Next Article in Journal
CNP-miR146a Decreases Inflammation in Murine Acute Infectious Lung Injury
Previous Article in Journal
Cavitation-Enhanced Drug Delivery and Immunotherapy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Imidazoles as Serotonin Receptor Modulators for Treatment of Depression: Structural Insights and Structure–Activity Relationship Studies

1
Department of Pharmaceutical Sciences, Gurukul Kangri (Deemed to Be University), Haridwar 249404, Uttarakhand, India
2
Department of Pharmaceutical Sciences, HNB Garhwal University, Chauras Campus, Srinagar Garhwal 246174, Uttarakhand, India
3
Amity Institute of Pharmacy, Amity University, Noida 201313, Uttar Pradesh, India
4
Department of Bioproduct and Biosystems, Aalto University, 02150 Espoo, Finland
5
Department of Chemistry, Graphic Era (Deemed to Be University), Dehradun 248002, Uttarakhand, India
*
Authors to whom correspondence should be addressed.
Pharmaceutics 2023, 15(9), 2208; https://doi.org/10.3390/pharmaceutics15092208
Submission received: 21 June 2023 / Revised: 13 August 2023 / Accepted: 19 August 2023 / Published: 26 August 2023
(This article belongs to the Special Issue Drug Development and Treatment for Neurological Disease)

Abstract

:
Serotoninergic signaling is identified as a crucial player in psychiatric disorders (notably depression), presenting it as a significant therapeutic target for treating such conditions. Inhibitors of serotoninergic signaling (especially selective serotonin reuptake inhibitors (SSRI) or serotonin and norepinephrine reuptake inhibitors (SNRI)) are prominently selected as first-line therapy for the treatment of depression, which benefits via increasing low serotonin levels and norepinephrine by blocking serotonin/norepinephrine reuptake and thereby increasing activity. While developing newer heterocyclic scaffolds to target/modulate the serotonergic systems, imidazole-bearing pharmacophores have emerged. The imidazole-derived pharmacophore already demonstrated unique structural characteristics and an electron-rich environment, ultimately resulting in a diverse range of bioactivities. Therefore, the current manuscript discloses such a specific modification and structural activity relationship (SAR) of attempted derivatization in terms of the serotonergic efficacy of the resultant inhibitor. We also featured a landscape of imidazole-based development, focusing on SAR studies against the serotoninergic system to target depression. This study covers the recent advancements in synthetic methodologies for imidazole derivatives and the development of new molecules having antidepressant activity via modulating serotonergic systems, along with their SAR studies. The focus of the study is to provide structural insights into imidazole-based derivatives as serotonergic system modulators for the treatment of depression.

1. Introduction

The discovery of serotonin dates back to the mid-19th century, when scientists explored the physiological implications of numerous chemical messengers in the body [1]. Serotonin was first isolated from the blood serum in 1937 by Vialli and Erspamer, Italian scientists. The group identified the highest reservoir of serotonin in the enterochromaffin cells of the gut system. Two other scientists, Rapport and Page, further crystallized the chemical and coined the name ‘serotonin’ in 1945. They and the former scientists found that this very chemical was capable of causing smooth muscle contraction. Owing to this, they named this chemical “serotonin,” derived from two words, ‘sero’ meaning presence in blood serum, and ‘tonin,’ meaning the ability to induce contraction in smooth muscles [2,3].
Additional research revealed that serotonin high-concentration reservoirs are also located in platelets and the central nervous system, in addition to the enterochromaffin cells of the gut [4]. This chemical was also discovered to modulate numerous physiological processes other than smooth muscle contraction and was found to be involved in regulating mood, sleep, appetite, and pain perception [5]. Later, as the research progressed, scientists were keen to discover the specific receptors through which serotonin exerts its physiological effects. It was in the 1970s that scientists discovered two significant classes of serotonin receptors, viz., 5-HT1 and 5-HT2, that were further divided into numerous subtypes [6]. The discovery of these receptors and their subtypes led to advancements in understanding their molecular biology, unraveling numerous physiological questions associated with their signaling [7]. The research also facilitated the scientist’s ability to explore the impact of serotonin in various pathological and physiological conditions and accordingly develop specific serotonin modulators. These modulators include but are not limited to selective agonists and antagonists, selective serotonin reuptake inhibitors (SSRIs), for treating associated conditions that include mental health conditions, including depression and anxiety disorders [8].
With the advancement in molecular neuroscience in the last 60 years, there has been a paradigm shift in developing antidepressant-based therapies [9]. Numerous studies are underway that attempt to recognize the other orchestrating partners with serotonin in combination with or independently associated with depression. Besides paving the way for new research in depression, the studies also expressed their concerns over the serious fears among the patients prescribed serotonin modulators and overmedicalized for years [10]. The US FDA approves numerous modulators for treating any imbalance in serotonin physiology, including (a) Selective serotonin reuptake inhibitors (SSRIs): Citalopram, Escitalopram, Fluoxetine, Fluvoxamine, Paroxetine, Sertraline, Vilazodone, and other drugs; (b) Serotonin and norepinephrine reuptake inhibitors (SNRIs): Venlafaxine, Duloxetine, Desvenlafaxine, and other drugs; (c) Tricyclic and tetracyclic antidepressants: Amitriptyline, Imipramine, Nortriptyline, Doxepin. Notably, more than 90% of these drugs belong to the nitrogen heterocyclic categories, which mainly include indole rings (Fluoxetine, Sertraline, Vilazodone, Vortioxetine); Benzimidazole (Vortioxetine); Benzazepine (Tianeptine); Benzothiazepine (Esmirtazapine); and other drugs [11].
Among the known nitrogen-containing heterocycles, imidazole is an essential heterocycle moiety explored for its biological and medicinal attributes. Imidazole is a widely explored five-membered aromatic heterocyclic compound found in synthetic and natural compounds [12]. Imidazole-containing molecules attach to a wide range of therapeutic targets thanks to their unique structural characteristics and electron-rich environment, resulting in a wide range of bioactivities [13,14,15].
The imidazole ring, as a heterocycle, is part of essential amino acids, including histidine (a histamine precursor). It is widely used in drugs such as antifungal agents (ketoconazole, clotrimazole) [16], antihistamines (cimetidine) [15], COX-inhibitors [17], as well as in other diseases (for example, anticancer [18], antibacterial, antitubercular [19], anti-inflammatory [20], antineuropathic [21], anti-Alzheimer [22,23], antihypertensive [24], antiviral [25], anti-obesity [26], and antiparasitic activity [27]). Significantly, due to facile structural substituting availability on the imidazole ring, various derived molecules were exploited for gaining potent anticancer activities, for example, topoisomerase inhibitors (as imine-amide imidazole conjugates targeting liver and lung cancer cells) [28]. Additionally, they have been employed in naturally derived compounds, such as MIM1 [29], Meiogynins [30,31,32], and synthetically designed Mcl-1 inhibitors (as in the form of imidazolidine-2,4-dione for chronic myelogenous leukemia cells, prostatic cancer cells, and breast cancer cells [33,34]). Furthermore, these compounds have been investigated for their potential in targeting insulin-linked cancer cells (insulin receptor-A and IGF-1R and their heterodimers [35,36]). They have also been explored for hormonal-based targeting of cancer cells (estrogen-based drugs targeting breast cancer [37,38,39,40,41]).
The application of imidazole rings is not limited to small molecule-based medicinal chemistry; aspects of coordination chemistry and targeted synthesis (with chemical biology applications) were explored as well. For example, medicinal coordination chemistry utilizing elements of supramolecular chemistry has recently gained interest [42]. In coordination chemistry, the imidazole ring presented broad applications with its superlative chemical feasibility, such as lone pairs of nitrogen in imidazole being available to facilitate making coordinate bonds with central metals, as observed in nature (in some of the essential biomolecules, hemoglobin, and hemocyanin), encouraging researchers to develop metal-containing medicinal compounds. The imidazole derivatized coordinate compounds additionally exhibit more advantages, such as lower incidences of cellular resistance, increased therapeutic efficacy, selective cellular targeting [13], and probing, as broadly summarized: (A) Imidazole-based supermolecules as anticancer agents (alkylating agents [43], noble metal complexes with anticancer activities [44,45], light-activated cytotoxicity of ruthenium-metal complexes [46]). (B) Imidazole-based coordinate compounds for organelles and cellular detection: (a) Boron pyridyl imidazole complex with a characteristic of twisted intra-molecular charge transfer (TICT) does specific detection of BSA levels (identifying denatured BSA versus native BSA) [47]; (b) Imidazole containing dinuclear Ru (II) complex is a lysosome-specific probe, accumulating (specifically) in the lysosomes and therefore assisting in identifying the HeLa cells from healthy HEK293 cells [48]; (c) Gram-negative bacteria differentiated from Gram-positive bacteria using imidazole containing dibenzimidazole-substituted pyridine [49]. (C) Cellular biomolecular and metal detection and probing: (a) Imidazole containing diphenyl derivatives [50] and anthraquinones detected the biological mercaptans (cysteine, homocysteine, and glutathione) levels [51]. (b) Imidazole containing fluorescent sensors detected approximately 4.70 × 10−7 mol/L adenine levels [52]. (c) Fluorescent probe-conjugated imidazole–pyridine pharmacophore structures demonstrated high sensitivity at a molar concentration of 3.38 μM for Ag (I), applicable to liver cellular imaging [53]. (d) Thiophene-derivatized imidazoles as on–off fluorescent reversible chemosensors detecting intracellular Pd (II) ions in living cells (to a level of 20 μg/mL) [54]. (e) Fluorescent probe-based peptide receptors detected intracellular Cu (I) ions in the Golgi apparatus (even in living A549 cells) [55]. (f) Tripyridyl imidazole molecule as a dual sensing probe of Hg (II) (pH 6–8) and Cu (II) (pH 3–11) ions detecting at 0.77 and 1.58 μM, respectively [56]. (g) Ruthenium (II) complexes containing imidazole as 1O2-responsive fluorescent probes as an “on–off”-type fluorescent pH sensor (pKa1 = 1.12 ± 0.15, pKa2 = 6.90 ± 0.24, pKa1* = 1.09, and pKa2* = 6.92) [57]. (h) Fluorescein dye containing imidazole as a colorimetric and fluorescent chemical sensor for fluoride detection with an application in live cells [58]. (i) Cu (II) complexed Bipyridyl imidazole derivative selectively detected the HS ions and was used to develop a fluorescence microplate assay [59]. (j) Fluorescent zinc complexes presented a high-potential HS/H2S fluorescence sensor and detection probes at cellular levels [60]. (k) Ir (III) containing a methylene-bridged benzimidazole-substituted complex displayed a high selectivity for pyrophosphate ions (H2P2O72−) with lower cellular HeLA cell toxicity [61]. (l) Fluorescent-active benzimidazole derivatives have high selectivity toward aqueous solubilized Ag (I) within (<30 s) [62]. (m) A reversible naphthalimide-based probe detected Hg (II) ions in a phosphate buffer over a wide pH range (7.0–10.0), which is highly applicable in biological experiments [63]. (n) Imidazolyl Schiff bases exhibit Zn (II) detection as low as 6.78 × 10−9 M compared to the recommended sensitivity detection according to the World Health Organization’s drinking water guidelines (7.6 × 10−5 M) [64]. (o) Imidazole-based anthracene structure detection limit of Zn (II): 1.0 × 10−9 M [65]. (p) Tridentate dibenzimidazole–pyridines have a detection limit of 3.09 × 10−7 M toward Zn (II) ions [66]. (q) Imidazole-based supermolecules as probes for Cu (II) detection (allyl-substituted imidazole derivative with a detection limit for cupric ion (Cu2+) of 1.01 nM [67]; dibenzimidazole derivative with a detection limit of 0.094 μM within a 1 s time period [68]; tetraphenyl ethylene-functionalized aryl imidazole-derivative detection limit of aqueous solubilized Cu (II) at 34.8 nM [69]). (r) Imidazole-based iron chemosensors (fluorescence “on–off” aniline-derived imidazole probe demonstrated high selectivity and sensitivity for ferric ions Fe (III) with a detection limit of 0.72 μM/L at 30 min) [70].
Several imidazole-based medicines have been widely employed in clinical trials to treat various disorders with significant therapeutic promise. Imidazole-containing drug research and development is becoming a more active and appealing issue in medicinal chemistry due to its substantial therapeutic usefulness. Therefore, this article is kept forth to illuminate the landscape of imidazole-based drug discovery and development, focusing on structure–activity relationship aspects (SAR) against serotonin receptors to target depression. The insights covered will compel medicinal chemists, synthetic chemists, biologists, and pharmacologists to explore and unravel this exciting area for the drug discovery of serotonin modulators.

