Introduction

Adrenergic receptors (ARs), which mediate physiological responses to the neurotransmitter noradrenaline (norepinephrine) and the hormone adrenaline (epinephrine), are family A G protein-coupled receptors (GPCRs). They mediate responses to sympathetic nervous system activation and are subdivided into α11A, α1B and α1D), α22A, α2B and α2c) and β (β1, β2, and β3) ARs. α1ARs are predominantly coupled to the heterotrimeric Gq/11 family of G proteins, leading to the activation of phospholipase C and the increase of cytosolic Ca2+. α1AAR was cloned from the bovine brain and initially designated α1CAR1. As the original α1AAR and α1DAR appeared to represent the same subtypes, these clones have been renamed α1AAR (formerly α1CAR), α1BAR (formerly α1BAR) and α1DAR (formerly α1AAR and α1DAR)2,3. α1ARs are expressed in a wide range of tissues including blood vessels, kidney, spleen, liver, brain and lower urinary tract2,3. In the periphery, postsynaptic α1AR activation mediates smooth muscle contraction, therefore the selective α1AAR agonist oxymetazoline4 is clinically used for the treatment of nasal congestion, whereas selective α1AAR antagonists such as tamsulosin and silodosin are prescribed to treat hypertension and benign prostatic hyperplasia5 (Supplementary Fig. 1). As α1AR plays a role in regulating synaptic plasticity and memory consolidation2, the α1AR antagonist prazosin is used to reduce nightmares and overall Post Traumatic Stress Disorder symptoms6 and has a potential for the preventing cytokine storm syndrome caused by the severe acute respiratory syndrome coronavirus 2, a leading cause of morbidity and mortality in coronavirus disease 20197.

Recent progress in the structural characterization of GPCRs including the adrenergic receptors has clarified the mechanism of ligand recognition and G protein activation. The βARs are extensively well-characterized GPCRs and a number of structures have been determined in the active and inactive states8,9,10,11,12,13. Moreover, recent structures of active and inactive α2AAR14,15, active α2BAR-Gi/o16 and inactive α2CAR17 demonstrated the subtype selectivity of ligand recognition between α2ARs and βARs. However, little is known about the structure and mechanism of activation of the α1AR subtypes. The only available structure is the inactive α1BAR structure18. Here we present three cryo-electron microscopy (cryo-EM) structures of active α1AAR bound to oxymetazoline and the endogenous agonist noradrenaline, along with inactive α1AAR bound to tamsulosin. We also discovered a nanobody (a single domain antibody) Nb29 that binds to the extracellular vestibule of the agonist-binding pocket. Antibodies against GPCRs have attracted particular interest for pharmaceutical applications19, and are useful research tools for stabilizing GPCR conformations for structural analysis20,21. Nevertheless, only a few class A GPCR structures in complex with the extracellular antibody are available22,23,24,25,26. These findings may guide the development of more effective drugs for the α1AAR.

Results

Structure determination of active and inactive α1AAR

We expressed human α1AAR in baculovirus-infected Spodoptera frugiperda (Sf9) insect cells. We constructed a variant of human α1AAR lacking residues 371–466 in the C-terminus, and three N-linked glycosylation sites (N7Q, N13Q and N22Q) in the N-terminus were mutated. To further stabilize the receptor, we discovered a conformationally selective nanobody from a library of synthetic nanobodies displayed on the surface of Saccharomyces cerevisiae27. After two rounds of magnetic-activated cell sorting (MACS) and four rounds of fluorescence-activated cell sorting (FACS) with oxymetazoline- and tamsulosin-bound α1AAR, we identified Nb29 as the most enriched clone (Fig. 1a and Supplementary Fig. 2; See Method). An on-yeast titration assay indicated that Nb29 has selectivity for oxymetazoline-bound α1AAR compared with apo, noradrenaline-, and antagonist (tamsulosin and phentolamine)-bound states (Fig. 1b and Supplementary Table 1a). In a ligand binding assay, Nb29 induced the left shift of the agonist competition curves for α1AAR over α1B- and α1DAR (Fig. 1c and Supplementary Table 1b). Although Nb29 on its own competed for the antagonist [3H]prazosin binding for α1AAR and might affect the competition binding results, we observed a larger left shift of oxymetazoline competition curves than those for noradrenaline, which is in agreement with the on-yeast titration result (Fig. 1d, e and Supplementary Table 1c).

Fig. 1: Selection and characterization of Nb29 as a conformationally selective nanobody for α1AAR.
figure 1

a Flow chart of the selection process of conformationally selective nanobodies from the yeast-displayed nanobody library. For rounds 1 and 2, 0.2 μM α1AAR bound to oxymetazoline (oxy) was used for selection and alexa647-labeled anti-FLAG M1 antibody (M1-647) or biotin-labeled anti-FLAG M1 antibody Fab fragment (M1-biotin) was used for the preclear. For the FACS selection, different combinations of counterselection were performed using oxymetazoline, tamsulosin (Tam), M1-488/647, and anti-HA antibodies. b On-yeast titration to estimate the affinity of Nb29 for α1AAR, evaluated by flow cytometry. The ratio of Nb29-displayed yeast cells bound purified α1AAR in the presence or absence of 500 μM ligands was analyzed. The data represent mean ± s.e.m. of n = 3 independent measurements. c 3H-prazosin radioligand competition binding of α1AR subtype for oxymetazoline in Sf9 membranes. Samples in the presence of Nb29 were used at 5 μM concentration of Nb29. The data represent mean ± s.e.m. of n = 3 independent measurements. d, e 3H-prazosin radioligand competition binding of the purified α1AAR-bound M1-Flag affinity resin for oxymetazoline (Oxy), noradrenaline (NA) or Nb29. The data represent mean ± s.e.m. of n = 3 (NA + 5 μM Nb29), and n = 6 (the others) independent measurements. Binding affinity values are provided in Supplementary Table 1. Source data are provided in the Source Data file.

