Significance
Nature has diversified functions of a single hormone to accomplish the concerted physiological reactions, but one particularly important question is how a single hormone–GPCR pair generate diverse signals, and dissecting the mechanisms underlying these questions could offer therapeutic routes for refractory diseases. In the present article, we illustrated the important role of the EP4-biased signal in treating AKI, and by resolving the agonist-induced Gs/Gi coupling selectivity of EP receptors, we revealed two hypotheses that account for a single hormone–induced functional diversity: 1) The single ligand–receptor interaction mode induces the different intracellular conformational states; 2) the receptor assumes different conformational states to recognize the same ligand, and further generates the different structural propagation paths.
Keywords: prostaglandin E2 receptor, Gs/Gi coupling selectivity, cryo-EM, propagating paths
Abstract
To accomplish concerted physiological reactions, nature has diversified functions of a single hormone at at least two primary levels: 1) Different receptors recognize the same hormone, and 2) different cellular effectors couple to the same hormone–receptor pair [R.P. Xiao, Sci STKE 2001, re15 (2001); L. Hein, J. D. Altman, B.K. Kobilka, Nature 402, 181–184 (1999); Y. Daaka, L. M. Luttrell, R. J. Lefkowitz, Nature 390, 88–91 (1997)]. Not only these questions lie in the heart of hormone actions and receptor signaling but also dissecting mechanisms underlying these questions could offer therapeutic routes for refractory diseases, such as kidney injury (KI) or X-linked nephrogenic diabetes insipidus (NDI). Here, we identified that Gs-biased signaling, but not Gi activation downstream of EP4, showed beneficial effects for both KI and NDI treatments. Notably, by solving Cryo-electron microscope (cryo-EM) structures of EP3-Gi, EP4-Gs, and EP4-Gi in complex with endogenous prostaglandin E2 (PGE2)or two synthetic agonists and comparing with PGE2-EP2-Gs structures, we found that unique primary sequences of prostaglandin E2 receptor (EP) receptors and distinct conformational states of the EP4 ligand pocket govern the Gs/Gi transducer coupling selectivity through different structural propagation paths, especially via TM6 and TM7, to generate selective cytoplasmic structural features. In particular, the orientation of the PGE2 ω-chain and two distinct pockets encompassing agonist L902688 of EP4 were differentiated by their Gs/Gi coupling ability. Further, we identified common and distinct features of cytoplasmic side of EP receptors for Gs/Gi coupling and provide a structural basis for selective and biased agonist design of EP4 with therapeutic potential.
To accomplish concerted physiological reactions and maintain whole-body homeostasis, single hormone may generate distinct effects at different organ/tissue locations. Nature has evolved a system to diversify hormone function at three primary levels: 1) Different receptors can recognize the same ligands (1–12), probably via variable amino acid composition at the receptor level; and 2) the same receptor–ligand pair can couple to different cellular effectors, in certain cases potentially dependent on different structural states with the same primary amino acid sequence for the particular receptor at the conformational level (additionally regulated by the presence of the effector or local effector concentrations); and 3) different alternative splicing or posttranslational modifications of G protein–coupled receptor (GPCR) may serve as a putative layer for the generations of diverse signals downstream of a single hormone–receptor pair engagement. Notably, examples of sensing the same hormone by different receptors or connection of the same receptor–ligand pair to different downstream effectors are often observed within the largest membrane receptor superfamily GPCRs, which account for more than one-third of all clinically used drug targets (13–16). Well-known instances include adrenal, dopamine, or prostaglandin E2 (PGE2) receptor systems (SI Appendix, Fig. S1 A and B) (1, 2, 4, 5, 17).
Mechanistic understanding of hormone function diversified at GPCR level requires the knowledge of how ligand engagement within the receptor pocket propagates to distinct intracellular rearrangements of the receptor for selective G protein/arrestin signaling. Crystal structures of AT1R have provided important mechanistic insights into the structural propagation paths underlying ligand-induced Gq/arrestin bias (18, 19). However, the structural basis for selective downstream signaling via plastic shaping of the hormone-binding pocket at the amino acid or conformational state levels, and mechanistic insights for the path delivering information for ligand binding to intracellular conformational states for GPCRs are still elusive. Moreover, biased ligands with the ability to selectively activate one of several downstream GPCR signaling pathways have the potential to avoid undesirable side effects (15, 20). Therefore, there is tremendous interest in understanding how the same or different ligands interact with their corresponding receptor pockets, and further transduce the binding signals to specific downstream effectors via selective propagating paths.
PGE2 is an unstable fatty acid hormone produced by cyclooxygenase (COX) and three PGE synthases (mPGESs) and is widely distributed throughout the human body (21, 22). In response to PGE2 stimulation, differential coupling of each EP receptor (EP1-EP4) to selective G proteins is fundamental to dictating their diverse functional outputs (SI Appendix, Fig. S1A) (17). In particular, EP4 can couple to both the excitatory Gs and the inhibitory Gi proteins to confer diverse functions such as renoprotection, cardioprotection, and antiinflammation, indicating the prevalence of signaling plasticity (23). For example, EP4 activation prevented the acute kidney injury (AKI) by reducing the expression levels of kidney injury molecule-1 (KIM-1) protein and other inflammatory factors such as IL-1β and TNF-α. Systemic administration of EP4 agonist reduced the serum creatinine (SCR) values and increased the survival rate in the rat model of AKI (Fig. 1A) (24). Moreover, the selective agonist ONO-AE1-329 induced EP4-Gs-cAMP signaling to promote the phosphorylation and plasma membrane translocation of aquaporin 2 (AQP2), thus providing an alternative strategy to treat NDI caused due to V2 vasopressin receptor (V2R) defects (25) (Fig. 1A). In addition, EP4-Gs signaling prevented the occurrence of glomerulonephritis, increased vasodilation, maintained vascular integrity, and showed cardioprotective effects (26–28). In contrast, the EP4-Gi pathway not only antagonizes cAMP accumulation, but also activates PI3K to inhibit the infiltration of eosinophils, resulting in an antiallergic inflammatory effect (29–31). Although the regulation of EP4 activity has therapeutic potential for the treatment of many diseases, the specific contribution of each specific G protein signaling, including both Gs and Gi, remains largely unknown (32–34). Therefore, it is important to understand the function and structural basis of the EP4-Gs and EP4-Gi signaling and to develop selective Gs/Gi-biased agonists of EP4, which will not only help dissect the function of biased EP4 signaling and the development of more precise drugs for a spectrum of diseases with EP4 as a promising therapeutic target, but also facilitate the elucidation of how a single hormone–receptor pair to recognize each other and transduce specific interactions via propagating path located in 7TM bundle for selective G protein subtype coupling in a broader context.
Fig. 1.
Functional characterization and biased properties of different EP4 agonists. (A) Schematic diagram showing that EP4 mediates membrane trafficking of AQP2, and restrains the KIM-1, inflammatory factor responses in AKI via the EP4-Gs pathway in renal. (B) Chemical structures of PGE2, Rivenprost, L902688, and KMN-80. The carboxylic acid ester of Rivenprost’s α chain can be hydrolyzed into carboxylic acid (35). (C) Schematic diagram of the effects of Rivenprost and L902688 on AKI mice. The Cdh16Cre+/Gnasflox/flox (Cko) mice represented the renal tubule-specific Gαs knockout mice, while the Gnasflox/flox (flox) mice were used as control. “√”, beneficial to treat AKI; “X”, not conducive to treat AKI; “XX”, accentuated the AKI. (D and E) The effects of Rivenprost on SCR (D) and BUN (E) levels of AKI mice (n = 8). Data are mean ± SEM. ***P < 0.001. Data were analyzed by Student’s t test. ###P < 0.001, &&&P < 0.001. Data were analyzed by one-way ANOVA with Tukey’s test. (F) The Emax of ΔBRET values of AQP2 trafficking stimulated with Rivenprost alone, Rivenprost and PTX (100 ng/mL), Rivenprost and NF449 (100 μM), or Rivenprost and MF498 (1 μM). Data are mean ± SEM, (n = 4).###P < 0.001. Data were analyzed by Student’s t test. ns, no significant difference. ***P < 0.001; **P < 0.01. Data were analyzed by one-way ANOVA with Tukey’s test. (G) Dextran permeability in response to L902688 (5 μM), Rivenprost (5 μM), Rivenprost (5 μM) and PTX (100 ng/mL), Rivenprost (5 μM) and NF449(100 μM), Rivenprost (5 μM) and MF498 (1 μM) in HUVECs. Data are mean ± SEM, (n = 4). ###P < 0.001; ns, no significant difference. Data were analyzed by Student’s t test. ***P < 0.001, **P < 0.01. Data were analyzed by one-way ANOVA with Tukey’s test.
Results
The Gs- and Gi-Biased Properties of Different EP4 Agonists.
We evaluated the biased properties of available EP4 agonists using endogenous ligand PGE2 as a reference. Among EP4 agonists, Rivenprost is a potential treatment for ulcerative colitis (36) and osteoporosis (37) (Fig. 1B and SI Appendix, Table S1). Whereas L902688 is Gi-biased, Rivenprost shows greater Gs-biased properties (SI Appendix, Fig. S1 C–F and Table S2). The separation window of the absolute value of biased factor β-value between Rivenprost and L902688 in human or mouse EP4 is more than 1 to 2, respectively, representing at least 10- to 100-fold difference once comparing the potency or efficacy on selectivity of two pathways between two ligands (38–41). Therefore, the Rivenprost and L902688 compound pairs may be good candidates for analysis of different pharmacological functions mediated by the Gs/Gi-biased properties of EP4.