Etiology of Depression, Structural and Mechanistic Insights

According to the World Health Organization, depression is a severe condition affecting millions of people worldwide and is one of the primary causes of disability [71,72]. Monoamine neurotransmitters like norepinephrine and serotonin (5-hydroxytryptamine, 5-HT) have been used as significant indicators of psychiatric disorders, such as anxiety and depression, for several decades [73,74]. Serotonin exists in both the CNS and PNS systems, which describe its nature as both an autacoid and a neurotransmitter [75,76]. It is released by serotonergic neurons. The serotonergic system is linked to the regulation of mood, emotion, and sleep, as well as a variety of behavioral and physiological activities [77]. It is one of the most studied and multifunctional biogenic amines among neurotransmitters. Depending on the physiological action, occurrence, agonist, and antagonist, it is categorized into various classes, among whom 5-HT1 (5-HT1A, 5-HT1B, 5-HT1D, 5-HT1E, and 5-HT1F), 5-HT2 (5-HT2A, 5-HT2B, and 5-HT2C), 5-HT3, 5-HT4, 5-HT5 (5-HT5A, 5-HT5B), 5-HT6, and 5-HT7 have been widely reported in the literature [78]. Serotonin is primarily found in three cell types: (a) serotonergic neurons in the central nervous system and the intestinal myenteric plexus [79], (b) enterochromaffin cells in the gastrointestinal tract mucosa, and (c) blood platelets [80].
Serotonin receptors are categorized as G protein-coupled receptors (GPCRs), which orchestrate the signaling mechanism of serotonin. Serotonin binds to its chief receptors (5-HT1 and 5-HT2 subtypes) within the transmembrane region of the receptor. This region comprises transmembrane helices classified as TM1–TM7 [81]. These transmembranes are further connected via intracellular and extracellular loops, forming a binding pocket that facilitates the formation of serotonin and its competitive modulators. These transmembrane domains are further surrounded by cholesterol molecules that, besides contributing to the receptor’s shape, are also contributing factors for receptor stability and proper receptor folding. The cholesterol composition within the cell membrane directly modulates receptor activity.
The essential amino acids that play a critical role in serotonin and ligand are embedded within the binding pocket. Considering the 5-HT1 subtype receptor, the key amino acid is aspartate (ASP), located in TM3 (ASP116). This amino acid mediates serotonin affinity via the formation of a salt bridge with the amine functionality of serotonin, thus stabilizing the ligand-receptor complex. Considering the 5-HT2 subtype, the key amino acid that is involved in stabilizing the ligand affinity is asparagine (ASN), located in TM7. This specific amino acid is known to interact with the carboxyl group of serotonins via the formation of the H-bond. Apart from these vital amino acids, Arginine (ARG) residues present in TM3 (ARG 134) and TM6 are also found to interact with carboxyl groups via H-bonding and electrostatic interactions. Glutamate (GLU), present in TM2 and TM7, is involved as a proton acceptor or donor required during protein interactions, besides participating in H-bonding and electrostatic interactions with ligands. Another amino acid, histidine (HIS), present in TM3 and TM5, is also associated with receptor activation and participates in ligand stability via the formation of an H-bond, a salt bridge, or by acting as a proton donor or acceptor. In addition to this, cysteine (CYS) residues present in TM2 and TM7 are involved in the stability of the receptor protein via the formation of disulfide bridges. They are vital for ligand recognition and receptor activation. The tryptophan (TRP) residues (particularly TRP 358) available at TM6 and TM7 also contribute to receptor activation along with receptor conformational stability and ligand binding primarily via hydrophobic interactions. Apart from amino acids, three water molecules (W1–W3) within the active domains are also vital in maintaining the stability and activation of the receptor. These water molecules interact with 5-HT (apoprotein), where W1 interacts with the hydroxy group and W2 with the indole ring, followed by the interaction of the primary amine with W3. W3 also interacts with ASP116 by forming an H-bond that is conserved in GPCRs (aminergic), whereas W2 is found to toggle the amino acid residue TRP358, which determines receptor activation [82].
Mechanistically, serotonin and norepinephrine are released in the synaptic cleft, where they activate the postsynaptic cleft and some reuptakes through the pump, where MAO breaks them and moves them back to the presynaptic neuron [83]. In the case of depression treatment using SSRIs/SNRIs as per the monoamine hypothesis, the therapeutic benefits are based on increasing low serotonin levels and norepinephrine (Figure 1) [84,85]. As the name indicates, SSRIs/SNRIs function by preventing serotonin/norepinephrine reuptake and thereby increasing activity. SSRIs block the serotonin transporter (SERT) at the presynaptic axon terminal [86], whereas SNRIs block the norepinephrine reuptake transporter [87]. The information-sending presynaptic cell in the brain releases neurotransmitters into the gap [88]. The neurotransmitters are subsequently identified by receptors on the recipient postsynaptic cell’s surface, which then relay the signal in response to the stimulus. More serotonin (5-hydroxytryptamine, or 5HT) remains in the synaptic cleft when SERT is blocked, which can stimulate postsynaptic receptors for extended periods [89]. Further, SNRIs and SSRIs boost serotonin and norepinephrine levels in the brain. Neurotransmitters, such as serotonin and norepinephrine, are chemical messengers that transfer messages from one region of the body to another. Following the transmission of a signal by neurotransmitters, cells in the brain typically take up these substances and store them for later use. SNRIs and SSRIs block serotonin and norepinephrine reuptake, resulting in higher levels of serotonin and norepinephrine in the synaptic cleft [90].

2. Modern Synthetic Methods for Substituted Imidazole Derivatives

Imidazole is reported to exhibit a broad range of applications in pharmaceutical and industrial applications [14]. For example, this organic framework is sought in many drug pharmacophores, such as angiotensin II inhibitors, anti-inflammatory agents [91], anticancer agents, and building blocks of naturally occurring products [13,92]. The imidazole ring is a well-observed ligand in metalloenzymes, and its imidazolium salts are also well-exploited to serve as excellent precursors of stable carbene ligands in various metal complexes [93,94]. The application of imidazolium salts to environmentally friendly ionic solvents is another example [95]. Thenrajan and coworkers explored the role of imidazolate-based bimetallic nickel-iron zeolitic fibers as sensors for serotonin neurotransmitters [96]. Various advancements reported in the synthesis of new imidazole derivatives with various bioactivities using different catalytic systems are described in Figure 2.
The use of metallic catalysts for imidazole synthesis has increased in recent years. This can be attributed to the improved percentage yields, lessening the time required for reaction, and ease of removal from a reaction mixture that make the technique appealing [93,97]. Because of their nontoxic, affordable, reusable, and eco-friendly properties, zinc (Zn)-based heterogeneous catalysts were well-exploited in a range of multicomponent reactions for synthesizing this organic framework. Marzouk et al. [98] developed a one-pot multicomponent synthesis of 1,2,4,5-tetrasubstituted imidazoles via reacting aromatic aldehydes, benzil, 1-amine-2-propanol, and ammonium acetate in the presence of nanoparticles of ZnFe2O4 catalyst. After 30–50 min, the condensation process with a metal catalyst yielded 87–96% multi-substituted imidazoles. For the production of substituted imidazole, Nejatianfar et al. [99] proposed a magnetically separable nanocatalyst based on copper (II) immobilized on guanidine epibromohydrin-functionalized c-Fe2O3@TiO2 (c-Fe2O3@TiO2-EGCu(II)) with a core–shell structure. Eidi et al. [100] used benzil, substituted aldehydes, and ammonium acetate in a condensation procedure to produce 2,4,5-trisubstituted imidazole conjugates. The best reaction conditions were discovered using 10 mg of catalyst and a 60 percent rate power of ultrasonic irradiation at 40 °C in ethanol. The method’s advantages include recovering the catalyst using an external magnetic field and reusing it for up to five runs without losing activity. Maleki et al. [101] developed a greener synthetic strategy for the formation of 2,4,5-trisubstituted imidazoles via condensation of 1,2-diketone, aromatic aldehydes, and ammonium acetate in the presence of mixed oxide (Fe3O4/SiO2) nanocatalyst, yielding up to 95%. Compared to traditional catalysts such as Fe3O4, the reaction utilizing Fe3O4/SiO2/urea nanoparticles took 50 min and yielded 95% trisubstituted imidazoles. However, designing and synthesizing trisubstituted scaffolds remain of keen interest among researchers [102,103]. These scaffolds provide a highly functional multitargeting scaffold [104,105], including dendrimers [106,107]. Girish et al. [108] used ZrO2-supported b-cyclodextrin as a reusable solid catalyst to synthesize 2,4,5-trisubstituted imidazoles and 1,2-disubstituted benzimidazoles under solvent-free conditions. Using a 40 mol% ZrO2-b-Cyclodextrin catalyst, the reaction was screened with various solvents, including water, DMF, ethanol, and solventless systems. The reaction went off without a hitch in a solvent-free environment, and the product was produced with good yields. Using ZrO2 nanoparticles as a reusable catalyst, Bajpai et al. [109] investigated the one-pot synthesis of multi-substituted imidazoles. The authors used isatin, aromatic aldehydes, and ammonium acetate as reactants in the presence of 15 mol% ZrO2 nanoparticles in a solvent-free environment to create new imidazoles.
Fang et al. [110] described a new method for cyclizing amido-nitriles to produce disubstituted imidazoles. The reaction conditions were moderate enough to include aryl halides, aromatic and saturated heterocycles, and other functional groups. The necessary 2,4-disubstituted NH-imidazoles were obtained by nickel-catalyzed addition to nitrile, followed by proto-demetallation, tautomerization, and dehydrative cyclization. Combining a C2–N3 fragment with an N1–C4–C5 unit has recently been examined as a two-bond disconnection for synthesizing imidazoles. For example, Shi et al. [111] employed this disconnection to make trisubstituted NH-imidazoles in the presence of zinc(II) chloride by reacting benzimidates with 2H-azirines. For the synthesis of 2-aminoimidazoles, Man et al. [112] utilized an approach where vinylazides were transformed in situ into 2H-azirines, which then interacted with cyanamide to create the required 2-aminoimidazoles in moderate to good yields under a range of conditions. Nitriles have also been employed as reagents in metal-free processes to synthesize substituted imidazoles with two bonds. Harisha et al. [113], for example, recently reported the formation of tri-substituted NH-imidazoles by reacting -azidoenones with substituted nitriles. In the absence of a metal catalyst, imidamides can be employed as starting materials for synthesizing imidazoles. In the presence of trifluoroacetic acid, Tian et al. [114] reported the synthesis of substituted imidazole by reacting imidamides with sulphoxonium ylides. At the first, second, and fourth positions, the resulting imidazoles were replaced.

3. SAR of Various Imidazole Derivatives

As a new generation of selective 5-HT7 receptor agonists, Hogendorf et al. developed and synthesized a series of 27 fluorinated 3-(1-alkyl-1H-imidazol-5-yl)-1H-indoles. By optimizing the halogen bond formation with Ser5.42 as the expected partner, a powerful and drug-like agonist, compound 1, 3-(1-ethyl-1H-imidazol-5-yl)-5-iodo-4-fluoro-1H-indole (Ki 5-HT7R = 4 nM), was discovered. Excellent water solubility, good selectivity over related CNS targets, high metabolic stability, oral bioavailability, and minimal cytotoxicity were all attributes of the molecule. After i.p. (2.5 mg/kg) treatment in mice, rapid absorption into the blood, a medium half-life, and a high peak concentration in the brain (Cmax = 1069 ng/g) were discovered. The antinociceptive effect reported in a mouse model of neuropathic pain suggests that 1 might be a long-sought tool chemical in the research of 5-HT7 receptor function, as well as a potential analgesic [115]. The detailed SAR analysis of this series of compounds is discussed in Figure 3.
As non-sulfonamide 5-HT6 receptor ligands, a new series of 3H-imidazo[4,5-b]pyridine and 3H-imidazo[4,5-c] pyridine derivatives were reported by Vanda et al. Compound 2 (2-ethyl-3-(3-fluorobenzyl)-7-(piperazin-1-yl)-3H-imidazo[4,5-b]pyridine) was identified as a strong 5-HT6 receptor partial inverse agonist in Gs signaling (Ki = 6 nM, IC50 = 17.6 nM) after in vitro testing. Compound 2 had good metabolic stability, a favorable cytochrome P450 isoenzyme profile (2D6, 3A4), did not impact PgP-protein binding, and had no mutagenesis effects. The bioavailability and blood-brain barrier (BBB) permeability of this compound further favors its candidature as an investigative molecule in clinical studies. Along with non-neurotoxicity, the use of a synergistic combination of inactive doses of compound 2 (0.1 mg/kg) and donepezil (0.3 mg/kg) to reverse scopolamine-induced memory deficits produced a synergistic effect. In the binding pocket of ligands for 5-HT6R, various small or bulky alkyl/aryl substituents were employed at the C2-position of the imidazo[4,5-b]pyridine moiety, and SAR studies were developed as discussed in Figure 4 [116].
Zagorska et al. developed and performed pharmacological screening of 2-fluoro and 3-trifluoromethylphenylpiperazinylalkyl derivatives of 1H-imidazo[2,1-f]purine-2,4(3H,8H)-dione (3au) for their antidepressant activity targeting serotonin (5-HT1A/5-HT7) receptors as well as phosphodiesterase (PDE4B and PDE10A). Although the results from in vitro studies revealed the synthesized compounds as potent 5-HT1A, 5-HT7, or dual 5-HT1A/5-HT7 receptor ligands, they had lower inhibitory potencies for PDE4B and PDE10A. Most of the compounds displayed selective 5-HT1A receptor affinity, which encouraged further preclinical investigation. Target compounds 3au was investigated for their metabolic stability and lipophilicity properties utilizing micellar electrokinetic chromatography (MEKC) technology and a human liver microsomes (HLM) model, which presented the target compounds as having moderate pharmacophoric features. During FST, compound 3i, 8-(5-(4-(2-fluorophenyl)piperazin-1-yl)pentyl)-1,3,7-trimethyl-1H-imidazo[2,1-f]purine-2,4(3H,8H)-dione, was found to possess the most prominent antidepressant activity at dosages of 2.5 mg/kg and 5 mg/kg. At a dosage of 2.5 mg/kg, it also showed an anxiolytic effect in the four-plate method. Molecular docking studies also revealed that fluorinated arylpiperazinylalkyl derivatives of 1H-imidazo[2,1-f]purine-2,4(3H,8H)-dione have major pharmacophoric features for the development of the antidepressant and anxiolytic compound. The SAR studies of this series of compounds are discussed in Figure 5 [117].
Arylpiperazine-substituted imidazole derivatives are one of the major classes explored in various neurological disorders [118,119,120]. Zagόrska et al. synthesized a novel series of arylpiperazinylalkyl purine-2,4-diones (4au) and purine-2,4,8-triones (5ah) and screened them in vitro for their serotonergic and dopaminergic receptor affinity. Compounds containing an imidazole ring with purine-2,4-diones in their structure have a high affinity for serotonin receptors (5-HT1A and 5-HT7) as well as dopamine receptors (D2) as compared to compounds with purine-2,4,8-trione. The resultant compounds displayed various CNS activities. During in vivo evaluation, it was discovered that compounds 4h and 4o operate as potential antidepressants, while compounds 4r and 4u act as both potential antidepressants and anxiolytic agents. It has been demonstrated by molecular docking studies that the presence of 1,3-dimethyl-(1H,8H)-imidazo[2,1-f]purine-2,4-dione at position C7 is required for receptor affinity and selectivity, particularly for 5-HT1A and 5-HT7. The SAR studies are discussed below in Figure 6 [121].
The same research group reported two new series of N-8-arylpiperazinylpropyl derivatives of 1,3-dimethyl-(1H,8H)-imidazo[2,1-f]purine-2,4-dione (6ai) and amide derivatives of 1,3-dimethyl-6,7-dihydroimidazo[2,1-f]purine-2,4-(1H,3H)-dione-7-carboxylic acid (7ac) and evaluated for their antidepressant potential using the FST model. During in vitro studies, compound 6h displayed the highest affinity for 5-HT1A receptors and a Ki value of 5.6 nM with high selectivity over 5-HT2A receptors. Further, in vivo evaluation of antidepressant activity revealed that compounds 6a, 8-[3-(N4-phenyl)-piperazin-N1-yl-propyl]-1,3-dimethyl-(1H,8H)-imidazo[2,1-f]purine-2,4-dione and 6b, 8-[3-(N4-20-metoxyphenyl)-piperazin-N1-yl-propyl]-1,3-dimethyl-(1H,8H)-imidazo[2,1-f]purine-2,4-dione possesses the most significant activity and reduces the immobility time in FST, similar to the drug imipramine. These long-chain arylpiperazine derivatives with a tricyclic moiety can be further explored for the development of new antidepressant compounds with minimal to no side effects. The SAR studies of these imidazole-based compounds are discussed in Figure 7 [122].
Tokgoz et al. reported a series of benzazole derivatives (8ah) and evaluated them for their antidepressant-like activities. Novel benzazole subordinate mixtures were synthesized by the reaction of respective 2-(benzazol-2-ylthio)acetohydrazide and 4-substituted benzaldehydes. The antidepressant-like activity of these compounds was assessed by TST and modified constrained swimming tests (MFST). During in vivo studies, at a dosage of 50 mg/kg, compounds 8a, 8b, 8e, and 8f (Figure 8) significantly reduced the immobility time in both TST and MFST. There was no change observed in the climbing duration, which demonstrates the selective antidepressant-like action of these compounds. Further, when pre-treated with serotonin synthesis inhibitors, i.e., p-chloro-phenylalanine methyl ester, NAN-190 (a 5-HT1A antagonist), ketanserin (a 5-HT2A/2C antagonist), and ondansetron (a 5-HT3 antagonist), these compounds showed opposite effects in mice, which shows that these compounds act via the serotonergic system for antidepressant-like activity. In addition, the locomotor exercises of the creatures were surveyed by an action confinement contraption [123].
Czopek et al. designed imidazolidine-2,4-dione derivatives using a computational ligand design approach for finding new dual-targeting (5HT1A receptor and serotonin transporter (SERT)) agents. Structural optimization of the initial 5HT1AR ligands gives a series of Mannich bases (9ah) with an aryl substituent at the 5-position of the imidazolidine-2,4-dione as lead molecules, having the best match with the SERT binding site too. However, the in vitro results were found to be contrary to the expected results for this newly synthesized series. Only compounds with the substituent 3-chlorophenylpiperazine (9c and 9g) showed significant affinities for both 5HT1AR and SERT, with Ki values of 80 nM and 166 nM (9c) and 76 nM and 278 nM (9g), respectively. The results demonstrated the high importance of imidazole substituted with an arylpiperazine moiety. The selected compounds 9c and 9g further showed partial agonist and antagonist properties at the 5HT1A and 5HT2A receptor sites, respectively, and had low affinity for α1 receptors. Such encouraging results further motivated the evaluation of the compounds for their antidepressant and anxiolytic properties. The studies demonstrated that compounds 9c and 9g significantly reduced the immobility time in mice in FST, and the antidepressant effect of 9c was found to be comparable to the reference drug imipramine. The compounds at the doses administered did not affect locomotor activity. The detailed SAR study of these compounds is discussed in Figure 9 [124].
Seo et al. investigated a series of imidazole cores containing arylpiperazine-4-carboxamide derivatives targeting 5-HT2A receptors and 5-HT transporters for the treatment of depressive disorders. Aa imidazole moiety provides good pharmacokinetic properties in drug development. The research group focused on developing core molecules around this moiety. The current series of compounds displayed potential IC50 values against 5-HT2A/2C and serotonin reuptake inhibition. Further, these compounds showed significant in vivo antidepressant-like effects in the FST. Based on the evaluation, compounds 10, 11, 12, 13, and 14 (Figure 10) were found to be more promising for antidepressant activity, and it was concluded that this imidazole series could be used as a promising tool for the development of new antidepressants [125].
One of the imidazole-based novel neuronal nitric oxide synthase (nNOS) inhibitors, i.e., 1-(2-trifluoromethylphenyl)-imidazole (15, Figure 10), is sought to synergize the action of other antidepressants via the serotonergic system. Ulak and colleagues reported that FST models augment the antidepressant activity of TCAs like imipramine, SSRIs like citalopram and fluoxetine, and the selective serotonin reuptake enhancer tianeptine. Unfortunately, the same results were not observed with the noradrenaline reuptake inhibitor reboxetine, which confirms the action of 15 via the serotonergic system [126]. Sherwin et al. further confirmed the effect and regional-specific modulation of 1-(2-trifluoromethylphenyl)-imidazole (15) [127].

