Next Article in Journal
The Role of Resolvins, Protectins and Marensins in Non-Alcoholic Fatty Liver Disease (NAFLD)
Next Article in Special Issue
Alternative Pre-mRNA Splicing of the Mu Opioid Receptor Gene, OPRM1: Insight into Complex Mu Opioid Actions
Previous Article in Journal
The Role of Cholesterol in Amyloidogenic Substrate Binding to the γ-Secretase Complex
Previous Article in Special Issue
Integrated Multi-Class Classification and Prediction of GPCR Allosteric Modulators by Machine Learning Intelligence
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Ligands of Adrenergic Receptors: A Structural Point of View

1
iHuman Institute, ShanghaiTech University, Shanghai 201210, China
2
School of Life Science and Technology, ShanghaiTech University, Shanghai 201210, China
*
Author to whom correspondence should be addressed.
Biomolecules 2021, 11(7), 936; https://doi.org/10.3390/biom11070936
Submission received: 25 May 2021 / Revised: 9 June 2021 / Accepted: 12 June 2021 / Published: 24 June 2021
(This article belongs to the Special Issue GPCRs: Structure, Biology and Potential Applications)

Abstract

:
Adrenergic receptors are G protein-coupled receptors for epinephrine and norepinephrine. They are targets of many drugs for various conditions, including treatment of hypertension, hypotension, and asthma. Adrenergic receptors are intensively studied in structural biology, displayed for binding poses of different types of ligands. Here, we summarized molecular mechanisms of ligand recognition and receptor activation exhibited by structure. We also reviewed recent advances in structure-based ligand discovery against adrenergic receptors.

1. Introduction

Adrenergic receptors (adrenoreceptors, ARs) are membrane proteins mediating the actions of epinephrine and norepinephrine. Epinephrine and norepinephrine (also called (–)-adrenaline and (–)-noradrenaline, respectively) are neurotransmitters of the sympathetic nervous system and the central nervous system, and function as hormones secreted by the adrenal medulla. There are nine members of adrenergic receptors in humans (Figure 1a), widely distributed throughout the body and playing key roles in many important physiological processes, such as response to stress, control of heart rate and blood pressure, and regulation of metabolism [1,2,3]. Besides epinephrine and norepinephrine as emergency medicines, adrenergic receptors are also the target of many medications in treatment of various conditions, including hypertension, hypotension, heart failure, arrhythmias, and asthma [1,4]. In the 21st century, the development of new drugs against adrenergic receptors has slowed down, but the once most-neglected member β3-adrenergic receptor is raising more attention than ever (Figure 1b).
The adrenergic receptors belong to the rhodopsin family/class A G protein-coupled receptors (GPCRs). The β2-adrenergic receptor (β2AR) is a prototype GPCR for studies of sequence, structure, and function: β2AR is the first GPCR cloned [5], the second GPCR was determined for atomic structure [6,7] (only after rhodopsin), and the first GPCR solved for the active state structure in a functional complex [8] (with Gs protein heterotrimer). β-arrestins, the cytoplasmic partners responsible for desensitization of most GPCRs, were named after the β2AR [9,10]. Turkey β1AR has been comprehensively studied in structural biology (in this article, all structure of β1AR refers to turkey β1AR unless noted). A complex of β1AR and β-arrestin-1 was reported in 2020 [11].
β2AR and β1AR are among the GPCRs with the largest numbers of structures obtained, providing rich information for ligand recognition and activation mechanisms (β2AR has 38 structures for 19 ligands and β1AR has 30 structures for 17 ligands). In 2019, structures of the three α2 adrenergic receptors were also reported, disclosing the structural basis for the receptor-type selectivity of ligands. In this review, we summarized current achievements of structural study in adrenergic receptors and advances in drug discovery based on these structures.

2. Adrenergic Receptors in the Aminergic Receptor Subfamily

The nine adrenergic receptors in human body can be classified into three major types: α1, α2, and β, each having three subtypes. Members within a major type are highly similar in sequence and function, but in different major types are distinct in these aspects (see next section for detail). All the adrenergic receptors belong to the aminergic receptor subfamily, which has 42 members, forming a single clade in the phylogenetic tree of GPCRs [18,19,20,21]. The aminergic receptors bind monoamine neurotransmitters, acetylcholine, or trace amines, and share common features in sequence, structure, and function (Figure 2): in the orthosteric site (binding site of endogenous ligand), a conserved D3.32 (Ballesteros–Weinstein numbering, BWN [22]) forms a salt bridge to the positively charged amino group in ligand, and Y7.43 and W7.40 stabilize this salt bridge through a hydrogen bond and π-π interactions (Figure 3b). Because of the shared ligand recognition mechanism, polypharmacology (drugs interacting with multiple targets) is common in aminergic receptors. For example, ergotamine (an alkaloid produced by fungus) can bind to 22 aminergic receptors with Ki less than 1 μM [23]. Other notable cases include clozapine [24] and apomorphine [25]. Therefore, selectivity is a major concern in drug discovery against aminergic receptors.
The endogenous ligands of adrenergic receptors, epinephrine and norepinephrine, are monoamine neurotransmitters. Monoamine (also include dopamine, serotonin, and histamine) contains an amino group connected to an aromatic ring by a two-carbon chain (Figure 2b). When binding to receptors, the aromatic ring forms π-π interactions to F6.51 and F6.52 (Figure 2a and Figure 3b). Among the monoamines, dopamine is most similar to epinephrine and norepinephrine in chemical structure: each contains a catechol (1,2-dihydroxybenzene) group (thus collectively referred to as catecholamines). The catecholamines are products of three successive steps in the tyrosine metabolism pathway. Their receptors recognize the catechol hydroxy groups by S5.42 and S5.46 (Figure 2a and Figure 3b), two serine residues in TM5 that appear in all members of adrenergic and dopamine receptors. These two positions are non-conserved in other aminergic receptors. Among the catecholamines, ligand–receptor recognition is not strictly specific: dopamine also acts on adrenergic receptors [4,26,27], while epinephrine and norepinephrine are reported to stimulate some dopamine receptor subtypes [28,29].

3. Major Types and Subtypes of Adrenergic Receptors

Adrenergic receptors are among the most studied GPCRs. Research on their roles in sympathetic nerve stimulation started at the beginning of the 20th century. In the pre-molecular biology era, major types and subtypes of adrenergic receptors were identified through their pharmacological characteristics and tissue locations. All the nine members were cloned by the early 1990s, before the human genome project was completed [1,32].
Major types largely correspond to the sequences and functions of adrenergic receptors. Members within a major type share 51–64% sequence identities at full-length, while in different major types, the values decrease to 30–40% (Figure 3a). This is the basis: adrenergic receptors in each major type have similar affinities for many ligands and mainly couple to a specific G protein type (Table 1). Notably, pocket residues of α1ARs and α2ARs are more similar, thus they share some ligands that do not target βARs. For the distinct cell and tissue responses of adrenergic receptors, medications targeting different major types have different indications (Table 1). Two residues at the ligand binding site, 6.55 [33] and 7.39 [34], are reported to affect the major type selectivity of ligands (Figure 2a and Figure 3b,c).
Even within a major type, adrenergic receptor subtypes may have distinct effects, due to their tissue distribution. This is exhibited in the β adrenergic receptors: β1AR is mainly expressed in heart and adipose tissue, thus its activation produces cardiac stimulation and increased lipolysis; β2AR is ubiquitously expressed but with high levels in the smooth muscle, especially of the airway, thus its main effect is bronchodilation; and β3AR is mainly expressed in adipose tissue, thus is responsible for increased lipolysis [35]. Therefore, while using β antagonists to treat cardiovascular diseases, β1-selective agents have less asthma adverse effects than non-selective agents [36]. β3-selective agonists are widely investigated as anti-obesity agents, though they still face many challenges [37]. The benefits of subtype-selective drugs are also found in the other two major types: α1A1D selective antagonist tamsulosin has less impact on blood pressure while treating benign prostatic hyperplasia [38]; and α2C selective antagonists are in development as potential treatments for multiple psychiatric diseases [39]. Positions 5.39 [40], 5.43 [41], 6.58 [42], 6.62 [43] (ECL3), 7.32 [42], and 7.35 [44,45] are reported to affect the subtype selectivity of ligands. These residues are at the extracellular edge of the pocket (Figure 2a and Figure 3b,c).

4. Mechanisms of β Adrenergic Receptors

4.1. Activation Mechanism

For a class A GPCR, the most significant conformational changes during activation are of the outward movement of the cytoplasmic end of TM6 and rearrangement of TM7 (Figure 4a). About 90 receptors [52] share a common mechanism: the conformational changes are mediated by the microswitch of the P5.50-I3.40-F6.44 motif, induced by an agonist. How agonists trigger the microswitch of the P-I-F motif largely varies in different receptor families. For β2AR or β1AR, an agonist forms a hydrogen bond to S5.46, thus pulling TM5 closer to TM3/6. An antagonist, in contrast, lacks this hydrogen bond and tends to occupy the space with hydrophobic moiety, thus preventing the movement of TM5 (Figure 4b). This mechanism is presented in all β21 agonists and antagonists (Figure 4c, representative ligands are listed in Table S1).
Binding of the orthosteric agonist and coupling of G protein have an effect on each other: agonists have higher affinities in β2AR with the presence of G protein than in the absence of G protein [53]. Structure studies of agonists binding in different states of β1AR disclosed the molecular basis: active state structures have smaller pockets when comparing to agonist-bound inactive state structures (Figure 4d) [54]. Inferred from the mechanism, antagonists tend to fit in more expanded orthosteric pockets. This was established in different types of class A GPCRs, including lipid receptors [55,56,57,58], peptide receptors [59,60,61], and nucleotide receptors [62,63,64,65].

4.2. Biased Signaling

Some agonists stimulate the β-arrestin signaling with higher or lower EC50 comparing to the G protein signaling (called biased signaling). For βARs, separately targeting β-arrestin pathways or G protein pathways is considered a therapeutic potential for diseases, including heart failure and asthma [78,79]. Structures of GPCR in complex with arrestin have been obtained in several receptors [80,81,82,83] including β1AR [11]. Compared to Gs-coupled β1AR, arrestin-coupled β1AR adopts a smaller degree of outward movement of TM6 and additional twist of TM7 (Figure 5a). Correspondingly, β1AR forms interactions to arrestin with more residues in TM2/7 and fewer residues in TM5/6 [11,66]. In accordance with this, NMR experiments of 19F-labeling on the cytoplasmic ends of TM6 and TM7 showed that arrestin-biased agonists induce more conformational change of TM7 than of TM6 for β2AR in solution [84]. Structures showed that arrestin-biased agonists carvedilol [85], bucindolo [86], formoterol [87], and BI-167107 [88] have an aryl-substituted alkyl tail attached to the amino N atom, and the bulky aromatic rings occupy an extended pocket formed by TM2/3/7 and ECL2 in β2AR or β1AR [8,11,41,66,75,89,90] (Figure 5). In contrast, Gs-biased agonist salmeterol occupies this space but forms distinct interactions with a long aryloxyalkyl tail [70]. However, there are also mysteries of arrestin-biased signaling: an approved β agonist drug, isoetharine, is arrestin-biased in β2AR [91] but Gs-biased in β1AR [92]. The structural mechanism remains to be explored.