To solve the structure, we formed the α1AAR complex with Nb29 and obtained the initial cryo-EM structure at 4 Å resolution, and found that Nb29 binds to the extracellular region of the α1AAR. To further stabilize the receptor, we used α1AAR construct fused with a minimal-T4 lysozyme (mT4L) in the intracellular loop 3 (ICL3)28 (Supplementary Fig. 3a), and formed a complex with engineered minimal Gsq protein (miniGsq) in which mini-Gs was substituted with the 15 residues of carboxyl-terminal α5 helix of Gq protein29 (Methods and Supplementary Fig. 3b, c). The α1AAR in complex with the heterotrimeric Gq/11 proteins was not stable enough for structure determination. Finally, we obtained the cryo-EM structures of active α1AAR bound with oxymetazoline (Nb29-α1AAR-miniGsq) at a global resolution of 2.9 Å (Fig. 2a, b; Supplementary Figs. 3d–f and 4a; Supplementary Table 2). The cryo-EM map allowed the model building of most of the regions with a clear electron density for the ligand. The map density of the extracellular region of α1AAR is relatively clear because of the bound Nb29, whereas there is poor density for the fused mT4L due to map refinement by masking out the mT4L. Subsequently, we also solved the Nb29-α1AAR-miniGsq complex bound with noradrenaline at a global resolution of 3.5 Å (Fig. 2c, d; Supplementary Figs. 3g–i and 4b; Supplementary Table 2). Although the map quality allowed the model building of the receptor with clear electron density for the ligand, the map density for Nb29 is weak. The weaker map density is in agreement with the observation that Nb29 preferentially binds to oxymetazoline-bound receptors over noradrenaline-bound receptors (Fig. 1b). The resolution of the Nb29-dissociated α1AAR-miniGsq complex is much lower (~6 Å) than that of the Nb29-bound complex (Supplementary Fig. 3h).

Fig. 2: Cryo-EM structures of oxymetazoline- and noradrenaline-bound Nb29–α1AAR-miniGsq and tamsulosin-bound α1AAR-Nb6 complexes.
figure 2

The cryo-EM density maps and structure models of the Nb29-α1AAR-miniGsq complexes bound to the agonists oxymetazoline (a, b) and noradrenaline (c, d), and α1AAR-Nb6 complex bound to the antagonist tamsulosin (e, f). The detergent micelle (a, c, e) and unmodelled mT4L (b) are shown in gray. The densities of the ligands (shown as sticks) are depicted as surfaces. Color code for the proteins is as follows: oxymetazoline-bound active α1AAR (blue), noradrenaline-bound active α1AAR (green), inactive α1AAR (orange), miniGsq (pink), Nb29 (yellow), and Nb6 (purple). Small molecules are colored as follows: oxymetazoline in magenta, noradrenaline in gray, and tamsulosin in cyan.

To solve the inactive α1AAR structure, our crystallographic and cryo-EM experiments using fusion protein and/or the other nanobodies selected from the synthetic nanobody library were unsuccessful. Thus, we utilized a recently described engineering strategy to enable the binding of nanobody 6 (Nb6) that engages the ICL3 of the inactive-state κ-opioid receptor (kOR)21. Based on the cryo-EM structure of engineered neurotensin receptor 1 (NTSR1)-Nb6 complex, we swapped the same site from C2055.59 of transmembrane (TM) 5 (T2475.59 of kOR) to G2756.38 of TM6 (L2776.38 of kOR) including intracellular loop (ICL) 3 (superscripts indicate Ballesteros-Weinstein numbering for GPCRs30). In addition, we introduced two thermostabilizing point mutations, S113R3.39 to mimic allosteric sodium ion binding31, and M115W3.41 to increase expression32. These mutations were used for solving other inactive GPCR structures such as prostaglandin E receptor EP4 for the R3.39 mutation22, and dopamine D2 receptor for double mutation of positions 3.39 and 3.4125. In radioligand binding studies, the α1AAR-kOR mutant showed enhancement of [3H]prazosin binding by Nb6, but did not significantly alter the binding affinities for the agonist oxymetazoline or the antagonist tamsulosin (Supplementary Fig. 5a, b). We formed the α1AAR-Nb6 complex and solved the cryo-EM structure of the inactive α1AAR bound to tamsulosin at a global resolution of 3.3 Å resolution. (Fig. 2e, f; Supplementary Fig. 5c–h; Supplementary Table 2). The cryo-EM map allowed the model building of most of the regions with a clear electron density for tamsulosin; the map density of the inactive α1AAR-Nb6 complex is relatively weak for ECL2 and well-defined in the cytoplasmic region including Nb6 binding domain where ICL3 and the cytoplasmic sides of TM5 and 6 were exchanged to those of the kOR. Although Nb6 binds to a similar site as in the kOR-Nb6 complex33 and locks the conformation of the TM5 and TM6 of α1AAR, we observed fewer polar interactions in our structure when compared with the kOR-Nb6 and NTSR1-Nb6 complexes21 (Supplementary Fig. 6a–e). The active α1AAR structures enable us to model a putative cholesteryl hemisuccinate (CHS) molecule bordering TM3-5 in the active structures, in contrast, we observed only a weak density in the inactive α1AAR structure, since the side chain of M115W3.41 overlaps the regions corresponding to the lipid tail of CHS (Supplementary Fig. 6g, h). The side chain of S113R3.39 is located in the putative sodium ion binding site as designed (Supplementary Fig. 6i, j)22,31.

Structural comparison of active and inactive states of α1AAR exhibits 14.5 Å outward displacement of an intracellular segment of TM6 that is characteristic of receptor activation (Fig. 3a–c). The TM6 movement is accompanied by a small rotation of the helix, as well as inward movements of TMs 3, 5, and 7 toward TM6. α1AAR also exhibits other characteristics of the activation of class A GPCRs9,14,16. We observed the displacement of the side chain of W2856.48 (Fig. 3d), a highly conserved residue that contributes to conformational changes associated with activation for some GPCRs. We also observe conformational changes in the conserved PIF (P1965.50 I1143.40 and F2606.44) interaction, as well as the NPxxY (N3227.49, P3237.50, and Y3267.53) and DRY (D1233.49, R1243.50 and Y1253.51) motifs (Fig. 3d, e). In the active α1AAR, R1243.50 forms hydrogen bond networks with Y2045.58, Y3267.53 and C3297.56 (Fig. 3e). These structural changes allow the C-terminal helix (α5 helix) of Gα to engage the receptor core, as described below. In the extracellular view, due to the Nb29 binding, the conformation of ECL2, TMs 4 and 7 in the active state is closer to the receptor core, as discussed later in detail.

Fig. 3: Structural comparison between active and inactive α1AAR.
figure 3

a Comparison of α1AAR between oxymetazoline-bound (blue), noradrenaline-bound (green), and inactive (orange) states viewed from the side. b Intracellular view of superposed α1AARs. Distances were measured between the Cα atoms of E2696.30 (E269L6.30 in α1AAR-kOR) in TM6, K2125.66 in TM5, and C3287.55 in TM7. c Extracellular view of superposed α1AARs. The maximum distances of the side chain displacement of W285 (position 7th carbon of the indole ring) are 2.4 Å between the noradrenaline-bound active state and the tamsulosin-bound inactive state, and 1.8 Å between the oxymetazoline-bound state and the tamsulosin-bound state. d Conformational change of PIF motif and toggle switch W2856.48. e Conformational change of DRY and NpxxY motifs. Conformational changes upon activation are shown with magenta arrows. Hydrogen bonds are shown as black dashed lines.

