Introduction

Epinephrine and norepinephrine are neurotransmitters of the sympathetic nervous system, and hormones secreted by the adrenal medulla1,2. They function through nine distinct human adrenergic receptor (AR) subtypes: α1-type (α1A, α1B, α1D), α2-type (α2A, α2B, α2C,) and β-type (β1, β2, and β3)3,4. Due to the lack of selective pharmacological agonists for α1-ARs, the therapeutic potential of α1-ARs has been largely unexplored5,6,7,8,9. However, mouse genetic studies with individual gene deletions have demonstrated distinct but overlapping physiological functions of α1-AR subtypes in the regulation of blood pressure, cardiac hypertrophy, vascular smooth muscle contraction, neurotransmission, learning and memory, and metabolism10,11,12,13,14,15,16,17. For example, the α1A-AR subtype is a vasopressor expressed in resistance arteries and is required for normal arterial blood pressure regulation11. In addition, α1B-AR in human coronary endothelial cells mediates vasodilation18,19.

α1-ARs couple to the Gq family of G-proteins, leading to the stimulation of phospholipase C-β to cleave phosphatidylinositol-4,5-bisphosphate into inositol-1,4,5-trisphosphate and 2-diacylglycerol20. The former promotes the release of Ca2+ from intracellular stores, while the latter activates protein kinase C21,22. While some pharmacological compounds exist that can target α1-ARs, there are only two structurally related α1A-specific agonists, and no α1B-AR and α1D-AR selective agonists have been reported17. A61603 is a high affinity, selective α1A-AR agonist which shows almost no activity at α1B-AR and α1D-AR7,23. Furthermore, there are no structures available for the active states of α1-ARs. In this paper, we use cryo-EM to determine the structures of α1A-AR/Gq signaling complexes with epinephrine or A61603. These structures reveal the molecular basis for the binding specificity of A61603 for α1A-AR, and the different conformations of epinephrine in interacting with α-ARs versus β-ARs. Gaussian accelerated molecular dynamics (GaMD) simulations and functional studies provide further insights into the mechanisms of specificity, ultimately enabling the validation of key sites that determine the ability of A61603 to specifically activate α1A-AR but not α1B-AR24,25.

Results

Cryo-EM structures of A61603–α1A-AR–Gq and epinephrine–α1A-AR–Gq signaling complexes

To understand the agonist specificity of α1-AR, we solved cryo-EM structures of human α1A-AR bound to A61603 (a synthetic specific agonist for α1A-AR) at 2.6 Å or epinephrine (an endogenous agonist for all ARs) at 3.0 Å, both in complex with its cognate signaling Gq heterotrimer (mini-Gαq12) (Fig. 1a, b, Supplementary Figs. 13, and Supplementary Table 1). Comparisons of these structures should provide insights into subtype-specific agonist binding. Overall, the structures of the α1A-AR–Gq complex in the presence of epinephrine or A61603 are similar (Fig. 1a, b). However, there are local conformational differences, especially in the ligand-binding pockets (see below). While some of the interacting residues are common to both ligands, epinephrine, and A61603 each make a unique set of interactions in the orthosteric ligand-binding pocket (see below).

Fig. 1: Cryo-EM structures of the complexes of A61603–α1A-AR–Gq and epinephrine–α1A-AR–Gq.
figure 1

a The density map, the model, and the ligand-binding pocket of A61603–α1A-AR–Gq are shown. b The density map, the model, and the ligand-binding pocket of epinephrine–α1A-AR–Gq are shown. A61603-bound α1A-AR is colored in purple. Epinephrine-bound α1A-AR is colored orange. Gαq in green. Gβ in blue. Gγ in yellow. ce Different views of the inactive state α1B-AR (gray; PDB 7B6W) and the active state of α1A-AR in complex with Gq (this work).

Since the A61603–α1A-AR–Gq and epinephrine–α1A-AR–Gq structures reveal the first active state conformation of α1-AR family receptors, we first investigated the structural basis of activation. For the α1-AR family, there is only one recently reported structure: the X-ray crystal structure of the inactive state of α1B-AR bound with the inverse agonist cyclazosin (PDB: 7B6W)26. We thus compared our active state structures of α1A-AR with this inactive state structure of α1B-AR. This comparison revealed characteristic conformational changes associated with class A GPCR activation (Fig. 1c–e, and Supplementary Fig. 4)27. Since the structures of α1A-AR bound with A61603 and epinephrine were similar, we focused our comparison on the A61603–α1A-AR–Gq complex. The overall root-mean-square deviation between the structures of the active α1A-AR and inactive α1B-AR is 1.6 Å over 198 Cα atoms. The largest structural changes upon activation occur on the cytoplasmic side of α1-AR (Fig. 1c–e), with an outward rotation of TM6 by ~10 Å (measured at the Cα of A6.33) (The superscript is the Ballesteros-Weinstein numbering28) (Fig. 1c, e), and an inward ~5 Å movement of TM7 (measured at the Cα of Y7.53) (Fig. 1c, e). In addition to these TM conformational changes, rearrangements of side chains of certain residues are observed as part of the α1-AR activation process. Just below the orthosteric ligand-pocket, the rotameric change of W6.48 (within the CWxP motif) denotes the opening of TM6 in class A GPCRs for G-protein engagement27 (Supplementary Fig. 4a). A layer below the CWxP motif, the rotation of F5.47 causes the change of L5.51, accompanied by the translational movement of V6.45, F6.44 (part of the PIF motif), and W6.48 to form new contacts (Supplementary Fig. 4b). Moving closer to the G-protein-interacting site below the PIF motif, the highly conserved NPxxY motif at the cytoplasmic end of TM7 is another key micro-switch of GPCR activation27 (Supplementary Fig. 4c). TM7 rotates around the NPxxY motif. This moves Y7.53 toward the position that was occupied by TM6 in the inactive structure (Supplementary Fig. 4c). Among the G-protein-interacting residues, the rearrangement of side chains in the highly conserved D(E)/RY motif in TM3 is critical for GPCR activation27 (Supplementary Fig. 4d). Therefore, α1-ARs undergo conformational changes propagating from the orthosteric ligand-binding site to the G-protein-interacting site during its activation.