4. Other Imidazole-Based Serotonin-Modulating Agents

In continuation of efforts to develop new serotonin reuptake inhibitors, Lauro and coworkers recently (2022) reported strychnidin-oxiran-napthalenol derivatives 16 and 17 (Figure 11), which were found to possess dual noradrenaline and serotonin reuptake inhibitory activity in docking models [128]. Although these results were only based on computational docking studies, there is no further proof regarding their clinical efficiency.
Czopek et al. synthesized a new series of compounds derived from 4-methoxy-1H-isoindole-1,3(2H)-dione as dual targeting ligands having affinity for serotonin receptors along phosphodiesterase 10A (18at). To understand the structure–activity relationship of compounds, 4-methoxy-1H-isoindole-1,3(2H)-dione derivatives (18at) conjugated with a variety of amine moieties, including imidazole and benzimidazole, were screened against both targets. In this study, compounds having a benzimidazole moiety conjugated (18g) (2-[4-(1H-benzimidazol-2-yl)butyl]-4-methoxy-1H-isoindole-1,3(2H)-dione) were found to possess the most balanced profile with affinity toward both targets based on in vitro experiments. The lead compound 18g was investigated for its safety profile and subjected to computational studies, where it should have good affinity and bind at the active site, with such interactions responsible for the inhibition of the PDE10A enzyme. The SAR studies of the current series of compounds against serotonin receptors are described in Figure 12 [129].
Zmudzki et al. investigated the influence of the modifications of the fused imidazole nucleus (xanthine) and the effect of the substituent in position eight on the affinity for serotonin 5-HT1A, 5-HT2A, 5-HT6, 5-HT7, and dopamine D2 receptors. For this purpose, this group reported different series of arylpiperazynylalkyl derivatives of 8-amino-3,7-dimethyl-1H-purine-2,6(3H,7H)-dione (total 26 compounds). They performed screening of these compounds in vitro against 5-HT1A, 5-HT2A, 5-HT6, 5-HT7, and dopamine D2 receptors. This study aimed to create a thorough structure–activity relationship profile via various substitutions at position eight with eight different amino moieties. In preliminary evaluation parameters, compounds 1923 (Figure 13) were selected as promising leads, which were further subjected to functional assays for the 5-HT1A and D2 receptors. The results demonstrated that these arylpiperazynylalkyl derivatives of 8-amino-3,7-dimethyl-1H-purine-2,6(3H,7H)-dione act as potential antagonists of 5-HT1A receptors while having agonistic, partial agonistic, or antagonistic activity for D2 receptors. The SAR analysis revealed that the lipophilic substituent at the 8th position plays a crucial role in affinity toward various serotonin receptors (especially 5-HT1A, 5-HT6, and 5-HT7). The removal of the lipophilic moiety caused a complete loss of affinity toward these receptors. Further, the compounds with lipophilic substituents, such as propoxy or N-ethylbenzylamino substituents, possess the most optimal affinity, which was further joined by dipropyloamino and piperidine-1-yl substituents due to their occupancy of similar volumes and similar lipophilic behavior [130].
Hogendorf et al. utilized aromatic basic groups like imidazoles or thiazoles as aminergic receptor ligands and designed compounds from the N-(1H imidazol-2-yl)acylamide chemotype (24at and 25ad). The synthesized compounds displayed high affinity for 5-HT6R and high selectivity over 5-HT1A, 5-HT2A, 5-HT7, and D2 receptors. In this series, compound 4-methyl-5-[1-(naphthalene-1-sulfonyl)-1H-indol-3-yl]-1H-imidazol-2-amine (24i) showed the reversal of scopolamine-induced cognitive impairment in rats. The replacement of the amine group of the 5-HT6 receptor with 2-aminoimidazole, which is its bioisostere, resulted in highly potent compounds with variable physicochemical behavior. Several studies were performed to analyze the basic character of compounds, demonstrating that lowering the basicity below par level compromised 5-HT6R affinity very slightly, although it enhanced the selectivity. The X-ray structure analysis with 5-HT6R homology of 24i and 25b (Figure 14) revealed the binding mode, which further evidenced the in vitro experiment results. The results from these experiments concluded that 2-amidoimidazole-based moieties incorporated with other pharmacophoric features could provide new insights and lead to the development of new ligands of aminergic receptors as potential drug molecules [131].
Bromidge et al. reported novel tricyclic benzoxazine derivatives (26az), which were designed via bioisosteric replacement of the metabolically prone N-methyl amide group with comparatively smaller heterocyclic moieties to give these new tricyclic benzoxazines. These novel compounds were also found to be potent 5-HT1A/B/D receptor antagonists with some 5-HT transporter activity. In this study, compound 26d (Figure 15) emerged as a prominent lead molecule, having 5-HT1A/B/D receptor antagonists with zero intrinsic activity and Ki values of 9.5, 8.8, and 9.8, respectively, with 7.5 against hSerT. This lead compound, 26d, displayed high selectivity over hERG potassium channels and a favorable pharmacokinetic profile during in vivo studies in the PD model. With a potentially balanced profile and encouraging results, compound 26d was used by this research group as a clinical candidate for further investigations as a quick-acting antidepressant/anxiolytic with minimal to no chances of adverse/side effects [132].

5. Imidazole-Based Drugs under Clinical Trials against Depression

There are various imidazole-based drug candidates in different clinical trials for depression. These are listed in Table 1, with clinical trial numbers per data available at clinicaltrials.gov. EVT101 is an orally active NMDR antagonist that entered a phase II clinical trial but was terminated in the initial phase, and the FDA put this drug on hold for further development [133]. Pentoxifylline is a PDE inhibitor that has completed various phase I/II trials as an adjuvant in major depressive disorders [134]. In preclinical studies, it has shown promising results, and pretreatment of pentoxifylline itself showed antidepressant-like activity in animal models [135,136]. Etomidate is well explored for its effects on major depressive disorders in electroconvulsive therapies [137]. It completed a study on its role as a neurorestorative agent (NCT02667353), and one clinical phase IV study status is unknown (NCT02924090), in which it was being explored for major depression. Leuprolide acetate is known for its diverse pharmacological activities [138]. It has completed phase II (NCT04051320) and phase IV (NCT01116401) clinical trials against perinatal depression and menopause depression, respectively.
RO4917523, also known as basimglurant (2-chloro-4-[1-(4-fluoro-phenyl)-2,5-dimethyl-1H-imidazol-4-ylethynyl]-pyridine), is a negative allosteric modulator of the mGlu5 receptor under clinical trials for development against depression [139]. It has completed phase II trials (NCT00809562, NCT01437657) and is under further investigation by Hoffmann-La Roche for treatment-resistant and major depressive disorders. Candesartan is an angiotensin receptor blocker mainly used for hypertension treatment but has also been found to reverse depression-like symptoms in preclinical studies in the initial phase [140,141]. But, in clinical trials, it was withdrawn from phase I in the early phase (NCT04430959), and phase IV studies were terminated due to the adverse effects in treated patients (NCT01794455). Its further development as an antidepressant is still questionable. Cimicoxib, one of the NSAIDs used for pain relief and inflammation in dogs, has completed a phase II trial in major depressive disorder patients (NCT00510822) conducted by Affectis Pharmaceuticals AG [142]. The study was based on the hypothesis that adjuvant COX-2 inhibition can help reverse major depressive symptoms in patients. Valganciclovir, one of the antiviral compounds, is under phase I/II clinical trial for the antiviral treatment of cytomegalovirus in depression (NCT04724447). However, various severe side effects associated with this drug are major hurdles in further development [143]. Another imidazole derivative, Midazolam, a known benzodiazepine medicine, is under various trials for comparison in depressive patients (NCT05026203, NCT02360280, NCT05383313, NCT05528718, NCT01700829), although its role in treatment is still under question [144]. Thus, it is mostly being used as a control group for comparison of other medicines and their effects [145,146].

6. Recently Granted Patents on Imidazole-Based Compounds

Due to their unique structural characteristics and electron-rich environment, imidazole-containing molecules have attracted the attention of several research groups. It led to the development of various fused/unfused molecules with a wide range of bioactivities. Because of its wide variety of actions, many patents have been granted in the last few years. The patents for imidazole derivatives with antidepressant activity are listed in Table 2.
After reviewing the patents related to imidazole-based compounds, it is clear that in the last two decades, very few patents have been granted related to imidazole-based compounds for their potential against depression. Thus, comprehensive options are available to be explored in the upcoming time to explore this moiety for developing new antidepressant molecules with more significant potential and better efficacy.

7. Conclusions and Future Perspective

One key issue that has been unaddressed for a long time is finding a rational design with high preferential selectivity for serotonergic isoforms. Researchers attempted to gain preferential selectivity by bringing specific fused heterocyclic structures, monosaccharides, macrocycles, glycopeptides, and dendrimer-based scaffolds [152,153,154,155,156,157]. However, similar scaffolds with similar modifications exhibit multitarget activities, severely limiting their direct clinical implications [158,159,160,161,162,163]. Therefore, there is considerable demand for identifying newer chemical modalities with enough structural features to incorporate into pharmacophores to bring about selective serotonergic activities.
Due to its various pharmacological properties, the imidazole moiety has emerged as a significant pharmacophoric characteristic in medicinal chemistry. In this review, we aim to discuss the SAR analyses of several imidazole derivatives as antidepressants that target serotonin (5-HT) receptors. We compiled some studies and summarized the structure–activity relationship of the fundamental nucleus in all of them. We have developed a pharmacophore from existing selective serotonin reuptake inhibitors (SSRIs) on the open web server Pharmit. The pharmacophore obtained from the server contains five features: two aromatic rings in the violet color sphere, two hydrogen bond acceptors in an orange sphere, and hydrophobic in the green sphere (Figure 16).
Talking about imidazole acceptability in other chemical biology-oriented approaches, it has shown tremendous potential. These include serving as pH-responsive potent antimicrobials [164], improving the photophysical characteristics of azo dyes [165,166,167], decreasing side toxicities [13,168,169,170], and successful chemical integration in polymeric materials as theragnostic applications (such as in zeolitic imidazolate framework applications in drug delivery) [171,172,173,174,175,176,177,178,179,180].
After reviewing the data, it was evident that the imidazole moiety plays a critical role in treating depression by acting on serotonin receptors. The literature survey revealed that imidazole is a less explored moiety in developing 5-HT receptors that modulate chemical entities for treating depressive disorders. The molecules being explored against 5-HT-bearing imidazole are still in their initial states. This chemical feature, along with encouraging pharmacokinetic features, can provide broad scope for further exploration. Computer-based drug design can further speed up the imidazole-based drug development process with target specificity and efficiency. The electron-rich feature of this five-membered chemical moiety can help it bind with serotonin receptors and transporters very quickly. This can be beneficial for improving the efficacy of molecules [181]. The development of imidazole-based molecules against various disease states has become a hot topic [22,42,182,183,184]. This moiety can further enter into the development of antidepressants with its enriched structural features. This review may be used to draw key conclusions regarding the imidazole nucleus and appropriate substitutes for it, which will aid in developing novel antidepressant medications based on the imidazole nucleus.

