Next Article in Journal
Synthesis and Studies of the Inhibitory Effect of Hydroxylated Phenylpropanoids and Biphenols Derivatives on Tyrosinase and Laccase Enzymes
Next Article in Special Issue
8-Hydroxyquinoline Glycoconjugates Containing Sulfur at the Sugar Anomeric Position—Synthesis and Preliminary Evaluation of Their Cytotoxicity
Previous Article in Journal
Synthesis and Characterization of New Series of 1,3-5-Triazine Hydrazone Derivatives with Promising Antiproliferative Activity
Previous Article in Special Issue
Theoretical and Experimental Investigations of Large Stokes Shift Fluorophores Based on a Quinoline Scaffold
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

One-Pot Multicomponent Synthesis and Bioevaluation of Tetrahydroquinoline Derivatives as Potential Antioxidants, α-Amylase Enzyme Inhibitors, Anti-Cancerous and Anti-Inflammatory Agents

Department of Pharmacy, Faculty of Biological Science, Quaid-I-Azam University, Islamabad 45320, Pakistan
*
Author to whom correspondence should be addressed.
Molecules 2020, 25(11), 2710; https://doi.org/10.3390/molecules25112710
Submission received: 26 May 2020 / Revised: 5 June 2020 / Accepted: 7 June 2020 / Published: 11 June 2020
(This article belongs to the Special Issue Synthesis and Application of Quinolines and Quinoline Derivatives)

Abstract

:
Mankind has always suffered from multiple diseases. Therefore, there has been a rigorous need in the field of medicinal chemistry for the design and discovery of new and potent molecular entities. In this work, thirteen tetrahydroquinoline derivatives were synthesized and evaluated biologically for their antioxidant, α-amylase enzyme inhibitory, anti-proliferative and anti-inflammatory activities. SF8 showed the lowest IC50 of 29.19 ± 0.25 µg/mL by scavenging DPPH free radicals. SF5 showed significant antioxidant activity in total antioxidant capacity (TAC) and total reducing power (TRP) assays. SF5 and SF9 showed the maximum inhibition of α-amylase enzyme i.e., 97.47% and 89.93%, respectively, at 200 µg/mL concentration. Five compounds were shortlisted to determine their anti-proliferative potential against Hep-2C cells. The study was conducted for 24, 48 and 72 h. SF8 showed significant results, having an IC50 value of 11.9 ± 1.04 µM at 72 h when compared with standard cisplatin (IC50 value of 14.6 ± 1.01 µM). An in vitro nitric oxide (NO) assay was performed to select compounds for in vivo anti-inflammatory activity evaluation. SF13 scavenged the NO level to a maximum of 85% at 50 µM concentration, followed by SF1 and SF2. Based on the NO scavenging assay results, in vivo anti-inflammatory studies were also performed and the results showed significant activity compared to the standard, acetylsalicylic acid (ASA).

Graphical Abstract

1. Introduction

The field of drug discovery has conventionally been a dynamic, innovation-driven and highly prosperous sector of the global industry. Revolutionary advancements in health disciplines and technology have helped researchers in the discovery and development of numerous drugs, which play a significant role in the pharmacotherapy of a range of ailments and diseases [1]. Innovations in medicinal chemistry are responsible for the discovery of many breakthrough remedies that have ultimately improved the life and health of human beings over the past century. Continuous efforts in the development and evaluation of new molecules are required to drive the discovery of new medicines [2]. Numerous studies have been focused on the treatment of medical problems such as cancer, oxidative stress, inflammatory disorders, diabetes, etc.
The human body produces free radicals in the form of reactive oxygen species (ROS), which regulate cellular processes, i.e., angiogenesis, platelet aggregation, signal transduction, etc. When produced in excess, free radicals can impose a harmful and deleterious impact on cellular components, i.e., lipids, proteins, DNA, etc. This may lead to many chronic diseases, such as diabetes, cancer and atherosclerosis [3]. Antioxidants can scavenge these free radicals and helps to regulate the progression and prevention of carcinogenesis [4,5]. Cancer is an enigma that has threatened human lives for ages. It has a major impact on society across the globe as a leading cause of death worldwide. Subsequently, the amount of resources spent on its treatment is increasing daily. Therefore, enormous efforts have been invested in the research and development of anti-cancer products both in industry and academia [6].
Diabetes mellitus is one of the most common and emerging diseases in the world [7]. α-Amylase is a key enzyme responsible for the hydrolysis of carbohydrates. Inhibiting α-amylase is a successful approach to controlling postprandial blood sugar within the acceptable range, which is a key factor in controlling diabetes mellitus (DM) [8,9,10].
Inflammation is an indication of many pathological conditions, such as rheumatoid arthritis, osteoarthritis, Alzheimer’s disease, hepatitis, cancer, pulmonary fibrosis and obesity-related diseases [11,12]. Lipopolysaccharide (LPS), a bacterial endotoxin, plays an important role during inflammation by regulating a cascade of cellular responses [13]. Throughout the inflammatory process, pro-inflammatory mediators, cytokines, nitric oxide (NO) and prostaglandin E2 are produced by the inducible isoform of nitric oxide synthase (NOS) and cyclooxygenase-2 (COX-2) [14]. NO is a free radical synthesized from L-arginine and catalyzed by iNOS [15]. Additionally, iNOS is closely associated with a variety of pathological conditions [16] and can act as a cytotoxic agent in pathological conditions [17]. Thus, the inhibition of the production of NO by iNOS can be an effective approach to treat and prevent inflammatory diseases [18].
The concept of multicomponent organic synthesis has received increased attention in organic chemistry since it provides access to diverse libraries of organic molecules [19]. Thus, multicomponent reactions have gained great importance, especially in medicinal chemistry, and are used for the preparation of novel therapeutic agents that have numerous applications [20]. Many biologically active compounds are obtained by using the multicomponent reaction (MCR) approach. The corresponding biological activities exhibited by organic molecules include antileishmanial [21], anti-inflammatory [22], ROCK inhibitors [23], neuroprotective agents [24], acetylcholinesterase inhibitors [25], antimicrobial [26,27,28,29,30], antioxidants [31], anti-tumor [32,33], anti-cancer [34,35,36,37,38,39] and so on.
Quinoline [40] (1-azanaphthalene) is a very significant nitrogen-containing heterocyclic ring structure in medicinal chemistry [41,42]. It has been integrated into many molecules, resulting in compounds with a broad spectrum of potential [43]. A literature survey revealed that quinoline derivatives possess a wide range of biological activities, such as anti-inflammatory, anti-malarial, anti-tumor, anti-hypertensive, anti-microbial, tyrosine PDGF-RTK inhibiting agents, anti-HIV and anti-tubercular, etc. [44,45,46]. Some of the promising compounds with a quinoline ring system are depicted in the Figure 1.
This diversity of biological effects is exhibited by benzofused six-membered heterocyclic rings. From a chemical point of view, the tetrahydroquinoline (THQ) moiety is a promising derivative of quinoline and a potential scaffold in medicinal chemistry due to its broad-spectrum biological profile, as mentioned in Figure 2. Helquinoline is a potent tetrahydroquinoline-derived antibiotic that has been extracted from Janibecter limosus [47]. In particular, tetrahydroquinoline exhibits a wide range of biological activities, for example, anti-HIV [48], antitrypanosomal [49], psychotropic [50], anti-inflammatory [51], antibacterial [52], antimalarial [53], antifungal [54] and antitumor activities [55].
The incorporation of a functional group into a pharmacophore is an attractive approach to the design and synthesis of new bioactive compounds. Considering the pharmacological potential of the tetrahydroquinoline moiety and our long-standing interest in exploring Mannich base chemistry [56], in the present study, we used a one-pot MCR approach to prepare a series of tetrahydroquinoline derivatives. The majority of the synthesized Mannich bases have not been reported in the literature to date. All compounds were subsequently evaluated in biological studies for their pharmacokinetic, antioxidant, α-amylase enzyme inhibition, cytotoxic and anti-inflammatory potential. The comprehensive biological profile of tetrahydroquinoline-derived N-Mannich bases by one-pot MCR reaction has not yet been thoroughly investigated. A literature survey demonstrated that there are no studies concerning the cytotoxic and anti-proliferative activities of tetrahydroquinoline-derived N-Mannich bases.

2. Results and Discussions

2.1. Chemistry

Multicomponent reactions (MCRs) are considered to be very important and powerful reactions in combinatorial and medicinal chemistry [57] because of the benefits they offer in terms of the synthesis of diverse structures that will lead to an increase in the economy of organic synthesis [58,59] by shortening the time span [60].
The reaction between aldehydes, amines or ammonia and heterocyclic acidic proton-containing pharmacophores is known as a Mannich reaction. It was reported for the first time by a German chemist, Carl Mannich, in 1912, for the synthesis of atropine glucoside [61,62].
The synthetic route used to obtain compounds (SF1SF13) is described in Scheme 1. N-Mannich products were prepared by a one-pot, three-component condensation reaction, performed under reflux conditions. Formaldehyde (S), amine (113), and tetrahydroquinoline (F) were placed in a pressure tube and dissolved in ethanol and a catalytic amount of HCl was added. The mixture was refluxed at 80 °C for ~5–7 h, leading to the formation of the N-Mannich base. The product precipitated from the reaction mixture and the formed solid was separated, washed and recrystallized by ethanol. For some of the products, an additional purification step by flash column chromatography was necessary. To establish the exact position of the substituents of the titled compound, we hypothesized that the reaction was carried out at position 1, where the acidic proton-H of the parent compound reacts with the oxygen of formaldehyde (Formalin 37%) and takes one -H from the amines, yielding a water molecule as a byproduct during the reflux condensation reaction.
The Mannich reaction is a condensation reaction where the substrate used is XH compounds, with X being any heteroatom (C, N, S, etc.) with nucleophilic properties. The reaction is the nucleophilic addition of an amine to a carbonyl group of an aldehyde, followed by dehydration to a Schiff base, which acts as an electrophile and reacts with a compound containing an acidic proton. The reaction is an example of an SN2 addition reaction [56]. From the structure of tetrahydroquinoline, it should be kept in mind that it contains one nucleophilic site, i.e., −NH. Under a basic environment, the nucleophile site gets deprotonated, which results in an increase in electronic condensation on the electronegative nitrogen atom (nucleophilic activation). Thus, the tetrahydroquinoline anion starts a nucleophilic attack on the electrophile and carries out an efficient SN2 displacement and intramolecular cyclization. All the tetrahydroquinoline-derived Mannich bases were characterized by various spectroanalytical techniques. In the IR spectrum of quinoline derivatives, the strong bonds observed at around 1600 cm−1 and 1505 cm−1 correspond to aromatic C=C stretching. The specific carbonyl peak shown by some synthetic compounds appeared at 1620–1680 cm−1. In proton NMR, aromatic protons were observed mostly downfield, near δ 6.52–7.7 ppm, as per the substitution pattern at various positions of the aromatic ring. The methylene proton of −N linkage (N-CH2-N) was resonated at around δ 3.52–5.44 ppm and gave a singlet (s). Carbon NMR further confirmed the structure of the synthesized compounds. In carbon NMR of synthetic compounds, most downfield carbon was of aromatic quinoline carbon observed near δ 115–145 ppm. The carbon of methylene imine linkage was observed at δ ~69 ppm. Solvent peaks for CDCl3 and DMSO-d6 were observed at δ 77.2 ppm and 39.5 ppm, respectively. All other aromatic signals were observed at the expected regions. In the mass spectral analysis of synthetic compounds (SF1SF13), the molecular ion peak for these compounds was observed as (M+) for all the compounds.

2.2. ADME Predictions

Synthetic compounds were evaluated for their in silico ADME for the assessment of pharmacokinetic parameters. A computational tool was used for the prediction of drug-likeness and molecular properties to obtain new drug-like leads. The values are presented in Table 1.
According to the biopharmaceutics drug disposition classification system (BDDCS), new molecules that are of large molecular weight, poorly water-soluble and lipophilic are categorized under BCS Class II. Lipinski et al. [63] demonstrated that molecules obtained via high-throughput screening (HTS) tend to have lipophilicity and greater molecular weights than leads in the pre-HTS era [64]. The Lipinski rule of five was developed to achieve the “druggability” of new molecular entities (NMEs) [65], which predicts that poor absorption or permeation is more likely when there are more than five H-bond donors, 10 H-bond acceptors, the molecular weight is greater than 500 and the calculated log p (CLog p) is greater than 5 [63,65]. In the current study, all the synthetic compounds were shown to have followed Lipinski’s rule of five, which demonstrates that these all fulfill the basic requirement for orally active drugs and tend to have low attrition rates during clinical trials.

2.3. Bioevaluation

2.3.1. Antioxidant Assays

DPPH Assay

The % free radical scavenging activity (%FRSA) of the synthetic compounds was evaluated by the discoloration of an unstable DPPH reagent to the stable yellow-colored 2,2-diphenyl-1-picrylhydrazine molecule by accepting one electron from a donor antioxidant [66]; the mechanism is shown in Scheme 2.
The results of the %FRSA of the synthetic compounds and their IC50 data are summarized in Table 2.
All the synthetic compounds were shown to scavenge the DPPH free radical to some extent. SF8, which has an electronegative atom, was shown to have the most potent effect, with an IC50 value of 29.19 ± 0.25 µg/mL, which was lower than the standard, ascorbic acid and quercetin, with values of 41.38 ± 0.34 and 41.64 ± 1.01 µg/mL, respectively. Among the other synthetic series, SF4, SF6 and SF12 exhibited good antioxidant potential against DPPH free radicals, with IC50 values of 29.79 ± 0.26, 35.89 ± 0.33 and 39.33 ± 0.28 µg/mL, respectively. These characteristic results could be because of the formation of the hydrogen bond between an amine and a lone pair of nitrogen present on an unstable DPPH free radical. The high activity of aromatic amines is because of their ability to form nitroxyl radicals [68]. The literature survey supports the results of SF6 and SF8 being good antioxidants because of the nature of their substituents, i.e., electron-withdrawing (chlorine in SF8) and electron-donating (piperazine moiety in SF6) groups attached to the heterocyclic core [69].