Different conformations of epinephrine bound to α-ARs and β-ARs

Epinephrine is a chiral endogenous full agonist common for all ARs. In the complex of epinephrine–α1A-AR–Gq, the para-hydroxyl group of the catechol ring of epinephrine forms H-bond with the hydroxyl side-chain of S1885.43 (Fig. 2a–c and Supplementary Fig. 5). The protonated amine of epinephrine forms a salt bridge with residue D1063.32 (Fig. 2a–c and Supplementary Fig. 5). Previous experiments have shown that this interaction is critical for both affinity and efficacy29. Y3167.42 and W3137.39 stabilize this salt bridge through a hydrogen bond between D1063.32 and Y3167.42, and π–π interactions between Y3167.42 and W3137.39 (Fig. 2b, c). The catechol ring forms π–π interactions with F2886.51 and F2896.52 (Fig. 2c). F3127.38 acts as a lid covering the catechol ring of epinephrine from the extracellular side (Fig. 2b). Furthermore, epinephrine adopted a binding conformation similar to the cryo-EM structure in the GaMD simulations (Supplementary Fig. 5a), with mostly <2 Å root-mean-square derivation (RMSD) (Supplementary Fig. 5b). The ligand maintained stable interactions with residues D1063.32 and S1885.43 in α1A-AR at 3.82 ± 0.33 Å and 3.10 ± 0.33 Å, respectively (Supplementary Fig. 5c, d).

Fig. 2: Interactions between epinephrine and α1A-AR.
figure 2

a Chemical structure of epinephrine. b Schematic diagram of the epinephrine-binding pocket of α1A-AR from the cryo-EM structure is shown. c Ligplot diagrammatic representation of interactions between epinephrine and α1A-AR. d Comparisons of the conformations of epinephrine bound with α1A-AR, β1-AR, and β2-AR.

The conformation of epinephrine in the complex with α1A-AR is different from that in the previously reported complexes with β-ARs (Fig. 2d and Supplementary Fig. 6). Currently there are no structures available for epinephrine in complex with any β-ARs and a G-protein. However, there are two structures of epinephrine bound to β1-AR or β2-AR with a nanobody 6B9 that stabilizes β-ARs in an active state30,31. In the complex of epinephrine–β1-AR–nanobody 6B9 and of epinephrine–β2-AR–nanobody 6B930,31, the β-carbon hydroxyl and the N-methyl group face opposite directions when compared with the conformation in complex with α1A-AR. The β-carbon hydroxyl faces the intracellular side, and the N-methyl group points toward TM3 when epinephrine was complexed with β-ARs (Fig. 2d, Supplementary Fig. 6). On the other hand, the β-carbon hydroxyl group faces the extracellular side, and the N-methyl group points to Y3167.42 on TM7, when epinephrine was in the complex with α1A-AR (Fig. 2b, Supplementary Fig. 6). These two different epinephrine conformations result from the rotation around the chemical bond linking the catechol ring and the β-carbon of epinephrine (Supplementary Fig. 6). In β-ARs, N3637.38 (in β1-AR) and N3127.38 (in β2-AR) interacted through hydrogen bonds with the β-carbon hydroxyl and the N-methyl group of epinephrine. In α1A-AR, it is F3127.38 at the equivalent position, and its side-chain did not form hydrogen bonds with epinephrine (Supplementary Fig. 6). The different conformations of epinephrine, when bound with β-ARs and α-ARs, might explain the higher affinity of epinephrine for β-ARs than α-ARs32. Together these data show that epinephrine adopts different conformations in order to interact with α and β families of ARs.

Specific interaction between A61603 and α1A-AR

Since A61603 is a selective agonist for α1A-AR23, we asked what residues mediated this specific interaction. In the complex of A61603–α1A-AR–Gq, S1885.43 forms multiple interactions with A61603 (Fig. 3a–c). The hydroxyl group of S1885.43 forms H-bonds with the hydroxy attached to the tetrahydronaphthalene group, the amide, and the oxygen atom of the methanesilfonamide group of A61603 (Fig. 3b, c). In addition, the side-chain of D1063.32 forms a salt bridge with the imidazoline group of A61603 (Fig. 3b, c), and the backbone amide hydrogen of A1895.44 interacts with the oxygen atom of the methanesilfonamide group of A61603 (Fig. 3b, c). There are also extensive hydrophobic interactions between A61603 and α1A-AR (Fig. 3c). These include V1073.33, C1103.36, I17845.52, W2856.48, F2886.51, F2896.52, M2926.55, F3127.38, G3157.41 and Y3167.42 (Fig. 3c). Since this is the first structure of A61603 bound to any proteins, we performed GaMD simulations to validate the ligand pose (Fig. 3d–g and Supplementary Fig. 7). In these simulations, the conformation of A61603 is similar to that observed in the cryo-EM structure (Fig. 3d, e). The imidazoline group of A61603 and residue D1063.32 in α1A-AR maintained the stable salt-bridge interaction during the GaMD simulations (Fig. 3f). Residue S1885.43 in α1A-AR formed hydrogen bonds with different atoms (e.g., N, O, and O2) in the methanesulfonamide group of A61603 in the GaMD simulations (Fig. 3g and Supplementary Fig. 7). Together, these data reveal the molecular interactions between A61603 and α1A-AR.

