Next Article in Journal
Galactoside-Based Molecule Enhanced Antimicrobial Activity through Acyl Moiety Incorporation: Synthesis and In Silico Exploration for Therapeutic Target
Previous Article in Journal
Cyclic Peptides in Pipeline: What Future for These Great Molecules?
Previous Article in Special Issue
Antibacterial Properties of Mesoporous Silica Nanoparticles Modified with Fluoroquinolones and Copper or Silver Species
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Indol-3-ylglyoxylamide as Privileged Scaffold in Medicinal Chemistry

1
Department of Pharmacy, University of Pisa, Via Bonanno 6, 56126 Pisa, Italy
2
Laboratory of Bioorganic Chemistry, National Institute of Diabetes and Digestive and Kidney Diseases (NIDDK), National Institutes of Health, 9000 Rockville Pike, Bethesda, MD 20892, USA
3
Department of Pharmacy, University of Salerno, 84084 Fisciano, Italy
*
Author to whom correspondence should be addressed.
These authors contributed equally to this study.
Pharmaceuticals 2023, 16(7), 997; https://doi.org/10.3390/ph16070997
Submission received: 12 June 2023 / Revised: 5 July 2023 / Accepted: 10 July 2023 / Published: 12 July 2023
(This article belongs to the Special Issue Privileged Structures as Leads in Medicinal Chemistry 2023)

Abstract

:
In recent years, indolylglyoxylamide-based derivatives have received much attention due to their application in drug design and discovery, leading to the development of a wide array of compounds that have shown a variety of pharmacological activities. Combining the indole nucleus, already validated as a “privileged structure,” with the glyoxylamide function allowed for an excellent template to be obtained that is suitable to a great number of structural modifications aimed at permitting interaction with specific molecular targets and producing desirable therapeutic effects. The present review provides insight into how medicinal chemists have elegantly exploited the indolylglyoxylamide moiety to obtain potentially useful drugs, with a particular focus on compounds exhibiting activity in in vivo models or reaching clinical trials. All in all, this information provides exciting new perspectives on existing data that can be useful in further design of indolylglyoxylamide-based molecules with interesting pharmacological profiles. The aim of this report is to present an update of collection data dealing with the employment of this moiety in the rational design of compounds that are able to interact with a specific target, referring to the last 20 years.

Graphical Abstract

1. Introduction

The concept of a “privileged structure” was proposed in 1988 by Evans and coworkers to define structural moieties able to furnish, with appropriate decorations, small molecules that are able to interact with different targets, receptors, or, more generally, proteins; actually, smart modifications of these structures could represent a fruitful strategy for the design and development of new receptor agonists/antagonists/allosteric modulators, enzyme inhibitors/activators, protein–protein interaction modulators, and so on [1].
The indole scaffold (Figure 1) emerges as one of the most important structural subunits for discovering novel drug candidates [2,3,4]. The evidence that indole constitutes a common structural feature of many biomolecules and natural products such as tryptophan, ergot alkaloids, and the neurotransmitter serotonin triggered extensive study of indole chemistry, leading to the development of a huge number of synthetic indole-based derivatives with biological activity. Many of them became drugs and others are useful for pharmacological purposes, but the majority became lead prototypes, validating the indole scaffold as a privileged structure [5].
A nitrogen lone pair is implicated in maintaining the aromaticity of the indole ring, with the nitrogen proton being very slightly acidic (pKa ~ 17) and able to engage in H-bond interactions with specific target proteins [5]. This interaction may be crucial to the anchoring of indole derivatives to certain molecular targets. As an example, Da Settimo and coworkers highlighted the crucial role of the indole N-H in a series of indolylglyoxylamides with activity on the central benzodiazepine receptor (BzR) [6]. Despite the literature indicating that the N-H group as a donor of hydrogen bonds was not required to have BzR affinity, it was shown that all derivatives featuring a benzofuran and benzothiophene nucleus possessed a lower affinity for the receptor compared to those bearing an indole nucleus [6].
In addition, the indole ring is able to establish noncovalent interactions due to its aromaticity, such as π–π stacking or cation–π interactions [7]. In this respect, two example studies of G protein-coupled receptors (GPCRs) were reported by Bondensgaard and coworkers [8] and Rad and colleagues [9], not only showing the crucial role of the indole scaffold in the molecular recognition of the target receptor but also, most importantly, proposing the biological validation of the indole substructure as a truly privileged scaffold for GPCR targets [8,9]. The concept of “biological validation” was used to distinguish “chemically privileged” substructures from privileged scaffolds that have overcome biological tests, gaining clinical significance [9].
The propensity of amides to establish hydrogen bonds has been extensively studied and exploited in crystal engineering [10]. In this respect, glyoxylamides provide a greater number of features with respect to a simple amide function. They introduce a greater degree of versatility to the system due to the possibility to form an additional hydrogen bond through the keto group and the variable glyoxylamide torsional angle [11].
The combination of the indole nucleus with the glyoxylamide function originated an excellent template, the indolylglyoxylamide moiety (Figure 1), which is suitable to a great number of structural modifications either to produce a specific therapeutic effect or to improve the pharmacokinetic (PK) profiles of biologically active molecules [12]. Actually, the synthetic ease and versatility of the indole nucleus and the potency attributed to glyoxylamides have led to the development of a wide array of indolylglyoxylamide compounds, which have shown different pharmacological effects in medicinal chemistry [12].
The purpose of this report is to provide an update of collection data dealing with the employment of indolylglyoxylamide moiety in the rational design of compounds able to interact with specific targets. Altogether, this information provides exciting new perspectives on existing data that can be useful in further design of indolylglyoxylamide-based molecules with interesting pharmacological profiles.
A synopsis of some of the applications of the indolylglyoxylamides in drug design encompassing different fields of biological activity is outlined in Figure 2.

2. Indolylglyoxylamides with Sedative–Hypnotic Properties

About 10% of the adult population suffers from insomnia, which is often a chronic condition. Insomnia increases with age, and women are more affected with respect to men. Only a few people seek medical help to treat insomnia, remaining unaware of behavioral and medical options available [13]. Barbiturates, often used as sedative hypnotics/anxiolytics, were the first generation of drugs to treat insomnia [14]. Benzodiazepines (Bzs), the second generation, have been used for the treatment of insomnia, and have, unlike barbiturates, lower potential for abuse and less danger of lethal overdose [15]. However, Bzs cause many adverse effects, including cognitive and psychomotor impairment, dependence, tolerance, hangover, and rebound insomnia [16]. Thus, a third generation of hypnotics, such as zolpidem, zopiclone, and zaleplon (Figure 3), was developed to obtain compounds endowed with sleep-inducing action combined with minor adverse effects, including amnesia and motor dysfunction [17]. During long-term treatment with zaleplon, eszopiclone, and a modified release formulation of zolpidem, neither tolerance during treatment nor limited rebound insomnia after therapy discontinuation was observed [18,19]. Unfortunately, when administered at higher doses, the same Bzs’ side effects were observed, whose severity depends on the specific drugs [20].
Bzs allosterically modulate the γ-aminobutyric acid (GABA) affinity for the type A GABA receptor (GABAAR), acting as agonists (positive allosteric modulators) endowed with anxiolytic, anticonvulsant, sedative-hypnotic, and myorelaxant effects.
The main Bz-sensitive GABAAR subtypes in the brain are α1β3γ2, α2β3γ2, α3β3γ2, and α5β3γ2, at variance with the α4β3γ2 and α6β3γ2 subtypes, which do not respond to Bzs [21,22,23,24]. The α subunit is the only one that influences the affinity and efficacy of BzR ligands, because the γ2 and β3 subunits do not vary [21,22,23,24].
Since 1980, numerous research groups have aimed at the development of new ligands with high affinity and selectivity for the different GABAA/BzR subtypes. In particular, selective α1 agonists should represented sedative–hypnotic agents without effects on learning or memory processes, selective α23 agonists should possess anxiolytic activity without sedation, and selective α5 inverse agonists should constitute cognitive enhancers lacking anxiogenic activity [24].
Structure–affinity relationships (SARs) of structurally different classes of BzR ligands were rationalized by Cooks’ research group by means of a pharmacophore/topological receptor model constituted by (i) hydrogen-bond acceptor (A2), (ii) hydrogen-bond donors (H1 and H2), (iii) four lipophilic pockets (L1, L2, L3, and LDi), and (iv) three sterically forbidden sites (S1, S2, and S3) [25].
Some years ago, Da Settimo and coworkers described a class of N-(benzyl)indol-3-ylglyoxylamide BzR ligands [26,27]. These compounds showed higher BzR α1-subtype affinity with respect to the α2 and α5 isoforms. SARs of this class indicated an interdependent effect of substituents at the 5-position and groups on the benzyl ring on α1 affinity. Affinity was favored when the benzyl ring was substituted with OH/OCH3 or halogens, depending on the substitution at the 5-position of the indole, in particular with a chlorine atom or with a nitro group, or not. Thus, whereas in the 5-unsubstitued series the most active compounds were those featuring an electron-attracting substituent (Cl), in the 5-Cl/NO2 series the activity was optimized with 3′,4′-(OH)2 or 3′,4′-(OCH3)2. These results evidenced the existence of two different binding modes, called A and B, for the interaction of these compounds within the receptor site (Figure 4) [27].
In a subsequent study, the authors exploited the difference in dimension between L2 and LDi lipophilic pockets to identify indolylglyoxylamides that were selective toward the different BzR subtypes [28]. Because the LDi and L2 clefts are wider in the α1 and α5 sites, respectively, compared to the other subtypes, full occupation of LDi or L2 may lead to α1 and α5 selective compounds, respectively [28,29]. Based on benzylindolylglyoxylamides 1 and 2 (Figure 5) as lead compounds [26], derivatives featuring different substituents on the benzylamide phenyl ring and compounds in which the benzyl moiety was replaced with alkyl groups were studied with the aim of probing the LDi and L2 pockets [28].
SARs evidenced how compounds featuring the same amide side chain showed different affinity depending on the substitution at the 5-position of the indole ring (NO2, H), confirming the ability of such compounds to adopt two binding poses (A and B) into the receptor binding site (Figure 5). The LDi cleft was wider than the L2 one; indeed, the former could host bulky 4′-substituents of 5-H indoles (binding mode B), whereas the latter could not accommodate the same 4′-substituents of the 5-NO2 counterparts (binding mode A). Some selected compounds were tested for their affinity at the α1, α2, and α5 BzR subtypes together with leads 1 and 2, and all derivatives showed fair affinity and selectivity for α1 with respect to the α2 and α5 subtypes. Compound 3 showed the best α1 affinity and selectivity, displaying Ki = 31.3 nM at the α1 isoform and no affinity at the α2 or α5 isoforms [28]. In line with Cook’s BzR pharmacophore model, the data presented by Da Settimo and coworkers indicated how the full occupancy of the LDi region by a bulky lipophilic group shifted the selectivity and affinity of the ligand towards the α1 subtype, making the interactions with the L1/L2 lipophilic pockets contribute less to the binding and selectivity to other receptor subtypes. A functional assay performed on 3 evidenced a similar efficacy with respect to the standard diazepam in co-application with GABA at the BzR α1 subtype (72 ± 2% and 78 ± 3% for 3 and diazepam, respectively), showing that this compound acts as a full agonist at the α1 subtype [28]. Finally, in vivo behavioral assays based on the observation of the spontaneous motor activity of mice identified 3 as a sedative–hypnotic agent, although less active with respect to zolpidem [28].

3. Indolylglyoxylamides with Anxiolytic-Like and Neuroprotective Properties

As described above, Bzs are the first-line treatment for anxiety disorders, eliciting their action by allosterically binding the GABAAR. If their binding causes an enhancement in GABA inhibitory action, they are called agonists, whereas when it causes a GABA action reduction, they are called partial agonists. Finally, if their binding does not produce any effects, they are called antagonists [30]. Bzs are also classified as full agonists (or full inverse agonists) and as partial agonists (or full inverse agonists), depending on their efficacy toward the different GABAAR subtypes.
As Bzs cause several unwanted side effects, research has focused on the identification of safer anxiolytic agents. Due to their differential localization in brain areas, the different GABAAR subtypes may be associated with distinct physiological effects, specifically (i) α1 subtype moderating sedation, (ii) α2/(α3) subtype mediating anxiolytic and myorelaxation actions, and (iii) α5 subtype being involved with cognition processes [24,31,32].
Based on this, to develop an anxioselective compound [33], either selective binding agonism to α2 (affinity-based selective agents) or agonism to the α2 subtype and antagonism to other isoforms (efficacy-based selective agents) is required. In addition, anxioselective compounds lacking sedation and dependence may interact with all four GABAAR subtypes as partial agonists [30].
With respect to nonselective Bzs, either affinity- or efficacy-based α2 selective agonists should conserve anxiolytic properties with minor collateral effects, namely, sedation, ataxia, tolerance, dependence, and impairment of cognitive processes [24,31]. Aiming at developing novel anxioselective agents, Da Settimo et al. [27,28,34,35] studied a huge number of indol-3-ylglyoxylamides, and the major result was the identification of compounds 4 and 5 (Figure 6), which showed the highest potency at the α2-BzR together with an interesting in vitro profile (full α2 agonism, α1 partial agonism/antagonism, low affinity to α5). Such compounds were also assayed in vivo for their effects on mouse anxiety by means of a light/dark (LD) apparatus. The LD test is based on the mouse being conflicted between its natural hostility to brightly lit areas and its spontaneous tendency to explore novel environments. The most reliable parameters to assess anxiolytic-like activity have been reported to be the time spent in the illuminated area and the number of transitions between the light and dark areas [36]. Compounds 4 and 5 administrated in mouse i.p. at a dose of 20 mg/kg each determined an increase in the time spent in the white compartment (TW) and the number of transfers from the illuminated to the dark area and vice versa (Tran) values, thus representing potential anxiolytic agents without sedative activity.
Recently, research has focused on ligands acting at different sites of the GABAAR in order to develop novel anxiolytic drugs endowed with improved safety profiles. Neurosteroids elicit anxiolytic effects favorable for memory, learning, and emotional processes without sedation by positively modulating GABA neurotransmission [37]. Of note, at the central nervous system (CNS) level, their concentration was enhanced by ligands of translocator protein (TSPO) [38], formerly known as the “peripheral benzodiazepine receptor,” such as diazepam, Ro5-4864, alpidem [39], PK11195 [40], and FGIN-1-27 [41,42]. TSPO ligands were shown to promote cholesterol translocation into mitochondria, where it is converted into pregnenolone, a steroid precursor [43]. In this connection, neurosteroidogenic TSPO ligands may be exploited to identify novel anxiolytic agents, avoiding the side effects typical of Bzs [39,44,45,46]. In light of the above, Taliani and coworkers [47,48,49,50] developed a class of N,N-dialkyl-2-phenylindol-3-ylglyoxylamides (PIGAs) potentially able to bind TSPO (Figure 7). Notably, these compounds were developed because of the structural similarity between indolylglyoxylamides, previously reported as BzR ligands with a sedative/hypnotic or anxiolytic/nonsedative profile [28,34,35], and 2-arylindol-3-acetamides, reported by Kozikowski et al. [42] to be TSPO selective high-affinity ligands. Numerous PIGAs showed subnanomolar affinity for TSPO together with the capacity to enhance pregnenolone concentrations in rat C6 glioma cells and in a human astroglioma cell line (U87MG) [47,48,49,51]. SARs of this class of compounds were rationalized by means of a pharmacophore/topological model composed of three lipophilic clefts (L1, L3, and L4) and an H-bond donor moiety (H1) (Figure 7). In detail, the subsequent interactions were supposed to be established: (i) an H-bond between the amide carbonyl and the H1 site, (ii) lipophilic interactions between the two hydrophobic groups R1 and R2 with the L3 and/or L4 clefts, and (iii) π-stacking interaction between the 2-phenyl group and the L1 pocket (Figure 7). Biological evaluations of a large number of PIGAs allowed the authors to delineate the following structural requirements for an optimal interaction with TSPO: (i) a double substitution on the amide nitrogen, (ii) an electron-withdrawing substituent at R3, and (iii) an electron withdrawing and very small substituent at R4. Finally, the presence of a substituent at the 7-position (R5) did not increase affinity (Figure 7).
In addition, a number of these compounds exhibited in vivo anxiolytic/nonsedative properties in rodents, which was related to their ability to stimulate neurosteroid production, which, in turn, caused a positive allosteric modulation of GABAAR [48,52].
A peculiar characteristic of several classes of TSPO ligands is the lack of correlation between the in vitro steroidogenic activity and binding affinity [53,54,55]. To resolve the gap in that connection, the residence time (RT) has been identified as the best parameter to correlate these two issues. Accordingly, the RT of some selected PIGAs, i.e., 612 (Figure 8), was quantified and the results highlighted a positive correlation between the steroidogenic efficacy and RT, unlike what observed between efficacy and affinity [51,56]. It was also observed that the main requirements to obtain high RTs and thus high-efficacy compounds were a highly lipophilic moiety at the 2-position of the indole together with at least one of the two N-alkyl groups on the amide nitrogen with a number of carbon atoms in the 1–3 range. In fact, PIGAs with in vivo anxiolytic-like properties in rats, namely, 7, 8, and 12 (Figure 8) [48,56], displayed long RT values, evidencing the important role played by RT in predicting the in vivo efficacy of TSPO ligands. To determine the structural reasons behind this difference in RT values, an enhanced sampling molecular dynamics simulation of compounds 6 and 12 (Figure 8), selected on the basis of their different kinetic parameters and efficacy in steroidogenesis, was carried out. The results clearly evidenced that slight structural differences of PIGAs had a significant impact on the unbinding energetics and thus on RT [57].
Very recently, compound 12 (Figure 8) was used to evaluate the immunomodulatory function of TSPO in human microglia [58]. Specifically, 12-pretreatment of interleukin-1β-activated human C20 microglial cells, an adult-derived immortalized cell line that shows the typical phenotypic and functional characteristics of physiological microglia, induced a phenotypic shift from M1 and M2 by reducing the release of pro-inflammatory (IL-8) cytokines and increasing the release of anti-inflammatory (IL-4) ones [58].
Compounds 8 and 12 (Figure 8) were also evaluated for their effect on inflammatory-based retinal neurodegeneration in an in vitro model of lipopolysaccharide (LPS)-induced degeneration in 661W cells, a photoreceptor-like cell line [59]. Both compounds modulated the inflammatory and apoptotic processes in 661W cells and reduce LPS-driven cytotoxicity, effects that were shown to be mediated by neurosteroids [59].
Compound 12 (Figure 8) was shown to be also able to reduce myelin oligodendrocyte glycoprotein experimental autoimmune encephalomyelitis (MOG-EAE) disease progression, severity, and neuropathological markers by (i) promoting neurosteroid allopregnenolone synthesis in the CNS and (ii) increasing the production of anti-inflammatory IL-10, evidencing its potential use as a tool against primary progressive multiple sclerosis (PPMS) or severe multiple sclerosis (MS) [60].
Finally, compound 12 (Figure 8) was also employed to investigate the impact of neurosteroidogenesis on cholesterol homeostasis in an in vitro human microglia model, with cholesterol accumulation usually being associated with the impairment of neuroinflammatory response, favoring neurodegenerative disease progression. The results evidenced the crucial role played by TSPO ligands, such as 12, in promoting neurosteroidogenesis, allowing for the restoration of cholesterol homeostasis and thus the maintenance of the correct functionality of microglia for the treatment of brain diseases related to neuroinflammation [61].
The biological activities of 12 are summarized in Table 1.

