Next Article in Journal
Hybrid Raman and Laser-Induced Breakdown Spectroscopy for Food Authentication Applications
Next Article in Special Issue
Design, Synthesis, Antifungal Activity, and 3D-QSAR Study of Novel Quinoxaline-2-Oxyacetate Hydrazide
Previous Article in Journal
Medical Applications and Advancement of Near Infrared Photosensitive Indocyanine Green Molecules
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Anticancer Evaluation of 4-Chloro-2-((5-aryl-1,3,4-oxadiazol-2-yl)amino)phenol Analogues: An Insight into Experimental and Theoretical Studies

1
Department of Pharmaceutical Chemistry, College of Pharmacy, Prince Sattam Bin Abdulaziz University, Al-Kharj 11942, Saudi Arabia
2
Department of Pharmaceutical Chemistry, College of Pharmacy, Taif University, P.O. Box 11099, Taif 21944, Saudi Arabia
3
Department of Pharmacognosy, College of Pharmacy, Taif University, P.O. Box 11099, Taif 21944, Saudi Arabia
4
Department of Chemistry, College of Science and Humanity Studies, Prince Sattam Bin Abdulaziz University, Al-Kharj 11942, Saudi Arabia
5
Department of Pharmaceutical Chemistry, Noida Institute of Engineering and Technology, (Pharmacy Institute), Knowledge Park-2, Greater Noida 201 306, India
6
Department of Chemistry, Bihar National College, Patna 800 004, India
7
Department of Pharmaceutical Chemistry, Maharishi Arvind College of Pharmacy, Jaipur 302 039, India
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(16), 6086; https://doi.org/10.3390/molecules28166086
Submission received: 30 June 2023 / Revised: 6 August 2023 / Accepted: 7 August 2023 / Published: 16 August 2023
(This article belongs to the Special Issue Research and Development of Antibacterial and Antitumor Drugs)

Abstract

:
We report herein the synthesis, docking studies and biological evaluation of a series of new 4-chloro-2-((5-aryl-1,3,4-oxadiazol-2-yl)amino)phenol analogues (6a-h). The new compounds were designed based on the oxadiazole-linked aryl core of tubulin inhibitors of IMC-038525 and IMC-094332, prepared in five steps and further characterized via spectral analyses. The anticancer activity of the compounds was assessed against several cancer cell lines belonging to nine different panels as per National Cancer Institute (NCI US) protocol. 4-Chloro-2-((5-(3,4,5-trimethoxyphenyl)-1,3,4-oxadiazol-2-yl)amino)phenol (6h) demonstrated significant anticancer activity against SNB-19 (PGI = 65.12), NCI-H460 (PGI = 55.61), and SNB-75 (PGI = 54.68) at 10 µM. The compounds were subjected to molecular docking studies against the active site of the tubulin–combretastatin A4 complex (PDB ID: 5LYJ); they displayed efficient binding and ligand 4h (with docking score = −8.030 kcal/mol) lay within the hydrophobic cavity surrounded by important residues Leu252, Ala250, Leu248, Leu242, Cys241, Val238, Ile318, Ala317, and Ala316. Furthermore, the antibacterial activity of some of the compounds was found to be promising. 4-Chloro-2-((5-(4-nitrophenyl)-1,3,4-oxadiazol-2-yl)amino)phenol (6c) displayed the most promising antibacterial activity against both Gram-negative as well as Gram-positive bacteria with MICs of 8 µg/mL and a zone of inhibition ranging from 17.0 ± 0.40 to 17.0 ± 0.15 mm at 200 µg/mL; however, the standard drug ciprofloxacin exhibited antibacterial activity with MIC values of 4 µg/mL.

1. Introduction

Oxadiazoles are five-membered heterocyclic compounds that contain two nitrogen atoms and one oxygen atom. On the basis of the arrangement of nitrogen and oxygen atoms, there are four different isomers of oxadiazoles [1]. The isomeric 1,3,4-oxadiazole ring is generally advantageous in terms of thermal and metabolic stabilities, water solubility, as well as lipophilicity profiles [2,3]. 1,3,4-Oxadiazoles act as a flat aromatic linker in some cases, giving the molecule a better orientation to bind to biological targets [3]. Many drugs such as raltegravir (anti-HIV) (A), furamizole (antibacterial) (B), tidazosin (α1-adrenergic receptor antagonist) (C), fenadiazole (hypnotic) (D), zibotentan (anticancer) (E), setileuton (MK-0633; LOX inhibitor) (F), nesapidil (vasodialator) (G), and AZD1979 (melanin concentration receptor 1 antagonist) (H) contain a 1,3,4-oxodiazole ring in their structure (Figure 1) [4,5,6,7,8,9,10,11]. The anticancer [12,13,14], antitubercular [15,16], anti-HIV [17], antimicrobial [18], anti-inflammatory [19], anticonvulsant [20], and antimalarial [14] activities, as well as many other biological activities [1,21,22,23,24] reported in the literature, have drawn a lot of attention to 1,3,4-oxadiazole analogues.
Cancer, on the other hand, is one of the most pressing health concerns today, with nearly 10 million deaths and 19.3 million new cases reported worldwide in 2020. The number of new cancer cases is expected to exceed 28.4 million by 2040 [25]. Breast (2.26 million), lung (2.21 million), colon and rectum (1.93 million), and prostate (1.41 million) cancers were the most common cancer types reported in 2020 [26]. Chemotherapy is the most widely used and well-known cancer treatment. It is, however, linked to a number of negative consequences, hence the search for new compounds is always admired. Inspired by the miraculous biological activities of 1,3,4-oxadiazoles, we have synthesized eight new oxadiazole analogues and studied their antiproliferative effect against the nearly five dozen cancer cell lines, as well as their antibacterial potentials.
The oxadiazole-linked aryl core (represented as bold blue bond) of IMC-038525 (tubulin inhibitor; IC50 = 0.39 ± 0.06 μM), IMC-094332 (tubulin inhibitor), and FABT (anticancer; ID50 = 6.2 ± 1.4 (HCV29T), 3.9 ± 1.1 (SW707) and 4.2 ± 1.2 (T47D) µg/mL, respectively) was used in the design strategy for the newer oxadiazole analogues shown in Figure 2 [27,28,29]. Similarly, FTAB is another anticancer compound, whose 1,3,4-thiadiazole ring was replaced with an isostere 1,3,4-oxadiazole in the title compounds (6a-h). Microtubules are highly dynamic structures made up of α- and β-tubulin heterodimers that are indispensable for cell division (mitosis) and growth [30]. The expression of specific tubulin isotypes has been linked to cancer, but the exact mechanism is still unknown. Many tubulin-binding chemotherapeutics target the tubulin family protein to suppress the dynamics of the mitotic spindle to cause mitotic arrest and cell death [31]. Because the synthesized compounds have the same pharmacophoric core as the tubulin inhibitors IMC-038525 and IMC-094332, they can be used as tubulin inhibitors. As a result, a putative mechanism of action of the target compounds (6a-h) as anticancer agents was discovered using a molecular docking approach.
The oxadiazoles were also reported as antibacterial and antifungal agents [16,18,32,33]. Hence, we further explored their antibacterial activity as well as their molecular docking studies against DNA gyrase, a potential target for many antibacterial drugs. Since some of the compounds exhibited promising antibacterial activity, molecular docking against DNA gyrase was studied for the title compounds (6a-h).