We next investigated the effects of Rivenprost and L902688 on AKI induced by cisplatin, a widely used chemotherapy drug for various solid malignant tumors (42, 43). However, with high mortality and morbidity, a consistent therapeutic treatment option to prevent cisplatin-induced AKI in patients remains unavailable. Importantly, the administration of 0.1 mg/kg Rivenprost, but not 0.1 mg/kg L902688, was found to significantly reduce the levels of serum creatinine (SCR), blood urea nitrogen (BUN), kidney injury molecule-1 (KIM-1) protein, and the inflammatory factor related to AKI, including IL6, TNF-α, and MCP1 in the kidney induced by intraperitoneal injection of cisplatin (30 mg/kg) (Fig. 1 C–E and SI Appendix, Fig. S3 A–L). The concentrations of these two compounds were administrated according to previous studies (SI Appendix, Table S3) and further examined by our bioactivity measurements (SI Appendix, Fig. S2). Notably, specific deficiency of the Gαs protein in renal tubes by Cdh16Cre+/Gnasflox/flox (SI Appendix, Fig. S3M) abolished the beneficial effects of Rivenprost on AKI induced by cisplatin (Fig. 1 C–E and SI Appendix, Fig. S3 A–E). In addition, L902688 treatments aggravated the SCR, BUN, and inflammatory responses of AKI by Gs deficiency in Cdh16Cre+/Gnasflox/flox mice (SI Appendix, Fig. S3 F–L). We further examined the effects of EP2 antagonist TG6-129, EP3 antagonist DG-041, and EP4 antagonist GW627368X to inspect the contributions of these EP receptors in preventing AKI after Rivenprost administration. The results indicated that the Gs-biased protective effect of Rivenprost in vivo was mainly through EP4 signaling, but not EP2 or EP3 (SI Appendix, Figs. S3 N–P and S4 A–H). These results suggested the essential role of the Gs signaling in renoprotection of EP4 activity, and the Gs-biased EP4 agonist may have better therapeutic potential than Gi-biased EP4 agonist for AKI treatments.
In addition to AKI, activation of EP4 was recently proposed as an alternative therapy for NDI (25, 44). We then screened EP4 agonists to explore their efficacy in regulating AQP2 trafficking to the plasma membrane using BRET assay (SI Appendix, Fig. S4I) (45). Notably, all four EP4 agonists promoted the plasma membrane AQP2 trafficking, and the maximal response of AQP2 trafficking by Rivenprost was significantly higher than with other EP4 ligands (SI Appendix, Fig. S4 J and K and Table S4). The application of either the selective EP4 antagonist MF498 or the Gs inhibitor NF449 significantly blocked AQP2 trafficking, whereas administration of the Gi inhibitor PTX promoted it (Fig. 1F and SI Appendix, Fig. S4 L–N and Table S4). Therefore, the Gs and Gi pathways of EP4 inversely regulate AQP2 trafficking, and a Gs-biased agonist may have greater therapeutic potential to treat NDI by promoting AQP2 trafficking.
Activation of EP4 may have protective effects against mechanically induced lung injury, atherosclerosis, or other inflammatory diseases by enhancing the endothelial barrier function (46–48). Rivenprost, but not L902688, markedly reduced FITC-dextran permeability in a Gs-dependent manner, indicating the potential of a barrier-protective effect mediated by EP4 (Fig. 1G and SI Appendix, Fig. S4 O and P). Taken together, these series of experiments indicated that Gs-biased EP4 agonist may have greater therapeutic potential for AKI, NDI, and inflammation-related diseases.
Overall Structures of Agonist-Bound EP3-Gi, EP4-Gs, and EP4-Gi Complexes.
We determined the structures of EP4-Gs-Nb35 in complex with PGE2, L902688, or Rivenprost; and EP4-Gi-ScFv16 in complex with PGE2 or L902688 by cryo-EM at an overall resolution of 3.1 Å, 3.3 Å, 3.2 Å, 3.1 Å, and 3.5 Å, respectively. To dissect how PGE2 induced differential G protein subtype coupling through engagements with different EP receptors, we additionally solved the PGE2-EP3-Gi structure at 3.5 Å resolution (Fig. 2 A–C and SI Appendix, Figs. S5–S8 and Tables S5 and S6). These structures, together with our recently solved PGE2-EP2-Gs-Nb35 structure (49), enabled a paralleling comparison of the interaction modes of the PGE2 inside of ligand pockets of these receptors (SI Appendix, Fig. S9 A and B). The overall structures of EP4-Gs complexed with agonists were highly similar in the receptor region, with similar X-shaped extracellular loop 2 (ECL2), but shorter ECL1 and different conformations of ECL3 compared with previously solved EP2 structure (49) (SI Appendix, Fig. S9C).
Fig. 2.
Cryo-EM structure of agonist-bound EP4-Gs, EP4-Gi and EP3-Gi complexes. (A) Cryo-EM density of the Rivenprost-EP4-Gs complex (Left), the L902688-EP4-Gi complex (Middle) and the PGE2-EP3-Gi complex (Right). (B) Vertical positions comparison of the agonists in EP4-Gs complexes, EP4-Gi complexes, and the EP3-Gi complex. The brown dashed line indicates the insertion depth of the E-ring of L902688 in the EP4-Gs complex (lower line) or EP4-Gi complex (upper line). (C) Schematic diagram showing that the nature diversifies hormone function at three primary levels: 1) the yellow box: different receptors can recognize the same ligands, probably via variable amino acid composition at the receptor level; and 2) the green box: the same receptor–ligand pair can couple to different cellular effectors, in certain cases potentially dependent on different structural states with the same primary amino acid sequence for the particular receptor at the conformational level (additionally regulated by the presence of the effector or local effector concentrations); 3) the blue dotted box: different alternative splicing or posttranslational modifications of GPCR may serve as a putative layer for the generations of diverse signals downstream of a single hormone–receptor pair engagement.
A striking difference was the binding modes of the same ligand between Gs- and Gi-coupled EP4 structures. The EM density results indicated that L902688 was completely enveloped and embedded deeply inside the receptor in the EP4-Gs complex, whereas the same ligand was relatively shallow within the EP4-Gi complex (Fig. 2B and SI Appendix, Fig. S8 E and F). Moreover, local configuration of PGE2 fitted with EM density in the EP4-Gs/Gi complex also exhibited notable differences (SI Appendix, Figs. S8 A and B and S9 A and B). The binding poses of these ligands inside of the EP4 pocket were supported by computational simulations (SI Appendix, Fig. S9 D–I). The E-ring of L902688 was inserted 5.7 Å deeper toward the transmembrane core in the EP4-Gs complex compared to the EP4-Gi structure (Fig. 2B). Engaging with the same endogenous hormone PGE2, the primary sequence within the orthosteric pocket of EP2, EP3, and EP4 showed marked differences, which may play vital roles for their different downstream effector (G protein subtypes) coupling (Fig. 2C). Moreover, in particular for EP4, the different conformational states of the ligand pocket and corresponding propagating paths connected to the intracellular surface of the receptor, may underlay mechanisms governing the distinct downstream signaling of Gs vs. Gi in response to endogenous PGE2 or synthetic agonist L902688 stimulation (Fig. 2C).
Primary Sequence Differences in EP Receptor Pocket Contributed to G Protein Subtype Coupling Selectivity.
The four human PGE2 GPCRs, including EP1-EP4, recognized the same endogenous prostaglandin lipids and were connected to distinct G protein subtypes (50) (Fig. 2C). Because our EM map has better resolution with no artificial truncation or mutations of EP4 compared to the 3.3 Å PGE2-EP4-Gs structure solved in parallel (SI Appendix, Fig. S10 A–H), we used our 3.1 Å PGE2-EP4-Gs structural model for further structural analysis.
PGE2 showed different binding configurations in all solved structures of EP receptors. The positions of the E-ring and ω-chain of PGE2 in the EP2-Gs complex were lower than those in EP4-Gs/Gi and EP3-Gi complexes (Fig. 3A and SI Appendix, Fig. S9 A and B), though the α-chain of PGE2 was at similar horizontal heights and recognized by the conserved Y2.65TECL2R7.40 motif and L/M3.32L7.36 motif (Fig. 3A and SI Appendix, Figs. S10 I–M and S11 A and B and Table S7). Totally 16 to 19 amino acids, with only six amino acids identical in primary sequence (Fig. 3A), participated in the interactions between EP receptors and PGE2. The different amino acids in primary sequence could serve as important determinants for selective interactions of the PGE2 within the ligand pocket of different EP receptors, and then connecting them to specific downstream G protein subtype couplings. For instance, the carboxy group of the PGE2 α-chain engaged a hydrogen bond with the S281.39 of EP2, which was replaced by the allelic P1.39 in EP3 or EP4. Compared with the bulged Q1032.54–Q3397.46 pair of EP3, the T822.54–S3087.46 pair of EP2 had smaller side chains, providing enough space for penetration of the E-ring and ω-chain of PGE2 inside the receptor core and establishing stable hydrogen bonds with the hydroxyl of PGE2. Moreover, whereas the steric packing by S1203.36-L3047.42 of EP2 restricted the placement of the ω-chain of PGE2 at a deep position, the allelic G1413.36–A3357.42 pair of EP3 and S1033.36–A3187.42 pair of EP4 provided less entrance and the ω-chain assumed a shallower pose (Fig. 3B and SI Appendix, Fig. S9 A and B). Alanine mutation of T822.54, T1233.39, S3087.46 of EP2 (49); Q1032.54, Q3397.46 of EP3 (51), and T692.54, P3227.46 of EP4 (Fig. 3C) significantly affected the Gs/Gi activity of different EP receptors. These results provided important evidence that the difference of the primary amino acid sequence in the ligand pocket was important for the Gs/Gi coupling selectivity of EP receptors by accommodating distinct ligand-binding configurations.
Fig. 3.
Different binding modes of endogenous PGE2 in distinct EP receptor subtypes. (A) Comparison of PGE2 binding sites between EP4, EP2, and EP3. The L/M3.32-L7.36 hydrophobic interactions and the Y2.65-TECL2-R7.40 motif are shown as blue- and red-filled circles, respectively. The F1023.35 engaged with PGE2 in the EP4-Gs complex is highlighted by orange-filled circles. Residues without interaction are shown as blank circles. Residues circled by the red dashed line constituted the different interaction modes with PGE2 in EP2 or EP3 compared to those in EP4, and the combined mutation of the allergic positions in EP4 that mutated to those of EP2 (M3.32T3.36L7.42S7.46) or of EP3 (Q2.54G3.36S3.39Q7.46) are shown. The F102, S103, and P322 in EP4 are surrounded by a blue dashed box. Ballesteros–Weinstein residue numbers are provided for reference. (B) The ω-chain of PGE2 showed differences in the ligand interaction pattern in EP receptor structures with magenta in EP2-Gs, blue in EP4-Gs, orange in EP4-Gi, and green in EP3-Gi. The red dashed line represents the same horizontal plane. (C) Mutagenesis studies showing the effects of residues on ligand-binding pocket. Heatmap of the ΔpEC50 (pEC50 of mutant - pEC50 of wild type) and Emax (normalized to 100% wild type) showed differences in the Gαs-Gγ/Gαi-Gγ dissociation assay. Data are mean ± SEM, (n = 3). (D) Comparison of the biased properties of the combined mutations of EP4. The bias factor (β value) of EP4-MTLS or EP4-QGSQ was calculated using the EP4-WT as the reference by the method in the Materials and Methods of Gαs-Gγ/Gαi–Gγ dissociation assay. Data are mean ± SEM, (n = 3).
