Abstract
The human adenosine A2B receptor (A2BR) is a class A G protein–coupled receptor that is involved in several major physiological and pathological processes throughout the body. A2BR recognizes its ligands adenosine and NECA with relatively low affinity, but the detailed mechanism for its ligand recognition and signaling is still elusive. Here, we present two structures determined by cryo–electron microscopy of A2BR bound to its agonists NECA and BAY60-6583, each coupled to an engineered Gs protein. The structures reveal conserved orthosteric binding pockets with subtle differences, whereas the selectivity or specificity can mainly be attributed to regions extended from the orthosteric pocket. We also found that BAY60-6583 occupies a secondary pocket, where residues V2506.51 and N2737.36 were two key determinants for its selectivity against A2BR. This study offers a better understanding of ligand selectivity for the adenosine receptor family and provides a structural template for further development of A2BR ligands for related diseases.
The Cryo-EM structures of A<sub>2B</sub>R–G<sub>s</sub> complexes provide structural basis for recognition of NECA and selectivity of BAY60-6583.
INTRODUCTION
G protein–coupled receptors (GPCRs) are the largest and most diverse family of membrane proteins in eukaryotes, with the common seven-transmembrane helix bundle (7TM) architecture (1). GPCRs transmit a variety of signals through cell membranes and regulate a wide range of physiological and pathological processes. Adenosine is an endogenous nucleoside ubiquitously distributed in mammalian organisms, which modulates different important physiological functions upon triggering of the downstream signaling of its receptors (2). Of the four adenosine receptors, adenosine A2A receptor (A2AR) and A2BR were originally found to interact with Gs/olf proteins, while A1R and A3R function in an inhibitory manner by interacting with the Gi/o subtypes (2). Recently, A2BR was also found to couple to the Gi subtype; thus, it may trigger downstream events in various cell types through different G proteins (3). The broad distribution and expression of adenosine receptors in various body tissues, as well as their role in controlling essential functions in the body, make them potential drug targets for treating various pathological diseases, such as heart disease, cancer, Parkinson’s disease, inflammation, and glaucoma (4–7). A2AR received extensive attention over the past few decades, and its crystal structure was determined in 2008 (8). From then, A2AR was considered a prototypical receptor within the GPCR superfamily, and structures with different type of ligands were determined by crystallography or cryo–electron microscopy (cryo-EM), revealing the inactive, intermediate, and fully active conformations (9–13). In contrast, neither inactive nor active state structures have yet been determined for A2BR.
Compared to the other three adenosine receptors, A2BR has a weaker affinity to the adenosine receptors’ common ligands, e.g., adenosine and its derivative, 5′-N-ethylcarboxamidoadenosine (NECA) (14). However, previous research has shown that there is an apparent up-regulation of A2BR expression along with a marked increase of the extracellular adenosine concentration under pathological conditions (5, 14, 15). Hence, A2BR was reported to be involved in several critical physiological and pathological processes, including glucose metabolism, angiogenesis induction, tumor growth, intestinal inflammation, myocardial ischemia, and acute lung and kidney injuries. Furthermore, accumulating evidence has demonstrated that A2BR can activate different intracellular signaling and exert physiological and pathological functions that are distinct from those of A2AR (3, 16). While several selective agonists or antagonists have been reported for A2BR (17–19), the leading agonist, BAY60-6583 [2-({6-amino-3,5-dicyano-4-[4-(cyclopropylmethoxy)phenyl]pyridin-2-yl}sulfanyl)acetamide] (20, 21), has been investigated for the treatment of atherosclerosis and cardiac diseases (22). In another example, the potent A2BR antagonist, PBF-1129, is now being evaluated in clinical trials for tumor immunotherapy. To understand the mechanism for low affinity of NECA and selectivity of BAY60-6583 against A2BR, we determined two cryo-EM structures of A2BR bound with NECA and BAY60-6583 with the presence of engineered trimeric Gs proteins. These structures, combined with cell-based assays, provide molecular details of the recognition specificity of ligands for the adenosine receptor family.
RESULTS
Structure determination
In preparing the homogeneous human A2BR-Gs complexes, we used the NanoBiT tethering strategy to facilitate the formation of stable complexes (fig. S1) (23). To facilitate the binding of a single-chain variable fragment (scFv16) (see Materials and Methods) (24), the Gαs was modified on the basis of mini-Gαs (11) with its αN replaced by the equivalent region of Gαi. Unless otherwise stated, the engineered G protein is henceforth referred to as Gs. NECA or BAY60-6583 was added during large-scale purification, and protein samples from size exclusion chromatography (SEC) were collected for the cryo-EM analysis. Despite the addition of scFv16 at the membrane extraction stage, it appeared to be dissociated partially during sample preparation, and the classification suggested the presence of two populations of particles for both complexes. In each case, the one without scFv16 generated better density overall in the receptor region, although in the NECA-bound complex, the scFv16 apparently stabilized the G proteins by binding to the Gβ–Gγ interface (fig. S2). Thus, we removed the scFv16 in the final A2BR–NECA–Gs and A2BR–BAY60-6583–Gs models, and their density maps were reconstructed to resolutions of 3.26 and 2.99 Å, respectively (figs. S2 and S3 and Table 1).
Table 1. Cryo-EM data collection and refinement statistics.
RMSD, root mean square deviation; PDB, Protein Data Bank.
| A2BR–NECA–Gs | A2BR–BAY60-6583–Gs | |
|---|---|---|
| Data collection and processing | ||
| Magnification | 130,000 | 130,000 |
| Voltage (kV) | 300 | 300 |
| Electron exposure (e−/Å2) | 60 | 60 |
| Defocus range (μm) | −1.2 ~ −1.8 | −1.2 ~ −1.8 |
| Pixel size (Å) | 0.96 | 0.96 |
| Symmetry imposed | C1 | C1 |
| Initial particle images (no.) | 1,546,966 | 1,063,086 |
| Final particle images (no.) | 165,307 | 190,323 |
| Map resolution (Å) | 3.26 | 2.99 |
| FSC threshold | 0.143 | 0.143 |
| Refinement | ||
| Initial model used (PDB ID) | 6GDG | 6GDG |
| Map sharpening B factor (Å2) | −123.2 | −114.8 |
| Model composition | ||
| Non-hydrogen atoms | 7048 | 7045 |
| Protein residues | 895 | 895 |
| Ligands | 1 | 1 |
| B factors (Å2) | ||
| Protein | 81.59 | 73.82 |
| Ligand | 56.51 | 65.88 |
| RMSD | ||
| Bond lengths (Å) | 0.003 | 0.003 |
| Bond angles (°) | 0.539 | 0.538 |
| Validation | ||
| MolProbity score | 1.31 | 1.37 |
| Clash score | 5.62 | 5.13 |
| Poor rotamers (%) | 0.26 | 0.27 |
| Ramachandran plot | ||
| Favored (%) | 98.07 | 97.5 |
| Allowed (%) | 1.93 | 2.5 |
| Disallowed (%) | 0 | 0 |
| PDB ID | 7XY7 | 7XY6 |
Overall structure
The A2BR–NECA–Gs and A2BR–BAY60-6583–Gs complexes are very similar to each other [Cα root mean square deviation (RMSD), 0.4 Å], and they share an essentially identical GPCR–Gs interface (Fig. 1). The human A2BR has a 55% sequence identity with A2AR and 42% with A1R; hence, it is expected that the active structures of human A2BR are similar to those of the other two members, with a Cα RMSD of 0.9 and 1.1 Å to A1R [Protein Data Bank (PDB) ID: 7LD3) (25) and A2AR (PDB ID: 6GDG) (26), respectively (fig. S4A). Although an inactive structure of A2BR is not available, the sequence and active structure comparisons of A2AR and A1R indicate that the key motifs and common features of class A receptor activation are also preserved in A2BR (fig. S4B) (27).