Orthosteric ligand-binding pocket of α1AAR

Extensive site-directed mutagenesis studies have identified amino acids that form the binding pocket of the α1AAR, including residues responsible for subtype selectivity34,35,36,37,38,39,40,41,42. Our structures largely confirm these observations. α1AAR structures bound to the endogenous agonist noradrenaline, selective partial agonist oxymetazoline, and selective antagonist tamsulosin are shown in Fig. 4. All three ligands form polar interactions with D1063.32 which is a highly conserved residue involved in ligand binding in all aminergic receptors (Fig. 4 and Supplementary Fig. 7). The binding pocket of noradrenaline is formed by residues in TMs 3, 5, 6, and 7 (Fig. 4a, e). The noradrenaline has two catechol hydroxyl groups. The para-hydroxyl forms a hydrogen bond with S1885.42, whereas meta-hydroxyl does not form polar interaction but is close to M2926.55, a unique residue among ARs (Supplementary Fig. 7). Previous mutagenesis34,35,36 and [13CεH]methionine labeling NMR studies37 support the role of S1885.42 and M2926.55 in ligand binding. The chiral β-hydroxyl forms a hydrogen bond with D1063.32 and the amino group of the noradrenaline forms cation-π stacking with the phenyl ring of F3127.39 and a hydrogen bond with the backbone carbonyl of F3127.39. Noradrenaline forms extensive non-polar interactions with highly conserved aromatic residues among ARs, including Y1845.38, F2886.51, and F2896.52.

Fig. 4: Orthosteric ligand-binding pocket of α1AAR.
figure 4

Side view of α1AAR ligand-binding site of noradrenaline (a), oxymetazoline (b), and tamsulosin (c). Residues within 4.1 Å distance of the ligands are shown in stick representation. Hydrogen bonds are shown as yellow dashed lines. d Comparison of residues involved in ligand-binding pockets. e–g Diagram of ligand interactions between α1AAR and the ligands. Residues within 4.1 Å distance of the ligands are shown.

The oxymetazoline binds in a similar site (Fig. 4b, f), but the para-catechol hydroxyl is replaced by tertiary butyl, leading to the lack of polar interaction with S1885.42, which may account for its partial agonism in receptor activation12,14,15. Tertiary-butyl group forms extensive van der Waals interaction with Y1845.38, V1855.39, S1885.42, M2926.55 and the backbone carbonyl of N179ECL2. Position 5.39 is a valine in α1AAR, α2AAR, and β3AR, but is an alanine/isoleucine in other subtypes of ARs (Supplementary Fig. 7). A previous mutagenesis study showed that V185A5.39 and M292L6.55 mutations resulted in decreased oxymetazoline binding but not noradrenaline binding, and the equivalent A204V5.39 and L312M6.55 mutants of α1BAR increased agonist binding35,36. In place of the chiral β-hydroxyl and the amino group of noradrenaline, oxymetazoline has an imidazoline ring which forms polar interactions with D1063.32 and C1103.36, π-π stacking with F3127.39, and aromatic interactions with W2856.48, F2886.51 and Y3167.42. In α1- and α2ARs, C1103.36 and F3127.39 are conserved residues and involved in ligand recognition for imidazoline-type agonists14,15,16. Consistent with this result, a mutagenesis study indicated that F312A7.39 and F312N7.39 (the equivalent residues for βARs) mutations decreased oxymetazoline binding but not adrenaline binding38. This report also demonstrated that F3087.35, a residue above the F3127.39, influences oxymetazoline binding, even though it is too far away for a direct interaction38.

The antagonist tamsulosin has two aromatic groups on each side of an ethyl-aminopropyl backbone (Fig. 4c, g). Similar to agonist binding, the ethylamine group of tamsulosin forms a hydrogen bond with D1063.32. The ethoxyphenoxy group forms van der Waals interaction with V1073.33, S1885.42, M2926.55, F2886.51 and F2896.52 (Supplementary Fig. 7). Unlike agonist binding, reorientation of F3127.39 enlarges the binding pocket and enables the antagonist to bind towards the extracellular vestibule, which is also called an exosite or secondary binding pocket with less conserved residues compared to the orthosteric pockets12,43 (Fig. 4d). The sulfonamide group forms a polar interaction with the backbone carbonyl of C17645.50 and a non-polar interaction with F3087.35. The methoxybenzene group forms non-polar interactions with F862.64, E872.65, W1023.28, F3127.39 and the backbone carbonyl of S832.61 (Fig. 4c and Supplementary Fig. 7). F862.64 is a unique residue to α1AAR relative to other ARs and was previously identified as a determinant for the interaction of the α1AAR with various antagonists including HEAT and prazosin39,40,41 (Supplementary Fig. 1). In addition, another mutation study indicated that three non-conserved residues (Q17745.51, I17845.52, N17945.53) in ECL2 are responsible for the α1AAR selectivity of phentolamine and WB4101 over α1BAR42, but are not involved in binding tamsulosin.

Ligand recognition of adrenergic receptor subtypes

Although all adrenergic receptors are activated by endogenous adrenaline and noradrenaline, their binding pockets are not identical. Comparisons of the key residues of the noradrenaline binding pockets in α1AAR, α2AAR14 and β1AR11 reveal similar but different mechanisms of noradrenaline recognition (Fig. 5a, b). Compared to M6.55 in α1AAR, Y6.55 in α2AAR (conserved in all α2ARs) forms a hydrogen bond with the meta-hydroxyl of the catechol ring, while the para-hydroxyl is involved in hydrogen bonds with S5.42 in both α1AAR and α2AAR. The rotamer of S5.42 differs between α1AAR and α2AAR, leading a distinct difference in the pose of the catechols. In α1AAR, the β-hydroxyl group interacts with D3.32, and the amino group forms a hydrogen bond and a cation-π interaction with F7.39(Figs. 4a and 5a). In contrast, only the amino group forms hydrogen bonds with both D3.32 and Y7.43 in α2AAR. The noradrenaline binding pose of β1AR is different from that of α1AAR (Fig. 5b). The meta-hydroxyl forms polar interaction networks with S5.42, S5.43 and N6.55, and the para-hydroxyl forms a hydrogen bond with S5.46 in β1AR. Previous α1AAR mutagenesis studies indicated that double mutation of S188A5.42 and S192A5.46 decreased agonist binding rather than S188A5.42 or S192A5.46 alone34. Non-aromatic N7.39 interacts with the amine group of noradrenaline through the polar interaction networks with D3.32 and Y7.43 in β1AR. Moreover, the bulkier F45.52 in ECL2 of β1AR (conserved in all βARs) forms a non-polar interaction with noradrenaline10,11.

