Abstract
The orexin system consists of the peptide transmitters orexin-A and -B and the G protein-coupled orexin receptors OX1 and OX2. Orexin receptors are capable of coupling to all four families of heterotrimeric G proteins, and there are also other complex features of the orexin receptor signaling. The system was discovered 25 years ago and was immediately identified as a central regulator of sleep and wakefulness; this is exemplified by the symptomatology of the disorder narcolepsy with cataplexy, in which orexinergic neurons degenerate. Subsequent translation of these findings into drug discovery and development has resulted to date in three clinically used orexin receptor antagonists to treat insomnia. In addition to sleep and wakefulness, the orexin system appears to be a central player at least in addiction and reward, and has a role in depression, anxiety and pain gating. Additional antagonists and agonists are in development to treat, for instance, insomnia, narcolepsy with or without cataplexy and other disorders with excessive daytime sleepiness, depression with insomnia, anxiety, schizophrenia, as well as eating and substance use disorders. The orexin system has thus proved an important regulator of numerous neural functions and a valuable drug target. Orexin prepro-peptide and orexin receptors are also expressed outside the central nervous system, but their potential physiological roles there remain unknown.
Significance Statement The orexin system was discovered 25 years ago and immediately emerged as an essential sleep–wakefulness regulator. This discovery has tremendously increased the understanding of these processes and has thus far resulted in the market approval of three orexin receptor antagonists, which promote more physiological aspects of sleep than previous hypnotics. Further, orexin receptor agonists and antagonists with different pharmacodynamic properties are in development since research has revealed additional potential therapeutic indications. Orexin receptor signaling is complex and may represent novel features.
I. Introduction
A. History of Orexins
The discovery of orexins was simultaneously reported by two research groups (de Lecea et al., 1998; Sakurai et al., 1998) in 1998. The group of Sutcliffe discovered two novel hypothalamic peptides with excitatory properties on hypothalamic neurons (de Lecea et al., 1998). They named these peptides hypocretin 1 and 2, with reference to their hypothalamic origin and sequence similarity with incretins. The approach of the group of Yanagisawa, in collaboration with SmithKline Beecham’s orphan receptor program, started with the deorphanization of the orphan receptor HFGAN72 upon isolation of two peptides from the rat brain; the peptides were shown to come from the same prepro-peptide (Sakurai et al., 1998). They identified another putative receptor gene sequence in an expressed sequence tag (EST) library; this receptor was also shown to bind and respond to these two peptides. They further mapped the genes for the prepro-peptide and the receptors to human chromosomes. They then investigated the tissue expression of the prepro-peptide and the receptor mRNAs, and demonstrated in rats that the prepro-peptide expression increased upon fasting, whereas cerebroventricular infusion of the mature peptides could stimulate food intake. Because of this coupling to appetite, the peptides were named orexins: The prepro-peptide became prepro-orexin (PPO) and the mature peptides orexin-A and -B, while the receptors became OX1 and OX2 orexin receptor. It was soon realized that the two studies described the same peptides, i.e., hypocretins and orexins referred to the same peptides, and the parallel nomenclatures coexist to date (see further discussion under I.B.).
The discovery of the orexin system led to a surge of research activities. The major findings were related to narcolepsy: a) canine hereditary narcolepsy is caused by inactivating OX2 receptor mutations (Lin et al., 1999; Hungs et al., 2001; Mignot, 2004), b) PPO knockout (KO) mice display a narcolepsy with cataplexy-like phenotype (Chemelli et al., 1999), and c) human patients with narcolepsy type 1 (NT1) show very low cerebrospinal fluid (CSF) orexin-A levels and, postmortem, strongly reduced numbers of orexin-producing neurons in the lateral hypothalamus (Nishino et al., 2000; Thannickal et al., 2000). Development of tools for orexin research started rapidly and has resulted in, e.g., diverse transgenic animals and virus vectors. Drug discovery was also initiated early on in both biotech and large pharma. Dozens of lead structures for antagonists have been developed, some of which are in different phases of drug development, and three dual orexin receptor antagonists (DORAs, i.e., orexin receptor subtype-nonselective antagonists) suvorexant (Merck Co., Inc.), lemborexant (Eisai Co., Ltd) and daridorexant (Idorsia Pharmaceuticals Ltd) (Cox et al., 2010; Yoshida et al., 2015; Treiber et al., 2017), have currently been registered in the US and/or Europe/Japan/Australia/Canada and are in clinical use to treat insomnia.
B. Orexin Nomenclature
The first published study (publication date January 6, 1998) reported the discovery of the peptides (hypocretin 1 and 2; the N-terminal extent of orexin-A/hypocretin 1 was not established), their prepro-peptide and its gene in mouse chromosome 11, as well as the neuroexcitatory property of the peptides (de Lecea et al., 1998). The second study (publication date February 20, 1998) deorphanized OX1 and identified and deorphanized OX2, both of which are class A G protein-coupled receptors (GPCRs) with relatively similar sequences, and reported the two native peptide ligands (orexin-A and -B), including their posttranslational modifications, and the corresponding prepro-peptide (prepro-orexin, PPO), and the rat, mouse and human PPO DNA sequences, mapped all three genes in the human genome, and demonstrated the connection of the peptides with aspects of appetite regulation (Sakurai et al., 1998). Taking into account the parallel work in the discovery of these peptides, it is clear that both groups and their members have contributed enormously to the field and continue to do so.
No agreement has been reached regarding the adoption of either nomenclature in its entirety or any third nomenclature, and thus both original nomenclatures exist in parallel, but orexin is the dominant term in the literature: Statistics from the 600 most recent orexin research and review papers (false hits and other types of publications removed) in PubMed (May 18, 2024) using the search “(orexin or hypocretin)” show that 85% use dominantly orexin, 12% dominantly hypocretin and 3% the two terms equally/mixed. Orexin is a MeSH term (but not hypocretin) and thus in PubMed hypocretin should link to orexin but not vice versa; however, the search “orexin or hypocretin” gives more hits than either term alone. Publications use the terms orexin and hypocretin and the abbreviations interchangeably and sometimes new nomenclatures are invented (for instance, orexin-1/2 receptor, type 1/2 orexin receptor, OXR1/2). Of note, OX 2/OX-2/OX2 (the name based on an antibody-secreting cell line; equivalent to CD200) and OXR1 (oxidation resistance 1) are unrelated genes/proteins, but sometimes these are wrongly associated with the orexin system even in PubMed. Further diversion in the nomenclature is caused by gene and protein naming by, e.g., UniProt and HUGO Gene Nomenclature Committee (HGNC).
NC-IUPHAR aims to standardize receptor and ligand nomenclature. The official NC-IUPHAR compromise recommendation is to call the gene products/peptides based on the orexin nomenclature (PPO, orexin-A/-B, OX1/OX2 orexin receptors) and the genes based on the hypocretin nomenclature (the human genes HCRT, HCRTR1/2) (Table 1) (Alexander et al., 2023; Aston-Jones et al., 2023). Use of other nomenclatures is strongly discouraged.
C. The Orexin System
1. Orexin Peptides
Orexin peptides include the precursor peptide prepro-orexin (PPO) and the active transmitters orexin-A and orexin-B (in mammals 33 and 28 amino acids, respectively) that are produced upon PPO cleavage in equal amounts (one of each from one molecule of PPO: amino acids 34–66 → orexin-A, amino acids 70–97 → orexin-B) (Sakurai et al., 1998). Both peptides are processed with the typical biological stabilization by C-terminal amidation. Orexin-A additionally contains an N-terminal pyroglutamate (cyclized from glutamine) and two intrachain disulfide bridges (Sakurai et al., 1998) (Fig. 1A). The enzymes responsible for the maturation of the orexin peptides are unknown, although the enzymes prohormone convertase 1/3 and 2 have been shown to be expressed in orexin neurons (Nilaweera et al., 2003; Helwig et al., 2006). There are, in principle, suitable basic dipeptides in the relevant places in PPO for prohormone convertase cleavage, though there are similar sites in prodynorphin also expressed in these neurons (IV.A.).
Orexin-A and -B are highly identical in their C-termini (Fig. 1A), suggested to be due to tandem duplication (Alvarez and Sutcliffe, 2002). The peptide sequences of the mammalian orthologs of each mature orexin peptide are either completely conserved or highly similar but begin to diverge in other tetrapod classes and fishes (Fig. 1, B and C) (Soya and Sakurai, 2020). The C-terminal rat PPO fragment (PPO98–131; in humans AGAEPAPRPCLGRRCSAPAAASVAPGGQSGI), the sequence of which does not show any similarity to orexin-A and -B, has been suggested to have biological activity in rats (Tsuneki et al., 2024); however, the sequence is highly divergent among different species which make the general applicability of the results low.
A few single nucleotide polymorphisms (SNPs) have been identified in human PPO. L16R in the signal peptide causes narcolepsy (Peyron et al., 2000) while K68R (within the putative recognition site for cleavage between orexin-A and orexin-B sequences) is associated with a fivefold increased risk for idiopathic hypersomnia (Miyagawa et al., 2022). Several other missense variants of the orexin-A and orexin-B parts of the PPO sequence have been reported using the NIH tool Variation viewer, but these represent no known phenotype and could also result from sequencing errors (Ericson and Haskell-Luevano, 2018).
PPO is expressed in a subset of hypothalamic neurons and potentially in some other neuron types in the central nervous system (CNS) (IV.A., IV.K.). PPO mRNA and protein have also been demonstrated in some tissues outside the CNS (e.g., kidney, pancreas, placenta, etc.; see IV.M. and Kukkonen, 2013), but no functional role has yet been found. The proximal Hcrt gene (PPO) promoter (3.2 kb) has been investigated in mouse (Sakurai et al., 1999; Moriguchi et al., 2002). This promoter directs the expression of the product of the following coding sequence specifically to the orexin neurons of the hypothalamus, and has been used to express enhanced green fluorescent protein (EGFP) (Li et al., 2002), cre recombinase (Matsuki et al., 2009) and many other proteins in orexin neurons in mouse (and rat), as reviewed in Table 2 of (Kukkonen, 2013) (but see also VI. for potential problems). There are different findings as to how early in prenatal development this promotor is active (van den Pol et al., 2001; Amiot et al., 2005; Faraco et al., 2006); this issue has bearing on the specificity of the expression (IV.). A shorter variant of the promoter (1.3 kb; Moriguchi et al., 2002) has been used in some viral vectors; it has been reported that the specificity of this promoter for orexin neurons may be compromised in mouse and rat (Kakava-Georgiadou et al., 2019) but not in zebrafish (Faraco et al., 2006).
2. Orexin Receptors
Two orexin receptors are found in humans, OX1 and OX2 (the gene names HCRTR1 and HCRTR2, respectively; Human Gene Nomenclature Committee IDs 4848 and 4849, respectively). OX1 and OX2 belong to the class A (rhodopsin-like) GPCR family, in which they form their own subfamily. The subtypes share high sequence identity: 64% for overall protein sequence, 80% when only the transmembrane regions are compared. By contrast, the identity of the full sequences is <34% when compared with other human GPCRs.
Orexin receptors are expressed in the CNS along the orexinergic projections (see Fig. 7 in Kukkonen et al., 2002). Some projections may exit the CNS to act on the pineal gland (IV.C.) and potentially the pituitary gland (Kukkonen, 2013). Orexin receptor peptide mRNA and protein have also been found at peripheral sites clearly outside the reach of central orexinergic projections (e.g., intestinal smooth muscle, pancreas, adipose tissue etc.; see IV.M. and Kukkonen, 2013), but no functional role has been unequivocally demonstrated. The findings regarding the peripheral orexin peptides and orexin receptors are presented under IV.B., -C., -L. and -M. and in Kukkonen, 2013.
The promoter regions of the orexin receptor genes have only been examined in a single study, in which EGFP was added in front of the mouse Hcrtr1 gene in a bacterial artificial chromosome construct which was then inserted in the mouse genome (Darwinkel et al., 2014). The EGFP fluorescence and staining followed the pattern reported in OX1 receptor mRNA expression studies; however, there is no information as to where the construct was inserted in the mouse genome and which parts of the promoter it used in the end. The mouse Hcrtr2 gene promoter 1.8 kb upstream of the first exon (1 A; part of the 5′-untranslated region) has been examined and several regulatory elements, e.g., for AP-1 (activating protein-1), MZF1 (myeloid zinc finger 1) and CREB (cAMP-response element-binding protein), have been identified (Chen and Randeva, 2010). There is a polymorphism 2.7 kb upstream of the first exon (exon 1 A; part of the 5′-untranslated region) of the human Hcrtr2 receptor gene that disrupts a potential binding site for the transcription factor TCF4. Heart failure patients with the minor allele (T) were reported to maintain a lower ejection fraction than those with the major allele (C) despite adequate intervention (Perez et al., 2015); likewise, the minor allele conferred a poorer prognosis after myocardial infarction (Wohlfahrt et al., 2023) (IV.M.). Mice devoid of the Hcrtr2 gene expression are more prone to diastolic heart failure, and orexin-A can improve the heart function of wild-type mice exposed to cardiac stress (angiotensin II and isoproterenol treatment for 4 weeks) (Perez et al., 2015). Nevertheless, the understanding of the role of orexins at the myocardial level is minimal (IV.M.).
The human HCRTR1 has one additional predicted transcript (https://www.ncbi.nlm.nih.gov/nuccore/XM_017001107) that would produce an OX1 receptor with a shorter and distinct C-terminus as compared with the consensus sequence (coding exons 3–9) due to skipped exon 9 including its stop codon; the extra amino acids come, instead, from the region 3′ from the last annotated exon. The human HCRTR2 gene has been reported to present with four mRNA splicing variants 5′ from the translated region, all coding for the same OX2 protein sequences (Chen and Randeva, 2010); however, Ensembl only recognizes two such variants. There is also a predicted transcript of HCRTR2 with a divergent C-terminus; the transcript (https://www.ncbi.nlm.nih.gov/nuccore/XM_017010798) shows a 17-amino acid-longer C-terminus than the consensus sequence (coding exons 2–8), which is caused by removal of the stop codon within exon 8 by splicing together with a 3′ sequence. A sequenced, alternatively spliced OX2 transcript (https://www.ncbi.nlm.nih.gov/nuccore/KC812500) leads to a receptor that first diverges and then gets truncated in the third intracellular loop. A skipped splice site in the beginning of intron 4–5 causes this truncation; the divergent piece comes from a retained intron. In this case, the mechanism or the significance of the retention is unknown to date. However, some similarly produced, truncated GPCRs have been described (Wise, 2012; Schöneberg and Liebscher, 2021). Alternative splicing of the mouse OX2 gene generates two expressed receptor variants; the sequences are the same except for 17 additional amino acids in the C-terminus of the longer variant (exons 1–7 and 1–8, respectively) (Chen and Randeva, 2004). The situation is analogous to the “normal” HCRTR2 gene transcript versus the predicted longer variant XM_017010798 (above). Little work has been dedicated to the functional role of the two splice variants in mice. Interestingly, the longer mouse Hcrtr2 transcript and the predicted longer human HCRTR2 transcript are homologous to the validated rat Hcrtr2 transcript. A predicted, shorter rat Hcrtr2 transcript is also available in NCBI.
Several SNPs in the coding region of the human receptors are known, but with no clear effects on orexin receptor expression, structure or function (Kukkonen, 2013; Thompson et al., 2014). Several additional SNPs can be identified in Variation viewer as tested in the present study, but the significance and even correctness of these are so far uncertain. Some SNPs show weak associations with disease phenotypes, such as cluster headache, depression and panic disorder (Thompson et al., 2014). For instance, Ile3086.39 variant of OX2 has been very weakly associated with self-reported characteristic as being a ‘morning-person’ (Jones et al., 2016) whereas the Val3086.39 variant is associated in some studies with a somewhat increased risk of cluster headache and Alzheimer's disease (AD) (Gallone et al., 2014; Thompson et al., 2014). Note, that the receptor amino acids are also allocated the Ballesteros–Weinstein numbers (in superscript), which are based on highly conserved motives in each transmembrane (TM) helix: The positions are named with the TM number followed by an ordinal number in relation to the most conserved residue in each helix, this being defined as 50 (Ballesteros and Weinstein, 1995).
The most notable endogenous orexin receptor mutations are the canarc mutations, i.e., the three OX2 mutations causing hereditary narcolepsy in dogs (IV.C.1.). While two of these are truncations producing obviously nonfunctional receptors, Glu541.32 is located at the junction of the N-terminus and TM1 in OX2. The single point mutant, Glu541.32Lys, found in narcoleptic dachshunds (Hungs et al., 2001), has been inserted in the human OX2 as well as part of the OxLight1 orexin sensor (VI.) (Duffet et al., 2022a). The experimental data show that this mutant dog receptor is expressed on the HEK293 human embryonic kidney cell surface like the wildtype receptor, but has lost its ability to bind radiolabeled orexin-A and -B in the usual low nanomolar range (Hungs et al., 2001; Duffet et al., 2022a), although the molecular dynamics simulations do not demonstrate a direct contact between Glu54 and the orexin peptides (Karhu et al., 2019; Hong et al., 2021). In functional studies, the receptor is activated and produces intracellular Ca2+ elevation upon stimulation with orexin-A with several hundred-fold lower potency than the wildtype receptor (Hungs et al., 2001). In the agonist-bound, putative active 3D structures of the wildtype human OX2, the N-terminus is lacking, while in the antagonist-bound, inactive structures of human OX2, the α-helical N-terminus can sometimes be seen, but is oriented in a different way in different complexes (Yin et al., 2015; Suno et al., 2018; Hong et al., 2021; Rappas et al., 2020; Asada et al., 2022; Yin et al., 2022). Thus, the functional role of this mutation is difficult to speculate upon. However, it has been shown that the receptor N-terminus is important for orexin-A binding to OX1 receptors (Yin et al., 2016).
3. The Diversity and Evolution of the Orexin System
The furthest species from human with experimental data of orexin peptides to date are lancelets (Wang et al., 2019a). A putative orexin receptor has been identified and recently also characterized in vase tunicate (Ciona intestinalis) (Fridmanis et al., 2007; Rinne et al., 2024). A corresponding peptide-ligand was proposed, showing ability to bind, but not activate C. intestinalis orexin receptor (Rinne et al., 2024). Predicted orexin precursors can be found throughout deuterostome species, all the way to the hemichordate Saccoglossus kowalevskii. There are no identified orexin precursors or orexin receptors in protostome species but the protostome allatotropin system has been suggested to correspond to the deuterostome orexin system (Horodyski et al., 2011; Mirabeau and Joly, 2013; Alzugaray et al., 2019).
The overall comparison among mammalian orthologs for each receptor subtype shows 75–100% amino acid sequence identity (comparison run in this study for 84 OX1 sequences and 80 OX2 sequences from Ensembl). OX2 is present throughout vertebrate species, while OX1 is present mostly in mammals, and thus OX1 has been suggested to result from a later duplication in the mammalian lineage (Wong et al., 2011; Alzugaray et al., 2019; Soya and Sakurai, 2020). Spotted gar (Lepisosteus oculatus), reedfish (Erpetoichtys calabaricus) and coelacanth (Latimeria chalumnae) possess two types of orexin receptors: One of these is clearly OX2, while the “OX1” was previously suggested to be a third orexin receptor type and not orthologous to mammalian OX1 (Cai et al., 2018). We performed here a phylogenetic reconstruction with the set (346 sequences in total from Ensembl or NCBI) of orexin and allatotropin receptor protein sequences from (Alzugaray et al., 2019) and also included reedfish and coelacanth putative OX1 amino acid sequences (spotted gar was included in the set) using the maximum likelihood method. In this, the coelacanth (XP 015203966), spotted gar (ENSLOCT00000001479.1) and reedfish (ENSECRT00000019324.1) putative OX1 cluster together with mammalian OX1 receptors with good branch support (Fig. 2). We further repeated the phylogenetic analysis with the neighbor-joining method (bootstrap × 500) with similar results (not shown). Additionally, we constructed alignment with TM regions only, and conducted the same analysis. Neighbor-joining method produced again similar result, but in the maximum likelihood tree the coelacanth OX1 was branched outside OX1 and OX2 receptors while the reedfish and spotted gar OX1 were clustered with the other OX1 receptors (not shown). We further investigated the gene synteny of human HCRTR1 and the putative Hcrtr1 in spotted gar and in coelacanth; indeed, there is strong evidence that they are orthologs. This, taken together with the chromosomal location of human HCRTR1 and HCRTR2 (Nakatani et al., 2007; C. A. Bergqvist & Dr. D. Larhammar, personal communication), supports the theory that OX1 and OX2 are ohnologs, i.e., they arose through the whole genome duplications (WGDs, 1R/2R) in early vertebrate evolution, and OX1 was later lost in other tetrapod and teleost lineages. For the WGD theory, see (Ohno, 1970; Simakov et al., 2022).
As indicated above, the protostome allatotropin system has been suggested to correspond to the deuterostome orexin system, implying that the orexin system has evolved in deuterostomes and the allatotropin system in protostomes from a common origin (Adami et al., 2011; Adami et al., 2012; Alzugaray et al., 2013). Allatotropin receptors and orexin receptors have similar primary structure, and they cluster together in phylogenetic analyses (Mirabeau and Joly, 2013; Alzugaray et al., 2019), which may indicate common ancestral origin. They also share the same signature motif E/DRWYAI in the second intracellular loop (Alzugaray et al., 2019). Allatotropin was initially identified as a regulator of the juvenile hormone production in insects (Kataoka et al., 1989), but several other functions have later been described (reviewed in Verlinden et al., 2015). Allatotropin has been strongly linked to feeding, and it is a known regulator of myotropic activity and the circadian clock (reviewed in Verlinden et al., 2015). The link between orexin and allatotropin peptides is less understood, and the mature orexins and allatotropins do not share sequence similarity (Fig. 1D). A proposed link is found in the putative orexin precursor of Saccoglossus kowalevskii, in which the orexin-A sequence is followed by a sequence of unknown function called “a cryptic peptide”, that also follows allatotropin in protostome prepro-allatotropins (Mirabeau and Joly, 2013).
II. Orexin Receptor Signaling
A. Primary Signal Transducers of Orexin Receptors
1. Heterotrimeric G Proteins
Orexin receptor signaling is complex, and the details are only briefly outlined here; other reviews provide more details (Kukkonen, 2013; Kukkonen and Leonard, 2014; Leonard and Kukkonen, 2014; Kukkonen, 2017; Kukkonen, 2019; Kukkonen and Turunen, 2021). The original publications are in most cases numerous and thus only the first or particularly significant ones are cited here.
The primary signal transducers of orexin receptors have been investigated in a few studies with native cells. [32P]GTP-azidoanilide binding to activated G proteins and identification of these with antibodies demonstrated the ability of (mostly) OX2 receptors to activate Gq, Gi and Gs (Karteris et al., 2001; Randeva et al., 2001; Karteris et al., 2005; reviewed in Kukkonen and Leonard, 2014; Leonard and Kukkonen, 2014) while orexin stimulation of [35S]GTPγS binding in rat brain slices (Bernard et al., 2002; Bernard et al., 2003) suggested coupling of the orexin receptors to the Gi/o subfamily G proteins. These methods require permeabilization of the cells before the receptor stimulation.
Most other studies conducted have not directly identified the G proteins but have “deduced” them from more easily measured downstream signals or based on previous studies. Neither type of extrapolation is straight-forward or recommended (Kukkonen, 2004) (II.B.). Orexin receptor signaling is usually assumed to be Gq subfamily-mediated. This is based on the original finding of the coupling of these receptors to Ca2+ elevation a) upon recombinant expression in HEK293 and CHO-K1 Chinese hamster ovary cells (Sakurai et al., 1998) and b) later on in several studies with PC12, neuro-2a and IMR-32 cells (Smart et al., 1999; Holmqvist et al., 2002; Zhu et al., 2003; Näsman et al., 2006; Putula et al., 2011) as well as c) findings of the robust phospholipase C (PLC) activation in a number of recombinant cell lines (HEK293, CHO-K1, neuro-2a) (Lund et al., 2000; Holmqvist et al., 2002; Johansson et al., 2007; Putula and Kukkonen, 2012). However, other G proteins can also activate PLC – and even other PLC isoforms than PLCβ usually associated with the Gq subfamily signaling (Bunney and Katan, 2011; Kadamur and Ross, 2013; Katan and Cockcroft, 2020). Indirect methods applied in CHO-K1 cells suggested involvement of Gi and Gs proteins in the signaling of the OX1 receptor (Holmqvist et al., 2005), and further evidence was obtained for both OX1 and OX2 upon use of the Gq/11/14 inhibitor UBO-QIC/FR900359 (Kukkonen, 2016a,b). Unfortunately, neither this inhibitor nor its close analog, YM-254890, have been applied to any endogenously orexin receptor-expressing cell.
Recently, high throughput methods to identify coupling of GPCRs to all heterotrimeric G protein subfamilies have been developed. Two studies assessed a set of GPCRs including human orexin receptors recombinantly expressed in HEK293 with orexin-A as agonist. One study used chimeric G proteins for detection (Inoue et al., 2019). In this study, members of all four heterotrimeric G protein subfamilies were activated via both orexin receptor subtypes (see https://gproteindb.org/signprot/couplings for simpler presentation of the data). Another study used other proteins for a readout of the G protein activity (except for Gs) but the experiments were nevertheless conducted using heterologous overexpression of endogenous G proteins (Avet et al., 2022). Here, OX2 was found to couple to all heterotrimeric G protein subfamilies except Gs. Fluorescently tagged mini-G proteins (Wan et al., 2018) tested demonstrated coupling of OX2 receptors to Gq but barely to members of other G protein families (Duffet et al., 2022a). A recent structural study had its starting point in the idea that OX2 receptors couple more potently to Gq than Gi and motivated this concept with different structural determinants of the interaction of activated OX2 receptor with Gi1 and a Gq-mimetic (mini-Gαs/q/iN/Gβ/Gγ) (Yin et al., 2022). However, there is no evidence for a difference in coupling of OX2 to Gi versus Gq, rather the opposite, and thus the difference in interaction has an unclear significance (Kukkonen, 2023).
Based on this, it may be assumed that orexin receptors likely possess the ability to couple to all heterotrimeric G protein subfamilies, but which G proteins they actually signal through in which tissue and cell type, remains unclear. The most significant knowledge gap is in the brain, the major target of orexin actions. It is generally not known how the promiscuous GPCRs select their primordial signal transducers in different cases, as the simplest explanation, expression levels of the signal transducers, is insufficient to fully explain this. To complicate matters, it cannot even be assumed that coupling of a given receptor in a given brain region or nucleus is homogeneous; the individual cell populations within a nucleus may be highly diverse, and therefore the receptor may couple differently in cells from the same nucleus (Eberwine and Bartfai, 2011; Okaty et al., 2020). More recently, a series of studies suggests the existence of more than 3000 cell subclusters in the brain based on transcriptomics data (see Neuroscience Multiomic Archive and the article collection https://www.science.org/collections/brain-cell-census). Potential dimerization of GPCRs sometimes changes the receptor signaling. For instance, coexpression of OX1 receptors together with κ opioid receptors has been reported to lead to dimerization and redirection of the signaling from the Gq and Gi families, respectively, to Gs (Chen et al., 2015). Regulation of the G protein-coupling via heterodimerization GPCR partners is one potential idea, but there may be other explanations for the altered signaling (III.B.).
2. β-Arrestins and Other GPCR-Interacting Proteins
Orexin receptors have been shown to interact with β-arrestins upon heterologous expression, as probably all GPCRs do. This interaction and its relation to orexin receptor signaling and trafficking are discussed under II.G. In addition to heterotrimeric G proteins and β-arrestins, interaction with dynein light chain Tctex-types 1 and 3, norbin/neurochondrin, periplakin and the protein phosphatase SHP-2 has been suggested for OX1 (Voisin et al., 2008; El Firar et al., 2009; Ward et al., 2009; Duguay et al., 2011; reviewed in Kukkonen and Leonard, 2014; Leonard and Kukkonen, 2014). The interaction with SHP-2 and β-arrestins is discussed under II.D. and II.G., respectively.