Fig. 1. Overall structure of A2BR–NECA–Gs and A2BR–BAY60-6583–Gs.
(A and B) Cryo-EM maps showing the disk-shaped micelle (left) and cartoon representation (right) of the A2BR–NECA–Gs (A) and A2BR–BAY60-6583–Gs (B) complexes. Densities and structures of NECA and BAY60-6583 are shown in the top-right corner of each complex structure.
The most pronounced difference between A2BR, A2AR, and A1R is in the extracellular loops (ECLs) (Fig. 2 and fig. S5). The ECL1 of A2BR is similar in both sequence and structure with the other two receptors. The ECL3 of A2BR is unique as it forms a short helix and misses an intraloop disulfide that is present in A2AR and A1R (Fig. 2, A and D). ECL2 is the most diversified region within the adenosine receptor family and that of A2BR is the longest. Here, at the N-terminal region, K147ECL2 forms a π-cation interaction with the F753.22 [Ballesteros-Weinstein numbering in superscript (28)] of ECL1. This is similar to A1R, where a hydrophobic interaction is preserved between L149ECL2 and F773.22, but differs from A2AR, which contains another interloop disulfide bond (C743.22–C146ECL2) (Fig. 2, B to D). The ECL2 of A2BR is only partially solved in both A2BR complexes, and a fragment of 153 to 167 is not solved in the A2BR models. The unmodeled fragment includes three cysteines, and we predict that one of these three cysteines should form a disulfide bond with C72ECL1 as there is clear density near C72ECL1 (fig. S6A). A similar disulfide bridge is also present in A2AR (C71ECL1–C159ECL2) to link ECL1 and the C-terminal region of ECL2 (Fig. 2B). Analyzing the sequence of A2BR species reveals that C72ECL1 and C166ECL2 are only partially conserved and present pairwise in six species, including human, while the C154ECL2 and C167ECL2 are invariantly conserved in all A2BR orthologs (fig. S6B). Thus, we speculated that the former and the latter pair of cysteines may be joined via disulfide bonding. Unlike the conserved C783.30–C171ECL2 disulfide bond that is important for GPCR’s function, these two A2BR-specific disulfide bonds may not play an essential role in its structural folding or ligand binding because individual mutations of these four cysteines did not significantly deteriorate the potency or affinity of its ligands (29) and, at times, even increased the agonist potency (30). Although the literature could not provide direct evidence of specific disulfide linkages, previous functional studies, together with our structures, indicate that these two potential disulfide bonds are partially flexible and may alter the receptor function indirectly by perturbing ligand binding or receptor equilibrium between different states.
Fig. 2. The extracellular loops of A2BR.
(A) Superposition of A2BR structures with A2AR (PDB ID: 6GDG) and A1R (PDB ID: 7LD3) from the extracellular view. The A2BR–NECA–Gs, A2BR–BAY60-6583–Gs, A2AR, and A1R structures are shown in cartoon and colored green, dark green, light blue, and brown, respectively. (B and C) Zoomed-in views of the extracellular loop 1 (ECL1) and ECL2 regions reveal differences between the three receptors. Disulfide bonds and key residues are labeled. (D) Sequence alignment of ECLs for adenosine receptors. This alignment is modified from a whole sequence alignment of these receptors (fig. S3). Key cysteines are labeled, and residues that form the helix are colored pink. The disordered ECL2 residues in A2BR and A2AR (6GDG) structures are shown with gray background.
The orthosteric pocket
The NECA binding mode in the A2BR–NECA–Gs structure is similar overall to previous A2AR-NECA and A1R-adenosine binding poses (25, 26). The conserved interactions between orthosteric pocket residues and NECA in the A2BR complex include: π-stacking with F173ECL2; hydrogen bonding with T893.36, N2546.55, S2797.42, and H2807.43; and hydrophobic interactions with V853.32, L863.33, M1825.38, V1915.47, W2476.48, V2506.51, and I2767.39 (Fig. 3A). All these residues are highly conserved within the adenosine receptor family with small variations in A3R (Fig. 3, B to C, and fig. S5). One notable position in A2BR is V2506.51 as its corresponding residue in the other three receptors is leucine. Because of this less bulky side chain, we found that, in the A2BR complex structure, the NECA rotates slightly, and the ribose moiety moves toward V2506.51 by ~1 Å to form hydrophobic contacts (Fig. 3B). Furthermore, the V2506.51 also appears to affect nearby residues H2516.52 and N2546.55 as their rotamer conformations are different from their counterparts in the A2AR and A1R structures (Fig. 3B). These subtle changes result in preserved hydrogen bond interactions between N2546.55 and the adenine moiety of NECA, whereas the previously identified NECA-H6.52 interaction in A2AR (PDB ID: 6GDG) is slightly compromised in A2BR. The importance of N2546.55 for NECA-mediated A2BR activation is confirmed by the cyclic AMP (cAMP) accumulation assay, showing that the N2546.55A mutation reduced the potency of NECA >20-fold (Fig. 3D). Meanwhile, the hydrophobic interaction contributed by V2506.51 was confirmed when the V2506.51A mutation significantly reduced the potency of NECA. We found that the potency of NECA is only slightly increased by the V2506.51L mutation, suggesting that the V/L6.51 is not the key determination for NECA’s selectivity against adenosine receptors (Fig. 3D).
Fig. 3. NECA binding pose.
(A) Detailed binding pose of NECA. NECA and key pocket residues are shown as sticks. Hydrogen bonds are shown as red dashed lines. (B) Comparison of A2BR–NECA–Gs structure (current study) with A2AR (PDB ID: 6GDG) and A1R (PDB ID: 7LD3) in the orthosteric pocket. Key movements of NECA and pocket residues are shown with black arrows. ADN, adenosine. (C) Sequence alignment of the orthosteric pocket residues. (D) Mutagenesis analysis of A2BR key residues on the potency of NECA by the cAMP assay. The maximum and minimum activation levels of wild-type (WT) A2BR were set to 100 and 0%, respectively. The X-fold median effective concentration (EC50) was calculated by dividing the EC50 of the mutant by the EC50 of WT. Data are shown as means ± SEM from at least three independent experiments. RLU, relative luminescence unit.
Another residue that gained our attention is E174ECL2. In several previous A2AR structures (8, 9, 12), the corresponding residue E169ECL2 forms a salt bridge with ECL3 residue, H264ECL3. It was suggested that this linkage forms a portion of the orthosteric binding pocket, specifically, to hold or contact with the moieties that extend from the adenine group, e.g., the phenylethylamino moiety of ZM241385 (8, 9) or the (2-carboxyethyl)phenylethylamino moiety of CGS21680 (12). These structures, together with mutagenesis and molecular dynamic (MD) simulation studies on A2AR, confirmed that the salt bridge plays a role in the dissociation of its orthosteric ligands (31, 32). In A2BR, a similar linkage was not observed as the ECL3 holds a different conformation, and a histidine is absent in the corresponding position. Furthermore, the side chain of E174ECL2 was not modeled in the A2BR–NECA structure, indicating that, although adjacent to the N6 of the adenosine group, the polar interaction of E174ECL2 may be negligible. In alignment with these structural findings, E174ECL2A displays comparable potency to wild-type (WT) A2BR in our cAMP assay; on the contrary, the corresponding E169ECL2 mutation in A2AR markedly reduced the potency of its ligands (33). Therefore, the E174ECL2 along with its counterparts in ECL3 significantly contribute to the pharmacological difference of NECA on A2BR.