Phosphomolybdenum-Based Total Antioxidant Capacity (TAC) and Total Reducing Power (TRP)

In the presence of an antioxidant, the phosphomolybdate ion gets reduced and a green phosphate–molybdate complex is formed, which is analyzed spectrophotometrically [70]. The highest TAC was found to be 285.46 ± 2.1 µg/AAE for SF5 bearing a methyl piperazine moiety, followed by SF7SF9 and SF12, with TAC values of 242.77 ± 2.31 µg/AAE, 212.71 ± 231 µg/AAE, 179.76 ± 1.6231 µg/AAE and 191.19 ± 1.1231 µg/AAE, respectively (Figure 3 and Figure 4 (standard curve)).
A TRP assay was carried out to evaluate the reducing potential of synthetic compounds. The method implicates reductones. These reductones are species with antioxidant potential that is believed to be due to their proton-donating capacity [71]. This results in a discontinuation of a free radical chain. The maximum value of reducing power was 104.7 ± 1.9 µg/AAE for SF5, followed by 103.18 ± 2.4 µg/AAE for SF7 and 84.57 ± 1.9 µg/AAE for SF12 (Figure 3 and Figure 4 (standard curve)). The results indicate that synthetic compounds have the potential to stabilize free radicals by donating electrons and exhibited reductive potential with the highest reducing power [72]. The lowest reducing potential was observed for SF1, i.e., 33.22 ± 1.5 µg/AAE. SF10 showed no activity in this assay. SF7, with a diphenylamine moiety, showed the highest activity in antioxidant assays due to the reason that secondary amines with an aromatic ring and nitrogen atom attached to them are good antioxidants [73] and are considered as singlet oxygen scavengers [74].

2.3.2. α- Amylase Inhibitory Assay

In the present research, the inhibition of the α-amylase enzyme was assessed in vitro. The results are depicted in Figure 5.
Among the tested compounds, SF5 showed the maximum % enzyme inhibition of 97.37%, followed by SF9 and SF7, with 93%, and 88.67%, respectively. SF2 showed a % enzyme inhibition close to the standard, acarbose, i.e., 79.97%. However, SF13, with an acetaminophen moiety, showed the lowest enzyme inhibition, i.e., 10.51%. The IC50 data were calculated by GraphPad Prism software. SF5 showed the most potent value compared to the standard, acarbose (38.05 ± 0.98 µg/mL), followed by SF9 and SF8, i.e., 18.55 ± 0.89 and 27.39 ± 1.43 µg/mL, respectively.

2.3.3. Evaluation of Cytotoxic Potential

Brine Shrimp Cytotoxicity Assay

Cytotoxicity via a brine shrimp lethality assay was studied to reveal the cytotoxicity potential of the synthetic compounds. Toxicity to brine shrimps has a direct correlation with anti-cancer activity [75]. The brine shrimp larvae correspond to the mammalian system [76] since the DNA-dependent RNA polymerases of Artemia salina have been reported to be similar to the mammalian type [77]. This test is not only used to predict the anti-cancer potential of corresponding compounds but also anti-microbial and pesticidal behavior [78].
Thirteen of the newly synthesized compounds were screened against brine shrimp nauplii. The cytotoxicity of synthetic compounds was determined by calculating the median lethal concentration, LD50 (concentration at which 50% of the nauplii died). The results are shown in Table 3, indicating that SF5, SF4, SF7 and SF3 have good activity against brine shrimp nauplii, in comparison to a standard, doxorubicin (124 ± 1.54 µg/mL), with LD50 values of 94 ± 1.06, 100 ± 1.78, 100.8 ± 1.14 and 102 ± 1.98 µg/mL, respectively, which corresponds to a high toxicity against cancer cells. SF1 and SF8 showed moderate toxicity, 120.8 ± 2.3 µg/mL and 123.4 ± 1.87 µg/mL, respectively. The rest of the compounds exhibited weaker activity and were considered to be safe or non-toxic. Further cytotoxicity of the compounds was evaluated using cancer cell lines (Hep-2C).

Cytotoxicity Against Raw Macrophages and Cancer Cell Lines (Hep-2C)

Before the screening studies for the % cell viability inhibition against Hep-2C, it was essential to perform cytotoxicity tests against raw macrophages to determine the toxic and non-toxic behavior of every compound. The compounds that showed the lowest IC50 are considered to be toxic, as compared to the reference drug, and are screened out against cancer cells. The mechanism is depicted in Scheme 3.
The % cell viability of the synthetic library was evaluated in vitro by measuring the mitochondrial dehydrogenase activity (MTT test) on raw macrophages. All the compounds were shown to have a dose-dependent dose response (Table 4).
A 100 µM concentration was considered to be toxic for compounds because of the least number of viable cells. To determine the concentration required to achieve a 50% inhibition of cells induced by each compound, IC50 values (µM) were determined and their results were compared with a cytotoxic drug, i.e., doxorubicin. SF1, SF4, SF5 and SF9 exhibited the most potent activity, with IC50 values of 6.232 ± 0.01, 7.208 ± 0.05, 6.181 ± 0.05 and 14.24 ± 0.26 µM, respectively, while the standard drug doxorubicin had an IC50 value of 17.08 ± 0.37 µM, conclusively, which was considered toxic for cells. The activity of the most potent compounds, SF1 and SF5, was three times more than that of doxorubicin. Piperidine-bearing compounds are known to have significant anti-cancer potential, as proved in the literature [79]. From the previous data, it is evident that the electronegativity of the substituents at the heterocyclic ring steer the magnitude of activity [80], as in the case of SF9 with an electronegative bromine atom in it. The method is not specific to anti-tumor activity. However, a positive correlation was found between the least cell viability and the cytotoxicity toward some cancer cell lines [81].

MTT Assay Against Hep-2C Cells

From the results of the brine shrimp assay and cell cytotoxicity, compounds that showed significant activity and the lowest IC50 values in both assays were shortlisted for further analysis against Hep-2C cells. SF1, SF4, SF5, SF7 and SF8 showed cytotoxicity against brine shrimp nauplii at IC50 values of 120 ± 2.3, 100 ± 1.78, 94 ± 1.06, 100.8 ± 1.14 and 123.4 ± 1.87 µg/mL, respectively. The same five compounds showed good cytotoxic profiles against raw macrophages at IC50 values of 6.232 ± 0.01, 7.208 ± 0.05, 6.181 ± 0.05, 18.6 ± 0.13 and 21.23 ± 0.33 µM, respectively.
The results were plotted against % cell viability of Hep-2C cells and sample concentration, while cisplatin was taken as a standard (Figure 6). From the achieved results, it was clear that SF1, SF4 and SF7 exhibited better activity against the cell lines than SF5 and SF8. All other compounds were not considered to have anti-cancer potential. The anti-cancer activity was determined at 24, 48 and 72 h (Figure 6). The anti-proliferative activity of Mannich bases is supposed to be due to the formation of enones [82].
IC50 values were determined by GraphPad software and plotted against time intervals (24 h, 48 h and 72 h) as summarized in Table 5. Previous data proved that heterocyclic compounds bearing piperidine in their structure are highly cytotoxic (SF1) [83]. Electronegative atoms impart moderate anti-tumor activity to compounds [16], as in SF8, which has a -chloro group in it. Electronegative or electron-withdrawing atoms will lead to an increase in cytotoxic potential (bromine, chlorine, fluorine, etc.). The structure–activity relationship (SAR) was established from these findings, which states that by derivatizing the already reported anti-cancerous lead molecule with secondary amines of cytotoxic potential will impart more cytotoxicity activity to the new product that has both nuclei. The observation of microscopic analysis of MTT stained cells and crystals is depicted in Figure 7.

2.3.4. Anti-Inflammatory Activity

Nitric Oxide Scavenging Assay (NO)

Nitric oxide (NO) is an indicator of inflammation. It is an important chemical mediator produced by endothelial cells, macrophages and neurons, and is involved in the regulation of many different physiological processes [84]. An excess amount of nitric oxide (NO) is associated with several diseases, i.e., inflammation, cancer and other pathological conditions. Therefore, NO production inhibition is a principle key factor for the screening of drugs with anti-inflammatory potential [85].
A Griess reagent, a spectrophotometric determination of nitrite level, was carried out to measure the nitrite level in the conditioned medium of macrophages treated with lipopolysaccharide (LPS). Sodium nitrite (NaNO2) was used as a standard compound for the standard curve (Figure 8). The NO concentrations were obtained using the standard curve equation: y = 0.0083x + 0.0017, R2 = 0.9999, where y is the absorbance at 540 nm and x is the NO concentration in µM. The inhibitory activity of the tested compounds towards NO generated by LPS-induced macrophages was obtained from the calculated value of x.
The highest concentration to determine the inhibition of NO production was considered to be 100 µM, and from the results summarized in Table 6, it is evident that the synthetic compounds showed the concentration-dependent response. Among the thirteen compounds, the ones that inhibited the NO production to a greater extent even at 1 µM were considered to be potent and were considered for further in vivo assays.
The results from this finding showed that SF13 has the highest NO inhibitory potential, with the maximum inhibition of 75% at 1 µM, followed by SF1 and SF2, with NO production at 1 µM concentration of 27.771 ± 0.45 and 32.231 ± 0.46, respectively (Table 6).
The higher NO inhibition by the synthetic compounds corresponds to the epidemiological data that suggest that lower incidences of certain inflammatory diseases, i.e., cancer, arthritis, diabetes and acquired immune deficiency syndrome (AIDS) [86,87]. These results indicate that SF13 and SF1SF3, with acetaminophen, piperidine, morpholine and pyrrolidine amine incorporated into the basic tetrahydroquinoline ring, may suppress NO generation to a maximum extent by inhibiting iNOS enzyme activity. It is considered that the tested compounds that have shown the maximum decrease in NO production have beneficial therapeutic effects in the management of inflammatory conditions. Based on the results of NO, compounds were selected for in vivo anti-inflammatory activity.

In Vivo Anti-Inflammatory Activity

Synthetic compounds that have significantly inhibited the % nitric oxide (NO) production were further assessed for in vivo anti-inflammatory activity. Dose optimization results showed that doses of 0.1 and 1 mg/kg did not show any significant decrease in inflammatory behavior. However, at 10 mg/kg there was a significant decrease in paw edema (Figure 9). That is why the 10 mg/kg dose was selected for further study.
The inhibitory potential of SF1SF3 and SF13 was evaluated by measuring paw thickness (Table 7). The synthetic compounds showed a significant decrease in paw edema with time duration as compared to the standard, acetylsalicylic acid (ASA). The results revealed that the compounds showed anti-inflammatory activity with meaningful statistical results. The results revealed that the compound with acetaminophen (SF13) significantly decreased the inflammation because of the acetaminophen moiety, which is categorized under non-steroidal anti-inflammatory drugs (NSAID) and has the potential to inhibit COX-1 and COX-2 enzymes involved in the production of prostaglandin [88]. Compounds SF1, SF2 and SF13, with piperidine, morpholine and acetaminophen, also showed significant results in comparison to ASA.

2.3.5. Structure-Activity Relationship (SAR)

The structure activity relationship is established from the findings of chemical moieties and their relationship to its biological activity. The results are summarized in Figure 10.

3. Materials and Methods

Unless otherwise noted, all commercially available compounds and solvents were purchased from Sigma-Aldrich and Merck, Germany. Analytical thin-layer chromatography (TLC) was performed with aluminum sheets, silica gel 60 F254 (Merck), by a solvent system, EtOAc: pentane in a 7:2 ratio and the products were visualized with UV irradiation (254 nm). Gallenkamp melting point apparatus was used to determine the melting points with open capillaries that and are uncorrected. FTIR was recorded by the KBr pellets method by the PerkinElmer spectrum using the attenuated total reflectance (ATR). Nuclear magnetic resonance (NMR) spectra were recorded on an Agilent VNMR 400 (1H NMR: 400 MHz, 13C NMR: 101 MHz) or an Agilent VNMR 600 (1H NMR: 600 MHz, 13C NMR: 151 MHz) spectrometer. The chemical shifts are given in parts per million (ppm) relative to the residual solvent peak of the nondeuterated solvent (CDCl3: 1H NMR: δ = 7.26 ppm; 13C NMR: δ = 77.00 ppm). The multiplicity was reported with the following abbreviations: s = singlet, d = doublet, t = triplet, m = multiplet, p = pentet, br = broad signal, dd = doublet of doublet, dt = doublet of triplet, ddt = doublet of doublet of triplet, td = triplet of doublet, tp = triplet of pentet, tdd = triplet of doublet of doublet. Mass spectra were recorded on a Finnigan SSQ 7000 spectrometer. The PE 2400 Series II CHNS/O Analyzer was used to determine the content of carbon, hydrogen and nitrogen in organic materials.