4. Indolylglyoxylamides with Antitumoral Properties

Cancer is the main cause of death worldwide, accounting for more than 19.3 million newly diagnosed cases. The widely recognized multistep process of carcinogenesis assumes that the accumulation of several mutations in an individual cell leads to significant modifications in its proliferation and differentiation behavior, reaching a malignant state with metastasis via benign intermediate stages [62]. Cancer therapy differs depending on the type and stage of the tumor, but it can be identified with common strategies, including surgery, which is the most suitable in the case of solid tumors, although it may favor the spreading of metastases [63]. Moreover, radiation (ionizing [64], thermal [65], and photodynamic [66]) can be used to treat tumors through what is known as radiotherapy, which, like surgery, is preferable for localized or isolated cancers [67]. Chemotherapy and immunotherapy are other possible ways of treating oncological diseases represented mainly by disseminated or systemic tumors [65]. The first is based on the use of “antineoplastic” substances, namely, chemical agents against new growth, whose cytotoxic action interferes with cellular synthesis, DNA and RNA functions, and life-sustaining proteins. The second one, discovered more recently, is among the most widely studied approaches due to its greater selectivity and effectiveness through immune stimulation by using monoclonal antibodies [68].
Despite massive research efforts, it has still not been possible to develop drugs able to trigger a marked prolongation of survival time or even a complete cure for widespread solid tumors. The main problems associated with cancer chemotherapy are the lack of, or poor response, to treatments currently available, severe side effects, and the onset in many tumors of a multi-drug resistance process, which can be considered the most severe complication of these diseases. Tumors that have developed resistance to more than one drug are said to be “multi-drug resistant” (MDR), and in this case there is little that can be done to halt or retard further progression of the disease.
Therefore, there is still a need to develop new drugs to overcome the aforementioned problems with the currently used treatments. Ideal anticancer drugs should be efficient against currently untreatable or poorly treatable tumors and against MDR tumors, with oral bioavailability and/or reduced side effects [69].
Among the indol-3-ylglyoxylamides, indibulin 13 (ZIO-301; D-24851, Figure 9) is an orally applied, synthetic, small molecule whose antitumor effect is based on microtubule destabilization [70].
Indibulin 13 provokes cell accumulation with condensed nuclei and abnormal mitotic spindles and cell cycle arrest at metaphase; it exhibits anticancer activity against a broad spectrum of human tumor cell lines and xenografts, including MDR tumor cells and taxane refractory tumors [71]. In preclinical studies, 13 did not show neurotoxicity—which is usually observed with other tubulin binding drugs, i.e., taxanes and vinca alkaloids—because of the lack of affinity for microtubules in differentiated neuronal cells with post-translationally modified tubulin [72]. Furthermore, 13 exhibits good bioavailability after oral administration and therapeutic treatment at nontoxic doses [71]. Therefore, this antimitotic drug exhibits a notable improved therapeutic index with respect to other tubulin binding agents, i.e., paclitaxel and vincristine [73]. However, further clinical tests in a phase I trial [74] of an oral drinking solution of 13 in 10% lactic acid evidenced an increase in the occurrence of nausea and vomiting [74]. In 2010, Oostendorp et al. tested a capsule formulation of 13 to improve its tolerability [75]. The protocol was modified to twice daily dosing with the aim of further increasing the systemic exposure upon oral administration of 13. Unfortunately, the results were not the expected ones, and the formulation still needs optimization.
With the aim of investigating whether the replacement of the aryl rings with different heterocycles at the 1- and/or 3-positions of 13 may increase the biological activity and confer potential advantages in terms of pharmacological and PK features [76,77], Li et al. developed a series of N-heterocyclic indolylglyoxylamides [78]. In general, the replacement of the p-chlorophenyl and pyridine rings of 13 with different heterocyclic moieties increases the solubility of the molecules in water. The anticancer effect of such derivatives was assayed in vitro against six human cancer cell lines (gastric NUGC3, murine leukemic P388, hepatocellular HepG2, breast MCF7 and its doxorubicin (adriamycin)-resistant MCF7/ADR, uterus MES-SA and its doxorubicin-resistant MES-SA/Dx5 subline) and the results clearly demonstrated an enhanced anticancer activity of these compounds with respect to 13 [78]. A crucial role played by steric and electronic factors of the heterocycle at the 1-position was evidenced, whereas for the 3-position, the presence of an isothiazolyl ring appeared to be an important feature. The most active compounds were 14 and 15 (Figure 10 and Figure 11). Both compounds exhibited a similar dose-dependent in vivo activity at oral doses up to 100 mg/kg when tested on a leukemic P388 cancer survival model in young female inbred DBA/2J mice, whereas 13 was more efficacious at an oral dose of 200 mg/kg. After a continuation of up to 9 days of daily oral administrations with 14 and 15, further prolongation of cancer survival was revealed without evident loss of body weight.
Compound 14 (BPR0C261, Figure 10) was further examined for its mechanisms behind its anticancer action. It was effectively found to be able to (i) inhibit tubulin polymerization by binding at the colchicine site, (ii) disrupt microtubule arrangement, (iii) arrest the cell cycle at the G2/M phase in cancer cells, and (iv) inhibit angiogenesis in various types of human cancer [79]. In addition, it was orally adsorbable in mice and exhibited good oral bioavailability (43%) in dogs and the ability to cross the human intestinal Caco-2 cell monolayer, suggesting good oral bioavailability in humans. Compound 14 exhibited antiproliferative properties in vivo against different types of tumors, and its combination with cis-platin synergistically prolonged the lifespans of the mice inoculated with murine leukemia cells. Several years later, 14 was shown to be able to modulate the radiation response on human non-small cell lung cancer (NSCLC) cells via p53-dependent and p53-indipendent pathways [80].
In this connection, two compounds of this series, 15 and 16, were assayed on NSCLC cell lines A549 (p53+/+) and H1299 (p53−/−) [81]. Compound 15 (Figure 11) showed a greater cytotoxic effect than 16 (Figure 12) on both cell lines, and this was likely due to its higher cell permeability deriving from its higher lipophilicity. Further biological evaluation highlighted a p53-indipendent antiproliferative mechanism for this class of compounds. Of note, both compounds caused similar patterns of G2/M phase arrest in both cancer cell lines. In addition, they enhanced the radiosensitivity of both cell lines. This effect was probably due to the increased proportion of cells in the G2/M phase, the most radiosensitive cell cycle phase [82].
Pursuing the interest in the development of orally active anticancer drugs, Li et al. showed that compound 17 (BPR0C305, Figure 13) was able to induce cell cycle arrest at the G2/M phase by interfering with microtubule assembly kinetics and to exert antimitotic and antiproliferative effects against human cancer cells. Additionally, 17 is orally adsorbable, is almost safe, and shows anticancer activity against leukemia and solid tumors in mice [83].
Following these studies, Colley et al. [84] in 2015 described a series of 13-related compounds with less aromatic rings and a greater degree of saturation, aiming to obtain compounds endowed with potent tubulin polymerization activity and improved physicochemical and PK properties. Starting from the suggestion that N1-unsubstituted indol-3-ylglyoxylamides can still maintain tubulin polymerization inhibitory activity, the authors first focused on the substitution of the p-chlorobenzyl group of 13 with an aliphatic moiety. All compounds were initially evaluated for their cytotoxicity against FaDu, an EGFR inhibitor insensitive (head and neck cancer) cell line derived from squamous cell carcinoma of the pharynx. As a general trend, the presence of various simple alkyl groups at the N1-position induced a loss of cytotoxic activity; however, the introduction of an OCH3 group to the pendant pyridine retrieved activity and changing the positions of these structural features caused an improvement in potency. Rational changes in the group at the 3-position (R2) revealed some clear trends in activity: (i) the p-methoxy group bounded to the pyridine ring played an essential role; (ii) the substitution of the p-methoxy with a p-ethoxy was tolerated, with a 2-fold increase in LC50, whereas the introduction of a second methoxy group was detrimental and (iii) replacement of the pyridine ring with a five-membered heterocycle appeared to be unproductive. The introduction of different groups at the 1-position (R1) did led to the desired increase in antiproliferative activity. The introduction of a sec-butyl group was more beneficial for the activity compared to the introduction of an isopropyl group. Changing the oxygen atom in the 2-methoxyethyl moiety with a methylene group did not influence activity, suggesting that the shape, rather than the polarity, of the R1 group determines the potency. The introduction of an additional α-methyl group to the R1 of the 2-methoxyethyl substituent resulted in the most potent compound of the series (18, Figure 13, LC50 against FaDu = 12 nM) [85], whereas introducing a β-tertiary center was not tolerated at all. In sum, although some of the most potent compounds featured simple and linear R1 chains, a degree of further substitution was tolerable or even advantageous, whereas too much substitution led to compounds with severely compromised or abolished cytotoxic activity. Finally, a subset of compounds was further evaluated, and it was discovered that they were effectively able to inhibit tubulin polymerization through binding at the colchicine site.
Guggilapu and coworkers developed a small library of C5-tethered indol-3-ylglyoxylamides as tubulin polymerization inhibitors [86]. All derivatives were tested for cytotoxic activity against different cancer cell lines (prostate (DU145, PC-3), lung (A549), and colon (HCT-15)), with the DU145 cell line highlighted as the most sensitive one. In particular, derivative 19 (Figure 13) was shown to be the most cytotoxic one and was further biologically evaluated, evidencing its ability to (i) inhibit cell migration, (ii) induce apoptosis, (iii) arrest the cell cycle at the G2/M phase in a dose-dependent manner, (iv) inhibit tubulin polymerization with an IC50 value of 0.40 μM, (v) induce the collapse of mitochondrial membrane potential (Δψm), (vi) increase ROS levels, and (vii) decrease anti-apoptotic Bcl-2 and increase pro-apoptotic Bax levels.
In a subsequent study, the same research group developed a class of compounds by merging the indol-3-ylglyoxylamide scaffold with the bioactive thiazole nucleus [87]. All compounds were first evaluated for antiproliferative activity against different cancer cell lines (prostate (DU145, PC-3), lung (A549), and colon (HCT-15)) and, also for this class, the DU145 cell line turned out to be the most sensitive one. The most active compound was 20, which showed IC50 = 93 nM against DU145, was safe with the healthy cell line RWPE-1, and inhibited tubulin polymerization with a percentage of 68.5%. Moreover, DU145 treatment with compound 20 led to (i) cell migration inhibition, (ii) apoptosis induction, (iii) G2/M phase cell cycle arrest in a concentration-dependent manner, (iv) the collapse of Δψm, and (v) an increase in ROS levels. Finally, molecular modeling studies performed on 20 evidenced the crucial role played by the glyoxyl moiety, which was able to establish three hydrogen bonds with active site residues of the colchicine binding site of the tubulin [87].
More recently [88], aiming at developing more water-soluble compounds, Brel and colleagues synthesized indibulin 13 (Figure 9) derivatives featuring a bisphosphonate moiety on the indole nitrogen. Assessment of the compounds’ antimitotic microtubule-destabilizing activity and cytotoxicity toward a panel of human cancer cells outlined the tetraacid derivative 21 (Figure 9) as the most active compound, which was more potent with respect to indibulin 13. The storage of stock solutions of its corresponding tetraethyl esters in 96% EtOH at room temperature provoked spontaneous hydrolysis of bisphosphonate and subsequent time-dependent activity enhancement. Compound 21 also showed significant cytotoxicity toward A549 and A375 human cancer cells. At physiological pH, 21 formed tetrasodium salt endowed with an aqueous solubility of at least 10 mg/mL, three orders of magnitude higher than that of indibulin 13.
In 2016, Tantak’s research group [89], encouraged by the good antitumoral activity shown by bisindole-based derivatives and by the crucial role played by the glyoxylamide in bioactive compounds [90,91], synthesized and tested a series of bis(indolyl)glyoxylamides against different human cancer cells. The most promising compounds, 22 and 23, are presented in Figure 13. Studies on the preliminary mechanism of action indicated the ability of these compounds to induce apoptosis in PC-3 cells by increasing cleaved PARP1 levels.
Very recently, Soni et al. [92] designed and synthesized a series of compounds by joining the indol-3-ylglyoxylamide moiety and the β-carboline scaffold to obtain cytotoxic agents acting as DNA intercalators. In particular, the β-carboline alkaloid skeleton was incorporated as a pharmacophoric planar aromatic ring for DNA intercalation [93,94] and the indole ring was introduced as a second pharmacophoric requirement of the hybrids. Lastly, the glyoxamide may have helped the molecule to assume the appropriate orientation in the active site (Figure 14).
When tested for their antiproliferative effects toward several human cancer cell lines, the most potent compound was 24 (Figure 14), with IC50 values ranging from 4.37 to 10.36 μM and good selectivity to cancer cells with respect to healthy ones. Deep biological evaluation revealed that 24 was able to (i) induce early and late apoptosis, (ii) bind DNA, and (iii) inhibit topoisomerases II (TopoII). The interaction of 24 with a DNA-TopoII complex was also studied in silico by molecular modeling analysis, revealing the crucial role of glyoxylamide moiety, which is implicated in a hydrogen bond with DNA. An in silico study also revealed promising drug-like properties of 24.
The indol-3-ylglyoxylamide scaffold was exploited by Taliani et al. to identify compounds useful for the multi-target treatment of glioblastoma multiforme (GBM), a particularly aggressive form of brain cancer. In fact, the pathogenesis of malignant gliomas was characterized by the deregulation of different intracellular signaling pathways, most of them converging to escape from cell death, which is a common feature of tumors and the main cause of treatment failure [95]. In this respect, apoptosis inducers have emerged as promising drugs for a large variety of tumors. In GBM, tumor suppressor protein p53 [96] and mitochondrial 18 kDa translocator protein (TSPO) [97] represent two attractive intracellular targets, both acting as apoptosis inducers. On this basis, indol-3-ylglyoxyldipeptides were rationally designed to dually modulate TSPO and p53 [98]. Specifically, the authors focused their attention on their in-house collection of 2-phenylindol-3-ylglyoxylamide TSPO ligands (PIGAs) [47,48,49], which were suitable to be easily structurally modified, and, aided by molecular modeling studies, planned adequate decoration of the basic scaffold, with the aim of reactivating p53 while retaining TSPO binding affinity. The result of the study was the identification of compound 25 (Figure 15), which was able to bind TSPO (Ki = 438 nM) and reactivate p53 functionality by inhibiting (IC50 = 11.65 nM) the dissociation from its physiological inhibitor, murine double minute 2 (MDM2) [98]. Compound 25 was also able to induce Δψm collapse and inhibit cell viability in GBM cells. Of note, 25 was more active with respect to the single target reference standards applied alone, nutlin-3 for MDM2-p53 and PK11195 for TSPO, and was as potent as the combination of them. This is consistent with the synergism resulting from the simultaneous targeting of TSPO and p53, which may represent a valuable anticancer therapy to treat GBM, since most GBM phenotypes maintain wild-type p53 and over-express TSPO and MDM2 [98].
Encouraged by these good results, the same research group started a lead optimization process and developed a small library of 2-phenylindol-3-ylglyoxylyldipeptide derivatives in their methyl ester form [99].
From this study, compound 26 (Figure 15) emerged as the most active compound in terms of TSPO binding affinity and p53-MDM2 interaction inhibition (TSPO: Ki = 87.2 nM, p53-MDM2: IC50 = 4.3 nM) [99]. Moreover, its effects were superior to those of the lead, 25. For these reasons, it was selected for further studies aimed at characterizing its biological effectiveness as an antitumoral agent. The results evidenced the ability of 26 to restore normal p53 activity and inhibit GBM cell growth by cell cycle arrest and apoptosis. Furthermore, 26 was ineffective at affecting the viability of a GBM cell line expressing mutant p53, whereas it was able to reduce the proliferation of glioma cancer stem cells, cells within the tumor that are resistant to therapies and responsible for GBM recurrence. Cell viability assays on non-tumoral human mesenchymal stem cells (MSCs) evidenced the ability of 26 to display its effect predominantly on tumor cells with respect to healthy ones [99].
In cancer therapy, reversible drugs may have several limitations, including the lack of ability to induce a therapeutic effect over time and the consequent activation of alternative signaling pathways that escape drug action, causing resistance [100]. In this connection, Taliani et al. [101] developed a dual target compound able to covalently bind MDM2 and TSPO (27, Figure 15). In detail, starting from the previously developed reversible derivative 25, the authors introduced the chemo-reactive isothiocyanate group at the 5-position of the indole ring, a position that is not implicated in the interaction with the target protein [98]. Then the ability of compound 27 to bind TSPO with a long-lasting binding profile was shown (Ki = 199 nM). Through this binding, 27 caused mitochondrial permeability transition pore opening and consequent Δψm dissipation in GBM cells. Both effects were shown to be maintained over time, unlike those observed for the reversible analogue 25. In addition, 27 bound MDM2 with an IC50 of 6.81 nM, inhibiting its interaction with p53; also in this case, the effect was sustained over time, thus demonstrating the covalent nature of such an interaction. Compound 27 was also able to inhibit GBM cell growth by arresting the cell cycle and inducing apoptosis. All of these effects appeared to be greater and more long lasting with respect to those elicited by 25. Moreover, the apoptosis appeared to be irreversible, hindering the cells from recovering proliferative activity after the drug was washed out. Overall, all these findings demonstrated that in GBM cells compound 27 acts as a dual-targeting and irreversible ligand, representing a useful alternative to overcome the time-limited effects of classical chemotherapies [101].
Platinum-based agents, i.e., cisplatin, carboplatin, and oxaliplatin, are some of the most used and active chemotherapeutic drugs currently prescribed to treat many types of human cancer [102]. However, behind their effectiveness, their use is limited by drug resistance and adverse effects [103]. Accordingly, several effective strategies, including the use of transition metals different from platinum, have been developed [104]. Among these, paddlewheel ruthenium-based complexes, featuring a direct metal–metal bond and a (II,III) mixed valence, emerged as very promising [105]. In this regard, a diruthenium (II,III) complex 28 (Ru2(25)4Cl) (Figure 15) was developed by reacting Ru2(μ-O2CCH3)4Cl with the dual agent 25 [98,106]. At variance with 25, 28 was ineffective in GBM because of its high stability, which is due to the steric hindrance provided by indolylglyoxylyl moiety that hampers the attack of the water molecules or other potential ligands at the Ru metal centers, inhibiting the release of ligand and, in turn, its anticancer effect. Thus, complex 29 (Ru2(26)4Cl) (Figure 15) was developed, featuring 26 (isomer of 25) [99], which coordinates the Ru2 core. Complex 29 was less stable and, in turn, more active as an antitumoral agent due to the enhanced availability of the Ru2 core to attack by nucleophiles with respect to 28 [107].
The same research group in 2023 developed a complex designed as an auranofin analogue obtained by conjugating a PIGA ligand [47,48,49] with the auranofin-derived cationic fragment [Au(PEt3)]+ [108]. Auranofin is a gold-based drug clinically used for the treatment of arthritis. Lastly, it took part in different drug repurposing strategies, and it was shown to be promising toward different tumors, such as ovarian cancer (OC), in particular by inhibiting thioredoxin reductase (TrxR), localized at the mitochondrial level [109]. In parallel, TSPO expression was altered in various multiple diseases, including OC [110]. In view of this, the rationale of the design is to exploit the high TSPO affinity of phenylindolylglyoxylamide moiety to drive the complex into the mitochondria, where the [Au(PEt3)]+ cation, after its release, can exert its anticancer function. Notably, complex 30 (Figure 15) exhibited enhanced antitumor activity toward two models of human OCs (i.e., A2780 line and SKOV-3) and much lower cytotoxicity toward a healthy cell line (i.e., HEK293). In addition, 30 was effective as a TrxR inhibitor, with EC50 in the low nanomolar range.

5. Indolylglyoxylamides with Antibacterial Properties

In 2008, Takhi and colleagues [111] developed a small library of 3-indolylglyoxylamides endowed with antibacterial properties structurally related to linezolid 31 and eperezolid 32 (Figure 16), oxazolidinone-based antibacterial agents effective against many susceptible and resistant Gram-positive organisms such as methicillin-resistant Staphylococcus aureus (MRSA), vancomycin-resistant enterococci, and penicillin-resistant Streptococcus Pneumoniae (PRSP) [112]. The oxazolidinones inhibited bacterial protein synthesis by binding to a site on the 50S ribosomal subunit, thus preventing the formation of a functional 70S initiation complex [113,114]. Despite that, with this single mechanism, no cross-resistance was expected between this class and other antibacterial families, although some cases of 31-resistant pathogens in clinical have been observed [115]. In this context, to increase the spectrum and potency of these oxazolidinones, Takhi’s research group exploited the indolylglyoxylamide moiety. Two series of compounds were synthesized: the first one by replacing the hydroxyacetamido group of 32 with various substituted 3-indolylglyoxylamides, and the second one by replacing the acetamide at the C-5 position with a variety of N- and O-linked groups and keeping the 5-bromoindole fixed on the left side. Almost all derivatives showed increased activity with respect to 31 against MRSA, Enterococcus faecalis (E. faecalis), and Enterococcus faecium (E. faecium). Among these compounds, the most active one was 5-bromoindole 33 (Figure 16), which showed 1-2-fold higher activity with respect to 31 against a panel of organisms. The MIC values against Staphylococcus aureus (S. aureus), E. faecalis, and E. faecium were in the range of 0.5–1 μg/mL. However, all derivatives were ineffective across a Gram-negative community acquired pneumonia pathogen-like Haemophilus influenzae (H. influenzae) [111].
In a subsequent phase, starting from 33, a class of molecules was developed in which the 5-bromoindole moiety was kept fixed, whereas the acetoamidomethyl side chain was replaced with various substituents, such as amide, carbamate, and azole. In general, almost all compounds were shown to be active across all the organisms investigated, except for the Gram-negative H. influenzae, as observed for the first series. N-linkage modifications at C-5 furnished the most active compounds of the series, such as 34 (Figure 16) (MIC = 0.5–2 μg/mL), which also demonstrated in vivo efficacy against S. aureus DRCC035 in a lethal murine infection model via oral route.
Singh and colleagues developed a series of N-1, C-3, and C-5 substituted bis-indoles [116], among which 35 (Figure 16) was the most active compound, being able to generate the highest zone of inhibition. Molecular modeling studies performed on the enzyme targeted by clinically used antimicrobial agents (TopoII, dihydrofolate reductase) evidenced that the glyoxylamide moiety of 35 plays a key role in such interactions [116].
In 2016, Tantak’s research group also made use of the indol-3-yl-glyoxylamide scaffold to design potential antibacterial agents [89]. In detail, encouraged by the good antimicrobial activity of bisindole-based derivatives [116,117], a series of bis(indolyl)glyoxylamides was synthesized and tested against two Gram-positive bacteria, Bacillus subtilis and S. aureus, and three Gram-negative bacteria, Escherichia coli (E. coli), Pseudomonas putida, and Klebsiella pneumonia (K. pneumoniae). The most promising compound was 36 (Figure 16), which showed significant antibacterial activity across all the investigated organisms, with MIC values ranging from 12.5 to 25 μg/mL, combined with high effectiveness in killing Gram-negative organisms.
Via a screening campaign on compounds from an in-house library, Li et al. [118] identified two 6-bromoindol-3-ylglyoxylamido-polyamines, 37 and 38 (Figure 16), which displayed antimicrobial activity across Gram-positive bacteria Staphylococcus intermedius (MIC 3.125 µM for both compounds) and S. aureus (MIC 6.25 and 3.125 µM, respectively), with 37 also being able to enhance the in vitro antibiotic activity of doxycycline toward the resistant Gram-negative Pseudomonas aeruginosa (P. aeruginosa). Starting from these results, a series of 6-bromoindol-3-ylglyoxylamido-polyamines was developed, leading to the discovery of compounds with enhanced antimicrobial activity. Compounds with a polyamine (PA) core were the most active within the series, reaching maximal potency with a PA3-10-3 core (39, Figure 16). Compound 39 also displayed the ability to enhance the antibiotic activity of doxycycline toward P. aeruginosa. Deep investigation of 37 revealed its ability to (i) restore antibiotic activity against a variety of Gram-negative bacteria, such as Escherichia coli (E. coli), K. pneumoniae, and Acinetobacter baumannii (A. baumannii); (ii) disrupt bacteria membrane integrity; and (iii) inhibit the efflux pump via depolarization of the membrane [119]. Unfortunately, 37 exhibited cytotoxicity and strong red blood cell hemolytic activity.
In the same year [120], the same researchers, using 37 as a lead compound, described a series of bromoindol-3-ylglyoxylamides in search of antibacterial agents lacking cytotoxic/hemolytic activity. All the compounds showed intrinsic activity across Gram-positive bacteria and the fungus Cryptococcus neoformans, with 40 (Figure 16) also displaying slight activity across Gram-negative bacteria like E. coli (MIC 25 µM) and K. pneumoniae (MIC 34.4 µM). Most derivatives also potentiated the activity of doxycycline against Gram-negative bacteria such as P. aeruginosa, A. baumannii, E. coli, and K. pneumoniae. This study allowed the authors to identify five antibacterial compounds (4044, Figure 16) without cytotoxic/hemolytic activity. Their mechanism of action consisted of disrupting the integrity of the bacterial membrane and inhibiting the efflux pump by means of depolarization of the membrane, as observed for the lead, 37.

