Next Article in Journal
The Diversity of Wolbachia across the Turtle Ants (Formicidae: Cephalotes spp.)
Previous Article in Journal
Helenus and Ajax, Two Groups of Non-Autonomous LTR Retrotransposons, Represent a New Type of Small RNA Gene-Derived Mobile Elements
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Structural and Functional Insights into CRF Peptides and Their Receptors

by
Minos-Timotheos Matsoukas
1,*,†,
Vasilis Panagiotopoulos
1,†,
Vlasios Karageorgos
2,†,
George P. Chrousos
3,
Maria Venihaki
4 and
George Liapakis
2,*
1
Department of Biomedical Engineering, School of Engineering, University of West Attica, 12243 Athens, Greece
2
Department of Pharmacology, Faculty of Medicine, University of Crete, 71003 Heraklion, Greece
3
University Research Institute of Maternal and Child Health and Precision Medicine and UNESCO, National and Kapodistrian University of Athens, Livadias 8, 11527 Athens, Greece
4
Department of Clinical Chemistry, Faculty of Medicine, University of Crete, 71003 Heraklion, Greece
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Biology 2024, 13(2), 120; https://doi.org/10.3390/biology13020120
Submission received: 30 December 2023 / Revised: 2 February 2024 / Accepted: 9 February 2024 / Published: 13 February 2024
(This article belongs to the Section Biophysics)

Abstract

:

Simple Summary

Corticotropin-releasing factor or hormone (CRF or CRH) belongs to the family of CRF peptide and non-peptide analogs (or CRF ligands), which play important roles in many physiologic and pathophysiologic conditions. Several of the CRF ligands have shown considerable therapeutic potential in the treatment of various diseases. The CRF ligands act by interacting with two types of receptors. This work describes the structure of CRF ligands and their receptors, as well as the mode of CRF ligand binding to receptors and the activation mechanism of the latter. Understanding the structural basis of CRF ligand binding and activation of their receptors opens avenues for the development of novel drugs targeting CRF receptors.

Abstract

Corticotropin-releasing factor or hormone (CRF or CRH) and the urocortins regulate a plethora of physiological functions and are involved in many pathophysiological processes. CRF and urocortins belong to the family of CRF peptides (CRF family), which includes sauvagine, urotensin, and many synthetic peptide and non-peptide CRF analogs. Several of the CRF analogs have shown considerable therapeutic potential in the treatment of various diseases. The CRF peptide family act by interacting with two types of plasma membrane proteins, type 1 (CRF1R) and type 2 (CRF2R), which belong to subfamily B1 of the family B G-protein-coupled receptors (GPCRs). This work describes the structure of CRF peptides and their receptors and the activation mechanism of the latter, which is compared with that of other GPCRs. It also discusses recent structural information that rationalizes the selective binding of various ligands to the two CRF receptor types and the activation of receptors by different agonists.

1. Introduction

The human corticotropin-releasing factor or hormone (CRF or CRH), also known as h/r CRF, because the human sequence (hCRF) is identical to that of its rat counterpart (rCRF), is a peptide consisting of 41 amino acids. h/rCRF (or in general CRF) belongs to a family of peptides (CRF peptide family) from several species, such as mammals, amphibians, and fish, which includes ovine CRF (oCRF), Sauvagine (SVG), Urotensin (UI), Urocortin I (UcnI), Urocortin II (UcnII) and Urocortin III (UcnIII) [1,2,3,4,5,6,7]. The peptides of the CRF family act by interacting with two types of G-protein-coupled CRF receptors (GPCRs), type 1 (CRF1R) and type 2 (CRF2R), and their action is modulated by the CRF-binding protein (CRF-BP), which binds and inactivates them [1,8,9].
The peptide is secreted by the hypothalamus and it is transported to the pituitary, where it is responsible for the release of adrenocorticotropic hormone (ACTH) [10,11]. Subsequently, ACTH stimulates the release of glucocorticoids from the adrenals [10,11]. Hypothalamic CRF is essential for homeostatic maintenance by regulating the function of the hypothalamic–pituitary–adrenal (HPA) axis and orchestrating various responses to stress, including autonomic, neuroendocrine, immunologic, and behavioral ones [10,11]. Moreover, CRF plays a role in stress and other physiological processes by regulating the cardiovascular, gastrointestinal, reproductive, and central nervous (CNS) systems [12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37].
CRF is implicated in the pathophysiology of many psychiatric disorders, including depression, anxiety, post-traumatic stress disorder (PTSD), and substance/alcohol abuse [34]. The effects of CRF on anxiety and depression are predominately mediated through its interaction with CRF1R [15,38,39,40]. Several non-peptide CRF1R-selective analogs have been used to treat psychiatric diseases in preclinical studies and clinical trials [34,41,42,43]. In general, CRF2R acts in a manner that counteracts the effects of CRF1R, having anxiolytic-like effects following administration of the CRF2R-selective UcnII or UcnIII in experimental animals [44,45,46]. However, activation of CRF2R could differentially affect depression and anxiety, depending on the region of the CNS [33]. In addition to their CNS effects, preclinical studies have shown that CRF1R-selective antagonists could be possibly effective in the treatment of abdominal and pelvic diseases, whereas previous studies have shown that UcnI has cardioprotective properties and UcnII and UcnIII are potent vasodilators [36,37,47,48,49,50]. Moreover, the CRF1R-selective antagonists, antalarmin and tildacerfont (or LY2371712), have been shown to decrease the progression of endometriosis and the excessive androgen production in congenital adrenal hyperplasia (allowing for glucocorticoid dose reduction), respectively [51,52].
The CRF receptors belong to the subfamily B1 of family B GPCRs [53]. Like all GPCRs, CRF receptors are plasma membrane proteins that contain seven alpha-helical transmembrane domains (TMs), and three extracellular (ELs) and three intracellular loops (ILs), as well as an extracellular amino-terminal domain (ECD) and an intracellular carboxyl–terminal region (C-domain) (Figure 1) [1].
The ECD is also named the N-domain, whereas their ELs along with their TMs comprise the J-domain (Figure 1) [1]. The TMs of CRF receptors are arranged such as to form a pocket into the phospholipid bilayer of the plasma membrane, named the binding pocket, the surface of which is accessible to the extracellular fluid and interacts with ligands [54]. The binding of CRF agonists to the extracellular domains of CRF receptors triggers conformational changes, which are associated with receptor activation and the subsequent stimulation of G-proteins, thus leading to a biological response (Figure 1). Importantly, activation of each type of CRF receptor could result in a diverse spectrum of biological responses, given that it could stimulate different signaling pathways by interacting with diverse G-proteins, such as Gαs, Gαi, Gαo, Gαq, and Gαz [55,56].
Even though CRF receptors and the other subfamily B1 GPCRs have a low sequence homology with family A GPCRs, all share a common structural architecture and several common activation mechanisms, suggesting that these characteristics have been conserved over the course of evolution, being very important for their function [53,57,58].

2. Pharmacological Properties of CRF Ligands and Their Receptors and Historical Overview

Pharmacological studies have shown that different CRF peptides display different binding affinities for the CRF1R and CRF2R (Table 1). The oCRF and all known non-peptide CRF analogs, including antalarmin (Ki = 9.7 nM, where Ki is the binding affinity of the ligand determined from heterologous displacement radioligand experiments [59]), bind selectively to CRF1R, whereas UcnII and UcnIII are CRF2R-selective ligands. Human UcnI (hUcnI) and SVG bind non-selectively to CRF1R and CRF2R. In addition, the synthetic CRF peptides, astressin, and α-helical CRF(9-41) (see below) interact with CRF1R and CRF2R with slightly different binding affinities [1].
The oCRF was first characterized and synthesized by Vale and his research team at Salk Institute, in 1981, whereas concurrently a similar peptide, named sauvagine, was isolated from the skin of Phyllomedusa sauvagei [3,10]. A year later, urotensin I, a CRF-like peptide with hypotensive properties, was isolated in Catostomus species [4]. Two years after the initial discovery of CRF, its rat counterpart was isolated by the same research team at Salk Institute, whereas concurrently a Japanese research group identified the gene encoding the human CRF [62,63]. Subsequently, a CRF-like peptide has been also identified and isolated from other species such as pig, fish, and frog [64,65,66]. Later on, and specifically in 1995, Vaughan et al. identified the rat UcnI, followed by the identification of its human counterpart and the homologous peptides UcnII and UcnIII from various species [2,5,6,7,67]. In addition to CRF peptides from different species, the CRF family includes synthetic peptides, such as stressin1-A, astressin, astressin-2B, antiSVG-30, and cortagine, and synthetic non-peptide CRF analogs, such as antalarmin. These CRF synthetic peptides are CRF1R-selective, CRF2R-selective, or non-selective agonists or antagonists and were derived from appropriate modifications of the maternal peptides, as reviewed by Liapakis et al. 2011 [1]. In addition to CRF peptide analogs, non-peptide CRF1R-selective antagonists were created as discussed below.

3. Structural Features of CRF Family Peptides

3.1. Peptide CRF Analogs

The peptides of the CRF family consist of three segments: a carboxy-terminal (C-segment), an amino-terminal (N-segment), and an internal (I-segment) one. A study from Grace et al. (2007) describes the NMR structures of six CRF peptide analogs (the antagonists astressin-2B and astressin-B, the agonist stressin1-A, and the natural ligands human UcnI, UcnII, and UcnIII) in solution (Figure 2) [68].

3.1.1. The N-Segment

The N-segment of CRF peptides is significant for their biological activity. Removal of the amino-terminal 23 or 28 amino acids from UI created biologically inactive analogs [69]. Similarly, deletion of the eight amino-terminal amino acids of oCRF resulted in the inactive peptide oCRF (9–41) [70]. This was likely due to the loss of the ability of the truncated peptide to produce a biological effect and not to its binding to CRF receptor, granted that it was able to antagonize the biological actions of CRF [70]. Further modifications of the truncated oCRF (9–41), which enhanced its α-helical structure, created the first CRF antagonist, named α-helical-CRF (9–41) [70]. Supportive evidence for the functional importance of the N-segment of CRF peptides has been provided by the study of Kornreich et al., which has shown that Ala mutation of most amino acids in this region had detrimental results on the biological activity of CRF [71].

3.1.2. The C-Segment

The C-segment is also very important for the function of CRF peptides and their binding to receptors. The CRF analog, YAM19, which consists of only the 12 C-terminal amino acids of CRF, bound with high affinity to a soluble form of the ECD of CRF1R [72]. In marked contrast, deletion of the two last amino acids of its C-segment, Ile40, and Ala41, largely decreased the biological potency of CRF [10,71].
Similarly, deletion of the amino acid at position 41 from a CRF-truncated antagonist decreased its antagonistic activity [73]. Furthermore, the removal of the five carboxyl-terminal amino acids (37–41) from UI or substitution of a free acid for the amidated carboxyl-terminal end of CRF reduced (or abolished) the biological activity of the peptides [10,69,71,74].
Supportive evidence for the functional importance of the C-segment of CRF peptides is provided by several studies. Truncation of the carboxyl-terminal segment of UI, resulting in a smaller fragment, containing residues 1–19, largely decreased the activity of the peptide [69]. Similar to UI, the stepwise deletion of C-terminal amino acids from astressin, thus resulting in different fragments having detrimental effects in the peptide binding [73]. Importantly, even the deletion of only the last C-terminal residue (Ile41P) largely reduced the binding affinity of the peptide [73]. In addition, Ala substitutions for Arg35P and Leu38P in the C-segment of CRF significantly reduced the potency and binding affinity of the peptide [10,71,74]. The superscripts of peptide residues represent their position in the peptide sequence followed by the letter p, which is the abbreviation of the peptide. For example, Arg35P indicates that Arg is the 35th amino acid in the peptide sequence, starting from the N-terminus. Substitution with alanine is used to determine the role of side chains because it removes all side-chain atoms past the β-carbon [75]. Consistent with the functional importance of residue 38, its replacement in peptide-1, a small 12-amino-acid N-truncated analog of CRF, by cyclohexylalanine (Cha) increased its affinity for CRF1R [76]. In contrast, the substitution of residue 38 by the smaller Phe did not increase its affinity, suggesting that the bulkier Cha maximized the hydrophobic interaction between residue 38 of the peptide and the receptor [76]. It is possible that in small peptides with fewer interactions with the CRF1R, the side chain of residue 38 plays a crucial role in this network of interactions.
The significant role of Arg35P in its interaction with the receptor is supported by the crystal structure of the CRF1R in complex with CRF, in which a network of interactions take place between Arg35P of CRF and receptor residues, as thoroughly discussed below [74].