3.1. General Procedure for the Synthesis of N-Mannich Bases of Tetrahydroquinoline

In a pressure tube equipped with a stirring bar, amines 113 (1.0 equiv, 0.5 mmol), formaldehyde S (2.0 equiv, 1 mmol) and tetrahydroquinoline F (1.0 equiv, 0.5 mmol) were added into 3 mL of ethanol and the mixture was heated to 80 °C in an oil bath under reflux conditions. Concentrated HCl (2 mol %) was added as a catalyst. The reaction progress and completion of the reaction was monitored by TLC. After ~5–7 h of refluxing, the reaction mixture was removed and cooled at room temperature. The reaction mixture was neutralized with a NaHCO3 solution and extracted with DCM (dichloromethane) (3 × 10 mL). The organic layer was dried over Na2SO4 and evaporated by a rotary evaporator. The product was washed and recrystallized by ethanol. For some of the compounds, an additional purification step of flash column chromatography was performed by using ethyl acetate: pentane in a 7:2 ratio.
The target compounds SF1SF13 were synthesized according to the synthetic route shown in Scheme 4.
1- (Piperidine-1-ylmethyl) 1, 2, 3, 4-tetrahahydroquinoline (SF1)
White solid; yield 82%; m.p. 130–137 °C; Rf = 0.89; IR (ATR) υmax 3397, 2850, 2327, 1710, 1613, 1437, 1191, 1042, 993 cm−1; 1H-NMR (600 MHz, CDCl3) δ ppm 6.94 (t, 1H), 6.91 (t, J = 4.7 Hz, 1H), 6.64 (dd, J = 7.0, 3.7 Hz, 1H), 6.60–6.54 (m, 1H), 4.70 (s, 2H), 3.27 (q, J = 5.5 Hz, 6H), 2.73 (t, 2H), 1.93 (td, J = 5.3, 2.6 Hz, 6H), 1.57 (t, 2H); 13C NMR (151 MHz, CDCl3) δ ppm 144.92, 127.11, 125.48, 63.50, 51.73, 47.07, 28.17, 25.92, 25.76, 22.35; EIMS m/z 230.2(M +), Anal. (C15H22N2): C 78.21, H 9.63, N 12.16, Calcd. C 78.27, H 9.61, N 12.19.
4- ((3, 4-Dihydroquinolin—1(2H)-yl) methyl) morpholine (SF2)
Light yellow solid; yield 84%; m.p. 142–148 °C; Rf = 0.63; IR (ATR) υmax 3390, 2834, 2322, 2083, 1898, 1663, 1611, 1503, 1304, 1049 cm−1; 1H NMR (600 MHz, CDCl3) δ ppm 7.04 (td, J = 7.9, 7.4, 1.6 Hz, 1H), 6.95 (dd, J = 7.4, 1.5 Hz, 1H), 6.88 (dd, J = 8.3, 1.1 Hz, 1H), 6.61 (td, J = 7.3, 1.1 Hz, 1H), 3.77 (s, 2H), 3.71 (t, 4H), 3.34 (t, 2H), 2.76 (t, J = 6.4 Hz, 2H), 2.50 (s, 4H), 1.93 (tt, 2H); 13C NMR (151 MHz, CDCl3) δ ppm 145.48, 129.03, 127.10, 126.88, 116.54, 67.01, 66.87, 52.00, 50.80, 27.94, 22.40; EIMS m/z 232 (M+), Anal. (C14H20N2O): C 72.38, H 8.68, N 12.06, Calcd: C 72.39, H 8.67, N 12.08.
1-(Pyrollidin-1-ylmethyl) -1, 2, 3, 4-tetrahydroquinoline (SF3)
Yellow solid; yield 80%; m.p. 115–117 °C; Rf = 0.86; IR (ATR) υmax 3388, 2921, 2323, 1990, 1656, 1501, 1300, 1153, 972 cm−1; 1H NMR (600 MHz, CDCl3) δ ppm 7.03 (td, 1H), 6.94 (dd, 1H), 6.86 (td, 1H), 6.60 (dd, 1H), 3.98 (s, 2H), 2.77 (t, J = 6.3 Hz, 2H), 2.60 (ddt, J = 6.7, 3.9, 2.1 Hz, 5H), 1.95 (t, 2H), 1.79 (ddd, J = 6.7, 4.6, 1.9 Hz, 5H); 13C NMR (151 MHz, CDCl3) δ ppm 145.48, 129.05, 126.90, 122.17, 116.16, 67.48, 51.84, 26.97, 23.53, 22.41; EIMS m/z 216 (M+), Anal. (C14H20N2): C 77.73, H 9.32, N 12.95, Calcd. C 77.77, H 9.31, N 12.97.
N-((3, 4-dihydroquinolin-1(2H)-yl), methyl)-N-propylpropan-1 amine (SF4)
Yellow powder; yield 79%; m.p. 123–132 °C; Rf = 0.74; IR (ATR) υmax 3841, 2831, 1685, 1510, 1304, 972 cm−1.; 1H NMR (600 MHz, CDCl3) δ ppm 6.98 (dd, 1H), 6.91 (td, 1H), 6.74 (dd, 1H), 6.66 (td, 1H), 4.71 (s, 2H), 3.29 (t, J = 11.4 Hz, 2H), 2.78–2.60 (m, 4H), 2.59 (t, 2H), 1.99–1.95 (m, 4H), 1.50 (t, 2H), 0.89 (tt, 6H); 13C NMR (151 MHz, CDCl3) δ ppm 145.95, 128.98, 127.17, 126.79, 122.44, 115.97, 67.48, 59.85, 46.94, 27.87, 22.41, 20.41, 11.82; EIMS m/z 246.22 (M+); Anal. (C16H26N2): C 77.99, H 10.64, N 11.37, Found: C 77.95, H 10.67, N 11.39.
1-(4-Methylpiperazin-1-yl) methyl)-1, 2, 3, 4 -tetrahydroquinoline (SF5)
Light yellow solid; yield 81%; m.p. 167 °C; Rf = 0.76; IR (ATR) υmax 3354, 3061, 1605, 1461, 1354, 1003, 949 cm−1; 1H NMR (600 MHz, CDCl3) δ ppm 7.01–6.99 (m, 1H), 6.88 (td, J = 7.5, 2.0 Hz, 1H), 6.56 (td, J = 7.5, 2.0 Hz, 1H), 6.37 (dd, J = 7.5, 2.0 Hz, 1H), 4.42 (s, 2H), 3.33 (t, J = 5.1 Hz, 2H), 2.90 (td, J = 6.2, 1.0 Hz, 2H), 2.51–2.47 (m, 9H), 2.24 (s, 3H), 2.08–2.04 (m, 2H); 13C NMR (151 MHz, CDCl3) δ ppm 145.62, 128.98, 126.89, 122.28, 116.37, 67.47, 54.95, 50.29, 48.41, 46.08, 27.91, 22.42; EIMS m/z 245.20 (M+); Anal. (C15H23N3): C 73.47, H 9.42, N 17.17, Found: C 73.43, H 9.45, N 17.13.
1-(Piperazine-1-yl methyl)-1, 2, 3, 4-tetrahydroquinoline (SF6)
Off-white crystals; yield 87%; m.p. 220–227 °C; Rf = 0.77; IR (ATR) υmax 3863, 3379, 2833, 2493, 2285, 2088, 1900, 1655, 1611, 1500, 1314, 1050 cm−1; 1H NMR (600 MHz, CDCl3) δ ppm 6.97–6.94 (m, 2H), 6.59 (dd, J = 7.4, 1.2 Hz, 1H), 6.47 (d, J = 1.1 Hz, 1H), 3.81 (s, 2H), 3.30–3.29 (m, 4H), 2.94–2.85 (m, 4H), 2.76 (t, J = 6.4 Hz, 4H), 2.17 (s, 1H), 1.95–1.93 (m, 2H); 13C NMR (151 MHz, CDCl3) δ ppm 145.48, 129.03, 127.10, 126.88, 122.36, 116.54, 67.01, 52.00, 50.80, 46.93, 28.09, 22.21; EIMS m/z 231.34 (M+). Anal. (C14H21N3): C 72.69, H 9.15, N 18.16, Found: C 72.66, H 9.19, N 18.21.
N-((3, 4-dihydroquinolin-1(2H)-yl) methyl)-N-phenylaniline (SF7)
Brown crystals; yield 82%; m.p. 243–248 °C; Rf = 0.81; IR (ATR) υmax 3378, 2921, 2692, 2494, 2285, 2110, 1904, 1794, 1592, 1311, 1174, 946 cm−1; 1H NMR (600 MHz, DMSO-d6) δ ppm) 7.24–7.20 (m, 8H), 7.05–6.99 (m, 3H), 6.88 (td, J = 7.5, 2.0 Hz, 1H), 6.56 (td, J = 7.5, 2.0 Hz, 1H), 6.38 (dd, J = 7.5, 2.0 Hz, 1H), 5.12 (s, 2H), 3.33 (t, J = 5.1 Hz, 2H), 2.90 (td, J = 6.3, 1.1 Hz, 2H), 2.06 (tt, J = 6.2, 5.1 Hz, 2H); 13C NMR (151 MHz, DMSO-d6) δ ppm 148.27, 129.32, 129.26, 129.13, 122.46, 122.22, 121.49, 121.45, 120.95, 120.68, 117.76, 111.17, 111.04, 89.72, 62.07, 55.77, 54.85, 51.42, 49.45, 40.02, 22.48, 22.38; EIMS m/z 314 (M+); Anal. (C22H22N2): C 84.04, H 7.05, N 8.91, Found: C 83.07, H 7.07, N 8.91.
3-Chloro-N-((3, 4-dihydroquinolin-1(2H)-yl) methyl) aniline (SF8)
Brown solid; yield 77%; m.p. 219 °C; Rf = 0.83; IR (KBr) υmax 3357, 2921, 2835, 2287, 2113, 1795, 1598, 1314, 1197, 989 cm−1; 1H NMR (600 MHz, DMSO-d6) δ ppm 7.09–7.03 (m, J = 27.7, 6.8 Hz, 3H), 7.01–6.96 (m, 2H), 6.74 (s, 1H), 6.69–6.65 (dd, 1H), 6.54 (d, J = 2.9 Hz, 1H), 4.70 (s, 2H), 4.35 (s, 1H), 3.40 (t, 2H), 2.80 (d, J = 12.5 Hz, 2H), 2.01–1.97 (m, 2H); 13C NMR (151 MHz, DMSO-d6) δ ppm 147.67, 134.84, 130.66, 127.57, 127.27, 124.27, 121.23, 115.97, 115.36, 114.92, 114.20, 58.76, 48.77, 27.00, 22.40; EIMS m/z 272.78(M+). Anal. (C16H17ClN2) C 70.45, H 6.28, N 10.27, Found: C 70.49, H 6.23, N 10.28.
4-Bromo-N-((3, 4-dihydroquinolin-1(2H)-yl) methyl) aniline (SF9)
Green solid; yield 80%; m.p. 200 °C; Rf = 0.81; IR (ATR) υmax 3631, 2834, 2287, 1902, 1612, 1313, 1194, 1070, 944 cm−1; 1H NMR (600 MHz, DMSO-d6) δ ppm 7.24–7.22 (m, 2H), 7.00 (ddt, J = 7.5, 2.0, 0.9 Hz, 1H), 6.88 (td, J = 7.5, 2.0 Hz, 1H), 6.67–6.65 (m, 2H), 6.56 (td, J = 7.5, 2.0 Hz, 1H), 6.38 (dd, J = 7.5, 2.0 Hz, 1H), 5.12 (s, 2H), 3.43 (s, 1H), 3.33 (t, J = 5.2 Hz, 2H), 2.90 (td, J = 6.1, 1.0 Hz, 2H), 2.06 (tt, J = 6.1, 5.2 Hz, 2H); 13C NMR (151 MHz, DMSO-d6) δ ppm 147.42, 144.96, 132.16, 127.52, 127.14, 124.22, 122.56, 119.33, 115.06, 114.31, 111.50, 61.78, 48.79, 26.98, 22.38; EIMS m/z 316 (M+). Anald. (C16H17BrN2) C 60.58, H 5.40, N 8.83, Found: C 60.54, H 5.42, N 8.86.
N-((3, 4-dihydroquinolin-1(2H)-yl) methyl)-N-phenylacetamide (SF10)
White crystals; yield 78%; m.p. 220 °C; Rf = 0.93; IR (ATR) υmax 3863, 3371, 2931, 2689, 2326, 2110, 1906, 1661, 1598, 1372, 1073, 884 cm−1; 1H NMR (600 MHz, CDCl3) δ ppm 7.49–7.47 (m, 2H), 7.37–7.28 (m, 3H), 7.00 (ddt, J = 7.5, 2.0, 1.0 Hz, 1H), 6.88 (td, J = 7.5, 2.0 Hz, 1H), 6.56 (td, J = 7.5, 2.0 Hz, 1H), 6.38 (dd, J = 7.5, 2.0 Hz, 1H), 5.44 (s, 2H), 3.33 (t, J = 5.1 Hz, 2H), 2.90 (td, J = 6.2, 1.0 Hz, 2H), 2.08–2.04 (m, 2H), 2.05 (s, 3H); 13C NMR (151 MHz, CDCl3) δ ppm 168.68, 142.24, 137.99, 128.92, 128.22, 128.11, 127.22, 125.55, 124.24, 119.97, 117.44, 49.53, 27.74, 22.57, 22.46; EIMS m/z 280 (M+). Anald. (C18H20N20) C 77.11, H 7.19, N 9.99, Found: C 77.13, H 7.19, N 9.96.
N-((3, 4-dihydroquinolin-1(2H)-yl) methyl)-N-ethylethanamine (SF11)
Off-white solid; yield 81%; m.p. 110 °C; Rf = 0.86; IR (ATR) υmax 3851, 3373, 2838, 2235, 2003, 1895, 1610, 1496, 1198, 1093 cm−1; 1H NMR (400 MHz, CDCl3) δ ppm 6.99–6.95 (dd, J = 8.2, 1.6 Hz, 1H), 6.93–6.87 (m, 1H), 6.85–6.80 (m, 1H), 6.53 (dt, J = 23.3, 7.3, 1.2 Hz, 1H), 3.81 (s, 2H), 3.26–3.18 (m, 2H), 2.69 (td, 2H), 2.60–2.47 (m, 4H), 1.86 (tt, 2H), 0.99–0.92 (m, 6H); 13C NMR (101MHz, CDCl3) δ ppm 145.93, 129.04, 127.14, 126.88, 122.46, 116.09, 46.98, 27.00, 22.46, 11.88; EIMS m/z 218 (M+). Anald. (C14H22N2) C 77.01, H 10.16, N 12.83, Found: C 77.04, H 10.13, N 12.83.
N-((3, 4-dihydroquinolin-1(2H)-yl) methyl)-4-methoxyaniline (SF12)
Pale yellow solid; yield 86%; m.p. 215 °C; Rf = 0.78; IR (ATR) υmax 3853, 3393, 2997, 2833, 2323, 2160, 2077, 1887, 1611, 1305, 1154, 1036, 975 cm−1; 1H NMR (600 MHz, CDCl3) δ ppm 6.97–6.87 (m, 4H), 6.73 (dd, J = 17.1, 9.0 Hz, 2H), 6.65 (dd, J = 8.4, 1.1 Hz, 1H), 6.54 (td, 1H), 4.60 (s, 2H), 4.32 (s, 1H), 3.67 (s, 3H), 3.21 (td, 2H), 2.69 (s, 1H), 1.87–1.84 (m, 2H);13C NMR (151 MHz, CDCl3) δ ppm 154.07, 144.97, 143.67, 128.92, 127.31, 124.21, 122.18, 120.93, 120.09, 114.47, 55.77, 55.56, 46.97, 26.98, 22.39; EIMS m/z 269 (M+). Anald. (C17H20N2O) C 76.09, H 7.51, N 10.54, Found: C 76.07, H 7.53, N s10.54.
N-((3, 4-dihydroquinolin-1(2H)-yl) methyl)-N-(4-hydroxyphenyl) acetamide (SF13)
White solid; yield 84%; m.p. 285 °C; Rf = 0.91; IR (ATR) υmax 3249, 2831, 2611, 2063, 1893, 1611, 1454, 1370, 1233, 1019, 833 cm−1; 1H NMR (400 MHz, DMSO-d6) δ ppm 9.95 (s, 1H), 7.08–7.06 (m, 2H), 7.00 (ddt, J = 7.5, 2.0, 1.0 Hz, 1H), 6.88 (td, J = 7.5, 2.0 Hz, 1H), 6.72–6.69 (m, 2H), 6.56 (td, J = 7.5, 2.0 Hz, 1H), 6.38 (dd, J = 7.5, 2.0 Hz, 1H), 5.44 (s, 2H), 3.33 (t, J = 5.2 Hz, 2H), 2.90 (td, J = 6.0, 1.0 Hz, 2H), 2.06 (tt, J = 6.0, 5.2 Hz, 2H), 2.03 (s, 3H); 13C NMR (101 MHz, DMSO-d6,) δ ppm 168.58, 153.46, 145.77, 130.06, 128.73, 127.15, 127.12, 125.40, 120.95, 115.54, 88.78, 49.27, 27.49, 22.64, 22.39; EIMS m/z 296 (M+); Anald. (C18H20N2O2) C 72.95, H 6.80, N 9.45, Found: C 72.93, H 6.82, N 9.44.
Figures S1–S26 (Supplementary Materials) report the 1H and 13C-NMR spectra of all the synthesized compounds.

