] [Google Scholar]
- 17.Venkatakrishnan, A. J. et al. Diverse GPCRs exhibit conserved water networks for stabilization and activation. Proc. Natl Acad. Sci. USA116, 3288–3293 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Koehl, A. et al. Structure of the μ-opioid receptor-Gi protein complex. Nature558, 547–552 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Mobbs, J. I. et al. Structures of the human cholecystokinin 1 (CCK1) receptor bound to Gs and Gq mimetic proteins provide insight into mechanisms of G protein selectivity. PLoS Biol.19, e3001295 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Shihoya, W. et al. Activation mechanism of endothelin ETB receptor by endothelin-1. Nature537, 363–368 (2016). [DOI] [PubMed] [Google Scholar]
- 21.Okuta, A., Tani, K., Nishimura, S., Fujiyoshi, Y. & Doi, T. Thermostabilization of the human endothelin type B receptor. J. Mol. Biol.428, 2265–2274 (2016). [DOI] [PubMed] [Google Scholar]
- 22.Shihoya, W. et al. X-ray structures of endothelin ETB receptor bound to clinical antagonist bosentan and its analog. Nat. Struct. Mol. Boil.24, 758–764 (2017). [DOI] [PubMed] [Google Scholar]
- 23.Liang, Y.-L. et al. Dominant negative G proteins enhance formation and purification of agonist-GPCR‑G Protein complexes for structure determination. ACS Pharmacol. Transl. Sci.1, 12–20 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Maeda, S. et al. Development of an antibody fragment that stabilizes GPCR/G protein complexes. Nat. Commun.9, 3712 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Doi, T., Kikuta, K. & Tani, K. Characterization of critical residues in the extracellular and transmembrane domains of the endothelin type-B receptor for propagation of the endothelin-1 signal. Biochemistry59, 1718–1727 (2020). [DOI] [PubMed] [Google Scholar]
- 26.Thom, C. et al. Structures of neurokinin 1 receptor in complex with Gq and Gs proteins reveal substance P binding mode and unique activation features. Sci. Adv.7, eabk2872 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Harris, J. A. et al. Selective G protein signaling driven by substance P–neurokinin receptor dynamics. Nat. Chem. Biol.18, 109–115 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Kato, H. E. et al. Conformational transitions of a neurotensin receptor 1-Gi1 complex. Nature572, 80–85 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Flock, T. et al. Universal allosteric mechanism for Gα activation by GPCRs. Nature524, 173–179 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Sun, D. et al. Probing Gαi1 protein activation at single-amino acid resolution. Nat. Struct. Mol. Biol.22, 686–694 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Ji, Y. et al. Structural basis of peptide recognition and activation of endothelin receptors. Nat. Commun.14, 1268 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Peng, S. et al. Structures of human gastrin-releasing peptide receptors bound to antagonist and agonist for cancer and itch therapy. Proc. Natl Acad. Sci. USA1ss20, e2216230120 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Sano, F. K., Akasaka, H., Shihoya, W. & Nureki, O. Cryo-EM structure of the endothelin-1-ETB-Gi complex. eLife12, e85821 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Dror, R. O. et al. Structural basis for nucleotide exchange in heterotrimeric G proteins. Science348, 1361–1365 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Yamaguchi, T., Arimoto-Tahara, I., Fujiyoshi, Y. & Doi, T. Characterization and application of monoclonal antibodies against human endothelin B receptor expressed in insect cells. Biotech. Lett.26, 293–299 (2004). [DOI] [PubMed] [Google Scholar]
- 36.Rosenthal, P. B. & Henderson, R. Optimal determination of particle orientation, absolute hand, and contrast loss in single-particle electron cryomicroscopy. J. Mol. Biol.333, 721–745 (2003). [DOI] [PubMed] [Google Scholar]
- 37.Yonekura, K., Maki-Yonekura, S., Naitow, H., Hamaguchi, Y. & Takaba, K. Machine learning-based real-time object locator/evaluator for cryo-EM data collection. Commun. Biol.4, 1044 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Zheng, S. Q. et al. MotionCor2: anisotropic correction of beam-induced motion for improved cryo-electron microscopy. Nat. Methods14, 331–332 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Rohou, A. & Grigorieff, N. Fast and accurate defocus estimation from electron micrographs. J. Struct. Biol.192, 216–221 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Wagner, T. & Raunser, S. The evolution of SPHIRE-crYOLO particle picking and its application in automated cryo-EM processing workflows. Commun. Biol.3, 61 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Zivanov, J. et al. New tools for automated high-resolution cryo-EM structure determination in RELION-3. eLife7, e42166 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Bai, X. C., Rajendra, E., Yang, G., Shi, Y. & Scheer, S. H. Sampling the conformational space of the catalytic subunit of human gamma-secretase. eLife4, e11182 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Pettersen, E. F. et al. UCSF Chimera-a visualization system for exploratory research and analysis. J. Comput. Chem.25, 1605–1612 (2004). [DOI] [PubMed] [Google Scholar]
- 44.Emsley, P., Lohkamp, B., Scott, W. G. & Cowtan, K. Features and development of Coot. Acta Cryst.D66, 486–501 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Adams, P. D. et al. PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta Cryst.D66, 213–221 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Chen, V. B. et al. MolProbity: all-atom structure validation for macromolecular crystallography. Acta Crystallogr. D. Biol. Crystallogr.66, 12–21 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.DeLano, W. L. The PyMOL molecular graphics system (DeLano Scientific, LCC, 2004).