3.1.3. The I-Segment

An important role for the CRF1R binding likely plays the I-segment of peptides containing amino acids 14-30. Beyermann et al. created the analog UcnI-EK (UEK) by replacing the I-segment of UcnI with a highly charged linker consisting exclusively of Glu acid and Lys, which were arranged in a way that side chains at positions i and i4 form salt bridges (EKEEKEKKRKE) (Figure 3) [77]. This arrangement resulted in helix stabilization. Stepwise shortening of this charged linker by deleting amino acids resulted in peptide analogs of various lengths, with or without a complete alpha-helical conformation (Table 2). Only peptides bearing a complete a-helical linker, independently of its length, are potent, suggesting that the orientation of the two N- and C-segments rather than the preservation of a specific distance between them is important for peptide function [77].
Supportive evidence for the functional significance of the I-segment of CRF peptides has been provided by previous studies, which have shown that a segment between residues 8 and 32 could form an α-helical structure, which is likely stabilized in less hydrophilic environments, such as the amphiphilic one of cell membrane containing the CRF binding sites [78,79]. In addition, enhancement of the α-helical structure of peptide after the substitution of α-helical preferring residues with several amino acids of oCRF, increased the biological potency of the peptide [70]. In contrast, replacing several amino acids of oCRF with their D-enantiomers drastically decreased its biopotency [80,81]. D-enantiomers destabilize the α-helical structures [82]. The α-helical structure of CRF peptides has been verified in an NMR study (Figure 2) [68].
The importance of the I-segment of CRF peptides is further supported by the study of Eckart et al. (2001) [83]. Specifically, the Ala22P of the Ala22P-Arg23P-Ala24P-Glu25P (ARAE) motif of the α-helical I-segment of CRF is very important for peptide binding to the CRF-BP. Mutation of the Ala22P to the corresponding Glu of SVG changed the phenotype of CRF to that of SVG, by largely reducing the binding affinity of peptide to CRF-BP.

3.2. Non-Peptide CRF Analogs

In addition to peptide ligands, many non-peptide antagonists have been created. Most of these analogs have a planar heterocyclic structure (mono-, bi- or tri-) in the center of the molecule [84,85,86,87] (Figure 4). Their heterocyclic ring contains a nitrogen atom, which could participate in hydrogen bonds and it is an important functional element of these analogs. In addition to this nitrogen atom, functionally important groups are a lipophilic one attached at the top of the core, and a lower aryl or heteroaryl ring (Figure 4). Another principal functional element is an ortho-substituent in the lower ring, which restricts this ring orthogonal to the plane of the core (Figure 4). This restriction could also be accomplished by adding an alkyl or alkoxy group next to its nitrogen atom into the core. Moreover, new compounds have been synthesized, bearing an acyclic central core with a nitrogen atom [88]. The important role of this nitrogen atom was suggested by its methylation abolished binding [88]. The latest CRF1R small-molecule antagonists reported are NGD9002 and NGD98-2 (Figure 4), which suppressed stimulation of motor activity of the colon induced by stress and visceral hypersensitivity in experimental animals [87]. However, small allosteric molecules have failed to reach clinical utility due to limitations arising from their similar chemical properties [89].

4. Receptors

4.1. The ECD

The extracellular amino-terminal domain (or ECD) of CRF1R plays a major role in ligand binding. Replacement of the ECD of CRF1R with the ECD of the receptor for the GH-releasing hormone (GRF-R) resulted in the loss of astressin and UcnI binding [90]. In marked contrast, replacement of ECD of the GRF-R with the ECD of CRF1R generated a receptor capable to bind astressin and UcnI [90]. Supportive evidence for the significance of ECD in CRF peptide binding is provided by previous studies, which have shown that a soluble form of the isolated ECD of CRF1R and CRF2R, and a receptor formed by substituting the ECD of the receptor for activin receptor (a single transmembrane protein) with the corresponding region of CRF1R or CRF2R, bind various CRF peptides with considerable affinities [91,92,93]. A crystal structure of the isolated ECD of CRF1R revealed that main chain atoms of Val97 form hydrogen bonds with the oxygen and nitrogen atoms of the C-terminal amide group (of residue 41) of CRF, demonstrating its important role in peptide function (Figure 5) [74]. In addition, Ile41P of CRF hydrophobically interacts with residues Leu50 and Ile51 of the ECD. Furthermore, hydrophobic interactions between Met38P of CRF and Tyr99 and Tyr77 of the ECD have been observed, as well as hydrophobic interactions along Phe72 of the ECD with Leu37P of CRF. The Asn34P of CRF could also possibly form a hydrogen bond with the side chain oxygen of Tyr77 of the ECD. Moreover, the nitrogen of the C-terminal amide interacts through a hydrogen bond with the backbone carbonyl of Met38 of CRF, which stabilizes the significant α-helical structure of the peptide [74] (Figure 5). Similarly, the amide group of the C-terminal residue 41 of astressin forms H-bonds with CRF2R residues including Val113 (Val97 in CRF1R) [94].
In addition to these amino acids, several Cys of the ECD of CRF1R and CRF2R form disulfide bridges that play an important functional role [74,92,93,95,96]. A reduction in these disulfide bonds in CRF1R using DTTs or their mutation decreased CRF binding [97]. These disulfide bonds hold important structural motifs in the ECD of CRF1R. These structural motifs are a short N-terminal α-helix followed by two anti-parallel β-sheets each with two β-strands (β1-β2 and β3-β4), and a short C-terminal α-helix, as revealed in the crystal structure of the ECD of CRF1R (Figure 5) [74]. A similar structure has been observed in the structure of the ECD of CRF2R, as determined in an NMR study, with some differences, such as the absence of a salt bridge between Arg85 and Asp49 in CRF1R, as observed in CRF2R [74,94].
These structural motifs of the receptor’s ECD are of vital importance for its function. Specifically, the β1-β2 loop of CRF1R, which is formed between the β1-β2 strands and interacts with the Met38P and Ile41P of CRF, is shifted closer to CRF upon peptide binding, rendering the Ile51 at the top of this loop a contact site of the peptide [74]. Structural rearrangements are also associated with peptide binding to the CRF2R [92,94].
In addition to its significance in ligand binding, the ECD is proposed to play a role in CRF receptor activation by interacting with the third extracellular loop (EL3) of the J-domain of receptor [98]. Specifically, Dore et al. proposed that the EL3 of CRF1R likely interacts with the ECD to stabilize an inactive conformation of the receptor, like the glucagon receptor [98,99]. Upon CRF binding to the ECD, the peptide destabilizes the receptor’s inactive state by affecting ECD-EL3 interaction before interacting with the J-domain to result in receptor activation. In accordance with this theory, a recent cryo-EM study of the active states of both CRF1R and CRF2R in complex with UcnI has shown that the ECDs do not interact with the TMs of receptors [100].

4.2. The J-Domain

The J-domain of CRF receptors plays an important role in ligand binding and receptor activation. It contains contact sites of ligands, and is responsible for the transmission of conformational changes, associated with agonist binding and receptor activation, to the intracellular regions of receptor, which subsequently stimulate the G-proteins, thus resulting in a biological response [101,102]. Interaction of non-peptide antagonists, such as antalarmin and CP-376395 with TM residues of the J-domain of CRF1R allosterically inhibit agonist binding and block the receptor activation-associated conformational changes, thus antagonizing the CRF biological effects [1,102,103]. In detail, the pyrimidine nitrogen of CP-376395 interacts through a hydrogen bond with Asn2835.50b of CRF1R, whereas the aryloxy group of ligand interacts with a hydrophobic pocket of receptor formed by Phe 2845.51b, Leu2875.54b, Ile2905.57b, Thr3166.42b, Leu3196.45b and Leu3206.46b [103] (Figure 6).
The exocyclic alkylamino moiety of CP-376395 interacts with Gly3246.50b, Phe2033.44b, Leu2805.47b, Leu3236.49b, and Tyr3276.53b of CRF1R [103]. The pattern of binding of the structurally related antalarmin is similar to that of CP-376395 [102]. The superscripts of receptor residues represent their positions in the TMs of the receptor, with the most conserved residue in each TM of subfamily B1 GPCRs to be assigned the position index .50, and this number is preceded by the TM number (TM1–TM7) [104]. For example, Phe2845.51b denotes Phe284 located in TM5, one residue after the most conserved residue, Asn2835.50b.
In addition to the interaction with the non-peptide analogs, the residues of the J-domain interact with the CRF peptides. An alanine mutagenesis study has shown that Trp259EL2 and Phe260EL2 of the CRF1R play role in receptor interaction with CRF and SVG [101]. The superscripts EL1, EL2, and EL3 of receptor residues represent the ELs of the receptor, in which these residues are located. This interaction is supported by the cryoelectron microscopy (cryo-EM) results of CRF1R in complex with the Gs protein and CRF (receptor structure at 2.91 Å resolution, pdb: 6P9X) [105]. Specifically, the backbone of Trp259EL2 of the receptor is linked through a water-coordinated hydrogen bond network with Asn1963.37b, Tyr2725.39b, and Asp2695.36b, whereas the backbone of Phe260EL2 interacts through an H-bond with Arg16P.
Similar to EL2, the EL1 and EL3 of CRF1R are also important for peptide binding. A photo-cross-linking study demonstrated that the amino-terminally located residues 17 and 22 of the UcnI analogs lay in close proximity to a region between Trp170EL1 to Glu179EL1 of CRF1R [106]. A different study has also shown that position 185 EL1 of human CRF2R (189EL1 of human CRF1R) plays a role in receptor function. Specifically, SVG and UI bound with increased affinities to CRF2R when arginine was at position 185EL1 compared to the presence of histidine at this position [107]. In addition, the substitution of Tyr346EL3, Phe347EL3, and Asn348EL3 of CRF1R by Ala significantly reduced the binding affinity of CRF [108]. Furthermore, a recent photocrosslinking study has revealed several crosslinking pairs between UcnI and CRF1R, namely the pairs Gln273EL2-Asp8p, Phe330EL3-Asp8p, Leu329EL3-His12p, Phe330EL3-His12p, Asn333EL3-His12p, Ile345EL3-His12p, Asn348EL3-His12p, Ser349EL3-His12p and Ser349EL3-Leu14p [109].
The pairs Gln273EL2-Asp8p, Asn348EL3-His12p, and Ile345EL3-His12p are also observed in the cryo-EM structure of CRF1R [100]. However, other pairs of crosslinking were not consistent with the Cryo-EM structure. Based on these results, the authors suggested that several regions of receptor containing crosslinked amino acids are subjected to large conformational changes during activation of the receptor, in contrast to the cryo-EM structure of receptor in complex with ligand and G-protein, which represents the most stable conformational active state [100].
In addition to the interaction of EL1 and EL2 with ligands, these regions contain two Cys, which are highly conserved among GPCRs, and form a disulfide bond that connects these loops, playing an important functional role [110,111]. Mutation of Cys188EL1 and/or the Cys258EL2 of CRF1R to different amino acids broke this disulfide bond and largely decreased the binding of CRF [97].
The CRF also interacts with TM helices of CRF1R, excluding TM4 [105]. Specifically, the peptide enters the top of the receptor core, and forms interactions with the top of TM1 and TM2/EL1. As the peptide goes deeper into the receptor, it forms additional interactions with TM1 and EL2 at the outer membrane level and TM3, TM5, and TM7 deep into the pocket (Figure 7). The N-segment of peptide forms a loop in the receptor which orients toward the extracellular side of the receptor and forms polar interactions mainly with TM6, and secondarily with TM5 and EL2 residues. Residues at positions 7 to 9 of the N-segment of CRF have been characterized to play a crucial role in peptide’s agonist activity [112]. Specifically, Ser7P forms an H-bond with the backbone of Phe3316.57b and Tyr3276.53b of the receptor (Figure 7). In addition, Leu8P forms hydrophobic interactions in the hydrophobic region of the receptor consisting of Tyr3276.53b, Phe2033.44b, Ile2775.44b, and Met2765.43b and Asp9P forms an H-bond interaction with Asn2735.40b, among others (Figure 7).
In addition to the structural information of CRF1R in complex with CRF, a recent Cryo-EM study has revealed the structure of both CRF1R and CRF2R obtained in complex with UcnI and Gs protein (Figure 8A and Figure 8B, respectively) [100]. UcnI interacts with the ELs and TM helices of CRF receptors, excluding TM4. Similarly, to the CRF peptide, UcnI enters the top of the receptor core, and forms interactions with the top of TM1 and TM2/EL1 of both receptors. The N-terminus of UcnI loops back up inside the receptors and forms interactions mainly with TM7. The amino acids 6 to 8 of the N-segment of UcnI have been shown to play a crucial role in the activity of the peptides. Specifically, Ser6p of UcnI forms an H-bond with Asn3487.42b of CRF1R (Figure 8C), which is also observed in the structure of CRF2R (Ser6p–3447.42b interaction) (Figure 8D). Moreover, Asp8p interacts with Tyr2725.39b and Gln2735.40b of CRF1R, similarly to its mode of interaction with Tyr2685.39b and Gln2695.40b in CRF2R (Figure 8C,D). An additional important feature of Asp8p, as revealed in the cryo-EM structure of CRF receptors, is its intramolecular electrostatic interaction with the Arg15p of UcnI, thus stabilizing the bound peptide conformation [100]. In the structure of the CRF1R-UcnI complex, Ile7p interacts with a hydrophobic region consisting of Tyr3276.53b, Phe2033.44b, Ile2775.44b, and Met2765.43, whereas the same hydrophobic region in CRF2R has a differentiation at position 2765.43b, where there is an isoleucine instead of a methionine (Figure 8C,D).

