Abstract
1,3,4-Oxadiazoles are a vital class of heterocyclic compounds known for their diverse biological activities. In this study, a series of eight novel 1,3,4-oxadiazolyl sulfide derivatives 4a–h were synthesized and characterized using IR, NMR, and elemental analysis. The antioxidant activity of these derivatives was evaluated via DPPH and ABTS assays, revealing promising radical scavenging capabilities Compound 4 h emerged as the most potent antioxidant with SC50 values of 9.88 µM (ABTS) and 12.34 µM (DPPH), outperforming standard antioxidants surpassing standard antioxidants such as ascorbic acid (SC₅₀ = 23.92 µM) and gallic acid (SC₅₀ = 21.24 µM). The impact of electron-donating and electron-withdrawing substituents on activity was demonstrated through a comprehensive structure-activity relationship (SAR) study. Molecular docking against α-glucosidase (PDB: 3W37) validated the potential of these compounds as enzyme inhibitors, with docking scores ranging from − 8.59 to -9.81 kcal/mol and similar modes of binding. Insights into electronic properties were obtained through density functional theory (DFT) calculations, emphasizing that compound with the lowest HOMO–LUMO energy gaps (4b) exhibited higher polarizability and enhanced reactivity, which correlates with their biological antioxidant performance. This integrated study underscores the therapeutic potential of these derivatives as antioxidants and enzyme inhibitors, offering paths for further drug development.
Supplementary Information
The online version contains supplementary material available at 10.1186/s13065-025-01678-w.
Keywords: 1,3,4-Oxadiazole; Antioxidant; Molecular docking; SAR-study; DFT.
Introduction
Normal aerobic cellular metabolism frequently produces free radicals. An excess of free radicals can harm biomolecules (lipids, proteins, and DNA) through oxidative stress, which can ultimately result in tissue damage and cellular death. Free radicals contribute to the onset of chronic inflammation, atherosclerosis, cancer, diabetes, cardiovascular illnesses, rheumatoid arthritis, stroke, and various other degenerative conditions in humans [1–5]. The body’s antioxidant system is essential for preventing any damage brought on by free radicals. Also, antioxidants are able to scavenge free radical species [6]. In order to combat environmental and clinical factors that have accumulated in the body, such as oxidative stress, radiation, pressure, smoking, and contaminated particles, a number of recent studies have looked into the administration of strong antioxidants [7, 8].
Therefore, it is recommended that we enhance the antioxidant content of our bodies through a natural, balanced diet. Correspondingly, medicinal chemistry attempts to create antioxidant agents that are practical and effective for clinical use [6, 7, 9]. Type 2 diabetes mellitus is one of the most prevalent chronic metabolic disorders, characterized by persistent hyperglycemia resulting from impaired insulin action or secretion [10–12]. A key factor in postprandial hyperglycemia is the enzyme α-glucosidase, which catalyzes the hydrolysis of dietary oligosaccharides and disaccharides into glucose [13]. Inhibition of α-glucosidase slows carbohydrate digestion and glucose absorption, thereby reducing postprandial blood glucose spikes. Clinically approved α-glucosidase inhibitors such as acarbose, miglitol, and voglibose demonstrate the therapeutic relevance of this approach in type 2 diabetes management [14]. Developing novel α-glucosidase inhibitors with concurrent antioxidant activity may therefore provide dual benefits—improving glycemic control while reducing oxidative stress, a key contributor to diabetic complications.
Heterocyclic pharmacophores such as 1,2,4-oxadiazole-substituted scaffolds were used to develop strong antiproliferative drugs [6, 15–17]. They have therefore lately been developed as potential medications to combat these various illnesses, including aging in aerobic organisms, inflammation, atherogenesis, and carcinogenesis [18, 19]. Because of their significant biological characteristics, 1,3,4-oxadiazoles are a special type of heterocyclic system that has been observed in recent decades. These nitrogen-containing heterocycles with five members are essential, particularly in medicinal chemistry [20–23]. 1,3,4-Oxadiazole and its derivatives have demonstrated a wide array of biological activities, including antitubercular [24, 25], anti-inflammatory [26], analgesic [27], anticancer [28], antimalarial [29], anti-allergic agent [30], antibacterial [31, 32], anticonvulsant [33], and vasodilator activities [34] (Fig. 1). The oxadiazole nucleus is a fundamental component of many marketed drugs, such as the anti-cancer drug zibotentan, the HIV integrase inhibitor raltegravir, and the hypertension drug nesapidil [35–37] (Fig. 1). Moreover, some 1,3,4-oxadiazole derivatives showed promising α-glucosidase inhibition [22, 32].
Fig. 1.
Commercial drugs with the privilege oxadiazole nucleus
The 1,3,4-oxadiazolyl sulphide derivative 1 exhibited a potent inhibitory effect against α-glucosidase and α-amylase (IC50 = 126.84 and 129.72 µM in comparison to acarbose IC50 = 53.82 µM), in addition to potent antioxidant scavenging activity (DPPH IC50 = 36.76 µM compared to reference ascorbic acid IC50 = 50.30 µM) [38] (Fig. 2). Also, The 1,3,4-oxadiazolyl sulphide derivative 2 showed α-glucosidase(IC50 = 15.85 µM in comparison to acarbose IC50 = 17.85 µM), in addition to a relevant antioxidant activity with an IC50 of 54.86 µM in comparison to gallic acid (IC50 = 26.23 µM) [39]. Interestingly, the The 1,3,4-oxadiazolyl sulphide derivative 3 significantly decreased the blood glucose level in diabetic rat [40].
Fig. 2.
Potent antidiabetic agents with α-glucosidase inhibition, antioxidant properties and design methodology of the novel compounds
Scientists have created a number of novel techniques in recent years for the synthesis of 1,3,4-oxadiazole derivatives. Using an acyl hydrazine derivative, Harish et al. have reported synthesizing 2-monosubstituted-1,3,4-oxadiazole from triethylorthoformate [41]. Modern synthetic methods for 1,3,4-oxadiazole, rely on the oxidation of acylhydrazones and the cyclization of hydrazides or acylthiosemicarbazides [42–46] using a range of reagents such as sulfuric acid, phosphorus oxychloride, or thionyl chloride, typically under harsh reaction conditions.
Stimulatingly, the literature review highlighted 1,3,4-oxadiazole sulphide derivatives exhibited significant antidiabetic and antioxidant activity. As such, his cores serve as promising framework for developing novel antioxidants. Building on our continued interest in this field, we designed and synthesized a range of derivatives featuring these core structures (Fig. 2), incorporating the mecaptoacetamide as a diverse substituent. The resulting compounds were assessed for their antioxidant activity, and the findings were further validated through molecular docking and DFT studies.