Fig. 3: Interactions between A61603 and α1A-AR.
figure 3

a Chemical structure of A61603. b Schematic diagram of the A61603-binding pocket of α1A-AR from the cryo-EM structure is shown. c Ligplot diagrammatic representation of interactions between A61603 and α1A-AR. d Comparison of the binding poses of A61603 in the cryo-EM structure (in green color) and from GaMD simulations (in blue color). e The A61603 RMSD by comparing the conformations of A61603 in the cryo-EM structure and from GaMD simulations. f, g Time courses of the distances between A61603 and specific residues of α1A-AR were calculated from the GaMD simulations: the CG atom of D1063.32 and the N2 atom of A61603 (f), and the OG atom of S1885.43 and the N atom of A61603 (g). Three independent 500 ns simulations are shown.

Molecular basis for the discrimination between α1A-AR and α1B-AR by A61603

To further probe the structural basis for the specificity of A61603 for α1A-AR, we compared the ligand-binding pockets of all three α1-AR subtypes (Supplementary Fig. 8). Among the ligand-binding pocket residues, α1A-AR and α1B-AR have almost identical composition with the exception of three residues: V1855.40 in α1A-AR but A2045.40 in α1B-AR, A1895.44 in α1A-AR and S2085.44 in α1B-AR, and M2926.55 in α1A-AR but L3146.55 in α1B-AR (Fig. 4a and Supplementary Fig. 8b).

Fig. 4: Participation of V185, A189, and M292 of α1A-AR in interacting with A61603.
figure 4

a Comparison of the ligand-binding pockets of α1A-AR and α1B-AR. b Hydrophobic interaction between the methanesulfonamide group of A61603 and the three residues V185, A189, and M292 of α1A-AR. c Steric clash between the docked A61603 and α1B-AR.

In the A61603–α1A-AR–Gq structure, these three residues form a hydrophobic surface surrounding the methanesulfonamide group of A61603 (Fig. 4b). To understand why A61603 could not bind to α1B-AR, we modeled a complex of α1B-AR and A61603 (Fig. 4c). Based on our cryo-EM structure of the active state of α1A-AR with A61603 and the structure of the inactive state of α1B-AR, the modeled A61603–α1B-AR complex showed a steric clash between the methanesulfonamide group of A61603 and L3146.55 (Fig. 4c). In α1D-AR, the corresponding residue is also a Leu residue (Supplementary Fig. 8b). Together, these observations suggest a molecular basis for the specific binding of A61603 to α1A-AR, but not α1B-AR and α1D-AR.

Based on insights from our structures and molecular modeling, we attempted to engineer α1A-AR variants that would lose the response to A61603, and α1B-AR mutants that would gain the response to A61603. These loss-of-function and gain-of-function studies should provide evidence for the molecular basis of A61603-binding specificity. Based on the above data, we focused on three residues V1855.40, A1895.44, and M2926.55. We generated point mutants of α1A-AR(V185A), α1A-AR(A189S), and α1A-AR(M292L) by mutating these residues to the corresponding residues in α1B-AR. Since these residues are involved in hydrophobic interactions with A61603 (Fig. 3), we also made a mutant with all three residues mutated α1A-AR(V185A, A189S, M292L). Since α1A-AR is coupled to Gq, we used calcium responses as a functional readout (Fig. 5). As a control, wild-type α1A-AR produced robust dose-dependent Ca2+ responses to A61603 with an EC50 of ~1 nM with a similar amplitude at saturating doses compared to saturating epinephrine (Fig. 5a, c). The triple mutant α1A-AR(V185A, A189S, M292L) reduced the apparent affinity of the A61603 response by more than 1000-fold (Fig. 5b, c). This triple mutant was still functional in response to epinephrine (Fig. 5b). The single-point mutants of V185A and M292L also decreased the α1A-AR response to A61603, while the A189S mutant showed wild-type-like responses (Fig. 5c).

Fig. 5: Functional studies of mutant α1A-AR and α1B-AR.
figure 5

a Representative Ca2+ response traces of cells expressing wild-type α1A-AR in response to different concentrations of A61603. b Representative Ca2+ response traces of cells expressing the triple mutant α1A-AR(V185A, A189S, M292L) in response to different concentrations of A61603 and a saturating concentration of epinephrine. c Dose–response curves for different α1A-AR constructs in response to A61603. n = 3 independent experiments. Data are presented as mean ± SD. d Representative Ca2+ response traces of cells expressing wild-type α1B-AR in response to different concentrations of A61603 and a saturating concentration of epinephrine. e Representative Ca2+ response traces of cells expressing the triple mutant α1B-AR(A204V, S208A, L314M) in response to different concentrations of A61603. f Dose–response curves for different α1B-AR constructs in response to A61603. Calcium imaging were analyzed using ImageJ and NIS-Elements Advance Research 5.2.6 software. Each single cell was manually selected as region of interest and the fluorescence intensity was normalized to the baseline before agonist application. An average trace (60-200 cells per recording) was used to quantify maximum fluorescence amplitude after drug application. n = 3 independent experiments. Data are presented as mean ± SD. Dose–response curve was done using nonlinear fit in GraphPad Prism.