2. Results

2.1. Chemistry

The synthesis of 2,5-disubstituted oxadiazoles was accomplished in five steps. The initial step involved the preparation of 4-chloro-2-nitrophenol (2) via the nitration of 4-chlorophenol (0.1 mol; ~12.86 g) with nitric acid (HNO3) and concentrated H2SO4. In the subsequent step, 4-chloro-2-nitrophenol (2) (75 mmol; 13.02 g) was reduced in the presence of Sn/HCl to obtain 2-amino-4-chlorophenol (3). 2-Amino-4-chlorophenol (3) (50 mmol; 7.18 g) was dissolved in 10 mL glacial acetic acid followed by the addition of 40 mL hot water. A solution of sodium cyanate (75 mmol; 4.88 g) in hot water was then mixed to a solution of 2-amino-4-chlorophenol (3) with continuous stirring with a magnetic stirrer for 30 min at room temperature to obtain 1-(5-chloro-2-hydroxyphenyl)urea (4), which was then refluxed with hydrazine hydrate (60 mmol; 3.00 mL) in absolute ethanol for 24 h to obtain N-(5-chloro-2-hydroxyphenyl)hydrazinecarboxamide (5) [34,35]. In the final step, an equimolar mixture of compound (5) (1 mmol; 201 mg) and aromatic aldehyde (1 mmol) was refluxed in water/ethanol solvent (2:1) and 10% mol of sodium bisulphite for 8–10 h to obtain the final product: 4-chloro-2-((5-aryl-1,3,4-oxadiazol-2-yl)amino)phenol analogues (6a-h) [36]. The reaction sequences are summarized in Scheme 1. The final compounds were characterized via spectroscopic techniques (nuclear magnetic resonance and mass spectral analysis). The 1H NMR spectra of the prototype compounds 6a showed two doublets at δ ppm 6.85 and 7.01 for the two aromatic protons (H3 and H4), respectively, while a singlet at δ ppm 7.98 was observed for another proton (H5) of the chlorophenyl ring. Two multiplets at δ ppm 7.13–7.31 (H3’ and H4’) and 7.70–7.75 (H2’ and H6’) were observed for four protons of the 4-fluorophenyl ring. Two singlets at δ ppm 8.76 and 11.05 were observed for the secondary amine (ArNH) and phenolic (ArOH) functions, respectively. The 13C NMR of the compound exhibited twelve peaks at δ ppm 152.60 (oxadiazole C), 148.66 (oxadiazole C), 148.11 (C–F), 144.54 (C−OH), 141.39 (C1), 128.64 (C2’ and C6’), 125.66 (C−Cl), 122.81 (C4), 122.41 (C1’), 117.17 (C6), 115.51 (C3), and 108.54 (C3’ and C5’). The molecular mass (ESI-MS) peak [M + 1]+ was observed at 306.01, while the isotopic peak [M + 3]+ was observed at 308.00. The HPLC chromatogram of one of the compounds (6h) was recorded at λ210 nm and a retention time of 4.19 min demonstrated a purity level of >99% (Please refer to the Supplementary Material, Figure S28).

2.2. Anticancer Activity

All the compounds were designed based on the pharmacophore of tubulin inhibitors, and their anticancer activity was accessed using the National Cancer Institute protocol (NCI US) reported elsewhere [37,38,39,40]. The title compounds were initially screened via one dose assay at 10 µM drug concentration against 56 NCI cancer cell lines, and their results in the form of growth percent (GP) and percent growth inhibition (PGI) are given in Table 1 and Table 2, respectively. The GP and PGI are related as PGI = 100 − GP. Compounds 6b, 6e, and 6g showed maximum sensitivity against PC-3 (PGI = 14.13), NCI-H460 (PGI = 19.55), and NCI-H522 (PGI = 11.4), respectively, while compound 6a was found to be sensitive against NCI-H522 (PGI = 23.32). Compound 6d displayed maximum sensitivity against breast cancer cell line, MCF7 (PGI = 24.79), and MDA-MB-468 (PGI = 26.01), while compound 6f demonstrated maximum sensitivity against UACC-62 (PGI = 21.25), NCI-H522 (PGI = 20.32), and HCT-116 (PGI = 20.15). Compound 6h showed maximum anticancer activity among the series and was found to be sensitive against SNB-19 (PGI = 65.12), NCI-H460 (PGI = 55.61), and SNB-75 (PGI = 54.68). Compound 6h demonstrated <50 to >40 PGI against eight cancer cell lines including NCI/ADR-RES (PGI = 48.28), NCI-H226 (PGI = 42.59), HOP-62 (PGI = 40.28), SF-295 (PGI = 45.26), SF-539 (PGI = 43.79), OVCAR-4 (PGI = 40.03), MDA-MB-231 (PGI = 42.58), and HS 578T (PGI = 47.22), while it demonstrated <40 to >20 PGI against nine cancer cell lines including HCT-116 (PGI = 35.59), H322M (PGI = 33.58), A459 (PGI = 34.85), SF-268 (PGI = 29.96), NCI- ACNH (PGI = 27.99), U251 (PGI = 26.04), SN 12C (PGI = 23.24), NCI-H23 (PGI = 20.82), and 786-O (PGI = 20.42) [Table 2]. The average PGI of individual panels was calculated and compared with the anticancer activity of the reference drug imatinib and their results are given in Table 3. Compound 6h exhibited far better anticancer activity against non-small cancer, colon, CNS, melanoma, ovarian, renal, and breast cancer cell lines. The data for imatinib were retrieved from the NCI website with NSC Code 759854 [37]. The structure–activity relationship was established based on the anticancer activity. The 3,4,5-trimethoxy (6h) substitution of the phenyl ring exhibited promising anticancer activity followed by 2-hydroxy (6f) substitution. The overall structure–activity relationship (SAR) can be represented as 3,4,5-trimethoxy > 2-hydroxy > 4-nitro > 4-hydroxy-3-methoxy substitution in the phenyl ring directly attached to the oxadiazole ring.

2.3. Molecular Docking

Some of the oxadiazole analogues displayed good anticancer activity against some of the cancer cell lines and the compounds were also designed based on the pharmacophore of tubulin inhibitors, hence a molecular docking study of tubulin was explored against the active site of tubulin–combretastatin A4 (PDB ID: 5LYJ). The molecular docking scores and interactions are given in Table 4. The molecular docking score of the ligands (6a-h) ranged from −7.617 to −8.389 kcal/mol and they displayed two types of interactions, including H-bonding and halogen bonding. Ligand 6d displayed maximum binding affinity with a docking score of −8.389 kcal/mol and exhibited a halogen bond interaction with the residue Ala317, while ligand 6e displayed a docking score of −7.617 kcal/mol with no significant interaction. Ligands 6d and 6f displayed a similar type of interaction with the residue Ala317 through a halogen bond interaction. The 2D and 3D interactions of ligands 6d and 6f are shown in Figure 3 and Figure 4, respectively. Ligands 6a and 6b displayed a H-bond interaction with the residue Asn258, while they showed a halogen bond interaction with the residues Ala317 and Asn349, respectively. Ligand 6h displayed no significant electrostatic interaction; however, the trimethoxy phenyl ring lies with the hydrophobic cavity of containing residues Leu252, Ala250, Leu248, Leu242, Cys241, Val238, Ile318, Ala317, and Ala316, as shown in Figure 5.

2.4. Antibacterial Activity

The antibacterial activity of compounds 6a-h against Gram-negative (E. coli, NCIM 5051 and P. aeruginosa, ATCC 10145) as well as Gram-positive (S. aureus, NCIM 2079 and B. subtilis, NCIM 2439) bacterial strains was assessed via a disc diffusion method and broth dilution method as per the reported protocol [41,42,43]. Compounds 6d, 6f, 6g, and 6h displayed less antibacterial activity with zones of inhibition ranging from 4.8 ± 0.75 to 6.7 ± 0.20 mm at 200 µg/mL drug concentrations and exhibited MICs ranging from 128 to 512 µg/mL against various bacterial strains. Compounds 6a and 6d displayed moderate antibacterial activity with MICs ranging between 16 and 64 µg/mL and zones of inhibition ranging between 13.0 ± 0.21 and 9.8 ± 0.25 mm at 200 µg/mL. Compound 6e displayed good antibacterial activity against S. aureus, E. coli, and P. aeruginosa with MICs of 8 µg/mL, while it inhibited B. subtilis with an MIC of 8 µg/mL and a zone of inhibition ranging between 14.0 ± 0.45 and 16.0 ± 0.26 mm at 200 µg/mL. Compound 6c displayed the most promising antibacterial activity against both the Gram-negative as well as Gram-positive bacteria with MICs of 8 µg/mL and zones of inhibition ranging from 17.0 ± 0.40 to 17.0 ± 0.15 mm at 200 µg/mL; however, the standard drug ciprofloxacin exhibited MICs of 8 µg/mL. The antibacterial activity in terms of zones of inhibition and MICs is given in Table 5. The 4-nitro substitution exhibited promising antibacterial activity followed by 2-chloro substitution in the phenyl ring attached to the oxadiazole ring. The SAR of the antibacterial activity can be represented as 4-nitro > 2-chloro > 4-fluoro > 4-chloro substitution in the phenyl ring. Molecular docking studies of the molecular target DNA gyrase were carried out to explore the binding mode of compounds against the 2-oxo-1,2-dihydroquinoline derivative binding active site of DNA gyrase (PDB ID; 6KZV). The results of the molecular docking studies are given in Table S1 and the 2D binding interactions of ligands 6c and 6e are given in Figure S1 (refer to Supplementary Materials). The antibacterial activity was supported by in silico studies, and the compounds (6c and 6e) that had the maximum docking score (of −6.200 and −6.424 kcal/mol, respectively) demonstrated promising antibacterial activity.

