Next Article in Journal
A Synergistic Effect between Plasma Dickkopf-1 and Obstructive Coronary Artery Disease on the Prediction of Major Adverse Cardiac Events in Patients with Angina: An Observational Study
Next Article in Special Issue
Structure/Function Studies on the Activation Motif of Two Non-Mammalian Mrap1 Orthologs, and Observations on the Phylogeny of Mrap1, Including a Novel Characterization of an Mrap1 from the Chondrostean Fish, Polyodon spathula
Previous Article in Journal
A Distinctive γδ T Cell Repertoire in NOD Mice Weakens Immune Regulation and Favors Diabetic Disease
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Ligands for Melanocortin Receptors: Beyond Melanocyte-Stimulating Hormones and Adrenocorticotropin

1
College of Animal Science and Technology, Anhui Agricultural University, Hefei 230061, China
2
Department of Anatomy, Physiology and Pharmacology, College of Veterinary Medicine, Auburn University, Auburn, AL 36849, USA
*
Author to whom correspondence should be addressed.
Submission received: 30 August 2022 / Revised: 25 September 2022 / Accepted: 28 September 2022 / Published: 1 October 2022

Abstract

:
The discovery of melanocortins in 1916 has resulted in more than 100 years of research focused on these peptides. Extensive studies have elucidated well-established functions of melanocortins mediated by cell surface receptors, including MSHR (melanocyte-stimulating hormone receptor) and ACTHR (adrenocorticotropin receptor). Subsequently, three additional melanocortin receptors (MCRs) were identified. Among these five MCRs, MC3R and MC4R are expressed primarily in the central nervous system, and are therefore referred to as the neural MCRs. Since the central melanocortin system plays important roles in regulating energy homeostasis, targeting neural MCRs is emerging as a therapeutic approach for treating metabolic conditions such as obesity and cachexia. Early efforts modifying endogenous ligands resulted in the development of many potent and selective ligands. This review focuses on the ligands for neural MCRs, including classical ligands (MSH and agouti-related peptide), nonclassical ligands (lipocalin 2, β-defensin, small molecules, and pharmacoperones), and clinically approved ligands (ACTH, setmelanotide, bremelanotide, and several repurposed drugs).

Graphical Abstract

1. Introduction

Melanocortin peptides, primarily consisting of α-, β-, and γ-melanocyte-stimulating hormones (α-, β-, γ-MSH) and adrenocorticotropin (ACTH), are elements of an ancient modulatory system [1]. Melanocortins are derived from tissue-specific posttranslational processing of the common precursor—proopiomelanocortin (POMC) [2]. The melanocortins are involved in pleiotropic physiological functions, including pigmentation, adrenal steroidogenesis, appetite regulation and energy homeostasis, energy metabolism, cardiovascular function, exocrine secretion, inflammation, immunomodulation, neuromuscular regeneration, sexual function, and temperature control [3,4,5,6,7,8]. Recent FDA approval of melanocortin-based drugs is expected to further promote the investigation of the melanocortin system and lead to new pharmacotherapies.
G protein-coupled receptors (GPCRs) have been tractable drug targets for decades, with over one-third of currently marketed drugs targeting GPCRs [9,10,11]. Of these, family A rhodopsin-like GPCRs are highly represented and continued drug discovery for this family of receptors may provide novel therapeutics for a wide range of diseases [12]. Melanocortin receptors (MCRs) are typical members of family A GPCRs. Extensive studies spanning several decades have elucidated the functions of melanocortins mediated by cell surface receptors named MSH receptor and ACTH receptor. In the early 1990s, with degenerate PCR cloning techniques, three additional related MCRs were cloned, named MC3R, MC4R, and MC5R, based on the chronological order of their characterization, with MSH receptor renamed as MC1R and ACTH receptor renamed as MC2R [1,3].
Among these receptors, MC3R and MC4R are expressed primarily in the central nervous system, and therefore referred to as the neural MCRs [13,14,15,16]. Analysis of mice lacking either Mc3r or Mc4r or both revealed that these two neural MCRs served important nonredundant roles in regulating energy homeostasis, with MC3R regulating feed efficiency and MC4R regulating both food intake and energy expenditure [17,18,19,20,21]. Targeting neural MCRs is emerging as a therapeutic approach for metabolic diseases and chronic inflammation, but until recently, drug development was limited by lack of selectivity and safety concerns [22,23,24].
Several endogenous ligands interact with the MCRs (Figure 1). Table 1 summarizes the pharmacological properties of these ligands at the human MCRs. Melanocortins have been studied for over a century, resulting in multiple drugs approved for treating several diseases [25,26,27]. This review focuses on the ligands for the two neural MCRs, including endogenous ligands (MSHs, agouti-related peptide (AgRP), and lipocalin 2), nonclassical ligands (β-defensin, small molecules, and pharmacoperones), and clinically approved ligands (ACTH, setmelanotide, and bremelanotide, and several repurposed drugs).

2. Endogenous Ligands

2.1. MSHs as Agonists

The endogenous agonists of neural MCRs are α-, β-, and γ-MSH, as well as ACTH, with the conserved tetrapeptide sequence His-Phe-Arg-Trp (HFRW) as the pharmacophore (Figure 1A), the minimally active sequence that possesses agonist activity in the classic frog and lizard skin bioassays [31,32]. Activation of the MCRs with MSH analogues indicated that the length of ACTH or MSH is different for a specific receptor subtype. Four amino acids (His-Phe-Arg-Trp) are the minimal MSH peptide for MC1R activation. The first 17 amino acids of ACTH are required for MC2R activation [33]. Four amino acids (His-Phe-Arg-Trp) in MSH are required for MC3R and MC5R activation [34]. Three amino acids (Phe-Arg-Trp) in MSH are required for MC4R activation [35]. Of the four major endogenous melanocortins, human MC1R and MC5R have the highest affinities for α-MSH, MC3R has the highest affinity for γ1-MSH, and MC4R has the highest affinity for β-MSH (Table 1) [28,29].
Pro-ACTH is cleaved by PC1 in the anterior pituitary corticotrophs to produce ACTH(1–39), which can be further processed through PC2 to produce ACTH(1–13)-NH2 and α-MSH [36]. ACTH is the only known endogenous ligand for MC2R, making it the sole agonist that stimulates all five MCRs. The full-length ACTH is 39 amino acids, of which ACTH(1–24) is hypothesized to be the functional domain. An anorectic effect was observed at 4 h after intracerebroventricular (ICV) administration of ACTH(1–24) at 0.05 μg/animal in mice and 10 μg/animal in rabbits [37]. A 4 μg/animal dose of ACTH(1–24) injected into the lateral ventricle or ventromedial hypothalamus also results in decreased food intake in rats [38].
The N-terminal 13 residues of ACTH are cleaved and modified through N-terminal acetylation and C-terminal carboxyamidation to produce α-MSH (Figure 1A). Acetylated α-MSH is more resistant to degradation than des-acetylated form [39]. At the human MCRs, α-MSH has been reported to possess single-digit nanomolar binding affinity only at the MC1R (Table 2). This peptide possesses nonselective subnanomolar to nanomolar potencies at the MC1R and MC3-5Rs (Table 3). Alanine scans of α-MSH have indicated the Met4, Phe7, Arg8, and Trp9 positions affect binding affinity and functional activity at the mouse MC1R and rat MC3R [40,41]. Expression of α-MSH in the central nervous system is predominantly in the hypothalamus, dispersed throughout the arcuate nucleus, the dorsomedial nucleus of the hypothalamus, and fibers in the medial preoptic, paraventricular, periventricular, and anterior hypothalamic nuclei [42,43]. ICV administration of α-MSH reduces food intake in rodents and intraperitoneal (IP) injection of α-MSH alters skin/hair coloration of humans and other mammals [44,45].
Unlike α-MSH, both termini of 22-residue β-MSH are unmodified (Figure 1A). Compared to the human MC3-5R, β-MSH has lower binding affinity at the MC1R (Table 2). Rodents lack the dibasic cleavage site for β-MSH and are therefore β-MSH-deficient [62]. However, human β-MSH reduces food intake in rats after ICV administration [63,64,65]. Although most studies on body weight regulation focused on α-MSH, β-MSH has also been shown to be important in the regulation of human energy homeostasis. The obesity phenotype in POMC mutation carriers resulting in β-MSH deficiency (Tyr5Cys, Tyr221Cys, and Arg236Gly) implies a significant effect of β-MSH on body weight [66,67,68], underscoring the importance of β-MSH as a physiologically relevant melanocortin ligand. In one study, human plasma β-MSH ranges from 20 to 110 ng/L [69]. In another study, cerebrospinal fluid samples obtained from 30 patients have a mean β-MSH concentration of 60.1 ng/L, compared with mean plasma concentration of 16.1 ng/L for normal adults [70]. Significant amounts of β-MSH were also observed in various regions of the brain [71].
The N-terminal domain of POMC contains three γ-MSH peptides: γ3-MSH (23-residue N-glycosylated peptide), γ2-MSH (N-terminal 12-residue cleaved from γ3-MSH), and γ1-MSH (N-terminal 11-residue cleaved from γ3-MSH with a C-terminal carboxyamidation). An alanine scanning of γ2-MSH indicated the importance of Met3, His5, Phe6, Arg7, and Trp8 for functional activity at the human MC3-5R, similar to residues in α-MSH that are important for activity [59]. γ-MSH is expressed in the pituitary and arcuate nucleus, as well as adrenal medulla [72,73,74,75]. Differences in receptor selectivity have been observed between species, where γ2-MSH is selective for the human MC3R over MC4R and MC5R, whereas there is no selectivity between the mouse MC3R and MC5R [59,76,77]. The >50-fold selectivity of three γ-MSHs for the MC3R over the MC4R led to several investigations of the role the MC3R might play in food intake regulation by administering γ-MSH ligands in vivo (Table 3). Interestingly, the γ-MSH sequence has been deleted from ancestral teleost pomc sequences [78], in parallel with the absence of the mc3r in some teleosts [79,80], providing support for coevolution of melanocortin ligand and the corresponding MCR.

2.2. AgRP as an Inverse Agonist

The MCRs are unique in having two endogenous antagonists, agouti (agouti-signaling protein, ASIP) and AgRP, the only endogenous GPCR antagonists known until 2018, when liver-expressed antimicrobial peptide 2 (LEAP-2) was identified as an endogenous antagonist of growth hormone secretagogue receptor 1a (GHS-R1a, ghrelin receptor) [81]. ASIP and AgRP contain a C-terminal tripeptide Arg-Phe-Phe (RFF) sequence postulated to be critical for binding and functional activity [82,83,84] (Figure 1B). Both proteins possess five C-terminal disulfide bonds, creating a similar cysteine knot conformation [82,84,85].
The active form of ASIP has been hypothesized to be ASIP(23–132), following cleavage of the N-terminal 22 residue signal peptide [85].This protein, produced by hair follicle, acts on melanocytes to inhibit α-MSH-induced eumelanin production by serving as an antagonist at the MC1R [86]. Indeed, ASIP is shown to be an inverse agonist for the MC1R [87]. Pharmacological studies demonstrated that ASIP is also a competitive antagonist at the MC4R, but did not interact with the MC3R [86]. A later study reported that ASIP inhibits the generation of cAMP stimulated by α-MSH at human MC1R and MC3-5R or by ACTH at human MC2R [53] (Table 3). Expression of ASIP is normally limited to the skin, where it affects pigmentation, while ubiquitous expression of ASIP causes other physiological functions, such as its expression in central nervous system resulting in obesity by inhibiting the function of MC4R [53,82,88].
AgRP consists of 112 amino acids (the gene product of 132 amino acids is processed by removal of the N-terminal 20-residue signal peptide). Despite these structural similarities, these two antagonists possess different pharmacological profiles at the MCRs. AgRP is not an antagonist at the MC1R but does interact with the centrally expressed MC3R and MC4R [51,82] (Table 3). Truncated and chimeric ASIP-AgRP ligands indicated that the C-terminal loop of ASIP is responsible for MC1R selectivity [89]. Ubiquitous expression of AgRP in the central nervous system results in increased food intake and obesity without altering pigmentation [82]. AgRP is predominantly expressed in the arcuate nucleus and binds to neural MCRs with nanomolar affinity [51,82]. In Pomc-null mice, administration of AgRP has delayed and long-lasting effects on energy balance and appetite control in a melanocortin signaling-independent way [90]. Recent studies in vivo using mouse model lacking neuronal MSH reported that basal intracellular cAMP levels in cells with MC3R or MC4R expression are decreased by AgRP administration [90], indicating AgRP functions as an inverse agonist at the constitutively active neural MCRs [91,92,93,94]. In addition, AgRP also acts as a biased agonist that activates the mitogen-activated protein kinase signaling pathway [94,95,96].
As in mammals, AgRP also plays critical roles in regulating food intake and energy homeostasis in teleosts [97], acting as an inverse agonist at neural MCRs of many fish [98,99,100,101,102,103]. We and others have shown that both Mc3r and Mc4r in fish have very high constitutive activities compared with human MC3R and MC4R, respectively [80,99,100,101,103,104,105,106,107,108,109,110,111]. Unlike mammals, two gene products (Agrp1 and Agrp2) of Agrp have been identified in several fish species, such as goldfish Carassius auratus [98], zebrafish Danio rerio [99], Atlantic salmon Salmo salar [112], seabass Dicentrarchus labrax [113], and pufferfish Takifugu rubripes [114] (reviewed in [115]). In seabass (Perciforme), long-term fasting increases hypothalamic expression of Agrp1 but decreases that of Agrp2 [113], suggesting an isoform-specific orexigenic action in fish.

2.3. Lipocalin 2

Lipocalin 2 was previously known as neutrophil gelatinase-associated protein exclusively secreted by adipose tissue (an adipokine) and associated with obesity and insulin resistance [116,117]. Recently, lipocalin 2 was shown to be also produced by osteoblasts, at levels much higher than white adipose tissue or other organs [54]. Recently, lipocalin 2 was identified as peripheral-derived novel potent polypeptide agonists for human MC1R, MC3R, and MC4R [54]. This peptide has been reported to possess nanomolar binding affinity at human MC1R, MC3R, and MC4R [54] (Table 2). Lipocalin 2 can activate MC4R in PVN neurons expressing neuropeptide Y Y1 receptor, suggesting that the osteoblast-derived peptide could cross the blood–brain barrier and suppresses appetite after binding to MC4R in hypothalamus [54,118]. In addition, loss- and gain-of-function experiments in mice demonstrate that lipocalin 2 maintains glucose homeostasis by enhancing brown fat activation and thermogenesis, inducing insulin secretion, and improving insulin sensitivity and glucose tolerance [54,117,119,120]. These studies demonstrated that lipocalin 2 links bone and the CNS for metabolic homeostasis of the whole body, and may be a key player for MC4R-mediated physiological functions in mammals.

3. Nonclassical Ligands

3.1. Defensin

β-Defensins are a rapidly evolving family of small secreted proteins thought to mediate responses to various and changing environmental stresses [121]. Mammalian β-defensins are expressed primarily in epithelial tissues, and usually induced in response to exposure to proinflammatory agents [122]. Initially recognized for their potential as “endogenous antibiotics,” some β-defensins can act as ligands for MCRs [123,124]. Canine β-defensin 3 (cBD103) solved a long-standing mystery regarding the dominant inheritance of black coats in domestic dogs through modulating melanocortin signaling [123]. A follow-up study conducted a structure–function analysis of the human orthologue of cBD103, human β-defensin 3 (hBD3), to determine the physicochemical properties of its high-affinity binding to MCRs [56] (Table 2). The unusual feature of melanocortin system is the modulation of the MCRs in the opposite direction by endogenous agonists (α-MSH, β-MSH, γ-MSH, and ACTH) and inverse agonists (ASIP and AgRP). Genetic and pharmacologic alstudies indicate that hBD3 acts as a neutral MCR antagonist capable of blocking the action of either stimulatory endogenous agonists such as α-MSH or inhibitory inverse agonists such as ASIP and AgRP [56]. Accordingly, central administration of hBD3 results in reduced food intake and decreased weight gain in rats [55,56]. Recently, human and mouse β-defensin 3 as well β-defensin 1 are shown to be full micromolar agonists at the MCRs, including MC3R and MC4R [55].
Since hBD1 is expressed in many tissues including brain, it was hypothesized that hBD1, which share the similar purported structure with hBD3, may interact with the neural MCRs involved in energy homeostasis [55]. Mouse and human β-defensin 1 are also shown to be micromolar MCR agonists [55]. We recently showed that cBD103ΔG23, the common variant of cBD103, increases ERK1/2 activation and cAMP accumulation when bound to the human or canine MC4R, suggesting that mutations in defensins may influence more than just the immune system [125].