Author Contributions

Conceptualization, K.K.G. and S.T.; methodology, K.K.G. and R.K.; resources, G.J. and B.K.; data curation, B.K.; writing—original draft preparation, K.K.G., S.T. and R.K.; writing—review and editing, B.K. and A.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

All contributing authors are thankful to their respective universities for necessary support.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

5-HT5-Hydroxytryptamine
α1RAlpha-1 Receptors
σ2RSigma-2 Receptors
ADMETAbsorption, Distribution, Metabolism, Excretion, Toxicity
BBBBlood Brain Barrier
BBPBlood Brain Permeability
CDKCyclin-dependent Kinase
CNSCentral Nervous System
D receptorDopamine Receptor
EAAT3Excitatory amino acid transporter 3
FSTForce swim test
hERGhuman Ether-à-go-go-Related Gene
HLMHuman Liver Microsomes
hSerTHuman Serotonin Transporter
IC50Half Maximal Inhibitory Concentration
I.P.Indian Pharmacopoeia
IL-17Interleukin-17
KiInhibition constant
MAOMono Amine Oxidase
MDCKMadin-Darby Canine Kidney Cells
MDRMultidrug Resistance Gene
MFSTModified Force Swim Test
MEKCMicellar Electrokinetic Chromatography
PDParkinson’s disease
PgpP-glycoprotein
PKPharmacokinetic
pKaAcid Dissociation Constant
PNSPeripheral Nervous System
PTZPentylenetetrazole
SARStructure–Activity Relationship
SERTSerotonin reuptake transporter
SSRISelective Serotonin Reuptake Inhibitors
TGFTransforming growth factor
TNFTumor necrosis factor
TSTTail Suspension Test