3.2. ADME Predictions

A computational study to predict the ADME (absorption, distribution, metabolism, excretion) properties of the synthesized compounds (SF1SF13) was performed using SwissADME software [89] to determine drug-likeness properties. We determined the following parameters: molecular weight (MW), molar refractivity logarithm of the partition coefficient (ilog PO/W), Alog P, HBA and HBD to forecast Lipinski’s rule of five (RO5).

3.3. Bioevaluation

3.3.1. Antioxidant Assay

DPPH Assay

The % FRSA (free radical scavenging activity) of the synthetic compounds was determined by using the method of Tai et al. [90] with slight modifications. The antioxidant potential was determined by detecting the capacity of the synthetic compounds to quench the DPPH free radicals. The synthetic compounds were taken in a 96-well plate at three-fold concentrations of 200 µg/mL, 66.6 µg/mL, 22.2 µg/mL and 7.4 µg/mL. DPPH free radicals were added to entire wells to make up the volume up to 200 µL. DMSO and ascorbic acid were taken as negative and positive controls, respectively. The absorbance was taken at 630nm using a microplate reader (ELx800 BioTek). The percentage of scavenging activity was calculated by the given formula:
% scavenging activity = (1 − Abs/Abc) × 100
Abs = sample’s absorbance; Abc = control’s absorbance.

Determination of Total Antioxidant Capacity (TAC)

The total antioxidant capacity was determined by using the phosphomolybdenum-based method [91]. The method is based on the reduction of Mo (VI) to Mo (V), leading to the formation of a green-colored phosphate–molybdenum complex that gives absorption at 630 nm. In 96-well plates, 20 µL of the sample was added with 180 µL TAC reagents in it and incubated at 95 °C for 90 min in a water bath and cooled at room temperature. The results were calculated as µg AAE/mg.

Total Reducing Power (TRP) Determination

The assay is based on the reduction of Fe+3 to Fe+2. The method described by Khan et al. was used with slight modifications. One hundred microliters of test samples (4 mg/mL DMSO) were taken in an Eppendrof tube and 200 μL of phosphate buffer (0.2 mol/L, pH 6.6) and 250 µL of 1% potassium ferricyanide (K3Fe CN)6) were added. This was allowed to incubate at 50 °C for 20 min. Later on, 200 μL of 10% trichloroacetic acid (TCA) was added to the mixture. The mixture was centrifuged at 3000 rpm at room temperature for 10 min. After centrifugation, 150 µL of supernatant were transferred to microplate with FeCl3 (50 μL, 0.1%) and the absorbance was taken at 630 nm [92]. The results were calculated as µg AAE/mg.

3.3.2. α- Amylase Enzyme Inhibitory Assay

An in vitro α-amylase enzyme inhibitory assay was performed by the standard protocol [93]. Four milligrams per 1 mL of sample was prepared in DMSO. In each well of a 96-well plate, synthetic compounds were added to a final concentration of 200 µg/mL. To this, 15 μL phosphate buffer, 25 mL α-amylase enzyme and 40 μL starch solution were added stepwise. The microplate was incubated at 50 °C for 30 min. Twenty microliters of HCl (1 M) and 90 μL iodine solution (0.254 g I2 and 4 g of KI in 1 L of distilled water) were added and the reading was noted at 540 nm. Acarbose and DMSO were taken as positive and negative controls, respectively. The formula used to calculate the percent enzyme inhibition is:
% Enzyme inhibition = [(As − An) ÷ (Ab − An] × 100
As = sample’s absorbance; An = negative’s absorbance; Ab = blank’s absorbance
Synthetic compounds with a % enzyme inhibition ≥ 50% were taken at three-fold concentrations of 200 µg/mL, 66.6 µg/mL, 22.2 µg/mL and 7.4 µg/mL and their IC50 was calculated.

3.3.3. Cytotoxicity Evaluation

Brine Shrimp Cytotoxicity Assay

Brine shrimp eggs were hatched in a shallow rectangular plastic dish and filled with artificial seawater, which was prepared with a commercial salt mixture. In a 96-well plate, 150 µL of artificial seawater was added and 10 nauplii were added into this. Test compounds (400, 200, 100 and 50 µg/mL) were added into each well and the final volume was made up by adding the artificial seawater. The well plate was left uncovered under a lamp. The number of surviving shrimps were counted and recorded after 24 h. The test was repeated in triplicate [94]. Using GraphPad Prism, the lethality concentration (LD50) was assessed at 95% confidence intervals. The percentage mortality (%M) was also calculated by dividing the number of dead nauplii by the total number and then multiplied by 100.

Cytotoxicity Against Raw Macrophages

The cytotoxic potential was assessed by using an MTT 3-(4,5-dimethyl thiazole-2-yl)-2,5- diphenyl tetrazolium bromide assay [95]. Briefly, macrophages were extracted from the peritoneal cavity of albino rats and plated at a density of 1 × 106 per well in a 96-well plate and incubated in a 5% CO2 incubator at 37 °C for 24 h. The cells were treated with various concentrations of the test compounds (100 µM, 50 µM, 10 µM and 1 µM) or vehicle alone. The test compounds were first dissolved in DMSO to make a 100 mM stock solution and then further dilutions were made from this stock solution. After 24 h of incubation, 20 µL of MTT (1 mg/mL in normal saline) was added into each well and incubated under the same conditions for 2 h. Mitochondrial succinate dehydrogenase converted MTT in live cells into purple formazan crystals. The formazan crystals were then solubilized in 100 µL DMSO, and the absorbance was measured at 570 nm. The % cell viability was calculated by the following formula;
% cell viability = (abssample − absblank)/(abscontrol − absblank) × 100

MTT Assay Against Hep-2C Cells

Hep-2C cell line experiments were performed in the Immunology Lab of the National Institute of Health (NIH), Islamabad, Pakistan. Human cervix carcinoma cells Hep-2C (ATCC HB-8065) were cultured in Dulbecco’s modified Eagle’s medium (DMEM), comprising 10% fetal calf serum (10% FCS) and supplemented with 2 mM L-glutamine, 100 U/mL penicillin, 100 µg/mL streptomycin and 1 mM Na pyruvate, at 37 °C in a humidified 5% CO2 incubator.
The tetrazolium dye 3-(4, 5-dimethyl thiazolyl-2)-2, 5-diphenyltetrazolium bromide (MTT) was used to assess the cytotoxic potential of the test compounds. In living cells, MTT is reduced into its insoluble purple product formazan, which is measured spectrophotometrically.
Hep-2C cells, in a density of 1 × 106 (> 90% cell viability), were cultured in a 96-well plate and treated with different concentrations of the test compounds for 24, 48 and 72 h. Ten microliters of MTT (1 mg per mL) were added per well followed by incubation for 4 h. Insoluble formazan crystals were dissolved by adding 100 µL of DMSO. Cells were then incubated for another 2 h. Absorbance was measured at 570 nm by a microplate reader [96]. Untreated Hep-2C cells were taken as controls and DMSO as a negative.
% cell viability = (abssample − absblank)/(abscontrol − absblank) × 100

3.3.4. Nitric Oxide Assay

Isolation of Peritoneal Macrophages and Measurement of Nitrite Production

The anti-inflammatory effect of the synthesized compounds in murine macrophages was evaluated by using the Griess reaction method, described previously [97]. In brief, 1 × 106 were plated in a 96-well plate and incubated in a 5% CO2 incubator at 37 °C for 24 h, pre-treated with different concentrations (100 µM, 50 µM, 10 µM and 1 µM) of synthetic compounds for another 2 h and challenged with LPS (1 µg/mL) for an additional 18 h. Equal volumes of Griess reagent (1% sulphanilamide in 50% phosphoric acid and 0.1% naphthyl ethylenediamine dihydrochloride in distilled water and vortexed) and culture medium were mixed and the absorbance was taken at 540 nm. Lipopolysaccharide (LPS) was taken as a blank and piroxicam was taken as a positive control. A t-test was applied to determine the significance of the results.

In Vivo Anti-Inflammatory Activity

At 7–8 weeks old, male BALB/C albino mice (29–35 g), were purchased from The National Institute of Health (NIH), Islamabad, Pakistan. Five animals were housed per group and placed in a controlled temperature and humidity-controlled room (22 °C and 66 ± 5%, respectively) in a 12 h light–dark cycle and provided with water and food ad libitum. Ethical approval was taken from the bio-ethical committee of Quaid-I-Azam University (Approval No. BEC-FBS-QAU2018-119) and all protocols were conducted following the Ethical Guide to the Code of Practice for the Housing and Care of Animals Act 1986 (ARRIVE (Animal Research: Reporting In Vivo Experiments) guidelines).
The anti-inflammatory activity of the synthesized compounds was evaluated by using the carrageenan-induced rat paw edema assay. To determine the dose response of tetrahydroquinoline derivatives, animals were first given a dose of 0.1, 1 and 10 mg/kg against 100 µL carrageenan (1% solution in normal saline)/paw. Treatment was given 60 min before carrageenan injection. Readings were taken at 4 h post carrageenan injection [98].
The in vivo anti-inflammatory activity was performed following the procedure described by Koksal, et al. [85] in 2017. Paw edema was induced by injecting 100 µL of 1% sterile carrageenan solution into the right hind paw of mice. Synthesized compounds (10 mg/kg), vehicle (saline with 2% DMSO) and drug (acetylsalicylic acid, ASA 10 mg/kg) were administered to the mice orally by gastric intubation 60 min before injecting carrageenan into the right hind paw. The decrease in edema was measured by vernier calipers every 2 h.