3.2. Small Molecules

Numerous small molecule ligands for the MCRs have been developed due to their physiological importance [26,126,127,128,129,130]. The earliest work featured β-turn-inducing thioether-cyclized peptidomimetics [131]. In 2002, the first small molecule with single-digit nanomolar potency at the MC4R, THIQ, was described [132] (Figure 2). In the same year, a thioether cyclized peptidomimetic scaffold that generated one ligand, JB1n, possessing submicromolar potency at the MC4R, was reported [133]. Following these initial reports, several classes of small molecule ligands have been disclosed and reviewed in the literature [26,128,134,135].
MC4R was heavily targeted by the pharmaceutical industry due to its role in obesity [17]. Since MC4R agonists decrease food intake in rodent models [88], small molecule MC4R agonists were investigated as potential therapeutics to promote weight loss. However, administration of purported MC4R-selective compounds identified several side effects including hypertension, increased heart rate, skin darkening, and sexual arousal [136,137,138,139]. To avoid undesirable side effects, targeting of the centrally located MC3R, which also plays a role in food intake and fat disposition [18,19,140], has emerged as a valuably alternative approach. As recently reported, pyrrolidine bis-cyclic guanidines have nanomolar agonist potency at the MC3R and over 10-fold selectivity for the MC3R over the MC4R through “unbiased” mixture-based high-throughput screening approaches [141]. Recently, Ericson et al. (2017) [26] and Yeo et al. (2021) [135] published two outstanding reviews that summarized recent small molecule ligands of MCRs at humans and rodents.
Many new, potent, and enzyme-resistant analogs of melanocortin peptides have been developed based on the extensive studies of the melanocortin peptide, α-MSH. These agonists include NDP-α-MSH ([Nle4-D-Phe7]-α-MSH), melanotan II (MTII, Ac-Nle-c [Asp-His-DPhe-Arg-Trp-Lys]-NH2), and SHU9119 (Ac-Nle-c [Asp-His-Dnal(2′)-Arg-Trp-Lys]-OH). NDP-MSH has nanomolar to subnanomolar potencies in MC1R and MC3-5R (Table 2). Thirty-four years after its discovery, NDP-MSH was approved in the European Union as a treatment for adult erythropoietic protoporphyria in 2014 [142]. MTII is a nonselective melanocortin agonist with nanomolar to subnanomolar potencies [143,144] (Table 2). Since its discovery, MTII has been used as an in vitro and in vivo probe, with central ICV administration of MTII inhibiting food intake in mice [88]. SHU9119 is a nanomolar to subnanomolar agonist in MC1R and MC5R with antagonist activity in MC3R and MC4R [61] (Table 2). As the first peptide ligand discovered with potent antagonist activity at the MC3R and MC4R, ICV administration of SHU9119 was shown to significantly increase food intake in mice [88,140].
As in mammals, melanocortin system has also been shown to be involved in regulation of food intake and energy homeostasis in lower vertebrates such as fish (reviewed in [27]). In fact, the two neural MCRs, Mc3r [79,80,102,145] and Mc4r [104,105,106,107,109,110,111,146,147], have also been identified and characterized in several piscine species. Although MC3R is well studied in mammals, the studies on physiology and pharmacology of teleost Mc3rs are still limited, potentially due to the absence of mc3r genes in some teleosts, such as fugu, medaka, and stickleback [114,148,149].
MC4R also plays an important role in regulating energy homeostasis in teleosts. Previous in vivo studies with MC4R ligands showed MC4R agonist (MTII) suppresses food intake whereas the MC4R antagonist (SHU9119) increases the food intake in rainbow trout and goldfish after ICV administration [104,150]. In rainbow trout, ICV administration of MC4R specific antagonist (HS024) is more potent than MC3R/MC4R antagonist (SHU9119) in increasing food intake [150]. In addition, Mexican cavefish (Astyanax mexicanus) mc4r mutations contribute to physiological adaptations to nutrient-poor conditions by increasing appetite, growth, and starvation resistance [151]. The study of regulation of energy homeostasis in teleosts is related to better economic return in aquaculture, therefore the regulation of energy homeostasis by piscine neural Mc3r and Mc4r deserves further exploration. The large variations in genetics, ecology, morphology, anatomy, and physiology of teleost species result in complex species-specific energy homeostasis regulatory mechanisms [152]. For example, a large proportion of teleosts continue to grow during the entire life span, which is in contrast to the determinate growth in mammals and model animals, such as zebrafish [152]. Whether the melanocortin system functions differently in different species is unknown.

3.2.1. THIQ

THIQ (Figure 2A), a tetrahydroisoquinolone ligand, is a synthetic small molecule compound developed by Merck. It was the first selective small molecule agonist of the MC4R with high affinity (IC50, 1.2 nM) and potency (EC50, 2.1 nM) [132]. Structure–function relationship study shows that THIQ competes with [Nle4,D-Phe7]-α-MSH (NDP-MSH) for binding to the MC4R, sharing several binding determinants with peptide agonists but also possessing unique binding determinants [153]. This small molecule MC4R agonist has been shown to attenuate brain inflammation and promote survival [154]. We showed that THIQ is a pharmacoperone of the MC4R, rescuing the cell surface expression and signaling of some intracellularly retained MC4R mutants [155].
THIQ binds to human MC4R orthosterically, but binds to fish Mc4r allosterically. In teleosts, including spotted scat (Scatophagus argus) [105], swamp eel (Monopterus albus) [107], and spotted sea bass (Lateolabrax maculatus) [109], THIQ fails to displace radiolabeled NDP-MSH but stimulates intracellular cAMP accumulation, suggesting that THIQ acts as an allosteric agonist in teleost Mc4r. THIQ exhibits dramatically different pharmacology at teleost and human MC4Rs, suggesting that for any potential use of small molecule ligands developed at human MC4R, pharmacological studies on Mc4r of the intended species need to be performed before any field trials in fish.

3.2.2. Ipsen 5i

Ipsen 5i (Figure 2B), as the representative member of the amino-benzimidazole series, was first described by Roubert and colleagues at Ipsen with a Ki of 2 nM at the human MC4R [156]. Although Ipsen 5i has a high affinity with the MC4R, it has relatively low functional antagonist potency (77 nM) [156], minimizing its antagonizing effect on NDP-MSH. We also demonstrated that Ipsen 5i is a high-affinity inverse agonist of MC4R competing with NDP-MSH for binding to the MC4R while low affinity for the MC3R (Ki, 400 nM) [157]. Although Ipsen5i decreases basal signaling at the classical Gs-cAMP pathway, it acts as an agonist in the mitogen-activated protein kinase pathway [95]. We reported that Ipsen 5i is a novel potent pharmacoperone of the MC4R, correcting trafficking and signaling of a significant portion (73%) of intracellularly retained mutants [158,159].
A recent study with two teleost Mc4r showed that Ipsen 5i serves as a neutral allosteric modulator at cAMP signaling pathway but allosteric agonists at the ERK1/2 signaling pathway [103]. In this study, we showed that Ipsen 5i can bind to either spotted scat or grass carp Mc4r allosterically, failing to compete with radiolabeled NDP-MSH [103]. With swamp eel Mc4r, Ipsen 5i also binds to allosteric site(s) but induces an approximately two-fold increase in intracellular cAMP level, revealing it to be a weak agonist on swamp eel Mc4r [107]. Therefore, Ipsen 5i can have different effects on the Mc4r from different fish.

3.2.3. ML00253764

The peptidomimetic compound ML00253764 (Figure 2C) was initially reported to function as an MC4R antagonist in reducing tumor-induced weight loss after peripheral administration with a low binding affinity (Ki, 160 nM) [160]. This was confirmed in another study using mice grafted with Lewis lung carcinoma tumors [161]. We subsequently showed that ML00253764 behaves as an inverse agonist at the Gs-cAMP pathway and meanwhile serves as an agonist in the mitogen-activated protein kinase pathway in wild-type and several constitutively active mutant human MC4Rs [93,95]. We recently showed that at spotted scat and grass carp Mc4r, ML00253764 serves as a neutral allosteric modulator at cAMP signaling but allosteric agonist at the ERK1/2 signaling pathway [103].

3.3. Pharmacoperones

Dysfunction in folding caused by genetic mutations in numerous genes causes protein conformational diseases [162]. Pharmacological chaperones (pharmacoperones, first coined by the late P. Michael Conn [163]), specifically bind to target protein, stabilizing the native conformation or facilitating the folding of nonnative intermediates into native conformation [164,165,166,167,168,169]. They can serve as novel therapeutics for treating genetic diseases caused by mutations in GPCR genes that result in misfolded mutant receptors [167,170,171,172,173,174,175,176,177]. So far, numerous ligands have been identified as pharmacoperones for GPCRs, such as SR121463A for arginine vasopressin V2 receptor [178], naltrexone for δ-opioid receptor [179], 11-cis-7-ring retinal for rhodopsin [180], Org 42599 for luteinizing hormone receptor [181], and IN3 for gonadotropin-releasing hormone receptor [182].
The MC4R is a critical regulator of energy homeostasis, regulating both food intake and energy expenditure [157]. Numerous MC4R mutations have been identified from 141,456 humans and have been registered in gnomAD v2.1.1 database (https://gnomad.broadinstitute.org/, accessed on 25 August 2022) [183,184]. Many functional studies of naturally occurring MC4R mutations showed that the most common defect is decreased or lack of cell surface expression [157,185].
ML00253764, the first pharmacoperone for MC4R, was identified by our team in 2009 [157,186]. Since then, seven more antagonists have been identified as pharmacoperones for the MC4R: MTHP, PPPone, MPCI, DCPMP, NBP, Ipsen 5i, Ipsen 17, and UM0130866 [158,159,176,187,188]. Antagonist-rescued proteins arrive at the plasma membrane in an antagonist-bound state, potentially preventing the receptor being activated by endogenous agonists. High-affinity antagonist pharmacoperones are generally difficult to dissociate from their target receptors, while low-affinity antagonist pharmacoperones are easy to dissociate. In contrast to antagonist pharmacoperones, agonist pharmacoperones do not need to dissociate from the target receptor. The small molecule potent agonist of the MC4R, THIQ, is also identified as a pharmacoperone, promoting the proper folding and subsequent signaling of intracellularly retained mutant MC4Rs [155]. Interestingly, all the nine antagonist and one agonist pharmacoperones of the MC4R increase the cell surface expression of the wild-type MC4R, indicating pharmacoperone therapy could be used to treat obese patients without any MC4R mutation [157,158,186,187]. Very recently, Bouvier’s group showed that a pharmacoperone can rescue the mutant MC4R, R165W, in knockin mice [176].

3.4. Ligands Identified from Animals

It was suggested that studying nature-derived peptides confer advantages for developing GPCR ligands as drug discover effort [189]. Full agonists were identified from frog peptides that can bind to human MC4R with good affinities, although EC50s are in the tens of micromolar range [190].
Animal venoms contain numerous chemicals with diverse biological activities. In a bank of 3597 toxins, 8% are found to interact with human MC4R [191]. Two peptides, lacking HFRW pharmacophore, were described that can bind to the four MCRs (except MC2R) with low micromolar affinities and activate the human MC1R, causing increased intracellular cAMP [191]. They can work synergistically with α-MSH, although the mechanism is not clear. These peptides are not related structurally to the endogenous agonists, the melanocortins, but have some similarities with AgRP and hBD3 [191]. Further characterization and optimization of these toxins might identify new tools for the study of diverse aspects of MCR pharmacology, such as biased signaling and high selectivity, potentially leading to new therapeutics.

4. Ligands Used in Human Medicine

4.1. ACTH

ACTH is an old forgotten melanocortin peptide that might potentially be rescued and repurposed for new indications [192]. ACTH received the US Food and Drug Administration (FDA) approval for use in humans in 1952, only three years after it was first tested in rheumatoid arthritis patients [193]. Philip S. Hench, Edward C. Kendall, and Tadeus Reichstein were awarded the Nobel Prize in 1950 in Physiology or Medicine for these discoveries on ACTH and adrenal hormones. However, highly efficient methods for glucocorticoid synthesis were subsequently developed and oral forms also became available, making glucocorticoid the treatment of choice to the detriment of ACTH. The use of this melanocortin peptide becomes very limited, usually regarded as a second choice when glucocorticoid therapy is not possible.
ACTH was the gold standard therapy for multiple sclerosis in the 1960s, but its use decreased with the advent of glucocorticoid [194]. H.P. Acthar® Gel is the only treatment available for infantile spasm, a medical condition usually diagnosed within the first year of life, consisting of seizures and mental retardation with poor prognosis [195,196]. A retrospective study on 181 gout patients reported positive response in 77.9% of patients within one day after ACTH injection, indicating that it is effective and safe for the treatment of gout [197,198]. The efficacy of ACTH in the treatment of nephrotic syndrome is well documented, as it improves proteinuria, reduces kidney inflammation, and corrects dyslipidemia, which could not be fully explained by induction of glucocorticoids [199]. It is worth noting that the extra-adrenal effects of ACTH were not known when it was first commercialized, and the relevance of these effects is only emerging recently [199,200,201].
Interestingly, exactly 50 years after ACTH was approved by the FDA, Getting et al. envisaged a novel mechanism of action and discovered that the anti-inflammatory effects of ACTH persist in adrenalectomized rats using a model of knee gout [202]. ACTH also reduces neutrophil infiltration in a model of crystal inflammation and decreases the production of the chemoattractant cytokine CXCL-1, an effect blocked when the MC3R/MC4R antagonist SHU9119 was used [202,203]. This cortisol-independent effect is mediated by the MC3R, which is expressed in immune cells and the brain, presenting this receptor as a novel therapeutic target for ACTH-like drugs devoid of cortisol-related side effects. These findings have renewed interest in ACTH therapy by reconsidering the use of ACTH for new indications and proposing innovative therapeutic targets like the melanocortin system for the development of new anti-inflammatory therapies [192].
In addition, the ability of ACTH to directly modulate local CNS inflammation has been reported in several in vitro studies. Using rat brain cultures containing oligodendroglia, astrocytes, and microglia, preincubated with cytotoxic agents, ACTH was shown to protect mature oligodendroglia and oligodendroglia progenitor cell from death induced by staurosporine, kainate, quinolinic acid, or reactive oxygen species [204,205]. Using rat glial cultures, the same group demonstrated that ACTH induces proliferation of oligodendroglia progenitor cells and accelerates differentiation of platelet-derived growth factor receptor-α to a later stage characterized by greater expansion of oligodendroglia myelin-like sheets compared to untreated cells [205]. Furthermore, the same group showed that ACTH also protects cultured rat forebrain neurons from excitotoxic, apoptotic, oxidative and inflammation related insults [206], but the specific MCR subtype(s) involved are not known. Since excitotoxic damage to neurons is an important cause to several experimental and clinical CNS diseases, it is reasonable to explain the therapeutic benefits of ACTH in several inflammatory animal models of CNS disorders [206]. However, the contribution from direct effects on oligodendroglia, oligodendroglia progenitor cell, or neurons within the brain and whether such beneficial actions can be observed in vivo in human patients remain to be investigated.