References

  1. Whitaker-Azmitia, P.M. The Discovery of Serotonin and its Role in Neuroscience. Neuropsychopharmacology 1999, 21, 2–8. [Google Scholar] [CrossRef]
  2. Sjoerdsma, A.; Palfreyman, M.G. History of serotonin and serotonin disorders. Ann. New York Acad. Sci. 1990, 600, 1–7. [Google Scholar] [CrossRef]
  3. Andrews, A.M. Celebrating serotonin. ACS Chem. Neurosci. 2012, 3, 644–645. [Google Scholar] [CrossRef]
  4. Maclean, J.A.; Schoenwaelder, S.M. Serotonin in platelets. In Serotonin; Elsevier: Amsterdam, The Netherlands, 2019; pp. 91–119. [Google Scholar]
  5. Jonnakuty, C.; Gragnoli, C. What do we know about serotonin? J. Cell. Physiol. 2008, 217, 301–306. [Google Scholar] [CrossRef]
  6. Berger, M.; Gray, J.A.; Roth, B.L. The expanded biology of serotonin. Annu. Rev. Med. 2009, 60, 355–366. [Google Scholar] [CrossRef]
  7. Kroeze, W.K.; Kristiansen, K.; Roth, B.L. Molecular biology of serotonin receptors-structure and function at the molecular level. Curr. Top. Med. Chem. 2002, 2, 507–528. [Google Scholar] [CrossRef]
  8. Edinoff, A.N.; Akuly, H.A.; Hanna, T.A.; Ochoa, C.O.; Patti, S.J.; Ghaffar, Y.A.; Kaye, A.D.; Viswanath, O.; Urits, I.; Boyer, A.G. Selective serotonin reuptake inhibitors and adverse effects: A narrative review. Neurol. Int. 2021, 13, 387–401. [Google Scholar] [CrossRef]
  9. Moncrieff, J.; Cooper, R.E.; Stockmann, T.; Amendola, S.; Hengartner, M.P.; Horowitz, M.A. The serotonin theory of depression: A systematic umbrella review of the evidence. Mol. Psychiatry 2022. [Google Scholar] [CrossRef]
  10. Kendrick, T.; Collinson, S. Antidepressants and the serotonin hypothesis of depression. BMJ 2022, 378, o1993. [Google Scholar] [CrossRef]
  11. Bhutani, P.; Joshi, G.; Raja, N.; Bachhav, N.; Rajanna, P.K.; Bhutani, H.; Paul, A.T.; Kumar, R. U.S. FDA Approved Drugs from 2015–June 2020: A Perspective. J. Med. Chem. 2021, 64, 2339–2381. [Google Scholar] [CrossRef]
  12. Hu, F.; Zhang, L.; Nandakumar, K.S.; Cheng, K. Imidazole Scaffold Based Compounds in the Development of Therapeutic Drugs. Curr. Top. Med. Chem. 2021, 21, 2514–2528. [Google Scholar] [CrossRef]
  13. Zhang, L.; Peng, X.M.; Damu, G.L.; Geng, R.X.; Zhou, C.H. Comprehensive review in current developments of imidazole-based medicinal chemistry. Med. Res. Rev. 2014, 34, 340–437. [Google Scholar] [CrossRef]
  14. Hossain, M.; Nanda, A.K. A review on heterocyclic: Synthesis and their application in medicinal chemistry of imidazole moiety. Science 2018, 6, 83–94. [Google Scholar] [CrossRef]
  15. Siwach, A.; Verma, P.K. Synthesis and therapeutic potential of imidazole containing compounds. BMC Chem. 2021, 15, 1–69. [Google Scholar] [CrossRef]
  16. Demchenko, S.; Lesyk, R.; Yadlovskyi, O.; Zuegg, J.; Elliott, A.G.; Drapak, I.; Fedchenkova, Y.; Suvorova, Z.; Demchenko, A. Synthesis, antibacterial and antifungal activity of new 3-aryl-5H-pyrrolo [1, 2-a] imidazole and 5H-imidazo [1, 2-a] azepine quaternary salts. Molecules 2021, 26, 4253. [Google Scholar] [CrossRef]
  17. Chahal, S.; Rani, P.; Kiran; Sindhu, J.; Joshi, G.; Ganesan, A.; Kalyaanamoorthy, S.; Mayank; Kumar, P.; Singh, R. Design and Development of COX-II Inhibitors: Current Scenario and Future Perspective. ACS Omega 2023, 8, 17446–17498. [Google Scholar] [CrossRef]
  18. Ali, I.; Lone, M.N.; Aboul-Enein, H.Y. Imidazoles as potential anticancer agents. MedChemComm 2017, 8, 1742–1773. [Google Scholar] [CrossRef]
  19. Dhameliya, T.M.; Patel, K.I.; Tiwari, R.; Vagolu, S.K.; Panda, D.; Sriram, D.; Chakraborti, A.K. Design, synthesis, and biological evaluation of benzo [d] imidazole-2-carboxamides as new anti-TB agents. Bioorganic Chem. 2021, 107, 104538. [Google Scholar] [CrossRef]
  20. Ali, E.M.; Abdel-Maksoud, M.S.; Hassan, R.M.; Mersal, K.I.; Ammar, U.M.; Se-In, C.; He-Soo, H.; Kim, H.-K.; Lee, A.; Lee, K.-T. Design, synthesis and anti-inflammatory activity of imidazol-5-yl pyridine derivatives as p38α/MAPK14 inhibitor. Bioorganic Med. Chem. 2021, 31, 115969. [Google Scholar] [CrossRef]
  21. Eliewi, A.; Al-Garawi, Z.; Al-Kazzaz, F.; Atia, A. Multi target-directed imidazole derivatives for neurodegenerative diseases. J. Phys. Conf. Ser. 2021, 1853, 012066. [Google Scholar] [CrossRef]
  22. Dhingra, A.K.; Chopra, B.; Jain, A.; Chaudhary, J. Imidazole: Multi-targeted therapeutic leads for the management of Alzheimer’s disease. Mini Rev. Med. Chem. 2022, 22, 1352–1373. [Google Scholar] [CrossRef] [PubMed]
  23. Akhtar, A.; Gupta, S.M.; Dwivedi, S.; Kumar, D.; Shaikh, M.F.; Negi, A. Preclinical Models for Alzheimer’s Disease: Past, Present, and Future Approaches. ACS Omega 2022, 7, 47504–47517. [Google Scholar] [CrossRef]
  24. Georgiou, N.; Gkalpinos, V.K.; Katsakos, S.D.; Vassiliou, S.; Tzakos, A.G.; Mavromoustakos, T. Rational Design and Synthesis of AT1R Antagonists. Molecules 2021, 26, 2927. [Google Scholar] [CrossRef] [PubMed]
  25. Hassan, A.Y.; El-Sebaey, S.A.; El Deeb, M.A.; Elzoghbi, M.S. Potential antiviral and anticancer effect of imidazoles and bridgehead imidazoles generated by HPV-Induced cervical carcinomas via reactivating the P53/pRb pathway and inhibition of CA IX. J. Mol. Struct. 2021, 1230, 129865. [Google Scholar] [CrossRef]
  26. Modh, P.G.; Patel, L.J. Synthesis, Drug Likeness and In-vitro Screening of Some Novel Quinazolinone Derivatives for Anti-Obesity Activity. J. Pharm. Res. Int. 2021, 33, 81–92. [Google Scholar] [CrossRef]
  27. Adeyemi, O.S.; Eseola, A.O.; Plass, W.; Kato, K.; Otuechere, C.A.; Awakan, O.J.; Atolani, O.; Otohinoyi, D.A.; Elebiyo, T.C.; Evbuomwan, I.O. The anti-parasite action of imidazole derivatives likely involves oxidative stress but not HIF-1α signaling. Chem.-Biol. Interact. 2021, 349, 109676. [Google Scholar] [CrossRef]
  28. Negi, A.; Alex, J.M.; Amrutkar, S.M.; Baviskar, A.T.; Joshi, G.; Singh, S.; Banerjee, U.C.; Kumar, R. Imine/amide–imidazole conjugates derived from 5-amino-4-cyano-N1-substituted benzyl imidazole: Microwave-assisted synthesis and anticancer activity via selective topoisomerase-II-α inhibition. Bioorganic Med. Chem. 2015, 23, 5654–5661. [Google Scholar] [CrossRef] [PubMed]
  29. Cohen, N.A.; Stewart, M.L.; Gavathiotis, E.; Tepper, J.L.; Bruekner, S.R.; Koss, B.; Opferman, J.T.; Walensky, L.D. A competitive stapled peptide screen identifies a selective small molecule that overcomes MCL-1-dependent leukemia cell survival. Chem. Biol. 2012, 19, 1175–1186. [Google Scholar] [CrossRef] [PubMed]
  30. Abou Samra, A.; Robert, A.; Gov, C.; Favre, L.; Eloy, L.; Jacquet, E.; Bignon, J.; Wiels, J.; Desrat, S.; Roussi, F. Dual inhibitors of the pro-survival proteins Bcl-2 and Mcl-1 derived from natural compound meiogynin A. Eur. J. Med. Chem. 2018, 148, 26–38. [Google Scholar] [CrossRef]
  31. Litaudon, M.; Bousserouel, H.; Awang, K.; Nosjean, O.; Martin, M.-T.; Dau, M.E.T.H.; Hadi, H.A.; Boutin, J.A.; Sevenet, T.; Gueritte, F. A dimeric sesquiterpenoid from a Malaysian Meiogyne as a new inhibitor of Bcl-xL/BakBH3 domain peptide interaction. J. Nat. Prod. 2009, 72, 480–483. [Google Scholar] [CrossRef]
  32. Negi, A.; Murphy, P.V. Natural products as Mcl-1 inhibitors: A comparative study of experimental and computational modelling data. Chemistry 2022, 4, 983–1009. [Google Scholar] [CrossRef]
  33. Wang, G.; Wang, Y.; Wang, L.; Han, L.; Hou, X.; Fu, H.; Fang, H. Design, synthesis and preliminary bioactivity studies of imidazolidine-2, 4-dione derivatives as Bcl-2 inhibitors. Bioorganic Med. Chem. 2015, 23, 7359–7365. [Google Scholar] [CrossRef] [PubMed]
  34. Negi, A.; Murphy, P.V. Development of Mcl-1 inhibitors for cancer therapy. Eur. J. Med. Chem. 2021, 210, 113038. [Google Scholar] [CrossRef] [PubMed]
  35. Negi, A.; Ramarao, P.; Kumar, R. Recent advancements in small molecule inhibitors of insulin–like growth factor-1 receptor (IGF-1R) tyrosine kinase as anticancer agents. Mini Rev. Med. Chem. 2013, 13, 653–681. [Google Scholar] [CrossRef] [PubMed]
  36. Peled, N.; Wynes, M.W.; Ikeda, N.; Ohira, T.; Yoshida, K.; Qian, J.; Ilouze, M.; Brenner, R.; Kato, Y.; Mascaux, C. Insulin-like growth factor-1 receptor (IGF-1R) as a biomarker for resistance to the tyrosine kinase inhibitor gefitinib in non-small cell lung cancer. Cell. Oncol. 2013, 36, 277–288. [Google Scholar] [CrossRef]
  37. Osmaniye, D.; Levent, S.; Sağlık, B.N.; Karaduman, A.B.; Özkay, Y.; Kaplancıklı, Z.A. Novel imidazole derivatives as potential aromatase and monoamine oxidase-B inhibitors against breast cancer. New J. Chem. 2022, 46, 7442–7451. [Google Scholar] [CrossRef]
  38. Çetiner, G.; Çevik, U.A.; Celik, I.; Bostancı, H.E.; Özkay, Y.; Kaplancıklı, Z.A. New imidazole derivatives as aromatase inhibitor: Design, synthesis, biological activity, molecular docking, and computational ADME-Tox studies. J. Mol. Struct. 2023, 1278, 134920. [Google Scholar] [CrossRef]
  39. Makar, S.; Saha, T.; Swetha, R.; Gutti, G.; Kumar, A.; Singh, S.K. Rational approaches of drug design for the development of selective estrogen receptor modulators (SERMs), implicated in breast cancer. Bioorganic Chem. 2020, 94, 103380. [Google Scholar] [CrossRef]
  40. Xie, B.; Xu, B.; Xin, L.; Wei, Y.; Guo, X.; Dong, C. Discovery of estrogen receptor α targeting caged hypoxia-responsive PROTACs with an inherent bicyclic skeleton for breast cancer treatment. Bioorganic Chem. 2023, 137, 106590. [Google Scholar] [CrossRef]
  41. Negi, A.; Kesari, K.K.; Voisin-Chiret, A.S. Estrogen Receptor-α Targeting: PROTACs, SNIPERs, Peptide-PROTACs, Antibody Conjugated PROTACs and SNIPERs. Pharmaceutics 2022, 14, 2523. [Google Scholar] [CrossRef]
  42. Li, S.-R.; Tan, Y.-M.; Zhang, L.; Zhou, C.-H. Comprehensive insights into medicinal research on imidazole-based supramolecular complexes. Pharmaceutics 2023, 15, 1348. [Google Scholar] [CrossRef]
  43. Ralhan, R.; Kaur, J. Alkylating agents and cancer therapy. Expert Opin. Ther. Pat. 2007, 17, 1061–1075. [Google Scholar] [CrossRef]
  44. Frezza, M.; Hindo, S.; Chen, D.; Davenport, A.; Schmitt, S.; Tomco, D.; Ping Dou, Q. Novel metals and metal complexes as platforms for cancer therapy. Curr. Pharm. Des. 2010, 16, 1813–1825. [Google Scholar] [CrossRef] [PubMed]
  45. Zhou, C.-H.; Zhang, Y.-Y.; Yan, C.-Y.; Wan, K.; Gan, L.-L.; Shi, Y. Recent researches in metal supramolecular complexes as anticancer agents. Anti-Cancer Agents Med. Chem. (Former. Curr. Med. Chem.-Anti-Cancer Agents) 2010, 10, 371–395. [Google Scholar] [CrossRef]
  46. Liu, P.; Jia, J.; Zhao, Y.; Wang, K.-Z. Recent advances on dark and light-activated cytotoxity of imidazole-containing ruthenium complexes. Mini Rev. Med. Chem. 2016, 16, 272–289. [Google Scholar] [CrossRef] [PubMed]
  47. Liu, C.; Yang, W.; Du, J.; Shen, P.; Yang, C. A Boron 2-(2′-pyridyl) Imidazole Fluorescence Probe for Bovine Serum Albumin: Discrimination over Other Proteins and Identification of Its Denaturation. Photochem. Photobiol. 2017, 93, 1414–1422. [Google Scholar] [CrossRef]
  48. Liu, H.-Y.; Zhang, S.-Q.; Cui, M.-C.; Gao, L.-H.; Zhao, H.; Wang, K.-Z. pH-sensitive near-IR emitting dinuclear ruthenium complex for recognition, two-photon luminescent imaging, and subcellular localization of cancer cells. ACS Appl. Bio Mater. 2020, 3, 5420–5427. [Google Scholar] [CrossRef]
  49. Zhu, Y.; Xu, C.; Wang, Y.; Chen, Y.; Ding, X.; Yu, B. Luminescent detection of the lipopolysaccharide endotoxin and rapid discrimination of bacterial pathogens using cationic platinum (ii) complexes. RSC Adv. 2017, 7, 32632–32636. [Google Scholar] [CrossRef]
  50. Okda, H.E.; El Sayed, S.; Ferreira, R.C.; Costa, S.P.; Raposo, M.M.; Martinez-Manez, R.; Sancenon, F. 4-(4, 5-Diphenyl-1H-imidazole-2-yl)-N, N-dimethylaniline-Cu (II) complex, a highly selective probe for glutathione sensing in water-acetonitrile mixtures. Dye. Pigment. 2018, 159, 45–48. [Google Scholar] [CrossRef]
  51. Zhao, C.; Kong, X.; Shuang, S.; Wang, Y.; Dong, C. An anthraquinone-imidazole-based colorimetric and fluorescent sensor for the sequential detection of Ag+ and biothiols in living cells. Analyst 2020, 145, 3029–3037. [Google Scholar] [CrossRef]
  52. Tian, F.; Jiang, X.; Dou, X.; Wu, Q.; Wang, J.; Song, Y. Design and synthesis of novel adenine fluorescence probe based on Eu (III) complexes with dtpa-bis (guanine) ligand. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2017, 179, 194–200. [Google Scholar] [CrossRef]
  53. Zhu, Y.-Y.; Sun, Q.; Shi, J.-W.; Xia, H.-Y.; Wang, J.-L.; Chen, H.-Y.; He, H.-F.; Shen, L.; Zhao, F.; Zhong, J. A novel triple substituted imidazole fluorescent sensor for Ag+ and its imaging in living cell and zebrafish. J. Photochem. Photobiol. A Chem. 2020, 389, 112244. [Google Scholar] [CrossRef]
  54. Suresh, S.; Bhuvanesh, N.; Raman, A.; Sugumar, P.; Padmanabhan, D.; Easwaramoorthi, S.; Ponnuswamy, M.N.; Kavitha, S.; Nandhakumar, R. Experimental and theoretical studies of imidazole based chemosensor for Palladium and their biological applications. J. Photochem. Photobiol. A Chem. 2019, 385, 112092. [Google Scholar] [CrossRef]
  55. Mehta, P.K.; Oh, E.-T.; Park, H.J.; Lee, K.-H. Ratiometric detection of Cu+ in aqueous buffered solutions and in live cells using fluorescent peptidyl probe to mimic the binding site of the metalloprotein for Cu+. Sens. Actuators B Chem. 2018, 256, 393–401. [Google Scholar] [CrossRef]
  56. Bhattacharya, A.; Mahata, S.; Bandyopadhyay, A.; Mandal, B.B.; Manivannan, V. Application of 2, 4, 5-tris (2-pyridyl) imidazole as ‘turn-off’fluorescence sensor for Cu (II) and Hg (II) ions and in vitro cell imaging. Luminescence 2022, 37, 883–891. [Google Scholar] [CrossRef] [PubMed]