In parallel, GaMD simulations were performed on α1A-AR(V185A), α1A-AR(A189S), α1A-AR(M292L), and α1A-AR(V185A, A189S, M292L) (Supplementary Fig. 9). During the GaMD simulations, RMSDs of A61603 in α1A-AR(V185A), α1A-AR(M292L), and α1A-AR(V185A, A189S, M292L) relative to the starting cryo-EM structure were 4.62 ± 1.68 Å, 3.21 ± 1.05 Å, and 3.99 ± 1.13 Å, respectively (Supplementary Fig. 9a, g, j). Hence, α1A-AR(V185A), α1A-AR(M292L), and α1A-AR(V185A, A189S, M292L) reduced the binding of A61603 to the receptor (Supplementary Fig. 9a, g, j). On the other hand, A61603 in α1A-AR(A189S) exhibited a smaller RMSD with lower fluctuations in the GaMD simulations (Supplementary Fig. 9d). The distances between the imidazoline group of A61603 and residue D1063.32 in α1A-AR(V185A), α1A-AR(A189S), α1A-AR(M292L), and α1A-AR(V185A, A189S, M292L) were 4.25 ± 0.98 Å, 3.43 ± 0.47 Å, 4.13 ± 1.33 Å, and 4.38 ± 1.27 Å, respectively (Supplementary Fig. 9b, e, h, k). Therefore, the interaction between the imidazoline group of A61603 and residue D1063.32 was disrupted in α1A-AR(V185A), α1A-AR(M292L), and α1A-AR(V185A, A189S, M292L), but still maintained in α1A-AR(A189S), during the simulations (Supplementary Fig. 9b, e, h, k). This is consistent with our functional data where V185A and M292L produced a large impairment of A61603 responses while A189S maintained WT-like properties. For further information, we performed molecular mechanics/generalized Born surface area (MM/GBSA) binding free energy calculations using GaMD trajectories33,34,35. Due to the inherent inaccuracy of entropy calculations, we focused on comparing only the enthalpy values (Supplementary Table 2). The A61603-binding free energy was less favorable in the α1A-AR mutants, i.e., −22.18 ± 0.73 kcal/mol for α1A-AR(V185A) and −11.52 ± 2.47 kcal/mol for α1A-AR(V185A, A189S, M292L), compared to −26.73 ± 0.22 kcal/mol for the wild-type α1A-AR. α1A-AR(A189S) exhibited a similar binding free energy of A61603 (−26.54 ± 1.89 kcal/mol) as the wild-type α1A-AR. However, an exception was observed in α1A-AR(M292L), where the binding free energy was −27.88 ± 1.53 kcal/mol. This likely resulted from the exclusion of the entropy and MM/GBSA free energy calculations of GaMD simulation frames without the energetic reweighting36. In general, the systems with a stronger functional response to A61603-binding exhibited more favorable binding free energy compared to those with weaker or no functional responses. Therefore, the GaMD simulation findings correlated well with the calcium response data.

Conversely, we attempted to convert α1B-AR from non-responsive to responsive to A61603. This type of exercise is usually difficult given that many residues are typically involved in ligand interactions. However, based on our above data, most of the agonist interacting residues are the same between α1A-AR and α1B-AR, except for the three residues (Supplementary Fig. 8), suggesting that it may be possible to replace the three distinct residues in α1B-AR by those in α1A-AR. We thus generated the corresponding triple mutant α1B-AR(A204V, S208A, L314M). Wild-type α1B-AR did not respond to A61603 at concentrations up to 10 μM, while it was responsive to epinephrine (Fig. 5d). Remarkably, the triple mutant α1B-AR(A204V, S208A, L314M) responded to A61603 in a dose-dependent manner with an EC50 around 10 nM (Fig. 5e, f). The single-point mutants α1B-AR(A204V), α1B-AR(S208A) or α1B-AR(L314M) alone did not respond to A61603 (Fig. 5f), indicating that more of a large-scale reshaping of the orthosteric site than can be achieved by a single mutation is needed to rescue the binding. In parallel, we have built a homology model of the active α1B-AR using the cryo-EM structure of A61603-bound α1A-AR as the template and conducted GaMD simulations on the wild-type α1B-AR, α1B-AR(S208A), and α1B-AR(A204V, S208A,L314M). In wild-type α1B-AR and α1B-AR(S208A), A61603 exhibited large RMSDs relative to the starting conformation, with reduced interactions between A61603 and α1B-AR (Supplementary Fig. 10). In contrast, A61603 in α1B-AR(A204V, S208A, L314M) exhibited low RMSD at ~2.5 Å, forming stable interactions with residues D1253.32 and S2075.43 in the receptor (Supplementary Fig. 10). α1B-AR(A204V, S208A, L314M) showed the strongest binding affinity (-28.21 ± 2.80 kcal/mol), followed by α1B-AR(S208A) (-25.81 ± 2.67 kcal/mol) and wild-type α1B-AR (-11.52 ± 2.47 kcal/mol) (Supplementary Table 2).

We also compared the root-mean-square fluctuations (RMSF) of A61603 in different receptor systems. In this context, larger RMSF values indicate weaker binding of A61603. The calculated RMSF values of A61603 in α1A-AR (WT), α1A-AR(V185A), α1A-AR(A189S), α1A-AR(M292L), and α1A-AR(V185A, A189S, M292L) were 1.46 ± 0.10 Å, 2.36 ± 0.52 Å, 1.10 ± 0.088 Å, 1.74 ± 0.39 Å, and 2.58 ± 0.33 Å, respectively (Supplementary Table 2). α1A-AR(A189S) (with a functional response to A61603) exhibited a similar RMSF value as wild-type α1A-AR, while the other α1A-AR mutants (with impaired functional responses to A61603) showed higher A61603 fluctuations. The RMSF values of A61603 in α1B-AR(A204V, S208A, L314M), α1B-AR(S208A) and wild-type α1B-AR were 0.98 ± 0.21 Å, 3.55 ± 1.97 Å, and 4.28 ± 0.30 Å, respectively (Supplementary Table 2). Additionally, we calculated the 2D free energy profiles of A61603 RMSD and the distance between the CG atom of D1063.32 and N2 atom of A61603, as well as the distance between the OG atom of S1885.43 and O2 atom of A61603 in both the α1A-AR and α1B-AR systems (Supplementary Fig. 11). In wild-type α1A-AR and α1A-AR(A189S), one low-energy state of A61603 with a narrow conformational space was sampled. Conversely, the other mutant α1A-ARs sampled larger conformational space with at least two distinct low-energy states. Among the α1B-AR systems, α1B-AR(A204V, S208A, L314M) sampled only one low-energy state, while α1B-AR(WT) and α1B-AR(S208A) sampled at least two low-energy states with much larger conformational space (Supplementary Fig. 12). These findings suggest that the systems with functional responses to A61603 binding are sampling only one stable A61603-binding pose, whereas the systems without functional responses to A61603 could explore larger conformational space, indicating reduced stability of A61603. Together, our studies present molecular insights into the mechanism underlying the discrimination between α1A-AR and α1B-AR by A61603. The observed differences in selectivity are consistent with a model that is affinity-based, as opposed to efficacy-based.