3. Discussion

The 1,3,4-oxadiazoles were designed using the pharmacophore found in some tubulin inhibitors (IMC-094322 and IMC-038525) and the anticancer agent FTAB. The current research aimed to create scaffolds with the basic pharmacophore of tubulin inhibitors (IMC-094322 and IMC-038525) and a low-molecular-weight small compound. Molecular docking studies were performed for the tubulin–combretastatin A4 binding site of the D-chain (PDB ID: 5LYJ). We discovered that the benzene ring substitution produced higher docking scores than IMC-094322 and IMC-038525; hence, we performed modifications in the benzene ring rather than the dioxane ring of IMC-094322 and IMC-038525. The oxadiazole analogues were prepared in satisfactory yields and characterized using spectroscopic techniques followed by their anticancer evaluation at 10 µM against nine panel of 56 cancer cell lines as per the NCI US protocol. Some of the compounds displayed promising anticancer activity against a few cancer cell lines. Compound 6h demonstrated significant anticancer activity against SNB-19 (CNS cancer), NCI-H460 (non-small-cell lung cancer), and SNB-75 (CNS cancer) with PGIs of 65.12, 55.61, and 54.68 percent, respectively at 10 µM. Compound 6h also displayed efficient binding against the tubulin–combretastatin A4 binding site of the D-chain with a docking score of −8.030 kcal/mol and a trimethoxy phenyl ring lying within the hydrophobic cavity surrounded by important residues Leu252, Ala250, Leu248, Leu242, Cys241, Val238, Ile318, Ala317, and Ala316. The antibacterial activities of the oxadiazoles were also assessed against Gram-negative (E. coli and P. aeruginosa) as well as Gram-positive (S. aureus and B. subtilis) bacterial strains via a disc diffusion method and broth dilution method as per the reported protocol. Compounds 6c and 6e demonstrated good bacterial inhibition against both the Gram-positive as well as Gram-negative bacterial strains. A molecular docking study was conducted for the title compounds (6a-h) against DNA gyrase, one of the attractive molecular targets for many antibacterial drugs. All the compounds exhibited docking scores ranging between −5.883 and −6.424 kcal/mol. The binding affinities of compounds 6c (docking score = −6.200 kcal/mol) and 6e (docking score = −6.424 kcal/mol) were found to be the highest among the series, and these compounds also possessed promising antibacterial activity.

4. Materials and Methods

4.1. Chemistry

4.1.1. Method for the Synthesis of 4-Chloro-2-nitrophenol (2)

An amount of 21 mL of concentrated H2SO4 and 15 mL of concentrated HNO3 was mixed in a 250 mL round-bottom flask with a few smaller pieces of porcelain. 4-Chlorophenol (100 mmol; 12.86 g) was added in portions of 2–3 g to the acidic mixture, shaking the reaction mixture after each addition. The temperature was controlled to no more than 55 °C by immersing the flask into the cold-water bath. The reaction mixture was then heated to 45 °C for 45 min with continuous shaking on magnetic stirrer [44]. The content of the flask was poured into the crushed ice with continuous stirring with a glass rod to obtain the precipitate. The precipitate was filtered, washed with water, and re-crystallized from methanol to obtain 4-chloro-2-nitrophenol (2). Colour, light yellow; yield, 82%; Mp 86–88 °C.

4.1.2. Method for the Synthesis of 2-Amino-4-chlorophenol (3)

4-Chloro-2-nitrophenol (2) (75 mmol; 13.02 g) was dissolved in concentrated HCl (50 mL) and 45 g granulated tin was added, and the reaction mixture was left to heat up in a boiling water bath (100 °C) for 60 min [44]. The reaction mixture was poured into the crushed ice to obtain solid precipitate, which was filtered, dried, and re-crystallized with absolute ethanol to obtain 2-amino-4-chlorophenol (3). Colour, brown; yield, 78%; Mp 142–144 °C.

4.1.3. Method for the Synthesis of 1-(5-Chloro-2-hydroxyphenyl)urea (4)

2-Amino-4-chlorophenol (3) (50 mmol; 7.18 g) was dissolved in 20 mL glacial acetic acid with heating and 40 mL hot water was added and 40 mL of NaCNO (75 mmol; 4.88 g) was added with continuous stirring at room temperature (25 °C) for 30 min. It was then allowed to stand for another 30 min and then dipped into cold-water bath to obtain precipitate, which was then filtered, dried, and re-crystallized with hot water to obtain light brown coloured precipitate of 1-(5-chloro-2-hydroxyphenyl)urea (4) [34,35]. Colour, light brown; yield, 88%; Mp 183–185 °C.

4.1.4. Method for the Synthesis of N-(5-Chloro-2-hydroxyphenyl)hydrazinecarboxamide (5)

1-(5-chloro-2-hydroxyphenyl)urea (4) (40 mmol; 7.44 g) was refluxed at 80 °C with hydrazine hydrate (60 mmol; 3.00 mL) in absolute ethanol for 24 h to obtain N-(5-chloro-2-hydroxyphenyl)hydrazinecarboxamide (5) [34,35]. Colour, creamy; yield, 85%; Mp 202–204 °C.