6. Indolylglyoxylamides with Anti-Inflammatory Properties

6.1. Phosphodiesterase PDE4 Inhibitors

Phosphodiesterases (PDEs) are enzymes responsible for the modulation of signal transduction, and they catalyze the degradation of cyclic nucleotides (cAMP and/or cGMP). Among the 11 families known, PDE4, PDE7, and PDE8 are selective for cAMP [121]. As the PDE4 isozyme is the most prominent cAMP metabolizing enzyme in immune and inflammatory cells, selective PDE4 inhibitors have been deeply investigated, with the aim of developing novel anti-inflammatory agents alternative to corticosteroids [122]. Potent anti-inflammatory effects in patients with inflammatory diseases such as asthma or chronic obstructive pulmonary disease (COPD) were observed following selective inhibition of PDE4 [123].
PDE4 inhibitors are more potent than corticosteroids in clinical use because they can influence diverse cell types involved in inflammatory diseases, including respiratory epithelial cells, smooth muscle cells, and submucosal glands. Furthermore, they lack the common side effects associated with corticosteroid therapy, namely, adverse reactions on the pituitary–hypophyseal axis and on bone density, among others [124]. However, the therapeutic use of PDE4 inhibitors so far has been limited by several side effects, such as nausea, vomiting, and headache.
Compound 45, AWD12-281 (Figure 17), was a highly selective PDE4 inhibitor [125,126] with a better safety profile than other PDE4 inhibitors, namely cilomilast and roflumilast [127,128], in clinical development. Additionally, it was demonstrated that when 45 was administered topically by inhalation in dogs, also at the highest suitable dose (15 mg/kg in 50% lactose blend), no emesis was produced. These data support its potential use for the topical treatment of asthma and COPD. It is characterized by low oral bioavailability and solubility, and it exerts robust and durable pharmacological effects after intratracheal administration in various animal models, indicating persistence in lung tissue. Compound 45 was capable of suppressing the production of cytokines in stimulated peripheral blood mononuclear cells (PBMCs): interleukin-2 (IL-2, phytohemagglutinin stimulation), IL-5 and IL-4 (anti-CD3/anti-CD28 costimulation), and lipopolysaccharide-stimulated release of tumor necrosis factor α (TNFα). The corresponding EC50s for 45 were within a narrow range (46–121 nM). Compound 45 also suppressed TNFα release in dispersed nasal polyps (EC50 = 111 nM) and in diluted whole blood (EC50 = 934 nM). High plasma protein binding may be responsible for the reduced activity in human blood. The molecule entered phase II clinical trials to evaluate its therapeutic potential in asthma, COPD, and allergic rhinitis. Even though the results are not available to the public, development was discontinued in 2006 due to poor efficacy [129]. Despite the halt, the good safety profile, along with the persistence in lung tissue, may serve as a good starting point for the development of novel clinical candidates for the aforementioned pathological conditions.

6.2. Phospholipase A2 Inhibitors

Phospholipases A2 (PLA2s) belong to the superfamily of phospholipases, enzymes that hydrolyze phospholipids. PLA2s are lipolytic enzymes responsible for the catalysis of the ester bond hydrolysis at the sn-2 position of glycerophospholipids, which generate free fatty acids, including arachidonic acid and lysophospholipids. This family comprises four predominant types of PLA2: secreted PLA2 (sPLA2), cytosolic Ca2+-dependent PLA2 (cPLA2), cytosolic Ca2+-independent PLA2 (iPLA2), and PAF-AH (platelet activating factor acetyl hydrolases). The other two types are lysosomal PLA2 (LPLA2) and adipose-PLA2 (AdPLA). Enzymatic functions are performed by means of a catalytic dyad/triad (His/Asp for sPLA2; Ser/Asp for cPLA2 and iPLA2; Ser/His/Asp for PAF-AH and LPLA2; His/Cys for AdPLA) [130].
sPLA2s are involved in several inflammatory diseases. The first indication came from high levels of group IIA (GIIA) sPLA2 found in the synovial fluid of patients with rheumatoid arthritis. Abnormal levels of sPLA2s were also found in the plasma or serum of patients with acute pancreatitis, septic shock, Crohn’s disease, and ulcerative colitis. Furthermore, sPLA2s seem to be involved in adult respiratory distress syndrome and inflammatory bowel disease [130].
Researchers at Lilly published a series of papers regarding indole-based derivatives as inhibitors of GIIA sPLA2 (referred to by the authors as “human nonpancreatic secretory phospholipase A2,” hnps-PLA2) [131]. Lead compound 46 (IC50 = 13.6 μM, Figure 18) was initially obtained by high-volume screening, and several rounds of optimization were performed, including the replacement of the acetate function first with an acetamide moiety and then with the α-ketoamide group [132].
Incorporation of the α-ketoamide group, yielding the indolylglyoxylamide moiety, was crucial, as delineated by compounds 4750 (Table 2), in which the introduction of a 4-oxyacetic acid group allowed optimal potency and selectivity to be reached [133].
Interactions of the lead compound 46 were confirmed by X-ray crystallography studies, which also allowed for a rationalization of the effective binding between the calcium ion in the active site of hGIIA and the carbonyl of the 4-substituent and the carboxamide carbonyl of the 3-glyoxylamide moiety of compound 50 (LY315920, varespladib) [134]. Introduction of the glyoxylamide moiety was instrumental to additional interactions in the active site, namely, the hydrogen bond formation between the carboxamide and His 48, and the new interaction between the ketone carbonyl and Phe 106 of the enzyme [133].
Several clinical trials were launched to evaluate the efficacy of Varespladib 50, also formulated as a methyl ester prodrug, for various diseases (i.e., sepsis-induced systemic inflammatory response syndrome, asthma, cardiovascular diseases), but the molecule did not demonstrate sufficient efficacy in either phase II or phase III [135,136,137,138,139,140,141].

6.3. p38α Inhibitors

p38α belongs to the well-known mitogen-activated protein (MAP) kinase family of serine/threonine protein kinases. It is widely expressed in endothelial, immune, and inflammatory cells and plays a pivotal role in the release of proinflammatory cytokines such as TNF-R, IL-1β, and IL-6 [142,143]. Compounds able to selectively block any one of these cytokines have proven efficacious for the treatment of pathological inflammatory conditions, including rheumatoid arthritis (RA), psoriasis, and inflammatory bowel disease [144]. Separate genes encode for the four isoforms (p38α, p38β, p38γ, and p38δ) of the p38 subfamily of MAP kinases. The p38α isoform was overactivated within inflamed tissues, which suggests that this target could be used for the treatment of these types of diseases [145,146].
Starting from the ability of p38α inhibitors to block the synthesis and release of proinflammatory cytokines, several pharmaceutical companies, like Merck, Vertex, Roche, and Scios (subsidiary of J&J), among others, invested in the development of novel therapeutic agents for the treatment of RA, inflammatory bowel disease, psoriasis, systemic lupus erythematosus, and other indications characterized by chronic inflammation. This led to the development of an impressive variety of chemically diverse competitive inhibitors with excellent drug-like properties. These compounds showed a high degree of selectivity despite most of them being ATP competitive inhibitors. This is because these compounds, although structurally different, position themselves in specific regions in or in proximity to the ATP binding site of p38α, which is characterized by amino acid sequences distinct from the majority of other human kinases. Being able to identify such selectivity hotspots is an important process to developing highly selective ATP competitive kinase inhibitors.
First- and second-generation p38α inhibitors, like 51 (Figure 19) [147], did not show high selectivity against other human kinases. However, third-generation inhibitors have recently been developed, demonstrating that selective p38α inhibition is possible. Three third-generation compounds, 52 [148], 53 [149], and 54 [150] (Figure 19), are highly kinase selective.
In a panel of 321 human protein kinases, Scios inhibitor 52 bound exclusively to p38α (KD = 0.48 nM) and its closely related isoform p38β (KD = 15 nM). Vertex inhibitor 53 bound to p38α (KD = 3.7 nM), p38β (KD = 17 nM) and five other kinases. Roche inhibitor 54 bound to p38α (KD = 1.3 nM) and to JNK2/3 and p38β, although less strongly (KD =16/19 nM and KD = 120 nM, respectively); in addition, it bound weakly to five other kinases [152]. Compound 52 displayed an IC50 = 9 nM against p38α. It was also able to inhibit LPS-induced TNFα release from human whole blood (IC50 ≈ 0.3 μM) and LPS-induced IL-1β release from human peripheral blood mononuclear cells (PBMCs) in a concentration-dependent manner [153]. A phase I study evaluted the safety, pharmacodynamics, and PK of 52 in healthy volunteers. A phase II study investigated the analgesic efficacy of 52 in acute postsurgical dental pain (263 subjects). Because it is well known that inflammatory mediators and cytokines such as TNFα and IL-1 can contribute to peripheral and central sensitization and also to modulating acute, chronic, and neuropathic pain, inhibition of p38 could be beneficial to managing pain [154,155]. This study showed for the first time the acute analgesic effects induced by inhibition of p38α. The molecule was evaluated in a phase II clinical trial in patients with active rheumatoid arthritis and showed no greater efficacy compared to the placebo [156].
Recently, a combination of compound 52 with remdesivir has been suggested to treat stage I, stage II, or stage III COVID-19 or COVID-19 cytokine storm [157].

7. Indolylglyoxylamides with Antiviral Properties

7.1. Viral NS2B/NS3 Protease Inhibitors

Dengue virus (DenV) is the infectious agent responsible for Dengue fever, a neglected tropical disease. It belongs to the Flaviviridae family, a family of viruses characterized by a positive single-stranded RNA genome [158,159]. Essential to the viral replication cycle is the NS2B/NS3 protease, a serine endoprotease member of the chymotrypsin family expressing the catalytic triad His51–Asp75–Ser135. The protease cleaves the polyprotein encoded by the viral genome into functional proteins, thus representing an interesting therapeutic target [160].
Klein and colleagues [161] synthesized several β,γ-unsaturated-α-ketoamides, and among these compounds, the most interesting derivative, 55 (Figure 20), was able to inhibit DenV replication in a cell-culture assay in a concentration-dependent manner and to induce a 1000-fold reduction in virus titer at noncytotoxic concentrations. SARs highlighted the role of the indolylglyoxylamide moiety, with the α-ketoamide warhead showing superior inhibitory activity compared to α-hydroxy and α-epoxy derivatives, and the indole portion binding to the S1 subsite. It should be speculated that 55 may tautomerize to 56 because of the double bond at the α,β-position with respect to the ketocarbonyl (Figure 20) and that both may contribute to the activity of the compound.
Zika virus (ZIKV) is another pathogen belonging to the Flaviviridae family and Flavivirus genus, like DenV. ZIKV infections are usually asymptomatic or characterized by mild symptoms, but during the latest outbreaks severe ZIKV-associated complications emerged (depending on the genetic variant of ZIKV), such as congenital birth defects, including microcephaly, and Guillain–Barre syndrome in adults [162]. In 2018, ZIKV infection was added to the Research and Development Blueprint by the WHO (the list of diseases recommended for intensive research) [163].
Del Rosario García-Lozano et al. investigated a series of piperazine-based small molecules to find inhibitors with broad activity against Flaviviridae [164]. By means of a privileged structure-based design, the authors coupled the central piperazine core with preferred medicinal chemistry structures, which has proven to possess multiple biological activities, namely, indol-3-yl-2-oxoacetyl [11,161,165,166,167], cinnamoyl [168,169,170,171,172,173,174,175,176], and quinoline-3-carbonyl groups [177,178,179,180,181,182,183,184]. They initially synthesized 26 acylpiperazine amide and urea derivatives (general formula 57, Figure 21), which were screened by a commercial HCV NS3/4A protease inhibition assay, resulting in inhibitory values above 70%, and then assessed in a cell-based antiviral assay to determine their potential activity to inhibit ZIKV and DenV replication in vitro. From the first round of testing, compound 58 (Figure 21), featuring an indolylglyoxylamide, emerged as an interesting molecule with dual activity against ZIKV (IC50 = 8.9 μM) and DenV (IC50 = 10.7 μM) and low toxicity (CC50 = 153 μM). Antiviral activity of this molecule is comparable to the protease inhibitor sofosbuvir 59 (ZIKV IC50 = 4.0 μM, DENV IC50 = 13.1 μM, CC50 > 200 μM, Figure 21), which was used as a reference compound.
In a subsequent round of optimization, the urea linker was functionalized with various aromatic substituents (general formula 60, Figure 21). In this second set of compounds, the indolylglyoxylamide derivatives 61 and 62 (Figure 21) performed worse than the cinnamoyl ones, 63 and 64 (Figure 21). Compound 61 showed activity only against DenV (IC50 = 7.3 μM), and the toxicity profile was not optimal (CC50 = 12 μM), whereas compound 62 was just toxic. Compound 64 was the best in terms of activity against ZIKV (IC50 = 1.9 μM) and DenV (IC50 = 1.4 μM), showing values lower than sofosbuvir 59. Although compound 64 was the most potent of the study, compound 58 showed a lower LogP value (2.17 vs. 5.15) and a minor fraction bound to plasma protein (78.00% vs. 91.97%) in the predicted ADME properties, confirming that the properties of the indolylglyoxylamide moiety are interesting for drug development.

7.2. HIV-1 Inhibitors

The human immunodeficiency virus (HIV) infection pandemic is now over 30 years old and continues to be a global health issue, with more than 37 million infected people (https://www.unaids.org/en/resources/fact-sheet, accessed on 10 June 2023). The utilization of highly active antiretroviral therapy (HAART) helped to impede the spread of HIV and transformed it into a chronic disease for many patients. However, continuous treatment with HAART can lead to limitations, such as the onset of drug resistance. In addition, for this reason, there is still a need for the development of new anti-retroviral agents that can target different steps of the replication cycle with improved tolerability and dosing schedules [185]. Among the numerous steps of the HIV infection cycle, the specific interaction between the membrane-bound HIV-1 glycoprotein 120 (gp120) and cluster of differentiation 4 (CD4), the primary attachment receptor for HIV-1, represents an attractive target for the development of novel inhibitors since the disruption of this interaction would likely curb HIV-1’s infectivity at a very early stage of the viral infection [186].
In this regard, a cell-based screening assay led to the identification of indolylglyoxylamide derivative 65 (Table 3), which was demonstrated to interfere with the gp120/CD4 interaction. Subsequent optimization studies on 65 generated molecules strongly able to inhibit HIV-1 infection in vitro [187,188,189,190], including indolylglyoxylamide 66 (Table 3), which possessed nanomolar EC50 values (4.0 and 4.9 nM against two different viral strains, C-C chemokine receptor type 5 (CCR5)-dependent JRFL and CXC-chemokine receptor 4 (CXCR4)-dependent LAI strains of HIV-1, respectively) and no cytotoxicity to the HeLa host cell line, with a 20-fold improvement in potency compared to 65 (EC50 = 86 ± 24 nM in CXCR4-dependent LAI strain of HIV-1) [188].
Despite the good antiviral profile, physicochemical properties of these first derivatives were associated with drug formulation and delivery issues. Moderate stability in human liver microsomes (HLM) and low aqueous solubility exerted by 66 may have complicated its preclinical and/or clinical development. In order to improve these features, nitrogen was introduced to the indole ring, generating four azaindole analogues of 66 (6770, Table 3) all possessing better PK and pharmaceutical profiles [191]. Interestingly, the indole position onto which the nitrogen was inserted seems to play a crucial role in potency, as exemplified by the 4-aza 67 and the 7-aza 70 isomers, which maintained the antiviral activity compared to 66, whereas incorporation of the nitrogen atom in a more exposed position of the core was detrimental for HIV-1 inhibitory activity (the 6-aza isomer 69 and the 5-aza analogue 68 were 5- and 100-fold less potent, respectively). Insertion of the nitrogen was instrumental to improving the metabolic stability of isomers 6770 with respect to 66 (half-life (t1/2) in HLM: 66 16.9 min; 6770 from 38.5 to >100 min. The presence of the azaindole ring may also be exploited to convert the compounds into the corresponding salts, facilitating their formulation [191]. The improved basicity displayed by the azaindoles seemed to correlate with their permeability across a Caco-2 monolayer at pH 6.5. Compound 7-azaindole 70 (pKa 2.0) should predominantly exist as a free base, thus possessing high permeability, whereas 4-azaindole 67 (pKa 5.0) could be protonated, leading to reduced permeability. For the same reason, a considerable amount of both 5-azaindole 68 (pKa 6.2) and 6-azaindole 69 (pKa 6.0) would be present as the pyridinium cation at pH 6.5, reducing their ability to penetrate across the Caco-2 membrane.
Additional development of these azaindoles led to the identification of clinical candidates: 7-azaindole HIV-1 attachment inhibitor BMS-378806 71 [192] and 6-azaindole derivative BMS-488043 72 (Table 3) [193]. Low permeability and moderate metabolic stability caused the plasma concentration of 71 to decrease after oral administration in humans, leading to a halt in its development, whereas 72 showed an improved in vivo PK profile in rat, dog, and monkey (71: t1/2 in HLM 37 min, Caco-2 permeability 51 nm/s; 72: t1/2 in HLM 100 min, Caco-2 permeability 178 nm/s) [191,194]. Compound 72 was able to reduce viremia in HIV-1-infected subjects when administered as a monotherapy, validating the use of HIV-1 inhibitors as a potential treatment of HIV-1 infection in vivo [195].
Capitalizing on these findings, a subsequent broad optimization campaign on compound 65 led to the identification of temsavir 73 (GSK2616713, Table 4), with enhanced antiviral activity against a spectrum of laboratory strains (Table 4) and good PK [167]. The ability of this compound to bind gp120 was demonstrated by mechanistic studies conducted on an X-ray structure of crystal complex 73/gp120, showing interactions at the interface between the inner and outer domains under the β20−β21 loop (Figure 22) [196]. Despite a predominance of hydrophobic interactions, crucial interactions involved the indolylglyoxylamide moiety, namely, the H-bonds between the backbone NH of W427 with the oxoamide carbonyl and the azaindole NH and the side chain of D11. The benzamide occupied the gp120 site that was also occupied by W427, pushing both W427 and the β20−β21 loop toward the CD4 binding loop and preventing CD4 binding (Figure 22) [167].
Despite the optimal profile of 73, the phosphonooxymethyl prodrug fostemsavir 74 (GSK3648934, Table 4) was synthesized to fix emerging problems linked to dissolution and solubility-limited absorption. Compound 74 successfully decreased the viral RNA level in patients compared with those receiving placebo during the first 8 days, with efficacy sustained through 48 weeks, in a phase III clinical trial in patients with limited therapeutic options [197]. Eventually, these results led to the approval of 74 from the Food and Drug Administration in July 2020 for patients with limited treatment options (FDA approves new HIV treatment for patients with limited treatment options (https://www.fda.gov/news-events/press-announcements/fda-approves-new-hiv-treatment-patients-limited-treatment-options (accessed 18 August 2020)).