5. The Two-Step Model of Ligand–Receptor Interaction

The CRF peptides interact with the CRF receptors according to a two-step model. In the first step, the C-segment of the peptides interacts with the ECD of the receptors. This interaction orients the peptides in such a way that their N-segment residues interact with the J-domain of the receptors in the second step. This model was built based on the experimental results from the study of Hoare et al. [113]. Previous pharmacological studies have also shown the interaction between the ECD of CRF receptors, and the C-segment of the peptides [72,74,90,94]. In addition, other studies have shown the interaction between the amino acids of the N-segment of CRF peptides with the J-domain of CRF receptors [101,106,114,115]. The last interaction is responsible for receptor activation [116]. Specifically, Nielsen et al. have shown that replacement of the ECD of CRF1R with the first 16 N-segment residues of CRF constitutively activated the receptor because it mimicked the first-step interaction and allowed the important for receptor activation second-step interaction [116]. The constitutive activity of this chimeric receptor, which lacks the ECD, was not blocked by astressin which binds to the ECD of CRF1R [116]. In contrast, the small non-peptide allosteric antagonist, antalarmin, which binds to the TMs of CRF1R, decreased the constitutive activity of this construct. Interestingly, the Ala mutation of Leu8p in the tethered N-segment of CRF abolished the constitutive activation of the chimeric receptor, suggesting the important role of this peptide residue in receptor–ligand interactions [116].

6. Structural Basis of Receptor Activation

The activation of CRF1R involves a complex and finely tuned set of structural interactions between different receptor residues, between peptide and receptor amino acids, and between receptor and G-protein residues. The high-resolution structures of CRF receptors provide a detailed view of these interactions. The receptor core, inclusive of all loops and the Gas domain, is resolved with great precision. This clarity allows for the accurate placement of side-chain rotamers and a deep understanding of the receptor’s structural conformation upon ligand binding.
Interactions of agonist peptides with the transmembrane helices (TM1–TM7) and extracellular regions of the receptor are characterized by a combination of hydrogen bonds, hydrophobic interactions, as described previously, and structural water molecules that stabilize the peptides within the receptor. The presence of structural water molecules in CRF receptor structures contributes significantly to the stability and specificity of peptide binding, by linking key residues such as E3527.46b and Y1953.36b that are essential for maintaining the receptors in an active conformation [105]. As revealed by the comparison of the inactive and active crystal structures of CRF receptors, their activation is associated with conformational changes, such as the reorganization of EL2, which includes an upward movement of TM4 and TM5 that repositions both EL2 and IL2 [105]. In addition, the computational data from a recent pharmacological study have proposed a movement of TM3 and TM5 of CRF1R during its activation [102]. Strengthening the interface between TM3 and TM5 by appropriately mutating their amino acids further stabilized the inactive state of the receptor [102].
The CRF1R-Gas protein interface is another critical aspect of receptor activation. The receptor makes extensive contacts with Gas, predominantly through hydrophobic interactions and hydrogen bonds across TM2, TM3, IL2, TM5, TM6, and the junction of TM7 and helix 8 [105]. Helix 8 is a structure of receptors located intracellularly after the end of TM7 and connects the TMs of the receptor with its C-domain. Specifically, Y391aH5 of Gas protein participates in hydrophobic interactions with H1552.50b, L2133.54b, Y2123.53b, and R1512.46b of CRF1R, in addition to the H-bond networks in which participates as described in the study of Liang et al. (Figure 9) [105]. E392aH5 of Gas protein interacts through a hydrogen bond with the main chain of S3688.48b located in helix 8, and L393aH5 of Gas protein participates in hydrophobic interactions with A3156.41b and L2945.61b of the receptor (Figure 9). In addition, the main chain oxygen of L394aH5 of Gas protein interacts with receptor residues, K3116.37b forming a hydrogen bond and L2985.65b by hydrophobic interactions. Residues from the alpha helix 5 of the Gas protein, and specifically, Q384aH5 make H-bond interactions with K2975.64b, K3345.64b, T258IL2 and L2553.58b. In addition, T220IL2 interacts with main chain atoms of I2173.58b, while I383aH5 and R380aH5 participate in hydrophobic interactions with T220IL2 and Y221IL2. Y221IL2 also forms a hydrogen bond with the main chain oxygen of F376aH5 in the Gas protein, whereas Q35aHN interacts with R2274.41b. Moreover, R385aH5 participates in hydrophobic interactions with K2975.64b (Figure 9). The superscripts IL1, IL2, and IL3 of receptor residues represent the intracellular loops (ILs) of the receptor, in which these residues are located. The superscript aH5 represents the alpha helix 5 of the Gas protein. Similar to CRF1R, the α5 helix of the Gαs protein extensively interacts with TM2, TM3, TM5, TM6, IL2, and IL3, and the TM7-H8 junction of the CRF2R, highlighting the extensive interface for G-protein coupling [117]. Specifically, Y391 in the Gas protein binds to a sub-pocket formed by R1482.46b, H1522.50b, and E2053.50b, Y2083.53b, L2093.54b of CRF2R. Other interface residues in the α5 helix of the Gαs protein include E392, forming polar contacts with K3106.40b and N3638.47b of the receptor, and Q390 forming a hydrogen bond with R1482.46b of the receptor. In addition, the C-terminal L394 of the Gas protein forms a charge interaction with K3076.37b of the TM6 of the receptor.
The IL2 of CRF receptors plays an important role in their interaction with different G-proteins, consistent with the ability of subfamily B1 GPCRs to engage with multiple G-proteins and their related signaling pathways. Conformational changes in IL2 of CRF receptors are key in differentiating their interactions with various G-proteins (Gs, G11, and Go). Specifically, the α5 helices of different G-proteins interact differently with the receptor [117]. The sharp kink in the middle of TM6 of CRF2R facilitates the formation of an open G-protein-binding pocket, allowing for the accommodation of the relatively large C-termini of the α5 helix of Gα subunits, particularly Gs [117].
The structural–functional knowledge derived from CRF receptors can provide valuable insights into the activation mechanisms of subfamily B1 GPCRs. Common structural features of CRF receptors and subfamily B1 GPCRs have been extensively studied. The formation of a cytoplasmic cavity by three intracellular loops (ILs) for G-protein coupling, the interaction of the α5 helix of Gαs proteins with TM2, TM3, and IL2, which is a highly conserved feature across subfamily B1 GPCRs, suggest a common mechanism in Gs protein coupling for these receptors. The ECD and ELs of subfamily B1 GPCRs, including CRF receptors, show the ability to regulate the binding of different peptides, indicating that the general principles of ligand recognition and binding are shared across this subfamily. Similar to CRF receptors, subfamily B1 GPCRs undergo conformational changes upon activation, such as the outward movement of TM6/ECL3/TM7, which are variable among different receptors but follow a common trend [57].
However, there are structural aspects of activation within the CRF receptor family which are not necessarily extrapolatable. These are the specific interactions at the residue level, and the diverse G-protein engagement. Differences between receptors of this family are likely reflective of the distinct dynamics of side chains and backbone conformations within individual receptors. Thus, although CRF receptors provide a valuable framework for understanding subfamily B1 GPCR activation, variations in ligand recognition, G-protein engagement, and structural dynamics of these receptors highlight the importance of studying each of them individually.

7. Molecular Mechanisms of Ligand Selectivity

7.1. Selectivity of Non-Peptide Antagonists

The crystal structure of the inactive state of CRF1R has revealed that a layer formed by the side chains of several residues, including His1993.40b and Met2765.43b, is located just above the bound CP-376395 [103]. Interestingly, although His1993.40b and Met2765.43b do not interact with the ligand, their mutation to the corresponding residues of CRF2R (V1953.40b and I2725.43b, respectively) largely reduces non-peptide antagonist binding to CRF1R [118]. The non-conserved His1993.40 and Met2765.43 interact with the conserved Tyr3276.53b and Phe2033.44b, possibly affecting the positioning of these two aromatic residues. It has been proposed that during ligand binding and dissociation Tyr3276.53b and Phe2033.44b change rotameric states, such as to allow the entrance and exit of the small non-peptide ligands from the CRF1R [98]. In contrast, the mutation of His1993.40b and Met2765.43b of CRF1R to alanine did not affect non-peptide ligand binding, since this small well-tolerated amino acid does not largely affect the conformation of receptor [102,119]. These conformational changes provide a theoretical rationale for the CRF1R-selectivity of non-peptide ligands given that all receptor residues directly interacting with these molecules are completely conserved in CRF2R. It is possible that in CRF2R, the corresponding Val1953.40b and Ile2725.43b restrict the aromatic residues, which correspond to Tyr3276.53b and Phe2033.44b of CRF1R, to a conformation that prohibits the access of non-peptide ligands to their binding sites [98].