4.1.5. General Method for the Synthesis of 4-Chloro-2-((5-aryl-1,3,4-oxadiazol-2-yl)amino)phenol Analogues (6a-h)

An equimolar mixture of N-(5-chloro-2-hydroxyphenyl)hydrazinecarboxamide (5) (1 mmol, 201 mg) and aromatic aldehyde (1 mmol) was dissolved in 5 mL ethanol, then 5 mL water was added and the reaction mixture was allowed to reflux at 100 °C for 8–10 h with an addition of 10% mol NaHSO3 (5 mL). The reaction mixture was then concentrated and poured into the crushed ice, filtered, dried, and recrystallized with ethanol to obtain the final product, 4-chloro-2-((5-aryl-1,3,4-oxadiazol-2-yl)amino)phenol analogues (6a-h) [36].
4-Chloro-2-((5-(4-fluorophenyl)-1,3,4-oxadiazol-2-yl)amino)phenol (6a): yield: 78%; Mp 102–104 °C; 1H NMR (300 MHz, DMSO-d6): δ ppm: 6.85 (d, 1H, J = 6.0 Hz, ArH), 7.01 (d, 1H, J = 6.1 Hz, ArH), 7.13–7.31 (m, 2H, ArH), 7.70–7.75 (m, 2H, ArH), 7.89 (s, 1H, ArH), 8.76 (s, 1H, NH), 11.05 (s, 1H, ArOH); 13C NMR (75 MHz, DMSO-d6): δ ppm: 152.60, 148.66, 148.11, 144.54, 141.39, 128.64, 125.66, 122.81, 122.41, 117.17, 115.51, 108.54; Anal. Calc. for C14H9ClFN3O2: C, 55.01; H, 2.97; N, 13.75, found: C, 54.98; H, 2.99; N, 13.70%. ESI-MS m/z = 306.01 (M + 1)+, 308.00 (M + 3)+.
4-Chloro-2-((5-(4-chlorophenyl)-1,3,4-oxadiazol-2-yl)amino)phenol (6b): yield: 72%; Mp 154–156 °C; 1H NMR (300 MHz, DMSO-d6): δ ppm: 6.91 (d, 1H, J = 7.5 Hz, ArH), 7.03 (d, 1H, J = 9.0 Hz, ArH), 7.40 (s, 1H, ArH), 7.57 (d, 2H, J = 7.8 Hz, ArH), 8.11 (d, 2H, 3.6 Hz, ArH), 8.75 (s, 1H, NH), 11.00 (s, 1H, ArOH); 13C NMR (75 MHz, DMSO-d6): δ ppm: 152.59, 148.53, 143.51, 135.69, 134.52, 129.97, 128.99, 127.79, 124.49, 124.21, 118.19, 116.89; Anal. Calc. for C14H9Cl2N3O2: C, 52.20; H, 2.82; N, 13.04, found: C, 52.15; H, 2.80; N, 13.01%. ESI-MS m/z = 322.14 (M + 1)+, 324.12 (M + 3)+.
4-Chloro-2-((5-(4-nitrophenyl)-1,3,4-oxadiaol-2-yl)amino)phenol (6c): yield: 83%; Mp 196–198 °C; 1H NMR (300 MHz, DMSO-d6): δ ppm: 6.86 (d, 1H, J = 6.1 Hz, ArH), 6.91 (d, 1H, J = 6.1 Hz, ArH), 7.92 (d, 2H, J = 6.3 Hz, ArH), 8.08 (s, 1H, ArH), 8.27 (d, 2H, J = 6.9 Hz, ArH), 8.76 (s, 1H, NH), 11.37 (s, 1H, ArOH); 13C NMR (75 MHz, DMSO-d6): δ ppm: 152.12, 147.57, 145.01, 140.47, 139.02, 128.04, 127.37, 124.12, 122.62, 122.06, 118.04, 115.63; Anal. Calc. for C14H9ClFN4O4: C, 50.54; H, 2.73; N, 16.84, found: C, 50.49; H, 2.75; N, 16.81%. ESI-MS m/z = 333.23 (M + 1)+, 335.25 (M + 3)+.
4-Chloro-2-((5-(4-methoxyphenyl)-1,3,4-oxadiazol-2-yl)amino)phenol (6d): yield: 81%; Mp 162–164 °C; 1H NMR (300 MHz, DMSO-d6): δ ppm: 3.83 (s, 3H, OCH3), 6.72 (d, 1H, J = 6.1 Hz, ArH), 6.99 (d, 1H, J = 6.3 Hz, ArH), 7.13 (d, 2H, J = 6.0 Hz, ArH), 7.32 (s, 1H, ArH), 8.09 (d, 2H, J = 6.1 Hz, ArH), 9.81 (s, 1H, NH), 10.42 (s, 1H, ArOH); 13C NMR (75 MHz, DMSO-d6): 152.65, 148.52, 147.66, 142.57, 135.39, 127.24, 124.45, 118.13, 116.99, 115.97, 114.83, 56.18; Anal. Calc. for C15H12ClN3O3: C, 56.70; H, 3.81; N, 13.23, found: C, 56.67; H, 3.79; N, 13.21%. ESI-MS m/z = 318.11 (M + 1)+, 320.10 (M + 3)+.
4-Chloro-2-((5-(2-chlorophenyl)-1,3,4-oxadiazol-2-yl)amino)phenol (6e): yield: 71%; Mp 182–184 °C; 1H NMR (300 MHz, DMSO-d6): δ ppm: 6.84 (d, 1H, J = 6.0 Hz, ArH), 6.86 (d, 1H, J = 6.1 Hz, ArH), 7.41–7.45 (m, 2H, ArH), 7.51–7.54 (m, 2H, ArH), 7.91–7.94 (m, 1H, ArH), 8.10 (s, 1H, ArH), 8.78 (s, 1H, NH), 11.26 (s, 1H, ArOH); 13C NMR (75 MHz, DMSO-d6): 152.63, 148.59, 142.51, 136.92, 132.22, 130.13, 129.31, 128.92, 127.71, 127.34, 124.41, 118.11, 116.91; Anal. Calc. for C14H9Cl2N3O2: C, 52.20; H, 2.82; N, 13.04, found: C, 52.17; H, 2.80; N, 13.00%. ESI-MS m/z = 322.13 (M + 1)+, 324.19 (M + 3)+.
4-Chloro-2-((5-(2-hydroxyphenyl)-1,3,4-oxadiazol-2-yl)amino)phenol (6f): yield: 82%; Mp 168–170 °C; 1H NMR (300 MHz, DMSO-d6): δ ppm: 6.85 (d, 1H, J = 6.0 Hz, ArH), 6.93 (d, 1H, J = 6.1 Hz, ArH), 7.14–7.69 (m, 4H, ArH), 8.14 (s, 1H, ArH), 9.00 (s, 1H, NH), 10.25 (s, 1H, ArOH), 10.97 (s, 1H, ArOH); 13C NMR (75 MHz, DMSO- d6): 152.63, 148.95, 147.14, 142.51, 135.61, 130.11, 127.72, 126.30, 124.42, 121.81, 118.13, 117.82, 116.91, 108.11; Anal. Calc. for C14H10ClN3O3: C, 55.37; H, 3.32; N, 13.84, found: C, 55.34; H, 3.30; N, 13.81%. ESI-MS m/z = 304.23 (M + 1)+, 306.23 (M + 3)+.
4-Chloro-2-((5-(4-hydroxy-3-methoxyphenyl)-1,3,4-oxadiazol-2-yl)amino)phenol (6g): yield: 88%; Mp 112–114 °C; 1H NMR (300 MHz, DMSO-d6): δ ppm: 3.83 (s, 3H, OCH3), 6.82 (d, 1H, J = 6.0 Hz, ArH), 6.85 (d, 1H, J = 6.1 Hz, ArH), 7.31 (s, 2H, J = 6.0 Hz, ArH), 7.86 (s, 1H, ArH), 8.15 (s, 1H, ArH), 9.80 (s, 1H, NH), 10.48 (s, 1H, ArOH), 10.88 (s, 1H, ArOH); 13C NMR (75 MHz, DMSO-d6): 152.59, 148.65, 148.10, 144.53, 141.39, 128.64, 125.67, 122.80, 121.40, 121.23, 117.17, 115.51, 115.47, 108.55, 55.40; Anal. Calc. for C15H12ClN3O4: C, 53.99; H, 3.62; N, 12.59, found: C, 53.87; H, 3.60; N, 12.57%. ESI-MS m/z = 334.09 (M + 1)+, 336.06 (M + 3)+.
4-Chloro-2-((5-(3,4,5-trimethoxyphenyl)-1,3,4-oxadiazol-2-yl)amino)phenol (6h): yield: 89%; Mp 180–182 °C; 1H NMR (300 MHz, DMSO- d6): δ ppm: 3.69 (s, 3H, OCH3), 3.83 (s, 6H, OCH3), 6.87 (d, 1H, J = 6.1 Hz, ArH), 7.03 (d, 1H, J = 6.0 Hz, ArH), 7.47 (s, 1H, ArH), 7.88 (s, 2H, ArH), 8.94 (s, 1H, NH), 11.10 (s, 1H, ArOH); 13C NMR (75 MHz, DMSO- d6): 153.20, 152.42, 144.54, 138.79, 129.72, 128.48, 122.72, 121.46, 117.05, 115.43, 104.61, 103.62, 60.15, 55.82; Anal. Calc. for C17H16ClN3O5: C, 54.05; H, 4.27; N, 11.12, found: C, 54.02; H, 4.25; N, 11.10%. ESI-MS m/z = 378.27 (M + 1)+, 378.28 (M + 3)+. HPLC purity: 99.129%.

4.2. Anticancer Activity

The anticancer activity of the compounds was assessed using the standard NCI US protocol at 10 µM drug concentration [37,38,39,40]. In our earlier published work, we provided a thorough explanation of the experimental process. [45].

4.3. Molecular Docking Studies

The tubulin–combretastatin A4 complex’s X-ray crystallographic structure, with a resolution of 2.40 Å and an R-value of 0.192 (observed), was retrieved from protein data bank (PDB) [46]. The combretastatin A4 binding site of D-chain was used to prepare grid generation and molecular docking as per reported protocol [47].
The E. coli DNA gyrase complex with 2-oxo-1,2-dihydroquinoline derivative’s X-ray crystal structure with a resolution of 2.40 Å, R-value 0.199 (obs.), was retrieved from protein data bank (https://www.rcsb.org/structure/6KZV accessed on 16 March 2023) [48]. The standard molecular docking protocol was used for docking studies [49].

4.4. Antibacterial Activity

The antibacterial activities of the compounds were tested against B. subtilis, E. coli, S. aureus, and P. aeruginosa bacterial strains via disc diffusion method and broth dilution method [41,42,43] (please refer to the Supplementary Materials).