4.3. Partial Agonism

Many βARs ligands in clinical use are partial agonists, which do not produce a full response, even at high concentration. NMR [84,93,94], single-molecule fluorescence resonance energy transfer [95], and double electron–electron resonance [94,96] results showed that partial agonists induce a smaller shift of conformational equilibrium of TM6, compared to full agonists. A well-identified key interaction affecting ligand efficacy (ratio of maximal response of the ligand to full response) is a hydrogen bond to N6.55 by a full agonist, but not a partial agonist (Figure 5b). Mutations of N6.55 and S5.43 (which forms hydrogen bond to N6.55) in β2AR reduce the efficacy of full agonist isoproterenol, strongly supporting that this hydrogen bond network is crucial to the full activation of βARs [70]. All full agonists can form a hydrogen bond to N6.55, either by 3-hydroxy in catechol (like epinephrine [31,43] or isoproterenol [54,66]) or groups mimicking catechol (such as formoterol [11] or BI-167107 [8,43,90]). Most partial agonists cannot form this hydrogen bond for replacement of the 3-hydroxy, including salmeterol [70], salbutamol [54,69], xamoterol [54], and clenbuterol. However, this cannot explain dobutamine, a partial agonist of β1AR and β2AR, reserving the catechol moiety and forming a hydrogen bond with N6.55 [54]. Therefore, there may be other receptor–ligand interactions affecting ligand efficacy to be discovered. NMR experiments showed that in β2AR, the chemical shift of M2.53 depends on the efficacy of the ligand [97], and dobutamine’s direct contact with TM2 (at 2.61, 2.64, and 2.65). Whether this is the basis of partial agonism of dobutamine remains to be investigated.

4.4. β12 Subtype Selectivity

β1AR and β2AR are highly similar in sequence (identity 57% at full-length and 70% at pocket residues) and structure (root mean square deviation 0.71 Å for inactive state and 1.00 Å for Gs-coupled active state). Despite their high homology, numerous agonists and antagonists selective for β1AR or β2AR have been discovered. Even endogenous agonist norepinephrine, small in size, thus all the interacting residues are the same in the two receptors, is approximately 10-fold selective for β1AR (while the other endogenous ligand epinephrine, with only one extra methyl group, is nonselective). Revealing of the molecular mechanisms of β12 subtype selectivity began only recently.
In β1AR and β2AR, although all the residues at the orthosteric site are the same, some residues at the edge of the pocket are different. Large ligands may directly interact with these residues. For example, salmeterol (>1000-fold selective for β2AR) has a long aryloxyalkyl tail, forming additional interactions to TM2, TM3, TM7, ECL2, and ECL3. Three of the interacting residues, while mutated to the corresponding residue in β1AR (H6.58N, K7.32D, and Y7.35F), largely reduce the affinity of salmeterol [70]. Residues at ECL2 and the extracellular end of TM6/7 sculpt the different shapes of pocket entrances in the two β adrenergic receptors: in β2AR, K7.32 forms a salt bridge to D45.51, covering the space among TM2/3/7; and in human β1AR, a different salt bridge is formed by R6.62 to E205 (ECL2), covering the space among TM3/5/6. Therefore, ligands potentially take different paths to enter β1AR and β2AR. The passage in β1AR is continuously negatively charged, thus preferred by norepinephrine, while epinephrine, a secondary amine, is less affected. This hypothesis explains the fast association rate of norepinephrine in β1AR and is supported by free energy profiles, calculated using metadynamics simulations [43].

5. Mechanisms of α2 Receptors

5.1. Ligand Binding

Structures of α2AAR [98] and α2CAR [99] in an inactive state and α2BAR [52] in an active state are disclosed. The binding of both agonist and antagonist in α2 adrenergic receptors involves more π-π or cation-π interactions and fewer hydrogen bonds (Figure 6a,b), comparing to βARs (Figure 3b). The expansion of π interaction network is mainly caused by F7.39, conserved in all α1ARs and α2ARs but is N7.39 in βARs. In the agonist-bound structures, F7.39 acts as a ‘lid’, covering the imidazole ring of ligand from the extracellular side (Figure 6c). Typical α agonists and some α antagonists contain positively charged planar groups, such as imidazole, imidazoline, and guanidine [100], which fit well in this closed pocket. In inactive structures, the ‘lid’ F7.39 is pushed up to another rotamer by antagonist (Figure 6d), making the pocket larger, thus able to accommodate bulky ligands. Some α ligands with such chemical features are in clinical use, including piperazine derivatives as α1 antagonists (e.g., prazosin) and indole alkaloids with multiple fused rings as α2 antagonists (e.g., yohimbine).

5.2. α/β Selectivity

The binding of α agonists and antagonists relies on π interactions involving F7.39, thus, is not favored in βARs with N7.39. β ligands, instead, largely prefer to retain ethanolamine moiety [101] to form dual hydrogen bonds to N7.39 (Figure 6c). Another feature of β ligands is a single large substituent on the amino N atom. A large group at this position is not favored in αARs as it is too large to fit in the closed pocket and too flexible to trigger the ‘switching lid’. Only one substitution is allowed because only primary and secondary amines can form a hydrogen bond to N7.39 in βARs. This is validated by the polypharmacology profile of ergotamine, a tertiary amine: the binding affinities are 0.3–3 nM in α2ARs, 10–30 nM in α1ARs, and ~100 nM in β2AR and >10,000 nM in β1AR/β3AR [23].

5.3. Alternative Activation Mechanism

Activation of β1AR/β2AR requires an agonist forming a hydrogen bond to S5.46. In α2ARs, the catecholamine agonists can form this hydrogen bond, but many α2 agonists, including dexmedetomidine and clonidine, do not have polar groups at this position. Mutation S5.42A or S5.46A in α2AAR reduce the activation of epinephrine, but not of dexmedetomidine and clonidine [102]. Structures show that dexmedetomidine activates α2B by forming π-π interactions to W6.48 and pushing it downward (Figure 6d). The downward movement of W6.48 is part of the common activation mechanism in class A GPCRs (Figure 4a). To effectively push W6.48, the ligand must compactly fill the pocket. This mechanism explains the functions of imidazole/imidazoline derivatives as α2 drugs: dexmedetomidine and clonidine are agonists for they fill the pocket, while tolazoline, without substituents on the benzen ring, thus being smaller in size, is an antagonist (Figure 6e). This series of imidazole/imidazoline derivatives is a special case in GPCR ligands, as normally, antagonists are larger than agonists because the pocket of inactive state is larger.

5.4. Pathway Selectivity

Position 6.55 is the only one in the orthosteric site that each major type of adrenergic receptor has a residue with a specific property: α1ARs, hydrophobic M6.55 (α1AAR) or L6.55 (α1BAR/α1DAR); α2ARs, aromatic Y6.55; βARs, polar N6.55. Interestingly, Y6.55 in α2ARs shows a distinct impact on different downstream pathways: α2ARs mainly functions through the Gi pathway, but they also couple to Gs. Mutation Y6.55N demolishes Gs signaling but retains Gi signaling in α2AAR, suggesting that this residue has a subtle impact on receptor conformation [98]. In both active and inactive structures of α2ARs, Y6.55 lies horizontally, and forms van der Waals interactions with ligands from the extracellular site. In active state α2BAR, Y6.55 forms a hydrogen bond with S5.42 (Figure 6a). This hydrogen bond may be comparable to the hydrogen bond between N6.55 and S5.43 in βARs, while N6.55 and S5.43 affect the efficacy of agonists [70]. Structures also exhibit other possible roles of Y6.55: first, Y6.55 defines the upper boundary of the pocket (Figure 6e), thus rejecting many β ligands with large groups at the catechol recognizing side (e.g., formoterol and carvedilol); second, Y6.55 blocks the TM3/5/6 space, leaving TM2/3/7 the only passage for ligands. This might be the basis for both epinephrine and norepinephrine having similar affinities in all three α2 adrenergic receptors [13,43].

5.5. α2A2C Subtype Selectivity

α2ARs are highly similar in sequence, but there are still α2 ligands with high subtype selectivity. The molecular docking of two α2C antagonists JP1302 (100 folds selectivity to α2AAR, 50 folds to α2BAR [103]) and OPC-28326 (300 folds to α2AAR, 50 folds to α2BAR [104]) suggested that the two compounds, large and long, interact with non-conserved extracellular residues. Mutagenesis showed that an interaction network among Y6.58 (unique to α2CAR), R6.62, and D45.54 are key to the functions of JP1302 and OPC-28326, while in α2AAR, mutations of a different interaction network involving R7.32 (unique to α2AAR) restore their functions [99]. Such impact of extracellular residues on subtype selectivity is comparable to the β12 selectivity case.

6. Allosteric Modulations of β2 Adrenergic Receptor

In a GPCR, a ligand may bind outside the orthosteric site and stabilize a specific state, and this phenomenon is called allosteric modulation [105,106]. An allosteric ligand increasing the signaling of orthosteric agonists is a positive allosteric modulator (PAM), and one decreasing the signaling is a negative allosteric modulator (NAM). Allosteric modulators have raised attention in drug discovery against GPCRs because targeting less-conserved allosteric sites may produce subtype selectivity. The most effective way to identify an allosteric site is determining the structure of the GPCR in complex with the allosteric modulator [107]. Structures of β2AR captured small molecules binding at different sites and exhibited how they affect the signaling (Figure 7). The cholesterol binding sites were thoroughly reviewed recently [108], thus is not discussed here.