Results and discussion
Chemistry
The multifunctional precursor 5-(4-aminophenyl)-1,3,4-oxadiazole-2-thiol (2) has been prepared according to the reported conditions [47] through a cyclization reaction of 4-aminobenzohydrazide (1) with carbon disulfide. Compound (2) has a great affinity to undergo alkylation reaction with numerous aromatic and heterocyclic 2-chloroacetamide derivatives (Scheme 1). Therefore, a series of linked sulfur scaffolds bearing a 1,3,4-oxadiazole moiety (4a-h) was constructed. The reaction started through nucleophilic attack of compound (2) on various aromatic and heterocyclic chloroacetamide derivatives (3a-h), namely, p-toluidine, p-anisidine, p-chloroaniline, p-aminoacetophenone, chloroacetamide thiazole, chloroacetamide ethoxybenzothiazole, chloroacetamide pyridine, and chloroacetamide antipyrine. The reaction proceeded by stirring a mixture of compound 2 and chloroacetamide derivatives (3a-h) in dry acetone and K2CO3 to afford the corresponding 1,3,4-oxadiazolyl sulfide scaffolds 4a–h, respectively (Scheme 1). The chemical structure of the newly synthesized 1,3,4-oxadiazolyl sulfide derivatives 4a–h was confirmed based on their elemental and spectral data. The distinctive absorption bands at 3405, 3319, and 3265 cm− 1 that correspond to the stretching vibrations of amino and imino groups and at 1702 and 1674 cm− 1 that correspond to two carbonyl groups (acetyl and amidic) were found in compound 4d’s infrared spectrum. Singlet signals at δ 2.52 and 4.29 ppm in the 1H-NMR spectrum were attributed to the protons of the methyl (CH3) and methylene (CH2) groups. The 13C-NMR spectrum showed distinct signals for the amidic and acetyl carbonyl groups at δ 160.90 and 165.75 ppm, respectively. The distinctive absorption bands at 3470, 3380, 3084, and 1671 cm− 1 in the infrared spectrum of 4f were found to correspond to the stretching vibrations of amino, imino, and amidic carbonyl groups, respectively. Methyl protons showed a triplet signal at δ 1.33 ppm in the 1H-NMR spectrum, while methylene protons showed quartet and singlet signals at δ 4.03 and 4.37 ppm. At δ 5.94 ppm, the amino group’s protons showed up as a singlet signal. Characteristic methyl and two methylene carbon signals were detected in the 13C-NMR spectrum at δ = 14.70, 35.56, and 63.61 ppm, respectively. The amidic carbonyl group carbon was represented by the typical signal at δ 166.16 ppm.
Scheme 1.
Synthesis of 1,3,4-oxadiazolyl sulfide derivatives 4a-h
In vitro antioxidant (DPPH and ABTS) radical scavenging activities
The synthetic 1,3,4-oxadiazolyl sulfide derivatives 4a-h were calculated for their in vitro antioxidant activity using DPPH and ABTS assays according to the previously described procedure [48]. The antioxidant capacity for the compounds was expressed as SC50 (µM) for both DPPH and ABTS assays. The radical scavenging activity results were calculated from the absorbance values of the oxidized ABTS.+ or DPPH.+, detected by color and used to calculate radical scavenging concentration SC50 (Table 1, Fig. S19, S20). The obtained results were compared to standard ascorbic acid with SC50 = 23.92 µM (for DPPH) and gallic acid with SC50 = 21.24 µM (for ABTS), respectively. The compounds showed DPPH & ABTS scavenging activities in the range of SC50 = 12.34 to 87.89 µM (for DPPH) and 9.88 to 74.42 µM (for ABTS), respectively. According to the ABTS assay, the best synthetic compound was 4 h, since it showed the lowest SC50 value of 9.88 µM. The scavenging activities for the rest of the compounds were arranged as 4 h < 4b < 4f < 4a < 4c < 4e < 4d < 4 g, with SC50 values 9.88, 12.84, 14.02, 17.48, 39.64, 55.95, 68.06 and 74.42 µM. While for the DPPH assay, 4 h, 4a, 4b and 4c showed superior scavenging activity than ascorbic acid. The compounds are arranged as 4 h < 4a < 4b < 4c < 4d < 4f < 4e < 4 g with SC50 values of 12.34, 12.63, 13.19, 22.38, 25.56, 26.20, 37.89, and 87.89 µM. Although the present study demonstrates the antioxidant and α-glucosidase inhibitory potential of oxadiazolyl sulfide derivatives through in vitro and in silico approaches, further validation is required.
Table 1.
ABTS & DPPH Inhibition of 1,3,4-oxadiazole derivatives 4a-h
| Compounds | ABTS (SC50 value, µM) |
DPPH data (SC50 value, µM) |
|---|---|---|
| 4a | 17.48 | 12.63 |
| 4b | 12.84 | 13.19 |
| 4c | 39.64 | 22.38 |
| 4d | 68.06 | 25.56 |
| 4e | 55.95 | 37.89 |
| 4f | 14.02 | 26.20 |
| 4 g | 74.42 | 87.89 |
| 4 h | 9.88 | 12.34 |
| Ascorbic acid | 23.92 | |
| Gallic acid | 21.24 |
Structure-activity relationship (SAR)
It was reported that 1,3,4-oxadiazole derivatives to have measurable DPPH/ABTS scavenging. Sulfur substituents (thioether / thioamide) frequently enhance radical reactivity hydrogen-atom transfer (HAT) and single-electron transfer (SET) pathways. The synthesized compounds generally exhibited almost equal activity against DPPH radicals compared to ABTS radicals.
The activity varies according to the substituents and the mechanistic drivers. Overall rank, 4 h derivative containing pyrazolone group anchored on the oxadiazole ring through mercapto acetamide linker exhibited the strongest antioxidant activities (SC50 = 12.34 µM for DPPH, SC50 = 9.88 µM for ABTS, respectively). This may be attributed to the strong resonance stabilization of the resulting radical across the compound with the privilege of the phenyl ring hanged on pyrazolone moiety. The proposed mechanism by which the synthesized compounds function as scavengers is illustrated in Figs. 2 and 3 [49]. Additionally, the structure–activity relationship (SAR) insights outlined below can be derived from the data presented in Table 1:
Fig. 3.
The proposed mechanism of compound 4b for the antioxidant DPPH assay
In p-substituted phenyl series, compound 4a,4b containing methyl and methoxy groups, improved the scavenging activity among the tested compounds (SC50 = 17.48 and 12.84 µM for DPPH and SC50 = 12.63 and 13.19 µM for ABTS, respectively). This can be attributed to the strong electron-donating properties of methyl and methoxy groups which stabilize the formed radical through resonance and inductive effects, which reduce the N-H bond energy and subsequently enhance radical scavenging activity. On the other hand, the antioxidant activity was diminished by the presence of Cl- or COCH3 group, as demonstrated by the SC50 values of compound 4d, 4c (22.38 and 25.56 µM for DPPH and 39.64 and 68.06 µM for ABTS). The e donation was hindered by the effect of electron withdrawing effect of the groups, destabilizing the radical intermediate.
Replacing the phenyl ring with heterocyclic thiazole ring 4e shows a slight decrease in the antioxidant activity (SC50 = 37.89 µM and 55.95 µM for DPPH and ABTS, respectively). The fused benzothiazole with ethoxy substituent derivative 4f shows imminent scavenging activity compared to the thiazole ring (SC50 = 26.20 µM and 14.02 µM for DPPH and ABTS, respectively). Providing larger conjugation and better radical delocalization than simple thiazole.