Distinct PGE2 Configurations within the Ligand Pocket in EP4-Gs and EP4-Gi Complex Structures.
Despite binding to the same EP4 receptor, the binding pose of PGE2 in the EP4-Gi and EP4-Gs showed marked differences (overall RMSD of ~1.2 Å for average atoms) (SI Appendix, Fig. S11C). Compared to the EP4-Gs complex, the carboxyl end of the α-chain of PGE2 in the EP4-Gi complex was rotated ~60° toward TM7, losing its hydrogen bond with T168ECL2. The E-ring of PGE2 in the EP4-Gi complex was rotated upward by ~60°, forming a new H-bond with T692.54, which was not observed in the EP4-Gs complex (SI Appendix, Fig. S11 A–C). Accompanied by E-ring rotation, the ω-chain swing ~18° toward TM7 and the atoms of C18-C20 were moved upward, forming additional contacts with I3157.39 but lost the hydrophobic interaction with F1023.35 in EP4-Gi compared with the EP4-Gs complex (SI Appendix, Fig. S11 C and D). Consistent with this observation, whereas T692.54A, I315A, and I315S showed greater effects on Gi activity more than Gs activity of EP4 in response to PGE2 stimulation, F1023.35contributed to the Gs more than Gi activity of EP4 in response to both PGE2 and synthetic agonist L902688 stimulation (Fig. 3C and SI Appendix, Fig. S11 E and F and Tables S8 and S9). These structural observations and pharmacological characterizations suggested that specific structural features within the orthosteric ligand-binding pocket of single receptor EP4 determine its downstream G protein subtype coupling in response to endogenous agonist PGE2 stimulation. Notably, because the current models of the PGE2 are on the edge of the noise of the cryo-EM maps and carboxyl groups tend to be sensitive to radiation damage, the real binding pose of PGE2 to EP4 may be different from the current models, which could be evaluated by other biophysical methods or technologies.
Orientations of the ω-Chain Correlated to G Protein Coupling Selectivity of the EP Receptor.
We noticed that the ω-chain in the EP receptor structures showed marked differences, including both the orientation and depth of insertion (49, 51) (Fig. 3B and SI Appendix, Fig. S11 G and H). In the PGE2-EP3-Gi complex structure, the ω-chain of PGE2 extended toward TM6 of the receptor and directly interacted with the “toggle switch” W2956.48 (Fig. 3B). However, the ω-chain of PGE2 was inserted deeply inside EP2 and had an orientation toward TM3 and TM7. The insertion depth of the ω-chain in both the PGE2 bound EP4-Gs and EP4-Gi complexes was shallower than that of PGE2-bound EP2, but deeper than that of PGE2-bound EP3 (Fig. 3B). Importantly, the binding orientations of the ω-chain in the EP4-Gs and EP4-Gi complexes were between those of EP2 and EP3. The ω-chain diverged at an angle of 31° or 35° in the EP4-Gs complex in correlation to the ω-chains of the PGE2-EP3 and PGE2-EP2 complexes, respectively (SI Appendix, Fig. S11G). Interestingly, the ω-chain in the EP4-Gi complex was oriented closer to the PGE2-EP3 complex, was separated by angles of 23° and 49°, correlating to the ω-chains of the PGE2-EP3 and PGE2-EP2 complexes, respectively (SI Appendix, Fig. S11H). Taken together, the conformational states of the ω-chain were separated by different orientations divided by the Gs/Gi transducer coupling abilities. Therefore, we described the PGE2-binding modes in different EP receptors as the “goldfish” model, that is, the α-chain and E-ring of PGE2 form the “trunk part” with relatively conserved interactions among the EP receptor members. The ω-chain serves as the tail part, which wiggles toward different orientations (SI Appendix, Fig. S11I).
It is worth to note that human EP2 specifically interacts with Gs protein, whereas human EP3 selectively couples with Gi protein (Fig. 2C). Combined mutations of EP4 pocket residues with allelic residues in EP2 (EP4-MTLS) rendered EP4 more biased toward Gs signaling, whereas the allelic mutations EP4 to EP3, by combined mutations of EP4-QGSQ enabled EP4 to be more biased toward Gi (Fig. 3D and SI Appendix, Fig. S11 J and K). These results indicate that the binding modes of the ω-chain of the endogenous ligand PGE2, as well as the different primary sequence between different EP receptor subtypes, are important determinants of EP receptor G protein coupling selectivity.
Structural Basis for Selective EP4 Agonism.
Selective activation of EP4 has therapeutic potential in the treatment of a spectrum of diseases, including acute and chronic kidney failure (24), glomerulonephritis (26), osteoporosis (37) and hypertension (27) etc. The cryo-EM structures of EP4-Gs complexes bound with Rivenprost and L902688 may provide structural basis for their selectivity.
Similar to PGE2, Rivenprost and L902688 consist of three parts, including the α-chain, E-ring, and ω-chain, which interact with three subpockets within the EP4 orthosteric sites, respectively. Both polar and hydrophobic groups are accommodated in the subpocket A of EP4 because of its plasticity, mostly via interactions with the Y802.65-T168ECL2-R3167.40 motif (SI Appendix, Fig. S12 A–E). The region B subpocket of EP4 accommodated the E-ring of Rivenprost or L902688. The carbonyl group of the E-ring of Rivenprost or L902688 forms a hydrogen bond with T762.61 of EP4 (SI Appendix, Fig. S12 C and D), which was substituted by V2.61 in all other EP family members (SI Appendix, Fig. S12E). Notably, region C is the tail region of the “goldfish” model, which shows high diversity between members of the EP family (SI Appendix, Fig. S11I). The polar residue S1033.36 donates H-bonds to both the -COCH3 group of the ω-chains of Rivenprost, whereas S1033.36 was substituted by G3.36 in EP1 and EP3 (Fig. 4 A and B and SI Appendix, Fig. S12 E–G). Moreover, EP4 had a smaller A3187.42 compared with the allelic L3047.42 in EP2, which enabled the positioning of the ether group (–COCH3) of Rivenprost toward TM3 and reached L993.32 in the Rivenprost-EP4 complex (Fig. 4 A and B). Importantly, allelic mutations of T762.61V or A3187.42L in EP4 caused a significant decrease in Rivenprost- or L902688-induced EP4 activity. Moreover, the reverse mutations V892.61T or L3047.42A in EP2 also decreased its response to Taprenepag, a selective agonist (Fig. 4C and SI Appendix, Fig. S12 H and I and Table S10). The S1033.36A mutation of EP4 (related to that of EP1 and EP3) caused significant reduction in EP4 activity in response to both Rivenprost and L902688 stimulation. Consistent with this observation, allelic mutations G1413.36S of EP3 decreased its activity in response to selective agonist Sulprostone (SI Appendix, Fig. S12 H and I and Table S10), highlighting that these key regions confer agonist selectivity.
Fig. 4.
The selective binding pockets of synthetic agonists in EP4. (A) Cutaway view of Rivenprost in EP4 receptors. The residues T762.61 and S1033.36 form a hydrogen bond with Rivenprost. (B) Sequence alignment of EP family members in ligand pocket. Note that T762.61 was only found in EP4; L3047.42 was only found in EP2; S3.36 is conserved in EP2 and EP4, but replaced by G3.36 in EP1 and EP3. (C) The decreasing folds of the equivalent mutations of T/V2.61, S/G3.36 and A/L7.42 between EP subfamily receptors in ligand binding. Data are mean ± sem.
Collectively, key amino acids in region B and region C of the three EP4 subpockets, in particular small and polar residues including T762.61, A3187.42, and S1033.36 which enabled H-bond formation and accommodated a larger group at region C at A3187.42 -S1033.36, contributed to the selective binding of agonists to EP4.
A G-Protein-Subtype-Dependent Agonist-Binding Pocket in EP4.
One striking and unexpected observation is the different binding modes of the same agonist (for both L902688 and PGE2) in the EP4 receptor in complex with different G protein subtypes (Fig. 5 A and B and SI Appendix, Fig. S9 A and B). Notably, the L902688 could be positioned in the EM map of L902688-EP4-Gs by two orientations and the MD analysis suggested both binding modes are stable (SI Appendix, Figs. S9 D and E and S13 A–C). In the binding mode 1, the tetrazole ring faced extracellular side and form polar contacts with Y802.65, T168ECL2, and R3167.40 (SI Appendix, Fig. S13A). Because replacing these polar residues with hydrophobic residues, including mutations of Y802.65F, T168ECL2V and R3167.40M, significantly decreased Gs activity downstream of EP4 in response to L902688 stimulation, we choose mode 1 for further structural analysis (SI Appendix, Fig. S13 D–F and Tables S8 and S9). The exact conformation could be further evaluated by more detailed Cryo-EM structure analysis in the future.
Fig. 5.