Discussion

α1-ARs are one of the three major subfamilies of ARs. Due to the lack of specific agonists, the physiological functions of α1-ARs are still not completely revealed, and their therapeutic potentials in treating cardiovascular, neurological, neuropsychiatric, and inflammatory disorders are not fully exploited. Here we have investigated the structural basis of activation and agonist interactions with α1-AR. Our data shows that M2926.55 and V1855.40 of α1A-AR define the specificity of A61603 for α1A-AR. Hydrophobic M2926.55 in α1A-AR corresponds to hydrophobic L6.55 in α1B-AR and α1D-AR, while aromatic Y6.55 in α2-AR and polar N6.55 in β-ARs (Supplementary Fig. 8b). Hydrophobic V1855.40 in α1A-AR corresponds to A5.40 in α1B-AR, α1D-AR, β1-AR and β2-AR, I5.40 in α2B-AR and α2C-AR, and V5.40 in α2A-AR and β3-AR (Supplementary Fig. 8b). Given that these residues form one hydrophobic surface (Fig. 4), this hydrophobic pocket should be explored for the design of AR subtype-specific ligands.

Comparisons of the two agonists (epinephrine and A61603) revealed that they share some conserved interactions while also displaying certain unique interactions (Supplementary Fig. 13). Epinephrine is a monoamine that contains an amino group linked to an aromatic ring by a two-carbon chain (Fig. 2a). In α1A-AR, β1-AR and β2-AR, the aromatic ring forms π–π interactions to F6.51 and F6.52 (Fig. 2b, c). The positively charged amino group forms a salt bridge to D3.32 (Fig. 2c). Although A61603 is not a monoamine, it uses the imidazoline group to form the same salt bridge with the conserved D3.32 (Fig. 3c and Supplementary Fig. 13). Its polycyclic aromatic ring consists of a benzene fused to cyclohexane (Fig. 3a). The benzene of A61603, as the catechol ring of epinephrine, forms π–π interactions to the same F6.51 and F6.52 (Fig. 3b, c, and Supplementary Fig. 13). Furthermore, the para-hydroxyl group of epinephrine and the hydroxyl group of A61603 interact with the same S5.43 (Figs. 2, 3 and Supplementary Fig. 13). As discussed above, the molecular basis for the receptor specificity lies in the hydrophobic interaction of the methenesulfonamide group of A61603 with the hydrophobic surface formed by M6.55 and V5.40, whereas the meta-hydroxyl group of epinephrine occupies this space (Figs. 2 and 3). The differences in selectivity are consistent with a model that is affinity-based, as opposed to efficacy-based. Together our results provide insights into the design and development of selective agonists targeting individual α1-AR subtypes, and polypharmacological agonists tailored for multiple related receptors.

Methods

Expression and purification of α1A-AR, Gq, and scFv16

The human α1A-AR construct used in the cryo-EM study was optimized by truncation of N-terminal residues 1-14, third intracellular loop (ICL3) residues 228-255, and C-terminal residues 351-466. A hemagglutinin (HA) signal peptide, a FLAG tag, and a T4 lysozyme were fused to its N-terminus, and a PreScission protease cleavage site, an eGFP, and an 8xHis tag were added to its C-terminus. This construct was subcloned into pFastBac1 vector for expression in Spodoptera frugiperda Sf9 insect cells. Sf9 cells were grown to 2 to 3 million cells per ml in ESF 921 protein-free medium (Expression Systems) before 50 ml of baculoviruses were added for infection. 72 h later, cells were harvested by centrifugation, flash frozen in liquid nitrogen, and stored at −80 °C until use. For α1A-AR purification, thawed cell pellets were lysed with 1% n-Dodecyl-β-d-Maltopyranoside (DDM, Anatrace) in a buffer containing 25 mM HEPES (pH 7.0), 350 mM NaCl, 100 μM A61603 (Cayman Chemicals) or epinephrine (MP Biomedicals), supplemented with protease inhibitors (0.5 μM PMSF, 2 μg/ml leupeptin, 0.8 μM aprotinin and 2 μM Pepstatin A, Goldbio) at 4 °C. Insoluble fractions were removed by ultracentrifugation at 142,000 × g for 1 h at 4 °C. The supernatant was then incubated with house-made GFP nanobody beads for at least 2 h, and washed with a buffer containing 25 mM HEPES (pH 7.0), 350 mM NaCl, 0.05% DDM, and 100 μM A61603 or epinephrine before PreScission protease was added to elute α1A-AR from GFP nanobody beads overnight at 4 °C. Eluted α1A-AR was further purified by size-exclusion chromatography using a Superdex 200 Increase 10/300 column (GE Healthcare) equilibrated with a buffer containing 25 mM HEPES (pH 7.0), 150 mM NaCl, 0.02% Lauryl Maltose Neopentyl Glycol (LMNG, Anatrace), and 100 μM A61603 or epinephrine. Peak fractions were pooled and concentrated for complex assembly.