B. Downstream Signal Transducers of Orexin Receptors
1. Phospholipase C
PLC practically refers to the phosphoinositide-specific phospholipase C, and is used in this meaning here as well. It comes in many isoforms stimulated by different factors (Bunney and Katan, 2011; Kadamur and Ross, 2013; Katan and Cockcroft, 2020). PLCβ is classically associated with GPCR signaling along the Gq subfamily pathway (Fig. 3, A and D), but GPCRs can, more or less directly, regulate any other PLC isoform as well (Bunney and Katan, 2011; Kadamur and Ross, 2013; Katan and Cockcroft, 2020). Upon hydrolysis of phosphatidylinositol-4,5-bisphosphate (PIP2) by PLC, the product is inositol-1,4,5-trisphosphate (IP3) − PIP2 + H2O → IP3 + diacylglycerol (DAG) (Fig. 3, A and D) − but lower phosphorylated species inositol monophosphate and bisphosphate, of unknown physiological role, are released when the substrate is phosphatidylinositol or phosphatidylinositol monophosphate, respectively (Fig. 3, A and D) (Katan and Cockcroft, 2020). Even though PIP2 is often referred to as the preferred PLC substrate (Katan and Cockcroft, 2020), no evidence is presented; in contrast, there are definite findings showing significant PLC activity on phosphatidylinositol and phosphatidylinositol monophosphate as well as high cellular levels of especially phosphatidylinositol (Harden et al., 1987; Homma et al., 1988; Gandhi et al., 1990; Zhou et al., 1993; Hwang et al., 1996; Zhang et al., 2013; Johansson et al., 2008). Endo-/sarcoplasmic reticulum IP3 receptor channels open in response to IP3 and thus give rise to intracellular Ca2+ elevation, which is usually followed by so-called capacitative Ca2+ influx, but Ca2+ elevation upon GPCR stimulation can also be achieved by other means (II.B.2.).
DAG, the other hydrolysis product of phosphoinositides, has plenty of targets (Carrasco and Merida, 2007) and it can also act as a source of other second messengers, i.e., phosphatidic acid (PA; II.B.5.), and a free fatty acid and 2-monoacylglycerol (MAG; see below) upon hydrolysis by diacylglycerol lipase (DAGL) (Fig. 3, A and D) (Merida et al., 2008; Fowler et al., 2017).
Orexin receptor signaling has been strongly associated with the Gq subfamily → Gαq → PLCβ → IP3 → Ca2+ release. This is deduced from the strong coupling of both human orexin receptor subtypes to Ca2+ elevation and PLC activation upon recombinant expression (II.A.1.), but has seldom been assessed in native cells. Since there are other ways for GPCRs to elevate intracellular Ca2+ concentration in addition to this pathway, Ca2+ elevation per se does not allow extrapolation of possible upstream pathways. The poor efficacy and selectivity and well-known toxicity of the small molecule PLC inhibitors (for instance, the most commonly used one, U73122; Taylor and Broad, 1998) preclude conclusions based on their effect on orexin signaling. In recombinant human OX1- and OX2-expressing CHO-K1 cells, the Gq/11/14 inhibitor UBO-QIC/FR900359 fully blocks orexin receptor simulation-induced PLC activity (Kukkonen, 2016a). PLC activity upon orexin receptor activation is also supported by the effect of protein kinase C (PKC) inhibitors on some responses in native neurons (see Leonard and Kukkonen, 2014, especially Supporting Information Table S1), though a) their selectivity is limited (Davies et al., 2000; Wu-Zhang and Newton, 2013), b) there are other sources for DAG (see below), and c) PKC can also be activated by other means (Zeng et al., 2012).
Bacteria express phosphatidylcholine-specific PLC and a similar activity has been suggested to be present in mammals, however without success in identifying the enzyme. A synthetic compound, D609 [O-(octahydro-4,7-methano-1H-inden-5-yl) carbonopotassium dithioate], has been suggested to inhibit this PLC, and has been often used to implicate involvement of the enzyme. D609 has also been used in orexin research (reviewed in Kukkonen, 2013). Sphingomyelin synthase classically catalyzes the reaction ceramide + phosphatidylcholine ↔ sphingomyelin + DAG. Sphingomyelin synthase is inhibited by D609, which has given one reasonable explanation to the reported mammalian “phosphatidylcholine-specific PLC activity” (Luberto and Hannun, 1998). Furthermore, both known mammalian sphingomyelin synthase enzymes (SMS1 and -2) were recently shown to exhibit phosphatidylcholine-specific PLC activity, i.e., phosphatidylcholine + H2O → phosphocholine + DAG (Chiang et al., 2021b), which is not surprising since mammalian phosphatidylethanolamine-specific PLC is also related to sphingomyelin synthase (Chiang et al., 2021a). This offers a satisfactory explanation to the issue for the time being. GPCRs, including orexin receptors, have also been suggested to signal via phosphatidylcholine-specific PLC based on the inhibitory effect of D609 (reviewed in Kukkonen, 2013). However, the regulation of sphingomyelin synthase remains undetermined, and no known domains for regulation are found in UniProt. Thus, the molecular mechanism of potential regulation of sphingomyelin synthase by orexin receptors (or any GPCR) remains unresolved. Sphingomyelin synthase 2 would be the major candidate for GPCR regulation as it is mainly found in the plasma membrane (Chen and Cao, 2017).
Both human orexin receptor subtypes couple to PLC activation in recombinant cell lines, as demonstrated by robust elevation of total inositol phosphates (and IP3 and DAG, as measured occasionally) (Lund et al., 2000; Holmqvist et al., 2002; Johansson et al., 2007; Johansson et al., 2008; Tang et al., 2008; Putula and Kukkonen, 2012). The exact type of PLC is not clear; a study with human OX1 receptors in CHO-K1 cells implicates two different PLC activities with specificities for different phosphoinositides (Johansson et al., 2008). DAG generated upon the PLC activity in these cells further leads to 2-arachidonoyl glycerol (2-AG) production – upon apparent DAGL activity – as well as arachidonic acid (AA) release from 2-AG hydrolysis (Turunen et al., 2012). 2-AG exits CHO-K1 cells and acts as an endocannabinoid (Jäntti et al., 2013). PLC activity upon orexin receptor stimulation has been directly measured in a few native cells and cell lines (Mazzocchi et al., 2001a; Randeva et al., 2001; Karteris et al., 2004; Karteris et al., 2005; Wenzel et al., 2009) and 2-AG production upon orexin receptor activation has been implicated in several brain regions (see II.B.6.).
2. Calcium
Cytosolic Ca2+ elevation can be accomplished a) by release from intracellular stores such as endoplasmic reticulum and b) by influx from the extracellular side (Fig. 3D). When investigating the orexin receptor-mediated (or any) Ca2+ elevation, it is important to distinguish these two and the different influx mechanisms, i.e., the receptor-operated and Na+/Ca2+ exchanger (NCX)-mediated influx from the more “trivial” Ca2+ influx caused by Ca2+ release (store-operated Ca2+ influx) or depolarization (influx via voltage-gated Ca2+ channels) (Fig. 3D) (Bootman and Bultynck, 2020; Kukkonen, 2023). Orexin receptors were originally assumed to be classical Gq subfamily-coupled receptors utilizing the cascade the Gq subfamily → Gαq → PLCβ → IP3 → IP3 receptor → Ca2+ release (Fig. 3D) (II.B.1.). This type of behavior is likely to take place in recombinant cells, but evidence from direct measurements or reliable inhibitors is scarce in native cells. Nevertheless, several mechanisms contribute to orexin-mediated Ca2+ elevation in both recombinant and native cells.
In several types of native neurons, orexin receptor stimulation causes inward current with decreased membrane resistance (Kukkonen and Leonard, 2014). In some cases, this has been ascribed to undefined nonselective cation channels (NSCCs), which are permeable to at least Na+ and K+, but potentially also to Ca2+, whereas in other cases NCX has been suggested to be involved (Fig. 3, B–D). The forward mode action of NCX takes in 3–4 Na+ for 1 Ca2+ extruded (Harper and Sage, 2016), and thus this activity indirectly indicates that there would be a preceding (orexin-induced) Ca2+ elevation in the neurons, though it does not pinpoint the precise source of this Ca2+ (Fig. 3C). In the opposite case, i.e., reverse mode action of NCX, there would be a preceding Na+ elevation, which NCX would try to amend, at the same time elevating intracellular Ca2+ (Fig. 3B) (Harper and Sage, 2016). This could make NSCCs with low Ca2+ permeability capable of allowing an indirect Ca2+ entry, which is difficult to distinguish from direct Ca2+ entry via the NSCCs. As discussed previously, distinction between NSCCs and NCX for orexin receptor-mediated responses is not possible with the commonly used tools, except in a few select cases (Kukkonen and Leonard, 2014; Leonard and Kukkonen, 2014). For instance, the typically used NCX inhibitor, KB-R7943, is not selective for the NCX reverse mode at the concentrations usually employed but it also inhibits both the forward mode of NCX (Iwamoto et al., 1996) and other ion transporters and channels, including transient receptor potential (TRP) channels (Kraft, 2007; Santo-Domingo et al., 2007; Liang et al., 2008; Barrientos et al., 2009; Storozhevykh et al., 2010; Brustovetsky et al., 2011; Abramochkin et al., 2013; Abramochkin and Vornanen, 2014; Wiczer et al., 2014; Sun et al., 2023a; Wu and Yu, 2023). Finally, other cation exchangers should be considered as potential sources of Ca2+ entry (Lytton, 2007; Khananshvili, 2022).
In recombinant cells, orexin receptors have been reported to activate NSCCs. In OX1-expressing CHO-K1 cells, the Ca2+ response at low concentrations of orexin-A (0.03–10 nM) apparently relies on Ca2+ influx (see, e.g., Lund et al., 2000; Larsson et al., 2005), the source of which has not been identified. Enzymatic degradation of (most of) the IP3 generated in the cytosol of OX1-expressing CHO-K1 cells totally blocks the apparent Ca2+ release while 50% of the receptor-operated Ca2+ influx remains (Ekholm et al., 2007), suggesting the influx does not absolutely require IP3. In contrast, this influx is almost fully blocked by the nonselective phospholipase A2 (PLA2) inhibitor methyl arachidonyl fluorophosphonate and the specific cytosolic (cPLA2) inhibitor pyrrophenone (II.B.4.), implicating involvement of cPLA2. The Gq/11/14 inhibitor UBO-QIC/FR900359 completely blocks all Ca2+ responses in these cells (Kukkonen, 2016a). In OX1-expressing HEK293 cells, low concentrations of orexin-A (1 nM) give rise to Ca2+ oscillations (Peltonen et al., 2009). Expression of a dominant-negative TRPC3-channel reduces the number of oscillating cells by more than 50% whereas the nonselective PLA2 inhibitor methyl arachidonyl fluorophosphonate totally blocks the oscillations. By contrast, Ca2+ elevation in OX1-expressing IMR-32 cells is inhibited by more than 50% by a dominant-negative TRPC6-channel (Näsman et al., 2006; Louhivuori et al., 2010). In agreement with the concept of co-operation between TRP channels and NCX (above), inhibition of the reverse mode action of NCX abrogates the Ca2+ influx in IMR-32 cells (Louhivuori et al., 2010). In all three mentioned cell lines, there is a potential involvement of DAG-sensitive TRPC channels in the response, which is also supported by the channel inhibitor and potentiator profile. An obvious question is how the cPLA2 products, AA and sn2-lysophosphatidylcholine (LPC), or their potential metabolites, might fit into the picture. LPC is known to activate TRPC5 (Flemming et al., 2006) and to some extent TRPC6 in endothelial cells (Chaudhuri et al., 2008). TRPC1 and -3 are activated by LPC in coronary artery smooth muscle cells (Wang et al., 2016). TRPC6 has been suggested to be activated by the AA metabolite 20-hydroxyeicosatetraenoic acid (20-HETE) (Basora et al., 2003). PLA2 activity increases TRPC6 trafficking to plasma membrane (Putta et al., 2021). These offer some potential connections, but currently there is no proof that the channels in CHO-K1 cells are either TRPC3 or -6. The potentially AA-regulated ARC channel should also be recognized as a candidate for orexin receptor regulation (Thompson and Shuttleworth, 2013; Putta et al., 2021).
In native neurons, there is currently only one report of a potential involvement of TRP channels in orexin responses: Based on the inhibitor profile, TRPC4/5 were suggested to be at play in thalamic midline neurons (Kolaj et al., 2014). Another study demonstrates expression of different TRPC channel mRNAs in the brainstem and hypothalamic orexin receptor target areas (Sergeeva et al., 2003). As indicated above, the potential activation mechanisms of the NSCCs reported to be activated upon orexin signaling in neurons have not been elucidated, and the knowledge of orexin receptor-mediated activation of lipid cascades is scarce in the CNS, except in the endocannabinoid context.
The hypothesis presented by Kukkonen et al. – based on significant but indirect evidence from extracellular Ca2+ chelation and inhibition of Ca2+ influx obtained in CHO-K1 cells – has been that Ca2+ influx is somehow important for coupling of the orexin receptors to other responses (Lund et al., 2000; Johansson et al., 2007; Ammoun et al., 2006a; Turunen et al., 2010; Jäntti et al., 2012), but the molecular mechanisms are unknown. As mentioned above, the effect of the Gq/11/14 inhibitor UBO-QIC/FR900359 suggests that the entire Ca2+ response to orexin receptor stimulation is dependent on the Gq subfamily signaling (Kukkonen, 2016a), confounding the situation. Another complicating issue is that the binding of 125I-orexin-A to (at least) OX1 receptors apparently requires extracellular Ca2+ (V.A.) (Putula et al., 2014). The reviews (Kukkonen and Leonard, 2014; Leonard and Kukkonen, 2014) present and discuss the evidence and the underlying ideas in more detail.
3. Adenylyl Cyclase
Adenylyl cyclase (AC) catalyzes the reaction ATP + H2O → cAMP + PPi. There are nine membrane-bound isoforms (AC1–9 or ACI–IX) of mammalian ACs, in addition to one soluble cytosolic isoform (sAC) (Halls and Cooper, 2017). Each membrane-bound AC isoform is regulated in a distinct way by GPCRs (Halls and Cooper, 2017). Classically, the membrane-bound ACs are regulated by Gαs (stimulation of all isoforms), forskolin (an experimentally used diterpene stimulator of all isoforms), Gαi (inhibition or no effect), Ca2+ (directly or via calmodulin; stimulation, inhibition or no effect), phosphorylation by the protein kinases protein kinase A and C (PKA and PKC, respectively; stimulation, inhibition or no effect) and calmodulin kinase (stimulation, inhibition or no effect) and by Gβγ (stimulation, inhibition or no effect), but there are probably additional regulatory mechanisms, including protein–protein complexes and other type of compartmentalization (Ostrom and Insel, 2004; Wang et al., 2009; West and Hanyaloglu, 2015; Wright et al., 2015; Halls and Cooper, 2017). It should be noted that every AC isoform is regulated in its own way by the different factors and that additive and synergistic effects between these typically occur. Since several factors can be stimulatory or inhibitory, AC stimulation or inhibition as such cannot be extrapolated to mean action of a specific factor: A common mistake is to consider AC stimulation to result from Gαs activation and AC inhibition to mean Gαi activation (Kukkonen, 2023). Cellular cAMP levels are additionally regulated by hydrolysis catalyzed by cyclic nucleotide phosphodiesterases, which are regulated by intracellular signals (Francis et al., 2011; Maurice et al., 2014); this also needs to be considered when assessing cAMP levels.
Orexin receptors interact with the classical G proteins directly involved in AC regulation (Gi, Gs) as demonstrated by direct evidence (II.A.1.), and both positive and negative regulation of the ACs can be seen under conditions that allow these G proteins to be distinguished (Mazzocchi et al., 2001b; Randeva et al., 2001; Holmqvist et al., 2003; Zhu et al., 2003; Holmqvist et al., 2005; Karteris et al., 2005; Tang et al., 2008; Kukkonen, 2016b; Rinne et al., 2018; Kukkonen, 2023). However, in most studies, the involvement of different signal transducers in reported AC regulation is not assessed. Orexin receptors could theoretically also use at least Gβγ, Ca2+ and protein kinases in their regulation of the AC isoforms. Due to the multiple potential regulators of cellular cAMP levels, orexin receptor activation-induced changes cannot be directly related to any specific G protein pathway unless the system is mapped, and the confounding elements are eliminated (Holmqvist et al., 2005; Kukkonen, 2016b; Kukkonen, 2023). Some orexin signaling cannot yet be explained based on the classical GPCR signaling and AC regulation. For instance, when the contribution from Ca2+- or Gβγ-regulated AC isoforms, PKC, Gαs and cyclic nucleotide phosphodiesterases have been eliminated, the orexin-A-induced AC inhibition is still fully sensitive to the Gi inhibitor pertussis toxin (OX1 and OX2), suggesting the classical Gαi-mediated AC inhibition. However, the response is at the same time partially (OX1) or fully (OX2) sensitive to the Gq/11/14 inhibitor UBO-QIC/FR900359 (Kukkonen, 2016a). This may be a reflection of an unknown AC regulation and/or illustrate the complicated nature of orexin receptor signaling.
4. Phospholipase A2
PLA2 hydrolyzes the sn2-fatty acid bond of membrane glycerophospholipids (GPLs): GPL + H2O → a fatty acid + 1-lyso-GPL (Fig. 3, A and D) (reviewed in Dennis et al., 2011; Astudillo et al., 2018). When the GPL is phosphatidylcholine, the lyso-GPL is LPC (Fig. 3, A and D). There are many isoforms of mammalian PLA2, which are divided into several subfamilies. PLA2 enzymes may serve in both signaling and membrane remodeling/homeostasis (Dennis et al., 2011; Astudillo et al., 2018). With respect to GPCR signaling, the most central subfamily may be cPLA2 (a.k.a. group IV PLA2): These are activated by (at least) intracellular Ca2+ elevation, and they, in particular cPLA2α (a.k.a. PLA2 IVA), release AA or other polyunsaturated fatty acids [at least eicosapentaenoic acid (EPA)] for, e.g., generation of eicosanoids (Astudillo et al., 2018). LPC can activate some TRP channels (TRPC5, TRPC6, TRPM8) (Kukkonen, 2011) and act as a precursor for lysophosphatidic acid (LPA) (II.B.5.) (Aoki et al., 2008).
Both human orexin receptors couple to cPLA2α activation in recombinant CHO-K1 cells, which thus constitutes one major pathway to orexin receptor-mediated AA release in these cells (Fig. 3, A and D) (Turunen et al., 2012; Kukkonen, 2016b). The other major part of the AA released comes from hydrolysis of 2-AG (II.B.1., II.B.6.) to AA and glycerol by MAG lipase (MAGL) (Fig. 3, A and D) (Turunen et al., 2012; Kukkonen, 2016b). The cPLA2α activity also seems to be required for the receptor-operated Ca2+ influx in these cells (II.B.2.) (Turunen et al., 2010; Turunen et al., 2012).
5. Phospholipase D
Phospholipase D (PLD) 1 and -2 hydrolyze phosphatidylcholine to generate PA and choline: phosphatidylcholine + H2O → PA + choline (Fig. 3, A and D) (reviewed in Bowling et al., 2021). PA acts as an intracellular messenger and as a precursor for other messengers such as DAG (Fig. 3A) (reviewed in Merida et al., 2008; Brindley and Pilquil, 2009; McDermott et al., 2020). Through a biophysical action, it affects membrane curvature (reviewed in Merida et al., 2008; Brindley and Pilquil, 2009; McDermott et al., 2020). Choline binds to σ1 receptors with a Ki = 500 μM and acts as an agonist (Brailoiu et al., 2019). PLD1/2 are regulated by GPCRs, but the mechanisms of regulation remain elusive. Indirect means of PLD activation by GPCRs include activation of monomeric G proteins of the Ras and Rho families, the protein kinases PKC, protein kinase N, ribosomal protein S6 kinases (RSK) and p38 mitogen-activated protein kinase (MAPK), and free unsaturated fatty acids (McDermott et al., 2020; Bowling et al., 2021). Decrease in PIP2 may, on the other hand, decrease PLD activity (McDermott et al., 2020; Bowling et al., 2021) while PLD activity increases phosphatidylinositol-4-phosphate 5-kinase (PIP5K) activity, which increases PIP2 production (Cockcroft, 2009; van den Bout and Divecha, 2009). Especially the Gq and G12/13 subfamilies are, thus, assumed to take part in the regulation of PLD activity, although other G proteins cannot be excluded.
The PLD family also includes members that do not hydrolyze phosphatidylcholine, namely PLD3–6 (McDermott et al., 2020). PLD6 (a.k.a. mitoPLD) was originally reported to take part in mitochondrial fission and fusion by hydrolyzing the mitochondrial GPL cardiolipin (Choi et al., 2006). PLD3, -4 and -6 have nuclease activity, while PLD5 may be catalytically inactive, but may have another function in humans (Ipsaro et al., 2012; Nishimasu et al., 2012; McDermott et al., 2020; Liu et al., 2021). N-acyl/arachidonoyl phosphatidylethanolamine-hydrolyzing PLD (NAPE-PLD), which produces the endocannabinoid N-arachidonoylethanolamine (anandamide), does not belong to the mammalian PLD family (Okamoto et al., 2009).
LysoPLDs hydrolyze LPC (or other lyso-GPLs): LPC + H2O → choline + LPA. LPA is a GPCR and nuclear receptor agonist (Kukkonen, 2011), and it has also been proposed to activate TRPA1 and TRPV1 channels (Kittaka et al., 2017). A known lysoPLD is autotaxin (Fig. 3 A) which hydrolyzes extracellular LPC (or other lyso-GPLs) (Aoki et al., 2008), producing extracellular LPA. LPA can also be generated from PA by PLA2 (or PLA1) activity (Aoki et al., 2008).
Both human orexin receptors couple to PLD1 activation in recombinant CHO-K1 cells (Johansson et al., 2008; Jäntti et al., 2012; Kukkonen, 2016b). The signaling between orexin receptors and PLD1 is not clear, but one activation mechanism could be phosphorylation of PLD1 by PKCδ (Jäntti et al., 2012). Interestingly, the OX2-mediated activation of PLD1 was found to be much weaker than that of OX1 (Kukkonen, 2016b).
6. Orexins and Endocannabinoids
Endocannabinoids are produced upon enzymatic activity and released to act as paracrine messengers (Kano et al., 2009; Muccioli, 2010). The most classical endocannabinoid cascade involves postsynaptic production and release of endocannabinoids and retrograde inhibition of excitatory or inhibitory neurotransmitter release via presynaptic CB1 receptors (Kano et al., 2009). Endocannabinoids act naturally also on CB2 cannabinoid receptors, primarily found on immune cells in the brain and periphery (Di Marzo and Petrosino, 2007; Abood et al., 2023). The major endocannabinoids are assumed to be anandamide and 2-AG (Kano et al., 2009; Muccioli, 2010; Fowler et al., 2017). Orexin signaling in the brain has been associated with 2-AG in the dorsal raphe (Haj-Dahmane and Shen, 2005; Haj-Dahmane and Shen, 2009), periaqueductal gray (Ho et al., 2011; Lee et al., 2016; Chen et al., 2018), ventral tegmental area (VTA) (Tung et al., 2016), hippocampus (Forte et al., 2021) and amygdala (Ten-Blanco et al., 2022). In addition, other cannabinoid receptor-independent functions of 2-AG (via, e.g., other receptors and ion channels) have been identified (Oz, 2006; Gantz and Bean, 2017; Soderstrom et al., 2017).
2-AG is usually thought to be produced from DAG coming from the PLC cascade (Fig. 3, A and D). DAG is hydrolyzed by the sn1-specific DAGLα/β and the MAG produced is 2-AG (Fig. 3, A and D), if the DAG contains arachidonic acid in sn2 position, as often is the case (Bisogno et al., 2003; Fowler et al., 2017). Activation of either orexin receptor can lead to generation of 2-AG in recombinant CHO-K1 cells, probably via the PLC pathway, but not the PLD1 pathway (Turunen et al., 2012). For orexin receptors, the cascade in the CNS was originally determined using CB1 receptor antagonists and agonists as well as DAGL and MAGL inhibitors (Haj-Dahmane and Shen, 2005; Ho et al., 2011; Lee et al., 2016; Morello et al., 2016; Tung et al., 2016; Chen et al., 2018) (II.C.). For example, Ho et al. (2011) demonstrated that OX1 activation in the rat periaqueductal gray suppressed GABAA receptor-mediated inhibitory currents. The orexin effect was blocked by a CB1 receptor antagonist and mimicked by a CB1 receptor agonist. The orexin effect was also blocked by inhibition of 2-AG production and enhanced by inhibition of 2-AG degradation. Direct measurement of 2-AG release upon orexin receptor activity has been reported in recombinant cells (Turunen et al., 2012; Jäntti et al., 2013) and in CNS neurons (Morello et al., 2016; Forte et al., 2021; Forte et al., 2022). Involvement of the PLC cascade in this response is as such logical, but the PLC inhibitor U73122 is a problematic compound (Taylor and Broad, 1998). It should be noted that the regulation of DAGL is not well understood, except for Ca2+ stimulation and putative nondisclosed phosphorylation sites (Bisogno et al., 2003; Reisenberg et al., 2012). Both presynaptic inhibition in the dorsal raphe (Haj-Dahmane and Shen, 2005; Haj-Dahmane and Shen, 2009) and presynaptic disinhibition in periaqueductal gray (Ho et al., 2011; Lee et al., 2016; Chen et al., 2018) and VTA (Tung et al., 2016) have been reported upon orexin receptor-mediated endocannabinoid generation. The potential physiological functions of this orexin signaling are not fully known, but the studies point to at least antinociception (IV.G.), addiction (IV.D.), and responses to high lipid diets (Lau et al., 2017; Fernandez-Rilo et al., 2023). Interestingly, it was suggested that peripheral median nerve stimulation enhances orexin release in the periaqueductal gray (Chen et al., 2018) and that this effect is independent of potential opioid tolerance (Lee et al., 2021).
2-AG signaling is terminated/modulated upon either by the action of the prostanoid-modifying enzymes cyclooxygenase 2 (COX-2) and lipoxygenase, generating glyceryl ester variants of their normal products (Urquhart et al., 2015), or hydrolysis by MAGL, ABHD6 or ABHD12 (Savinainen et al., 2012). The latter enzymes release AA which has a signaling role in itself and acts as a substrate for the generation of series 2 prostanoids (Bos et al., 2004; Meves, 2008; Shuttleworth, 2009; Kukkonen, 2011). A third possible fate of 2-AG is phosphorylation to an sn2-LPA (Kanoh et al., 1986; Bektas et al., 2005; Sato et al., 2016).
There are also other potential endocannabinoids, of which anandamide is the most well-known. These have not been associated with orexin receptor signaling.
C. Regulation of Electrical Activity
Orexin receptor activation was found early on to be excitatory to native CNS neurons (de Lecea et al., 1998; reviewed in Kukkonen and Leonard, 2014; Leonard and Kukkonen, 2014). This is usually accomplished by postsynaptic depolarization mediated by K+ channel closure (Horvath et al., 1999b; Ivanov and Aston-Jones, 2000; Hwang et al., 2001; Bayer et al., 2002; reviewed in Kukkonen and Leonard, 2014; Leonard and Kukkonen, 2014) or activation of NSCCs or NCX in reverse mode (Eriksson et al., 2001; Brown et al., 2002; Burlet et al., 2002; Liu et al., 2002; Wu et al., 2002; Yang and Ferguson, 2002; Burdakov et al., 2003; Yang and Ferguson, 2003; reviewed in Kukkonen and Leonard, 2014; Leonard and Kukkonen, 2014). It is rarely possible to distinguish between NSCC and NCX by the methods used (II.B.2.). It should be noted that the membrane conductance-decreasing (the first) and -increasing (the two latter) actions produce different effects on the overall excitability of the neurons (Schöne and Burdakov, 2017). Tandem-pore K+ family channels (e.g., TASK1/3; Doroshenko and Renaud, 2009) or inward rectifier channels (KIR or more specifically here GPCR-regulated KIR, GIRK) are likely candidates for the PKC-mediated inhibition and the latter are also inhibited by reduced PIP2; however, the channel types have usually not been directly identified. The precise types of NSCCs and NCXs are usually also not known, nor are the signals that regulate them. One study suggests TRPC4/5 channels as the NSCCs involved (Kolaj et al., 2014), but no other study has attempted to identify them. During recent years, more selective TRP channel inhibitors have become available (Bon et al., 2021; Koivisto et al., 2021).