8. Indolylglyoxylamides with Antileishmania Properties

Leishmaniasis is a parasitic disease caused by the Leishmania parasite, which is spread to humans through the bite of infected sand flies. This parasite exists in two different forms: the motile flagellated form (promastigotes), which can be found in the gut of the vector, and the non-flagellated form (amastigotes), which is in the mammalian host and it is responsible for acute disease [198,199]. There are three types of cutaneous leishmaniasis: visceral leishmaniasis (VL), the most lethal if left untreated; cutaneous leishmaniasis (CL); and mucocutaneous leishmaniasis (MCL). According to the World Health Organization (WHO), 12 million people are infected worldwide, and in the next few years a dramatic increase is expected of about 2 million new cases per year, out of which 500,000 will be cases of VL. Today, more than 1 billion people live in areas endemic for leishmaniasis and are at risk of infection. An estimated 30,000 new cases of VL and more than 1 million new cases of CL occur annually (https://www.who.int/news-room/fact-sheets/detail/leishmaniasis). Unfortunately, there are few treatment options for leishmaniasis, and the side effects can be very serious. In view of the foregoing facts, there is still an exigent need to develop novel antileishmanial agents.
Chauhan et al. investigated 8,9-dihydrocoscinamide B (75, Figure 23) for its anti-leishmaniasis properties. This molecule is the reduced derivative of coscinamide B, a bis-indole ketoamide isolated from a marine sponge possessing cytoprotective activity against HIV. They also synthesized four analogues maintaining the coscinamide scaffold (76, Figure 23) and a small series of indolylglyoxylamide derivatives bearing different amines as substituents of the ketoamide moiety (77, Figure 23) [200].
The molecules were tested for their inhibitory capacity in both luciferase-promastigote and luciferase-amastigote systems at 10 μg/mL concentration, highlighting the better anti-leishmaniasis activity of coscinamide derivatives compared to the differently substituted indolylglyoxylamides. Particularly, 8,9-dihydrocoscinamide B 75 and the methylated derivative 78 (Figure 23) showed the most potent inhibitory percentage in both systems (100 and 99% in promastigote system; 97.72 and 97.78% in amastigote system) [200].
In a subsequent study, Chauhan et al. synthesized a series of indolylglyoxylamide derivatives, introducing a carboxylic moiety at the alpha position to the ketoamide nitrogen in the form of either a methyl ester 79 or carboxylic acid 80 (Figure 24). Moreover, prompted by the good anti-leishmaniasis activity shown by some β-carboline derivatives [201,202,203], another series of indolylglyoxylamides 81 bearing various tetrahydro-β-carboline moieties was developed (Figure 24) [204].
Most methyl ester derivatives showed low activity and selectivity when tested for activity in vitro against transgenic L. donovani amastigotes. They also showed an unfavorable toxicity profile. Hydrolysis of methyl ester to carboxylic acid led to a complete loss of activity.
On the contrary, indolylglyoxylamide derivatives 80 displayed good biological activity, with IC50 values in the range of 3.79–8.04 µM, thus representing promising candidates for the development of new antileishmanial drugs. The antileishmanial activity was influenced by varying the substitution pattern on the pendant phenyl ring of tetrahydro-β-carbolines, with electron-withdrawing groups more potent than electron-donating ones (81). The effect of a withdrawing group on the indole ring was studied by introducing a bromine substituent at position 5, but this led to an increase in toxicity without significantly impacting the activity. Different isomers of tetrahydro-β-carbolines did not affect the activity. Compound 82 (Figure 24), bearing an ethyl group at the para position of the pendant phenyl ring, appeared to be the molecule with the best profile, with an IC50 value of 5.17 µM, toxicity above 100 μM, and a high selectivity index of 31.48—12- and 5-fold greater than the standard drugs pentamidine and sodium stibogluconate (SSG), respectively [204].

9. Indolylglyoxylamides with Antiprion Properties

The indolylglyoxylamide moiety, with its ability to form noncovalent interactions, has found application in the field of antiprion agents, with remarkably potent compounds having been generated. Prion diseases, or transmissible spongiform encephalopathies (TSEs), are a group of progressive neurodegenerative diseases that affect both humans and animals. In TSEs, normal cellular prion protein (PrPC) is converted into an insoluble aggregate conformer PrPSc, in which “Sc” stands for scrapie, the prion disease of sheep and goats, which is thought to be infectious. The death of neuronal cells in TSEs is caused by these aggregates, resulting in vacuolization and the characteristic spongiform degeneration of brain tissue. Despite the high levels of expression in neurons and conservation among mammalian species, the physiological role of PrPC is still generally unclear, even though it appears to be crucially involved in neuroprotection, cell adhesion, and iron metabolism [205,206]. Thompson et al. designed and synthesized a vast series of indol-3-ylglyoxylamides with a general structure 83 (Figure 25). The choice of this structure, obtained from a scrapie-infected mouse brain (SMB) cell line screening assay, was determined by the extensive array of drug candidates bearing this moiety in either the clinical or preclinical stage with different biological activities [2,207].
The compounds’ capability to inhibit PrPSc formation was tested in a prion-infected cell line (SMB) of mesodermal origins, and nanomolar activity was demonstrated only by indol-3-ylglyoxylamide derivatives, featuring a aniline moiety substituted at the para position with an aromatic heterocycle bearing at least one hydrogen-bond acceptor (84 and 85, EC50 6 nM and 1 nM, respectively, Figure 25) [207].
Substitutions at the C-4 to C-7 positions of the indole ring (86, Figure 25) were investigated, which emphasized that only modification at C-6 can improve antiprion activity. Specifically, the introduction of strongly electron-withdrawing groups at C-6 delivered compounds with an optimal antiprion effect (compounds 87 and 88, EC50 6.1 nM and 1.2 nM, respectively, Figure 25) [208]. Toxicity profiles were determined in a zebrafish model with more than half of the tested compounds showing no effect on zebrafish survival, including the most potent candidates. Derivatization at R1 with methyl or morpholine was not tolerated, yielding a mortality rate of at least 20%. Interestingly, all derivatives substituted at the 6-position displayed enhanced microsomial stability, suggesting that this position is likely a metabolic site for unsubstituted compounds [208].
A subsequent series of antiprion agents was synthesized by Thompson et al., first focusing on the para substitution of the aniline and then modifying the ketoamide functionality [209]. The results from this investigation corroborated the previous findings that the best phenyl substituent at the para position (R1) of indol-3-ylglyoxylamide derivatives is a 5-member aromatic ring containing at least two heteroatoms. Additional heteroatoms can be tolerated if at least one is oxygen; further modification of the heterocycle is generally detrimental [209].
Detailed modifications were carried out on the glyoxylamide moiety, producing compounds with reduced potency: (1) formation of 3-(aminoacetyl)indoles and indole-3-acetamides by replacing either carbonyl with a methylene group, (2) the introduction of a maleimide bridge, and (3) an increment in the distance between the two carbonyls via the introduction of one or two carbon spacers. After analyzing the two series lacking either the amide carbonyl or the α-keto carbonyl, it was seen that the latter produced a more serious reduction in activity, suggesting its more substantial role in conferring potency to the molecules [209].

10. Others

10.1. A2B Allosteric Modulators

Among adenosine receptors (ARs), the A2B subtype displays lower affinity for adenosine with respect to the other subtypes (A1, A2A, and A3), and it is therefore triggered only at micromolar concentrations of adenosine. Under physiological conditions, extracellular concentration of adenosine is about 100 nM, whereas under stress conditions, it increases to micromolar levels, thus also activating A2BAR. A2BAR is involved in many physio-pathological conditions, including inflammatory processes, angiogenesis, glucose metabolism, tumor growth, and modulation of the function of cardiac fibroblasts, which are involved in many cardiovascular diseases, like chronic heart failure. For these reasons, A2BAR represents an important pharmacological target for a variety of diseases, such as diabetes, tumors, cardiovascular diseases, and pulmonary fibrosis.
In this context, Taliani and colleagues employed indol-3-ylglyoxylamide to discover novel AR ligands that, surprisingly, acted as A2BAR-positive allosteric modulators: compounds 89 and 90 (Figure 26) [210]. Compound 90, selected as representative for a deeper biological characterization, showed the ability to increase the efficacy but not the potency of A2B agonists, such as adenosine, NECA and BAY60-6583 [211], and the ability to promote MSC differentiation to osteoblasts and bone formation [212].
Subsequently, starting from the basic structures of 89 and 90, a structurally related series of compounds was developed by the same authors [213]. When evaluated for mineralization activity in the presence of the orthosteric agonist BAY60-6583, the most effective compound was 91 (Figure 26), which showed higher activity with respect to 90. Of note, compound 91 also significantly enhanced mineralization in the absence of BAY60-6583, thus highlighting its potential use as an anti-osteoporosis agent, as it was able to increase mineralization in experimental conditions close to physio-pathological ones. Finally, functional assays were performed, which confirmed allosteric modulator behavior at A2BAR for 91. Also in this case, 91 enhanced the efficacy of NECA without affecting its potency.

10.2. Carbonic Anhydrase Activators

Recently, an involvement of carbonic anhydrase (CA) activation in cognitive and memory disorders was demonstrated [214,215]. In fact, significantly reduced levels of different cerebral isoforms of α-CAs were found in the hippocampus of patients with Alzheimer’s disease. Moreover, following the administration of the activator phenylalanine to experimental animals, a pharmacological improvement in synaptic efficacy, spatial learning, and memory was produced. All of these observations suggest that these enzymes may represent a potential therapeutic target to identify useful agents to treat neurodegenerative diseases. In particular, CA activators (CAAs) act by restoring the active form of the enzyme, working as an alternative or additional site for proton transfer, accelerating the reaction. Exhaustive SARs evidenced a lipophilic aromatic ring substituted with a proton-shuttling moiety by means of a flexible chain as basic structural requirements to develop active CAAs.
In this connection, a series of indolylglyoxylamides featuring a basic protonable or a polar group at the 3- or 5-position was developed (9295, Figure 27) [216]. The choice of this moiety came from the longstanding experience of the research group with the development of indolylglyoxylamides with in vivo and/or in vitro activity [4,35,51,52,57,217] together with the widely recognized role of the alpha-keto amide moiety as a privileged motif in medicinal chemistry [11].
Biological assays evidenced that the CA isoform most sensitive to activation by all compounds was the cytosolic isoform VII, also known as the brain-associated isoform, as it is mainly expressed only in the brain [218].
The most active compounds were indole-based derivatives featuring at the 3-position a carboxamide or an ethyl ester moiety and the protonable/polar group at the 5-position. These latter compounds increased the release of brain-derived neurotrophic factor (BDNF), acting on the CNS by increasing the growth and differentiation of new neurons [219].

11. Conclusions

The aim of this review was to emphasize to medicinal chemists how the indolylglyoxylamide scaffold may serve as a versatile option to develop compounds with different therapeutic applications. Because of its suitability for a vast number of different modifications that enable interaction with specific molecular targets and produce desirable therapeutic effects, indolyglyoxylamide functionality may be considered a unique combination of privileged structures in medicinal chemistry, namely, the indole ring and the α-ketoamide moiety.
The indole ring features ease of functionalization and the ability both to engage H-bond interactions by means of its nitrogen proton and to establish π-π stacking or cation-π interactions due to its aromaticity. The α-ketoamide moiety may be exploited to modulate the conformation of a lead compound by increasing or decreasing its structural rigidity and to confer additional ability to establish hydrogen bonds in order to improve its potency and/or selectivity to a specific target. As an example, Da Settimo et al. were able to modulate the affinity of indolylglyoxylamide derivatives towards BzR subtypes by varying substituents at the 5-position of the indole and on the ketoamide nitrogen [26,27,28,29,34,35].
Furthermore, the α-ketoamide moiety possesses a unique pharmacokinetic profile showing improved membrane permeability compared to α-ketoacids and enhanced stability toward plasma esterases than α-ketoesters, together with a higher resistance against proteolytic cleavage. As reported by del Rosario García-Lozano et al., the α-ketoamide derivative was not the most potent in terms of antiviral activity against Dengue and Zika viruses but presented a superior predicted pharmacokinetic profile [164].
To increase the potency and/or to enhance the selectivity of a lead compound, as well as to improve its pharmacokinetic properties and/or reduce its toxicity and acquire novel chemical space to secure intellectual property, medicinal chemists may exploit bioisosterism to rationally modify its structure. The new bioisosteric molecule may show structural changes in molecular size, shape, electronic distribution, polarizability, pKa, or dipole that can deeply ameliorate the pharmacological activity of the original compound. In this view, the indolylglyoxylamide moiety may be fruitfully used by medicinal chemists to bioisosterically replace a tricyclic heterocycle to modulate its conformation by decreasing its structural rigidity and, possibly, to confer a new capacity to establish stronger and/or more selective interactions with a target protein. Furthermore, the two α-ketoamide electron-rich oxygen atoms may represent further points of interaction with a target, thus playing a crucial role in enhancing the affinity and selectivity of the compound for a specific protein, especially if the protein is prone to forming hydrogen bonds. In this vein, as an example, Primofiore et al. bioisosterically substituted the β-carboline scaffold with the indolylglyoxylamide one, obtaining highly potent BzR ligands [220]. The indolylglyoxylamide scaffold, having a more flexible structure than a completely planar β-carboline, assumes a planar or pseudoplanar conformation in the interaction with the receptor site, as demonstrated by theoretical calculations showing the indole aromatic ring and the glyoxylamide moiety to lie approximately in the same plane. These results confirm how medicinal chemists may take advantage of the bioisosteric substitution of a tricyclic heteroaromatic scaffold with the more flexible indolylglyoxylamide to develop novel lead compounds with improved pharmacological properties.
We are aware that in silico methods such as virtual screening (VS) may rapidly provide a set of biologically active compounds in an economically efficient manner and that, very recently, the employment of artificial intelligence methods such as deep and machine learning has added value in small-molecule drug discovery. However, in our opinion, one of the strategies that can aid medicinal chemists in the discovery of new drugs in a shorter amount of time with respect to other strategies is the exploitation of privileged structures and molecular fragments that are able to interact with more than one target. In this overview, we evidenced how the indolylglyoxylamide fragment was exploited in numerous applications to generate biologically active compounds and clinical candidates, primarily as sedative/hypnotics, anxiolytics, antitumorals, antibacterials, antivirals, anti-inflammatories, and antiprions, suggesting its prominent place in drug development. These data unquestionably support how the use of privileged scaffolds in drug design provides access to better success rates by overcoming the intrinsic hurdles connected with time-consuming drug development that rely on structures that are not yet validated.
To conclude, the authors believe that this overview, highlighting indolylglyoxylamide as a unique moiety and emphasizing its peculiar role as valuable option within future drug discovery programs, may aid medicinal chemistry to develop novel agents that are increasingly potent, efficient, and safe.