4.2. Setmelanotide

Obesity has reached epidemic proportions in developed countries and represents a major public health challenge. The pandemic of obesity is growing at an alarming rate, with an estimated direct and indirect cost of $150 billion per year in the US alone, and one out of five deaths is linked in one way or another to obesity [207]. Obesity is a main risk factor for many diseases, such as type II diabetes, cardiovascular diseases, fatty liver, and cancer, among others [208,209]. For many obese individuals, behavioral and lifestyle modifications are insufficient for long-term weight-loss maintenance and thus depend on pharmacological interventions that alter either appetite or absorption of calories. Until 2014, FDA-approved weight loss medications for obesity management include orlistat, lorcaserin, liraglutide, bupropion-naltrexone, and phentermine-topiramate for long-term treatment; and benzphetamine, phentermine, phendimetrazine, and diethylpropion for short-term use [210].
Over the past few decades, concerted efforts have been made to translate melanocortin ligands into clinical therapies [26,135]. Both α- and β-MSHs were injected into humans in 1961 [44]. Since then, several oral small molecules and injectable peptide agonists of MC4R have been clinically evaluated for treating obesity, including oral small molecule MK-0493 (Merck) [211] and peptide agonists LY2112688 (Eli Lilly) [138], MC4-NN-0453 (Novo Nordisk) [212], AZD2820 (Astra Zeneca) [213], and setmelanotide (Rhythm Pharmaceuticals) [214,215,216,217,218]. However, most MC4R agonist drug candidates developed as potential therapeutics for treating obesity failed in the clinic despite preclinical efficacy, due to lack of efficacy and side effects. The exception up to now has been setmelanotide (Figure 3A), a cyclic octapeptide that has a high potency for the two neutral MCRs (EC50 at MC4R is 0.032 nM and EC50 at MC3R is 0.69 nM) [57]. This synthetic peptide was approved by FDA at the end of 2020 for treating obesity caused by genetic defects in pro-opiomelanocortin (POMC), leptin receptor (LEPR), or proprotein convertase subtilisin/kexin type 1 (PCSK1) [219].
In a 28-day Phase 1b clinical trial, setmelanotide led to moderate weight loss in obese people with MC4R deficiency [220]. The efficacy of setmelanotide was more pronounced in obese subjects with POMC and LEPR deficiency, with 80% of POMC-deficient subjects and 45% of LEPR-deficient subjects losing >10% of their weight within approximately one year [215,217]. In addition, compared to other clinically tested MC4R agonists, setmelanotide does not elicit undesirable side effects such as cardiovascular effects, with only minor adverse events including hyperpigmentation, nausea/vomiting, penile erection, and injection site reactions [214,215,217].
The safety of setmelanotide might be caused by differential signaling pathways [215,221]. Setmelanotide, as a mimetic of α-MSH, initiates similar signaling as α-MSH in Gαs and β-arrestin 2 pathways [222]. Setmelanotide also potently activates Gαq signaling through MC4R, which cannot be efficiently antagonized by AgRP [215,223]. It was found that 100 nM of AgRP cannot antagonize setmelanotide-stimulated phospholipase C activation (using the NFAT reporter gene assay) while stimulation by α-MSH is antagonized [215]. The MC4R-mediated Gαq signaling has been shown to reduce food intake without affecting cardiovascular activity [224]. Therefore, the biased ligands for MC4R that selectively activate Gαq signaling to decrease food intake but not trigger Gαs signaling with cardiovascular side effects might be better drug candidates for obesity treatment [21].

4.3. Bremelanotide (PT-141)

In addition to obesity therapies, the other potential use of MC4R agonists as therapeutics for erectile dysfunction and premenopausal hypoactivity sexual desire disorder has also been investigated, including melanotan II (University of Arizona) [225], bremelanotide (Palatin Inc) [226], and PF-00446687 (Pfizer Inc) [227].
Bremelanotide (PT-141) (Figure 3B) is a MC4R agonist approved in June of 2019 by the FDA for erectile dysfunction in men and hypoactivity sexual desire disorder in premenopausal women, as characterized by low sexual desire that causes marked distress or interpersonal difficulty [228,229]. Bremelanotide was developed by Palatin Technologies (trade name Vyleesi) as a self-administered and on-demand subcutaneous therapy for female sexual dysfunction [230]. MC4R, expressed on neurons in many areas of the central nervous system and some peripheral tissues, are thought to be the most important for sexual function [139,231,232]. As a synthetic peptide analogue of α-MSH, bremelanotide has a high potency for the MC4R (EC50 of 0.25 nM), making it possible to modulate brain pathways involved in sexual response [232,233].
The most common adverse events of bremelanotide are nausea, flushing and headache, with no marked dosage dependence [226,234]. Meanwhile, bremelanotide has limited drug–drug interactions, including no clinically significant interactions with alcohol [235]. The only other FDA-approved treatment indicated for hypoactivity sexual desire disorder is flibanserin, which was approved in 2015. Recently, clinical practices revealed that bremelanotide may offer advantages over flibanserin for certain individuals [235]. Nevertheless, further studies are needed to review the use of this drug in menopausal and postmenopausal populations and to determine its place in therapy of these patients.

4.4. Drug Repurposing

MC3R and MC4R are inextricably linked to obesity. However, research on the old drugs repurposing for new therapeutic uses for the melanocortin system is scarce. The FDA-approved drug fingolimod, a sphingosine 1-phosphate receptor agonist, is used in monotherapy for the treatment of relapsing-remitting multiple sclerosis [236,237]. Recently, fingolimod is demonstrated to exert antiangiogenic activity in diabetic retinopathy not only through the sphingosine 1 phosphate receptor but also through activating MC1R and MC5R [238]. However, in the study, no binding and signaling data were provided; only in silico analysis was done; and AgRP was used as MC1R antagonist [238], which is indeed an antagonist for the MC3R and MC4R. Pharmacology data from cells expressing MC1R and MC5R are needed to confirm the conclusions reached in this study.
Another case is the marketed drug fenoprofen. Long known for the antiarthritic activities and efficacy in rheumatoid arthritis and osteoarthritis, fenoprofen is used as a nonsteroidal anti-inflammatory drug and inhibits cyclooxygenase [239]. Montero-Melendez et al. demonstrated that fenoprofen displays antiarthritic actions on cartilage integrity and synovitis mediated by MC3R, with fenoprofen serving as a positive allosteric modulator, revealing an important contribution of MC3R in the therapeutic actions of fenoprofen in arthritis [240]. In agreement with this study, we showed that positive allosteric modulator activity and biased signaling induced by fenoprofen are observed not only at wild-type MC3R, but also at naturally occurring MC3R mutants, suggesting that this biased allosteric enhancer action might constitute a novel therapeutic opportunity for obese patients harboring these mutations [241].
Another recent study suggested that metformin, used as first-line treatment for type 2 diabetes mellitus, inhibits the activation of MC2R and MC3R [242]. At the MC3R, metformin does not change the sensitivity to α-MSH stimulation, but decreases the maximal response. Since no binding experiment was done, it is not clear whether metformin binds to the MC3R orthosterically or allosterically.

5. Conclusions

Numerous melanocortin ligands have been developed in the 70 years since the sequences of the first endogenous ligands were elucidated. While much of the early focus was on the development of compounds that alter pigmentation, the cloning and structure analysis of the MCRs led to the development of potent and selective ligands. Targeting neural MCRs (MC3R and MC4R) is emerging as a therapeutic approach for metabolic diseases and chronic inflammation. This review summarized the advances in the ligands for neural MCRs, including classical endogenous ligands, nonclassical ligands such as small molecules and pharmacoperones, and some ligands approved for human medicine. Development of novel nonclassical ligands and old drugs repurposed for targeting the neural MCRs to treat obesity and related diseases are of imminent importance to fulfill a highly unmet medical need in the future.

Author Contributions

X.-C.Y. and Y.-X.T. conceived the study. X.-C.Y. drafted the manuscript. Y.-X.T. supervised and finalized the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the National Natural Science Foundation of China, grant 32002399 (to XCY).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors upon request, without undue reservation.

Conflicts of Interest

The authors declare that there is no conflict of interest that would prejudice the impartiality of this study. The funders had no role in the design of the study, in the collection, analyses, or interpretation of data, in the writing of the manuscript, or in the decision to publish the results.