In recombinant CHO-K1, HEK293 and IMR-32 cells, involvement of TRPC channels (Larsson et al., 2005; Näsman et al., 2006; Peltonen et al., 2009; Louhivuori et al., 2010) as well as NCX (Louhivuori et al., 2010) in OX1 responses has been suggested, but the regulatory mechanisms may vary in different cell types (II.B.2.). Lipid signaling cascades are one potential regulatory mechanism for NSCCs (Kukkonen, 2011; Kukkonen, 2014), as also discussed and examined in a few original studies; the engagement of orexin receptors in the regulation of these cascades (PLC, PLD, PLA2, DAGL) has been discussed in the context of the experimental findings above. The regulation of NCX is largely unknown. As indicated (II.B.2.), the forward mode action (Na+ in, Ca2+ out) that is depolarizing can be precipitated by Ca2+ influx by any means (Fig. 3C), while the reverse mode action (Na+ out, Ca2+ in) could come from Na+ influx (Fig. 3B), but the other forces to consider are the Na+ and Ca2+ gradients and the membrane potential; the direction of NCX action is principally in energetic downstream direction along the net force determined by the relevant two ionic gradients and the membrane potential (Harper and Sage, 2016; Khananshvili, 2022). The number of Na+ ions transported can vary, and therefore this factor also affects the energetics of the transport (Khananshvili, 2022). The cell surface levels of the transporters and the velocity of the reaction and even its directionality may be additionally regulated by Na+ and Ca2+ themselves, phosphorylation, proteolysis and interactions with other proteins or other molecules or ions etc. (Harper and Sage, 2016; Khananshvili, 2022). Of the known orexin receptor-regulated signaling components, PKA, PKC and PIP2 have been clearly identified. It should be noted that, in addition to NCX, other members of the Ca2+/cation antiporter family can also take part in the cation transport (Khananshvili, 2022). Finally, even voltage-gated Ca2+ channels have been implicated in the depolarizing action of orexins (Chrobok et al., 2017).
Neuronal effects of orexin receptor signaling are not limited to simple excitation. Orexin receptor stimulation can affect (reduce or enhance) afterhyperpolarization mediated by different types of K+ channels [putatively KCa2/KCa3.1 (SK), KNa1 (Slick/Slack), Kv4.3 (A) and unidentified types; faster or slower activation of the channel, altered open time] which may affect action potentials and other signals in multiple ways, e.g., by increasing the refractory time, changing the pattern of impulse firing, enhancing Ca2+ influx etc. (Zhang et al., 2010; Kolaj et al., 2014; Ishibashi et al., 2016; reviewed in Leonard and Kukkonen, 2014). The mechanisms used by orexin receptors to regulate these channels are putatively Ca2+ as well as PKA and PKC activities, but in most cases they are not known. PIP2 levels, affected directly by PLC and indirectly by PLD, may be one relevant channel regulator (Hille et al., 2015). Recently, orexin receptors were also shown to enhance the persistent Na+ current, INaP (Tanaka et al., 2021), which is assumed to be carried by the voltage-gated Na+ channels (Nav) and to act as a central regulator of cellular excitability (Lin and Baines, 2015; Stafstrom, 2007).
Recombinant CHO-K1 cells are depolarized by orexin receptor stimulation upon activation of an NSCC suggested to be a TRPC channel (Larsson et al., 2005). In human OX1-expressing HEK293 cells, TRPC3/6 channels are suggested to contribute to the orexin-A induced Ca2+ oscillations (Peltonen et al., 2009) and in OX1-expressing IMR-32 cells to reverse mode action of NCX, as in Fig. 3B (Louhivuori et al., 2010). Both human orexin receptor subtypes inhibit GPCR-regulated inward rectifier K+ (GIRK1 and -2) channels in recombinant HEK293 cells and depolarize the cells (Hoang et al., 2003).
Activation of human orexin receptors is capable of causing production of the endocannabinoid 2-AG (Fig. 3, A and D) as first demonstrated in recombinant CHO-K1 cells and then in native CNS neurons; significant additional information comes from the innovative utilization of inhibitors of relevant enzymes in several studies (II.B.6.). 2-AG is likely to act as a central mediator of orexin responses in the CNS. This implies that orexin receptors engage the PLC pathway also in the CNS, including its other characteristic messages such as PIP2 decrease, DAG and IP3 increase and intracellular Ca2+ release (II.B.1., II.B.2.). Interestingly, 2-AG may also act on ion channels and on the postsynaptic side (Oz, 2006; Gantz and Bean, 2017; Soderstrom et al., 2017), which may cloud the distinction between “direct” and “indirect” effects of orexin receptor signaling.
Presynaptic orexin actions have also been reported (van den Pol et al., 1998; Li et al., 2002; Smith et al., 2002; Davis et al., 2003; Lambe and Aghajanian, 2003; Borgland et al., 2008). Ca2+ elevation has been seen in a few studies (reviewed in Kukkonen and Leonard, 2014; Leonard and Kukkonen, 2014). Upon longer exposure to exogenous orexins, plastic effects on neuronal signaling, for instance long-term potentiation, take place (Selbach et al., 2004; Borgland et al., 2006; Chen et al., 2008; Selbach et al., 2010; Lu et al., 2016).
D. Cell Plasticity and Death in Non-Neuronal Cells
Long-term activation (24–96 hours of continuous exposure) of both orexin receptor subtypes has been shown to induce programmed cell death in recombinant cells, cancer cell lines and native cancer cells in culture. The findings and the mechanisms have been reviewed in detail in Kukkonen, 2013. In short, two different mechanisms, though potentially not mutually exclusive, have been reported in two different CHO cell clones. One relies on orexin receptor activation of a Gq subfamily G protein → an unidentified Src family kinase → orexin receptor phosphorylation → engagement of the protein phosphatase SHP-2 (Voisin et al., 2008; El Firar et al., 2009). The other is mediated via the p38 MAPK pathway (Ammoun et al., 2006b). These mechanisms have not been investigated further.
Regardless of the signaling cascades, it is very interesting that some carcinoma cells express orexin receptors (primary cells: colorectal carcinoma; cell lines: colorectal carcinoma, glioma, pancreatic carcinoma, prostate carcinoma, and some others suggested but not verified) which couple to (putative) programmed cell death (Rouet-Benzineb et al., 2004; Voisin et al., 2006; Voisin et al., 2011; Bieganska et al., 2012; Alexandre et al., 2014; Dayot et al., 2018; Huan et al., 2023). A peculiar finding reports cell death in the AsPC-1 pancreatic carcinoma cell line upon exposure to either orexin-A or the DORA almorexant (Dayot et al., 2018). As such it is not impossible in the world of biased agonism that a compound is an antagonist for one pathway and an agonist for another. However, the cell death upon orexin receptor activation has also been reported to rely on Gq (Voisin et al., 2008), but as yet, there is no evidence that almorexant acts as a Gq (or other) pathway agonist in any direct orexin receptor assay. Thus, the mechanism of the effect of almorexant may be distinct from action on orexin receptors.
Several orexin receptor agonists are in various clinical development stages for narcolepsy etc. (V.B.2.), raising the question of whether such ligands could also activate the cell death response. It is currently unknown whether native neurons also express the cell death response following long-term orexin receptor activation. Orexin receptor agonists for the indications would only be administered diurnally to consolidate daytime waking, and thus the receptor stimulation would not be continuous. Also, the current data from limited animal or human experiments with small molecule orexin receptor agonists does not indicate any demise of orexin receptor-expressing neurons, as concluded from the retained response (II.G.). However, a potential sustained nonmedical use (e.g., misuse) of orexin receptor agonists might carry risk.
In contrast to the cell death-stimulating effects of orexins, other studies associate them with promotion of cell survival and plasticity (reviewed in Kukkonen, 2013). Some studies suggest that orexin receptor stimulation increases cancer cell proliferation (see, e.g., Suo et al., 2018). Part of the explanation may arise from the fact that cancers are heterogeneous, and even the established, defined cancer cell lines tend to be subject to geno- and phenotypic drift. More recent studies suggest a protective role of orexins in, e.g., drug-induced kidney injury (Jo et al., 2022), though most such studies have focused on neuronal cells (IV.F.). No long-reaching conclusions about the actual role of orexins or molecular mechanisms can be drawn from such limited studies.
E. Other Orexin Receptor Signaling
Orexin receptor signaling has been suggested to activate a number of protein kinases, e.g., PKC, extracellular signal-regulated kinase (ERK), p38 and p90RSK, in studies using either direct or indirect methods (Milasta et al., 2005; Ammoun et al., 2006a,b; Johansson et al., 2008; Wang et al., 2014). The primary signaling cascades may be extrapolated from the various kinase activities, but the approach is not generally valid.
AMP-activated protein kinase (AMPK) is a central regulator of cellular energy metabolism and cell growth. Orexin-A activates AMPK in cultured rat arcuate nucleus neurons via L-type voltage-gated Ca2+ channel-mediated Ca2+ influx (Wu et al., 2013); the simplest explanation is that orexin-A induces depolarization via some of the mechanisms discussed above (II.C.). Inhibition of AMPK abrogates orexin-mediated increase in neuropeptide Y production and food intake; however, the AMPK inhibitor was introduced intracerebroventricularly (icv) (Wu et al., 2013). Phosphoinositide-3-kinase (PI3K) → phosphoinositide-dependent kinase 1 (PDK1) → protein kinase B (PKB/Akt) cascade has been suggested for orexin receptor signaling in both direct and indirect studies (Mazzocchi et al., 2001a; Ammoun et al., 2006a; Göncz et al., 2008), although the activation mechanisms of this cascade are unclear. In the recombinant hypothalamic neuronal cell lines N41/OX1 and N41/OX2 (and likewise recombinant OX1- or OX2-expressing HEK293 cells), orexins activate mTORC1 (mammalian target of rapamycin complex 1) in a PI3K- and PKB- (and ERK-) independent manner. Instead, the signal is suggested to be relayed via Ca2+ influx and the lysosomal pathway. However, this is not specific to orexin receptors since any Ca2+ elevation mimics it (Wang et al., 2014). mTOR activation has also been associated with corticosteroid synthesis in the cell lines H295R and Y-1 (Zhang et al., 2024a). mTOR and AMPK are often put in opposition, but orexin receptors are known to activate even otherwise apparently counteracting pathways such as Gi and Gs or ERK and p38 (see above). PKB is also shown to be activated in rat embryonic cortical neuronal cultures (Sokolowska et al., 2014); it is unclear, how PKB phosphorylation could be so strong if only a small subset of cortical neurons expresses orexin receptors, as is suggested (IV.A.).
Normal brown adipose tissue (BAT) development in mice has been suggested to require OX1 receptor signaling via p38 and BMP-7 (bone morphogenic protein-7) (Sellayah et al., 2011; Sellayah and Sikder, 2012). Interaction of orexin and bone morphogenic protein signaling pathways has also been reported in other studies (Fujita et al., 2018; Fujisawa et al., 2019; Fujisawa et al., 2021).
F. Biased Signaling
Where there is a receptor with several primordial signaling pathways, there is potential for biased signaling. For orexin receptors, the question has been raised in several reviews (e.g., Kukkonen, 2012; Leonard and Kukkonen, 2014; Kukkonen, 2019; Kukkonen and Turunen, 2021; Dale et al., 2022). Unfortunately, primary studies investigating the issue, even with regard to the physiological transmitters orexin-A and orexin-B, are rare. Since the original study (Sakurai et al., 1998), orexin-A is believed to be a more potent agonist on OX1 than orexin-B while the ligands would be equipotent on OX2. Information about the relative potencies of these two ligands for different responses in the same CHO-K1 cell clones has been collected in Table 1 of (Kukkonen, 2013). The analysis indicates that the EC50(orexin-B)/EC50(orexin-A) for OX1 varies between 1.6–18-fold – depending on the response measured – while in the literature, regarding different cell lines and clones, this has been suggested to be as much as 83-fold (Sakurai et al., 1998). The potencies for orexin-A, orexin-B and the putatively OX2-selective synthetic agonist, Ala11, d-Leu15-orexin-B, at OX1 and OX2 receptors in CHO-K1 and HEK293 cell lines have also been compared with respect to Ca2+ elevation. In general, the preference of orexin-B and Ala11, d-Leu15-orexin-B for OX2 was confirmed although the results in the two cell lines were significantly different (Putula et al., 2011). Along the same lines, the apparent OX2 receptor-selectivity of Ala11-d-Leu15-orexin-B is highly variable in different cellular and physiological studies (Asahi et al., 2003; Putula et al., 2011; Yamamoto et al., 2022). Note, that the potential off-target effects of Ala11-d-Leu15-orexin-B have not been assessed.
The small molecule agonist Nag 26 was compared for OX1 and OX2 receptor-mediated responses; some indication of biased signaling was found (Rinne et al., 2019). Firazorexton/TAK-994 has been compared with native orexin peptides for different responses mediated by the OX2 receptor and shown to be equally potent and efficacious as orexin-B with respect to PLC activation, but clearly less potent (or efficacious) than orexin-B with respect to β-arrestin recruitment and ERK and CREB (cAMP response element-binding protein) phosphorylation (Ishikawa et al., 2023); however, this may alternatively be explained by lower intrinsic efficacy of firazorexton/TAK-994.
Thus, there is limited evidence of biased agonism for the few orexin receptor agonists that have been studied, but the primary studies are limited and more extensive work is required, especially with a greater number of agonists. The novel methods to directly measure G protein and other primary pathways (II.A.1.) would be preferred although it should be recognized that these studies are not always in perfect agreement either. Independently of the results of such studies, the use of agonists for orexin receptor classification in experimental work is not recommended; antagonist studies are more conclusive, especially when determining the rank orders of antagonistic potency.
G. Orexin Receptor Desensitization and Trafficking
GPCR stimulation is classically associated with receptor desensitization, internalization and downregulation, though these processes affect different GPCRs to different degrees; in certain cases, such phenomena may actually result in a redirection of the receptor signaling, most notably upon the action of β-arrestin as an adaptor protein for both internalization and signaling (Shenoy and Lefkowitz, 2011). Kinases such as GRKs (GPCR kinases) and adaptor proteins β-arrestin-1 and -2 are often but not always associated with these processes (Shenoy and Lefkowitz, 2011; Magalhaes et al., 2012; Moo et al., 2021). In recombinant HEK293 cells, both orexin receptor subtypes interact with both isoforms of recombinant β-arrestin upon orexin-A challenge; the interaction of OX2 with both β-arrestin isoforms is proposed to be more sustained than that of OX1 (Dalrymple et al., 2011). After a prolonged agonist challenge, the receptors and β-arrestin internalize together (Evans et al., 2001; Dalrymple et al., 2011). The C-termini of both orexin receptors are at least partially involved in this interaction (Milasta et al., 2005; Dalrymple et al., 2011; Jaeger et al., 2014); whether interaction with β-arrestin is required for internalization remains unclear due to contradictory data (Milasta et al., 2005). Blocking the interaction with β-arrestin-2 using receptor mutagenesis shortens the time of sustained ERK activation by OX1 receptors, whereas the Ca2+ response is apparently not affected (Milasta et al., 2005). When quantified, the decrease in cell surface expression after orexin stimulation was quite substantial for both OX1 and OX2 receptors in HEK293 cells (>50%), but much more limited for OX2 receptors in CHO-K1 cells (∼25%) (Dalrymple et al., 2011; Ward et al., 2011; Kumagai et al., 2015). However, there was no attempt to quantify the potential functional effect of this on receptor signaling. CHO-K1 cells expressing either orexin receptor subtype can be stimulated repeatedly with orexin-A with some minutes of washout between stimulations without any apparent decrease in the Ca2+ response (Johansson et al., 2007; Kukkonen et al., unpublished). OX1 receptor-expressing CHO-K1 cells can be stimulated with 100 nM orexin-A continuously for 60 minutes with little apparent decrease in the Ca2+ signal (Ammoun et al., 2006a). OX1 receptors give a much more sustained ERK activation in CHO-K1 cells (Ammoun et al., 2006a) than in HEK293 cells (Milasta et al., 2005; Dalrymple et al., 2011). It is difficult to draw firm conclusions, but the combined evidence from the studies thus appears to indicate that orexin receptor internalization is stronger in HEK293 than in CHO-K1 cells.
The results from cell lines tell us little about the therapeutically central question, whether a drug will retain its clinical efficacy and potency upon long-term use. The apparent sensitization or desensitization on a systemic level cannot be traced back to a specific mechanism but is a sum of effects on different levels. Orexin peptides and small molecular agonists ameliorate narcolepsy symptoms in mouse models and human patients (V.B.2.). Repeated orexin receptor stimulation with small molecule orexin receptor agonists has been tested on the systemic level in a few studies. No reduction in the effect was observed in narcoleptic PPO-KO mice upon YNT-185 administration during the active periods of three consecutive days (Irukayama-Tomobe et al., 2017). After a two-week-treatment of narcoleptic PPO-ataxin-3 mice (in which the orexinergic neurons die postnatally) during the daily active periods, the responses to TAK-925 and TAK-994 were either not at all or only mildly reduced (Evans et al., 2022; Ishikawa et al., 2022; Ishikawa et al., 2023). In NT1 patients (IV.C.), a similar protocol has not been applied but the effect of TAK-994 was sustained for an 8-week-treatment period (Dauvilliers et al., 2023). Along the same lines, unpublished clinical data presented at the recent 2023 World Sleep Congress with various new OX2 agonists demonstrate sustained effects on various daily sleep parameters in narcolepsy patient populations, implying that receptor desensitization is not an issue in patient populations with narcolepsy type 1 or 2 (NT1 and NT2, respectively) or other excessive daytime hypersomnias (see also Jacobson et al., 2022). This is reminiscent of that observed with chronic somatostatin receptor agonists such as octreotide, used to treat acromegaly and other endocrine disorders, given over several decades (Weckbecker et al., 2003; Hoyer and Bartfai, 2012). It should also be noted that ectopic overexpression of PPO gave an amelioration of narcoleptic symptoms in PPO-ataxin-3 mice, though the persistence of the effect was not defined in the study (Mieda et al., 2004). Nevertheless, the findings on the cellular and systemic level are quite similar.
Compensatory effects may also be seen upon long-term antagonist treatment. However, orexin receptor antagonists in clinical use or testing retain their efficacy without the need to increase the dose during long-term treatment: Suvorexant, lemborexant and daridorexant were the longest ever tested hypnotics in phase 3 studies, some of which lasted for up to 14 months (Hoyer et al., 2020; Jacobson et al., 2022).
III. Orexin Receptor Structure
A. Receptor 3D Structures
Human orexin receptor 3D structures were initially constructed by homology modeling based on known GPCR crystal structures and used for modeling of the potential binding modes for receptor ligands and molecular dynamics simulations (Malherbe et al., 2010; Tran et al., 2011; Heifetz et al., 2012; Karhu et al., 2015; Turku et al., 2016; Turku et al., 2019). The first (inactive) orexin receptor 3D structure (OX2–suvorexant) was published in 2015, followed by OX1 structures in 2016 (OX1–SB-674942 and OX1–suvorexant) and then other bound ligands (Yin et al., 2015; Yin et al., 2016; Suno et al., 2018; Hellmann et al., 2020; Rappas et al., 2020; Hong et al., 2021; Asada et al., 2022; Yin et al., 2022) [Protein Data Bank (PDB) codes for OX1: 4ZJ8, 4ZJC, 6V9S, 6TQ9, 6TQ7, 6TQ6, 6TQ4, 6TP6, 6TP4, 6TP3, 6TOT, 6TOS, 6TOD, 6TO7; PDB codes for OX2: 4S0V, 5WS3, 5WQC, 7L1U, 7L1V, 6TPG, 6TPJ, 6TPN, 7SR8, 7SQO, 7XRR]. The studies have been mostly performed using X-ray crystallography except for two, where cryo-electron microscopy (cryo-EM) was used. The latter investigated the putatively active receptor conformations: one with the OX2 receptor in complex with orexin-B and the small molecule agonist compound 1 (PDB: 7L1U and 7L1V, respectively) and the other with OX2 receptor in complex with a Gq-mimetic (mini-Gαs/q/iN/Gβ/Gγ) or Gi1 and a small molecule agonist danavorexton/TAK-925 (PDB: 7SR8 and 7SQO, respectively) (Hong et al., 2021; Yin et al., 2022).
When bound to OX2, only the C-terminal part of orexin-B (N20–M28) had high enough density and was included in the final structure (Hong et al., 2021). This visible peptide region adopts extended conformation and does not retain the α-helix that is seen in the nuclear magnetic resonance (NMR) solvent structures of orexin-A and -B (see V.A.), which have been used in molecular modeling studies (Heifetz et al., 2012; Karhu et al., 2015; Karhu et al., 2019). In contrast, the sodium dodecyl sulfate (SDS) micelle-bound NMR structure of orexin-A C-terminus is less well defined (leaving a possibility of conformation freedom) and orexin-B is unstructured in its extreme C-terminus; thus the micelle-bound peptide structures (Miskolzie and Kotovych, 2003; Miskolzie et al., 2003) are somewhat more similar to the secondary structure of the receptor-bound orexin-B than are those in aqueous solvent (V.A.). The peptide bound to OX2 is in contact with TM2–TM7 as well as the extracellular loops 2 and 3 (ECL2 and -3) and the N-terminus (Hong et al., 2021). The peptide residues G24–M28 bind deep in the binding pocket complementing the space. I25 is sandwiched, M28 forms lipophilic contacts with OX2, and there are hydrogen bonds between T27 and Asn3246.55 and Gln1343.32 as well as L26 and His3507.39, likely through a water molecule. Note that the orexin peptide amino acids are here given with single letter codes and the receptor amino acids with three letter codes. Because of the high similarity of the C-terminal region of orexin-A and orexin-B, it is likely that they form similar interactions in this region. The region N20–A23 has several polar interactions with the widened upper part of the binding pocket (Hong et al., 2021). The N-terminal regions of orexin-B and the extracellular regions of OX2 receptor were of low resolution (low density) in cryo-EM. The model was complemented by molecular dynamics simulations, where the N-terminal part of the peptide adopts an α-helical conformation, and forms contacts with the α-helical N-terminus of OX2 as well as the ECL2, indicating a role of extracellular receptor parts in orexin peptide binding (Hong et al., 2021), although the N-terminal part of neither orexin peptide is highly significant for receptor binding (V.A.).
The main differences between the putative active and inactive conformations of OX2 receptors are in TM5–TM7, in a manner that has been described as the conserved mechanism of GPCR activation (Manglik and Kruse, 2017; Weis and Kobilka, 2018), in addition to several microswitches such as Asn-Pro-x-x-Tyr (a.k.a. NPxxY) motif and the agonist connector motif in the base of the orthosteric binding pocket (Pro-Val-Phe in orexin receptors) (Hong et al., 2021).
Both putative active structures with small molecule agonists (danavorexton/TAK-925 and compound 1; Fig. 4B2) are well aligned – despite the use of the nanobody Sb51 to obtain the compound 1-bound structure – showing a root mean square deviation of atomic positions (RMSD) of 0.6 Å when overall structures are compared (Hong et al., 2021; Yin et al., 2022). The level of difference between the peptide-bound structure and compound 1-bound structure is also similar, having an RMSD of 0.54 Å, when the overall structures (distinguishable α-carbons in the receptors and G proteins) are compared; the main difference is seen in the ECL sites in contact with Sb51 (Hong et al., 2021). Inactive OX1 (SB-674042-bound, 4ZJC) and OX2 structures (suvorexant-bound, 4S0V) also align well, having an RMSD of 0.4 Å over the distinguishable α-carbons of the receptors, mostly differing in the ECL2 and N-terminus (Yin et al., 2016).
These studies show that the small molecule binding sites in OX2 and OX1 are highly similar, differing only in two residues: Thr1112.61 versus Ser1032.61 and Thr1353.33 versus Ala1273.33, which have been suggested to be important for subtype selectivity, possibly through water network interactions (Malherbe et al., 2010; Tran et al., 2011; Yin et al., 2015; Yin et al., 2016; Rappas et al., 2020; Hong et al., 2021; Yin et al., 2022). These positions contribute to the selectivity of danavorexton/TAK-925 (>10 000-fold OX2 receptor-selective) as determined by reciprocal mutations: The potency at OX1S103T, A127T was more than 300-fold higher as compared with wildtype OX1 while the potency at OX2T111S, T135A was 500-fold lower as compared with wildtype OX2 (Yin et al., 2022). Interestingly, mutating Thr1112.61 of OX2 to Ser or Ala did not have any effect on the binding of the DORA lemborexant (Fig. 4A3), while Thr1353.33Ala mutation led to a 10-fold decreased affinity (Asada et al., 2022). Among wildtype receptors, lemborexant prefers OX2 only 2.3-fold (Beuckmann et al., 2017; Asada et al., 2022). Being a smaller residue, Ala1273.33 allows two configurations for lemborexant as compared with one for threonine, which may explain the difference in the affinity in favor of threonine (Asada et al., 2022). The Ala1273.33Thr mutation in the OX1 receptor reduces the binding affinity of the OX1-selective SB-674042 20-fold (Malherbe et al., 2010). The position 3.33 may also explain the selectivity of SB-674042 for OX1: The larger residue threonine (as compared with alanine) in OX2 creates a steric clash with SB-674042, and reciprocal mutations increase the affinity of OX2 while decreasing the affinity of OX1, whereas the binding of DORA-12 (Fig. 4A3) is not affected (Malherbe et al., 2010; Tran et al., 2011; Yin et al., 2016). A recent study presented an NMR analysis of a mutated OX2 receptor labeled with 13C-labeled methionine upon binding of the OX2 orexin receptor-selective antagonist (2-SORA) EMPA and the DORA suvorexant (Imamura et al., 2024). The analysis revealed conformational heterogeneity of the suvorexant-bound complex, in which suvorexant adopted multiple poses and the whole complex acquired multiple states. In contrast, EMPA was more rigidly fitted into OX2 binding pocket, showing strong hydrogen bonds to T1353.33 and P1313.29. In molecular dynamics simulations, residues responsible of subtype-specificity, T1112.61 and T1353.33, were affected differently: Suvorexant showed higher affinity toward T1112.61, while EMPA toward T1353.33, which was in line with the conclusion of heterogeneity caused by T1353.33 (Imamura et al., 2024).
Asn3246.55 of OX2 plays an important role for the orexin-B binding mode (hydrogen bonding) and receptor activation, while its role in danavorexton/TAK-925 binding is less prominent (Hong et al., 2021; Yin et al., 2022). The Asn3246.55Ala mutation is better tolerated by danavorexton/TAK-925 than orexin-B with respect to PLC activation (Yin et al., 2022). Orexin-B, compound 1 and danavorexton/TAK-925 share an important contact with Gln1343.32, which is critical for many GPCRs (Venkatakrishnan et al., 2013): Alanine mutation in this position markedly reduces responses to orexin-A, orexin-B and danavorexton/TAK-925 (Ca2+ or PLC activity) (Malherbe et al., 2010; Hong et al., 2021; Yin et al., 2022). Residues at the base of the binding pocket (Gln1343.32 as well as Tyr3176.48) show a clear effect in the β-arrestin recruitment assay for danavorexton/TAK-925. Both residues undergo large conformational changes between inactive and active receptor forms: For instance, Gln1343.32 projects upwards in the active and downward in the inactive structures (Yin et al., 2015; Yin et al., 2016; Rappas et al., 2020; Hong et al., 2021; Asada et al., 2022; Yin et al., 2022). Interestingly, Gln126/1343.32 (OX1 and OX2 position numbers, respectively) also adopts different conformations between OX1 and OX2 upon the DORA lemborexant binding: In both cases, lemborexant forms a hydrogen bond to Gln126/1343.32, but the bond is to the sidechain nitrogen in OX1 and to the sidechain oxygen in OX2. Gln1263.32Ala reduces the affinity of OX1 for SB-674042 50-fold (Malherbe et al., 2010); Gln126 has the largest surface contact with SB-674042, when compared with any other residue (Yin et al., 2016). Mutation of Gln1343.32Ala in OX2 reduces the affinity for the DORA lemborexant 50-fold, but only fourfold for the DORA suvorexant whereas the binding of the OX2-selective antagonist EMPA is not affected (Asada et al., 2022).