Fig. 5: Comparisons of ligand-binding pockets for the adrenergic receptors.
figure 5

a Noradrenaline binding between α1AAR (green sticks with green hydrogen bonds) and α2AAR (yellow sticks with yellow hydrogen bonds; PDB ID: 7EJ0). b Noradrenaline binding between α1AAR (green sticks) and β1AR (light pink with yellow hydrogen bonds; PDB ID: 7BU6). c Oxymetazoline binding between α1AAR (blue sticks with black hydrogen bonds) and α2AAR (wheat sticks with yellow hydrogen bonds; PDB ID: 7EJK). Aromatic interaction is shown as an orange dashed line. d Orthosteric and extracellular vestibule of α1AAR bound to noradrenaline (gray) and tamsulosin (cyan). e Comparison of antagonists binding to α1AAR (orange sticks with yellow hydrogen bonds, tamsulosin is colored cyan), and α1BAR (gray sticks with blue hydrogen bonds and pink ligand; PDB ID: 7B6W). Note that F7.39 of the α1BAR is mutated to L7.39 for stabilization. f Comparison of antagonists binding between α2AAR (light orange sticks with yellow hydrogen bonds; ligand is a colored gray stick; PDB ID: 6KUX) and β2AR (cyan sticks with cyan hydrogen bonds and green colored ligand; PDB ID: 2RH1).

We next compare the binding pose of imidazoline-type partial agonist oxymetazoline in α1AAR and α2AAR14 (Fig. 5c). The oxymetazoline presents high selectivity for α1AAR and α2AAR over other αARs4. As mentioned before, oxymetazoline replaces the para-hydroxyl with the hydrophobic tertial-butyl group which no longer forms the polar interaction but engages in hydrophobic interactions with partially conserved V5.39 in α1AAR. In α2AAR, oxymetazoline also does not form polar interaction with TM5, but C5.43 is involved in the hydrophobic interaction. This position is A1895.43 in α1AAR, while S5.43or C5.43 in other ARs (Supplementary Fig. 7). In both α1AAR and α2AAR, the imidazoline ring is stabilized by π-π stacking with F7.39 along with polar interaction with D3.32, in contrast, C3.36 (conserved in all αARs and V in βARs) forms weak polar interaction with the imidazoline ring only in α1AAR. Although most of the residues have the same orientation between noradrenaline and oxymetazoline binding in α1AAR, the orientation of Y6.55 is shifted in α2AAR. This Y6.55 is involved in Gi/o-biased signaling over β-arrestin recruitment for oxymetazoline among the other agonists such as noradrenaline, brimonidine and dexmedetomidine in α2AAR14.

Compared to α1AR agonists, the α1AR antagonists have higher subtype-selectivity because they extend to the extracellular vestibule (Fig. 5d). As mentioned above, inactive α1AAR bound to tamsulosin reveals that unique (F862.64, and M2926.55) and partially conserved (S832.61, E872.65, W1023.28, I17845.52, and F3127.39) residues are involved in subtype selectivity (Fig. 3c, f and Supplementary Fig. 7). Recent crystal structure of inactive α1BAR bound to its selective inverse agonist (+)-cyclazosin, along with the chimeric α1BAR-α2CAR mutagenesis studies indicated that non-conserved residues L2.64, W3.28, A3.29, V45.52, and L6.55 are important for the selectivity in α1BAR18 (Supplementary Fig. 7). The (+)-cyclazosin is a derivative of prazosin in which piperazinyl quinazoline scaffold is introduced in a bulky cycloaliphatic group (Supplementary Fig. 1), leading to 100–1000 fold selectivity for α1ARs over α2ARs, and a slight preference for α1BAR over α1AAR. When comparing the α1AAR and α1BAR (Fig. 5e), both the tamsulosin and (+)-cyclazosin extend into the extracellular vestibule. Tamsulosin interacts with F862.64 more closely than the furan group of (+)-cyclazosin, which is consistent with the binding selectivity44, however, a stabilizing mutation of F7.39 to L7.39 in the α1BAR structure might affect the (+)cyclazosin binding mode18. In contrast to α1ARs, the antagonists of α2AAR and β2AR do not interact with TM2 (Fig. 5e, f)8,15. The positions 2.64, 3.28, 3.29, and 45.52 are different from α1ARs and likely involved in the ligand selectivity. In addition, residues M6.55 in α1AAR and N6.55 in β2AR allow antagonist interactions with the extracellular side of TM6, in contrast to the bulkier Y6.55 in α2AAR.

It is known that the other α1AAR ligands such as A61603 (agonist) and silodosin (antagonist) have high selectivity for α1AAR (Supplementary Fig. 1)37,44. In these compounds, the phenyl rings corresponding to catechol have much bulkier substituents, suggesting that they may exhibit selectivity through interaction with α1AAR unique residues such as M2926.55, A1895.43 and the non-conserved residue V1855.39.

Structural insight into Nb29 binding

Nb29 binds to the extracellular side of α1AAR which is topologically distinct from the orthosteric agonist pocket (Fig. 2). This site has been shown to bind to allosteric modulators for muscarinic receptors45,46,47,48. We do observe a left shift of the agonist competition binding curves in the presence of Nb29 (Fig. 1c–e); however, these experiments are complicated by the fact that Nb29 is a competitive inhibitor of the radioligand [3H] prazosin. In cell signaling assays, Nb29 exhibits no agonist activity on its own, has no effect on EC50 for oxymetazoline or noradrenaline, and slightly reduces the maximum efficacy of α1AAR activation (Supplementary Fig. 8a–d), suggesting that Nb29 appears to antagonize receptor activation or possibly block the ligand entry into the orthosteric pocket. It should be noted that the radioligand competition assay was performed in equilibrium and the agonists had a longer incubation time to access the orthosteric pocket than in the signaling assay. In both assays, the effects of Nb29 are larger for oxymetazoline compared with noradrenaline, which is consistent with Nb29’s binding selectivity towards the oxymetazoline-bound state of the α1AAR in the titration assay (Fig. 1b–e and Supplementary Fig. 8). Thus, Nb29 might be considered a weak positive allosteric modulator (PAM) or a neutral allosteric modulator, given that it binds to a known allosteric binding pocket in other GPCRs.

