Next Article in Journal
Sedimentation of a Charged Soft Sphere within a Charged Spherical Cavity
Previous Article in Journal
Hierarchical Y Zeolite-Based Catalysts for VGO Cracking: Impact of Carbonaceous Species on Catalyst Acidity and Specific Surface Area
Previous Article in Special Issue
Synthesis and Biological Evaluation of New Compounds with Nitroimidazole Moiety
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and In Silico Analysis of New Polyheterocyclic Molecules Derived from [1,4]-Benzoxazin-3-one and Their Inhibitory Effect against Pancreatic α-Amylase and Intestinal α-Glucosidase

1
Laboratory of Molecular Chemistry, Materials and Catalysis (LCMMC), Faculty of Sciences and Technology, Sultan Moulay Slimane University, P.O. Box 523, Beni-Mellal 23000, Morocco
2
Laboratory of Bioresources, Biotechnology, Ethnopharmacology and Health, Faculty of Sciences, Mohammed First University, Boulevard Mohamed VI, P.O. Box 717, Oujda 60000, Morocco
3
Laboratory of Biology and Health, Faculty of Sciences, Ibn Tofail University, Kenitra 14000, Morocco
4
Oriental Center for Water and Environmental Sciences and Technologies (COSTE), Mohammed Premier University, Oujda 60000, Morocco
5
Laboratory of Organic and Physical Chemistry, Applied Bioorganic Chemistry Team, Faculty of Sciences, Ibnou Zohr University, Agadir 80000, Morocco
6
Biological Engineering Laboratory, Faculty of Sciences and Techniques, Sultan Moulay Slimane University, Beni Mellal 23000, Morocco
7
Laboratoires TBC, Laboratory of Pharmacology, Pharmacokinetics, and Clinical Pharmacy, Faculty of Pharmaceutical and Biological Sciences, P.O. Box 83, F-59000 Lille, France
8
Department of Pharmacognosy, College of Pharmacy, King Saud University, P.O. Box 2457, Riyadh 11451, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Molecules 2024, 29(13), 3086; https://doi.org/10.3390/molecules29133086
Submission received: 27 May 2024 / Revised: 22 June 2024 / Accepted: 24 June 2024 / Published: 28 June 2024

Abstract

:
This study focuses on synthesizing a new series of isoxazolinyl-1,2,3-triazolyl-[1,4]-benzoxazin-3-one derivatives 5a5o. The synthesis method involves a double 1,3-dipolar cycloaddition reaction following a “click chemistry” approach, starting from the respective [1,4]-benzoxazin-3-ones. Additionally, the study aims to evaluate the antidiabetic potential of these newly synthesized compounds through in silico methods. This synthesis approach allows for the combination of three heterocyclic components: [1,4]-benzoxazin-3-one, 1,2,3-triazole, and isoxazoline, known for their diverse biological activities. The synthesis procedure involved a two-step process. Firstly, a 1,3-dipolar cycloaddition reaction was performed involving the propargylic moiety linked to the [1,4]-benzoxazin-3-one and the allylic azide. Secondly, a second cycloaddition reaction was conducted using the product from the first step, containing the allylic part and an oxime. The synthesized compounds were thoroughly characterized using spectroscopic methods, including 1H NMR, 13C NMR, DEPT-135, and IR. This molecular docking method revealed a promising antidiabetic potential of the synthesized compounds, particularly against two key diabetes-related enzymes: pancreatic α-amylase, with the two synthetic molecules 5a and 5o showing the highest affinity values of 9.2 and 9.1 kcal/mol, respectively, and intestinal α-glucosidase, with the two synthetic molecules 5n and 5e showing the highest affinity values of −9.9 and −9.6 kcal/mol, respectively. Indeed, the synthesized compounds have shown significant potential as antidiabetic agents, as indicated by molecular docking studies against the enzymes α-amylase and α-glucosidase. Additionally, ADME analyses have revealed that all the synthetic compounds examined in our study demonstrate high intestinal absorption, meet Lipinski’s criteria, and fall within the required range for oral bioavailability, indicating their potential suitability for oral drug development.

1. Introduction

The benzoxazine structure has been thoroughly investigated in both academic and industrial settings. These compounds demonstrate a wide array of biological activities, indicating that the benzoxazine core could be a valuable scaffold in pharmaceutical research and therapeutic applications, including antifungal [1], antidiabetic [2], antimicrobial [3,4], anticancer [5,6], anti-inflammatory [7], antioxidant [8], antiviral [9], and antiherpetic [10] properties. Additionally, a literature review has identified several 1,4-benzoxazine and [1,4]-benzoxazinone-based compounds (Figure 1) in the developmental stage as potential new medications. For example, antibacterial agent A acts as an inhibitor of bacterial histidine protein kinase [11], the derivative B of [1,4]-benzoxazine appears to be a new neuroprotective agent, effective in a brain injury model [12], the derivative of [1,4]-benzoxazine C demonstrated comparable activity against four human cancer cell lines: MCF-7 (breast), A549 (lung), HeLa (cervical), and PC3 (prostate), when compared to the standard etoposide [13], and the [1,4]-benzoxazine D derivative is a protective agent in tissue culture and in vivo models of neurodegeneration [14].
Heterocyclic compounds containing 1,2,3-triazole have garnered the attention of various researchers due to their facile synthesis through the 1,3-dipolar azide–alkyne cycloaddition reaction with a Cu(I) catalyst [15]. This aromatic five-membered heterocyclic moiety has been extensively explored in medicinal chemistry owing to its stability against metabolic degradation, its high dipole moment, and its resistance to different chemical environments, such as oxidative/reductive conditions and acid/base hydrolysis. The combination of 1,2,3-triazole with other molecules, such as chalcone, opens promising prospects for medicinal applications [16,17,18,19]. Organic compounds containing the 1,2,3-triazole core have been identified in several drugs, demonstrating a diversity of biological activities, including anti-HIV [20], antitumor and antibacterial [21], anti-inflammatory [22], antimalarial [23], antifungal [24,25], anticancer [26,27], antimicrobial [28,29], antioxidant [30,31], and antiviral [32]. The 1,4-disubstituted 1,2,3-triazole displays a bio-isosteric effect because its planarity and length are comparable to those of an amide bond.
Isoxazoline is an innovative heterocycle that is gaining increasing interest due to its diverse pharmacological activities, making it a primary focus for several research groups worldwide. Primarily recognized for its potential in synthesizing new antibacterial agents [33], it also possesses other highly exciting biological activities, such as antidiabetic [34,35], anticancer [36], antioxidant [37], antimicrobial [38], anti-inflammatory [39], antifungal [40], antiviral [41], and anti-Alzheimer properties [42]. The development of novel isoxazoline derivatives continues to be a major focus in medical research.
Diabetes mellitus is a persistent metabolic disorder marked by elevated levels of glucose in the bloodstream resulting from a dysfunction in the production or function of insulin [43]. Diabetes management is based on regulating blood glucose levels as closely as possible to normal physiological levels to prevent the development of chronic diabetic complications such as retinopathy, nephropathy, and neurological and cardiovascular diseases [44]. Among the strategies for treating diabetic patients is the administration of medications endowed with inhibitory effects on the enzymatic activity of the α-amylase and α-glucosidase enzymes [45,46].
Due to the significant pharmaceutical and biological activities observed with isoxazoline, 1,2,3-triazole, and [1,4]-benzoxazin-3-one, various approaches have been developed to access these molecular structures: [1,4]-benzoxazin-3-one [47,48], 1,2,3-triazole [49], and isoxazoline [50]. On our part, we carried out the first step of the 1,3-dipolar cycloaddition under catalytic conditions to selectively synthesize the 1,2,3-triazole 1,4-disubstituted compound with a good yield, following the method described in the literature [51,52]. Subsequently, we improved the yield of the second step of the cycloaddition [53]. We devised a novel method for synthesizing these [1,4]-benzoxazin-3-one derivatives, achieving satisfactory yields. This endeavor aims to enhance various biological activities such as antidiabetic, anticancer, antioxidant, antiviral, anti-inflammatory, antimicrobial, antifungal, and more. Each of the three components of these newly synthesized molecules, namely [1,4]-benzoxazin-3-one, 1,2,3-triazole, and isoxazoline, exhibit these activities individually. Consequently, the contribution of each of these compounds through its biological activity could eventually lead to very interesting biological activities. After synthesizing these active molecules, we tested their effect on the enzymatic activity of the two enzymes involved in carbohydrate digestion, namely α-amylase and α-glycosidase, in silico.

2. Results

2.1. Chemical Synthesis of Isoxazolinyl-1,2,3-triazolyl-[1,4]-benzoxazin-3-one Derivatives 5a5o

To generate novel heterocyclic systems, we present the synthesis of the isoxazolinyl-1,2,3-triazolyl-[1,4]-benzoxazin-3-one derivatives 5a5o. This was achieved through a double 1,3-dipolar cycloaddition reaction employing a “click chemistry” strategy, starting from the respective [1,4]-benzoxazin-3-ones [51,52]. To synthesize the 1,2,3-triazole motif 3, the 1,3-dipolar cycloaddition was carried out with good yield using the product 4-(prop-2-yn-1-yl)-2H-[1,4]-benzoxazin-3-one 1, which contains the propargylic moiety as the dipolarophile, and the allylic azide 2 synthesized through nucleophilic substitution of allyl bromide with sodium azide as the 1,3-dipole. This method is widely employed for obtaining allylic azides [54]. The 1,3-dipolar cycloaddition reaction was conducted at room temperature, under catalytic conditions, in the presence of ( CuSO 4 · 5 H 2 O and sodium ascorbate) as the catalyst. For the last step of this work, we were mainly interested in the reactivity of nitriloxides concerning exocyclic carbon-carbon double bonds by carrying out a second 1,3-dipolar cycloaddition, this time between the 4-[(1 -allyl-1H-1,2,3-triazol-4-yl)methyl]-2H-[1,4]-benzoxazine-3-one 3 prepared in the previous step, containing the allylic part and various oximes 4 (4-methylphenyloxime; 2-nitrophenyloxime; 3-nitrophenyloxime; 4-nitrophenyloxime; 4-chlorophenyloxime; benzyloxime; furan-2-yloxime; styryloxime; pyridin-2-yloxime; 1-methyl-1H-pyrrol-2-yloxime; 4-bromophenyloxime; 4-dimethylaminophenyloxime; 4-methoxyphenyloxime; 3-methoxyphenyloxime; 4-fluorophenyloxime) prepared by condensation of different aldehydes and hydroxylamine, following the process described in the literature [53]. However, the dehydrohalogenation of the various oximes by 24° chlorometric bleach generates oximes 4, which react with the dipolarophile 3 in a two-phase medium (water/chloroform) at a temperature that varies between −5 and 0 °C for 4 h, to lead, respectively, to the cycloadducts 5a5o with good yield.
All the synthesized compounds were characterized utilizing proton nuclear magnetic resonance (1H NMR) and carbon-13 nuclear magnetic resonance (13C NMR) spectroscopy (see the experimental section) (Table 1).
We see that these 1,3-dipole cycloaddition reactions are totally regioselective because the direction of attack of the dipole is unique (Figure 2). This reaction led to the synthesis of isoxazoline, which constitutes the third heterocyclic component and also exhibits very interesting biological activities.