  57. Yin, H.; Cheng, F.; Liu, Z.; He, C.; Yang, Y.; Wang, K. Preparation, characterization, pH titration, and electrochemical properties of an anthracene-bridged binuclear ruthenium complex containing imidazole. J. Coord. Chem. 2019, 72, 2957–2967. [Google Scholar] [CrossRef]
  58. Niu, L.; Liu, J.; Gao, S.; Gao, J.; Zhou, Y.; Liu, S.; Ma, C.; Zhao, Y. Fluoride ions detection in aqueous media by unprecedented ring opening of fluorescein dye: A novel multimodal sensor for fluoride ions and its utilization in live cell imaging. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2023, 287, 122001. [Google Scholar] [CrossRef]
  59. Zhou, L.; Chen, Y.; Shao, B.; Cheng, J.; Li, X. Recent advances of small-molecule fluorescent probes for detecting biological hydrogen sulfide. Front. Chem. Sci. Eng. 2022, 16, 34–63. [Google Scholar] [CrossRef]
  60. Strianese, M.; Brenna, S.; Ardizzoia, G.A.; Guarnieri, D.; Lamberti, M.; D’Auria, I.; Pellecchia, C. Imidazo-pyridine-based zinc (ii) complexes as fluorescent hydrogen sulfide probes. Dalton Trans. 2021, 50, 17075–17085. [Google Scholar] [CrossRef] [PubMed]
  61. Rabha, M.; Sen, B.; Sheet, S.K.; Aguan, K.; Khatua, S. Cyclometalated iridium (iii) complex of a 1, 2, 3-triazole-based ligand for highly selective sensing of pyrophosphate ion. Dalton Trans. 2022, 51, 11372–11380. [Google Scholar] [CrossRef] [PubMed]
  62. Lai, Q.; Liu, Q.; He, Y.; Zhao, K.; Wei, C.; Wojtas, L.; Shi, X.; Song, Z. Triazole-imidazole (TA-IM) derivatives as ultrafast fluorescent probes for selective Ag+ detection. Org. Biomol. Chem. 2018, 16, 7801–7805. [Google Scholar] [CrossRef] [PubMed]
  63. Li, G.; Gao, G.; Cheng, J.; Chen, X.; Zhao, Y.; Ye, Y. Two new reversible naphthalimide-based fluorescent chemosensors for Hg2+. Luminescence 2016, 31, 992–996. [Google Scholar] [CrossRef] [PubMed]
  64. Gopalakrishnan, A.K.; Angamaly, S.A.; Pradeep, S.D.; Madhusoodhanan, D.T.; Manoharan, D.K.; Mohanan, P.V. A Novel Imidazole Bound Schiff Base as Highly Selective “Turn-on” Fluorescence Sensor for Zn2+ and Colorimetric Kit for Co2+. J. Fluoresc. 2022, 32, 189–202. [Google Scholar] [CrossRef]
  65. Pandith, A.; Uddin, N.; Choi, C.H.; Kim, H.-S. Highly selective imidazole-appended 9, 10-N, N′-diaminomethylanthracene fluorescent probe for switch-on Zn2+ detection and switch-off H2PO4− and CN− detection in 80% aqueous DMSO, and applications to sequential logic gate operations. Sens. Actuators B Chem. 2017, 247, 840–849. [Google Scholar] [CrossRef]
  66. Kırpık, H.; Kose, M.; Ballı, J.N. Tridentate benzimidazole ligand and its metal complexes: Synthesis, characterization, photo physical and sensor properties. Appl. Organomet. Chem. 2020, 34, e5992. [Google Scholar] [CrossRef]
  67. Mahnashi, M.H.; Mahmoud, A.M.; Alkahtani, S.A.; Ali, R.; El-Wekil, M.M. A novel imidazole derived colorimetric and fluorometric chemosensor for bifunctional detection of copper (II) and sulphide ions in environmental water samples. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2020, 228, 117846. [Google Scholar] [CrossRef]
  68. Pan, J.; Yu, J.; Qiu, S.; Zhu, A.; Liu, Y.; Ban, X.; Li, W.; Yu, H.; Li, L. A novel dibenzimidazole-based fluorescent probe with high sensitivity and selectivity for copper ions. J. Photochem. Photobiol. A Chem. 2021, 406, 113018. [Google Scholar] [CrossRef]
  69. Yin, Y.; Zhang, S.; He, X.; Xu, X.; Zhang, G.; Yang, L.; Kong, L.; Yang, J. A novel tetraphenylethylene-functionalized arylimidazole AIEgen for detections of picric acid and Cu2+. Chem. Pap. 2021, 75, 6297–6306. [Google Scholar] [CrossRef]
  70. Shi, Y.; Chen, X.; Mi, Z.; Zheng, R.; Fan, J.; Gu, Q.; Zhang, Y. A New Tetrasubstituted Imidazole Based Difunctional Probe for UV-spectrophotometric and Fluorometric Detecting of Fe 3+ Ion in Aqueous Solution. Chem. Res. Chin. Univ. 2019, 35, 200–208. [Google Scholar] [CrossRef]
  71. Daly, M.; Sutin, A.R.; Robinson, E. Depression reported by US adults in 2017–2018 and March and April 2020. J. Affect. Disord. 2021, 278, 131–135. [Google Scholar] [CrossRef]
  72. Sahu, B.; Bhatia, R.; Kaur, D.; Choudhary, D.; Rawat, R.; Sharma, S.; Kumar, B. Design, synthesis and biological evaluation of oxadiazole clubbed piperazine derivatives as potential antidepressant agents. Bioorganic Chem. 2023, 136, 106544. [Google Scholar] [CrossRef]
  73. Kumar, B.; Prakash Gupta, V.; Kumar, V. A perspective on monoamine oxidase enzyme as drug target: Challenges and opportunities. Curr. Drug Targets 2017, 18, 87–97. [Google Scholar] [CrossRef] [PubMed]
  74. Kumar, B.; Mantha, A.K.; Kumar, V. Recent developments on the structure–activity relationship studies of MAO inhibitors and their role in different neurological disorders. RSC Adv. 2016, 6, 42660–42683. [Google Scholar] [CrossRef]
  75. Priyadarshini, R.; Raj, G.M. Histamine, Serotonin, Bradykinin, and the Ergot Alkaloids. In Introduction to Basics of Pharmacology and Toxicology: Volume 2: Essentials of Systemic Pharmacology: From Principles to Practice; Springer: Berlin/Heidelberg, Germany, 2021; Volume 2, p. 283. [Google Scholar]
  76. Paul, A.; Anandabaskar, N.; Mathaiyan, J.; Raj, G.M. Introduction to Basics of Pharmacology and Toxicology: Volume 2: Essentials of Systemic Pharmacology: From Principles to Practice; Springer Nature: Berlin/Heidelberg, Germany, 2021. [Google Scholar]
  77. Charvériat, M.; Guiard, B.P. Serotonergic neurons in the treatment of mood disorders: The dialogue with astrocytes. In Progress in Brain Research; Elsevier: Amsterdam, The Netherlands, 2021; Volume 259, pp. 197–228. [Google Scholar]
  78. Park, J.; Jeong, W.; Yun, C.; Kim, H.; Oh, C.-M. Serotonergic Regulation of Hepatic Energy Metabolism. Endocrinol. Metab. 2021, 36, 1151. [Google Scholar] [CrossRef]
  79. Shah, P.A.; Park, C.J.; Shaughnessy, M.P.; Cowles, R.A. Serotonin as a mitogen in the gastrointestinal tract: Revisiting a familiar molecule in a new role. Cell. Mol. Gastroenterol. Hepatol. 2021, 12, 1093–1104. [Google Scholar] [CrossRef]
  80. Kanova, M.; Kohout, P. Serotonin—Its Synthesis and Roles in the Healthy and the Critically Ill. Int. J. Mol. Sci. 2021, 22, 4837. [Google Scholar] [CrossRef]
  81. Zweckstetter, M.; Dityatev, A.; Ponimaskin, E. Structure of serotonin receptors: Molecular underpinning of receptor activation and modulation. Signal Transduct. Target. Ther. 2021, 6, 243. [Google Scholar] [CrossRef]
  82. 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] [PubMed]
  83. Healy, D. Serotonin and depression. BMJ 2015, 350, h1771. [Google Scholar] [CrossRef] [PubMed]
  84. Murphy, S.E.; Capitão, L.P.; Giles, S.L.; Cowen, P.J.; Stringaris, A.; Harmer, C.J. The knowns and unknowns of SSRI treatment in young people with depression and anxiety: Efficacy, predictors, and mechanisms of action. Lancet Psychiatry 2021, 8, 824–835. [Google Scholar] [CrossRef]
  85. Zoega, H.; Kieler, H.; Nørgaard, M.; Furu, K.; Valdimarsdottir, U.; Brandt, L.; Haglund, B. Use of SSRI and SNRI antidepressants during pregnancy: A population-based study from Denmark, Iceland, Norway and Sweden. PLoS ONE 2015, 10, e0144474. [Google Scholar] [CrossRef] [PubMed]
  86. Stahl, M.S.; Lee-Zimmerman, C.; Cartwright, S.; Ann Morrissette, D. Serotonergic drugs for depression and beyond. Curr. Drug Targets 2013, 14, 578–585. [Google Scholar] [CrossRef] [PubMed]
  87. Zheng, G.; Xue, W.; Wang, P.; Yang, F.; Li, B.; Li, X.; Li, Y.; Yao, X.; Zhu, F. Exploring the inhibitory mechanism of approved selective norepinephrine reuptake inhibitors and reboxetine enantiomers by molecular dynamics study. Sci. Rep. 2016, 6, 1–13. [Google Scholar] [CrossRef] [PubMed]
  88. Lovinger, D.M. Communication networks in the brain: Neurons, receptors, neurotransmitters, and alcohol. Alcohol Res. Health 2008, 31, 196–214. [Google Scholar]
  89. Nichols, D.E.; Nichols, C.D. Serotonin receptors. Chem. Rev. 2008, 108, 1614–1641. [Google Scholar] [CrossRef]
  90. He, J.H.; Liu, R.P.; Peng, Y.M.; Guo, Q.; Zhu, L.B.; Lian, Y.Z.; Hu, B.L.; Fan, H.H.; Zhang, X.; Zhu, J.H. Differential and paradoxical roles of new-generation antidepressants in primary astrocytic inflammation. J. Neuroinflamm. 2021, 18, 47. [Google Scholar] [CrossRef]
  91. Sharma, S.; Kumar, D.; Singh, G.; Monga, V.; Kumar, B. Recent advancements in the development of heterocyclic anti-inflammatory agents. Eur. J. Med. Chem. 2020, 200, 112438. [Google Scholar] [CrossRef]
  92. Singh, K.; Bhatia, R.; Kumar, B.; Singh, G.; Monga, V. Design strategies, chemistry and therapeutic insights of multi-target directed ligands as antidepressant agents. Curr. Neuropharmacol. 2022, 20, 1329–1358. [Google Scholar] [CrossRef]
  93. Kerru, N.; Bhaskaruni, S.V.; Gummidi, L.; Maddila, S.N.; Maddila, S.; Jonnalagadda, S.B. Recent advances in heterogeneous catalysts for the synthesis of imidazole derivatives. Synth. Commun. 2019, 49, 2437–2459. [Google Scholar] [CrossRef]
  94. Naowarojna, N.; Cheng, R.; Lopez, J.; Wong, C.; Qiao, L.; Liu, P. Chemical modifications of proteins and their applications in metalloenzyme studies. Synth. Syst. Biotechnol. 2021, 6, 32–49. [Google Scholar] [CrossRef]
  95. Kaur, R.; Kumar, B.; Dwivedi, A.R.; Kumar, V. Regioselective alkylation of 1, 2, 4-triazole using ionic liquids under microwave conditions. Green Process. Synth. 2016, 5, 233–237. [Google Scholar] [CrossRef]
  96. Thenrajan, T.; Sankar, S.S.; Kundu, S.; Wilson, J. Bimetallic nickel iron zeolitic imidazolate fibers as biosensing platform for neurotransmitter serotonin. Colloid Polym. Sci. 2022, 300, 223–232. [Google Scholar] [CrossRef]
  97. Kamijo, S.; Yamamoto, Y. Recent progress in the catalytic synthesis of imidazoles. Chem. Asian J. 2007, 2, 568–578. [Google Scholar] [CrossRef] [PubMed]
  98. Marzouk, A.A.; Abu-Dief, A.M.; Abdelhamid, A.A. Hydrothermal preparation and characterization of ZnFe2O4 magnetic nanoparticles as an efficient heterogeneous catalyst for the synthesis of multi-substituted imidazoles and study of their anti-inflammatory activity. Appl. Organomet. Chem. 2018, 32, e3794. [Google Scholar] [CrossRef]
  99. Nejatianfar, M.; Akhlaghinia, B.; Jahanshahi, R. Cu (II) immobilized on guanidinated epibromohydrin-functionalized γ-Fe2O3@ TiO2 (γ-Fe2O3@ TiO2-EG-Cu (II)): A highly efficient magnetically separable heterogeneous nanocatalyst for one-pot synthesis of highly substituted imidazoles. Appl. Organomet. Chem. 2018, 32, e4095. [Google Scholar] [CrossRef]
  100. Eidi, E.; Kassaee, M.Z.; Nasresfahani, Z. Synthesis of 2, 4, 5-trisubstituted imidazoles over reusable CoFe2O4 nanoparticles: An efficient and green sonochemical process. Appl. Organomet. Chem. 2016, 30, 561–565. [Google Scholar] [CrossRef]
  101. Maleki, A.; Alrezvani, Z.; Maleki, S. Design, preparation and characterization of urea-functionalized Fe3O4/SiO2 magnetic nanocatalyst and application for the one-pot multicomponent synthesis of substituted imidazole derivatives. Catal. Commun. 2015, 69, 29–33. [Google Scholar] [CrossRef]
  102. Kazemi, M. Reusable nanomagnetic catalysts in synthesis of imidazole scaffolds. Synth. Commun. 2020, 50, 2095–2113. [Google Scholar] [CrossRef]
  103. Negi, A.; Mirallai, S.I.; Konda, S.; Murphy, P.V. An improved method for synthesis of non-symmetric triarylpyridines. Tetrahedron 2022, 121, 132930. [Google Scholar] [CrossRef]
  104. Bertrand, J.; Dostálová, H.; Kryštof, V.; Jorda, R.; Delgado, T.; Castro-Alvarez, A.; Mella, J.; Cabezas, D.; Faúndez, M.; Espinosa-Bustos, C. Design, Synthesis, In Silico Studies and Inhibitory Activity towards Bcr-Abl, BTK and FLT3-ITD of New 2, 6, 9-Trisubstituted Purine Derivatives as Potential Agents for the Treatment of Leukaemia. Pharmaceutics 2022, 14, 1294. [Google Scholar] [CrossRef]
  105. Georgiou, M.; Lougiakis, N.; Tenta, R.; Gioti, K.; Baritaki, S.; Gkaralea, L.-E.; Deligianni, E.; Marakos, P.; Pouli, N.; Stellas, D. Discovery of New 1, 4, 6-Trisubstituted-1 H-pyrazolo [3, 4-b] pyridines with Anti-Tumor Efficacy in Mouse Model of Breast Cancer. Pharmaceutics 2023, 15, 787. [Google Scholar] [CrossRef] [PubMed]
  106. Li, X.; Naeem, A.; Xiao, S.; Hu, L.; Zhang, J.; Zheng, Q. Safety challenges and application strategies for the use of dendrimers in medicine. Pharmaceutics 2022, 14, 1292. [Google Scholar] [CrossRef] [PubMed]
  107. Skwarecki, A.S.; Nowak, M.G.; Milewska, M.J. Synthetic strategies in construction of organic low molecular-weight carrier-drug conjugates. Bioorganic Chem. 2020, 104, 104311. [Google Scholar] [CrossRef] [PubMed]
  108. Girish, Y.R.; Kumar, K.S.S.; Thimmaiah, K.N.; Rangappa, K.S.; Shashikanth, S. ZrO2-β-cyclodextrin catalyzed synthesis of 2, 4, 5-trisubstituted imidazoles and 1, 2-disubstituted benzimidazoles under solvent free conditions and evaluation of their antibacterial study. RSC Adv. 2015, 5, 75533–75546. [Google Scholar] [CrossRef]
  109. Bajpai, S.; Singh, S.; Srivastava, V. Nano zirconia catalysed one-pot synthesis of some novel substituted imidazoles under solvent-free conditions. RSC Adv. 2015, 5, 28163–28170. [Google Scholar] [CrossRef]
  110. Fang, S.; Yu, H.; Yang, X.; Li, J.; Shao, L. Nickel-Catalyzed Construction of 2, 4-Disubstituted Imidazoles via C–C Coupling and C− N Condensation Cascade Reactions. Adv. Synth. Catal. 2019, 361, 3312–3317. [Google Scholar] [CrossRef]
  111. Shi, S.; Xu, K.; Jiang, C.; Ding, Z. ZnCl2-Catalyzed [3+ 2] Cycloaddition of Benzimidates and 2 H-Azirines for the Synthesis of Imidazoles. J. Org. Chem. 2018, 83, 14791–14796. [Google Scholar] [CrossRef]
  112. Man, L.; Copley, R.C.; Handlon, A.L. Thermal and photochemical annulation of vinyl azides to 2-aminoimidazoles. Org. Biomol. Chem. 2019, 17, 6566–6569. [Google Scholar] [CrossRef]