The Nb29 binding interactions are almost identical in oxymetazoline- and noradrenaline-bound Nb29-α1AAR complexes, but the map resolution is relatively poor in the noradrenaline-bound state (Fig. 1 and Supplementary Figs. 3 and 4). Thus, we used the oxymetazoline-bound state for structural analysis. Nanobodies consist of three complementarity-determining regions (CDRs). The relatively long CDR3 interacts with a broad range of residues from ECL2 and with E3057.32 at the top of TM7 (Supplementary Figs. 9a–c and 10a, b). Among them, seven amino acids (R166, Q167, E171, T174, Q177, N179, and E3057.32) are non-conserved residues in α1AR subtypes, suggesting that these residues are involved in nanobody specificity (Supplementary Figs. 7, 9a–c). Nb29 binding also stabilized the polar interaction network between ECL2 and R963.22, which is not observed in the inactive α1AAR structure without Nb29 (Supplementary Fig. 9d–f). Four residues of CDR3 (Y100, R101, D102 and H103) bind to the extracellular vestibule from the agonist-binding pocket (Fig. 6a). R101 of Nb29 forms charge networks with E180ECL2 and E3057.32, and a cation-π interaction with F3087.35 of α1AAR, stabilizing the inward conformation of TM7 (Fig. 6a, b). As mentioned above, F3087.35 and the close-lid conformation of F7.39 is important for the agonist binding in αARs (Fig. 4)38, the π-π stacking of oxymetazoline with F3127.39 might contribute to the Nb29 binding selectivity compared with the cation-π stacking of noradrenaline with F3127.39.

Fig. 6: Structural comparison of the Nb29 binding site.
figure 6

Nb29 binding sites of (a) Nb29-α1AAR-oxymetazoline; b Nb29-α1AAR-oxymetazoline and inactive α1AAR; c LY2119620-M2R-iperoxo (PDB ID: 4MQT), M2R-iperoxo (PDB ID: 4MQS) and inactive M2R (PDB ID: 3UON); d MT7-M1R-atropine (PDB ID: 6WJC) and inactive M1R (PDB ID: 5CXV); e β2AR-salmeterol (PDB ID: 6MXT) and inactive β2AR (PDB ID: 2RH1); f Nb29-α1AAR-oxymetazoline and β2AR-salmeterol (PDB ID: 6MXT). Polar and aromatic interactions are shown as cyan and orange dashed lines, respectively. Conformational changes are shown with cyan arrows. g–j Surface representations of top views of Nb29-α1AAR-oxymetazoline, inactive α1AAR-tamsulosin, LY2119620-M2R-iperoxo (PDB ID: 4MQT) and β2AR-salmeterol (PDB ID: 6MXT).

The residues at position 7.35 have also been identified as critical residues for both PAM and negative allosteric modulator (NAM) binding of muscarinic acetylcholine receptors (MRs) by stabilizing the extracellular side of TM7 either in inward or outward conformations, respectively45,46. The LY2119620, a PAM for M2R, binds to the extracellular vestibule and changes the conformation of W4227.35 by an aromatic stacking, whereas those of the other residues are almost identical between the LY2119620-M2R-iperoxo and M2R-iperoxo complexes (Fig. 6c and Supplementary Fig. 10c, d)45. In contrast to this PAM, a peptide toxin MT7, the NAM for M1R activation, stabilizes the outward displacement of TMs 6 and 7 through interactions with W4007.3546 (Fig. 6d and Supplementary Fig. 10e, f). Moreover, a mutagenesis study indicated that F3307.35 in α1BAR is involved in the allosteric binding for conotoxin ρ-TIA, a selective NAM for α1BAR among α1ARs49. The Nb29 binding site also overlaps with the aryloxyalkyl tail of the selective agonist salmeterol binding in β2AR (Fig. 6e, f)12. The smaller N7.39 in β2AR enables the salmeterol to extend into the extracellular vestibule to make aromatic interactions with F194ECL2, H2966.58 and Y3087.35, which are unique to β2AR (Supplementary Fig. 7). Among aminergic receptors, the aromatic amino acid at position 7.39 is observed in αARs (F7.39), muscarinic receptors (Y7.39) and histamine H3 and H4 receptors (F7.39)12,43,45,50. The selective agonists targeting the extracellular vestibule of αARs are limited compared to other GPCRs such as βARs and muscarinic receptors12,13,43,50 (Fig. 6g–j and Supplementary Fig. 1).

Subsequently, we compared the nanobody binding site with available class A GPCR structures in complex with extracellular nanobodies or antibody Fab fragments (Supplementary Fig. 10). Consistent with the review for GPCR antibodies19, peptide-binding GPCRs are more frequently targeted by antibodies because they have relatively large binding pockets compared with the small-molecule binding GPCRs such as aminergic GPCRs. Only two structures in complex with extracellular nanobodies have been reported in class A GPCRs, which are the apelin receptor (APJ)24 and the orexin receptor 2 (OX2R)51, although more structures of the intracellular binding nanobodies have been published for stabilizing the GPCR active conformations as G protein mimetics, such as Nb9-8 for M2R10,20,45 (Supplementary Fig. 10c–j). Anti-APJ nanobody JN241 antagonizes APJ through extensive interactions with extracellular loops of APJ and the insertion of CDR3 into the peptide-binding site24, whereas anti-OX2R nanobody Sb51 is positioned above the small-molecule agonist and partially overlaps with natural-peptide orexin B binding site51. In contrast to nanobodies, conventional antibodies and Fab fragments consist of heavy (CDRs H1-3) and light (CDRs L1-3) chains (Supplementary Fig. 10k–r). The antibody for protease-activated receptor 2 (PAR2) behaved as an antagonist by blocking ligand access from the extracellular region by both heavy and light chains (H1, H3, L2, and L3)26. In the angiotensin II type 2 receptor (AT2)23, EP422, and D2R structures25, their antibodies allosterically enhance the ligand binding. The antibody for AT2 bound to the ECL1 and β-hairpin motif of ECL2 to stabilize the peptide agonist binding pocket, while the antibody for EP4 stabilizes the occluded β-hairpin of ECL2, leading to enhanced antagonist binding. In the D2R-Fab3089 structure, the CDR-H2 stabilizes the antagonist spiperone which binds toward the extracellular vestibule.

Taken together, the Nb29 structure provides insights into allosteric binding and antibody recognition for α1AAR. Nb29 covers the extracellular surface of α1AAR like anti-APJ nanobody JN241 and anti-PAR2 Fab, whereas the CDR3 loop of Nb29 binds to a similar site of PAM for M2R (Fig. 6 and Supplementary Fig. 10). It should be noted that nanobodies are amenable to optimization due to the single variable domain. For example, the G protein-mimicking nanobody for β2AR was optimized by directed evolution to increase its affinity to the receptor10; and the APJ nanobody antagonist JN241was rationally engineered into an APJ agonist by structure-guided site-directed mutation of CDR324.