6.1. PAM and NAM at Lipid Interface of TM3/4/5

The arrangement of seven TM helices leads to a groove at the lipid interface formed by TM3/4/5 in all classes of GPCRs. In β2AR, both PAM and NAM binds in this groove are identified by crystal structures (Figure 7b). The PAM and the NAM locate at different heights and stabilize different states through specific interactions: the PAM Cmpd-6FA binds close to the cytoplasmic end and interacts with ICL2 and adjacent residues on TM3 and TM4 [110]. At this region, GPCRs interact with multiple segments of Gα, and Cmpd-6FA stabilizes conformation of the receptor. The NAM AS408 binds close to the center of the receptor and interacts with TM3 and TM5 [111]. AS408 forms a hydrogen bond to the main-chain O atom of V5.45, thus stabilizing the inactive conformation of TM5 and the P-I-F microswitch motif. Both the PAM and the NAM are selective for β2AR over β1AR; for positive modulating, two key residues (F3.52/L3.52 and K4.41/R4.41 in β2AR/β1AR) are identified. Since functions of ICL2 and TM5 are common in activation process of class A GPCRs, it is not surprising that allosteric modulators binding at the same site were also discovered in other receptors: PAMs AP8 [112] and Cpd-1 [113] of free fatty acid receptor 1, PAM LY3154207 [114] of dopamine receptor D1, and NAMs NDT9513727 [115,116] and avancopan [117] of C5a anaphylatoxin chemotactic receptor 1.

6.2. NAM at Cytoplasmic Surface

GPCRs use the cytoplasmic surface to interact with intracellular partners, including G proteins and β-arrestins. Small molecules binding at this site were mainly discovered in chemokine receptors [116,117,118,119], while Cmpd-15PA in β2AR [120] is the only case outside the chemokine subfamily (Figure 7c). All these small molecules are NAMs. Cmpd-15PA has little collisions with G protein but stabilizes the inactive state of the receptor through extensive interactions with TM1, TM2, TM6, TM7, H8, and ICL1. Most notably, the hydrogen bond to T6.36 and π-π interaction to Y7.53 block the outward movement of TM6 and rearrangement of TM7. Although β2AR and β1AR only differ in one position (F1.60 in β2AR and T1.60 in β1AR) among all residues interacting with Cmpd-15PA, the Cmpd-15 (parent compound of Cmpd-15PA with comparable effects) decreases the EC50 of orthosteric agonist in only β2AR but not β1AR [121].

7. Structure-Based Ligand Discovery of Adrenergic Receptors

7.1. Virtual Screening with Molecular Docking

Structure-based virtual screening can identify ligands with novel scaffolds and chemical types [122,123]. Since structures of β2AR are reported, new antagonists and agonists were discovered through virtual screening with molecular docking. Docking screening identified a new antagonist with benzhydrylpiperazine moiety having high sub-nM affinity (Ki = 0.311 nM) [124]. Screening of the lead-like subset of the ZINC database [125] identified 5 hit inverse agonists (among 25 tested compounds) with Ki < 40 μM, while the strongest had Ki = 9 nM [126]. Using an active state structure, docking screening of the lead-like and fragment-like subsets of the ZINC database identified 6 agonists (4 full and 2 partial, among 22 tested compounds) [88].

7.2. Rational Design

Structures can guide modification of known ligands to achieve new function. Two designed β agonists, analogs of BI-167107 and epinephrine but that can form disulfide bond, covalently bind to β2AR with designed mutation H2.64C. They were obtained for the crystal structure in the agonist-bound inactive state [127] and nanobody-coupled active state [128], respectively. Adding of a methyl group to cyanopindolol (partial agonist of β1AR and β2AR) at a position close to S5.46 further reduced its efficacy [77]. Inspired by structures of β adrenergic receptors, five compounds based on atypical antipsychotic drug aripiprazole were designed and shown to have different effect on dopamine receptor D2 (one unbiased agonist, two arrestin-biased agonists, and two antagonists) [129].

7.3. Predicting Actions of Ligands

Molecular docking aims to identify compounds with high affinity but does not distinguish between an agonist and antagonist. Molecular interaction fingerprints derived from structures of a GPCR with various ligands have been developed to predict the actions of ligand. When only the first inactive β2AR structure was available, the molecular interaction fingerprint was generated based on docking poses of ligands [130]. The newest version of fingerprint was developed based on 31 structures of β1AR and β2AR [131]. Molecular docking followed by molecular interaction fingerprint (derived from 13 structures) screening identified 26 hit agonists (among 63 tested compounds) of β2AR with EC50 < 100 μM, while the strongest had EC50 38 nM [132].

7.4. Accelerated Structure Determination

Structure-based drug design against GPCRs relies on a large number of solved structures binding various ligands. Since 2018, many structures of active state GPCRs in functional complexes have been solved with the rapidly developed cryogenic electron microscopy (cryo-EM) technique. A remarkable case is dopamine receptor D1: no structure was reported before February 2021, and then 11 structures were released within two months [113,133,134,135]. For adrenergic receptors, up to April 2021, a total of eight cryo-EM structures were published: one is arrestin-coupled β1AR and the others are G protein-coupled (four β2AR, Gs [8,68,69]; one β1AR, Gs [66]; two α2BAR, Gi and Go [52]).
Crystallography is still powerful for the structural study of adrenergic receptors. Thermostabilized turkey β1AR can co-crystallize with small, fragment-like ligands [89]. Crystal structures of human β1AR were first reported in 2020, in both a nanobody-coupled active state and inactive state, exhibiting a detailed difference from turkey β1AR [43]. For β2AR, structures in complex with eight ligands were obtained using a newly developed transient ligand exchange method termed Complex-LCP (Crystallization of Membrane Proteins using transient Ligand EXchange in LCP). The Complex-LCP method relies on the serial femtosecond crystallography (SFX) data-collection approach, using extremely bright and short pulses generated by an X-ray free-electron laser (XFEL), thus collects data from small crystals at room temperature [75].

8. Conclusions

Structures of adrenergic receptors exhibited mechanisms of ligand-induced effects, including agonism/antagonism, partial agonism, and biased signaling. A preliminary picture of ligand function in β adrenergic receptors was generated. The picture in α2 is still incomplete, while that in α1 remains to be uncovered. Interpreting of these mechanisms would help to discover new drugs with desired effects against adrenergic receptors.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/biom11070936/s1, Table S1: Representative ligands used in structural studies of adrenergic receptors.