Author Contributions

Conceptualization, E.B. (Elisabetta Barresi), M.R., S.T. and F.D.S.; writing—original draft preparation, E.B. (Elisabetta Barresi) and M.R.; writing—review and editing, E.B. (Elisabetta Barresi), M.R. and S.T.; visualization, E.B. (Emma Baglini), M.V., S.S. and V.P.; supervision, S.T. and F.D.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the University of Pisa, grant numbers PRA_2022_19 and PRA_2022_60.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Evans, B.E.; Rittle, K.E.; Bock, M.G.; DiPardo, R.M.; Freidinger, R.M.; Whitter, W.L.; Lundell, G.F.; Veber, D.F.; Anderson, P.S.; Chang, R.S.L.; et al. Methods for Drug Discovery: Development of Potent, Selective, Orally Effective Cholecystokinin Antagonists. J. Med. Chem. 1988, 31, 2235–2246. [Google Scholar] [CrossRef] [PubMed]
  2. de Sa Alves, F.; Barreiro, E.; Manssour Fraga, C. From Nature to Drug Discovery: The Indole Scaffold as a “Privileged Structure”. Mini Rev. Med. Chem. 2009, 9, 782–793. [Google Scholar] [CrossRef] [PubMed]
  3. Salerno, S.; Barresi, E.; Baglini, E.; Poggetti, V.; Da Settimo, F.; Taliani, S. Target-Based Anticancer Indole Derivatives for the Development of Anti-Glioblastoma Agents. Molecules 2023, 28, 2587. [Google Scholar] [CrossRef] [PubMed]
  4. Taliani, S.; Da Settimo, F.; Martini, C.; Laneri, S.; Novellino, E.; Greco, G. Exploiting the Indole Scaffold to Design Compounds Binding to Different Pharmacological Targets. Molecules 2020, 25, 2331. [Google Scholar] [CrossRef]
  5. Shimazaki, Y.; Yajima, T.; Takani, M.; Yamauchi, O. Metal Complexes Involving Indole Rings: Structures and Effects of Metal–Indole Interactions. Coord. Chem. Rev. 2009, 253, 479–492. [Google Scholar] [CrossRef]
  6. Da Settimo, F.; Lucacchini, A.; Marini, A.M.; Martini, C.; Primofiore, G.; Senatore, G.; Taliani, S. Isosteric Replacement of the Indole Nucleus by Benzothiophene and Benzofuran in a Series of Indolylglyoxylylamine Derivatives with Partial Agonist Activity at the Benzodiazepine Receptor. Eur. J. Med. Chem. 1996, 31, 951–956. [Google Scholar] [CrossRef]
  7. Bianucci, A.M.; Da Settimo, A.; Da Settimo, F.; Primofiore, G.; Martini, C.; Giannaccini, G.; Lucacchini, A. Benzodiazepine Receptor Affinity and Interaction of Some N-(Indol-3-Ylglyoxylyl)Amine Derivatives. J. Med. Chem. 1992, 35, 2214–2220. [Google Scholar] [CrossRef]
  8. Bondensgaard, K.; Ankersen, M.; Thøgersen, H.; Hansen, B.S.; Wulff, B.S.; Bywater, R.P. Recognition of Privileged Structures by G-Protein Coupled Receptors. J. Med. Chem. 2004, 47, 888–899. [Google Scholar] [CrossRef]
  9. Rad, R.; Mracec, M.; Mracec, M.; Oprea, T. The Privileged Structures Hypothesis for G Protein-Coupled Receptors—Some Preliminary Results. Rev. Roum. Chim. 2007, 52, 853–858. [Google Scholar]
  10. Desiraju, G.R. Supramolecular Synthons in Crystal Engineering—A New Organic Synthesis. Angew. Chemie Int. Ed. English 1995, 34, 2311–2327. [Google Scholar] [CrossRef]
  11. Robello, M.; Barresi, E.; Baglini, E.; Salerno, S.; Taliani, S.; Settimo, F. Da The Alpha Keto Amide Moiety as a Privileged Motif in Medicinal Chemistry: Current Insights and Emerging Opportunities. J. Med. Chem. 2021, 64, 3508–3545. [Google Scholar] [CrossRef]
  12. Etter, M.C. Encoding and Decoding Hydrogen-Bond Patterns of Organic Compounds. Acc. Chem. Res. 1990, 23, 120–126. [Google Scholar] [CrossRef]
  13. Morin, C.M.; Jarrin, D.C. Epidemiology of Insomnia: Prevalence, Course, Risk Factors, and Public Health Burden. Sleep Med. Clin. 2022, 17, 173–191. [Google Scholar] [CrossRef]
  14. Mitler, M.M. Nonselective and Selective Benzodiazepine Receptor Agonists—Where Are We Today? Sleep 2000, 23, S39–S47. [Google Scholar]
  15. Barbera, J.; Shaprio, C. Benefit-Risk Assessment of Zaleplon in the Treatment of Insomnia. Drug Saf. 2005, 28, 301–318. [Google Scholar] [CrossRef]
  16. Sanger, D.J.; Benavides, J.; Perrault, G.; Morel, E.; Cohen, C.; Joly, D.; Zivkovic, B. Recent Developments in the Behavioral Pharmacology of Benzodiazepine (ω) Receptors: Evidence for the Functional Significance of Receptor Subtypes. Neurosci. Biobehav. Rev. 1994, 18, 355–372. [Google Scholar] [CrossRef]
  17. Terzano, M.G.; Rossi, M.; Palomba, V.; Smerieri, A.; Parrino, L. New Drugs for Insomnia: Comparative Tolerability of Zopiclone, Zolpidem and Zaleplon. Drug Saf. 2003, 26, 261–282. [Google Scholar] [CrossRef]
  18. Krystal, A.D.; Walsh, J.K.; Laska, E.; Caron, J.; Amato, D.A.; Wessel, T.C.; Roth, T. Sustained Efficacy of Eszopiclone Over 6 Months of Nightly Treatment: Results of a Randomized, Double-Blind, Placebo-Controlled Study in Adults with Chronic Insomnia. Sleep 2003, 26, 793–799. [Google Scholar] [CrossRef]
  19. Roth, T.; Soubrane, C.; Titeux, L.; Walsh, J.K.; on behalf of the Zoladult Study Group. Efficacy and Safety of Zolpidem-MR: A Double-Blind, Placebo-Controlled Study in Adults with Primary Insomnia. Sleep Med. 2006, 7, 397–406. [Google Scholar] [CrossRef]
  20. Farber, R.H.; Burke, P.J. Post-Bedtime Dosing with Indiplon in Adults and the Elderly: Results from Two Placebo-Controlled, Active Comparator Crossover Studies in Healthy Volunteers. Curr. Med. Res. Opin. 2008, 24, 837–846. [Google Scholar] [CrossRef]
  21. Sieghart, W.; Sperk, G. Subunit Composition, Distribution and Function of GABA-A Receptor Subtypes. Curr. Top. Med. Chem. 2005, 2, 795–816. [Google Scholar] [CrossRef] [PubMed]
  22. Möhler, H.; Fritschy, J.M.; Rudolph, U. A New Benzodiazepine Pharmacology. J. Pharmacol. Exp. Ther. 2002, 300, 2–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Whiting, P.J. GABA-A Receptor Subtypes in the Brain: A Paradigm for CNS Drug Discovery? Drug Discov. Today 2003, 8, 445–450. [Google Scholar] [CrossRef] [PubMed]
  24. Da Settimo, F.; Taliani, S.; Trincavelli, M.; Montali, M.; Martini, C. GABAA/Bz Receptor Subtypes as Targets for Selective Drugs. Curr. Med. Chem. 2007, 14, 2680–2701. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, W.; Koehler, K.F.; Zhang, P.; Cook, J.M. Development of a Comprehensive Pharmacophore Model for the Benzodiazepine Receptor. Drug Des. Discov. 1995, 12, 193–248. [Google Scholar]
  26. Da Settimo, A.; Primofiore, G.; Da Settimo, F.; Marini, A.M.; Novellino, E.; Greco, G.; Martini, C.; Giannaccini, G.; Lucacchini, A. Synthesis, Structure-Activity Relationships, and Molecular Modeling Studies of N-(Indol-3-Ylglyoxylyl)Benzylamine Derivatives Acting at the Benzodiazepine Receptor. J. Med. Chem. 1996, 39, 5083–5091. [Google Scholar] [CrossRef]
  27. Primofiore, G.; Da Settimo, F.; Taliani, S.; Marini, A.M.; Novellino, E.; Greco, G.; Lavecchia, A.; Besnard, F.; Trincavelli, L.; Costa, B.; et al. Novel N-(Arylalkyl)Indol-3-Ylglyoxylylamides Targeted as Ligands of the Benzodiazepine Receptor: Synthesis, Biological Evaluation, and Molecular Modeling Analysis of the Structure-Activity Relationships. J. Med. Chem. 2001, 44, 2286–2297. [Google Scholar] [CrossRef]
  28. Primofiore, G.; Taliani, S.; Da Settimo, F.; Marini, A.M.; La Motta, C.; Simorini, F.; Patrizi, M.P.; Sergianni, V.; Novellino, E.; Greco, G.; et al. Novel N-Substituted Indol-3-Ylglyoxylamides Probing the LDi and L1/L2 Lipophilic Regions of the Benzodiazepine Receptor Site in Search for Subtype-Selective Ligands. J. Med. Chem. 2007, 50, 1627–1634. [Google Scholar] [CrossRef]
  29. He, X.; Huang, Q.; Ma, C.; Yu, S.; McKernan, R.; Cook, J.M. Pharmacophore/Receptor Models for GABA(A)/BzR Alpha2beta3gamma2, Alpha3beta3gamma2 and Alpha4beta3gamma2 Recombinant Subtypes. Included Volume Analysis and Comparison to Alpha1beta3gamma2, Alpha5beta3gamma2, and Alpha6beta3gamma2 Subtypes. Drug Des. Discov. 2000, 17, 131–171. [Google Scholar]
  30. Atack, J.R. Anxioselective Compounds Acting at the GABA(A) Receptor Benzodiazepine Binding Site. Curr. Drug Targets CNS Neurol. Disord. 2003, 2, 213–232. [Google Scholar] [CrossRef]
  31. Atack, J.R. The Benzodiazepine Binding Site of GABAA Receptors as a Target for the Development of Novel Anxiolytics. Expert Opin. Investig. Drugs 2005, 14, 601–618. [Google Scholar] [CrossRef]
  32. Sigel, E.; Ernst, M. The Benzodiazepine Binding Sites of GABAA Receptors. Trends Pharmacol. Sci. 2018, 39, 659–671. [Google Scholar] [CrossRef]
  33. Skolnick, P. Anxioselective Anxiolytics: On a Quest for the Holy Grail. Trends Pharmacol. Sci. 2012, 33, 611–620. [Google Scholar] [CrossRef] [Green Version]
  34. Primofiore, G.; Da Settimo, F.; Marini, A.M.; Taliani, S.; La Motta, C.; Simorini, F.; Novellino, E.; Greco, G.; Cosimelli, B.; Ehlardo, M.; et al. Refinement of the Benzodiazepine Receptor Site Topology by Structure-Activity Relationships of New N-(Heteroarylmethyl)Indol-3- Ylglyoxylamides. J. Med. Chem. 2006, 49, 2489–2495. [Google Scholar] [CrossRef]
  35. Taliani, S.; Cosimelli, B.; Da Settimo, F.; Marini, A.M.; La Motta, C.; Simorini, F.; Salerno, S.; Novellino, E.; Greco, G.; Cosconati, S.; et al. Identification of Anxiolytic/Nonsedative Agents among Indol-3- Ylglyoxylamides Acting as Functionally Selective Agonists at the γ-Aminobutyric Acid-A (GABAA) A2 Benzodiazepine Receptor. J. Med. Chem. 2009, 52, 3723–3734. [Google Scholar] [CrossRef]
  36. Bourin, M.; Hascoët, M. The Mouse Light/Dark Box Test. Eur. J. Pharmacol. 2003, 463, 55–65. [Google Scholar] [CrossRef]
  37. Olsen, R.W. GABAA Receptor: Positive and Negative Allosteric Modulators. Neuropharmacology 2018, 136, 10–22. [Google Scholar] [CrossRef]
  38. Papadopoulos, V.; Baraldi, M.; Guilarte, T.R.; Knudsen, T.B.; Lacapère, J.J.; Lindemann, P.; Norenberg, M.D.; Nutt, D.; Weizman, A.; Zhang, M.R.; et al. Translocator Protein (18 KDa): New Nomenclature for the Peripheral-Type Benzodiazepine Receptor Based on Its Structure and Molecular Function. Trends Pharmacol. Sci. 2006, 27, 402–409. [Google Scholar] [CrossRef]
  39. Taliani, S.; Da Settimo, F.; Da Pozzo, E.; Chelli, B.; Martini, C. Translocator Protein Ligands as Promising Therapeutic Tools for Anxiety Disorders. Curr. Med. Chem. 2009, 16, 3359–3380. [Google Scholar] [CrossRef]
  40. Le Fur, G.; Vaucher, N.; Perrier, M.L.; Flamier, A.; Benavides, J.; Renault, C.; Dubroeucq, M.C.; Guérémy, C.; Uzan, A. Differentiation between Two Ligands for Peripheral Benzodiazepine Binding Sites, [3H]R05-4864 and [3H]PK 11195, by Thermodynamic Studies. Life Sci. 1983, 33, 449–457. [Google Scholar] [CrossRef]
  41. Romeo, E.; Auta, J.; Kozikowski, A.P.; Ma, D.; Papadopoulos, V.; Puia, G.; Costa, E.; Guidotti, A. 2-Aryl-3-Indoleacetamides (FGIN-1): A New Class of Potent and Specific Ligands for the Mitochondrial DBI Receptor (MDR). J. Pharmacol. Exp. Ther. 1992, 262, 971–978. [Google Scholar] [PubMed]
  42. Kozikowski, A.P.; Ma, D.; Brewer, J.; Sun, S.; Costa, E.; Romeo, E.; Guidotti, A. Chemistry, Binding Affinities, and Behavioral Properties of a New Class of “Antineophobic” Mitochondrial DBI Receptor Complex (MDRC) Ligands. J. Med. Chem. 1993, 36, 2908–2920. [Google Scholar] [CrossRef] [PubMed]
  43. Costa, B.; Da Pozzo, E.; Martini, C. Translocator Protein and Steroidogenesis. Biochem. J. 2018, 475, 901–904. [Google Scholar] [CrossRef] [PubMed]
  44. Papadopoulos, V.; Lecanu, L.; Brown, R.C.; Han, Z.; Yao, Z.X. Peripheral-Type Benzodiazepine Receptor in Neurosteroid Biosynthesis, Neuropathology and Neurological Disorders. Neuroscience 2006, 138, 749–756. [Google Scholar] [CrossRef] [PubMed]
  45. Nothdurfter, C.; Rammes, G.; Baghai, T.C.; Schüle, C.; Schumacher, M.; Papadopoulos, V.; Rupprecht, R. Translocator Protein (18 KDa) as a Target for Novel Anxiolytics with a Favourable Side-Effect Profile. J. Neuroendocrinol. 2012, 24, 82–92. [Google Scholar] [CrossRef]
  46. Nothdurfter, C.; Rupprecht, R.; Rammes, G. Recent Developments in Potential Anxiolytic Agents Targeting GABAA/BzR Complex or the Translocator Protein (18 kDa) (TSPO). Curr. Top. Med. Chem. 2012, 12, 360–370. [Google Scholar] [CrossRef]
  47. Primofiore, G.; Da Settimo, F.; Taliani, S.; Simorini, F.; Patrizi, M.P.; Novellino, E.; Greco, G.; Abignente, E.; Costa, B.; Chelli, B.; et al. N,N-Dialkyl-2-Phenylindol-3-Ylglyoxylamides. A New Class of Potent and Selective Ligands at the Peripheral Renzodiazepine Receptor. J. Med. Chem. 2004, 47, 1852–1855. [Google Scholar] [CrossRef]
  48. Da Settimo, F.; Simorini, F.; Taliani, S.; La Motta, C.; Marini, A.M.; Salerno, S.; Bellandi, M.; Novellino, E.; Greco, G.; Cosimelli, B.; et al. Anxiolytic-like Effects of N,N-Dialkyl-2-Phenylindol-3-Ylglyoxylamides by Modulation of Translocator Protein Promoting Neurosteroid Biosynthesis. J. Med. Chem. 2008, 51, 5798–5806. [Google Scholar] [CrossRef]
  49. Barresi, E.; Bruno, A.; Taliani, S.; Cosconati, S.; Da Pozzo, E.; Salerno, S.; Simorini, F.; Daniele, S.; Giacomelli, C.; Marini, A.M.; et al. Deepening the Topology of the Translocator Protein Binding Site by Novel N,N-Dialkyl-2-Arylindol-3-Ylglyoxylamides. J. Med. Chem. 2015, 58, 6081–6092. [Google Scholar] [CrossRef]
  50. Barresi, E.; Robello, M.; Costa, B.; Da Pozzo, E.; Baglini, E.; Salerno, S.; Da Settimo, F.; Martini, C.; Taliani, S. An Update into the Medicinal Chemistry of Translocator Protein (TSPO) Ligands. Eur. J. Med. Chem. 2021, 209, 112924. [Google Scholar] [CrossRef]
  51. Da Pozzo, E.; Giacomelli, C.; Costa, B.; Cavallini, C.; Taliani, S.; Barresi, E.; Da Settimo, F.; Martini, C. TSPO PIGA Ligands Promote Neurosteroidogenesis and Human Astrocyte Well-Being. Int. J. Mol. Sci. 2016, 17, 1028. [Google Scholar] [CrossRef]
  52. Costa, B.; Da Pozzo, E.; Chelli, B.; Simola, N.; Morelli, M.; Luisi, M.; Maccheroni, M.; Taliani, S.; Simorini, F.; Da Settimo, F.; et al. Anxiolytic Properties of a 2-Phenylindolglyoxylamide TSPO Ligand: Stimulation of In Vitro Neurosteroid Production Affecting GABAA Receptor Activity. Psychoneuroendocrinology 2011, 36, 463–472. [Google Scholar] [CrossRef]
  53. Tu, L.N.; Morohaku, K.; Manna, P.R.; Pelton, S.H.; Butler, W.R.; Stocco, D.M.; Selvaraj, V. Peripheral Benzodiazepine Receptor/Translocator Protein Global Knock-out Mice Are Viable with No Effects on Steroid Hormone Biosynthesis. J. Biol. Chem. 2014, 289, 27444–27454. [Google Scholar] [CrossRef] [Green Version]
  54. Morohaku, K.; Pelton, S.H.; Daugherty, D.J.; Butler, W.R.; Deng, W.; Selvaraj, V. Translocator Protein/Peripheral Benzodiazepine Receptor Is Not Required for Steroid Hormone Biosynthesis. Endocrinology 2014, 155, 89–97. [Google Scholar] [CrossRef] [Green Version]
  55. M Scarf, A.; M Auman, K.; Kassiou, M. Is There Any Correlation between Binding and Functional Effects at the Translocator Protein (TSPO) (18 KDa)? Curr. Mol. Med. 2012, 12, 387–397. [Google Scholar] [CrossRef]
  56. Costa, B.; Da Pozzo, E.; Giacomelli, C.; Barresi, E.; Taliani, S.; Da Settimo, F.; Martini, C. TSPO Ligand Residence Time: A New Parameter to Predict Compound Neurosteroidogenic Efficacy. Sci. Rep. 2016, 6, 18164. [Google Scholar] [CrossRef] [Green Version]
  57. Bruno, A.; Barresi, E.; Simola, N.; Da Pozzo, E.; Costa, B.; Novellino, E.; Da Settimo, F.; Martini, C.; Taliani, S.; Cosconati, S. Unbinding of Translocator Protein 18 KDa (TSPO) Ligands: From In Vitro Residence Time to In Vivo Efficacy via in Silico Simulations. ACS Chem. Neurosci. 2019, 10, 3805–3814. [Google Scholar] [CrossRef]
  58. Germelli, L.; Da Pozzo, E.; Giacomelli, C.; Tremolanti, C.; Marchetti, L.; Wetzel, C.H.; Barresi, E.; Taliani, S.; Settimo, F.D.; Martini, C.; et al. De Novo Neurosteroidogenesis in Human Microglia: Involvement of the 18 Kda Translocator Protein. Int. J. Mol. Sci. 2021, 22, 3115. [Google Scholar] [CrossRef]
  59. Corsi, F.; Baglini, E.; Barresi, E.; Salerno, S.; Cerri, C.; Martini, C.; Da Settimo Passetti, F.; Taliani, S.; Gargini, C.; Piano, I. Targeting TSPO Reduces Inflammation and Apoptosis in an In Vitro Photoreceptor-Like Model of Retinal Degeneration. ACS Chem. Neurosci. 2022, 13, 3188–3197. [Google Scholar] [CrossRef]
  60. Tremolanti, C.; Cavallini, C.; Meyer, L.; Klein, C.; Da Pozzo, E.; Costa, B.; Germelli, L.; Taliani, S.; Patte-Mensah, C.; Mensah-Nyagan, A.G. Translocator Protein Ligand PIGA1138 Reduces Disease Symptoms and Severity in Experimental Autoimmune Encephalomyelitis Model of Primary Progressive Multiple Sclerosis. Mol. Neurobiol. 2022, 59, 1744–1765. [Google Scholar] [CrossRef]
  61. Angeloni, E.; Germelli, L.; Marchetti, L.; Da Pozzo, E.; Tremolanti, C.; Wetzel, C.H.; Baglini, E.; Taliani, S.; Da Settimo, F.; Martini, C.; et al. The Human Microglial Surveillant Phenotype Is Preserved by de Novo Neurosteroidogenesis through the Control of Cholesterol Homeostasis: Crucial Role of 18 KDa Translocator Protein. Biochim. Biophys. Acta Mol. Basis Dis. 2023, 1869, 166751. [Google Scholar] [CrossRef] [PubMed]
  62. Chhikara, B.S.; Parang, K. Global Cancer Statistics 2022: The Trends Projection Analysis. Chem. Biol. Lett. 2023, 10, 451. [Google Scholar]
  63. Ganesh, K.; Massagué, J. Targeting Metastatic Cancer. Nat. Med. 2021, 27, 34–44. [Google Scholar] [CrossRef] [PubMed]
  64. Mothersill, C.; Seymour, C.B. Mechanisms and Implications of Genomic Instability and Other Delayed Effects of Ionizing Radiation Exposure. Mutagenesis 1998, 13, 421–426. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Charmsaz, S.; Collins, D.M.; Perry, A.S.; Prencipe, M. Novel Strategies for Cancer Treatment: Highlights from the 55th IACR Annual Conference. Cancers 2019, 11, 1125. [Google Scholar] [CrossRef] [Green Version]