2.2. Computational Analysis Using Molecular Docking

Molecular docking, an advanced computational technique, is frequently utilized to offer valuable insights into the molecular mechanisms of pharmacologically active substances. In this study, molecular docking was employed to unveil the potential mechanism of action associated with the fifteen synthetic molecules’ pancreatic α-amylase and intestinal α-glucosidase activities.

2.2.1. In Silico, Inhibitory Activity of Synthetic Molecules on α-Amylase Activity

The provided data, which comprise binding affinity values, imply that the molecule under study potentially exhibits either a heightened or diminished affinity toward the specified target in comparison with a native ligand, namely acarbose, if a decrease in binding energy correlates with an increase in compound affinity (Table 2). The active sites of pancreatic α-amylase predominantly feature amino acid residues such as Glu A:233 and Asp A:197, A:300, alongside pivotal residues like Arg A:195 and A:337; Trp A:58, A:284, A:203, and A:59; His A:101, A:201, and A:299; Phe A:298, A:265, and A:295; Asn A:298; Gly A:306; Ala A:307; and Tyr A:62 [55,56,57].
Within this context, it is observed that all the examined molecules, except 5n, exhibit lower free binding energy values compared with the native ligand, suggesting potent inhibitory potential. Molecules 5a and 5o demonstrate the lowest free binding energy values, standing at 9.2 and 9.1 kcal/mol, respectively. Notably, both 5a and 5o establish electrostatic bonds with amino acid residues surrounding the protein’s active site, primarily in the forms of Pi-sigma, Pi-Pi stacked, and Pi-alkyl interactions (Figure 3). Nevertheless, it is noteworthy that 5o additionally forms a conventional hydrogen bond with the amino acid residue His A:202. The findings of the computational analysis indicate that the observed antihyperglycemic effects and inhibition of pancreatic α-amylase can be ascribed to these molecules.

2.2.2. In Silico, Inhibitory Activity of Synthetic Molecules on α-Glucosidase Activity

The data presented, in the form of binding affinity values, suggest that the molecule under investigation may exhibit either a heightened or diminished affinity for the specified target in comparison with the native ligand (acarbose), assuming a decrease in binding energy correlates with an increase in the compound’s affinity (Table 3). The active sites of α-glucosidase are primarily surrounded by the amino acid residues Trp A:376, Asp A:404, Leu A:405, Ile A:441, Trp A:481, Asp A:518, Met A:519, Arg A:600, Trp A:613, Asp A:616, Phe A:649, and His A:674 [58].
Our observations within this framework indicate that all examined molecules exhibit a significant free binding energy compared with the native ligand, ranging from −8.8 to −9.6 kcal/mol (Table 2). Specifically, compounds 5e and 5n display the lowest values of free binding energy, at −9.9 and −9.6 kcal/mol, respectively. It is noteworthy that these molecules establish hydrogen bonds (interactions between a hydrogen atom bonded to an electronegative atom and a neighboring electronegative atom) and electrostatic bonds (interactions between oppositely charged entities) with the amino acid residues surrounding the protein’s active site, primarily in the forms of conventional hydrogen bonds, Pi-sigma interactions (bonds between a pi electron and a sigma atom), Pi-Pi stacked interactions (interactions between pi systems), and Pi-alkyl interactions (interactions between a pi system and alkyl groups). Specifically, compound 5n forms four conventional hydrogen bonds with the amino acid residues Tyr A:360, Met A:363, Arg A:608, and Glu A:866 (Figure 4A), while compound 5e forms five hydrogen bonds with Tyr A:360, Met A:363, His A:584, Arg A:608, and Glu A:866 (Figure 4B) from the active site of α-glucosidase.
The computational findings suggest that these molecules may contribute to the observed antihyperglycemic effects and the inhibition of pancreatic α-glucosidase, underscoring the significance of hydrogen bonding interactions in modulating enzymatic activity.

2.3. ADME Analysis

In silico ADME studies are essential to advancing pharmaceutical development by offering a cost-effective method for predicting how a drug will act within the body [59]. Utilizing computer models for early pharmacokinetic assessment, these studies enable the swift selection of drug candidates and streamline development processes. They help minimize the risk of adverse side effects, decrease the likelihood of drug development failures, and increase the chances of clinical success, making them an invaluable tool in modern drug discovery and development [60]. In order to comply with Lipinski’s rule of five and Veber’s rule, compounds deemed suitable for oral drug development should typically not exceed one violation of the following criteria: (1) no more than 10 hydrogen bond acceptors (nitrogen or oxygen atoms), (2) an octanol-water partition coefficient log P (MLogP) < 5, (3) a molecular mass < 500 daltons, and (4) no more than 5 hydrogen bond donors [61]. In our study, we observed that all the compounds examined met Lipinski’s criteria, indicating their potential suitability for oral drug development.
The blood–brain barrier (BBB) serves as a crucial barrier between the systemic circulation and the central nervous system, protecting the brain through both biochemical processes, like enzyme reactions, and physical mechanisms such as active expulsion systems [62]. Our investigation revealed that synthetic molecules are unable to cross this barrier, because they have a TPSA > 79 Å2, as shown in Table 4 and illustrated in the yellow portion of Figure 5. No molecules disperse in this yellow area. Additionally, most of the compounds analyzed were determined to be P-glycoprotein non-substrates (PGP-), except for 5f, 5l, 5j, 5n, 5i, and 5d, which were determined to be PGP+ (Figure 4). A molecule that is strongly absorbed by the intestine offers significant advantages in terms of bioavailability, efficacy, convenience, and tolerance, making it a promising candidate for the development of oral medications [63]. Notably, all the synthetic compounds examined in our study demonstrate high intestinal absorption. Cytochrome P450 is a crucial enzyme for detoxification, primarily located in the liver [64]. Our analysis identified that the majority of the compounds are neither inhibitors nor substrates of CYP450 enzymes, particularly CYP1A2, except 5i, 5m, 5g, 5a, 5e, and 5k, which were identified as CYP1A2 inhibitors (Table 3). This finding implies a lower likelihood of medication metabolism disruption, thereby strengthening the safety profile of the synthetic compounds. Figure 6 displays the bioavailability profiles of these drugs. The pink zone in these radar graphs corresponds to the oral bioavailability space. A chemical’s characteristics must entirely fall within this specified region to qualify as drug-like [65]. In the present study, all the synthetic compounds meet the required range for oral bioavailability, suggesting their potential as drug candidates.

3. Materials and Methods

3.1. General

Merck-60 silica gel (230–400 mesh E) was employed for column chromatography. Melting points for compounds 3 and 5a5o were measured using a Kofler bench (FST, Beni Mellal, Morocco). Reaction progress was tracked with thin-layer chromatography (TLC) on aluminum plates coated with silica gel 60 F254 (E. Merck). Nuclear magnetic resonance (NMR) spectra were obtained on a Varian Unity Plus spectrometer (CNRST, Rabat, Morocco)at 500 MHz for 1H NMR and at 125.76 MHz for 13C NMR. Chemical shifts were given in parts per million (ppm), with coupling constants (J) noted in Hertz (Hz). The signals were characterized as s (singlet), d (doublet), t (triplet), and m (multiplet), and tetramethylsilane Si CH 3 4 was used as the reference.

3.2. Procedure for the Preparation of Compound 3 by “Click Chemistry” (CuAAC)

A solution of 1 mmol of compound 1 and 2 mmol of 3-azidoprop-1-ene 2 in 8 mL of methanol was prepared, to which 1 mmol of sodium ascorbate and 1 mmol of CuSO4 · 5 H 2 O dissolved in 7 mL of distilled water were added. The reaction mixture was stirred at room temperature for 3 h and monitored using TLC. Following filtration and concentration under decreased pressure, the resultant substance was then submitted to column chromatography on silica gel, utilizing a mixture of ethyl acetate and hexane at a ratio of 3 to 7 as the eluting solvent. Compound 3 was obtained with a good yield of 81%.
4-[(1-Allyl-1H-1,2,3-triazol-4-yl)methyl]-2H-[1,4]-benzoxazin-3-one (3): Brown oil, yield (81%); 1H NMR (500 MHz, DMSO-d6): δ 7.95 (s, 1H), 6.95–7.29 (m, 4H), 5.92–6.01 (m, 1H), 5.20, 5.08 (dd, 2H, J = 10, 17 Hz), 5.11 (s, 2H), 4.94 (d, 2H, J = 6 Hz), 4.66 (s, 2H). 13C NMR (125 MHz, DMSO-d6): δ 164.62, 145.35, 128.95, 128.90, 124.21, 124.12, 133.16, 123.20, 119.31, 117.06, 116.31, 67.67, 52.23, 36.7 (See Figures S1–S3 (Supplementary Materials: spectrum 1H, 13C NMR, and DEPT-135 for 3)).

