Next Article in Journal
The Content of Phenolic Compounds and Organic Acids in Two Tagetes patula Cultivars Flowers and Its Dependence on Light Colour and Substrate
Previous Article in Journal
The Effect of pH and Sodium Caseinate on the Aqueous Solubility, Stability, and Crystallinity of Rutin towards Concentrated Colloidally Stable Particles for the Incorporation into Functional Foods
Previous Article in Special Issue
Effective Use of Empirical Data for Virtual Screening against APJR GPCR Receptor
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Novel Molecular Targets of Antidepressants

by
Małgorzata Jarończyk
* and
Jarosław Walory
National Medicines Institute, 30/34 Chełmska Str., 00-725 Warsaw, Poland
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(2), 533; https://doi.org/10.3390/molecules27020533
Submission received: 6 December 2021 / Revised: 10 January 2022 / Accepted: 12 January 2022 / Published: 14 January 2022

Abstract

:
Antidepressants target a variety of proteins in the central nervous system (CNS), the most important belonging to the family of G-protein coupled receptors and the family of neurotransmitter transporters. The increasing number of crystallographic structures of these proteins have significantly contributed to the knowledge of their mechanism of action, as well as to the design of new drugs. Several computational approaches such as molecular docking, molecular dynamics, and virtual screening are useful for elucidating the mechanism of drug action and are important for drug design. This review is a survey of molecular targets for antidepressants in the CNS and computer based strategies to discover novel compounds with antidepressant activity.

1. Introduction

Major depressive disorder (MDD) is a serious and, in severe cases, a lethal disorder. Despite extensive research, the knowledge about the pathophysiology, specific mechanisms, and regulatory pathways underlying the disease remains limited [1]. Based on clinical effects and molecular mechanism of antidepressant drugs, the monoamine hypotheses were proposed more than 50 years ago, suggesting that depression pathophysiology is associated with deficiencies of the monoamine neurotransmitter serotonin (5-HT), dopamine (DA), and norepinephrine (NE) [2]. However, it is evident that monoamine-deficiency only partly explains the pathogenesis, and other neurotransmitters including acetylcholine, glutamate, and gamma-amino butyric acid (GABA) have also been connected to the etiology of depression. The monoamine deficiency may rather be a consequence than a cause of the disorder [3,4]. Although numerous antidepressant drugs are on the market, effective management of MDD is still problematic. Several drugs such as selective serotonin reuptake inhibitors (SSRI) and serotonin-norepinephrine reuptake inhibitors (SNRI), older tricyclic antidepressants (TCA), and a growing number of other types of drugs are used as antidepressants. However, delayed onset of action and undesirable side effects are common. Furthermore, approximately 50% of depressed patients achieve complete remission despite optimized treatment, including trials of multiple drugs with and without simultaneous psychotherapy. Therefore, there is an urgent need to develop new approaches to obtain more effective, safer, and faster antidepressants.
A better understanding of pathogenic processes involved in depression is needed to develop improved therapeutics. The monoamine-depletion hypothesis alone cannot fully explain the pathogenesis of depression. Depression is a complex disease associated with dysregulation of the immune system and the hypothalamic-pituitary-adrenal (HPA) axis, as well as neurotransmitter and neurotrophic systems imbalance [5,6]. The “inflammatory hypothesis” of depression was proposed more than two decades ago and suggests that inflammatory processes are involved in the onset and maintenance of MDD [7,8]. The increased concentrations of circulating pro-inflammatory cytokines such as interleukin-6 (IL-6), tumor necrosis factor-α (TNF-α), and interleukin-1β (IL-1β) were found in depressed patients. IL-1β is one of the most potent pro-inflammatory cytokines secreted by activated inflammatory cells that was found to play an important role in depressive-like behaviors as well as stress-related cellular actions. In the central nervous system (CNS), a key player involved in the secretion of biologically active IL-1β is P2X7R, an ATP-gated ion channel present on immune cells. Pharmacological targeting of P2X7R might have the potential as a future therapy for the treatment of a subset of patients with depression [9]. Hyperactivity of the HPA-axis has also been implicated in the pathophysiology of depression. There are two glucocorticoid receptors in the brain that control the HPA axis: the mineralocorticoid receptor (MR) and the glucocorticoid receptor (GR). Hence, targeting these receptors by antidepressants was found to normalize the HPA-axis and is considered a promising strategy for the development of novel antidepressants drugs [10]. Reduced levels of brain-derived neurotrophic factor (BDNF) lead to depression, according to the neurotrophic hypothesis of depression [11,12]. Other neurotrophic/growth factors linked to depression include the vascular endothelial growth factor (VEGF), the fibroblast growth factor 2 (FGF2), and the insulin-like growth factor 1 (IGF-1).
G-protein coupled receptors (GPCRs) and neurotransmitter transporters play well-established roles in the pathophysiology of depression and are targets for present antidepressant drugs, and are also recognized targets in the search for more specific and effective pharmacological approaches. In the present review, we will focus on GPCRs and neurotransmitter transporters in the CNS as targets for antidepressants.

2. GPCR as Targets for Antidepressants

GPCRs represent the largest class of membrane receptors in humans, containing seven membrane-spanning helices. They transduce extracellular signals into the cell interior through coupling to heterotrimeric G proteins and activation of different signaling pathways. Members of GPCRs from the different receptor families have been implicated in depression (see review [13]). The most important GPCR targets, antidepressant compounds and signaling pathways activated by target proteins, are listed in Table 1.
The first GPCR structure at an atomic resolution was the structure of bovine rhodopsin in its inactive state (Figure 1), which was released in 2000 [14]. X-ray crystallography and cryo-electron microscopy (cryo-EM) technology enabled for GPCR structure determination and significantly deepened our knowledge of molecular mechanisms of signal transduction. At present, 455 structures representing 82 different receptors, are deposited in the Protein Data Bank (PDB) [15]. Structural studies of GPCRs have provided insight on the arrangements of transmembrane domains, the location of orthosteric, allosteric, bitopic, as well as biased ligand binding sites, and conformational changes upon GPCR activation and inactivation [16]. Structural knowledge of GPCRs have also given the possibility of structure-based drug design for obtaining new and improved compounds targeting the receptors.

2.1. Serotonin Receptors

Serotonin receptors are found throughout central and peripheral nervous systems in the brain, mainly in regions involved in the neurobiology of anxiety and depression. There are 7 families of 5-HT receptors (5-HT1–5-HT7), which are further subdivided into 14 distinct receptor subtypes. Except for the 5-HT3 receptor, which is a ligand-gated ion channel, all known 5-HT receptors are G-protein coupled [17].
One of the most important and extensively studied 5-HT1 subtypes is the 5-HT1A receptor due to its implication in the pathophysiology of several neuropsychiatric disorders including anxiety and MDD [18]. The 5-HT1A receptors are distributed in the limbic, cortical, and dorsal and median raphe nucleus. The 5-HT1A receptors couple to the Gi/Go pathways, inhibiting the adenylyl cyclase to reduce cyclic adenosine monophosphate (cAMP) level and activating the G-protein inward rectifying potassium (GIRK) channels [19]. 5-HT1A function as presynaptic autoreceptors and postsynaptic heteroreceptors and signal to diverse and sometimes opposing pathways [19]. Identifying biased 5-HT1A ligands that preferentially activate one pathway over another may offer novel strategies for depression treatment [20]. Several 5-HT1A receptor agonists such as buspirone and tandospirone are medications approved to treat anxiety and depression [21]. Moreover, biased ligands of 5-HT1A such as F-15599 and F-13714 activate selectively postsynaptic heteroreceptors and presynaptic autoreceptors, respectively [22,23].
Until recently, the atomistic structure of the 5-HT1A receptor was not resolved. Thus, homology models of 5-HT1A receptor were built based on the bovine rhodopsin template [14], which after a while was replaced by the β2AR crystal structure [24]. Later, the β1-adrenergic [25] and the A2A adenosine [26] receptor crystal structures were also used as templates. However, during the last years the crystal structure of the 5-HT1B receptor has been the most used template for 5-HT1A receptor models [27]. Recently, the cryo-EM structures of the 5-HT1A in complex with the Gi subunit were determined at a resolution of 3.0–3.1 Å [28]. These structures reveal insight in the primary activity of the 5-HT1A receptor and the mode of drug recognition.

2.2. Dopamine Receptors

The dopamine system plays an important role in the pathogenesis of depression, and efficacy of dopamine receptor ligands in the treatment of human depression have been reported [29]. Dopamine receptors are grouped into two families: D1-like receptors (D1- and D5-receptors) and D2-like receptors (D2-, D3-, and D4-receptors). The D2-like receptors (D2/D3) play a key role in the response to antidepressant treatment [30]. This group of receptors have been found mainly in the striatum, amygdala, cerebral cortex, hippocampus, and pituitary gland [31]. They couple to Gαi/o proteins and primarily inhibit adenylate cyclase.
Aripiprazole is a D2 partial agonist that was initially approved for the treatment of schizophrenia, and later as an augmenting agent in MDD [32]. The brexpiprazole is a partial D2 agonist that has lower intrinsic activity than aripiprazole but higher 5-HT1A/2A receptor binding affinity [33]. The compounds with high affinity and occupancy of both D2 and D3 dopamine receptors may be effective in the treatment of depressive disorders and schizophrenia [34]. Cariprazine is a potent dopamine D2/D3 receptor partial agonist that was found to be effective in the treatment of depression [35]. Other receptor interactions are also implicated in the antidepressant effects of cariprazine, notably 5-HT1A receptor agonism.
The resolved crystal structures of the D3 and D2 receptors are helpful for understanding their mechanisms of action and their key interactions with the ligand in the binding site [36,37]. Studies to identify potent, novel, and selective dopamine D2 and D3 receptor ligands have been performed and novel potent compounds have been synthesized and tested using in vitro and in silico methods [38].

2.3. Opioid Receptors

A novel target for the treatment of depression may be the endogenous opioid system, which consists of mu (μ), kappa (κ), and delta (δ) opioid receptors (MORs, KORs, DORs) and the non-opioid receptor, nonciceptin (NOP), previously referred to as opioid-like 1 receptor. Several studies concerning the role of opioid receptors in depression treatment have emerged [39,40,41]. Opioid receptors are widely distributed in the hippocampus, nucleus accumbens, prefrontal cortex, amygdala, claustrum, thalamus, hypothalamus, ventral tegmental area and dorsal raphe nucleus [42]. Opioid receptors are coupled to inhibitory heterotrimeric Gαi/o proteins and also stimulate G protein-independent signaling pathways, notably via β-arrestins.
Buprenorphine is a partial agonist of μ opioid receptor and κ opioid receptors, and has activity at DOR and NOP receptors [43]. In preclinical studies, the antidepressant effects of buprenorphine were shown to be mediated via κ opioid receptors [44]. Furthermore, in vivo studies and behavioral tests indicated that nalmefene (NMF), a partial κ opioid receptor agonist and potent μ opioid receptor antagonist, has antidepressant activity [45]. Tianeptine is a MOR agonist and activate signaling pathways different to that of morphine [46]. A specific blockade of NOP receptors has been suggested to induce antidepressant-like action in preclinical tests [47]. The NOP receptor antagonist BTRX-246040 (known as LY-2940094) displays antidepressant-like effects in rodents models [48,49].
High-resolution crystal structures providing insight into molecular determinants required for ligand binding to opioid receptors were resolved: μ receptor [50,51], δ receptor [52], κ receptor [53], and NOP receptor [54,55]. Recently a 3.1-Å resolution X-ray structure of the KOR in the activated state bound to the high-affinity agonist MP1104 was reported [56]. Computational studies using the crystal structure of NOP receptor gave insight into the binding mode of two novel NOP antagonists, one selective (BTRX-246040), and one unselective (AT-076) antagonist [57].