Replacing the phenyl ring with pyridine ring 4 g markedly exhibit the weakest antioxidant activity (SC50 = 87.89 µM and 74.42 µM for DPPH and ABTS, respectively). The pyridine electron practice the withdrawing effect lowering the radical donation ability.
Molecular docking
Molecular docking was conducted to examine the interactions and binding behavior of the synthesized compounds with the α-glucosidase enzyme. This will directly contribute to the SAR and biological relevance of the findings reported. The enzyme crystallographic coordinates were retrieved from www.rcsb.org using the PDB ID of 3w37 for α-glucosidase [50, 51]. The docking protocol was validated, proving its reliability for the intended study. This was evident in Fig. 4, where the alignment between the native co-crystallized ligand and its re-docked pose within the active pocket showed a minimal RMSD value of 0.5233 Å. The docking analysis demonstrated that the 1,3,4-oxadiazolyl sulfide derivatives 4a-h exhibited significant potential against the α-glucosidase enzyme. These compounds consistently formed three key interactions with three identical amino acids, which are likely crucial for stabilizing the ligand inside the active site. The docking scores (kcal/mol) and comprehensive binding interactions are outlined in Table 2, with binding energies ranging from − 8.59 to -9.81 kcal/mol.
Fig. 4.
SAR of the synthesized 1,3,4-oxadiazolyl sulfide derivatives 4a-h
Table 2.
The molecular Docking results for the synthesized 1,3,4-oxadiazolyl sulfide derivatives targeting the protein structure with PDB ID: 3W37
| Code | S (energy score) (Kcal/mol) |
Moiety | Residue | Interactions | Distance (A) |
|---|---|---|---|---|---|
| 4a | – 8.73 |
N-aniline Oxadiazole ring N-Oxadiazole N-carboxamide Ph-ring Ph-ring |
Asp 357 Arg 552 Met 470 Asp 232 Ile 233 Ala 234 |
H-donor π –cation H-donor H-donor π-H interaction π-H interaction |
3.13 4.23 3.34 3.44 4.52 4.55 |
| 4b | – 9.07 |
N-aniline N-Oxadiazole Ph-ring |
Asp 357 Arg 552 Ile 233 |
H-donor H-acceptor π-H interaction |
2.80 3.25 4.45 |
| 4c | – 8.77 |
N-aniline Oxadiazole ring N-Oxadiazole N-carboxamide Ph-ring Ph-ring |
Asp 357 Arg 552 Met 470 Asp 232 Ile 233 Ala 234 |
H-donor π-cation H-donor H-donor π-H interaction π-H interaction |
3.14 4.23 3.34 3.43 4.52 4.56 |
| 4d | – 9.00 |
N-aniline Oxadiazole ring N-Oxadiazole N-carboxamide Ph-ring Ph-ring Ph-ring |
Asp 357 Arg 552 Met 470 Asp 232 Ile 233 Ile 233 Ala 234 |
H-donor π-cation H-donor H-donor π-H interaction π-H interaction π-H interaction |
3.14 4.23 3.34 3.35 4.48 4.71 4.60 |
| 4e | – 8.64 |
N-aniline Oxadiazole ring N-Oxadiazole S-thiazole |
Asp 357 Arg 552 Met 470 Asp 232 |
H-donor π-cation H-donor H-donor |
3.08 4.27 3.48 3.48 |
| 4f | – 9.35 |
N-aniline Oxadiazole ring N-Oxadiazole Thiazole ring Thiazole ring Thiazole ring N-carboxamide S-thiazole |
Asp 357 Arg 552 Met 470 Ile 233 Ile 233 Ala 234 Asp 232 Asp 232 |
H-donor π -cation H-donor π-H interaction π-H interaction π-H interaction H-donor H-donor |
3.15 4.23 3.34 4.39 4.72 4.53 3.25 3.58 |
| 4 g | – 8.59 |
N-aniline Oxadiazole ring N-Oxadiazole |
Asp 357 Arg 552 Met 470 |
H-donor π-cation H-donor |
3.11 4.46 3.48 |
| 4 h | – 9.81 |
N-aniline Oxadiazole ring N-Oxadiazole O-pyrazolone Pyrazolone ring |
Asp 357 Arg 552 Met 470 Ala 234 Ile 233 |
H-donor π-cation H-donor H-acceptor π-H interaction |
3.19 4.28 3.40 3.57 4.33 |
The main scaffold in all synthesized compounds oxadiazole ring, is hydrogen bonded to Arg552 as presented in Fig. 5 and Figs. S21-S28. The N-atom of the oxadiazole ring attached to Met 470 through a hydrogen bond in all derivatives except 4b. All derivatives show the same π–cation interaction between the amino group and Asp 357. The NH of the carboxamide group shows a hydrogen bond interaction with Asp 232 in 4a, 4c, 4d, and 4f. The phenyl ring in derivatives 4a-d shows a number of hydrophobic interactions with two amino acids present in the active site, Ile 233 and Ala 234. In derivative 4e, the thiazole ring is attached to Asp 232 present in the active site through a hydrogen bond, while derivative 4f shows enforced binding to the thiazole ring through hydrophobic interactions with Ile 233 and Ala 234. Also, in derivative 4 h, the pyrazolone ring is attached to the active site through a hydrogen bond between the O-atom of the pyrazolone ring with Ala 234 and a hydrophobic bond between the pyrazolone ring and Ile 233. The alignment of all derivatives 4a-h inside the active pocket of α-glucosidase, showing similar binding mode of action (Fig. 6).
Fig. 5.

The 2D figure shows the overlay between the cocrystallized ligand (acarbose) (red) and redocked acarbose (green)
Fig. 6.
The binding interaction of the 1,3,4-oxadiazolyl sulfide derivative 4f with (PDB ID: 3W37)
Frontier molecular orbitals (FMOs)
The 1,3,4-oxadiazole derivatives’ 4a-h electrical features and structural differences are determined by DFT simulations. An application of molecular orbital theory is Frontier Molecular Orbitals (FMOs), it can determine the electronic properties and the chemical reactivity of the synthesized derivatives [52, 53]. Through examining Fig. 7, Fig. S29-S36 and Table 3, optimized structures, HOMO and LUMO (highest occupied and lowest unoccupied molecular orbitals) distributions of the investigated molecules can be deduced.
Fig. 7.
3D visualization for the alignment of docked poses of 1,3,4-oxadiazolyl sulfide derivatives 4a-h inside the active pocket of (PDB ID: 3W37) A without showing a receptor. B showing binding with receptor
Table 3.
Optimized structures, electron density of the 1,3,4-oxadiazolyl sulfide derivatives 4a-h

As presented in Table 4, the findings were reflected in the energy values (EHOMO and ELUMO). The EHOMO values were relatively similar, ranging from 6.21 to 5.78 eV, while the ELUMO values fell within the range of 1.84 to 1.52 eV. Furthermore, the 4-acetylphenyl derivative demonstrated the highest EHOMO, where 4d >4c >4a >4b. The values of the HOMO–LUMO energy gap (ΔEH−L) ranged between 4.46 − 4.07 eV. They are arranged in the order: 4f < 4 h < 4c < 4d < 4 g < 4a < 4e < 4b. The HOMO–LUMO energy gap correlate the relative low ΔE value of derivate 4b with its scavenging activity in both ABTS and DPPH with immenant SC50 value 12.84 and 13.19 µM. A small band gap means the compound is more polarizable, linked to low kinetic stability and high chemical reactivity, and the compound is termed soft [54]. 4b derivative is the softest compound in the series (Fig. 8).