3.4. Statistical Analysis

The experimental results were expressed as mean ± SEM. Each test was performed in triplicate. The results were statistically analyzed by t-tests and ANOVA at a 95% confidence level (p < 0.05). GraphPad Prism software was used to calculate the half maximal inhibitory concentration (IC50). A value of p < 0.05 was chosen as the criterion for statistical significance.

4. Conclusions

A series of N-Mannich bases, containing a tetrahydroquinoline nucleus, were synthesized and evaluated biologically for their drug-likeness properties, antioxidant potential, α-amylase enzyme inhibition, cytotoxicity and in vivo anti-inflammatory properties. The results of these assays were found to be encouraging and promising. The compounds exhibited significant scavenging DPPH free radical activity, i.e., SF8 had the lowest IC50 = 29.19 ± 0.29 µg/mL and SF5 was the most potent in inhibiting α-amylase enzyme, i.e., 97.47%. The compounds were successfully evaluated for their cytotoxic activities on the raw macrophages and anti-tumor activity using the cervix carcinoma cell line (Hep-2C). Compounds SF1 and SF7 revealed cytotoxicity to the cervix carcinoma cell line. These results revealed that the cytotoxic potential of new compounds can be a starting point for further surveying them as chemotherapeutic agents in cancer treatment. Moreover, the compounds were further tested on animal models to evaluate their anti-inflammatory behavior. Moderate to potent anti-inflammatory activity of these derivatives containing tetrahydroquinoline was observed, which is comparable with the standard drug, acetylsalicylic acid. As a conclusion, the current study gave us the scope for further work in this area for future developments through derivatization to design more potent compounds. It felt necessary from the results of the current work that there is a need for further advanced studies, at least on the few synthetic compounds which are found to be superior and to produce a rational quantitative structure–activity relationship (QSAR) mapping.

Supplementary Materials

The following are available online at https://www.mdpi.com/1420-3049/25/11/2710/s1, Figure S1–S26 represent the 1H NMR and 13C NMR of the synthetic compounds.

Author Contributions

Funding acquisition, S.F.; methodology, S.F., A.M. and A.G.; resources, I.-U.-H.; supervision, N.U.; writing—original draft, S.F.; writing—review and editing, S.F., N.U. and I.-U.-H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Higher Education Commission (HEC) of Pakistan, grant number, PIN NO. 315-11229-2MD3-036.

Acknowledgments

Samra Farooq was financially supported by the Indigenous Ph.D. scholarship (PIN NO. 315-11229-2MD3-036) granted by the Higher Education Commission (HEC) of Pakistan. We are grateful to the National Institute of Health (NIH) for providing Hep-2C cell lines. The authors acknowledge Prof. Carsten Bolm (RWTH Aachen University, Germany) for the equipment and assistance with the acquisition of the characterization analysis.

Conflicts of Interest

The authors declare that there was no conflict of interest and are responsible for the content and writing of the paper.

Abbreviations

The following abbreviations are used in this manuscript.
THQTetrahydroquinoline
ASAAcetylsalicylic acid
FRSAFree radical scavenging activity
TACTotal antioxidant capacity
TRPTotal reducing power
NONitric oxide
SEMStandard error of the mean
SDStandard deviation
DMSODimethylsulfoxide
KBrPotassium bromide
DMDiabetes mellitus
ROSReactive oxygen species
COX-I/IICyclooxygenase I/II
MCRMulticomponent reaction
CDCl3Deuterated chloroform
ADMEAbsorption, distribution, metabolism, excretion
AAEAscorbic acid equivalent
LDLethal dose
TLCThin-layer chromatography
DCMDichloromethane
SARStructure–activity relationship
RO5Rule of five (Lipinski’s five pints)
MTT3-(4,5-Dimethylthiazol-2-yl)-2,5-diphenyl tetrazolium bromide
DPPH2, 2-Diphenyl-1-picrylhydrazine