3.3. General Procedure for the Synthesis of Compounds 5a5o

To a solution of 4-[(1-allyl-1H-1,2,3-triazol-4-yl)methyl]-2H-[1,4]-benzoxazin-3-one 3 (1 mmol) in 8 mL of chloroform, 3 equivalents of oximes: (4-methylphenyloxime; 2-nitrophenyloxime; 3-nitrophenyloxime; 4-nitrophenyloxime; 4-chlorophenyloxime; benzyloxime; furan-2-yloxime; styryloxime; pyridin-2-yloxime; 1-methyl-1H-pyrrol-2-yloxime; 4-bromophenyloxime; 4-dimethylaminophenyloxime; 4-methoxyphenyloxime; 3-methoxyphenyloxime; 4-fluorophenyloxime) were added with vigorous stirring. The mixture was brought to a temperature between −5 and 0 °C, and then 15 mL of NaOCl sodium hypochlorite (bleach 24°) was added dropwise. The progress of the reaction was checked using thin-layer chromatography (TLC). After 9 h of stirring, the organic layer was separated and dried with Na 2 SO 4 . Subsequently, the solvent was evaporated using reduced pressure. The obtained residue was purified using column chromatography on silica gel utilizing a gradient of hexane and ethyl acetate.
4-[(1-[(3-(4-Methylphenyl)-4,5-dihydroisoxazolin-5-yl)methyl]-1H-1,2,3-triazol-4-yl)methyl]-2H-[1,4]-benzoxazin-3-one (5a): Colorless solid, yield (82%); m.p. 191–193 °C; IR (KBr, vmax/cm−1): (C=O lactame str.). 1H NMR (500 MHz, DMSO-d6): δ 8.02 (s, 1H), 6.96–7.45 (m, 8H), 5.10 (s, 2H), 5.02–5.07 (m, 1H), 4.65 (s, 2H), 4.52 (dd, 2H, J = 6.5, 14.5 Hz), 3.50, 3.20 (dd, 2H, J = 6.5, 17 Hz), 2.30 (s, 3H). 13C NMR (125 MHz, DMSO-d6): δ 164.66, 157.23, 145.37, 142.88, 140.54, 128.92, 126.68, 129.86, 127.10, 125.03, 124.18, 123.15, 117.09, 116.27, 79.22, 67.50, 52.84, 37.99, 36.97, 21.61. Elemental analysis calculated (%) for C22H21N5O3·1/10 H2O: C 65.21, H 5.27, N 17.28, O 12.24; found C 65.50, H 5.25, N 17.36, O 11.90 (See Figures S4–S7 (Supplementary Materials: spectrum 1H, 13C NMR, DEPT-135, and IR for 5a)).
4-[(1-[(3-(2-Nitrophenyl)-4,5-dihydroisoxazolin-5-yl)methyl]-1H-1,2,3-triazol-4-yl)methyl]-2H-[1,4]-benzoxazin-3-one (5b): Colorless solid, yield (81%); m.p. 152–154 °C; 1H NMR (500 MHz, DMSO-d6): δ 8.04 (s, 1H), 6.96–7.99 (m, 8H), 5.15–5.20 (m, 1H), 5.12 (s, 2H), 4.66 (s, 2H), 4.56 (dd, 2H, J = 6.5, 14.5 Hz), 3.52, 3.16 (dd, 2H, J = 6.5, 17.5 Hz). 13C NMR (125 MHz, DMSO-d6): δ 164.64, 155.27, 148.41, 145.35, 142.94, 128.96, 123.45, 133.73, 131.77, 131.11, 125.08, 124.85, 124.17, 123.16, 117.04, 116.31, 79.63, 67.61, 52.72, 38.82, 36.76. Elemental analysis calculated (%) for C21H18N6O5·1/10H2O·3/20C6H14: C 58.56, H 4.56, N 18.71, O 18.17; found C 58.06, H 4.18, N 19.35, O 18.42 (See Figures S8–S10 (Supplementary Materials: spectrum 1H, 13C NMR, and DEPT-135 for 5b)).
4-[(1-[(3-(3-Nitrophenyl)-4,5-dihydroisoxazolin-5-yl)methyl]-1H-1,2,3-triazol-4-yl)methyl]-2H-[1,4]-benzoxazin-3-one (5c): Pale yellow solid, yield (83%); m.p. 196–198 °C; 1H NMR (500 MHz, DMSO-d6): δ 8.03 (s, 1H), 6.92–8.28 (m, 8H), 5.16–5.20 (m, 1H), 5.09 (s, 2H), 4.64 (s, 2H), 4.58 (dd, 2H, J = 6.5, 14.5 Hz), 3.63, 3.29 (dd, 2H, J = 6.5, 17 Hz). 13C NMR (125 MHz, DMSO-d6): δ 164.61, 156.20, 148.45, 145.54, 142.93, 128.86, 123.06, 133.35, 130.97, 125.20, 125.07, 124.17, 123.16, 121.42, 116.93, 116.25, 79.96, 67.59, 52.80, 37.59, 36.74. Elemental analysis calculated (%) for C21H18N6O5·1/10H2O·1/5C6H14: C 58.80, H 4.67, N 18.53, O 18.00; found C 58.06, H 4.18, N 19.35, O 18.42 (See Figures S11–S13 (Supplementary Materials: spectrum 1H, 13C NMR, and DEPT-135 for 5c)).
4-[(1-[(3-(4-Nitrophenyl)-4,5-dihydroisoxazolin-5-yl)methyl]-1H-1,2,3-triazol-4-yl)methyl]-2H-[1,4]-benzoxazin-3-one (5d): Colorless solid, yield (82%); m.p. 262–264 °C; IR (KBr, vmax/cm−1): 1678 (C=O lactame str.). 1H NMR (500 MHz, DMSO-d6): δ 8.02 (s, 1H), 6.96–8.20 (m, 8H), 5.17–5.22 (m, 1H), 5.08 (s, 2H), 4.64 (s, 2H), 4.60 (dd, 2H, J = 6.5, 14 Hz), 3.57, 3.29 (dd, 2H, J = 6.5, 17 Hz). 13C NMR (125 MHz, DMSO-d6): δ 164.64, 156.41, 145.19, 142.83, 135.26, 128.89, 128.3, 129.40, 128.83, 125.07, 124.04, 123.07, 116.94, 116.14, 79.46, 67.50, 52.71, 37.81, 36.66. Elemental analysis calculated (%) for C21H18N6O5·1/10CH3CO2C2H5: C 57.99, H 4.28, N 18.96, O 18.77; found C 58.06, H 4.18, N 19.35, O 18.42 (See Figures S14–S17 (Supplementary Materials: spectrum 1H, 13C NMR, DEPT-135, and IR for 6d)).
4-[(1-[(3-(4-Chlorophenyl)-4,5-dihydroisoxazolin-5-yl)methyl]-1H-1,2,3-triazol-4-yl)methyl]-2H-[1,4]-benzoxazin-3-one (5e): White powder, yield (83%); m.p. 172–174 °C; 1H NMR (500 MHz, DMSO-d6): δ 8.01 (s, 1H), 6.93–7.57 (m, 8H), 5.10 (s, 2H), 5.10–5.14 (m, 1H), 4.64 (s, 2H), 4.54 (dd, 2H, J = 6.5, 14.5 Hz), 3.52, 3.20 (dd, 2H, J = 6.5, 17 Hz). 13C NMR (125 MHz, DMSO-d6): δ 164.66, 156.50, 145.39, 142.89, 135.34, 128.94, 128.34, 129.40, 128.83, 125.09, 124.22, 123.15, 117.03, 116.26, 79.46, 67.60, 52.78, 37.81, 36.70. Elemental analysis calculated (%) for C21H18ClN5O3·1/10H2O: C 59.26, H 4.31, N 16.45, O 11.65; found C 59.51, H 4.28, N 16.52, O 11.32 (See Figures S18–S20 (Supplementary Materials: spectrum 1H, 13C NMR, and DEPT-135 for 5e)).
4-[(1-[(3-Benzyl-4,5-dihydroisoxazolin-5-yl)methyl]-1H-1,2,3-triazol-4-yl)methyl]-2H-[1,4]-benzoxazin-3-one (5f): Colorless solid, yield (78%); m.p. 142–144 °C; 1H NMR (500 MHz, DMSO-d6): δ 7.94 (s, 1H), 6.95–7.29 (m, 9H), 5.11 (s, 2H), 4.82–4.87 (m, 1H), 4.66 (s, 2H), 4.38 (dd, 2H, J = 6.5, 14 Hz), 3.52 (s, 2H), 3.35, 2.95 (dd, 2H, J = 6.5, 17.5 Hz). 13C NMR (125 MHz, DMSO-d6): δ 164.68, 158.64, 145.39, 142.94, 136.36, 128.84, 129.89, 129.16, 127.35, 124.95, 124.25, 123.17, 117.03, 116.32, 78.12, 60.90, 67.62, 52.72, 38.96, 36.80. Elemental analysis calculated (%) for C22H21N5O3·1/10H2O·1/20C6H14: C 65.40, H 5.39, N 17.10, O 12.11; found C 65.50, H 5.25, N 17.36, O 11.90 (See Figures S21–S23 (Supplementary Materials: spectrum 1H, 13C NMR, and DEPT-135 for 5f)).
4-[(1-[(3-(Furan-2-yl)-4,5-dihydroisoxazolin-5-yl)methyl]-1H-1,2,3-triazol-4-yl)methyl]-2H-[1,4]-benzoxazin-3-one (5g): Brown oil, yield (80%); IR (KBr, vmax/cm−1): 1682 (C=O lactame str.). 1H NMR (500 MHz, DMSO-d6): δ 8.00 (s, 1H), 6.59–7.26 (m, 7H), 5.11 (s, 2H), 5.00–5.06 (m, 1H), 4.66 (s, 2H), 4.53 (dd, 2H, J = 6.5, 14 Hz), 3.45, 3.13 (dd, 2H, J = 6.5, 17.5 Hz). 13C NMR (125 MHz, DMSO-d6): δ 164.65, 149.17, 145.37, 144.39, 142.90, 128.92, 145.95, 125.01, 124.12, 123.16, 117.03, 116.30, 113.89, 112.52, 78.87, 67.82, 52.73, 38.14, 36.71. Elemental analysis calculated (%) for C22H21N5O3·1/12H2O: C 59.92, H 4.54, N 18.39, O 17.15; found C 60.15, H 4.52, N 18.46, O 16.87 (See Figures S24–S27 (Supplementary Materials: spectrum 1H, 13C NMR, DEPT-135, and IR for 5g)).
4-[(1-[(3-Styryl-4,5-dihydroisoxazolin-5-yl)methyl]-1H-1,2,3-triazol-4-yl)methyl]-2H-[1,4]-benzoxazin-3-one (5h): Pale yellow solid, yield (82%); m.p. 174–176 °C; 1H NMR (500 MHz, DMSO-d6): δ 8.01 (s, 1H), 6.92–7.57 (m, 9H), 6.89, 6.98 (d, 2H, J = 16.5 Hz), 5.11 (s, 2H), 5.00–5.05 (m, 1H), 4.66 (s, 2H), 4.52 (dd, 2H, J = 6.5, 14.5 Hz), 3.34, 3.05 (dd, 2H, J = 6.5, 17.5 Hz). 13C NMR (125 MHz, DMSO-d6): δ 164.68, 158.38, 145.40, 142.90, 136.13, 128.62, 137.69, 129.35, 128.88, 127.68, 124.99, 124.16, 123.14, 117.58, 116.93, 116.26, 79.12, 67.61, 52.87, 39.08, 36.80. Elemental analysis calculated (%) for C23H21N5O3·1/11H2O: C 66.23, H 5.12, N 16.79, O 11.86; found C 66.49, H 5.09, N 16.86, O 11.55 (See Figures S28–S30 (Supplementary Materials: spectrum 1H, 13C NMR, and DEPT-135 for 5h)).
4-[(1-[(3-(Pyridin-2-yl)-4,5-dihydroisoxazolin-5-yl)methyl]-1H-1,2,3-triazol-4-yl)methyl]-2H-[1,4]-benzoxazin-3-one (5i): Yellow oil, yield (79%); IR (KBr, vmax/cm−1): 1672 (C=O lactame str.). 1H NMR (500 MHz, DMSO-d6): δ 8.01 (s, 1H), 6.93–7.81 (m, 8H), 5.15 (s, 2H), 5.08–5.13 (m, 1H), 4.64 (s, 2H), 4.57 (dd, 2H, J = 6.5, 14.5 Hz), 3.54, 3.20 (dd, 2H, J = 6.5, 17 Hz). 13C NMR (125 MHz, DMSO-d6): δ 164.65, 149.17, 145.31, 144.37, 142.71, 128.75, 145.75, 125.01, 123.88, 122.69, 117.03, 116.21, 114.54, 113.77, 112.38, 79.02, 67.62, 52.60, 38.36, 36.55. Elemental analysis calculated (%) for C20H18N6O3·1/10H2O·1/20C6H14: C 61.49, H 4.80, N 21.20, O 12.51; found C 61.53, H 4.65, N 21.53, O 12.29 (See Figures S31–S34 (Supplementary Materials: spectrum 1H, 13C NMR, DEPT-135, and IR for 5i)).
4-[(1-[(3-(1-Methyl-1H-pyrrol-2-yl)-4,5-dihydroisoxazolin-5-yl)methyl]-1H-1,2,3-triazol-4-yl)methyl]-2H-[1,4]-benzoxazin-3-one (5j): Colorless solid, yield (78%); m.p. 170–172 °C; IR (KBr, vmax/cm−1): 1678 (C=O lactame str.). 1H NMR (500 MHz, DMSO-d6): δ 8.00 (s, 1H), 6.03–7.26 (m, 7H), 5.11 (s, 2H), 4.89–4.94 (m, 1H), 4.65 (s, 2H), 4.48 (dd, 2H, J = 6.5, 14.5 Hz), 3.69 (s, 3H), 3.46, 3.14 (dd, 2H, J = 6.5, 17.5 Hz). 13C NMR (125 MHz, DMSO-d6): δ 164.86, 151.18, 145.39, 142.89, 128.97, 122.03, 128.71, 124.90, 124.20, 123.15, 117.04, 116.28, 115.11, 108.42, 77.03, 56.16, 67.59, 52.06, 38.35, 36.59. Elemental analysis calculated (%) for C20H20N6O3·1/10H2O: C 60.94, H 5.16, N 21.32, O 12.58; found C 61.21, H 5.14, N 21.42, O 12.23 (See Figures S35–S38 (Supplementary Materials: spectrum 1H, 13C NMR, DEPT-135, and IR for 5j)).
4-[(1-[(3-(4-Bromophenyl)-4,5-dihydroisoxazolin-5-yl)methyl]-1H-1,2,3-triazol-4-yl)methyl]-2H-[1,4]-benzoxazin-3-one (5k): White powder, yield (81%); m.p. 183–185 °C; 1H NMR (500 MHz, DMSO-d6): δ 8.01 (s, 1H), 6.93–7.51 (m, 8H), 5.10 (s, 2H), 4.98–5.03 (m, 1H), 4.65 (s, 2H), 4.51 (dd, 2H, J = 6.5, 14 Hz), 3.49, 3.15 (dd, 2H, J = 6.5, 17 Hz). 13C NMR (125 MHz, DMSO-d6): δ 164.62, 156.60, 145.29, 142.80, 132.28, 129.10, 128.66, 132.33, 129.04, 125.06, 124.19, 123.15, 117.04, 116.26, 79.47, 67.60, 52.78, 37.81, 36.69. Elemental analysis calculated (%) for C21H18BrN5O3·1/20H2O·1/20C6H14: C 54.03, H 4.00, N 14.79, O 10.31; found C 53.86, H 3.87, N 14.95, O 10.25 (See Figures S39–S41 (Supplementary Materials: spectrum 1H, 13C NMR, and DEPT-135 for 5k)).
4-[(1-[(3-(4-Dimethylaminophenyl)-4,5-dihydroisoxazolin-5-yl)methyl]-1H-1,2,3-triazol-4-yl)methyl]-2H-[1,4]-benzoxazin-3-one (5l): Brown oil, yield (80%); IR (KBr, vmax/cm−1): 1677 (C=O lactame str.). 1H NMR (500 MHz, DMSO-d6): δ 8.01 (s, 1H), 6.66–7.38 (m, 8H), 5.09–5.13 (m, 1H), 4.96 (s, 2H), 4.65 (s, 2H), 4.49 (dd, 2H, J = 6.5, 14.5 Hz), 3.44, 3.12 (dd, 2H, J = 6.5, 17.5 Hz), 2.91 (s, 6H). 13C NMR (125 MHz, DMSO-d6): δ 164.66, 156.93, 151.87, 145.40, 142.79, 128.92, 116.30, 128.30, 124.98, 124.20, 123.18, 117.04, 116.46, 112.14, 78.34, 67.55, 53.14, 38.37, 36.73, 40.25. Elemental analysis calculated (%) for C23H24N6O3·1/20H2O·1/20C6H14·1/25C4H8O2: C 63.86, H 5.74, N 19.05, O 11.35; found C 63.88, H 5.59, N 19.43, O 11.10 (See Figures S42–S45 (Supplementary Materials: spectrum 1H, 13C NMR, DEPT-135, and IR for 5l)).
4-[(1-[(3-(4-Methoxyphenyl)-4,5-dihydroisoxazolin-5-yl)methyl]-1H-1,2,3-triazol-4-yl)methyl]-2H-[1,4]-benzoxazin-3-one (5m): White powder, yield (82%); m.p. 153–155 °C; IR (KBr, vmax/cm−1): 1679 (C=O lactame str.). 1H NMR (500 MHz, DMSO-d6): δ 8.01 (s, 1H), 6.93–7.50 (m, 8H), 5.11 (s, 2H), 5.01–5.05 (m, 1H), 4.65 (s, 2H), 4.51 (dd, 2H, J = 6.5, 14.5 Hz), 3.76 (s, 3H), 3.49, 3.15 (dd, 2H, J = 6.5, 17 Hz). 13C NMR (125 MHz, DMSO-d6): δ 164.95, 156.63, 161.25, 145.34, 142.85, 128.75, 121.86, 128.94, 125.02, 124.19, 123.16, 117.03, 116.30, 114.74, 78.87, 55.95, 67.61, 52.77, 38.25, 36.64. Elemental analysis calculated (%) for C22H21N5O4·1/10H2O: C 62.73, H 5.07, N 16.63, O 15.57; found C 63.00, H 5.05, N 16.70, O 15.26 (See Figures S46–S49 (Supplementary Materials: spectrum 1H, 13C NMR, DEPT-135, and IR for 5m)).
4-[(1-[(3-(3-Methoxyphenyl)-4,5-dihydroisoxazolin-5-yl)methyl]-1H-1,2,3-triazol-4-yl)methyl]-2H-[1,4]-benzoxazin-3-one (5n): Colorless solid, yield (83%); m.p. 182–184 °C; IR (KBr, vmax/cm−1): 1674 (C=O lactame str.). 1H NMR (500 MHz, DMSO-d6): δ 8.03 (s, 1H), 6.96–7.41 (m, 8H), 5.11 (s, 2H), 5.04–5.07 (m, 1H), 4.65 (s, 2H), 4.55 (dd, 2H, J = 6.5, 14 Hz), 3.74 (s, 3H), 3.53, 3.18 (dd, 2H, J = 6.5, 17.5 Hz). 13C NMR (125 MHz, DMSO-d6): δ 164.68, 157.22, 159.86, 145.41, 142.92, 130.73, 128.97, 130.50, 125.06, 124.18, 123.14, 119.67, 117.05, 116.61, 116.27, 112.12, 79.22, 55.74, 67.67, 52.78, 38.03, 36.73. Elemental analysis calculated (%) for C22H21N5O4·1/15H2O·1/20C6H14: C 63.03, H 5.18, N 16.48, O 15.31; found C 63.00, H 5.05, N 16.70, O 15.26 (See Figures S50–S53 (Supplementary Materials: spectrum 1H, 13C NMR, DEPT-135, and IR for 5n)).
4-[(1-[(3-(4-Fluorophenyl)-4,5-dihydroisoxazolin-5-yl)methyl]-1H-1,2,3-triazol-4-yl)methyl]-2H-[1,4]-benzoxazin-3-one (5o): Colorless solid, yield (81%); m.p. 132–134 °C; 1H NMR (500 MHz, DMSO-d6): δ 8.01 (s, 1H), 6.93–7.90 (m, 8H), 5.10 (s, 2H), 5.05–5.09 (m, 1H), 4.64 (s, 2H), 4.54 (dd, 2H, J = 6.5, 14.5 Hz), 3.52, 3.20 (dd, 2H, J = 6.5, 17.5 Hz). 13C NMR (125 MHz, DMSO-d6): δ 167.33, 156.43, 164.60, 145.37, 142.86, 128.94, 125.04, 130.69, 129.71, 124.18, 123.15, 117.03, 116.47, 115.54, 79.27, 67.61, 52.78, 38.05, 36.73. Elemental analysis calculated (%) for C21H18FN5O3·1/12H2O: C 61.68, H 4.48, N 17.13, O 12.06; found C 61.91, H 4.45, N 17.19, O 11.78 (See Figures S54–S56 (Supplementary Materials: spectrum 1H, 13C NMR, and DEPT-135 for 5o)).