  66. Shams, M.; Owczarczak, B.; Manderscheid-Kern, P.; Bellnier, D.A.; Gollnick, S.O. Development of Photodynamic Therapy Regimens That Control Primary Tumor Growth and Inhibit Secondary Disease. Cancer Immunol. Immunother. 2015, 64, 287–297. [Google Scholar] [CrossRef] [Green Version]
  67. Allen, C.; Her, S.; Jaffray, D.A. Radiotherapy for Cancer: Present and Future. Adv. Drug Deliv. Rev. 2017, 109, 1–2. [Google Scholar] [CrossRef]
  68. Yahya, E.B.; Alqadhi, A.M. Recent Trends in Cancer Therapy: A Review on the Current State of Gene Delivery. Life Sci. 2021, 269, 119087. [Google Scholar] [CrossRef]
  69. Bukowski, K.; Kciuk, M.; Kontek, R. Mechanisms of Multidrug Resistance in Cancer Chemotherapy. Int. J. Mol. Sci. 2020, 21, 3233. [Google Scholar] [CrossRef]
  70. Jordan, M.A. Mechanism of Action of Antitumor Drugs That Interact with Microtubules and Tubulin. Curr. Med. Chem. Anticancer Agents 2002, 2, 1–17. [Google Scholar] [CrossRef]
  71. Bacher, G.; Nickel, B.; Emig, P.; Vanhoefer, U.; Seeber, S.; Shandra, A.; Klenner, T.; Beckers, T. D-24851, a Novel Synthetic Microtubule Inhibitor, Exerts Curative Antitumoral Activity In Vivo, Shows Efficacy toward Multidrug-Resistant Tumor Cells, and Lacks Neurotoxicity. Cancer Res. 2001, 61, 392–399. [Google Scholar]
  72. Windebank, A.J. Chemotherapeutic Neuropathy: Current Opinion in Neurology. Curr. Opin. Neurol. 1999, 12, 565–571. [Google Scholar] [CrossRef]
  73. Rowinsky, E.K. The Development and Clinical Utility of the Taxane Class of Antimicrotubule Chemotherapy Agents. Annu. Rev. Med. 1997, 48, 353–374. [Google Scholar] [CrossRef]
  74. Kuppens, I.E.L.M.; Witteveen, P.O.; Schot, M.; Schuessler, V.M.; Daehling, A.; Beijnen, J.H.; Voest, E.E.; Schellens, J.H.M. Phase I Dose-Finding and Pharmacokinetic Trial of Orally Administered Indibulin (D-24851) to Patients with Solid Tumors. Invest. New Drugs 2007, 25, 227–235. [Google Scholar] [CrossRef]
  75. Oostendorp, R.L.; Witteveen, P.O.; Schwartz, B.; Vainchtein, L.D.; Schot, M.; Nol, A.; Rosing, H.; Beijnen, J.H.; Voest, E.E.; Schellens, J.H.M. Dose-Finding and Pharmacokinetic Study of Orally Administered Indibulin (D-24851) to Patients with Advanced Solid Tumors. Invest. New Drugs 2010, 28, 163–170. [Google Scholar] [CrossRef]
  76. Sorgel, F.; Kinzig, M. Pharmacokinetics of Gyrase Inhibitors, Part 2: Renal and Hepatic Elimination Pathways and Drug Interactions. Am. J. Med. 1993, 94, 56S–69S. [Google Scholar] [CrossRef]
  77. Hagen, S.E.; Domagala, J.; Gajda, C.; Lovdahl, M.; Tait, B.D.; Wise, E.; Holler, T.; Hupe, D.; Nouhan, C.; Urumov, A.; et al. 4-Hydroxy-5,6-Dihydropyrones as Inhibitors of HIV Protease: The Effect of Heterocyclic Substituents at C-6 on Antiviral Potency and Pharmacokinetic Parameters. J. Med. Chem. 2001, 44, 2319–2332. [Google Scholar] [CrossRef]
  78. Li, W.T.; Hwang, D.R.; Chen, C.P.; Shen, C.W.; Huang, C.L.; Chen, T.W.; Lin, C.H.; Chang, Y.L.; Chang, Y.Y.; Lo, Y.K.; et al. Synthesis and Biological Evaluation of N-Heterocyclic Indolyl Glyoxylamides as Orally Active Anticancer Agents. J. Med. Chem. 2003, 46, 1706–1715. [Google Scholar] [CrossRef]
  79. Hu, C.B.; Chen, C.P.; Yeh, T.K.; Song, J.S.; Chang, C.Y.; Chuu, J.J.; Tung, F.F.; Ho, P.Y.; Chen, T.W.; Lin, C.H.; et al. BPR0C261 Is a Novel Orally Active Antitumor Agent with Antimitotic and Anti-Angiogenic Activities. Cancer Sci. 2011, 102, 182–191. [Google Scholar] [CrossRef]
  80. Leu, J.D.; Lin, S.T.; Chen, C.T.; Chang, C.A.; Lee, Y.J. BPR0C261, An Analogous of Microtubule Disrupting Agent D-24851 Enhances the Radiosensitivity of Human Non-Small Cell Lung Cancer Cells via P53-Dependent and P53-Independent Pathways. Int. J. Mol. Sci. 2022, 23, 14083. [Google Scholar] [CrossRef]
  81. Vogelstein, B.; Lane, D.; Levine, A.J. Surfing the P53 Network. Nature 2000, 408, 307–310. [Google Scholar] [CrossRef] [PubMed]
  82. Choy, H. Taxanes in Combined Modality Therapy for Solid Tumors. Crit. Rev. Oncol. Hematol. 2001, 37, 237–247. [Google Scholar] [CrossRef] [PubMed]
  83. Li, W.T.; Yeh, T.K.; Song, J.S.; Yang, Y.N.; Chen, T.W.; Lin, C.H.; Chen, C.P.; Shen, C.C.; Hsieh, C.C.; Lin, H.L.; et al. BPR0C305, an Orally Active Microtubule-Disrupting Anticancer Agent. Anticancer. Drugs 2013, 24, 1047–1057. [Google Scholar] [CrossRef] [PubMed]
  84. Colley, H.E.; Muthana, M.; Danson, S.J.; Jackson, L.V.; Brett, M.L.; Harrison, J.; Coole, S.F.; Mason, D.P.; Jennings, L.R.; Wong, M.; et al. An Orally Bioavailable, Indole-3-Glyoxylamide Based Series of Tubulin Polymerization Inhibitors Showing Tumor Growth Inhibition in a Mouse Xenograft Model of Head and Neck Cancer. J. Med. Chem. 2015, 58, 9309–9333. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Chen, Y.J.; Huang, W.C.; Wei, Y.L.; Hsu, S.C.; Yuan, P.; Lin, H.Y.; Wistuba, I.I.; Lee, J.J.; Yen, C.J.; Su, W.C.; et al. Elevated BCRP/ABCG2 Expression Confers Acquired Resistance to Gefitinib in Wild-Type EGFR-Expressing Cells. PLoS ONE 2011, 6, e21428. [Google Scholar] [CrossRef] [PubMed]
  86. Guggilapu, S.D.; Lalita, G.; Reddy, T.S.; Prajapti, S.K.; Nagarsenkar, A.; Ramu, S.; Brahma, U.R.; Lakshmi, U.J.; Vegi, G.M.N.; Bhargava, S.K.; et al. Synthesis of C5-Tethered Indolyl-3-Glyoxylamide Derivatives as Tubulin Polymerization Inhibitors. Eur. J. Med. Chem. 2017, 128, 1–12. [Google Scholar] [CrossRef]
  87. Guggilapu, S.D.; Guntuku, L.; Reddy, T.S.; Nagarsenkar, A.; Sigalapalli, D.K.; Naidu, V.G.M.; Bhargava, S.K.; Bathini, N.B. Synthesis of Thiazole Linked Indolyl-3-Glyoxylamide Derivatives as Tubulin Polymerization Inhibitors. Eur. J. Med. Chem. 2017, 138, 83–95. [Google Scholar] [CrossRef]
  88. Brel, V.K.; Artyushin, O.I.; Chuprov-Netochin, R.N.; Leonov, S.V.; Semenova, M.N.; Semenov, V.V. Synthesis and Biological Evaluation of Indolylglyoxylamide Bisphosphonates, Antimitotic Microtubule-Targeting Derivatives of Indibulin with Improved Aqueous Solubility. Bioorg. Med. Chem. Lett. 2020, 30, 127635. [Google Scholar] [CrossRef]
  89. Tantak, M.P.; Gupta, V.; Nikhil, K.; Arun, V.; Singh, R.P.; Jha, P.N.; Shah, K.; Kumar, D. Sequential One-Pot Synthesis of Bis(Indolyl)Glyoxylamides: Evaluation of Antibacterial and Anticancer Activities. Bioorg. Med. Chem. Lett. 2016, 26, 3167–3171. [Google Scholar] [CrossRef] [Green Version]
  90. Tantak, M.P.; Wang, J.; Singh, R.P.; Kumar, A.; Shah, K.; Kumar, D. 2-(3′-Indolyl)-N-Arylthiazole-4-Carboxamides: Synthesis and Evaluation of Antibacterial and Anticancer Activities. Bioorg. Med. Chem. Lett. 2015, 25, 4225–4231. [Google Scholar] [CrossRef] [Green Version]
  91. Venkataramana Reddy, P.O.; Tantak, M.P.; Valdez, R.; Singh, R.P.; Singh, O.M.; Sadana, R.; Kumar, D. Synthesis and Biological Evaluation of Novel Carbazolyl Glyoxamides as Anticancer and Antibacterial Agents. RSC Adv. 2016, 6, 9379–9386. [Google Scholar] [CrossRef]
  92. Soni, J.P.; Nikitha Reddy, G.; Rahman, Z.; Sharma, A.; Spandana, A.; Phanindranath, R.; Dandekar, M.P.; Nagesh, N.; Shankaraiah, N. Synthesis and Cytotoxicity Evaluation of DNA-Interactive β-Carboline Indolyl-3-Glyoxamide Derivatives: Topo-II Inhibition and in Silico Modelling Studies. Bioorg. Chem. 2023, 131, 106313. [Google Scholar] [CrossRef]
  93. Tokala, R.; Thatikonda, S.; Vanteddu, U.S.; Sana, S.; Godugu, C.; Shankaraiah, N. Design and Synthesis of DNA-Interactive β-Carboline–Oxindole Hybrids as Cytotoxic and Apoptosis-Inducing Agents. ChemMedChem 2018, 13, 1909–1922. [Google Scholar] [CrossRef] [Green Version]
  94. Jadala, C.; Sathish, M.; Reddy, T.S.; Reddy, V.G.; Tokala, R.; Bhargava, S.K.; Shankaraiah, N.; Nagesh, N.; Kamal, A. Synthesis and In Vitro Cytotoxicity Evaluation of β-Carboline-Combretastatin Carboxamides as Apoptosis Inducing Agents: DNA Intercalation and Topoisomerase-II Inhibition. Bioorg. Med. Chem. 2019, 27, 3285–3298. [Google Scholar] [CrossRef]
  95. M O’Boyle, N.; J Meegan, M. Designed Multiple Ligands for Cancer Therapy. Curr. Med. Chem. 2011, 18, 4722–4737. [Google Scholar] [CrossRef]
  96. Villalonga-Planells, R.; Coll-Mulet, L.; Martínez-Soler, F.; Castaño, E.; Acebes, J.J.; Giménez-Bonafé, P.; Gil, J.; Tortosa, A. Activation of P53 by Nutlin-3a Induces Apoptosis and Cellular Senescence in Human Glioblastoma Multiforme. PLoS ONE 2011, 6, e18588. [Google Scholar] [CrossRef] [Green Version]
  97. Shoukrun, R.; Veenman, L.; Shandalov, Y.; Leschiner, S.; Spanier, I.; Karry, R.; Katz, Y.; Weisinger, G.; Weizman, A.; Gavish, M. The 18-KDa Translocator Protein, Formerly Known as the Peripheral-Type Benzodiazepine Receptor, Confers Proapoptotic and Antineoplastic Effects in a Human Colorectal Cancer Cell Line. Pharmacogenet. Genom. 2008, 18, 977–988. [Google Scholar] [CrossRef]
  98. Daniele, S.; Taliani, S.; Da Pozzo, E.; Giacomelli, C.; Costa, B.; Trincavelli, M.L.; Rossi, L.; La Pietra, V.; Barresi, E.; Carotenuto, A.; et al. Apoptosis Therapy in Cancer: The First Single-Molecule Co-Activating P53 and the Translocator Protein in Glioblastoma. Sci. Rep. 2014, 4, 4749. [Google Scholar] [CrossRef] [Green Version]
  99. Daniele, S.; La Pietra, V.; Barresi, E.; Di Maro, S.; Da Pozzo, E.; Robello, M.; La Motta, C.; Cosconati, S.; Taliani, S.; Marinelli, L.; et al. Lead Optimization of 2-Phenylindolylglyoxylyldipeptide Murine Double Minute (MDM)2/Translocator Protein (TSPO) Dual Inhibitors for the Treatment of Gliomas. J. Med. Chem. 2016, 59, 4526–4538. [Google Scholar] [CrossRef]
  100. Juchum, M.; Günther, M.; Laufer, S.A. Fighting Cancer Drug Resistance: Opportunities and Challenges for Mutation-Specific EGFR Inhibitors. Drug Resist. Updat. 2015, 20, 12–28. [Google Scholar] [CrossRef]
  101. Daniele, S.; Barresi, E.; Zappelli, E.; Marinelli, L.; Novellino, E.; Settimo, F.D.; Taliani, S.; Trincavelli, M.L.; Martini, C.; Daniele, S.; et al. Long Lasting MDM2/Translocator Protein Modulator: A New Strategy for Irreversible Apoptosis of Human Glioblastoma Cells. Oncotarget 2016, 7, 7866–7884. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Wang, D.; Lippard, S.J. Cellular Processing of Platinum Anticancer Drugs. Nat. Rev. Drug Discov. 2005, 4, 307–320. [Google Scholar] [CrossRef] [PubMed]
  103. Oun, R.; Moussa, Y.E.; Wheate, N.J. The Side Effects of Platinum-Based Chemotherapy Drugs: A Review for Chemists. Dalt. Trans. 2018, 47, 6645–6653. [Google Scholar] [CrossRef] [PubMed]
  104. Johnstone, T.C.; Suntharalingam, K.; Lippard, S.J. The Next Generation of Platinum Drugs: Targeted Pt(II) Agents, Nanoparticle Delivery, and Pt(IV) Prodrugs. Chem. Rev. 2016, 116, 3436–3486. [Google Scholar] [CrossRef] [Green Version]
  105. Aquino, M.A.S. Diruthenium and Diosmium Tetracarboxylates: Synthesis, Physical Properties and Applications. Coord. Chem. Rev. 1998, 170, 141–202. [Google Scholar] [CrossRef]
  106. Barresi, E.; Tolbatov, I.; Pratesi, A.; Notarstefano, V.; Baglini, E.; Daniele, S.; Taliani, S.; Re, N.; Giorgini, E.; Martini, C.; et al. A Mixed-Valence Diruthenium(Ii,Iii) Complex Endowed with High Stability: From Experimental Evidence to Theoretical Interpretation. Dalt. Trans. 2020, 49, 14520–14527. [Google Scholar] [CrossRef]
  107. Barresi, E.; Tolbatov, I.; Marzo, T.; Zappelli, E.; Marrone, A.; Re, N.; Pratesi, A.; Martini, C.; Taliani, S.; Da Settimo, F.; et al. Two Mixed Valence Diruthenium(II,III) Isomeric Complexes Show Different Anticancer Properties. Dalt. Trans. 2021, 50, 9643–9647. [Google Scholar] [CrossRef]
  108. Chiaverini, L.; Baglini, E.; Mannelli, M.; Poggetti, V.; Da Settimo, F.; Taliani, S.; Gamberi, T.; Barresi, E.; La Mendola, D.; Marzo, T. A Complex Bearing TSPO PIGA Ligand Coordinated to the [Au(PEt3)]+ Pharmacophore Is Highly Cytotoxic against Ovarian Cancer Cells. BioMetals 2023, 36, 1–8. [Google Scholar] [CrossRef]
  109. Landini, I.; Massai, L.; Cirri, D.; Gamberi, T.; Paoli, P.; Messori, L.; Mini, E.; Nobili, S. Structure-Activity Relationships in a Series of Auranofin Analogues Showing Remarkable Antiproliferative Properties. J. Inorg. Biochem. 2020, 208, 111079. [Google Scholar] [CrossRef]
  110. Nutma, E.; Ceyzériat, K.; Amor, S.; Tsartsalis, S.; Millet, P.; Owen, D.R.; Papadopoulos, V.; Tournier, B.B. Cellular Sources of TSPO Expression in Healthy and Diseased Brain. Eur. J. Nucl. Med. Mol. Imaging 2021, 49, 146–163. [Google Scholar] [CrossRef]
  111. Takhi, M.; Singh, G.; Murugan, C.; Thaplyyal, N.; Maitra, S.; Bhaskarreddy, K.M.; Amarnath, P.V.S.; Mallik, A.; Harisudan, T.; Trivedi, R.K.; et al. Novel and Potent Oxazolidinone Antibacterials Featuring 3-Indolylglyoxamide Substituents. Bioorg. Med. Chem. Lett. 2008, 18, 5150–5155. [Google Scholar] [CrossRef]
  112. Clemett, D.; Markham, A. Linezolid. Drugs 2000, 59, 815–827. [Google Scholar] [CrossRef]
  113. Shinabarger, D.L.; Marotti, K.R.; Murray, R.W.; Lin, A.H.; Melchior, E.P.; Swaney, S.M.; Dunyak, D.S.; Demyan, W.F.; Buysse, J.M. Mechanism of Action of Oxazolidinones: Effects of Linezolid and Eperezolid on Translation Reactions. Antimicrob. Agents Chemother. 1997, 41, 2132–2136. [Google Scholar] [CrossRef] [Green Version]
  114. Kloss, P.; Xiong, L.; Shinabarger, D.L.; Mankin, A.S. Resistance Mutations in 23 S RRNA Identify the Site of Action of the Protein Synthesis Inhibitor Linezolid in the Ribosomal Peptidyl Transferase Center. J. Mol. Biol. 1999, 294, 93–101. [Google Scholar] [CrossRef] [Green Version]
  115. Johnson, A.P.; Tysall, L.; Stockdale, M.W.; Woodford, N.; Kaufmann, M.E.; Warner, M.; Livermore, D.M.; Asboth, F.; Allerberger, F.J. Emerging Linezolid-Resistant Enterococcus Faecalis and Enterococcus Faecium Isolated from Two Austrian Patients in the Same Intensive Care Unit. Eur. J. Clin. Microbiol. Infect. Dis. 2002, 21, 751–754. [Google Scholar]
  116. Singh, P.; Verma, P.; Yadav, B.; Komath, S.S. Synthesis and Evaluation of Indole-Based New Scaffolds for Antimicrobial Activities—Identification of Promising Candidates. Bioorg. Med. Chem. Lett. 2011, 21, 3367–3372. [Google Scholar] [CrossRef]
  117. Mielczarek, M.; Devakaram, R.V.; Ma, C.; Yang, X.; Kandemir, H.; Purwono, B.; Black, D.S.; Griffith, R.; Lewis, P.J.; Kumar, N. Synthesis and Biological Activity of Novel Bis-Indole Inhibitors of Bacterial Transcription Initiation Complex Formation. Org. Biomol. Chem. 2014, 12, 2882–2894. [Google Scholar] [CrossRef]
  118. Li, S.A.; Cadelis, M.M.; Sue, K.; Blanchet, M.; Vidal, N.; Brunel, J.M.; Bourguet-Kondracki, M.L.; Copp, B.R. 6-Bromoindolglyoxylamido Derivatives as Antimicrobial Agents and Antibiotic Enhancers. Bioorganic Med. Chem. 2019, 27, 2090–2099. [Google Scholar] [CrossRef] [Green Version]
  119. Pieri, C.; Borselli, D.; Di Giorgio, C.; De Méo, M.; Bolla, J.M.; Vidal, N.; Combes, S.; Brunel, J.M. New Ianthelliformisamine Derivatives as Antibiotic Enhancers against Resistant Gram-Negative Bacteria. J. Med. Chem. 2014, 57, 4263–4272. [Google Scholar] [CrossRef]
  120. Cadelis, M.M.; Pike, E.I.W.; Kang, W.; Wu, Z.; Bourguet-Kondracki, M.L.; Blanchet, M.; Vidal, N.; Brunel, J.M.; Copp, B.R. Exploration of the Antibiotic Potentiating Activity of Indolglyoxylpolyamines. Eur. J. Med. Chem. 2019, 183, 111708. [Google Scholar] [CrossRef]
  121. Conti, M.; Jin, S.L. The Molecular Biology of Cyclic Nucleotide Phosphodiesterases. Prog. Nucleic Acid Res. Mol. Biol. 1999, 63, 1–38. [Google Scholar] [CrossRef] [PubMed]
  122. Torphy, T.J.; Stadel, J.M.; Burman, M.; Cieslinski, L.B.; McLaughlin, M.M.; White, J.R.; Livi, G.P. Coexpression of Human CAMP-Specific Phosphodiesterase Activity and High Affinity Rolipram Binding in Yeast. J. Biol. Chem. 1992, 267, 1798–1804. [Google Scholar] [CrossRef] [PubMed]
  123. Compton, C.H.; Gubb, J.; Nieman, R.; Edelson, J.; Amit, O.; Bakst, A.; Ayres, J.G.; Creemers, J.P.H.M.; Schultze-Werninghaus, G.; Brambilla, C.; et al. Cilomilast, a Selective Phosphodiesterase-4 Inhibitor for Treatment of Patients with Chronic Obstructive Pulmonary Disease: A Randomised, Dose-Ranging Study. Lancet 2001, 358, 265–270. [Google Scholar] [CrossRef] [PubMed]
  124. Cave, A.; Arlett, P.; Lee, E. Inhaled and Nasal Corticosteroids: Factors Affecting the Risks of Systemic Adverse Effects. Pharmacol. Ther. 1999, 83, 153–179. [Google Scholar] [CrossRef]
  125. Marx, D.; Tassabehji, M.; Heer, S.; Hüttenbrink, K.B.; Szelenyi, I. Modulation of TNF and GM-CSF Release from Dispersed Human Nasal Polyp Cells and Human Whole Blood by Inhibitors of Different PDE Isoenzymes and Glucocorticoids. Pulm. Pharmacol. Ther. 2002, 15, 7–15. [Google Scholar] [CrossRef]
  126. Kuss, H.; Hoefgen, N.; Johanssen, S.; Kronbach, T.; Rundfeldt, C. In Vivo Efficacy in Airway Disease Models of N-(3,5-Dichloropyrid-4-Yl)-[1-(4-Fluorobenzyl)-5-Hydroxy-Indole-3-Yl]-Glyoxylic Acid Amide (AWD 12-281), a Selective Phosphodiesterase 4 Inhibitor for Inhaled Administration. J. Pharmacol. Exp. Ther. 2003, 307, 373–385. [Google Scholar] [CrossRef] [Green Version]
  127. Burnouf, C.; Pruniaux, M.-P. Recent Advances in PDE4 Inhibitors as Immunoregulators and Anti- Inflammatory Drugs. Curr. Pharm. Des. 2005, 8, 1255–1296. [Google Scholar] [CrossRef]
  128. Barnette, M.S. Phosphodiesterase 4 (PDE4) Inhibitors in Asthma and Chronic Obstructive Pulmonary Disease (COPD). Prog. Drug Res. 1999, 53, 193–229. [Google Scholar] [CrossRef]
  129. Phillips, J.E. Inhaled Phosphodiesterase 4 (PDE4) Inhibitors for Inflammatory Respiratory Diseases. Front. Pharmacol. 2020, 11, 259. [Google Scholar] [CrossRef] [Green Version]
  130. Dennis, E.A.; Cao, J.; Hsu, Y.H.; Magrioti, V.; Kokotos, G. Phospholipase A2 Enzymes: Physical Structure, Biological Function, Disease Implication, Chemical Inhibition, and Therapeutic Intervention. Chem. Rev. 2011, 111, 6130–6185. [Google Scholar] [CrossRef] [Green Version]