The secondary pocket
In the A2BR–BAY60-6583–Gs structure, the ligand occupies an overlapped position with NECA but with some unique features (Fig. 4). Similar to NECA, BAY60-6583 also forms hydrogen bonds with T893.36 and N2546.55, as well as van der Waals interactions with previously mentioned residues (Fig. 4A). The cyclopropyl group of BAY60-6583 points deep into the orthosteric pocket and makes close contacts with L863.33, V1915.47, W2476.48, and H2516.52. In the middle region, the phenyl and pyridine moieties of BAY60-6583 partially overlapped with the ribose and adenine groups of NECA, respectively. Unlike the ribose group of NECA, BAY60-6583 does not form hydrogen bond with S2797.42 and H2807.43, and the hydrophobic phenyl moiety of BAY60-6583 is located 2 to 3 Å further away from these TM7 residues. These structural differences are aligned with the functional data showing that the S7.42 and H7.43 mutations greatly decreased the potency and affinity of the ribose-containing compounds on all adenosine family receptors (34–37), in contrast to a slight increase on the potency of BAY60-6583 on A2BR by the S2797.42A mutation (35).
Fig. 4. BAY60-6583 binding pose.
(A and B) The detailed binding pose of BAY60-6583 (A) and comparison with NECA (B). The ligands and key pocket residues are shown as sticks. Hydrogen bonds are shown as red dashed lines. Key movements of the secondary pocket residues are shown with black arrows. In (A), the NECA structure is superimposed for comparison (50% transparency). (C) Comparison of A2BR structure with A2AR (PDB ID: 6GDG) and A1R (PDB ID: 7LD3) in the orthosteric pocket and secondary pocket. ADN, adenosine. (D) Mutagenesis analysis of A2BR orthosteric and secondary pocket residues on the potency of BAY60-6583 by the cAMP assay. (E) The potency of BAY60-6583 on the pocket mutations of A2AR. In (C) and (D), the maximum and minimum activation levels of WT receptor were set to 100 and 0%, respectively. The X-fold EC50 was calculated as in Fig. 3. Data are shown as means ± SEM from at least three independent experiments.
There are two cyano modifications on the pyridine moiety of BAY60-6583: one cyano group hydrogen bonds to the N2546.55 of A2BR, and the other points to an empty void embraced by TM2, TM3, and TM6. Another amino modification on the pyridine moiety points upward and forms a week hydrophilic interaction with E174ECL2. The cAMP accumulation assay confirmed that the potency of BAY60-6583 is ~2-fold reduced by the E174ECL2A mutant (Fig. 4D), in contrast to a negligible effect of E174ECL2A on NECA. These assay results also aligned with the successful modeling of E174ECL2 in the A2BR–BAY60-6583–Gs complex and may be a consequence of a stronger polar interaction between E174ECL2 and BAY60-6583. In contrast to the assay results of NECA, the potency of BAY60-6583 is markedly reduced by the V2506.51L mutation but only moderately affected by the V2506.51A mutation. These distinct mutation results are compatible with the precise differences in the binding pose of these two molecules. We further conducted the reverse mutation on A2AR and found that the L2496.51V mutant can increase the potency of BAY60-6583 on A2AR by >714-fold (Fig. 4E). These results thus identify V2506.51 as a key determinant for the high selectivity of BAY60-6583 toward A2BR.
BAY60-6583 contains an acetamide group on the C1 position of the pyridine moiety that points toward a secondary pocket composed of the extracellular tips of TM1, TM2, and TM7 (Fig. 4, B and C). Density of the acetamide group is relatively weak but allows us to model the polar tail toward Y101.35 and N2737.36 of the receptor. Three nearby receptor residues, Y101.35, S682.65, and N2737.36, are well modeled. Among them, S682.65 points in the opposite direction and did not form a hydrogen bond with BAY60-6583, while Y101.35 and N2737.36 face the acetamide group and within hydrogen bond distance of the polar tail on BAY60-6583. It is also worth mentioning that comparison of the A2BR–BAY60-6583–Gs and A2BR–NECA–Gs complex structures reveals interesting subtle main-chain or side-chain movement of these residues in both the orthosteric and secondary pockets (Fig. 4B). Pronounced differences include the three residues in the secondary pocket that move away from the pocket to accommodate the acetamide group of BAY60-6583 (Fig. 4B). Aligned with the structural findings, the cAMP assay revealed that the S682.65A mutant does not affect the potency, whereas the Y101.35F and N2737.36A caused an ~8- and ~6-fold decrease in potency, respectively, suggesting that these two residues contribute to the recognition of BAY60-6583 (Fig. 4D). As a control, the mutant Y101.35F of the secondary pocket does not affect the activity of NECA (Fig. 3D). Y101.35 is absolutely conserved, but N2737.36 is only present in A2BR. In the other three adenosine receptors, the corresponding position of 7.36 is a tyrosine. Thus, we mutated the N2737.36 to tyrosine on A2BR, and the compromised activity indicated that the bulky side chain of Y7.36 may interfere with the acetamide group of BAY60-6583 (Fig. 4, C and D). Aligned with these results, the reverse mutant (Y2717.36N) on A2AR can significantly increase the potency of BAY60-6583 (Fig. 4E). These results identify N2737.36 as another determinant for BAY60-6583’s selectivity against A2BR.
The A2BR–G interface
Like other GPCRs that featured a larger outward shift of TM6 when coupled to Gs protein, TM6 of A2BR occupies an orientation similar to the TM6 of A2AR but is distinct from the counterpart of A1R, which coupled to Gi (Fig. 5A). Consistently, the A2BR–Gs interface buries 2610 Å2 of the total surface area, which is very close to the A2AR–Gs interface (2572 Å2), and both are much larger than the A1R–Gi interface (2086 Å2). Despite its similarity with the A2AR–Gs interface, there are several interesting features of the A2BR–Gs interface. The A2BR–Gs interface is distinctive in that helix8 tilts toward cytosolic and makes extra contacts with the β subunit of G proteins, whereas, in A2AR and A1R, it points toward the cell membrane and shows a lack of contact with the G proteins (Fig. 5A). A close-up view of the structures shows that the Y2927.55 of A2BR changes its rotamer and points toward TM1, and hydrogen bonds to the backbone of G241.49 (Fig. 5B). Consequently, the side chain of Y2927.55 partially occupies the position of F3018.54, which likely pushes the downward shift of helix8.
Fig. 5. A2BR-Gs interface.
(A) Superposition of A2BR–NECA–Gs (current study) with A2AR (PDB ID: 6GDG) and A1R (PDB ID: 7LD3) structures on receptor. (B) Comparison of the TM7-helix8 regions of A2BR, A2AR, and A1R showing the key differences on Y2927.55 and nearby residues. (C) The A2BR helix8–Gβ interface. (D) Mutagenesis analysis of A2BR helix8 residues on the potency of NECA by the cAMP assay. The maximum and minimum activation levels of WT A2BR were set to 100 and 0%, respectively. The X-fold EC50 was calculated by dividing the EC50 of the mutant by the EC50 of WT. Data are shown as means ± SEM from at least three independent experiments. (E) The receptor–α5 interface in A2BR. Interacting residues are shown as sticks and are labeled. Color codes are indicated at the top-left corner.