Lemborexant was also uniquely sandwiched by Pro1313.29 and His3507.39 in OX2 while other small molecule antagonists (EMPA: 5WQC, suvorexant: 4S0V, HTL6641: 6TPN) formed a water-mediated hydrogen bond with His3507.39 (Yin et al., 2015; Suno et al., 2018; Rappas et al., 2020). For compound 1 binding, the contact is not water-mediated (Hong et al., 2021). Mutation of His3507.39 to alanine was deleterious for lemborexant and suvorexant – but not EMPA – binding to OX2 (Asada et al., 2022). His3447.39 in OX1 has also been described to be important for SB-674042 binding by forming an aromatic contact with the ligand (Malherbe et al., 2010; Yin et al., 2016).
B. Receptor Dimers or Oligomers
GPCRs have been proposed to form di- or oligomeric homo- or heteromeric complexes (reviewed in Milligan, 2009; Farran, 2017). The current view is that all GPCRs of the (originally defined) class A and class C make homo- or heteromeric complexes, and possibly all GPCRs do. Assay methods do not reveal the exact number of protomers and thus the complexes are often referred to as dimers. Dimerization is an obligatory feature for (possibly all) class C receptors (Jones et al., 1998; Kaupmann et al., 1998; White et al., 1998; Kuner et al., 1999; Wu et al., 2014; Mao et al., 2020; Papasergi-Scott et al., 2020; Gao et al., 2021), while the case for class A receptors is unclear. For class A GPCRs, heterodimerization has been demonstrated to affect properties such as trafficking, ligand affinity and potency and receptor signaling pathways, but in the vast majority of cases the significance of dimerization is unknown (Maggio et al., 2007; Milligan, 2009; Gomes et al., 2016; Gaitonde and Gonzalez-Maeso, 2017). Both human orexin receptor subtypes have been demonstrated to form homo- and heteromeric complexes with each other and with every GPCR tested in recombinant setting. Complex partners determined to date for the OX1 receptor include apelin, CB1, CCK1, CRF1, CRF2, ghrelin, GPR103 and κ opioid receptors, and for the OX2 receptor, 5-HT1A, CB1 and GPR103 receptors (reviewed in Kukkonen, 2017; Kukkonen and Turunen, 2021). The non-GPCR σ1 and σ2 receptors have also been suggested to complex with orexin receptors (Navarro et al., 2015; Navarro et al., 2019). Heterodimerization of OX1 and κ opioid receptors has been proposed to shift the signaling from Gq and Gi, respectively, of the homomeric receptors to Gs (Chen et al., 2015). However, another study demonstrated a functional interaction between OX1 and κ opioid receptors (Robinson and McDonald, 2015), providing an alternative explanation; this is similar to the question regarding dimerization versus functional interaction of OX1 and CB1 receptors (Jäntti et al., 2013). So far, the only reported complex with a likely physiological significance is that between OX1, CRF1 and σ1 receptors in the rat VTA (Navarro et al., 2015). However, there are also issues with this study regarding the in vitro and in vivo methodology for testing of the inhibition of dimerization.
A fully reliable demonstration of dimerization/oligomerization is not simple. Colocalization demonstrated with high-resolution microscopy does not prove anything else but colocalization, though comobility may be indicative for complexes. Bioluminescence and fluorescence resonance energy transfer (BRET and FRET, respectively) may be slightly more specific, but not all receptor complexes work well in these assays. Bimolecular fluorescence complementation (BiFC) has the built-in problems that the two pieces of the fluorescent protein have affinity for each other and thus promote dimerization, and that the complex formation is irreversible (Kerppola, 2006). Importantly, these methods are difficult to apply to native receptors. Antibody-based methods such as the proximity ligation assay also only prove colocalization and are totally dependent on the antibodies' specificities (generally low for orexin receptors). Native gel electrophoresis may miss dynamic interactions and is not able to determine the molecular weights of the complexes. Physiological rather than molecular interactions need to be excluded (Turunen et al., 2012; Jäntti et al., 2013; Robinson and McDonald, 2015).
As concerns orexin receptor complexes, there is thus little evidence of any physiological significance, although some of the findings may be interesting as such. In most cases, the receptors reported to form complexes with orexin receptors upon recombinant overexpression have not been demonstrated to be expressed in the same native cells. The specificity of cell-permeating peptides used to inhibit dimerization is by no means clear among GPCRs or other proteins (see, e.g., Ekokoski et al., 2010) and there is no unequivocal theoretical framework to support specificity. A single study has tackled the question of whether the orexin receptor (homomeric) complexes are stable or transient (Xu et al., 2011). The issues with these methods are additionally discussed in Kukkonen, 2017, 2019.
IV. Orexin Physiology
Orexin neurons of the lateral hypothalamus receive a variety of inputs encoding information about, for instance, circadian rhythm, metabolic and energy status, as well as positive and negative emotions to guide their physiological output. Orexin neurons express multiple receptor and ion channel types (Fig. 5). They integrate this information and direct the output to multiple neuromodulatory regions influencing physiological functions such as wakefulness and arousal, feeding, motivated behaviors for natural or drug rewards, antinociception, neuroinflammatory and cell death responses, as well as sympathetic, cardiovascular and respiratory responses (Fig. 6). Yet it is still unclear to which degree subdivisions of orexin neurons with functional specialization exist. The physiological effects of orexin as well as pathophysiological implications are discussed in the sections below.
The cre–lox recombination system is an essential tool in modern physiological research, and also many of the physiological roles of the orexin system described in this review have been revealed upon expression of other constructs in these neurons utilizing Hcrt promoter-driven cre expression. However, there are potential issues with this method, discussed under VI.
A. Orexin Neurons and Their Projections
Orexin neurons are found in the perifornical, dorsal and lateral hypothalamic areas, as determined in several mammalian species using antibodies against PPO, detection of PPO mRNA, labeling of the orexin neurons with the EGFP gene under the Hcrt promoter (Li et al., 2002) and tracing of the orexin neurons from target areas (more recent studies: Ciriello et al., 2013; Ciriello and Caverson, 2014; Garcia-Garcia et al., 2013; Yokota et al., 2016; reviewed in Sakurai, 2007; Kukkonen, 2013; Sakurai et al., 2021). Isolated orexin neurons have been reported in other nearby regions by peptide detection (reviewed in Kukkonen et al., 2002) and recently, based on PPO mRNA, even on clearly separate CNS regions (see below). In humans, the number of orexin neurons is estimated at 80 000 at most and just a few thousand or hundred in rats and mice (Peyron et al., 1998; Harrison et al., 1999; Fronczek et al., 2005; Thannickal et al., 2000; Fronczek et al., 2007; Thannickal et al., 2009; Obukuro et al., 2010; Black et al., 2018; Gonzalez and Prehn, 2018; Berteotti et al., 2021) (see also IV.C.1. and IV.D. for the observed changes in the neuron count). In the lateral hypothalamus, orexin neurons are interspersed among other neuron types, e.g., melanin-concentrating hormone (MCH)-ergic neurons (Elias et al., 1998; reviewed in Kukkonen, 2013). There has been considerable interest in finding molecularly or functionally distinct subpopulations of orexin neurons but with few notable results so far. It has been suggested that rats have anatomically distinct neuronal populations (medial versus lateral), based on the projection sites, whereas in mice, the neurons with different projection sites appear intermixed (reviewed in Sagi et al., 2021). In addition to tracing studies, functional and expression analyses in mice identify potential subsets of neurons, though these also appear intermingled (Schöne et al., 2011; Dalal et al., 2013; Mickelsen et al., 2017). Lateral hypothalamic neurons – between populations or within them – have been suggested to be minimally connected with each other (Burdakov and Karnani, 2020), although this may not align with previous anatomical and functional studies (reviewed in Kukkonen, 2013). If true (and this would not be unusual in the brain in, e.g., the hypothalamus or dorsal raphe nucleus), it would allow subpopulations of orexin neurons (or even individual neurons) to act independently in different functions.
No other unique marker for orexin neurons, other than PPO, has thus far been unequivocally reported (Dalal et al., 2013; Seifinejad et al., 2019; reviewed in Kukkonen, 2013). Orexin neurons express, in addition to orexins, other neuropeptides/neurotransmitters, e.g., dynorphin (and other prodynorphin products in different degrees), neurotensin, glutamate and amylin (these neuromodulators are present in most, if not all neurons) as well as galanin and GABA (present in a fraction of neurons) (Ciriello et al., 2013; Furutani et al., 2013; Muschamp et al., 2014; Apergis-Schoute et al., 2015; Li et al., 2015; Dergacheva et al., 2016a; Mickelsen et al., 2017; reviewed in Kukkonen, 2013); these cotransmitters may support the actions of each other (Schöne et al., 2014; Concetti and Burdakov, 2021). The neurons also express neuronal activity-regulated pentraxin II (NARP/NPTX2) (Reti et al., 2002), an early intermediate gene that plays a critical role in guiding synaptic plasticity.
Orexin neurons are regulated by many transmitter systems, and the neuronal pathways for some of these have been identified by retrograde or anterograde tracing while others remain speculative. Orexin neurons express many different types of receptors based either on functional studies (reviewed in Kukkonen, 2013) or mRNA expression studies (Harthoorn et al., 2005; Dalal et al., 2013). Figure 5 summarizes the transmitters regulating orexin neurons. Note that the studies cannot identify which regulatory systems apply to all orexin neurons versus only subpopulations of neurons, as would seem sensible from the functional perspective. Some studies have suggested that orexin neurons also contain orexin receptors (see Kukkonen, 2013) while others dispute this view (Vassalli et al., 2015). A recent study demonstrates that about 5% of the orexin-A-positive cells express either OX1 or OX2 receptor mRNA (Tsuneoka and Funato, 2024).
Dietary cues represent another central regulator of the activity of orexin neurons. Fasting increases PPO mRNA expression (Sakurai et al., 1998). Glucose directly hyperpolarizes the neurons via unidentified leak-type K+ channels and hypoglycemia depolarizes them (reviewed in Kukkonen, 2013). This has been suggested to give rise to the compensatory adrenal catecholamine secretion via stimulation of OX2 receptors in the rostral ventrolateral medulla (Korim et al., 2016) and (partially) to pancreatic glucagon secretion via OX1 receptors in the dorsal vagal motor nucleus (Paranjape et al., 2007), though these suggestions are only based on the effects of orexin receptor antagonists. Other studies have demonstrated that the glucose-mediated hyperpolarization of orexin neurons is only transient and thus may represent a sensor for the change in glucose levels rather than for absolute glucose levels (Williams et al., 2008). In contrast, lactate and nonessential amino acids depolarize orexin neurons (Parsons and Hirasawa, 2010; Karnani et al., 2011a; Viskaitis et al., 2022). Orexin neurons are also pH-sensitive, responding to acid with depolarization (IV.I.).
An early study found that the proximal promoter region (spanning 3.2 kB upstream of the Hcrt gene) was sufficient to direct the expression to orexin neurons (I.C.) (Sakurai et al., 1999; Moriguchi et al., 2002). Several transcription factors and other proteins are suggested either to repress (interferon-α and IGFBP3; Waleh et al., 2001; Honda et al., 2009) or to promote PPO expression (FOXA2, NR6A1, EBF2 and PLAGL1; Silva et al., 2009; Tanaka et al., 2010; Sanchez-Garcia et al., 2018; Tanaka et al., 2019). On the network level, orexin neurons have been suggested to be activated by, for instance, hunger (IV.B.), wakefulness (IV.C.), anticipation (at least in motor activity context; IV.D. and IV.J.), pain (IV.G.), stress and anxiety (Cohen et al., 2020; Yamashita et al., 2021), reward (IV.D.) (Hassani et al., 2016) and imminent cataplexy (in PPO-KO mice; Zhou et al., 2022) while inflammation may silence orexin signaling of these neurons (Grossberg et al., 2011; Seifinejad et al., 2023).
The activity of orexin neurons and the orexin production/release have been suggested to show diurnal rhythmicity in animal models (Taheri et al., 2000; Yoshida et al., 2001; Martinez et al., 2002). The original orexin measurements were performed using antibodies and are thus somewhat uncertain, given that these antibodies may possibly also detect the parent peptide PPO, orexin-A fragments or altogether different molecules (VI.), but measurements with mass spectrometry-based methods or biosensors, which are not subject to these liabilities, have also recently confirmed the diurnal rhythmicity of orexin release (Hopkins et al., 2021; Duffet et al., 2022a; Narita et al., 2023).
Potential targets for orexin neurons have been mapped by the methods described above as well as by detection of synapse formation, orexin receptor mRNA expression, and functional testing. These aspects have been thoroughly reviewed previously (see, e.g., Sakurai, 2007; Kukkonen, 2013; Sakurai et al., 2021) and thus are not specifically dealt with here. In general, orexinergic projections from hypothalamus span all major parts of the brain, though only specific sites. Early studies reported very sparse innervation of the cerebellum (Kukkonen et al., 2002) and thus the role of orexins there was questioned (Sakurai, 2007). More recent studies have identified some potential orexin actions there as well (Yu et al., 2010; Nisimaru et al., 2013; Ciriello and Caverson, 2014; Zhang et al., 2024b), but another recent study finds no orexin receptor mRNA in the cerebellum (Tsuneoka and Funato, 2024).
Studies suggest that orexinergic projections to the neocortex are scarce though significant. Originally, only layer 6b was pointed out (Bayer et al., 2004) and thus most studies have focused on it (Wenger Combremont et al., 2016; Hay et al., 2015; Zolnik et al., 2024), yet a recent study demonstrates orexin receptor mRNA expression in several cortical layers (Tsuneoka and Funato, 2024). 3H-EMPA autoradiography was used to map OX2 receptor expression in the rat brain and peripheral tissues (Mitsukawa and Kimura, 2022); the results are in general agreement with findings of receptor mRNA in situ hybridization (Marcus et al., 2001; Trivedi et al., 1998), although higher receptor densities are detected by autoradiography, in particular in some cortical areas. By contrast, some ex vivo studies using isolated rodent cortical neurons suggest that a large proportion of total cortical neurons respond to orexins (Urbanska et al., 2012; Sokolowska et al., 2014; Palomba et al., 2020). It is possible that the culture of the embryonic or neonatal neurons prior to experiments changes the properties of the neurons or selects particular neuron subpopulations (Urbanska et al., 2012; Sokolowska et al., 2014; Palomba et al., 2020). Nevertheless, there may be more orexin activity in neocortex than previously assumed as revealed in the recent mapping of the peptidergic neuronal networks in the human prefrontal cortex by mRNA analysis (Zhong et al., 2022), demonstrating wide-spread expression of OX2 mRNA and limited expression of OX1 mRNA in these regions. The authors also suggested that PPO mRNA is expressed in pyramidal glutamatergic projection neurons and in GABAergic interneurons, in addition to the putative “classical” projection neurons from the hypothalamus; it is unclear whether the Hcrt promoter acts in the same way in these neurons as in the hypothalamus. Thus far there is, however, no other evidence to support these findings (i.e., orexin peptide or receptor expression, orexin receptor responses). In transgenic rodents, ectopic PPO expression may occur extrahypothalamically due to other processes (VI.). Altogether, only very few cells and regions seem to express mRNA for both orexin receptor subtypes (Tsuneoka and Funato, 2024).
There is certainly stronger evidence for some connections than for others. Additional confusion is generated by some methodological issues. Simple stimulation of neurons by exogenous ligands does not prove that there is endogenous signaling via the receptors, and many studies have not properly distinguished between pre- and postsynaptic effects. In essence, tetrodotoxin is not sufficient to eliminate all potential effects on presynaptic terminals, though it blocks action potentials and thus input to transmitter release. We have here altogether excluded the information based on the use of orexin receptor antibodies since they are likely to produce misleading results due the generally poor immunogenicity of GPCR epitopes (see, e.g., Dahl et al., 2023); specific evidence as to this has been obtained with orexin receptors (Kukkonen, 2012; Kukkonen, 2013). Ideally, validation of such antibodies should be performed in cells, tissues and/or animals from which the target receptor has been knocked out/down, but such controls are still not commonly performed.
B. Orexins in Feeding and Metabolism
Originally, orexin receptors were associated with appetite stimulation and arousal and maintenance of wakefulness. Appetite regulation was first reported by Sakurai and coworkers, since fasting increased brain PPO mRNA expression and icv injection of orexin-A or -B increased short-term feeding in fed rats (Sakurai et al., 1998). A major reason for testing appetite regulation originated from the hypothalamic expression of orexins. Orexin-A induced feeding more potently than orexin-B and thus the receptor of interest was thought to be OX1 (Sakurai et al., 1998). Several studies replicated such findings in both rats and mice (reviewed in Kukkonen et al., 2002). As described above, orexin neurons are regulated by glucose, lactate and nonessential amino acids. A more detailed mapping of the orexin projections and orexin receptor mRNA distribution identified orexinergic innervation of feeding-associated nuclei such as the arcuate nucleus (Broberger et al., 1998; Elias et al., 1998; Peyron et al., 1998). Furthermore, injection of orexin-A into specific brain sites, such as the arcuate nucleus, the perifornical hypothalamic area, the rostral portion of the lateral hypothalamus, the paraventricular nucleus of the hypothalamus (PVN), and the nucleus accumbens, but not the substantia nigra, increased food intake (Dube et al., 1999; Kotz et al., 2002; Thorpe and Kotz, 2005; Kotz et al., 2008; Yang et al., 2018). Orexin neurons are also regulated by feeding-associated neuronal and endocrine input (IV.A.) (Figs. 7 and 9 in Kukkonen et al., 2002; Fig. 8 in Kukkonen, 2013).
Orexins were soon found to increase metabolic rate in rodents (Lubkin and Stricker-Krongrad, 1998), largely through activation of BAT via second order neurons (Oldfield et al., 2002; Berthoud et al., 2005; Zheng et al., 2005; Tupone et al., 2011; Martins et al., 2016). Orexins also increase locomotor activity (Hagan et al., 1999; Yamanaka et al., 2003a; Garau et al., 2020; Karnani et al., 2020). Thus, at least in rodents, orexins could also lead to increased energy expenditure. This was further validated by ectopic overexpression of PPO which protected mice from diet-induced obesity by burning excess calories in BAT (Funato et al., 2009). Interestingly, orexins produced by the placenta have been proposed to be required for prenatal BAT development in mice (Sellayah et al., 2011), although this view has been challenged (Kakizaki et al., 2019). Overweight and obesity are more common in human NT1 patients resulting in an average 10–20% higher body mass index (BMI) than in healthy controls (Kok et al., 2003; Scammell, 2015; Bassetti et al., 2019) although their basal metabolic rate is similar to BMI-matched controls (Fronczek et al., 2008; Dahmen et al., 2009) – assuming the comparison with BMI-matched controls is relevant. It should, however, be kept in mind that the loss of orexin neurons would not only affect orexins, but also other neurotransmitter systems involved in feeding and weight control, including the transmitters coexpressed in orexin neurons. Of note, the effects on weight gain in transgenic rodent models of deficient orexin signaling under normal and high-calorie food intake are not fully consistent with the human situation and may be mouse strain- and sex-dependent (Hara et al., 2005; Fujiki et al., 2006; Kakizaki et al., 2019).
Cold is a central trigger for BAT metabolic activity. PPO-ataxin-3 rats showed reduced thermogenesis upon cold exposure which may be a result of reduction in both BAT activation and locomotor activity as compared with normal controls (Mohammed et al., 2016). By contrast, in a study of PPO-ataxin-3 and PPO-KO mice, locomotor activity of both strains was normal and PPO-KO mice showed normal thermogenesis (Takahashi et al., 2013). Along these lines, human NT1 patients show normal BAT activation upon cold exposure (Enevoldsen et al., 2018). In addition to methodological differences between these studies, it remains possible that the role of orexins in BAT regulation in rodents and humans is different, since the structure and function of BAT are also distinct in adult rodents and humans (Wu et al., 2012; de Jong et al., 2019). Furthermore, different mouse models of orexin- or orexin neuron-deficiency have produced inconsistent results (Kukkonen, 2013; Takahashi et al., 2013; Inutsuka et al., 2014; Mohammed et al., 2016).
It has been suggested that central orexins regulate peripheral energy metabolism not only via BAT and locomotor activity, but may also involve hormonal regulators such as glucagon, insulin and adrenocorticotropic hormone, although the physiological significance remains unclear (Kukkonen, 2013). More recently, it was shown that a high fat diet induces steatohepatitis, hepatic inflammation and hepatocellular carcinoma more strongly in PPO-KO mice than in wildtype mice (Tsuneki et al., 2022). The occurrence of hepatic inflammation was largely abolished by icv orexin supplementation whereas obesity and fatty liver were not affected (Tsuneki et al., 2022). On the other hand, increased orexin signaling by optogenetic stimulation of orexin neurons reduced glucose tolerance in obese mice on a high fat diet (Xiao et al., 2021). Importantly, in the latter study, the expression of channelrhodopsin-2 in orexin neurons was (partially) validated, as the specificity of expression for transgenes under the proximal orexin promotor can strongly vary (see section VI. for discussion). Indeed, some ectopic expression was detected and not all expression was quantitated, introducing an aspect with which to question these results. However, the questionability of the results with orexin–cre lines in general is much higher when there is no attempt to assess the efficiency and specificity of the expression, as is the case in most studies (VI.). Finally, peripheral tumor-induced metabolic (higher plasma glucose, lower glucose tolerance) and sleep disturbances co-occur with abnormal activity of orexin neurons and are inhibited by blocking orexin receptors with the DORA almorexant (Fig. 4A3) (Borniger et al., 2018).
Overall, based on transgenic animal models as well as animal and human studies with orexin receptor antagonists (V.B.1.), it seems appetite regulation is not a major physiological role of orexins, or at least not one that can be easily manipulated, and thus appetite suppression appears unlikely to be a major therapeutic indication for orexin receptor antagonists (V.B.1.). This may be partially explained by the complicated relationship between appetite, metabolism and reward in orexin signaling (see below), or altogether lower significance of orexins in the regulation of appetite. While orexin neurons respond to energy shortage and orexins obviously increase feeding behaviors in the short term, the role of orexins in this may be rather in activating vigilance and seeking of essential nutrients (Yamanaka et al., 2003a; Linehan and Hirasawa, 2022; Viskaitis et al., 2022; reviewed in Peleg-Raibstein et al., 2023). Nevertheless, there are ongoing clinical trials with OX1 receptor-selective antagonists (1-SORAs), such as JNJ-61393215, CVN45502 or CVN7666 (Fig. 4A1) in feeding-related conditions.
Studies with orexin receptor antagonists with regard to appetite and feeding are described under V.B.1. For the role of orexins in the reward and motivated behavior associated with food ingestion, please see IV.D.
C. Orexins in Wakefulness and Sleep
The central role of orexins in the regulation of wakefulness and sleep emerged soon after the discovery of orexins. Central application of orexin-A icv or locally at structures involved in sleep–wake control increased wakefulness in rats (Hagan et al., 1999; Bourgin et al., 2000; Methippara et al., 2000; Piper et al., 2000). Soon after the discovery of the orexin system, hereditary canine narcolepsy was found to result from three distinct OX2-inactivating Hcrtr2 mutations in three breeds of dogs (Lin et al., 1999; Hungs et al., 2001). Along the same lines, PPO-KO mice did not lose weight as expected based on the hypothesis of reduced appetite, but instead suffered from narcolepsy including cataplectic attacks (Chemelli et al., 1999), i.e., rapid, transient loss of skeletal muscle tone (except for respiratory function and eye movement). Cataplexy is typically induced by strong positive or negative emotions (laughter, fear) in humans suffering from narcolepsy with cataplexy, a.k.a. NT1. The cataplectic features of narcoleptic dogs and PPO-KO mice expanded to the investigation of the role of the orexin system in human NT1 subjects. Unlike in dogs, there is no evidence for OX2 mutations in human NT1 but, instead, the patients have low orexin-A levels in their CSF and lack of orexin expression in the lateral hypothalamus (IV.C.1.). Further studies in rodents demonstrated that narcolepsy could also be induced by postnatal genetic destruction of the orexin neurons in mice and rats via the toxin ataxin-3 expressed under the Hcrt promoter (Hara et al., 2001; Beuckmann et al., 2004; Mieda et al., 2004), postnatal toxic destruction of the orexin target neurons by orexin-B conjugated with the toxin saporin (Gerashchenko et al., 2001) or by deletion of both Hcrtr1 and Hcrtr2 genes (double-KO) in mice (Willie et al., 2003). Interestingly, OX1-KO mice show only minimal sleep–wake dysfunction, whereas OX2-KO mice show a modest sleep phenotype with very rare cataplectic attacks (Willie et al., 2003), which contrasts with the receptor double-KO where cataplexy is relatively easily triggered (Kisanuki et al., 2001; Willie et al., 2003; Sakurai, 2007; Kalogiannis et al., 2011; Mang et al., 2012), like in the PPO-KO mice (Chemelli et al., 1999; Espana et al., 2007; Clark et al., 2009; Morawska et al., 2011; Oishi et al., 2013). Of note, selective knockout of OX2 in dopaminergic neurons gives paradoxical hyperarousal (Bandarabadi et al., 2023).
While the symptoms of NT1 patients and the narcoleptic phenotype in animals were major indicators of a central role of orexins in the regulation of sleep and wakefulness, converging evidence was also obtained from healthy animals, i.e., centrally applied orexin-A increased wakefulness and reduced sleep (Hagan et al., 1999; Bourgin et al., 2000; Methippara et al., 2000; Piper et al., 2000). This was also supported by mapping of the orexinergic projections to wake-promoting nuclei (Trivedi et al., 1998; Date et al., 1999; Horvath et al., 1999b; Mondal et al., 1999; Marcus et al., 2001). Further studies have refined and expanded these findings. Optogenetic activation of orexin neurons in mice stimulates wakefulness while their silencing induces sleep (Adamantidis et al., 2007; Tsunematsu et al., 2011; Carter et al., 2012; Tsunematsu et al., 2012; Tsunematsu et al., 2013). Similarly, optogenetic stimulation of vasopressinergic neurons of the hypothalamic paraventricular nucleus increases wakefulness via activation of orexin neurons (Islam et al., 2022). However, it should be noted that optogenetic constructs needed for the light-mediated control of the neurons may have additional effects: High expression of archeorhodopsin 3 has been reported to silence orexin neurons even in the absence of illumination (Williams et al., 2019).
Exogenous expression of PPO results in a partial recovery from narcoleptic-like symptoms in PPO-ataxin-3 mice (Mieda et al., 2004; Liu et al., 2011b) and exogenous stimulation with synthetic OX2 receptor agonists improves the symptoms in these and PPO-KO mice (Evans et al., 2022; Yamamoto et al., 2022) as well as in NT1 and NT2 patients (Evans et al., 2022). On the other hand, orexin receptor antagonists, both DORAs and 2-SORAs, are effective in inducing and maintaining sleep (V.B.1.). Evidence for the role of orexins in sleep-wakefulness circuitry has been reviewed previously (Sakurai, 2007; Scammell et al., 2017; Sakurai et al., 2021; Seifinejad et al., 2021; Jacobson et al., 2022). In short, orexin neurons are most active during wakefulness, somewhat active during nonrapid eye movement (NREM) sleep and cease to fire upon transition to rapid eye movement (REM) sleep – except for a subpopulation that is weakly active during REM sleep (Ito et al., 2023). Orexin neurons stimulate the wake-active aminergic neurons of the brain stem. They are themselves inhibited by these neurons and the preoptic area (POA) GABAergic NREM sleep-active neurons as sleep pressure develops. In addition, there is reciprocal inhibition between the aminergic neurons and the GABAergic neurons. During wakefulness, orexinergic drive activates aminergic neurons, which increases the inhibition of POA and allows cortical activation. Orexin neurons also regulate the wake-active/rapid eye movement (REM) sleep-active cholinergic neurons of the brain stem and the basal forebrain. The neuronal circuitry involved in sleep regulation, including the effects of orexins, has been mathematically modeled (Sorooshyari et al., 2015). With increased sleep drive, the activity of the POA increases, suppressing the activity of orexin neurons and the aminergic neurons. When orexin is lost in NT1, the reciprocal inhibition between the aminergic neurons and POA is disrupted and becomes a chaotic oscillator. The regulation of REM sleep is less clear (Scammell et al., 2017; Sakurai et al., 2021; Seifinejad et al., 2021), and the role of orexin receptor subtypes in REM versus NREM sleep (or inhibition of these) is different (Mieda et al., 2011; Clark et al., 2020; Sun et al., 2021; Ito et al., 2023). There is also a potential link between orexins and melatonin: a) orexinergic neurons project to the pineal gland, b) there are orexin receptors in the pineal gland and c) exogenous orexins inhibit or stimulate melatonin expression, depending on other stimuli (Mikkelsen et al., 2001; Fabris et al., 2004; Zhang et al., 2005; Appelbaum et al., 2009; Zieba et al., 2011; Kirsz et al., 2020). It has also been suggested that there are inhibitory melatonin receptors on orexinergic neurons, but the evidence is only based on receptor antibodies and no direct functional effects have yet been demonstrated (Sharma et al., 2018).