References

  1. Tao, Y.X. Melanocortin receptors. Biochim. Biophys. Acta 2017, 1863, 2411–2413. [Google Scholar] [CrossRef]
  2. Smith, A.I.; Funder, J.W. Proopiomelanocortin processing in the pituitary, central nervous system, and peripheral tissues. Endocr. Rev. 1988, 9, 159–179. [Google Scholar] [CrossRef]
  3. Gantz, I.; Fong, T.M. The melanocortin system. Am. J. Physiol. Endocrinol. Metab. 2003, 284, E468–E474. [Google Scholar] [CrossRef] [PubMed]
  4. Cone, R.D. Studies on the physiological functions of the melanocortin system. Endocr. Rev. 2006, 27, 736–749. [Google Scholar] [CrossRef] [PubMed]
  5. Carmo, J.M.D.; da Silva, A.A.; Wang, Z.; Fang, T.; Aberdein, N.; de Lara, C.E.P.; Hall, J.E. Role of the brain melanocortins in blood pressure regulation. Biochim. Biophys. Acta (BBA) -Mol. Basis Dis. 2017, 1863, 2508–2514. [Google Scholar] [CrossRef] [PubMed]
  6. Wang, W.; Guo, D.Y.; Lin, Y.J.; Tao, Y.X. Melanocortin regulation of inflammation. Front. Endocrinol. 2019, 10. [Google Scholar] [CrossRef]
  7. Xu, Y.; Guan, X.; Zhou, R.; Gong, R. Melanocortin 5 receptor signaling pathway in health and disease. Cell. Mol. Life Sci. 2020, 77, 3831–3840. [Google Scholar] [CrossRef]
  8. Ji, L.Q.; Hong, Y.; Tao, Y.X. Melanocortin-5 receptor: Pharmacology and its regulation of energy metabolism. Int. J. Mol. Sci. 2022, 23, 8727. [Google Scholar] [CrossRef]
  9. Rask-Andersen, M.; Almén, M.S.; Schiöth, H.B. Trends in the exploitation of novel drug targets. Nat. Rev. Drug Discov. 2011, 10, 579–590. [Google Scholar] [CrossRef]
  10. Sriram, K.; Insel, P.A. G protein-coupled receptors as targets for approved drugs: How many targets and how many drugs? Mol. Pharmacol. 2018, 93, 251–258. [Google Scholar] [CrossRef]
  11. Slosky, L.M.; Caron, M.G.; Barak, L.S. Biased allosteric modulators: New frontiers in GPCR drug discovery. Trends Pharmacol. Sci. 2021, 42, 283–299. [Google Scholar] [CrossRef] [PubMed]
  12. Wold, E.A.; Chen, J.; Cunningham, K.A.; Zhou, J. Allosteric modulation of Class A GPCRs: Targets, agents, and emerging concepts. J. Med. Chem. 2018, 62, 88–127. [Google Scholar] [CrossRef] [PubMed]
  13. Gantz, I.; Konda, Y.; Tashiro, T.; Shimoto, Y.; Miwa, H.; Munzert, G.; Watson, S.; DelValle, J.; Yamada, T. Molecular cloning of a novel melanocortin receptor. J. Biol. Chem. 1993, 268, 8246–8250. [Google Scholar] [CrossRef]
  14. Roselli-Rehfuss, L.; Mountjoy, K.G.; Robbins, L.S.; Mortrud, M.T.; Low, M.J.; Tatro, J.B.; Entwistle, M.L.; Simerly, R.B.; Cone, R.D. Identification of a receptor for γ melanotropin and other proopiomelanocortin peptides in the hypothalamus and limbic system. Proc. Natl. Acad. Sci. USA 1993, 90, 8856–8860. [Google Scholar] [CrossRef] [PubMed]
  15. Gantz, I.; Miwa, H.; Konda, Y.; Shimoto, Y.; Tashiro, T.; Watson, S.J.; DelValle, J.; Yamada, T. Molecular cloning, expression, and gene localization of a fourth melanocortin receptor. J. Biol. Chem. 1993, 268, 15174–15179. [Google Scholar] [CrossRef]
  16. Mountjoy, K.G.; Mortrud, M.T.; Low, M.J.; Simerly, R.B.; Cone, R.D. Localization of the melanocortin-4 receptor (MC4-R) in neuroendocrine and autonomic control circuits in the brain. Mol. Endocrinol. 1994, 8, 1298–1308. [Google Scholar] [CrossRef]
  17. Huszar, D.; A Lynch, C.; Fairchild-Huntress, V.; Dunmore, J.H.; Fang, Q.; Berkemeier, L.R.; Gu, W.; A Kesterson, R.; A Boston, B.; Cone, R.D.; et al. Targeted disruption of the melanocortin-4 receptor results in obesity in mice. Cell 1997, 88, 131–141. [Google Scholar] [CrossRef]
  18. Chen, A.S.; Marsh, D.J.; Trumbauer, M.E.; Frazier, E.G.; Guan, X.M.; Yu, H.; Rosenblum, C.I.; Vongs, A.; Feng, Y.; Cao, L.; et al. Inactivation of the mouse melanocortin-3 receptor results in increased fat mass and reduced lean body mass. Nat. Genet. 2000, 26, 97–102. [Google Scholar] [CrossRef]
  19. Butler, A.A.; Kesterson, R.A.; Khong, K.; Cullen, M.J.; Pelleymounter, M.A.; Dekoning, J.; Baetscher, M.; Cone, R.D. A unique metabolic syndrome causes obesity in the melanocortin-3 receptor-deficient mouse. Endocrinology 2000, 141, 3518–3521. [Google Scholar] [CrossRef]
  20. Balthasar, N.; Dalgaard, L.T.; Lee, C.E.; Yu, J.; Funahashi, H.; Williams, T.; Ferreira, M.; Tang, V.; McGovern, R.A.; Kenny, C.D.; et al. Divergence of melanocortin pathways in the control of food intake and energy expenditure. Cell 2005, 123, 493–505. [Google Scholar] [CrossRef]
  21. Yang, L.K.; Tao, Y.X. Biased signaling at neural melanocortin receptors in regulation of energy homeostasis. Biochim. Biophys. Acta (BBA) -Mol. Basis Dis. 2017, 1863, 2486–2495. [Google Scholar] [CrossRef] [PubMed]
  22. Catania, A.; Gatti, S.; Colombo, G.; Lipton, J.M. Targeting melanocortin receptors as a novel strategy to control inflammation. Pharmacol. Rev. 2004, 56, 1–29. [Google Scholar] [CrossRef] [PubMed]
  23. Getting, S.J.; Riffo-Vasquez, Y.; Pitchford, S.; Kaneva, M.; Grieco, P.; Page, C.P.; Perretti, M.; Spina, D. A role for MC3R in modulating lung inflammation. Pulm. Pharmacol. Ther. 2008, 21, 866–873. [Google Scholar] [CrossRef] [PubMed]
  24. Spaccapelo, L.; Bitto, A.; Galantucci, M.; Ottani, A.; Irrera, N.; Minutoli, L.; Altavilla, D.; Novellino, E.; Grieco, P.; Zaffe, D. Melanocortin MC4 receptor agonists counteract late inflammatory and apoptotic responses and improve neuronal functionality after cerebral ischemia. Eur. J. Pharmacol. 2011, 670, 479–486. [Google Scholar] [CrossRef] [PubMed]
  25. Dores, R.M. Hypothesis and Theory: Revisiting views on the co-evolution of the melanocortin receptors and the accessory proteins, MRAP1 and MRAP2. Front. Endocrinol. 2016, 7, 79. [Google Scholar] [CrossRef]
  26. Ericson, M.D.; Lensing, C.J.; Fleming, K.A.; Schlasner, K.N.; Doering, S.R.; Haskell-Luevano, C. Bench-top to clinical therapies: A review of melanocortin ligands from 1954 to 2016. Biochim. Biophys. Acta (BBA) -Mol. Basis Dis. 2017, 1863, 2414–2435. [Google Scholar] [CrossRef]
  27. Tao, Y.X. Mutations in melanocortin-4 receptor: From fish to men. Prog. Mol. Biol. Transl. Sci. 2022, 189, 215–257. [Google Scholar] [CrossRef]
  28. Schiöth, H.B.; Muceniece, R.; Wikberg, J.E.S. Characterisation of the melanocortin 4 receptor by radioligand binding. Pharmacol. Toxicol. 1996, 79, 161–165. [Google Scholar] [CrossRef]
  29. Schiöth, H.B.; Muceniece, R.; Wikberg, J.E.; Chhajlani, V. Characterisation of melanocortin receptor subtypes by radioligand binding analysis. Eur. J. Pharmacol. Mol. Pharmacol. 1995, 288, 311–317. [Google Scholar] [CrossRef]
  30. Kievit, P.; Halem, H.; Marks, D.L.; Dong, J.Z.; Glavas, M.M.; Sinnayah, P.; Pranger, L.; Cowley, M.A.; Grove, K.L.; Culler, M.D. Chronic treatment with a melanocortin-4 receptor agonist causes weight loss, reduces insulin resistance, and improves cardiovascular function in diet-induced obese rhesus macaques. Diabetes 2013, 62, 490–497. [Google Scholar] [CrossRef]
  31. Hruby, V.J.; Wilkes, B.C.; Hadley, M.E.; Al-Obeidi, F.; Sawyer, T.K.; Staples, D.J.; de Vaux, A.E.; Dym, O.; Castrucci, A.M.D.; Hintz, M.F.; et al. α-Melanotropin: The minimal active sequence in the frog skin bioassay. J. Med. Chem. 1987, 30, 2126–2130. [Google Scholar] [CrossRef] [PubMed]
  32. Castrucci, A.M.; Hadley, M.; Sawyer, T.; Wilkes, B.; Al-Obeidi, F.; Staples, D.; de Vaux, A.; Dym, O.; Hintz, M.; Riehm, J.; et al. α-melanotropin: The minimal active sequence in the lizard skin bioassay. Gen. Comp. Endocrinol. 1989, 73, 157–163. [Google Scholar] [CrossRef]
  33. Yang, Y.; Mishra, V.; Crasto, C.J.; Chen, M.; Dimmitt, R.; Harmon, C.M. Third transmembrane domain of the adrenocorticotropic receptor is critical for ligand selectivity and potency. J. Biol. Chem. 2015, 290, 7685–7692. [Google Scholar] [CrossRef] [PubMed]
  34. Chen, M.; Aprahamian, C.J.; Celik, A.; Georgeson, K.E.; Garvey, W.T.; Harmon, C.M.; Yang, Y. Molecular characterization of human melanocortin-3 receptor ligand−receptor interaction. Biochemistry 2006, 45, 1128–1137. [Google Scholar] [CrossRef]
  35. Yang, Y.K.; Fong, T.M.; Dickinson, C.J.; Mao, C.; Li, J.Y.; Tota, M.R.; Mosley, R.; Van der Ploeg, A.L.H.T.; Gantz, I. Molecular determinants of ligand binding to the human melanocortin-4 receptor. Biochemistry 2000, 39, 14900–14911. [Google Scholar] [CrossRef]
  36. Zhou, A.; Bloomquist, B.; Mains, R. The prohormone convertases PC1 and PC2 mediate distinct endoproteolytic cleavages in a strict temporal order during proopiomelanocortin biosynthetic processing. J. Biol. Chem. 1993, 268, 1763–1769. [Google Scholar] [CrossRef]
  37. Bertolini, A.; Poggioli, R.; Vergoni, A.V. Cross-species comparison of the ACTH-induced behavioral syndrome. Ann. N. Y. Acad. Sci. 1988, 525, 114–129. [Google Scholar] [CrossRef]
  38. Vergoni, A.V.; Poggioli, R.; Bertolini, A. Corticotropin inhibits food intake in rats. Neuropeptides 1986, 7, 153–158. [Google Scholar] [CrossRef]
  39. O'Donohue, T.L.; Handelmann, G.E.; Chaconas, T.; Miller, R.L.; Jacobowitz, D.M. Evidence that N-acetylation regulates the behavioral activity of α-MSH in the rat and human central nervous system. Peptides 1981, 2, 333–344. [Google Scholar] [CrossRef]
  40. Sahm, U.G.; Olivier, G.W.; Branch, S.K.; Moss, S.H.; Pouton, C.W. Synthesis and biological evaluation of α-MSH analogues substituted with alanine. Peptides 1994, 15, 1297–1302. [Google Scholar] [CrossRef]
  41. Sahm, U.; Qarawi, M.; Olivier, G.; Ahmed, A.; Branch, S.; Moss, S.; Pouton, C. The melanocortin (MC3) receptor from rat hypothalamus: Photoaffinity labelling and binding of alanine-substituted α-MSH analogues. FEBS Lett. 1994, 350, 29–32. [Google Scholar] [CrossRef]
  42. Elias, C.F.; Saper, C.B.; Maratos-Flier, E.; Tritos, N.A.; Lee, C.; Kelly, J.; Tatro, J.B.; Hoffman, G.E.; Ollmann, M.M.; Barsh, G.S.; et al. Chemically defined projections linking the mediobasal hypothalamus and the lateral hypothalamic area. J. Comp. Neurol. 1998, 402, 442–459. [Google Scholar] [CrossRef]
  43. Jacobowitz, D.M.; O’Donohue, T.L. alpha-Melanocyte stimulating hormone: Immunohistochemical identification and mapping in neurons of rat brain. Proc. Natl. Acad. Sci. USA 1978, 75, 6300–6304. [Google Scholar] [CrossRef] [PubMed]
  44. Lerner, A.B.; McGuire, J.S. Effect of alpha- and beta-melanocyte stimulating hormones on the skin colour of man. Nature 1961, 189, 176–179. [Google Scholar] [CrossRef]
  45. Millington, G.; Tung, Y.; Hewson, A.; O’rahilly, S.; Dickson, S. Differential effects of α-, β-and γ2-melanocyte-stimulating hormones on hypothalamic neuronal activation and feeding in the fasted rat. Neuroscience 2001, 108, 437–445. [Google Scholar] [CrossRef]
  46. Schiöth, H.B.; Muceniece, R.; Mutulis, F.; Prusis, P.; Lindeberg, G.; Sharma, S.D.; Hruby, V.J.; Wikberg, J.E. Selectivity of cyclic [D-Nal7] and [D-Phe7] substituted MSH analogues for the melanocortin receptor subtypes. Peptides 1997, 18, 1009–1013. [Google Scholar] [CrossRef]
  47. Yan, L.Z.; Hsiung, H.M.; Heiman, M.L.; Gadski, R.A.; Emmerson, P.J.; Hertel, J.; Flora, D.; Edwards, P.; Smiley, D.; Zhang, L.; et al. Structure-activity relationships of beta-MSH derived melanocortin-4 receptor peptide agonists. Curr. Top. Med. Chem. 2007, 7, 1052–1067. [Google Scholar]
  48. Conde-Frieboes, K.; Thøgersen, H.; Lau, J.F.; Sensfuss, U.; Hansen, T.K.; Christensen, L.; Spetzler, J.; Olsen, H.B.; Nilsson, C.; Raun, K.; et al. Identification and in vivo and in vitro characterization of long acting and melanocortin 4 receptor (MC4-R) selective α-melanocyte-stimulating hormone (α-MSH) analogues. J. Med. Chem. 2012, 55, 1969–1977. [Google Scholar] [CrossRef]
  49. Bednarek, M.A.; MacNeil, T.; Tang, R.; Fong, T.M.; Cabello, M.A.; Maroto, A.M.; Teran, A. Potent and selective agonists of human melanocortin receptor 5: Cyclic analogues of α-melanocyte-stimulating hormone. J. Med. Chem. 2007, 50, 2520–2526. [Google Scholar] [CrossRef]
  50. Haskell-Luevano, C.; Nikiforovich, G.; Sharma, S.D.; Yang, Y.K.; Dickinson, C.; Hruby, V.J.; Gantz, I. Biological and conformational examination of stereochemical modifications using the template melanotropin peptide, Ac-Nle-c[Asp-His- Phe-Arg-Trp-Ala-Lys]-NH2, on human melanocortin receptors. J. Med. Chem. 1997, 40, 1738–1748. [Google Scholar] [CrossRef]
  51. Yang, Y.K.; Thompson, D.A.; Dikinson, C.J.; Wilken, J.; Barsh, G.S.; Kent, S.B.; Gantz, I. Characterization of Agouti-related protein binding to melanocortin receptors. Mol. Endocrinol. 1999, 13, 148e155. [Google Scholar] [CrossRef] [PubMed]
  52. Kiefer, L.L.; Ittoop, O.R.R.; Bunce, K.; Truesdale, A.T.; Willard, D.H.; Nichols, J.S.; Blanchard, S.G.; Mountjoy, K.; Chen, A.W.-J.; Wilkison, W.O. Mutations in the carboxyl terminus of the agouti protein decrease agouti inhibition of ligand binding to the melanocortin receptors. Biochemistry 1997, 36, 2084–2090. [Google Scholar] [CrossRef] [PubMed]
  53. Yang, Y.K.; Ollmann, M.M.; Wilson, B.D.; Dickinson, C.; Yamada, T.; Barsh, G.S.; Gantz, I. Effects of recombinant agouti-signaling protein on melanocortin action. Mol. Endocrinol. 1997, 11, 274–280. [Google Scholar] [CrossRef] [PubMed]