Despite of weak densities, the Y2998.52, K3038.56, and R3078.60 of helix8 point toward the Gβ residues D312, H311, and A309, respectively, in the A2BR–Gβ interface (Fig. 5C). We thus introduced mutations on helix8 of A2BR to validate the potential function of the helix8–Gβ interface. The cAMP assay results showed that the mutant R3078.60A decreased the potency of NECA by 4-fold, while other mutants did not significantly deteriorate the potency (Fig. 5D). These results suggested that the A2BR–Gβ interface is determined by the very C-terminal region of helix8. Furthermore, the insusceptibility of the potency by the F3018.54A mutation suggested that the tilted conformation of helix8 is an intrinsic feature and may not relate to the contacts between Y2927.55 and F3018.54. Such a tilted helix8 is frequently seen in class B GPCRs (38, 39).
The tilt of helix8 in A2BR also induces 2 to 4 Å downward movement of G protein subunits. Nevertheless, the receptor–Gαs interface of A2BR is quite similar to A2AR, and most of the key interactions are preserved, e.g., the π-cation interaction of R1033.50 and Y374H5.23 [uppercase refers to common G protein numbering (40)], and the hydrophobic network by L376H5.25 and TM3 and TM5 residues (I1073.54, I2055.61, V2085.64, and A2095.65) (Fig. 5E). There are also extensive hydrophilic interactions between α5 residues (Q367H5.16, R368H5.17, R372H5.21, and L377H5.26) and the A2BR residues from ICL2, ICL3, and helix8 (Fig. 5E). Therefore, these results are consistent with the fact that the receptor–Gαs interactions are largely conserved between A2BR and A2AR despite being disturbed by the tilted helix8, and the tiny vibration of the interface is also frequently seen in other subfamilies of Gs-coupled receptors.
DISCUSSION
Our complex structures provided a precise structural template for previous site-directed mutagenesis and structure-activity relationship studies (14, 35, 41, 42). The A2BR structures reveal a highly conserved orthosteric binding pocket with diversity mainly coming from the extracellular regions, including the ECL2 and ECL3 residues. We found that the ECL2 and ECL3 of A2BR may not form a similar salt bridge as in A2AR (E169ECL2–H264ECL3) because of sequence and structural differences. Importance of this salt bridge has previously been demonstrated on A2AR structurally: The ligand ZM241385 occupied the orthosteric pocket and its phenylethylamino group formed contacts with the salt bridge when crystallized under acidic condition (8), whereas under basic condition (pH > 8), the salt bridge was broken, and the phenylethylamino group of ZM241385 pointed to a reverse direction (43). In addition, it was shown that agonist binding to A2AR is pH sensitive with higher affinities observed at a lower pH (44, 45). These structural and functional findings suggested that, for A2AR, the key determinant for the sensitivity is probably the H264ECL3 considering its pKa (where Ka is the acid dissociation constant) value of ~7 because, at lower pH, the formed E169ECL2–H264ECL3 linkage may function like a lid over the orthosteric pocket and thus enhanced the residence time of its ligand and increased the binding affinity. Therefore, it is our best speculation for A2BR that, without such kind of linkage, the ligand NECA thus binds with fast-on/fast-off kinetics and weak affinity (46).
Besides the ECLs, the subtle differences of the residues within the orthosteric pocket, including but not limited to the V/L6.51 and its nearby residues H6.52 and N6.55, may further contribute to the selectivity. We found that the potency of NECA is slightly increased by the A2AR/A1R/A3R-mimicing mutation V2506.51L, reminiscent of a previous study that suggested L6.51 as a hotspot for A2AR’s selective antagonist ZM241385 (47). The phenomenon that a ligand adopts slightly different conformations in homologous receptors is not unusual and has been illustrated, for example, in the d-lysergic acid diethylamide binding poses in serotonin receptors HTR2A and HTR2B (48, 49). In this study, we also identified V2506.51 and N2737.36 as two key determinants for the selectivity of BAY60-6583 against A2BR. The distinct effects of V250L6.51 and V250A6.51 on BAY60-6583 and NECA aligned well with their binding details (Figs. 3D and 4D). N2737.36 is uniquely located in the secondary binding pocket of A2BR. The secondary pocket occupied by BAY60-6583 is similar to previously determined binding modes of DU172-bound A1R and XAC-bound A2AR (fig. S8A), whereas most other adenosine receptor ligands occupy only the orthosteric pocket (fig. S8B). Compared to the position of BAY60-6583 in the A2BR pockets, the two A2AR selective agonists, UK-432097 and CGS21680, extend upward from the orthosteric pocket and interact with the ECL2–ECL3 salt bridge that is uniquely present in A2AR. Without such a linkage in A2BR, the A2BR agonist BAY60-6583 likely extends horizontally toward TM1/2/7 and occupies the secondary pocket, which contains an aliphatic residue N2737.36.
Although having a similar level of potency, BAY60-6583 was previously reported as a partial agonist because its efficacy is much lower than that of NECA or adenosine (20). Our cAMP assay confirmed that BAY60-6583 shares quantitatively equipotent potency with NECA (3.67 versus 8.85 nM) but a much lower efficacy (~5%) compared to NECA (fig. S7). Although the NECA- and BAY60-6583–bound structures revealed a nearly identical active conformation of A2BR, the binding details may shed light on the structural basis of their distinct efficacies. The GPCR activation pathway is featured by shrinking the middle region and expanding the intercellular region of the 7TM upon ligand binding, a process that is mainly facilitated by outward movement of TM6 and inward movement of TM7 (27). Using A2AR as an example, compared to the antagonist-bound inactive state, the TM7 moves toward the agonist, and the S7.42 and H7.43 hydrogen bond to the ligand ribose group in the agonist-bound conformation, thus stabilizing the active state of TM7 to accommodate downstream G proteins (8, 10). For A2BR, the consensus ligand NECA makes conserved contacts with the corresponding TM7 residues, whereas the selective ligand BAY60-6583 misses these polar contacts. Hence, our structures, together with previous functional results (34–37), suggest that, on A2BR, the NECA (or adenosine) can efficiently trigger the conformational change toward the active state through these hydrogen bond interactions, whereas, for BAY60-6583, the transition efficacy is relatively low as a result of the absent polar contacts. In contrast, for those tested residues in the secondary pocket, their decreased efficacies are likely because the mutations indirectly affected the conformational transition that involves all helices from inactive to active conformation.
We further hypothesize that it is feasible to enhance the efficacy of BAY60-6583 by modifying the cyano group on C2 of the pyridine moiety that faces the lower edge void of the secondary pocket (fig. S8, C and D). This proton-accepting cyano group is relatively incompatible with nearby residues A602.57, A642.61, V853.32, and H2807.43 (distances, 3.5 to 6.8 Å). Thus, it is possible to increase the efficacy by replacing the cyano group with a proton-accepting group to hydrogen bond to the previously mentioned H2807.43, a residue that is critical for the conformation transition of the adenosine receptors toward active state. Relatedly, an analog of BAY60-6583 that cyclized the C1 and C2 derivatives (compound 52) is more rigid, and its activity is almost abolished (50). From our docking model, the phenyl group of compound 52 rotates and becomes incompatible with the nearby nonpolar residues, and its polar acetamide tail interferes with A642.57 of TM2 (fig. S8E).
With the A2BR active structure in hand, we are now able to map the structure of the cancer-related mutations that have been pharmacologically characterized (51). All these mutations are located far from the ligand-binding pocket (fig. S9), so they may function by affecting receptor folding and stability, the equilibrium between inactive and active states, or the coupling efficacy with the downstream adaptors. Of note, our last modeled residue in the receptor structure is Y308. Thus, the dysfunction caused by L310 mutations is possibly through vibrating the preceding helix8 that was involved in Gβ interactions, or the following C311, which is potentially a palmitoylation site.