The phenotype of narcoleptic canines and orexin receptor KO-mice puts OX2 in the focus for regulation of wakefulness and sleep. It is, yet, possible that, at least in rodents, both orexin receptors affect sleep, as seen in the orexin receptor knockout mice and by comparing the effects of OX1 or OX2 antagonism alone and in combination on sleep. However, the effect of OX1 elimination/inhibition on sleep is only seen when OX2 is inhibited at the same time (Willie et al., 2003; Dugovic et al., 2009; Mang et al., 2012; Betschart et al., 2013; Hoyer et al., 2013; Dugovic et al., 2014; Gotter et al., 2016; Brooks et al., 2023). Various pharmaceutical companies have thus been developing either 2-SORAs (Fig. 4A2) or DORAs (Fig. 4A3). There are currently three DORAs on the market as hypnotics, namely suvorexant, lemborexant and daridorexant (Fig. 4A3), while further DORAs as well as a few 2-SORAs are under development. The evidence and the ideas regarding antagonism of orexin receptor subtypes versus REM and NREM sleep are presented under V.B.1.b. The ligand discovery has significantly contributed to the understanding of the respective role of the two orexin receptor subtypes in the regulation of sleep and wakefulness.
Insomnia and reduced sleep quality become more frequent in humans and animals upon aging. A recent mouse study suggests this is caused by hyperactivity of the orexin neurons (Li et al., 2022c), despite the concurrent loss of orexin neurons. One of the major functions of sleep, in addition to memory consolidation, is suggested to be “clearing” of toxic waste products from the brain by the glymphatic system (reviewed in Rasmussen et al., 2018). Note that a recent study (Miao et al., 2024) disputes the view of the positive impact of sleep on brain clearance. The substances cleared from the brain include (phospho-)tau and amyloid β, linking sleep and AD (Rasmussen et al., 2018), and consequently the orexin system via regulation of sleep and wakefulness. Wakefulness, and especially sleep deprivation, increases soluble amyloid β in the brain interstitial fluid and CSF, whereas the DORA almorexant reduces this burden (Kang et al., 2009). Limited human studies with respect to NT1 do not allow clear conclusions to be drawn about the relationship between orexin signaling and AD (Gabelle et al., 2019; Dauvilliers, 2021; Lucey et al., 2023), and the orexin receptor antagonists have not yet been on the market long enough to test this concept. The potential interplay of orexins, neuronal degeneration and sleep disturbances is discussed under IV.F.
1. Orexins and Human Narcolepsy
Narcolepsy is a disorder affecting the regulation of sleep and wakefulness (Bassetti et al., 2019; Mahoney et al., 2019). In humans, it is expressed in two forms, narcolepsy type 1 (NT1) and type 2 (NT2). Classic symptoms of the fulminant form, NT1, include excessive daytime sleepiness and dysregulation of various features of REM sleep, such as short REM sleep latency after falling asleep, hypnagogic and hypnopompic hallucinations, sleep paralysis and cataplectic attacks (Bassetti et al., 2019; Mahoney et al., 2019). Encephalitides can present with similar symptoms, but narcolepsy diagnosis usually requires that the symptoms are limited to the regulation of sleep–wakefulness. Narcolepsy is nearly always a sporadic disease with young adult onset in humans. 98% of patients have the HLA DQB1*0602 genotype (Juji et al., 1984; Mignot et al., 1994; Scammell, 2015), which is the main reason why narcolepsy is considered an autoimmune disease (Scammell, 2015; Bassetti et al., 2019). In contrast to the autoimmune or other types of encephalitides, narcolepsy is often slow to progress to symptoms other than excessive daytime sleepiness and may only be expressed after several years in its fulminant form including cataplexy. Human NT1 patients show very low levels of CSF orexin-A (Nishino et al., 2000). In those few studies that have assessed postmortem brains of NT1 patients, a severe decrease of orexin-producing neurons in the hypothalamus has been reported (Thannickal et al., 2000; Thannickal et al., 2003; Blouin et al., 2005; Thannickal et al., 2009; Shan et al., 2022). According to the current hypothesis, the orexin neurons of the hypothalamus are thus thought to degenerate by an autoimmune process. It was recently demonstrated that corticotropin-releasing hormone production in the paraventricular nucleus of hypothalamus is also concomitantly lost (Shan et al., 2022). Furthermore, it was proposed that orexin-producing neurons are not lost but are unable to produce PPO due to epigenetic silencing caused by inflammation, and that the expression of dynorphin and corticotropin-releasing hormone in the hypothalamus is affected in the same way (Seifinejad et al., 2023). This concept could work in line of the idea that some orexin-producing neurons may express immunohistochemically subdetectable levels of PPO resulting in apparently different orexin neuron counts under different conditions (McGregor et al., 2017; Thannickal et al., 2018; Fragale et al., 2021a; McGregor et al., 2024) (see also IV.D.).
In addition to the hereditary narcolepsy (I.C.2.), dogs also suffer from a sporadic form (Tonokura et al., 2007); in these animals, orexin production is significantly reduced as in human NT1 (Ripley et al., 2001), suggesting a similar pathogenesis. There are polymorphisms in human OX1 and OX2 genes (HCRTR1 and -2, respectively) but these are not associated with narcolepsy (Peyron et al., 2000; Olafsdottir et al., 2001). One reported mutation of the human PPO gene (HCRT) apparently causes narcolepsy (Peyron et al., 2000).
Either cataplexy or a CSF orexin-A level below 110 pg/ml is required for a NT1 diagnosis – in addition to a set of other symptoms (American Psychiatric Association, 2013; American Academy of Sleep Medicine, 2014). Either finding is sufficient for the diagnosis, although the low CSF orexin-A levels and cataplexy should go hand in hand (Scammell, 2015; Bassetti et al., 2019). However, there are, on the one hand, patients with cataplexy in spite of apparently normal orexin-A levels, and on the other, patients with reduced CSF orexin-A levels and narcolepsy symptoms but no cataplexy. Association of low CSF orexin-A with major dysfunction of the orexin system that should result in cataplexy may thus not be fully valid, but also the issues with CSF orexin determinations may contribute to this (VI.). The diagnosis of NT2 requires the set of narcolepsy symptoms to be present, but the CSF orexin-A levels must be normal and cataplexy must not be exhibited (American Psychiatric Association, 2013; American Academy of Sleep Medicine, 2014).
With regard to mechanisms underlying cataplexy, it seems logical that the extended amygdala (central nucleus of the amygdala, basolateral amygdala and bed nucleus of stria terminalis) plays a central role in initiating cataplexy as observed in functional imaging studies performed in rodent models of narcolepsy and NT1 patients (Ballotta et al., 2021; Sardar et al., 2023). Moreover, activation of the nucleus accumbens, by optogenetic or chemogenetic stimulation, facilitates induction, but not maintenance of cataplectic events in rodents (Su et al., 2020). However, chemogenetic inhibition of the nucleus accumbens inhibits cataplexy while optogenetic inhibition does not (Su et al., 2020; Kawashima et al., 2023). Orexin neurons have been shown to increase their firing rate before initiation of cataplectic episode but to decrease their firing rate below the baseline during the episode in PPO-KO mice (Zhou et al., 2022).
OX2 agonists have been investigated as a therapeutic strategy for narcolepsy. Indeed, they have been found to ameliorate symptoms of orexin deficiency including those of narcolepsy in animal models, for instance by increasing wakefulness and activity and reducing cataplexy (see V.B.2.), cementing the role of OX2 receptor signaling in this situation.
In 2009, the world became subject to the H1N1/09 influenza A virus pandemic ("swine flu"). Soon after the start of the vaccinations, reports of rapidly developing fulminant narcolepsy among children, adolescents and young adults began to appear, particularly in Sweden and Finland (Swedish Medical Products Agency, 2011; Nohynek et al., 2012). The cases were soon linked to the Pandemrix vaccine (Swedish Medical Products Agency, 2011; Nohynek et al., 2012). Among the HLA DQB1*0602-positive children and adolescents in Sweden, the risk for developing narcolepsy was estimated to be elevated 49-fold by the vaccination (Hallberg et al., 2019), though the absolute incidence was still very low: The overall incidence was in total 1 per 18400 vaccinations and among HLA DQB1*0602-positive individuals 1 per 4500 vaccinations (Hallberg et al., 2019). The risk increase varied between different countries (Sarkanen et al., 2018). All the genotyped children/adolescents, at least in Finland, had the HLA DQB1*0602 genotype and all tested also had low CSF orexin-A levels (Partinen et al., 2012). However, the clinical picture of the vaccination-associated narcolepsy was atypical: The onset was very rapid and dramatic, usually culminating in cataplexy, and in 50% of the cases psychiatric symptoms were also reported (Partinen et al., 2012). No increased risk of narcolepsy was seen for vaccines other than Pandemrix (Sarkanen et al., 2018), and thus the focus was turned to specific vaccine constituents. The adjuvant AS03 was initially suspected, but dismissed, since another vaccine, Arepanrix, also contained this adjuvant (Vaarala et al., 2014). The focus turned to the antigens used in the vaccines – which also distinguishes Pandemrix and Arepanrix. In 2013, it was reported that H1N1/09 hemagglutinin, contained in the vaccine, harbored an epitope with homology to PPO and that the patients who developed narcolepsy after Pandemrix vaccinations had CD4+ T cells reactive to PPO (De la Herran-Arita et al., 2013) but this report was soon retracted. In 2015, another group reported that H1N1/09 nucleoprotein, contained in Pandemrix, harbored an epitope with higher homology to the N-terminus of the OX2 receptor and lower homology to the N-terminus of the OX1 receptor, and that the patients who developed narcolepsy after Pandemrix vaccination had highly selective antibodies against the OX2 receptor (Ahmed et al., 2015). Several studies using different patient sample sets and methods have attempted to reproduce the latter finding, but with no success (Bergman et al., 2014; Giannoccaro et al., 2017; Luo et al., 2017; Melen et al., 2020); instead, the specificity of some of the original methods have been questioned. Furthermore, it also seems unlikely that antibodies against orexin receptors would cause the loss of orexin neurons. Thus, it remains unclear what the causal mechanism (if any) underlying the association of Pandemrix vaccination and elevated narcolepsy risk is. The investigations as to this mechanism have a specific scientific value concerning narcolepsy and a general one as regards production of vaccines. No medical product is risk-free, and we should note that the swine flu vaccines saved lives.
There is a historically earlier example of an epidemic disease with some features common with narcolepsy, namely encephalitis lethargica. Since its outbreak occurred during the Spanish flu pandemic, it is still debated whether there is any connection between these two (da Mota Gomes, 2020; Brigo and Vogrig, 2023). It is interesting to note that the causative agent behind Spanish flu was also an H1N1 strain. It is clear that the symptoms of encephalitis lethargica were for a large part very different from narcolepsy (da Mota Gomes, 2020; Brigo and Vogrig, 2023) – but even in the Pandemrix-associated narcolepsy, the symptoms are wider than in the sporadic narcolepsy (Partinen et al., 2012).
The autoimmune hypothesis for both sporadic and Pandemrix-associated NT1 remains current. For decades before the swine flu, autoantibodies and autoreactive T cells have been sought in the blood samples of patients with narcolepsy. Both healthy individuals and patients may have antibodies that stain diverse neuronal populations in the brain, possibly even orexin neurons, and thus no clear conclusions can be drawn from these data (Bergman et al., 2014; personal communication with Dr. Krister Eriksson). Autoreactive T cells have been investigated in a few studies with interesting but thus far inconclusive results (Latorre et al., 2018; Luo et al., 2021; Vuorela et al., 2021). Molecular mimicry of orexin-A and -B peptides by H1N1 hemagglutinin and nucleoprotein peptides was suggested (Luo et al., 2018). However, there is no primary sequence homology between these sequences. The findings were instead based on crossreactivity of T cells and some association with the HLA DQB1*0602 and T cell receptor polymorphisms predisposing to NT1; thus, the mimicry would rather be based on how the peptides are presented and recognized, although a further study found that there was no crossreactivity (Luo et al., 2022). T cell receptor polymorphisms exposing for narcolepsy have been reported in other studies as well (Ollila et al., 2023). H1N1 hemagglutinin has been heterologously expressed in mouse orexin neurons where it can, maybe less surprisingly, induce neuroinflammation and death of orexin neurons in a process requiring both CD4+ and CD8+ T cells, when exposed to “autoreactive” T cells or Pandemrix vaccine (Bernard-Valnet et al., 2016; Bernard-Valnet et al., 2022). Interestingly, H1N1 itself has been shown to induce encephalopathy (including cell death in the areas involved in the regulation of sleep and wakefulness) in mice devoid of B and T cells (Tesoriero et al., 2016). Recently, it was shown that mice vaccinated with Pandemrix had reduced PPO mRNA expression in the lateral hypothalamic area, thought there was no indication of inflammation and none of the mice in the small group (24 Pandemrix-treated animals) developed narcolepsy (Pagh-Berendtsen et al., 2024).
Unique antigens on orexin neurons have been searched for as potential epitopes for specific immunological attack but with little success (IV.A.). In some cases (Pandemrix involved or not), there have been preceding infections – such are known to trigger (other) autoimmune diseases – but the evidence is limited (Aran et al., 2009; Ambati et al., 2015). As noted above, the hypothesis of the autoimmune pathology of narcolepsy – also other than that caused by Pandemrix – is primarily based on the strong overrepresentation of HLA DQB1*0602 and HLA DRB1*15:01 carriers among affected individuals (of certain ethnic backgrounds), yet the vast majority of the carriers of these HLA molecules never develop narcolepsy. Studies of association of narcolepsy with other single gene polymorphisms report low odds ratios (<2) (Hallmayer et al., 2009; Ouyang et al., 2020; Miyagawa and Tokunaga, 2019). Regardless of the disease etiology and pathogenesis, orexin receptor agonists are expected to be useful therapeutic agents for the treatment of narcolepsy symptoms such as excessive daytime sleepiness (V.B.2.). In contrast, all DORAs are currently contra-indicated in narcolepsy mainly due to the risk of increased cataplexy, as seen in one animal study (Mahoney et al., 2020), and at low frequency in clinical studies (V.B.1.).
D. Orexins in Motivation and Addiction
The orexin system appears to play a major role in motivation and addiction. Activation of orexin receptors is associated with the motivation to work for different kinds of palatable foods (Thorpe et al., 2005; Borgland et al., 2009; Choi et al., 2010; Kay et al., 2014; Castro et al., 2016; Terrill et al., 2016). Injection of orexin-A into the nucleus accumbens shell increases motivation for food in a progressive ratio task, i.e., in which the effort (i.e., number of lever presses) required to receive a food reward, such as a sucrose pellet, progressively increases (Thorpe and Kotz, 2005; Choi et al., 2010). The 1-SORA SB-334867 has been reported to reduce motivation for (Thorpe and Kotz, 2005; Nair et al., 2008; Borgland et al., 2009) and consumption of (Borgland et al., 2009; Choi et al., 2010) palatable food rewards in a progressive ratio task but a recent study utilizing three different 1-SORAs (including SB-334867) questions this with respect to sucrose (Bergamini et al., 2024). Orexin neurons are activated in drug- or food-paired contexts (Harris et al., 2005) or in response to stimuli that predict the availability of food (Campbell et al., 2017). Furthermore, optogenetic stimulation of orexin terminals in the VTA, a region that encodes the significance of motivationally relevant stimuli, increases dopamine release in the nucleus accumbens as well as preference for contexts associated with that optical stimulation (Thomas et al., 2022). Orexin neuronal activity also correlates with reward anticipation in a Go/No-Go task (Tyree et al., 2023). In this study, optogenetic stimulation of orexin neurons prior to the cue increased premature responses in the Go/NoGo task, and the effect was blocked by an undisclosed 1-SORA suggesting a role of orexin in impulse control. In several studies, human NT1 patients were shown to have a reduced ability in specific decision-making tasks as compared with NT2 patients or controls, but whether this is associated with impulsivity is debated (Bayard et al., 2011; Dimitrova et al., 2011; Bayard et al., 2013). On the other hand, orexin signaling has been associated with risk avoidance via the nucleus accumbens (Blomeley et al., 2018).
Given the role of orexin signaling in the motivation for food, orexin receptor antagonists have been proposed as a pharmacotherapy for binge eating disorder. In animal models systemic administration of 1-SORAs and DORAs, but not 2-SORAs, reduced binge-like consumption of palatable food at doses that do not affect homeostatic eating (Piccoli et al., 2012; Alcaraz-Iborra et al., 2014; Rorabaugh et al., 2014; Olney et al., 2015; Vickers et al., 2015). These studies suggested that OX1 antagonism may be effective at reducing binge eating in humans, but the results of a phase 2 clinical trial (NCT04753164; see https://clinicaltrials.gov/) with the 1-SORA, ACT-539313 (Fig. 4A1), have been disappointing (V.B.1.).
A role for orexin signaling in addiction was first suggested from behavioral studies showing that orexin signaling can reinstate extinguished place preference paired with morphine and that inhibition of OX1 can block reinstatement of place preference or drug self-administration in rats (Boutrel et al., 2005; Harris et al., 2005). Orexin receptor activation contributes to the reinforcement, spatial and contextual memory and stress-associated drug relapse (reviewed in Baimel and Borgland, 2017; James et al., 2017; Fragale et al., 2021b). Many preclinical studies claim a major involvement of OX1 receptors in drug seeking, although some studies implicate OX2 over OX1 for alcohol self-administration, conditioned place preference, and reinstatement (Shoblock et al., 2011). Addiction is associated with neural circuit dysfunction characterized by serial changes in synaptic transmission in the VTA. Morphine or cocaine-induced neuroplasticity within the VTA requires OX1 signaling at dopamine neurons (Baimel and Borgland, 2017). Interestingly, elimination of OX2 receptors in dopamine neurons led to faster acquisition of an operant task but more premature and perseverative responses, suggesting impaired inhibitory control, while motivation remained unchanged (Bandarabadi et al., 2023). NT1 patients had normal reinforcement learning when tested in a task known to be dopamine-dependent. However, they had decreased learning to avoid losses, suggesting that individuals with NT1 had reduced sensitivity to punishment, an effect likely linked to altered vigilance, as this was also seen in other central disorders of hypersomnolesence (Strauss et al., 2024) and even reported in controls deprived of sleep (Gerhardsson et al., 2021). Furthermore, exposure to cocaine, morphine or stress increases the activity of orexin neurons which may, in turn, further drive orexin inputs to the VTA, modulating motivated behavior while simultaneously promoting arousal states leading to insomnia, a common comorbidity of substance use (Fragale et al., 2021b). Indeed, an OX1 receptor-preferring agonist, YNT-3708 (Fig. 4B1), increases conditioned place preference for cocaine (Iio et al., 2023). Thus, concern has been raised regarding the safety of using orexin receptor activators (e.g., in narcolepsy and other sleep disorders), whereas orexin receptor antagonists (especially 1-SORAs) may be protective (Fragale et al., 2021b; Zlebnik et al., 2021). However, current agonists in development for NT1 and other sleep disorders are targeted specifically to the OX2 and there is little reported research activity on OX1 agonists to date (V.B.2.).
DORAs have proved effective in animal models in decreasing cocaine-related impulsivity (Gentile et al., 2018b), cocaine motivation (Gentile et al., 2018a) and opioid seeking (Illenberger et al., 2023). Furthermore, Illenberger et al. (2023) recently showed that suvorexant blocked oxycodone seeking in male Wistar rats, while reducing it in females which tended to self-administer oxycodone twice as much as males. This idea has initiated several clinical trials for off-label use of the DORAs suvorexant and lemborexant (Fig. 4A3) for the treatment of opioid, cocaine, alcohol, or nicotine use disorders (e.g., NCT04229095, NCT03937986, NCT03999099, NCT04818086, NCT05145764) and substance use disorders comorbid with sleep disorders (NCT03897062, NCT05458609, NCT04287062). While most trials have not been completed yet, early results using suvorexant have been somewhat surprising. In a small study of non-treatment-seeking cocaine users (n = 7), suvorexant enhanced the reinforcing effects of low-dose intravenous cocaine and increased the motivation to choose cocaine over monetary rewards (Stoops et al., 2022). At clinical doses, the abuse potential of suvorexant and lemborexant appears low, although there may be small differences in abuse potential at higher doses of these drugs (Schoedel et al., 2016; Born et al., 2017; Moline et al., 2023). In a study examining the efficacy of suvorexant on reducing opioid withdrawal-related insomnia, there was an increase in sleep duration and no abuse potential of suvorexant (Huhn et al., 2022). Some studies have shown an increase in number of orexin-expressing neurons in opioid users and animal models of opioid or cocaine use disorder (Thannickal et al., 2018; James et al., 2019; Fragale et al., 2021a). This effect is likely due to increased orexin peptide expression levels in individual neurons allowing these neurons to show above the detection threshold. In addition, decrease in orexin cell body size has been observed in opioid users (Thannickal et al., 2018). Interestingly, even the number of orexin-expressing fibers within the VTA is increased after morphine treatment and repeated suvorexant treatment during morphine administration prevents this increase (McGregor et al., 2024). Therefore, it will be interesting to evaluate the effects of DORAs in pending clinical trials for substance use disorders as they are completed and published in peer reviewed journals. The 1-SORA JNJ-61393215 (Fig. 4A1) is being developed, for instance, for addiction. No data on its impact in addiction setting have been published yet, but a recent phase 1 study reported no sleepiness or other significant CNS side effects (Brooks et al., 2023).
E. Orexins and Other Psychiatric Disorders
A potential role for orexins in anxiety and depression is not immediately apparent, yet it seems reasonable to associate orexins with anxiety-like behavior because they increase vigilance, activity and stress response (reviewed in Kukkonen, 2013) and may reduce risk-taking (Blomeley et al., 2018) in rodent models. However, they also increase motivated behaviors (IV.D.) that may counteract stress (reviewed in Peleg-Raibstein and Burdakov, 2021). Whether the orexin system is involved in innate anxiety is not clear, and animal models of anxiety are not without their pitfalls (reviewed in Jacobson and Cryan, 2010; Peleg-Raibstein and Burdakov, 2021). Nevertheless, central administration of orexins increases anxiety-like behavior (reviewed in Kukkonen, 2013), whereas OX1 antagonism reduces anxiety in a variety of anxiety and stress paradigms in both rodents and humans (Flores et al., 2015; Bonaventure et al., 2017; Salvadore et al., 2020; Yaeger et al., 2022; Ten-Blanco et al., 2023). On the other hand, there are also instances where orexin shows anxiolytic effects via OX2 activation (Zhang et al., 2024b; reviewed in Kukkonen, 2013). In a posttraumatic stress model in rats, orexin-A, applied before the stressful stimulus, reduced the development of stress-related behavior whereas almorexant did the opposite (Cohen et al., 2020). Hypothalamic deep brain stimulation was identified as a potential anxiolytic intervention in a mouse model, and it was suggested that the action requires orexinergic neuronal activity (Li et al., 2022a). Insulin-like growth factor I (IGF-1) activates orexin neurons and loss of this signaling by IGF-1 receptor-KO in these neurons exposes animals to stronger anxiety response (Zegarra-Valdivia et al., 2020; Fernandez de Sevilla et al., 2022). Escitalopram, used commonly in both major depressive disorder (MDD) and in some anxiety disorders, was recently shown to upregulate orexin-A expression (Golyszny et al., 2022). These discrepancies may relate to a) different etiology of anxiety forms, b) potentially different roles of orexin signaling in the different stages (e.g., exposure to the anxiogenic stimulus versus the later symptomatic phase) and c) issues with the animal models. A few clinical studies have been performed or initiated to investigate the effect of orexin receptor antagonism in anxiety and stress, sometimes in connection with sleep. (However, note that when searching at https://clinicaltrials.gov/, the majority of clinical studies with orexin receptor antagonists do not contain orexin or hypocretin as a searchable term.) Interestingly, suvorexant was shown to reduce anticipatory anxiety in healthy humans (Gorka et al., 2022), raising the question whether this links to the role of orexins in anticipation (Donegan et al., 2022b) or anxiety (or both). The 1-SORA compound 56 (Fig. 4A1) has been reported to reduce stress-induced hyperarousal in rats without any hypnotic effects (Bonaventure et al., 2015) and another 1-SORA, JNJ-54717793 (Fig. 4A1), reduces panic attacks in rats (Bonaventure et al., 2017). Currently, the 1-SORA JNJ-612393215 (Fig. 4A1) is undergoing clinical trials for anxiety disorders (Brooks et al., 2023).
Animal models of depression suffer from similar problems as the models of anxiety (Cryan and Holmes, 2005; Jacobson and Cryan, 2007), and the results for orexins in this domain are not clear. In general, DORAs and to some extent 2-SORAs, but not 1-SORAs, are antidepressive, but the pharmacological investigations with SORAs and DORAs aiming at amelioration of the symptoms of are not unequivocal (V.B.1.c.). By contrast, it has been proposed that depression in humans correlates with lower CSF orexin levels (reviewed in Peleg-Raibstein and Burdakov, 2021) – based on the (unreliable) antibody-based orexin measurements in the brain or in the CSF (VI.). Some rodent studies suggest that orexin signaling itself has an antidepressant effect (Lutter et al., 2008; Scott et al., 2011). Insomnia is a common symptom in MDD and it is well-known that good sleep improves mood (reviewed in Riemann et al., 2020). Hence the question: Could improved sleep, by contributing to mood elevation, hinder the development of MDD (reviewed in Riemann et al., 2020)? In agreement, insomnia and/or hyperarousal are thought to increase the risk of depression (and metabolic disorders), providing potential links between depression and a hyperactive orexin system (reviewed in Khan and Aouad, 2017; Riemann et al., 2020). Some experimental evidence supports the involvement of orexins in this interplay (Nakamura and Nagamine, 2017), and there could be bidirectional interactions, with depression favoring insomnia and vice versa, resulting in a vicious cycle. With respect to psychosis, increased dopaminergic activity and putative behavioral correlates have been reported to be reduced by orexin receptor antagonists (Elam et al., 2021; Perez and Lodge, 2021). Cerevance has a 1-SORA, CVN766, under development for a variety of psychiatric conditions (V.B.1.c.).
Finally, both suvorexant and lemborexant have been found to effectively prevent delirium – which is considered a somatic rather than psychiatric condition – in small groups of hospitalized patients (Adams et al., 2020; Matsuoka et al., 2022, 2023).
F. Orexins in Inflammation and Neuronal Cell Plasticity and Death
The role of orexins in neuroinflammation and neurodegeneration including Parkinson's disease (PD) and AD is actively investigated. The questions are interesting and a potential role of orexins in such processes would not be surprising based on their other reported effects on cell plasticity, death and survival (see, e.g., Rouet-Benzineb et al., 2004; Ammoun et al., 2006a; Ammoun et al., 2006b; Sellayah and Sikder, 2012) (II.C–D.) However, many studies suffer from potential methodological issues, as discussed in Kukkonen and Turunen, 2021, and although exogenous orexin stimulation is generally suggested to be protective, conclusions are often difficult to draw. For instance, orexin-A is reported to alleviate neuronal cell death in experimental models of intracerebral hemorrhage (Kitamura et al., 2010). Different mechanisms such as induction of hypoxia-induced factor-1α (Yuan et al., 2011), reduced inflammation (Xiong et al., 2013), inhibition of endoplasmic reticulum stress (Xu et al., 2021) and inhibition of autophagy (Zhang et al., 2022) have been proposed, but the causalities are difficult to demonstrate. A chapter of its own are the cellular model systems. A very popular model for PD research with focus on orexins, SH-SY5Y neuroblastoma cells, are sympathetic ganglion-derived (hence peripheral, not central) noradrenergic (not dopaminergic) neurons, in which no expression of orexin receptors has been demonstrated to date. However, there also are potentially more interesting neuroprotective findings for orexin peptides in PD (Guerreiro et al., 2015; Liu et al., 2018b), though there are still confounding factors in these. One such study reports a functional motor improvement by exogenous orexins in the 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP)-induced PD mouse model (Wang et al., 2019b).