7.2. Selectivity of Peptide Agonists

Residues at the N-segment of the CRF and related peptides play an important role in CRF receptor selectivity. Specifically, the replacement of the motif Thr11p-Phe12p-His13p of CRF with the corresponding motifs of UcnII and UcnIII, Pro11P-Ile12P-Gly13P and Pro11P-Thr12P-Asn13P, respectively, conferred to peptide CRF2R-selectivity [120]. Supportive evidence for the structural importance of the Thr11p-Phe12p-His13p motif of CRF in receptor selectivity is provided by the study of Isfort et al. [121]. Specifically, Phe, Leu, Ile, Thr, Gln, His and Tyr at position 12 and Phe, Gln, Trp, Tyr, Val, Ile, Leu and 2-naphthylalanine at position 13 are the preferable substitutions for CRF2R selectivity. In addition to the N-region, amino acids at the C-region of the CRF and related peptides are significant for CRF receptor selectivity. Importantly, Arg35p and the acidic residue (Asp or Glu) at position 39 are not conserved in the CRF2R-selective UcnII and Ucn III, which have an Ala at these positions, suggesting that they could play an important role in the selective binding of these peptides to CRF2R. Simultaneous replacement of Arg34p (Arg35p of CRF) and Asp38p (Glu39p of CRF) of SVG by the corresponding Ala35p and Ala39p of UcnII increased the CRF2R-selectivity obtained after the substitution of Ser10p of SVG (which corresponds to Thr11p of CRF) by the corresponding Pro11p of UcnII and UcnIII, [122]. Similarly, the corresponding substitutions in the UcnI increased CRF2R-selectivity, whereas in the CRF it resulted in a decrease in binding to both receptor subtypes (EC50 > 100 nM for CRF2R and EC50 > 1000 nM for CRF1R), suggesting that UcnI, SVG, and CRF bind to CRF receptor with slightly different modes, which might be differentially affected by the same modification of peptides [122].
The molecular determinants of peptide selectivity could be determined by examining the existing structures of CRF1R and CRF2R and comparing the amino acids at positions 11-13 of the N-segment of CRF (Thr11P-Phe12P-His13P), UcnI (Thr11P-Phe12P-His13P), UcnII (Pro11P-Ile12P-Gly13P) and UcnIII (Pro11P-Thr12P-Asn13P) bound to the two CRF receptor types. Given that Pro11P is an alpha-helix breaker, these residues abate the α-helicity, leading to an impairment of binding to CRF1R [120]. Moreover, a comparison of the electrostatic surface potentials of the ECD for both CRF receptors indicates that Arg35P (positively charged amino acid) present in CRF and UcnI is compatible for interaction with Glu104 of CRF1R as well as the corresponding Pro100 in CRF2R. In contrast, UcnII and UcnIII have Ala35P, the hydrophobic surface of which may interact only with Pro100 of CRF2R and not with Glu104 of CRF1R, which could provide an additional determinant of peptide selectivity [123].
Receptor selectivity could also be attributed to modifications of different sets of amino acids at positions 30, 31, 33, 34, and 35 and/or to a change in the conformation of peptide by introducing lactam bridges at different positions of CRF [124,125]. For example, Rivier et al. have shown that the introduction of a Glu32p-Lys35p lactam bridge into a CRF analog with the eleven N-terminal residues deleted (creating the analog, cyclo(32-35)[DPhe12,Nle21,38,Glu32,Lys35]- hCRF(12-41)) yielded a CRF2R-selective ligand [124]. In marked contrast, the introduction of a Glu30p-Lys33p lactam bridge in a CRF truncated analog created the non-selective astressin ([cyclo(30-33)[DPhe(12),Nle(21),Glu(30), Lys(33),Nle(38)]hCRF((12-41))]) [125]. Furthermore, salt bridges that are formed in the Glu31p–Glu34p region of CRF analogs could play a crucial role in peptide selectivity. The CRF analog Stressin1-A is a CRF1R-selective peptide with quite similar sequence as CRF (cyclo(31-34)[DPhe12,Nle21,38,Glu31,Lys34]Ac-hCRF(4-41)). In Stressin1-A, residues Glu31p and Lys34p form a lactam bridge resulting in a 130-fold selectivity increase towards the CRF1R [126]. Interestingly, the linear counterpart of stresin-1A, (linear[DPhe12, Nle21,38,Glu31,Lys34]-Ac-hCRF(4-41)), which has Glu and Lys at positions 31 and 34, respectively, displayed CRF1R selectivity, similar to stressin-1 A, whereas the linear analog [DPhe12,Nle21,38,Glu30,Lys33]-Ac-hCRF(4-41), which has Glu and Lys at positions 30 and 33, respectively, was non-selective [126].
Differential truncation of peptides could also confer the selective binding of peptides to the two CRF receptor types. Thus, although the introduction of a cyclic constraint between Glu32p and Lys35p of a truncated CRF analog created the CRF2R-selective ligand (cyclo(32-35)[DPhe12, Nle21,38,Glu32, Lys35]- hCRF(12-41)), the same modification in a lengthier CRF analog created the non-selective ligand cyclo(32-35)[DPhe12,Nle21,38, Glu32,Lys 35]-Ac-hCRF(4-41) [124,126].

8. Concluding Remarks

In this exploration of CRF peptides and their receptors, we have delved into the intricate structural and functional nuances that govern their roles in diverse physiologic and pathologic states. The elucidation of the crystal structures of CRF receptors, particularly CRF1R and CRF2R, has been pivotal in enhancing our understanding of their activation mechanisms and ligand specificity.
The detailed insights into the interaction dynamics between CRF peptides and their receptors underscore the sophistication of ligand–receptor binding, ligand selectivity for the CRF1R and CRF2R, and activation processes. In addition, these interactions, characterized by specific amino acid interactions, and conformational changes, highlight the intricacy of G-protein-coupled-receptor (GPCR) signaling.
Moreover, the structural–functional insights gained from CRF receptors provide a valuable framework for understanding the function of the broader subfamily B1 GPCRs. While certain aspects of CRF receptor activation and ligand binding can be generalized to other receptors in this subfamily, the unique features of each receptor must be appreciated. This knowledge underscores the potential for developing targeted therapies that harness the specific characteristics of each receptor within this subfamily.
The revelation of the molecular mechanisms underlying receptor activation and ligand selectivity has profound implications for therapeutic applications. Understanding the structural basis of receptor activation opens avenues for the development of novel drugs targeting CRF receptors. These drugs hold significant promise in the treatment of a range of conditions, from psychiatric disorders such as depression and anxiety, to those of the cardiovascular, gastrointestinal, and immune systems. As we continue to deepen our understanding of these molecular mechanisms, we edge closer to unlocking the full therapeutic potential of targeting CRF receptors and their related signaling pathways.