Table 4.
The HOMO energy (EHOMO), LUMO energy (ELUMO), HOMO-LUMO energy gap (Egap) in eV, electronegativity (χ), global hardness (η), softness (δ), and electrophilicity (ω)
| Molecules | EHOMO eV | ELUMO eV | Egap | χ | η | δ | ω |
|---|---|---|---|---|---|---|---|
| 4a | – 6.09 | – 1.71 | 4.38 | 3.90 | 2.19 | 0.46 | 3.46 |
| 4b | – 5.78 | – 1.71 | 4.07 | 3.74 | 2.03 | 0.49 | 3.45 |
| 4c | – 6.16 | – 1.71 | 4.45 | 3.93 | 2.23 | 0.45 | 3.47 |
| 4d | – 6.21 | – 1.78 | 4.42 | 3.99 | 2.21 | 0.45 | 3.61 |
| 4e | – 6.14 | – 1.84 | 4.30 | 3.99 | 2.15 | 0.47 | 3.71 |
| 4f | – 5.90 | – 1.44 | 4.46 | 3.67 | 2.23 | 0.45 | 3.01 |
| 4 g | – 6.08 | – 1.68 | 4.41 | 3.88 | 2.20 | 0.45 | 3.42 |
| 4 h | – 5.98 | – 1.52 | 4.45 | 3.75 | 2.23 | 0.45 | 3.16 |
Fig. 8.
FMO energies of 1,3,4-oxadiazolyl sulfide derivatives 4a-h
Experimental
Chemistry
The supplementary data contains information on all devices utilized in the chemistry section’s analytical experiments. Compounds 2 [47] and 3a-h [55–62] were prepared as reported.
Synthesis of 2-((5-(4-aminophenyl)-1,3,4-oxadiazol-2-yl)thio)-N- acetamide derivatives 4a-h:
A suspension of chloroacetamide-based compounds 3a-h (10 mmol), 2-mercapto-1,3,4-oxadiazole compound 2 (1.93 g, 10 mmol), and potassium carbonate (1.38 g, 10 mmol) was stirred overnight in 30 mL dry acetone. The resulting solid was obtained by pouring the mixture into ice water. After collection, the solid underwent recrystallization using absolute ethanol, yielding the corresponding pure oxadiazole sulfide compounds 4a-h.
2-((5-(4-Aminophenyl)-1,3,4-oxadiazol-2-yl)thio)-N-(p-tolyl)acetamide (4a)
Yellow crystals; yield 44%; m.p. = 230–231 °C. TLC solvent [pet. ether: ethyl acetate (1:3)]. IR (ν/cm− 1): 3356, 3300, 3190 (NH2, NH), 1665 (C = O). 1H NMR (500 MHz) (DMSO): δ 2.23 (s, 3 H, CH3), 4.29 (s, 2 H, S-CH2), 7.11 (d, J = 8.50 Hz, 2 H), 7.44 (t, J = 8.50 Hz, 3 H), 7.70–7.78 (m, 3 H, Ar-H, NH2), 7.84 (d, J = 8.50 Hz, 1H, ), 7.90 (d, J = 9.00 Hz, 1H), 10.39 ppm (s, 1H, N-H). 13C NMR (125 MHz): δ 20.71, 40.40, 56.20, 113.72, 119.35 (2 C), 120.32 (2 C), 120.42, 126.86, 127.05, 127.78 (2 C), 128.03, 129.37, 129.39, 129.44 ppm. Anal. Calcd for C17H16N4O2S (340.10): C, 59.98; H, 4.74; N, 16.46%. Found: C, 59.86; H, 4.76; N, 16.58%. HPLC: rt 14.41 min (purity 97,24%).
2-((5-(4-Aminophenyl)-1,3,4-oxadiazol-2-yl)thio)-N-(4 methoxyphenyl)acetamide (4b)
White crystals; yield 35%; m.p. = 240–241 °C. TLC solvent [pet. ether: ethyl acetate (1:2)]. IR (ν/cm− 1): 3368, 3062 (NH2, NH), 1634 (C = O). 1H NMR (500 MHz) (DMSO): δ 3.71 (s, 3 H, O-CH3), 4.21 (s, 2 H, S-CH2), 5.96 (s, 2 H, NH2), 6.62 (d, J = 8.50 Hz, 2 H), 6.80–6.93 (m, 2 H), 7.44–7.52 (m, 2 H), 7.56 (d, J = 6.50 Hz, 2 H), 10.26 ppm (s, 1H, N-H). 13C NMR (125 MHz): δ 36.67, 55.30, 113.68 (4 C), 114.09 (2 C), 120.89 (2 C), 128.05 (3 C), 155.74 (2 C), 165.01 ppm (2 C). Anal. Calcd for C17H16N4O3S (356.09): C, 57.29; H, 4.53; N, 15.72%. Found: C, 57.48; H, 4.55; N, 15.84%. HPLC: rt 13,38 min (purity 93,62%).
2-((5-(4-Aminophenyl)-1,3,4-oxadiazol-2-yl)thio)-N-(4-chlorophenyl)acetamide (4c)
Golden yellow crystals; yield 60%; m.p. = 228–229 °C. TLC solvent [pet. ether: ethyl acetate (1:2)]. IR (ν/cm− 1): 3343, 3308 (NH2, NH), 1650 (C = O). 1H NMR (500 MHz) (DMSO): δ 4.32 (s, 2 H, S-CH2), 7.31–7.43 (m, 4 H), 7.50–7.62 (m, 4 H), 7.76 (s, 2 H), 10.56 ppm (s, 1H, NH). 13C NMR (125 MHz): δ 36.50, 114.02 (2 C), 121.23 (2 C), 121.34, 123.17, 127.77, 128.39, 129.11, 129.26 (2 C), 138.28, 166.50, 167.34, 168.18 ppm. Anal. Calcd for C16H13ClN4O2S (360.04): C, 53.26; H, 3.63; N, 15.53%. Found: C, 53.29; H, 3.65; N, 15.48%.%. HPLC: rt 14.75 min (purity 93,87%).