The expression constructs of heterotrimeric mini-GαqiN–Gβ1–Gγ2 and scFv16 were kindly provided by Dr. Bryan L Roth (University of North Caroline at Chapel Hill)37. For the expression of heterotrimeric mini-GαqiN–Gβ1–Gγ2, a single virus that encodes all three subunits, including mini-GαqiN, Gβ1, and Gγ2, with a 6xHis tag fused to the N-terminus of Gβ1, was used to infect Sf9 cells. 48 h post infection, cells were harvested by centrifugation, flash frozen in liquid nitrogen, and stored at −80 °C. Cell pellets were thawed in a lysis buffer containing 20 mM HEPES (pH 7.5), 100 mM NaCl, 2 mM β-mercaptoethanol, 1 mM MgCl2, 0.1 mM GDP, 0.2% (v/v) Triton X-100, supplemented with protease inhibitors (0.5 μM PMSF, 2 μg/ml leupeptin, 0.8 μM aprotinin and 2 μM Pepstatin A) at 4 °C and insoluble fractions were removed by centrifugation at 142,000 × g for 1 h. Supernatant was collected and incubated with HisPur Ni-NTA Resin (Thermo Fisher Scientific) for 2 h at 4 °C. Resin was then washed with 20 mM imidazole in the lysis buffer, and eluted with 250 mM imidazole in the lysis buffer. Eluted heterotrimeric G-protein was further purified by size-exclusion chromatography using a Superdex 200 Increase 10/300 column (GE Healthcare) equilibrated with a buffer containing 20 mM HEPES (pH 7.5), 100 mM NaCl, 0.02% LMNG, 0.1 mM TCEP, 1 mM MgCl2, and 10 μM GDP. Peak fractions were pooled and concentrated for complex assembly.

scFv16 was expressed and purified from Sf9 cells as a secreted protein37. 72 h after infection, Sf9 cell culture media expressing scFv16 was collected and PH balanced to pH 8.0 by addition of Tris buffer. 10 mM calcium chloride was then added to quench the chelating agents with stirring at room temperature for 1 h. The precipitate was removed by centrifugation at 8000 × g for 30 min. The supernatant was incubated with HisPur Ni-NTA Resin with stirring for 4 h at 4 °C. The Ni-NTA resin was then collected with centrifugation at 800 g and washed with 20 mM imidazole in a buffer containing 20 mM HEPES (pH 7.5) and 100 mM NaCl before 200 mM imidazole in the same buffer was used to elute the protein. The C-terminal 8xHis tag was removed by the treatment of PreScission protease overnight at 4 °C. The protein was then further purified by size-exclusion chromatography using a Superdex 200 Increase 10/300 column (GE Healthcare) equilibrated with a buffer containing 20 mM HEPES (pH 7.5) and 100 mM NaCl. Peak fractions were pooled and concentrated for complex assembly.

Protein complex assembly and purification

To assemble the α1A-AR–Gq–scFv16 complex, α1A-AR, and heterotrimeric Gq were mixed at a 1:1.5 molar ratio and incubated for 30 min at room temperature followed by the treatment of 0.4 U Apyrase (New England Biolabs) for another 30 min at room temperature. A 1.5 molar access of scFv16 was then added to the mixture and incubated overnight at 4 °C. The next day, the mixture was centrifuged at 16,000 × g for 10 min to remove any precipitants. The supernatant was then loaded onto a Superdex 200 Increase 10/300 column equilibrated with 10 mM HEPES pH 7.0, 100 mM NaCl, 0.1 mM TCEP, 0.02% LMNG, and 30 μM A61603 or epinephrine. The elution fractions from a single peak containing pure α1A-AR–Gq–scFv16 complex concentrated to 2.5 mg/ml and used directly for making cryo-EM grids.

Cryo-EM data collection

3.5 µL of α1A-AR–Gq–scFv16 complex at a concentration of 2.5 mg/ml was applied to glow-discharged 400 mesh gold Quantifoil R1.2/1.3 holey carbon grids (Quantifoil Micro Tools) and vitrified using a Vitrobot Mark IV (Thermo Fisher Scientific/FEI) at 22 °C and 100% humidity. Micrographs were collected on a 300 kV Titan Krios electron microscope (Thermo Fisher Scientific/FEI) with a Gatan K3 direct electron detector (Gatan, Inc.) in super-resolution mode at a nominal ×81,000 and ×64,000 magnification for the A61603 and epinephrine-bound complexes, respectively. For the A61603-bound complex, 7,773 movies in the defocus range of −1.0 to −1.8 μm were recorded with a total accumulated dose of 50 e2. For the epinephrine-bound complex, 10,486 movies in the defocus range of −1.0 to −1.8 μm were recorded with a total accumulated dose of 52 e2.

Image processing, 3D reconstructions, modeling, and refinement

Super-resolution movies were aligned, two-times Fourier cropped, and dose-weighted using MotionCor2 implemented in Relion 4.0-beta38,39,40,41,42,43,44. The effects of contrast-transfer function were estimated with CTFFIND v4.1.845. For the A61603 dataset, Relion 2D template-based auto-picking was used to pick particles and the resulting particle stacks were two-times Fourier cropped and processed through 2D classification in CryoSparc v3.3.1 to remove false positives, receptor alone, or G-proteins alone particles46 (Supplementary Fig. 1). A stack of 1,497,830 intact complex particles were then subjected to 3D classification in Relion. For the epinephrine dataset, both Relion 2D and 3D template-based auto-picking was used to pick particles, and the resulting particle stacks from both 2D and 3D template-based auto-picking were separately two-times Fourier cropped and processed through 2D classification in CryoSparc v3.3.1 to remove false positives, receptor alone, or G-proteins alone particles46 (Supplementary Fig. 2). After Relion 3D classification, the intact complex particles stacks from 2D and 3D template-based auto-picking were selected and combined, and duplicate particles were removed. The resulting stack of 880,531 intact complex particles were then subjected to another round of 3D classification in Relion. Stacks of 773,300 (A61603) and 782,191 (epinephrine) particles that went to high resolution were selected after 3D classification and further polished by Bayesian polishing in Relion. 3D variability analysis in cryoSPARC revealed the twisting of the α1A-AR transmembrane region around a vertical axis, perpendicular to the membrane in both A61603 and epinephrine datasets. Subsets of 360,489 (A61603) and 219,834 (epinephrine) particles obtained by 3D variability analysis that yielded higher resolution in the transmembrane region were selected and non-uniform refinement of these particles in CryoSparc yielded maps with 2.6 Å-resolution (A61603) and 3.0 Å-resolution (epinephrine) using the 0.143 Fourier Shell Correlation criterion. These consensus stacks were also subjected to focused refinement of α1A-AR and Gαq region in CryoSparc. The focused refinement map showed significant improvement in the α1A-AR region compared to the consensus map. Models were built starting from AlphaFold predicted structure of inactive α1A-AR (AF-P35348-F1) and scFv16 bound Gq heterotrimer (PDB code 6WHA) in Coot v0.9.1, which combined with consensus and focused refinement maps, were used to generate composite maps in Phenix dev-469447,48. The resulting maps were super-sampled in Coot to 0.856 Å per pixel (A61603) and 0.8608 Å (epinephrine) with 320-voxel boxes. The A61603 and epinephrine bound α1A-AR–Gq–scFv16 models were real-space refined against the composite maps in Phenix, and a work/free half-map pair was used to ensure against over-fitting (Supplementary Figs. 1 and 2). The binding poses of A61603 and epinephrine were initially determined by density maps as well as their geometry restraints and were further supported by MD simulations49. All model statistics were validated using MolProbity50.