3.4. Molecular Docking Analysis

The molecular docking analysis was conducted following the guidelines outlined in the reference [66,67,68]. The crystalline structures of α-amylase (PDB ID: 1SMD) and α-glycosidase (PDB ID: 5NN5) were obtained from the RCSB protein database (http://www.rcsb.org/pdb) (accessed on 3 March 2024), established at the Brookhaven National Laboratory in 1971. The removal of water molecules was achieved using AutoDock Tools v1.5.7, while also incorporating polar hydrogens and Kollman charges; co-crystallized ligands were excluded; and the protein was saved in the “pdbqt” format. The two-dimensional configuration of each ligand was converted to the three-dimensional configuration using Avogadro version 1.2.0 software, as depicted in [69,70]. Using AutoDock Tools (version 1.5.6), the final pdbqt file of the ligand was obtained. The grid box representing the docking search space was enlarged to better fit the active binding site. The coordinates of the grid box for the two enzymes, α-amylase and α-glycosidase, were defined as follows: for α-amylase, the centers (x, y, and z) were set at 8.349, 58.705, and 19.096, while for α-glucosidase, the centers (x, y, and z) were fixed at 1.591, −26.522, and 87.364, with a uniform grid box size maintained at 40. The results for the docked ligand complexes were expressed as ΔG binding energy values in kcal/mol. Acarbose, an agent with a history of 30 years in treating type 2 diabetes, is utilized to prevent postprandial hyperglycemia by blocking carbohydrate digestion in the small intestine. In this computational section of the investigation, acarbose was employed as the native ligand. The process of generating 2D molecular interaction diagrams and examining protein-ligand binding interactions was carried out using Discovery Studio 4.1 (Dassault Systems Biovia, San Diego, CA, USA).