Together, our structures reveal interesting similarities and diversities among adenosine receptors, thus representing an important complement to the previously determined A2AR and A1R structures. The current research suggests that residues of the orthosteric pocket and the ECLs together contribute to the potency and affinity differences between A2BR and the other adenosine receptors. This study also identified key determinants for the specificity of A2BR agonist BAY60-6583, illuminating the path forward for future discovery of even more potent A2BR ligands for related diseases.
MATERIALS AND METHODS
Construct cloning, expression, and purification of A2BR-Gs complexes
For the receptor, human A2BR (residues 2 to 332), containing an N-terminal thermostabilized apocytochrome b562RIL fusion protein and a C-terminal LgBiT protein, was cloned into pFastBac1 vector, with hemagglutinin signal peptide followed by FLAG and 10× His tags at the very N terminus. For the G protein components, mini-GαsIN, equal to mini-Gs but with its αN replaced by the counterpart of Gi (to facilitate binding of scFv16, which binds to the interface between Gβγ subunits and Gαi subunit) (52), was fused to bovine Gγ2 at its N terminus (Gγ–mini-GsIN). Human Gβ1, followed by SmBiT at its C terminus, was cloned into pFastBac-Dual together with Gγ–mini-GαsIN (fig. S1). Sf9 insect cells were infected with viruses of the receptor (A2BR-LgBiT) and Gγ–mini-GαsIN–Gβ–SmBiT in the ratio of 1:2 for 72 hours at 27°C. The cell pellets were lysed by Dounce homogenization in 20 mM Hepes (pH 7.5), 100 mM NaCl, 5 mM MgCl2, 5 mM CaCl2, 10% glycerol, EDTA-free protease inhibitor cocktail (Bimake), apyrase (25 mU/ml; New England Biolabs), 200 μM Tris(2-carboxyethyl)phosphine (TCEP) (Thermo Fisher Scientific), 20 μM agonist NECA (Abcam) or BAY60-6583 (MedChemExpress), and scFv16 (15 μg/ml), incubated at 4°C for 3 hours, and then centrifuged at 30,000 rpm for 45 min to collect the membranes. The washed membranes were solubilized in 20 mM Hepes (pH 7.5), 100 mM NaCl, 5 mM MgCl2, 5 mM CaCl2, EDTA-free protease inhibitor cocktail, 1% (w/v) lauryl maltose neopentyl glycol (LMNG; Anatrace), 0.2% (w/v) cholesteryl hemisuccinate (CHS; Sigma-Aldrich), 20 μM agonist (NECA or BAY60-6583), apyrase (25 mU/ml), and 150 μM TCEP and incubated at 4°C for 3 hours. The supernatant was collected by centrifugation at 30,000 rpm for 45 min and then incubated with TALON immobilized metal affinity chromatography resin (Clontech) with an addition of 20 mM imidazole at 4°C overnight. The resin was collected by centrifugation at 1000g for 5 min, packed into a gravity flow column, and washed with 20 column volumes of buffer containing 20 mM Hepes (pH 7.5), 100 mM NaCl, 2 mM MgCl2, 2 mM CaCl2, 5 μM agonist (NECA or BAY60-6583), 0.01% (w/v) LMNG, 0.002% (w/v) CHS, and 20 mM imidazole and eluted using buffer containing 20 mM Hepes (pH 7.5), 100 mM NaCl, 2 mM MgCl2, 2 mM CaCl2, 100 μM agonist (NECA or BAY60-6583), 0.01% (w/v) LMNG, 0.002% (w/v) CHS, and 300 mM imidazole. The elution was collected and concentrated with an Amicon Ultra Centrifugal Filter [molecular weight cutoff (MWCO), 100 kDa] and subjected to SEC using a Superdex 6 10/300 GL column (GE Healthcare) with running buffer, containing 20 mM Hepes (pH 7.5), 100 mM NaCl, 2 mM MgCl2, 2 mM CaCl2, 100 μM TCEP, 5 μM agonist (NECA or BAY60-6583), 0.01% (w/v) LMNG, and 0.002% (w/v) CHS. The fractions of monomeric protein complex were pooled and concentrated with an Amicon Ultra Centrifugal Filter (MWCO, 100 kDa) to 3 to 5 mg/ml for cryo-EM studies.
Cryo-EM sample preparation and image acquisition
Samples (3 μl) of A2BR–NECA–Gs and A2BR–BAY60-6583–Gs were applied to Au grids (Quantifoil, 300 mesh AU R1.2/1.3) that were glow-discharged for 40 s using Gatan Solarus (950). Grids were subsequently plunge-frozen by Vitrobot Mark IV (Thermo Fisher Scientific) at 4°C and 100% humidity.
Cryo-EM datasets were collected on the Krios G4 Cryo–Transmission Electron Microscope equipped with the Falcon 4 Direct Electron Detector (Thermo Fisher Scientific) in superresolution mode operating at 300 kV accelerating voltage with 10-eV slit width. Movies were taken under Energy-filtered transmission electron microscopy (EFTEM) nanoprobe mode, with 50-μm C2 aperture and calibrated magnification of ×130,000, corresponding to a pixel size of 0.96 Å. Each movie was taken in electron-event representation (EER) format with a total dose of 60 electrons per Å2. EPU software (Thermo Fisher Scientific, v.1.11.0) was used to collect data with a defocus range of −1.2 to −1.8 μm.
Cryo-EM data processing and three-dimensional reconstruction
A total of 10,274 movies of the A2BR–NECA–Gs sample and 18,412 movies of the A2BR–BAY60-6583–Gs sample were imported into RELION 4.0 (53) to perform motion correction. The resulting .mrc files were imported into CryoSPARC 3.3.1 for further processing (54). Contrast transfer function (CTF) parameters were determined by patch CTF estimation (multi). After that, automated particle picking and rounds of two-dimensional classification were performed.
For the A2BR–NECA–Gs sample, selected subsets with 1,546,966 particles were used to generate the initial model. After rounds of heterogeneous refinement, models of A2BR–NECA–Gs with 165,307 particles and A2BR–NECA–Gs-scFv16 with 188,487 particles were generated. Subsequently, homogeneous refinement, nonuniform refinement, and local refinement were performed on these two models. Last, A2BR–NECA–Gs with a 3.26-Å map and A2BR–NECA–Gs-scFv16 with a 3.19-Å map were determined by gold standard Fourier shell correlation (FSC) using the 0.143 criterion.
For the A2BR–BAY60-6583–Gs sample, selected subsets with 1,063,086 particles were used to generate the initial model. After rounds of heterogeneous refinement, a model with 190,323 particles was generated to perform homogeneous refinement, nonuniform refinement, and local refinement. Because the scFv16 only generated better density in the G protein region, but not in the receptor region in the A2BR–NECA–Gs-scFv16 complex, we thus removed the scFv16 from the data process and refinement for the A2BR–BAY60-6583–Gs complex. The final 2.99-Å map was determined by gold standard FSC using the 0.143 criterion.