Author Contributions

Investigation, Y.W. and L.Z.; writing—original draft preparation, Y.W.; writing—review and editing, S.Z.; visualization, Y.W.; supervision, S.Z.; funding acquisition, S.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Key R&D Program of China, grant number 2018YFA0507000 (S.Z.) and National Natural Science Foundation of China, grant number 31971178 (S.Z.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bylund, D.B.; Eikenberg, D.C.; Hieble, J.P.; Langer, S.Z.; Lefkowitz, R.J.; Minneman, K.P.; Molinoff, P.B.; Ruffolo, R.R.; Trendelenburg, U. International Union of Pharmacology nomenclature of adrenoceptors. Pharmacol. Rev. 1994, 46, 121–136. [Google Scholar]
  2. Ciccarelli, M.; Santulli, G.; Pascale, V.; Trimarco, B.; Iaccarino, G. Adrenergic receptors and metabolism: Role in develop-ment of cardiovascular disease. Front Physiol. 2013, 4, 265. [Google Scholar] [CrossRef] [Green Version]
  3. Bylund, D.B. Adrenergic Receptors. In Encyclopedia of Biological Chemistry, 2nd ed.; Lennarz, W.J., Lane, M.D., Eds.; Academic Press: Waltham, MA, USA, 2013; pp. 57–60. [Google Scholar]
  4. Farzam, K.; Kidron, A.; Lakhkar, A.D. Adrenergic Drugs; StatPearls: Treasure Island, FL, USA, 2021. [Google Scholar]
  5. Kobilka, B.K.; Dixon, R.A.; Frielle, T.; Dohlman, H.G.; Bolanowski, M.A.; Sigal, I.S.; Yang-Feng, T.L.; Francke, U.; Caron, M.G.; Lefkowitz, R.J. cDNA for the human beta 2-adrenergic receptor: A protein with multiple membrane-spanning do-mains and encoded by a gene whose chromosomal location is shared with that of the receptor for platelet-derived growth factor. Proc. Natl. Acad. Sci. USA 1987, 84, 46–50. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Cherezov, V.; Rosenbaum, D.M.; Hanson, M.A.; Rasmussen, S.G.; Thian, F.S.; Kobilka, T.S.; Choi, H.J.; Kuhn, P.; Weis, W.I.; Kobilka, B.K.; et al. High-resolution crystal structure of an engineered human beta2-adrenergic G protein-coupled receptor. Science 2007, 318, 1258–1265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Rasmussen, S.G.; Choi, H.J.; Rosenbaum, D.M.; Kobilka, T.S.; Thian, F.S.; Edwards, P.C.; Burghammer, M.; Ratnala, V.R.; Sanishvili, R.; Fischetti, R.F.; et al. Crystal structure of the human beta2 adrenergic G-protein-coupled receptor. Nature 2007, 450, 383–387. [Google Scholar] [CrossRef] [PubMed]
  8. Rasmussen, S.G.; DeVree, B.T.; Zou, Y.; Kruse, A.C.; Chung, K.Y.; Kobilka, T.S.; Thian, F.S.; Chae, P.S.; Pardon, E.; Calinski, D.; et al. Crystal structure of the beta2 adrenergic receptor-Gs protein complex. Nature 2011, 477, 549–555. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Lohse, M.J.; Benovic, J.L.; Codina, J.; Caron, M.G.; Lefkowitz, R.J. beta-Arrestin: A protein that regulates beta-adrenergic re-ceptor function. Science 1990, 248, 1547–1550. [Google Scholar] [CrossRef] [PubMed]
  10. Attramadal, H.; Arriza, J.; Aoki, C.; Dawson, T.; Codina, J.; Kwatra, M.; Snyder, S.; Caron, M.; Lefkowitz, R. Beta-arrestin2, a novel member of the arrestin/beta-arrestin gene family. J. Biol. Chem. 1992, 267, 17882–17890. [Google Scholar] [CrossRef]
  11. Lee, Y.; Warne, T.; Nehme, R.; Pandey, S.; Dwivedi-Agnihotri, H.; Chaturvedi, M.; Edwards, P.C.; Garcia-Nafria, J.; Leslie, A.G.W.; Shukla, A.K.; et al. Molecular basis of beta-arrestin coupling to formoterol-bound beta1-adrenoceptor. Nature 2020, 583, 862–866. [Google Scholar] [CrossRef] [PubMed]
  12. Hone, K.; Obika, K.; Foglar, R.; Tsujimoto, G. Selectivity of the imidazoline α-adrenoceptor agonists (oxymetazoline and cirazoline) for human cloned α1-adrenoceptor subtypes. Br. J. Pharmacol. 1995, 116, 1611–1618. [Google Scholar] [CrossRef]
  13. Jasper, J.R.; Lesnick, J.D.; Chang, L.K.; Yamanishi, S.S.; Chang, T.K.; Hsu, S.A.; Daunt, D.A.; Bonhaus, D.W.; Eglen, R.M. Lig-and efficacy and potency at recombinant alpha2 adrenergic receptors: Agonist-mediated [35S]GTPgammaS binding. Biochem. Pharmacol. 1998, 55, 1035–1043. [Google Scholar] [CrossRef]
  14. Hoffmann, C.; Leitz, M.R.; Oberdorf-Maass, S.; Lohse, M.J.; Klotz, K.N. Comparative pharmacology of human be-ta-adrenergic receptor subtypes--characterization of stably transfected receptors in CHO cells. Naunyn Schmiedebergs Arch. Pharmacol. 2004, 369, 151–159. [Google Scholar] [CrossRef]
  15. Armstrong, J.F.; Faccenda, E.; Harding, S.D.; Pawson, A.J.; Southan, C.; Sharman, J.L.; Campo, B.; Cavanagh, D.R.; Alexander, S.P.H.; Davenport, A.P.; et al. The IUPHAR/BPS Guide to PHARMACOLOGY in 2020: Extending immunopharmacology content and introducing the IUPHAR/MMV Guide to MALARIA PHARMACOLOGY. Nucleic Acids Res. 2019, 48, D1006–D1021. [Google Scholar] [CrossRef]
  16. Kooistra, A.J.; Mordalski, S.; Pándy-Szekeres, G.; Esguerra, M.; Mamyrbekov, A.; Munk, C.; Keserű, G.M.; Gloriam, E.D. GPCRdb in 2021: Integrating GPCR sequence, structure and function. Nucleic Acids Res. 2021, 49, D335–D343. [Google Scholar] [CrossRef]
  17. Wishart, D.S.; Feunang, Y.D.; Guo, A.C.; Lo, E.J.; Marcu, A.; Grant, J.R.; Sajed, T.; Johnson, D.; Li, C.; Sayeeda, Z.; et al. DrugBank 5.0: A Major Update to the DrugBank Database for 2018. Nucleic Acids Res. 2018, 46, D1074–D1082. [Google Scholar] [CrossRef]
  18. Surgand, J.-S.; Rodrigo, J.; Kellenberger, E.; Rognan, D. A chemogenomic analysis of the transmembrane binding cavity of human G-protein-coupled receptors. Proteins Struct. Funct. Bioinform. 2005, 62, 509–538. [Google Scholar] [CrossRef] [PubMed]
  19. Katritch, V.; Cherezov, V.; Stevens, R.C. Structure-Function of the G Protein–Coupled Receptor Superfamily. Annu. Rev. Pharmacol. Toxicol. 2013, 53, 531–556. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Flock, T.; Hauser, A.; Lund, N.; Gloriam, D.E.; Balaji, S.; Babu, M.M. Selectivity determinants of GPCR–G-protein binding. Nat. Cell Biol. 2017, 545, 317–322. [Google Scholar] [CrossRef] [PubMed]
  21. Gloriam, D.E.; Foord, S.M.; Blaney, F.E.; Garland, S. Definition of the G Protein-Coupled Receptor Transmembrane Bundle Binding Pocket and Calculation of Receptor Similarities for Drug Design. J. Med. Chem. 2009, 52, 4429–4442. [Google Scholar] [CrossRef]
  22. Ballesteros, J.A.; Weinstein, H. [19] Integrated methods for the construction of three-dimensional models and computational probing of structure-function relations in G protein-coupled receptors. In Methods in Neurosciences; Sealfon, S.C., Ed.; Academic Press: Waltham, MA, USA, 1995; Volume 25, pp. 366–428. [Google Scholar]
  23. Peng, Y.; McCorvy, J.D.; Harpsøe, K.; Lansu, K.; Yuan, S.; Popov, P.; Qu, L.; Pu, M.; Che, T.; Nikolajsen, L.F.; et al. 5-HT2C Receptor Structures Reveal the Structural Basis of GPCR Polypharmacology. Cell 2018, 172, 719–730.e14. [Google Scholar] [CrossRef] [Green Version]
  24. Roth, B.L.; Sheffler, D.J.; Kroeze, W.K. Magic shotguns versus magic bullets: Selectively non-selective drugs for mood dis-orders and schizophrenia. Nat. Rev. Drug Discov. 2004, 3, 353–359. [Google Scholar] [CrossRef] [Green Version]
  25. Butini, S.; Nikolic, K.; Kassel, S.; Brückmann, H.; Filipic, S.; Agbaba, D.; Gemma, S.; Brogi, S.; Brindisi, M.; Campiani, G.; et al. Polypharmacology of dopamine receptor ligands. Prog. Neurobiol. 2016, 142, 68–103. [Google Scholar] [CrossRef]
  26. Baker, J.G. The selectivity of beta-adrenoceptor agonists at human beta1-, beta2- and beta3-adrenoceptors. Br. J. Pharmacol. 2010, 160, 1048–1061. [Google Scholar] [CrossRef] [Green Version]
  27. Lei, S. Cross interaction of dopaminergic and adrenergic systems in neural modulation. Int. J. Physiol. Pathophysiol. Pharmacol. 2014, 6, 137–142. [Google Scholar]
  28. Lanau, F.; Zenner, M.T.; Civelli, O.; Hartman, D.S. Epinephrine and norepinephrine act as potent agonists at the recombi-nant human dopamine D4 receptor. J. Neurochem. 1997, 68, 804–812. [Google Scholar] [CrossRef]
  29. Sánchez-Soto, M.; Bonifazi, A.; Cai, N.S.; Ellenberger, M.P.; Newman, A.H.; Ferré, S.; Yano, H. Evidence for Noncanonical Neurotransmitter Activation: Norepinephrine as a Dopamine D2-Like Receptor Agonist. Mol. Pharmacol. 2016, 89, 457–466. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Katoh, K.; Standley, D.M. MAFFT multiple sequence alignment software version 7: Improvements in performance and usability. Mol. Biol. Evol. 2013, 30, 772–780. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Ring, A.M.; Manglik, A.; Kruse, A.C.; Enos, M.D.; Weis, W.I.; Garcia, K.C.; Kobilka, B.K. Adrenaline-activated structure of beta2-adrenoceptor stabilized by an engineered nanobody. Nature 2013, 502, 575–579. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Insel, P.A. Adrenergic Receptors — Evolving Concepts and Clinical Implications. N. Engl. J. Med. 1996, 334, 580–585. [Google Scholar] [CrossRef]
  33. Pooput, C.; Rosemond, E.; Karpiak, J.; Deflorian, F.; Vilar, S.; Costanzi, S.; Wess, J.; Kirk, K.L. Structural basis of the selectiv-ity of the beta(2)-adrenergic receptor for fluorinated catecholamines. Bioorg. Med. Chem. 2009, 17, 7987–7992. [Google Scholar] [CrossRef] [Green Version]