N-(4-Acetylphenyl)-2-((5-(4-aminophenyl)-1,3,4-oxadiazol-2-yl)thio)acetamide (4d)
White crystals; yield 78%; m.p. = 234–235 °C. TLC solvent [pet. ether: ethyl acetate (1:3)]. IR (ν/cm− 1): 3405, 3319, 3265 (NH2, NH), 1702, 1674 (2 C = O). 1H NMR (500 MHz) (DMSO): δ 2.52 (s, 3 H, CH3), 4.29 (s, 2 H, S-CH2), 5.92 (s, 2 H, NH2), 6.61 (d, J = 8.00 Hz, 2 H), 7.56 (d, J = 9.00 Hz, 2 H), 7.7 (d, J = 9.00 Hz, 2 H), 7.93 (d, J = 8.50 Hz, 2 H), 10.72 (s, 1H, N-H). 13C NMR (125 MHz): δ 26.46, 36.83, 109.17, 113.48 (2 C), 116.07, 118.43 (2 C), 127.90 (2 C), 129.58 (2 C), 132.06, 142.92, 152.40, 160.90, 165.75, 196.52 ppm. MS, m/z (%) 368 (M+, 38.30%), 359.81 (61.03%), 346.07 (90.57%), 326.35 (60.67%), 314.64 (94.85%), 308.35 (52.42%), 283.01 (63.09%), 249.68 (50.82%), 220.58 (80.31%), 193.95 (51.34%), 184.95 (62.58%), 174.87 (72.68%), 100.99 (92.16%), 92.00 (100.00%). Anal. Calcd for C18H16N4O3S (368.09): C, 58.68; H, 4.38; N, 15.21%. Found: C, 58.61; H, 4.40; N, 15.26%. HPLC: rt 14,55 min (purity 99,32%).
2-((5-(4-Aminophenyl)-1,3,4-oxadiazol-2-yl)thio)-N-(thiazol-2-yl)acetamide (4e)
White crystals; yield 78%; m.p. = 200–201 °C. TLC solvent [pet. ether: ethyl acetate (1:2)]. IR (ν/cm− 1): 3463, 3357, 3225 (NH2, NH), 1690 (C = O). 1H NMR (500 MHz) (DMSO): δ 4.32 (s, 2 H, S-CH2), 6.61 (d, J = 8.50 Hz, 2 H), 7.24 (d, J = 3.00 Hz, 1H), 7.48 (d, J = 3.00 Hz, 1H), 7.55 (d, J = 9.00 Hz, 2 H), 12.51 ppm (s, 1H, N-H). 13C NMR (125 MHz): δ 35.46, 109.57, 113.94 (2 C), 114.19, 128.16 (2 C), 137.99, 152.32, 157.92, 160.94, 165.84, 166.36 ppm. Anal. Calcd for C13H11N5O2S2 (333.38): C, 46.84; H, 3.33; N, 21.01%. Found: C, 46.89; H, 3.35; N, 21.11%. HPLC: rt 13,06 min (purity 100%).
2-((5-(4-Aminophenyl)-1,3,4-oxadiazol-2-yl)thio)-N-(6-ethoxybenzo[d]thiazol-2-yl)acetamide (4f)
Gray crystals; yield 60%; m.p. = 240–241 °C. TLC solvent [pet. ether: ethyl acetate (1:1)]. IR (ν/cm− 1): 3470, 3380, 3084 (NH2, NH), 1671 (C = O). 1H NMR (500 MHz) (DMSO): δ 1.33 (t, J = 7.00 Hz, 3 H, -OCH2CH3), 4.03 (q, J = 7.00 Hz, 2 H, -OCH2CH3), 4.37 (s, 2 H, S-CH2), 5.94 (s, 2 H, NH2), 6.60 (d, J = 8.00 Hz, 2 H), 7.00–7.03 (m, 1H), 7.54 (d, J = 2.50 Hz, 1H), 7.66 (d, J = 9.00 Hz, 2 H), 7.64 (d, J = 9.00 Hz, 1H), 12.62 ppm (s, 1H, N-H). 13C NMR (125 MHz): δ 14.70, 35.56, 63.61, 105.36, 113.56 (2 C), 115.44 (3 C), 121.32, 127.95 (2 C), 132.77, 142.49, 152.35, 155.51, 160.63, 166.16, 166.33 ppm. Anal. Calcd for C19H17N5O3S2 (427.08): C, 53.38; H, 4.01; N, 16.38%. Found: C, 53.57; H, 4.03; N, 16.26%. HPLC: rt 14,40 min (purity 99,86%).
2-((5-(4-Aminophenyl)-1,3,4-oxadiazol-2-yl)thio)-N-(pyridin-2-yl)acetamide (4 g)
Brown crystals; yield 71%; m.p. = 195–196 °C. TLC solvent [pet. ether: ethyl acetate (1:2)]. IR (ν/cm− 1): 3463, 3308, 3195 (NH2, NH), 1690 (C = O). 1H NMR (500 MHz) (DMSO): δ 4.37 (s, 2 H, S-CH2), 5.93 (s, 2 H, NH2), 6.61 (d, J = 8.50 Hz, 2 H), 7.12 (t, J = 6.50 Hz, 1H), 7.56 (d, J = 9.00 Hz, 1H), 7.76–7.79 (m, 2 H, Ar-H), 8.02 (s, 1H, Ar-H), 8.32 (d, J = 4.50 Hz, 1H, Ar-H), 10.88 ppm (s, 1H, NH). 13C NMR (125 MHz): δ 36.6, 109.36, 113.65, 113.68, 120.03, 128.10 (2 C), 138.56, 148.26 (2 C), 151.69, 152.52, 161.07, 166.22, 166.33 ppm. Anal. Calcd for C15H13N5O2S (327.08): C, 55.04; H, 4.00; N, 21.39%. Found: C, 55.12; H, 4.05; N, 21.45%. HPLC: rt 16,40 min (purity 89,94%).
2-((5-(4-Aminophenyl)-1,3,4-oxadiazol-2-yl)thio)-N-(1,5-dimethyl-3-oxo-2-phenyl-2,3-dihydro-1 H-pyrazol-4-yl)acetamide (4 h)
White crystals; yield 30%; m.p. = 225–226 °C. TLC solvent [pet. ether: ethyl acetate (1:1)]. IR (ν/cm− 1): 3412, 3295, 3181 (NH2, NH), 1693 (2 C = O). 1H NMR (500 MHz) (DMSO): δ 2.08 (s, 3 H, CH3), 3.03 (s, 3 H, CH3), 4.27 (s, 2 H, S-CH2), 7.25–7.38 (m, 3 H), 7.41–7.55 (m, 2 H), 7.77 (d, J = 9.00 Hz, 1H), 7.95 (d, J = 9.00 Hz, 2 H), 9.58 (s, 2 H, NH2), 10.43 (s, 1H, N-H). 13C NMR (100 MHz): δ 14.28, 24.58, 39.28, 113.58, 114.93 (2 C), 115.00, 115.10, 118.63, 127.48, 127.83, 129.03, 129.14 (2 C), 132.16, 132.33, 147.91, 161.22, 162.98, 164.70, 169.44 ppm. Anal. Calcd for for C21H20N6O3S (436.13): C, 57.79; H, 4.62; N, 19.25%. Found: C, 57.98; H, 4.64; N, 19.13%. HPLC: rt 15,60 min (purity 95,43%).
In vitro antioxidant (DPPH and ABTS) radical scavenging activities
The radical scavenging activity of the targeting oxadiazole sulfide compounds was tested using ABTS or DPPH, referring to the reported methods [48, 63]. The details of the experimental methods are provided in the Supporting Information.
Molecular docking
Docking experiments were performed via Molecular Operating Environment (MOE) docking program version 2019. Crystal structures of α-glucosidase, PDB: 3W37, were selected for docking.