Gaussian accelerated molecular dynamics (GaMD)

GaMD is an enhanced sampling method that works by adding a harmonic boost potential to reduce the system energy barriers24,25. When the system potential \(V(\mathop{r}\limits^{ \rightharpoonup })\) is lower than a reference energy E, the modified potential \({V}^{*}(\mathop{r}\limits^{ \rightharpoonup })\) of the system is calculated as:

$${V}^{*}\left(\mathop{r}\limits^{ \rightharpoonup }\right)=V\left(\mathop{r}\limits^{ \rightharpoonup }\right)+\Delta V\left(\mathop{r}\limits^{ \rightharpoonup }\right)$$
$$\Delta V\left(\mathop{r}\limits^{ \rightharpoonup }\right)=\left\{\begin{array}{c}\frac{1}{2}k{\left(E-V\left(\mathop{r}\limits^{ \rightharpoonup }\right)\right)}^{2},\,V\left(\mathop{r}\limits^{ \rightharpoonup }\right) \, < \, E\\ 0,\, V\left(\mathop{r}\limits^{ \rightharpoonup }\right)\ge E,\hfill \end{array}\right.$$
(1)

where k is the harmonic force constant. The two adjustable parameters E and k are automatically determined on three enhanced sampling principles. First, for any two arbitrary potential values \({v}_{1}(\mathop{r}\limits^{ \rightharpoonup })\) and \({v}_{2}(\mathop{r}\limits^{ \rightharpoonup })\) found on the original energy surface, if \({V}_{1}(\mathop{r}\limits^{ \rightharpoonup }) \, < \, {V}_{2}(\mathop{r}\limits^{ \rightharpoonup })\), \(\Delta V\) should be a monotonic function that does not change the relative order of the biased potential values; i.e., \({V}_{1}^{*}(\mathop{r}\limits^{ \rightharpoonup }) \, < \, {V}_{2}^{*}(\mathop{r}\limits^{ \rightharpoonup })\). Second, if \({V}_{1}(\mathop{r}\limits^{ \rightharpoonup }) \, < \, {V}_{2}(\mathop{r}\limits^{ \rightharpoonup })\), the potential difference observed on the smoothened energy surface should be smaller than that of the original; i.e., \({V}_{2}^{*}(\mathop{r}\limits^{ \rightharpoonup }){-V}_{1}^{*}(\mathop{r}\limits^{ \rightharpoonup }) \, < \, {V}_{2}(\mathop{r}\limits^{ \rightharpoonup }){-V}_{1}(\mathop{r}\limits^{ \rightharpoonup })\). By combining the first two criteria and plugging in the formula of \({V}^{*}(\mathop{r}\limits^{ \rightharpoonup })\) and \(\Delta V\), we obtain

$${V}_{\max }\le E\le {V}_{\min }+\frac{1}{k},$$
(2)

Where \({V}_{\min }\) and \({V}_{\max }\) are the system minimum and maximum potential energies. To ensure that Eq. 2 is valid, k has to satisfy: \(k\le 1/({V}_{\max }-{V}_{\min })\). Let us define: \(k={k}_{0} \, \cdot \, 1/({V}_{\max }-{V}_{\min })\), then \(0{ \, < \, k}_{0}\le 1\). Third, the standard deviation (SD) of \(\Delta V\) needs to be small enough (i.e., narrow distribution) to ensure accurate reweighting using cumulant expansion to the second order: \({\sigma }_{\Delta V}=k(E-{V}_{{avg}}){\sigma }_{V}\le {\sigma }_{0}\), where \({V}_{{avg}}\) and \({\sigma }_{V}\) are the average and SD of \(\Delta V\) with \({\sigma }_{0}\) as a user-specified upper limit (e.g., \({10k}_{B}T\)) for accurate reweighting. When E is set to the lower bound \(E={V}_{\max }\) according to Eq. 2, \({k}_{0}\) can be calculated as

$${k}_{0}={{\min }}\left(1.0,\,{k}_{0}^{{\prime} }\right)=\min \left(1.0,\frac{{\sigma }_{0}}{{\sigma }_{V}} \, \cdot \, \frac{{V}_{\max }-{V}_{\min }}{{V}_{\max }-{V}_{{{{{\rm{avg}}}}}}}\right),$$
(3)

Alternatively, when the threshold energy E is set to its upper bound \(E={V}_{\min }+1/k\), \({k}_{0}\) is set to:

$${k}_{0}={k}_{0}^{\prime\prime}\equiv \left(1-\frac{{\sigma }_{0}}{{\sigma }_{V}}\right) \, \cdot \, \frac{{V}_{\max }-{V}_{\min }}{{V}_{{{{{\rm{avg}}}}}}-{V}_{\min }},$$
(4)

If \({k}_{0}^{\prime\prime}\) is calculated between 0 and 1. Otherwise, \({k}_{0}\) is calculated using Eq. 3.