- 48.Pettersen, E. F. et al. UCSF ChimeraX: Structure visualization for researchers, editors, and developers. Protein Sci.30, 70–82 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Martí-Renom, M. A. et al. Comparative protein structure modeling of genes and genomes. Annu. Rev. Biophys. Biomol. Struct.29, 291–325 (2000). [DOI] [PubMed] [Google Scholar]
- 50.Jo, S., Kim, T., Iyer, V. G. & Im, W. CHARMM-GUI: A web-based graphical user interface for CHARMM. J. Comput. Chem.29, 1859–1865 (2008). [DOI] [PubMed] [Google Scholar]
- 51.Huang, J. & MacKerell, A. D. CHARMM36 all-atom additive protein force field: Validation based on comparison to NMR data. J. Comput. Chem.34, 2135–2145 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Huang, J. et al. CHARMM36m: An improved force field for folded and intrinsically disordered proteins. Nat. Methods14, 71–73 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Klauda, J. B. et al. Update of the CHARMM all-atom additive force field for lipids: validation on six lipid types. J. Phys. Chem. B114, 7830–7843 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Jorgensen, W. L., Chandrasekhar, J., Madura, J. D., Impey, R. W. & Klein, M. L. Comparison of simple potential functions for simulating liquid water. J. Chem. Phys.79, 926–935 (1983). [Google Scholar]
- 55.Hoover, W. G. Canonical dynamics: Equilibrium phase-space distributions. Phys. Rev. A31, 1695–1697 (1985). [DOI] [PubMed] [Google Scholar]
- 56.Nosé, S. A molecular dynamics method for simulations in the canonical ensemble. Mol. Phys.52, 255–268 (1984). [Google Scholar]
- 57.Parrinello, M. & Rahman, A. Crystal Structure and Pair Potentials: A Molecular-Dynamics Study. Phys. Rev. Lett.45, 1196–1199 (1980). [Google Scholar]
- 58.Darden, T., York, D. & Pedersen, L. Particle mesh Ewald: An N⋅log(N) method for Ewald sums in large systems. J. Chem. Phys.98, 10089–10092 (1993). [Google Scholar]
- 59.Hess, B., Bekker, H., Berendsen, H. J. C. & Fraaije, J. G. E. M. LINCS: A linear constraint solver for molecular simulations. J. Comput. Chem.18, 1463–1472 (1997). [Google Scholar]
- 60.Hess, B. P-LINCS: A parallel linear constraint solver for molecular simulation. J. Chem. Theory Comput.4, 116–122 (2008). [DOI] [PubMed] [Google Scholar]
- 61.Abraham, M. J. et al. Gromacs: High performance molecular simulations through multi-level parallelism from laptops to supercomputers. SoftwareX1, 19–25 (2015). [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
The map and model generated in this study have been deposited in the EMDB and PDB with accession codes: EMD-38741 and PDB-8XWQ for the ET-1-bound ETBR-wild-type Gi1-scFv16 complex, EMD-38740 and PDB-8XWP for the ET-1-bound ETBR–DNGi1-scFv16 complex and EMD-60404 and PDB-8ZRT for the focused 3D refinement of ETBR in the ET-1-bound ETBR–DNGi1-scFv16 complex. All other study data including uncropped gel images are included in the article and/or Supplementary Data.