Model building and refinement for A2BR structures
Structures of A2BR predicted by AlphaFold (55, 56) and A2AR-miniGs (PDB ID: 6GDG) (26) were used as the starting models for model building and refinement against the electron density map of A2BR–NECA–Gs and A2BR–BAY60-6583–Gs. Models were docked into the EM density maps using Chimera (57). Subsequently, iterative manual adjustments and rebuilding were performed in Coot (58) and phenix.real_space_refine in Phenix (59). The final refinement statistics are provided in Table 1. The structure figures were prepared by Chimera and PyMOL (www.pymol.org).
cAMP assay
For the cAMP assay, the codon-optimized WT and mutant A2BR and A2AR genes were cloned into a customized pLEXm vector with N-terminal signal peptide, Flag tag, and C-terminal His10 tags. Human embryonic kidney 293T cells were cultured and preseeded into 96-well plates in Dulbecco’s modified Eagle’s medium supplemented with 10% (v/v) fetal bovine serum, penicillin (50 IU/ml), and streptomycin (50 μg/ml). After 36 hours, the cells were transiently transfected with WT or mutant receptors in combination with cAMP response element reporter at the ratio of 1:1 (w/w) using polyethylenimine (ratio of 1:3). After 12 hours of culture at 37°C and 5% CO2, the cells were stimulated with serial-diluted concentrations of agonists (NECA, 10.24 pM to 4 μM; BAY60-6583, 10.24 pM to 4 μM). Stimulated cAMP accumulation was measured by the Steady-Glo Luciferase Assay System (Promega). Briefly, after 4 hours of stimulation, the medium was discarded, and cells were lysed using a passive lysis buffer (Promega) and transferred into 96-well assay plates. Assay substrate was added to each well following the supplier’s instructions. Plates were incubated in the dark at room temperature for 5 min, and the glow intensity is measured by a LUMIstar Omega reader (BMG Labtech). Relative luminescence unit (RLU) was acquired by subtracting intensity of the well with apo receptor. Cell surface expression for WT and each mutant was monitored by flow cytometry. In brief, the expressed cells were treated by trypsin, resuspended in phosphate-buffered saline (PBS), and incubated with mouse anti-Flag (M2–fluorescein isothiocyanate) antibody (Sigma-Aldrich) for 15 min in the dark, and then, a ninefold excess of PBS was added to cells. Last, the surface expression data were measured by detecting the percentage of cells sorted by fluorescence of fluorescein isothiocyanate using BD FACSCanto II (BD Biosciences).
Acknowledgments
Funding: This work was supported by the National Key Research and Development Program of China (2018YFA0507001 and 2018YFA0507000), the National Natural Science Foundation of China (grant 32171215 to G.S.), the basic research program of Science and Technology Commission of Shanghai Municipality (21JC1402400), and the National Science Fund for Distinguished Young Scholars 32022038 (T.H.). We thank L. Wu for useful comments on EM data process. We thank the Instruments Sharing Platform of School of Life Sciences, East China Normal University.
Author contributions: Y.C. optimized constructs, expressed and purified the complex proteins for cryo-EM studies, and edited the initial manuscript. J.Z. undertook cryo-EM data collection and three-dimensional reconstruction and optimized the initial structure models. Y.W. helped in cell culture and molecular cloning and undertook the cAMP assay. Y.X. optimized conditions for GPCR–G protein complex preparation at early stage and assisted with the cAMP assay. W.Lu analyzed the assay data and assisted with compound preparation. W.Liu and M.L. oversaw the data analysis and edited the manuscript. T.H. supervised the cryo-EM data collection and structural modeling and edited the manuscript. G.S. supervised the whole project, analyzed the data, and wrote the manuscript.
Competing interests: The authors declare that they have no competing interests.
Data and materials availability: Atomic coordinates and structure factors for the A2BR–BAY60-6583–Gs and A2BR–NECA–Gs structures have been deposited in the Protein Data Bank with identification codes 7XY6 and 7XY7, and the corresponding electron microscopy maps have been deposited in the Electron Microscopy Data Bank under accession codes EMD-33512 and EMD-33513. All data needed to evaluate the conclusions in the paper are present in the paper and/or the Supplementary Materials.
Supplementary Materials
This PDF file includes:
Figs. S1 to S9
REFERENCES AND NOTES
- 1.R. Fredriksson, M. C. Lagerström, L.-G. Lundin, H. B. Schiöth, The G-protein-coupled receptors in the human genome form five main families. phylogenetic analysis, paralogon groups, and fingerprints. Mol. Pharmacol. 63, 1256–1272 (2003). [DOI] [PubMed] [Google Scholar]
- 2.B. B. Fredholm, A. P. IJzerman, K. A. Jacobson, K. N. Klotz, J. Linden, International Union of Pharmacology. XXV. Nomenclature and classification of adenosine receptors. Pharmacol. Rev. 53, 527–552 (2001). [PMC free article] [PubMed] [Google Scholar]
- 3.Z. G. Gao, A. Inoue, K. A. Jacobson, On the G protein-coupling selectivity of the native A2B adenosine receptor. Biochem. Pharmacol. 151, 201–213 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.M. R. Blackburn, C. O. Vance, E. Morschl, C. N. Wilson, Adenosine receptors and inflammation. Handb. Exp. Pharmacol. ,215–269 (2009). [DOI] [PubMed] [Google Scholar]
- 5.P. A. Borea, S. Gessi, S. Merighi, F. Vincenzi, K. Varani, Pharmacology of adenosine receptors: The state of the art. Physiol. Rev. 98, 1591–1625 (2018). [DOI] [PubMed] [Google Scholar]
- 6.B. N. Cronstein, M. Sitkovsky, Adenosine and adenosine receptors in the pathogenesis and treatment of rheumatic diseases. Nat. Rev. Rheumatol. 13, 41–51 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.S. Gessi, S. Merighi, V. Sacchetto, C. Simioni, P. A. Borea, Adenosine receptors and cancer. Biochim. Biophys. Acta 1808, 1400–1412 (2011). [DOI] [PubMed] [Google Scholar]
- 8.V. P. Jaakola, M. T. Griffith, M. A. Hanson, V. Cherezov, E. Y. Chien, J. R. Lane, A. P. Ijzerman, R. C. Stevens, The 2.6 angstrom crystal structure of a human A2A adenosine receptor bound to an antagonist. Science 322, 1211–1217 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.W. Liu, E. Chun, A. A. Thompson, P. Chubukov, F. Xu, V. Katritch, G. W. Han, C. B. Roth, L. H. Heitman, A. P. IJzerman, V. Cherezov, R. C. Stevens, Structural basis for allosteric regulation of GPCRs by sodium ions. Science 337, 232–236 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.F. Xu, H. Wu, V. Katritch, G. W. Han, K. A. Jacobson, Z. G. Gao, V. Cherezov, R. C. Stevens, Structure of an agonist-bound human A2A adenosine receptor. Science 332, 322–327 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.B. Carpenter, R. Nehmé, T. Warne, A. G. Leslie, C. G. Tate, Structure of the adenosine A2A receptor bound to an engineered G protein. Nature 536, 104–107 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.G. Lebon, P. C. Edwards, A. G. Leslie, C. G. Tate, Molecular determinants of CGS21680 binding to the human adenosine A2A receptor. Mol. Pharmacol. 87, 907–915 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.M. Cui, Q. Zhou, Y. Xu, Y. Weng, D. Yao, S. Zhao, G. Song, Crystal structure of a constitutive active mutant of adenosine A(2A) receptor. IUCrJ 9, 333–341 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.M. W. Beukers, H. den Dulk, E. W. van Tilburg, J. Brouwer, A. P. Ijzerman, Why are A2B receptors low-affinity adenosine receptors? Mutation of Asn273 to Tyr increases affinity of human A2B receptor for 2-(1-Hexynyl)adenosine. Mol. Pharmacol. 58, 1349–1356 (2000). [DOI] [PubMed] [Google Scholar]
- 15.P. A. Borea, S. Gessi, S. Merighi, K. Varani, Adenosine as a multi-signalling guardian angel in human diseases: When, where and how does it exert its protective effects? Trends Pharmacol. Sci. 37, 419–434 (2016). [DOI] [PubMed] [Google Scholar]
- 16.M. Klinger, M. Freissmuth, C. Nanoff, Adenosine receptors: G protein-mediated signalling and the role of accessory proteins. Cell. Signal. 14, 99–108 (2002). [DOI] [PubMed] [Google Scholar]
- 17.P. G. Baraldi, M. A. Tabrizi, F. Fruttarolo, R. Romagnoli, D. Preti, Recent improvements in the development of A2B adenosine receptor agonists. Purinergic Signal 5, 3–19 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.A. Saini, R. Patel, S. Gaba, G. Singh, G. D. Gupta, V. Monga, Adenosine receptor antagonists: Recent advances and therapeutic perspective. Eur. J. Med. Chem. 227, 113907 (2022). [DOI] [PubMed] [Google Scholar]
- 19.C. E. Müller, K. A. Jacobson, Recent developments in adenosine receptor ligands and their potential as novel drugs. Biochim. Biophys. Acta 1808, 1290–1308 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.S. Hinz, S. K. Lacher, B. F. Seibt, C. E. Müller, BAY60-6583 acts as a partial agonist at adenosine A2B receptors. J. Pharmacol. Exp. Ther. 349, 427–436 (2014). [DOI] [PubMed] [Google Scholar]
- 21.Z. G. Gao, R. Balasubramanian, E. Kiselev, Q. Wei, K. A. Jacobson, Probing biased/partial agonism at the G protein-coupled A2B adenosine receptor. Biochem. Pharmacol. 90, 297–306 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.U. Rosentreter, R. Henning, M. Bauser, T. Kramer, A. Vaupel, W. Hubsch, K. Dembowsky, O. Salcher-Schraufstatter, J. P. Stasch, T. Krahn, E. Perzborn, CA 2386147A1 (2001).