  54. Mosialou, I.; Shikhel, S.; Liu, J.M.; Maurizi, A.; Luo, N.; He, Z.; Huang, Y.; Zong, H.; Friedman, R.A.; Barasch, J.; et al. MC4R-dependent suppression of appetite by bone-derived lipocalin 2. Nature 2017, 543, 385–390. [Google Scholar] [CrossRef] [PubMed]
  55. Ericson, M.D.; Singh, A.; Tala, S.R.; Haslach, E.M.; Dirain, M.L.S.; Schaub, J.W.; Flores, V.; Eick, N.; Lensing, C.J.; Freeman, K.T.; et al. Human β-defensin 1 and β-defensin 3 (mouse ortholog mBD14) function as full endogenous agonists at select melanocortin receptors. J. Med. Chem. 2018, 61, 3738–3744. [Google Scholar] [CrossRef]
  56. Nix, M.A.; Kaelin, C.B.; Ta, T.; Weis, A.; Morton, G.J.; Barsh, G.S.; Millhauser, G.L. Molecular and functional analysis of human β-defensin 3 action at melanocortin receptors. Chem. Biol. 2013, 20, 784–795. [Google Scholar] [CrossRef]
  57. Durek, T.; Cromm, P.M.; White, A.M.; Schroeder, C.I.; Kaas, Q.; Weidmann, J.; Ahmad Fuaad, A.; Cheneval, O.; Harvey, P.J.; Daly, N.L.; et al. Development of novel melanocortin receptor agonists based on the cyclic peptide framework of sunflower trypsin inhibitor-1. J. Med. Chem. 2018, 61, 3674–3684. [Google Scholar] [CrossRef]
  58. Bednarek, M.A.; MacNeil, T.; Tang, R.; Fong, T.M.; Cabello, M.A.; Maroto, M.; Teran, A. Potent and selective peptide agonists of α-melanocyte stimulating hormone (αMSH) action at human melanocortin receptor 5; their synthesis and biological evaluation in vitro. Chem. Biol. Drug Des. 2007, 69, 350–355. [Google Scholar] [CrossRef]
  59. Grieco, P.; Balse-Srinivasan, P.; Han, G.; Weinberg, D.; MacNeil, T.; Van der Ploeg, L.H.; Hruby, V.J. Synthesis and biological evaluation on hMC3, hMC4 and hMC5 receptors of gamma-MSH analogs substituted with L-alanine. J. Pept. Res. 2002, 59, 203–210. [Google Scholar] [CrossRef]
  60. Qu, H.; Cai, M.; Mayorov, A.V.; Grieco, P.; Zingsheim, M.; Trivedi, D.; Hruby, V.J. Substitution of arginine with proline and proline derivatives in melanocyte-stimulating hormones leads to slectivity for human melanocortin 4 receptor. J. Med. Chem. 2009, 52, 3627–3635. [Google Scholar] [CrossRef]
  61. Hruby, V.J.; Lu, D.; Sharma, S.D.; Castrucci, A.L.; Kesterson, R.A.; al-Obeidi, F.A.; Hadley, M.E.; Cone, R.D. Cyclic lactam α-melanotropin analogues of Ac-Nle4-c[Asp5, D-Phe7,Lys10] α-melanocyte-stimulating hormone-(4-10)-NH2 with bulky aromatic amino acids at position 7 show high antagonist potency and selectivity at specific melanocortin receptors. J. Med. Chem. 1995, 38, 3454–3461. [Google Scholar] [CrossRef] [PubMed]
  62. Bennett, H. Biosynthetic fate of the amino-terminal fragment of pro-opiomelanocortin within the intermediate lobe of the mouse pituitary. Peptides 1986, 7, 615–622. [Google Scholar] [CrossRef]
  63. Abbott, C.R.; Rossi, M.; Kim, M.-S.; AlAhmed, S.H.; Taylor, G.M.; A Ghatei, M.; Smith, D.M.; Bloom, S.R. Investigation of the melanocyte stimulating hormones on food intake: Lack of evidence to support a role for the melanocortin-3-receptor. Brain Res. 2000, 869, 203–210. [Google Scholar] [CrossRef]
  64. Kask, A.; Rägo, L.; Wikberg, J.E.; Schiöth, H.B. Differential effects of melanocortin peptides on ingestive behaviour in rats: Evidence against the involvement of MC3 receptor in the regulation of food intake. Neurosci. Lett. 2000, 283, 1–4. [Google Scholar] [CrossRef]
  65. Harrold, J.A.; Widdowson, P.S.; Williams, G. β-MSH: A functional ligand that regulated energy homeostasis via hypothalamic MC4-R? Peptides 2003, 24, 397–405. [Google Scholar] [CrossRef]
  66. Challis, B.G.; Pritchard, L.E.; Creemers, J.W.; Delplanque, J.; Keogh, J.M.; Luan, J.; Wareham, N.J.; Yeo, G.S.H.; Bhattacharyya, S.; Froguel, P.; et al. A missense mutation disrupting a dibasic prohormone processing site in pro-opiomelanocortin (POMC) increases susceptibility to early-onset obesity through a novel molecular mechanism. Hum. Mol. Genet. 2002, 11, 1997–2004. [Google Scholar] [CrossRef]
  67. Biebermann, H.; Castaneda, T.R.; van Landeghem, F.; von Deimling, A.; Escher, F.; Brabant, G.; Hebebrand, J.; Hinney, A.; Tschop, M.H.; Gruters, A.; et al. A role for β-melanocyte-stimulating hormone in human body-weight regulation. Cell Metab. 2006, 3, 141–146. [Google Scholar] [CrossRef]
  68. Lee, Y.S.; Challis, B.G.; Thompson, D.A.; Yeo, G.S.; Keogh, J.M.; Madonna, M.E.; Wraight, V.; Sims, M.; Vatin, V.; Meyre, D.; et al. A POMC variant implicates β-melanocyte-stimulating hormone in the control of human energy balance. Cell Metab. 2006, 3, 135–140. [Google Scholar] [CrossRef]
  69. Abe, K.; Nicholson, W.E.; Liddle, G.W.; Orth, D.N.; Island, D.P. Normal and abnormal regulation of β-MSH in man. J. Clin. Investig. 1969, 48, 1580–1585. [Google Scholar] [CrossRef]
  70. Smith, A.; Shuster, S. Immunoreactive beta-melanocyte-stimulating hormone in cerebrospinal fluid. Lancet 1976, 307, 1321–1322. [Google Scholar] [CrossRef]
  71. Shuster, S.; Smith, A.; Plummer, N.; Thody, A.; Clark, F. Immunoreactive beta-melanocyte-stimulating hormone in cerebrospinal fluid and plasma in hypopituitarism: Evidence for an extrapituitary origin. BMJ 1977, 1, 1318–1319. [Google Scholar] [CrossRef] [PubMed]
  72. Shibasaki, T.; Ling, N.; Guillemin, R. Pituitary immunoreactive γ-melanotropins are glycosylated oligopeptides. Nature 1980, 285, 416–417. [Google Scholar] [CrossRef] [PubMed]
  73. Osamura, R.; Komatsu, N.; Watanabe, K.; Nakai, Y.; Tanaka, I.; Imura, H. Immunohistochemical and immunocytochemical localization of γ-melanocyte stimulating hormone (γ-MSH)-like immunoreactivity in human and rat hypothalamus. Peptides 1982, 3, 781–787. [Google Scholar] [CrossRef]
  74. Fodor, M.; Sluiter, A.; Frankhuijzen-Sierevogel, A.; Wiegant, V.M.; Hoogerhout, P.; De Wildt, D.J.; Versteeg, D.H. Distribution of Lys-γ2-melanocyte-stimulating hormone-(Lys-γ2-MSH)-like immunoreactivity in neuronal elements in the brain and peripheral tissues of the rat. Brain Res. 1996, 731, 182–189. [Google Scholar] [CrossRef]
  75. Denef, C.; Lu, J.; Swinnen, E. γ-MSH Peptides in the pituitary. Ann. New York Acad. Sci. 2003, 994, 123–132. [Google Scholar] [CrossRef] [PubMed]
  76. Grieco, P.; Balse, P.M.; Weinberg, D.; MacNeil, T.; Hruby, V.J. D-amino acid scan of γ-melanocyte-stimulating hormone: Importance of Trp8 on human MC3 receptor selectivity. J. Med. Chem. 2000, 43, 4998–5002. [Google Scholar] [CrossRef]
  77. Joseph, C.G.; Yao, H.; Scott, J.W.; Sorensen, N.B.; Marnane, R.N.; Mountjoy, K.G.; Haskell-Luevano, C. γ2-Melanocyte stimulation hormone (γ2-MSH) truncation studies results in the cautionary note that γ2-MSH is not selective for the mouse MC3R over the mouse MC5R. Peptides 2010, 31, 2304–2313. [Google Scholar] [CrossRef]
  78. Dores, R.M.; Baron, A.J. Evolution of POMC: Origin, phylogeny, posttranslational processing, and the melanocortins. Ann. N. Y. Acad. Sci. 2011, 1220, 34–48. [Google Scholar] [CrossRef]
  79. Klovins, J.; Haitina, T.; Ringholm, A.; Lowgren, M.; Fridmanis, D.; Slaidina, M.; Stier, S.; Schioth, H.B. Cloning of two melanocortin (MC) receptors in spiny dogfish: MC3 receptor in cartilaginous fish shows high affinity to ACTH-derived peptides while it has lower preference to γ-MSH. Eur. J. Biochem. 2004, 271, 4320–4331. [Google Scholar] [CrossRef]
  80. Ji, R.L.; Huang, L.; Wang, Y.; Liu, T.; Fan, S.-Y.; Tao, M.; Tao, Y.X. Topmouth culter melanocortin-3 receptor: Regulation by two isoforms of melanocortin-2 receptor accessory protein 2. Endocr. Connect. 2021, 10, 1489–1501. [Google Scholar] [CrossRef]
  81. Ge, X.; Yang, H.; Bednarek, M.A.; Galon-Tilleman, H.; Chen, P.; Chen, M.; Lichtman, J.S.; Wang, Y.; Dalmas, O.; Yin, Y.; et al. LEAP2 is an endogenous antagonist of the ghrelin receptor. Cell Metab. 2018, 27, 461–469.e6. [Google Scholar] [CrossRef] [PubMed]
  82. Ollmann, M.M.; Wilson, B.D.; Yang, Y.K.; Kerns, J.A.; Chen, Y.; Gantz, I.; Barsh, G.S. Antagonism of central melanocortin receptors in vitro and in vivo by Agouti-related protein. Science 1997, 278, 135–138. [Google Scholar] [CrossRef]
  83. Kiefer, L.L.; Veal, J.M.; Mountjoy, K.G.; Wilkison, W.O. Melanocortin receptor binding determinants in the Agouti protein. Biochemistry 1998, 37, 991–997. [Google Scholar] [CrossRef] [PubMed]
  84. Tota, M.R.; Smith, T.S.; Mao, C.; MacNeil, T.; Mosley, R.T.; Van der Ploeg, A.L.H.T.; Fong, T.M. Molecular interaction of Agouti protein and Agouti-related protein with human melanocortin receptors. Biochemistry 1998, 38, 897–904. [Google Scholar] [CrossRef] [PubMed]
  85. Willard, D.H.; Bodnar, W.; Harris, C.; Kiefer, L.; Nichols, J.S.; Blanchard, S.; Hoffman, C.; Moyer, M.; Burkhart, W. Agouti structure and function: Characterization of a potent alpha-melanocyte stimulating hormone receptor antagonist. Biochemistry 1995, 34, 12341–12346. [Google Scholar] [CrossRef]
  86. Lu, D.; Willard, D.; Patel, I.R.; Kadwell, S.; Overton, L.; Kost, T.; Luther, M.; Chen, W.; Woychik, R.; Wilkison, W.O.; et al. Agouti protein is an antagonist of the melanocyte-stimulating-hormone receptor. Nature 1994, 371, 799–802. [Google Scholar] [CrossRef]
  87. Siegrist, W.; Drozdz, R.; Cotti, R.; Willard, D.H.; Wilkison, W.O.; Eberle, A.N. Interactions of α-melanotropin and Agouti on B16 melanoma cells: Evidence for inverse agonism of Agouti. J. Recept. Signal Transduct. 1997, 17, 75–98. [Google Scholar] [CrossRef]
  88. Fan, W.; Boston, B.A.; Kesterson, R.A.; Hruby, V.J.; Cone, R.D. Role of melanocortinergic neurons in feeding and the agouti obesity syndrome. Nature 1997, 385, 165–168. [Google Scholar] [CrossRef]
  89. Patel, M.P.; Fabersunne, C.S.C.; Yang, Y.-K.; Kaelin, C.B.; Barsh, G.S.; Millhauser, G.L. Loop-swapped chimeras of the Agouti-related protein and the Agouti signaling protein identify contacts required for melanocortin 1 receptor selectivity and antagonism. J. Mol. Biol. 2010, 404, 45–55. [Google Scholar] [CrossRef]
  90. Tolle, V.; Low, M.J. In vivo evidence for inverse agonism of Agouti-related peptide in the central nervous system of proopiomelanocortin-deficient mice. Diabetes 2008, 57, 86–94. [Google Scholar] [CrossRef]
  91. Nijenhuis, W.A.J.; Oosterom, J.; Adan, R.A.H. AgRP(83–132) acts as an inverse agonist on the human-melanocortin-4 receptor. Mol. Endocrinol. 2001, 15, 164–171. [Google Scholar] [CrossRef] [PubMed]
  92. Haskell-Luevano, C.; Monck, E.K. Agouti-related protein functions as an inverse agonist at a constitutively active brain melanocortin-4 receptor. Regul. Pept. 2001, 99, 1–7. [Google Scholar] [CrossRef]
  93. Tao, Y.X.; Huang, H.; Wang, Z.-Q.; Yang, F.; Williams, J.N.; Nikiforovich, G.V. Constitutive activity of neural melanocortin receptors. Methods Enzymol. 2010, 484, 267–279. [Google Scholar] [CrossRef] [PubMed]
  94. Tao, Y.X. Constitutive activity in melanocortin-4 receptor. Adv. Pharmacol. 2014, 70, 135–154. [Google Scholar] [CrossRef]
  95. Mo, X.L.; Tao, Y.X. Activation of MAPK by inverse agonists in six naturally occurring constitutively active mutant human melanocortin-4 receptors. Biochim. Biophys. Acta -Mol. Basis Dis. 2013, 1832, 1939–1948. [Google Scholar] [CrossRef]
  96. Yang, Z.; Tao, Y.X. Biased signaling initiated by agouti-related peptide through human melanocortin-3 and -4 receptors. Biochim. Biophys. Acta 2016, 1862, 1485–1494. [Google Scholar] [CrossRef] [PubMed]
  97. Volkoff, H. The neuroendocrine regulation of food intake in fish: A review of current knowledge. Front. Neurosci. 2016, 10, 540. [Google Scholar] [CrossRef] [PubMed]
  98. Cerdá-Reverter, J.M.; Peter, R. Endogenous melanocortin antagonist in fish: Structure, brain mapping, and regulation by fasting of the goldfish agouti-related protein gene. Endocrinology 2003, 144, 4552–4561. [Google Scholar] [CrossRef]
  99. Song, Y.; Golling, G.; Thacker, T.L.; Cone, R.D. Agouti-related protein (AGRP) is conserved and regulated by metabolic state in the zebrafish, Danio rerio. Endocrine 2003, 22, 257–266. [Google Scholar] [CrossRef]
  100. Sanchez, E.; Rubio, V.C.; Thompson, D.; Metz, J.; Flik, G.; Millhauser, G.L.; Cerdá-Reverter, J.M. Phosphodiesterase inhibitor-dependent inverse agonism of agouti-related protein on melanocortin 4 receptor in sea bass (Dicentrarchus labrax). Am. J. Physiol. Regul. Integr. Comp. Physiol. 2009, 296, R1293–R1306. [Google Scholar] [CrossRef]
  101. Renquist, B.J.; Zhang, C.; Williams, S.Y.; Cone, R.D. Development of an assay for high-throughput energy expenditure monitoring in the zebrafish. Zebrafish 2013, 10, 343–352. [Google Scholar] [CrossRef] [PubMed]
  102. Yang, L.K.; Zhang, Z.R.; Wen, H.S.; Tao, Y.X. Characterization of channel catfish (Ictalurus punctatus) melanocortin-3 receptor reveals a potential network in regulation of energy homeostasis. Gen. Comp. Endocrinol. 2019, 277, 90–103. [Google Scholar] [CrossRef] [PubMed]
  103. Yang, Z.; Liang, X.F.; Li, G.L.; Tao, Y.X. Biased signaling in fish melanocortin-4 receptors (MC4Rs): Divergent pharmacology of four ligands on spotted scat (Scatophagus argus) and grass carp (Ctenopharyngodon idella) MC4Rs. Mol. Cell. Endocrinol. 2020, 515, 110929. [Google Scholar] [CrossRef] [PubMed]
  104. Cerdá-Reverter, J.M.; Ringholm, A.; Schiöth, H.B.; Peter, R. Molecular cloning, pharmacological characterization, and brain mapping of the melanocortin 4 receptor in the goldfish: Involvement in the control of food intake. Endocrinology 2003, 144, 2336–2349. [Google Scholar] [CrossRef]