  131. Dillard, R.D.; Bach, N.J.; Draheim, S.E.; Berry, D.R.; Carlson, D.G.; Chirgadze, N.Y.; Clawson, D.K.; Hartley, L.W.; Johnson, L.M.; Jones, N.D.; et al. Indole Inhibitors of Human Nonpancreatic Secretory Phospholipase A2. 1. Indole-3-Acetamides. J. Med. Chem. 1996, 39, 5119–5136. [Google Scholar] [CrossRef]
  132. Dillard, R.D.; Bach, N.J.; Draheim, S.E.; Berry, D.R.; Carlson, D.G.; Chirgadze, N.Y.; Clawson, D.K.; Hartley, L.W.; Johnson, L.M.; Jones, N.D.; et al. Indole Inhibitors of Human Nonpancreatic Secretory Phospholipase A2. 2. Indole-3-Acetamides with Additional Functionality. J. Med. Chem. 1996, 39, 5137–5158. [Google Scholar] [CrossRef]
  133. Draheim, S.E.; Bach, N.J.; Dillard, R.D.; Berry, D.R.; Carlson, D.G.; Chirgadze, N.Y.; Clawson, D.K.; Hartley, L.W.; Johnson, L.M.; Jones, N.D.; et al. Indole Inhibitors of Human Nonpancreatic Secretory Phospholipase A2. 3. Indole-3-Glyoxamides. J. Med. Chem. 1996, 39, 5159–5175. [Google Scholar] [CrossRef]
  134. Mihelich, E.D.; Schevitz, R.W. Structure-Based Design of a New Class of Anti-Inflammatory Drugs: Secretory Phospholipase A2 Inhibitors, SPI. Biochim. Biophys. Acta Mol. Cell Biol. Lipids 1999, 1441, 223–228. [Google Scholar] [CrossRef]
  135. Abraham, E.; Naum, C.; Bandi, V.; Gervich, D.; Lowry, S.F.; Wunderink, R.; Schein, R.M.; Macias, W.; Skerjanec, S.; Dmitrienko, A.; et al. Efficacy and Safety of LY315920Na/S-5920, a Selective Inhibitor of 14-KDa Group IIA Secretory Phospholipase A2, in Patients with Suspected Sepsis and Organ Failure. Crit. Care Med. 2003, 31, 718–728. [Google Scholar] [CrossRef]
  136. Snyder, D.W.; Bach, N.J.; Dillard, R.D.; Draheim, S.E.; Carlson, D.G.; Fox, N.; Roehm, N.W.; Armstrong, C.T.; Chang, C.H.; Hartley, L.W.; et al. Pharmacology of LY315920/S-5920, [[3-(Aminooxoacetyl)-2-Ethyl-1-(Phenylmethyl)-1H-Indol-4-Yl]Oxy]Acetate, a Potent and Selective Secretory Phospholipase A2 Inhibitor: A New Class of Anti-Inflammatory Drugs, SPI. J. Pharmacol. Exp. Ther. 1999, 288, 1117–1124. [Google Scholar]
  137. Rosenson, R.S.; Fraser, H.; Trias, J.; Hislop, C. Varespladib Methyl in Cardiovascular Disease. Expert Opin. Investig. Drugs 2010, 19, 1245–1255. [Google Scholar] [CrossRef]
  138. Bradley, J.D.; Dmitrienko, A.A.; Kivitz, A.J.; Gluck, O.S.; Weaver, A.L.; Wiesenhutter, C.; Myers, S.L.; Sides, G.D. A Randomized, Double-Blinded, Placebo-Controlled Clinical Trial of LY333013, a Selective Inhibitor of Group II Secretory Phospholipase A 2, in the Treatment of Rheumatoid Arthritis. J. Rheumatol. 2005, 32, 417–423. [Google Scholar]
  139. Bowton, D.L.; Dmitrienko, A.A.; Israel, E.; Zeiher, B.G.; Sides, G.D. Impact of a Soluble Phospholipase A2 Inhibitor on Inhaled Allergen Challenge in Subjects with Asthma. J. Asthma 2005, 42, 65–71. [Google Scholar] [CrossRef]
  140. Rosenson, R.S.; Hislop, C.; McConnell, D.; Elliott, M.; Stasiv, Y.; Wang, N.; Waters, D.D. Effects of 1-H-Indole-3-Glyoxamide (A-002) on Concentration of Secretory Phospholipase A2 (PLASMA Study): A Phase II Double-Blind, Randomised, Placebo-Controlled Trial. Lancet 2009, 373, 649–658. [Google Scholar] [CrossRef]
  141. Nicholls, S.J.; Kastelein, J.J.P.; Schwartz, G.G.; Bash, D.; Rosenson, R.S.; Cavender, M.A.; Brennan, D.M.; Koenig, W.; Jukema, J.W.; Nambi, V.; et al. Varespladib and Cardiovascular Events in Patients With an Acute Coronary Syndrome: The VISTA-16 Randomized Clinical Trial. JAMA 2014, 311, 252–262. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Schett, G.; Zwerina, J.; Firestein, G. The P38 Mitogen-Activated Protein Kinase (MAPK) Pathway in Rheumatoid Arthritis. Ann. Rheum. Dis. 2008, 67, 909–916. [Google Scholar] [CrossRef] [PubMed]
  143. Westra, J.; C Limburg, P. P38 Mitogen-Activated Protein Kinase (MAPK) in Rheumatoid Arthritis. Mini-Reviews Med. Chem. 2006, 6, 867–874. [Google Scholar] [CrossRef] [PubMed]
  144. Schett, G.; Stach, C.; Zwerina, J.; Voll, R.; Manger, B. How Antirheumatic Drugs Protect Joints from Damage in Rheumatoid Arthritis. Arthritis Rheum. 2008, 58, 2936–2948. [Google Scholar] [CrossRef]
  145. Korb, A.; Tohidast-Akrad, M.; Cetin, E.; Axmann, R.; Smolen, J.; Schett, G. Differential Tissue Expression and Activation of P38 MAPK α, β, γ, and δ Isoforms in Rheumatoid Arthritis. Arthritis Rheum. 2006, 54, 2745–2756. [Google Scholar] [CrossRef]
  146. Schett, G.; Tohidast-Akrad, M.; Smolen, J.S.; Schmid, B.J.; Steiner, C.W.; Bitzan, P.; Zenz, P.; Redlich, K.; Xu, Q.; Steiner, G. Activation, Differential Localization, and Regulation of the Stress-Activated Protein Kinases, Extracellular Signal-Regulated Kinase, c-Jun N-Terminal Kinase, and P38 Mitogen-Activated Protein Kinase, in Synovial Tissue and Cells in Rheumatoid Arthritis. Arthritis Rheum. 2000, 43, 2501–2512. [Google Scholar] [CrossRef]
  147. Pargellis, C.; Tong, L.; Churchill, L.; Cirillo, P.F.; Gilmore, T.; Graham, A.G.; Grob, P.M.; Hickey, E.R.; Moss, N.; Pav, S.; et al. Inhibition of P38 MAP Kinase by Utilizing a Novel Allosteric Binding Site. Nat. Struct. Biol. 2002, 9, 268–272. [Google Scholar] [CrossRef]
  148. Mavunkel, B.J.; Chakravarty, S.; Perumattam, J.J.; Dugar, S.; Lu, Q.; Liang, X. Indole-Type Derivatives as Inhibitors of P38 Kinase; Scios LLC.: Titusville, NJ, USA, 2000. [Google Scholar]
  149. Snoonian, J.R.; Oliver-Shaffer, P.-A. Processes for the Preparation Of N-Heteroaryl-N-Aryl-Amines by Reacting an N-Aryl Carbamic Acid Ester with a Halo-Heteroaryl And Analogous Processes; Vertex Pharmaceuticals Incorporated: Boston, MA, USA, 2004. [Google Scholar]
  150. Hill, R.J.; Dabbagh, K.; Phippard, D.; Li, C.; Suttmann, R.T.; Welch, M.; Papp, E.; Song, K.W.; Chang, K.C.; Leaffer, D.; et al. Pamapimod, a Novel P38 Mitogen-Activated Protein Kinase Inhibitor: Preclinical Analysis of Efficacy and Selectivity. J. Pharmacol. Exp. Ther. 2008, 327, 610–619. [Google Scholar] [CrossRef] [Green Version]
  151. Kuma, Y.; Sabio, G.; Bain, J.; Shpiro, N.; Márquez, R.; Cuenda, A. BIRB796 Inhibits All P38 MAPK Isoforms In Vitro and In Vivo. J. Biol. Chem. 2005, 280, 19472–19479. [Google Scholar] [CrossRef] [Green Version]
  152. Goldstein, D.M.; Kuglstatter, A.; Lou, Y.; Soth, M.J. Selective P38α Inhibitors Clinically Evaluated for the Treatment of Chronic Inflammatory Disorders. J. Med. Chem. 2010, 53, 2345–2353. [Google Scholar] [CrossRef]
  153. Nikas, S.N.; Drosos, A.A. SCIO-469 Scios Inc. Curr. Opin. Investig. Drugs 2004, 5, 1205–1212. [Google Scholar]
  154. Ji, R.R. Peripheral and Central Mechanisms of Inflammatory Pain, with Emphasis on MAP Kinases. Curr. Drug Targets. Inflamm. Allergy 2004, 3, 299–303. [Google Scholar] [CrossRef]
  155. Schindler, J.F.; Monahan, J.B.; Smith, W.G. P38 Pathway Kinases as Anti-Inflammatory Drug Targets. J. Dent. Res. 2007, 86, 800–811. [Google Scholar] [CrossRef]
  156. Genovese, M.C.; Cohen, S.B.; Wofsy, D.; Weinblatt, M.E.; Firestein, G.S.; Brahn, E.; Strand, V.; Baker, D.G.; Tong, S.E. A 24-Week, Randomized, Double-Blind, Placebo-Controlled, Parallel Group Study of the Efficacy of Oral SCIO-469, a P38 Mitogen-Activated Protein Kinase Inhibitor, in Patients with Active Rheumatoid Arthritis. J. Rheumatol. 2011, 38, 846–854. [Google Scholar] [CrossRef]
  157. Rescher, S.L.U.; Brunotte, L.; Faist, A.; Schloer, S.M. P38-Inhibitors for the Treatment of Coronavirus Infections and/or COVID-19 Cytokine Storm. International Application No. PCT/EP2022/068724, 12 January 2023. [Google Scholar]
  158. Leitmeyer, K.C.; Vaughn, D.W.; Watts, D.M.; Salas, R.; Villalobos, I.; de Chacon; Ramos, C.; Rico-Hesse, R. Dengue Virus Structural Differences That Correlate with Pathogenesis. J. Virol. 1999, 73, 4738–4747. [Google Scholar] [CrossRef] [Green Version]
  159. Wu, H.; Bock, S.; Snitko, M.; Berger, T.; Weidner, T.; Holloway, S.; Kanitz, M.; Diederich, W.E.; Steuber, H.; Walter, C.; et al. Novel Dengue Virus NS2B/NS3 Protease Inhibitors. Antimicrob. Agents Chemother. 2015, 59, 1100–1109. [Google Scholar] [CrossRef] [Green Version]
  160. Li, Z.; Zhang, J.; Li, H. Flavivirus NS2B/NS3 Protease: Structure, Function, and Inhibition. In Viral Proteases and Their Inhibitors; Academic Press: Cambridge, MA, USA, 2017; pp. 163–188. [Google Scholar] [CrossRef]
  161. Steuer, C.; Gege, C.; Fischl, W.; Heinonen, K.H.; Bartenschlager, R.; Klein, C.D. Synthesis and Biological Evaluation of α-Ketoamides as Inhibitors of the Dengue Virus Protease with Antiviral Activity in Cell-Culture. Bioorg. Med. Chem. 2011, 19, 4067–4074. [Google Scholar] [CrossRef]
  162. Nie, S.; Yao, Y.; Wu, F.; Wu, X.; Zhao, J.; Hua, Y.; Wu, J.; Huo, T.; Lin, Y.L.; Kneubehl, A.R.; et al. Synthesis, Structure-Activity Relationships, and Antiviral Activity of Allosteric Inhibitors of Flavivirus NS2B-NS3 Protease. J. Med. Chem. 2021, 64, 2777–2800. [Google Scholar] [CrossRef]
  163. Mehand, M.S.; Al-Shorbaji, F.; Millett, P.; Murgue, B. The WHO R&D Blueprint: 2018 Review of Emerging Infectious Diseases Requiring Urgent Research and Development Efforts. Antiviral Res. 2018, 159, 63–67. [Google Scholar] [CrossRef]
  164. del Rosario García-Lozano, M.; Dragoni, F.; Gallego, P.; Mazzotta, S.; López-Gómez, A.; Boccuto, A.; Martínez-Cortés, C.; Rodríguez-Martínez, A.; Pérez-Sánchez, H.; Manuel Vega-Pérez, J.; et al. Piperazine-Derived Small Molecules as Potential Flaviviridae NS3 Protease Inhibitors. In Vitro Antiviral Activity Evaluation against Zika and Dengue Viruses. Bioorg. Chem. 2023, 133, 106408. [Google Scholar] [CrossRef]
  165. Kling, A.; Jantos, K.; Mack, H.; Hornberger, W.; Drescher, K.; Nimmrich, V.; Relo, A.; Wicke, K.; Hutchins, C.W.; Lao, Y.; et al. Discovery of Novel and Highly Selective Inhibitors of Calpain for the Treatment of Alzheimer’s Disease: 2-(3-Phenyl-1H-Pyrazol-1-Yl)-Nicotinamides. J. Med. Chem. 2017, 60, 7123–7138. [Google Scholar] [CrossRef] [PubMed]
  166. Dorababu, A. Indole—A Promising Pharmacophore in Recent Antiviral Drug Discovery. RSC Med. Chem. 2020, 11, 1335–1353. [Google Scholar] [CrossRef] [PubMed]
  167. Meanwell, N.A.; Krystal, M.R.; Nowicka-Sans, B.; Langley, D.R.; Conlon, D.A.; Eastgate, M.D.; Grasela, D.M.; Timmins, P.; Wang, T.; Kadow, J.F. Inhibitors of HIV-1 Attachment: The Discovery and Development of Temsavir and Its Prodrug Fostemsavir. J. Med. Chem. 2018, 61, 62–80. [Google Scholar] [CrossRef] [PubMed]
  168. Ruwizhi, N.; Aderibigbe, B.A. Cinnamic Acid Derivatives and Their Biological Efficacy. Int. J. Mol. Sci. 2020, 21, 5712. [Google Scholar] [CrossRef]
  169. Abd El-Raouf, O.M.; El-Sayed, E.S.M.; Manie, M.F. Cinnamic Acid and Cinnamaldehyde Ameliorate Cisplatin-Induced Splenotoxicity in Rats. J. Biochem. Mol. Toxicol. 2015, 29, 426–431. [Google Scholar] [CrossRef]
  170. Wang, R.; Yang, W.; Fan, Y.; Dehaen, W.; Li, Y.; Li, H.; Wang, W.; Zheng, Q.; Huai, Q. Design and Synthesis of the Novel Oleanolic Acid-Cinnamic Acid Ester Derivatives and Glycyrrhetinic Acid-Cinnamic Acid Ester Derivatives with Cytotoxic Properties. Bioorg. Chem. 2019, 88, 102951. [Google Scholar] [CrossRef]
  171. Amano, R.; Yamashita, A.; Kasai, H.; Hori, T.; Miyasato, S.; Saito, S.; Yokoe, H.; Takahashi, K.; Tanaka, T.; Otoguro, T.; et al. Cinnamic Acid Derivatives Inhibit Hepatitis C Virus Replication via the Induction of Oxidative Stress. Antivir. Res. 2017, 145, 123–130. [Google Scholar] [CrossRef]
  172. Chen, Y.; Li, Z.; Pan, P.; Lao, Z.; Xu, J.; Li, Z.; Zhan, S.; Liu, X.; Wu, Y.; Wang, W.; et al. Cinnamic Acid Inhibits Zika Virus by Inhibiting RdRp Activity. Antiviral Res. 2021, 192, 105117. [Google Scholar] [CrossRef]
  173. Nitsche, C.; Behnam, M.A.M.; Steuer, C.; Klein, C.D. Retro Peptide-Hybrids as Selective Inhibitors of the Dengue Virus NS2B-NS3 Protease. Antivir. Res. 2012, 94, 72–79. [Google Scholar] [CrossRef]
  174. Nitsche, C.; Steuer, C.; Klein, C.D. Arylcyanoacrylamides as Inhibitors of the Dengue and West Nile Virus Proteases. Bioorg. Med. Chem. 2011, 19, 7318–7337. [Google Scholar] [CrossRef]
  175. Deng, J.; Li, N.; Liu, H.; Zuo, Z.; Liew, O.W.; Xu, W.; Chen, G.; Tong, X.; Tang, W.; Zhu, J.; et al. Discovery of Novel Small Molecule Inhibitors of Dengue Viral NS2B-NS3 Protease Using Virtual Screening and Scaffold Hopping. J. Med. Chem. 2012, 55, 6278–6293. [Google Scholar] [CrossRef]
  176. Guo, S.; Zhen, Y.; Zhu, Z.; Zhou, G.; Zheng, X. Cinnamic Acid Rescues Behavioral Deficits in a Mouse Model of Traumatic Brain Injury by Targeting MiR-455-3p/HDAC2. Life Sci. 2019, 235, 116819. [Google Scholar] [CrossRef]
  177. Faucher, A.M.; Bailey, M.D.; Beaulieu, P.L.; Brochu, C.; Duceppe, J.S.; Ferland, J.M.; Ghiro, E.; Gorys, V.; Halmos, T.; Kawai, S.H.; et al. Synthesis of BILN 2061, an HCV NS3 Protease Inhibitor with Proven Antiviral Effect in Humans. Org. Lett. 2004, 6, 2901–2904. [Google Scholar] [CrossRef]
  178. Li, F.; Lee, E.M.; Sun, X.; Wang, D.; Tang, H.; Zhou, G.C. Design, Synthesis and Discovery of Andrographolide Derivatives against Zika Virus Infection. Eur. J. Med. Chem. 2020, 187, 111925. [Google Scholar] [CrossRef]
  179. Barbosa-Lima, G.; Moraes, A.M.; da Araújo, A.S.; da Silva, E.T.; de Freitas, C.S.; Vieira, Y.R.; Marttorelli, A.; Neto, J.C.; Bozza, P.T.; de Souza, M.V.N.; et al. 2,8-Bis(Trifluoromethyl)Quinoline Analogs Show Improved Anti-Zika Virus Activity, Compared to Mefloquine. Eur. J. Med. Chem. 2017, 127, 334–340. [Google Scholar] [CrossRef]
  180. Balasubramanian, A.; Teramoto, T.; Kulkarni, A.A.; Bhattacharjee, A.K.; Padmanabhan, R. Antiviral Activities of Selected Antimalarials against Dengue Virus Type 2 and Zika Virus. Antiviral Res. 2017, 137, 141–150. [Google Scholar] [CrossRef]
  181. de la Guardia, C.; Stephens, D.E.; Dang, H.T.; Quijada, M.; Larionov, O.V.; Lleonart, R. Antiviral Activity of Novel Quinoline Derivatives against Dengue Virus Serotype 2. Molecules 2018, 23, 672. [Google Scholar] [CrossRef] [Green Version]
  182. Kaptein, S.J.F.; Vincetti, P.; Crespan, E.; Rivera, J.I.A.; Costantino, G.; Maga, G.; Neyts, J.; Radi, M. Identification of Broad-Spectrum Dengue/Zika Virus Replication Inhibitors by Functionalization of Quinoline and 2,6-Diaminopurine Scaffolds. ChemMedChem 2018, 13, 1371–1376. [Google Scholar] [CrossRef]
  183. Beesetti, H.; Tyagi, P.; Medapi, B.; Krishna, V.S.; Sriram, D.; Khanna, N.; Swaminathan, S. A Quinoline Compound Inhibits the Replication of Dengue Virus Serotypes 1-4 in Vero Cells. Antivir. Ther. 2018, 23, 385–394. [Google Scholar] [CrossRef]
  184. Zhao, J.; Zhang, Y.; Wang, M.; Liu, Q.; Lei, X.; Wu, M.; Guo, S.; Yi, D.; Li, Q.; Ma, L.; et al. Quinoline and Quinazoline Derivatives Inhibit Viral RNA Synthesis by SARS-CoV-2 RdRp. ACS Infect. Dis. 2021, 7, 1535–1544. [Google Scholar] [CrossRef]
  185. Vella, S.; Palmisano, L. The Global Status of Resistance to Antiretroviral Drugs. Clin. Infect. Dis. 2005, 41, S239–S246. [Google Scholar] [CrossRef] [PubMed]
  186. Qian, K.; Morris-Natschke, S.L.; Lee, K.H. HIV Entry Inhibitors and Their Potential in HIV Therapy. Med. Res. Rev. 2009, 29, 369–393. [Google Scholar] [CrossRef] [PubMed]
  187. Wang, T.; Zhang, Z.; Wallace, O.B.; Deshpande, M.; Fang, H.; Yang, Z.; Zadjura, L.M.; Tweedie, D.L.; Huang, S.; Zhao, F.; et al. Discovery of 4-Benzoyl-1-[(4-Methoxy-1H-Pyrrolo [2,3-b]Pyridin-3-Yl)Oxoacetyl]-2-(R) -Methylpiperazine (BMS-378806): A Novel HIV-1 Attachment Inhibitor That Interferes with CD4-Gp120 Interactions. J. Med. Chem. 2003, 46, 4236–4239. [Google Scholar] [CrossRef] [PubMed]
  188. Meanwell, N.A.; Wallace, O.B.; Fang, H.; Wang, H.; Deshpande, M.; Wang, T.; Yin, Z.; Zhang, Z.; Pearce, B.C.; James, J.; et al. Inhibitors of HIV-1 Attachment. Part 2: An Initial Survey of Indole Substitution Patterns. Bioorg. Med. Chem. Lett. 2009, 19, 1977–1981. [Google Scholar] [CrossRef]
  189. Meanwell, N.A.; Wallace, O.B.; Wang, H.; Deshpande, M.; Pearce, B.C.; Trehan, A.; Yeung, K.S.; Qiu, Z.; Wright, J.J.K.; Robinson, B.A.; et al. Inhibitors of HIV-1 Attachment. Part 3: A Preliminary Survey of the Effect of Structural Variation of the Benzamide Moiety on Antiviral Activity. Bioorg. Med. Chem. Lett. 2009, 19, 5136–5139. [Google Scholar] [CrossRef]
  190. Wang, T.; Kadow, J.F.; Zhang, Z.; Yin, Z.; Gao, Q.; Wu, D.; Parker, D.D.G.; Yang, Z.; Zadjura, L.; Robinson, B.A.; et al. Inhibitors of HIV-1 Attachment. Part 4: A Study of the Effect of Piperazine Substitution Patterns on Antiviral Potency in the Context of Indole-Based Derivatives. Bioorg. Med. Chem. Lett. 2009, 19, 5140–5145. [Google Scholar] [CrossRef]
  191. Wang, T.; Yin, Z.; Zhang, Z.; Bender, J.A.; Yang, Z.; Johnson, G.; Yang, Z.; Zadjura, L.M.; D’Arienzo, C.J.; Parker, D.D.G.; et al. Inhibitors of Human Immunodeficiency Virus Type 1 (HIV-1) Attachment. 5. An Evolution from Indole to Azaindoles Leading to the Discovery of 1-(4-Benzoylpiperazin-1-Yl)-2-(4,7-Dimethoxy-1H-Pyrrolo[2,3-c]-Pyridin-3-Yl) Ethane-1,2-Dione (BMS-488043), a Drug. J. Med. Chem. 2009, 52, 7778–7787. [Google Scholar] [CrossRef]
  192. Lin, P.F.; Blair, W.; Wang, T.; Spicer, T.; Guo, Q.; Zhou, N.; Gong, Y.F.; Wang, H.G.H.; Rose, R.; Yamanaka, G.; et al. A Small Molecule HIV-1 Inhibitor That Targets the HIV-1 Envelope and Inhibits CD4 Receptor Binding. Proc. Natl. Acad. Sci. USA 2003, 100, 11013–11018. [Google Scholar] [CrossRef]
  193. Ho, H.-T.; Fan, L.; Nowicka-Sans, B.; McAuliffe, B.; Li, C.-B.; Yamanaka, G.; Zhou, N.; Fang, H.; Dicker, I.; Dalterio, R.; et al. Envelope Conformational Changes Induced by Human Immunodeficiency Virus Type 1 Attachment Inhibitors Prevent CD4 Binding and Downstream Entry Events. J. Virol. 2006, 80, 4017. [Google Scholar] [CrossRef] [Green Version]