2.4. Glutamate Receptors

Targeting the glutamatergic system may be a promising strategy for developing new treatments in mood disorders [58]. The glutamate receptors are divided into two groups: the ionotropic glutamate receptors and G-protein coupled metabotropic glutamate receptors (mGluRs). The first group includes N-methyl-D-aspartate (NMDA), α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) and kainite receptors. The mGluRs are located presynaptically and postsynaptically, and are classified into three subgroups based on sequence similarity, G-protein coupling, and ligand selectivity. Receptors of subgroup I (mGlu1 and mGlu5) couple to Gq/G11 and activate phospholipase Cβ, while receptors of subgroup II (mGlu2 and mGlu3) and subgroup III (mGlu4, mGlu6–8) couple to Gi/o proteins, thus leading to adenylyl cyclase inhibition. Subgroup II of mGluRs is an attractive target for the development of novel antidepressants, as confirmed by preclinical studies [59]. The mGluR2 is expressed only in cerebellar cortex and olfactory bulbs while the mGluR3 is extensively detected in dentate gyrus, cerebral cortex, striatum, substantia nigra pars reticulata, olfactory tubercle, and lateral septal nucleus [60]. Preclinical studies indicated that two mGlu2/3 receptor antagonists (MGS0039 and LY341495) produced antidepressant effects [61,62]. Moreover, the negative allosteric modulator of mGlu2/3 receptors, RO4491533, was found to have antidepressant-like effects in mice in the forced swim test (FST) [63].
All mGlu receptors exist as homodimers possessing an extracellular Venus flytrap (VFT) domains, which is linked via cysteine rich-domains (CRDs) to their 7-transmembrane (TM) domain. The 7TM domain of mGlu receptors is the binding site of various non-endogenous allosteric ligands that can modulate signaling either on their own or in conjunction with orthosteric ligands, thereby acting as positive allosteric modulators (PAMs), or negative allosteric modulators (NAMs). Crystal structures of mGlu1 [64] and mGlu5 [65] 7TM domains show that the ligand binding site of allosteric modulators largely overlap with that of orthosteric ligands in family A GPCRs. Computational and experimental studies of mGlu2 allosteric modulators based on the crystal templates of the mGlu1 [64] and mGlu5 [65] were performed to study binding modes of known NAMs and PAMs [66].

2.5. Orphan Receptors

The ‘orphan’ GPCRs are examples of genes without known functions. The expressed proteins share the structural similarity of seven transmembrane helices with other GPCRs but are called ‘orphans’ because their endogenous ligands have not been identified yet. Recent findings show that orphan GPCRs (oGPRs) may be implicated in depression and be putative targets for new drug development [67,68]. No crystal structures of the oGPRs are available yet, although homology models are deposited in the GPCR database [69].
GPR26 is an orphan GPCR expressed in the hippocampus, amygdala, and thalamus of the human brain [70]. GPR26 is coupled to Gs and activates the adenylyl cyclase pathway. Behavioral tests on GPR26 knockout mice indicated that GPR26 is important in regulating anxiety- and depression-like behaviors [71].
GPR56 is involved in a number of biological functions such as myelin formation, neurogenesis, and oligodendrocyte development. In vivo experiments indicated reduction of GPR56 expression in the prefrontal cortex (PFC) and dorsal hippocampus, which can be reversed after antidepressant treatment [72]. It was found that protein kinase B (AKT), glycogen synthase kinase 3 (GSK3), and eukaryotic initiation factor 4E (EIF4) pathways involved in depression were upregulated in cells after GPR56 agonists treatment such as peptides P7 “TYFAVLM-NH2” and P19 “TYFAVLMQLSPALVPAELL-NH2” [72]. In summary, it is suggested that GPR56 may represent a potential molecular target for the treatment of depression.
GPR158 is highly upregulated in PFC area and plays an important role in the regulation of depression [73]. GPR158 lacks the ligand binding VFT domain, but it has conserved amino acids involved in G protein binding. The plasma membrane GPR158 acts as an anchor for regulating G protein signaling 7 (RGS7) complexes, thus may modulate the signaling of other GPCRs.

2.6. Trace Amine-Associated Receptor

In the CNS, trace amines (TAs) such as tyramine, β-phenylethylamine (β-PEA), octopamine, and tryptamine were found to play an important role as neurotransmitters. Their dysregulations are implicated in the pathophysiology of neuropsychiatric disorders such as depression and schizophrenia. Clinical studies reported decreased urinary excretion of β-PEA in patients with depressive disorders and PEA administration produced long-lasting relief of depression in a patient population [74,75]. In 2001, the trace-amine-associated receptors (TAARs) were discovered and classified into the three subfamilies of TAARs 1–4, TAAR5, and TAARs 6–9 [76]. The most studied of these receptors is TAAR1, which could be activated by different biogenic amines and psychoactive compounds [77]. TAAR1 is widely expressed across the mammalian brain, particularly in cerebral cortex and hippocampus and signals through stimulatory Gs proteins to elevate intracellular cAMP levels and stimulate inwardly rectifying K+ channels [78,79]. In behavioral animal studies, the partial agonists of TAAR1, RO5203648 demonstrated potential antidepressant-like properties [80]. Furthermore, the studies of RO5263397, TAAR1 agonist demonstrated a response in the forced swim test, a rodent model of depression-like behaviors [81].
The second most studied member of the TAARs family, TAAR5, was initially identified as a putative neurotransmitter receptor (PNR) and expression of TAAR5 mRNA was reported in the human amygdala, the hippocampus, the caudate nucleus, the thalamus, the hypothalamus, and the substantia nigra and skeletal muscles [82]. Recent animal studies revealed involvement of this receptor in the pathogenesis of neuropsychiatric diseases [83]. The TAAR5 knockout mice had elevated levels of dopamine and its dopamine metabolites and altered levels of serotonin in the brain [84]. The human TAAR5 signals through Gs, the Gq/11, or G12/13-dependent mitogen-activated protein kinase pathways. Trimethylamine is one of the known agonists of TAAR5, but until now antagonists have not been discovered.
Crystal structures of trace amine receptors have not been resolved yet. However, homology models of TAAR1 were built based on the available crystal structures of human β2-adrenoreceptor to identify novel ligands of this receptor and key amino acids responsible for ligand binding [85,86]. The murine and human homology models of TAAR5 were built based on the crystal structures of human β2-adrenoreceptor and virtual screening was performed in order to find novel ligands [87]. These studies identified two compounds acting as murine TAAR5 antagonists, which were further validated by experimental studies.

2.7. Cannabinoid Receptors

The importance of the endogenous cannabinoid system (ECS) in depression, composed of cannabinoid receptor types 1 and 2 (CB1R and CB2R), has been confirmed in preclinical and clinical studies. Bambico et al. [88] reported that at low doses, the CB1R agonist WIN55212-2 exerts potent antidepressant-like properties in the rat FST. An increase of CB1R density in the pre-frontal cortex and concomitant mediated signaling suggest a role of the endocannabinoid system in the etiology of depression [89]. Animal studies based on a mouse model with CB2R deletion in dopamine neurons showed that CB2Rs in dopamine neurons play a role in modulating depression- and anxiety-like behaviors [90,91]. The CB1 receptors are located in the central nervous system, particularly in the hippocampus, prefrontal cortex, basal ganglia, cerebellum, amygdala, spinal cord, and mesolimbic nuclei [92]. The CB2 receptors are mainly distributed in the immune system (the spleen, tonsils, immune system cells), but can also be found within the CNS in microglia [93]. CB1 and CB2 receptors are both Gi/o-coupled GPCRs, and their activation leads to the inhibition of adenylyl cyclase and reduction in the production of cAMP [94]. Since both cannabinoid receptors play an important role in mood regulation (see review [95]), studies of searching for novel pharmacological agents is needed. As an example, antidepressant-like behavioral properties of CB1R agonists such as Δ9-tetrahydrocannabinol (THC) and rimonabant, terpineol were confirmed in olfactory bulbectomy studies [96]. Among agents acting through CB2R, β-caryophyllene (BCP), a naturally available sesquiterpene, is pharmacologically a selective agonist for CB2 receptors that produced an antidepressant–like effect shown in the animal model of depression [97].
Recently the crystal structures of CB1 [98,99] and CB2 [100] receptors were published. The role of cannabinoid receptor signaling pathway in the antidepressant-like effect of terpineol was confirmed by using pharmacological and molecular modeling approaches [101]. Terpineol shares structural similarity with cannabinoid receptor ligands and showed a coherent predicted binding mode mainly against CB1 and CB2 receptors during docking analyses [101].

2.8. Cholinergic Receptors

In 1980, Janowsky et al. [102] and Dilsaver et al. [103] postulated that the cholinergic system is involved in the pathophysiology of depression. The role of both muscarinic and nicotinic cholinergic receptors in the mood regulation, based on human and animal studies, has been reviewed by Dulawa and Janovsky [104]. Nicotinic receptors are ligand gated ion channels, whereas muscarinic receptors belong to the super family of GPCRs. Activation of the muscarinic M1, M3, and M5 receptors stimulates a cascade of interactions, including G protein Gq/11 and phospholipase C activation, resulting in formation of the second messenger inositol triphosphate (IP3) from phospholipids, which subsequently induces a release of calcium.
The muscarinic M2 and M4 receptors couple through Gi and Go proteins, giving inhibition of adenylate cyclase activity and reduced formation of cAMP. The muscarinic antagonist scopolamine has been reported to have antidepressant-like effects in patients suffering from unipolar depression [105]. At first, it was found that scopolamine exerts its antidepressant effect by blockade of the M1 receptor, increasing the mechanistic target of rapamycin complex 1 (mTORC1) signaling in pre-frontal cortex [106]. Furthermore, the involvement of the M2 receptor in mediating antidepressant-like effects of scopolamine by increasing the expression of BDNF and activating the mTORC1 signaling pathway was confirmed [107]. Other antagonists such as VU0255035 also demonstrate their antidepressant effect by interaction with M1 or M2 receptors.
For better understanding of the molecular mechanism of the action of muscarinic receptors, knowledge of the crystal structure of these receptors is needed. So far, the crystal structures of the M2 and M3 muscarinic receptors and the M1 and M4 muscarinic receptors have been resolved [108,109,110].

2.9. GABA Receptors

The γ-Aminobutyric acid (GABA) is the principal neurotransmitter mediating neural inhibition in the brain. There are two types of GABAergic receptors: the ionotropic receptors GABA-A and GABA-C, and the metabotropic receptor GABA-B. Located both pre- and postsynaptically, GABA-B receptors influence cAMP production through coupling to Gi and Go proteins. GABA-B receptors are formed by dimerization of two homologous subunits GABAB(1) and GABAB(2), composed of three domains: a long extracellular N-terminal domain called the Venus fly-trap domain (VFT), which contains the orthosteric binding site for GABA; a heptahelical transmembrane domain (7TM); and a C-terminal intracellular tail. The role of GABA-B receptors in depression was first proposed by Lloyd [111] and confirmed in several studies. In preclinical studies, the knockout mice lacking the GABAB(1) subunit demonstrated decreased immobility in FST while no difference in mice behavior was observed in tail suspension test (TST) [112]. Baclofen, a known agonist of the GABA-B receptor was found to be effective in the treatment of posttraumatic stress disorder in the clinical studies [113]. The GABA-B antagonist CGP56433A exerted an antidepressant-like effect by decreasing the immobility in the FST [112]. Other GABA-B antagonists such as CGP36742 and CGP51176 were also effective in inducing antidepressant-like effects in FST [114]. The positive allosteric modulators GS39783 and CGP7930 were found to modulate activity of GABA-B by decreasing the affinity of antagonist radioligand in saturation experiments [115].
Recently, the GABA-B receptor structure and activity and interactions with compounds were reviewed by Evenseth et al. [116]. The published structures resolved by cryo-electron microscopy [117,118,119] provide insight into GABA-B receptor mechanism of action and information of allosteric and orthosteric ligand binding sites and may lead to design new potent antidepressants.

2.10. Neurokinin Receptors

Neuropeptides such as substance P (SP) and neurokinin A and neurokinin B belong to the family of neurotransmitters known as tachykinin. There are three subtypes of neurokinin receptors: NK1, NK2, and NK3. Neuroanatomical studies demonstrated that the predominant tachykinin receptor is NK1R in the human brain, whereas the expression of NK2 or NK3 receptors is either weak or absent. The NK1Rs are widely expressed throughout the CNS in areas including the amygdala, hypothalamus, hippocampus, and striatum [120]. These regions are involved in regulating the stress response and controlling affective behavior, such as depression and anxiety. Neurokinin-1 receptor signals through the Gαq-protein and induces activation of phospholipase C followed by production of IP3, leading to elevation of intracellular calcium as a second messenger. Further, cyclic AMP is stimulated by NK1R coupled to the Gαs-protein. In the clinical studies of patients with MDD, Kramer et al. [121,122] observed antidepressant effects of the NK1 receptor antagonists, MK-869 (aprepitant) and L-759274.
Recently, the crystal structure of hNK1R bound to the antagonist L760735, a close analog of aprepitant, was solved [123]. In addition, three high-resolution crystal structures of the human NK1 receptor bound to two small-molecule antagonist therapeutics—aprepitant, netupitant, and the progenitor antagonist CP-99,994 were determined [124]. Conformational changes at the helix II-VII interface upon antagonist binding were observed. The understanding of molecular structures provides opportunity to develop ligands that will selectively target different neurokinin receptors.