References

  1. Patwardhan, B.; Vaidya, A.D.; Chorghade, M.; Joshi, S.P. Reverse pharmacology and systems approaches for drug discovery and development. Curr. Bioact. Compd. 2008, 4, 201–212. [Google Scholar] [CrossRef]
  2. Campos, K.R.; Coleman, P.J.; Alvarez, J.C.; Dreher, S.D.; Garbaccio, R.M.; Terrett, N.K.; Tillyer, R.D.; Truppo, M.D.; Parmee, E.R. The importance of synthetic chemistry in the pharmaceutical industry. Science 2019, 363, eaat0805. [Google Scholar] [CrossRef]
  3. Afsah, E.M.; Fadda, A.A.; Hammouda, M.M. Synthesis and antioxidant activity of 2-indolinone bis (Mannich bases) and related compounds. Mon. Chem. Chem. Mon. 2016, 147, 2009–2016. [Google Scholar] [CrossRef]
  4. Kattappagari, K.K.; Teja, C.R.; Kommalapati, R.K.; Poosarla, C.; Gontu, S.R.; Reddy, B.V.R. Role of antioxidants in facilitating the body functions: A review. J. Orofac. Sci. 2015, 7, 71. [Google Scholar] [CrossRef]
  5. Garcia, E.J.; Oldoni, T.L.C.; Alencar, S.M.d.; Reis, A.; Loguercio, A.D.; Grande, R.H.M. Antioxidant activity by DPPH assay of potential solutions to be applied on bleached teeth. Braz. Dent. J. 2012, 23, 22–27. [Google Scholar] [CrossRef]
  6. Demirci, S.; Demirbaş, N. Anticancer activities of novel Mannich bases against prostate cancer cells. Med. Chem. Res. 2019, 28, 1945–1958. [Google Scholar] [CrossRef]
  7. Aschner, P. Current concepts of diabetes mellitus. Int. Ophthalmol. Clin. 1998, 38, 1–10. [Google Scholar] [CrossRef]
  8. Sõukand, R.; Pieroni, A.; Biró, M.; Dénes, A.; Dogan, Y.; Hajdari, A.; Kalle, R.; Reade, B.; Mustafa, B.; Nedelcheva, A. An ethnobotanical perspective on traditional fermented plant foods and beverages in Eastern Europe. J. Ethnopharmacol. 2015, 170, 284–296. [Google Scholar] [CrossRef]
  9. Etxeberria, U.; de la Garza, A.L.; Campión, J.; Martinez, J.A.; Milagro, F.I. Antidiabetic effects of natural plant extracts via inhibition of carbohydrate hydrolysis enzymes with emphasis on pancreatic alpha amylase. Expert Opin. Ther. Targets 2012, 16, 269–297. [Google Scholar] [CrossRef] [Green Version]
  10. Balan, K.; Ratha, P.; Prakash, G.; Viswanathamurthi, P.; Adisakwattana, S.; Palvannan, T. Evaluation of invitro α-amylase and α-glucosidase inhibitory potential of N2O2 schiff base Zn complex. Arab. J. Chem. 2017, 10, 732–738. [Google Scholar] [CrossRef] [Green Version]
  11. Arora, R.K.; Kaur, N.; Bansal, Y.; Bansal, G. Novel coumarin–benzimidazole derivatives as antioxidants and safer anti-inflammatory agents. Acta Pharm. Sin. B 2014, 4, 368–375. [Google Scholar] [CrossRef] [Green Version]
  12. Madar, J.M.; Shastri, L.A.; Shastri, S.L.; Holiyachi, M.; Naik, N.; Kulkarni, R.; Shaikh, F.; Sungar, V. Design, synthesis, characterization, and biological evaluation of pyrido [1, 2-a] pyrimidinone coumarins as promising anti-inflammatory agents. Synth. Commun. 2018, 48, 375–386. [Google Scholar] [CrossRef]
  13. Yin, L.-L.; Zhang, W.-Y.; Li, M.-H.; Shen, J.-K.; Zhu, X.-Z. CC05, a novel anti-inflammatory compound, exerts its effect by inhibition of cyclooxygenase-2 activity. Eur. J. Pharmacol. 2005, 520, 172–178. [Google Scholar] [CrossRef]
  14. Zhang, X.; Cao, J.; Zhong, L. Hydroxytyrosol inhibits pro-inflammatory cytokines, iNOS, and COX-2 expression in human monocytic cells. Naunyn-Schmiedebergs Arch. Pharmacol. 2009, 379, 581. [Google Scholar] [CrossRef]
  15. Hseu, Y.-C.; Wu, F.-Y.; Wu, J.-J.; Chen, J.-Y.; Chang, W.-H.; Lu, F.-J.; Lai, Y.-C.; Yang, H.-L. Anti-inflammatory potential of Antrodia camphorata through inhibition of iNOS, COX-2 and cytokines via the NF-κB pathway. Int. Immunopharmacol. 2005, 5, 1914–1925. [Google Scholar] [CrossRef]
  16. Sharma, J.; Al-Omran, A.; Parvathy, S. Role of nitric oxide in inflammatory diseases. Inflammopharmacology 2007, 15, 252–259. [Google Scholar] [CrossRef]
  17. Bogdan, C. Nitric oxide and the immune response. Nat. Immunol. 2001, 2, 907–916. [Google Scholar] [CrossRef]
  18. Aktan, F. iNOS-mediated nitric oxide production and its regulation. Life Sci. 2004, 75, 639–653. [Google Scholar] [CrossRef]
  19. Insuasty, D.; Castillo, J.; Becerra, D.; Rojas, H.; Abonia, R. Synthesis of Biologically Active Molecules through Multicomponent Reactions. Molecules 2020, 25, 505. [Google Scholar] [CrossRef] [Green Version]
  20. Cioc, R.C.; Ruijter, E.; Orru, R.V. Multicomponent reactions: Advanced tools for sustainable organic synthesis. Green Chem. 2014, 16, 2958–2975. [Google Scholar] [CrossRef]
  21. Fields, E.K. The synthesis of esters of substituted amino phosphonic acids1a. J. Am. Chem. Soc. 1952, 74, 1528–1531. [Google Scholar] [CrossRef]
  22. Lal, J.; Gupta, S.K.; Thavaselvam, D.; Agarwal, D.D. Synthesis and pharmacological activity evaluation of curcumin derivatives. Chin. Chem. Lett. 2016, 27, 1067–1072. [Google Scholar] [CrossRef]
  23. Dayal, N.; Mikek, C.G.; Hernandez, D.; Naclerio, G.A.; Chu, E.F.Y.; Carter-Cooper, B.A.; Lapidus, R.G.; Sintim, H.O. Potently inhibiting cancer cell migration with novel 3H-pyrazolo [4, 3-f] quinoline boronic acid ROCK inhibitors. Eur. J. Med. Chem. 2019, 180, 449–456. [Google Scholar] [CrossRef]
  24. Tenti, G.; Parada, E.; León, R.; Egea, J.; Martínez-Revelles, S.; Briones, A.M.; Sridharan, V.; López, M.G.; Ramos, M.T.; Menéndez, J.C. New 5-unsubstituted dihydropyridines with improved CaV1. 3 selectivity as potential neuroprotective agents against ischemic injury. J. Med. Chem. 2014, 57, 4313–4323. [Google Scholar] [CrossRef]
  25. Pourshojaei, Y.; Abiri, A.; Eskandari, R.; Dourandish, F.; Eskandari, K.; Asadipo, A. Synthesis, biological evaluation, and computational studies of novel fused six-membered O-containing heterocycles as potential acetylcholinesterase inhibitors. Comput. Biol. Chem. 2019, 80, 249–258. [Google Scholar] [CrossRef]
  26. Lakshmi, N.V.; Thirumurugan, P.; Noorulla, K.; Perumal, P.T. InCl3 mediated one-pot multicomponent synthesis, anti-microbial, antioxidant and anticancer evaluation of 3-pyranyl indole derivatives. Bioorg. Med. Chem. Lett. 2010, 20, 5054–5061. [Google Scholar] [CrossRef]
  27. Murlykina, M.V.; Sakhno, Y.I.; Desenko, S.M.; Konovalova, I.S.; Shishkin, O.V.; Sysoiev, D.A.; Kornet, M.N.; Chebanov, V.A. Features of switchable multicomponent heterocyclizations of salicylic aldehydes and 5-aminopyrazoles with pyruvic acids and antimicrobial activity of the reaction products. Tetrahedron 2013, 69, 9261–9269. [Google Scholar] [CrossRef]
  28. Darandale, S.N.; Pansare, D.N.; Mulla, N.A.; Shinde, D.B. Green synthesis of tetrahydropyrimidine analogues and evaluation of their antimicrobial activity. Bioorg. Med. Chem. Lett. 2013, 23, 2632–2635. [Google Scholar] [CrossRef]
  29. Aouali, M.; Mhalla, D.; Allouche, F.; El Kaim, L.; Tounsi, S.; Trigui, M.; Chabchoub, F. Synthesis, antimicrobial and antioxidant activities of imidazotriazoles and new multicomponent reaction toward 5-amino-1-phenyl [1, 2, 4] triazole derivatives. Med. Chem. Res. 2015, 24, 2732–2741. [Google Scholar] [CrossRef]
  30. Kenchappa, R.; Bodke, Y.D.; Chandrashekar, A.; Telkar, S.; Manjunatha, K.; Sindhe, M.A. Synthesis of some 2, 6-bis (1-coumarin-2-yl)-4-(4-substituted phenyl) pyridine derivatives as potent biological agents. Arab. J. Chem. 2017, 10, S1336–S1344. [Google Scholar] [CrossRef] [Green Version]
  31. Ezzatzadeh, E.; Hossaini, Z.; Varasteh Moradi, A.; Salimifard, M.; Afshari-Sharif Abad, S. Copper iodide and ZnO nanoparticles catalyzed multicomponent synthesis of 1, 3-cyclopentadiene: Study of antioxidant activity. Can. J. Chem. 2019, 97, 270–276. [Google Scholar] [CrossRef]
  32. Kouznetsov, V.V.; Puentes, C.O.; Bohorquez, A.R.; Zacchino, S.A.; Sortino, M.; Gupta, M.; Vazquez, Y.; Bahsas, A.; Amaro-Luis, J. A straightforward synthetic approach to antitumoral pyridinyl substituted 7H-indeno [2, 1-c] quinoline derivatives via three-component imino Diels-Alder reaction. Lett. Org. Chem. 2006, 3, 300–304. [Google Scholar] [CrossRef]
  33. Roopan, S.M.; Khan, F.N.; Jin, J.S.; Kumar, R.S. An efficient one pot–three component cyclocondensation in the synthesis of 2-(2-chloroquinolin-3-yl)-2, 3-dihydroquinazolin-4 (1H)-ones: Potential antitumor agents. Res. Chem. Intermed. 2011, 37, 919. [Google Scholar] [CrossRef]
  34. Nagaraju, B.; Kovvuri, J.; Babu, K.S.; Adiyala, P.R.; Nayak, V.L.; Alarifi, A.; Kamal, A. A facile one pot CC and CN bond formation for the synthesis of spiro-benzodiazepines and their cytotoxicity. Tetrahedron 2017, 73, 6969–6976. [Google Scholar] [CrossRef]
  35. Islam, M.S.; Ghawas, H.M.; El-Senduny, F.F.; Al-Majid, A.M.; Elshaier, Y.A.; Badria, F.A.; Barakat, A. Synthesis of new thiazolo-pyrrolidine–(spirooxindole) tethered to 3-acylindole as anticancer agents. Bioorg. Chem. 2019, 82, 423–430. [Google Scholar] [CrossRef]
  36. Wu, L.; Liu, Y.; Li, Y. Synthesis of Spirooxindole-O-Naphthoquinone-Tetrazolo [1, 5-a] Pyrimidine Hybrids as Potential Anticancer Agents. Molecules 2018, 23, 2330. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Liu, X.-L.; Feng, T.-T.; Jiang, W.-D.; Yang, C.; Tian, M.-Y.; Jiang, Y.; Lin, B.; Zhao, Z.; Zhou, Y. Molecular hybridization-guided 1, 3-dipolar cycloaddition reaction enabled pyrimidine-fused spiropyrrolidine oxindoles synthesis as potential anticancer agents. Tetrahedron Lett. 2016, 57, 4411–4416. [Google Scholar] [CrossRef]
  38. Sudhapriya, N.; Perumal, P.; Balachandran, C.; Ignacimuthu, S.; Sangeetha, M.; Doble, M. Synthesis of new class of spirocarbocycle derivatives by multicomponent domino reaction and their evaluation for antimicrobial, anticancer activity and molecular docking studies. Eur. J. Med. Chem. 2014, 83, 190–207. [Google Scholar] [CrossRef]
  39. Liao, S.-R.; Du, L.-J.; Qin, X.-C.; Xu, L.; Wang, J.-F.; Zhou, X.-F.; Tu, Z.-C.; Li, J.; Liu, Y.-H. Site selective synthesis of cytotoxic 1, 3, 6-trisubstituted 3, 6-diunsaturated (3Z, 6Z)-2, 5-diketopiperazines via a one-pot multicomponent method. Tetrahedron 2016, 72, 1051–1057. [Google Scholar] [CrossRef]
  40. Abadi, A.H.; Hegazy, G.H.; El-Zaher, A.A. Synthesis of novel 4-substituted-7-trifluoromethylquinoline derivatives with nitric oxide releasing properties and their evaluation as analgesic and anti-inflammatory agents. Bioorg. Med. Chem. 2005, 13, 5759–5765. [Google Scholar] [CrossRef]
  41. Liu, J.; Ming, B.; Gong, G.-H.; Wang, D.; Bao, G.-L.; Yu, L.-J. Current research on anti-breast cancer synthetic compounds. RSC Adv. 2018, 8, 4386–4416. [Google Scholar] [CrossRef] [Green Version]
  42. Husain, A.; Maaz, M.; Ahmad Ansari, K.; Ahmad, A.; Rashid, M. Synthesis and microbiological evaluation of mannich bases derived from 4, 6-diacetylresorcinol. J. Chil. Chem. Soc. 2010, 55, 332–334. [Google Scholar] [CrossRef] [Green Version]
  43. Keri, R.S.; Patil, S.A. Quinoline: A promising antitubercular target. Biomed. Pharmacother. 2014, 68, 1161–1175. [Google Scholar] [CrossRef] [PubMed]
  44. Eswaran, S.; Adhikari, A.V.; Shetty, N.S. Synthesis and antimicrobial activities of novel quinoline derivatives carrying 1, 2, 4-triazole moiety. Eur. J. Med. Chem. 2009, 44, 4637–4647. [Google Scholar] [CrossRef] [PubMed]
  45. Castillo, J.-C.; Jiménez, E.; Portilla, J.; Insuasty, B.; Quiroga, J.; Moreno-Fuquen, R.; Kennedy, A.R.; Abonia, R. Application of a catalyst-free Domino Mannich/Friedel-Crafts alkylation reaction for the synthesis of novel tetrahydroquinolines of potential antitumor activity. Tetrahedron 2018, 74, 932–947. [Google Scholar] [CrossRef]
  46. Nandagokula, C.; Poojary, B.; Vittal, S.; Shenoy, S.; Shetty, P.; Tangavelu, A. Synthesis, characterization, and biological evaluation of some N-aryl hydrazones and their 2, 3-disubstituted-4-thiazolidinone derivatives. Med. Chem. Res. 2013, 22, 253–266. [Google Scholar] [CrossRef]
  47. Asolkar, R.N.; Schroeder, D.; Heckmann, R.; Lang, S.; Wagner-Doebler, I.; Laatsch, H. Helquinoline, a new tetrahydroquinoline antibiotic from Janibacter limosus Hel 1. J. Antibiot. 2004, 57, 17–23. [Google Scholar] [CrossRef] [Green Version]
  48. Gudmundsson, K.S.; Boggs, S.D.; Catalano, J.G.; Svolto, A.; Spaltenstein, A.; Thomson, M.; Wheelan, P.; Jenkinson, S. Imidazopyridine-5, 6, 7, 8-tetrahydro-8-quinolinamine derivatives with potent activity against HIV-1. Bioorg. Med. Chem. Lett. 2009, 19, 6399–6403. [Google Scholar] [CrossRef] [PubMed]
  49. Fotie, J.; Kaiser, M.; Delfin, D.A.; Manley, J.; Reid, C.S.; Paris, J.-M.; Wenzler, T.; Maes, L.; Mahasenan, K.V.; Li, C. Antitrypanosomal activity of 1, 2-dihydroquinolin-6-ols and their ester derivatives. J. Med. Chem. 2010, 53, 966–982. [Google Scholar] [CrossRef]
  50. Lukevics, E.; Segal, I.; Zablotskaya, A.; Germane, S. Synthesis and neurotropic activity of novel quinoline derivatives. Molecules 1997, 2, 180–185. [Google Scholar] [CrossRef] [Green Version]
  51. Faber, K.; StÚckler, H.; Kappe, T. Non-steroidal antiinflammatory agents. 1. Synthesis of 4-hydroxy-2-oxo-1, 2-dihydroquinolin-3-yl alkanoic acids by the wittig reaction of quinisatines. J. Heterocycl. Chem. 1984, 21, 1177–1181. [Google Scholar] [CrossRef]
  52. Jarvest, R.L.; Armstrong, S.A.; Berge, J.M.; Brown, P.; Elder, J.S.; Brown, M.J.; Copley, R.C.; Forrest, A.K.; Hamprecht, D.W.; O’Hanlon, P.J. Definition of the heterocyclic pharmacophore of bacterial methionyl tRNA synthetase inhibitors: Potent antibacterially active non-quinolone analogues. Bioorg. Med. Chem. Lett. 2004, 14, 3937–3941. [Google Scholar] [CrossRef]
  53. Bendale, P.; Olepu, S.; Suryadevara, P.K.; Bulbule, V.; Rivas, K.; Nallan, L.; Smart, B.; Yokoyama, K.; Ankala, S.; Pendyala, P.R. Second generation tetrahydroquinoline-based protein farnesyltransferase inhibitors as antimalarials. J. Med. Chem. 2007, 50, 4585–4605. [Google Scholar] [CrossRef] [Green Version]
  54. Gholap, A.R.; Toti, K.S.; Shirazi, F.; Kumari, R.; Bhat, M.K.; Deshpande, M.V.; Srinivasan, K.V. Synthesis and evaluation of antifungal properties of a series of the novel 2-amino-5-oxo-4-phenyl-5, 6, 7, 8-tetrahydroquinoline-3-carbonitrile and its analogues. Bioorg. Med. Chem. 2007, 15, 6705–6715. [Google Scholar] [CrossRef]
  55. Liou, J.-P.; Wu, Z.-Y.; Kuo, C.-C.; Chang, C.-Y.; Lu, P.-Y.; Chen, C.-M.; Hsieh, H.-P.; Chang, J.-Y. Discovery of 4-amino and 4-hydroxy-1-aroylindoles as potent tubulin polymerization inhibitors. J. Med. Chem. 2008, 51, 4351–4355. [Google Scholar] [CrossRef] [PubMed]
  56. Tramontini, M. Advances in the chemistry of Mannich bases. Synthesis 1973, 1973, 703–775. [Google Scholar] [CrossRef]
  57. Malinakova, H.C. Multicomponent Reaction Sequences Using Palladacyclic Complexes. In Palladacycles; Elsevier: Amsterdam, The Netherlands, 2019; pp. 263–295. [Google Scholar]
  58. Paprocki, D.; Madej, A.; Koszelewski, D.; Brodzka, A.; Ostaszewski, R. Multicomponent reactions accelerated by aqueous micelles. Front. Chem. 2018, 6, 502. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Malinakova, H.C. Recent advances in the discovery and design of multicomponent reactions for the generation of small-molecule libraries. Rep. Org. Chem. 2015, 5, 75–90. [Google Scholar] [CrossRef] [Green Version]
  60. Weber, L. The application of multi-component reactions in drug discovery. Curr. Med. Chem. 2002, 9, 2085–2093. [Google Scholar] [CrossRef] [PubMed]
  61. Mannich, C.; Krösche, W. Ueber ein kondensationsprodukt aus formaldehyd, ammoniak und antipyrin. Arch. Pharm. 1912, 250, 647–667. [Google Scholar] [CrossRef] [Green Version]
  62. Blicke, F. The M annich Reaction. Org. React. 2004, 1, 303–341. [Google Scholar]
  63. Lipinski, C.A.; Lombardo, F.; Dominy, B.W.; Feeney, P.J. Experimental and computational approaches to estimate solubility and permeability in drug discovery and development settings. Adv. Drug Deliv. Rev. 1997, 23, 3–25. [Google Scholar] [CrossRef]
  64. Benet, L.Z.; Hosey, C.M.; Ursu, O.; Oprea, T.I. BDDCS, the rule of 5 and drugability. Adv. Drug Deliv. Rev. 2016, 101, 89–98. [Google Scholar] [CrossRef] [PubMed]
  65. Lipinski, C.A. Drug-like properties and the causes of poor solubility and poor permeability. J. Pharmacol. Toxicol. Methods 2000, 44, 235–249. [Google Scholar] [CrossRef]
  66. Rahman, L.; Shinwari, Z.K.; Iqrar, I.; Rahman, L.; Tanveer, F. An assessment on the role of endophytic microbes in the therapeutic potential of Fagonia indica. Ann. Clin. Microbiol. Antimicrob. 2017, 16, 53. [Google Scholar] [CrossRef]
  67. Seelolla, G.; Cheera, P.; Ponneri, V. Synthesis, antimicrobial and antioxidant activities of novel series of cinnamamide derivatives having morpholine moiety. Med. Chem. 2014, 4, 778–783. [Google Scholar] [CrossRef] [Green Version]
  68. Katbab, A.; Scott, G. Mechanisms of antioxidant action: The involvement of nitroxyl radicals in the antifatigue action of secondary amines. Eur. Polym. J. 1981, 17, 559–565. [Google Scholar] [CrossRef]
  69. Mistry, B.; Patel, R.V.; Keum, Y.-S.; Kim, D.H. Synthesis of N-Mannich bases of berberine linking piperazine moieties revealing anticancer and antioxidant effects. Saudi J. Biol. Sci. 2017, 24, 36–44. [Google Scholar] [CrossRef] [Green Version]
  70. Prieto, P.; Pineda, M.; Aguilar, M. Spectrophotometric Quantitation of Antioxidant Capacity through the Formation of a Phosphomolybdenum Complex: Specific Application to the Determination of Vitamin E. Anal. Biochem. 1999, 269, 337–341. [Google Scholar] [CrossRef]
  71. Kanzler, C.; Haase, P.T.; Schestkowa, H.; Kroh, L.W. Antioxidant properties of heterocyclic intermediates of the Maillard reaction and structurally related compounds. J. Agric. Food Chem. 2016, 64, 7829–7837. [Google Scholar] [CrossRef]
  72. Singh, N.; Rajini, P. Free radical scavenging activity of an aqueous extract of potato peel. Food Chem. 2004, 85, 611–616. [Google Scholar] [CrossRef]
  73. Semon, W.L. Antioxidant. U.S. Patent No. 2,166,223, 18 July 1939. [Google Scholar]
  74. Shahidi, F.; Janitha, P.; Wanasundara, P. Phenolic antioxidants. Crit. Rev. Food Sci. Nutr. 1992, 32, 67–103. [Google Scholar] [CrossRef]
  75. El-Gohary, N.; Shaaban, M. Antimicrobial and antiquorum-sensing studies. Part 2: Synthesis, antimicrobial, antiquorum-sensing and cytotoxic activities of new series of fused [1, 3, 4] thiadiazole and [1, 3] benzothiazole derivatives. Med. Chem. Res. 2014, 23, 287–299. [Google Scholar] [CrossRef]