3.5. ADME Studies

Understanding pharmacokinetic properties, including absorption, distribution, metabolism, and excretion (ADME), is essential for comprehending how a substance acts in the body [71]. These stages describe the journey of a substance from absorption to elimination. Computational tools have become indispensable for predicting the ADME characteristics of molecules, assessing their ability to cross cellular barriers and interact with essential transporters and enzymes for absorption and excretion, and determining their metabolic stability [59]. In our approach to molecule evaluation, we have chosen to use the SwissADME platform (available online: www.swissadme.ch, accessed on 10 April 2024) [61]. This platform allows us to thoroughly examine the physicochemical attributes of synthetic molecules, their potential as therapeutic agents, and their pharmacokinetic properties, thus providing a comprehensive understanding of their ADME profile [72].

4. Conclusions

A series of novel polyheterocyclic molecules, incorporating all three heterocycles [1,4]-benzoxazin-3-one, 1,2,3-triazole, and isoxazoline, were synthesized with high yields. This synthesis involved a double 1,3-dipolar cycloaddition reaction. Initially, a 1,3-dipolar cycloaddition reaction of the “click chemistry” type was conducted in a one-pot process at room temperature using (CuSO4 · 5 H 2 O and sodium ascorbate) as catalysts. Subsequently, the second cycloaddition reaction was carried out between the allylic part of compound 3 and various oximes at a temperature ranging from −5 to 0 °C. These molecules exhibit potential biological activities, as demonstrated by their testing on α-amylase and α-glucosidase. The study reveals that the majority of the synthesized compounds (5a5o) exhibit favorable binding affinities compared with the native ligand acarbose, suggesting potent inhibitory potential against both enzymes. Particularly, compounds 5a and 5o demonstrate notable interactions with amino acid residues surrounding the active site of α-amylase, while compounds 5n and 5e exhibit strong interactions with the active site of α-glucosidase. Additionally, ADME analyses suggest that most of the synthetic compounds have promising pharmacokinetic profiles for potential drug development. These findings support the potential therapeutic efficacy of the synthesized molecules in combating hyperglycemia by targeting key enzymes involved in carbohydrate metabolism. Further studies could involve in vitro and in vivo experiments to validate the antidiabetic potential of the synthesized compounds. Additionally, structural modifications could be explored to enhance the potency and specificity of these compounds for potential drug development purposes.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29133086/s1, Figures S1–S56: Spectra of compounds 3, 5a, 5b, 5c, 5d, 5e, 5f, 5g, 5h, 5i, 5j, 5k, 5l, 5m, 5n, 5o.

Author Contributions

Conceptualization, M.E., A.I., R.A.M., M.B. and M.C.; methodology, S.L, M.E., A.I., A.F. and A.B.; software, N.K.S., S.L. and D.A.; validation, M.B., O.M.N., F.C. and M.C.; formal analysis, M.B. and D.A; investigation, M.C.; resources, M.E. and A.I.; data curation, N.K.S. and S.L.; writing—original draft preparation, A.F., A.B., M.E. and A.I.; writing—review and editing, M.B., R.A.M. and B.E.; visualization, M.B., B.E., R.A.M. and M.C.; supervision, M.C. and F.C.; funding acquisition, O.M.N. and R.A.M. All authors have read and agreed to the published version of the manuscript.

Funding

Researchers from King Saud University in Riyadh, Saudi Arabia provided support for project number RSP2024R119. Faculty of Sciences and Technology, Sultan Moulay Slimane University, provided support for the analyses and the supply of chemicals.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The article contains the original contributions that were made in the study. For additional queries, please contact the corresponding author.

Acknowledgments

The authors extend their appreciation to Researchers Supporting Project Number (RSP2024R119), King Saud University, Riyadh, Saudi Arabia, for funding this work. The authors express their gratitude to the National Centre for Scientific and Technical Research (CNRST) of Morocco for providing access to the technical facilities of the UATRS Division.

Conflicts of Interest

The authors affirm that the research was carried out without any commercial or financial associations that could be seen as a possible conflict of interest.