- 23.J. Duan, D. D. Shen, X. E. Zhou, P. Bi, Q. F. Liu, Y. X. Tan, Y. W. Zhuang, H. B. Zhang, P. Y. Xu, S. J. Huang, S. S. Ma, X. H. He, K. Melcher, Y. Zhang, H. E. Xu, Y. Jiang, Cryo-EM structure of an activated VIP1 receptor-G protein complex revealed by a NanoBiT tethering strategy. Nat. Commun. 11, 4121 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.S. Maeda, A. Koehl, H. Matile, H. Hu, D. Hilger, G. F. X. Schertler, A. Manglik, G. Skiniotis, R. J. P. Dawson, B. K. Kobilka, Development of an antibody fragment that stabilizes GPCR/G-protein complexes. Nat. Commun. 9, 3712 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.C. J. Draper-Joyce, R. Bhola, J. Wang, A. Bhattarai, A. T. N. Nguyen, I. Cowie-Kent, K. O’Sullivan, L. Y. Chia, H. Venugopal, C. Valant, D. M. Thal, D. Wootten, N. Panel, J. Carlsson, M. J. Christie, P. J. White, P. Scammells, L. T. May, P. M. Sexton, R. Danev, Y. Miao, A. Glukhova, W. L. Imlach, A. Christopoulos, Positive allosteric mechanisms of adenosine A1 receptor-mediated analgesia. Nature 597, 571–576 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.J. García-Nafría, Y. Lee, X. Bai, B. Carpenter, C. G. Tate, Cryo-EM structure of the adenosine A2A receptor coupled to an engineered heterotrimeric G protein. eLife 7, e35946 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Q. Zhou, D. Yang, M. Wu, Y. Guo, W. Guo, L. Zhong, X. Cai, A. Dai, W. Jang, E. I. Shakhnovich, Z. J. Liu, R. C. Stevens, N. A. Lambert, M. M. Babu, M. W. Wang, S. Zhao, Common activation mechanism of class A GPCRs. eLife 8, e50279 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.J. A. Ballesteros, H. Weinstein, [19] 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]
- 29.A. C. Schiedel, S. Hinz, D. Thimm, F. Sherbiny, T. Borrmann, A. Maass, C. E. Müller, The four cysteine residues in the second extracellular loop of the human adenosine A2B receptor: Role in ligand binding and receptor function. Biochem. Pharmacol. 82, 389–399 (2011). [DOI] [PubMed] [Google Scholar]
- 30.M. C. Peeters, Q. Li, G. J. van Westen, A. P. Ijzerman, Three “hotspots” important for adenosine A2B receptor activation: A mutational analysis of transmembrane domains 4 and 5 and the second extracellular loop. Purinergic Signal 8, 23–38 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.D. Guo, A. C. Pan, R. O. Dror, T. Mocking, R. Liu, L. H. Heitman, D. E. Shaw, I. J. AP, Molecular basis of ligand dissociation from the adenosine A2A receptor. Mol. Pharmacol. 89, 485–491 (2016). [DOI] [PubMed] [Google Scholar]
- 32.B. Carpenter, G. Lebon, Human adenosine A2A receptor: Molecular mechanism of ligand binding and activation. Front. Pharmacol. 8, 898 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.W. Jespers, A. C. Schiedel, L. H. Heitman, R. M. Cooke, L. Kleene, G. J. P. van Westen, D. E. Gloriam, C. E. Müller, E. Sotelo, H. Gutiérrez-de-Terán, Structural mapping of adenosine receptor mutations: Ligand binding and signaling mechanisms. Trends Pharmacol. Sci. 39, 75–89 (2018). [DOI] [PubMed] [Google Scholar]
- 34.J. Kim, J. Wess, A. M. van Rhee, T. Schöneberg, K. A. Jacobson, Site-directed mutagenesis identifies residues involved in ligand recognition in the human A2A adenosine receptor. J. Biol. Chem. 270, 13987–13997 (1995). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.D. Thimm, A. C. Schiedel, F. F. Sherbiny, S. Hinz, K. Hochheiser, D. C. Bertarelli, A. Maass, C. E. Müller, Ligand-specific binding and activation of the human adenosine A2B receptor. Biochemistry 52, 726–740 (2013). [DOI] [PubMed] [Google Scholar]
- 36.K. K. Palaniappan, Z. G. Gao, A. A. Ivanov, R. Greaves, H. Adachi, P. Besada, H. O. Kim, A. Y. Kim, S. A. Choe, L. S. Jeong, K. A. Jacobson, Probing the binding site of the A1 adenosine receptor reengineered for orthogonal recognition by tailored nucleosides. Biochemistry 46, 7437–7448 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Z. G. Gao, H. T. Duong, T. Sonina, S. K. Kim, P. Van Rompaey, S. Van Calenbergh, L. Mamedova, H. O. Kim, M. J. Kim, A. Y. Kim, B. T. Liang, L. S. Jeong, K. A. Jacobson, Orthogonal activation of the reengineered A3 adenosine receptor (neoceptor) using tailored nucleoside agonists. J. Med. Chem. 49, 2689–2702 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.C. de Graaf, G. Song, C. Cao, Q. Zhao, M. W. Wang, B. Wu, R. C. Stevens, Extending the structural view of class B GPCRs. Trends Biochem. Sci. 42, 946–960 (2017). [DOI] [PubMed] [Google Scholar]
- 39.Z. Cong, Y. L. Liang, Q. Zhou, S. Darbalaei, F. Zhao, W. Feng, L. Zhao, H. E. Xu, D. Yang, M. W. Wang, Structural perspective of class B1 GPCR signaling. Trends Pharmacol. Sci. 43, 321–334 (2022). [DOI] [PubMed] [Google Scholar]
- 40.T. Flock, C. N. J. Ravarani, D. Sun, A. J. Venkatakrishnan, M. Kayikci, C. G. Tate, D. B. Veprintsev, M. M. Babu, Universal allosteric mechanism for Gα activation by GPCRs. Nature 524, 173–179 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.M. C. Peeters, G. J. van Westen, D. Guo, L. E. Wisse, C. E. Müller, M. W. Beukers, A. P. Ijzerman, GPCR structure and activation: An essential role for the first extracellular loop in activating the adenosine A2B receptor. FASEB J. 25, 632–643 (2011). [DOI] [PubMed] [Google Scholar]