  113. Harisha, M.B.; Dhanalakshmi, P.; Suresh, R.; Kumar, R.R.; Muthusubramanian, S.; Bhuvanesh, N. TMSOTf-Catalysed Synthesis of 2, 4, 5-Trisubstituted Imidazoles from Vinyl Azides and Nitriles. ChemistrySelect 2019, 4, 2954–2958. [Google Scholar] [CrossRef]
  114. Tian, Y.; Qin, M.; Yang, X.; Zhang, X.; Liu, Y.; Guo, X.; Chen, B. Acid-catalyzed synthesis of imidazole derivatives via N-phenylbenzimidamides and sulfoxonium ylides cyclization. Tetrahedron 2019, 75, 2817–2823. [Google Scholar] [CrossRef]
  115. Hogendorf, A.S.; Hogendorf, A.; Popiołek-Barczyk, K.; Ciechanowska, A.; Mika, J.; Satała, G.; Walczak, M.; Latacz, G.; Handzlik, J.; Kieć-Kononowicz, K. Fluorinated indole-imidazole conjugates: Selective orally bioavailable 5-HT7 receptor low-basicity agonists, potential neuropathic painkillers. Eur. J. Med. Chem. 2019, 170, 261–275. [Google Scholar] [CrossRef]
  116. Vanda, D.; Soural, M.; Canale, V.; Chaumont-Dubel, S.; Satała, G.; Kos, T.; Funk, P.; Fülöpová, V.; Lemrova, B.; Koczurkiewicz, P. Novel non-sulfonamide 5-HT6 receptor partial inverse agonist in a group of imidazo [4, 5-b] pyridines with cognition enhancing properties. Eur. J. Med. Chem. 2018, 144, 716–729. [Google Scholar] [CrossRef] [PubMed]
  117. Zagórska, A.; Bucki, A.; Kołaczkowski, M.; Siwek, A.; Głuch-Lutwin, M.; Starowicz, G.; Kazek, G.; Partyka, A.; Wesołowska, A.; Słoczyńska, K. Synthesis and biological evaluation of 2-fluoro and 3-trifluoromethyl-phenyl-piperazinylalkyl derivatives of 1 H-imidazo [2, 1-f] purine-2, 4 (3 H, 8 H)-dione as potential antidepressant agents. J. Enzym. Inhib. Med. Chem. 2016, 31, 10–24. [Google Scholar] [CrossRef] [PubMed]
  118. Kumar, B.; Kumar, N.; Thakur, A.; Kumar, V.; Kumar, R.; Kumar, V. A Review on the Arylpiperazine Derivatives as Potential Therapeutics for the Treatment of Various Neurological Disorders. Curr. Drug Targets 2022, 23, 729–751. [Google Scholar] [CrossRef] [PubMed]
  119. Kumar, R.R.; Sahu, B.; Pathania, S.; Singh, P.K.; Akhtar, M.J.; Kumar, B. Piperazine, a Key Substructure for Antidepressants: Its Role in Developments and Structure-Activity Relationships. ChemMedChem 2021, 16, 1878–1901. [Google Scholar] [CrossRef]
  120. Kumar, R.R.; Kumar, V.; Kaur, D.; Nandi, N.K.; Dwivedi, A.R.; Kumar, V.; Kumar, B. Investigation of indole-3-piperazinyl derivatives as potential antidepressants: Design, synthesis, in-vitro, in-vivo and in-silico analysis. ChemistrySelect 2021, 6, 11276–11284. [Google Scholar] [CrossRef]
  121. Zagórska, A.; Kołaczkowski, M.; Bucki, A.; Siwek, A.; Kazek, G.; Satała, G.; Bojarski, A.J.; Partyka, A.; Wesołowska, A.; Pawłowski, M. Structure–activity relationships and molecular studies of novel arylpiperazinylalkyl purine-2, 4-diones and purine-2, 4, 8-triones with antidepressant and anxiolytic-like activity. Eur. J. Med. Chem. 2015, 97, 142–154. [Google Scholar] [CrossRef]
  122. Zagórska, A.; Jurczyk, S.; Pawłowski, M.; Dybała, M.; Nowak, G.; Tatarczyńska, E.; Nikiforuk, A.; Chojnacka-Wójcik, E. Synthesis and preliminary pharmacological evaluation of imidazo [2, 1-f] purine-2, 4-dione derivatives. Eur. J. Med. Chem. 2009, 44, 4288–4296. [Google Scholar] [CrossRef]
  123. Tokgöz, G.; Demir Özkay, Ü.; Osmaniye, D.; Turan Yücel, N.; Can, Ö.D.; Kaplancıklı, Z.A. Synthesis of Novel Benzazole Derivatives and Evaluation of Their Antidepressant-Like Activities with Possible Underlying Mechanisms. Molecules 2018, 23, 2881. [Google Scholar] [CrossRef]
  124. Czopek, A.; Kołaczkowski, M.; Bucki, A.; Byrtus, H.; Pawłowski, M.; Siwek, A.; Bojarski, A.J.; Bednarski, M.; Wróbel, D.; Wesołowska, A. Novel mannich bases, 5-arylimidazolidine-2, 4-dione derivatives with dual 5-HT1A receptor and serotonin transporter affinity. Arch. Der Pharm. 2013, 346, 98–109. [Google Scholar] [CrossRef]
  125. Seo, H.J.; Park, E.-J.; Kim, M.J.; Kang, S.Y.; Lee, S.H.; Kim, H.J.; Lee, K.N.; Jung, M.E.; Lee, M.; Kim, M.-S. Design and synthesis of novel arylpiperazine derivatives containing the imidazole core targeting 5-HT2A receptor and 5-HT transporter. J. Med. Chem. 2011, 54, 6305–6318. [Google Scholar] [CrossRef] [PubMed]
  126. Ulak, G.; Mutlu, O.; Akar, F.Y.; Komsuoğlu, F.I.; Tanyeri, P.; Erden, B.F. Neuronal NOS inhibitor 1-(2-trifluoromethylphenyl)-imidazole augment the effects of antidepressants acting via serotonergic system in the forced swimming test in rats. Pharmacol. Biochem. Behav. 2008, 90, 563–568. [Google Scholar] [CrossRef]
  127. Sherwin, E.; Gigliucci, V.; Harkin, A. Regional specific modulation of neuronal activation associated with nitric oxide synthase inhibitors in an animal model of antidepressant activity. Behav. Brain Res. 2017, 316, 18–28. [Google Scholar] [CrossRef] [PubMed]
  128. Lauro, F.V.; Francisco, D.C.; Ricardo, G.; Heidari, A.; Maria, L.R.; Virginia, M.A.M.; Magdalena, A.R.; Tomas, L.G. Design and synthesis of two Strychnidin-oxiran-naphthalenol derivatives and their theoretical evaluation as noradrenaline and serotonin reuptake inhibitors. Vietnam J. Chem. 2022, 60, 245–256. [Google Scholar]
  129. Czopek, A.; Partyka, A.; Bucki, A.; Pawłowski, M.; Kołaczkowski, M.; Siwek, A.; Głuch-Lutwin, M.; Koczurkiewicz, P.; Pękala, E.; Jaromin, A. Impact of N-Alkylamino Substituents on Serotonin Receptor (5-HTR) Affinity and Phosphodiesterase 10A (PDE10A) Inhibition of Isoindole-1, 3-dione Derivatives. Molecules 2020, 25, 3868. [Google Scholar] [CrossRef]
  130. Żmudzki, P.; Satała, G.; Chłoń-Rzepa, G.; Bojarski, A.J.; Kazek, G.; Siwek, A.; Gryboś, A.; Głuch-Lutwin, M.; Wesołowska, A.; Pawłowski, M. Structure–5-HT/D2 Receptor Affinity Relationship in a New Group of 1-Arylpiperazynylalkyl Derivatives of 8-Dialkylamino-3, 7-dimethyl-1H-purine-2, 6 (3H, 7H)-dione. Arch. Der Pharm. 2016, 349, 774–784. [Google Scholar] [CrossRef] [PubMed]
  131. Hogendorf, A.S.; Hogendorf, A.; Kurczab, R.; Kalinowska-Tłuścik, J.; Popik, P.; Nikiforuk, A.; Krawczyk, M.; Satała, G.; Lenda, T.; Knutelska, J. 2-Aminoimidazole-based antagonists of the 5-HT6 receptor–A new concept in aminergic GPCR ligand design. Eur. J. Med. Chem. 2019, 179, 1–15. [Google Scholar] [CrossRef]
  132. Bromidge, S.M.; Arban, R.; Bertani, B.; Bison, S.; Borriello, M.; Cavanni, P.; Dal Forno, G.; Di-Fabio, R.; Donati, D.; Fontana, S. Design and Synthesis of Novel Tricyclic Benzoxazines as Potent 5-HT1A/B/D Receptor Antagonists Leading to the Discovery of 6-{2-[4-(2-methyl-5-quinolinyl)-1-piperazinyl] ethyl}-4 H-imidazo [5, 1-c][1, 4] benzoxazine-3-carboxamide (GSK588045). J. Med. Chem. 2010, 53, 5827–5843. [Google Scholar] [CrossRef]
  133. Wilkinson, S.T.; Sanacora, G. A new generation of antidepressants: An update on the pharmaceutical pipeline for novel and rapid-acting therapeutics in mood disorders based on glutamate/GABA neurotransmitter systems. Drug Discov. Today 2019, 24, 606–615. [Google Scholar] [CrossRef]
  134. El-Haggar, S.M.; Eissa, M.A.; Mostafa, T.M.; El-Attar, K.S.; Abdallah, M.S. The phosphodiesterase inhibitor pentoxifylline as a novel adjunct to antidepressants in major depressive disorder patients: A proof-of-concept, randomized, double-blind, placebo-controlled trial. Psychother. Psychosom. 2018, 87, 331–339. [Google Scholar] [CrossRef]
  135. Bah, T.M.; Kaloustian, S.; Rousseau, G.; Godbout, R. Pretreatment with pentoxifylline has antidepressant-like effects in a rat model of acute myocardial infarction. Behav. Pharmacol. 2011, 22, 779–784. [Google Scholar] [CrossRef]
  136. Yasrebi, S.-O.; Momtazmanesh, S.; Moghaddam, H.S.; Shahmansouri, N.; Mehrpooya, M.; Arbabi, M.; Ghazizadeh-Hashemi, F.; Akhondzadeh, S. Pentoxifylline for treatment of major depression after percutaneous coronary intervention or coronary artery bypass grafting: A randomized, double-blind, placebo-controlled trial. J. Psychosom. Res. 2021, 150, 110635. [Google Scholar] [CrossRef] [PubMed]
  137. Abdollahi, M.H.; Izadi, A.; Hajiesmaeili, M.R.; Ghanizadeh, A.; Dastjerdi, G.; Hosseini, H.A.; Ghiamat, M.M.; Abbasi, H.R. Effect of etomidate versus thiopental on major depressive disorder in electroconvulsive therapy, a randomized double-blind controlled clinical trial. J. ECT 2012, 28, 10–13. [Google Scholar] [CrossRef] [PubMed]
  138. Wilson, A.C.; Vadakkadath Meethal, S.; Bowen, R.L.; Atwood, C.S. Leuprolide acetate: A drug of diverse clinical applications. Expert Opin. Investig. Drugs 2007, 16, 1851–1863. [Google Scholar] [CrossRef]
  139. Lindemann, L.; Porter, R.H.; Scharf, S.H.; Kuennecke, B.; Bruns, A.; von Kienlin, M.; Harrison, A.C.; Paehler, A.; Funk, C.; Gloge, A. Pharmacology of basimglurant (RO4917523, RG7090), a unique metabotropic glutamate receptor 5 negative allosteric modulator in clinical development for depression. J. Pharmacol. Exp. Ther. 2015, 353, 213–233. [Google Scholar] [CrossRef] [PubMed]
  140. Stedenfeld, K.A.; Clinton, S.M.; Kerman, I.A.; Akil, H.; Watson, S.J.; Sved, A.F. Candesartan reverses depression-like behavior in a rodent model of depression. FASEB J. 2010, 24, 1052.3. [Google Scholar] [CrossRef]
  141. Luo, C.; Fan, H.; Li, S.; Zou, Y. Therapeutic of Candesartan and Music Therapy in Diabetic Retinopathy with Depression in Rats. Evid. -Based Complement. Altern. Med. 2021, 2021, 5570356. [Google Scholar] [CrossRef]
  142. Rizvi, S.J.; Kennedy, S.H. Emerging drugs for major depressive disorder: An update. Expert Opin. Emerg. Drugs 2012, 17, 285–294. [Google Scholar] [CrossRef]
  143. Ar, M.; Ozbalak, M.; Tuzuner, N.; Bekoz, H.; Ozer, O.; Ugurlu, K.; Tabak, F.; Ferhanoglu, B. Severe bone marrow failure due to valganciclovir overdose after renal transplantation from cadaveric donors: Four consecutive cases. Transplant. Proc. 2009, 41, 1648–1653. [Google Scholar] [CrossRef]
  144. Treggiari-Venzi, M.; Borgeat, A.; Fuchs-Buder, T.; Gachoud, J.-P.; Suter, P.M. Overnight sedation with midazolam or propofol in the ICU: Effects on sleep quality, anxiety and depression. Intensive Care Med. 1996, 22, 1186–1190. [Google Scholar] [CrossRef]
  145. Grunebaum, M.F.; Galfalvy, H.C.; Choo, T.-H.; Keilp, J.G.; Moitra, V.K.; Parris, M.S.; Marver, J.E.; Burke, A.K.; Milak, M.S.; Sublette, M.E. Ketamine for rapid reduction of suicidal thoughts in major depression: A midazolam-controlled randomized clinical trial. Am. J. Psychiatry 2018, 175, 327–335. [Google Scholar] [CrossRef]
  146. Grunebaum, M.F.; Ellis, S.P.; Keilp, J.G.; Moitra, V.K.; Cooper, T.B.; Marver, J.E.; Burke, A.K.; Milak, M.S.; Sublette, M.E.; Oquendo, M.A. Ketamine versus midazolam in bipolar depression with suicidal thoughts: A pilot midazolam-controlled randomized clinical trial. Bipolar Disord. 2017, 19, 176–183. [Google Scholar] [CrossRef] [PubMed]
  147. Albert, K.; Bin, L.; Ray, M.D.; Michael, S.J.; Deyi, Z. Imidazole Carboxamides. Indian Patent 271995, 18 March 2016. [Google Scholar]
  148. Lee, J.; Seo, H.J.; Kang, S.Y.; Park, E.-J.; Kim, M.J.; Lee, S.H.; Kim, J.Y.; Kim, J.; Jung, M.E.; Kim, H.J.; et al. Arylpiperazine-Containing Imidazole 4-carboxamide Derivatives and Pharmaceutical Composition Comprising Same. U.S. Patent US 8,835,436, 16 September 2014. [Google Scholar]
  149. Ceccarelli, S.M.; Jagasia, R.; Jakob-roetne, R.; Wichmann, J. Benzimidazoles as CNS Active Agents. U.S. Patent US 20,150,203,472A, 23 July 2015. [Google Scholar]
  150. Schwartz, J.-C.; Lecomte, J.-M. Combination Product Comprising an Antagonist or Inverse Agonist of Histamine Receptor H3 and an Antipsychotic and Antidepressant agent, and Use Thereof for the Preparation of a Medicament That Prevents the Adverse Effects of Psychotropic Drugs. U.S. Patent US8,106,041B2, 31 January 2012. [Google Scholar]
  151. Thurkauf, A.; Horvath, R.F.; Yuan, J.; Peterson, J.M. Certain 4-Aminomethyl-2-substituted Imidazole Derivatives and 2-Aminomethyl-4-substituted Imidazole Derivatives; New Classes of Dopamine Receptor Subtype Specific Ligands. U.S. Patent US6,797,824B2, 28 September 2004. [Google Scholar]
  152. Borroto-Escuela, D.O.; Li, X.; Tarakanov, A.O.; Savelli, D.; Narváez, M.; Shumilov, K.; Andrade-Talavera, Y.; Jimenez-Beristain, A.; Pomierny, B.; Díaz-Cabiale, Z. Existence of brain 5-HT1A–5-HT2A isoreceptor complexes with antagonistic allosteric receptor–receptor interactions regulating 5-HT1A receptor recognition. ACS Omega 2017, 2, 4779–4789. [Google Scholar] [CrossRef] [PubMed]
  153. Bricker, B.A.; Voshavar, C.; Onyameh, E.K.; Gonela, U.M.; Lin, X.; Swanson, T.L.; Kozell, L.B.; Schmachtenberg, J.L.; Bloom, S.H.; Janowsky, A.J. Enantiomeric Separation, Absolute Configuration by X-ray Crystallographic Analysis, and Functional Evaluation of Enantiomers of the Dual Ligand, SYA0340 at 5-HT1A and 5-HT7A Receptors. ACS Omega 2023, 8, 21736–21744. [Google Scholar] [CrossRef]
  154. Ostrowska, K.; Leśniak, A.; Karczyńska, U.; Jeleniewicz, P.; Głuch-Lutwin, M.; Mordyl, B.; Siwek, A.; Trzaskowski, B.; Sacharczuk, M.; Bujalska-Zadrożny, M. 6-Acetyl-5-hydroxy-4, 7-dimethylcoumarin derivatives: Design, synthesis, modeling studies, 5-HT1A, 5-HT2A and D2 receptors affinity. Bioorganic Chem. 2020, 100, 103912. [Google Scholar] [CrossRef] [PubMed]
  155. Zhou, J.; Reidy, M.; O’Reilly, C.; Jarikote, D.V.; Negi, A.; Samali, A.; Szegezdi, E.; Murphy, P.V. Decorated macrocycles via ring-closing double-reductive amination. Identification of an apoptosis inducer of leukemic cells that at least partially antagonizes a 5-HT2 receptor. Org. Lett. 2015, 17, 1672–1675. [Google Scholar] [CrossRef]
  156. Negi, A.; Reilly, C.O.; Jarikote, D.V.; Zhou, J.; Murphy, P.V. Multi-targeting protein-protein interaction inhibitors: Evolution of macrocyclic ligands with embedded carbohydrates (MECs) to improve selectivity. Eur. J. Med. Chem. 2019, 176, 292–309. [Google Scholar] [CrossRef]