  105. Li, J.T.; Yang, Z.; Chen, H.P.; Zhu, C.H.; Deng, S.P.; Li, G.L.; Tao, Y.X. Molecular cloning, tissue distribution, and pharmacological characterization of melanocortin-4 receptor in spotted scat, Scatophagus argus. Gen. Comp. Endocrinol. 2016, 230-231, 143–152. [Google Scholar] [CrossRef]
  106. Li, L.; Yang, Z.; Zhang, Y.P.; He, S.; Liang, X.F.; Tao, Y.X. Molecular cloning, tissue distribution, and pharmacological characterization of melanocortin-4 receptor in grass carp (Ctenopharyngodon idella ). Domest. Anim. Endocrinol. 2016, 59, 140–151. [Google Scholar] [CrossRef]
  107. Yi, T.L.; Yang, L.K.; Ruan, G.L.; Yang, D.Q.; Tao, Y.X. Melanocortin-4 receptor in swamp eel (Monopterus albus): Cloning, tissue distribution, and pharmacology. Gene 2018, 678, 79–89. [Google Scholar] [CrossRef]
  108. Rao, Y.; Chen, R.; Zhang, Y.; Tao, Y.X. Orange-spotted grouper melanocortin-4 receptor: Modulation of signaling by MRAP2. Gen. Comp. Endocrinol. 2019, 284, 113234. [Google Scholar] [CrossRef]
  109. Zhang, K.Q.; Hou, Z.S.; Wen, H.S.; Li, Y.; Qi, X.; Li, W.J.; Tao, Y.X. Melanocortin-4 receptor in spotted sea bass, Lateolabrax maculatus: Cloning, tissue distribution, physiology, and pharmacology. Front. Endocrinol. 2019, 10, 705. [Google Scholar] [CrossRef]
  110. Tao, M.; Ji, R.L.; Huang, L.; Fan, S.Y.; Liu, T.; Liu, S.J.; Tao, Y.X. Regulation of melanocortin-4 receptor pharmacology by two isoforms of melanocortin receptor accessory protein 2 in topmouth culter (Culter alburnus). Front. Endocrinol. 2020, 11, 538. [Google Scholar] [CrossRef]
  111. Wen, Z.Y.; Liu, T.; Qin, C.J.; Zou, Y.C.; Wang, J.; Li, R.; Tao, Y.X. MRAP2 interaction with melanocortin-4 receptor in SnakeHead (Channa argus). Biomolecules 2021, 11, 481. [Google Scholar] [CrossRef] [PubMed]
  112. Murashita, K.; Kurokawa, T.; Ebbesson, L.O.; Stefansson, S.O.; Rønnestad, I. Characterization, tissue distribution, and regulation of agouti-related protein (AgRP), cocaine- and amphetamine-regulated transcript (CART) and neuropeptide Y (NPY) in Atlantic salmon (Salmo salar). Gen. Comp. Endocrinol. 2009, 162, 160–171. [Google Scholar] [CrossRef] [PubMed]
  113. Agulleiro, M.J.; Cortés, R.; Leal, E.; Ríos, D.; Sánchez, E.; Cerdá-Reverter, J.M. Characterization, tissue distribution and regulation by fasting of the agouti family of peptides in the sea bass (Dicentrarchus labrax). Gen. Comp. Endocrinol. 2014, 205, 251–259. [Google Scholar] [CrossRef] [PubMed]
  114. Klovins, J.; Haitina, T.; Fridmanis, D.; Kilianova, Z.; Kapa, I.; Fredriksson, R.; Gallo-Payet, N.; Schioth, H.B. The melanocortin system in Fugu: Determination of POMC/AGRP/MCR gene repertoire and synteny, as well as pharmacology and anatomical distribution of the MCRs. Mol. Biol. Evol. 2004, 21, 563–579. [Google Scholar] [CrossRef]
  115. Cerdá-Reverter, J.M.; Agulleiro, M.J.; R, R.G.; Sánchez, E.; Ceinos, R.; Rotllant, J. Fish melanocortin system. Eur. J. Pharmacol. 2011, 660, 53–60. [Google Scholar] [CrossRef]
  116. Flower, D.R. The lipocalin protein family: Structure and function. Biochem. J. 1996, 318, 1–14. [Google Scholar] [CrossRef]
  117. Yan, Q.W.; Yang, Q.; Mody, N.; Graham, T.E.; Hsu, C.-H.; Xu, Z.; Houstis, N.E.; Kahn, B.B.; Rosen, E.D. The adipokine lipocalin 2 is regulated by obesity and promotes insulin resistance. Diabetes 2007, 56, 2533–2540. [Google Scholar] [CrossRef]
  118. Petropoulou, P.-I.; Mosialou, I.; Shikhel, S.; Hao, L.; Panitsas, K.; Bisikirska, B.; Luo, N.; Bahna, F.; Kim, J.; Carberry, P.; et al. Lipocalin-2 is an anorexigenic signal in primates. eLife 2020, 9. [Google Scholar] [CrossRef]
  119. Guo, H.; Jin, D.; Zhang, Y.; Wright, W.; Bazuine, M.; Brockman, D.A.; Bernlohr, D.A.; Chen, X. Lipocalin-2 deficiency impairs thermogenesis and potentiates diet-induced insulin resistance in mice. Diabetes 2010, 59, 1376–1385. [Google Scholar] [CrossRef]
  120. Zhang, Y.; Guo, H.; Deis, J.A.; Mashek, M.G.; Zhao, M.; Ariyakumar, D.; Armien, A.G.; Bernlohr, D.A.; Mashek, D.G.; Chen, X. Lipocalin 2 regulates brown fat activation via a nonadrenergic activation mechanism. J. Biol. Chem. 2014, 289, 22063–22077. [Google Scholar] [CrossRef]
  121. Pazgier, M.; Hoover, D.M.; Yang, D.; Lu, W.; Lubkowski, J. Human β-defensins. Cell. Mol. Life Sci. CMLS 2006, 63, 1294–1313. [Google Scholar] [CrossRef] [PubMed]
  122. Lehrer, R.I. Primate defensins. Nat. Rev. Genet. 2004, 2, 727–738. [Google Scholar] [CrossRef] [PubMed]
  123. Candille, S.I.; Kaelin, C.B.; Cattanach, B.M.; Yu, B.; Thompson, D.A.; Nix, M.A.; Kerns, J.A.; Schmutz, S.M.; Millhauser, G.L.; Barsh, G.S. A β-defensin mutation causes black coat color in domestic dogs. Science 2007, 318, 1418–1423. [Google Scholar] [CrossRef] [PubMed]
  124. Kaelin, C.B.; I Candille, S.; Yu, B.; Jackson, P.; A Thompson, D.; A Nix, M.; Binkley, J.; Millhauser, G.L.; Barsh, G.S. New ligands for melanocortin receptors. Int. J. Obes. 2008, 32, S19–S27. [Google Scholar] [CrossRef]
  125. Aono, S.; Dennis, J.C.; He, S.; Wang, W.; Tao, Y.X.; Morrison, E.E. Exploring pleiotropic functions of canine β-defensin 103: Nasal cavity expression, antimicrobial activity, and melanocortin receptor activity. Anat. Rec. 2019, 304, 210–221. [Google Scholar] [CrossRef]
  126. Benoit, S.C.; Schwartz, M.W.; Lachey, J.L.; Hagan, M.M.; Rushing, P.A.; Blake, K.A.; Yagaloff, K.A.; Kurylko, G.; Franco, L.; Danhoo, W.; et al. A novel selective melanocortin-4 receptor agonist reduces food intake in rats and mice without producing aversive consequences. J. Neurosci. 2000, 20, 3442–3448. [Google Scholar] [CrossRef]
  127. Mutulis, F.; Yahorava, S.; Mutule, I.; Yahorau, A.; Liepinsh, E.; Kopantshuk, S.; Veiksina, S.; Tars, K.; Belyakov, S.; Mishnev, A.; et al. New substituted piperazines as ligands for melanocortin receptors. Correlation to the X-ray structure of “THIQ”. J. Med. Chem. 2004, 47, 4613–4626. [Google Scholar] [CrossRef]
  128. Todorovic, A.; Haskell-Luevano, C. A review of melanocortin receptor small molecule ligands. Peptides 2005, 26, 2026–2036. [Google Scholar] [CrossRef]
  129. Hess, S.; Linde, Y.; Ovadia, O.; Safrai, E.; Shalev, D.E.; Swed, A.; Halbfinger, E.; Lapidot, T.; Winkler, I.; Gabinet, Y.; et al. Backbone cyclic peptidomimetic melanocortin-4 receptor agonist as a novel orally administrated drug lead for treating obesity. J. Med. Chem. 2008, 51, 1026–1034. [Google Scholar] [CrossRef]
  130. He, S.; Ye, Z.; Dobbelaar, P.H.; Sebhat, I.K.; Guo, L.; Liu, J.; Jian, T.; Lai, Y.; Franklin, C.L.; Bakshi, R.K.; et al. Spiroindane based amides as potent and selective MC4R agonists for the treatment of obesity. Bioorganic Med. Chem. Lett. 2010, 20, 4399–4405. [Google Scholar] [CrossRef]
  131. Haskell-Luevano, C.; Rosenquist, A.; Souers, A.; Khong, K.C.; Ellman, J.A.; Cone, R.D. Compounds that activate the mouse melanocortin-1 receptor identified by screening a small molecule library based upon the beta-turn. J. Med. Chem. 1999, 42, 4380–4387. [Google Scholar] [CrossRef] [PubMed]
  132. Sebhat, I.K.; Martin, W.J.; Ye, Z.; Barakat, K.; Mosley, R.T.; Johnston, D.B.; Bakshi, R.; Palucki, B.; Weinberg, D.H.; MacNeil, T.; et al. Design and pharmacology of N-[(3R)-1,2,3,4-tetrahydroisoquinolinium-3-ylcarbonyl]-(1R)-1-(4-chlorobenzyl)-2-[4-cyclohexyl-4-(1H-1,2,4-triazol-1-ylmethyl)piperidin-1-yl]-2-oxoethylamine (1), a potent, selective, melanocortin subtype-4 receptor agonist. J. Med. Chem. 2002, 45, 4589–4593. [Google Scholar] [CrossRef] [PubMed]
  133. Bondebjerg, J.; Xiang, Z.; Bauzo, R.M.; Haskell-Luevano, C.; Meldal, M. A solid-phase approach to mouse melanocortin receptor agonists derived from a novel thioether cyclized peptidomimetic scaffold. J. Am. Chem. Soc. 2002, 124, 11046–11055. [Google Scholar] [CrossRef]
  134. Ujjainwalla, F.; Sebhat, I.K. Small molecule ligands of the human melanocortin-4 receptor. Curr. Top. Med. Chem. 2007, 7, 1068–1084. [Google Scholar] [CrossRef]
  135. Yeo, G.S.; Chao, D.H.M.; Siegert, A.-M.; Koerperich, Z.M.; Ericson, M.D.; Simonds, S.E.; Larson, C.M.; Luquet, S.; Clarke, I.; Sharma, S.; et al. The melanocortin pathway and energy homeostasis: From discovery to obesity therapy. Mol. Metab. 2021, 48, 101206. [Google Scholar] [CrossRef] [PubMed]
  136. Martin, W.J.; McGowan, E.; E Cashen, D.; Gantert, L.T.; E Drisko, J.; Hom, G.J.; Nargund, R.; Sebhat, I.; Howard, A.D.; Van der Ploeg, L.H.; et al. Activation of melanocortin MC4 receptors increases erectile activity in rats ex copula. Eur. J. Pharmacol. 2002, 454, 71–79. [Google Scholar] [CrossRef]
  137. Hadley, M.E. Discovery that a melanocortin regulates sexual functions in male and female humans. Peptides 2005, 26, 1687–1689. [Google Scholar] [CrossRef]
  138. Greenfield, J.R.; Miller, J.W.; Keogh, J.M.; Henning, E.; Satterwhite, J.H.; Cameron, G.S.; Astruc, B.; Mayer, J.P.; Brage, S.; See, T.C.; et al. Modulation of blood pressure by central melanocortinergic pathways. N. Engl. J. Med. 2009, 360, 44–52. [Google Scholar] [CrossRef]
  139. Van der Ploeg, L.H.T.; Martin, W.J.; Howard, A.D.; Nargund, R.P.; Austin, C.P.; Guan, X.; Drisko, J.; Cashen, D.; Sebhat, I.; Patchett, A.A.; et al. A role for the melanocortin 4 receptor in sexual function. Proc. Natl. Acad. Sci. USA 2002, 99, 11381–11386. [Google Scholar] [CrossRef]
  140. Irani, B.G.; Xiang, Z.; Yarandi, H.N.; Holder, J.R.; Moore, M.C.; Bauzo, R.M.; Proneth, B.; Shaw, A.M.; Millard, W.J.; Chambers, J.B.; et al. Implication of the melanocortin-3 receptor in the regulation of food intake. Eur. J. Pharmacol. 2011, 660, 80–87. [Google Scholar] [CrossRef]
  141. Doering, S.R.; Freeman, K.; Debevec, G.; Geer, P.; Santos, R.G.; Lavoi, T.M.; Giulianotti, M.A.; Pinilla, C.; Appel, J.R.; Houghten, R.A.; et al. Discovery of nanomolar melanocortin-3 receptor (MC3R)-selective small molecule pyrrolidine bis-cyclic guanidine agonist compounds via a high-throughput “unbiased” screening campaign. J. Med. Chem. 2021, 64, 5577–5592. [Google Scholar] [CrossRef] [PubMed]
  142. Luger, T.A.; Böhm, M. An α-MSH analog in erythropoietic protoporphyria. J. Investig. Dermatol. 2015, 135, 929–931. [Google Scholar] [CrossRef] [PubMed]
  143. Al-Obeidi, F.; Castrucci, A.M.D.L.; Hadley, M.E.; Hruby, V.J. Potent and prolonged-acting cyclic lactam analogs of alpha-melanotropin: Design based on molecular dynamics. J. Med. Chem. 1989, 32, 2555–2561. [Google Scholar] [CrossRef] [PubMed]
  144. Al-Obeidi, F.; Hadley, M.E.; Pettitt, B.M.; Hruby, V.J. Design of a new class of superpotent cyclic α-melanotropins based on quenched dynamic simulations. J. Am. Chem. Soc. 1989, 111, 3413–3416. [Google Scholar] [CrossRef]
  145. Logan, D.W.; Bryson-Richardson, R.J.; Taylor, M.S.; Currie, P.; Jackson, I.J. Sequence characterization of teleost fish melanocortin receptors. Ann. N. Y. Acad. Sci. 2003, 994, 319–330. [Google Scholar] [CrossRef]
  146. Jangprai, A.; Boonanuntanasarn, S.; Yoshizaki, G. Characterization of melanocortin 4 receptor in Snakeskin Gourami and its expression in relation to daily feed intake and short-term fasting. Gen. Comp. Endocrinol. 2011, 173, 27–37. [Google Scholar] [CrossRef]
  147. Wang, M.; Chen, Y.; Zhu, M.; Xu, B.; Guo, W.; Lyu, Y.; Zhang, C. Pharmacological modulation of melanocortin-4 receptor by melanocortin receptor accessory protein 2 in Nile tilapia. Gen. Comp. Endocrinol. 2019, 282, 113219. [Google Scholar] [CrossRef]
  148. Logan, D.W.; Bryson-Richardson, R.J.; E Pagán, K.; Taylor, M.S.; Currie, P.D.; Jackson, I.J. The structure and evolution of the melanocortin and MCH receptors in fish and mammals. Genomics 2003, 81, 184–191. [Google Scholar] [CrossRef]
  149. Selz, Y.; Braasch, I.; Hoffmann, C.; Schmidt, C.; Schultheis, C.; Schartl, M.; Volff, J.-N. Evolution of melanocortin receptors in teleost fish: The melanocortin type 1 receptor. Gene 2007, 401, 114–122. [Google Scholar] [CrossRef]
  150. Schjolden, J.; Schiöth, H.B.; Larhammar, D.; Winberg, S.; Larson, E.T. Melanocortin peptides affect the motivation to feed in rainbow trout (Oncorhynchus mykiss). Gen. Comp. Endocrinol. 2009, 160, 134–138. [Google Scholar] [CrossRef]
  151. Aspiras, A.C.; Rohner, N.; Martineau, B.; Borowsky, R.L.; Tabin, C.J. Melanocortin 4 receptor mutations contribute to the adaptation of cavefish to nutrient-poor conditions. Proc. Natl. Acad. Sci. USA 2015, 112, 9668–9673. [Google Scholar] [CrossRef] [PubMed]
  152. Rønnestad, I.; Gomes, A.S.; Murashita, K.; Angotzi, R.; Jönsson, E.; Volkoff, H. Appetite-controlling endocrine systems in teleosts. Front. Endocrinol. 2017, 8, 73. [Google Scholar] [CrossRef] [PubMed]
  153. Yang, Y.; Cai, M.; Chen, M.; Qu, H.; McPherson, D.; Hruby, V.; Harmon, C.M. Key amino acid residues in the melanocortin-4 receptor for nonpeptide THIQ specific binding and signaling. Regul. Pept. 2009, 155, 46–54. [Google Scholar] [CrossRef]
  154. Muceniece, R.; Zvejniece, L.; Vilskersts, R.; Liepinsh, E.; Baumane, L.; Kalvinsh, I.; Wikberg, J.E.; Dambrova, M. Functional evaluation of THIQ, a melanocortin 4 receptor agonist, in models of food intake and Inflammation. Basic Clin. Pharmacol. Toxicol. 2007, 101, 416–420. [Google Scholar] [CrossRef] [PubMed]