System setup and simulation analysis

The epinephrine–α1A-AR–Gq and A61603–α1A-AR–Gq cryo-EM structures were used for setting up simulation systems to explore the stability of the complex. The missing resides in the Helix 8 of α1A-AR (KKAFQNVLR, residue number 334-342) were added by SWISS-MODEL51. The initial models of the A61603-bound mutants of α1A-AR(V185A), α1A-AR(A189S), α1A-AR(M292L), and α1A-AR(V185A, A189S, M292L) were built by VMD based on the A61603–α1A-AR–Gq cryo-EM structure. All six simulation systems were prepared with the CHARMM-GUI web server for using the membrane protein input generator52. The receptor was inserted into a palmitoyl-oleoyl-phosphatidyl-choline (POPC) bilayer. All chain termini were capped with neutral patches (acetyl and methylamide). All the disulfide bonds in the complexes that were resolved in the cryo-EM structures were maintained in the simulations. The systems were solvated in 0.15 M NaCl solution at temperature 310 K. The AMBER ff14SB and AMBER LIPID 21 parameter sets were used for the receptor and lipids53,54. The GAFF2 parameters with RESP charges were used for epinephrine and A6160355. For each of the complex systems, initial energy minimization, thermalization, and 20 ns cMD equilibration were performed before GaMD simulations. A cutoff distance of 9 Å was used for the van der Waals and short-range electrostatic interactions and the long-range electrostatic interactions were computed with the particle-mesh Ewald summation method56. A 2-fs integration time step was used for all MD simulations. The SHAKE algorithm was applied to all hydrogen-containing bonds. With all other atoms fixed, the lipid tails were energy minimized for 5000 steps using the conjugate gradient algorithm and melted with a constant number, volume, and temperature (NVT) run for 0.5 ns at 310 K. The six systems were further equilibrated using a constant number, pressure, and temperature (NPT) run at 1 atm and 310 K for 1 ns with 1.0 kcal/(mol Å2) harmonic position restraints applied to the protein and ligand atoms. Then, conventional MD simulations were performed on each system for 20 ns at 1 atm pressure and 310 K with a constant ratio constraint applied on the lipid bilayer in the X-Y plane. The GaMD module implemented in the GPU version of AMBER22 was then applied to perform the simulations24,57. GaMD simulations included an 8-ns short cMD simulation used to collect the potential statistics for calculating GaMD acceleration parameters, and a 48-ns equilibration after adding the boost potential. Finally, three independent 500-ns GaMD simulations with randomized initial atomic velocities were performed for the epinephrine–α1A-AR–Gq, A61603–α1A-AR–Gq, and A61603-bound mutant α1A-AR–Gq complexes. The average and SD of the system potential energies were calculated every 800,000 steps (1.6 ns). All GaMD simulations were run at the “dual-boost” level by setting the reference energy to the lower bound. One boost potential was applied to the dihedral energetic term and the other to the total potential energetic term. The upper limit of the boost potential SD, σ0 was set to 6.0 kcal/mol for both the dihedral and the total potential energetic terms.

CPPTRAJ was used to analyze the GaMD simulations58. Root-mean-square fluctuations (RMSFs) were calculated for the agonist A61603, averaged over three independent GaMD simulations. Interactions between epinephrine and the receptor were measured by the distance between the CG atom of D106 and N1 atom of epinephrine, and the OG atom of S188 and O2 atom of epinephrine. Interactions between A61603 and the receptor were measured by the distance between the CG atom of D106 and the N2 atom of A61603, the N atom of A189 and the O2 atom of A61603, the OG atom S188 and the N atom of A61603, the OG atom of S188 and the O atom of A61603, and the OG atom of S188 and the O2 atom of A61603. Time courses of these reaction coordinates obtained from the GaMD simulation were plotted in Fig. 3, Supplementary Figs. 5, 7, 9, and 10.

Calcium imaging

HEK293T cells were purchased from ATCC (CRL-11268; CRL-1573), authenticated by Bio-Synthesis, Inc. and routinely tested negative for mycoplasma. Cells were maintained in DMEM supplemented with 10% fetal bovine serum (FBS) in 5% CO2 at 37oC in humidified incubator. HEK293T cells were seeded on poly-L-lysine coated glass coverslips (18 mm) one day before transfection. GCaMP 6 f (0.2 μg per well) and α1A-AR plasmid (0.5 µg per well) were used to transfect cells using Lipofectamine 2000 in 12-well plates. Cells were protected with 10 μM Prazosin 6–8 hours after transfection. Calcium imaging experiments were performed 24 hours after transfection. Cells were placed on an inverted Nikon Eclipse Ti2-E microscope equipped with an Andor Zyla 5.5 sCMOS camera, Lumencor SOLA-SE II light engine, and CFI Nikon Plan Apo Lambda ×20 with 0.75 NA. GCaMP6f was excited using 488 nm SOLA LED light. Cells were continuously perfused with extracellular (EX) solution (containing (in mM): 135 NaCl, 5.4 KCl, 10 HEPES, 2 CaCl2, 1 MgCl2, pH = 7.4) utilizing a gravity-driven perfusion system at room temperature. Drugs (A61603 or epinephrine) were dissolved in EX solution. Calcium imaging was analyzed using ImageJ 1.52a and NIS-Elements Advance Research 5.2.6. Each single cell was manually selected as region of interest and the fluorescence intensity was normalized to the baseline before agonist application. An averaged trace (60–200 cells per recording) was used to quantify maximum fluorescence amplitude after drug application. Dose–response curve was done using nonlinear fit in GraphPad Prism.

Quantification and statistical analysis

In Fig. 5, the Ca2+ assays were repeated three times, and the data are represented as mean ± SD of the three independent experiments. The analysis was done using the log(agonist) vs. response function of Prism 9 (GraphPad) as indicated in the figure legends. Cryo-EM data collection and refinement statistics are listed in Supplementary Table 1.

Reporting summary

Further information on research design is available in the Nature Portfolio Reporting Summary linked to this article.