A G-protein-subtype-dependent agonist binding mode in EP4. (A) The distinct configurations and positions of L902688 in the EP4-Gs and EP4-Gi complex. The red arrows indicate the movement of corresponding L902688 regions. Hydrogen bonds and polar interactions are shown by red dashed lines. (B) Interactional differences of EP4 in L902688 pocket between Gs and Gi signaling complex. (C) Schematic diagram of the ternary complex model (52–54). The model involves the interaction of the hormone (ligand), the receptor (EP4), and the downstream effector (G protein). The KLo represents the ligand interacting with the free receptor to constitute a low-affinity state, which then coupled with G protein, representing as KC. The KG represents that the free receptor coupled with G protein, and the KHi represents the ligand binding to such a precoupled receptor-G protein complex. The scheme also shows the formation of the high-affinity ternary complex for effective signaling. (D and E) The competition binding curves of L902688 in response to wild-type EP4 receptor, EP4-Gαi/EP4-Gαs fusion protein (EP4-Gi/EP4-Gs) and EP4-Gαi/EP4-Gαs with A318L mutation (EP4-A318L-Gi/EP4-A318L-Gs). L902688 showed two different binding states to EP4-Gi/EP4-Gs, containing a high-affinity state (KHi) and a low-affinity state (KLo). Gα fusion induced a fraction of EP4 into the high-affinity state, at 34% and 47% in EP4-Gs and EP4-Gi, and a large shift of KHi. In contrast, the KLo of both EP4-Gs and EP4-Gi are similar to wild-type EP4. A318L mutant weakened L902688 KHi and KLo, both in EP4-Gs and EP4-Gi. Importantly, the high-affinity fraction reduced from 34% in EP4-Gs to 9% in EP4-A318L-Gs chimeric mutation, while EP4-A318L-Gi chimeric mutation showed no reduction. Data are mean ± SEM, (n = 3).
Importantly, in contrast to the deep insertion inside of the 7TM bundle with an overall inversion C configuration in the EP4-Gs complex, L902688 assumed a Z shape and was bound to the orthosteric pocket with a much shallower position in the EP4-Gi complex (Fig. 5A and SI Appendix, Fig. S13G). These two binding modes of L902688 shared 11 common interacting residues but showed 10 different contacts (Fig. 5B and SI Appendix, Figs. S12C and S13F).
The α-chain of L902688 moved upward by 6.2 Å in the EP4-Gi complex compared with that in the EP4-Gs complex. The tetrazole ring constituted a polar network with the Y2.65TECL2R7.40 motif in the L902688-EP4-Gs complex structure. In contrast, the tetrazole ring of L902688 only parallelly packed against T792.64 in the EP4-Gi complex (Fig. 5A and SI Appendix, Fig. S13 E and F). In the EP4-Gi complex, the E-ring of L902688 was lifted by 5.7 Å, occupying the position of the tetrazole ring of the α-chain for L902688 in the EP4-Gs complex (Fig. 5A). The hydrogen bond between the carbonyl group of the E-ring and T762.61 was substituted by hydrophobic interaction with W169ECL2 (Fig. 5A and SI Appendix, Fig. S13 H and I). An upward rotation of the indole ring of W169ECL2 was observed in the EP4-Gi complex to accommodate these different binding poses (SI Appendix, Fig. S13 H and I). Accompanied by the upward movement of the α-chain and the E-ring, the phenyl ring of the ω-chain of L902688 was moved upward by ~2 Å and traverse by 4.4 Å; thus, the ω-chain was removed from the pocket created between TM3 and TM7, losing the interactions with S1033.36, A3187.42, and P3227.46 in EP4-Gi compared to the EP4-Gs complex (SI Appendix, Fig. S13 H and I).
These observations suggest that distinct binding modes of L902688 may determine the coupling difference observed with downstream G protein subtypes. Consistent with this hypothesis, mutations of W169ECL2L substantially impaired both Gs and Gi activity of EP4 in response to L902688 stimulation, with a larger effect on Gi. Conversely, mutations of S1033.36A showed greater effects on L902688-induced Gs activity than that for Gi (Fig. 3C and SI Appendix, Tables S8 and S9).
Our structures revealed that the same ligand could bind to receptor-G protein complexes with distinct configurations, which are determinants of G protein subtype selectivity. These different binding modes could occur with free receptors before G protein binding, or were created after G protein association alone. The classical ternary complex model states that the G protein may play an important role in shaping the ligand-binding pocket of the receptor (Fig. 5C) (52, 53, 55); however, direct evidence for distinct ligand-binding poses induced by coupling of different G proteins has not been achieved previously at atomic resolution. Next, we combined the radio-ligand binding assay and the EP4-Gs/Gi chimeras to inspect the agonist-induced high-affinity states for EP4-Gs/Gi protein coupling (53–56). Notably, two ligand-binding sites with different affinities were identified in the competition binding curves of both the EP4-Gs and EP4-Gi chimeras in response to L902688 engagement, competing with [3H]-PGE2. The goodness of fit of the competition binding curves by one or two sites was evaluated using the residual variance with the ratio of the sum of squares of residuals (RSS) divided by degrees of freedom (df) (Fig. 5D and SI Appendix, Fig. S13 J and K and Table S11) (57). The low affinity (KLo) of both EP4-Gs and EP4-Gi is similar to EP4 wild type, suggesting the affinity state of KLo represents the binding of the L902688 with the EP4 receptor alone. Conversely, the high-affinity (KHi) state should represent the state of binding of the L902688 to the EP4-Gs or EP4-Gi complexes. Importantly, the L906288-induced high-affinity fraction was reduced from 34% in EP4-Gs to 9% in EP4-A318L-Gs chimeric mutation. In contrast, L906288-induced high-affinity fractions in EP4-Gi chimera and EP4-A318L-Gi chimeric mutation are quite similar. These results are consistent with our hypothesis that introducing a larger residue at the A318 position (A318L) could specifically block the EP4-Gs ligand-binding pocket but had not much effect on the EP4-Gi ligand-binding pocket (Fig. 5 D and E). We therefore speculated that precoupling of EP4 to Gs or Gi may have significant effects on their ligand pocket conformations.
Distinct Pocket Configurations and Propagating Paths of EP4 Gs- and Gi-Biased Activation Mechanisms.
In the Rivenprost-EP4-Gs and L902688-EP4-Gs complex structures, the agonists were inserted approximately 6 Å deeper into the transmembrane core compared with that of the EP4-Gi complex structures, by placing the ω-chain between TM3 and TM7 (SI Appendix, Fig. S13L). The ω-chain of Rivenprost and L902688 was packed with P3227.46 and S1033.36 in the EP4-Gs complex (Fig. 6 A and B). In contrast, L902688 sat in a much shallower position and did not contact either P3227.46 or S1033.36 in the L902688-EP4-Gi complex (Figs. 5A and 6B). Importantly, mutations in S1033.36A or P3227.46A mainly impaired Gs but not Gi activity downstream of EP4 activation as observed in G protein dissociation assays in response to both L902688 and PGE2 (Fig. 3C and SI Appendix, Tables S8 and S9). These results, along with the mutational effect of F102A (Fig. 3C), indicate that specific structural features with in the ligand pocket, including that of F1023.36, S1033.36, and P3227.46, play potential common roles in determination of Gs- vs. Gi-biased property of different EP4 agonists. These specific structural features could serve as potential structural guidance for the development of biased ligands of EP4.
Fig. 6.
Distinct configurations and propagating paths of Gs and Gi-bias activation mechanisms of EP4. (A) Schematic diagram representing the propagating paths that contribute to the Gs/Gi bias decision mechanism. The ligand pocket in L902688-EP4-Gs (magenta) was deeper than that in L902688-EP4-Gi (blue). The propagating paths started with residues constituting two different ligand-binding pockets and extended to the Gs/Gi protein interface region. The residues in TM2, TM3, TM6, and TM7 are shown as brown, blue, magenta, and green circles, respectively. (B and C) The different structural rearrangement of the ω-chain of L902688 in the EP4-Gi and EP4-Gs complex. (D–F) Key residues along the propagating paths connected the ligand-binding pocket to the cytoplasmic side by comparison of L902688-EP4-Gs (green); L902688-EP4-Gi (blue); ONO-AE3-208-EP4 complexes (gray). (G) Effects of the residues’ mutations along the propagating path in response to L902688. The heat map is filled according to the ΔpEC50 and Emax (100% wild type). “X”, no detectable signal. Data are mean ± SEM, (n = 3). (H and I) The competition binding curves of EP4-Gαs (H) and EP4-Gαi fusion protein (I) with V320A, N321A, and D325R mutants in response to L902688. Data are mean ± SEM, (n = 3). (J) The FlAsH-BRET responses of EP4 intracellular region between the ICL2 (S2) or TM7 terminus (S4) and the C terminus in response to the Gs/Gi signal. Data are mean ± SEM, (n = 3 or 4). The concentration (M) of Rivenprost and L902688, and the mutants along the propagating path are shown left. The Gs and Gi signal’s effects were surrounded by red and green dotted lines, respectively. **P < 0.01; ns, no significant difference. Data were analyzed by one-way ANOVA with Tukey’s test.
After aligning with TM3, binding of L902688 in the EP4-Gs complex enabled the inward movement of TM7, as observed in a 2.5 Å position shift of P3227.46 and G1063.39 compared to the inactive ONO-AE3-208-EP4 structure (Fig. 6 B and C). Moreover, the binding of L902688 in the EP4-Gi complex induced a further inward rotation of TM7, with a 2.3 Å shift of P3227.46 (Fig. 6C). The interaction between the phenyl ring of L902688 and S1033.36/P3227.46 promoted the packing between V3207.44 and C2846.47 and the building of the polar network constituted by the side chain of C2846.47, the main chain carbonyl of V2806.43, and the side chain of N3217.45. These rearrangements further led to the rotation of the carboxylate group of D3257.49, enabling a polar interaction with the main chain carbonyl of N3217.45 and readjustment of the packing between V2816.44, I1103.43, Y3297.53, and L612.46 (Fig. 6 D–F and SI Appendix, Fig. S14 A and B). These structural rearrangements were propagated to the changes in conformational locks, including the D7.49P7.50xxY7.53 and the ion lock E3.49R3.50Y3.51 motifs on the cytoplasmic side. Moreover, the packing patterns of R1173.50, Y3297.53, M2706.33, and L2746.37showed notable differences in Gs or Gi-coupled EP4 structures (Fig. 6 E and F). These structural alterations enabled a ~1.4 Å outward tilting of TM6 in the L902688-EP4-Gi complex structure compared to that in the Rivenprost/L902688-EP4-Gs complexes (SI Appendix, Fig. S14C), which enabled accommodation of different G protein subtype C-tails (SI Appendix, Fig. S14D).