5. Conclusions

We discussed the design, synthesis, and anticancer activity of new 4-chloro-2-((5-aryl-1,3,4-oxadiazol-2-yl)amino)phenol analogues (6a-h) in the current investigation. The compounds were designed based on the pharmacophoric features of tubulin inhibitors (IMC-094322 and IMC-038525). Therefore, the compounds were further subjected to molecular docking studies against the active site of the tubulin–combretastatin A4 complex (PDB ID: 5LYJ); they displayed efficient binding with docking scores ranging between −7.617 and −8.389 kcal/mol and demonstrated two types of binding interactions: H-bond and halogen bond interactions. The compounds’ binding affinities against the active site of the tubulin–combretastatin A4 complex were found to be higher than those of the tubulin inhibitors IMC-094322 and IMC-038525 in the molecular docking studies, with the exception of compound 6e. The compounds were prepared efficiently in five steps in good yields. The anticancer activity of the compounds was evaluated at 10 µM drug concentration as per the standard protocol. 4-Chloro-2-((5-(3,4,5-trimethoxyphenyl)-1,3,4-oxadiazol-2-yl)amino)phenol (6h) demonstrated significant anticancer activity against some of the cancer cell lines, including SNB-19, NCI-H460, and SNB-75 at 10 µM. Compound 6h was discovered to be highly sensitive to the CNS cancer cell line SNB-19 (PGI = 65.12%), but less sensitive to the renal cancer cell line A498 (PGI = −30.65%). The average anticancer activity of compound 6h against nine panels of cancer cell lines was compared with that of the reference drug imatinib at 10 µM. It was discovered that compound 6h displayed significantly greater anticancer activity against non-small cancer, colon, CNS, melanoma, ovarian, renal, and breast cancer cell lines than imatinib. Based on anticancer activity, the SAR was concluded. It was discovered that the electron-donating group, such as 1,3,4-trimethoxy (compound 6h; mean GP = 81.49%) and 2-hydroxy (compound 6f; mean GP = 96.45%) substitution, as well as the electron-withdrawing 4-nitro (compound 6c; mean GP = 96.97%) substitution on the phenyl ring demonstrated significant anticancer activity compared to other substitutions. The 4-chloro (compound 6b; mean GP = 98.2%) substitution showed slightly more anticancer activity than 2-chloro (compound 6e; mean GP = 100.79%) and 4-fluoro (compound 6a; mean GP = 100.91%) substitution on the phenyl ring. Additionally, the 1,3,4-oxadiazoles’ antibacterial activities are well documented in the literature. The antibacterial activity of the compounds was studied against Gram-negative (E. coli and P. aeruginosa) as well as Gram-positive (S. aureus and B. subtilis) bacterial strains via the disc diffusion method and broth dilution method. The compounds 4-chloro-2-((5-(4-nitrophenyl)-1,3,4-oxadiazol-2-yl)amino)phenol (6c) and 4-chloro-2-((5-(2-chlorophenyl)-1,3,4-oxadiazol-2-yl)amino)phenol (6e) displayed the most promising antibacterial activities against both the Gram-negative as well as Gram-positive bacteria. Compound 6c displayed the most significant antibacterial activities with MICs of 8 µg/mL and a zone of inhibition ranging from 17.0 ± 0.40 to 17.0 ± 0.15 mm at 200 µg/mL; however, the standard drug ciprofloxacin exhibited MICs of 4 µg/mL. The electron-withdrawing group, such as 4-nitro and 2-chloro substitutions on the phenyl ring, demonstrated the most promising antibacterial activity among the series; 4-chloro and 4-fluoro substitutions exhibited moderate antibacterial activity, while the electron-releasing group, such as 4-methoxy, 2-hydroxy, 4-hydroxy-3-methoxy, and 1,3,4-trimethoxy substitutions on the phenyl ring, displayed less antibacterial activity. A molecular docking study was performed for the title compounds (6a-h) against DNA gyrase (PDB ID: 6KZV), one of the most appealing targets for many antibacterial agents, and it was discovered that both compounds 6c and 6e had efficient binding affinity with docking scores of −6.200 and −6.424 kcal/mol, respectively. The results of the molecular docking studies revealed a correlation between docking scores and antibacterial activity. We have reported the anticancer and antibacterial activities of 1,3,4-oxadiazoles in the present investigation. The bioactivities of the compounds were found to be significant and some of the in vitro experimental results were found to be in good agreement with in silico studies. The current discovery could pave the way for future breakthroughs in cancer and bacterial disease treatment. The biological activities of these oxadiazoles can be increased through structural modification. In our laboratory, we are currently conducting additional research into the anticancer and antibacterial potential of modified oxadiazoles.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/molecules28166086/s1, Figures S1 and S2: 2D and 3D binding interactions of ligands 6c and 6e against DNA gyrase; Figures S3–S10: Anticancer data of compounds 6a-h against 56 cancer cell lines; Figures S11–S27: 1H NMR, 13C NMR and mass spectra of some of the synthetic compounds (6a-h); Figure S28: HPLC chromatogram of compound 6h; Table S1: The molecular docking studies of oxadiazoles against DNA gyrase (PDB ID: 6KZV).

Author Contributions

Conceptualization, O.A., M.J.A., S. and A.A. (Amena Ali); methodology, M.J.A. and A.A. (Abuzer Ali); software, A.A. (Amena Ali) and M.J.A.; validation, M.J.A. and O.A.; formal analysis, O.A., A.A. (Abuzer Ali) and M.A.B.; investigation, M.J.A. and M.A.B.; resources, A.A. (Abuzer Ali) and M.F.A.; data curation, M.J.A. and M.A.B.; writing—original draft preparation, M.J.A. and A.A. (Amena Ali); writing—review and editing, A.S.A.A., M.J.A., M.F.A., M.A.A. (Manal A. Alossaimi) and M.A.A. (Mubarak A. Alamri); visualization, M.J.A.; supervision, M.J.A.; project administration, M.J.A. and A.A. (Amena Ali); funding acquisition, A.A. (Amena Ali). All authors have read and agreed to the published version of the manuscript.

Funding

Amena Ali would like to acknowledge the Deanship of Scientific Research, Taif University, for funding this work.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The Supplementary Material contains the information that supports the findings of this study.