As concerns AD, there is more direct evidence regarding orexins. Brain interstitial fluid amyloid β and tau levels are aligned with the sleep–wake cycle, increasing during wakefulness and decreasing during sleep in mice (IV.C.) (Kang et al., 2009; Holth et al., 2019). Brain interstitial fluid amyloid β levels and accumulation in plaques in the brain are increased by sleep deprivation and icv orexin-A whereas they are decreased by the DORA almorexant (Fig. 4A3) below the background levels, as investigated in the Swedish amyloid precursor protein (APP) mutation and APP+presenilin-1 mutation mouse models (Tg2576 and APPswe/PS1dE9 mice, respectively) (Kang et al., 2009). Intranasal orexin-A worsens disease symptoms and increases amyloid β plaque accumulation in the 3 × Tg mouse model of amyloidogenesis and tauopathy (harboring the previous two mutations + the tau P301L mutation), potentially via increased β-secretase 1 (BACE1) and decreased neprilysin expression (Li et al., 2023). Both tau and amyloid β rodent models show hyperarousal (reviewed in Kron et al., 2024). Orexin neuron count and activation are elevated in the APPswe/PSldE9 mice (Zhao et al., 2022), while the tau P301L mutation mice show a selective reduction of OX1 mRNA in the locus coeruleus (Keenan et al., 2021). These two datasets appear conflicting, but might be explained by OX1 downregulation counteracting orexin system hyperactivity.
Some studies find elevated CSF orexin-A levels in AD patients, while others report decreased levels: A meta-analysis failed to detect a significant change in orexin levels in AD (Treu and Plante, 2021). The numbers of orexin-producing neurons and their projections are reduced in AD (Fronczek et al., 2012; Kasanuki et al., 2014; Oh et al., 2019). The apparently conflicting results may be explained by different disease stages, i.e., potentially an increase in early stage and reduction later upon the demise of orexin neurons. However, further consideration is required, given that the immunological methods regularly used to detect CSF orexin levels may lack specificity (VI.). The impact of orexins versus other transmitter systems in the sleep–wakefulness cycle is also difficult to pinpoint in the presence of the generalized neurodegeneration in advanced AD. In some studies, correlations of CSF orexin-A with CSF amyloid precursor protein fragments, tau and phospho-tau levels have been determined (Liguori et al., 2014). Currently, reduction in CSF amyloid β42 (or the ratio amyloid β42/β40) and increase in CSF tau and phospho-T181-tau are associated with AD in clinical diagnostics, though the causality behind this is not fully clear (Olsson et al., 2016; Delaby et al., 2022; Gorelick, 2023).
In summary, orexins stimulate wakefulness which may reduce the glymphatic clearance of amyloid β fragments and (phospho-)tau, which may increase plaque and neurofibrillary tangle formation. One hypothesis is that part of the pathology in AD and frontotemporal dementia may result from distortion of the sleep–wakefulness cycle upon which amyloid β fragments and/or (phospho-)tau, among others, are no longer effectively cleared from the brain by the glymphatic system, but this has not been validated. Orexin neuron overactivity may be seen in AD but it is unclear whether this is (part of) the cause or a result of the AD pathogenesis – or both (reviewed in Wang and Holtzman, 2020). Aging-enhanced orexinergic drive (Li et al., 2022c) should thus be considered with regard to the glymphatic system, should the relationship between sleep and glymphatic clearance hold true (Miao et al., 2024). Elderly patients with NT1 have been reported to have less amyloid plaque formation in the brain than age-matched controls (Gabelle et al., 2019) and the DORA suvorexant (Fig. 4A3) has been reported to have some positive effect in a mouse model of AD and in human AD patients (Herring et al., 2020; Zhou et al., 2020; Lucey et al., 2023). The I308V variant of OX2 has been reported to confer a slightly increased risk for AD (Gallone et al., 2014). However, the studies are thus far too preliminary and limited to allow conclusions, and the orexin receptor antagonists have not been on the market long enough to allow assessment in a cohort context. Further research in this area is thus clearly required. It should be noted that SH-SY5Y cells have been used as a cellular model in AD research as well, which is equally questionable as in PD.
Several studies address the role of orexins in immune regulation in the CNS or in the periphery. For instance, orexin-A reduces the expression and exosomal delivery of PD-L1 in colon cancer cell lines, which in turn promotes T cell activity against the cancer cells (Wen et al., 2022). Intranasally administered orexin-A reduces mRNA for inflammatory markers in a rat model of cardiac arrest-induced neuroinflammation (Modi et al., 2017). Intrathecally administered orexin-A reduces diabetic neuropathy in streptozocin-induced diabetes model in rats (Niknia et al., 2019). Icv orexin-A ameliorates experimental autoimmune encephalopathy in mice (Becquet et al., 2019). Palmitic acid activates the BV2 microglial cell line whereas orexin-A reduces this activity and the expression of inflammatory markers (Duffy et al., 2015). However, these are isolated findings and thus do not allow definitive conclusions about the physiological role of orexins in these processes or the molecular mechanisms involved.
G. Orexins and Antinociception
Orexin receptor activation, likely via OX1 receptors, is antinociceptive in animal models (Inutsuka et al., 2016; Iio et al., 2023; Nakamoto et al., 2023; reviewed in Kukkonen, 2013), possibly via endocannabinoid-mediated retrograde presynaptic inhibition (see II.B.6) especially in the periaqueductal gray (Ho et al., 2011; Lee et al., 2016; Chen et al., 2018; Lee et al., 2021). In contrast, orexin-induced analgesia is independent of opioids and opioid receptors, as assessed using naloxone (Bingham et al., 2001; Cheng et al., 2003; Kajiyama et al., 2005; Mobarakeh et al., 2005; Lee et al., 2016). Orexin neurons can be activated upon painful stimuli and their activity may be a normal physiological response to counteract pain (Zhu et al., 2002; Watanabe et al., 2005; Inutsuka et al., 2016). The novel OX1-selective agonist (R)-YNT-3708 (Fig. 4B1) was reported to be antinociceptive in the mouse tail-flick test (Iio et al., 2023). On the other hand, activation of orexin neurons has been suggested to be involved in itch response, since they are activated by intradermal chloroquine, a histamine-independent pruritogen, and ablation or inhibition of orexin neurons reduces acute and chronic itch; the periaqueductal gray has been associated with this as well (Kaneko et al., 2022; Kaneko et al., 2024). Interestingly, one major group of analgesics, opioids, can induce itch via μ receptor-dependent central mechanisms, i.e., independent of peripheral mechanisms such as histamine (Ikoma et al., 2006; Buddenkotte and Steinhoff, 2010). Potential linking of orexin and opioid systems in itch has not been assessed, but orexins are linked to opioids on the level of reward and addiction (IV.D.).
H. Orexins and Epilepsy
High doses of icv orexin-A can induce behavioral seizures, although this has not been scrutinized with EEG recordings (Ida et al., 1999). REM sleep is antiepileptic due to its attenuation of cortical synchronization, and the kindling model of epilepsy is enhanced by REM sleep deprivation (Frauscher et al., 2016; Ng, 2017). Orexins contribute to increased excitability in the brain (II.C.), hence a potential, though weak, link has been suggested between orexin activity and epilepsy. Blocking of either orexin receptor subtype reduces pentylenetetrazol-induced seizures in rats (Ni et al., 2014; Asadi et al., 2018). Hippocampal optogenetic stimulation-induced seizure intensity in mice is reduced by high-dose (likely nonselective between orexin receptors) SB-334867 (Fig. 4A1). Furthermore, it was found that orexin neuron preseizure activity correlated positively with the seizure intensity and that inhibition of this activity optogenetically or by deep brain stimulation reduced seizure intensity and/or probability (Li et al., 2024). Though sleep deprivation is known to promote epileptic seizures (Roliz and Kothare, 2023), these data indicate that the role of orexins in seizure activity is distinct from their role in sleep and wakefulness. Cannabidiol, used to treat a small subset of epilepsy patients (Franco et al., 2021), has been suggested to act as an orexin receptor antagonist (Vitale et al., 2021), although its affinity for orexin receptors appears too low to explain its physiological effects.
I. Sympathetic, Cardiovascular and Respiratory Regulation by Orexins
Orexin neurons project to brain stem sites regulating sympathetic outflow in rodents (reviewed in Kukkonen et al., 2002; Kukkonen, 2013). This leads to a) increased metabolic activity (at least partly by activation of BAT; IV.B.), and b) cardiovascular effects, e.g., increased mean arterial pressure and heart rate (Narai et al., 2024; reviewed in Kukkonen et al., 2002; Kukkonen, 2013). Both orexin receptor subtypes have been implicated (Shahid et al., 2012; Beig et al., 2015). Recently, intranasal orexin-A administration in humans was shown to increase sympathetic vasoconstrictor action, suggested to take place via changed baroreflex sensitivity, though without concomitant changes in blood pressure or heart rate (Meusel et al., 2022). NT1 patients may have higher heart rate than controls during sleep (Berteotti and Silvani, 2018) and cardiovascular disorders (and risk factors for such) are altogether more frequent in NT1 than in control populations (Jennum et al., 2021). However, other somatic and psychiatric comorbidities are also more frequent in NT1 and thus the causalities are not evident (Jennum et al., 2021). For findings concerning the metabolic rate in NT1 patients and the role of orexins in the heart, see IV.B. and IV.M., respectively.
Several studies point to a role for orexins in respiratory regulation. Orexin neurons project to the pons, medulla and spinal sites responsible for regulation of breathing in rats, exogenous orexins stimulate breathing via these sites, even via direct effect on motoneurons, and PPO-KO mice show impaired central hypercapnic response, which causes central sleep apnea (Young et al., 2005; Dutschmann et al., 2007; Nakamura et al., 2007; Kuwaki, 2008; Yokota et al., 2016). Both orexin receptors have been associated with respiratory regulation (Corcoran et al., 2010; Beig et al., 2015). However, injection of orexin-B or Nag 26 (Fig. 4B2) into the Kölliker–Fuse nucleus of anesthetized mice slowed down the respiratory rate (Varga et al., 2022), while orexin-B within spinal cord attenuated the respiratory depression (suppression of cervical C4 motor root activity) caused by sevoflurane, propofol and remifentanil in a rat medulla–spinal cord preparation (Umezawa et al., 2015).
Within the CNS, respiration is thought to be regulated by the sensing of carbon dioxide upon its acidifying effect (Sharabi et al., 2009; Nattie and Li, 2012). Orexin neurons are sensitive to extracellular pH (Burdakov et al., 2006; Williams et al., 2007). They express the acid-sensitive ion channels ASIC1a (Song et al., 2012) and TASK1 and -3 (Burdakov et al., 2006; Gonzalez et al., 2009b; Wang et al., 2021), although the specific role of each channel type is unclear and some results appear contradictory: ASIC inhibition fully blocks the acid-induced increase in respiratory drive (Song et al., 2012), TASK1 and -2 inhibition has a partial effect (Wang et al., 2021) while TASK1/3-KO does not affect the acid response on respiration (Gonzalez et al., 2009b). Since anandamide, bupivacaine and ruthenium red, used in (Wang et al., 2021), are not selective TASK inhibitors, one conclusion may be that the acid sensitivity of orexin neurons is ASIC1a-dependent (Song et al., 2012), but this is contrasted by other findings suggesting inhibition of a leak-like K+ conductance by lowered pH (Williams et al., 2007; Gonzalez et al., 2009b). Part of the response to acidification may come via orexin-mediated enhancement of acid responses of other nuclei (Lazarenko et al., 2011) or vice versa. Intrathecally administered orexin-A sensitizes the respiratory response to hypoxia in rats, which has been interpreted to originate from sensitized peripheral chemoreceptor output (Kim et al., 2016). It has also been suggested that the hypoxia-stimulated C1 cell group could activate orexin neurons (Kim et al., 2016). In contrast, hypoxia alone (weakly) or combined hypoxia and hypercapnia (more strongly) have been reported to inhibit orexin neurons in rat hypothalamic slice preparations (Dergacheva and Mendelowitz, 2017).
In summary, orexin neurons display acid sensitivity themselves and may also have contacts with other neurons with acid-, CO2- or O2-sensitivity. Whether orexins then increase or decrease respiration seems to depend on the preparation and conditions in each experimental setting, and thus the physiological role of orexins in the regulation of respiration is unclear. It should be noted that NT1 patients do not have a reduced hypercapnic response (Han et al., 2010) and the DORAs suvorexant and daridorexant (Fig. 4A3) in clinical use do not affect breathing in chronic obstructive pulmonary disease (with reduced hypercapnic response) or obstructive sleep apnea (Sun et al., 2015; Sun et al., 2016; Boof et al., 2021a; Boof et al., 2021b), contrasting with the phenotype of PPO-KO mice and the effect observed in neonatal rats (Nakamura et al., 2007; Kuwaki, 2008; Spinieli et al., 2021).
J. Orexins in Premotor and Motor Control
A few studies have identified orexinergic innervation of lower motoneurons to skeletal muscles and potential excitation of these synapses (Zhang and Luo, 2002; Xi et al., 2003) (see also V.I.). Excitation of brain stem motor nuclei has also been demonstrated (Zhang et al., 2011; Liu et al., 2018a). Lack of orexinergic tone contributes to the lack of muscle tonus during REM sleep and in cataplexy (Xi et al., 2002; Burgess and Peever, 2013). Orexins have been recently linked to anticipation and initiation of voluntary movements (Concetti and Burdakov, 2021; Donegan et al., 2022b) and even motor adaptation (Donegan et al., 2022a), though this may be difficult to distinguish from motor control (Tyree et al., 2023). On the other hand, orexins have been suggested to be centrally involved in the cardiovascular adaptation for locomotion (Narai et al., 2024).
K. Other Putative Orexin Functions in the CNS
Retinal signaling involving at least bipolar, amacrine and ganglion cells (including the melanopsin-expressing ones) is modified by orexins, as indicated by the endogenous expression of the peptides and the receptors, the responses to exogenous orexin-A and the impact of the DORA TCS1102 (Fig. 4A3) on the endogenous signaling (Liu et al., 2011a; Zheng et al., 2015; Ruan et al., 2021; Zhou et al., 2021). Interestingly, orexin signaling has been suggested to modulate processing of visual information in the brain (Chrobok et al., 2017; Chrobok et al., 2021b) and potentially the pupillary light reflex (Chrobok et al., 2021a; Grujic et al., 2023). Disturbances in the visual system are not reported as side-effects of the three clinically available DORAs, suggesting that the visual system is not a major target of orexins or that the DORAs in use do not enter the eye in a sufficient degree. However, the development of the orexin receptor agonist JZP-441 was recently halted partially due to visual disturbances (V.B.2.).
As described under IV.D., orexin signaling contributes to contextual memory with respect to addictive drug use. In addition, it has been reported that orexin signaling is central for “normal” memory functions, as especially investigated with respect to avoidance and spatial memory (see, for instance Jaeger et al., 2002; Telegdy and Adamik, 2002; Aitta-Aho et al., 2016; Akbari et al., 2007; Liao et al., 2024). Thus, increased orexinergic activity during wakefulness plays a role in memory acquisition whereas decreased orexinergic activity during sleep allows for memory consolidation (IV.C.). The memory functions promoted by orexins include social memory (Yang et al., 2013), and social interaction may be promoted by orexins (Abbas et al., 2015; Faesel et al., 2021; Dawson et al., 2023). This links to the psychiatric conditions, especially anxiety and depression, and the actual functions are in many cases difficult to distinguish in animal models (IV.D.).
L. Pituitary Regulation by Orexins
Isolated pituitary cells and cell lines express orexin receptor mRNA and respond to orexins; both increased and decreased release of hormones has been observed (reviewed in Kukkonen, 2013). However, a central question is how the orexins could reach pituitary endocrine cells. As pointed out in Kukkonen, 2013 potential physiological orexin actions could take place via regulation of the release of the known hypothalamic releasing factors and factors alike, but these do not explain the reported direct effects of exogenous orexins on the pituitary gland. There is no evidence based on extensive phase 3 clinical trials or postmarket assessments that DORAs affect pituitary functions.
M. Other Potential Orexin Functions in the Periphery
A few peripheral tissues/organs have been reported to express orexin receptor or PPO protein or mRNA or respond to ex vivo orexin stimulation. While the receptor protein expression data based on receptor antibodies are uncertain (IV.A.), other data may be more convincing. 3H-EMPA autoradiography suggested the presence of very low levels of OX2 in the rat adrenal cortex (Mitsukawa and Kimura, 2022), whereas OX2 mRNA levels and ex vivo functional assays suggested significant receptor levels (Jöhren et al., 2001; Karteris et al., 2005). Another study reported expression of OX2 mRNA and potential functional responses in the rat testis (Karteris et al., 2004), where, however, no 3H-EMPA binding was detected (Mitsukawa and Kimura, 2022). It should be noted that mismatches between mRNA and protein levels are not uncommon.
OX1 receptors were found in rat osteoblast precursor cells (Ziolkowska et al., 2007). PPO-overexpressing mice and OX1-KO mice were shown to have a higher bone mass, whereas PPO-KO mice and OX2-KO mice appeared to have lower bone mass (Wei et al., 2014). OX1 was suggested to promote the adipocyte differentiation instead of osteoblast differentiation by suppressing local ghrelin production. The local PPO production in bone was proposed to come from mesenchymal stem cells, osteoblasts and bone marrow adipocytes, since these expressed PPO mRNA. The OX2-dependent effect was instead suggested to require central OX2 signaling-mediated suppression of leptin expression. However, the causality of ghrelin and leptin in these processes was not verified (Wei et al., 2014). In contrast, OX1 receptor activation was reported to increase osteoblastic differentiation of the mouse cell line MC3T3-E1 via p38 MAPK activity and hypoxia-induced factor-1α induction and to inhibit hypoxia-induced reduction in osteogenic factors (Han et al., 2018; Gu et al., 2022). Subcutaneous orexin-A injections reversed hypoxia-induced bone mass loss in mice (Gu et al., 2022). Thus, there are only a few studies of the role of orexins in bone remodeling and the studies are in apparent contradiction as concerns the role of OX1.
Placentally produced orexins have been suggested to contribute to prenatal development of BAT in mice, though these data have been questioned (IV.B.). An SNP upstream of human HCRTR2 gene conferred a poorer prognosis in systolic heart dysfunction and after myocardial infarction (Perez et al., 2015; Wohlfahrt et al., 2023) (III.). While the reported lower protein levels for this polymorphism are not reliable (see above), OX2-KO mice were more prone to diastolic dysfunction. Wildtype mice were also partially protected by orexin-A infusion in the angiotensin II and isoproterenol chemical stress model of heart failure (Perez et al., 2015). Rat cardiomyocytes were reported to express OX2 receptor mRNA and isolated cells contracted upon orexin-B or Ala11, d-Leu15-orexin-B challenge (Patel et al., 2018). Orexin-B and/or Ala11, d-Leu15-orexin-B gave partial protection against the ischemic injury in the isolated rat heart and in intact rats and mice (Patel et al., 2018; Jiao et al., 2024). Even human heart muscle samples were shown to respond to orexin-B. It was recently suggested that the effect of exogenous orexin stimulation (here Ala11-d-Leu15-orexin-B) is inhibition of sensory neurons which further inhibit excessive sympathetic activation (Jiao et al., 2024). However, none of the studies has identified any endogenous orexin source. PPO-KO mice have higher numbers of circulating monocytes and neutrophiles (McAlpine et al., 2019); the authors suggest that the link is via hypothalamically produced orexins that are released in the blood stream where they act upon bone marrow to suppress production of colony-stimulating factor 1.
Any putative peripheral orexin signaling has the problem of PPO production. Are orexins coming from a peripheral source and used a) locally or b) systemically via circulation, or would central orexin neurites c) target peripheral tissues directly or d) release orexins in the circulation? There is little evidence for c except for the pineal gland (IV.C.), and thus this possibility might be excluded. As concerns circulating orexin (alternatives b and d), the release of orexins would need to be quite high. Using radio- and enzyme immunoassay kits, highly variable plasma levels of orexin-A have been measured in both human and rodents. For instance, a recent very high impact paper reports about 20 pg/ml in mice (McAlpine et al., 2019), which corresponds to 5.6 pmol/l and total of 6.2 fmol (the plasma volume = 25 g × 80 μl blood/g × 0.55 = 1.1 ml). The reported numbers of orexin neurons in the mouse brain are quite variable, e.g., ranging between 440 and 2300 (Obukuro et al., 2010; Black et al., 2018; Gonzalez and Prehn, 2018; Berteotti et al., 2021). Assuming there were 540 neurons (the median of these studies), every neuron would have to produce 0.011 fmol of PPO to reach this a plasma concentration. 125I-orexin-A and 125I-orexin-B have half-lives of 30.7 minutes and 4.4 minutes, respectively, in the mouse blood (Kastin and Akerstrom, 1999). Keeping a steady-state concentration of 5.6 pmol/l of orexin-A would thus require an additional constant production, processing, transport and release of 0.14 fmol/min orexin-A, which is equal to 155,000 molecules/min/neuron. In some human studies plasma levels below 0.3 pmol/l are reported while other studies report levels as high as 700 pmol/l (summarized in Mäkelä et al., 2018). The metabolism of orexin-A in the human plasma has not been investigated but its half-life in rats has been suggested to be similar to mice (Ehrström et al., 2004) and thus we may assume it to be similar even in humans. The number of orexin neurons in the human brain can be approximated with 80 000 (Kukkonen et al., 2002). Thus the corresponding values for 5.6 pmol/l in human plasma would be 0.19 fmol/neuron + 350 fmol/min/neuron (2 600 000 molecules/min/neuron). We can compare this to human pancreatic β-cells. The reports are not fully consistent but, for instance, the normal fasting plasma insulin level in the clinical tests (Helsinki University Hospital Laboratory; Helsinki, Finland) is 12–120 pmol/l. Assuming a half-life of 5 minutes, this would correspond to 2800–28,000 molecules/min/cell (in all 109 β-cells), which is similar to published studies (Polonsky et al., 1988; Ramchandani et al., 2010; Knopp et al., 2019). It, thus, appears that either the central orexin neurons or the peripheral tissues (PPO expression has been reported in gastrointestinal tract, adrenal gland, epididymis, testis, pancreas, kidney and placenta; Kukkonen, 2013) would have to be quite efficient (neuro)endocrine organs, which seems unlikely. Surprisingly, more than hundredfold higher circulating orexin-A levels than 5.6 pmol/l have been reported (e.g., 790 pmol/l; Goldstein et al., 2021), making the comparison even more absurd at the current level of understanding. The most likely explanation lies in the failure of the antibody-based methods in orexin detection (VI.), but even other factors need to be considered, for instance the much higher number of orexin neurons in relation to the blood volume in rodents (which have been the subject for most studies), potentially different pharmacokinetics of orexin-A in different species and maybe pulsatile secretion, but still there are central questions to be answered before the alternatives b and c would make any sense. Thus, a paracrine signaling (alternative a) would be most logical, though there is a mismatch of the receptor and peptide expression (above).
The final significant issue of peripheral orexin receptors concerns their expression in cancer cells (II.D.), which currently rather has bearing on pharmacology than physiology.
In summary, there are potentially very interesting findings of the peripheral role of orexins and orexin receptors. However, the findings are isolated and, as viewed in the light of our current understanding, both inexplainable and contradictory. There is also evidence for a mismatch between orexin mRNA expression and functional responses to exogenous orexins on the one hand and peripheral PPO peptide or mRNA expression on the other (Kukkonen, 2013); in most cases, there is no clear source for orexin peptides and thus the receptor stimulation is uncertain. Thus, alternative explanations for the orexin receptor expression should be considered, such as roles for orexins under certain conditions (induced expression?), other ligands for the receptors, constitutively active receptors, other roles for the receptors (and the peptides), or an altogether superfluous role of the peripheral orexin receptors (could these be evolutionary relics?). Clearly, more research is required. One source of information will be the potential phase 4 trials and other postmarketing data on DORAs.
V. Orexin Receptor Ligands and Drug Discovery
A. Orexin Receptor Agonist Peptides
The original study that reported both human orexin receptors and native peptides (Sakurai et al., 1998) demonstrated that both OX1 and OX2 were activated by orexin-A and -B when recombinantly expressed in CHO-K1 cells. While the peptides had equal affinity and potency for OX2, orexin-A showed an approximately 100-fold higher potency for OX1 compared with orexin-B, i.e., orexin-B was deemed OX2-selective whereas orexin-A did not distinguish between OX2 and OX1 receptors. However, later studies have demonstrated that the ratio of the peptide potencies on OX2 depends on the response measured, pointing at biased signaling (II.F.).
The C-termini of orexin-A and -B are highly conserved (Fig. 1), and studies involving peptide truncation and amino acid exchange have confirmed its central role in orexin receptor binding and activation (Darker et al., 2001; Okumura et al., 2001; Ammoun et al., 2003; Lang et al., 2004; Lang et al., 2006) (see below). In 2021, a study presented an orexin-B-bound human OX2 cryo-EM structure, revealing for the first time the structure of a potentially active orexin receptor state (Hong et al., 2021). As expected, the C-terminal region of orexin-B bound deep in the binding pocket, occupying the earlier described small molecule binding site (see III.A.), in a complementary shape. The structure of the peptide was extended, meaning that the α-helical solution structure (see below) of orexin-B's C-terminus is not preserved in the binding mode. Because of the low resolution of the structure on the extracellular site, the N-terminal part of orexin-B was excluded from the final structure.
Successive truncation of the orexin peptides from the N-terminal end is initially well tolerated, but the potency starts to gradually decrease when reaching beyond the first 12 or so amino acids (Darker et al., 2001; Okumura et al., 2001; Ammoun et al., 2003; Lang et al., 2004; Lang et al., 2006). The shortest peptides acting in the nanomolar range have been described to be composed of the 15–19 C-terminal amino acids (Darker et al., 2001; Ammoun et al., 2003; Lang et al., 2004; German et al., 2013), while an orexin-A variant as short as 7 amino acids has been demonstrated to have moderate activity on OX2 at 10 μM concentration (Ammoun et al., 2003); the potency is naturally also dependent on the receptor expression levels and the cellular background. By contrast, C-terminal manipulation of the peptides is not tolerated (Darker et al., 2001; Lang et al., 2004).
In most studies, reduction of the disulfide bridges or substitution of the cysteines with alanines leads to 10-fold or less decrease in the potency for OX1 and OX2 (Okumura et al., 2001; Lang et al., 2004; unpublished studies by Kukkonen, Rinne et al.). However, potential spontaneous (re)formation of the disulfide bridges has not been investigated in any study. Orexin-A modifications are generally better tolerated by OX2 than OX1 (Darker et al., 2001; Okumura et al., 2001; Holmqvist et al., 2002; Ammoun et al., 2003; Zhu et al., 2003; Lang et al., 2004; Lang et al., 2006; Johansson et al., 2007; Putula et al., 2011). Substitution with alanine, d-amino acid, etc. in orexin-A and -B has also offered little conclusive information (Darker et al., 2001; Ammoun et al., 2003; Lang et al., 2004; Lang et al., 2006). The putative OX2-selective agonist peptide Ala11, d-Leu15-orexin-B may represent an exception (Asahi et al., 2003), though its actual selectivity for OX2 is still debated and may be signaling pathway-dependent (Asahi et al., 2003; Putula et al., 2011; Yamamoto et al., 2022). Nevertheless, Ala11, d-Leu15-orexin-B is frequently used to support OX2-mediated effects. The use of agonists of limited selectivity is generally not recommended in pharmacological studies as there is always potential for biased signaling, hence the use of combinations of an agonist and at least two antagonists (one 1-SORA and one 2-SORA) should be the norm (Kukkonen, 2019).