2.11. Cholecystokinin Receptors

Cholecystokinin (CCK) is a gut–brain peptide that has been implicated in stress and anxiety disorders [125]. There are two types of receptors: CCK1 (CCK-A) receptors are present mainly in peripheral tissues and in discrete brain regions, while CCK2 (CCK-B) receptors predominate in the central nervous system. The CCKR2 is widely distributed in the brain, especially in the cerebral cortex, nucleus accumbens, caudate nucleus, hippocampus, amygdala, substantia nigra, and ventral tegmental area [126,127]. The neuronal co-localization of CCK receptors and the dopaminergic system reveal involvement of both systems in neuropsychiatric and CNS disorders [128]. Signal transduction at CCK1 receptors is mediated via Gs protein while at CCK1R via Gq protein and to a lesser extent with Gi.
A reduction in the suppression of motility in shocked mice was observed after treatment with L-365,260, a CCK2R antagonist, which strongly suggests that CCK2R antagonists are able to induce antidepressant-like effects [129]. The blockade of CCK2R by the selective antagonist CI-988 prevented an increase in serum corticosterone levels in animal tests and increased immobility time in the FST [130].
Recently, the cryo-EM structures of CCK1R in complex with sulfated cholecystokinin-8 (CCK-8) coupled to different G-proteins were determined [131]. Furthermore, the crystal structures of CCK1R in complex with antagonists (devazepide and lintitript) and CCK-8, and two cryo-EM structures of CCK2R with gastrin coupled to Gi and Gq proteins were resolved [132]. They are the basis for studying the mechanism of ligand binding using molecular modeling methods to develop new antidepressants.

2.12. G-Proteins and G-Protein Regulating Proteins

G-proteins play an important role in transducing signals from receptors into cell and may be involved in the pathogenesis and treatment of mood disorders (see review [133]). Firstly, the alterations in the concentration or function of G proteins in peripheral blood elements and in postmortem tissues of patients with bipolar and other mood disorders were observed after treatment with lithium [134,135]. Further studies with other antidepressant do not confirm changes in G protein expression [136,137].
Regulators of G protein signaling (RGS) proteins control GPCR and linked G protein signaling. They are implicated in the CNS disorders (see reviews [138,139]). Among RGS proteins, RGS2 was found to be expressed in brain areas important in the pathogenesis of anxiety and depression such as the hippocampus, amygdala, cerebral cortex, hypothalamus, and raphe nucleus [140]. The crystal structure of RGS2 in complex with Gαq was resolved [141]. The studies on RGS7 show that the loss of striatal RGS7 induces an anxiolytic-like and antidepressant-like phenotype [142,143]. The crystal structure of RGS7-Gβ5 dimer has been resolved in 2018 [144]. Other regulatory proteins such as RGS4, RGS6, and RGS8 are widely expressed in the brain regions involved in depression [145,146]. Further studies are needed to understand the potential role of RGS proteins in major depressive disorders and their mechanism of action.

3. Neurotransmitter Transporters as Targets for Antidepressants

Several transporters in the solute carrier 6 (SLC6) family mediate the uptake of released neurotransmitters from the extracellular space into neurons and glial cells. They are widely expressed in the mammalian brain and play an important role in regulating neurotransmitter signaling [147,148]. The three types of cell membrane monoamine transporters (MATs) are the dopamine transporter (DAT), norepinephrine transporter (NET), and the serotonin transporter (SERT). These transporters mediate the uptake of monoamine neurotransmitters dopamine, norepinephrine, and serotonin, respectively, from the extracellular space into the intracellular compartment [149]. They are the main targets for the present antidepressants and several drugs of abuse [150].
Human MATs are large integral membrane proteins consisting of about 600 amino acids that contain 12 transmembrane α-helices (TM1 to TM12) connected by intra- and extracellular loops (ILs and ELs). Crystal structures of transporter bound with substrate and inhibitors have advanced our understanding of the mechanism of action of MATs and help in the discovery of novel antidepressants. The first structure elucidated was a bacterial homologue of the human monoamine transporters, the leucine transporter (LeuT) bound to its substrate in an occluded conformation (Figure 2) [151]. Furthermore, several crystal structures of LeuT in different conformational states were reported [152,153,154,155]. In addition, crystal structures of the Drosophila melanogaster dopamine transporter (dDAT) [156] and human serotonin transporter [157] in complex with different transport blockers and in altered conformational states were resolved and have given remarkable insights into the inhibitory mechanisms of these transporters.
Selective serotonin reuptake inhibitors, serotonin-norepinephrine reuptake inhibitors, and serotonin-norepinephrine-dopamine reuptake inhibitors (SNDRI) are widely used antidepressants [158]. Furthermore, antidepressants that inhibit both norepinephrine and serotonin transporters have been developed. It has been suggested that these SNRIs have improved antidepressant efficacy and faster onset of action in comparison to SSRIs. Among these SNRIs are duloxetine and venlafaxine that increase 5-HT and NE levels specifically in the prefrontal cortex area of the brain as well as the dopamine levels. Selective norepinephrine reuptake inhibitors (NRIs) such as reboxetine and nisoxetine are also in use for depression treatment and attention deficit hyperactivity disorder (ADHD). Another group of antidepressant are dual NET/DAT inhibitors (e.g., nomifensin), while triple reuptake inhibitors (SERT/DAT/NET inhibitors, e.g., indatraline and mazindol) are being examined for their efficacy in depression and other CNS-disorders [158].

3.1. Serotonin Transporter

The SERT protein plays an important role in the serotoninergic system. The highest levels of this transporter are found in the raphe nuclei and the cerebellum, basal ganglia, and thalamus, followed by the hippocampus and the prefrontal cortex [159]. The first-generation of MAT inhibitors include the tricyclic antidepressants such as imipramine, clomipramine, desipramine, and amitriptyline [160]. They exhibit their antidepressant effects by increasing synaptic levels of 5-HT and NE via inhibition of SERT and NET. SSRIs are considered as the second generation of antidepressants, and block the reuptake of serotonin from the synaptic cleft. The most commonly prescribed SSRIs are paroxetine, fluoxetine, citalopram, and sertraline [161,162]. Despite their good safety profile and efficacy, the major disadvantage of SSRIs are their slow onset of action and side effects such as anxiety, sleep disturbances, and sexual dysfunction.
The recently reported X-ray structures of SERT in complex with (S)-citalopram, paroxetine, sertraline, or fluvoxamine now provides knowledge to help understand the structure activity relationship and selectivity of compounds relative to other monoamine transporters [157,163]. The cryo-EM structures of hSERT in complex with ibogaine in various conformational states were recently resolved [164]. The conformational dynamics of SERT transport function and inhibition is now extensively studied by using computational studies and hydrogen-deuterium exchange (HDX) experiments [165,166].

3.2. Dopamine Transporter

DAT belongs to MAT and controls neurotransmitter dopamine homeostasis in the brain by reuptake of DA. DAT is widely distributed through the brain in areas of dopaminergic activity, including the striatum and substantial nigra [167]. Numerous abused as well as clinically important drugs have important pharmacological interactions with DAT. Bupropion is a dopamine reuptake inhibitor and also an inhibitor of the norepinephrine transporter [168].
However, the molecular mechanism of action of the dopamine transporter has been elusive due to the lack of X-ray structure. The first reported was the crystal structure of the Drosophila melanogaster dopamine transporter (dDAT) in complex with TCA nortriptyline, then the X-ray structure of dDAT in complex with nisoxetine and reboxetine was resolved [156,169]. Furthermore, another dDAT crystal structure in complex with its substrate DA, as well as psychostimulants cocaine and amphetamine, was published [170]. Computational studies based on the crystal structures gave insight into the mechanism of interaction of drugs with DAT [171]. In addition, the potential significance of orphenadrine (ORPH), an inhibitor of NET reuptake, as a repurposable hDAT-inhibitor, was highlighted [171]. Recently, computational methods including molecular docking simulation and pharmacokinetics study were used for the identification of novel DAT inhibitors [124].

3.3. Norepinephrine Transporter

NET is a MAT which mediates the uptake of norepinephrine and is a drug target in major depression [172]. Radioligand studies indicated that binding of [3H] nisoxetine in the locus coeruleus (LC) is characteristic of binding to NET [173]. In vivo studies using positron emission tomography (PET) have reported higher norepinephrine transporter availability in the thalamus and its sub-regions in patients with major depressive disorder [172]. Reboxetine and nisoxetine are potent and selective NET inhibitors which have been successfully developed to treat depression and attention deficit hyperactivity disorder (ADHD).
For the development of more active and/or selective compounds, the knowledge of the crystal structure of the NET is needed for understanding the interactions with its ligands. However, the crystal structure of hNET is not available yet. Recent studies on molecular mechanism of action using homology models of hNET indicated amino acid residues responsible for inhibitor binding and may help to understand the molecular interactions between NET and its inhibitors [174,175].

4. In Silico Methods in Drug Design

The advances in X-ray crystallography and cryo-EM methods have given new structural knowledge about the structure and function of GPCRs and MATs in the CNS and have given the opportunity of using target based (structure based) approaches for designing new compounds with a putative therapeutic value.
The basis for drug design is studying the action mechanism of a protein by determination of the correct binding conformation of small molecule ligands in the protein. Molecular docking is one of the most commonly used methods for predicting the conformation of small-molecule ligands within the appropriate target binding site with a high degree of accuracy [176]. Docking can provide theoretical calculations for target-ligand binding conformation and binding affinity scores, making it useful for both initial hit compound screening and computational analysis of lead compound binding patterns.
Understanding the structural dynamics and mechanisms of various GPCR signaling pathways is critical for the design of GPCR targeted ligands. GPCRs play critical roles in cellular signal transduction. The structural dynamics of GPCR signaling are supported by strong evidence that these receptors exist in multiple conformational states, contrary to their initial understanding as simple ON/OFF molecular switches [177]. Different agonists or allosteric modulators can stabilize different conformational states of the receptor, resulting in a signaling bias towards a specific G protein subtype or β-arrestin-mediated signaling pathways. Molecular dynamic simulations of GPCRs may demonstrate several “intermediate” conformational states that differ from the crystallographically observed active and inactive states. During substrate translocation, the monoamine transporters undergo conformational changes as well [178]. The established binding modes and behavior of neurotransmitters within MATs can greatly aid in structure-based development of novel drugs as well as in the ongoing optimization of existing drugs [179].
Virtual screening of large and chemically diverse compound libraries using computational methods is one of the most commonly used strategies in drug discovery [180]. Virtual screening approaches are divided into ligand-based and structure-based. The structure-based virtual screening (SBVS) method makes use of the target protein crystal structure, whereas the ligand-based virtual screening (LBVS) method makes use of the structural information and physicochemical properties of the chemical scaffold of known active and inactive molecules. Virtual screening techniques were employed to design small molecule modulators of monoamine transporters to orthosteric and allosteric sites [181]. The virtual screening approach can be used for drug repurposing (or repositioning) [182]. The drug repurposing is the methodology of developing new pharmacological indications for existing drugs and has been found as a cost-effective strategy. Among the now 13 reported drugs that have been repurposed for the treatment of depression or bipolar depression are ketamine, dextromethorphan, and scopolamine [183,184].
The main reasons for the high rates of drug failure in the later drug development stage are undesirable pharmacokinetics and toxicity of drug candidates. Pharmacokinetic properties such as absorption, distribution, metabolism, and elimination/excretion properties (ADME) as well as toxicity are important in the drug discovery process. Therefore, there is a need for predictive tools that can eliminate inappropriate compounds to save time and money. For ADMET parameters such as bioavailability, aqueous solubility, intestinal permeability, blood-brain barrier penetration, metabolism, and toxicity, several computational approaches are useful [185].