Acknowledgments

The researchers would like to acknowledge the Deanship of Scientific Research, Taif University, for funding this work.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Khalilullah, H.; Khan, S.; Ahsan, M.J.; Ahmed, B. 1,3,4-Oxadiazole: A Biologically Active Scaffold. Mini-Rev. Med. Chem. 2012, 12, 789–801. [Google Scholar] [CrossRef] [PubMed]
  2. Ahsan, M.J. 1,3,4-Oxadiazole containing compounds as therapeutic targets for cancer therapy. Mini-Rev. Med. Chem. 2022, 22, 144–197. [Google Scholar] [CrossRef] [PubMed]
  3. Boström, J.; Hogner, A.; Llinàs, A.; Wellner, E.; Plowright, A.T. Oxadiazoles in Medicinal Chemistry. J. Med. Chem. 2012, 55, 1817–1830. [Google Scholar] [CrossRef]
  4. Cocohoba, J.; Dong, B.J. Raltegravir: The First HIV Integrase Inhibitor. Clin. Ther. 2008, 30, 1747–1765. [Google Scholar] [CrossRef]
  5. Vardan, S.; Smulyan, H.; Mookherjee, S.; Eich, R. Effects of tiodazosin, a new antihypertensive, hemodynamics and clinical variables. Clin. Pharm. Ther. 1983, 34, 290–296. [Google Scholar] [CrossRef]
  6. Schlecker, R.; Thieme, P.C. The synthesis of antihypertensive 3-(1,3,4-oxadiazol-2-yl)phenoxypropanolahines. Tetrahedron 1988, 44, 3289–3294. [Google Scholar] [CrossRef]
  7. Ogata, M.; Atobe, H.; Kushida, H.; Yamamoto, K. In Vitro Sensitivity of Mycoplasmas Isolated From Various Animals and Sewage to Antibiotics and Nitrofurans. J. Antibiot. 1971, 24, 443–451. [Google Scholar] [CrossRef] [PubMed]
  8. James, N.D.; Growcott, J.W. Zibotentan. Drugs Fut. 2009, 34, 624–633. [Google Scholar] [CrossRef]
  9. Ducharme, Y.; Blouin, M.; Brideau, C.; Chateauneuf, A.; Gareau, Y.; Grimm, E.L.; Juteau, H.; Laliberte, S.; MacKay, B.; Masse, F.; et al. The Discovery of Setileuton, a Potent and Selective 5-Lipoxygenase Inhibitor. ACS Med. Chem. Lett. 2010, 1, 170–174. [Google Scholar] [CrossRef]
  10. Yarmohammadi, E.; Beyzaei, H.; Aryan, R.; Moradi, A. Ultrasound-assisted, low-solvent and acid/base-free synthesis of 5-substituted 1,3,4-oxadiazole-2-thiols as potent antimicrobial and antioxidant agents. Mol. Divers. 2021, 25, 2367–2378. [Google Scholar] [CrossRef]
  11. Johansson, A.; Löfberg, C.; Antonsson, M.; Von Unge, S.; Hayes, M.A.; Judkins, R.; Ploj, K.; Benthem, L.; Lindén, D.; Brodin, P. Discovery of (3-(4-(2-Oxa-6-Azaspiro[3.3]Heptan-6-Ylmethyl)Phenoxy)Azetidin-1-Yl)(5-(4-Methoxyphenyl)-1,3,4-Oxadiazol-2-Yl)Methanone (AZD1979), a melanin concentrating hormone receptor 1 (MCHr1) antagonist with favourable physicochemical properties. J. Med. Chem. 2016, 59, 2497–2511. [Google Scholar] [CrossRef] [PubMed]
  12. Yousef, T.A.; Alhamzani, A.G.; Abou-Krisha, M.A.; Kanthimathi, G.; Raghu, M.S.; Kumar, K.Y.; Prashanth, M.K.; Jeon, B.H. Synthesis, molecular docking study and anticancer activity of novel 1,3,4-oxadiazole derivatives as potential tubulin inhibitors. Heliyon 2023, 9, e13460. [Google Scholar] [CrossRef]
  13. Nieddu, V.; Pinna, G.; Marchesi, I.; Sanna, L.; Asproni, B.; Pinna, G.A.; Bagella, L.; Murineddu, G. Synthesis and Antineoplastic Evaluation of Novel Unsymmetrical 1,3,4-Oxadiazoles. J. Med. Chem. 2016, 59, 10451–10469. [Google Scholar] [CrossRef]
  14. Verma, G.; Chashoo, G.; Ali, A.; Khan, M.F.; Akhtar, W.; Ali, I.; Akhtar, M.; Alam, M.M.; Shaquiquzzaman, M. Synthesis of pyrazole acrylic acid based oxadiazole and amide derivatives as antimalarial and anticancer agents. Bioorg. Chem. 2018, 77, 106–124. [Google Scholar] [CrossRef] [PubMed]
  15. Karabanovich, G.; Zemanová, J.; Smutný, T.; Székely, R.; Šarkan, M.; Centárová, I.; Vocat, A.; Pávková, I.; Čonka, P.; Němeček, J.; et al. Development of 3,5-Dinitrobenzylsulfanyl-1,3,4-oxadiazoles and thiadiazoles as selective antitubercular agents active against replicating and nonreplicating Mycobacterium tuberculosis. J. Med. Chem. 2016, 59, 2362–2380. [Google Scholar] [CrossRef]
  16. Irfan, A.; Faisal, S.; Zahoor, A.F.; Noreen, R.; Al-Hussain, S.A.; Tuzun, B.; Javaid, R.; Elhenawy, A.A.; Zaki, M.E.A.; Ahmad, S.; et al. In Silico Development of Novel Benzofuran-1,3,4-Oxadiazoles as Lead Inhibitors of M. tuberculosis Polyketide Synthase 13. Pharmaceuticals 2023, 16, 829. [Google Scholar] [CrossRef]
  17. Khan, M.H.; Hameed, S.; Akhtar, T.; Al-Masoudi, W.A.; Jones, P.G.; Pannecouque, C. Synthesis, crystal structure, and molecular docking study of new thiazdiazole and thiazole analogues as potential anti-HIV agents. Med. Chem. Res. 2016, 25, 2399–2409. [Google Scholar] [CrossRef]
  18. Khalilullah, H.; Shamshir, K.M.; Shivli, N.; Bahar, A. Synthesis, characterization and antimicrobial activity of benzodioxane ring containing 1,3,4-oxadiazole derivatives. Arab. J. Chem. 2016, 2, 1029–1035. [Google Scholar] [CrossRef]
  19. Rana, S.M.; Islam, M.; Saeed, H.; Rafique, H.; Majid, M.; Aqeel, M.T.; Imtiaz, F.; Ashraf, Z. Synthesis, Computational Studies, Antioxidant and Anti-Inflammatory Bio-Evaluation of 2,5-Disubstituted-1,3,4-Oxadiazole Derivatives. Pharmaceuticals 2023, 16, 1045. [Google Scholar] [CrossRef]
  20. Tabatabai, S.A.; Lashkari, S.B.; Zarrindast, M.R.; Golibeikian, M.; Shafiee, A. Design, Synthesis and Anticonvulsant Activity of 2-(2-Phenoxy)phenyl-1,3,4-oxadiazole Derivatives. Iranian J. Pharm. Sci. 2013, 12, 105–111. [Google Scholar]
  21. Vaidya, A.; Jain, S.; Jain, P.; Jain, P.; Tiwari, N.; Jain, R.; Jain, R.; Jain, A.K.; Agrawal, R.K. Synthesis and Biological Activities of Oxadiazole Derivatives: A Review. Mini-Rev. Med. Chem. 2016, 16, 825–845. [Google Scholar] [CrossRef]
  22. Salahuddin; Mazumdar, A.; ShaharYar, M.; Mazumdar, R.; Chakraborthy, G.S.; Ahsan, M.J.; Rahman, M. Updates on synthesis and biological activities of 1,3,4-oxadiazole: A review. Synth. Comm. 2017, 47, 1805–1847. [Google Scholar] [CrossRef]
  23. Verma, G.; Khan, M.F.; Akhtar, W.; Alam, M.M.; Akhter, M.; Shquiquzzaman, M. A Review Exploring Therapeutic Worth of 1,3,4-Oxadiazole Tailored Compounds. Mini-Rev. Med. Chem. 2019, 6, 477–509. [Google Scholar] [CrossRef] [PubMed]
  24. El-Masry, R.M.; Kadry, H.H.; Taher, A.T.; Abou-Seri, S.M. Comparative Study of the Synthetic Approaches and Biological Activities of the Bioisosteres of 1,3,4-Oxadiazoles and 1,3,4-Thiadiazoles over the Past Decade. Molecules 2022, 27, 2709. [Google Scholar] [CrossRef]
  25. Sung, H.; Ferley, J.; Siegel, R.L.; Laversanne, M.; Soerjomartaram, I.; Jemal, A.; Bray, F. Global cancer statistics 2020: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef] [PubMed]
  26. WHO. WHO Cancer Reports, 2020; WHO: Geneva, Switzerland, 2020; ISBN 978-92-4-000129-9. [Google Scholar]
  27. Tuma, M.C.; Malikzay, A.; Ouyang, X.; Surguladze, D.; Fleming, J.; Mitelman, S.; Camara, M.; Finnerty, B.; Doody, J.; Chekler, E.L.P.; et al. Antitumor activity of IMC-038525, a novel oral tubulin polymerization inhibitor. Trans. Oncol. 2010, 3, 318–325. [Google Scholar] [CrossRef]
  28. Ouyang, X.; Chen, X.; Piatnitski, E.L.; Kiselyov, A.S.; He, H.Y.; Mao, Y.; Pattaropong, V.; Yu, Y.; Kim, K.H.; Kincaid, J.; et al. Synthesis and structure-activity relationships of 1,2,4-triazoles as a novel class of potent tubulin polymerization inhibitors. Bioorg. Med. Chem. Lett. 2005, 15, 5154–5159. [Google Scholar] [CrossRef] [PubMed]
  29. Juszczak, M.; Matysiak, J.; Szeliga, M.; Pozarowski, P.; Niewiadomy, A.; Albrecht, J.; Rzeski, W. 2-Amino-1,3,4-thiadiazole derivative (FABT) inhibits extracellular signal-regulated kinase pathway and induces cell cycle arrest in human non-small lung carcinoma cells. Bioorg. Med. Chem. Lett. 2012, 22, 5466–5469. [Google Scholar] [CrossRef] [PubMed]
  30. Parker, A.L.; Kavallaris, M.; McCarroll, J.A. Microtubules and their role in cellular stress in cancer. Front. Oncol. 2014, 4, 153. [Google Scholar] [CrossRef] [PubMed]
  31. Binarová, P.; Tuszynski, J. Tubulin: Structure, Functions and Roles in Disease. Cells 2019, 8, 1294. [Google Scholar] [CrossRef]
  32. Dawbaa, S.; Nuha, D.; Evren, A.E.; Cankilic, M.Y.; Yurttas, L.; Turan, G. New oxadiazole/triazole derivatives with antimicrobial and antioxidant properties. J. Mol. Struct. 2023, 1282, 135213. [Google Scholar] [CrossRef]
  33. Salahuddin; Mazumder, A.; Shaharyar, M. Synthesis, antibacterial and anticancer evaluation of 5-substituted (1,3,4-oxadiazol-2-yl)quinoline. Med. Chem. Res. 2015, 24, 2514–2528. [Google Scholar] [CrossRef]
  34. Ahsan, M.J. Rationale design, synthesis and anticancer activity of 2,5-disubstituted-1,3,4-oxadiazole analogues. ChemistrySelect 2016, 1, 4713–4720. [Google Scholar] [CrossRef]
  35. Agarwal, M.; Singh, V.; Sharma, S.K.; Sharma, P.; Ansari, M.Y.; Jadav, S.S.; Yasmin, S.; Sreenivasulu, R.; Hassan, M.Z.; Saini, V.; et al. Design and synthesis of new 2,5-disubstituted-1,3,4- oxadiazole analogues as anticancer agents. Med. Chem. Res. 2016, 25, 2289–2303. [Google Scholar] [CrossRef]
  36. Sangshetti, J.N.; Chabukswar, A.R.; Shinde, B. Microwave assisted one pot synthesis of some novel 2,5-disubstituted 1,3,4-oxadiazoles as antifungal agents. Bioorg. Med. Chem. Lett. 2011, 21, 444–448. [Google Scholar] [CrossRef]
  37. Development Therapeutic Program NCI/NIH (2014). Available online: http://dtp.nci.nih.gov (accessed on 20 January 2022).
  38. Boyd, M.R.; Paull, K.D. Some practical considerations and applications of the National Cancer Institute in vitro anticancer drug discovery screen. Drug Dev. Res. 1995, 34, 91–109. [Google Scholar] [CrossRef]
  39. Monks, A.; Scudiero, D.; Skehan, P.; Shoemaker, R.; Paull, K.; Vistica, D.; Hose, C.; Langley, J.; Cronise, P.; Vaigro-Wolff, A.; et al. Feasibility of a highflux anticancer drug screening using a diverse panel of cultured human tumor cell lines. J. Nat. Cancer Inst. 1991, 83, 757–766. [Google Scholar] [CrossRef]
  40. Shoemaker, R.H. The NCI60 human tumour cell line anticancer drug screen. Nat. Rev. Cancer 2006, 6, 813–823. [Google Scholar] [CrossRef]
  41. Collins, A.H. Microbiological Methods, 2nd ed.; Butterworth: London, UK, 1976. [Google Scholar]
  42. Cruickshank, R.; Duguid, J.P.; Marmion, B.P.; Swain, R.H.A. Medicinal Microbiology, 12th ed.; Churchill Livingstone: London, UK, 1975; Volume 2, p. 196. [Google Scholar]
  43. Ahsan, M.J.; Ali, A.; Ali, A.; Afzal, O.; Salahuddin; Yusuf, M.; Altamimi, A.S.A.; Sharma, O.; Alossaimi, M.A.; Bakht, M.A. Synthesis of New 4’-(Substituted phenyl)spiro[indoline-3,3’-[1,2,4]triazolidine]-2,5’-diones as Antimicrobial, Antitubercular, and Antifungal Agents: An Insight into the ADME and Toxicity Prediction as well as in-silico Molecular Docking Studies. J. Mol. Struct. 2023, 1290, 135846. [Google Scholar] [CrossRef]