  157. Foley, C.A.; Al-Issa, Y.A.; Hiller, K.P.; Mulcahy, S.P. Synthesis and Structure–Activity Relationships of 1-Aryl-β-carbolines as Affinity Probes for the 5-Hydroxytryptamine Receptor. ACS Omega 2019, 4, 9807–9812. [Google Scholar] [CrossRef]
  158. Singh, D.; Singh, P.; Srivastava, P.; Kakkar, D.; Pathak, M.; Tiwari, A.K. Development and Challenges in the discovery of 5-HT1A and 5-HT7 receptor ligands. Bioorganic Chem. 2023, 131, 106254. [Google Scholar] [CrossRef]
  159. Chakraborty, S.; Rakshit, J.; Bandyopadhyay, J.; Basu, S. Multi-target inhibition ability of neohesperidin dictates its neuroprotective activity: Implication in Alzheimer’s disease therapeutics. Int. J. Biol. Macromol. 2021, 176, 315–324. [Google Scholar] [CrossRef]
  160. Tzschentke, T.M. Measuring reward with the conditioned place preference paradigm: A comprehensive review of drug effects, recent progress and new issues. Prog. Neurobiol. 1998, 56, 613–672. [Google Scholar] [CrossRef]
  161. Ibrahim, M.M.; Gabr, M.T. Multitarget therapeutic strategies for Alzheimer’s disease. Neural Regen. Res. 2019, 14, 437. [Google Scholar]
  162. Jenwitheesuk, E.; Horst, J.A.; Rivas, K.L.; Van Voorhis, W.C.; Samudrala, R. Novel paradigms for drug discovery: Computational multitarget screening. Trends Pharmacol. Sci. 2008, 29, 62–71. [Google Scholar] [CrossRef]
  163. Negi, A.; Zhou, J.; Sweeney, S.; Murphy, P.V. Ligand design for somatostatin receptor isoforms 4 and 5. Eur. J. Med. Chem. 2019, 163, 148–159. [Google Scholar] [CrossRef]
  164. Dantas, D.; Ribeiro, A.I.; Carvalho, F.; Gil-Martins, E.; Silva, R.; Remião, F.; Zille, A.; Cerqueira, F.; Pinto, E.; Dias, A.M. Red-shifted and pH-responsive imidazole-based azo dyes with potent antimicrobial activity. Chem. Commun. 2023, 59, 2791–2794. [Google Scholar] [CrossRef] [PubMed]
  165. Slassi, S.; Aarjane, M.; El-Ghayoury, A.; Allain, M.; Amine, A. Synthesis, crystal structure, photoisomerization, and DFT studies of novel azo compounds based on imidazole. J. Phys. Org. Chem. 2023, 36, e4486. [Google Scholar] [CrossRef]
  166. Negi, A.; Kieffer, C.; Voisin-Chiret, A.S. Azobenzene photoswitches in proteolysis targeting chimeras: Photochemical control strategies and therapeutic benefits. ChemistrySelect 2022, 7, e202200981. [Google Scholar] [CrossRef]
  167. Negi, A.; Kesari, K.K.; Voisin-Chiret, A.S. Light-Activating PROTACs in Cancer: Chemical Design, Challenges, and Applications. Appl. Sci. 2022, 12, 9674. [Google Scholar] [CrossRef]
  168. Narasimhan, B.; Sharma, D.; Kumar, P. Biological importance of imidazole nucleus in the new millennium. Med. Chem. Res. 2011, 20, 1119–1140. [Google Scholar] [CrossRef]
  169. Negi, A.; Voisin-Chiret, A.S. Strategies to reduce the on-target platelet toxicity of Bcl-xL inhibitors: PROTACs, SNIPERs and prodrug-based approaches. ChemBioChem 2022, 23, e202100689. [Google Scholar] [CrossRef]
  170. Singh, P.K.; Negi, A.; Gupta, P.K.; Chauhan, M.; Kumar, R. Toxicophore exploration as a screening technology for drug design and discovery: Techniques, scope and limitations. Arch. Toxicol. 2016, 90, 1785–1802. [Google Scholar] [CrossRef] [PubMed]
  171. Wang, Q.; Sun, Y.; Li, S.; Zhang, P.; Yao, Q. Synthesis and modification of ZIF-8 and its application in drug delivery and tumor therapy. RSC Adv. 2020, 10, 37600–37620. [Google Scholar] [CrossRef] [PubMed]
  172. Zhao, H.; Gong, L.; Wu, H.; Liu, C.; Liu, Y.; Xiao, C.; Liu, C.; Chen, L.; Jin, M.; Gao, Z. Development of Novel Paclitaxel-Loaded ZIF-8 Metal-Organic Framework Nanoparticles Modified with Peptide Dimers and an Evaluation of Its Inhibitory Effect against Prostate Cancer Cells. Pharmaceutics 2023, 15, 1874. [Google Scholar] [CrossRef] [PubMed]
  173. Asadi, V.; Marandi, A.; Kardanpour, R.; Tangestaninejad, S.; Moghadam, M.; Mirkhani, V.; Mohammadpoor-Baltork, I.; Mirzaei, R. Carbonic Anhydrase-Embedded ZIF-8 Electrospun PVA Fibers as an Excellent Biocatalyst Candidate. ACS Omega 2023, 8, 17809–17818. [Google Scholar] [CrossRef]
  174. Xia, Y.; Hong, Y.; Geng, R.; Li, X.; Qu, A.; Zhou, Z.; Zhang, Z. Amine-functionalized ZIF-8 as a fluorescent probe for breath volatile organic compound biomarker detection of lung cancer patients. ACS Omega 2020, 5, 3478–3486. [Google Scholar] [CrossRef]
  175. Costa, B.A.; Abuçafy, M.P.; Barbosa, T.W.L.; da Silva, B.L.; Fulindi, R.B.; Isquibola, G.; da Costa, P.I.; Chiavacci, L.A. ZnO@ ZIF-8 nanoparticles as nanocarrier of ciprofloxacin for antimicrobial activity. Pharmaceutics 2023, 15, 259. [Google Scholar] [CrossRef] [PubMed]
  176. Negi, A.; Kesari, K.K. Chitosan nanoparticle encapsulation of antibacterial essential oils. Micromachines 2022, 13, 1265. [Google Scholar] [CrossRef]
  177. Mujtaba, M.; Negi, A.; King, A.W.; Zare, M.; Kuncova-Kallio, J. Surface modifications of nanocellulose for drug delivery applications; a critical review. Curr. Opin. Biomed. Eng. 2023, 100475. [Google Scholar] [CrossRef]
  178. Ramos, V.C.; Reyes, C.B.G.; García, G.M.; Quesada, M.I.S.; Barrero, F.J.M.-C.; Rábago, J.J.S.; Polo, M.S. ZIF-8 and Its Magnetic Functionalization as Vehicle for the Transport and Release of Ciprofloxacin. Pharmaceutics 2022, 14, 2546. [Google Scholar] [CrossRef]
  179. Gao, Z.; Mansor, M.H.; Winder, N.; Demiral, S.; Maclnnes, J.; Zhao, X.; Muthana, M. Microfluidic-Assisted ZIF-Silk-Polydopamine Nanoparticles as Promising Drug Carriers for Breast Cancer Therapy. Pharmaceutics 2023, 15, 1811. [Google Scholar] [CrossRef]
  180. Franco, A.; Negi, A.; Luque, R.; Carrillo-Carrión, C. Selectivity Control in the Oxidative Ring-Opening of Dimethylfuran Mediated by Zeolitic-Imidazolate Framework-8 Nanoparticles. ACS Sustain. Chem. Eng. 2021, 9, 8090–8096. [Google Scholar] [CrossRef]
  181. Tolomeu, H.V.; Fraga, C.A.M. Imidazole: Synthesis, Functionalization and Physicochemical Properties of a Privileged Structure in Medicinal Chemistry. Molecules 2023, 28, 838. [Google Scholar] [CrossRef]
  182. Zheng, X.; Ma, Z.; Zhang, D. Synthesis of imidazole-based medicinal molecules utilizing the van leusen imidazole synthesis. Pharmaceuticals 2020, 13, 37. [Google Scholar] [CrossRef]
  183. Andrei, G.Ş.; Andrei, B.F.; Roxana, P.R. Imidazole derivatives and their antibacterial activity—A mini-review. Mini Rev. Med. Chem. 2021, 21, 1380–1392. [Google Scholar] [CrossRef]
  184. Parwani, D.; Bhattacharya, S.; Rathore, A.; Mallick, C.; Asati, V.; Agarwal, S.; Rajoriya, V.; Das, R.; Kashaw, S.K. Current insights into the chemistry and antitubercular potential of benzimidazole and imidazole derivatives. Mini Rev. Med. Chem. 2021, 21, 643–657. [Google Scholar] [CrossRef]
Figure 1. Role of serotonin receptor modulators in depression.
Figure 1. Role of serotonin receptor modulators in depression.
Pharmaceutics 15 02208 g001
Figure 2. Recently developed synthetic methodologies for imidazole and its derivatives.
Figure 2. Recently developed synthetic methodologies for imidazole and its derivatives.
Pharmaceutics 15 02208 g002
Figure 3. SAR studies of fluorinated 3-(1-alkyl-1H-imidazol-5-yl)-1H-indoles.
Figure 3. SAR studies of fluorinated 3-(1-alkyl-1H-imidazol-5-yl)-1H-indoles.
Pharmaceutics 15 02208 g003
Figure 4. SAR studies of 3H-imidazo[4,5-b] pyridine and 3H-imidazo[4,5-c] pyridine derivatives.
Figure 4. SAR studies of 3H-imidazo[4,5-b] pyridine and 3H-imidazo[4,5-c] pyridine derivatives.
Pharmaceutics 15 02208 g004
Figure 5. SAR studies of trifluoromethylphenylpiperazinylalkyl derivatives of 1H-imidazo[2,1-f]purine-2,4(3H,8H)-dione.
Figure 5. SAR studies of trifluoromethylphenylpiperazinylalkyl derivatives of 1H-imidazo[2,1-f]purine-2,4(3H,8H)-dione.
Pharmaceutics 15 02208 g005
Figure 6. SAR studies of arylpiperazinylalkyl purine-2,4-diones.
Figure 6. SAR studies of arylpiperazinylalkyl purine-2,4-diones.
Pharmaceutics 15 02208 g006
Figure 7. SAR studies of imidazo[2,1-f]purine-2,4-dione derivatives.
Figure 7. SAR studies of imidazo[2,1-f]purine-2,4-dione derivatives.
Pharmaceutics 15 02208 g007
Figure 8. SAR studies of benzazole derivatives.
Figure 8. SAR studies of benzazole derivatives.
Pharmaceutics 15 02208 g008
Figure 9. SAR studies of imidazolidine-2,4-dione derivatives.
Figure 9. SAR studies of imidazolidine-2,4-dione derivatives.
Pharmaceutics 15 02208 g009
Figure 10. SAR studies of imidazole core containing arylpiperazine-4-carboxamide.
Figure 10. SAR studies of imidazole core containing arylpiperazine-4-carboxamide.
Pharmaceutics 15 02208 g010
Figure 11. Strychnidin-oxiran-napthalenol derivatives.
Figure 11. Strychnidin-oxiran-napthalenol derivatives.
Pharmaceutics 15 02208 g011
Figure 12. SAR studies of imidazole-containing 4-methoxy-1H-isoindole-1,3(2H)-dione derivatives.
Figure 12. SAR studies of imidazole-containing 4-methoxy-1H-isoindole-1,3(2H)-dione derivatives.
Pharmaceutics 15 02208 g012
Figure 13. Structure of arylpiperazynylalkyl derivatives of 8-amino-3,7-dimethyl-1H-purine-2,6(3H,7H)-dione.
Figure 13. Structure of arylpiperazynylalkyl derivatives of 8-amino-3,7-dimethyl-1H-purine-2,6(3H,7H)-dione.
Pharmaceutics 15 02208 g013
Figure 14. SAR studies of imidazole N-(1H imidazol-2-yl)acylamide derived imidazole derivatives.
Figure 14. SAR studies of imidazole N-(1H imidazol-2-yl)acylamide derived imidazole derivatives.
Pharmaceutics 15 02208 g014
Figure 15. SAR studies of novel tricyclic benzoxazine derivatives.
Figure 15. SAR studies of novel tricyclic benzoxazine derivatives.
Pharmaceutics 15 02208 g015
Figure 16. Pharmacophore generated using open web server Pharmit from existing SSRIs. It represents the presence of two aromatic rings (violet sphere), two hydrogen bond acceptors (orange sphere), and a hydrophobic region (green sphere).
Figure 16. Pharmacophore generated using open web server Pharmit from existing SSRIs. It represents the presence of two aromatic rings (violet sphere), two hydrogen bond acceptors (orange sphere), and a hydrophobic region (green sphere).
Pharmaceutics 15 02208 g016
Table 1. Imidazole-based drugs under clinical trials against depression.
Table 1. Imidazole-based drugs under clinical trials against depression.
S. No.CompoundStructureClinical Trial StatusClinical Trial Number
1EVT101Pharmaceutics 15 02208 i001Phase 2
Terminated
NCT01128452
2PentoxifyllinePharmaceutics 15 02208 i002Phases 1 and 2
Completed
NCT05324735
Early Phase 1
Completed
NCT04417049
Not Applicable
Completed
NCT03554447
Phase 1
Completed
NCT05271084
3EtomidatePharmaceutics 15 02208 i003Not Applicable
Completed
NCT02667353
Phase 4
Unknown
NCT02924090
4Leuprolide Acetate Not Applicable
Completed
NCT01762943
Phase 2
Completed
NCT04051320
Phase 4
Completed
NCT01116401
5RO4917523Pharmaceutics 15 02208 i004Phase 2
Completed
NCT00809562
Phase 2
Completed
NCT01437657
6CandesartanPharmaceutics 15 02208 i005Early Phase 1
Withdrawn
NCT04430959
Phase 4
Terminated
NCT01794455
7CimicoxibPharmaceutics 15 02208 i006Phase 2
Completed
NCT00510822
8ValganciclovirPharmaceutics 15 02208 i007Phases 1 and 2
Recruited
NCT04724447
9MidazolamPharmaceutics 15 02208 i008Phase 1
Terminated
NCT04082858
Phase 4
Recruiting
NCT05026203
Phase 2
Completed
NCT02360280
Phase 2
Recruiting
NCT05383313
Not yet recruitingNCT05528718
Phase 4
Completed
NCT01700829
Phase 3
Active, not recruiting
NCT03889756
Phase 4
Recruiting
NCT04220125
Phase 3
Recruiting
NCT04939649
Table 2. Recently granted patents on imidazole-based compounds with antidepressant activities.
Table 2. Recently granted patents on imidazole-based compounds with antidepressant activities.
Patentee Name/InventorsTitlePatent No.YearRef.
Albert K, Bin L, Ray MD, Michael SJ, and Deyi ZImidazole carboxamidesIndian patent No. 2719952016[147]
Lee J, Seo HJ, Kang SY, Park EJ, Kim MJ, Lee SH, Kim JY, Kim J, Jung ME, Kim HJ, and Kim MSArylpiperazine-containing imidazole 4-carboxamide derivatives and a pharmaceutical composition comprising the sameUS 8,835,4362014[148]
Ceccarelli SM, Jagasia R, Jakob-roetne R, and Wichmann JBenzimidazoles as CNS active agents.US 20,150,203,472A2014[149]
Schwartz JC and Lecomte JMCombination product comprising an antagonist or inverse agonist of histamine receptor H3 and an antipsychotic and antidepressant agent, and use thereof for the preparation of a medicament that prevents the adverse effects of psychotropic drugs. US 8,106,0412012[150]
Thurkauf A, Horvath RF, Yuan J, and Peterson JMCertain 4-aminomethyl-2-substituted imidazole derivatives and 2-aminomethyl-4-substituted imidazole derivatives; new classes of dopamine receptor subtype-specific ligands.US 6,797,8242004[151]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Goel, K.K.; Thapliyal, S.; Kharb, R.; Joshi, G.; Negi, A.; Kumar, B. Imidazoles as Serotonin Receptor Modulators for Treatment of Depression: Structural Insights and Structure–Activity Relationship Studies. Pharmaceutics 2023, 15, 2208. https://doi.org/10.3390/pharmaceutics15092208

AMA Style

Goel KK, Thapliyal S, Kharb R, Joshi G, Negi A, Kumar B. Imidazoles as Serotonin Receptor Modulators for Treatment of Depression: Structural Insights and Structure–Activity Relationship Studies. Pharmaceutics. 2023; 15(9):2208. https://doi.org/10.3390/pharmaceutics15092208

Chicago/Turabian Style

Goel, Kapil Kumar, Somesh Thapliyal, Rajeev Kharb, Gaurav Joshi, Arvind Negi, and Bhupinder Kumar. 2023. "Imidazoles as Serotonin Receptor Modulators for Treatment of Depression: Structural Insights and Structure–Activity Relationship Studies" Pharmaceutics 15, no. 9: 2208. https://doi.org/10.3390/pharmaceutics15092208

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

Article Metrics

Back to TopTop