5. Conclusions

GPCRs and MATs are the most important present therapeutic targets for developing novel drugs for the CNS disorders. GPCRs adopt a large number of conformational states upon ligand binding that activate various signaling pathways. The most important is linking the activation of various signaling pathways with the pharmacological activity of the drug, thus investigating the role of individual signaling pathways in the pathogenesis of depression. MATs also undergo local conformational changes in response to ligand binding and releasing, such as the closing of a gate on one side while opening a gate on the other side, or by forming a transient occluded state by closing both ends. However, the mechanisms of depression are still not well understood, and thus studies on the mechanism of protein actions and signaling pathways activated in the CNS are needed and justified to find new selective antidepressants.

Author Contributions

Conceptualization, M.J., J.W.; original draft preparation, M.J.; review and editing, M.J. and J.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financed by internal funding from the National Medicines Institute (DS8/2021).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Azam, S.; Haque, M.E.; Jakaria, M.; Jo, S.-H.; Kim, I.-S.; Choi, D.-K. G-Protein-Coupled Receptors in CNS: A Potential Therapeutic Target for Intervention in Neurodegenerative Disorders and Associated Cognitive Deficits. Cells 2020, 9, 506. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Mulinari, S. Monoamine Theories of Depression: Historical Impact on Biomedical Research. J. Hist. Neurosci. 2012, 21, 366–392. [Google Scholar] [CrossRef] [PubMed]
  3. Blier, P. Neurotransmitter Targeting in the Treatment of Depression. J. Clin. Psychiatry 2013, 74, 19–24. [Google Scholar] [CrossRef] [PubMed]
  4. Delgado, P.L. Depression: The case for a monoamine deficiency. J. Clin. Psychiatry 2000, 61 (Suppl. 6), 7–11. [Google Scholar]
  5. Nedic Erjavec, G.; Sagud, M.; Nikolac Perkovic, M.; Svob Strac, D.; Konjevod, M.; Tudor, L.; Uzun, S.; Pivac, N. Depression: Biological markers and treatment. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2021, 105, 110139. [Google Scholar] [CrossRef]
  6. Cai, S.; Huang, S.; Hao, W. New hypothesis and treatment targets of depression: An integrated view of key fi ndings. Neurosci. Bull. 2015, 31, 61–74. [Google Scholar] [CrossRef] [Green Version]
  7. Maes, M.; Bosmans, E.; Suy, E.; Vandervorst, C.; De Jonckheere, C.; Raus, J. Immune Disturbances during Major Depression: Upregulated Expression of Interleukin-2 Receptors. Neuropsychobiology 1990, 24, 115–120. [Google Scholar] [CrossRef]
  8. Maes, M. Evidence for an immune response in major depression: A review and hypothesis. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 1995, 19, 11–38. [Google Scholar] [CrossRef]
  9. Deussing, J.M.; Arzt, E. P2X7 Receptor: A Potential Therapeutic Target for Depression? Trends Mol. Med. 2018, 24, 736–747. [Google Scholar] [CrossRef]
  10. Keller, J.; Gomez, R.; Williams, G.; Lembke, A.; Lazzeroni, L.; Murphy, G.M.; Schatzberg, A.F. HPA axis in major depression: Cortisol, clinical symptomatology and genetic variation predict cognition. Mol. Psychiatry 2017, 22, 527–536. [Google Scholar] [CrossRef]
  11. Castren, E.; Voikar, V.; Rantamaki, T. Role of neurotrophic factors in depression. Curr. Opin. Pharmacol. 2007, 7, 18–21. [Google Scholar] [CrossRef]
  12. Duman, R.S.; Li, N. A neurotrophic hypothesis of depression: Role of synaptogenesis in the actions of NMDA receptor antagonists. Philos. Trans. R. Soc. B Biol. Sci. 2012, 367, 2475–2484. [Google Scholar] [CrossRef] [Green Version]
  13. Mantas, I.; Saarinen, M.; Xu, Z.-Q.D.; Svenningsson, P. Update on GPCR-based targets for the development of novel antidepressants. Mol. Psychiatry 2021, 1–25. [Google Scholar] [CrossRef]
  14. Palczewski, K. Crystal Structure of Rhodopsin: A G Protein-Coupled Receptor. Science 2000, 289, 739–745. [Google Scholar] [CrossRef] [Green Version]
  15. Berman, H.M. The Protein Data Bank. Nucleic Acids Res. 2000, 28, 235–242. [Google Scholar] [CrossRef] [Green Version]
  16. Yang, D.; Zhou, Q.; Labroska, V.; Qin, S.; Darbalaei, S.; Wu, Y.; Yuliantie, E.; Xie, L.; Tao, H.; Cheng, J.; et al. G protein-coupled receptors: Structure- and function-based drug discovery. Signal Transduct. Target. Ther. 2021, 6, 7. [Google Scholar] [CrossRef]
  17. Hoyer, D.; Hannon, J.P.; Martin, G.R. Molecular, pharmacological and functional diversity of 5-HT receptors. Pharmacol. Biochem. Behav. 2002, 71, 533–554. [Google Scholar] [CrossRef]
  18. Savitz, J.; Lucki, I.; Drevets, W.C. 5-HT1A receptor function in major depressive disorder. Prog. Neurobiol. 2009, 88, 17–31. [Google Scholar] [CrossRef] [Green Version]
  19. Albert, P.R.; Vahid-Ansari, F. The 5-HT1A receptor: Signaling to behavior. Biochimie 2019, 161, 34–45. [Google Scholar] [CrossRef]
  20. Becker, G.; Bolbos, R.; Costes, N.; Redouté, J.; Newman-Tancredi, A.; Zimmer, L. Selective serotonin 5-HT1A receptor biased agonists elicitdistinct brain activation patterns: A pharmacoMRI study. Sci. Rep. 2016, 6, 26633. [Google Scholar] [CrossRef] [Green Version]
  21. Heiser, J.F.; Wilcox, C.S. Serotonin 5-HT1A Receptor Agonists as Antidepressants. CNS Drugs 1998, 10, 343–353. [Google Scholar] [CrossRef]
  22. Newman-Tancredi, A.; Martel, J.-C.; Assié, M.-B.; Buritova, J.; Lauressergues, E.; Cosi, C.; Heusler, P.; Bruins Slot, L.; Colpaert, F.; Vacher, B.; et al. Signal transduction and functional selectivity of F15599, a preferential post-synaptic 5-HT 1A receptor agonist. Br. J. Pharmacol. 2009, 156, 338–353. [Google Scholar] [CrossRef] [Green Version]
  23. Głuch-Lutwin, M.; Sałaciak, K.; Gawalska, A.; Jamrozik, M.; Sniecikowska, J.; Newman-Tancredi, A.; Kołaczkowski, M.; Pytka, K. The selective 5-HT1A receptor biased agonists, F15599 and F13714, show antidepressant-like properties after a single administration in the mouse model of unpredictable chronic mild stress. Psychopharmacology 2021, 238, 2249–2260. [Google Scholar] [CrossRef]
  24. Cherezov, V.; Rosenbaum, D.M.; Hanson, M.A.; Rasmussen, S.G.F.; Thian, F.S.; Kobilka, T.S.; Choi, H.-J.; Kuhn, P.; Weis, W.I.; Kobilka, B.K.; et al. High-Resolution Crystal Structure of an Engineered Human 2-Adrenergic G Protein-Coupled Receptor. Science 2007, 318, 1258–1265. [Google Scholar] [CrossRef] [Green Version]
  25. Warne, T.; Serrano-Vega, M.J.; Baker, J.G.; Moukhametzianov, R.; Edwards, P.C.; Henderson, R.; Leslie, A.G.W.; Tate, C.G.; Schertler, G.F.X. Structure of a β1-adrenergic G-protein-coupled receptor. Nature 2008, 454, 486–491. [Google Scholar] [CrossRef] [Green Version]
  26. Jaakola, V.-P.; Griffith, M.T.; Hanson, M.A.; Cherezov, V.; Chien, E.Y.T.; Lane, J.R.; IJzerman, A.P.; Stevens, R.C. The 2.6 Angstrom Crystal Structure of a Human A2A Adenosine Receptor Bound to an Antagonist. Science 2008, 322, 1211–1217. [Google Scholar] [CrossRef] [Green Version]
  27. Wang, C.; Jiang, Y.; Ma, J.; Wu, H.; Wacker, D.; Katritch, V.; Han, G.W.; Liu, W.; Huang, X.-P.; Vardy, E.; et al. Structural Basis for Molecular Recognition at Serotonin Receptors. Science 2013, 340, 610–614. [Google Scholar] [CrossRef] [Green Version]
  28. Xu, P.; Huang, S.; Zhang, H.; Mao, C.; Zhou, X.E.; Cheng, X.; Simon, I.A.; Shen, D.-D.; Yen, H.-Y.; Robinson, C.V.; et al. Structural insights into the lipid and ligand regulation of serotonin receptors. Nature 2021, 592, 469–473. [Google Scholar] [CrossRef]
  29. Dunlop, B.W.; Nemeroff, C.B. The Role of Dopamine in the Pathophysiology of Depression. Arch. Gen. Psychiatry 2007, 64, 327. [Google Scholar] [CrossRef]
  30. Gershon, A.A.; Vishne, T.; Grunhaus, L. Dopamine D2-Like Receptors and the Antidepressant Response. Biol. Psychiatry 2007, 61, 145–153. [Google Scholar] [CrossRef]
  31. Larisch, R.; Klimke, A.; Vosberg, H.; Löffler, S.; Gaebel, W.; Müller-Gärtner, H.-W. In VivoEvidence for the Involvement of Dopamine-D2Receptors in Striatum and Anterior Cingulate Gyrus in Major Depression. Neuroimage 1997, 5, 251–260. [Google Scholar] [CrossRef] [PubMed]
  32. Berman, R. Augmentation treatment in major depressive disorder: Focus on aripiprazole. Neuropsychiatr. Dis. Treat. 2008, 4, 937. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Citrome, L. Brexpiprazole: A new dopamine D2 receptor partial agonist for the treatment of schizophrenia and major depressive disorder. Drugs Today 2015, 51, 397. [Google Scholar] [CrossRef] [PubMed]
  34. Veselinović, T.; Paulzen, M.; Gründer, G. Cariprazine, a new, orally active dopamine D 2/3 receptor partial agonist for the treatment of schizophrenia, bipolar mania and depression. Expert Rev. Neurother. 2013, 13, 1141–1159. [Google Scholar] [CrossRef]
  35. Duric, V.; Banasr, M.; Franklin, T.; Lepack, A.; Adham, N.; Kiss, B.; Gyertyán, I.; Duman, R.S. Cariprazine Exhibits Anxiolytic and Dopamine D3 Receptor-Dependent Antidepressant Effects in the Chronic Stress Model. Int. J. Neuropsychopharmacol. 2017, 20, 788–796. [Google Scholar] [CrossRef] [Green Version]
  36. Chien, E.Y.T.; Liu, W.; Zhao, Q.; Katritch, V.; Won Han, G.; Hanson, M.A.; Shi, L.; Newman, A.H.; Javitch, J.A.; Cherezov, V.; et al. Structure of the Human Dopamine D3 Receptor in Complex with a D2/D3 Selective Antagonist. Science 2010, 330, 1091–1095. [Google Scholar] [CrossRef] [Green Version]
  37. Wang, S.; Che, T.; Levit, A.; Shoichet, B.K.; Wacker, D.; Roth, B.L. Structure of the D2 dopamine receptor bound to the atypical antipsychotic drug risperidone. Nature 2018, 555, 269–273. [Google Scholar] [CrossRef]