  44. Furniss, B.S.; Hannaford, A.J.; Smith, P.W.G.; Tatchell, A.R. Vogels’ Textbook of Practical Organic Chemistry, 5th ed.; Longman Group UK Limited: London, UK, 1989. [Google Scholar]
  45. Ali, A.; Ali, A.; Tahir, A.; Bakht, M.A.; Salahuddin; Ahsan, M.J. Molecular Engineering of Curcumin, an Active Constituent of Curcuma longa L. (Turmeric) of the Family Zingiberaceae with Improved Antiproliferative Activity. Plants 2021, 10, 1559. [Google Scholar] [CrossRef] [PubMed]
  46. X-Ray Crystallographic Structure of Tubulin-Combrestatin A4 Complex. Available online: https://www.rcsb.org/structure/5lyj (accessed on 12 January 2023).
  47. Gaspari, R.; Prota, A.E.; Bargsten, K.; Cavalli, A.; Steinmetz, M.O. Structural Basis of cis- and trans-Combretastatin Binding to Tubulin. Chem 2017, 2, 102–113. [Google Scholar] [CrossRef]
  48. X-Ray Crystal of Cytochrome Bacterial DNA Gyrase (PDB ID: 6KZV). Available online: https://www.rcsb.org/structure/6KZV (accessed on 16 March 2023).
  49. Ushiyama, F.; Amada, H.; Takeuchi, T.; Tanaka-Yamamoto, N.; Kanazawa, H.; Nakano, K.; Mima, M.; Masuko, A.; Takata, I.; Hitaka, K.; et al. Lead Identification of 8-(Methylamino)-2-oxo-1,2-dihydroquinoline Derivatives as DNA Gyrase Inhibitors: Hit-to-Lead Generation Involving Thermodynamic Evaluation. ACS Omega 2020, 5, 10145–10159. [Google Scholar] [CrossRef] [PubMed]
Figure 1. 1,3,4-Oxadiazoles containing drugs.
Figure 1. 1,3,4-Oxadiazoles containing drugs.
Molecules 28 06086 g001
Figure 2. Design of new 2,5-disubstituted-1,3,4-oxadiazoles.
Figure 2. Design of new 2,5-disubstituted-1,3,4-oxadiazoles.
Molecules 28 06086 g002
Scheme 1. Reagents and conditions: (i) HNO3, H2SO4, 45 °C, 1 h, 82% yield; (ii) Sn, HCl, 100 °C, 1 h, 78%, yield; (iii) NaCNO, CH3COOH; 25 °C, 30 min, 88% yield; (iv) NH2NH2.H2O, C2H5OH, 80 °C, 24 h, 85% yield; (v) ArCHO, NaHSO3, C2H5OH-H2O (1:2), 100 °C, 8–10 h, 71–89% yield.
Scheme 1. Reagents and conditions: (i) HNO3, H2SO4, 45 °C, 1 h, 82% yield; (ii) Sn, HCl, 100 °C, 1 h, 78%, yield; (iii) NaCNO, CH3COOH; 25 °C, 30 min, 88% yield; (iv) NH2NH2.H2O, C2H5OH, 80 °C, 24 h, 85% yield; (v) ArCHO, NaHSO3, C2H5OH-H2O (1:2), 100 °C, 8–10 h, 71–89% yield.
Molecules 28 06086 sch001
Figure 3. Two-dimensional binding interaction of ligands 6d and 6f against the active site of tubulin–combretastatin A4 (PDB ID: 5LYJ).
Figure 3. Two-dimensional binding interaction of ligands 6d and 6f against the active site of tubulin–combretastatin A4 (PDB ID: 5LYJ).
Molecules 28 06086 g003
Figure 4. Three-dimensional binding interaction of ligands 6d (violet) and 6f (green)with the active site of tubulin–combretastatin A4 (PDB ID: 5LYJ).
Figure 4. Three-dimensional binding interaction of ligands 6d (violet) and 6f (green)with the active site of tubulin–combretastatin A4 (PDB ID: 5LYJ).
Molecules 28 06086 g004
Figure 5. Two-dimensional binding interaction of ligand 6h with the active site of tubulin–combretastatin A4 (PDB ID: 5LYJ).
Figure 5. Two-dimensional binding interaction of ligand 6h with the active site of tubulin–combretastatin A4 (PDB ID: 5LYJ).
Molecules 28 06086 g005
Table 1. The anticancer activity of oxadiazoles at 10 µM in terms of growth percent (GP).
Table 1. The anticancer activity of oxadiazoles at 10 µM in terms of growth percent (GP).
PanelCell Lines6a6b6c6d6e6f6g6h
LeukemiaCCRF-CEM95.3190.6197.4198.199.5298.7683.8495.14
HL-60(TB)91.9496.58103.0398.47103.499.51119.06111.48
K-56293.0191.6196.7494.0597.9692.24100.4495.61
MOLT-493.2394.59100.1595.6496.0696.1991.33103.1
RPMI-822692.597.0891.8392.2792.8493.7692.45102.57
SR95.7390.5393.798.0493.42102.11101.2698.87
Non-small cell lung cancerA549/ATCC94.3887.1592.8893.3794.3589.3793.0365.15
EKVX102.369491.2599.04105.196.0691.1382
HOP-62105.99102.7497.05101.95105.83104.4294.459.72
HOP-92120.92119.71100.76117.37117.06119.62119.37115.91
NCI-H226105.72107.8399.0897.66109.2899.2593.8657.41
NCI-H23101.3697.2189.12100.3896.6795.1393.6679.18
NCI-H322M99.7998.1597.07101.97101.5398.01102.3966.42
NCI-H460104.48101.02104.92105.2580.4599.18102.1644.39
NCI-H52276.6886.1187.3385.9482.2779.6888.682.71
Colon cancerHCC-299899104.62105.4112.79106.1295.16107.1786
HCT-116104.58103.4399.699.1789.7279.8593.3964.41
HCT-1595.6795.2593.2495.6291.2887.9795.7893.96
HT2999.6392.6198.93101.0798.0388.79102.9189.66
KM1299.899.7699.97100.2198.1499.38100.0599.39
SW-620103.1895.56101.61106.5696.9998.1597.4687.4
CNS cancerSF-268101.69103.2398.799.96102.5995.496.3270.04
SF-29594.7494.5493.9495.8697.9790.9496.154.74
SF-53996.75102.6794.3697.0997.6199.3596.0356.21
SNB-1996.8695.1793.4995.69101.7896.14101.7334.88
SNB-75106104.388.1110.01102.5385.2595.745.32
U251108.42101.95106.8101.81110.3898.86101.2973.96
MelanomaLOX IMVI105.5696.6392.71105.8195.8899.5993.2393.26
MALME-3M99.5694.6995.3198.6799.0396.8591.7487.03
M14101.31101.56104.02108.7296.97102.72101.42102.06
MDA-MB-43598.992.79103.33103.9695.5997.75108.7489.29
SK-MEL-28103.13107.19101.02105.24109.05101.02118.1199.89
SK-MEL-598.1899.3497.5197.81100.1896.24100.7594.83
UACC-25797.6787.4895.1102.94102.2495.0789.7489.39
UACC-628686.9887.0686.489.9378.7583.795.39
Ovarian cancerIGROV1106.78110.0198.18103.66112.2999.07111.3895.15
OVCAR-3108.43106.72105.56108.82109.51107.84113.1193.72
OVCAR-4106.2392.66105.72112.65107.8798.4697.7759.97
OVCAR-596.14100.9395.0596.83103.3895.3597.8699.22
NCI/ADR-RES100.2794.74100.36103.6999.898.1297.9351.72
SK-OV-3117.64102.6299.89112.57115.3105.56104.7181.81
Renal cancer786–094.9598.08100.57100.28108.0892.93111.4579.58
A498113.51103.52117.23121.53127.23104.1130.65101.82
ACHN106.84107.22103.61105.68101.8106.06103.1572.01
CAKI-1103.1295.6487.1194.9998.3988.8992.0181.17
RXF 393112.18107.36108.62112.07117.594.92106.8590.15
SN 12C92.5996.7991.49101.31101.8297.7395.376.76
UO-3192.0493.5385.6592.2104.9284.5381.9982.82
Prostate cancerPC-394.3685.8789.6690.9792.4588.8782.2189.9
DU-145104.69102.62106.63106.77101.93104.47109.3187.7
Breast cancerMCF788.7583.9265.5482.1975.2185.9684.2381.18
MDA-MB-231107.8697.2595.3193.6104.7999.0290.4957.42
HS 578T111.42100.53100.25105.9109.39103.8593.6552.78
BT-549115.53112.05108.47111.75101.27104.77111.8698
T-47D106.36102.0895.87111.22102.1897.9683.0381.98
MDA-MB-468101.4290.6473.9987.291.04100.03108.1982.26
Mean Growth Percent100.9198.296.97101.09100.7996.4598.8581.49
Table 2. The comparative anticancer activity of oxadiazoles on percent growth inhibition (PGI) scale.
Table 2. The comparative anticancer activity of oxadiazoles on percent growth inhibition (PGI) scale.
CompoundsCell Line (Percent Growth Inhibition; % GI)
69.99 to 50.0049.99 to 40.0039.99 to 20.0019.99 to 10.00
6aNCI-H522 (23.32)CCRF-CEM (16.16)
UACC-62 (14)
MCF7 (11.25)
6bMCF7 (16.08)
PC-3 (14.13)
NCI-H522 (13.89)
UACC-62 (13.02)
A549 (12.85)
UACC-257 (12.52)
6cMCF7 (34.46)UO-31 (14.35)
UACC-62 (12.94)
CAKI-1 (12.89)
NCI-H522 (12.67)
NCI-H23 (10.88)
PC-3 (10.34)
6dMDA-MB-468 (26.01)
MCF7 (24.79)
MCF7 (17.81)
NCI-H522 (14.06)
UACC-62 (13.06)
MDA-MB-468 (12.08)
6eNCI-H460 (19.55)
NCI-H522 (17.73)
HCT-116 (10.28)
UACC-62 (10.07)
6fUACC-62 (21.25)
NCI-H522 (20.32)
HCT-116 (20.15)
UO-31 (15.47)
SNB-75 (14.75)
MCF7 (14.04)
HCT-15 (12.03)
HT29 (11.21)
PC-3 (11.13)
CAKI-1 (11.11)
A549 (10.63)
6gUO-31 (18.01)
PC-3 (17.79)
T-47D (16.97)
UACC-62 (16.3)
MCF7 (15.77)
NCI-H522 (11.4)
UACC-257 (10.26)
6hSNB-19 (65.12)
NCI-H460 (55.61)
SNB-75 (54.68)
NCI/ADR-RES (48.28)
HS 578T (47.22)
SF-295 (45.26)
SF-539 (43.79)
NCI-H226 (42.59)
MDA-MB-231 (42.58)
HOP-62 (40.28)
OVCAR-4 (40.03)
A459 (34.85)
HCT-116 (35.59)
NCI-H322M (33.58)
SF-268 (29.96)
ACNH (27.99)
U251 (26.04)
SN 12C (23.24)
NCI-H23 (20.82)
786-O (20.42)
CAKI-1 (18.83)
MCF7 (18.82)
SK-OV-3 (18.19)
T-47D (18.02)
MDA-MB-468 (17.74)
NCI-H522 (17.29)
UO-31 (17.18)
HCC-2998 (14.0)
MALME-3M (12.97)
DU-145 (12.3)
MDA-MB-435 (10.71)
UACC-257 (10.61)
HT29 (10.34)
PC-3 (10.1)
(−) None of the cell lines demonstrated this level of growth inhibition at 10 µM.
Table 3. The comparative anticancer activity of oxadiazoles and imatinib at 10 µM in terms of average PGI.
Table 3. The comparative anticancer activity of oxadiazoles and imatinib at 10 µM in terms of average PGI.
Panels6a6b6c6d6e6f6g6hImatinib
Leukemia6.386.52.863.912.82.911.94−1.139
Non-Small Cancer Cell−1.290.684.50−0.330.832.142.3827.4615.68
Colon Cancer−0.311.460.21−2.573.298.450.5413.195.34
CNS Cancer−1.290.684.50−0.330.832.142.3827.465.8
Melanoma1.214.172.99−1.191.394.001.576.11−0.87
Ovarian Cancer−5.92−1.28−0.79−6.37−8.03−0.73−3.7919.74−7.16
Renal Cancer−2.18−0.310.82−4.01−8.534.41−3.0616.533.25
Prostate Cancer0.485.761.861.132.813.334.2411.212.5
Breast Cancer−5.222.2610.091.362.691.404.7624.3912.15
Table 4. The molecular docking studies of oxadiazoles against tubulin.
Table 4. The molecular docking studies of oxadiazoles against tubulin.
S. No.CompoundPDB ID: 5LYJ
Docking ScoreEmodel ScoreInteraction
16a−8.233−68.717H-bond (Asn258, 2.52 Å); halogen bond (Ala317, 3.17 Å)
26b−8.247−71.236H-bond (Asn258, 2.52 Å); halogen bond (Asn349, 3.43 Å)
36c−8.356−74.181H-bond (Asn258, 2.58 Å)
46d−8.389−73.546Halogen bond (Ala317, 3.35 Å)
56e−7.617−63.671
66f−8.073−71.277Halogen bond (Ala317, 3.13 Å)
76g−8.266−74.933H-bond (Asn349, 2.39 Å); halogen bond (Ala317, 3.12 Å)
86h−8.030−71.237
9IMC-094322−7.669−42.387
10IMC-038525−7.941−52.924π-Cationic (Lys352, 6.46 Å)
(−) In molecular docking, no electrostatic interaction was seen.
Table 5. The antibacterial activity of oxadiazoles.
Table 5. The antibacterial activity of oxadiazoles.
S. No.CompoundZone of Inhibition in mm at 200 µg/mLMinimum Inhibitory Concentration (µg/mL)
S. aureusB. subtilisE. coliP. aeruginosaS. aureusB. subtilisE. coliP. aeruginosa
16a12.0 ± 0.3611.0 ± 0.2513.0 ± 0.2112.0 ± 0.2616321616
26b11.0 ± 0.659.9 ± 0.3011.0 ± 0.369.8 ± 0.2532643264
36c17.0 ± 0.4017.0 ± 0.2117.0 ± 0.1517.0 ± 0.218888
46d5.5 ± 0.504.8 ± 0.755.6 ± 0.325.1 ± 0.15256512256256
56e16.0 ± 0.4914.0 ± 0.4516.0 ± 0.2616.0 ± 0.4083288
66f5.2 ± 0.255.0 ± 0.105.4 ± 0.215.8 ± 0.10256256256256
76g5.6 ± 0.355.0 ± 0.235.8 ± 0.305.6 ± 0.21256256256256
86h6.7 ± 0.205.7 ± 0.296.0 ± 0.216.0 ± 0.23128256128128
9Ciprofloxacin18.0 ± 0.6517.0 ± 0.2618.0 ± 0.3218.0 ± 0.314444
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