Notably, introducing a bulky side chain by replacing G1063.39 with R abolished EP4 activation as the change blocks the structural rearrangement between TM7 and TM3 (58). Moreover, mutations along the propagating path that connected the ligand-binding pocket to the cytoplasmic region and showed greater conformational differences between the Gs- and Gi-coupled EP4 receptors bound with the same agonist, such as C2846.47A, V3207.44A, N3217.45A, P3227.46A, and D3257.49A, significantly impaired Gs activity more than influence on Gi activity downstream of EP4 activation. D3257.49R mutation completely abolished Gs activity, whereas C2846.47R mutation abolished Gi activity downstream of EP4 activation (Fig. 6G and SI Appendix, Fig. S14E and Tables S8 and S9). We then inspected the contributions of the proposed conformational propagating path on high-affinity states of the agonist-EP4-G protein ternary complex exploiting the radio-ligand assay toward wild-type EP4 and EP4-Gs/Gi chimeras. Importantly, the L906288-induced high-affinity fractions of EP4-V320A-Gs, EP4-N321A-Gs, and EP4-D325R-Gs were reduced from 34 ± 6% in EP4-WT-Gs to 18 ± 4%, 24 ± 2% and 13 ± 3%, respectively, but showed significantly less effects on the high-affinity fraction of EP4-Gi chimeras (from 47 ± 10% to 37 ± 5%, 44 ± 5% and 34 ± 6% respectively (Fig. 6 H and I and SI Appendix, Fig. S14 F–H and Table S11). These results suggest that the conformational differences between key propagating path residues along TM3 and TM6-TM7, such as microswitches including V3207.44/C2846.47/N3217.45 and D3257.49, played important roles in the Gs/Gi coupling selectivity (Fig. 6A).
Connecting the Structural Propagating Path to Cytoplasmic Configuration of EP4 for G-Subtype Coupling Selectivity.
To provide further insights into the structural rearrangements localized in cytoplasmic side governing selective G protein subtype coupling, we generated a series of FlAsH-BRET sensors by incorporating FlAsH motifs into the three intracellular loops (ICLs, S1-S3) and TM7 terminus (S4), and the luciferase domain at the C terminus (59–63) of EP4 (SI Appendix, Fig. S14 I and J). We noticed that 10−6 M or 10−7 M Rivenprost only activate the Gs, but not the Gi downstream of EP4 wild type. Conversely, 10−6 M L902688 showed only Gi activity, but not Gs, after engagement with D325R mutant of EP4. In addition, 10−8 M L902688 or 10−6 M L902688 only activate D325A or F102A mutants of EP4, respectively (SI Appendix, Figs. S1C, S11F, and S14E). We therefore exploited the combined conditions of these agonists’ concentrations and appropriate EP4 mutants vs. EP4 wild type to profile potential structural features located at the cytoplasmic side of EP4 correlating to Gs vs. Gi coupling. Notably, the Gs activation conditions were specifically reported by the S2 sensor, which suggested that the separation between the C terminus and ICL2 (S2) in response to agonists stimulation. On the contrary, the Gi activation conditions were captured by the S4 sensor, which indicated the separation between the TM7 terminus (S4) and C terminus (Fig. 6J and SI Appendix, Fig. S14 K and L). Moreover, two propagating path mutations, the V320A and N321A, which mainly decreased Gs but not Gi signaling of EP4, showed significant reduction in the S2 signal in response to 10−6 M Rivenprost stimulation. However, they showed no observed changes on conformation reported by S4 compared with D325R mutation in response to 10−6 M L902688 stimulation (Fig. 6J). These results suggest that ligands of EP4 can induce specific intracellular configuration, which were associated with the residues in the ligand pocket and propagating path, to determine the Gs/Gi coupling selectivity.
Common Features and Determinants of Gs/Gi Coupling Selectivity of EP4.
Three EP4-Gs complexes have larger buried surfaces than that of the two EP4-Gi complexes (SI Appendix, Table S12). Although 22 to 25 residues of EP4 are involved in direct interaction with the Gs protein, only 15 to 16 residues directly contact Gi (SI Appendix, Fig. S15A). Importantly, 13 residues including F542.39, E1163.49, R1173.50, Y127ICL2, and M2135.68 of EP4 constitute the common structural determinants of EP4 for coupling of both Gs and Gi (SI Appendix, Fig. S15 B–E and Tables S8 and S9).
The different interaction modes between EP4 and biased agonists were propagated to distinct conformational changes located at the cytoplasmic sides of the TM bundles, These structural alterations enabled a larger separation of TM3 and TM6, and created a larger 7TM cavity in EP4-Gi complexes than EP4-Gs complexes (SI Appendix, Fig. S14D), which was distinct from previous results for GCGR-Gs/Gi complex structures (64), where a larger cavity was required to accommodate Gs coupling than Gi (SI Appendix, Fig. S15F). Compared with the EP4-Gs complex, the α5 Helix of Gαi was rotated ~11° toward ICL2 (SI Appendix, Fig. S14D), enabling a “hook” configuration of the α5 helical end bending toward TM6 in the EP4-Gi complex, similar to previously solved CB1-Gi or GPR97-Go complex structures (63, 65). However, the EP4-Gs complex showed marked structural differences compared to those of classical Gs-coupled receptors, such as β2AR-Gs and DRD1-Gs complexes (64, 66, 67) (SI Appendix, Fig. S15 G and H), with the α5 Helix of the Gαs inserting into the cavity between ICL1 and Helix 8 by a spiral extension (SI Appendix, Fig. S15 I–K). These structural comparisons indicated that changes in TM6 configuration play critical roles in accommodating different G protein subtypes. In addition, the distinct interactional packing modes between EP4 and Gαs or Gαi in G protein interface also contributed to the selective G protein coupling of EP4 (SI Appendix, Fig. S16 A–F). Taken together, different agonists of EP4 could allosterically regulate the conformational states of the cytoplasmic ends of TM3 and TM5-TM7 to selectively regulate its coupling with the Gs or Gi subtype. Although Y130ICL2 is a determinant for specific Gi contact, E2676.30 and T3358.49, V3368.50 contributes to selective Gs coupling in the EP4-Gs and EP4-Gi complexes.
Determinants for Gs/Gi Selectivity in the EP Family Receptors.
Comparisons of interfaces and sequence alignment results suggested that Y552.40 (H682.40 in EP2 and L892.40 in EP3), F2175.72 (S2295.70 in EP2 and K2635.72 in EP3), and E2676.30 (T2776.30 in EP3) are different residues in EP3 structures, indicating that these residues may participate in Gs/Gi coupling selectivity (SI Appendix, Figs. S15A and S16 G–K). Importantly, the substitution of F2175.72 in EP4 or S2295.70 in EP2 by K2635.72 in EP3 caused loss of its favored cation-π or polar interactions with Gs, and introduced repulsion at this equivalent position in EP3 (SI Appendix, Fig. S16 G and H). The replacement of E2676.30 in EP2 and EP4 by T2776.30 in EP3 caused loss of the preferred charge interaction with R385H5.17 (SI Appendix, Fig. S16 G and H). Further, allelic mutations in EP2 and EP4 in corresponding positions significantly decreased Gs activity of EP2 or EP4 (SI Appendix, Fig. S16 L–N and Tables S8 and S9). Therefore, these three residues, and TM2, TM5, and TM6 are determinants of the Gs coupling activity of EP family members.
The EP3 and EP4 coupled with Gi via distinct mechanisms, shared few common structural features. Though the interaction mediated by Y130ICL2 with Gi is specific for EP4 (SI Appendix, Figs. S16C and S17 A and B), the M2706.33/T2806.33 of EP4/EP3 underwent external rotation to create the space to accommodate the insertion of the α5 Helix of Gi, but not for the corresponding allelic residues H2626.33 of EP2 (SI Appendix, Fig. S17C). These structural observations were further supported by mutagenesis effects (SI Appendix, Figs. S16 M and N and S17D and Tables S8 and S9).
Collectively, our studies identified common structural determinants of EP receptors for Gs coupling, including allelic residues Y/H/L2.40, S5.70/F5.72/K5.72, E/T6.30 in TM2 and TM5-TM6. In contrast, the EP3 and EP4 coupled with Gi via distinct mechanisms, whereas the YICL2 and M6.33 in ICL2 and TM6 of EP4 played important roles in Gi coupling, the extended helix of TM6 in EP3 provided important Gi interface for efficient coupling (SI Appendix, Fig. S17 E and F).
Discussion
In this study, using the PGE2-EP receptor system as a model (49, 51), we provide preliminary insights into how single hormones are recognized by different GPCR members to enable signal transduction to distinct downstream effectors. Importantly, the residues constituting the ω-chain-binding pockets showed notable diversity, enabling the separation of ω-chains by different orientations related to their Gs/Gi transducer coupling abilities. Among the 15 residues involved in ω-chain recognition in the EP receptors, 11 showed sequence-specific differences. Importantly, replacements of the ω-chain recognition residues of EP4 to the corresponding EP2 or EP3 allelic residues, created a more Gs-biased EP4 mutant or more Gi-biased EP4 mutant, respectively (Fig. 3 D and E). Therefore, the primary sequence differences in the ligand-binding pocket are determinants of PGE2 recognition by different EP receptors that generate selective Gs/Gi subtype coupling abilities (Fig. 2C).
One particularly important question is how a single hormone–receptor pair generate diverse signals. Our works suggested that two hypotheses may account for this diversity: 1) The single ligand-receptor interaction mode indices two different intracellular conformational states; and 2) the receptor assumes different conformational states to recognize the same ligand, and further generates variable intracellular conformations through different structural propagation paths (Fig. 2C). Recent structural studies of the glucagon-GCGR complexes suggested the first model that the specific coupling of different G protein subtypes was governed by individual conformational states of intracellular loops with the same glucagon interaction mode (64). In contrast, our studies indicated that the EP4 receptor recognized both endogenous PGE2 and the synthesized agonist L902688 through two different conformations, and correlated their selective coupling to Gs or Gi effectors. The PGE2 assumes two different poses with distinct ω-chain conformations in the EP4-Gs or EP4-Gi complexes. The different interaction patterns of PGE2 within the EP4 ligand pocket caused positional shifts and rotamer changes of specific residues in 7TM of EP4, which determined its downstream effector coupling selectivity. This conclusion was particularly supported by the observation that the two distinct pockets of EP4 accommodate L902688 to couple to Gs or Gi downstream effectors. Notably, the E-ring of L902688 was lifted by ~6 Å in the EP4-Gi complex compared with the EP4-Gs complex, and only 11 out of 21 total contact residues alone showed common interactions with L902688 in different binding states. The biochemical data supported these proposed mechanisms. Collectively, the above results demonstrate that the distinct binding modes of the EP4 pocket via a single hormone or synthetic agonist determine the selective coupling between the Gs and Gi effector proteins via a common structural mechanism with defined key motifs.