Selectivity of G protein interactions with adrenergic receptors

Noradrenaline- and oxymetazoline-bound α1AAR-miniGsq complexes are almost identical at the interfaces with miniGsq, as observed in α2AAR-Go complexes with different agonists14. Thus, we used the oxymetazoline-bound active state for structural analysis. Of the 15 residues at the carboxyl-terminal α5-helix of miniGsq (Fig. 7a)29, seven residues are specific to Gq, including KH5.12, LH5.16, QH5.17, NH5.19, E H5.22, NH5.24, and VH5.26 (superscript, CGN G protein numbering system52); five residues [DH5.13, IH5.15, LH5.20, YH5.23, and LH5.25] are conserved between Gs and Gq, and the other three residues [IH5.14, MH5.18 and RH5.21] are located on the opposite side of the interface. When comparing the interactions of α1AAR-miniGsq complexes with α2AAR-Go14 and β2AR-Gs complexes9, the α5-helix of miniGsq is slightly shifted towards helix 8 of α1AAR (Fig. 7b–h). In the α1AAR-miniGsq complex, α1AAR forms six hydrogen bonds with the residues corresponding to Gq: between R2135.67 and QH5.17, between the backbone carbonyl of G1273.53 and NH5.19, between T2736.36 and the backbone carbonyl of NH5.24; the side chain of NH5.24 forms hydrogen bond networks with the backbone carbonyl of C3287.55, the backbone carbonyl of S3308.47, and the side chain of Q3318.48 (Fig. 7c, d). These residues of α1AAR are not conserved in α2ARs and βARs (Supplementary Fig. 11). The R3.50 forms cation-π stacking interaction with YH5.23, which is also observed in the β2AR-Gs complex, but not in α2AAR-Go complex as this position is CH5.23 in Go protein (Fig. 7c–h). In the α2AAR-Go complex, polar interactions are observed between S3.53 and NH5.19, and hydrophobic interactions are predominant (Fig. 7e, f). In the β2AR-Gs complex, one side of the α5-helix of the Gs protein forms a cluster of hydrogen-bond interactions with TM3 (I3.54 and T3.55) and TM5 (E5.64, Q5.68, and K5.71), leading the α5-helix to shift towards TM5 (Fig. 7b, g, h). In addition to α5-helix, the N-terminal helix of Gα subunits are also involved in GPCR-G protein interactions9,14. The previous study indicates that polybasic cluster at the C terminus of M1R, which is conserved among most Gq/11-coupling GPCRs, interacts with the G-protein Gα11/β interface53. However, our structure lacks these regions and there are currently no other active structures of α1ARs. While α1AAR is predominantly coupled to Gq/11 proteins, a few studies have identified G12/1354 and β-arrestin signaling pathways2,55. Further studies will be required to better understand the mechanism of activation and selectivity of α1AR signaling.

Fig. 7: Comparison of the receptor-G protein binding interfaces of the adrenergic receptor subtypes.
figure 7

a Sequence alignments of the carboxyl-terminal α5-helix of G proteins subtypes. b Superimposition of the binding interfaces of α1AAR-miniGsq complexes with α2AAR-Go (PDB ID: 7EJ0) and β2AR -Gs (PDB ID: 3SN6) complexes. The receptors are used for alignment. The proteins are colored as follows: α1AAR (cobalt), miniGsq (pink), α2AAR (yellow), Go (cyan), β2AR (green), and Gs (blue purple). Detailed polar interactions and the equivalent residues of α1AAR-miniGsq (c, d), α2AAR-Go (e, f), and β2AR -Gs (g, h). The polar interactions are shown as black dashed lines.

Discussion

Here, we present three cryo-EM structures of α1AAR in both active and inactive states. These structures reveal several structural aspects of α1AAR. The ligand-binding modes of the endogenous agonist noradrenaline and the imidazoline-type agonist oxymetazoline demonstrated a key aromatic interaction involving F3127.39, which is conserved in αARs, and distinct ligand recognition by the unique residue M2926.55. The inactive α1AAR bound to tamsulosin reveals the subtype selectivity of antagonist binding pockets involving F862.64. Our results also provide structural insights into nanobody recognition for α1AAR. Nb29 binds to the extracellular vestibule of α1AAR and the cationic residue of CDR3 stabilizes F3087.35 which is the equivalent binding site of the positive allosteric modulator for M2R. Finally, our active α1AAR structures provide insight into G protein binding selectivity by comparisons with α1AAR-miniGsq, α2AAR-Go, and β2AR-Gs structures. Together with our α1AAR structures and previously published structures of α2ARs and βARs, the active and inactive structures of the major subtypes of the adrenergic receptor family have been determined. These results should facilitate the design of more selective and effective therapeutic drugs targeting both orthosteric and allosteric sites in this receptor family.

Methods

Construction

C-terminus truncated human α1AAR (residues 1–370, full length: 466) was modified by mutation of the N-linked glycosylation sites to glutamine (N7Q, N13Q and N22Q), N-terminal addition of the hemagglutinin signal peptide, FLAG-tag epitope, and C-terminal addition of the 8 × His-tag. For the active α1AAR structure, residues 223–261 of intracellular loop 3 were replaced with minimal T4L28. For the inactive α1AAR structure, we swapped residues from C2055.59 to G2756.38 with residues from T2475.59 to L2776.38 of kOR21. In addition, we introduced two thermostabilizing point mutations S113R3.39 and M115W3.41 which were previously used for structure determination of several inactive-state GPCRs22,25,31. The primers used in this study were obtained from Rui Biotech (Beijing, China), Xianghong Biotech (Beijing, China) or Genewiz (Beijing, China). The DNA sequencing analysis was performed at the Rui Biotech (Beijing, China).

Expression and purification of α1AAR

Recombinant baculovirus was generated using the Bac-to-Bac Baculovirus Expression System (Thermo Fisher Scientific). Sf9 insect cells at a cell density of 4 × 106 cells/ml in ESF-921 insect media (Expressions Systems) with 20 μg/ml gentamycin and 1 μM ligand were infected with baculovirus and shaken at 27 °C for 2 days. Cells were harvested by centrifugation and stored at −80 °C until use.