  155. Huang, H.; Tao, Y.X. A small molecule agonist THIQ as a novel pharmacoperone for intracellularly retained melanocortin-4 receptor mutants. Int. J. Biol. Sci. 2014, 10, 817–824. [Google Scholar] [CrossRef] [PubMed]
  156. Poitout, L.; Brault, V.; Sackur, C.; Bernetiere, S.; Camara, J.; Plas, P.; Roubert, P. Identification of a novel series of benzimidazoles as potent and selective antagonists of the human melanocortin-4 receptor. Bioorg. Med. Chem. Lett. 2007, 17, 4464–4470. [Google Scholar] [CrossRef]
  157. Tao, Y.X. The Melanocortin-4 receptor: Physiology, pharmacology, and pathophysiology. Endocr. Rev. 2010, 31, 506–543. [Google Scholar] [CrossRef]
  158. Tao, Y.X.; Huang, H. Ipsen 5i is a novel potent pharmacoperone for intracellularly retained melanocortin-4 receptor mutants. Front. Endocrinol. 2014, 5, 131. [Google Scholar] [CrossRef]
  159. Huang, H.; Wang, W.; Tao, Y.X. Pharmacological chaperones for the misfolded melanocortin-4 receptor associated with human obesity. Biochim. Biophys. Acta (BBA) -Mol. Basis Dis. 2017, 1863, 2496–2507. [Google Scholar] [CrossRef]
  160. Vos, T.J.; Caracoti, A.; Che, J.L.; Dai, M.; Farrer, C.A.; Forsyth, N.E.; Drabic, S.V.; Horlick, R.A.; Lamppu, D.; Yowe, D.L.; et al. Identification of 2-[2-[2-(5-bromo-2- methoxyphenyl)-ethyl]-3-fluorophenyl]-4,5-dihydro-1H-imidazole (ML00253764), a small molecule melanocortin 4 receptor antagonist that effectively reduces tumor-induced weight loss in a mouse model. J. Med. Chem. 2004, 47, 1602–1604. [Google Scholar] [CrossRef]
  161. Nicholson, J.R.; Kohler, G.; Schaerer, F.; Senn, C.; Weyermann, P.; Hofbauer, K.G. Peripheral administration of a melanocortin 4-receptor inverse agonist prevents loss of lean body mass in tumor-bearing mice. J. Pharmacol. Exp. Ther. 2006, 317, 771–777. [Google Scholar] [CrossRef] [PubMed]
  162. Valastyan, J.S.; Lindquist, S. Mechanisms of protein-folding diseases at a glance. Dis. Model. Mech. 2014, 7, 9–14. [Google Scholar] [CrossRef] [PubMed]
  163. Conn, P.M.; Ulloa-Aguirre, A.; Janovick, J.A. “Pharmacoperone”: What’s in a word? Pharmacol. Res. 2013, 83, 1–2. [Google Scholar] [CrossRef]
  164. Cohen, F.E.; Kelly, J.W. Therapeutic approaches to protein-misfolding diseases. Nature 2003, 426, 905–909. [Google Scholar] [CrossRef]
  165. Conn, P.J.; Lindsley, C.W.; Meiler, J.; Niswender, C.M. Opportunities and challenges in the discovery of allosteric modulators of GPCRs for treating CNS disorders. Nat. Rev. Drug Discov. 2014, 13, 692–708. [Google Scholar] [CrossRef] [PubMed]
  166. Loo, T.W.; Clarke, D.M. Chemical and pharmacological chaperones as new therapeutic agents. Expert Rev. Mol. Med. 2007, 9, 1–18. [Google Scholar] [CrossRef] [PubMed]
  167. Tao, Y.X.; Conn, P.M. Chaperoning G protein-coupled receptors: From cell biology to therapeutics. Endocr. Rev. 2014, 35, 602–647. [Google Scholar] [CrossRef]
  168. Tao, Y.X.; Conn, P.M. Pharmacoperones as novel therapeutics for diverse protein conformational diseases. Physiol. Rev. 2018, 98, 697–725. [Google Scholar] [CrossRef]
  169. Marinko, J.T.; Huang, H.; Penn, W.D.; Capra, J.A.; Schlebach, J.P.; Sanders, C.R. Folding and misfolding of human membrane proteins in health and disease: From single molecules to cellular proteostasis. Chem. Rev. 2019, 119, 5537–5606. [Google Scholar] [CrossRef]
  170. Bernier, V.; Lagacé, M.; Bichet, D.-G.; Bouvier, M. Pharmacological chaperones: Potential treatment for conformational diseases. Trends Endocrinol. Metab. 2004, 15, 222–228. [Google Scholar] [CrossRef]
  171. Ulloa-Aguirre, A.; Janovick, J.A.; Brothers, S.; Conn, P.M. Pharmacologic rescue of conformationally-defective proteins: Implications for the treatment of human disease. Traffic 2004, 5, 821–837. [Google Scholar] [CrossRef] [PubMed]
  172. Conn, P.M.; Ulloa-Aguirre, A.; Ito, J.; Janovick, J.A. G protein-coupled receptor trafficking in health and disease: Lessons learned to prepare for therapeutic mutant rescue in vivo. Pharmacol. Rev. 2007, 59, 225–250. [Google Scholar] [CrossRef] [PubMed]
  173. Conn, P.M.; Ulloa-Aguirre, A. Trafficking of G-protein-coupled receptors to the plasma membrane: Insights for pharmacoperone drugs. Trends Endocrinol. Metab. 2010, 21, 190–197. [Google Scholar] [CrossRef] [PubMed]
  174. Hou, Z.S.; Ulloa-Aguirre, A.; Tao, Y.X. Pharmacoperone drugs: Targeting misfolded proteins causing lysosomal storage-, ion channels-, and G protein-coupled receptors-associated conformational disorders. Expert Rev. Clin. Pharmacol. 2018, 11, 611–624. [Google Scholar] [CrossRef] [PubMed]
  175. Tao, Y.X. Molecular chaperones and G protein-coupled receptor maturation and pharmacology. Mol. Cell. Endocrinol. 2020, 511, 110862. [Google Scholar] [CrossRef] [PubMed]
  176. René, P.; Lanfray, D.; Richard, D.; Bouvier, M. Pharmacological chaperone action in humanized mouse models of MC4R-linked obesity. JCI Insight 2021, 6. [Google Scholar] [CrossRef]
  177. Ulloa-Aguirre, A.; Zariñán, T.; Gutiérrez-Sagal, R.; Tao, Y.X. Targeting trafficking as a therapeutic avenue for misfolded GPCRs leading to endocrine diseases. Front. Endocrinol. 2022, 13, 934685. [Google Scholar] [CrossRef]
  178. Morello, J.P.; Salahpour, A.; Laperrière, A.; Bernier, V.; Arthus, M.F.; Lonergan, M.; Petäjä-Repo, U.; Angers, S.; Morin, D.; Bichet, D.G.; et al. Pharmacological chaperones rescue cell-surface expression and function of misfolded V2 vasopressin receptor mutants. J Clin Invest 2000, 105, 887–895. [Google Scholar] [CrossRef]
  179. Petäjä-Repo, U.E.; Hogue, M.; Bhalla, S.; Laperrière, A.; Morello, J.P.; Bouvier, M. Ligands act as pharmacological chaperones and increase the efficiency of δ opioid receptor maturation. EMBO J. 2002, 21, 1628–1637. [Google Scholar] [CrossRef]
  180. Noorwez, S.M.; Kuksa, V.; Imanishi, Y.; Zhu, L.; Filipek, S.; Palczewski, K.; Kaushal, S. Pharmacological chaperone-mediated in vivo folding and stabilization of the P23H-opsin mutant associated with autosomal dominant retinitis pigmentosa. J. Biol. Chem. 2003, 278, 14442–14450. [Google Scholar] [CrossRef]
  181. Newton, C.L.; Whay, A.M.; McArdle, C.A.; Zhang, M.; van Koppen, C.J.; van de Lagemaat, R.; Segaloff, D.L.; Millar, R.P. Rescue of expression and signaling of human luteinizing hormone G protein-coupled receptor mutants with an allosterically binding small-molecule agonist. Proc. Natl. Acad. Sci. USA 2011, 108, 7172–7176. [Google Scholar] [CrossRef] [PubMed]
  182. Conn, P.M.; Ulloa-Aguirre, A. Pharmacological chaperones for misfolded gonadotropin-releasing hormone receptors. Adv. Pharmacol. 2011, 62, 109–141. [Google Scholar] [CrossRef] [PubMed]
  183. Lek, M.; Karczewski, K.J.; Minikel, E.V.; Samocha, K.E.; Banks, E.; Fennell, T.; O’Donnell-Luria, A.H.; Ware, J.S.; Hill, A.J.; Cummings, B.B.; et al. Analysis of protein-coding genetic variation in 60,706 humans. Nature 2016, 536, 285–297. [Google Scholar] [CrossRef]
  184. Liu, T.; Ji, R.L.; Tao, Y.X. Naturally occurring mutations in G protein-coupled receptors associated with obesity and type 2 diabetes mellitus. Pharmacol. Ther. 2022, 234, 108044. [Google Scholar] [CrossRef]
  185. Wang, Z.Q.; Tao, Y.X. Functional studies on twenty novel naturally occurring melanocortin-4 receptor mutations. Biochim. et Biophys. Acta (BBA) -Mol. Basis Dis. 2011, 1812, 1190–1199. [Google Scholar] [CrossRef] [PubMed]
  186. Fan, Z.C.; Tao, Y.X. Functional characterization and pharmacological rescue of melanocortin-4 receptor mutations identified from obese patients. J. Cell Mol. Med. 2009, 13, 3268–3282. [Google Scholar] [CrossRef]
  187. René, P.; Le Gouill, C.; Pogozheva, I.D.; Lee, G.; Mosberg, H.I.; Farooqi, I.S.; Valenzano, K.J.; Bouvier, M. Pharmacological chaperones restore function to MC4R mutants responsible for severe early-onset obesity. J. Pharmacol. Exp. Ther. 2010, 335, 520–532. [Google Scholar] [CrossRef]
  188. Wang, X.H.; Wang, H.M.; Zhao, B.L.; Yu, P.; Fan, Z.C. Rescue of defective MC4R cell-surface expression and signaling by a novel pharmacoperone Ipsen 17. J. Mol. Endocrinol. 2014, 53, 17–29. [Google Scholar] [CrossRef]
  189. Muratspahić, E.; Freissmuth, M.; Gruber, C.W. Nature-derived peptides: A growing niche for GPCR ligand discovery. Trends Pharmacol. Sci. 2019, 40, 309–326. [Google Scholar] [CrossRef]
  190. Do, E.U.; Jo, E.B.; Choi, G.; Piao, L.Z.; Shin, J.; Seo, M.-D.; Kang, S.-J.; Lee, B.-J.; Kim, K.H.; Kim, J.B.; et al. Melanocortin 4 receptors interact with antimicrobial frog peptide analogues. Biochem. Biophys. Res. Commun. 2006, 343, 1094–1100. [Google Scholar] [CrossRef]
  191. Reynaud, S.; Ciolek, J.; Degueldre, M.; Saez, N.J.; Sequeira, A.F.; Duhoo, Y.; Bras, J.L.A.; Meudal, H.; Cabo Diez, M.; Fernandez Pedrosa, V.; et al. A venomics approach coupled to high-throughput toxin production strategies identifies the first venom-derived melanocortin receptor agonists. J. Med. Chem. 2020, 63, 8250–8264. [Google Scholar] [CrossRef] [PubMed]
  192. Montero-Melendez, T. ACTH: The forgotten therapy. Semin. Immunol. 2015, 27, 216–226. [Google Scholar] [CrossRef] [PubMed]
  193. Hench, P.S.; Kendall, E.C.; Slocumb, C.H.; Polley, H.F. The effect of a hormone of the adrenal cortex (17-hydroxy-11-dehydrocorticosterone: Compound E) and of pituitary adrenocortical hormone in arthritis: Preliminary report. Ann. Rheum. Dis. 1949, 8, 97–104. [Google Scholar] [CrossRef] [PubMed]
  194. Miller, H.; Newell, D.; Ridley, A. Multiple sclerosis treatment of acute exacerbations with corticotropin (A.C.T.H.). Lancet 1961, 278, 1120–1122. [Google Scholar] [CrossRef]
  195. Gettig, J.; Cummings, J.P.; Matuszewski, K.H.P. Acthar gel and cosyntropin review: Clinical and financial implications. Formul. Manag. 2009, 34, 250–257. [Google Scholar]
  196. Stafstrom, C.E.; Arnason, B.G.W.; Baram, T.Z.; Catania, A.; Cortez, M.A.; Glauser, T.A.; Pranzatelli, M.R.; Riikonen, R.; Rogawski, M.; Shinnar, S.; et al. Treatment of infantile spasms. J. Child Neurol. 2011, 26, 1411–1421. [Google Scholar] [CrossRef] [PubMed]
  197. Khanna, R.; Flanagan, J.J.; Feng, J.; Soska, R.; Frascella, M.; Pellegrino, L.J.; Lun, Y.; Guillen, D.; Lockhart, D.J.; Valenzano, K.J. The pharmacological chaperone AT2220 increases recombinant human acid α-glucosidase uptake and glycogen reduction in a mouse model of Pompe disease. PLoS ONE 2012, 7, e40776. [Google Scholar] [CrossRef]
  198. Daoussis, D.; Antonopoulos, I.; Yiannopoulos, G.; Andonopoulos, A.P. ACTH as first line treatment for acute gout in 181 hospitalized patients. Jt. Bone Spine 2013, 80, 291–294. [Google Scholar] [CrossRef]
  199. Gong, R. The renaissance of corticotropin therapy in proteinuric nephropathies. Nat. Rev. Nephrol. 2011, 8, 122–128. [Google Scholar] [CrossRef]
  200. Berg, A.-L.; Arnadottir, M. ACTH-induced improvement in the nephrotic syndrome in patients with a variety of diagnoses. Nephrol. Dial. Transplant. 2004, 19, 1305–1307. [Google Scholar] [CrossRef]
  201. Bomback, A.S.; A Tumlin, J.; Baranski, J.; E Bourdeau, J.; Besarab, A.; Appel, A.S.; Radhakrishnan, J.; Appel, G.B. Treatment of nephrotic syndrome with adrenocorticotropic hormone (ACTH) gel. Drug Des. Dev. Ther. 2011, 5, 147–153. [Google Scholar] [CrossRef] [PubMed]
  202. Getting, S.J.; Christian, H.C.; Flower, R.J.; Perretti, M. Activation of melanocortin type 3 receptor as a molecular mechanism for adrenocorticotropic hormone efficacy in gouty arthritis. Arthritis Care Res. 2002, 46, 2765–2775. [Google Scholar] [CrossRef] [PubMed]
  203. Getting, S.J.; Flower, R.J.; Perretti, M. Agonism at melanocortin receptor type 3 on macrophages inhibits neutrophil influx. Agents Actions 1999, 48, 140–141. [Google Scholar] [CrossRef]
  204. Benjamins, J.A.; Nedelkoska, L.; Bealmear, B.; Lisak, R.P. ACTH protects mature oligodendroglia from excitotoxic and inflammation-related damagein vitro. Glia 2013, 61, 1206–1217. [Google Scholar] [CrossRef] [PubMed]
  205. Benjamins, J.A.; Nedelkoska, L.; Lisak, R.P. Adrenocorticotropin hormone 1-39 promotes proliferation and differentiation of oligodendroglial progenitor cells and protects from excitotoxic and inflammation-related damage. J. Neurosci. Res. 2014, 92, 1243–1251. [Google Scholar] [CrossRef]
  206. Lisak, R.P.; Nedelkoska, L.; Bealmear, B.; Benjamins, J.A. Melanocortin receptor agonist ACTH 1–39 protects rat forebrain neurons from apoptotic, excitotoxic and inflammation-related damage. Exp. Neurol. 2015, 273, 161–167. [Google Scholar] [CrossRef]
  207. Masters, R.K.; Reither, E.N.; Powers, D.A.; Yang, Y.C.; Burger, A.E.; Link, B.G. The impact of obesity on US mortality levels: The importance of age and cohort factors in population estimates. Am. J. Public Health 2013, 103, 1895–1901. [Google Scholar] [CrossRef]
  208. Kopelman, P.G. Obesity as a medical problem. Nature 2000, 404, 635–643. [Google Scholar] [CrossRef]
  209. Baik, I.; Ascherio, A.; Rimm, E.B.; Giovannucci, E.; Spiegelman, D.; Stampfer, M.J.; Willett, W.C. Adiposity and mortality in men. Am. J. Epidemiol. 2000, 152, 264–271. [Google Scholar] [CrossRef]