  76. Solis, P.N.; Wright, C.W.; Anderson, M.M.; Gupta, M.P.; Phillipson, J.D. A microwell cytotoxicity assay using Artemia salina (brine shrimp). Planta Med. 1993, 59, 250–252. [Google Scholar] [CrossRef]
  77. Birndorf, H.C.; D’Alessio, J.; Bagshaw, J.C. DNA-dependent RNA-polymerases from Artemia embryos: Characterization of polymerases I and II from nauplius larvae. Dev. Biol. 1975, 45, 34–43. [Google Scholar] [CrossRef]
  78. Sanchez, C.; Gupta, M.; Vasquez, M.; Montenegro, G. Bioessay with brine Artemia to predict antibacterial and pharmacologic activity. Rev. Med. Panama 1993, 18, 62–69. [Google Scholar]
  79. Das, S.; da Silva, C.J.; Silva, M.d.M.; Dantas, M.D.d.A.; de Fátima, Â.; Ruiz, A.L.T.G.; da Silva, C.M.; de Carvalho, J.E.; Santos, J.C.; Figueiredo, I.M. Highly functionalized piperidines: Free radical scavenging, anticancer activity, DNA interaction and correlation with biological activity. J. Adv. Res. 2018, 9, 51–61. [Google Scholar] [CrossRef]
  80. Abdo, N.Y.M.; Kamel, M.M. Synthesis and anticancer evaluation of 1, 3, 4-oxadiazoles, 1, 3, 4-thiadiazoles, 1, 2, 4-triazoles and Mannich bases. Chem. Pharm. Bull. 2015, 63, 369–376. [Google Scholar] [CrossRef] [Green Version]
  81. Wiji Prasetyaningrum, P.; Bahtiar, A.; Hayun, H. Synthesis and cytotoxicity evaluation of novel asymmetrical mono-carbonyl analogs of curcumin (AMACs) against Vero, HeLa, and MCF7 Cell Lines. Sci. Pharm. 2018, 86, 25. [Google Scholar] [CrossRef] [Green Version]
  82. Kumar, G.S.; Prasad, Y.R.; Mallikarjuna, B.; Chandrashekar, S. Synthesis and pharmacological evaluation of clubbed isopropylthiazole derived triazolothiadiazoles, triazolothiadiazines and mannich bases as potential antimicrobial and antitubercular agents. Eur. J. Med. Chem. 2010, 45, 5120–5129. [Google Scholar] [CrossRef] [PubMed]
  83. Dimmock, J.R.; Vashishtha, S.C.; Quail, J.W.; Pugazhenthi, U.; Zimpel, Z.; Sudom, A.M.; Allen, T.M.; Kao, G.Y.; Balzarini, J.; De Clercq, E. 4-(β-Arylvinyl)-3-(β-arylvinylketo)-1-ethyl-4-piperidinols and Related Compounds: A Novel Class of Cytotoxic and Anticancer Agents. J. Med. Chem. 1998, 41, 4012–4020. [Google Scholar] [CrossRef]
  84. Rao, U.M.; Ahmad, B.A.; Mohd, K.S. In vitro nitric oxide scavenging and anti-inflammatory activities of different solvent extracts of various parts of Musa paradisiaca. Malays. J. Anal. Sci. 2016, 20, 1191–1202. [Google Scholar] [CrossRef]
  85. Koksal, M.; Ozkan-Dagliyan, I.; Ozyazici, T.; Kadioglu, B.; Sipahi, H.; Bozkurt, A.; Bilge, S.S. Some Novel Mannich Bases of 5-(3, 4-Dichlorophenyl)-1, 3, 4-oxadiazole-2 (3H)-one and Their Anti-Inflammatory Activity. Arch. Pharm. 2017, 350, 1700153. [Google Scholar] [CrossRef] [PubMed]
  86. Ames, B.N.; Shigenaga, M.K.; Hagen, T.M. Oxidants, antioxidants, and the degenerative diseases of aging. Proc. Natl. Acad. Sci. USA 1993, 90, 7915–7922. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Chu, Y.-F.; Sun, J.; Wu, X.; Liu, R.H. Antioxidant and antiproliferative activities of common vegetables. J. Agric. Food Chem. 2002, 50, 6910–6916. [Google Scholar] [CrossRef] [PubMed]
  88. Simmons, D.L.; Wagner, D.; Westover, K. Nonsteroidal anti-inflammatory drugs, acetaminophen, cyclooxygenase 2, and fever. Clin. Infect. Dis. 2000, 31, S211–S218. [Google Scholar] [CrossRef] [Green Version]
  89. SwissADME Software. Available online: http://swissadme.ch/index.php (accessed on 8 January 2020).
  90. Tai, Z.; Cai, L.; Dai, L.; Dong, L.; Wang, M.; Yang, Y.; Cao, Q.; Ding, Z. Antioxidant activity and chemical constituents of edible flower of Sophora viciifolia. Food Chem. 2011, 126, 1648–1654. [Google Scholar] [CrossRef]
  91. Aguilar Urbano, M.; Pineda Priego, M.; Prieto, P. Spectrophotometric quantitation of antioxidant capacity through the formation of a phosphomolybdenum complex: Specific application to the determination of vitamin E1. Anal. Biochem. 2013, 269, 337–341. [Google Scholar]
  92. Khan, K.; Fatima, H.; Taqi, M.M.; Zia, M.; Mirza, B. Phytochemical and in vitro biological evaluation of Artemisia scoparia Waldst. & Kit for enhanced extraction of commercially significant bioactive compounds. J. Appl. Res. Med. Aromat. Plants 2015, 2, 77–86. [Google Scholar] [CrossRef]
  93. Kim, J.-S.; Kwon, C.-S.; SoN, K.H. Inhibition of alpha-glucosidase and amylase by luteolin, a flavonoid. Biosci. Biotechnol. Biochem. 2000, 64, 2458–2461. [Google Scholar] [CrossRef] [PubMed]
  94. Meyer, B.; Ferrigni, N.; Putnam, J.; Jacobsen, L.; Nichols, D.J.; McLaughlin, J.L. Brine shrimp: A convenient general bioassay for active plant constituents. Planta Med. 1982, 45, 31–34. [Google Scholar] [CrossRef] [PubMed]
  95. Khan, S.; Shin, E.M.; Choi, R.J.; Jung, Y.H.; Kim, J.; Tosun, A.; Kim, Y.S. Suppression of LPS-induced inflammatory and NF-κB responses by anomalin in RAW 264.7 macrophages. J. Cell. Biochem. 2011, 112, 2179–2188. [Google Scholar] [CrossRef]
  96. Zhang, H.; Wang, X.; Mao, J.; Huang, Y.; Xu, W.; Duan, Y.; Zhang, J. Synthesis and biological evaluation of novel benzofuroxan-based pyrrolidine hydroxamates as matrix metalloproteinase inhibitors with nitric oxide releasing activity. Bioorg. Med. Chem. 2018, 26, 4363–4374. [Google Scholar] [CrossRef] [PubMed]
  97. Ahn, K.S.; Noh, E.J.; Zhao, H.L.; Jung, S.H.; Kang, S.S.; Kim, Y.S. Inhibition of inducible nitric oxide synthase and cyclooxygenase II by Platycodon grandiflorum saponins via suppression of nuclear factor-κB activation in RAW 264.7 cells. Life Sci. 2005, 76, 2315–2328. [Google Scholar] [CrossRef] [PubMed]
  98. Zeeshan, S.; Naveed, M.; Khan, A.; Atiq, A.; Arif, M.; Ahmed, M.N.; Kim, Y.S.; Khan, S. N-Pyrazoloyl and N-thiopheneacetyl hydrazone of isatin exhibited potent anti-inflammatory and anti-nociceptive properties through suppression of NF-κB, MAPK and oxidative stress signaling in animal models of inflammation. Inflamm. Res. 2019, 68, 613–632. [Google Scholar] [CrossRef] [PubMed]
Sample Availability: Samples of the compounds are available from the authors.
Figure 1. Selected examples of promising compounds with a quinoline ring system.
Figure 1. Selected examples of promising compounds with a quinoline ring system.
Molecules 25 02710 g001
Figure 2. Selected examples of a promising tetrahydroquinoline ring system containing derivatives and their pharmacological activities.
Figure 2. Selected examples of a promising tetrahydroquinoline ring system containing derivatives and their pharmacological activities.
Molecules 25 02710 g002
Scheme 1. One-pot multicomponent synthesis of 1, 2, 3, 4-tetrahydroquinoline Mannich base derivatives.
Scheme 1. One-pot multicomponent synthesis of 1, 2, 3, 4-tetrahydroquinoline Mannich base derivatives.
Molecules 25 02710 sch001
Scheme 2. Mechanism of DPPH with an antioxidant with a transferable hydrogen radical [67].
Scheme 2. Mechanism of DPPH with an antioxidant with a transferable hydrogen radical [67].
Molecules 25 02710 sch002
Figure 3. Ascorbic acid equivalent (AAE) antioxidant values, expressed as µg/mL. The symbol ### indicates the significance of results where p < 0.01, ## p < 0.05.
Figure 3. Ascorbic acid equivalent (AAE) antioxidant values, expressed as µg/mL. The symbol ### indicates the significance of results where p < 0.01, ## p < 0.05.
Molecules 25 02710 g003
Figure 4. Ascorbic acid calibration curves (y = 0.0198x − 0.1257, R2 = 0.987) for total antioxidant capacity (TAC) and (y = 0.0436x + 1.7841, R2 = 0.9858) for total reducing power (TRP) estimations.
Figure 4. Ascorbic acid calibration curves (y = 0.0198x − 0.1257, R2 = 0.987) for total antioxidant capacity (TAC) and (y = 0.0436x + 1.7841, R2 = 0.9858) for total reducing power (TRP) estimations.
Molecules 25 02710 g004
Figure 5. % inhibition of α-amylase enzyme by a library of synthetic compounds and a reference amylase inhibitor, acarbose (values are expressed as ± SEM, n = 3), ### indicates the significance of results where p < 0.01, ## p < 0.05.
Figure 5. % inhibition of α-amylase enzyme by a library of synthetic compounds and a reference amylase inhibitor, acarbose (values are expressed as ± SEM, n = 3), ### indicates the significance of results where p < 0.01, ## p < 0.05.
Molecules 25 02710 g005
Scheme 3. Conversion of MTT into formazan crystals by mitochondrial reductase enzyme.
Scheme 3. Conversion of MTT into formazan crystals by mitochondrial reductase enzyme.
Molecules 25 02710 sch003
Figure 6. A 3-(4, 5-dimethylthiazol-2-yl)-2, 5-diphenyltetrazolium bromide (MTT) spectrophotometric assay of compounds (SF1, SF4, SF5, SF7, SF8) at 50, 25, 12.5 and 6.25 µM concentration in Hep-2C cells. Data are reported as mean ± SD of each compound tested in triplicate. The symbols ###/## indicates significance at p < 0.01 and p < 0.05, respectively.
Figure 6. A 3-(4, 5-dimethylthiazol-2-yl)-2, 5-diphenyltetrazolium bromide (MTT) spectrophotometric assay of compounds (SF1, SF4, SF5, SF7, SF8) at 50, 25, 12.5 and 6.25 µM concentration in Hep-2C cells. Data are reported as mean ± SD of each compound tested in triplicate. The symbols ###/## indicates significance at p < 0.01 and p < 0.05, respectively.
Molecules 25 02710 g006
Figure 7. Characteristic images of Hep-2C cells before (A) and after (B) MTT treatment. Cells with MTT solution showed dark purple crystal formations inside of the cells.
Figure 7. Characteristic images of Hep-2C cells before (A) and after (B) MTT treatment. Cells with MTT solution showed dark purple crystal formations inside of the cells.
Molecules 25 02710 g007
Figure 8. NaNO2 standard curve used to obtain NO concentration.
Figure 8. NaNO2 standard curve used to obtain NO concentration.
Molecules 25 02710 g008
Figure 9. The dose response of synthesized compounds at 0.1, 1 and 10 mg/kg i.p. carrageenan-induced (100 µL/paw) acute inflammatory model in mice (n = 5). To determine the effect of pretreatment on carrageenan-induced paw edema, the following effects were measured 4 h post carrageenan injection. The data are given as mean ± SEM (standard error of the mean), * p < 0.05, ** p < 0.01, *** p < 0.001 and ### indicates significant differences from the carrageenan-treated group.
Figure 9. The dose response of synthesized compounds at 0.1, 1 and 10 mg/kg i.p. carrageenan-induced (100 µL/paw) acute inflammatory model in mice (n = 5). To determine the effect of pretreatment on carrageenan-induced paw edema, the following effects were measured 4 h post carrageenan injection. The data are given as mean ± SEM (standard error of the mean), * p < 0.05, ** p < 0.01, *** p < 0.001 and ### indicates significant differences from the carrageenan-treated group.
Molecules 25 02710 g009
Figure 10. Structure–activity relationship of tetrahydroquinoline ring derivatized with thirteen amines.
Figure 10. Structure–activity relationship of tetrahydroquinoline ring derivatized with thirteen amines.
Molecules 25 02710 g010
Scheme 4. Synthesis of compounds SF1SF13. Reaction conditions: S (2.0 equiv), 113 (1.0 equiv), F (1.0 equiv); 3 mL solvent, conc. HCl 2 mol %, 80 °C (oil bath). Yields obtained after purification of product.
Scheme 4. Synthesis of compounds SF1SF13. Reaction conditions: S (2.0 equiv), 113 (1.0 equiv), F (1.0 equiv); 3 mL solvent, conc. HCl 2 mol %, 80 °C (oil bath). Yields obtained after purification of product.
Molecules 25 02710 sch004aMolecules 25 02710 sch004b
Table 1. Lipinski’s rule of five for the drug-likeness of all synthesized compounds (SF1SF13).
Table 1. Lipinski’s rule of five for the drug-likeness of all synthesized compounds (SF1SF13).
Comp. NoMW 1
(< 500)
HBA 2
(< 10)
HBD 3
(< 5)
ilog (Po/w) 4
(< 5)
Mlog(Po/w) 5
(< 5)
Lipinski Violation
SF1230103.023.21No
SF2232.32202.692.07No
SF3216.32102.782.95No
SF4246.39103.413.46No
SF5245.36202.942.33No
SF6231.34212.672.07No
SF7314.42003.464.89No
SF8272.77013.083.98No
SF9317.22013.174.10No
SF10280.36102.823.30No
SF11218.34103.112.95No
SF12268.35113.213.10No
SF13296.36212.612.81No
1 MW = molecular weight, 2 HBA = hydrogen bond acceptor, 3 HBD = hydrogen bond donor, 4 ilog (Po/w): octanol–water partition coefficient, 5 Mlog (Po/w): Moriguchi LogP (octanol–water partition).
Table 2. DPPH free radical scavenging activity of synthetic compounds with IC50 values.
Table 2. DPPH free radical scavenging activity of synthetic compounds with IC50 values.
Compound% DPPH Scavenging Activity at Different ConcentrationsIC50 (µg/mL) a
200 µg/mL66.66 µg/mL22.22 µg/mL7.4 µg/mL
SF162.45 ± 0.1250.08 ± 0.1827.76 ± 0.1111.91 ± 0.3244.22 ± 0.25
SF271.85 ± 0.2758.39 ± 0.2736.61 ± 0.2217.44 ± 0.2642.55 ± 0.36
SF327.19 ± 0.3118.32 ± 0.919.85 ± 0.183.76 ± 0.12552.75 ± 0.28
SF458.58 ± 0.3353.65 ± 0.1441.63 ± 0.2818.12 ± 0.2029.79 ± 0.26
SF545.24 ± 0.2828.77 ± 0.1417.39 ± 0.167.76 ± 0.2356.63 ± 0.32
SF666.61 ± 0.1350.32 ± 0.1837.75 ± 0.318.81 ± 0.2135.89 ± 0.33
SF764.48 ± 0.1850.48 ± 0.05426.76 ± 0.179.28 ± 0.1344.4 ± 0.29
SF862.35 ± 0.1452.12 ± 0.3650.17 ± 0.3227.98 ± 0.2429.19 ± 0.25
SF967.46 ± 0.1450.28 ± 0.2219.35 ± 0.117.64 ± 0.2650.76 ± 0.37
SF1049.73 ± 0.1940.77 ± 0.3131.89 ± 0.2123.45 ± 0.2847.81 ± 0.27
SF1143.93 ± 0.2830.9 ± 0.1215.54 ± 0.097.76 ± 0.1853.63 ± 0.13
SF1264.20 ± 0.3154.10 ± 0.1731.80 ± 0.4211.24 ± 0.1139.33 ± 0.28
SF1319.34 ± 0.3411.78 ± 0.15 s4.44 ± 0.181.98 ± 0.2060.78 ± 0.36
Blank0000--
Ascorbic Acid89.01 ± 0.1876.63 ± 0.1952.12 ± 0.3131.35 ± 0.2341.38 ± 0.34
Quercetin81.43 ± 0.2169.81 ± 0.1758.31 ± 0.2842.34 ± 0.3641.64 ± 1.01
a required concentration of the tested compounds to scavenge 50% of the DPPH radicals; values are mean ± SEM (standard error of the mean, n = 3).
Table 3. Cytotoxicity of compounds (SF1SF13) using brine shrimp lethality bioassay.
Table 3. Cytotoxicity of compounds (SF1SF13) using brine shrimp lethality bioassay.
CompoundsLD50
(µg/mL) ± SEM 1
SF1120.8 ± 2.3
SF2141.4 ± 2.1
SF3102 ± 1.98
SF4100 ± 1.78
SF594 ± 1.06
SF6121.2 ± 1.09
SF7100.8 ± 1.14
SF8123.4 ± 1.87
SF9139 ± 0.87
SF10196.1 ± 1.12
SF11200 ± 2.11
SF12128 ± 1.87
SF13204 ± 1.67
Doxorubicin124 ± 1.54
1 SEM = standard error of the mean, n = 3.
Table 4. Cell cytotoxicity of synthetic compounds at different concentrations.
Table 4. Cell cytotoxicity of synthetic compounds at different concentrations.
Compound% Cell Viability at Different ConcentrationsIC50 (µM) a
100 µM50 µM10 µM1 µM
SF144.66 ± 0.547 ± 0.8448 ± 1.3458 ± 0.766.23 ± 0.01 ***
SF250 ± 1.5260 ± 0.9170 ± 0.9176 ± 0.4626.80 ± 0.30
SF341 ± 1.0772 ± 1.3478 ± 0.7483 ± 0.6329.13 ± 0.50
SF439 ± 0.8842 ± 0.8753 ± 0.6379 ± 0.927.20 ± 0.05 ***
SF560 ± 1.2067 ± 0.8474 ± 0.4298 ± 1.016.18 ± 0.05 ***
SF660 ± 1.2165 ± 0.7687 ± 0.8291 ± 0.9923.37 ± 0.17
SF758 ± 0.8384 ± 0.7795 ± 0.9696 ± 0.8918.64 ± 0.13 **
SF852 ± 1.0068 ± 0.7983 ± 0.7897 ± 1.0421.23 ± 0.33 **
SF982 ± 0.9684 ± 1.2191 ± 0.9198 ± 0.8714.24 ± 0.26 **
SF1076 ± 1.2080 ± 1.2495 ± 0.93100 ± 0.8824.82 ± 0.269
SF1141 ± 0.8362 ± 1.2082 ± 0.8495 ± 0.9126.36 ± 0.31
SF1241 ± 0.8657 ± 0.8984 ± 0.8787 ± 0.8235.10 ± 0.50
SF1360 ± 1.2473 ± 1.0182 ± 0.6685 ± 0.9341.43 ± 0.69
Control1001001001000
Doxorubicin17 ± 0.8333 ± 1.0140 ± 0.8751 ± 1.3417.08 ± 0.37 ***
a Values are mean ± SEM (standard error of the mean, n = 3, *** indicates significance at p < 0.01 and ** indicates significance at p < 0.05.
Table 5. IC50 data of synthetic compounds at 24, 48 and 72 h.
Table 5. IC50 data of synthetic compounds at 24, 48 and 72 h.
CompoundIC50 (µM) 1
(Mean ± SEM)
Hep-2C Cells
24 h48 h72 h
SF127.98 ± 1.0721.82 ± 1.114.82 ± 1.07 **
SF423.94 ± 1.0721.15 ± 1.215.53 ± 1.03 **
SF521.3 ± 1.0918.02 ± 1.0516.78 ± 1.07
SF726.68 ± 1.1021.57 ± 1.0516.81 ± 1.03
SF820.23 ± 1.0715.73 ± 1.0811.90 ± 1.04 **
THQ41.27 ± 1.438.54 ± 1.2331.98 ± 1.12
Cisplatin19.12 ± 1.0617.47 ± 1.0714.63 ± 1.01 **
1 Values are calculated as standard error of the mean (SEM), ** indicates significance at p < 0.05.
Table 6. Nitric oxide scavenging activity of synthesized compounds at different concentrations.
Table 6. Nitric oxide scavenging activity of synthesized compounds at different concentrations.
CompoundConcentration (µM) and % NO Production
(Mean ± SEM) 1
10050101
Piroxicam13.351 ± 0.5416.457 ± 0.4120.710 ± 0.2425.35 ± 0.45
SF117.313 ± 0.3523.687 ± 0.4126.0406 ± 0.4327.77 ± 0.46 ***
SF215.345 ± 0.2922.091 ± 0.3327.093 ± 0.3632.23 ± 0.34 ***
SF319.67 ± 0.3723.810 ± 0.3929.571 ± 0.4233.17 ± 0.47 **
SF420.226 ± 0.2832.045 ± 0.3139.172 ± 0.3843.03 ± 0.41
SF521.924 ± 0.3828.644 ± 0.4443.036 ± 0.4551.43 ± 0.48
SF652.437 ± 0.4961.110 ± 0.5472.333 ± 0.5977.30 ± 0.62
SF717.455 ± 0.4521.382 ± 0.4863.122 ± 0.6169.71 ± 0.65
SF815.031 ± 0.3324.111 ± 0.4129.351 ± 0.4439.40 ± 0.48
SF949.073 ± 0.4555.601 ± 0.5158.541 ± 0.5566.97 ± 0.61
SF1033.842 ± 0.3342.487 ± 0.5458.512 ± 0.5962.40 ± 0.65
SF1125.701 ± 0.4433.701 ± 0.5339.799 ± 0.6142.97 ± 0.66
SF1223.203 ± 0.5526.522 ± 0.57832.310 ± 0.5949.70 ± 0.61
SF1314.494 ± 0.4415.572 ± 0.4722.153 ± 0.5225.10 ± 0.59 ***
1 Data are presented as standard error of the mean (SEM), n = 3, ***/** indicate significance at p < 0.01 and p < 0.05, respectively.
Table 7. Effect of acute treatment of synthetic compounds (10 mg/kg) on carrageenan-induced inflammation (n = 5).
Table 7. Effect of acute treatment of synthetic compounds (10 mg/kg) on carrageenan-induced inflammation (n = 5).
CompoundsTime After Carrageenan Injection 1
0 h2 h4 h6 h
Control2.04 ± 0.072.05 ± 0.062.07 ± 0.042.10 ± 0.02
Carrageenan (100 µL)2.11 ± 0.072.57 ± 0.08 ###2.75 ± 0.07 ###2.86 ± 0.05 ###
Acetyl salicylic acid (10 mg/kg)2.14 ± 0.082.26 ± 0.05 ***2.32 ± 0.04 ***2.39 ± 0.02 ***
SF1 (10 mg/kg)2.06 ± 0.072.13 ± 0.03 ***2.18 ± 0.06 ***2.25 ± 0.05 ***
SF2 (10 mg/kg)2.05 ± 0.052.11 ± 0.06 ***2.17 ± 0.03 ***2.23 ± 0.07 ***
SF3 (10 mg/kg)2.05 ± 0.062.16 ± 0.05 ***2.22 ± 0.02 ***2.28 ± 0.07 ***
SF13 (10 mg/kg)2.04 ± 0.072.18 ± 0.03 ***2.26 ± 0.06 ***2.31 ± 0.07 ***
1 Readings were measured every 2 h post carrageenan administration from 0 to 6 h. All values are expressed as mean ± SEM (n = 5). The data are given as mean ± SD, *** p < 0.001 and ### indicate significant differences from the carrageenan-treated group.