To purify the protein, receptor-expressing Sf9 cells were lysed by resuspension in a buffer containing 20 mM Tris-HCI pH 7.5, 1 mM EDTA, 10 µM ligand, 160 μg/ml benzamidine, 100 μg/ml leupeptin. The cell membranes were centrifuged at 10,000 g for 20 min at 4 °C. Receptor was extracted from cell membranes with solubilization buffer of 20 mM HEPES pH 7.5, 1% dodecyl maltoside (DDM), 0.03% cholesterol hemisuccinate (CHS), 0.2% Na cholate, 750 mM NaCl, and 30% glycerol. Iodoacetamide (2 mg ml−1) was added to block reactive cysteines at this stage. Nickel-NTA agarose was added to the solubilized receptor without prior centrifugation. After stirring for 2 h at 4 °C, receptor-bound nickel resin was washed and poured into a glass column, and the receptor was eluted in 20 mM HEPES pH 7.5, 0.1% DDM, 0.03% CHS, 0.02% Na cholate, 750 mM NaCl, 10 µM ligand and 250 mM imidazole. Nickel resin-purified receptor was bound to M1-Flag affinity resin with 2 mM CaCl2. Following extensive washing, detergent was gradually exchanged from DDM to 0.01% lauryl maltose neopentyl glycol (MNG). The receptor was eluted with 0.2 mg/ml Flag peptide and 5 mM EDTA and further purified by size exclusion chromatography (SEC) on a Sephadex S200 increase column (Cytiva) in a buffer of 20 mM HEPES pH 7.5, 0.01% MNG, 0.001% CHS, 100 mM NaCl and 10 µM ligand. The purified receptor was concentrated with a 50 kDa cutoff Amicon centrifugal filters (Millipore).

Discovery of the conformationally selective α1AAR nanobody

The synthetic nanobody library displayed on the surface of BJ5465 yeast strain was obtained from Drs. A. C. Kruse (Harvard University) and A. Manglik (University of California San Francisco)27. The yeast cells were recovered in tryptophane dropout (-Trp) medium [prepared by Yeast Synthetic Drop-out Medium Supplements without tryptophane (sigma) and Yeast Nitrogen Base without amino acids (BD Difco) at pH 6.0] with 2% (w/w) glucose at 30 °C, and the nanobody was induced by -Trp medium with 2% (w/w) galactose at 25 °C. Expression levels of nanobody were estimated by staining with anti-HA antibody (Cell Signaling Tech) and analyzing by flow cytometry with an Accuri C6 (BD Biosciences)

Induced yeast cells were washed and resuspended in a selection buffer (20 mM HEPES pH 7.5, 150 mM NaCl, 0.05% MNG, 0.005% CHS, 2.8 mM CaCl2, 0.1% (w/v) bovine serum albumin and 5 mM maltose). Nanobody clones against the purified FLAG-tagged α1AAR (C-terminus truncated after residue 370) bound to oxymetazoline were enriched by two rounds of magnetic-activated cell sorting (MACS) and four rounds of fluorescence-activated cell sorting (FACS) (See Fig. 1a). For the first round of the MACS, 5 × 109 yeast cells were precleared by incubating with Alexa Fluor-647 conjugated anti-FLAG M1 antibody (M1-647, prepared by anti-FLAG M1 antibody and Alexa Fluor 647-NHS ester) and anti-Alexa Fluor-647 microbeads (Miltenyi) and passed LD column (Miltenyi) to remove nonspecific nanobody. Flowed-through yeast cells were washed with the selection buffer, then incubated with 0.2 µM α1AAR bound to oxymetazoline, the antibody and the microbeads. After incubation at 4 °C for 30 min, yeast cells were loaded on the LD column, washed with the selection buffer and the eluted yeast cells (3.4 × 106 cells) by plunger were expanded and used in a subsequent round of MACS. The second round of MACS was performed similarly to the first, but beginning with 4 × 108 yeast cells and using biotin-conjugated anti-FLAG M1 antibody Fab fragment (anti-FLAG M1 antibody was digested by papain then labeled with biotin-NHS ester), streptavidin microbeads (Miltenyi) and LS column (Miltenyi) were used, and 5 × 106 yeast cells were eluted.

Subsequently, we performed four rounds of FACS by FACSAria II (BD Biosciences) (See Supplementary Fig. 2). For the selection rounds 3 and 6, yeast cells were stained with Alexa Fluor-488 or −647 conjugated anti-HA antibody (Cell Signaling Tech) and 0.1 µM FLAG-tagged α1AAR with anti-FLAG M1-647 or −488. For the selection rounds 4 and 5, in order to enrich for conformational selective nanobodies, yeast cells were stained with two different populations of α1AAR labeled with anti-FLAG M1-488 and −647 fluorophores, one bound with oxymetazoline and another bound to tamsulosin. Staining yeast cells for each round of FACS experiments were the following; 5 × 107 cells for round 3 and 1 × 107 cells for rounds 4–6. After round 6, the sorted yeast cells were diluted and plated on -Trp agar plates. Single clones were sequenced and cloned into the periplasmic expression vector pET26b, containing an N-terminal pelB signal sequence and a C-terminal histidine tag, and transformed into BL21(DE3) Escherichia coli. Cells were induced in Terrific Broth medium with 2 mM MgCl2, 0.1% glucose and 50 µg/ml kanamycin at an OD600 of 0.7 with 1 mM IPTG and incubated with shaking at 25 °C for 20 h. Periplasmic protein was obtained by osmotic shock in a buffer containing 0.2 M Tris pH 8.0, 0.5 mM EDTA and 0.5 M sucrose at 4 °C for 1 h, then diluted 4 times and incubated for another one hour. The lysate was centrifuged and the supernatant was purified by Ni-NTA resin and size-exclusion chromatography.

For the on-yeast titration assay, Nb29-displayed yeast cells were stained with the Alexa Fluor-647 conjugated anti-HA antibody and several concentrations of purified α1AAR fused at the C-terminus to an enhanced green fluorescent protein in the presence or absence of 500 μM ligands in the selection buffer. Yeast cells were analyzed by Accuri C6 and the ratio of double-positive yeast cells among anti-HA positive cells was calculated.

Purification of the Nb29-α1AAR-miniGsq and the α1AAR-Nb6 complexes

We modified the expression and purification method of miniGsq from the previous report29,56, in which miniGsq was used instead of miniGq for the expression in Escherichia coli. The pET21a plasmid encoding miniGsq and N-terminal histidine tag was transformed into BL21(DE3) Escherichia coli. Cells were induced in Terrific Broth medium at an OD600 of 0.6 with 1 mM IPTG and incubated with shaking at 25 °C for 20 h. Cells were harvested and lysed by sonication in a buffer containing 40 mM Hepes pH 7.5, 100 mM NaCl, 10 mM imidazole, 10% glycerol, 5 mM MgCl2, 50 μM guanosine diphosphate (GDP), 100 μM dithiothreitol, 160 μg/ml benzamidine, and 100 μg/ml leupeptin. The lysate was centrifuged at 10,000 g for 20 min at 4 °C and the supernatant was immobilized by Ni-NTA resin. The eluate was further purified by size-exclusion chromatography in a buffer containing 10 mM HEPES pH 7.5, 100 mM NaCl, 1 mM MgCl2, 10 μM GDP and 100 μM tris(2-carboxyethyl)phosphine (TCEP).