  34. Suryanarayana, S.; Daunt, D.; Von Zastrow, M.; Kobilka, B. A point mutation in the seventh hydrophobic domain of the alpha 2 adrenergic receptor increases its affinity for a family of beta receptor antagonists. J. Biol. Chem. 1991, 266, 15488–15492. [Google Scholar] [CrossRef]
  35. Hieble, J.P. Adrenergic Receptors. In Encyclopedia of Neuroscience; Squire, L.R., Ed.; Academic Press: Oxford, MS, USA, 2009; pp. 135–139. [Google Scholar]
  36. Morales, D.R.; Jackson, C.; Lipworth, B.J.; Donnan, P.T.; Guthrie, B. Adverse respiratory effect of acute beta-blocker exposure in asthma: A systematic review and meta-analysis of randomized controlled trials. Chest 2014, 145, 779–786. [Google Scholar] [CrossRef]
  37. Arch, J.R. Challenges in beta(3)-Adrenoceptor Agonist Drug Development. Ther. Adv. Endocrinol. Metab. 2011, 2, 59–64. [Google Scholar] [CrossRef] [Green Version]
  38. Schwinn, D.A.; Price, D.T.; Narayan, P. α1-Adrenoceptor Subtype Selectivity and Lower Urinary Tract Symptoms. Mayo Clin. Proc. 2004, 79, 1423–1434. [Google Scholar] [CrossRef] [Green Version]
  39. Uys, M.M.; Shahid, M.; Harvey, B.H. Therapeutic Potential of Selectively Targeting the α2C-Adrenoceptor in Cognition, Depression, and Schizophrenia—New Developments and Future Perspective. Front. Psychiatry 2017, 8, 144. [Google Scholar] [CrossRef] [Green Version]
  40. Hwa, J.; Graham, R.M.; Perez, D.M. Identification of Critical Determinants of α1-Adrenergic Receptor Subtype Selective Agonist Binding. J. Biol. Chem. 1995, 270, 23189–23195. [Google Scholar] [CrossRef] [Green Version]
  41. Laurila, J.M.; Xhaard, H.; Ruuskanen, J.O.; Rantanen, M.J.; Karlsson, H.K.; Johnson, M.S.; Scheinin, M. The second extracellu-lar loop of alpha2A-adrenoceptors contributes to the binding of yohimbine analogues. Br. J. Pharmacol. 2007, 151, 1293–1304. [Google Scholar] [CrossRef] [Green Version]
  42. Baker, J.G.; Proudman, R.G.; Hill, S.J. Salmeterol’s extreme beta2 selectivity is due to residues in both extracellular loops and transmembrane domains. Mol. Pharmacol. 2015, 87, 103–120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Xu, X.; Kaindl, J.; Clark, M.J.; Hubner, H.; Hirata, K.; Sunahara, R.K.; Gmeiner, P.; Kobilka, B.K.; Liu, X. Binding pathway determines norepinephrine selectivity for the human beta1AR over beta2AR. Cell Res. 2021, 31, 569–579. [Google Scholar] [CrossRef] [PubMed]
  44. Woo, A.Y.; Jozwiak, K.; Toll, L.; Tanga, M.J.; Kozocas, J.A.; Jimenez, L.; Huang, Y.; Song, Y.; Plazinska, A.; Pajak, K.; et al. Tyrosine 308 is necessary for ligand-directed Gs protein-biased signaling of beta2-adrenoceptor. J. Biol. Chem. 2014, 289, 19351–19363. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Michino, M.; Beuming, T.; Donthamsetti, P.; Newman, A.H.; Javitch, J.A.; Shi, L. What Can Crystal Structures of Aminergic Receptors Tell Us about Designing Subtype-Selective Ligands? Pharmacol. Rev. 2015, 67, 198–213. [Google Scholar] [CrossRef] [Green Version]
  46. Dean, I.J.; Reddivari, A.K.R. Alpha 1 Receptor Agonists; StatPearls Publishing: Treasure Island, FL, USA, 2021. [Google Scholar]
  47. Nachawati, D.; Patel, J. Alpha Blockers; StatPearls Publishing: Treasure Island, FL, USA, 2021. [Google Scholar]
  48. Giovannitti, J.A.; Thoms, S.M.; Crawford, J.J. Alpha-2 Adrenergic Receptor Agonists: A Review of Current Clinical Applications. Anesthesia Prog. 2015, 62, 31–38. [Google Scholar] [CrossRef] [Green Version]
  49. Morales, A. Yohimbine in erectile dysfunction: The facts. Int. J. Impot. Res. 2000, 12, S70–S74. [Google Scholar] [CrossRef] [Green Version]
  50. García-Sevilla, A.J.; Ventayol, P.; Pérez, V.; Rubovszky, G.; Puigdemont, D.; Ferrer-Alcón, M.; Andreoli, A.; Guimón, J.; Alvarez, E. Regulation of Platelet α2A-Adrenoceptors, Gi Proteins and Receptor Kinases in Major Depression: Effects of Mirtazapine Treatment. Neuropsychopharmacology 2003, 29, 580–588. [Google Scholar] [CrossRef]
  51. Dezsi, C.A.; Szentes, V. The Real Role of beta-Blockers in Daily Cardiovascular Therapy. Am. J. Cardiovasc. Drugs 2017, 17, 361–373. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Yuan, D.; Liu, Z.; Kaindl, J.; Maeda, S.; Zhao, J.; Sun, X.; Xu, J.; Gmeiner, P.; Wang, H.-W.; Kobilka, B.K. Activation of the α2B adrenoceptor by the sedative sympatholytic dexmedetomidine. Nat. Chem. Biol. 2020, 16, 507–512. [Google Scholar] [CrossRef] [PubMed]
  53. DeVree, B.; Mahoney, J.P.; Vélez-Ruiz, G.A.; Rasmussen, S.G.F.; Kuszak, A.; Edwald, E.; Fung, J.-J.; Manglik, A.; Masureel, M.; Du, Y.; et al. Allosteric coupling from G protein to the agonist-binding pocket in GPCRs. Nat. Cell Biol. 2016, 535, 182–186. [Google Scholar] [CrossRef] [Green Version]
  54. Warne, T.; Edwards, P.C.; Doré, A.S.; Leslie, A.G.W.; Tate, C.G. Molecular basis for high-affinity agonist binding in GPCRs. Science 2019, 364, 775–778. [Google Scholar] [CrossRef]
  55. Hua, T.; Vemuri, K.; Pu, M.; Qu, L.; Han, G.W.; Wu, Y.; Zhao, S.; Shui, W.; Li, S.; Korde, A.; et al. Crystal Structure of the Human Cannabinoid Receptor CB1. Cell 2016, 167, 750–762.e14. [Google Scholar] [CrossRef] [Green Version]
  56. Hua, T.; Li, X.; Wu, L.; Iliopoulos_Tsoutsouvas, C.; Wang, Y.; Wu, M.; Shen, L.; Brust, C.A.; Nikas, S.P.; Song, F.; et al. Activation and Signaling Mechanism Revealed by Cannabinoid Receptor-Gi Complex Structures. Cell 2020, 180, 655–665.e18. [Google Scholar] [CrossRef] [PubMed]
  57. Nojima, S.; Fujita, Y.; Kimura, K.T.; Nomura, N.; Suno, R.; Morimoto, K.; Yamamoto, M.; Noda, T.; Iwata, S.; Shigematsu, H.; et al. Cryo-EM Structure of the Prostaglandin E Receptor EP4 Coupled to G Protein. Structure 2021, 29, 252–260.e6. [Google Scholar] [CrossRef]
  58. Toyoda, Y.; Morimoto, K.; Suno, R.; Horita, S.; Yamashita, K.; Hirata, K.; Sekiguchi, Y.; Yasuda, S.; Shiroishi, M.; Shimizu, T.; et al. Ligand binding to human prostaglandin E receptor EP4 at the lipid-bilayer interface. Nat. Chem. Biol. 2019, 15, 18–26. [Google Scholar] [CrossRef] [PubMed]
  59. Deluigi, M.; Klipp, A.; Klenk, C.; Merklinger, L.; Eberle, S.A.; Morstein, L.; Heine, P.; Mittl, P.R.E.; Ernst, P.; Kamenecka, T.M.; et al. Complexes of the neurotensin receptor 1 with small-molecule ligands reveal structural determinants of full, partial, and inverse agonism. Sci. Adv. 2021, 7, eabe5504. [Google Scholar] [CrossRef]
  60. Fenalti, G.; Giguere, P.M.; Katritch, V.; Huang, X.P.; Thompson, A.A.; Cherezov, V.; Roth, B.L.; Stevens, R.C. Molecular control of delta-opioid receptor signalling. Nature 2014, 506, 191–196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Claff, T.; Yu, J.; Blais, V.; Patel, N.; Martin, C.; Wu, L.; Han, G.W.; Holleran, B.J.; Van der Poorten, O.; White, K.L.; et al. Elucidating the active delta-opioid receptor crystal structure with peptide and small-molecule agonists. Sci. Adv. 2019, 5, eaax9115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Cheng, R.K.; Segala, E.; Robertson, N.; Deflorian, F.; Doré, A.S.; Errey, J.C.; Fiez-Vandal, C.; Marshall, F.H.; Cooke, R.M. Structures of Human A 1 and A 2A Adenosine Receptors with Xanthines Reveal Determinants of Selectivity. Structure 2017, 25, 1275–1285.e4. [Google Scholar] [CrossRef] [PubMed]
  63. Draper-Joyce, C.J.; Khoshouei, M.; Thal, D.M.; Liang, Y.-L.; Nguyen, A.T.N.; Furness, S.G.B.; Venugopal, H.; Baltos, J.-A.; Plitzko, J.M.; Danev, R.; et al. Structure of the adenosine-bound human adenosine A1 receptor–Gi complex. Nat. Cell Biol. 2018, 558, 559–563. [Google Scholar] [CrossRef]
  64. Zhang, K.; Zhang, J.; Gao, Z.-G.; Zhang, D.; Zhu, L.; Han, G.W.; Moss, S.M.; Paoletta, S.; Kiselev, E.; Lu, W.; et al. Structure of the human P2Y12 receptor in complex with an antithrombotic drug. Nat. Cell Biol. 2014, 509, 115–118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Zhang, J.; Zhang, K.; Gao, Z.-G.; Paoletta, S.; Zhang, D.; Han, G.W.; Li, T.; Ma, L.; Zhang, W.; Müller, C.E.; et al. Agonist-bound structure of the human P2Y12 receptor. Nat. Cell Biol. 2014, 509, 119–122. [Google Scholar] [CrossRef]
  66. Su, M.; Zhu, L.; Zhang, Y.; Paknejad, N.; Dey, R.; Huang, J.; Lee, M.Y.; Williams, D.; Jordan, K.D.; Eng, E.T.; et al. Structural Basis of the Activation of Heterotrimeric Gs-Protein by Isoproterenol-Bound beta1-Adrenergic Receptor. Mol. Cell 2020, 80, 59–71 e54. [Google Scholar] [CrossRef]
  67. Warne, T.; Moukhametzianov, R.; Baker, J.G.; Nehme, R.; Edwards, P.C.; Leslie, A.G.; Schertler, G.F.; Tate, C.G. The structural basis for agonist and partial agonist action on a beta(1)-adrenergic receptor. Nature 2011, 469, 241–244. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Zhang, Y.; Yang, F.; Ling, S.; Lv, P.; Zhou, Y.; Fang, W.; Sun, W.; Zhang, L.; Shi, P.; Tian, C. Single-particle cryo-EM structural studies of the beta2AR-Gs complex bound with a full agonist formoterol. Cell Discov. 2020, 6, 45. [Google Scholar] [CrossRef]