DFT calculations
The compounds were subjected to DFT simulations to verify the intended geometries of the molecules under inquiry using Gaussian 09 W suite program [64]. The DFT calculations were carried out utilizing the Becke3–Lee–Yang– Parr (B3LYP) exchange–correlation functional [65–67] with standard 6–311 + + G(dp) basis set. The HOMO–LUMO plots and mep data were obtained using the GaussView program [68].
Conclusion
The synthesized 1,3,4-oxadiazolyl sulfide derivatives 4a-h demonstrated notable antioxidant activities, with compound 4f exhibiting superior radical scavenging capabilities in both DPPH and ABTS assays. The structure-activity relationship analysis highlighted the role of electron-donating and withdrawing groups in modulating activity. Molecular docking studies revealed significant binding interactions with α-glucosidase, supporting their potential as enzyme inhibitors. Furthermore, DFT calculations provided a deeper understanding of the electronic properties of the derivatives, correlating their HOMO-LUMO gaps with biological activity. These findings collectively establish the synthesized derivatives as promising candidates as antioxidants and enzyme inhibitors against diabetes. Future investigations in cellular and in vivo models will be necessary to evaluate the anti-interference ability of these derivatives in the presence of diverse bioactive substances within complex biological systems.
Supplementary Information
Below is the link to the electronic supplementary material.
Author contributions
**Norhan A. Khalaf** : Synthesis, methodology, characterization of compounds, Writing – original draft. **Gehad E. Said: ** Writing – review & editing, Validation, Supervision, Investigation, Formal analysis, Data curation. **Ehab Abdel-Latif: ** Writing – review & editing, Validation, Supervision, Investigation. **Heba M. Metwally**: Writing – original draft, review & editing, Supervision, Investigation, Formal analysis, Data curation, modelling and docking studies.
Funding
Open access funding provided by The Science, Technology & Innovation Funding Authority (STDF) in cooperation with The Egyptian Knowledge Bank (EKB).
Data availability
The datasets used and/or analysed during the current study are available from the corresponding author on reasonable request.
Declarations
Ethics approval and consent to participate
Not applicable.
Consent for publication
Not applicable.
Competing interests
The authors declare no competing interests.
Footnotes
Publisher’s note
Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Contributor Information
Gehad E. Said, Email: gehadsaid@mans.edu.eg
Heba M. Metwally, Email: hebama@mans.edu.eg
References
- 1.Fridovich I. Fundamental aspects of reactive oxygen species. Ann N Y Acad Sci. 1999;893:13–8. [DOI] [PubMed] [Google Scholar]
- 2.Huie RE, Padmaja S. The reaction of NO with superoxide. Free Radic Res Commun. 1993;18(4):195–9. [DOI] [PubMed] [Google Scholar]
- 3.Cheng X, Huang J, Li H, Zhao D, Liu Z, Zhu L, Zhang Z, Peng W. Quercetin: A promising therapy for diabetic encephalopathy through Inhibition of hippocampal ferroptosis. Phytomedicine. 2024;126:154887–901. [DOI] [PubMed] [Google Scholar]
- 4.Wei Y, Xu S, Wu Z, Zhang M, Bao M, He B. Exploring the causal relationships between type 2 diabetes and neurological disorders using a Mendelian randomization strategy. Medicine. 2024;103(46):e40412–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Zhang N, Wang Y, Li W, Wang Y, Zhang H, Xu D, Chen R, Tang L, Tang H. Association between serum vitamin D level and cardiovascular disease in Chinese patients with type 2 diabetes mellitus: A cross-sectional study. Sci Rep. 2025;15(1):6454–64. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Tang L, Wang Y, Gong X, Xiang J, Zhang Y, Xiang Q, Li J. Integrated transcriptome and metabolome analysis to investigate the mechanism of intranasal insulin treatment in a rat model of vascular dementia. Front Pharmacol. 2023;14:1182803–14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Bing P, Zhou W, Tan S. Study on the mechanism of astragalus polysaccharide in treating pulmonary fibrosis based on Drug-Target-Pathway network. Front Pharmacol. 2022;13:865065–75. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Yang D, Li K, Zhang Q, Yu T, Liu L, Wang Y, Kang S, Song S, Xu B, Liu R. Sex and age influence the relationship between serum creatinine/cystatin C and carotid plaque in patients with type 2 diabetes mellitus. Angiology. 2024;0(0):1–9. [DOI] [PubMed] [Google Scholar]
- 9.Dai B, Wu Q, Zeng C, Zhang J, Cao L, Xiao Z, Yang M. The effect of Liuwei Dihuang Decoction on PI3K/Akt signaling in liver of T2DM rats. J Ethnopharmacol. 2016;192:382–9. [DOI] [PubMed] [Google Scholar]
- 10.Banday MZ, Sameer AS, Nissar S. Pathophysiology of diabetes: an overview. Avicenna J Med. 2020;10(4):174–88. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Tomic D, Shaw JE, Magliano DJ. The burden and risks of emerging complications of diabetes mellitus. Nat Rev Endocrinol. 2022;18(9):525–39. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Dragan L, Alexia P, Ioanna Z, Haralambos G, Andreas P, Athanasios M. The growing epidemic of diabetes mellitus. Curr Vasc Pharmacol. 2020;18(2):104–9. [DOI] [PubMed] [Google Scholar]
- 13.Nair SS, Kavrekar V, Mishra A. In vitro studies on alpha-amylase and alpha-glucosidase inhibitory activities of selected plant extracts. Eur J Exp Biol. 2013;3(1):128–32. [Google Scholar]
- 14.Dahlén AD, Dashi G, Maslov I, Attwood MM, Jonsson J, Trukhan V, Schiöth HB. Trends in antidiabetic drug discovery: FDA approved drugs, new drugs in clinical trials and global sales. Front Pharmacol. 2022;12:807548–54. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Lukin A, Karapetian R, Ivanenkov Y, Krasavin M. Discovery, privileged 1,2,4-oxadiazoles in anticancer drug design: novel 5-aryloxymethyl-1,2,4-oxadiazole leads for prostate cancer therapy. Lett Drug Des Discov. 2016;13(3):198–204. [Google Scholar]
- 16.Kumar G, Saroha B, Arya P, Ghosh S, Kumari B, Nassare VD, Raghav N, Kumar S. 1,2,3-Triazole clubbed and Dichloro substituted novel Aurones as potential anticancer agents targeting digestive enzymes: Design, synthesis, DFT, ADME and molecular Docking studies. J Mol Struct. 2025;1319:139460–73. [Google Scholar]