References

  1. Śmist, M.; Kwiecień, H.; Krawczyk, M. Synthesis and Antifungal Activity of 2H-1,4-Benzoxazin-3(4H)-One Derivatives. J. Environ. Sci. Health B 2016, 51, 393–401. [Google Scholar] [CrossRef] [PubMed]
  2. Madhavan, G.R.; Chakrabarti, R.; Anantha Reddy, K.; Rajesh, B.M.; Balraju, V.; Bheema Rao, P.; Rajagopalan, R.; Iqbal, J. Dual PPAR-α and -γ Activators Derived from Novel Benzoxazinone Containing Thiazolidinediones Having Antidiabetic and Hypolipidemic Potential. Bioorg. Med. Chem. 2006, 14, 584–591. [Google Scholar] [CrossRef] [PubMed]
  3. de Bruijn, W.J.C.; Hageman, J.A.; Araya-Cloutier, C.; Gruppen, H.; Vincken, J.P. QSAR of 1,4-Benzoxazin-3-One Antimicrobials and Their Drug Design Perspectives. Bioorg. Med. Chem. 2018, 26, 6105–6114. [Google Scholar] [CrossRef] [PubMed]
  4. Mehdiyeva, G.M. Synthesis and Antimicrobial Activity of 3-Substituted 8-Propenylbenzo[e][1,3]Oxazines. Russ. J. Appl. Chem. 2022, 95, 277–283. [Google Scholar] [CrossRef]
  5. Nagaraju, A.; Kumar Nukala, S.; Narasimha Swamy Thirukovela, T.; Manchal, R. In Vitro Anticancer and In Silico Studies of Some 1,4-Benzoxazine-1,2,4-Oxadiazole Hybrids. Chem. Sel. 2021, 6, 3318–3321. [Google Scholar] [CrossRef]
  6. Song, T.; Lee, M.; Bae, I.; Byun, J.Y.; Ahn, Y.G.; Kim, Y.H.; Chun, Y.J. Synthesis and Evaluation of a 3,4-Dihydro-2H-Benzoxazine Derivative as a Potent CDK9 Inhibitor for Anticancer Therapy. Bull. Korean Chem. Soc. 2021, 42, 416–419. [Google Scholar] [CrossRef]
  7. Mandzyuk, L.Z.; Matiychuk, V.S.; Chaban, T.I.; Bodnarchuk, O.V.; Matiychuk, J.E.; Obushak, M.D. Spiro Derivatives of 1,10b-Dihydro-5H-Pyrazolo[1,5-c][1,3]-Benzoxazines and Their Antimicrobial, Anti-Inflammatory, and Antioxidant Activity. Chem. Heterocycl. Compd. 2020, 56, 1485–1490. [Google Scholar] [CrossRef]
  8. Hammouda, M.B.; Ahmad, I.; Hamdi, A.; Dbeibia, A.; Patel, H.; Bouali, N.; Hamadou, W.S.; Hosni, K.; Ghannay, S.; Alminderej, F.; et al. Design, Synthesis, Biological Evaluation and in Silico Studies of Novel 1,2,3-Triazole Linked Benzoxazine-2,4-Dione Conjugates as Potent Antimicrobial, Antioxidant and Anti-Inflammatory Agents. Arab. J. Chem. 2022, 15, 104226. [Google Scholar] [CrossRef]
  9. Krasnov, V.P.; Musiyak, V.V.; Levit, G.L.; Gruzdev, D.A.; Andronova, V.L.; Galegov, G.A.; Orshanskaya, I.R.; Sinegubova, E.O.; Zarubaev, V.V.; Charushin, V.N. Synthesis of Pyrimidine Conjugates with 4-(6-Amino-Hexanoyl)-7,8-Difluoro-3,4-Dihydro-3-Methyl-2H-[1,4] Benzoxazine and Evaluation of Their Antiviral Activity. Molecules 2022, 27, 4236. [Google Scholar] [CrossRef]
  10. Vozdvizhenskaya, O.; Andronova, V.L.; Galegov, G.; Levit, G.L.; Krasnov, V.P.; Charushin, V.N. Synthesis and Antiherpetic Activity of Novel Purine Conjugates with 7,8-Difluoro-3-Methyl-3,4-Dihydro-2H-1,4-Benzoxazine. Chem. Heterocycl. Compd. 2021, 57, 490–497. [Google Scholar] [CrossRef]
  11. Akhter, M.; Husain, A.; Akhter, N.; Y Khan, M.S. Synthesis, antiinflammatory and antimicrobial activity of some new 1-(3-Phenyl-3, 4-Dihydro-2H-1, 3-Benzoxazin-6-yl)-ethanone derivatives. Indian J. Pharm. Sci. 2011, 73, 101–104. [Google Scholar] [CrossRef] [PubMed]
  12. Largeron, M.; Mesples, B.; Gressens, P.; Cecchelli, R.; Spedding, M.; Le Ridant, A.; Fleury, M.B. The neuroprotective activity of 8-alkylamino-1,4-benzoxazine antioxidants. Eur. J. Pharmacol. 2001, 424, 189–194. [Google Scholar] [CrossRef] [PubMed]
  13. Benarjee, V.; Saritha, B.; Gangadhar, K.H.; Sailaja, B.B.V. Synthesis of some new 1,4-benzoxazine-pyrazoles in water as EGFR targeting anticancer agents. J. Mol. Struct. 2022, 1265, 133188. [Google Scholar] [CrossRef]
  14. Wang, L.; Ankati, H.; Akubathini, S.K.; Balderamos, M.; Storey, C.A.; Patel, A.V.; Price, V.; Kretzschmar, D.; Biehl, E.R.; D’Mello, S.R. Identification of novel 1, 4-benzoxazine compounds that are protective in tissue culture and in vivo models of neurodegeneration. J. Neurosci. Res. 2010, 88, 1970–1984. [Google Scholar] [CrossRef]
  15. Buchanan, D.; Pham, A.M.; Singh, S.K.; Panda, S.S. Molecular Hybridization of Alkaloids Using 1,2,3-Triazole-Based Click Chemistry. Molecules 2023, 28, 7593. [Google Scholar] [CrossRef]
  16. Bhukal, A.; Kumar, V.; Kumar, L.; Lal, K. Recent Advances in Chalcone-Triazole Hybrids as Potential Pharmacological Agents. Results Chem. 2023, 6, 101173. [Google Scholar] [CrossRef]
  17. Yadav, M.; Lal, K.; Kumar, A.; Kumar, A.; Kumar, D. Indole-Chalcone Linked 1,2,3-Triazole Hybrids: Facile Synthesis, Antimicrobial Evaluation and Docking Studies as Potential Antimicrobial Agents. J. Mol. Struct. 2022, 1261, 132867. [Google Scholar] [CrossRef]
  18. Sharma, M.K.; Parashar, S.; Chahal, M.; Lal, K.; Pandya, N.U.; Om, H. Antimicrobial and In-Silico Evaluation of Novel Chalcone and Amide-Linked 1,4-Disubstituted 1,2,3 Triazoles. J. Mol. Struct. 2022, 1257, 132632. [Google Scholar] [CrossRef]
  19. Yadav, M.; Lal, K.; Kumar, A.; Singh, P.; Vishvakarma, V.K.; Chandra, R. Click Reaction Inspired Synthesis, Antimicrobial Evaluation and in Silico Docking of Some Pyrrole-Chalcone Linked 1,2,3-Triazole Hybrids. J. Mol. Struct. 2023, 1273, 134321. [Google Scholar] [CrossRef]
  20. Jiang, X.; Wu, G.; Zalloum, W.A.; Meuser, M.E.; Dick, A.; Sun, L.; Chen, C.H.; Kang, D.; Jing, L.; Jia, R.; et al. Discovery of Novel 1,4-Disubstituted 1,2,3-Triazole Phenylalanine Derivatives as HIV-1 Capsid Inhibitors. RSC Adv. 2019, 9, 28961–28986. [Google Scholar] [CrossRef]
  21. Al-Taweel, S.; Al-Saraireh, Y.; Al-Trawneh, S.; Alshahateet, S.; Al-Tarawneh, R.; Ayed, N.; Alkhojah, M.; AL-Khaboori, W.; Zereini, W.; Al-Qaralleh, O. Synthesis and Biological Evaluation of Ciprofloxacin—1,2,3-Triazole Hybrids as Antitumor, Antibacterial, and Antioxidant Agents. Heliyon 2023, 9, e22592. [Google Scholar] [CrossRef]
  22. Shafi, S.; Mahboob Alam, M.; Mulakayala, N.; Mulakayala, C.; Vanaja, G.; Kalle, A.M.; Pallu, R.; Alam, M.S. Synthesis of Novel 2-Mercapto Benzothiazole and 1,2,3-Triazole Based Bis-Heterocycles: Their Anti-Inflammatory and Anti-Nociceptive Activities. Eur. J. Med. Chem. 2012, 49, 324–333. [Google Scholar] [CrossRef]
  23. Ravindar, L.; Hasbullah, S.A.; Rakesh, K.P.; Hassan, N.I. Pyrazole and Pyrazoline Derivatives as Antimalarial Agents: A Key Review. Eur. J. Pharm. Sci. 2023, 183, 106365. [Google Scholar] [CrossRef]
  24. Nehra, N.; Tittal, R.K.; Ghule Vikas, D.; Naveen; Lal, K. Synthesis, Antifungal Studies, Molecular Docking, ADME and DNA Interaction Studies of 4-Hydroxyphenyl Benzothiazole Linked 1,2,3-Triazoles. J. Mol. Struct. 2021, 1245, 131013. [Google Scholar] [CrossRef]
  25. Yan, W.; Wang, X.; Li, K.; Li, T.X.; Wang, J.J.; Yao, K.C.; Cao, L.L.; Zhao, S.S.; Ye, Y.H. Design, Synthesis, and Antifungal Activity of Carboxamide Derivatives Possessing 1,2,3-Triazole as Potential Succinate Dehydrogenase Inhibitors. Pestic. Biochem. Physiol. 2019, 156, 160–169. [Google Scholar] [CrossRef] [PubMed]
  26. Xu, Z.; Zhao, S.J.; Liu, Y. 1,2,3-Triazole-Containing Hybrids as Potential Anticancer Agents: Current Developments, Action Mechanisms and Structure-Activity Relationships. Eur. J. Med. Chem. 2019, 183, 111700. [Google Scholar] [CrossRef] [PubMed]
  27. Murthy, I.S.; Sreenivasulu, R.; Alluraiah, G.; Ramesh Raju, R. Design, Synthesis, and Anticancer Activity of 1,2,3-Triazole Linked 1,2-Isoxazole-Imidazo[4,5-b]Pyridine Derivatives. Russ. J. Gen. Chem. 2019, 89, 1718–1723. [Google Scholar] [CrossRef]
  28. Xu, Z. 1,2,3-Triazole-Containing Hybrids with Potential Antibacterial Activity against Methicillin-Resistant Staphylococcus Aureus (MRSA). Eur. J. Med. Chem. 2020, 206, 112686. [Google Scholar] [CrossRef]
  29. Ellouz, M.; Sebbar, N.K.; Fichtali, I.; Ouzidan, Y.; Mennane, Z.; Charof, R.; Mague, J.T.; Urrutigoïty, M.; Essassi, E.M. Synthesis and Antibacterial Activity of New 1,2,3-Triazolylmethyl-2H-1,4-Benzothiazin-3(4H)-One Derivatives. Chem. Cent. J. 2018, 12, 123. [Google Scholar] [CrossRef] [PubMed]
  30. Tan, W.; Li, Q.; Li, W.; Dong, F.; Guo, Z. Synthesis and Antioxidant Property of Novel 1,2,3-Triazole-Linked Starch Derivatives via “Click Chemistry”. Int. J. Biol. Macromol. 2016, 82, 404–410. [Google Scholar] [CrossRef]
  31. Shaikh, M.H.; Subhedar, D.D.; Khan, F.A.K.; Sangshetti, J.N.; Shingate, B.B. 1,2,3-Triazole Incorporated Coumarin Derivatives as Potential Antifungal and Antioxidant Agents. Chin. Chem. Lett. 2016, 27, 295–301. [Google Scholar] [CrossRef]
  32. El-Sayed, W.A.; Khalaf, H.S.; Mohamed, S.F.; Hussien, H.A.; Kutkat, O.M.; Amr, A.E. Synthesis and Antiviral Activity of 1,2,3-Triazole Glycosides Based Substituted Pyridine via Click Cycloaddition. Russ. J. Gen. Chem. 2017, 87, 2444–2453. [Google Scholar] [CrossRef]
  33. Kanzouai, Y.; Chalkha, M.; Hadni, H.; Laghmari, M.; Bouzammit, R.; Nakkabi, A.; Benali, T.; Tüzün, B.; Akhazzane, M.; El Yazidi, M.; et al. Design, Synthesis, in-Vitro and in-Silico Studies of Chromone-Isoxazoline Conjugates as Anti-Bacterial Agents. J. Mol. Struct. 2023, 1293, 136205. [Google Scholar] [CrossRef]
  34. Nie, J.-P.; Qu, Z.-N.; Chen, Y.; Chen, J.-H.; Jiang, Y.; Jin, M.-N.; Yu, Y.; Niu, W.-Y.; Duan, H.-Q.; Qin, N. Discovery and Anti-Diabetic Effects of Novel Isoxazole Based Flavonoid Derivatives. Fitoterapia 2020, 142, 104499. [Google Scholar] [CrossRef] [PubMed]
  35. Phongphane, L.; Radzuan, S.N.M.; Bakar, M.H.A.; Omar, M.T.C.; Supratman, U.; Harneti, D.; Wahab, H.A.; Azmi, M.N. Synthesis, Biological Evaluation, and Molecular Modelling of Novel Quinoxaline-Isoxazole Hybrid as Anti-Hyperglycemic. Comput. Biol. Chem. 2023, 106, 107938. [Google Scholar] [CrossRef] [PubMed]
  36. Bernal, C.C.; Vesga, L.C.; Mendez-Sánchez, S.C.; Romero Bohórquez, A.R. Synthesis and Anticancer Activity of New Tetrahydroquinoline Hybrid Derivatives Tethered to Isoxazoline Moiety. Med. Chem. Res. 2020, 29, 675–689. [Google Scholar] [CrossRef]
  37. Alshamari, A.; Al-Qudah, M.; Hamadeh, F.; Al-Momani, L.; Abu-Orabi, S. Synthesis, Antimicrobial and Antioxidant Activities of 2-Isoxazoline Derivatives. Molecules 2020, 25, 4271. [Google Scholar] [CrossRef] [PubMed]
  38. Ismail, A.H.; Abdula, A.M.; Tomi, I.H.R.; Al-Daraji, A.H.R.; Baqi, Y. Synthesis, Antimicrobial Evaluation and Docking Study of Novel 3,5-Disubstituted-2-Isoxazoline and 1,3,5-Trisubstituted-2-Pyrazoline Derivatives. Med. Chem. 2019, 17, 462–473. [Google Scholar] [CrossRef]
  39. Mota, F.V.B.; De Araújo Neta, M.S.; De Souza Franco, E.; Bastos, I.V.G.A.; Da Araújo, L.C.C.; Da Silva, S.C.; De Oliveira, T.B.; Souza, E.K.; De Almeida, V.M.; Ximenes, R.M.; et al. Evaluation of Anti-Inflammatory Activity and Molecular Docking Study of New Aza-Bicyclic Isoxazoline Acylhydrazone Derivatives. Medchemcomm 2019, 10, 1916–1925. [Google Scholar] [CrossRef]
  40. Zhang, T.; Dong, M.; Zhao, J.; Zhang, X.; Mei, X. Synthesis and Antifungal Activity of Novel Pyrazolines and Isoxazolines Derived from Cuminaldehyde. J. Pestic. Sci. 2019, 44, 181–185. [Google Scholar] [CrossRef]
  41. Quadrelli, P.; Vazquez Martinez, N.; Scrocchi, R.; Corsaro, A.; Pistarà, V. Syntheses of Isoxazoline-Carbocyclic Nucleosides and Their Antiviral Evaluation: A Standard Protocol. Sci. World J. 2014, 2014, 492178. [Google Scholar] [CrossRef] [PubMed]
  42. Filali, I.; Bouajila, J.; Znati, M.; Bousejra-El Garah, F.; Ben Jannet, H. Synthesis of New Isoxazoline Derivatives from Harmine and Evaluation of Their Anti-Alzheimer, Anti-Cancer and Anti-Inflammatory Activities. J. Enzym. Inhib. Med. Chem. 2015, 30, 371–376. [Google Scholar] [CrossRef] [PubMed]
  43. Kumar, R.; Saha, P.; Sahana, S.; Dubey, A. A review on diabetes mellitus: Type1 & type2. World J. Pharm. Pharm. Sci. 2020, 9, 838–850. [Google Scholar] [CrossRef]
  44. Sheard, N.F.; Clark, N.G.; Brand-miller, J.C.; Franz, M.J.; Xavier Pi-sunyer, F.; Mayer-davis, E.; Kulkarni, K.; Geil, P. Dietary Carbohydrate (Amount and Type) in the Prevention and Management of Diabetes A Statement by the American Diabetes Association. Diabetes Care 2004, 27, 2266–2271. [Google Scholar] [CrossRef] [PubMed]
  45. Ben Lamine, J.; Boujbiha, M.A.; Dahane, S.; Cherifa, A.B.; Khlifi, A.; Chahdoura, H.; Yakoubi, M.T.; Ferchichi, S.; El Ayeb, N.; Achour, L. α-Amylase and α-Glucosidase Inhibitor Effects and Pancreatic Response to Diabetes Mellitus on Wistar Rats of Ephedra Alata Areal Part Decoction with Immunohistochemical Analyses. Environ. Sci. Pollut. Res. 2019, 26, 9739–9754. [Google Scholar] [CrossRef] [PubMed]
  46. Rasouli, H.; Hosseini-Ghazvini, S.M.B.; Adibi, H.; Khodarahmi, R. Differential α-Amylase/α-Glucosidase Inhibitory Activities of Plant-Derived Phenolic Compounds: A Virtual Screening Perspective for the Treatment of Obesity and Diabetes. Food Funct. 2017, 8, 1942–1954. [Google Scholar] [CrossRef] [PubMed]