  69. Yang, F.; Ling, S.; Zhou, Y.; Zhang, Y.; Lv, P.; Liu, S.; Fang, W.; Sun, W.; Hu, A.L.; Zhang, L.; et al. Different conformational responses of the β2-adrenergic receptor-Gs complex upon binding of the partial agonist salbutamol or the full agonist isoprenaline. Natl. Sci. Rev. 2020, nwaa284. [Google Scholar] [CrossRef]
  70. Masureel, M.; Zou, Y.; Picard, L.P.; van der Westhuizen, E.; Mahoney, J.P.; Rodrigues, J.; Mildorf, T.J.; Dror, R.O.; Shaw, D.E.; Bouvier, M.; et al. Structural insights into binding specificity, efficacy and bias of a beta2AR partial agonist. Nat. Chem. Biol. 2018, 14, 1059–1066. [Google Scholar] [CrossRef]
  71. Warne, T.; Serrano-Vega, M.J.; Baker, J.G.; Moukhametzianov, R.; Edwards, P.C.; Henderson, R.; Leslie, A.G.; Tate, C.G.; Schertler, G.F. Structure of a beta1-adrenergic G-protein-coupled receptor. Nature 2008, 454, 486–491. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Christopher, J.A.; Brown, J.; Dore, A.S.; Errey, J.C.; Koglin, M.; Marshall, F.H.; Myszka, D.G.; Rich, R.L.; Tate, C.G.; Tehan, B.; et al. Biophysical fragment screening of the beta1-adrenergic receptor: Identification of high affinity arylpiperazine leads using structure-based drug design. J. Med. Chem. 2013, 56, 3446–3455. [Google Scholar] [CrossRef] [PubMed]
  73. Hanson, M.A.; Cherezov, V.; Griffith, M.T.; Roth, C.B.; Jaakola, V.-P.; Chien, E.Y.; Velasquez, J.; Kuhn, P.; Stevens, R.C. A Specific Cholesterol Binding Site Is Established by the 2.8 Å Structure of the Human β2-Adrenergic Receptor. Structure 2008, 16, 897–905. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Wacker, D.; Fenalti, G.; Brown, M.A.; Katritch, V.; Abagyan, R.; Cherezov, V.; Stevens, R.C. Conserved binding mode of human beta2 adrenergic receptor inverse agonists and antagonist revealed by X-ray crystallography. J. Am. Chem. Soc. 2010, 132, 11443–11445. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Ishchenko, A.; Stauch, B.; Han, G.W.; Batyuk, A.; Shiriaeva, A.; Li, C.; Zatsepin, N.; Weierstall, U.; Liu, W.; Nango, E.; et al. Toward G protein-coupled receptor structure-based drug design using X-ray lasers. IUCrJ 2019, 6, 1106–1119. [Google Scholar] [CrossRef] [PubMed]
  76. Moukhametzianov, R.; Warne, T.; Edwards, P.C.; Serrano-Vega, M.J.; Leslie, A.G.; Tate, C.G.; Schertler, G.F. Two distinct conformations of helix 6 observed in antagonist-bound structures of a beta1-adrenergic receptor. Proc. Natl. Acad. Sci. USA 2011, 108, 8228–8232. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Sato, T.; Baker, J.; Warne, T.; Brown, G.A.; Leslie, A.G.; Congreve, M.; Tate, C.G. Pharmacological Analysis and Structure Determination of 7-Methylcyanopindolol-Bound beta1-Adrenergic Receptor. Mol. Pharmacol. 2015, 88, 1024–1034. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Whalen, E.J.; Rajagopal, S.; Lefkowitz, R.J. Therapeutic potential of beta-arrestin- and G protein-biased agonists. Trends Mol. Med. 2011, 17, 126–139. [Google Scholar] [CrossRef] [Green Version]
  79. Mangmool, S.; Parichatikanond, W.; Kurose, H. Therapeutic Targets for Treatment of Heart Failure: Focus on GRKs and beta-Arrestins Affecting betaAR Signaling. Front. Pharmacol. 2018, 9, 1336. [Google Scholar] [CrossRef]
  80. Kang, Y.; Zhou, X.E.; Gao, X.; He, Y.; Liu, W.; Ishchenko, A.; Barty, A.; White, T.A.; Yefanov, O.; Han, G.W.; et al. Crystal structure of rhodopsin bound to arrestin by femtosecond X-ray laser. Nat. Cell Biol. 2015, 523, 561–567. [Google Scholar] [CrossRef] [Green Version]
  81. Yin, W.; Li, Z.; Jin, M.; Yin, Y.-L.; De Waal, P.W.; Pal, K.; Yin, Y.; Gao, X.; He, Y.; Gao, J.; et al. A complex structure of arrestin-2 bound to a G protein-coupled receptor. Cell Res. 2019, 29, 971–983. [Google Scholar] [CrossRef] [PubMed]
  82. Huang, W.; Masureel, M.; Qu, Q.; Janetzko, J.; Inoue, A.; Kato, H.E.; Robertson, M.J.; Nguyen, K.C.; Glenn, J.S.; Skiniotis, G.; et al. Structure of the neurotensin receptor 1 in complex with beta-arrestin 1. Nature 2020, 579, 303–308. [Google Scholar] [CrossRef] [PubMed]
  83. Staus, D.P.; Hu, H.; Robertson, M.J.; Kleinhenz, A.L.W.; Wingler, L.M.; Capel, W.D.; Latorraca, N.R.; Lefkowitz, R.J.; Skiniotis, G. Structure of the M2 muscarinic receptor-beta-arrestin complex in a lipid nanodisc. Nature 2020, 579, 297–302. [Google Scholar] [CrossRef] [PubMed]
  84. Liu, J.J.; Horst, R.; Katritch, V.; Stevens, R.C.; Wuthrich, K. Biased signaling pathways in beta2-adrenergic receptor characterized by 19F-NMR. Science 2012, 335, 1106–1110. [Google Scholar] [CrossRef] [Green Version]
  85. Wisler, J.W.; DeWire, S.M.; Whalen, E.J.; Violin, J.D.; Drake, M.T.; Ahn, S.; Shenoy, S.K.; Lefkowitz, R.J. A unique mechanism of beta-blocker action: Carvedilol stimulates beta-arrestin signaling. Proc. Natl. Acad. Sci. USA 2007, 104, 16657–16662. [Google Scholar] [CrossRef] [Green Version]
  86. Galandrin, S.; Oligny-Longpre, G.; Bonin, H.; Ogawa, K.; Gales, C.; Bouvier, M. Conformational rearrangements and signaling cascades involved in ligand-biased mitogen-activated protein kinase signaling through the beta1-adrenergic receptor. Mol. Pharmacol. 2008, 74, 162–172. [Google Scholar] [CrossRef]
  87. Rajagopal, S.; Ahn, S.; Rominger, D.H.; Gowen-MacDonald, W.; Lam, C.M.; DeWire, S.M.; Violin, J.D.; Lefkowitz, R.J. Quantifying Ligand Bias at Seven-Transmembrane Receptors. Mol. Pharmacol. 2011, 80, 367–377. [Google Scholar] [CrossRef] [Green Version]
  88. Weiss, D.R.; Ahn, S.; Sassano, M.F.; Kleist, A.; Zhu, X.; Strachan, R.; Roth, B.L.; Lefkowitz, R.J.; Shoichet, B.K. Conformation guides molecular efficacy in docking screens of activated beta-2 adrenergic G protein coupled receptor. ACS Chem. Biol. 2013, 8, 1018–1026. [Google Scholar] [CrossRef]
  89. Warne, T.; Edwards, P.C.; Leslie, A.G.; Tate, C.G. Crystal structures of a stabilized beta1-adrenoceptor bound to the biased agonists bucindolol and carvedilol. Structure 2012, 20, 841–849. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Rasmussen, S.G.; Choi, H.J.; Fung, J.J.; Pardon, E.; Casarosa, P.; Chae, P.S.; Devree, B.T.; Rosenbaum, D.M.; Thian, F.S.; Kobilka, T.S.; et al. Structure of a nanobody-stabilized active state of the beta(2) adrenoceptor. Nature 2011, 469, 175–180. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Drake, M.T.; Violin, J.D.; Whalen, E.J.; Wisler, J.W.; Shenoy, S.K.; Lefkowitz, R.J. beta-arrestin-biased agonism at the beta2-adrenergic receptor. J. Biol. Chem. 2008, 283, 5669–5676. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Lukasheva, V.; Devost, D.; Le Gouill, C.; Namkung, Y.; Martin, R.D.; Longpre, J.M.; Amraei, M.; Shinjo, Y.; Hogue, M.; Lagace, M.; et al. Signal profiling of the beta1AR reveals coupling to novel signalling pathways and distinct phenotypic responses mediated by beta1AR and beta2AR. Sci. Rep. 2020, 10, 8779. [Google Scholar] [CrossRef]
  93. Manglik, A.; Kim, T.H.; Masureel, M.; Altenbach, C.; Yang, Z.; Hilger, D.; Lerch, M.T.; Kobilka, T.S.; Thian, F.S.; Hubbell, W.L.; et al. Structural Insights into the Dynamic Process of beta2-Adrenergic Receptor Signaling. Cell 2015, 161, 1101–1111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Imai, S.; Yokomizo, T.; Kofuku, Y.; Shiraishi, Y.; Ueda, T.; Shimada, I. Structural equilibrium underlying ligand-dependent activation of beta2-adrenoreceptor. Nat. Chem. Biol. 2020, 16, 430–439. [Google Scholar] [CrossRef] [PubMed]
  95. Lerch, M.T.; Matt, R.A.; Masureel, M.; Elgeti, M.; Kumar, K.K.; Hilger, D.; Foys, B.; Kobilka, B.K.; Hubbell, W.L. Viewing rare conformations of the beta2 adrenergic receptor with pressure-resolved DEER spectroscopy. Proc. Natl. Acad. Sci. USA 2020, 117, 31824–31831. [Google Scholar] [CrossRef]
  96. Gregorio, G.G.; Masureel, M.; Hilger, D.; Terry, D.S.; Juette, M.; Zhao, H.; Zhou, Z.; Perez-Aguilar, J.M.; Hauge, M.; Mathiasen, S.; et al. Single-molecule analysis of ligand efficacy in beta2AR-G-protein activation. Nature 2017, 547, 68–73. [Google Scholar] [CrossRef]
  97. Kofuku, Y.; Ueda, T.; Okude, J.; Shiraishi, Y.; Kondo, K.; Maeda, M.; Tsujishita, H.; Shimada, I. Efficacy of the beta(2)-adrenergic receptor is determined by conformational equilibrium in the transmembrane region. Nat. Commun. 2012, 3, 1045. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Qu, L.; Zhou, Q.; Xu, Y.; Guo, Y.; Chen, X.; Yao, D.; Han, G.W.; Liu, Z.-J.; Stevens, R.C.; Zhong, G.; et al. Structural Basis of the Diversity of Adrenergic Receptors. Cell Rep. 2019, 29, 2929–2935.e4. [Google Scholar] [CrossRef] [Green Version]
  99. Chen, X.; Xu, Y.; Qu, L.; Wu, L.; Han, G.W.; Guo, Y.; Wu, Y.; Zhou, Q.; Sun, Q.; Chu, C.; et al. Molecular Mechanism for Ligand Recognition and Subtype Selectivity of α2C Adrenergic Receptor. Cell Rep. 2019, 29, 2936–2943.e4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Ruffolo, R.R.; Bondinell, W.; Hieble, J.P. .alpha.- and .beta.-Adrenoceptors: From the Gene to the Clinic. 2. Structure-Activity Relationships and Therapeutic Applications. J. Med. Chem. 1995, 38, 3681–3716. [Google Scholar] [CrossRef]