Author Contributions

Conceptualization, M.-T.M. and G.L.; software, M.-T.M. and V.P.; writing—original draft preparation, M.-T.M., V.P., V.K., M.V. and G.L.; writing—review and editing, M.-T.M., V.P., V.K., G.P.C., M.V. and G.L.; funding acquisition, G.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Special Account for Research Funds of University of Crete (SARF UoC) grant 10674 (George Liapakis).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Liapakis, G.; Venihaki, M.; Margioris, A.; Grigoriadis, D.; Gkountelias, K. Members of CRF family and their receptors: From past to future. Curr. Med. Chem. 2011, 18, 2583–2600. [Google Scholar] [CrossRef] [PubMed]
  2. Vaughan, J.; Donaldson, C.; Bittencourt, J.; Perrin, M.H.; Lewis, K.; Sutton, S.; Chan, R.; Turnbull, A.V.; Lovejoy, D.; Rivier, C.; et al. Urocortin, a mammalian neuropeptide related to fish urotensin I and to corticotropin-releasing factor. Nature 1995, 378, 287–292. [Google Scholar] [CrossRef] [PubMed]
  3. Montecucchi, P.C.; Henschen, A. Amino acid composition and sequence analysis of sauvagine, a new active peptide from the skin of Phyllomedusa sauvagei. Int. J. Pept. Protein Res. 1981, 18, 113–120. [Google Scholar] [CrossRef] [PubMed]
  4. Lederis, K.; Letter, A.; McMaster, D.; Moore, G.; Schlesinger, D. Complete amino acid sequence of urotensin I, a hypotensive and corticotropin-releasing neuropeptide from Catostomus. Science 1982, 218, 162–165. [Google Scholar] [CrossRef] [PubMed]
  5. Reyes, T.M.; Lewis, K.; Perrin, M.H.; Kunitake, K.S.; Vaughan, J.; Arias, C.A.; Hogenesch, J.B.; Gulyas, J.; Rivier, J.; Vale, W.W.; et al. Urocortin II: A member of the corticotropin-releasing factor (CRF) neuropeptide family that is selectively bound by type 2 CRF receptors. Proc. Natl. Acad. Sci. USA 2001, 98, 2843–2848. [Google Scholar] [CrossRef] [PubMed]
  6. Hsu, S.Y.; Hsueh, A.J. Human stresscopin and stresscopin-related peptide are selective ligands for the type 2 corticotropin-releasing hormone receptor. Nat. Med. 2001, 7, 605–611. [Google Scholar] [CrossRef]
  7. Lewis, K.; Li, C.; Perrin, M.H.; Blount, A.; Kunitake, K.; Donaldson, C.; Vaughan, J.; Reyes, T.M.; Gulyas, J.; Fischer, W.; et al. Identification of urocortin III, an additional member of the corticotropin-releasing factor (CRF) family with high affinity for the CRF2 receptor. Proc. Natl. Acad. Sci. USA 2001, 98, 7570–7575. [Google Scholar] [CrossRef]
  8. Potter, E.; Behan, D.P.; Fischer, W.H.; Linton, E.A.; Lowry, P.J.; Vale, W.W. Cloning and characterization of the cDNAs for human and rat corticotropin releasing factor-binding proteins. Nature 1991, 349, 423–426. [Google Scholar] [CrossRef]
  9. Behan, D.P.; Khongsaly, O.; Ling, N.; De Souza, E.B. Urocortin interaction with corticotropin-releasing factor (CRF) binding protein (CRF-BP): A novel mechanism for elevating ‘free’ CRF levels in human brain. Brain Res. 1996, 725, 263–267. [Google Scholar]
  10. Vale, W.; Spiess, J.; Rivier, C.; Rivier, J. Characterization of a 41-residue ovine hypothalamic peptide that stimulates secretion of corticotropin and β-endorphin. Science 1981, 213, 1394–1397. [Google Scholar] [CrossRef]
  11. Chrousos, G.P. The hypothalamic-pituitary-adrenal axis and immune-mediated inflammation. N. Engl. J. Med. 1995, 332, 1351–1362. [Google Scholar] [CrossRef]
  12. Owens, M.J.; Nemeroff, C.B. Physiology and pharmacology of corticotropin-releasing factor. Pharmacol. Rev. 1991, 43, 425–473. [Google Scholar]
  13. Gold, P.W.; Chrousos, G.P. Organization of the stress system and its dysregulation in melancholic and atypical depression: High vs. low CRH/NE states. Mol. Psychiatry 2002, 7, 254–275. [Google Scholar] [CrossRef]
  14. Keck, M.E.; Holsboer, F. Hyperactivity of CRH neuronal circuits as a target for therapeutic interventions in affective disorders. Peptides 2001, 22, 835–844. [Google Scholar] [CrossRef]
  15. Reul, J.M.; Holsboer, F. Corticotropin-releasing factor receptors 1 and 2 in anxiety and depression. Curr. Opin. Pharmacol. 2002, 2, 23–33. [Google Scholar] [CrossRef]
  16. Dautzenberg, F.M.; Kilpatrick, G.J.; Hauger, R.L.; Moreau, J. Molecular biology of the CRH receptors—In the mood. Peptides 2001, 22, 753–760. [Google Scholar] [CrossRef] [PubMed]
  17. Behan, D.P.; Heinrichs, S.C.; Troncoso, J.C.; Liu, X.J.; Kawas, C.H.; Ling, N.; De Souza, E.B. Displacement of corticotropin releasing factor from its binding protein as a possible treatment for Alzheimer’s disease. Nature 1995, 378, 284–287. [Google Scholar] [CrossRef] [PubMed]
  18. De Souza, E.B.; Whitehouse, P.J.; Kuhar, M.J.; Price, D.L.; Vale, W.W. Reciprocal changes in corticotropin-releasing factor (CRF)-like immunoreactivity and CRF receptors in cerebral cortex of Alzheimer’s disease. Nature 1986, 319, 593–595. [Google Scholar] [CrossRef] [PubMed]
  19. Makrigiannakis, A.; Zoumakis, E.; Kalantaridou, S.; Coutifaris, C.; Margioris, A.N.; Coukos, G.; Rice, K.C.; Gravanis, A.; Chrousos, G.P. Corticotropin-releasing hormone promotes blastocyst implantation and early maternal tolerance. Nat. Immunol. 2001, 2, 1018–1024. [Google Scholar] [CrossRef] [PubMed]
  20. Chrousos, G.P.; Torpy, D.J.; Gold, P.W. Interactions between the hypothalamic-pituitary-adrenal axis and the female reproductive system: Clinical implications. Ann. Intern. Med. 1998, 129, 229–240. [Google Scholar] [CrossRef] [PubMed]
  21. Martinez, V.; Rivier, J.; Wang, L.; Tache, Y. Central injection of a new corticotropin-releasing factor (CRF) antagonist, astressin, blocks CRF- and stress-related alterations of gastric and colonic motor function. J. Pharmacol. Exp. Ther. 1997, 280, 754–760. [Google Scholar] [PubMed]
  22. Orth, D.N. Corticotropin-releasing hormone in humans. Endocr. Rev. 1992, 13, 164–191. [Google Scholar] [PubMed]
  23. Parkes, D.G.; Weisinger, R.S.; May, C.N. Cardiovascular actions of CRH and urocortin: An update. Peptides 2001, 22, 821–827. [Google Scholar] [CrossRef] [PubMed]
  24. Venihaki, M.; Dikkes, P.; Carrigan, A.; Karalis, K.P. Corticotropin-releasing hormone regulates IL-6 expression during inflammation. J. Clin. Investig. 2001, 108, 1159–1166. [Google Scholar] [CrossRef] [PubMed]
  25. Venihaki, M.; Majzoub, J.A. Animal models of CRH deficiency. Front. Neuroendocrinol. 1999, 20, 122–145. [Google Scholar] [CrossRef] [PubMed]
  26. Wang, L.; Martinez, V.; Rivier, J.E.; Tache, Y. Peripheral urocortin inhibits gastric emptying and food intake in mice: Differential role of CRF receptor 2. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2001, 281, R1401–R1410. [Google Scholar] [CrossRef]
  27. Martinez, V.; Wang, L.; Rivier, J.E.; Vale, W.; Tache, Y. Differential actions of peripheral corticotropin-releasing factor (CRF), urocortin II, and urocortin III on gastric emptying and colonic transit in mice: Role of CRF receptor subtypes 1 and 2. J. Pharmacol. Exp. Ther. 2002, 301, 611–617. [Google Scholar] [CrossRef]
  28. Coste, S.C.; Kesterson, R.A.; Heldwein, K.A.; Stevens, S.L.; Heard, A.D.; Hollis, J.H.; Murray, S.E.; Hill, J.K.; Pantely, G.A.; Hohimer, A.R.; et al. Abnormal adaptations to stress and impaired cardiovascular function in mice lacking corticotropin-releasing hormone receptor-2. Nat. Genet. 2000, 24, 403–409. [Google Scholar] [CrossRef]
  29. Coste, S.C.; Quintos, R.F.; Stenzel-Poore, M.P. Corticotropin-releasing hormone-related peptides and receptors: Emergent regulators of cardiovascular adaptations to stress. Trends Cardiovasc. Med. 2002, 12, 176–182. [Google Scholar] [CrossRef]
  30. Karalis, K.; Sano, H.; Redwine, J.; Listwak, S.; Wilder, R.L.; Chrousos, G.P. Autocrine or paracrine inflammatory actions of corticotropin-releasing hormone in vivo. Science 1991, 254, 421–423. [Google Scholar] [CrossRef]
  31. Koob, G.F.; Bloom, F.E. Corticotropin-releasing factor and behavior. Fed. Proc. 1985, 44, 259–263. [Google Scholar]
  32. Kempuraj, D.; Papadopoulou, N.G.; Lytinas, M.; Huang, M.; Kandere-Grzybowska, K.; Madhappan, B.; Boucher, W.; Christodoulou, S.; Athanassiou, A.; Theoharides, T.C. Corticotropin-releasing hormone and its structurally related urocortin are synthesized and secreted by human mast cells. Endocrinology 2004, 145, 43–48. [Google Scholar] [CrossRef]
  33. Dedic, N.; Chen, A.; Deussing, J.M. The CRF Family of Neuropeptides and their Receptors—Mediators of the Central Stress Response. Curr. Mol. Pharmacol. 2018, 11, 4–31. [Google Scholar] [CrossRef]
  34. Sanders, J.; Nemeroff, C. The CRF System as a Therapeutic Target for Neuropsychiatric Disorders. Trends Pharmacol. Sci. 2016, 37, 1045–1054. [Google Scholar] [CrossRef] [PubMed]
  35. Makrigiannakis, A.; Vrekoussis, T.; Zoumakis, E.; Navrozoglou, I.; Kalantaridou, S.N. CRH Receptors in Human Reproduction. Curr. Mol. Pharmacol. 2018, 11, 81–87. [Google Scholar] [CrossRef] [PubMed]
  36. Basman, C.; Agrawal, P.; Knight, R.; Saravolatz, L.; McRee, C.; Chen-Scarabelli, C.; Narula, J.; Scarabelli, T. Cardioprotective Utility of Urocortin in Myocardial Ischemia- Reperfusion Injury: Where do We Stand? Curr. Mol. Pharmacol. 2018, 11, 32–38. [Google Scholar] [CrossRef] [PubMed]
  37. Pagan-Busigo, J.E.; Lopez-Carrasquillo, J.; Appleyard, C.B.; Torres-Reveron, A. Beyond depression and anxiety; a systematic review about the role of corticotropin-releasing hormone antagonists in diseases of the pelvic and abdominal organs. PLoS ONE 2022, 17, e0264909. [Google Scholar] [CrossRef]
  38. Arborelius, L.; Owens, M.J.; Plotsky, P.M.; Nemeroff, C.B. The role of corticotropin-releasing factor in depression and anxiety disorders. J. Endocrinol. 1999, 160, 1–12. [Google Scholar] [CrossRef]
  39. Muller, M.B.; Zimmermann, S.; Sillaber, I.; Hagemeyer, T.P.; Deussing, J.M.; Timpl, P.; Kormann, M.S.; Droste, S.K.; Kuhn, R.; Reul, J.M.; et al. Limbic corticotropin-releasing hormone receptor 1 mediates anxiety-related behavior and hormonal adaptation to stress. Nat. Neurosci. 2003, 6, 1100–1107. [Google Scholar] [CrossRef] [PubMed]
  40. Timpl, P.; Spanagel, R.; Sillaber, I.; Kresse, A.; Reul, J.M.; Stalla, G.K.; Blanquet, V.; Steckler, T.; Holsboer, F.; Wurst, W. Impaired stress response and reduced anxiety in mice lacking a functional corticotropin-releasing hormone receptor 1. Nat. Genet. 1998, 19, 162–166. [Google Scholar] [CrossRef] [PubMed]
  41. Zobel, A.W.; Nickel, T.; Kunzel, H.E.; Ackl, N.; Sonntag, A.; Ising, M.; Holsboer, F. Effects of the high-affinity corticotropin-releasing hormone receptor 1 antagonist R121919 in major depression: The first 20 patients treated. J. Psychiatr. Res. 2000, 34, 171–181. [Google Scholar] [CrossRef]
  42. Ising, M.; Zimmermann, U.S.; Kunzel, H.E.; Uhr, M.; Foster, A.C.; Learned-Coughlin, S.M.; Holsboer, F.; Grigoriadis, D.E. High-affinity CRF1 receptor antagonist NBI-34041: Preclinical and clinical data suggest safety and efficacy in attenuating elevated stress response. Neuropsychopharmacology 2007, 32, 1941–1949. [Google Scholar] [CrossRef]