  210. Yanovski, S.Z.; Yanovski, J.A. Long-term drug treatment for obesity: A systematic and clinical review. JAMA 2014, 311, 74–86. [Google Scholar] [CrossRef]
  211. Krishna, R.; Gumbiner, B.; Stevens, C.; Musser, B.; Mallick, M.; Suryawanshi, S.; Maganti, L.; Zhu, H.; Han, T.H.; Scherer, L.; et al. Potent and selective agonism of the melanocortin receptor 4 with MK-0493 does not induce weight loss in obese human subjects: Energy intake predicts lack of weight loss efficacy. Clin. Pharmacol. Ther. 2009, 86, 659–666. [Google Scholar] [CrossRef] [PubMed]
  212. Royalty, J.E.; Konradsen, G.; Eskerod, O.; Wulff, B.S.; Hansen, B.S. Investigation of safety, tolerability, pharmacokinetics, and pharmacodynamics of single and multiple doses of a long-acting α-MSH analog in healthy overweight and obese subjects. J. Clin. Pharmacol. 2013, 54, 394–404. [Google Scholar] [CrossRef] [PubMed]
  213. Skowronski, A.A.; Morabito, M.V.; Mueller, B.R.; Lee, S.; Hjorth, S.; Lehmann, A.; Watanabe, K.; Zeltser, L.M.; Ravussin, Y.; Rosenbaum, M.; et al. Effects of a novel MC4R agonist on maintenance of reduced body weight in diet-induced obese mice. Obesity 2013, 22, 1287–1295. [Google Scholar] [CrossRef] [PubMed]
  214. Kühnen, P.; Clement, K.; Wiegand, S.; Blankenstein, O.; Gottesdiener, K.; Martini, L.L.; Mai, K.; Blume-Peytavi, U.; Gruters, A.; Krude, H. Proopiomelanocortin deficiency treated with a melanocortin-4 receptor agonist. N. Engl. J. Med. 2016, 375, 240–246. [Google Scholar] [CrossRef]
  215. Clément, K.; Biebermann, H.; Farooqi, I.S.; Van der Ploeg, L.; Wolters, B.; Poitou, C.; Puder, L.; Fiedorek, F.; Gottesdiener, K.; Kleinau, G.; et al. MC4R agonism promotes durable weight loss in patients with leptin receptor deficiency. Nat. Med. 2018, 24, 551–555. [Google Scholar] [CrossRef]
  216. Eneli, I.; Xu, J.Y.; Fiedorek, F.; Webster, M.; McCagg, A.; Ayers, K.; Van der Ploeg, L.; Garfield, A.; Estrada, E. Tracing the effect of the melanocortin-4 receptor pathway in obesity: Study design and mthodology of the TEMPO Registry. Horm. Res. Paediat. 2018, 90, 335. [Google Scholar] [CrossRef]
  217. Clément, K.; Akker, E.V.D.; Argente, J.; Bahm, A.; Chung, W.K.; Connors, H.; De Waele, K.; Farooqi, I.S.; Gonneau-Lejeune, J.; Gordon, G.; et al. Efficacy and safety of setmelanotide, an MC4R agonist, in individuals with severe obesity due to LEPR or POMC deficiency: Single-arm, open-label, multicentre, phase 3 trials. Lancet Diabetes Endocrinol. 2020, 8, 960–970. [Google Scholar] [CrossRef]
  218. Haws, R.; Brady, S.; Davis, E.; Fletty, K.; Yuan, G.; Gordon, G.; Stewart, M.; Yanovski, J. Effect of setmelanotide, a melanocortin-4 receptor agonist, on obesity in Bardet-Biedl syndrome. Diabetes, Obes. Metab. 2020, 22, 2133–2140. [Google Scholar] [CrossRef]
  219. Markham, A. Setmelanotide: First approval. Drugs 2021, 81, 397–403. [Google Scholar] [CrossRef]
  220. Collet, T.H.; Dubern, B.; Mokrosinski, J.; Connors, H.; Keogh, J.M.; de Oliveira, E.M.; Henning, E.; Poitou-Bernert, C.; Oppert, J.M.; Tounian, P. Evaluation of a melanocortin-4 receptor (MC4R) agonist (Setmelanotide) in MC4R deficiency. Mol. Metab. 2017, 6, 1321–1329. [Google Scholar] [CrossRef]
  221. Kleinau, G.; Heyder, N.A.; Tao, Y.X.; Scheerer, P. Structural complexity and plasticity of signaling regulation at the melanocortin-4 receptor. Int. J. Mol. Sci. 2020, 21, 5728. [Google Scholar] [CrossRef] [PubMed]
  222. Israeli, H.; Degtjarik, O.; Fierro, F.; Chunilal, V.; Gill, A.K.; Roth, N.J.; Botta, J.; Prabahar, V.; Peleg, Y.; Chan, L.F.; et al. Structure reveals the activation mechanism of the MC4 receptor to initiate satiation signaling. Science 2021, 372, 808–814. [Google Scholar] [CrossRef] [PubMed]
  223. Sharma, S.; Garfield, A.S.; Shah, B.; Kleyn, P.; Ichetovkin, I.; Moeller, I.H.; Mowrey, W.R.; Van der Ploeg, L.H. Current mechanistic and pharmacodynamic understanding of melanocortin-4 receptor activation. Molecules 2019, 24, 1892. [Google Scholar] [CrossRef] [PubMed]
  224. Li, Y.Q.; Shrestha, Y.B.; Pandey, M.; Chen, M.; Kablan, A.; Gavrilova, O.; Offermanns, S.; Weinstein, L.S. Gq/11α and Gsα mediate distinct physiological responses to central melanocortins. J. Clin. Investig. 2015, 126, 40–49. [Google Scholar] [CrossRef]
  225. Wessells, H.; Gralnek, D.; Dorr, R.; Hruby, V.J.; E Hadley, M.; Levine, N. Effect of an alpha-melanocyte stimulating hormone analog on penile erection and sexual desire in men with organic erectile dysfunction. Urology 2000, 56, 641–646. [Google Scholar] [CrossRef]
  226. Clayton, A.H.; Althof, S.E.; Kingsberg, S.; DeRogatis, L.R.; Kroll, R.; Goldstein, I.; Kaminetsky, J.; Spana, C.; Lucas, J.; Jordan, R.; et al. Bremelanotide for female sexual dysfunctions in premenopausal women: A randomized, placebo-controlled dose-finding trial. Womens Health 2016, 12, 325–337. [Google Scholar] [CrossRef]
  227. Lansdell, M.I.; Hepworth, D.; Calabrese, A.; Brown, A.D.; Blagg, J.; Burring, D.J.; Wilson, P.; Fradet, D.; Brown, T.B.; Quinton, F.; et al. Discovery of a selective small-molecule melanocortin-4 receptor agonist with efficacy in a pilot study of sexual dysfunction in humans. J. Med. Chem. 2010, 53, 3183–3197. [Google Scholar] [CrossRef]
  228. Dhillon, S.; Keam, S.J. Bremelanotide: First approval. Drugs 2019, 79, 1599–1606. [Google Scholar] [CrossRef]
  229. Pfaus, J.G.; Shadiack, A.; Van Soest, T.; Tse, M.; Molinoff, P. Selective facilitation of sexual solicitation in the female rat by a melanocortin receptor agonist. Proc. Natl. Acad. Sci. USA 2004, 101, 10201–10204. [Google Scholar] [CrossRef]
  230. Molinoff, P.B.; Shadiack, A.M.; Earle, D.; Diamond, L.E.; Quon, C.Y. PT-141: A melanocortin agonist for the treatment of sexual dysfunction. Ann. N. Y. Acad. Sci. 2003, 994, 96–102. [Google Scholar] [CrossRef]
  231. Renquist, B.J.; Lippert, R.N.; Sebag, J.A.; Ellacott, K.L.; Cone, R.D. Physiological roles of the melanocortin MC3 receptor. Eur. J. Pharmacol. 2011, 660, 13–20. [Google Scholar] [CrossRef] [PubMed]
  232. Clayton, A.; Lucas, J.; Jordan, R.; Spana, C.; Pfaus, J. 221 Bremelanotide: A review of its neurobiology and treatment efficacy for HSDD. J. Sex. Med. 2017, 14, S62–S63. [Google Scholar] [CrossRef]
  233. Pfaus, J.; Giuliano, F.; Gelez, H. Bremelanotide: An overview of preclinical CNS effects on female sexual function. J. Sex. Med. 2007, 4, 269–279. [Google Scholar] [CrossRef] [PubMed]
  234. Kingsberg, S.A.; Clayton, A.H.; Portman, D.; Williams, L.A.; Krop, J.; Jordan, R.; Lucas, J.; Simon, J.A. Bremelanotide for the treatment of hypoactive sexual desire disorder: Two randomized Phase 3 trials. Obstet. Gynecol. 2019, 134, 899–908. [Google Scholar] [CrossRef] [PubMed]
  235. Mayer, D.; Lynch, S.E. Bremelanotide: New drug approved for treating hypoactive sexual desire disorder. Ann. Pharmacother. 2020, 54, 684–690. [Google Scholar] [CrossRef]
  236. Cohen, J.A.; Barkhof, F.; Comi, G.; Hartung, H.P.; Khatri, B.O.; Montalban, X.; Pelletier, J.; Capra, R.; Gallo, P.; Izquierdo, G.; et al. Oral fingolimod or intramuscular interferon for relapsing multiple sclerosis. N. Engl. J. Med. 2010, 362, 402–415. [Google Scholar] [CrossRef]
  237. Kappos, L.; Radue, E.W.; O’Connor, P.; Polman, C.; Hohlfeld, R.; Calabresi, P.; Selmaj, K.; Agoropoulou, C.; Leyk, M.; Zhang-Auberson, L.; et al. A placebo-controlled trial of oral fingolimod in relapsing multiple sclerosis. N. Engl. J. Med. 2010, 362, 387–401. [Google Scholar] [CrossRef]
  238. Gesualdo, C.; Balta, C.; Platania, C.B.M.; Trotta, M.C.; Herman, H.; Gharbia, S.; Rosu, M.; Petrillo, F.; Giunta, S.; Della Corte, A.; et al. Fingolimod and diabetic retinopathy: A drug repurposing study. Front. Pharmacol. 2021, 12. [Google Scholar] [CrossRef]
  239. Goodman, L.S. Goodman and Gilman’s the Pharmacological Basis of Therapeutics; McGraw-Hill: New York, NY, USA, 1996. [Google Scholar]
  240. Montero-Melendez, T.; Forfar, R.A.E.; Cook, J.M.; Jerman, J.C.; Taylor, D.L.; Perretti, M. Old drugs with new skills: Fenoprofen as an allosteric enhancer at melanocortin receptor 3. Cell. Mol. Life Sci. 2016, 74, 1335–1345. [Google Scholar] [CrossRef]
  241. Yuan, X.C.; Tao, Y.X. Fenoprofen—An old drug rediscovered as a biased allosteric enhancer for melanocortin receptors. ACS Chem. Neurosci. 2018, 10, 1066–1074. [Google Scholar] [CrossRef]
  242. Parween, S.; Rihs, S.; Flück, C.E. Metformin inhibits the activation of melanocortin receptors 2 and 3 in vitro: A possible mechanism for its anti-androgenic and weight balancing effects in vivo? J. Steroid Biochem. Mol. Biol. 2020, 200, 105684. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Endogenous ligands for melanocortin receptors. (A) Schematic diagram of POMC and sequences of four agonists; (B) sequences of two antagonists. The HFRW motif is shown in blue. The RFF motif is highlighted in red.
Figure 1. Endogenous ligands for melanocortin receptors. (A) Schematic diagram of POMC and sequences of four agonists; (B) sequences of two antagonists. The HFRW motif is shown in blue. The RFF motif is highlighted in red.
Biomolecules 12 01407 g001
Figure 2. Chemical structures of THIQ (A), Ipsen 5i (B), and ML00253764 (C).
Figure 2. Chemical structures of THIQ (A), Ipsen 5i (B), and ML00253764 (C).
Biomolecules 12 01407 g002
Figure 3. Amino acid sequences and chemical structures of setmelanotide (A) and bremelanotide (B). The functional motif is shown in blue.
Figure 3. Amino acid sequences and chemical structures of setmelanotide (A) and bremelanotide (B). The functional motif is shown in blue.
Biomolecules 12 01407 g003
Table 1. Pharmacological properties of endogenous ligands at human melanocortin receptors.
Table 1. Pharmacological properties of endogenous ligands at human melanocortin receptors.
LigandsBinding Affinity [28,29]Activity [30]
MC1Rα-MSH > β-MSH > γ3-MSH > ACTH > γ1-MSH > γ2-MSHγ1-MSH > α-MSH > γ2-MSH > γ3-MSH > β-MSH
MC3Rγ1-MSH > γ3-MSH > β-MSH > γ2-MSH > α-MSH > ACTHγ2-MSH > γ3-MSH > γ1-MSH > α-MSH > β-MSH
MC4Rβ-MSH > α-MSH > ACTH > γ1-MSH > γ3-MSH > γ2-MSHα-MSH > β-MSH > γ2-MSH > γ1-MSH > γ3-MSH
MC5Rα-MSH > β-MSH > ACTH > γ1-MSH > γ2-MSH= γ3-MSHα-MSH > β-MSH > γ1-MSH > γ2-MSH > γ3-MSH
Table 2. Binding affinities of melanocortin ligands at human melanocortin receptors.
Table 2. Binding affinities of melanocortin ligands at human melanocortin receptors.
PeptidesMC1RMC3RMC4RMC5RIndexReference
α-MSH0.033420.76418240Ki (nM)[28]
0.2331.59007160Ki (nM)[46]
0.3215.541.4332Ki (nM)[30]
0.9427.7430.37125.3Ki (nM)[47]
1.54626150Ki (nM)[48]
3.91919120IC50 (nM)[49]
5.9750.0438.7557IC50 (nM)[50]
β-MSH1.1713.437614,400Ki (nM)[28]
0.8910.648.1876.9Ki (nM)[47]
0.86423.219.9306Ki (nM)[30]
γ1-MSH2.687.0629,00042,600Ki (nM)[28]
2.75111300552Ki (nM)[30]
γ2-MSH11.217.7>100,000>100,000Ki (nM)[28]
20.857.262503250Ki (nM)[30]
γ3-MSH1.3910.933,500>100,000Ki (nM)[28]
0.4195.84117336Ki (nM)[30]
ACTH2.586.969317,000Ki (nM)[28]
AgRP-11.2925.6IC50 (nM)[51]
ASIP2319570>1000Ki (nM)[52]
0.476.40.141.16Ki (nM)[53]
Lipocalin 286.9682.1351.39-Kd (nM)[54]
hBD17000 >100,000 IC50 (nM)[55]
hBD326,000 >100,000 IC50 (nM)[55]
42 110 Ki (nM)[56]
Setmelanotide3.9102.1430Ki (nM)[30]
5.51851910IC50 (nM)[57]
Bremelanotide3.422029190IC50 (nM)[57]
NDP-MSH0.0460.780.300.48Ki (nM)[48]
0.1090.6800.6200.891Ki (nM)[30]
MTII0.2201.523IC50 (nM)[57]
0.27242.6623.1Ki (nM)[30]
0.363.10.181.2IC50 (nM)[58]
0.68634.16.646.1Ki (nM)[46]
6.41760.2517Ki (nM)[48]
SHU91190.620.230.070.065IC50 (nM)[58]
0.7141.20.361.12Ki (nM)[46]
Table 3. Functional pharmacology of melanocortin ligands at human melanocortin receptors.
Table 3. Functional pharmacology of melanocortin ligands at human melanocortin receptors.
PeptidesMC1RMC3RMC4RMC5RReference
EC50 (nM)pA2EC50 (nM)pA2EC50 (nM)pA2EC50 (nM)pA2
α-MSH0.057 0.669 0.21 0.807 [31]
1.01 1.04 4.7 10.5 [30]
3.4 1.1 1.9 16 [58]
β-MSH4.24 1.59 6.62 23.7 [30]
γ1-MSH0.932 0.734 48.5 101 [30]
γ2-MSH3.05 0.487 93.7 500 [30]
- 1 55 200 [59]
γ3-MSH1.34 0.465 42.2 233 [30]
AgRP 8.4 8.6 [51]
ASIP 9.3 8.2 9.9 8.9[53]
Lipocalin 21.52 1.83 1.41 [54]
hBD17400 >100,000 21,000 >100,000 [55]
hBD3400 35% @100 μM 2600 45% @100 μM [55]
Setmelanotide5.8 5.3 0.27 1600 [30]
0.26 0.69 <0.032 - [57]
Bremelanotide0.095 2.4 0.25 - [57]
NDP-MSH0.462 0.109 0.075 0.253 [30]
MTII0.2 0.51 0.05 5.33 [30]
0.3 1.3 2.9 3.3 [60]
0.32 1.1 0.26 2.3 [58]
SHU91190.036 Partial agonist8.3 9.30.434 [61]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yuan, X.-C.; Tao, Y.-X. Ligands for Melanocortin Receptors: Beyond Melanocyte-Stimulating Hormones and Adrenocorticotropin. Biomolecules 2022, 12, 1407. https://0-doi-org.brum.beds.ac.uk/10.3390/biom12101407

AMA Style

Yuan X-C, Tao Y-X. Ligands for Melanocortin Receptors: Beyond Melanocyte-Stimulating Hormones and Adrenocorticotropin. Biomolecules. 2022; 12(10):1407. https://0-doi-org.brum.beds.ac.uk/10.3390/biom12101407

Chicago/Turabian Style

Yuan, Xiao-Chen, and Ya-Xiong Tao. 2022. "Ligands for Melanocortin Receptors: Beyond Melanocyte-Stimulating Hormones and Adrenocorticotropin" Biomolecules 12, no. 10: 1407. https://0-doi-org.brum.beds.ac.uk/10.3390/biom12101407

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