  38. Elek, M.; Djokovic, N.; Frank, A.; Oljacic, S.; Zivkovic, A.; Nikolic, K.; Stark, H. Synthesis, in silico, and in vitro studies of novel dopamine D 2 and D 3 receptor ligands. Arch. Pharm. 2021, 354, 2000486. [Google Scholar] [CrossRef]
  39. Browne, C.A.; Lucki, I. Targeting opioid dysregulation in depression for the development of novel therapeutics. Pharmacol. Ther. 2019, 201, 51–76. [Google Scholar] [CrossRef]
  40. Browne, C.A.; Jacobson, M.L.; Lucki, I. Novel Targets to Treat Depression: Opioid-Based Therapeutics. Harv. Rev. Psychiatry 2020, 28, 40–59. [Google Scholar] [CrossRef]
  41. Browne, C.A.; Wulf, H.; Lucki, I. Kappa Opioid Receptors in the Pathology and Treatment of Major Depressive Disorder; Springer: Cham, Switzerland, 2021; Volume 271, pp. 493–524. [Google Scholar] [CrossRef]
  42. Lutz, P.-E.; Kieffer, B.L. Opioid receptors: Distinct roles in mood disorders. Trends Neurosci. 2013, 36, 195–206. [Google Scholar] [CrossRef] [Green Version]
  43. Emrich, H.; Vogt, P.; Herz, A.; Kissling, W. Antidepressant Effects of Buprenorphine. Lancet 1982, 320, 709. [Google Scholar] [CrossRef]
  44. Falcon, E.; Browne, C.A.; Leon, R.M.; Fleites, V.C.; Sweeney, R.; Kirby, L.G.; Lucki, I. Antidepressant-like Effects of Buprenorphine are Mediated by Kappa Opioid Receptors. Neuropsychopharmacology 2016, 41, 2344–2351. [Google Scholar] [CrossRef]
  45. Browne, C.A.; Smith, T.; Lucki, I. Behavioral effects of the kappa opioid receptor partial agonist nalmefene in tests relevant to depression. Eur. J. Pharmacol. 2020, 872, 172948. [Google Scholar] [CrossRef]
  46. Samuels, B.A.; Nautiyal, K.M.; Kruegel, A.C.; Levinstein, M.R.; Magalong, V.M.; Gassaway, M.M.; Grinnell, S.G.; Han, J.; Ansonoff, M.A.; Pintar, J.E.; et al. The Behavioral Effects of the Antidepressant Tianeptine Require the Mu-Opioid Receptor. Neuropsychopharmacology 2017, 42, 2052–2063. [Google Scholar] [CrossRef] [Green Version]
  47. Gavioli, E.C.; Holanda, V.A.D.; Ruzza, C. NOP Ligands for the Treatment of Anxiety and Mood Disorders; In The Nociceptin/Orphanin FQ Peptide Receptor; Springer: Cham, Switzerland, 2018; pp. 233–257. [Google Scholar] [CrossRef]
  48. Witkin, J.M.; Wallace, T.L.; Martin, W.J. Therapeutic Approaches for NOP Receptor Antagonists in Neurobehavioral Disorders: Clinical Studies in Major Depressive Disorder and Alcohol Use Disorder with BTRX-246040 (LY2940094). In The Nociceptin/Orphanin FQ Peptide Receptor; Springer: Cham, Switzerland, 2018; Volume 254, pp. 399–415. [Google Scholar] [CrossRef]
  49. Post, A.; Smart, T.S.; Krikke-Workel, J.; Dawson, G.R.; Harmer, C.J.; Browning, M.; Jackson, K.; Kakar, R.; Mohs, R.; Statnick, M.; et al. A Selective Nociceptin Receptor Antagonist to Treat Depression: Evidence from Preclinical and Clinical Studies. Neuropsychopharmacology 2016, 41, 1803–1812. [Google Scholar] [CrossRef] [Green Version]
  50. Manglik, A.; Kruse, A.C.; Kobilka, T.S.; Thian, F.S.; Mathiesen, J.M.; Sunahara, R.K.; Pardo, L.; Weis, W.I.; Kobilka, B.K.; Granier, S. Crystal structure of the µ-opioid receptor bound to a morphinan antagonist. Nature 2012, 485, 321–326. [Google Scholar] [CrossRef] [Green Version]
  51. Huang, W.; Manglik, A.; Venkatakrishnan, A.J.; Laeremans, T.; Feinberg, E.N.; Sanborn, A.L.; Kato, H.E.; Livingston, K.E.; Thorsen, T.S.; Kling, R.C.; et al. Structural insights into µ-opioid receptor activation. Nature 2015, 524, 315–321. [Google Scholar] [CrossRef] [Green Version]
  52. Granier, S.; Manglik, A.; Kruse, A.C.; Kobilka, T.S.; Thian, F.S.; Weis, W.I.; Kobilka, B.K. Structure of the δ-opioid receptor bound to naltrindole. Nature 2012, 485, 400–404. [Google Scholar] [CrossRef] [Green Version]
  53. Wu, H.; Wacker, D.; Mileni, M.; Katritch, V.; Han, G.W.; Vardy, E.; Liu, W.; Thompson, A.A.; Huang, X.-P.; Carroll, F.I.; et al. Structure of the human κ-opioid receptor in complex with JDTic. Nature 2012, 485, 327–332. [Google Scholar] [CrossRef]
  54. Thompson, A.A.; Liu, W.; Chun, E.; Katritch, V.; Wu, H.; Vardy, E.; Huang, X.-P.; Trapella, C.; Guerrini, R.; Calo, G.; et al. Structure of the nociceptin/orphanin FQ receptor in complex with a peptide mimetic. Nature 2012, 485, 395–399. [Google Scholar] [CrossRef]
  55. Miller, R.L.; Thompson, A.A.; Trapella, C.; Guerrini, R.; Malfacini, D.; Patel, N.; Han, G.W.; Cherezov, V.; Caló, G.; Katritch, V.; et al. The Importance of Ligand-Receptor Conformational Pairs in Stabilization: Spotlight on the N/OFQ G Protein-Coupled Receptor. Structure 2015, 23, 2291–2299. [Google Scholar] [CrossRef] [Green Version]
  56. Mafi, A.; Kim, S.-K.; Goddard, W.A. The atomistic level structure for the activated human κ-opioid receptor bound to the full Gi protein and the MP1104 agonist. Proc. Natl. Acad. Sci. USA 2020, 117, 5836–5843. [Google Scholar] [CrossRef] [Green Version]
  57. Longa, S.D.; Arcovito, A. “In silico” study of the binding of two novel antagonists to the nociceptin receptor. J. Comput. Aided Mol. Des. 2018, 32, 385–400. [Google Scholar] [CrossRef]
  58. Mathews, D.C.; Henter, I.D.; Zarate, C.A. Targeting the Glutamatergic System to Treat Major Depressive Disorder. Drugs 2012, 72, 1313–1333. [Google Scholar] [CrossRef]
  59. Niciu, M.J.; Ionescu, D.F.; Richards, E.M.; Zarate, C.A. Glutamate and its receptors in the pathophysiology and treatment of major depressive disorder. J. Neural Transm. 2014, 121, 907–924. [Google Scholar] [CrossRef]
  60. Wright, R.A.; Johnson, B.G.; Zhang, C.; Salhoff, C.; Kingston, A.E.; Calligaro, D.O.; Monn, J.A.; Schoepp, D.D.; Marek, G.J. CNS distribution of metabotropic glutamate 2 and 3 receptors: Transgenic mice and [3H]LY459477 autoradiography. Neuropharmacology 2013, 66, 89–98. [Google Scholar] [CrossRef]
  61. Chaki, S. mGlu2/3 Receptor Antagonists as Novel Antidepressants. Trends Pharmacol. Sci. 2017, 38, 569–580. [Google Scholar] [CrossRef]
  62. Musazzi, L. Targeting metabotropic glutamate receptors for rapid-acting antidepressant drug discovery. Expert Opin. Drug Discov. 2021, 16, 147–157. [Google Scholar] [CrossRef]
  63. Campo, B.; Kalinichev, M.; Lambeng, N.; El Yacoubi, M.; Royer-Urios, I.; Schneider, M.; Legrand, C.; Parron, D.; Girard, F.; Bessif, A.; et al. Characterization of an mGluR2/3 Negative Allosteric Modulator in Rodent Models of Depression. J. Neurogenet. 2011, 25, 152–166. [Google Scholar] [CrossRef]
  64. Wu, H.; Wang, C.; Gregory, K.J.; Han, G.W.; Cho, H.P.; Xia, Y.; Niswender, C.M.; Katritch, V.; Meiler, J.; Cherezov, V.; et al. Structure of a Class C GPCR Metabotropic Glutamate Receptor 1 Bound to an Allosteric Modulator. Science 2014, 344, 58–64. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Doré, A.S.; Okrasa, K.; Patel, J.C.; Serrano-Vega, M.; Bennett, K.; Cooke, R.M.; Errey, J.C.; Jazayeri, A.; Khan, S.; Tehan, B.; et al. Structure of class C GPCR metabotropic glutamate receptor 5 transmembrane domain. Nature 2014, 511, 557–562. [Google Scholar] [CrossRef] [PubMed]
  66. Pérez-Benito, L.; Doornbos, M.L.J.; Cordomí, A.; Peeters, L.; Lavreysen, H.; Pardo, L.; Tresadern, G. Molecular Switches of Allosteric Modulation of the Metabotropic Glutamate 2 Receptor. Structure 2017, 25, 1153–1162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Watkins, L.R.; Orlandi, C. Orphan G Protein Coupled Receptors in Affective Disorders. Genes 2020, 11, 694. [Google Scholar] [CrossRef] [PubMed]
  68. Albert, P.R. Orphans to the rescue: Orphan G-protein coupled receptors as new antidepressant targets. J. Psychiatry Neurosci. 2020, 45, 301–303. [Google Scholar] [CrossRef]
  69. Kooistra, A.J.; Mordalski, S.; Pándy-Szekeres, G.; Esguerra, M.; Mamyrbekov, A.; Munk, C.; Keserű, G.M.; Gloriam, D.E. GPCRdb in 2021: Integrating GPCR sequence, structure and function. Nucleic Acids Res. 2021, 49, D335–D343. [Google Scholar] [CrossRef]
  70. Jones, P.; Nawoschik, S.; Sreekumar, K.; Uveges, A.; Tseng, E.; Zhang, L.; Johnson, J.; He, L.; Paulsen, J.; Bates, B. Tissue distribution and functional analyses of the constitutively active orphan G protein coupled receptors, GPR26 and GPR78. Biochim. Biophys. Acta Gen. Subj. 2007, 1770, 890–901. [Google Scholar] [CrossRef]
  71. Zhang, L.-L.; Wang, J.-J.; Liu, Y.; Lu, X.-B.; Kuang, Y.; Wan, Y.-H.; Chen, Y.; Yan, H.-M.; Fei, J.; Wang, Z.-G. GPR26-deficient mice display increased anxiety- and depression-like behaviors accompanied by reduced phosphorylated cyclic AMP responsive element-binding protein level in central amygdala. Neuroscience 2011, 196, 203–214. [Google Scholar] [CrossRef]
  72. Belzeaux, R.; Gorgievski, V.; Fiori, L.M.; Lopez, J.P.; Grenier, J.; Lin, R.; Nagy, C.; Ibrahim, E.C.; Gascon, E.; Courtet, P.; et al. GPR56/ADGRG1 is associated with response to antidepressant treatment. Nat. Commun. 2020, 11, 1635. [Google Scholar] [CrossRef]
  73. Sutton, L.P.; Orlandi, C.; Song, C.; Oh, W.C.; Muntean, B.S.; Xie, K.; Filippini, A.; Xie, X.; Satterfield, R.; Yaeger, J.D.W.; et al. Orphan receptor GPR158 controls stress-induced depression. eLife 2018, 7, e33273. [Google Scholar] [CrossRef]
  74. Wolf, M.E.; Mosnaim, A.D. Phenylethylamine in neuropsychiatric disorders. Gen. Pharmacol. Vasc. Syst. 1983, 14, 385–390. [Google Scholar] [CrossRef]
  75. Sabelli, H.; Fink, P.; Fawcett, J.; Tom, C. Sustained antidepressant effect of PEA replacement. J. Neuropsychiatry Clin. Neurosci. 1996, 8, 168–171. [Google Scholar] [CrossRef]
  76. Pei, Y.; Asif-Malik, A.; Canales, J.J. Trace Amines and the Trace Amine-Associated Receptor 1: Pharmacology, Neurochemistry, and Clinical Implications. Front. Neurosci. 2016, 10, 148. [Google Scholar] [CrossRef] [Green Version]