  43. Zorrilla, E.P.; Koob, G.F. Progress in corticotropin-releasing factor-1 antagonist development. Drug Discov. Today 2010, 15, 371–383. [Google Scholar] [CrossRef]
  44. Valdez, G.R.; Inoue, K.; Koob, G.F.; Rivier, J.; Vale, W.; Zorrilla, E.P. Human urocortin II: Mild locomotor suppressive and delayed anxiolytic-like effects of a novel corticotropin-releasing factor related peptide. Brain Res. 2002, 943, 142–150. [Google Scholar] [CrossRef] [PubMed]
  45. Valdez, G.R.; Zorrilla, E.P.; Rivier, J.; Vale, W.W.; Koob, G.F. Locomotor suppressive and anxiolytic-like effects of urocortin 3, a highly selective type 2 corticotropin-releasing factor agonist. Brain Res. 2003, 980, 206–212. [Google Scholar] [CrossRef] [PubMed]
  46. Venihaki, M.; Sakihara, S.; Subramanian, S.; Dikkes, P.; Weninger, S.C.; Liapakis, G.; Graf, T.; Majzoub, J.A. Urocortin III, A Novel Murine Brain Neuropeptide of the Corticotropin-Releasing Hormone Family: Modulation by Stress and Attenuation of Some Anxiety-Like Behaviors. J. Neuroendocrinol. 2004, 16, 411–422. [Google Scholar] [CrossRef] [PubMed]
  47. Venkatasubramanian, S.; Griffiths, M.E.; McLean, S.G.; Miller, M.R.; Luo, R.; Lang, N.N.; Newby, D.E. Vascular effects of urocortins 2 and 3 in healthy volunteers. J. Am. Heart Assoc. 2013, 2, e004267. [Google Scholar] [CrossRef] [PubMed]
  48. Bale, T.L.; Hoshijima, M.; Gu, Y.; Dalton, N.; Anderson, K.R.; Lee, K.F.; Rivier, J.; Chien, K.R.; Vale, W.W.; Peterson, K.L. The cardiovascular physiologic actions of urocortin II: Acute effects in murine heart failure. Proc. Natl. Acad. Sci. USA 2004, 101, 3697–3702. [Google Scholar] [CrossRef] [PubMed]
  49. Tache, Y.; Larauche, M.; Yuan, P.Q.; Million, M. Brain and Gut CRF Signaling: Biological Actions and Role in the Gastrointestinal Tract. Curr. Mol. Pharmacol. 2018, 11, 51–71. [Google Scholar] [CrossRef]
  50. Rivier, J.E. Prospective Clinical Applications of CRF Peptide Antagonists. Curr. Mol. Pharmacol. 2017, 10, 264–269. [Google Scholar] [CrossRef]
  51. Torres-Reverón, A.; Rivera-Lopez, L.L.; Flores, I.; Appleyard, C.B. Antagonizing the corticotropin releasing hormone receptor 1 with antalarmin reduces the progression of endometriosis. PLoS ONE 2018, 13, e0197698. [Google Scholar] [CrossRef] [PubMed]
  52. Sarafoglou, K.; Barnes, C.N.; Huang, M.; Imel, E.A.; Madu, I.J.; Merke, D.P.; Moriarty, D.; Nakhle, S.; Newfield, R.S.; Vogiatzi, M.G.; et al. Tildacerfont in Adults with Classic Congenital Adrenal Hyperplasia: Results from Two Phase 2 Studies. J. Clin. Endocrinol. Metab. 2021, 106, e4666–e4679. [Google Scholar] [CrossRef]
  53. Harmar, A.J. Family-B G-protein-coupled receptors. Genome. Biol. 2001, 2, REVIEWS3013. [Google Scholar] [CrossRef]
  54. Gkountelias, K.; Papadokostaki, M.; Javitch, J.A.; Liapakis, G. Exploring the binding site crevice of a family B G protein-coupled receptor, the type 1 corticotropin releasing factor receptor. Mol. Pharmacol. 2010, 78, 785–793. [Google Scholar] [CrossRef]
  55. Hillhouse, E.W.; Grammatopoulos, D.K. The molecular mechanisms underlying the regulation of the biological activity of corticotropin-releasing hormone receptors: Implications for physiology and pathophysiology. Endocr. Rev. 2006, 27, 260–286. [Google Scholar] [CrossRef]
  56. Grammatopoulos, D.K.; Ourailidou, S. CRH Receptor Signalling: Potential Roles in Pathophysiology. Curr. Mol. Pharmacol. 2017, 10, 296–310. [Google Scholar] [CrossRef] [PubMed]
  57. Cong, Z.; Liang, Y.-L.; Zhou, Q.; Darbalaei, S.; Zhao, F.; Feng, W.; Zhao, L.; Xu, H.E.; Yang, D.; Wang, M.-W. Structural perspective of class B1 GPCR signaling. Trends Pharmacol. Sci. 2022, 43, 321–334. [Google Scholar] [CrossRef] [PubMed]
  58. Bockaert, J.; Pin, J.P. Molecular tinkering of G protein-coupled receptors: An evolutionary success. EMBO J. 1999, 18, 1723–1729. [Google Scholar] [CrossRef]
  59. Islam, M.R.; Teleb, M.; Karageorgos, V.; Sakellaris, S.; Papadopoulos, M.; Pirmettis, I.; Fronczek, F.R.; Liapakis, G.; Fahmy, H. Design, synthesis, structural optimization, SAR, in silico prediction of physicochemical properties and pharmacological evaluation of novel & potent thiazolo[4,5-d]pyrimidine corticotropin releasing factor (CRF) receptor antagonists. Eur. J. Pharm. Sci. 2022, 169, 106084. [Google Scholar] [CrossRef]
  60. Grigoriadis, D.E.; Liu, X.J.; Vaughn, J.; Palmer, S.F.; True, C.D.; Vale, W.W.; Ling, N.; De Souza, E.B. 125I-Tyro-sauvagine: A novel high affinity radioligand for the pharmacological and biochemical study of human corticotropin-releasing factor 2 α receptors. Mol. Pharmacol. 1996, 50, 679–686. [Google Scholar]
  61. Dautzenberg, F.M.; Py-Lang, G.; Higelin, J.; Fischer, C.; Wright, M.B.; Huber, G. Different binding modes of amphibian and human corticotropin-releasing factor type 1 and type 2 receptors: Evidence for evolutionary differences. J. Pharmacol. Exp. Ther. 2001, 296, 113–120. [Google Scholar]
  62. Rivier, J.; Spiess, J.; Vale, W. Characterization of rat hypothalamic corticotropin-releasing factor. Proc. Natl. Acad. Sci. USA 1983, 80, 4851–4855. [Google Scholar] [CrossRef]
  63. Shibahara, S.; Morimoto, Y.; Furutani, Y.; Notake, M.; Takahashi, H.; Shimizu, S.; Horikawa, S.; Numa, S. Isolation and sequence analysis of the human corticotropin-releasing factor precursor gene. EMBO J 1983, 2, 775–779. [Google Scholar] [CrossRef]
  64. Patthy, M.; Horvath, J.; Mason-Garcia, M.; Szoke, B.; Schlesinger, D.H.; Schally, A.V. Isolation and amino acid sequence of corticotropin-releasing factor from pig hypothalami. Proc. Natl. Acad. Sci. USA 1985, 82, 8762–8766. [Google Scholar] [CrossRef]
  65. Okawara, Y.; Morley, S.D.; Burzio, L.O.; Zwiers, H.; Lederis, K.; Richter, D. Cloning and sequence analysis of cDNA for corticotropin-releasing factor precursor from the teleost fish Catostomus commersoni. Proc. Natl. Acad. Sci. USA 1988, 85, 8439–8443. [Google Scholar] [CrossRef] [PubMed]
  66. Stenzel-Poore, M.P.; Heldwein, K.A.; Stenzel, P.; Lee, S.; Vale, W.W. Characterization of the genomic corticotropin-releasing factor (CRF) gene from Xenopus laevis: Two members of the CRF family exist in amphibians. Mol. Endocrinol. 1992, 6, 1716–1724. [Google Scholar]
  67. Donaldson, C.J.; Sutton, S.W.; Perrin, M.H.; Corrigan, A.Z.; Lewis, K.A.; Rivier, J.E.; Vaughan, J.M.; Vale, W.W. Cloning and characterization of human urocortin. Endocrinology 1996, 137, 2167–2170, Erratum in Endocrinoogy 1996, 137, 3896. [Google Scholar] [CrossRef] [PubMed]
  68. Grace, C.R.; Perrin, M.H.; Cantle, J.P.; Vale, W.W.; Rivier, J.E.; Riek, R. Common and divergent structural features of a series of corticotropin releasing factor-related peptides. J. Am. Chem. Soc. 2007, 129, 16102–16114. [Google Scholar] [CrossRef] [PubMed]
  69. Ohta, N.; Mochizuki, T.; Hoshino, M.; Jun, L.; Kobayashi, H.; Yanaihara, N. Adrenocorticotropic hormone-releasing activity of urotensin I and its fragments in vitro. J. Pept. Res. 1997, 50, 178–183. [Google Scholar] [CrossRef] [PubMed]
  70. Rivier, J.; Rivier, C.; Vale, W. Synthetic competitive antagonists of corticotropin-releasing factor: Effect on ACTH secretion in the rat. Science 1984, 224, 889–891. [Google Scholar] [CrossRef]
  71. Kornreich, W.D.; Galyean, R.; Hernandez, J.F.; Craig, A.G.; Donaldson, C.J.; Yamamoto, G.; Rivier, C.; Vale, W.; Rivier, J. Alanine series of ovine corticotropin releasing factor (oCRF): A structure-activity relationship study. J. Med. Chem. 1992, 35, 1870–1876. [Google Scholar] [CrossRef]
  72. Mesleh, M.F.; Shirley, W.A.; Heise, C.E.; Ling, N.; Maki, R.A.; Laura, R.P. NMR structural characterization of a minimal peptide antagonist bound to the extracellular domain of the corticotropin-releasing factor1 receptor. J. Biol. Chem. 2007, 282, 6338–6346. [Google Scholar] [CrossRef] [PubMed]
  73. Rijkers, D.T.; Kruijtzer, J.A.; van Oostenbrugge, M.; Ronken, E.; den Hartog, J.A.; Liskamp, R.M. Structure-activity studies on the corticotropin releasing factor antagonist astressin, leading to a minimal sequence necessary for antagonistic activity. Chembiochem 2004, 5, 340–348. [Google Scholar] [CrossRef] [PubMed]
  74. Pioszak, A.A.; Parker, N.R.; Suino-Powell, K.; Xu, H.E. Molecular recognition of corticotropin-releasing factor by its G-protein-coupled receptor CRFR1. J. Biol. Chem. 2008, 283, 32900–32912. [Google Scholar] [CrossRef]
  75. Weiss, G.A.; Watanabe, C.K.; Zhong, A.; Goddard, A.; Sidhu, S.S. Rapid mapping of protein functional epitopes by combinatorial alanine scanning. Proc. Natl. Acad. Sci. USA 2000, 97, 8950–8954. [Google Scholar] [CrossRef]
  76. Yamada, Y.; Mizutani, K.; Mizusawa, Y.; Hantani, Y.; Tanaka, M.; Tanaka, Y.; Tomimoto, M.; Sugawara, M.; Imai, N.; Yamada, H.; et al. New class of corticotropin-releasing factor (CRF) antagonists: Small peptides having high binding affinity for CRF receptor. J. Med. Chem. 2004, 47, 1075–1078. [Google Scholar] [CrossRef]
  77. Beyermann, M.; Rothemund, S.; Heinrich, N.; Fechner, K.; Furkert, J.; Dathe, M.; Winter, R.; Krause, E.; Bienert, M. A role for a helical connector between two receptor binding sites of a long-chain peptide hormone. J. Biol. Chem. 2000, 275, 5702–5709. [Google Scholar] [CrossRef]
  78. Pallai, P.V.; Mabilia, M.; Goodman, M.; Vale, W.; Rivier, J. Structural homology of corticotropin-releasing factor, sauvagine, and urotensin I: Circular dichroism and prediction studies. Proc. Natl. Acad. Sci. USA 1983, 80, 6770–6774. [Google Scholar] [CrossRef] [PubMed]
  79. Dathe, M.; Fabian, H.; Gast, K.; Zirwer, D.; Winter, R.; Beyermann, M.; Schumann, M.; Bienert, M. Conformational differences of ovine and human corticotropin releasing hormone. A CD, IR, NMR and dynamic light scattering study. Int. J. Pept. Protein Res. 1996, 47, 383–393. [Google Scholar] [CrossRef]
  80. Heinrich, N.; Meyer, M.R.; Furkert, J.; Sasse, A.; Beyermann, M.; Bonigk, W.; Berger, H. Corticotropin-releasing factor (CRF) agonists stimulate testosterone production in mouse leydig cells through CRF receptor-1. Endocrinology 1998, 139, 651–658. [Google Scholar] [CrossRef]
  81. Rivier, J.; Rivier, C.; Galyean, R.; Miranda, A.; Miller, C.; Craig, A.G.; Yamamoto, G.; Brown, M.; Vale, W. Single point D-substituted corticotropin-releasing factor analogues: Effects on potency and physicochemical characteristics. J. Med. Chem. 1993, 36, 2851–2859. [Google Scholar] [CrossRef]
  82. Rothemund, S.; Krause, E.; Beyermann, M.; Dathe, M.; Bienert, M.; Hodges, R.S.; Sykes, B.D.; Sonnichsen, F.D. Peptide destabilization by two adjacent D-amino acids in single-stranded amphipathic α-helices. Pept. Res. 1996, 9, 79–87. [Google Scholar]
  83. Eckart, K.; Jahn, O.; Radulovic, J.; Tezval, H.; Werven, L.; Spiess, J. A single amino acid serves as an affinity switch between the receptor and the binding protein of corticotropin-releasing factor: Implications for the design of agonists and antagonists. Proc. Natl. Acad. Sci. USA 2001, 98, 11142–11147. [Google Scholar] [CrossRef]