- 17.Kumar G, Saroha B, Kumar S, Kumari B, Arya P, Raghav N, Ghosh S, Nassare VD. 1,2,3-Triazole-based new Aurones as anticancer agents with the capability to target extracellular digestive enzymes. ChemistrySelect. 2024;9(44):e202403885. [Google Scholar]
- 18.Tyagi YK, Kumar A, Raj HG, Vohra P, Gupta G, Kumari R, Kumar P, Gupta RK. Synthesis of novel amino and acetyl amino-4-methylcoumarins and evaluation of their antioxidant activity. Eur J Med Chem. 2005;40(4):413–20. [DOI] [PubMed] [Google Scholar]
- 19.Karegoudar P, Karthikeyan MS, Prasad DJ, Mahalinga M, Holla BS, Kumari NS. Synthesis of some novel 2,4-disubstituted thiazoles as possible antimicrobial agents. Eur J Med Chem. 2008;43(2):261–7. [DOI] [PubMed] [Google Scholar]
- 20.Vaidya A, Pathak D, Shah K. 1,3,4-Oxadiazole and its derivatives: A review on recent progress in anticancer activities. Chem Biol Drug Des. 2021;97(3):572–91. [DOI] [PubMed] [Google Scholar]
- 21.Gupta O, Chawla G, Pradhan T. 1,3,4-Oxadiazole scaffold in antidiabetic drug discovery: an overview. Mini-Rev Med Chem. 2024;24(20):1800–21. [DOI] [PubMed] [Google Scholar]
- 22.Liang J, He Y, Huang C, Ji F, Zhou X, Yin Y. The regulation of selenoproteins in diabetes: A new way to treat diabetes. Curr Pharm Des. 2024;30(20):1541–7. [DOI] [PubMed] [Google Scholar]
- 23.Jaishree B, Basavarajaiah S, Karunakar P, Dhonnar SL, Nagesh G, Punithkumar M, Bhargav P, Vishwas Aradhya KJ. Exploration of 1,3,4-oxadiazoles engrafted with Indole and phthalimide scaffolds as multi-target peroxidase, acetylcholinesterase, and butyrylcholinesterase inhibitors: Synthesis, DFT calculations, and molecular Docking studies. ChemistrySelect. 2025;10(6):e202404925. [Google Scholar]
- 24.Zarghi A, Tabatabai SA, Faizi M, Ahadian A, Navabi P, Zanganeh V, Shafiee A. Synthesis and anticonvulsant activity of new 2-substituted-5-(2-benzyloxyphenyl)-1,3,4-oxadiazoles. Bioorg. Med Chem Lett. 2005;15(7):1863–5. [DOI] [PubMed] [Google Scholar]
- 25.Harish Kumar K, Sridhar BT, Karunakar P, Nagesh GY, Gupta N, Jisha SP, Basavarajaiah SM. Investigation of novel 2-(Chloromethyl)‐5‐(3, 5‐Disubstituted‐1H‐Indol‐2‐yl)‐1, 3, 4‐Oxadiazole derivatives as in Vitro, and in Silico bioactivity potential: Anti‐Inflammatory, Anti‐TB and antioxidant activities study. ChemistrySelect. 2024;9(33):e202401756. [Google Scholar]
- 26.Omar FA, Mahfouz NM, Rahman MA. Design, synthesis and antiinflammatory activity of some 1, 3, 4-oxadiazole derivatives. Eur J Med Chem. 1996;31(10):819–25. [DOI] [PubMed] [Google Scholar]
- 27.Husain A, Ajmal M. Synthesis of novel 1,3,4-oxadiazole derivatives and their biological properties. Acta Pharm. 2009;59(2):223–33. [DOI] [PubMed] [Google Scholar]
- 28.Sengupta P, Dash DK, Yeligar VC, Murugesh K, Rajalingam D, Singh J. Maity Tk. Evaluation of anticancer activity of some 1,3,4-oxadiazole derivatives. Indian J Chem. 2008;47B:460–2. [Google Scholar]
- 29.Adimule V, Jagadeesha Gowda AH, Nandi SS, Bowmik D. Antimalarial activity of novel class of 1,3-benzoxaborole derivatives containing 1,3,4-oxadiazole moiety. Drug Dev Malar. 2022;285–302.
- 30.Dinneswara Reddy G, Park S-J, Cho HM, Kim T-J, Lee ME. Antiallergic activity of new sila-substituted 1,3,4-oxadiazoles in vitro and in vivo. J Med Chem. 2012;55(14):6438–44. [DOI] [PubMed] [Google Scholar]
- 31.Holla BS, Gonsalves R, Shenoy S. Synthesis and antibacterial studies of a new series of 1, 2-bis (1, 3, 4-oxadiazol-2-yl) Ethanes and 1, 2-bis (4-amino-1, 2, 4-triazol-3-yl) Ethanes. Eur J Med Chem. 2000;35(2):267–71. [DOI] [PubMed] [Google Scholar]
- 32.Sardar A, Abid O, Khan S, Hussain R, Daud S, Rehman W, Aziz T, Shah BA, Alharbi M, Alasmari AF. Identification of in vitro α-glucosidase and urease inhibitory effect, and in Silico studies of naproxen-derived 1,3,4-oxadiazole-based Schiff-base derivatives. J Mol Struct. 2024;1305:137712–26. [Google Scholar]
- 33.Wang S, Liu H, Wang X, Lei K, Li G, Li J, Liu R, Quan Z. Synthesis of 1,3,4-oxadiazole derivatives with anticonvulsant activity and their binding to the GABAA receptor. Eur J Med Chem. 2020;206:112672–87. [DOI] [PubMed] [Google Scholar]
- 34.Atmaram UA, Roopan SM. Biological activity of oxadiazole and thiadiazole derivatives. Appl Microbiol Biotechnol. 2022;106(9):3489–505. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Singh S, Kumar S, Sharma LK, Singh RKP. A novel electroorganic synthesis of some 2-amino-5-substituted-1,3,4-oxadiazoles at the platinum electrode. Chem Inf. 2010;41(19).
- 36.Lotfi B, Mustafa B, Leila B, Salima M. Electrocyclization of semicarbazone; A novel route of green synthesis of 2,5-disubstituted-1,3,4-oxadiazoles. Int J Electrochem Sci. 2011;6(6):1991–2000. [Google Scholar]
- 37.El-Sayed WA, Ali OM, Hendy HA, Abdel-Rahman AAH. Synthesis and antimicrobial activity of new 2,5-disubstituted 1,3,4-oxadiazoles and 1,2,4-triazoles. Chin J Chem. 2012;30(1):77–83. [Google Scholar]
- 38.Gani RS, Kudva AK, Timanagouda K, Mujawar SBH, Joshi SD, Raghu SV. Synthesis of novel 5-(2,5-bis(2,2,2-trifluoroethoxy)phenyl)-1,3,4-oxadiazole-2-thiol derivatives as potential glucosidase inhibitors. Bioorg Chem. 2021;114:105046–57. [DOI] [PubMed] [Google Scholar]
- 39.Cele N, Awolade P, Seboletswe P, Olofinsan K, Islam MS, Singh P. α-Glucosidase and α-amylase inhibitory potentials of quinoline–1,3,4-oxadiazole conjugates bearing 1,2,3-triazole. Pharm. 2022;15(8):1035–65. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Mohammad BD, Baig MS, Bhandari N, Siddiqui FA, Khan SL, Ahmad Z, Khan FS, Tagde P, Jeandet P. Heterocyclic compounds as dipeptidyl peptidase-IV inhibitors with special emphasis on oxadiazoles as potent anti-diabetic agents. Molecules. 2022;27(18):6001–18. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Harish KP, Mohana KN, Mallesha L. Synthesis of novel 1-[5-(4-methoxy-phenyl)-[1, 3, 4] oxadiazol-2-yl]-piperazine derivatives and evaluation of their in vivo anticonvulsant activity. Eur J Med Chem. 2013;65:276–83. [DOI] [PubMed] [Google Scholar]
- 42.Bondock S, Adel S, Etman HA, Badria FA. Synthesis and antitumor evaluation of some new 1,3,4-oxadiazole-based heterocycles. Eur J Med Chem. 2012;48:192–9. [DOI] [PubMed] [Google Scholar]
- 43.Majhi S, Manickam S. Semisynthesis of bioactive compounds and their biological activities. Elsevier; 2023.