  77. Schwartz, M.D.; Canales, J.J.; Zucchi, R.; Espinoza, S.; Sukhanov, I.; Gainetdinov, R.R. Trace amine-associated receptor 1: A multimodal therapeutic target for neuropsychiatric diseases. Expert Opin. Ther. Targets 2018, 22, 513–526. [Google Scholar] [CrossRef]
  78. Borowsky, B.; Adham, N.; Jones, K.A.; Raddatz, R.; Artymyshyn, R.; Ogozalek, K.L.; Durkin, M.M.; Lakhlani, P.P.; Bonini, J.A.; Pathirana, S.; et al. Trace amines: Identification of a family of mammalian G protein-coupled receptors. Proc. Natl. Acad. Sci. USA 2001, 98, 8966–8971. [Google Scholar] [CrossRef] [Green Version]
  79. Lindemann, L.; Ebeling, M.; Kratochwil, N.A.; Bunzow, J.R.; Grandy, D.K.; Hoener, M.C. Trace amine-associated receptors form structurally and functionally distinct subfamilies of novel G protein-coupled receptors. Genomics 2005, 85, 372–385. [Google Scholar] [CrossRef]
  80. Revel, F.G.; Moreau, J.-L.; Gainetdinov, R.R.; Ferragud, A.; Velázquez-Sánchez, C.; Sotnikova, T.D.; Morairty, S.R.; Harmeier, A.; Groebke Zbinden, K.; Norcross, R.D.; et al. Trace Amine-Associated Receptor 1 Partial Agonism Reveals Novel Paradigm for Neuropsychiatric Therapeutics. Biol. Psychiatry 2012, 72, 934–942. [Google Scholar] [CrossRef]
  81. Espinoza, S.; Leo, D.; Sotnikova, T.D.; Shahid, M.; Kääriäinen, T.M.; Gainetdinov, R.R. Biochemical and Functional Characterization of the Trace Amine-Associated Receptor 1 (TAAR1) Agonist RO5263397. Front. Pharmacol. 2018, 9, 645. [Google Scholar] [CrossRef]
  82. Zeng, Z.; Fan, P.; Rand, E.; Kyaw, H.; Su, K.; Madike, V.; Carter, K.C.; Li, Y. Cloning of a Putative Human Neurotransmitter Receptor Expressed in Skeletal Muscle and Brain. Biochem. Biophys. Res. Commun. 1998, 242, 575–578. [Google Scholar] [CrossRef]
  83. Espinoza, S.; Sukhanov, I.; Efimova, E.V.; Kozlova, A.; Antonova, K.A.; Illiano, P.; Leo, D.; Merkulyeva, N.; Kalinina, D.; Musienko, P.; et al. Trace Amine-Associated Receptor 5 Provides Olfactory Input Into Limbic Brain Areas and Modulates Emotional Behaviors and Serotonin Transmission. Front. Mol. Neurosci. 2020, 13, 18. [Google Scholar] [CrossRef]
  84. Efimova, E.V.; Kozlova, A.A.; Razenkova, V.; Katolikova, N.V.; Antonova, K.A.; Sotnikova, T.D.; Merkulyeva, N.S.; Veshchitskii, A.S.; Kalinina, D.S.; Korzhevskii, D.E.; et al. Increased dopamine transmission and adult neurogenesis in trace amine-associated receptor 5 (TAAR5) knockout mice. Neuropharmacology 2021, 182, 108373. [Google Scholar] [CrossRef]
  85. Lam, V.M.; Rodríguez, D.; Zhang, T.; Koh, E.J.; Carlsson, J.; Salahpour, A. Discovery of trace amine-associated receptor 1 ligands by molecular docking screening against a homology model. MedChemComm 2015, 6, 2216–2223. [Google Scholar] [CrossRef]
  86. Cichero, E.; Espinoza, S.; Gainetdinov, R.R.; Brasili, L.; Fossa, P. Insights into the Structure and Pharmacology of the Human Trace Amine-Associated Receptor 1 (hTAAR1): Homology Modelling and Docking Studies. Chem. Biol. Drug Des. 2013, 81, 509–516. [Google Scholar] [CrossRef]
  87. Cichero, E.; Espinoza, S.; Tonelli, M.; Franchini, S.; Gerasimov, A.S.; Sorbi, C.; Gainetdinov, R.R.; Brasili, L.; Fossa, P. A homology modelling-driven study leading to the discovery of the first mouse trace amine-associated receptor 5 (TAAR5) antagonists. MedChemComm 2016, 7, 353–364. [Google Scholar] [CrossRef]
  88. Bambico, F.R.; Katz, N.; Debonnel, G.; Gobbi, G. Cannabinoids Elicit Antidepressant-Like Behavior and Activate Serotonergic Neurons through the Medial Prefrontal Cortex. J. Neurosci. 2007, 27, 11700–11711. [Google Scholar] [CrossRef]
  89. Hungund, B.L.; Vinod, K.Y.; Kassir, S.A.; Basavarajappa, B.S.; Yalamanchili, R.; Cooper, T.B.; Mann, J.J.; Arango, V. Upregulation of CB1 receptors and agonist-stimulated [35S]GTPγS binding in the prefrontal cortex of depressed suicide victims. Mol. Psychiatry 2004, 9, 184–190. [Google Scholar] [CrossRef]
  90. Liu, Q.-R.; Canseco-Alba, A.; Zhang, H.-Y.; Tagliaferro, P.; Chung, M.; Dennis, E.; Sanabria, B.; Schanz, N.; Escosteguy-Neto, J.C.; Ishiguro, H.; et al. Cannabinoid type 2 receptors in dopamine neurons inhibits psychomotor behaviors, alters anxiety, depression and alcohol preference. Sci. Rep. 2017, 7, 17410. [Google Scholar] [CrossRef]
  91. Onaivi, E.S.; Ishiguro, H.; Gong, J.-P.; Patel, S.; Meozzi, P.A.; Myers, L.; Perchuk, A.; Mora, Z.; Tagliaferro, P.A.; Gardner, E.; et al. Brain Neuronal CB2 Cannabinoid Receptors in Drug Abuse and Depression: From Mice to Human Subjects. PLoS ONE 2008, 3, e1640. [Google Scholar] [CrossRef] [Green Version]
  92. Howlett, A.C. International Union of Pharmacology. XXVII. Classification of Cannabinoid Receptors. Pharmacol. Rev. 2002, 54, 161–202. [Google Scholar] [CrossRef]
  93. Gong, J.-P.; Onaivi, E.S.; Ishiguro, H.; Liu, Q.-R.; Tagliaferro, P.A.; Brusco, A.; Uhl, G.R. Cannabinoid CB2 receptors: Immunohistochemical localization in rat brain. Brain Res. 2006, 1071, 10–23. [Google Scholar] [CrossRef]
  94. Howlett, A.C.; Qualy, J.M.; Khachatrian, L.L. Involvement of Gi in the inhibition of adenylate cyclase by cannabimimetic drugs. Mol. Pharmacol. 1986, 29, 307–313. [Google Scholar] [PubMed]
  95. Paudel, P.; Ross, S.; Li, X.-C. Molecular Targets of Cannabinoids Associated with Depression. Curr. Med. Chem. 2021, 28. [Google Scholar] [CrossRef] [PubMed]
  96. ElBatsh, M.M.; Moklas, M.A.A.; Marsden, C.A.; Kendall, D.A. Antidepressant-like effects of Δ9-tetrahydrocannabinol and rimonabant in the olfactory bulbectomised rat model of depression. Pharmacol. Biochem. Behav. 2012, 102, 357–365. [Google Scholar] [CrossRef] [PubMed]
  97. Bahi, A.; Al Mansouri, S.; Al Memari, E.; Al Ameri, M.; Nurulain, S.M.; Ojha, S. β-Caryophyllene, a CB2 receptor agonist produces multiple behavioral changes relevant to anxiety and depression in mice. Physiol. Behav. 2014, 135, 119–124. [Google Scholar] [CrossRef]
  98. Hua, T.; Vemuri, K.; Nikas, S.P.; Laprairie, R.B.; Wu, Y.; Qu, L.; Pu, M.; Korde, A.; Jiang, S.; Ho, J.-H.; et al. Crystal structures of agonist-bound human cannabinoid receptor CB1. Nature 2017, 547, 468–471. [Google Scholar] [CrossRef]
  99. Krishna Kumar, K.; Shalev-Benami, M.; Robertson, M.J.; Hu, H.; Banister, S.D.; Hollingsworth, S.A.; Latorraca, N.R.; Kato, H.E.; Hilger, D.; Maeda, S.; et al. Structure of a Signaling Cannabinoid Receptor 1-G Protein Complex. Cell 2019, 176, 448–458. [Google Scholar] [CrossRef] [Green Version]
  100. Li, X.; Hua, T.; Vemuri, K.; Ho, J.-H.; Wu, Y.; Wu, L.; Popov, P.; Benchama, O.; Zvonok, N.; Locke, K.; et al. Crystal Structure of the Human Cannabinoid Receptor CB2. Cell 2019, 176, 459–467. [Google Scholar] [CrossRef] [Green Version]
  101. Vieira, G.; Cavalli, J.; Gonçalves, E.C.D.; Braga, S.F.P.; Ferreira, R.S.; Santos, A.R.S.; Cola, M.; Raposo, N.R.B.; Capasso, R.; Dutra, R.C. Antidepressant-Like Effect of Terpineol in an Inflammatory Model of Depression: Involvement of the Cannabinoid System and D2 Dopamine Receptor. Biomolecules 2020, 10, 792. [Google Scholar] [CrossRef]
  102. Janowsky, D.; Davis, J.; El-Yousef, M.K.; Sekerke, H.J. A cholinergic-adrenergic hypothesis of mania and depression. Lancet 1972, 300, 632–635. [Google Scholar] [CrossRef]
  103. Dilsaver, S.C. Cholinergic mechanisms in depression. Brain Res. Rev. 1986, 11, 285–316. [Google Scholar] [CrossRef] [Green Version]
  104. Dulawa, S.C.; Janowsky, D.S. Cholinergic regulation of mood: From basic and clinical studies to emerging therapeutics. Mol. Psychiatry 2019, 24, 694–709. [Google Scholar] [CrossRef]
  105. Furey, M.L.; Drevets, W.C. Antidepressant Efficacy of the Antimuscarinic Drug Scopolamine. Arch. Gen. Psychiatry 2006, 63, 1121. [Google Scholar] [CrossRef]
  106. Navarria, A.; Wohleb, E.S.; Voleti, B.; Ota, K.T.; Dutheil, S.; Lepack, A.E.; Dwyer, J.M.; Fuchikami, M.; Becker, A.; Drago, F.; et al. Rapid antidepressant actions of scopolamine: Role of medial prefrontal cortex and M1-subtype muscarinic acetylcholine receptors. Neurobiol. Dis. 2015, 82, 254–261. [Google Scholar] [CrossRef] [Green Version]
  107. Liu, S.; Shi, D.; Sun, Z.; He, Y.; Yang, J.; Wang, G. M2-AChR Mediates Rapid Antidepressant Effects of Scopolamine Through Activating the mTORC1-BDNF Signaling Pathway in the Medial Prefrontal Cortex. Front. Psychiatry 2021, 12, 601985. [Google Scholar] [CrossRef]
  108. Thal, D.M.; Sun, B.; Feng, D.; Nawaratne, V.; Leach, K.; Felder, C.C.; Bures, M.G.; Evans, D.A.; Weis, W.I.; Bachhawat, P.; et al. Crystal structures of the M1 and M4 muscarinic acetylcholine receptors. Nature 2016, 531, 335–340. [Google Scholar] [CrossRef] [Green Version]
  109. Haga, K.; Kruse, A.C.; Asada, H.; Yurugi-Kobayashi, T.; Shiroishi, M.; Zhang, C.; Weis, W.I.; Okada, T.; Kobilka, B.K.; Haga, T.; et al. Structure of the human M2 muscarinic acetylcholine receptor bound to an antagonist. Nature 2012, 482, 547–551. [Google Scholar] [CrossRef] [Green Version]
  110. Kruse, A.C.; Hu, J.; Pan, A.C.; Arlow, D.H.; Rosenbaum, D.M.; Rosemond, E.; Green, H.F.; Liu, T.; Chae, P.S.; Dror, R.O.; et al. Structure and dynamics of the M3 muscarinic acetylcholine receptor. Nature 2012, 482, 552–556. [Google Scholar] [CrossRef] [Green Version]
  111. Lloyd, K.G.; Zivkovic, B.; Scatton, B.; Morselli, P.L.; Bartholimi, G. The gabaergic hypothesis of depression. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 1989, 13, 341–351. [Google Scholar] [CrossRef]