  84. Gilligan, P.J.; Robertson, D.W.; Zaczek, R. Corticotropin releasing factor (CRF) receptor modulators: Progress and opportunities for new therapeutic agents. J. Med. Chem. 2000, 43, 1641–1660. [Google Scholar] [CrossRef] [PubMed]
  85. Grigoriadis, D.E.; Haddach, M.; Ling, N.; Saunders, J. The CRF Receptor: Structure, Function and Potential for Therapeutic Intervention. Curr. Med. Chem. Cent. Nerv. Syst. Agents 2001, 1, 63–97. [Google Scholar] [CrossRef]
  86. Chen, Y.L.; Braselton, J.; Forman, J.; Gallaschun, R.J.; Mansbach, R.; Schmidt, A.W.; Seeger, T.F.; Sprouse, J.S.; Tingley, F.D., 3rd; Winston, E.; et al. Synthesis and SAR of 2-aryloxy-4-alkoxy-pyridines as potent orally active corticotropin-releasing factor 1 receptor antagonists. J. Med. Chem. 2008, 51, 1377–1384. [Google Scholar] [CrossRef]
  87. Million, M.; Zhao, J.-F.; Luckey, A.; Czimmer, J.; Maynard, G.D.; Kehne, J.; Hoffman, D.C.; Taché, Y. The Newly Developed CRF1-Receptor Antagonists, NGD 98-2 and NGD 9002, Suppress Acute Stress-Induced Stimulation of Colonic Motor Function and Visceral Hypersensitivity in Rats. PLoS ONE 2013, 8, e73749. [Google Scholar] [CrossRef]
  88. Molteni, V.; Penzotti, J.; Wilson, D.M.; Termin, A.P.; Mao, L.; Crane, C.M.; Hassman, F.; Wang, T.; Wong, H.; Miller, K.J.; et al. N-phenylphenylglycines as novel corticotropin releasing factor receptor antagonists. J. Med. Chem. 2004, 47, 2426–2429. [Google Scholar] [CrossRef] [PubMed]
  89. Williams, J.P. Corticotropin-releasing factor 1 receptor antagonists: A patent review. Expert Opin. Ther. Pat. 2013, 23, 1057–1068. [Google Scholar] [CrossRef] [PubMed]
  90. Perrin, M.H.; Sutton, S.; Bain, D.L.; Berggren, W.T.; Vale, W.W. The first extracellular domain of corticotropin releasing factor-R1 contains major binding determinants for urocortin and astressin. Endocrinology 1998, 139, 566–570. [Google Scholar] [CrossRef] [PubMed]
  91. Perrin, M.H.; Fischer, W.H.; Kunitake, K.S.; Craig, A.G.; Koerber, S.C.; Cervini, L.A.; Rivier, J.E.; Groppe, J.C.; Greenwald, J.; Moller Nielsen, S.; et al. Expression, purification, and characterization of a soluble form of the first extracellular domain of the human type 1 corticotropin releasing factor receptor. J. Biol. Chem. 2001, 276, 31528–31534. [Google Scholar] [CrossRef]
  92. Perrin, M.H.; DiGruccio, M.R.; Koerber, S.C.; Rivier, J.E.; Kunitake, K.S.; Bain, D.L.; Fischer, W.H.; Vale, W.W. A soluble form of the first extracellular domain of mouse type 2β corticotropin-releasing factor receptor reveals differential ligand specificity. J. Biol. Chem. 2003, 278, 15595–15600. [Google Scholar] [CrossRef]
  93. Klose, J.; Fechner, K.; Beyermann, M.; Krause, E.; Wendt, N.; Bienert, M.; Rudolph, R.; Rothemund, S. Impact of N-terminal domains for corticotropin-releasing factor (CRF) receptor-ligand interactions. Biochemistry 2005, 44, 1614–1623. [Google Scholar] [CrossRef] [PubMed]
  94. Grace, C.R.; Perrin, M.H.; Gulyas, J.; Digruccio, M.R.; Cantle, J.P.; Rivier, J.E.; Vale, W.W.; Riek, R. Structure of the N-terminal domain of a type B1 G protein-coupled receptor in complex with a peptide ligand. Proc. Natl. Acad. Sci. USA 2007, 104, 4858–4863. [Google Scholar] [CrossRef] [PubMed]
  95. Hofmann, B.A.; Sydow, S.; Jahn, O.; van Werven, L.; Liepold, T.; Eckart, K.; Spiess, J. Functional and protein chemical characterization of the N-terminal domain of the rat corticotropin-releasing factor receptor 1. Protein Sci. 2001, 10, 2050–2062. [Google Scholar] [CrossRef]
  96. Grace, C.R.; Perrin, M.H.; DiGruccio, M.R.; Miller, C.L.; Rivier, J.E.; Vale, W.W.; Riek, R. NMR structure and peptide hormone binding site of the first extracellular domain of a type B1 G protein-coupled receptor. Proc. Natl. Acad. Sci. USA 2004, 101, 12836–12841. [Google Scholar] [CrossRef] [PubMed]
  97. Qi, L.J.; Leung, A.T.; Xiong, Y.; Marx, K.A.; Abou-Samra, A.B. Extracellular cysteines of the corticotropin-releasing factor receptor are critical for ligand interaction. Biochemistry 1997, 36, 12442–12448. [Google Scholar]
  98. Dore, A.S.; Bortolato, A.; Hollenstein, K.; Cheng, R.K.Y.; Read, R.J.; Marshall, F.H. Decoding Corticotropin-Releasing Factor Receptor Type 1 Crystal Structures. Curr. Mol. Pharmacol. 2017, 10, 334–344. [Google Scholar] [CrossRef]
  99. Koth, C.M.; Murray, J.M.; Mukund, S.; Madjidi, A.; Minn, A.; Clarke, H.J.; Wong, T.; Chiang, V.; Luis, E.; Estevez, A.; et al. Molecular basis for negative regulation of the glucagon receptor. Proc. Natl. Acad. Sci. USA 2012, 109, 14393–14398. [Google Scholar] [CrossRef]
  100. Ma, S.; Shen, Q.; Zhao, L.H.; Mao, C.; Zhou, X.E.; Shen, D.D.; de Waal, P.W.; Bi, P.; Li, C.; Jiang, Y.; et al. Molecular Basis for Hormone Recognition and Activation of Corticotropin-Releasing Factor Receptors. Mol. Cell 2020, 77, 669–680.e4. [Google Scholar] [CrossRef]
  101. Gkountelias, K.; Tselios, T.; Venihaki, M.; Deraos, G.; Lazaridis, I.; Rassouli, O.; Gravanis, A.; Liapakis, G. Alanine scanning mutagenesis of the second extracellular loop of type 1 corticotropin-releasing factor receptor revealed residues critical for peptide binding. Mol. Pharmacol. 2009, 75, 793–800. [Google Scholar] [CrossRef] [PubMed]
  102. Spyridaki, K.; Matsoukas, M.T.; Cordomi, A.; Gkountelias, K.; Papadokostaki, M.; Mavromoustakos, T.; Logothetis, D.E.; Margioris, A.N.; Pardo, L.; Liapakis, G. Structural-Functional Analysis of the Third Transmembrane Domain of the Corticotropin-releasing Factor Type 1 Receptor: Role in Activation and Allosteric Antagonism. J. Biol. Chem. 2014, 289, 18966–18977. [Google Scholar] [CrossRef] [PubMed]
  103. Hollenstein, K.; Kean, J.; Bortolato, A.; Cheng, R.K.; Dore, A.S.; Jazayeri, A.; Cooke, R.M.; Weir, M.; Marshall, F.H. Structure of class B GPCR corticotropin-releasing factor receptor 1. Nature 2013, 499, 438–443. [Google Scholar] [CrossRef] [PubMed]
  104. Wootten, D.; Simms, J.; Miller, L.J.; Christopoulos, A.; Sexton, P.M. Polar transmembrane interactions drive formation of ligand-specific and signal pathway-biased family B G protein-coupled receptor conformations. Proc. Natl. Acad. Sci. USA 2013, 110, 5211–5216. [Google Scholar] [CrossRef] [PubMed]
  105. Liang, Y.L.; Belousoff, M.J.; Zhao, P.; Koole, C.; Fletcher, M.M.; Truong, T.T.; Julita, V.; Christopoulos, G.; Xu, H.E.; Zhang, Y.; et al. Toward a Structural Understanding of Class B GPCR Peptide Binding and Activation. Mol. Cell 2020, 77, 656–668.e655. [Google Scholar] [CrossRef] [PubMed]
  106. Kraetke, O.; Holeran, B.; Berger, H.; Escher, E.; Bienert, M.; Beyermann, M. Photoaffinity cross-linking of the corticotropin-releasing factor receptor type 1 with photoreactive urocortin analogues. Biochemistry 2005, 44, 15569–15577. [Google Scholar] [CrossRef] [PubMed]
  107. Dautzenberg, F.M.; Huber, G.; Higelin, J.; Py-Lang, G.; Kilpatrick, G.J. Evidence for the abundant expression of arginine 185 containing human CRF(2α) receptors and the role of position 185 for receptor-ligand selectivity. Neuropharmacology 2000, 39, 1368–1376. [Google Scholar] [CrossRef] [PubMed]
  108. Sydow, S.; Flaccus, A.; Fischer, A.; Spiess, J. The role of the fourth extracellular domain of the rat corticotropin-releasing factor receptor type 1 in ligand binding. Eur. J. Biochem. 1999, 259, 55–62. [Google Scholar] [CrossRef]
  109. Coin, I.; Katritch, V.; Sun, T.; Xiang, Z.; Siu, F.Y.; Beyermann, M.; Stevens, R.C.; Wang, L. Genetically encoded chemical probes in cells reveal the binding path of urocortin-I to CRF class B GPCR. Cell 2013, 155, 1258–1269. [Google Scholar] [CrossRef]
  110. Strader, C.D.; Fong, T.M.; Tota, M.R.; Underwood, D.; Dixon, R.A. Structure and function of G protein-coupled receptors. Annu. Rev. Biochem. 1994, 63, 101–132. [Google Scholar] [CrossRef]
  111. Savarese, T.M.; Fraser, C.M. In vitro mutagenesis and the search for structure-function relationships among G protein-coupled receptors. Biochem. J. 1992, 283 Pt 1, 1–19. [Google Scholar] [CrossRef]
  112. Cordomi, A.; Liapakis, G.; Matsoukas, M.T. Understanding Corticotropin Releasing Factor Receptor (CRFR) Activation Using Structural Models. Curr. Mol. Pharmacol. 2017, 10, 325–333. [Google Scholar] [CrossRef]
  113. Hoare, S.R.; Fleck, B.A.; Gross, R.S.; Crowe, P.D.; Williams, J.P.; Grigoriadis, D.E. Allosteric ligands for the corticotropin releasing factor type 1 receptor modulate conformational states involved in receptor activation. Mol. Pharmacol. 2008, 73, 1371–1380. [Google Scholar] [CrossRef]
  114. Assil-Kishawi, I.; Abou-Samra, A.B. Sauvagine cross-links to the second extracellular loop of the corticotropin-releasing factor type 1 receptor. J. Biol. Chem. 2002, 277, 32558–32561. [Google Scholar] [CrossRef] [PubMed]
  115. Hoare, S.R.; Sullivan, S.K.; Schwarz, D.A.; Ling, N.; Vale, W.W.; Crowe, P.D.; Grigoriadis, D.E. Ligand affinity for amino-terminal and juxtamembrane domains of the corticotropin releasing factor type I receptor: Regulation by G-protein and nonpeptide antagonists. Biochemistry 2004, 43, 3996–4011. [Google Scholar] [CrossRef] [PubMed]
  116. Nielsen, S.M.; Nielsen, L.Z.; Hjorth, S.A.; Perrin, M.H.; Vale, W.W. Constitutive activation of tethered-peptide/corticotropin-releasing factor receptor chimeras. Proc. Natl. Acad. Sci. USA 2000, 97, 10277–10281. [Google Scholar] [CrossRef]
  117. Zhao, L.-H.; Lin, J.; Ji, S.-Y.; Zhou, X.E.; Mao, C.; Shen, D.-D.; He, X.; Xiao, P.; Sun, J.; Melcher, K.; et al. Structure insights into selective coupling of G protein subtypes by a class B G protein-coupled receptor. Nat. Commun. 2022, 13, 6670. [Google Scholar] [CrossRef] [PubMed]
  118. Liaw, C.W.; Grigoriadis, D.E.; Lorang, M.T.; De Souza, E.B.; Maki, R.A. Localization of agonist- and antagonist-binding domains of human corticotropin-releasing factor receptors. Mol. Endocrinol. 1997, 11, 2048–2053. [Google Scholar] [CrossRef] [PubMed]
  119. Cunningham, B.C.; Wells, J.A. High-resolution epitope mapping of hGH-receptor interactions by alanine-scanning mutagenesis. Science 1989, 244, 1081–1085. [Google Scholar] [CrossRef] [PubMed]
  120. Jahn, O.; Tezval, H.; van Werven, L.; Eckart, K.; Spiess, J. Three-amino acid motifs of urocortin II and III determine their CRF receptor subtype selectivity. Neuropharmacology 2004, 47, 233–242. [Google Scholar] [CrossRef]
  121. Isfort, R.J.; Wang, F.; Tscheiner, M.; Donnelly, E.; Bauer, M.B.; Lefever, F.; Hinkle, R.T.; Mazur, A.W. Discovery of corticotropin releasing factor 2 receptor selective sauvagine analogues for treatment of skeletal muscle atrophy. J. Med. Chem. 2005, 48, 262–265. [Google Scholar] [CrossRef] [PubMed]