NMR structures of orexin-A and orexin-B both in aqueous solution (Lee et al., 1999; Kim et al., 2004; Takai et al., 2006) and bound to sodium dodecyl sulfate (SDS) micelles have been published (Miskolzie and Kotovych, 2003; Miskolzie et al., 2003). The set of NMR structures of orexin-A show bent or straight conformations, while orexin-B always adopts a bent conformation. In solution, the orexin-B structure is composed of two α-helical regions (L7–G19 and A23–M28; 1CQ0), while orexin-A is composed of three (C6–G9, L16–A23 and N25–T32; 1WSO) or two α-helical regions (C14–H21 and N25–T32; 1R02). In micelles, orexin-A is composed of the α-helices D5–Q9 and L16–G22 and orexin-B of α-helices L7–S18 and A22–L26. While the secondary structure of the C-terminus of orexin-A is unclear (the last 11 residues), the end of the C-terminus in orexin-B (T27 and M28) is clearly nonhelical (Miskolzie and Kotovych, 2003; Miskolzie et al., 2003). The micelle-bound structures, with unstructured C-termini, were assumed to mimic the membrane-bound structures of the peptides; thus, the conformational freedom was suggested to potentially have a role in receptor binding via the membrane interface (Miskolzie and Kotovych, 2003; Miskolzie et al., 2003) (see below). In the NMR structures, the α-helical regions are connected by a kink or a linker. It was hypothesized that the truncation of peptides destabilizes their α-helical structures, and thus leads to a decrease in the binding affinity (Karhu et al., 2018). Chemically stabilized ("stapled") α-helical parts of C-terminal orexin-A fragments were thus examined for orexin receptor activation, but all stapled peptides were found to be less potent than the nonstapled peptide of the same length, likely due to a reduced binding affinity (Karhu et al., 2018). The reason for the lower potency remained unclear at the time, but it was speculated that the staple itself or a too rigid peptide conformation hindered the receptor binding or activation. Later the 3D-structure of orexin-B-bound OX2 revealed that the C-terminal region of orexin-B does not retain α-helical conformation in the binding mode (Hong et al., 2021), offering a likely explanation to the behavior of the stapled peptides (see III.A.).
Orexins are amphipathic peptides, possessing hydrophilicity and hydrophobicity on the opposite sides of the α-helix 1 (Lee et al., 1999; Miskolzie and Kotovych, 2003; Miskolzie et al., 2003; Kim et al., 2004; Takai et al., 2006). A recent study investigating the binding of orexin-A to synthetic lipid bilayer vesicles (Ball et al., 2023) demonstrated that the α-helix content of orexin-A increases when bound to the membrane and that the C-terminal helix aligns in parallel to the surface of the membrane. The significance of this observation is unclear, but the authors speculated that this would “prepare” the peptide for receptor binding (Ball et al., 2023), as was also proposed for the NMR structures of the micelle-bound orexin peptides (see above). The corresponding L20A and L15A mutations in N-terminally truncated orexin-A and orexin-B variants, respectively, reduce the potencies by more than 10-fold (Darker et al., 2001; Ammoun et al., 2003; Lang et al., 2004; German et al., 2013) which has been suggested to be caused by reduced membrane lipid binding of the lipophilic helix side (Miskolzie and Kotovych, 2003; Miskolzie et al., 2003), though it is unclear why the –CH3 sidechain would be so much less lipophilic than –CH(CH3)2. However, the C-terminus of orexin-B loses its α-helical conformation when bound to the reported active receptor conformation (Hong et al., 2021) and the same is seen with micelle-bound orexin-A (see III.A.). In essence, structural information gained from studies with unbound peptides is of limited utility, since the same peptide can likely adopt multiple conformations depending on the target receptor, the complexed accessory proteins involved in the signaling apparatus, etc.
An apparent requirement of extracellular Ca2+ for the binding of 125I-orexin-A to OX1 has been reported, which suggests a similar effect on native orexin-A (Putula et al., 2014). It is unclear whether the total binding capacity or the affinity is affected (Lund et al., 2000; Johansson et al., 2007; Putula et al., 2014). It also remains unclear whether the property is related to the peptide or the receptor, but Ca2+ did not have any effect on the circular dichroism (CD) spectrum of orexin-A or its stability at different temperatures (Putula et al., 2014). If Ca2+ affects the receptor, it is difficult to determine whether the effect is intracellular or extracellular, since manipulation of the extracellular Ca2+ levels affects the intracellular levels, especially close to the plasma membrane. The fast Ca2+ buffer BAPTA (10–20 mM) has been introduced inside the cells in patch-clamp studies of orexin receptor signaling (Liu et al., 2002; Burdakov et al., 2003; Acuna-Goycolea and van den Pol, 2009; Kargar et al., 2015; Ishibashi et al., 2016; Usui et al., 2019; Jin et al., 2021). Its impact ranges from no effect to almost total inhibition of the response. All studies have used high concentrations of orexins (0.1–1 μM); thus, it is not possible to judge whether there is any effect of BAPTA on the orexin concentration–response relationship in those studies in which BAPTA was not effective.
B. Small Molecule Orexin Receptor Ligands
Most drug discovery efforts in both industry and academia have focused on orexin receptor antagonists. The original indications for therapeutic intervention using orexin receptor antagonists were obesity and insomnia, illnesses that affect a large proportion of the total population and are in urgent need of novel therapies. Conversely, the obvious indication for orexin receptor agonists, narcolepsy with cataplexy, is rare – although many of the recent new drug approvals are for rare or very rare diseases. Small molecule nonpeptide agonists are generally much more difficult to discover/develop than antagonists, especially in the absence of a good pharmacophore model, as is often the case with neuropeptides and their receptors (Hoyer and Bartfai, 2012). Thus, there is a great number of orexin receptor antagonist structures published, with many antagonists undergoing clinical studies and three DORAs on the market. Unfortunately, only a few antagonist structures have been thoroughly characterized. Much fewer agonist structures have been published, and two are in clinical trials. These advances are summarized below. Further details on orexin receptor drug discovery and development can be found in recent reviews (see, e.g., Perrey and Zhang, 2020; Jacobson et al., 2022; Bonifazi et al., 2023).
1. Small Molecule Orexin Receptor Antagonists
Drug discovery for the orexin system has been mostly propelled by pharmaceutical and biotechnological industry. Orexin receptor antagonists are broadly divided into subtype-selective orexin receptor antagonists, i.e., 1-SORAs and 2-SORAs, and nonselective antagonists, DORAs; the limits for the selectivity ratio for each class are not defined, but most DORAs display some level of OX2-selectivity. However, different orexin receptor antagonists have very different association and dissociation kinetics (Mould et al., 2014), and this affects the determined apparent affinity and inhibition constants. For instance, almorexant was, after a shorter incubation, apparently nonselective at mouse, rat and human OX1/OX2 receptor-driven calcium elevation, yet after 4 hours of incubation, it became up to 20-fold OX2-selective at rat and human receptors due to its slow dissociation (Mang et al., 2012); similarly, other DORAs may show some OX2-selectivity at equilibrium depending on the species investigated (Callander et al., 2013).
There are number of expert reviews on drug development that have been published as the field continues to evolve (e.g., Boss et al., 2009; Coleman and Renger, 2010; Lebold et al., 2013; Boss, 2014; Roecker et al., 2016).
a. Orexin receptor antagonists in suppression of appetite and feeding
The original indication for the company that was apparently fastest on the track, GlaxoSmithKline (at the time SmithKline Beecham), who funded the deorphanization efforts of Masashi Yanagisawa’s group, was obesity. For various reasons, e.g., findings in Sakurai et al., 1998, the receptor of interest became OX1. However, none of the 1-SORAs of the SB-series (Fig. 4A1) reached the clinic – probably due to both chemical and physiological issues – but they [and especially the first ever orexin receptor antagonist published, SB-334867 (Haynes et al., 2000; Smart et al., 2001)] have been important tools in numerous physiological research projects in helping to determine the orexin receptor subtype involved: As of May 18, 2024, there were 556 hits for “(SB-334867 or SB334867 or SB 334867 or SB-334,867)” in PubMed. However, SB-334867 is not very selective for OX1 (less than 100-fold) and may be unstable (reviewed in Kukkonen, 2019). Thus, some original receptor subtype determinations based only on SB-334867 may not be valid, especially in vivo. As a rule, one should never rely on a single molecule to determine the receptor subtype involvement (V.A.).
ACT-539313 (Williams et al., 2024) (Fig. 4A1), another 1-SORA, was tested in binge eating disorder, but failed to meet primary endpoints in a phase 2 clinical trial (NCT04753164); pursuit of the indication by Idorsia has reportedly stopped for this compound (IV.D.). There is little evidence from any of the large phase 3 clinical trials that have been conducted with the DORAs suvorexant, lemborexant or daridorexant (Fig. 4A3) in adult patients – of which a significant proportion have been elderly and/or overweight or obese, particularly in the female cohorts – to suggest that chronic dosing of DORAs affects food intake, weight or BMI.
Thus, it seems appetite regulation is not a major physiological role of orexins as discussed under IV.B., and, consequently, appetite suppression has been considered unlikely as a major therapeutic aim for orexin receptor antagonists, be it 1-SORAs or DORAs, as also evidenced by antagonist data in humans. Nevertheless, there are currently ongoing investigations with 1-SORAs, such as JNJ-61393215 (Salvadore et al., 2020) (Fig. 4A1), CVN45502 (even in clinical trials for obesity; Schneeberger et al., 2022) and CVN766 (Glen et al., 2024), which may give further information on the role of OX1 in appetite, feeding and metabolism in humans (see V.B.1.c.).
b. Orexin receptor antagonists in disorders of sleep and wakefulness
Instead of appetite and feeding, Actelion had insomnia as the primary indication. Actelion reported phase 2 studies with the DORA almorexant (Fig. 4A3) in insomnia in 2007 (reviewed in Hoyer and Jacobson, 2013). GlaxoSmithKline and Actelion formed a partnership to develop almorexant, but its clinical development was stopped in phase 3 in 2011. Incidentally, GlaxoSmithKline had its own DORA, SB-649868 (Fig. 4A3), in development as a hypnotic, but it was stopped after completing phase 2 studies for insomnia (Hoyer and Jacobson, 2013). Several pharmaceutical companies are actively developing 2-SORAs (Fig. 4A2) – based on the apparently more crucial role of OX2 in the regulation of sleep and wakefulness (IV.C.) – or DORAs (Fig. 4A3). The feasibility of the DORA approach in insomnia was finally demonstrated by the FDA approval of suvorexant (and subsequently lemborexant and daridorexant) (Fig. 4A3). DORAs enhance primarily REM sleep, with little or no effect on NREM sleep (Clark et al., 2020).
No 2-SORA is yet approved, but seltorexant (JNJ-42847922; Fig. 4A2) is in advanced clinical trials for insomnia and MDD with insomnia (De Boer et al., 2018; Recourt et al., 2019). Merck had a 2-SORA program, including MK-1064 (Roecker et al., 2014; Gotter et al., 2016), that has been suspended (Fig. 4A2). An interesting finding in clinical trials is that seltorexant and MK-1064 (Fig. 4A2) also predominantly increase REM sleep, like DORAs, at least in healthy volunteers (Gotter et al., 2016; De Boer et al., 2018), although another 2-SORA, JNJ-48816274 (Letavic et al., 2015) (Fig. 4A2), increases both REM and NREM sleep in the circadian phase advance model of insomnia (Revell et al., 2022). In rodents, in contrast, 2-SORAs enhance balanced/physiological sleep, i.e., both NREM and REM are increased in similar proportions to that of normal sleep, whereas DORAs dominantly enhance REM sleep (Betschart et al., 2013; Hoyer et al., 2013; Keenan et al., 2022). For example, IPSU (Betschart et al., 2013) (Fig. 4A2), a moderately OX2-selective antagonist, increased sleep in a physiologically balanced manner during the active phase, whereas suvorexant (Fig. 4A3) preferentially stimulated REM sleep. In addition, during the inactive phase, IPSU did not affect sleep (when mice were already asleep), whereas suvorexant administered at this time still increased REM sleep overall and as a proportion of total sleep (Betschart et al., 2013). Some studies even suggest that OX1 antagonism actually attenuates the NREM sleep-promoting effect of OX2 antagonism (Dugovic et al., 2009; Betschart et al., 2013; Dugovic et al., 2014). However, there are currently insufficient published clinical data on 2-SORAs to compare DORAs and 2-SORAs from the perspective of efficacy on human sleep promotion, maintenance or architecture (Clark et al., 2020). Please also note that the pharmacological distinction between DORAs and 2-SORAs may not be that clearcut (see above).
Whatever the clinical outcome, it is evident that hypnotics inducing a physiological sleep profile and architecture – which is important for the cognitive abilities (especially learning and memory) and long term brain health (Boyce et al., 2016; Poe, 2017; Sara, 2017; Spruyt, 2024) – are likely to be more desirable for the treatment of insomnia than current hypnotics (i.e., benzodiazepines, Z-drugs and different types of central H1 histamine receptor antagonizing drugs such as antidepressants, antipsychotics or first generation antihistamines), which suppress REM sleep (Hoyer and Jacobson, 2013; Hoyer et al., 2013; Hoyer et al., 2020; Scammell et al., 2019). However, in some disorders, such as MDD, REM sleep proportion may be inappropriately increased, and thus further promotion of REM sleep at the cost of NREM may not be ideal. Adequate amount of NREM deep sleep may be required for proper function of the glymphatic system (IV.C. and F.). There also are other important parameters of sleep architecture to consider in addition to the gross amounts of time spent in various sleep stages.
Safety and toxicological investigations of DORAs and 2-SORA have primarily demonstrated somnolence as a common side-effect (Na et al., 2024), as is expected for hypnotics. Cataplexy, sleep paralysis or muscle weakness has long been considered a potential safety risk of DORAs (Tafti, 2007) and is supported to some degree by preclinical studies (Black et al., 2013; Mahoney et al., 2020). However, these and other effects were very rare in the extensive phase 3 trials that lasted up to 14 months and collectively included thousands of patients and healthy controls (Hoyer et al., 2020), but DORAs are nevertheless contraindicated in NT1. The FDA also recommends that DORAs are administered shortly (30 minutes) before bedtime as a safety measure. The 2-SORA MK-1064 did not induce cataplexy in the food-induced cataplexy test in dogs, even at very high doses (Gotter et al., 2016). The DORAs suvorexant and daridorexant (Fig. 4A3) (lemborexant not investigated) also appear safe in chronic obstructive pulmonary disease and obstructive sleep apnea (Sun et al., 2015; Sun et al., 2016; Boof et al., 2021a; Boof et al., 2021b). There is evidence that suvorexant, when compared with benzodiazepines, does not increase hip fractures in hospitalized elderly (65–84 years old) patients (Saito et al., 2023) and it is generally agreed that DORAs are safe and efficacious in elderly patients (Herring et al., 2020; Fietze et al., 2022), and compare positively with the more traditional Z-drugs, e.g., zolpidem (Hoyer et al., 2020; De Crescenzo et al., 2022; Jacobson et al., 2022). It will be interesting to see which advantages if any [increased slow wave sleep (SWS), antidepressant effects?], the 2-SORAs will bring to the clinic.
Overall, with regard to sleep–wakefulness symptoms of primary insomnia or other indications, DORAs have clearly added to the diversity of hypnotic tools, which now provides clinicians with an unprecedented ability to manipulate sleep–wake architecture in a precision medicine manner (Jacobson et al., 2022), not only in idiopathic insomnia but also in other sleep disorders; for instance, lemborexant improved irregular sleep-wake rhythm disorder (ISWRD), which is a common sleep disorder in, e.g., AD dementia patients, in a phase 2b study in AD dementia (Moline et al., 2021). The treatment of insomnias and potentially other sleep–wake disorders have been recently reviewed (Hoyer et al., 2020; Jacobson et al., 2022; Muehlan et al., 2023).
c. Orexin receptor antagonists in addiction
As discussed in section IV.D., numerous studies have reported a role for orexin signaling in behaviors related to alcohol, opioid, and psychostimulant seeking in animal models, especially as it pertains to cue or contextual driven relapse (Fragale et al., 2021b). As such, several clinical trials have been initiated in various populations of substance use disorders to determine if administration of DORAs can reduce drug use. Suvorexant was recently trialled for normalizing sleep disturbances during abstinence in patients with opioid use disorder. Suvorexant treatment was associated with clinically relevant improvement in sleep duration as well as a reduction in posttaper opioid withdrawal symptom severity, and reduced opioid craving (Huhn et al., 2022). Ongoing clinical trials are also addressing the efficacy of suvorexant in sleep efficiency and opioid abstinence in outpatient groups (NCT04262193), as an adjuvant to buprenorphine treatment in fentanyl users (NCT05145764), and the effects of sleep and stress in early abstinence from opioid use (NCT04287062). Furthermore, suvorexant is under investigation for use in alcohol use disorder (NCT05656534, NCT04229095), and nicotine addiction (NCT04234997). The efficacy of lemborexant on sleep when used in combination with opioids, including opioid agonist therapies such as buprenorphine, is being assessed with secondary measures of opioid craving and withdrawal, although results have not yet been reported (NCT04818086). Similarly, the combination of naltrexone and lemborexant on alcohol craving and insomnia is under investigation (NCT05458609). 1-SORAs may offer advantages over the DORAs for relapse prevention as presumably they have reduced sedative properties and selectively target OX1 receptors, which are primarily implicated in drug seeking. However, to date, 1-SORAs may be too early in clinical development to have registered clinical trials for addiction-related indications.
d. Other indications for, or combinatorial use of, orexin receptor antagonist
Cerevance’s CVN45502 (Schneeberger et al., 2022) and CVN766 (Glen et al., 2024) (Fig. 4A1) are claimed to have > 1000-fold OX1-selectivity. Psychosis is classically associated with elevated dopamine signaling, which can be reduced by all categories of orexin receptor antagonist while only a 1-SORA and two DORAs, but not a 2-SORA, were reported to reduce induced “psychotic behavior” in rats (Elam et al., 2021; Perez and Lodge, 2021). CVN766 is being developed for schizophrenia (phase 1/2; phase 1 study data at https://www.cerevance.com/media/oral-presentation-of-phase-1-data-on-cvn766-at-american-society-of-clinical-psychopharmacology-ascp), and possibly for anxiety, binge eating, obesity, substance use disorders, and Prader–Willi Syndrome as suggested on the Cerevance website. Many atypical antipsychotic medications produce increased body weight and metabolic disorder; co-therapy with orexin receptor antagonists may reduce this, in addition to the possible antipsychotic effects through modulation of e.g., the dopaminergic system.
A number of other OX1 antagonists are effective in rodent and human panic anxiety models, and thus some, e.g., JNJ-61393215, are in development for anxiety disorders. Please see IV.E. for details. DORAs (almorexant and TCS1102), 1-SORAs (SB-334867 and SB-674042) and 2-SORAs (LSN2424100, MK-1064 and TCS-OX2-29; note that the selectivity of TCS-OX2-29 has not been demonstrated in any publication) (Haynes et al., 2000; Porter et al., 2001; Hirose et al., 2003; Langmead et al., 2004; Brisbare-Roch et al., 2007; Bergman et al., 2008; Fitch et al., 2014; Roecker et al., 2014) (Fig. 4A) have been assessed in rodent depression models for their potential to treat MDD. Antidepressant-like effects have been seen with DORAs and to some extent with 2-SORAs, but not with 1-SORAs (which are more efficacious in anxiety models) (reviewed in Fagan et al., 2023). However, not all reports are consistent: for instance, in some cases, DORAs or 1-SORAs increased depressive-like behavior, e.g., in the forced swim test, while in other cases they had an antidepressant action (reviewed in Fagan et al., 2023). The different effects may depend on the experimental conditions; one needs to take into account confounding factors such as the general increase in wakefulness and motor activity induced by orexins (Hagan et al., 1999; Nakamura et al., 2000) and the opposite by orexin receptor antagonists, as well as the effects of both depression and orexin receptor antagonists on sleep architecture, all of which may distort the results. The DORA filorexant (Fig. 4A3) (Connor et al., 2017) and the 2-SORA seltorexant (Fig. 4A2) (Brooks et al., 2019; Recourt et al., 2019; Savitz et al., 2021) were assessed in placebo-controlled trials in MDD patients, either as monotherapies or as augmentation to serotonin-selective (SSRI) or serotonin–noradrenaline reuptake inhibitors (SNRI) with up to 28 days of treatment. Only the phase 2b seltorexant trial showed superiority to placebo, and only at 20 mg, whereas 10 and 40 mg did not reach statistical significance. In other studies significant effects were observed in some depression scales, but not in others. Nevertheless, several seltorexant phase 2/3 trials are in progress in MDD and associated insomnia. Some additional effects or orexin receptor antagonists are described in chapter IV.
e. Other considerations
Little attention has been paid to the potential of inverse agonism by orexin receptor antagonists. Inverse agonism of a ligand can only be demonstrated when the receptor displays constitutive activity, and it is assumed that most GPCR “antagonists” are actually inverse agonists (Kenakin, 2011). To our knowledge, there are no reports of inverse agonism for orexin receptor ligands or constitutive activity of orexin receptors, and no constitutively active orexin receptor mutants have been constructed. Inverse or partial agonism could explain subtle differences that have been observed with orexin receptor antagonists in vivo, for instance in the situations where no endogenous orexin is present, yet the “antagonist” produces an effect in vivo; for instance, the DORA lemborexant (but not another DORA, almorexant) (Fig. 4A3) unexpectedly appears to reduce the number (but not duration) of cataplexy bouts in PPO-KO mice (Mahoney et al., 2020), although this finding requires replication since DORAs do not affect the sleep–wake architecture of PPO-KO mice (Mang et al., 2012; Mahoney et al., 2020). The question of allosteric modulation by orexin receptor antagonists (Kukkonen, 2020) has not yet been properly investigated either.
2. Small Molecule Orexin Receptor Agonists
As described under V.A., the orexin peptides were modified, for different purposes, for instance to achieve orexin receptor subtype-selective agonists or a pharmacophore model for small molecule agonist discovery. For the latter purpose, the results amount to nil. During the last 10 or so years studies or patents reporting weak agonists or allosteric enhancers slowly started emerging (Yanagisawa, 2010; Turku et al., 2016; Leino et al., 2018; Turku et al., 2019). The first significantly active series, with the leads Nag 26 and YNT-185 (Fig. 4B2), was published in 2014 as a result of a high throughput screen of approximately 250 000 compounds and modification of the hit compound (Nagahara et al., 2015); this is a highly remarkable achievement for an academic laboratory. YNT-185 was shown to be active against narcolepsy symptoms in the PPO-KO mouse model (Irukayama-Tomobe et al., 2017), providing a proof of concept, but it never progressed to clinical studies. Notably, the orexin peptides had been tested in rodents even earlier and found to produce the same effect (Mieda et al., 2004; Liu et al., 2011b) (II.G.). Takeda has produced a series of potent and highly OX2-selective agonists (Fujimoto et al., 2017), of which danavorexton/TAK-925 (Yukitake et al., 2019) (Fig. 4B2) was the lead for NT1, and the first compound to be tested in humans. That indication has been abandoned (due to insignificant oral bioavailability), but the compound is still in development for postanesthesia recovery (Suzuki et al., 2024). Development of another closely related lead molecule, firazorexton/TAK-994 (Dauvilliers et al., 2023; Ishikawa et al., 2023) (Fig. 4B2), was discontinued during phase 2/3 studies in October 2021 due to adverse hepatic events. Another OX2-selective agonist, TAK-861, is undergoing clinical development (see the ClinicalTrials identifiers NCT05816382 and NCT05687916 as well as the WHO International Clinical Trials Registry Platform IDs JPRN-jRCT2031230135, EUCTR2022-002966-34-FI and NL-OMON53459, etc.). After the recent phase 2b study results, Takeda has announced that it will proceed to phase 3 for NT1 but not NT2 (https://www.takeda.com/newsroom/newsreleases/2024/Takeda-Intends-to-Rapidly-Initiate-the-First-Global-Phase-3-Trials-of-TAK-861-an-Oral-Orexin-Agonist-in-Narcolepsy-Type%201-in-First-Half-of-Fiscal-Year-2024/). The structure of TAK-861 has not been published; this is according to the increasing trend of publishing the structure of the drug candidate when clinical studies have already started, and thus extending the patent life and ensuring exclusive access to the compound (see, e.g., Brisbare-Roch et al., 2007).
Merck has reported compound 1 (Fig. 4B2) (Hong et al., 2021), probably based on exploration of the chemical space around Nag 26 and YNT-185, in addition to another series (Sabnis, 2020). The chemical space around the original Nagahara hits (Nagahara et al., 2015) has been further explored (Zhang et al., 2021; Hino et al., 2022; Iio et al., 2022; Iio et al., 2023). Most of the reported agonists show some level of OX2 receptor selectivity, although not as marked as that of Nag 26 and YNT-185 as described in the original publications (Nagahara et al., 2015; Irukayama-Tomobe et al., 2017); however, a lower OX2 versus OX1-selectivity for Nag 26 than originally reported has also been demonstrated (Rinne et al., 2018). The most recent developments around the “Nagahara backbone” are the largely nonselective agonist RTOXA-43 (twofold OX2-selective as normalized to orexin-A) (Zhang et al., 2021) and the first OX1-selective agonists (R)-YNT-3708 (4.9-fold OX1-selective as normalized to orexin-A) (Iio et al., 2023) and TOL-27 (at least 100-fold OX1-selective as normalized to orexin-A) (Leino et al., 2023); TOL-27, although OX1-selective, has a very low potency (Fig. 4B2–3). (R)-YNT-3708 had antinociceptive activity in the mouse tail-flick test and was active in the conditioned place preference test in control mice, while these effects were absent in OX1-KO mice (Iio et al., 2023).
A recent patent reports a novel series of macrocyclic orexin receptor agonists (Armacost et al., 2021). For OX2, the compounds are equally efficacious as orexin-A in the inositol phosphate mobilization assay and some have a potency in subnanomolar range, but there is no comparison with the potency of the orexin peptides or between receptor subtypes (Armacost et al., 2021). The same comparison is lacking from another novel series of agonists (Wang et al., 2024) based on the low efficacy, toxic prototype agonist Yan 7874 (Yanagisawa, 2010; Turku et al., 2017).
The obvious indication for orexin receptor agonists is narcolepsy, with expansion to other disturbances of sleep and wakefulness, such as excessive daytime sleepiness in a range of disorders. In agreement with the current idea of OX2 being mainly involved with wakefulness, OX2-selective agonists YNT-185, danavorexton (TAK-925), firazorexton (TAK-994) and Ala11-d-Leu15-orexin-B increase wakefulness and reduce narcolepsy symptoms in rodents (Irukayama-Tomobe et al., 2017; Yukitake et al., 2019; Yamamoto et al., 2022; Ishikawa et al., 2023; Sun et al., 2023b). In human NT1 patients, danavorexton/TAK-925 and firazorexton/TAK-994 improve wakefulness as measured by e.g., increased sleep latency and maintenance of wakefulness (Evans et al., 2022; Dauvilliers et al., 2023), and danavorexton/TAK-925 additionally in NT2, idiopathic hypersomnia and excessive daytime sleepiness in obstructive sleep apnea (Evans et al., 2022; Bogan et al., 2023; Mignot et al., 2023). Instead of an orthosteric agonist, a positive allosteric OX2 receptor modulator might be suitable under conditions where the orexin tone is reduced but not totally absent. However, drug discovery for positive allosteric modulators of orexin receptor is relatively underdeveloped (or undisclosed), although potential chemical starting points have been reported for OX1 (Leino et al., 2018; Turku et al., 2019) and OX2 (Lee et al., 2010). Future research in this area is indicated.
Absolute selectivity is problematic for all GPCR agonists. One major issue is that agonists may show biased signaling and thus their potency and efficacy for different primary signals may differ, which may be further enhanced by cell-specific amplification mechanisms. In essence, a ligand may well behave as a full agonist at one pathway and as a partial agonist, antagonist or inverse agonist at another pathway (Kenakin, 2011). The other issue relates to the relative potency. Upon increased expression (in recombinant systems), the agonist potency easily increases as receptor reserve increases (Black and Leff, 1983; Kenakin, 1997). Thus, one cannot directly compare the responses of different receptor subtypes even in the same cellular background. True agonist affinity is difficult to measure reliably, and it is a relatively useless parameter for agonist action. Thus, in the case of orexin receptors, we suggest that the only way to compare agonist potencies is to normalize them to a reference compound; the potency of orexin-A has been used to that effect. However, no-one really knows whether the potency of orexin-A is equal for the two orexin receptor subtypes or, especially, for different responses. The final issue is how to relate responses measured in recombinant (and frequently receptor-overexpressing) systems, to those observed for native receptors.