Share and Cite

MDPI and ACS Style

Farooq, S.; Mazhar, A.; Ghouri, A.; Ihsan-Ul-Haq; Ullah, N. One-Pot Multicomponent Synthesis and Bioevaluation of Tetrahydroquinoline Derivatives as Potential Antioxidants, α-Amylase Enzyme Inhibitors, Anti-Cancerous and Anti-Inflammatory Agents. Molecules 2020, 25, 2710. https://doi.org/10.3390/molecules25112710

AMA Style

Farooq S, Mazhar A, Ghouri A, Ihsan-Ul-Haq, Ullah N. One-Pot Multicomponent Synthesis and Bioevaluation of Tetrahydroquinoline Derivatives as Potential Antioxidants, α-Amylase Enzyme Inhibitors, Anti-Cancerous and Anti-Inflammatory Agents. Molecules. 2020; 25(11):2710. https://doi.org/10.3390/molecules25112710

Chicago/Turabian Style

Farooq, Samra, Aqsa Mazhar, Areej Ghouri, Ihsan-Ul-Haq, and Naseem Ullah. 2020. "One-Pot Multicomponent Synthesis and Bioevaluation of Tetrahydroquinoline Derivatives as Potential Antioxidants, α-Amylase Enzyme Inhibitors, Anti-Cancerous and Anti-Inflammatory Agents" Molecules 25, no. 11: 2710. https://doi.org/10.3390/molecules25112710

Article Metrics

Back to TopTop