- 44.Wang X, Li Z, Zhang Z, Da Y. Synthesis of 2-(4-chlorobenzoylamido)-5-aryloxymethyl-1,3,4-oxadiazoles under microwave irradiation. Synth Commun. 2001;31(12):1907–11. [Google Scholar]
- 45.Wang X, Li Z, Wei B, Yang J. Synthesis of 2-(4-methoxylphenyloxy-acetylamido)-5-aryloxymethyl-1,3,4-oxadiazoles under microwave irradiation. Synth Commun. 2002;32(7):1097–103. [Google Scholar]
- 46.Srivastava SK, Pathak RP, Bahel SC. Synthesis of schiff bases, 2,5-disubstituted 1,3,4-oxadiazoles and 2,3-disubstituted 4-thiazolidinones as antifungal agents. Chem Inf. 1992;23(25).
- 47.Tafer A, Benalia M, Djedid M, Bouchareb H, Al-Dujaili AH. Synthesis of four liquid crystals and study of alkyl chain effect on the nematic range for application in GC. Mol Cryst Liq Cryst. 2018;665(1):10–9. [Google Scholar]
- 48.González-Palma I, Escalona-Buendía HB, Ponce-Alquicira E, Téllez-Téllez M, Gupta VK, Díaz-Godínez G, Soriano-Santos J. Evaluation of the antioxidant activity of aqueous and methanol extracts of pleurotus ostreatus in different growth stages. Front Microbiol. 2016;7:1099–108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Sonmez F, Gunesli Z, Kurt BZ, Gazioglu I, Avci D, Kucukislamoglu M. Synthesis, antioxidant activity and SAR study of novel spiro-isatin-based schiff bases. Mol Divers. 2019;23(4):829–44. [DOI] [PubMed] [Google Scholar]
- 50.Shulgau Z, Palamarchuk IV, Sergazy S, Urazbayeva A, Ramankulov Y, Kulakov IV. Synthesis, computational study, and in vitro α-glucosidase inhibitory action of 1,3,4-thiadiazole derivatives of 3-aminopyridin-2(1H)-ones. Pharmaceutics. 2024;17(3):377–82. [DOI] [PMC free article] [PubMed]
- 51.Tagami T, Yamashita K, Okuyama M, Mori H, Yao M, Kimura A. Molecular basis for the recognition of long-chain substrates by plant α-glucosidases. J Biol Chem. 2013;288(26):19296–303. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Sajan D, Joseph L, Vijayan N, Karabacak M. Natural bond orbital analysis, electronic structure, non-linear properties and vibrational spectral analysis of L-histidinium bromide monohydrate: A density functional theory. Spectrochim Acta - A: Mol Biomol Spectrosc. 2011;81(1):85–98. [DOI] [PubMed] [Google Scholar]
- 53.Bulat FA, Chamorro E, Fuentealba P, Toro-Labbe A. Condensation of frontier molecular orbital Fukui functions. J Phys Chem A. 2004;108(2):342–9. [Google Scholar]
- 54.Shanty AA, Raghu KG, Mohanan PV. Synthesis, characterization: spectral and theoretical, molecular Docking and in vitro studies of copper complexes with HIV RT enzyme. J Mol Struct. 2019;1197:154–63. [Google Scholar]
- 55.Shaaban HG. Synthesis of some heterocyclic compounds derived from 2-chloro-Np-tolylacetamide. Al-Mustansiriyah J Sci. 2016;27(4):1–5. [Google Scholar]
- 56.Bhat IK, Chaithanya SK, Satyanarayana PD, Kalluraya B. The synthesis and antimicrobial study of some Azetidinone derivatives with the para-anisidine moiety. J Serb Chem Soc. 2007;72(5):437–42. [Google Scholar]
- 57.Chauhan PM, Thummar SN, Chikhalia KHJ. Design, synthesis, bioactivity, and computational studies of morpholine-clubbed coumarinyl acetamide and cinnamide derivatives. J Indian Chem Soc. 2018;15:1261–77. [Google Scholar]
- 58.Ren J-L, Zhang X-Y, Yu B, Wang X-X, Shao K-P, Zhu X-G, Liu H-M. Discovery of novel AHLs as potent antiproliferative agents. Eur J Med Chem. 2015;93:321–9. [DOI] [PubMed] [Google Scholar]
- 59.Koçak Aslan E, Sezer A, Tüylü Küçükkılınç T, Palaska E. Novel 1,2,4-triazole derivatives containing the naphthalene moiety as selective butyrylcholinesterase inhibitors: Design, synthesis, and biological evaluation. Arch Pharm. 2024;357(11):e2400406. [DOI] [PubMed] [Google Scholar]
- 60.Patel RV, Patel JK, Kumari P, Chikhalia KH. Combination of bioactive moieties with different heteroatom(s): application of the Suzuki cross-coupling reaction. Heteroat Chem. 2012;23(4):399–410. [Google Scholar]
- 61.Bhanushali U, Rajendran S, Sarma K, Kulkarni P, Chatti K, Chatterjee S, Ramaa CS. 5-Benzylidene-2,4-thiazolidenedione derivatives: Design, synthesis and evaluation as inhibitors of angiogenesis targeting VEGFR-2. Bioorg Chem. 2016;67:139–47. [DOI] [PubMed] [Google Scholar]
- 62.Turan-Zitouni G, Sıvacı M, Kılıç FS, Erol K. Synthesis of some triazolyl-antipyrine derivatives and investigation of analgesic activity. Eur J Med Chem. 2001;36:685–9. [DOI] [PubMed] [Google Scholar]
- 63.Leaves L, Leaves LJA. Antioxidant activity by DPPH radical scavenging method of ageratum conyzoides. J Ethnobiol Ethnomed. 2014;1(4):244–9. [Google Scholar]
- 64.Frisch M, et al. Gaussian 09 W. Wallingford, CT. USA: Gaussian. Inc.; 2009. [Google Scholar]
- 65.Becke AD. Density-functional thermochemistry. III. The role of exact exchange. J Chem Phys. 1993;98(7):5648–52. [Google Scholar]
- 66.Lee C, Yang W, Parr RG. Development of the Colle-Salvetti correlation energy formula into a functional of the electron density. Phys Rev B. 1988;37(2):785–9. [DOI] [PubMed] [Google Scholar]
- 67.Perdew JP, Wang Y. Pair-distribution function and its coupling constant average for the spin-polarized electron gas. Phys Rev B. 1992;46(20):12947–54. [DOI] [PubMed] [Google Scholar]
- 68.Dennington R, Keith T, Millam J, Gaussview. version 5. Shawnee Mission, KS: Semichem Inc; 2009.
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
The datasets used and/or analysed during the current study are available from the corresponding author on reasonable request.