C. Labeled and Other Types of Special Ligands
125I-orexin-A has been the “traditional” radioligand as it was the only one commercially available for years, and 125I-orexin-B is also available. Agonists are theoretically less suitable as radioligands, and orexins (and their labeled variants) have additional biophysical and chemical properties that make them even less optimal (Kukkonen, 2012). Some 3H-labeled antagonists have become available more recently; the current commercial selection contains at least 3H-SB-674042, 3H-almorexant, 3H-EMPA and 3H-TCS1102/-BBAC, which had originally been custom-labeled by pharmaceutical companies for internal purposes. Radiolabeled orexin receptor antagonists have only been characterized to a limited extent, with some exceptions (Langmead et al., 2004; Mould et al., 2014). 125I-orexin-A is also used in the standard radioimmunoassay for measurement of orexin peptides in the CSF for research purposes and in narcolepsy diagnostics. Radiolabeling (3H) of the agonists danavorexton/TAK-925 and firazorexton/TAK-994 has also been reported (Ishikawa et al., 2022; Ishikawa et al., 2023).
There have been a few attempts to generate orexin receptor-specific small molecule positron emission tomography (PET) tracers. Such tracers might be useful in detecting either orexin receptor levels (when able to effectively compete away endogenous orexins) or orexinergic neuronal activity (when competed away by endogenous orexin peptides). There may be additional advantages and disadvantages of antagonist and agonist probes (Colom et al., 2019). The currently published orexin receptor positron emission tomography probes, when taken up in the brain, show rather high nonspecific binding as identified by, for instance, labeling of areas known not to express significant amounts of orexin receptors. Thus, positron emission tomographic imaging cannot currently be reliably applied in orexin research.
Orexin peptides have been fluorescently custom-labeled (Darker et al., 2001; Evans et al., 2001; Karteris et al., 2001), and both fluorescent and biotin-labeled orexin-A and -B are also commercially available, but have not been widely used. A time-resolved fluorescence resonance energy transfer-based binding assay utilizing labeled orexin-A and orexin receptors is also commercially available.
Nonradioactive heavy labels (13C, 15N) have been used in custom-made internal orexin-A and -B standards for mass spectrometry-based detection assays (VI.) (Hirtz et al., 2016; Bårdsen et al., 2019; Lindström et al., 2021; Hopkins et al., 2021; Ball et al., 2023; Narita et al., 2023) and in NMR (III.A., V.A.). A caged version of orexin-B – that can be uncaged upon a UV or near-UV light pulse – has recently been developed (Duffet et al., 2022b). This can be used in discrete temporal or regional settings which may allow for increased probing of the functions of orexin-B in the modulation of synaptic transmission and other physiological processes. A photoactivatable/-inactivatable orexin-B6–28 version (photorexin) has also been reported; however, the achieved degree of photoisomerization does not appear to be able to fully turn the compound off (Prischich et al., 2024).
VI. Tools and Methods for Orexin Research
The current orexin receptor ligands, including those for specific purposes are presented under V. Unfortunately, not all interesting receptor ligands are commercially available. As concerns the investigations of the orexin receptor signaling – as any signaling studies – it is important to acknowledge that the pharmacological tools have their selectivity problems as discussed under II. While molecular biological tools may have potentially higher selectivity, they also have their own issues. The novel ways of measuring receptor-mediated activation of heterotrimeric G proteins (or other signal transducers) are very promising but not fully error-proof; for instance, different methods may give different results for (currently) unexplainable reasons. When it comes to proving the physiological role of a particular signaling pathway, there are fewer good methods, and interactions occur between the intracellular signal transduction pathways (II.A.1., II.B.).
The previous genetic tools (transgenic animals and viral vectors) were presented in Tables 2 and 3 in Kukkonen, 2013. Novel genetic tools that have emerged since then as well as some unintentionally omitted from these tables are presented in Table 2 of the current review. Two types of tools are worth discussing in particular. OxLight1 is a genetically encoded mutated OX2 receptor that fluoresces upon binding of orexin-A or -B, due to a conformational change in the receptor leading to enhancement of a circularly permutated green fluorescent protein module (Duffet et al., 2022a). The OxLight1 biosensor can detect orexin concentrations in the mid–high nanomolar range, which allows monitoring of orexin dynamics in animals engaged in behaviors that increase orexin release, as demonstrated (Duffet et al., 2022a). On the chemical side, a photocaged orexin-B has been developed (Duffet et al., 2022b) that will allow for spatiotemporally controlled activation of endogenous orexin receptors (see V.C.).
Transgenic cre driver lines have been pivotal in advancing neuroscience research. Some potential issues of these should, however, be recognized. Cre recombinase (or other genes) are often expressed as random knock-in under some reduced proximal promoter fragment which does not always faithfully replicate the full promoter activity and thus not the endogenous gene expression. It should, though, be noted that sometimes the exogenous protein expression is accomplished by targeted insertion within the original gene which preserves nearly all endogenous regulation. However, even this may constitute a problem if the promoter is activated in different places at different timepoints during development (see below). Transgenic mouse lines produced by mating a cre-expressing line with a floxed line may not consistently show 100% recombination. Proximal 3.2 kB Hcrt promoter–cre driver lines (or other genes under this promoter) (Matsuki et al., 2009; Inutsuka et al., 2014; Giardino et al., 2018) are not immune to these problems. In a mouse line with Hcrt promoter–cre and loxP, recombination takes place somatically in every generation. The Hcrt promoter seems to be active in mice during neurogenesis and neural migration (E11–E14) (Amiot et al., 2005). If, during that time, the promoter is activated by developmentally active regulatory factors and sequences – different from postnatal regulation – this may lead to targeting of floxed alleles according to a different schedule than intended and potentially recombination that appear ectopic in the postnatal perspective. This possibility has not been investigated for the orexin system, but there are anecdotal reports of loss of specificity of cre expression in orexin neurons over multiple generations of mouse breeding. The mechanisms are unclear, but may be related to epigenetic modifications of the Hcrt promotor (Dr. L. de Lecea, personal communication). For other cre systems, similar problems have even been reported (Müller-Komorowska et al., 2020; Xiao et al., 2021; Parrish et al., 2023), and other similar recombination systems like flp–frt are subject to the same issues. Taken together, it is important to repeatedly validate the selectivity and efficiency of expression for transgenes expressed using cre driver lines under the Hcrt promotor, and to keep these potential issues in mind when interpreting the data utilizing such recombination systems (Table 2). This may have far-reaching consequences for orexin studies as, for instance, optogenetic studies have used the proximal Hcrt promoter for cre expression. If epigenetic mechanisms or other factors compromise the specificity of the PPO promoter variant used, the tetracyclin-regulated promoter systems may be affected (Table 2). It would be interesting to investigate whether the transgene expression is different in the cre mouse lines in which the 3.2 kB proximal Hcrt promoter is used (Matsuki et al., 2009; Inutsuka et al., 2014) as compared to the ones utilizing the native Hcrt promoter (Giardino et al., 2018). Use of recombining (e.g., floxed) viruses together with enzyme (e.g., cre)-expressing animals is less prone to the problems associated with development since they can be introduced after the developmental period, but have the specific problem of transduction efficiency.
In the clinic, the only “research” tool used in connection with orexins is the measurement of orexin-A levels in CSF samples in narcolepsy diagnostics using radioimmunoassay. This method for orexin-A measurement is problematic because the antibodies used might also detect PPO, degradation products of orexin-A and potentially unrelated molecules (Sakai et al., 2019); thus, the measured “orexin-A” may instead reflect a mixture of unknown peptides and peptide fragments including those without activity at orexin receptors (Kukkonen, 2021). Lack of knowledge about the physiological degradation mechanisms of the orexin peptides and the identity of the fragments (or potential other peptides) further complicates the matter. The nonvalidated radioimmunoassay used for CSF orexin-A measurements may be sufficient for the (very crude) narcolepsy diagnostics, but any extension beyond this is not well-founded, and even this use is based on empirical rather than clear-cut mechanistic evidence. Sandwich enzyme immunoassay should be more specific for orexin-A than radioimmunoassay; the orexin-A levels were indeed found much lower in the single study where comparison with radioimmunoassay was performed, but the results in narcolepsy versus control were much more variable as well (Ono et al., 2018). The results have been very similar in the studies with mass spectrometric methods versus radioimmunoassay (Hirtz et al., 2016; Bårdsen et al., 2019; Lindström et al., 2021); thus the other methods are currently inferior to radioimmunoassay in narcolepsy diagnostics – maybe, paradoxically, because they are much more specific for orexin-A. The mass spectrometric methods will nevertheless – once detection of the orexin breakdown products has been developed as well – be very useful in the research of the significance of the orexin metabolism and, subsequently, in all cases where a disorder of orexin peptides or their breakdown is suspected (Kukkonen, 2021). Whether orexin receptors, for instance OxLight1, could be harnessed as an orexin bioassay, remains to be shown. These questions are set to become increasingly important with the rapidly advancing field of biosensors and neural probes (see, e.g., Li et al., 2022b).
VII. Summary, Conclusions and Future Perspectives
The discovery of the orexin/hypocretin system has provided multiple insights into the circuitry and molecular mechanisms underlying, for instance, arousal, reward and addiction, eating behaviors, fear and anxiety, and cognition in health and disease over the last 26 years since its discovery in 1998. The orexin system is most clearly involved in the regulation of sleep and wakefulness. The loss of orexin neurons – or peptide expression – in the human lateral hypothalamus leads to NT1, a sleep–wake disorder characterized by intrusion of REM sleep features into wakefulness, and this has also been demonstrated in several animal models of disruption of orexin action. These findings together with the ability of orexin receptor stimulation to enhance wakefulness have provided evidence for the essential and nonredundant nature of the orexin system in sleep–wake regulation and stability, and have triggered the discovery and development of new hypnotics. The phase 2 clinical data of the DORA almorexant for insomnia were presented in 2007, less than 9 years after the discovery of the orexin system and 7 years after the demonstration of a definitive link between orexin loss and NT1. This is one of the fastest discovery-to-translation timelines in drug development history. Almorexant was stopped during phase 3 trials in 2011 due to safety issues, as occasionally happens with CNS drug candidates, but suvorexant was approved for insomnia in 2014, lemborexant in 2019 and daridorexant in 2022. Importantly, DORAs represent the first new and effective treatment of insomnia since the registration of Z drugs and are supported by very extensive phase 3 studies that were conducted to assure market approval. It is important to acknowledge that DORAs produce a closer replication of physiological sleep than classic hypnotics by promoting REM sleep, a component of normal sleep architecture important for memory consolidation. In contrast, all other hypnotics on the market suppress REM sleep, at least to some degree. It also appears that DORAs are safe hypnotics in sleep apnea and chronic obstructive pulmonary disease. There are more DORAs in drug discovery pipelines and expansion of indications on the horizon, especially related to promising early clinical results in substance use disorders. On the other hand, it is still early days for 2-SORAs, such as seltorexant, which may offer some advantages over DORAs in select patient groups. Clinical testing of seltorexant is currently focused primarily on insomnia with MDD. It is logical to combine the two principles (antidepressant and hypnotic), since MDD, generalized anxiety and others mental health disorders are commonly accompanied by insomnia/sleep disturbances. This is also the case for several neurodegenerative disorders, such as AD, frontotemporal dementia, and PD. In MDD, the dual action is represented in currently prescribed drugs, e.g., agomelatin and mirtazapine. On the other hand, sedation is a common side-effect of many types of drugs used in psychiatry, especially anxiolytics and antipsychotics. Regarding OX1, there are several approaches to target stress, anxiety and panic-related disorders, e.g., with the 1-SORAs JNJ-61293215 and ACT-539313, and eating disorders, obesity, and different forms of addiction, primarily with, e.g., CVN45502. Furthermore, the 1-SORA CVN766 is in clinical trials for schizophrenia. Active research is ongoing for these and other indications, and there are suitable molecules available for academic research as well. Altogether, it will be interesting to see which application subtype-selective and nonselective orexin receptor antagonists will find in humans among (or in addition to) the many potential ones suggested by rodent studies.
An interesting point relates to the combination of cognitive and sleep impairments caused by sleep deprivation, aging, and importantly, neurodegenerative diseases. Based on current data, DORAs are safe and efficacious in adults and elderly populations, but naturally long-term data are missing so far. Restoring physiological sleep with DORAs or 2-SORAs in such populations may help to improve memory consolidation, on the one hand, and perhaps increase the elimination of toxic waste via the glymphatic system, on the other. The latter, if valid, could potentially slow the progression of neurodegenerative, proteinopathic disorders common to older people, although clinical evidence for this is currently lacking.
Orexin replacement/supplementation is a promising therapeutic avenue for the treatment of narcolepsy − in which it has a more direct action on the mechanism behind the symptoms than the current treatment regimes, which are essentially composed of daytime use of nonselective stimulants and nighttime use of hypnotics – as well as for other sleep–wake disorders, where excessive daytime sleepiness is a major symptom. Finding the most effective method of delivering orexin, its peptide analogs or small molecule receptor agonists into the brain remains as an obstacle for developing replacement/supplementation into a viable therapy. Since peptides do not readily cross the blood–brain barrier, traditional peripheral administration routes are unlikely to be effective. Intrathecal orexin is effective in treating narcolepsy symptoms in mice and is a potential option for human NT1 patients, although this is a highly invasive method. Intranasal orexin-A appears to cross the blood–brain barrier, is minimally invasive, and improves narcolepsy symptoms in mice and to some extent in humans (Calva and Fadel, 2020); however, no validated clinical studies have been performed as yet. The obvious alternative is to develop nonpeptide orexin receptor agonists – OX2 is the current focal point – that can cross the blood–brain barrier when administered orally. After intravenous infusion, the OX2 receptor agonist danavorexton/TAK-925 was found to cross the blood–brain barrier and to promote wakefulness in NT1. This led to the development of an orally active analog, firazorexton/TAK-994, which increases wakefulness and decreases sleep fragmentation and cataplexy in two narcolepsy mouse models and humans. Although the development of firazorexton/TAK-994 has been stopped, TAK-861 is already moving toward phase 3 trials for NT1. JZP-441 (structure unpublished) from Jazz Pharmaceuticals and Sumitomo Pharma was also undergoing phase 1 clinical trials for NT1 (NCT05651152 and NCT05720494), but the biotech web media (see, e.g., https://sleepreviewmag.com/sleep-treatments/pharmaceuticals/emerging-compounds/jazz-pauses-orexin-agonist-trial-reports-visual-disturbances/) have reported that the development would be stopped due to adverse effects – at the time of this review the company has not released any communiqué. These findings are promising proofs of concept; however, the long-term effectiveness and safety in narcolepsy patients has not been established. Meanwhile, danavorexton/TAK-925 is also still in clinical development to treat post-anesthesia recovery as an intravenous formulation (Suzuki et al., 2024). Because orexin receptors are also expressed outside the brain, peripheral side-effects will need to be determined. Along these lines, there are potentially promising peripheral effects of orexin receptor agonists in cancer. As concerns the CNS, it would be interesting to see the effects of orexin receptor agonists on metabolism, nociceptive signaling, and neurologic and psychiatric parameters in humans.
Several unknowns and challenges remain in the orexin field. The development of assays for accurate measurement of orexin in physiological fluids is essential to accelerate the progress of the research of the role of the orexin system in different physiological processes and disorders and the subsequent application to clinic. The highly specific mass spectrometric assays have not been developed to the resolution assumably required (orexin breakdown fragments) and the current PET ligands show too low specificity to be useful. Another problematic area in orexin research is the receptor signaling. Orexin receptors, like many other GPCRs, can couple promiscuously to various G proteins and other proteins interacting with GPCRs. The molecular details of orexin receptor signaling have not been exposed but the glimpses offered suggest unusual features that await to be discovered. However, to reach meaningful results, understanding of the general signaling mechanisms of GPCRs is mandatory.
In summary, the developments in the orexin field over the last 25 years, since the orexin system was discovered, are truly remarkable at both the preclinical and clinical levels. Although significant questions and shortcomings still remain, several pharmacological and genetic tools are already available and new tools are being developed to solve these issues and further characterize the role of the orexin system in human health and disease.
Note Added in Proof: An incorrect statement regarding a co-development agreement between Merck and Cerevance was accidentally published in the Fast Forward version that appeared online June 20, 2024. The statement has now been corrected.
Data Availability
The paper is exclusively based on published data. The data used to generate Fig. 2 as well as the other comparisons among mammalian species were acquired from Ensembl and NCBI.
Authorship Contributions
Performed data analysis: Borgland, Hoyer, Jacobson, Kukkonen, Rinne.
Created figures: Borgland, Kukkonen, Rinne.
Wrote or contributed to the writing of the manuscript: Borgland, Hoyer, Jacobson, Kukkonen, Rinne.
Footnotes
- Received June 8, 2023.
- Accepted June 6, 2024.
This work was supported by the Magnus Ehrnrooth Foundation, the Liv & Hälsa Foundation and Finska Läkaresällskapet and the European Cooperation in Science and Technology (COST) through CA18133 ERNEST (to J.P.K.).
J.P.K. has received an honorarium from Idorsia Pharmaceuticals Ltd in 2022 for participation in the symposium Orexin Science Summit and writing a review on orexin signaling; S.L.B. had a limited term contract with Septerna, Inc.; L.H.J. has consulted for Eisai Ltd (2019) and received research funding support from Merck Co., Inc., and Boehringer Ingelheim; and D.H. and M.K.R. declare no conflicts of interest.
Abbreviations
- 1-SORA and 2-SORA
- selective orexin receptor antagonist with preference for OX1 and OX2, respectively
- 2-AG
- 2-arachidonoyl glycerol
- AA
- arachidonic acid
- AC
- adenylyl cyclase
- ACT-335827
- (R)-2-[(S)-1-(3,4-dimethoxybenzyl)-6,7-dimethoxy-3,4-dihydroisoquinolin-2(1H)-yl]-N-isopropyl-2-phenylacetamide
- ACT-539313
- (R)-(3-[3-(2H-1,2,3-triazol-2-yl)benzyl]morpholino)-[4-methyl-2-(2H-1,2,3-triazol-2-yl)phenyl]methanone
- AD
- Alzheimer's disease
- almorexant
- ACT-078573; (R)-2-([S]-6,7-dimethoxy-1-[4-(trifluoromethyl)phenethyl]-3,4-dihydroisoquinolin-2[1H)-yl]-N-methyl-2-phenylacetamide
- anandamide
- N-arachidonoylethanolamine a.k.a. arachidonoylethanolamide
- ARN-776
- (2R,3S)-N-ethyl-2-[([(3S)-3-isopropylcyclohexyl]oxy) methyl]-3-(methylsulfonamido)piperidine-1-carboxamide
- BAT
- brown adipose tissue
- BMI
- body mass index
- CHO-K1 cells
- A Chinese hamster ovary cell line
- cryo-EM
- cryo-electron microscopy
- CNS
- central nervous system
- compound 1
- 3'-(N-[3-(2-[2-(2H-1,2,3-triazol-2-yl)benzamido]ethyl) phenyl]sulfamoyl)-4'-methoxy-N, N-dimethyl-(1,1'-biphenyl)-3-carboxamide
- compound 56
- (3-ethoxy-6-methylpyridin-2-yl)-([1R,4S,6R]-4-[([5-(trifluoromethyl)pyrimidin-2-yl]amino)methyl]-3-azabicyclo[4.1.0]heptan-3-yl)methanone
- Cp-1
- (R)-2-([S]-6,7-dimethoxy-1-(2-[6-(trifluoromethyl)pyridin-3-yl]ethyl)-3,4-dihydroisoquinolin-2[1H]-yl)-N-methyl-2-phenylacetamide
- CSF
- cerebrospinal fluid
- CVN45502
- N-[(1S,2S)-2-methyl-2-([5-(trifluoromethyl)pyrazin-2-yl] amino)cyclopentyl]-3-(triazol-2-yl)pyridine-2-carboxamide
- D609
- O-(octahydro-4,7-methano-1H-inden-5-yl) carbonopotassium dithioate
- DAG and MAG
- di- and monoacylglycerol, respectively
- DAGL and MAGL
- di- and monoacylglycerol lipase, respectively
- danavorexton
- TAK-925; methyl (2R,3S)-3-(methanesulfonamido)-2-[(4-phenylcyclohexyl)oxymethyl]piperidine-1-carboxylate
- daridorexant
- ACT-541468; (S)-[2-(5-chloro-4-methyl-1H-benzo[d]imidazol-2-yl)-2-methylpyrrolidin-1-yl]-[5-methoxy-2-(2H-1,2,3-triazol-2-yl)phenyl]methanone
- DORA
- dual orexin receptor antagonist, i.e. with no major preference for either receptor subtype
- DORA-12
- (R)-[2-(2H-1,2,3-triazol-2-yl)phenyl]-[4-(6-fluoroquinazolin-2-yl)-7-methyl-1,4-diazepan-1-yl]methanone
- ECL
- extracellular loop (of a GPCR)
- EMPA
- N-ethyl-2-[(6-methoxypyridin-3-yl)-(2-methylphenyl) sulfonylamino]-N-(pyridin-3-ylmethyl)acetamide
- ERK
- extracellular signal-regulated kinase
- EGFP
- enhanced green fluorescent protein
- filorexant
- MK-6096; [(2R,5R)-5-[(5-fluoropyridin-2-yl)oxymethyl]-2-methylpiperidin-1-yl]-(5-methyl-2-pyrimidin-2-ylphenyl)methanone
- firazorexton
- TAK-994; N-[(2S,3S)-2-([3-(3,5-difluorophenyl)-2-fluorophenyl] methyl)-1-(2-hydroxy-2-methylpropanoyl)pyrrolidin-3-yl]methanesulfonamide
- GPCR
- G protein-coupled receptor
- (lyso-)GPL
- (lyso-)glycerophospholipid
- HEK293
- a human embryonic kidney cell line
- icv
- intracerebroventricular(ly)
- IP3
- inositol-1,4,5-trisphosphate
- IPSU
- 2-[(1H-indol-3-yl)methyl]-9-(4-methoxypyrimidin-2-yl)-2,9-diazaspiro[5.5]undecan-1-one
- JH112
- [(2S)-2-[(2S)-butan-2-yl]-4-(5-chloro-1,3-benzoxazol-2-yl)-1,4-diazepan-1-yl]-[5-methyl-2-(triazol-2-yl)phenyl] methanone
- JNJ-10397049
- 1-(2,4-dibromophenyl)-3-[(4S,5S)-2,2-dimethyl-4-phenyl-1,3-dioxan-5-yl]urea
- JNJ-54717793
- [3-fluoro-2-(pyrimidin-2-yl)phenyl] ([1S,2R,4R]-2-[(5-[trifluoromethyl]pyrazin-2-yl)amino]-7-azabicyclo[2.2.1]heptan-7-yl)methanone
- JNJ-61393215
- [3-fluoro-2-(pyrimidin-2-yl)phenyl]-([1S,4R,6R]-6-[(5-[trifluoromethyl)pyridin-2-yl)oxy)-2-azabicyclo[2.2.1] heptan-2-yl)methanone
- KO
- knockout
- lemborexant
- E-2006; (1R,2S)-2-([(2,4-dimethylpyrimidin-5-yl)oxy] methyl)-2-(3-fluorophenyl)-N-(5-fluoropyridin-2-yl)cyclopropane-1-carboxamide
- LPA
- lysophosphatidic acid
- LSN2424100
- 4-fluoro-N-(1H-imidazol-2-ylmethyl)-N-(2-phenylphenyl)benzenesulfonamide
- LPC
- lysophosphatidylcholine
- MAPK
- mitogen-activated protein kinase
- MK-1064
- 5-(5-chloropyridin-3-yl)-N-[(5,6-dimethoxypyridin-2-yl)methyl]-2-pyridin-2-ylpyridine-3-carboxamide
- MDD
- major depressive disorder
- Nag 26
- N-[2-(3-[(5-[3-(dimethylcarbamoyl)phenyl]-2-methoxyphenyl)sulfonylamino]anilino)ethyl]-3-methylbenzamide
- NCX
- Na+–Ca2+-exchanger
- NMR
- nuclear magnetic resonance
- NREM (sleep)
- non-rapid eye movement (sleep)
- NSCC
- non-selective cation channels
- NT1 and NT2
- narcolepsy type 1 and 2, respectively
- PA
- phosphatidic acid
- PD
- Parkinson's disease
- PDB
- protein data bank
- PIP2
- phosphatidylinositol-4,5-bisphosphate
- PKA and PKC
- protein kinase A and C, respectively
- PLA2, PLC and PLD
- phospholipase A2, C and D, respectively
- PPO
- prepro-orexin
- REM (sleep)
- rapid eye movement (sleep)
- RMSD
- root mean square deviation (of atomic positions)
- RTOXA-43
- 4'-methoxy-N-methyl-3'-(N-[3-([2-(3-methylbenzamido) ethyl]amino)phenyl]sulfamoyl)-N-(pyridin-4-ylmethyl)-(1,1'-biphenyl)-3-carboxamide
- SB-334867
- 1-(2-methyl-1,3-benzo[d]oxazol-6-yl)-3-(1,5-naphthyridin-4-yl)urea
- SB-408124
- 1-(6,8-difluoro-2-methylquinolin-4-yl)-3-[4-(dimethylamino)phenyl]urea
- SB-649868
- (S)-N-[(1-[5-(4-fluorophenyl)-2-methylthiazole-4-carbonyl] piperidin-2-yl)methyl]benzofuran-4-carboxamide
- SB-674042
- (S)-[5-(2-fluorophenyl)-2-methyl-1,3-thiazol-4-yl]-[(2S)-2-[(5-phenyl-1,3,4-oxadiazol-2-yl)methyl]pyrrolidin-1-yl]methanone
- seltorexant
- JNJ-42847922; [5-(4,6-dimethylpyrimidin-2-yl) hexahydropyrrolo[3,4-c]pyrrol-2(1H)-yl][2-fluoro-6-(2H-1,2,3-triazol-2-yl)phenyl]methanone
- SNP
- single nucleotide polymorphism
- suvorexant
- MK-4305; [(7R)-4-(5-chloro-1,3-benzoxazol-2-yl)-7-methyl-1,4-diazepan-1-yl]-[5-methyl-2-(triazol-2-yl)phenyl] methanone
- SWS
- slow wave sleep
- TCS1102
- BBAC; (S)-N-[(1,1'-biphenyl)-2-yl]-1-(2-[(1-methyl-1H-benzo[d]imidazol-2-yl)thio]acetyl)pyrrolidine-2-carboxamide
- TCS-OX2-29
- (S)-1-[6,7-dimethoxy-3,4-dihydroisoquinolin-2(1H)-yl]-3,3-dimethyl-2-[(pyridin-4-ylmethyl)amino]butan-1-one
- TM
- transmembrane helix (of a GPCR)
- TRP (channel)
- transient receptor potential (channel)
- TOL-27
- N,6-dimethyl-2-(N-[3-([2-(3-methylbenzamido)ethyl]amino) phenyl]sulfamoyl)azulene-1-carboxamide
- UBO-QIC/FR900359
- L-threonine,(3R)-N-acetyl-3-hydroxy-L-leucyl-(aR)-a-hydroxybenzenepropanoyl-2,3-idehydro-N-methylalanyl-L-alanyl-N-methyl-L-alanyl-(3R)-3-[([2S,3R]-3-hydroxy-4-methyl-1-oxo-2-[(1-oxopropyl)amino]pentyl)oxy]-L-leucyl-N, O-dimethyl-,(7→1)-lactone (9CI)
- VTA
- ventral tegmental area
- YNT-185
- 2-(dimethylamino)-N-[2-(3-[(5-[3-(dimethylcarbamoyl) phenyl]-2-methoxyphenyl)sulfonylamino]anilino)ethyl] benzamide
- (R)-YNT-3708
- (E)-3-[4-methoxy-3-([(8R)-8-([2-(3-methoxyphenyl)acetyl]-methylamino)-5,6,7,8-tetrahydronaphthalen-2-yl]sulfamoyl)phenyl]-N-pyridin-4-ylprop-2-enamide
- Copyright © 2024 by The Author(s)
This is an open access article distributed under the CC BY-NC Attribution 4.0 International license.