  47. Ilaš, J.; Anderluh, P.Š.; Dolenc, M.S.; Kikelj, D. Recent Advances in the Synthesis of 2H-1,4-Benzoxazin-3-(4H)-Ones and 3,4-Dihydro-2H-1,4-Benzoxazines. Tetrahedron 2005, 61, 7325–7348. [Google Scholar] [CrossRef]
  48. Fang, L.; Zuo, H.; Li, Z.B.; He, X.Y.; Wang, L.Y.; Tian, X.; Zhao, B.X.; Miao, J.Y.; Shin, D.S. Synthesis of Benzo[b][1,4]Oxazin-3(4H)-Ones via Smiles Rearrangement for Antimicrobial Activity. Med. Chem. Res. 2011, 20, 670–677. [Google Scholar] [CrossRef]
  49. Totobenazara, J.; Burke, A.J. New Click-Chemistry Methods for 1,2,3-Triazoles Synthesis: Recent Advances and Applications. Tetrahedron Lett. 2015, 56, 2853–2859. [Google Scholar] [CrossRef]
  50. Velikorodov, A.V.; Sukhenko, L.T. Synthesis and Antimicrobial Activity of 3,5-Disubstituted Isoxazolines and Isoxazoles with Carbamate Groups. Pharm. Chem. J. 2003, 37, 22–24. [Google Scholar] [CrossRef]
  51. Praveena Devi, C.H.B.; Vijay, K.; Hari Babu, B.; Adil, S.F.; Mujahid Alam, M.; Vijjulatha, M.; Ansari, M.B. CuSO4/Sodium Ascorbate Catalysed Synthesis of Benzosuberone and 1,2,3-Triazole Conjugates: Design, Synthesis and in Vitro Anti-Proliferative Activity. J. Saudi Chem. Soc. 2019, 23, 980–991. [Google Scholar] [CrossRef]
  52. Sebbar, N.K.; Mekhzoum, M.E.M.; Essassi, E.M.; Zerzouf, A.; Talbaoui, A.; Bakri, Y.; Saadi, M.; Ammari, L. El Novel 1,4-Benzothiazine Derivatives: Synthesis, Crystal Structure, and Anti-Bacterial Properties. Res. Chem. Intermed. 2016, 42, 6845–6862. [Google Scholar] [CrossRef]
  53. Kheira, N.; Labd, M.; Mokhtar, E.; Abdallah, B.; Zakaria, M.; Mague, T. Synthesis, DFT Study and Antibacterial Activity of Some Isoxazoline Derivatives Containing 1,4-Benzothiazin-3-One Nucleus Obtained Using 1,3-Dipolar Cycloaddition Reaction. Iran. J. Chem. Chem. Eng. 2020, 39, 53–67. [Google Scholar]
  54. Carlson, A.S.; Calcanas, C.; Brunner, R.M.; Topczewski, J.J. Regiocontrolled Wacker Oxidation of Cinnamyl Azides. Org. Lett. 2018, 20, 1604–1607. [Google Scholar] [CrossRef]
  55. Ramasubbu, N.; Ragunath, C.; Mishra, P.J.; Thomas, L.M.; Gyémánt, G.; Kandra, L. Human Salivary A-amylase Trp58 Situated at Subsite− 2 Is Critical for Enzyme Activity. Eur. J. Biochem. 2004, 271, 2517–2529. [Google Scholar] [CrossRef] [PubMed]
  56. Ragunath, C.; Manuel, S.G.A.; Venkataraman, V.; Sait, H.B.R.; Kasinathan, C.; Ramasubbu, N. Probing the Role of Aromatic Residues at the Secondary Saccharide-Binding Sites of Human Salivary α-Amylase in Substrate Hydrolysis and Bacterial Binding. J. Mol. Biol. 2008, 384, 1232–1248. [Google Scholar] [CrossRef] [PubMed]
  57. Hsiu, J.; Fischer, E.H.; Stein, E.A. Alpha-Amylases as Calcium-Metalloenzymes. II. Calcium and the Catalytic Activity. Biochemistry 1964, 3, 61–66. [Google Scholar] [CrossRef] [PubMed]
  58. Shah, M.; Rahman, H.; Khan, A.; Bibi, S.; Ullah, O.; Ullah, S.; Ur Rehman, N.; Murad, W.; Al-Harrasi, A. Identification of α-Glucosidase Inhibitors from Scutellaria Edelbergii: ESI-LC-MS and Computational Approach. Molecules 2022, 27, 1322. [Google Scholar] [CrossRef]
  59. Ferreira, L.L.G.; Andricopulo, A.D. ADMET Modeling Approaches in Drug Discovery. Drug Discov. Today 2019, 24, 1157–1165. [Google Scholar] [CrossRef]
  60. Srivastava, V.; Yadav, A.; Sarkar, P. Molecular Docking and ADMET Study of Bioactive Compounds of Glycyrrhiza Glabra against Main Protease of SARS-CoV2. Mater. Today Proc. 2020, 49, 2999–3007. [Google Scholar] [CrossRef]
  61. Zrouri, H.; Nasr, F.A.; Parvez, M.K.; Alahdab, A. Exploring Medicinal Herbs’ Therapeutic Potential and Molecular Docking Analysis for Compounds as Potential Inhibitors of Human Acetylcholinesterase in Alzheimer’s. Medicina 2023, 59, 1812. [Google Scholar] [CrossRef] [PubMed]
  62. Domínguez, A.; Suárez-Merino, B.; Goñi-de-Cerio, F. Nanoparticles and Blood-Brain Barrier: The Key to Central Nervous System Diseases. J. Nano Nanotechnol. 2014, 14, 766–779. [Google Scholar] [CrossRef] [PubMed]
  63. Stillhart, C.; Vučićević, K.; Augustijns, P.; Basit, A.W.; Batchelor, H.; Flanagan, T.R.; Gesquiere, I.; Greupink, R.; Keszthelyi, D.; Koskinen, M.; et al. European Journal of Pharmaceutical Sciences Impact of Gastrointestinal Physiology on Drug Absorption in Special Populations—An UNGAP Review. Eur. J. Pharm. Sci. 2020, 147, 105280. [Google Scholar] [CrossRef] [PubMed]
  64. Taşçıoğlu, N.; Saatçi, Ç.; Emekli, R.; Tuncel, G.; Eşel, E.; Dundar, M. Investigation of Cytochrome P450 CYP1A2, CYP2D6, CYP2E1 and CYP3A4 Gene Expressions and Polymorphisms in Alcohol Withdrawal. Klin. Psikiyatr. Derg. 2021, 24, 298–306. [Google Scholar] [CrossRef]
  65. Daina, A.; Michielin, O.; Zoete, V. SwissADME: A Free Web Tool to Evaluate Pharmacokinetics, Drug-Likeness and Medicinal Chemistry Friendliness of Small Molecules. Sci. Rep. 2017, 7, 42717. [Google Scholar] [CrossRef] [PubMed]
  66. Kandsi, F.; Elbouzidi, A.; Lafdil, F.Z.; Meskali, N.; Azghar, A.; Addi, M.; Hano, C.; Maleb, A.; Gseyra, N. Antibacterial and Antioxidant Activity of Dysphania ambrosioides (L.) Mosyakin and Clemants Essential Oils: Experimental and Computational Approaches. Antibiotics 2022, 11, 482. [Google Scholar] [CrossRef] [PubMed]
  67. Elbouzidi, A.; Taibi, M.; Laarej, S.; El Hassania, L.; Haddou, M.; Hachlafi, N.E.; Naceiri Mrabti, H.; Baraich, A.; Bellaouchi, R.; ASEHRAOU, A. Chemical Profiling of Volatile Compounds of the Essential Oil of Grey-Leaved Rockrose (Cistus albidus L.) and Its Antioxidant, Anti-Inflammatory, Antibacterial, Antifungal, and Anticancer Activity In Vitro and In Silico. Front. Chem. 2024, 12, 1334028. [Google Scholar] [CrossRef] [PubMed]
  68. Ouahabi, S.; Loukili, E.H.; Elbouzidi, A.; Taibi, M.; Bouslamti, M.; Nafidi, H.-A.; Salamatullah, A.M.; Saidi, N.; Bellaouchi, R.; Addi, M.; et al. Pharmacological Properties of Chemically Characterized Extracts from Mastic Tree: In Vitro and In Silico Assays. Life 2023, 13, 1393. [Google Scholar] [CrossRef]
  69. Rajendran, P.; Rathinasabapathy, R.; Chandra Kishore, S.; Bellucci, S. Computational-Simulation-Based Behavioral Analysis of Chemical Compounds. J. Compos. Sci. 2023, 7, 196. [Google Scholar] [CrossRef]
  70. Cooper, A.K.; Oliver-Hoyo, M.T. Creating 3D Physical Models to Probe Student Understanding of Macromolecular Structure. Biochem. Mol. Biol. Educ. 2017, 45, 491–500. [Google Scholar] [CrossRef]
  71. Van de Waterbeemd, H.; Gifford, E. ADMET in Silico Modelling: Towards Prediction Paradise? Nat. Rev. Drug Discov. 2003, 2, 192–204. [Google Scholar] [CrossRef] [PubMed]
  72. Aja, P.M.; Agu, P.C.; Ezeh, E.M.; Awoke, J.N.; Ogwoni, H.A.; Deusdedit, T.; Ekpono, E.U.; Igwenyi, I.O.; Alum, E.U.; Ugwuja, E.I.; et al. Prospect into Therapeutic Potentials of Moringa Oleifera Phytocompounds against Cancer Upsurge: De Novo Synthesis of Test Compounds, Molecular Docking, and ADMET Studies. Bull. Natl. Res. Cent. 2021, 45, 99. [Google Scholar] [CrossRef]
Figure 1. Examples of some bioactive molecules derived from [1,4]-benzoxazine-3-one.
Figure 1. Examples of some bioactive molecules derived from [1,4]-benzoxazine-3-one.
Molecules 29 03086 g001
Figure 2. Synthesis of the novel isoxazolinyl-1,2,3-triazolyl-[1,4]-benzoxazin-3-one derivatives 5a5o.
Figure 2. Synthesis of the novel isoxazolinyl-1,2,3-triazolyl-[1,4]-benzoxazin-3-one derivatives 5a5o.
Molecules 29 03086 g002
Figure 3. Two-dimensional schemes of the interactions with the amino acid residues of the two potent synthetic molecules, (A) 5a, (B) 5o, and the native ligand, Acarbose (C).
Figure 3. Two-dimensional schemes of the interactions with the amino acid residues of the two potent synthetic molecules, (A) 5a, (B) 5o, and the native ligand, Acarbose (C).
Molecules 29 03086 g003
Figure 4. Two-dimensional schemes of the interactions with the amino acid residues of the two potent synthetic molecules, (A) 5n, (B) 5e, and the native ligand, Acarbose (C).
Figure 4. Two-dimensional schemes of the interactions with the amino acid residues of the two potent synthetic molecules, (A) 5n, (B) 5e, and the native ligand, Acarbose (C).
Molecules 29 03086 g004
Figure 5. BOILED-Egg Model of the GI absorption and BBB permeability of synthetic molecules (1) 5k, (2) 5e, (3) 5o, (4) 5a, (5) 5f, (6) 5l, (7) 5j, (8) 5n, (9) 5g, (10) 5i, (11) 5d. PGP-: non-substrate of P-glycoprotein, PGP+: P-glycoprotein substrate.
Figure 5. BOILED-Egg Model of the GI absorption and BBB permeability of synthetic molecules (1) 5k, (2) 5e, (3) 5o, (4) 5a, (5) 5f, (6) 5l, (7) 5j, (8) 5n, (9) 5g, (10) 5i, (11) 5d. PGP-: non-substrate of P-glycoprotein, PGP+: P-glycoprotein substrate.
Molecules 29 03086 g005
Figure 6. Bioavailability radar of synthetic molecules. The pink region corresponds to the ideal range for each characteristic in terms of oral bioavailability (Lipophilicity, solubility, molecular weight, saturation and flexibility).
Figure 6. Bioavailability radar of synthetic molecules. The pink region corresponds to the ideal range for each characteristic in terms of oral bioavailability (Lipophilicity, solubility, molecular weight, saturation and flexibility).
Molecules 29 03086 g006
Table 1. 1H NMR and 13C NMR spectra exhibit some characteristic signals of the synthesized compounds 5a5o.
Table 1. 1H NMR and 13C NMR spectra exhibit some characteristic signals of the synthesized compounds 5a5o.
1H NMR (ppm)13C NMR (ppm)
CH TriazolCH2 IsoxazolineC=OC=N
5a8.023.20, 3.50164.66157.23
5b8.043.16, 3.52164.64155.27
5c8.033.29, 3.63164.61156.20
5d8.023.29, 3.57164.64156.41
5e8.013.20, 3.52164.66156.50
5f7.942.95, 3.35164.68158.64
5g8.003.13, 3.45164.65149.17
5h8.013.05, 3.34164.68158.38
5i8.013.20, 3.54164.65149.17
5j8.003.14, 3.46164.86151.18
5k8.013.15, 3.49164.62156.60
5l8.013.12, 3.44164.66156.93
5m8.013.15, 3.49164.95156.63
5n8.033.18, 3.53164.68157.22
5o8.013.20, 3.52167.33156.43
Table 2. H-bonds, binding energy, and interacting amino acids of phytocompounds found in the synthetic molecules 5a5o target protein.
Table 2. H-bonds, binding energy, and interacting amino acids of phytocompounds found in the synthetic molecules 5a5o target protein.
Compoundsα-Amylase Protein (PDB ID: 1SMD)
Affinity
(kcal/mol)
H-Bonding
Acarbose 1−7.8Tyr A:2, Gln A:7, Ser A:289, Ser A:289, Asp A:402, Gly A:334
5a−9.2 *-
5b−8.7 *-
5c−8.2 *Gln A:63, Arg A:195, His A:299
5d−9.0 *His A:201
5e−9 *-
5f−8.4 *-
5g−8.3 *His A:201
5h−8.2 *-
5i−8.6 *-
5j−7.9 *-
5k−8.2 *-
5l−8.8 *-
5m−8.8 *Lys A:200
5n−7.7Asp A:63
5o−9.1 *His A:202
1 Acarbose, a native ligand of α-amylase; * The potent ligands in comparison with the native ligand.
Table 3. H-bonds, binding energy, and interacting amino acids of phytocompounds found in the synthetic molecules 5a5o target protein.
Table 3. H-bonds, binding energy, and interacting amino acids of phytocompounds found in the synthetic molecules 5a5o target protein.
Compoundsα-Glucosidase Protein (PDB ID: 5NN5)
Affinity
(kcal/mol)
H-Bonding
Acarbose 1−7.2Asp A:356, Met A:363, Glu A: 866, Arg A:608
5a−9.5 *Tyr A:360, Met A:363, Arg A:608, Glu A:866
5b−9.5 *Tyr A:360, Met A:363, Arg A:608, Glu A:866
5c−9.5 *Tyr A:360, Met A:363, Arg A:608, Glu A:866
5d−8.9 *Tyr A:360, Glu A:866
5e−9.9 *Tyr A:360, Met A:363, His A: 584, Arg A:608, Glu A:866
5f−9 *Arg A:608, Glu A:866
5g−9.1 *Tyr A:360, Met A:363, Arg A:608, Glu A:866
5h−9.1 *Tyr A:360, Met A:363, Arg A:608, Glu A:866
5i−9.5 *Tyr A:360, Met A:363, Arg A:608, Glu A:866
5j−8.8 *Tyr A: 360, Glu A:866
5k−9.5 *-
5l−8.7 *Tyr A:360, Arg A: 866
5m−9.1 *Tyr A:360, Arg A: 608
5n−9.6 *Tyr A:360, Met A:363, Arg A:608, Glu A:866
5o−9.4 *Tyr A:360, Arg A:608, Glu A:866
1 Acarbose, a native ligand of α-amylase; * The potent ligands in comparison with the native ligand.
Table 4. Evaluation of the pharmacokinetic properties (ADME) of the Synthetic Compounds.
Table 4. Evaluation of the pharmacokinetic properties (ADME) of the Synthetic Compounds.
Physicochemical PropertiesLipophilicityDruglikenessPharmacokinetics
CompoundsMW g/molHBAHBDTPSA
Å2
Rotatable BondsM logPW logPLipinski’sVerber’sGI
Absorption
BBB
Permeation
CYP1A2
Inhibitor
5a403.46081.851.91.700HighNoYes
5b434.480131.261.81.200HighNoNo
5c434.480131.261.81.200HighNoNo
5d434.480131.261.81.200HighNoNo
5e423.86081.852.22.100HighNoYes
5f403.46081.861.91.600HighNoNo
5g379.37094.950.51.000HighNoYes
5h415.46081.862.12.000HighNoNo
5i390.47094.750.70.800HighNoYes
5j392.46086.750.80.800HighNoNo
5k468.36081.852.32.200HighNoYes
5l432.46085.061.61.500HighNoNo
5m419.47097.061.41.400HighNoYes
5n419.47091.0761.471.4900HighNoNo
5o407.47081.852.42.000HighNoYes
MW: molecular weight; HBD: Hydrogen-Bond Donors; HBA: Hydrogen-Bond Acceptors; WLogP: Lipophilicity; MLogP: Octanol/water partition coefficient; TPSA: Topological polar surface area; BBB: Blood-Brain Barrier; GI: Gastrointestinal.
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