  112. Mombereau, C.; Kaupmann, K.; Froestl, W.; Sansig, G.; van der Putten, H.; Cryan, J.F. Genetic and Pharmacological Evidence of a Role for GABAB Receptors in the Modulation of Anxiety- and Antidepressant-Like Behavior. Neuropsychopharmacology 2004, 29, 1050–1062. [Google Scholar] [CrossRef]
  113. Drake, R.G.; Davis, L.L.; Cates, M.E.; Jewell, M.E.; Ambrose, S.M.; Lowe, J.S. Baclofen Treatment for Chronic Posttraumatic Stress Disorder. Ann. Pharmacother. 2003, 37, 1177–1181. [Google Scholar] [CrossRef]
  114. Nowak, G.; Partyka, A.; Pałucha, A.; Szewczyk, B.; Wierońska, J.M.; Dybała, M.; Metz, M.; Librowski, T.; Froestl, W.; Papp, M.; et al. Antidepressant-like activity of CGP 36742 and CGP 51176, selective GABA B receptor antagonists, in rodents. Br. J. Pharmacol. 2006, 149, 581–590. [Google Scholar] [CrossRef] [Green Version]
  115. Urwyler, S.; Gjoni, T.; Koljatic, J.; Dupuis, D. Mechanisms of allosteric modulation at GABA receptors by CGP7930 and GS39783: Effects on affinities and efficacies of orthosteric ligands with distinct intrinsic properties. Neuropharmacology 2005, 48, 343–353. [Google Scholar] [CrossRef]
  116. Evenseth, L.S.M.; Gabrielsen, M.; Sylte, I. The GABAB Receptor—Structure, Ligand Binding and Drug Development. Molecules 2020, 25, 3093. [Google Scholar] [CrossRef]
  117. Mao, C.; Shen, C.; Li, C.; Shen, D.-D.; Xu, C.; Zhang, S.; Zhou, R.; Shen, Q.; Chen, L.-N.; Jiang, Z.; et al. Cryo-EM structures of inactive and active GABAB receptor. Cell Res. 2020, 30, 564–573. [Google Scholar] [CrossRef]
  118. Shaye, H.; Ishchenko, A.; Lam, J.H.; Han, G.W.; Xue, L.; Rondard, P.; Pin, J.-P.; Katritch, V.; Gati, C.; Cherezov, V. Structural basis of the activation of a metabotropic GABA receptor. Nature 2020, 584, 298–303. [Google Scholar] [CrossRef]
  119. Papasergi-Scott, M.M.; Robertson, M.J.; Seven, A.B.; Panova, O.; Mathiesen, J.M.; Skiniotis, G. Structures of metabotropic GABAB receptor. Nature 2020, 584, 310–314. [Google Scholar] [CrossRef]
  120. Rigby, M.; O’Donnell, R.; Rupniak, N.M.J. Species differences in tachykinin receptor distribution: Further evidence that the substance P (NK1) receptor predominates in human brain. J. Comp. Neurol. 2005, 490, 335–353. [Google Scholar] [CrossRef]
  121. Kramer, M.S.; Cutler, N.; Feighner, J.; Shrivastava, R.; Carman, J.; Sramek, J.J.; Reines, S.A.; Liu, G.; Snavely, D.; Wyatt-Knowles, E.; et al. Distinct Mechanism for Antidepressant Activity by Blockade of Central Substance P Receptors. Science 1998, 281, 1640–1645. [Google Scholar] [CrossRef] [PubMed]
  122. Kramer, M.S.; Winokur, A.; Kelsey, J.; Preskorn, S.H.; Rothschild, A.J.; Snavely, D.; Ghosh, K.; Ball, W.A.; Reines, S.A.; Munjack, D.; et al. Demonstration of the Efficacy and Safety of a Novel Substance P (NK1) Receptor Antagonist in Major Depression. Neuropsychopharmacology 2004, 29, 385–392. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Yin, J.; Chapman, K.; Clark, L.D.; Shao, Z.; Borek, D.; Xu, Q.; Wang, J.; Rosenbaum, D.M. Crystal structure of the human NK 1 tachykinin receptor. Proc. Natl. Acad. Sci. USA 2018, 115, 13264–13269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Babatunde Olasupo, S.; Uzairu, A.; Shallangwa, G.A.; Uba, S. Unveiling novel inhibitors of dopamine transporter via in silico drug design, molecular docking, and bioavailability predictions as potential antischizophrenic agents. Future J. Pharm. Sci. 2021, 7, 63. [Google Scholar] [CrossRef]
  125. Morley, J.E. The ascent of cholecystokinin (CCK)—From gut to brain. Life Sci. 1982, 30, 479–493. [Google Scholar] [CrossRef]
  126. Beinfeld, M.C.; Meyer, D.K.; Eskay, R.L.; Jensen, R.T.; Brownstein, M.J. The distribution of cholecystokinin immunoreactivity in the central nervous system of the rat as determined by radioimmunoassay. Brain Res. 1981, 212, 51–57. [Google Scholar] [CrossRef]
  127. Beinfeld, M.C.; Meyer, D.K.; Brownstein, M.J. Cholecystokinin in the central nervous system. Peptides 1981, 2, 77–79. [Google Scholar] [CrossRef]
  128. Kõks, S.; Volke, V.; Veraksits, A.; Rünkorg, K.; Sillat, T.; Abramov, U.; Bourin, M.; Huotari, M.; Männistö, P.; Matsui, T.; et al. Cholecystokinin 2 receptor-deficient mice display altered function of brain dopaminergic system. Psychopharmacology 2001, 158, 198–204. [Google Scholar] [CrossRef]
  129. Derrien, M.; Durieux, C.; Roques, B.P. Antidepressant-like effects of CCKB antagonists in mice: Antagonism by naltrindole. Br. J. Pharmacol. 1994, 111, 956–960. [Google Scholar] [CrossRef] [Green Version]
  130. Becker, C.; Zeau, B.; Rivat, C.; Blugeot, A.; Hamon, M.; Benoliel, J.-J. Repeated social defeat-induced depression-like behavioral and biological alterations in rats: Involvement of cholecystokinin. Mol. Psychiatry 2008, 13, 1079–1092. [Google Scholar] [CrossRef] [Green Version]
  131. Liu, Q.; Yang, D.; Zhuang, Y.; Croll, T.I.; Cai, X.; Dai, A.; He, X.; Duan, J.; Yin, W.; Ye, C.; et al. Ligand recognition and G-protein coupling selectivity of cholecystokinin A receptor. Nat. Chem. Biol. 2021, 17, 1238–1244. [Google Scholar] [CrossRef]
  132. Zhang, X.; He, C.; Wang, M.; Zhou, Q.; Yang, D.; Zhu, Y.; Feng, W.; Zhang, H.; Dai, A.; Chu, X.; et al. Structures of the human cholecystokinin receptors bound to agonists and antagonists. Nat. Chem. Biol. 2021, 17, 1230–1237. [Google Scholar] [CrossRef]
  133. Senese, N.B.; Rasenick, M.M.; Traynor, J.R. The Role of G-proteins and G-protein Regulating Proteins in Depressive Disorders. Front. Pharmacol. 2018, 9, 1289. [Google Scholar] [CrossRef] [Green Version]
  134. Avissar, S. The role of G proteins in the psychobiology and treatment of affective disorders and their integration with the neurotransmitter hypothesis. Curr. Psychiatry Rep. 1999, 1, 148–153. [Google Scholar] [CrossRef] [PubMed]
  135. Avissar, S.; Schreiber, G. The involvement of guanine nucleotide binding proteins in the pathogenesis and treatment of affective disorders. Biol. Psychiatry 1992, 31, 435–459. [Google Scholar] [CrossRef]
  136. Chen, J.; Rasenick, M.M. Chronic antidepressant treatment facilitates G protein activation of adenylyl cyclase without altering G protein content. J. Pharmacol. Exp. Ther. 1995, 275, 509–517. [Google Scholar] [PubMed]
  137. Chen, J.; Rasenick, M.M. Chronic Treatment of C6 Glioma Cells with Antidepressant Drugs Increases Functional Coupling Between a G Protein (GS) and Adenylyl Cyclase. J. Neurochem. 2002, 64, 724–732. [Google Scholar] [CrossRef]
  138. O’Brien, J.B.; Wilkinson, J.C.; Roman, D.L. Regulator of G-protein signaling (RGS) proteins as drug targets: Progress and future potentials. J. Biol. Chem. 2019, 294, 18571–18585. [Google Scholar] [CrossRef] [Green Version]
  139. Muma, N.A. RGS proteins: Impact on the treatment of depression and anxiety. Int. J. Neuropsychopharmacol. 2012, 15, 1199–1200. [Google Scholar] [CrossRef] [Green Version]
  140. Ingi, T.; Aoki, Y. Expression of RGS2, RGS4 and RGS7 in the developing postnatal brain. Eur. J. Neurosci. 2002, 15, 929–936. [Google Scholar] [CrossRef]
  141. Nance, M.R.; Kreutz, B.; Tesmer, V.M.; Sterne-Marr, R.; Kozasa, T.; Tesmer, J.J.G. Structural and Functional Analysis of the Regulator of G Protein Signaling 2-Gαq Complex. Structure 2013, 21, 438–448. [Google Scholar] [CrossRef] [Green Version]
  142. Orlandi, C.; Sutton, L.P.; Muntean, B.S.; Song, C.; Martemyanov, K.A. Homeostatic cAMP regulation by the RGS7 complex controls depression-related behaviors. Neuropsychopharmacology 2019, 44, 642–653. [Google Scholar] [CrossRef]
  143. Sutton, L.P.; Khalatyan, N.; Savas, J.N.; Martemyanov, K.A. Striatal RGS7 Regulates Depression-Related Behaviors and Stress-Induced Reinstatement of Cocaine Conditioned Place Preference. eNeuro 2021, 8, e0365-20. [Google Scholar] [CrossRef]
  144. Patil, D.N.; Rangarajan, E.S.; Novick, S.J.; Pascal, B.D.; Kojetin, D.J.; Griffin, P.R.; Izard, T.; Martemyanov, K.A. Structural organization of a major neuronal G protein regulator, the RGS7-Gβ5-R7BP complex. eLife 2018, 7, e42150. [Google Scholar] [CrossRef]
  145. Kobayashi, Y.; Takemoto, R.; Yamato, S.; Okada, T.; Iijima, M.; Uematsu, Y.; Chaki, S.; Saito, Y. Depression-resistant Phenotype in Mice Overexpressing Regulator of G Protein Signaling 8 (RGS8). Neuroscience 2018, 383, 160–169. [Google Scholar] [CrossRef]
  146. Ahlers, K.E.; Chakravarti, B.; Fisher, R.A. RGS6 as a Novel Therapeutic Target in CNS Diseases and Cancer. AAPS J. 2016, 18, 560–572. [Google Scholar] [CrossRef] [Green Version]
  147. Aggarwal, S.; Mortensen, O.V. Overview of Monoamine Transporters. Curr. Protoc. Pharmacol. 2017, 79, 1–17. [Google Scholar] [CrossRef]
  148. Lin, L.; Yee, S.W.; Kim, R.B.; Giacomini, K.M. SLC transporters as therapeutic targets: Emerging opportunities. Nat. Rev. Drug Discov. 2015, 14, 543–560. [Google Scholar] [CrossRef] [Green Version]
  149. Cheng, M.H.; Bahar, I. Monoamine transporters: Structure, intrinsic dynamics and allosteric regulation. Nat. Struct. Mol. Biol. 2019, 26, 545–556. [Google Scholar] [CrossRef]
  150. Iversen, L. Neurotransmitter transporters and their impact on the development of psychopharmacology. Br. J. Pharmacol. 2006, 147, S82–S88. [Google Scholar] [CrossRef] [Green Version]
  151. Yamashita, A.; Singh, S.K.; Kawate, T.; Jin, Y.; Gouaux, E. Crystal structure of a bacterial homologue of Na+/Cl-dependent neurotransmitter transporters. Nature 2005, 437, 215–223. [Google Scholar] [CrossRef]
  152. Singh, S.K.; Yamashita, A.; Gouaux, E. Antidepressant binding site in a bacterial homologue of neurotransmitter transporters. Nature 2007, 448, 952–956. [Google Scholar] [CrossRef]