- 42.B. F. Seibt, A. C. Schiedel, D. Thimm, S. Hinz, F. F. Sherbiny, C. E. Müller, The second extracellular loop of GPCRs determines subtype-selectivity and controls efficacy as evidenced by loop exchange study at A2 adenosine receptors. Biochem. Pharmacol. 85, 1317–1329 (2013). [DOI] [PubMed] [Google Scholar]
- 43.A. S. Doré, N. Robertson, J. C. Errey, I. Ng, K. Hollenstein, B. Tehan, E. Hurrell, K. Bennett, M. Congreve, F. Magnani, C. G. Tate, M. Weir, F. H. Marshall, Structure of the adenosine A2A receptor in complex with ZM241385 and the xanthines XAC and caffeine. Structure 19, 1283–1293 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.R. Askalan, P. J. Richardson, Role of histidine residues in the adenosine A2A receptor ligand binding site. J. Neurochem. 63, 1477–1484 (1994). [DOI] [PubMed] [Google Scholar]
- 45.C. R. Hiley, F. E. Bottrill, J. Wamock, P. J. Richardson, Effects of pH on responses to adenosine, CGS 21680, carbachol and nitroprusside in the isolated perfused superior mesenteric arterial bed of the rat. Br. J. Pharmacol. 116, 2641–2646 (1995). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.S. Hinz, W. M. Alnouri, U. Pleiss, C. E. Müller, Tritium-labeled agonists as tools for studying adenosine A2B receptors. Purinergic Signal 14, 223–233 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.X. Wang, W. Jespers, R. Prieto-Díaz, M. Majellaro, A. P. IJzerman, G. J. P. van Westen, E. Sotelo, L. H. Heitman, H. Gutiérrez-de-Terán, Identification of V6.51L as a selectivity hotspot in stereoselective A2B adenosine receptor antagonist recognition. Sci. Rep. 11, 14171 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.D. Wacker, S. Wang, J. D. McCorvy, R. M. Betz, A. J. Venkatakrishnan, A. Levit, K. Lansu, Z. L. Schools, T. Che, D. E. Nichols, B. K. Shoichet, R. O. Dror, B. L. Roth, Crystal structure of an LSD-bound human serotonin receptor. Cell 168, 377–389.e12 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.D. Cao, J. Yu, H. Wang, Z. Luo, X. Liu, L. He, J. Qi, L. Fan, L. Tang, Z. Chen, J. Li, J. Cheng, S. Wang, Structure-based discovery of nonhallucinogenic psychedelic analogs. Science 375, 403–411 (2022). [DOI] [PubMed] [Google Scholar]
- 50.M. Betti, D. Catarzi, F. Varano, M. Falsini, K. Varani, F. Vincenzi, D. Dal Ben, C. Lambertucci, V. Colotta, The aminopyridine-3,5-dicarbonitrile core for the design of new non-nucleoside-like agonists of the human adenosine A2B receptor. Eur. J. Med. Chem. 150, 127–139 (2018). [DOI] [PubMed] [Google Scholar]
- 51.X. Wang, W. Jespers, B. J. Bongers, M. C. C. Habben Jansen, C. M. Stangenberger, M. A. Dilweg, H. Gutiérrez-de-Terán, A. P. IJzerman, L. H. Heitman, G. J. P. van Westen, Characterization of cancer-related somatic mutations in the adenosine A2B receptor. Eur. J. Pharmacol. 880, 173126 (2020). [DOI] [PubMed] [Google Scholar]
- 52.R. Nehmé, B. Carpenter, A. Singhal, A. Strege, P. C. Edwards, C. F. White, H. Du, R. Grisshammer, C. G. Tate, Mini-G proteins: Novel tools for studying GPCRs in their active conformation. PLOS ONE 12, e0175642 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.S. H. W. Scheres, RELION: Implementation of a Bayesian approach to cryo-EM structure determination. J. Struct. Biol. 180, 519–530 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.A. Punjani, J. L. Rubinstein, D. J. Fleet, M. A. Brubaker, cryoSPARC: Algorithms for rapid unsupervised cryo-EM structure determination. Nat. Methods 14, 290–296 (2017). [DOI] [PubMed] [Google Scholar]
- 55.J. Jumper, R. Evans, A. Pritzel, T. Green, M. Figurnov, O. Ronneberger, K. Tunyasuvunakool, R. Bates, A. Žídek, A. Potapenko, A. Bridgland, C. Meyer, S. A. A. Kohl, A. J. Ballard, A. Cowie, B. Romera-Paredes, S. Nikolov, R. Jain, J. Adler, T. Back, S. Petersen, D. Reiman, E. Clancy, M. Zielinski, M. Steinegger, M. Pacholska, T. Berghammer, S. Bodenstein, D. Silver, O. Vinyals, A. W. Senior, K. Kavukcuoglu, P. Kohli, D. Hassabis, Highly accurate protein structure prediction with AlphaFold. Nature 596, 583–589 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.M. Varadi, S. Anyango, M. Deshpande, S. Nair, C. Natassia, G. Yordanova, D. Yuan, O. Stroe, G. Wood, A. Laydon, A. Žídek, T. Green, K. Tunyasuvunakool, S. Petersen, J. Jumper, E. Clancy, R. Green, A. Vora, M. Lutfi, M. Figurnov, A. Cowie, N. Hobbs, P. Kohli, G. Kleywegt, E. Birney, D. Hassabis, S. Velankar, AlphaFold Protein Structure Database: Massively expanding the structural coverage of protein-sequence space with high-accuracy models. Nucleic Acids Res. 50, D439–d444 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.E. F. Pettersen, T. D. Goddard, C. C. Huang, G. S. Couch, D. M. Greenblatt, E. C. Meng, T. E. Ferrin, UCSF Chimera—A visualization system for exploratory research and analysis. J. Comput. Chem. 25, 1605–1612 (2004). [DOI] [PubMed] [Google Scholar]
- 58.P. Emsley, B. Lohkamp, W. G. Scott, K. Cowtan, Features and development of Coot. Acta Crystallogr. D Biol. Crystallogr. 66, 486–501 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.P. D. Adams, P. V. Afonine, G. Bunkóczi, V. B. Chen, I. W. Davis, N. Echols, J. J. Headd, L. W. Hung, G. J. Kapral, R. W. Grosse-Kunstleve, A. J. McCoy, N. W. Moriarty, R. Oeffner, R. J. Read, D. C. Richardson, J. S. Richardson, T. C. Terwilliger, P. H. Zwart, PHENIX: A comprehensive Python-based system for macromolecular structure solution. Acta Crystallogr. D Biol. Crystallogr. 66, 213–221 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Figs. S1 to S9