The Nb29-α1AAR-miniGsq complex was prepared by mixing with the purified α1AAR bound the agonist (oxymetazoline or noradrenaline), Nb29 and miniGsq in a 1:1.2:1.2 molar ratio and supplemented with apyrase. After incubation for 2 h at room temperature, the mixtures were purified by size exclusion in SEC buffer containing 20 mM HEPES pH 7.5, 100 mM NaCl, 0.002% MNG, 0.0002% CHS, 2 mM MgCl2, 10 μM agonist and 100 μM TCEP. Fractions containing the complex were concentrated to 5–10 mg/ml using a 50 kDa molecular weight cutoff Amicon Ultra concentrator. For the α1AAR-Nb6 complexes, Nb6 from gene synthesis33 was prepared as the same as nanobody purification described above. Purified α1AAR-kOR swap construct and Nb6 were mixed in a 1:1.5 molar ratio in the presence of tamsulosin and incubated on ice for 30 min. The mixture was further purified by size exclusion in SEC buffer containing 20 mM HEPES pH 7.5, 100 mM NaCl, 0.002% MNG, 0.0002% CHS, 2 mM MgCl2, and 10 μM ligand. Fractions containing the complex were concentrated using a 50 kDa molecular weight cutoff Amicon Ultra concentrator.

Ligand binding assay

Ligand binding assays were performed with Sf9 cell membrane or purified α1AAR-bound M1-Flag affinity resin. Receptor-expressing cells were harvested and homogenized in a binding buffer containing 20 mM Tris–HCl (pH 7.5) and 100 mM NaCl. After centrifugation, the pellet was homogenized in binding buffer and used as the membrane fraction in binding assays. The purified α1AAR-bound M1-Flag affinity resin was resuspended in binding buffer containing 20 mM HEPES pH 7.5, 100 mM NaCl, 0.01% MNG, 0.001% CHS and 2 mM CaCl2.

Radioligand binding assay was performed using [3H]prazosin (PerkinElmer). Receptors were incubated for 1 h at room temperature with various concentrations of ligand in a total volume of 100 µl. After the reaction, the mixture was trapped on Whatman GF/B glass filters. Bound and free radioligands were separated by washing with ice-cold binding buffer. Radioactivity was measured on a MicroBeta2 liquid scintillation counter (PerkinElmer). All binding assay measurements were analyzed using the Prism 9 software (GraphPad).

Cryo-EM sample preparation and data acquisition

Two of the Nb29–α1AAR-miniGsq complexes were collected cryo-EM data at Tsinghua University, and the α1AAR-Nb6 complex was performed data acquisition by the Shuimu Bioscience (Beijing). The purified protein complexes were concentrated to 5–10 mg/mL. 4 μl of sample were applied to the glow-discharged holey carbon grids (Au R1.2/1.3, 300 mesh) purchased from Quantifoil for the Nb29-α1AAR-miniGsq complexes, and from Zhongjingkeyi Technology (Beijing) for the α1AAR-Nb6 complex. The grids were blotted for 3.0 s and flash-frozen in liquid ethane cooled by liquid nitrogen with Vitrobot (Mark IV, Thermo Fisher Scientific) before being transferred to a 300 kV Titan Krios microscope equipped with Gatan K3 Summit detector and a GIF Quantum energy filter (slit width 20 eV) or Falcon-4 detector and no energy filter. AutoEMation was used for the fully automated data collection in Tsinghua University57. The total dose of each stack was about 50 e2. All frames in each stack were aligned and summed using the whole-image motion correction program MotionCor258 and binned to a pixel size of 1.083 Å/1.098 Å/0.860 Å for Nb29-α1AAR-miniGsq bound to oxymetazoline/ Nb29-α1AAR-miniGsq bound to noradrenaline/α1AAR-Nb6 bound to tamsulosin datasets, respectively. The defocus value of each image, which was set from −1.3 to −1.8 μm during data collection, was determined by Gctf59.

Cryo-EM data processing

For Nb29-α1AAR-miniGsq bound to oxymetazoline/ Nb29-α1AAR-miniGsq bound to noradrenaline/ α1AAR-Nb6 bound to tamsulosin datasets, 700/1099/1911 dose-weighted micrographs were imported into cryoSPARC and CTF parameters were estimated by using patch-CTF, respectively. 1,161,201/2,597,118/2,334,500 particles picked by blob picker or template picker were extracted and subjected to 2D classification. 697,792/774,350/885,066 particles remained to generate the initial model by Ab-Initio Reconstruction and perform the following iterative rounds of heterogeneous refinement. After non-uniform refinement and local refinement, 359,833/ 393,438/ 285,284 particles yield the maps which reached the resolutions at 2.92 Å/3.52 Å/3.35 Å.

Model building and refinement

The atomic coordinate of the Nb29-α1AAR-miniGsq and the α1AAR-Nb6 complexes was generated by combining homology modeling and de novo model building. An initial structure model for the active α1AAR was predicted by the homology model from GPCRdb (gpcrdb.org)60, α1AAR-kOR was generated by AlphaFold261 and the structure model of Nb29 was predicted by the homology model from swiss-model62. The initial models of miniGsq and Nb6 were imported from miniGs (PDB code: 5G53)29 and Nb6 (PDB code: 6VI4)33, respectively. The cryo-EM model was docked into the electron microscopy density map using UCSF Chimera63, followed by iterative manual adjustment and rebuilding in PHENIX64 and COOT65. The structures were refined against the corresponding map using PHENIX and COOT in real space with secondary structure and geometry restraints. Figures were created using the PyMOL Molecular Graphics System v.2.4.0 (Schrӧdinger, LLC), UCSF Chimera and the UCSF Chimera X1.3 package.

Glo-sensor signaling assay

To determine the signaling profile of Nb29 against α1AAR, we used a cAMP Glo-sensor kit (Promega) with an engineered Gsq protein in which 15 residues at the C-terminus of Gs protein were replaced with those of Gq protein. The Gsq protein could be activated by the α1AAR and stimulates intracellular cAMP production. In brief, pGloSensor™−22F plasmid, receptor plasmid, and Gsq plasmid were transfected into HEK293T cells. At 24 h after transfection, the cells were switched into a CO2-Independent Medium (Gibco) and incubated with GloSensorTM cAMP reagent. The mixture was then transferred to a 96-well white plate. The 96-well plate is placed at 37 °C in the dark for 1 h, then placed at room temperature in the dark for 1 h before use. The luminescence signal was measured by EnsightTM plate reader (PerkinElmer) around 10–15 min after the addition of the agonist and/or Nb29. The result curves were calculated and fitted by GraphPad Prism 9.

Reporting summary

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