  101. Ruffolo, R.R., Jr.; Spradlin, T.A.; Pollock, G.D.; Waddell, J.E.; Murphy, P.J. Alpha and beta adrenergic effects of the stereoi-somers of dobutamine. J. Pharmacol. Exp. Ther. 1981, 219, 447–452. [Google Scholar] [PubMed]
  102. Pauwels, P.J.; Colpaert, F.C. Disparate ligand-mediated Ca(2+) responses by wild-type, mutant Ser(200)Ala and Ser(204)Ala alpha(2A)-adrenoceptor: G(alpha15) fusion proteins: Evidence for multiple ligand-activation binding sites. Br. J. Pharmacol. 2000, 130, 1505–1512. [Google Scholar] [CrossRef] [Green Version]
  103. Sallinen, J.; Hoglund, I.; Engstrom, M.; Lehtimaki, J.; Virtanen, R.; Sirvio, J.; Wurster, S.; Savola, J.M.; Haapalinna, A. Phar-macological characterization and CNS effects of a novel highly selective alpha2C-adrenoceptor antagonist JP-1302. Br. J. Pharmacol. 2007, 150, 391–402. [Google Scholar] [CrossRef] [Green Version]
  104. Sun, B.; Lockyer, S.; Li, J.; Chen, R.; Yoshitake, M.; Kambayashi, J.I. OPC-28326, a selective femoral vasodilator, is an al-pha2C-adrenoceptor-selective antagonist. J. Pharmacol. Exp. Ther. 2001, 299, 652–658. [Google Scholar]
  105. Wootten, D.; Christopoulos, A.; Sexton, P.M. Emerging paradigms in GPCR allostery: Implications for drug discovery. Nat. Rev. Drug Discov. 2013, 12, 630–644. [Google Scholar] [CrossRef]
  106. Thal, D.M.; Glukhova, A.; Sexton, P.M.; Christopoulos, A. Structural insights into G-protein-coupled receptor allostery. Nat. Cell Biol. 2018, 559, 45–53. [Google Scholar] [CrossRef]
  107. Wu, Y.; Tong, J.; Ding, K.; Zhou, Q.; Zhao, S. GPCR Allosteric Modulator Discovery. Adv. Exp. Med. Biol. 2019, 1163, 225–251. [Google Scholar] [CrossRef]
  108. Sarkar, P.; Chattopadhyay, A. Cholesterol interaction motifs in G protein-coupled receptors: Slippery hot spots? Wiley Interdiscip. Rev. Syst. Biol. Med. 2020, 12, e1481. [Google Scholar] [CrossRef]
  109. Oswald, C.; Rappas, M.; Kean, J.; Doré, A.S.; Errey, J.C.; Bennett, K.; Deflorian, F.; Christopher, J.A.; Jazayeri, A.; Mason, J.S.; et al. Intracellular allosteric antagonism of the CCR9 receptor. Nat. Cell Biol. 2016, 540, 462–465. [Google Scholar] [CrossRef]
  110. Liu, X.; Masoudi, A.; Kahsai, A.W.; Huang, L.Y.; Pani, B.; Staus, D.P.; Shim, P.J.; Hirata, K.; Simhal, R.K.; Schwalb, A.M.; et al. Mechanism of beta2AR regulation by an intracellular positive allosteric modulator. Science 2019, 364, 1283–1287. [Google Scholar] [CrossRef] [PubMed]
  111. Liu, X.; Kaindl, J.; Korczynska, M.; Stossel, A.; Dengler, D.; Stanek, M.; Hubner, H.; Clark, M.J.; Mahoney, J.; Matt, R.A.; et al. An allosteric modulator binds to a conformational hub in the beta2 adrenergic receptor. Nat. Chem. Biol. 2020, 16, 749–755. [Google Scholar] [CrossRef]
  112. Lu, J.; Byrne, N.; Wang, J.; Bricogne, G.; Brown, F.K.; Chobanian, H.R.; Colletti, S.L.; Di Salvo, J.; Thomas-Fowlkes, B.; Guo, Y.; et al. Structural basis for the cooperative allosteric activation of the free fatty acid receptor GPR40. Nat. Struct. Mol. Biol. 2017, 24, 570–577. [Google Scholar] [CrossRef] [PubMed]
  113. Ho, J.D.; Chau, B.; Rodgers, L.; Lu, F.; Wilbur, K.L.; Otto, K.A.; Chen, Y.; Song, M.; Riley, J.P.; Yang, H.-C.; et al. Structural basis for GPR40 allosteric agonism and incretin stimulation. Nat. Commun. 2018, 9, 1–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Zhuang, Y.; Krumm, B.; Zhang, H.; Zhou, X.E.; Wang, Y.; Huang, X.P.; Liu, Y.; Cheng, X.; Jiang, Y.; Jiang, H.; et al. Mecha-nism of dopamine binding and allosteric modulation of the human D1 dopamine receptor. Cell Res. 2021, 31, 593–596. [Google Scholar] [CrossRef] [PubMed]
  115. Robertson, N.; Rappas, M.; Doré, A.S.; Brown, J.; Bottegoni, G.; Koglin, M.; Cansfield, J.; Jazayeri, A.; Cooke, R.M.; Marshall, F.H. Structure of the complement C5a receptor bound to the extra-helical antagonist NDT9513727. Nat. Cell Biol. 2018, 553, 111–114. [Google Scholar] [CrossRef] [PubMed]
  116. Liu, H.; Kim, H.R.; Deepak, R.N.V.K.; Wang, L.; Chung, K.Y.; Fan, H.; Wei, Z.; Zhang, C. Orthosteric and allosteric action of the C5a receptor antagonists. Nat. Struct. Mol. Biol. 2018, 25, 472–481. [Google Scholar] [CrossRef]
  117. Zheng, Y.; Qin, L.; Zacarías, N.V.O.; De Vries, H.; Han, G.W.; Gustavsson, M.; Dabros, M.; Zhao, C.; Cherney, R.J.; Carter, P.; et al. Structure of CC chemokine receptor 2 with orthosteric and allosteric antagonists. Nat. Cell Biol. 2016, 540, 458–461. [Google Scholar] [CrossRef] [Green Version]
  118. Jaeger, K.; Brünle, S.; Weinert, T.; Guba, W.; Muehle, J.; Miyazaki, T.; Weber, M.; Furrer, A.; Haenggi, N.; Tetaz, T.; et al. Structural Basis for Allosteric Ligand Recognition in the Human CC Chemokine Receptor 7. Cell 2019, 178, 1222–1230.e10. [Google Scholar] [CrossRef] [Green Version]
  119. Liu, K.; Wu, L.; Yuan, S.; Wu, M.; Xu, Y.; Sun, Q.; Li, S.; Zhao, S.; Hua, T.; Liu, Z.-J. Structural basis of CXC chemokine receptor 2 activation and signalling. Nat. Cell Biol. 2020, 585, 1–9. [Google Scholar] [CrossRef]
  120. Liu, X.; Ahn, S.; Kahsai, A.W.; Meng, K.C.; Latorraca, N.R.; Pani, B.; Venkatakrishnan, A.J.; Masoudi, A.; Weis, W.I.; Dror, R.O.; et al. Mechanism of intracellular allosteric beta2AR antagonist revealed by X-ray crystal structure. Nature 2017, 548, 480–484. [Google Scholar] [CrossRef]
  121. Ahn, S.; Kahsai, A.W.; Pani, B.; Wang, Q.T.; Zhao, S.; Wall, A.L.; Strachan, R.T.; Staus, D.P.; Wingler, L.M.; Sun, L.D.; et al. Al-losteric “beta-blocker” isolated from a DNA-encoded small molecule library. Proc. Natl. Acad. Sci. USA 2017, 114, 1708–1713. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Shoichet, B.K.; Kobilka, B.K. Structure-based drug screening for G-protein-coupled receptors. Trends Pharmacol. Sci. 2012, 33, 268–272. [Google Scholar] [CrossRef] [Green Version]
  123. Yang, D.; Zhou, Q.; Labroska, V.; Qin, S.; Darbalaei, S.; Wu, Y.; Yuliantie, E.; Xie, L.; Tao, H.; Cheng, J.; et al. G pro-tein-coupled receptors: Structure- and function-based drug discovery. Signal Transduct. Target. Ther. 2021, 6, 7. [Google Scholar] [CrossRef] [PubMed]
  124. Topiol, S.; Sabio, M. Use of the X-ray structure of the Beta2-adrenergic receptor for drug discovery. Bioorg. Med. Chem. Lett. 2008, 18, 1598–1602. [Google Scholar] [CrossRef]
  125. Irwin, J.; Shoichet, B.K. ZINC—A Free Database of Commercially Available Compounds for Virtual Screening. J. Chem. Inf. Model. 2005, 45, 177–182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Kolb, P.; Rosenbaum, D.M.; Irwin, J.J.; Fung, J.J.; Kobilka, B.K.; Shoichet, B.K. Structure-based discovery of beta2-adrenergic receptor ligands. Proc. Natl. Acad. Sci. USA 2009, 106, 6843–6848. [Google Scholar] [CrossRef] [Green Version]
  127. Rosenbaum, D.M.; Zhang, C.; Lyons, J.A.; Holl, R.; Aragao, D.; Arlow, D.H.; Rasmussen, S.G.; Choi, H.J.; Devree, B.T.; Suna-hara, R.K.; et al. Structure and function of an irreversible agonist-beta(2) adrenoceptor complex. Nature 2011, 469, 236–240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Weichert, D.; Kruse, A.C.; Manglik, A.; Hiller, C.; Zhang, C.; Hübner, H.; Kobilka, B.K.; Gmeiner, P. Covalent agonists for studying G protein-coupled receptor activation. Proc. Natl. Acad. Sci. USA 2014, 111, 10744–10748. [Google Scholar] [CrossRef] [Green Version]
  129. McCorvy, J.D.; Butler, K.V.; Kelly, B.; Rechsteiner, K.; Karpiak, J.; Betz, R.M.; Kormos, B.L.; Shoichet, B.K.; Dror, R.O.; Jin, J.; et al. Structure-inspired design of beta-arrestin-biased ligands for aminergic GPCRs. Nat. Chem. Biol. 2018, 14, 126–134. [Google Scholar] [CrossRef] [PubMed]
  130. de Graaf, C.; Rognan, D. Selective structure-based virtual screening for full and partial agonists of the beta2 adrenergic re-ceptor. J. Med. Chem. 2008, 51, 4978–4985. [Google Scholar] [CrossRef] [PubMed]
  131. Kooistra, A.J.; Leurs, R.; de Esch, I.J.; de Graaf, C. Structure-Based Prediction of G-Protein-Coupled Receptor Ligand Func-tion: A beta-Adrenoceptor Case Study. J. Chem. Inf. Model. 2015, 55, 1045–1061. [Google Scholar] [CrossRef]
  132. Kooistra, A.J.; Vischer, H.F.; McNaught-Flores, D.; Leurs, R.; de Esch, I.; De Graaf, C. Function-specific virtual screening for GPCR ligands using a combined scoring method. Sci. Rep. 2016, 6, 28288. [Google Scholar] [CrossRef] [Green Version]
  133. Zhuang, Y.; Xu, P.; Mao, C.; Wang, L.; Krumm, B.; Zhou, X.E.; Huang, S.; Liu, H.; Cheng, X.; Huang, X.P.; et al. Structural in-sights into the human D1 and D2 dopamine receptor signaling complexes. Cell 2021, 184, 931–942 e918. [Google Scholar] [CrossRef]
  134. Xiao, P.; Yan, W.; Gou, L.; Zhong, Y.-N.; Kong, L.; Wu, C.; Wen, X.; Yuan, Y.; Cao, S.; Qu, C.; et al. Ligand recognition and allosteric regulation of DRD1-Gs signaling complexes. Cell 2021, 184, 943–956.e18. [Google Scholar] [CrossRef]
  135. Sun, B.; Feng, D.; Chu, M.L.; Fish, I.; Kelm, S.; Lebon, F.; Lovera, S.; Valade, A.; Wood, M.; Ceska, T.; et al. PDB 7JOZ. Available online: https://www.rcsb.org/structure/7JOZ (accessed on 24 May 2021).