Afzal, O.; Ali, A.; Ali, A.; Altamimi, A.S.A.; Alossaimi, M.A.; Bakht, M.A.; Salahuddin; Alamri, M.A.; Ahsan, M.F.; Ahsan, M.J. Synthesis and Anticancer Evaluation of 4-Chloro-2-((5-aryl-1,3,4-oxadiazol-2-yl)amino)phenol Analogues: An Insight into Experimental and Theoretical Studies. Molecules 2023, 28, 6086. https://doi.org/10.3390/molecules28166086

AMA Style

Afzal O, Ali A, Ali A, Altamimi ASA, Alossaimi MA, Bakht MA, Salahuddin, Alamri MA, Ahsan MF, Ahsan MJ. Synthesis and Anticancer Evaluation of 4-Chloro-2-((5-aryl-1,3,4-oxadiazol-2-yl)amino)phenol Analogues: An Insight into Experimental and Theoretical Studies. Molecules. 2023; 28(16):6086. https://doi.org/10.3390/molecules28166086

Chicago/Turabian Style

Afzal, Obaid, Amena Ali, Abuzer Ali, Abdulmalik Saleh Alfawaz Altamimi, Manal A. Alossaimi, Md Afroz Bakht, Salahuddin, Mubarak A. Alamri, Md. Faiyaz Ahsan, and Mohamed Jawed Ahsan. 2023. "Synthesis and Anticancer Evaluation of 4-Chloro-2-((5-aryl-1,3,4-oxadiazol-2-yl)amino)phenol Analogues: An Insight into Experimental and Theoretical Studies" Molecules 28, no. 16: 6086. https://doi.org/10.3390/molecules28166086

Article Metrics

Back to TopTop