  194. Yang, Z.; Zadjura, L.; D’Arienzo, C.; Marino, A.; Santone, K.; Klunk, L.; Greene, D.; Lin, P.F.; Colonno, R.; Wang, T.; et al. Preclinical Pharmacokinetics of a Novel HIV-1 Attachment Inhibitor BMS-378806 and Prediction of Its Human Pharmacokinetics. Biopharm. Drug Dispos. 2005, 26, 387–402. [Google Scholar] [CrossRef]
  195. Hanna, G.J.; Lalezari, J.; Hellinger, J.A.; Wohl, D.A.; Nettles, R.; Persson, A.; Krystal, M.; Lin, P.; Colonno, R.; Grasela, D.M. Antiviral Activity, Pharmacokinetics, and Safety of BMS-488043, a Novel Oral Small-Molecule HIV-1 Attachment Inhibitor, in HIV-1-Infected Subjects. Antimicrob. Agents Chemother. 2011, 55, 722–728. [Google Scholar] [CrossRef] [Green Version]
  196. Pancera, M.; Lai, Y.T.; Bylund, T.; Druz, A.; Narpala, S.; O’Dell, S.; Schön, A.; Bailer, R.T.; Chuang, G.Y.; Geng, H.; et al. Crystal Structures of Trimeric HIV Env with Entry Inhibitors BMS-378806 and BMS-626529. Nat. Chem. Biol. 2017, 13, 1115. [Google Scholar] [CrossRef] [Green Version]
  197. Kozal, M.; Aberg, J.; Pialoux, G.; Cahn, P.; Thompson, M.; Molina, J.-M.; Grinsztejn, B.; Diaz, R.; Castagna, A.; Kumar, P.; et al. Fostemsavir in Adults with Multidrug-Resistant HIV-1 Infection. N. Engl. J. Med. 2020, 382, 1232–1243. [Google Scholar] [CrossRef]
  198. Akopyants, N.S.; Kimblin, N.; Secundino, N.; Patrick, R.; Peters, N.; Lawyer, P.; Dobson, D.E.; Beverley, S.M.; Sacks, D.L. Demonstration of Genetic Exchange during Cyclical Development of Leishmania in the Sand Fly Vector. Science 2009, 324, 265–268. [Google Scholar] [CrossRef] [Green Version]
  199. Rogers, M.; Kropf, P.; Choi, B.S.; Dillon, R.; Podinovskaia, M.; Bates, P.; Müller, I. Proteophosophoglycans Regurgitated by Leishmania-Infected Sand Flies Target the L-Arginine Metabolism of Host Macrophages to Promote Parasite Survival. PLoS Pathog. 2009, 5, e1000555. [Google Scholar] [CrossRef] [Green Version]
  200. Gupta, L.; Talwar, A.; Nishi; Palne, S.; Gupta, S.; Chauhan, P.M.S. Synthesis of Marine Alkaloid: 8,9-Dihydrocoscinamide B and Its Analogues as Novel Class of Antileishmanial Agents. Bioorg. Med. Chem. Lett. 2007, 17, 4075–4079. [Google Scholar] [CrossRef]
  201. Kumar, A.; Katiyar, S.B.; Gupta, S.; Chauhan, P.M.S. Syntheses of New Substituted Triazino Tetrahydroisoquinolines and β-Carbolines as Novel Antileishmanial Agents. Eur. J. Med. Chem. 2006, 41, 106–113. [Google Scholar] [CrossRef]
  202. Costa, E.V.; Pinheiro, M.L.B.; Xavier, C.M.; Silva, J.R.A.; Amaral, A.C.F.; Souza, A.D.L.; Barison, A.; Campos, F.R.; Ferreira, A.G.; Machado, G.M.C.; et al. A Pyrimidine-β-Carboline and Other Alkaloids from Annona Foetida with Antileishmanial Activity. J. Nat. Prod. 2006, 69, 292–294. [Google Scholar] [CrossRef]
  203. Kumar, R.; Khan, S.; Verma, A.; Srivastava, S.; Viswakarma, P.; Gupta, S.; Meena, S.; Singh, N.; Sarkar, J.; Chauhan, P.M.S. Synthesis of 2-(Pyrimidin-2-Yl)-1-Phenyl-2,3,4,9-Tetrahydro-1H-β-Carbolines as Antileishmanial Agents. Eur. J. Med. Chem. 2010, 45, 3274–3280. [Google Scholar] [CrossRef]
  204. Chauhan, S.S.; Gupta, L.; Mittal, M.; Vishwakarma, P.; Gupta, S.; Chauhan, P.M.S. Synthesis and Biological Evaluation of Indolyl Glyoxylamides as a New Class of Antileishmanial Agents. Bioorganic Med. Chem. Lett. 2010, 20, 6191–6194. [Google Scholar] [CrossRef]
  205. Singh, A.; Mohan, M.L.; Isaac, A.O.; Luo, X.; Petrak, J.; Vyoral, D.; Singh, N. Prion Protein Modulates Cellular Iron Uptake: A Novel Function with Implications for Prion Disease Pathogenesis. PLoS ONE 2009, 4, e4468. [Google Scholar] [CrossRef]
  206. Hu, W.; Kieseier, B.; Frohman, E.; Eagar, T.N.; Rosenberg, R.N.; Hartung, H.P.; Stüve, O. Prion Proteins: Physiological Functions and Role in Neurological Disorders. J. Neurol. Sci. 2008, 264, 1–8. [Google Scholar] [CrossRef] [PubMed]
  207. Thompson, M.J.; Borsenberger, V.; Louth, J.C.; Judd, K.E.; Chen, B. Design, Synthesis, and Structure—Activity Relationship of Indole-3-Glyoxylamide Libraries Possessing Highly Potent Activity in a Cell Line Model of Prion Disease. J. Med. Chem. 2009, 52, 7503–7511. [Google Scholar] [CrossRef] [PubMed]
  208. Thompson, M.J.; Louth, J.C.; Ferrara, S.; Jackson, M.P.; Sorrell, F.J.; Cochrane, E.J.; Gever, J.; Baxendale, S.; Silber, B.M.; Roehl, H.H.; et al. Discovery of 6-Substituted Indole-3-Glyoxylamides as Lead Antiprion Agents with Enhanced Cell Line Activity, Improved Microsomal Stability and Low Toxicity. Eur. J. Med. Chem. 2011, 46, 4125–4132. [Google Scholar] [CrossRef]
  209. Thompson, M.J.; Louth, J.C.; Ferrara, S.; Sorrell, F.J.; Irving, B.J.; Cochrane, E.J.; Meijer, A.J.H.M.; Chen, B. Structure–Activity Relationship Refinement and Further Assessment of Indole-3-Glyoxylamides as a Lead Series against Prion Disease. ChemMedChem 2011, 6, 115–130. [Google Scholar] [CrossRef]
  210. Taliani, S.; Trincavelli, M.L.; Cosimelli, B.; Laneri, S.; Severi, E.; Barresi, E.; Pugliesi, I.; Daniele, S.; Giacomelli, C.; Greco, G.; et al. Modulation of A2B Adenosine Receptor by 1-Benzyl-3-Ketoindole Derivatives. Eur. J. Med. Chem. 2013, 69, 331–337. [Google Scholar] [CrossRef]
  211. Trincavelli, M.L.; Giacomelli, C.; Daniele, S.; Taliani, S.; Cosimelli, B.; Laneri, S.; Severi, E.; Barresi, E.; Pugliesi, I.; Greco, G.; et al. Allosteric Modulators of Human A2B Adenosine Receptor. Biochim. Biophys. Acta Gen. Subj. 2014, 1840, 1194–1203. [Google Scholar] [CrossRef]
  212. Trincavelli, M.L.; Daniele, S.; Giacomelli, C.; Taliani, S.; Da Settimo, F.; Cosimelli, B.; Greco, G.; Novellino, E.; Martini, C. Osteoblast Differentiation and Survival: A Role for A2B Adenosine Receptor Allosteric Modulators. Biochim. Biophys. Acta Mol. Cell Res. 2014, 1843, 2957–2966. [Google Scholar] [CrossRef]
  213. Barresi, E.; Giacomelli, C.; Marchetti, L.; Baglini, E.; Salerno, S.; Greco, G.; Da Settimo, F.; Martini, C.; Trincavelli, M.L.; Taliani, S. Novel Positive Allosteric Modulators of A2B Adenosine Receptor Acting as Bone Mineralisation Promoters. J. Enzyme Inhib. Med. Chem. 2021, 36, 286–294. [Google Scholar] [CrossRef]
  214. Supuran, C.T. Carbonic Anhydrases: Novel Therapeutic Applications for Inhibitors and Activators. Nat. Rev. Drug Discov. 2008, 7, 168–181. [Google Scholar] [CrossRef]
  215. Supuran, C.T. Carbonic Anhydrase Activators. Future Med. Chem. 2018, 10, 561–573. [Google Scholar] [CrossRef]
  216. Barresi, E.; Ravichandran, R.; Germelli, L.; Angeli, A.; Baglini, E.; Salerno, S.; Marini, A.M.; Costa, B.; Da Pozzo, E.; Martini, C.; et al. Carbonic Anhydrase Activation Profile of Indole-Based Derivatives. J. Enzyme Inhib. Med. Chem. 2021, 36, 1783–1797. [Google Scholar] [CrossRef]
  217. Santoro, A.; Mattace Raso, G.; Taliani, S.; Da Pozzo, E.; Simorini, F.; Costa, B.; Martini, C.; Laneri, S.; Sacchi, A.; Cosimelli, B.; et al. TSPO-Ligands Prevent Oxidative Damage and Inflammatory Response in C6 Glioma Cells by Neurosteroid Synthesis. Eur. J. Pharm. Sci. 2016, 88, 124–131. [Google Scholar] [CrossRef]
  218. Ruusuvuori, E.; Huebner, A.K.; Kirilkin, I.; Yukin, A.Y.; Blaesse, P.; Helmy, M.; Jung Kang, H.; El Muayed, M.; Christopher Hennings, J.; Voipio, J.; et al. Neuronal Carbonic Anhydrase VII Provides GABAergic Excitatory Drive to Exacerbate Febrile Seizures. EMBO J. 2013, 32, 2275–2286. [Google Scholar] [CrossRef] [Green Version]
  219. Matthews, T.A.; Abel, A.; Demme, C.; Sherman, T.; Pan, P.W.; Halterman, M.W.; Parkkila, S.; Nehrke, K. Expression of the CHOP-Inducible Carbonic Anhydrase CAVI-b Is Required for BDNF-Mediated Protection from Hypoxia. Brain Res. 2014, 1543, 28–37. [Google Scholar] [CrossRef] [Green Version]
  220. Primofiore, G.; Marini, A.M.; Settimo, F.D.; Martini, C.; Bardellini, A.; Giannaccini, G.; Lucacchini, A. Specific Inhibition of Benzodiazepine Receptor Binding by Some N-(Indo1-3-Ylglyoxylyl)Amino Acid Derivatives: Stereoselective Interactions. J. Med. Chem. 1989, 32, 2514–2518. [Google Scholar] [CrossRef]
Figure 1. Structure of the basic indolylglyoxylamide scaffold.
Figure 1. Structure of the basic indolylglyoxylamide scaffold.
Pharmaceuticals 16 00997 g001
Figure 2. Graph representing the application of indolylglyoxylamides with different biological activity.
Figure 2. Graph representing the application of indolylglyoxylamides with different biological activity.
Pharmaceuticals 16 00997 g002
Figure 3. Structures of zaleplon, zoplicone, and zolpidem.
Figure 3. Structures of zaleplon, zoplicone, and zolpidem.
Pharmaceuticals 16 00997 g003
Figure 4. Hypothetical binding modes A and B of indolylglyoxylamide BzR ligands [27] within the framework of Cook’s pharmacophore/topological model [25]. A2 = hydrogen-bond acceptor; H1 and H2 = hydrogen-bond donors; L1, L2, L3, and LDi = lipophilic clefts; S1 and S2 = sterically forbidden sites; X = OH, OCH3, Cl.
Figure 4. Hypothetical binding modes A and B of indolylglyoxylamide BzR ligands [27] within the framework of Cook’s pharmacophore/topological model [25]. A2 = hydrogen-bond acceptor; H1 and H2 = hydrogen-bond donors; L1, L2, L3, and LDi = lipophilic clefts; S1 and S2 = sterically forbidden sites; X = OH, OCH3, Cl.
Pharmaceuticals 16 00997 g004
Figure 5. Indolyl-3-glyoxylamides 13 as sedative–hypnotic agents.
Figure 5. Indolyl-3-glyoxylamides 13 as sedative–hypnotic agents.
Pharmaceuticals 16 00997 g005
Figure 6. Indol-3-ylglyoxylamides 4 and 5 as anxioselective agents.
Figure 6. Indol-3-ylglyoxylamides 4 and 5 as anxioselective agents.
Pharmaceuticals 16 00997 g006
Figure 7. SAR of N,N-dialkyl-2-phenylindol-3-ylglyoxylamides (PIGAs) in the pharmacophore/topological model.
Figure 7. SAR of N,N-dialkyl-2-phenylindol-3-ylglyoxylamides (PIGAs) in the pharmacophore/topological model.
Pharmaceuticals 16 00997 g007
Figure 8. N,N-dialkyl-2-phenylindol-3-ylglyoxylamides (PIGAs) 6-12 as anxiolytic and neuroprotective agents.
Figure 8. N,N-dialkyl-2-phenylindol-3-ylglyoxylamides (PIGAs) 6-12 as anxiolytic and neuroprotective agents.
Pharmaceuticals 16 00997 g008
Figure 9. Structure of indibulin 13.
Figure 9. Structure of indibulin 13.
Pharmaceuticals 16 00997 g009
Figure 10. Structure of BPR0C261 14.
Figure 10. Structure of BPR0C261 14.
Pharmaceuticals 16 00997 g010
Figure 11. Structure of BPR0C123 15.
Figure 11. Structure of BPR0C123 15.
Pharmaceuticals 16 00997 g011
Figure 12. Structure of BPR0C259 16.
Figure 12. Structure of BPR0C259 16.
Pharmaceuticals 16 00997 g012
Figure 13. Indol-3-ylglyoxylamides 1723 as antitumor agents.
Figure 13. Indol-3-ylglyoxylamides 1723 as antitumor agents.
Pharmaceuticals 16 00997 g013
Figure 14. Design of β-carboline indol-3-yl-glyoxylamide hybrids as antitumor agents.
Figure 14. Design of β-carboline indol-3-yl-glyoxylamide hybrids as antitumor agents.
Pharmaceuticals 16 00997 g014
Figure 15. Indol-3-ylglyoxylamides 2530 as antitumor agents.
Figure 15. Indol-3-ylglyoxylamides 2530 as antitumor agents.
Pharmaceuticals 16 00997 g015
Figure 16. Linezolid 31, eperezolid 32, and indol-3-ylglyoxylamides 3344 as antibacterial agents.
Figure 16. Linezolid 31, eperezolid 32, and indol-3-ylglyoxylamides 3344 as antibacterial agents.
Pharmaceuticals 16 00997 g016
Figure 17. Structure of indolylglyoxylamide PDE4 inhibitor 45.
Figure 17. Structure of indolylglyoxylamide PDE4 inhibitor 45.
Pharmaceuticals 16 00997 g017
Figure 18. Structure of the first indole derivative synthesized by Lilly as sPLA2 inhibitor 46 [131].
Figure 18. Structure of the first indole derivative synthesized by Lilly as sPLA2 inhibitor 46 [131].
Pharmaceuticals 16 00997 g018
Figure 19. Structure of derivatives 5154 as p38α inhibitors. a IC50 values taken from Kuma et al. [151]. b IC50 value taken from Goldstein et al. [152]. c IC50 values taken from Hill et al. [150].
Figure 19. Structure of derivatives 5154 as p38α inhibitors. a IC50 values taken from Kuma et al. [151]. b IC50 value taken from Goldstein et al. [152]. c IC50 values taken from Hill et al. [150].
Pharmaceuticals 16 00997 g019
Figure 20. Structure and biological activity of derivative 55 and its possible tautomer, 56.
Figure 20. Structure and biological activity of derivative 55 and its possible tautomer, 56.
Pharmaceuticals 16 00997 g020
Figure 21. General structure of the first series of acyl piperazine amide and urea derivatives 57 and of optimized urea derivatives 60 and the biological activity of 58, sofosbuvir 59, indolylglyoxylamides 6162, and cinnamoyl derivatives 6364 as antiviral agents [164].
Figure 21. General structure of the first series of acyl piperazine amide and urea derivatives 57 and of optimized urea derivatives 60 and the biological activity of 58, sofosbuvir 59, indolylglyoxylamides 6162, and cinnamoyl derivatives 6364 as antiviral agents [164].
Pharmaceuticals 16 00997 g021
Figure 22. X-ray structure of the co-crystal of the gp120/73 complex [167]. Reproduced from reference [167]. Copyright 2017 American Chemical Society; https://doi.org/10.2210/pdb5U7O/pdb accessed on 1 May 2023.
Figure 22. X-ray structure of the co-crystal of the gp120/73 complex [167]. Reproduced from reference [167]. Copyright 2017 American Chemical Society; https://doi.org/10.2210/pdb5U7O/pdb accessed on 1 May 2023.
Pharmaceuticals 16 00997 g022
Figure 23. Structures and biological activities of 8,9-dihydrocoscinamide B 75 and its derivatives 7678 as leishmaniasis agents.
Figure 23. Structures and biological activities of 8,9-dihydrocoscinamide B 75 and its derivatives 7678 as leishmaniasis agents.
Pharmaceuticals 16 00997 g023
Figure 24. Structures and biological activities of standard drugs pentamidine and sodium stibogluconate, novel tryptophan 7980, and tetrahydro-β-carboline derivatives 8182 as leishmaniasis agents [204].
Figure 24. Structures and biological activities of standard drugs pentamidine and sodium stibogluconate, novel tryptophan 7980, and tetrahydro-β-carboline derivatives 8182 as leishmaniasis agents [204].
Pharmaceuticals 16 00997 g024
Figure 25. Indolylglyoxylamides 8388 as antiprion agents [207,208].
Figure 25. Indolylglyoxylamides 8388 as antiprion agents [207,208].
Pharmaceuticals 16 00997 g025
Figure 26. (A) Structure of adenosine, NECA, and BAY60-6583; (B) indol-3-ylglyoxylamides 8991 as A2BAR-positive allosteric modulator agents.
Figure 26. (A) Structure of adenosine, NECA, and BAY60-6583; (B) indol-3-ylglyoxylamides 8991 as A2BAR-positive allosteric modulator agents.
Pharmaceuticals 16 00997 g026
Figure 27. Indol-3-yl-glyoxylamides 9295 as carbonic anhydrase activator agents.
Figure 27. Indol-3-yl-glyoxylamides 9295 as carbonic anhydrase activator agents.
Pharmaceuticals 16 00997 g027
Table 1. Biological activities of compound 12.
Table 1. Biological activities of compound 12.
Pharmaceuticals 16 00997 i001
ActivityCell Line/Animal ModelReference
Immunomodulation Human C20 microglial cells[58]
Modulation of inflammatory-based retinal neurodegenerationLPS-induced degeneration in 661W cells[59]
Neuroprotection Female mouse model of primary progressive multiple sclerosis[60]
Maintenance of the correct functionality of microglia in neuroinflammationHuman microglia C20 and HMC3 cells[61]
Table 2. GIIA and group IB (GIB) PLA2 inhibition by indol-3-ylglyoxylamides 4750. Data are taken from Draheim et al. [133].
Table 2. GIIA and group IB (GIB) PLA2 inhibition by indol-3-ylglyoxylamides 4750. Data are taken from Draheim et al. [133].
Pharmaceuticals 16 00997 i002
cpdR1R2hGIIA PLA2
(μM)
hGIB PLA2
(μM)
pGIB PLA2
(μM)
472-(C6H5)C6H4CH2CH30.006 ± 0.0010.3640.097
483-(C6H5)C6H4CH2CH30.009 ± 0.0010.570.007
49C6H5CH2CH30.011 ± 0.0040.7610.015
50----0.009 ± 0.0010.2280.048
Table 3. Biological activity of HIV-1 attachment inhibitors 6572. Data are taken from references [167,191].
Table 3. Biological activity of HIV-1 attachment inhibitors 6572. Data are taken from references [167,191].
Pharmaceuticals 16 00997 i003
cpdEC50 (nM)CC50 (μM)
6586 ± 24 (LAI)145 ± 23
664.0 (JR-FL)
4.9 (LAI)
200
671.52 (JR-FL)>300 (n = 2)
68575.9 (JR-FL)>300 (n = 2)
6921.6 (JR-FL)>300 (n = 2)
701.7 ± 1.6 (JR-FL, n = 11)280
711.47 ± 0.63 (JR-FL)
2.68 ± 1.64 (LAI)
>300
720.88 ± 0.46 (JR-FL, n = 56)
1.15 (LAI)
>300
Table 4. Activity in cellulo of 73 against laboratory strains of HIV-1. Data are taken from Ref. [167].
Table 4. Activity in cellulo of 73 against laboratory strains of HIV-1. Data are taken from Ref. [167].
Pharmaceuticals 16 00997 i004
Co-Receptor TropismVirusEC50 (nM)
CCR5JR-FL0.4 ± 0.1
SF-1620.5 ± 0.2
Bal1.7 ± 0.5
CXCR4LAI0.7 ± 0.4
NL4−32.2 ± 0.6
MN14.8 ± 5.2
IIIb16.2 ± 1.7
RF>2000
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

Barresi, E.; Robello, M.; Baglini, E.; Poggetti, V.; Viviano, M.; Salerno, S.; Da Settimo, F.; Taliani, S. Indol-3-ylglyoxylamide as Privileged Scaffold in Medicinal Chemistry. Pharmaceuticals 2023, 16, 997. https://doi.org/10.3390/ph16070997

AMA Style

Barresi E, Robello M, Baglini E, Poggetti V, Viviano M, Salerno S, Da Settimo F, Taliani S. Indol-3-ylglyoxylamide as Privileged Scaffold in Medicinal Chemistry. Pharmaceuticals. 2023; 16(7):997. https://doi.org/10.3390/ph16070997

Chicago/Turabian Style

Barresi, Elisabetta, Marco Robello, Emma Baglini, Valeria Poggetti, Monica Viviano, Silvia Salerno, Federico Da Settimo, and Sabrina Taliani. 2023. "Indol-3-ylglyoxylamide as Privileged Scaffold in Medicinal Chemistry" Pharmaceuticals 16, no. 7: 997. https://doi.org/10.3390/ph16070997

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