Ellouz, M.; Ihammi, A.; Baraich, A.; Farihi, A.; Addichi, D.; Loughmari, S.; Sebbar, N.K.; Bouhrim, M.; A. Mothana, R.; M. Noman, O.; et al. Synthesis and In Silico Analysis of New Polyheterocyclic Molecules Derived from [1,4]-Benzoxazin-3-one and Their Inhibitory Effect against Pancreatic α-Amylase and Intestinal α-Glucosidase. Molecules 2024, 29, 3086. https://doi.org/10.3390/molecules29133086

AMA Style

Ellouz M, Ihammi A, Baraich A, Farihi A, Addichi D, Loughmari S, Sebbar NK, Bouhrim M, A. Mothana R, M. Noman O, et al. Synthesis and In Silico Analysis of New Polyheterocyclic Molecules Derived from [1,4]-Benzoxazin-3-one and Their Inhibitory Effect against Pancreatic α-Amylase and Intestinal α-Glucosidase. Molecules. 2024; 29(13):3086. https://doi.org/10.3390/molecules29133086

Chicago/Turabian Style

Ellouz, Mohamed, Aziz Ihammi, Abdellah Baraich, Ayoub Farihi, Darifa Addichi, Saliha Loughmari, Nada Kheira Sebbar, Mohamed Bouhrim, Ramzi A. Mothana, Omar M. Noman, and et al. 2024. "Synthesis and In Silico Analysis of New Polyheterocyclic Molecules Derived from [1,4]-Benzoxazin-3-one and Their Inhibitory Effect against Pancreatic α-Amylase and Intestinal α-Glucosidase" Molecules 29, no. 13: 3086. https://doi.org/10.3390/molecules29133086

Article Metrics

Back to TopTop