  153. Zhou, Z.; Zhen, J.; Karpowich, N.K.; Goetz, R.M.; Law, C.J.; Reith, M.E.A.; Wang, D.-N. LeuT-Desipramine Structure Reveals How Antidepressants Block Neurotransmitter Reuptake. Science 2007, 317, 1390–1393. [Google Scholar] [CrossRef] [Green Version]
  154. Zhou, Z.; Zhen, J.; Karpowich, N.K.; Law, C.J.; Reith, M.E.A.; Wang, D.-N. Antidepressant specificity of serotonin transporter suggested by three LeuT–SSRI structures. Nat. Struct. Mol. Biol. 2009, 16, 652–657. [Google Scholar] [CrossRef] [Green Version]
  155. Quick, M.; Winther, A.-M.L.; Shi, L.; Nissen, P.; Weinstein, H.; Javitch, J.A. Binding of an octylglucoside detergent molecule in the second substrate (S2) site of LeuT establishes an inhibitor-bound conformation. Proc. Natl. Acad. Sci. USA 2009, 106, 5563–5568. [Google Scholar] [CrossRef] [Green Version]
  156. Penmatsa, A.; Wang, K.H.; Gouaux, E. X-ray structure of dopamine transporter elucidates antidepressant mechanism. Nature 2013, 503, 85–90. [Google Scholar] [CrossRef] [Green Version]
  157. Coleman, J.A.; Green, E.M.; Gouaux, E. X-ray structures and mechanism of the human serotonin transporter. Nature 2016, 532, 334–339. [Google Scholar] [CrossRef] [Green Version]
  158. Xue, W.; Fu, T.; Zheng, G.; Tu, G.; Zhang, Y.; Yang, F.; Tao, L.; Yao, L.; Zhu, F. Recent Advances and Challenges of the Drugs Acting on Monoamine Transporters. Curr. Med. Chem. 2020, 27, 3830–3876. [Google Scholar] [CrossRef]
  159. Gurevich, E. Comparison of [3H]Paroxetine and [3H]Cyanoimipramine for Quantitative Measurement of Serotonin Transporter Sites in Human Brain. Neuropsychopharmacology 1996, 14, 309–323. [Google Scholar] [CrossRef] [Green Version]
  160. Stahl, M.; Lee-Zimmerman, C.; Cartwright, S.; Ann Morrissette, D. Serotonergic Drugs for Depression and Beyond. Curr. Drug Targets 2013, 14, 578–585. [Google Scholar] [CrossRef]
  161. Butler, S.; Meegan, M. Recent Developments in the Design of Anti-Depressive Therapies: Targeting the Serotonin Transporter. Curr. Med. Chem. 2008, 15, 1737–1761. [Google Scholar] [CrossRef]
  162. Sanchez, C.; Reines, E.H.; Montgomery, S.A. A comparative review of escitalopram, paroxetine, and sertraline: Are they all alike? Int. Clin. Psychopharmacol. 2014, 29, 185. [Google Scholar] [CrossRef]
  163. Coleman, J.A.; Gouaux, E. Structural basis for recognition of diverse antidepressants by the human serotonin transporter. Nat. Struct. Mol. Biol. 2018, 25, 170–175. [Google Scholar] [CrossRef]
  164. Coleman, J.A.; Yang, D.; Zhao, Z.; Wen, P.-C.; Yoshioka, C.; Tajkhorshid, E.; Gouaux, E. Serotonin transporter–ibogaine complexes illuminate mechanisms of inhibition and transport. Nature 2019, 569, 141–145. [Google Scholar] [CrossRef] [PubMed]
  165. Möller, I.R.; Slivacka, M.; Nielsen, A.K.; Rasmussen, S.G.F.; Gether, U.; Loland, C.J.; Rand, K.D. Conformational dynamics of the human serotonin transporter during substrate and drug binding. Nat. Commun. 2019, 10, 1687. [Google Scholar] [CrossRef] [PubMed]
  166. Hellsbergid, E.; Ecker, G.F.; Stary-Weinzingerid, A.; Forrestid, L.R. A structural model of the human serotonin transporter in an outward-occluded state. PLoS ONE 2019, 14, e0217377. [Google Scholar] [CrossRef] [Green Version]
  167. Laasonen-Balk, T.; Kuikka, J.; Viinamӓki, H.; Husso-Saastamoinen, M.; Lehtonen, J.; Tiihonen, J. Striatal dopamine transporter density in major depression. Psychopharmacology 1999, 144, 282–285. [Google Scholar] [CrossRef] [PubMed]
  168. Stahl, S.M.; Pradko, J.F.; Haight, B.R.; Modell, J.G.; Rockett, C.B.; Learned-Coughlin, S. A Review of the Neuropharmacology of Bupropion, a Dual Norepinephrine and Dopamine Reuptake Inhibitor. Prim. Care Companion J. Clin. Psychiatry 2004, 6, 159–166. [Google Scholar] [CrossRef] [Green Version]
  169. Penmatsa, A.; Wang, K.H.; Gouaux, E. X-ray structures of Drosophila dopamine transporter in complex with nisoxetine and reboxetine. Nat. Struct. Mol. Biol. 2015, 22, 506–508. [Google Scholar] [CrossRef] [Green Version]
  170. Wang, K.H.; Penmatsa, A.; Gouaux, E. Neurotransmitter and psychostimulant recognition by the dopamine transporter. Nature 2015, 521, 322–327. [Google Scholar] [CrossRef] [Green Version]
  171. Cheng, M.H.; Block, E.; Hu, F.; Cobanoglu, M.C.; Sorkin, A.; Bahar, I. Insights into the Modulation of Dopamine Transporter Function by Amphetamine, Orphenadrine, and Cocaine Binding. Front. Neurol. 2015, 6, 134. [Google Scholar] [CrossRef] [Green Version]
  172. Moriguchi, S.; Yamada, M.; Takano, H.; Nagashima, T.; Takahata, K.; Yokokawa, K.; Ito, T.; Ishii, T.; Kimura, Y.; Zhang, M.-R.; et al. Norepinephrine Transporter in Major Depressive Disorder: A PET Study. Am. J. Psychiatry 2017, 174, 36–41. [Google Scholar] [CrossRef]
  173. Klimek, V.; Stockmeier, C.; Overholser, J.; Meltzer, H.Y.; Kalka, S.; Dilley, G.; Ordway, G.A. Reduced Levels of Norepinephrine Transporters in the Locus Coeruleus in Major Depression. J. Neurosci. 1997, 17, 8451–8458. [Google Scholar] [CrossRef] [Green Version]
  174. Góral, I.; Łątka, K.; Bajda, M. Structure Modeling of the Norepinephrine Transporter. Biomolecules 2020, 10, 102. [Google Scholar] [CrossRef] [Green Version]
  175. Jha, P.; Ragnarsson, L.; Lewis, R.J. Structure-Function of the High Affinity Substrate Binding Site (S1) of Human Norepinephrine Transporter. Front. Pharmacol. 2020, 11, 217. [Google Scholar] [CrossRef]
  176. Meng, X.-Y.; Zhang, H.-X.; Mezei, M.; Cui, M. Molecular Docking: A Powerful Approach for Structure-Based Drug Discovery. Curr. Comput. Aided-Drug Des. 2011, 7, 146–157. [Google Scholar] [CrossRef]
  177. Park, S.-H. Ensemble of G Protein-Coupled Receptor Active States. Curr. Med. Chem. 2012, 19, 1146–1154. [Google Scholar] [CrossRef]
  178. Grouleff, J.; Ladefoged, L.K.; Koldsø, H.; Schiøtt, B. Monoamine transporters: Insights from molecular dynamics simulations. Front. Pharmacol. 2015, 6, 235. [Google Scholar] [CrossRef] [Green Version]
  179. Manepalli, S.; Surratt, C.K.; Madura, J.D.; Nolan, T.L. Monoamine Transporter Structure, Function, Dynamics, and Drug Discovery: A Computational Perspective. AAPS J. 2012, 14, 820–831. [Google Scholar] [CrossRef] [Green Version]
  180. Lyne, P.D. Structure-based virtual screening: An overview. Drug Discov. Today 2002, 7, 1047–1055. [Google Scholar] [CrossRef]
  181. Mortensen, O.V.; Kortagere, S. Designing modulators of monoamine transporters using virtual screening techniques. Front. Pharmacol. 2015, 6, 223. [Google Scholar] [CrossRef] [Green Version]
  182. Park, K. A review of computational drug repurposing. Transl. Clin. Pharmacol. 2019, 27, 59. [Google Scholar] [CrossRef] [Green Version]
  183. Ebada, M.E. Drug repurposing may generate novel approaches to treating depression. J. Pharm. Pharmacol. 2017, 69, 1428–1436. [Google Scholar] [CrossRef] [Green Version]
  184. Caban, A.; Pisarczyk, K.; Kopacz, K.; Kapuśniak, A.; Toumi, M.; Rémuzat, C.; Kornfeld, A. Filling the gap in CNS drug development: Evaluation of the role of drug repurposing. J. Mark. Access Health Policy 2017, 5, 1299833. [Google Scholar] [CrossRef]
  185. Cheng, F.; Li, W.; Liu, G.; Tang, Y. In Silico ADMET Prediction: Recent Advances, Current Challenges and Future Trends. Curr. Top. Med. Chem. 2013, 13, 1273–1289. [Google Scholar] [CrossRef]
Figure 1. The structure of bovine rhodopsin (PDB ID: 1F88). The helices are colored from N-terminus (blue) to C-terminus (red).
Figure 1. The structure of bovine rhodopsin (PDB ID: 1F88). The helices are colored from N-terminus (blue) to C-terminus (red).
Molecules 27 00533 g001
Figure 2. The structure of leucine transporter (PDB ID: 2A65). The helices are colored from N-terminus (blue) to C-terminus (red).
Figure 2. The structure of leucine transporter (PDB ID: 2A65). The helices are colored from N-terminus (blue) to C-terminus (red).
Molecules 27 00533 g002
Table 1. GPCR targets, compounds with antidepressant activity, and signaling pathways activated by the target proteins.
Table 1. GPCR targets, compounds with antidepressant activity, and signaling pathways activated by the target proteins.
Receptor FamilyTargetCompoundsSignaling Pathways
Serotonin receptors 5-HT1ABuspirone,
Tandospirone,
F-15599,
F-13714
Gi/Go pathways (cAMP)
Dopamine receptorsD2, D3Aripiprazole,
Brexpiprazole,
Cariprazine
Gi/Go pathways (cAMP)
Opioid receptorsMOR, DOR, KORBuprenorphine,
Nalmefene,
Tianeptine,
BTRX-246040
Gi/Go pathways (cAMP)
Glutamate receptorsmGlu2, mGlu3MGS0039,
LY341495,
RO4491533
Gi/Go pathways (cAMP)
Orphan receptorsGPR26, GPR56, GPR158, P7,
P19
Gs pathways (cAMP)
AKT, GSK3,
EIF4 upregulation
RGS7 regulation
Trace amine associated receptorsTAAR1 TAAR5RO5203648,
RO5263397,
Trimethylamine
Gs pathways (cAMP)
Cannabinoid receptorsCB1, CB2THC,
Rimonabant,
Terpineol,
β-caryophyllene
Gi/Go pathways (cAMP)
Cholinergic receptorsM1, M2Scopolamine,
VU0255035
Gi/Go pathways (cAMP)
Gq pathways
GABA receptorsGABA BCGP56433A,
CGP36742,
CGP51176,
GS39783,
CGP7930
Gi/Go pathways (cAMP)
Neurokinin receptorsNK1Aprepitant,
L-759274
Gs pathways (cAMP)
Cholecystokinin receptorsCCK2L-365,260,
CI-988
Gq pathways
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jarończyk, M.; Walory, J. Novel Molecular Targets of Antidepressants. Molecules 2022, 27, 533. https://doi.org/10.3390/molecules27020533

AMA Style

Jarończyk M, Walory J. Novel Molecular Targets of Antidepressants. Molecules. 2022; 27(2):533. https://doi.org/10.3390/molecules27020533

Chicago/Turabian Style

Jarończyk, Małgorzata, and Jarosław Walory. 2022. "Novel Molecular Targets of Antidepressants" Molecules 27, no. 2: 533. https://doi.org/10.3390/molecules27020533

Article Metrics

Back to TopTop