Mechanistically, different ligand–receptor pocket-binding modes dictate conformational states of the intracellular region of the receptor to enable selective downstream effector coupling, which lies at the heart of receptor biology and biased drug development targeting GPCRs. Results from recent studies have indicated that the potentially different interactions with L1123.36 and Y2927.43, and the subsequent conformational change in the side chains of N1113.35, N2957.46, and N2987.49 serve as the key microswitches that determine the preferential recruitment of Gq/β-arrestin to AT1R (18, 19). However, the mechanism by which the structural propagation path enables a receptor to differentiate Gs/Gi signaling remains unclear. Our study shows that the shallower ligand interaction mode of the ligand in the EP4-Gi complex compared to the EP4-Gs complex caused the overall inward movement of TM7 and a larger separation of the cytoplasmic end of TM3 and TM6 to accommodate the C-terminal end of the α5 Helix of Gi. Consistent with this observation, our FlAsH-BRET biosensor indicated that separation between TM7 end and C terminus of EP4 is a structural feature reporting Gi coupling conformation, strengthened the key role of TM7 movement in Gi/Gs selectivity. In particular, direct interactions of the agonist with S1033.36 and P3227.46 in the pocket, as well as the rotamer changes for R1173.50, V3207.44/C2846.47/N3217.45/D3257.49, and Y3297.53/M2706.33/L2746.37 were identified as key microswitches for the conformational propagating path that connects ligand binding to the coupling of different G protein subtypes. Therefore, our results delineate how nature has delicately designed the conformational propagation path of a GPCR for translating ligand binding to selective cellular functions, and this property may be widely used in other ligand-GPCR systems.
In addition, it has been reported that the alternative C-terminal tail splicing isoforms of Bos EP3 determined its G-protein subtype coupling specificity (68, 69). For human EP3s, we can only detect Gi activation downstream of human EP3 isoform 4 in response to PGE2 stimulation. In contrast, human EP3 isoform 6 could activate the Gs, whereas the isoform 7 could activate both the Gs and Gi signaling (SI Appendix, Fig. S17 G–J). In adrenergic receptor system, the PKA-mediated phosphorylation of β2-AR was reported to decrease the affinity of the receptor for Gs and increase its affinity for Gi (1, 3, 70). Furthermore, the PKA-mediated phosphorylation of β1-AR was reported to enable the transition between the Gs to Gi signaling (71). Therefore, the different alternative splicing or posttranslational modifications of GPCR may serve as another hypothesis for generating diverse signals downstream of a single hormone–receptor pair engagement (Fig. 2C).
In addition to insights into the structural basis underlying the biased preference for signaling via hormones and GPCRs, our studies showed that 1) notably, the functional analysis of Gs vs. Gi-biased signaling suggests that an EP4 agonist with greater Gs bias showed better therapeutic potential to treat AKI, NDI, and vascular inflammation. Though with potential side effects for abdominal discomfort and increased risk of cardiovascular disease, the therapeutic approaches using EP4-biased agonists may be developed as precise medicine for specific patients, such as patients with NDI and obesity (72–74); 2) specific residue combinations were only found in EP4 receptor, but not in other EP family members, providing H-bond formation potentials and enabling accommodation of larger chemical groups that could be used to structurally design selective agonists for EP4-targeted therapy. 3) Common and distinct residues contacting with endogenous or synthetic agonists of EP4, such as PGE2 and L902688, served as determinants for Gs/Gi coupling specificity. 4) We identified the common EP4 activation mechanism and the different activation paths of the EP4 receptor for selective Gs/Gi coupling. 5) Although 13 residues of EP4 function as common structural features for EP4 for coupling to both Gs and Gi proteins, Y130ICL2 was identified as a determinant for specific Gi contact, and E2676.30 and T3358.49-V3368.50 contributed to the selective Gs coupling of EP4. 6) Finally, different residue components on the cytoplasmic side served as determinants of the Gs/Gi selectivity among EP4 family members.
Materials and Methods
Cisplatin-Induced AKI Model.
The cisplatin-induced AKI model was performed as previously described (75, 76) with details shown in SI Appendix files.
SCR and BUN Assays.
High-performance liquid chromatography (HPLC) was performed to measure SCR and BUN levels using the Agilent 1100 HPLC system. The details are shown in SI Appendix files.
Protein Expression and Complexes’ Formation.
Recombinant baculovirus of the protein for cryo-EM was generated using the Bac-to-Bac Baculovirus Expression System, and expressed in sf9 cell. The complexes for cryo-EM were reorganization in vitro. The details of the method are shown in SI Appendix files.
Cryo-EM Data Collection and Processing.
The detailed methods of cryo-EM data collection and processing are shown in SI Appendix files.
Gαs- Gγ/Gαi–Gγ Dissociation Assay.
The Gαs-Gγ/Gαi-Gγ dissociation performed in HEK293T or Cos-7 cells are shown in SI Appendix files.
Radioligand-Binding Assay.
Radioligand-binding assays were performed by using cell membrane from COS-7 transfected with EPs receptors. The details of the method are shown in SI Appendix files.
Structure Codes.
The codes in Protein Data Bank (PDB) and Electron Microscopy Data Bank (EMDB) databases of different structures are shown as follows: PDB ID 8GDB, EMD-29945 for the PGE2-EP4-Gs complex; PDB ID 8GDA, EMD-29944 for the L902688-EP4-Gs complex; PDB ID 8GCM, EMD-29935 for the L902688-EP4-Gi complex, and PDB ID 8GD9, EMD-29943 for the Rivenprost-EP4-Gs complex; PDB ID 8GCP, EMD-29940 for the PGE2-EP4-Gi complex; PDB ID 8GDC, EMD-29946 for the PGE2-EP3-Gi.
Statistical Analysis.
All data are representative of at least three independent experiments. The data are presented as mean ± SEM, and the Student’s t test or one-way ANOVA with Tukey’s test was used for comparisons among groups by GraphPad Prism 7.0.
Supplementary Material
Appendix 01 (PDF)
Acknowledgments
This work was supported by the National Key R&D Program of China Grant (2019YFA0904200 to J.-P.S. and P.X.; 2022YFC2702600 for F.Y.; 2022SA0062 to Y.D.), the National Science Fund for Distinguished Young Scholars Grant (81825022 to J.-P.S.), the National Science Fund for Excellent Young Scholars Grant (82122070 to F.Y.), the State Key Program of National Natural Science Foundation of China (32130055 to J.-P.S.), Beijing Natural Science Foundation (Z200019 to J.-P.S. and 7172113 to B.-X.Y.), the National Natural Science Foundation of China Grant (91939301 to Z.Li and J.-P.S.; 92057121 to X.Y.; 31971195 to P.X.; 81974083 to B.-X.Y.; 82104272 to Y.-L.J.; 32000850 to Z.Liu; 81970606 to X.Z.). We would like to thank the Cryo-Electron Microscopy Center of Southern University of Science and Technology and Cryo-EM Center of Kobilka Institute of Innovative Drug Discovery of Chinese University of Hong Kong for supporting our project; We would like to thank the High-performance Computing Platform for Cryo-Electron Microscopy of Shandong University and the High-performance Computing Platform of Peking University for data processing support; we also thank the Translational Medicine Sharing Platform of Advanced Medical Research Institute of Shandong University for the Mithras L940 microplate reader support (Berthold Technologies) and Research Platform of the Education Ministry affiliated Key Laboratory of Experimental Teratology of Shandong University for the MicroBeta TriLux scintillation counter support (PerkinElmer). We thank the Cryo-EM Facility Center of Southern University of Science and Technology for providing technical support during EM image acquisition. We also thank HPC-Service Station in Cryo-EM center of Southern University of Science and Technology for the calculation resource.
Author contributions
X.Y. and J.-P.S. designed research; S.-M.H., M.-Y.X., L.L., J.M., M.-W.W., Y.-L.J., K.C., C.Z., S.C., X.W., J.-L.W., S.-C.G., C.-X.Q., Q.-T.H., B.-Y.C., C.X., S.G., Y.X., X.C., and J.-P.S. performed research; S.-M.H., M.-Y.X., L.L., J.M., M.-W.W., Y.-L.J., K.C., L.T., C.Z., Y.L., Z.Y., W.K., S.L., Z. Li, P.X., F.Y., X.Y., Y.-F.G., X.Z., Z. Liu, B.-X.Y., Y.D., and J.-P.S. analyzed data; and S.-M.H. and J.-P.S. wrote the paper.
Competing interests
The authors declare no competing interest.
Footnotes
This article is a PNAS Direct Submission. R.K.S. is a guest editor invited by the Editorial Board.
Contributor Information
Xiaoyan Zhang, Email: xyzhang@hsc.ecnu.edu.cn.
Zhongmin Liu, Email: liuzm@sustech.edu.cn.
Bao-Xue Yang, Email: baoxue@bjmu.edu.cn.
Yang Du, Email: yangdu@cuhk.edu.cn.
Jin-Peng Sun, Email: sunjinpeng@sdu.edu.cn.
Data, Materials, and Software Availability
The cryo-EM density maps and corresponding atomic coordinates of the different G protein-bound EP receptors complexes data have been deposited in Electron Microscopy Data Bank and PDB [The atomic coordinates and the cryo-EM density maps have been validated in Protein Data Bank (PDB) and Electron Microscopy Data Bank (EMDB) databases under accession numbers PDB ID 8GDB (77), EMD-29945 (78) for the PGE2-EP4-Gs complex; PDB ID 8GDA (79), EMD-29944 (80) for the L902688-EP4-Gs complex; PDB ID 8GCM (81), EMD-29935 (82) for the L902688-EP4-Gi complex, and PDB ID 8GD9 (83), EMD-29943 (84) for the Rivenprost-EP4-Gs complex; PDB ID 8GCP (85), EMD-29940 (86) for the PGE2-EP4-Gi complex; PDB ID 8GDC (87), EMD-29946 (88) for the PGE2-EP3-Gi. All the other data, associated protocols, code, and materials are detailed in SI Appendix.
Supporting Information
References
- 1.Xiao R. P., Beta-adrenergic signaling in the heart: Dual coupling of the beta2-adrenergic receptor to G(s) and G(i) proteins. Sci. STKE 2001, re15 (2001). [DOI] [PubMed] [Google Scholar]
- 2.Hein L., Altman J. D., Kobilka B. K., Two functionally distinct alpha2-adrenergic receptors regulate sympathetic neurotransmission. Nature 402, 181–184 (1999). [DOI] [PubMed] [Google Scholar]
- 3.Daaka Y., Luttrell L. M., Lefkowitz R. J., Switching of the coupling of the beta2-adrenergic receptor to different G proteins by protein kinase A. Nature 390, 88–91 (1997). [DOI] [PubMed] [Google Scholar]
- 4.Beaulieu J. M., Gainetdinov R. R., The physiology, signaling, and pharmacology of dopamine receptors. Pharmacol. Rev. 63, 182–217 (2011). [DOI] [PubMed] [Google Scholar]
- 5.Rosengren A. H., et al. , Overexpression of alpha2A-adrenergic receptors contributes to type 2 diabetes. Science 327, 217–220 (2010). [DOI] [PubMed] [Google Scholar]
- 6.Petersen S. C., et al. , The adhesion GPCR GPR126 has distinct, domain-dependent functions in Schwann cell development mediated by interaction with laminin-211. Neuron 85, 755–769 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.An W., et al. , Progesterone activates GPR126 to promote breast cancer development via the Gi pathway. Proc. Natl. Acad. Sci. U.S.A. 119, e2117004119. (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Beaulieu J. M., et al. , An Akt/beta-arrestin 2/PP2A signaling complex mediates dopaminergic neurotransmission and behavior. Cell 122, 261–273 (2005). [DOI] [PubMed] [Google Scholar]
- 9.Yan Y., et al. , Dopamine controls systemic inflammation through inhibition of NLRP3 inflammasome. Cell 160, 62–73 (2015). [DOI] [PubMed] [Google Scholar]
- 10.Schultz W., Dopamine reward prediction-error signalling: A two-component response. Nat. Rev. Neurosci. 17, 183–195 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Link R. E., et al. , Cardiovascular regulation in mice lacking alpha2-adrenergic receptor subtypes b and c. Science 273, 803–805 (1996). [DOI] [PubMed] [Google Scholar]
- 12.Bachman E. S., et al. , betaAR signaling required for diet-induced thermogenesis and obesity resistance. Science 297, 843–845 (2002). [DOI] [PubMed] [Google Scholar]
- 13.Santos R., et al. , A comprehensive map of molecular drug targets. Nat. Rev. Drug Discov. 16, 19–34 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Hauser A. S., et al. , Pharmacogenomics of GPCR drug targets. Cell 172, 41–54.e19 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Smith J. S., Lefkowitz R. J., Rajagopal S., Biased signalling: From simple switches to allosteric microprocessors. Nat. Rev. Drug Discov. 17, 243–260 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Davenport A. P., Scully C. C. G., de Graaf C., Brown A. J. H., Maguire J. J., Advances in therapeutic peptides targeting G protein-coupled receptors. Nat. Rev. Drug Discov. 19, 389–413 (2020). [DOI] [PubMed] [Google Scholar]
- 17.Markovic T., Jakopin Z., Dolenc M. S., Mlinaric-Rascan I., Structural features of subtype-selective EP receptor modulators. Drug Discov. Today 22, 57–71 (2017). [DOI] [PubMed] [Google Scholar]
- 18.Suomivuori C. M., et al. , Molecular mechanism of biased signaling in a prototypical G protein-coupled receptor. Science 367, 881–887 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Wingler L. M., et al. , Angiotensin and biased analogs induce structurally distinct active conformations within a GPCR. Science 367, 888–892 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Schmid C. L., et al. , Bias factor and therapeutic window correlate to predict safer opioid analgesics. Cell 171, 1165–1175.e1113 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Samuelsson B., Morgenstern R., Jakobsson P. J., Membrane prostaglandin E synthase-1: A novel therapeutic target. Pharmacol. Rev. 59, 207–224 (2007). [DOI] [PubMed] [Google Scholar]
- 22.Chopra S., et al. , IRE1alpha-XBP1 signaling in leukocytes controls prostaglandin biosynthesis and pain. Science 365, eaau6499 (2019). [DOI] [PubMed] [Google Scholar]
- 23.Yokoyama U., Iwatsubo K., Umemura M., Fujita T., Ishikawa Y., The prostanoid EP4 receptor and its signaling pathway. Pharmacol. Rev. 65, 1010–1052 (2013). [DOI] [PubMed] [Google Scholar]
- 24.Vukicevic S., et al. , Role of EP2 and EP4 receptor-selective agonists of prostaglandin E(2) in acute and chronic kidney failure. Kidney Int. 70, 1099–1106 (2006). [DOI] [PubMed] [Google Scholar]
- 25.Li J. H., et al. , A selective EP4 PGE2 receptor agonist alleviates disease in a new mouse model of X-linked nephrogenic diabetes insipidus. J. Clin. Invest. 119, 3115–3126 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Aringer I., et al. , Agonism of prostaglandin E2 receptor 4 ameliorates tubulointerstitial injury in nephrotoxic serum nephritis in mice. J. Clin. Med. 10, 832–848 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Xu H., et al. , Endothelial cell prostaglandin E2 receptor EP4 is essential for blood pressure homeostasis. JCI Insight 5, e138505 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Xu H., et al. , VSMC-specific EP4 deletion exacerbates angiotensin II-induced aortic dissection by increasing vascular inflammation and blood pressure. Proc. Natl. Acad. Sci. U.S.A. 116, 8457–8462 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Luschnig-Schratl P., et al. , EP4 receptor stimulation down-regulates human eosinophil function. Cell Mol. Life Sci. 68, 3573–3587 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Fujino H., Xu W., Regan J. W., Prostaglandin E2 induced functional expression of early growth response factor-1 by EP4, but not EP2, prostanoid receptors via the phosphatidylinositol 3-kinase and extracellular signal-regulated kinases. J. Biol. Chem. 278, 12151–12156 (2003). [DOI] [PubMed] [Google Scholar]
- 31.Fujino H., Regan J. W., EP(4) prostanoid receptor coupling to a pertussis toxin-sensitive inhibitory G protein. Mol. Pharmacol. 69, 5–10 (2006). [DOI] [PubMed] [Google Scholar]
- 32.Patankar J. V., et al. , E-type prostanoid receptor 4 drives resolution of intestinal inflammation by blocking epithelial necroptosis. Nat. Cell Biol. 23, 796–807 (2021). [DOI] [PubMed] [Google Scholar]
- 33.Kabashima K., et al. , The prostaglandin receptor EP4 suppresses colitis, mucosal damage and CD4 cell activation in the gut. J. Clin. Invest. 109, 883–893 (2002). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Pan M. R., Hou M. F., Chang H. C., Hung W. C., Cyclooxygenase-2 up-regulates CCR7 via EP2/EP4 receptor signaling pathways to enhance lymphatic invasion of breast cancer cells. J. Biol. Chem. 283, 11155–11163 (2008). [DOI] [PubMed] [Google Scholar]
- 35.Maruyama T., et al. , Design and synthesis of a selective EP4-receptor agonist. Part 3: 16-phenyl-5-thiaPGE(1) and 9-beta-halo derivatives with improved stability. Bioorg. Med. Chem. 10, 1743–1759 (2002). [DOI] [PubMed] [Google Scholar]
- 36.Nakase H., et al. , Effect of EP4 agonist (ONO-4819CD) for patients with mild to moderate ulcerative colitis refractory to 5-aminosalicylates: A randomized phase II, placebo-controlled trial. Inflamm. Bowel. Dis. 16, 731–733 (2010). [DOI] [PubMed] [Google Scholar]
- 37.Ninomiya T., et al. , Prostaglandin E(2) receptor EP(4)-selective agonist (ONO-4819) increases bone formation by modulating mesenchymal cell differentiation. Eur. J. Pharmacol. 650, 396–402 (2011). [DOI] [PubMed] [Google Scholar]
- 38.Ma L., et al. , Structural basis and molecular mechanism of biased GPBAR signaling in regulating NSCLC cell growth via YAP activity. Proc. Natl. Acad. Sci. U.S.A. 119, e2117054119. (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Du Y. Q., et al. , Endogenous lipid-GPR120 signaling modulates pancreatic islet homeostasis to different extents. Diabetes 71, 1454–1471 (2022). [DOI] [PubMed] [Google Scholar]
- 40.Yang F., et al. , Structural basis of GPBAR activation and bile acid recognition. Nature 587, 499–504 (2020). [DOI] [PubMed] [Google Scholar]
- 41.Rajagopal S., et al. , Quantifying ligand bias at seven-transmembrane receptors. Mol. Pharmacol. 80, 367–377 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Jang H. S., et al. , Proximal tubule cyclophilin D regulates fatty acid oxidation in cisplatin-induced acute kidney injury. Kidney Int. 97, 327–339 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Bellomo R., Kellum J. A., Ronco C., Acute kidney injury. Lancet 380, 756–766 (2012). [DOI] [PubMed] [Google Scholar]
- 44.Gao M., et al. , Disruption of prostaglandin E2 receptor EP4 impairs urinary concentration via decreasing aquaporin 2 in renal collecting ducts. Proc. Natl. Acad. Sci. U.S.A. 112, 8397–8402 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Liu C. H., et al. , Arrestin-biased AT1R agonism induces acute catecholamine secretion through TRPC3 coupling. Nat. Commun. 8, 14335 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Omori K., Kida T., Hori M., Ozaki H., Murata T., Multiple roles of the PGE2 -EP receptor signal in vascular permeability. Br. J. Pharmacol. 171, 4879–4889 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Konya V., et al. , Endothelial E-type prostanoid 4 receptors promote barrier function and inhibit neutrophil trafficking. J. Allergy. Clin. Immunol. 131 (532–540), e531–532 (2013). [DOI] [PubMed] [Google Scholar]
- 48.Oskolkova O., et al. , Prostaglandin E receptor-4 receptor mediates endothelial barrier-enhancing and anti-inflammatory effects of oxidized phospholipids. FASEB J. 31, 4187–4202 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Qu C., et al. , Ligand recognition, unconventional activation, and G protein coupling of the prostaglandin E2 receptor EP2 subtype. Sci. Adv. 7 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Norel X., et al. , International Union of Basic and Clinical Pharmacology. CIX. Differences and Similarities between Human and Rodent Prostaglandin E2 Receptors (EP1-4) and Prostacyclin Receptor (IP): Specific Roles in Pathophysiologic Conditions. Pharmacol. Rev. 72, 910–968 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Morimoto K., et al. , Crystal structure of the endogenous agonist-bound prostanoid receptor EP3. Nat. Chem. Biol. 15, 8–10 (2019). [DOI] [PubMed] [Google Scholar]
- 52.Samama P., Cotecchia S., Costa T., Lefkowitz R. J., A mutation-induced activated state of the beta 2-adrenergic receptor. Extending the ternary complex model. J. Biol. Chem. 268, 4625–4636 (1993). [PubMed] [Google Scholar]
- 53.De Lean A., Stadel J. M., Lefkowitz R. J., A ternary complex model explains the agonist-specific binding properties of the adenylate cyclase-coupled beta-adrenergic receptor. J. Biol. Chem. 255, 7108–7117 (1980). [PubMed] [Google Scholar]
- 54.Strachan R. T., et al. , Divergent transducer-specific molecular efficacies generate biased agonism at a G protein-coupled receptor (GPCR). J. Biol. Chem. 289, 14211–14224 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.DeVree B. T., et al. , Allosteric coupling from G protein to the agonist-binding pocket in GPCRs. Nature 535, 182–186 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Fu Y., et al. , Cartilage oligomeric matrix protein is an endogenous beta-arrestin-2-selective allosteric modulator of AT1 receptor counteracting vascular injury. Cell Res. 31, 773–790 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Kent R. S., De Lean A., Lefkowitz R. J., A quantitative analysis of beta-adrenergic receptor interactions: Resolution of high and low affinity states of the receptor by computer modeling of ligand binding data. Mol. Pharmacol. 17, 14–23 (1980). [PubMed] [Google Scholar]
- 58.Toyoda Y., et al. , Ligand binding to human prostaglandin E receptor EP4 at the lipid-bilayer interface. Nat. Chem. Biol. 15, 18–26 (2019). [DOI] [PubMed] [Google Scholar]
- 59.Xiao P., et al. , Tethered peptide activation mechanism of the adhesion GPCRs ADGRG2 and ADGRG4. Nature 604, 771–778 (2022). [DOI] [PubMed] [Google Scholar]
- 60.Ping Y. Q., et al. , Structural basis for the tethered peptide activation of adhesion GPCRs. Nature 604, 763–770 (2022). [DOI] [PubMed] [Google Scholar]
- 61.Lee M. H., et al. , The conformational signature of beta-arrestin2 predicts its trafficking and signalling functions. Nature 531, 665–668 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Nuber S., et al. , beta-Arrestin biosensors reveal a rapid, receptor-dependent activation/deactivation cycle. Nature 531, 661–664 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Ping Y. Q., et al. , Structures of the glucocorticoid-bound adhesion receptor GPR97-Go complex. Nature 589, 620–626 (2021). [DOI] [PubMed] [Google Scholar]
- 64.Qiao A., et al. , Structural basis of Gs and Gi recognition by the human glucagon receptor. Science 367, 1346–1352 (2020). [DOI] [PubMed] [Google Scholar]
- 65.Hua T., et al. , Activation and signaling mechanism revealed by cannabinoid receptor-Gi complex structures. Cell 180, 655–665.e618 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Rasmussen S. G., et al. , Crystal structure of the beta2 adrenergic receptor-Gs protein complex. Nature 477, 549–555 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Xiao P., et al. , Ligand recognition and allosteric regulation of DRD1-Gs signaling complexes. Cell 184, 943–956.e918 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Namba T., et al. , Alternative splicing of C-terminal tail of prostaglandin E receptor subtype EP3 determines G-protein specificity. Nature 365, 166–170 (1993). [DOI] [PubMed] [Google Scholar]
- 69.Irie A., et al. , Third isoform of the prostaglandin-E-receptor EP3 subtype with different C-terminal tail coupling to both stimulation and inhibition of adenylate cyclase. Eur. J. Biochem. 217, 313–318 (1993). [DOI] [PubMed] [Google Scholar]
- 70.Zamah A. M., Delahunty M., Luttrell L. M., Lefkowitz R. J., Protein kinase A-mediated phosphorylation of the beta 2-adrenergic receptor regulates its coupling to Gs and Gi. Demonstration in a reconstituted system. J. Biol. Chem. 277, 31249–31256 (2002). [DOI] [PubMed] [Google Scholar]
- 71.Martin N. P., Whalen E. J., Zamah M. A., Pierce K. L., Lefkowitz R. J., PKA-mediated phosphorylation of the beta1-adrenergic receptor promotes Gs/Gi switching. Cell Signal 16, 1397–1403 (2004). [DOI] [PubMed] [Google Scholar]
- 72.Yoshida K., et al. , Stimulation of bone formation and prevention of bone loss by prostaglandin E EP4 receptor activation. Proc. Natl. Acad. Sci. U.S.A. 99, 4580–4585 (2002). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Babaev V. R., et al. , Macrophage EP4 deficiency increases apoptosis and suppresses early atherosclerosis. Cell Metab. 8, 492–501 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Pan Y., et al. , Cyclooxygenase-2 in adipose tissue macrophages limits adipose tissue dysfunction in obese mice. J. Clin. Invest. 132, e152391 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Fang W., et al. , Gpr97 exacerbates AKI by mediating Sema3A signaling. J. Am. Soc. Nephrol. 29, 1475–1489 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Li Q., et al. , NLRC5 deficiency protects against acute kidney injury in mice by mediating carcinoembryonic antigen-related cell adhesion molecule 1 signaling. Kidney Int. 94, 551–566 (2018). [DOI] [PubMed] [Google Scholar]
- 77.Huang S., et al. , 8GDB, Cryo-EM Structure of the Prostaglandin E2 Receptor 4 Coupled to G Protein. Protein Data Bank. https://www.rcsb.org/structure/unreleased/8GDB. Deposited 3 March 2023.
- 78.Huang S., et al. , EMD-29945, Cryo-EM Structure of the Prostaglandin E2 Receptor 4 Coupled to G Protein. Electron Microscopy Data Bank. https://www.ebi.ac.uk/emdb/EMD-29945. Deposited 3 March 2023.
- 79.Huang S., et al. , 8GDA, Cryo-EM Structure of the Prostaglandin E2 Receptor 4 Coupled to G Protein. Protein Data Bank. https://www.rcsb.org/structure/unreleased/8GDA. Deposited 3 March 2023.
- 80.Huang S., et al. , EMD-29944, Cryo-EM Structure of the Prostaglandin E2 Receptor 4 Coupled to G Protein. Electron Microscopy Data Bank. https://www.ebi.ac.uk/emdb/EMD-29944. Deposited 3 March 2023.
- 81.Huang S., et al. , 8GCM, Cryo-EM Structure of the Prostaglandin E2 Receptor 4 Coupled to G Protein. Protein Data Bank. https://www.rcsb.org/structure/unreleased/8GCM. Deposited 3 March 2023.
- 82.Huang S., et al. , EMD-29935, Cryo-EM Structure of the Prostaglandin E2 Receptor 4 Coupled to G Protein. Electron Microscopy Data Bank. https://www.ebi.ac.uk/emdb/EMD-29935. Deposited 3 March 2023.
- 83.Huang S., et al. , 8GD9, Cryo-EM Structure of the Prostaglandin E2 Receptor 4 Coupled to G Protein. Protein Data Bank. https://www.rcsb.org/structure/unreleased/8GD9. Deposited 3 March 2023.
- 84.Huang S., et al. , EMD-29943, Cryo-EM Structure of the Prostaglandin E2 Receptor 4 Coupled to G Protein. Electron Microscopy Data Bank. https://www.ebi.ac.uk/emdb/EMD-29943. Deposited 3 March 2023.
- 85.Huang S., et al. , 8GCP, Cryo-EM Structure of the Prostaglandin E2 Receptor 4 Coupled to G Protein. Protein Data Bank. https://www.rcsb.org/structure/unreleased/8GCP. Deposited 3 March 2023.
- 86.Huang S., et al. , EMD-29940, Cryo-EM Structure of the Prostaglandin E2 Receptor 4 Coupled to G Protein. Electron Microscopy Data Bank. https://www.ebi.ac.uk/emdb/EMD-29940. Deposited 3 March 2023.
- 87.Huang S., et al. , 8GDC, Cryo-EM Structure of the Prostaglandin E2 Receptor 3 Coupled to G Protein. Protein Data Bank. https://www.rcsb.org/structure/unreleased/8GDC. Deposited 3 March 2023.
- 88.Huang S., et al. , EMD-29946, Cryo-EM Structure of the Prostaglandin E2 Receptor 3 Coupled to G Protein. Electron Microscopy Data Bank. https://www.ebi.ac.uk/emdb/EMD-29946. Deposited 3 March 2023.
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Appendix 01 (PDF)
Data Availability Statement
The cryo-EM density maps and corresponding atomic coordinates of the different G protein-bound EP receptors complexes data have been deposited in Electron Microscopy Data Bank and PDB [The atomic coordinates and the cryo-EM density maps have been validated in Protein Data Bank (PDB) and Electron Microscopy Data Bank (EMDB) databases under accession numbers PDB ID 8GDB (77), EMD-29945 (78) for the PGE2-EP4-Gs complex; PDB ID 8GDA (79), EMD-29944 (80) for the L902688-EP4-Gs complex; PDB ID 8GCM (81), EMD-29935 (82) for the L902688-EP4-Gi complex, and PDB ID 8GD9 (83), EMD-29943 (84) for the Rivenprost-EP4-Gs complex; PDB ID 8GCP (85), EMD-29940 (86) for the PGE2-EP4-Gi complex; PDB ID 8GDC (87), EMD-29946 (88) for the PGE2-EP3-Gi. All the other data, associated protocols, code, and materials are detailed in SI Appendix.