  122. Mazur, A.W.; Wang, F.; Tscheiner, M.; Donnelly, E.; Isfort, R.J. Determinants of corticotropin releasing factor. Receptor selectivity of corticotropin releasing factor related peptides. J. Med. Chem. 2004, 47, 3450–3454. [Google Scholar] [CrossRef] [PubMed]
  123. Pal, K.; Swaminathan, K.; Xu, H.E.; Pioszak, A.A. Structural basis for hormone recognition by the Human CRFR2{α} G protein-coupled receptor. J. Biol. Chem. 2010, 285, 40351–40361. [Google Scholar] [CrossRef] [PubMed]
  124. Rivier, J.; Gulyas, J.; Kirby, D.; Low, W.; Perrin, M.H.; Kunitake, K.; DiGruccio, M.; Vaughan, J.; Reubi, J.C.; Waser, B.; et al. Potent and long-acting corticotropin releasing factor (CRF) receptor 2 selective peptide competitive antagonists. J. Med. Chem. 2002, 45, 4737–4747. [Google Scholar] [CrossRef]
  125. Gulyas, J.; Rivier, C.; Perrin, M.; Koerber, S.C.; Sutton, S.; Corrigan, A.; Lahrichi, S.L.; Craig, A.G.; Vale, W.; Rivier, J. Potent, structurally constrained agonists and competitive antagonists of corticotropin-releasing factor. Proc. Natl. Acad. Sci. USA 1995, 92, 10575–10579. [Google Scholar] [CrossRef]
  126. Rivier, J.; Gulyas, J.; Kunitake, K.; DiGruccio, M.; Cantle, J.P.; Perrin, M.H.; Donaldson, C.; Vaughan, J.; Million, M.; Gourcerol, G.; et al. Stressin1-A, a potent corticotropin releasing factor receptor 1 (CRF1)-selective peptide agonist. J. Med. Chem. 2007, 50, 1668–1674. [Google Scholar] [CrossRef]
Figure 1. Model of CRF1R receptor (blue) in complex with the CRF (orange) and the three subunits of G-proteins, namely the Gαs protein (green), Gβ protein (brown), and Gγ protein (red). This generalized model of receptor has been created based on the crystal structure of the J-Domain of CRF1R in complex with CRF peptide and G-proteins (PDB: 6P9X) and the crystal structure of the ECD domain of CRF1R in complex with the CRF (PDB: 3EHU). The place between the two dotted lines is the lipid bilayer of the plasma membrane.
Figure 1. Model of CRF1R receptor (blue) in complex with the CRF (orange) and the three subunits of G-proteins, namely the Gαs protein (green), Gβ protein (brown), and Gγ protein (red). This generalized model of receptor has been created based on the crystal structure of the J-Domain of CRF1R in complex with CRF peptide and G-proteins (PDB: 6P9X) and the crystal structure of the ECD domain of CRF1R in complex with the CRF (PDB: 3EHU). The place between the two dotted lines is the lipid bilayer of the plasma membrane.
Biology 13 00120 g001
Figure 2. Structures of CRF analog peptides Astressin-2B (PDB: 2RM9), Astressin-B (PDB: 2RMD), Stressin1-A (PDB: 2RME), UcnI (PDB: 2RMF), UcnII (PDB: 2RMG) and UcnIII (PDB: 2RMH). The C-segment of the peptides is depicted in blue color, whereas the I-segment is depicted in green color and the N-segment is depicted in orange color [68].
Figure 2. Structures of CRF analog peptides Astressin-2B (PDB: 2RM9), Astressin-B (PDB: 2RMD), Stressin1-A (PDB: 2RME), UcnI (PDB: 2RMF), UcnII (PDB: 2RMG) and UcnIII (PDB: 2RMH). The C-segment of the peptides is depicted in blue color, whereas the I-segment is depicted in green color and the N-segment is depicted in orange color [68].
Biology 13 00120 g002
Figure 3. Model of CRF1R receptor (green) in complex with UcnI-EK (cyan). Salt bridges between peptide amino acids are shown in pink sticks.
Figure 3. Model of CRF1R receptor (green) in complex with UcnI-EK (cyan). Salt bridges between peptide amino acids are shown in pink sticks.
Biology 13 00120 g003
Figure 4. Structures of reported small molecule antagonists of CRF1R [84,85,86,87].
Figure 4. Structures of reported small molecule antagonists of CRF1R [84,85,86,87].
Biology 13 00120 g004
Figure 5. Crystal structure of the ECD (cyan) of human CRF1R (PDB: 3EHU) in complex with CRF (orange). Among the interactions between ECD and peptide, functionally important ones are the H-bonds (yellow sticks) between the main chain atoms of Val97, and the oxygen and nitrogen atoms of the C-terminal amide group (of residue 41) of CRF.
Figure 5. Crystal structure of the ECD (cyan) of human CRF1R (PDB: 3EHU) in complex with CRF (orange). Among the interactions between ECD and peptide, functionally important ones are the H-bonds (yellow sticks) between the main chain atoms of Val97, and the oxygen and nitrogen atoms of the C-terminal amide group (of residue 41) of CRF.
Biology 13 00120 g005
Figure 6. Allosteric binding of CP-376395 (C21H30N2O) to CRF1R (PDB: 4K5Y). Receptor residues are denoted by superscripts indicating their positions in receptor transmembrane domains (TMs). In subfamily B1 GPCRs, the most conserved residue in each TM is labeled as position index .50, preceded by the TM number (TM1-TM7). For example, Phe2845.51b denotes Phe284 located in TM5, one residue after the most conserved residue, Asn2835.50b.
Figure 6. Allosteric binding of CP-376395 (C21H30N2O) to CRF1R (PDB: 4K5Y). Receptor residues are denoted by superscripts indicating their positions in receptor transmembrane domains (TMs). In subfamily B1 GPCRs, the most conserved residue in each TM is labeled as position index .50, preceded by the TM number (TM1-TM7). For example, Phe2845.51b denotes Phe284 located in TM5, one residue after the most conserved residue, Asn2835.50b.
Biology 13 00120 g006
Figure 7. Molecular interactions of crucial CRF amino acids (orange) with CRF1R (light blue) (PDB: 6P9X). CRF forms an inserted loop consisting of residues Ser7P, Leu8P, and Asp9P, to accommodate accordingly and form selective polar and hydrophobic interactions with amino acids of mainly TMs 3, 5, and 6. Polar interactions are depicted in yellow dashes.
Figure 7. Molecular interactions of crucial CRF amino acids (orange) with CRF1R (light blue) (PDB: 6P9X). CRF forms an inserted loop consisting of residues Ser7P, Leu8P, and Asp9P, to accommodate accordingly and form selective polar and hydrophobic interactions with amino acids of mainly TMs 3, 5, and 6. Polar interactions are depicted in yellow dashes.
Biology 13 00120 g007
Figure 8. (A) Cryo-EM structure of UcnI in complex with TMD of CRF1R (PDB: 6PB0). (B) Cryo-EM structure of UcnI in complex with TMD of CRF2R (PDB: 6PB1). (C) Molecular interactions of crucial UcnI amino acids (magenta) with CRF1R intracellular residues (wheat). (D) Molecular interactions of UcnI amino acids (pink) with CRF2R (green). Polar interactions are shown in yellow dashes.
Figure 8. (A) Cryo-EM structure of UcnI in complex with TMD of CRF1R (PDB: 6PB0). (B) Cryo-EM structure of UcnI in complex with TMD of CRF2R (PDB: 6PB1). (C) Molecular interactions of crucial UcnI amino acids (magenta) with CRF1R intracellular residues (wheat). (D) Molecular interactions of UcnI amino acids (pink) with CRF2R (green). Polar interactions are shown in yellow dashes.
Biology 13 00120 g008
Figure 9. Molecular interactions of Gas protein (green) with CRF1R (cyan) (PDB: 6P9X). CRF (orange) and the three subunits of G-proteins, Gβ protein (brown), and Gγ protein (red) are also shown as part of the general complex in the left side.
Figure 9. Molecular interactions of Gas protein (green) with CRF1R (cyan) (PDB: 6P9X). CRF (orange) and the three subunits of G-proteins, Gβ protein (brown), and Gγ protein (red) are also shown as part of the general complex in the left side.
Biology 13 00120 g009
Table 1. Amino acid sequences of peptides of the CRF family, their percentage of identity to h/rCRF, and their binding affinities to two CRF receptor types. Highly conserved residues are in bold. Ki is the binding affinity of peptides mostly determined from heterologous displacement radioligand experiments.
Table 1. Amino acid sequences of peptides of the CRF family, their percentage of identity to h/rCRF, and their binding affinities to two CRF receptor types. Highly conserved residues are in bold. Ki is the binding affinity of peptides mostly determined from heterologous displacement radioligand experiments.
PeptideAmino Acid SequenceKi (nM)% h/rCRF Identity
CRF1R CRF2R
h/rCRF 1AgonistSEEPPISLDLTFHLLREVLEMARAEQLAQQAHSNRKLMEII1.9 b31 a100
oCRFAgonist-CRF1RSQEPPISLDLTFHLLREVLEMTKADQLAQQAHSNRKLLDIA1.2 b 1.1.2 b185 a82.9
UIAgonistNDDPPISIDLTFHLLRNMIEMARIENEREQAGLNRKYLDEV0.4 b2.2 a53.7
SVGAgonistZGPPISIDLSLELLRKMIEIEKQEKEKQQAANNRLLLDTI0.7 b4.3 a45.0
α-helCRF 2AntagonistDLTFHLLREMLEMAKAEQEAEQAALNRLLLEEA23.7 b96 a67.9
AstressinAntagonistHLLREVLEBARAEQLAQEAHKNRKLBEII15.4 b1.5 b86.2
rUcnI 1AgonistDDPPLSIDLTFHLLRTLLELARTQSQRERAEQNRIIFDSV0.3 b0.3 b45.0
hUcnI 1AgonistDNPSLSIDLTFHLLRTLLELARTQSQRERAEQNRIIFDSV0.4 b0.3 b42.5
hUcnII 1Agonist-CRF2RIVLSLDVPIGLLQILLEQARARAAREQATTNARILARV>100 c1.7 c34.2
mUcnII 1Agonist-CRF2RVILSLDVPIGLLRILLEQARYKAARNQAATNAQILAHV>100 c2.1 c34.2
hUcnII I 1Agonist-CRF2RFTLSLDVPTNIMNLLFNIAKAKNLRAQAAANAHLMAQI>100 c22 c31.6
mUcnIII 1Agonist-CRF2RFTLSLDVPTNIMNILFNIDKAKNLRAKAAANAQLMAQI>100 c5 c26.3
1 The abbreviations h/r, r, h, and m before the peptides refer to human/rat (see main text), rat, human and mouse, respectively; 2 α-helCRF is the α-helical CRF (9–41); a Grigoriadis et al., 1996 [60], b Dautzenberg et al., 2001 [61], c Lewis et al., 2001 [7].
Table 2. Amino acid sequences of chimeric peptides, created by linking the N-segment (DDPPLSIDLTFHLLRTLDEI) and the C-segment (QNRKLLDEV) of UcnI with charged linkers (in bold) of various lengths. The EC50 are the biopotencies of peptides (Adopted and modified by the study of Beyermann et al., 2000 [77]).
Table 2. Amino acid sequences of chimeric peptides, created by linking the N-segment (DDPPLSIDLTFHLLRTLDEI) and the C-segment (QNRKLLDEV) of UcnI with charged linkers (in bold) of various lengths. The EC50 are the biopotencies of peptides (Adopted and modified by the study of Beyermann et al., 2000 [77]).
LigandEC50 (nM)
Chimeric Peptides
N-region-E K E E K E K K R K E-C-region 0.6
N-region-E K E K E K K R K E-C-region25
N-region-E K K E K K R K E-C-region8.5
N-region-E K E K K R K E-C-region0.9
N-region-E K K K R K E-C-region12
N-region-E K K R K E-C-region50
N-region-E K R K E-C-region30
N-region-E K K E-C-region4.6
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Matsoukas, M.-T.; Panagiotopoulos, V.; Karageorgos, V.; Chrousos, G.P.; Venihaki, M.; Liapakis, G. Structural and Functional Insights into CRF Peptides and Their Receptors. Biology 2024, 13, 120. https://doi.org/10.3390/biology13020120

AMA Style

Matsoukas M-T, Panagiotopoulos V, Karageorgos V, Chrousos GP, Venihaki M, Liapakis G. Structural and Functional Insights into CRF Peptides and Their Receptors. Biology. 2024; 13(2):120. https://doi.org/10.3390/biology13020120

Chicago/Turabian Style

Matsoukas, Minos-Timotheos, Vasilis Panagiotopoulos, Vlasios Karageorgos, George P. Chrousos, Maria Venihaki, and George Liapakis. 2024. "Structural and Functional Insights into CRF Peptides and Their Receptors" Biology 13, no. 2: 120. https://doi.org/10.3390/biology13020120

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