Figure 1. Adrenergic receptors as drug targets. (a) Binding affinities [12,13,14] of endogenous agonists in the nine adrenergic receptors. (b) Number of new drugs (approved or first marketing) for the nine adrenergic receptors in each decade. Data retrieved from Guide to PHARMACOLOGY [15] (www.guidetopharmacology.org, accessed on 25 April 2021), GPCRdb [16] (gpcrdb.org, accessed on 25 April 2021), and DrugBank [17] (go.drugbank.com, accessed on 25 April 2021).
Figure 1. Adrenergic receptors as drug targets. (a) Binding affinities [12,13,14] of endogenous agonists in the nine adrenergic receptors. (b) Number of new drugs (approved or first marketing) for the nine adrenergic receptors in each decade. Data retrieved from Guide to PHARMACOLOGY [15] (www.guidetopharmacology.org, accessed on 25 April 2021), GPCRdb [16] (gpcrdb.org, accessed on 25 April 2021), and DrugBank [17] (go.drugbank.com, accessed on 25 April 2021).
Biomolecules 11 00936 g001
Figure 2. Mechanisms of ligand recognition by adrenergic receptors. (a) Key residues for protein structure maintaining, receptor activation, and endogenous ligand recognition, marked with color on snake plot of β2AR. x.50 are Ballesteros–Weinstein numbering (BWN) [22] of the most conserved residues in each segment. Sequence of N-terminus, ICL3, and C-terminus are not shown because these regions do not contain key residues and are very long. (b) Chemical structures of monoamine neurotransmitters (shown in protonated state) and acetylcholine. Moieties colored in consistent with interacting residues in panel (a).
Figure 2. Mechanisms of ligand recognition by adrenergic receptors. (a) Key residues for protein structure maintaining, receptor activation, and endogenous ligand recognition, marked with color on snake plot of β2AR. x.50 are Ballesteros–Weinstein numbering (BWN) [22] of the most conserved residues in each segment. Sequence of N-terminus, ICL3, and C-terminus are not shown because these regions do not contain key residues and are very long. (b) Chemical structures of monoamine neurotransmitters (shown in protonated state) and acetylcholine. Moieties colored in consistent with interacting residues in panel (a).
Biomolecules 11 00936 g002
Figure 3. Comparison of adrenergic receptor major types and subtypes. (a) Sequence identity matrices of adrenergic receptors at full-length sequence (based on multiple sequence alignment generated with MAFFT v7.299b [30]) or pocket residues (33 residues: 2.61,64–65; 3.28–29,32–33,36–37; 45.50–54; 5.38–39,42–43,46; 6.48,51–52,55,58–59,62; 7.30,32,35–36,39–40,43). ‘Pocket’ is the cavity at the extracellular side of a GPCR. (b) Key interactions of endogenous ligand epinephrine in orthosteric site of β2AR/epinephrine (PDB: 4LDO [31], coloring scheme the same to Figure 2a). (c) Alignment of key residues for moiety recognition or type selectivity (residue BWN colored in the same scheme to Figure 2a and Figure 3b) in all the adrenergic receptors members (generated from GPCRdb plot [16], residues colored for property: red, negatively charged; blue, positively charged; yellow, hydrophobic; blue, aromatic; magenta, hydrophilic; pink, glycine; gray, proline).
Figure 3. Comparison of adrenergic receptor major types and subtypes. (a) Sequence identity matrices of adrenergic receptors at full-length sequence (based on multiple sequence alignment generated with MAFFT v7.299b [30]) or pocket residues (33 residues: 2.61,64–65; 3.28–29,32–33,36–37; 45.50–54; 5.38–39,42–43,46; 6.48,51–52,55,58–59,62; 7.30,32,35–36,39–40,43). ‘Pocket’ is the cavity at the extracellular side of a GPCR. (b) Key interactions of endogenous ligand epinephrine in orthosteric site of β2AR/epinephrine (PDB: 4LDO [31], coloring scheme the same to Figure 2a). (c) Alignment of key residues for moiety recognition or type selectivity (residue BWN colored in the same scheme to Figure 2a and Figure 3b) in all the adrenergic receptors members (generated from GPCRdb plot [16], residues colored for property: red, negatively charged; blue, positively charged; yellow, hydrophobic; blue, aromatic; magenta, hydrophilic; pink, glycine; gray, proline).
Biomolecules 11 00936 g003
Figure 4. Mechanism of agonism in β adrenergic receptors. (a) Structures of β2AR in active/inactive states (PDB IDs: 3SN6 [8], 2RH1 [6]). (b) Interactions of agonist/antagonist to S5.42/S5.46 and microswitch of P5.50-I3.40-F6.44 motif. (c) Binding poses of agonists/antagonists in β1AR/β2AR [6,8,11,31,43,54,66,67,68,69,70,71,72,73,74,75,76,77]. Structures were aligned based on Cα atoms of 7 residues: 3.32; 5.42,46; 6.51–52; 7.40,43. (d) Binding poses of isoproterenol in β1 active/inactive states (PDB IDs: 7JJO [66], 2Y03 [67]).
Figure 4. Mechanism of agonism in β adrenergic receptors. (a) Structures of β2AR in active/inactive states (PDB IDs: 3SN6 [8], 2RH1 [6]). (b) Interactions of agonist/antagonist to S5.42/S5.46 and microswitch of P5.50-I3.40-F6.44 motif. (c) Binding poses of agonists/antagonists in β1AR/β2AR [6,8,11,31,43,54,66,67,68,69,70,71,72,73,74,75,76,77]. Structures were aligned based on Cα atoms of 7 residues: 3.32; 5.42,46; 6.51–52; 7.40,43. (d) Binding poses of isoproterenol in β1 active/inactive states (PDB IDs: 7JJO [66], 2Y03 [67]).
Biomolecules 11 00936 g004
Figure 5. Mechanisms of effects of ligands on β adrenergic receptors. (a) Structures of β1AR coupled to Gs or arrestin (PDB IDs: 7JJO [66], 6TKO [11]) (b) Sketch map of key ligand interactions in βARs. Formoterol in β2AR was used for the drawing.
Figure 5. Mechanisms of effects of ligands on β adrenergic receptors. (a) Structures of β1AR coupled to Gs or arrestin (PDB IDs: 7JJO [66], 6TKO [11]) (b) Sketch map of key ligand interactions in βARs. Formoterol in β2AR was used for the drawing.
Biomolecules 11 00936 g005
Figure 6. Mechanisms in α2 adrenergic receptors. (a) Agonist recognition (dexmedetomidine in α2BAR, PDB ID: 6K41 [52]). (b) Antagonist recognition (RS 79948 in α2CAR, PDB ID: 6KUW [99]). (c) Comparison of agonist recognition in α2ARs and βARs (isoproterenol in β1AR, PDB ID: 6H7J [54]). (d) Conformational changes in active/inactive α2ARs. (e) Binding poses of agonist clonidine and antagonist torazoline, predicted by molecular docking (Glide program in Schrodinger platform 2019-2 [38]).
Figure 6. Mechanisms in α2 adrenergic receptors. (a) Agonist recognition (dexmedetomidine in α2BAR, PDB ID: 6K41 [52]). (b) Antagonist recognition (RS 79948 in α2CAR, PDB ID: 6KUW [99]). (c) Comparison of agonist recognition in α2ARs and βARs (isoproterenol in β1AR, PDB ID: 6H7J [54]). (d) Conformational changes in active/inactive α2ARs. (e) Binding poses of agonist clonidine and antagonist torazoline, predicted by molecular docking (Glide program in Schrodinger platform 2019-2 [38]).
Biomolecules 11 00936 g006
Figure 7. Allosteric modulating of β2AR. (a) Positions of allosteric sites (including cholesterol binding sites) identified in structures. The orthosteric site is marked by endogenous agonist epinephrine (EPI.). Positions of cholesterol are from a crystal structure in an inactive state (PDB ID: 2RH1 [6]). (b) PAM Cmpd-6FA and NAM AS408 in lipid interface at TM3/4/5 (PDB IDs: 6N48 [106] and 6OBA [107]). (c) Cytoplasmic NAM Cmpd-15PA compared to active state β2AR (PDB IDs: 5X7D [109] and 3SN6 [8]).
Figure 7. Allosteric modulating of β2AR. (a) Positions of allosteric sites (including cholesterol binding sites) identified in structures. The orthosteric site is marked by endogenous agonist epinephrine (EPI.). Positions of cholesterol are from a crystal structure in an inactive state (PDB ID: 2RH1 [6]). (b) PAM Cmpd-6FA and NAM AS408 in lipid interface at TM3/4/5 (PDB IDs: 6N48 [106] and 6OBA [107]). (c) Cytoplasmic NAM Cmpd-15PA compared to active state β2AR (PDB IDs: 5X7D [109] and 3SN6 [8]).
Biomolecules 11 00936 g007
Table 1. Major types of adrenergic receptors.
Table 1. Major types of adrenergic receptors.
Major TypePrimary PathwayMain EffectsIndications of AgonistsIndications of Antagonists
α1Gq/11smooth muscle contractionvasodilatory shock, hypotension [46]hypertension, benign prostatic hyperplasia [47]
α2Gi/oinhibition of norepinephrine releasehypertension,
pain and panic disorders [48]
erectile dysfunction [49], depression [50]
βGsβ1: cardiac stimulation
β2: bronchodilation
β3: increased lipolysis
cardiogenic shock, heart failure, asthma, overactive bladder [4]heart failure, arrhythmias, hypertension [51]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wu, Y.; Zeng, L.; Zhao, S. Ligands of Adrenergic Receptors: A Structural Point of View. Biomolecules 2021, 11, 936. https://doi.org/10.3390/biom11070936

AMA Style

Wu Y, Zeng L, Zhao S. Ligands of Adrenergic Receptors: A Structural Point of View. Biomolecules. 2021; 11(7):936. https://doi.org/10.3390/biom11070936

Chicago/Turabian Style

Wu, Yiran, Liting Zeng, and Suwen Zhao. 2021. "Ligands of Adrenergic Receptors: A Structural Point of View" Biomolecules 11, no. 7: 936. https://doi.org/10.3390/biom11070936

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop