Next Article in Journal
Betulonic Acid, as One of the Active Components of the Celastrus orbiculatus Extract, Inhibits the Invasion and Metastasis of Gastric Cancer Cells by Mediating Cytoskeleton Rearrangement In Vitro
Previous Article in Journal
Molecular Mechanisms of Amylin Turnover, Misfolding and Toxicity in the Pancreas
Previous Article in Special Issue
Novel Pyridothienopyrimidine Derivatives: Design, Synthesis and Biological Evaluation as Antimicrobial and Anticancer Agents
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Exploring the Synergistic Anticancer Potential of Benzofuran–Oxadiazoles and Triazoles: Improved Ultrasound- and Microwave-Assisted Synthesis, Molecular Docking, Hemolytic, Thrombolytic and Anticancer Evaluation of Furan-Based Molecules

1
Department of Chemistry, Government College University Faisalabad, Faisalabad 38000, Pakistan
2
Department of Zoology, Government College University Faisalabad, Faisalabad 38000, Pakistan
3
Department of Pharmaceutical Chemistry, Riphah International University, Islamabad 44000, Pakistan
4
Department of Histology, Embryology and Cytophysiology, Medical University of Lublin, Radziwiłłowska 11, 20-080 Lublin, Poland
5
Department of Chemistry, Siedlce University of Natural Sciences and Humanities, 3-Go Maja 54, 08-110 Siedlce, Poland
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(3), 1023; https://doi.org/10.3390/molecules27031023
Submission received: 26 December 2021 / Revised: 27 January 2022 / Accepted: 27 January 2022 / Published: 2 February 2022
(This article belongs to the Special Issue Heterocyclic Compounds: Design, Synthesis, and Applications)

Abstract

:
Ultrasound- and microwave-assisted green synthetic strategies were applied to furnish benzofuran–oxadiazole 5ag and benzofuran–triazole 7ah derivatives in good to excellent yields (60–96%), in comparison with conventional methods (36–80% yield). These synthesized derivatives were screened for hemolysis, thrombolysis and anticancer therapeutic potential against an A549 lung cancer cell line using an MTT assay. Derivatives 7b (0.1%) and 5e (0.5%) showed the least toxicity against RBCs. Hybrid 7f showed excellent thrombolysis activity (61.4%) when compared against reference ABTS. The highest anticancer activity was displayed by the 5d structural hybridwith cell viability 27.49 ± 1.90 and IC50 6.3 ± 0.7 μM values, which were considerably lower than the reference drug crizotinib (IC50 8.54 ± 0.84 μM). Conformational analysis revealed the spatial arrangement of compound 5d, which demonstrated its significant potency in comparison with crizotinib; therefore, scaffold 5d would be a promising anticancer agent on the basis of cytotoxicity studies, as well as in silico modeling studies.

1. Introduction

In recent years, natural- and synthetic furan-based (Figure 1) based chemotherapeutic agents have attracted the attention of medicinal chemists to develop novel pharmacological agents with diverse pharmacophores andhave expanded the scope and treatment of various diseases [1,2,3].
Heterocyclic benzofurans are the basic structural units of biologically active families of natural products such as egonol, homoegonol and moracin [4]. Natural and synthetic furan derivatives exhibit profound physiological and chemotherapeutic potential against a wide variety of pathogens by displaying antimicrobial, antioxidant [5], antitubulin [6], anti-inflammatory [7], antiviral [8], antihyperglycemic [9], analgesic [10,11], anticancer [11,12], antifungal [13] and antipyretic activities [14]. The therapeutic profile of furan-based molecules has attracted medicinal researchers to design and develop anticancer agents which could play a pivotal role in cancer therapy [1,15,16]. In addition to this, oxadiazole and triazole ring systems, which are part of many natural products, have become the focus of pharmacologists and medicinal chemists on account of their medicinal and pharmacology activities, especially in the field of oncology [17,18].
Cancer is a notably complex, prominent and lethal disease which poses a serious human health problem; ithasresulted in 7.6 million deaths globally, and this number is expected to reach 13 million by 2030 [19,20,21]. Naturally occurring benzofuran derivatives such as denthyrsin and HIBE (1-(6-hydroxy-2-isopropenyl-1-benzofuran-5-yl)-1-ethanone) displayed cytotoxic potential against ovarian cancer and human breast cancer cell lines [22,23]. Sulfonamide scaffolds of benzofuran–imidazopyridines showed anticancer activity against ovarian cancer, MCF-7, lung cancer and colon cancer [24]. Benzofuran–oxadiazole hybrids exhibited significant cytotoxicity against pancreatic cancer and colon cancer cell lines in vitro [25]. The 2-aryl benzo furan–appended 4-aminoquinazoline structural motifs significantly inhibited the epidermal growth factor receptortyrosine kinase phosphorylation in vitro against human lung cancer, cervical cancer, colorectal adenocarcinoma and hepatocellular carcinoma cell lines [26]. The benzofuran–carboxylic acid hybrids proved to be the best antiproliferative agents against breast cancer cell lines and significant inhibitors of carbonic anhydrases [27]. Thiazolodin-4-one benzofuran derivatives exhibited remarkable antitumor activity against human HEPG2 cell lines [28]. Different furan-based anticancer agents are listed in Figure 2 [1].
Continuing our previous research work towards the synthesis of benzofuran-based oxadiazole and triazole hybrids using a conventional approach [29], here, our group applied green synthetic strategies such as ultrasonic- and microwave-assisted synthetic approaches to achieve environmentally friendly synthesis with maximum production of benzofuran–oxadiazole and triazolestructural hybrids over shorter reaction times. The synthesized molecules were further evaluated for their hemolytic, thrombolytic and anticancer potential.

2. Materials and Methods

2.1. Chemistry

Ultrasonic irradiation and microwave-assisted experiments were performed, respectively, in an ultrasonic cleaner bath (Model 1510, 115 v, 1.9 L) with a mechanical timer and heater switch and at a frequency of 47 kHz, and with microwave apparatus (Model EA-180M) witha frequency of 2450 Hz and power consumption of 1150 watt. Analytical-grade solvents and reagents were used; these were purchased from Merck, Alfa Aesar and Sigma-Aldrich through local suppliers. The reaction progress was monitored using pre-coated silica gel plates. The synthesized scaffolds were purified using the column chromatographic technique and by the process of recrystallization in ethanol and methanol solvents.

2.2. General Ultrasound- and Microwave-Assisted Synthetic Protocols for Benzofuran–Oxadiazole Hybrids (5ag) and Benzofuran–Triazole Derivatives (7ah)

Method A: Ultrasound-assisted method [30]. The substituted S-alkylated oxadiazole- and triazole-based benzofuran derivatives were afforded by dissolving 5-(benzofuran-2-yl)-1,3,4-oxadiazole-2-thiol 3 (0.03 g, 0.137 mmol) and 5-(benzofuran-2-yl)-4-phenyl-4H-1,2,4-triazole-3-thiol 6 (0.03 g, 0.103 mmol) inacetonitrile (15 mL). Pyridine (0.213 mmol) was added, and the reaction mixture was stirred for 15 min at 0 °C. The substituted bromoacetanilides 4ag (0.24 mmol) were added, and the reaction mixture was sonicated at 40 °C for 30 min, as shown in Scheme 1. The reaction was monitored viathin-layer chromatography. On completion of the reaction, petroleum ether was added to the mixture with continuous stirring to obtain the final products in the form of precipitates, which were filtered, washed with distilled water and purified.
Method B: Microwave-assisted method [31,32,33]. Benzofuran–oxadiazole hybrid 3 (0.03 g, 0.137 mmol) and 5-(benzofuran-2-yl)-4-phenyl-4H-1,2,4-triazole-3-thiol 6 (0.03 g, 0.103 mmol) were dissolved in DMF (25 mL). Pyridine (0.213 mmol) was added to the reaction mixtureand stirred for 15 min at 0 °C. The substituted bromoacetanilide derivatives 4ag (0.24 mmol) werethen added, and the reaction mixture was irradiated in a microwave oven for 60–70s, respectively, as depicted in Scheme 1. After completion of the reaction, petroleum ether was added to the reaction mixture with continuous stirring to obtain the final products in the form of precipitates. The precipitates were filtered, washed with distilled waterand purified by recrystallization.The advantages of these preparatory protocols are simplicity, very short reaction times, generality and the elaboration of substituted benzofuran–oxadiazole and benzofuran–triazole with high to excellent yields compared toconventional synthetic approaches and microwave methods already cited for generally synthetic approaches for oxadiazole and triazole derivatives, and specifically benzofuran–oxadiazole and benzofuran–triazole scaffolds [29,34,35,36,37,38,39].

2.3. Biological Evaluation

Oxadiazole- and triazole-based benzofuran derivatives were evaluated for hemolysis, thrombolysis and anticancer activity.

2.3.1. Hemolysis Assay

A fresh blood sample (5 mL) from a healthy donor was collected in an EDTA tube. The blood was transferred to microcentrifuge tubes and centrifugedfor 5 min at 1000 rpm to obtain the red blood cells (RBCs). The supernatant was then discarded, and the RBCpellet was washed three times with phosphate-buffered saline (PBS). The RBC pellet was collected after washing and 20µL of sample solution in DMSO was added. The tubes were incubated at 37 °C for 60 min. Having been removed from the incubator, the tubes were recentrifuged at 13,000 rpm for 5 min. The supernatant was collected and diluted with chilled PBS solution. Absorbance was recordedat 517 nm. ABTS was used as a positive control, while DMSO was employed as a negative control in this protocol. The experiment wasconducted in triplicate, and the RBC lysis percentage was calculated using this formula [40]:
%   Age   hemolysis   = Absorbance   of   sample Absorbance   of   negative   control   × 100 Absorbance   of   positive   control

2.3.2. Thrombolysis Assay

The thrombolytic assay was performedaccording tomethodologyin the literature [41]. A blood sample (3 mL) was collected from a healthy human donor, and 500 µL was transferred to pre-weighed, clean Eppendorf tubes. The tubes containing the blood wereweighed again andincubatedfor 1 h at 37 °C to induce clot formation. The serum was then discarded, and the tube containing theclot was weighed. An amount of 40 µL of sample solution in DMSO was added to the clot.The tubes were again incubated for 3 h at 37 °C, and the lysis resultswere observed. ABTS was used as a positivecontrol, while DMSO was used as a negative control in this assay. The experiment was performed in triplicate, and the lysis percentage was calculated using the formula below:
%   Age   clot   lysis   = Initial   clot   weight Final   clot   weight   × 100 Initial   clot   weight

2.3.3. MTT Assay

Preparation of Cell Culture

The human lung cancer cell line A549 was cultured in Dulbecco’s modified Eagle medium (DMEM), composed of 10% FBS (fetal bovine serum), 100 units/mL of penicillin (1%) and 100 μg/mL of streptomycin (1%), with 5% CO2 (carbon dioxide) at 37 °C in a moistened atmosphere. The A549 cell line was treated with synthesized compounds in a final concentration of DMSO (dimethyl sulfoxide) of less than 1%.

Determination of Cell Viability

The cytotoxic therapeutic potential of synthesized structural hybrids was evaluated usinga standard MTT (3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide) assay. The cells of theA549 cell line were grown, overnight, in a microculture of 96-well plates. The A549 cells were supplemented with different concentrations of compounds for 48 h and further incubated with 20 μL of MTT mixture (5 mg/mL) for 4 h at 37 °C. Following this, 150 μL of control DMSO was added to formazan crystals.The percentage of cell viability was determined viaquantified absorbance in a microplate readerat 490 nm wavelength [42].

2.3.4. Molecular Docking Studies

The compound 5d with a higher pharmacological therapeutic potential was subjected to in silico studies to delineate the mechanism of anticancer activity against lung cancer A549cell lines. The molecular docking studies of these synthesized compounds at a targeted protein were carried out using auto dock Vina 1.1.2 software intermitted with PyRx. The runs of docking were achieved with grid box coordinates of −7.65, −28.27 and 43.34 for x, y and z, respectively. Considering the important role of signaling parameters in cancer, receptors of ALK (anaplastic lymphoma kinase) present an established target to evaluate the anticancer therapeutic potential of synthesized compounds [43,44]. To accomplish computational studies against this target, the crystallographic structure of anaplastic lymphoma kinase in conjunctionwith crizotinib having PDB ID: 2XP2 was obtained from the RCSB protein data bank; it was saved in Pdb formatafter the removal of water molecules and the co-crystallization ofthe ligand with the addition of polar hydrogen. The structure of the co-crystallized ligand crizotinib has been nominated as a potential therapeutic moiety against ALK targets;the structure of the synthesized compound 5d was selected because of its marked experimental results. The structures were drawn on Chemdraw 20 professional software (chemoffice) and saved as a mol file [45,46]. Following this, energy minimization and polar hydrogen addition were performed, and structures were saved in Pdb format through Bio via Discovery Studio Visualizer (DSV), as shown in Figure 3. The results in the form of pictures were processed through the latest versions of Pymol and Ligplot plus software.

2.4. Statistical Data

The statistical data analysis was performed by using prism. All the measurements were carried out in triplicate, and the results are depicted as mean ±SD.

3. Results and Discussions

3.1. Chemistry

3.1.1. Ultrasound- and Microwave-Assisted Synthesis of Oxadiazole-Based Benzofuran Derivatives (5ag)

Ultrasound- and microwave-assisted green reactions are environmentally friendly, time saving, economical and good yield methodologies, and were adopted as depicted in Scheme 1. The conventional approach of synthesizing benzofuran–oxadiazole was already reported [29] as affording compounds in moderate to good (36–80%) yields but required a longer period of time to complete the reaction. This study was initiated to increase the yield and reduce the time of synthesized derivatives by applying ultrasound- and microwave-assisted approaches to achieve benzofuran–oxadiazole derivatives. In the ultrasound- and microwave-assisted approaches, the scaffold 5-(benzofuran-2-yl)-1,3,4-oxadiazole-2-thiol 3 was coupled with substituted bromoacetanilide derivatives 4ah to synthesize corresponding S-alkylated products of benzofuran–oxadiazole 5ag in good to excellent yields, as depicted in Table 1. The ultrasound-assisted strategy yielded products of 60–86% in 30 min at 40 °C, compared withthe conventional method, which afforded products of 36–80% in 24 h. The microwave approach was much more efficient in that the maximum yield of products (69–94%) was obtained in the very short duration of 60 s. The benzofuran–oxadiazole derivative 5d was achieved withthe maximum yield, while the derivative 5f was obtained withthe least yield via ultrasonication and microwave irradiation approaches, as mentioned in Table 1.

3.1.2. Ultrasound- and Microwave-Assisted Synthesis of Triazole-Based Benzofuran Derivatives (7ah)

Benzofuran–triazole derivatives were synthesized via green synthetic methodologies, namelyultrasound- and microwave-assisted protocols. Benzofuran–triazole derivatives (7ah) were obtained in low to moderate (38–79%) yields in 36 h by utilizing conventional methodology [28], while the yield was increased (60–90%) in the ultrasound-assisted synthetic protocol, in which the scaffold 5-(benzofuran-2-yl)-4-phenyl-4H-1,2,4-triazole-3-thiol 6 was coupled with substituted bromoacetanilide derivatives 4ah to synthesize corresponding S-alkylated benzofuran–triazole derivatives 7ah in good to excellent yield at 40 °Cwithin 30 min. Similarly, the microwave approach was even more efficient in that the S-alkylated benzofuran–triazole structural motifs were obtained in maximum yield (68–96%) within the very short duration of 70 s.The benzofuran–triazole derivative 7h was achieved in the maximum yield (90% and 96%) in ultrasound- and microwave-assisted synthetic approaches, while the derivative 7c was obtained in the least yield(60% and 68%) in ultrasound-assisted andmicrowaveirradiation approaches, as shown in Table 2.

3.2. Hemolytic Activity

Benzofuran–triazole derivative 7b exhibited the least cytotoxicity (0.1%) among all the synthesized derivatives, whereas benzofuran–triazole derivative 7g (23.4%) and benzofuran–oxadiazole 5b (22.12%) showed the highest toxicity, as depicted in Table 3. Benzofuran-oxadiazole 5d (5.02%), 5g (4.86%) andbenzofuran–triazole scaffolds such as 7a, 7d and 7f displayed moderate results (6.13–15.7%). Benzofuran-oxadiazole hybrids 5a, 5c, 5e and 5f showed 0.5–3.7% hemolysis. The data from Table 3 also demonstrate thatbenzofuran–triazole derivatives 7b (0.1%), 7c (2.15%) and 7e (3.11%) showed the least cytotoxicity among all the benzofuran–triazole derivatives against the positive control. The results demonstrate that the nature and position of the functional groups have a significant effect on the hemolytic potential of synthesized derivatives.

3.3. Thrombolytic Activity

All the synthesized analogues were evaluated for thrombolytic potential. Most of the benzofuran–oxadiazole and benzofuran–triazole derivatives displayed a mild to moderate thrombolytic effect when compared with the positive control ABTS. The benzofuran-triazole derivative 7f showed the highest and best thrombolysis (61.4%) among all the synthesized hybrids when compared against the reference ABTS. The benzofuran–oxadiazole derivative 5a exhibited significantly good lysis activity (56.8%) among all the benzofuran–oxadiazole scaffolds. The benzofuran–oxadiazole derivative 5g and benzofuran–triazole derivative 7h showed the least lysis (48.3% and 48.1%), respectively. Other derivatives such as 5b (50.7%), 5c (52.8%), 5d (53.5%), 5e (52.4%), 5f (56.5%), 7a (52.2%), 7b (52.5%), 7c (54.0%), 7d (56.2%), 7e (59.1%) and 7g (49.06%) showed a mild thrombolytic effect, as described in Table 3. Furthermore, it wasfound that functional groups with a negative inductive effect have a positive influence on thrombolysis.

3.4. Anticancer Activity

The cytotoxic perspective of all the synthesized target derivatives 5ag and 7ah was evaluated against lung cancer cell line A549 in comparison with reference standard drugs crizotinib [47,48,49,50,51,52,53,54] and cisplatin [55,56,57,58,59,60] by determining cell viability usingan MTT assay, as shown in Table 3. The best and most noteworthy cytotoxicity was displayed by benzofuran–oxadiazole hybrid 5d, with cell viability 27.49 ± 1.90 and IC50 6.3 ± 0.7 μM, which was lower than the other derivatives. The benzofuran–triazole scaffold 7h had slightly less cytotoxic potential, with a cell viability of 29.29 ± 3.98 and IC50 10.9 ± 0.94 μM compared with the derivative 5d. The benzofuran–oxadiazole and triazole derivatives 5e, 7d and 7g, with cell viabilities 34.47 ± 2.19, 39.12 ± 2.21 and 36.26 ± 0.41, respectively, displayed good inhibition potential against the A549 cancercell line. The benzofuran structural hybrids 5b, 5c, 5f, 5g, 7a, 7c and 7e (cell viability = 45.99 ± 4.22, 43.7 ± 0.94, 43.67 ± 4.43, 41.45 ± 4.10, 49.8 ± 1.06, 44.52 ± 5.01 and 44.72 ± 0.84) showed moderate anticancer therapeutic potential against lung cancer cells. The target benzofuran derivatives 5a and 7b exhibited the least cytotoxic potential, with cell viability 64.1 ± 1.72 and 57.62 ± 4.94, respectively. The benzofuran derivative 7f was found to be inactive against the A549 lung cancer cell line with a high value of cell viability (99.1 ± 5.04), as depicted in Table 3.

3.5. Structure–Activity Relationship (SAR)

The structure–activity relationship ofhemolysis, thrombolysis and the anticancer potential of benzofuran–oxadiazole and triazole hybrids can be interpreted on the basis of substitution patterns present on the phenyl ring of N-(substituted-phenyl)-acetamide. The electron-donating ethoxy group atthe para position of the phenyl ring in benzofuran–oxadiazole compound 5e showed 0.5% toxicity against RBCs, whereas benzofuran–triazole derivative 7b exhibited 0.1% toxicity, which was significantly lower than the other derivatives. On the other hand, compounds 5b and 7g (22.12% and 23.4%) proved to be the most toxic compounds among the benzofuran–oxadiazole and benzofuran–triazole series, respectively (Figure 4a).
In the case of thrombolytic activity, the benzofuran–oxadiazole scaffold 5a with unsubstituted phenyl displayed better thrombolytic potential (56.8%) compared withdiethyl amine moiety containing the 5g benzofuran–oxadiazole hybrid. The phenyl group increased the thrombolytic potential of 5a, while the diethylamine resulted in a decrease in thrombolytic activity (48.3%) of the 5g hybrid. The benzofuran–triazole hybrid 7f was the most active thrombolytic agent among all the screened derivatives of benzofuran–oxadiazole and benzofuran–triazole, displaying the highest thrombolysis potential (61.4%) due to the presence of two methyl groups adjacently attached on the phenyl ring at ortho para positions. The benzofuran derivative 7h showed the least thrombolytic potential (48.1%) compared with all thetested compounds. The thrombolysis activity of 7h was significantly decreased due to the presence of two chloro EWD groups adjacently attached on the phenyl ring at the ortho and para positions, as depicted in Figure 4b.
The benzofuran–oxadiazole derivative 5d exhibited excellent anticancer activity (cell viability 27.49 ± 1.90% and IC50 6.3 ± 0.7 μM) compared withall the screened derivatives, due to the presence of a methoxy group at the meta position on the phenyl ring;because of the presence of the ethoxy group at the para position on the phenyl ring, however, the benzofuran–oxadiazole derivative 5e displayedmoderate anticancer activity (cell viability 34.47 ± 2.19% and IC50 17.9 ± 0.46 μM) in comparison with the reference drugs crizotinib and cisplatin (28.22 ± 3.88 and 15.34 ± 2.98 μM), respectively. The structural hybrid 5d is more potent than crizitonib and showed less cytotoxic potential than cisplatin. Due to the presence of two chloro EWD groups adjacently attached on the phenyl ring at the ortho and para positions, benzofuran–triazole scaffold 7h exhibited slightly less potency (cell viability 29.29 ± 3.98% and IC50 10.9 ± 0.94 μM) against lung cancer cell line A549 compared with 5d scaffold, which was the most active compound among all the screened derivatives. The benzofuran–triazole hybrid 7f was an inactive compound (cell viability 99.1%) against the A549 cell line due to the presence of two methyl groups adjacently attached on the phenyl ring at the ortho and para positions (Figure 4c).

3.6. Computational Modeling Studies of the Most Active Compound 5d

The synthesized compound 5d (B) and co-crystallized ligand crizotinib were docked against the anaplastic lymphoma kinase (ALK) receptors to understand better the mechanisms of attachment and inhibition. The compound-docked conformations were interpreted by binding interactions, energies, polar bonding and ligand–receptor binding.
The molecular comprehension of docking experimentssuggest that compound 5d (B) is more prominent as it produces excellent results with binding energies of −9.925 Kcal/mol in the best binding mode. This exceeds the average energy threshold of −8.985 Kcal/mol given by crizotinib, thereby commending the improved affinity of synthesized compound 5d (B) against the ALK receptor (Table 4).
The active binding site of an ALK receptor prominently consists of ALA: 1252, ARG: 1253, ASN: 1254, CYS: 1255, LEU: 1256, LEU: 1257, THR: 1258, CYS: 1259 and PRO: 1260. Crizotinib was found to have noticeable contacts with these active site residues, but compound 5d (B) displayed more effective binding at these active amino acid AA residues, representing more effective working of this synthesized chemical moiety. The binding of compound 5d (B) and crizotinib with receptors are elaborated in Figure 5. The phenyl and heterocyclic rings of compound 5d (B) showed π–sigma interaction with LEU A: 1256 and VAL A: 1130, while the NH group gave conventional hydrogen bonds with GLY A: 1201. Another prominent interaction appeared with GLU A: 1210 through π–anion linkage.
Compound 5d(B) was docked to determine its interaction with the subject protein and exhibited interesting results. It gave the negative binding energies of −9.925 Kcal/mol in its best binding mode. The prominent amino acid residues were VAL A: 1130, ALA A: 1148, GLY A: 1201, ASP A: 1203, GLU A: 1210, LEU A: 1256 and PRO A: 1260, as shown in Figure 6.
Evaluation through docking studies provides good evidence that synthesized compound 5d presents excellent interaction inside the binding pocket of a receptor, showing its better anticancer potential as compared withthe co-crystallized ligand crizotinib. These computational studies are reliable and appropriate for performing binding interaction prediction. The conformational analysis revealed thespatial arrangement of compound 5d, justifying it to be more potent, in comparison with crizotinib. Interestingly, compound 5d interactions were more realistic, with better results. Therefore, scaffold 5d would be a promising anticancer agent on the basis of both cytotoxicity and in silico studies.

4. Conclusions

The target S-alkylated structural motifs of benzofuran–oxadiazole and –triazole were synthesized in good to excellent yields by utilizing ultrasound- and microwave-assisted green synthetic protocols. The compounds were evaluated for anticancer, hemolytic and thrombolytic activities, and it wasrevealed that most of tested derivatives displayed significant hemolytic, thrombolytic and anticancer activities. The hemolysis assay indicated that benzofuran–triazole derivative 7b and benzofuran–oxadiazole 5e proved to be the least toxic compounds among all the synthesized derivatives. The benzofuran–oxadiazole hybrid 5d with cell viability 27.49 ± 1.90 and IC50 (6.3 ± 0.7 μM) values was recognized as the most potent anticancer candidate against lung cancer cell line A549 compared with the reference drugs crizotinib (IC50 8.54 ± 0.84 μM) and cisplatin with (IC50 3.88 ± 0.76 μM). The docking studies explored the mode of action and inhibition mechanism of compound 5d againstcancer cells; the results of the comprehensive docking analysis of compound 5d are consistent and in agreement with biological diagnostic findings. Overall, the current studies suggest that benzofuran-linked oxadiazole and triazole hybrids are capable of being established as lead compounds in cancer therapy; in particular, scaffold 5d could be a promising anticancer agent on the basis of cytotoxicity andin silico studies. The SAR studies suggested that more modifications to oxadiazole and triazole derivatives of benzofuran may lead to advanced anticancer candidates in cancer therapy.

Author Contributions

Conceptualization: A.F.Z.; biological research: A.R.; docking studies; R.Z.; synthesis: A.I.; literature collection and analysis: S.F.; writing—original draft preparation: A.I. and S.F.; writing—review and editing: K.K.-M. and M.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not Applicable.

Informed Consent Statement

Not Applicable.

Data Availability Statement

Not Applicable.

Acknowledgments

The authors are thankful for and acknowledge the Government College University Faisalabad, Pakistan for providing research facilities to carry out this research.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Napiórkowska, M.; Cieślak, M.; Barańska, K.J.; Golińska, K.K.; Nawrot, B. Synthesis of new derivatives of benzofuran as potential anticancer agents. Molecules 2019, 24, 1529. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Aslam, S.N.; Stevenson, P.C.; Phythian, S.J.; Veitch, N.C.; Hall, D.R. Synthesis of cicerfuran, an antifungal benzofuran, and some related analogues. Tetrahedron 2006, 62, 4214–4226. [Google Scholar] [CrossRef]
  3. Clive, D.L.J.; Stoffman, E.J.L. Total synthesis of (-)-conocarpan and assignment of the absolute configuration by chemical methods. Chem. Commun. 2007, 21, 2151–2153. [Google Scholar] [CrossRef] [PubMed]
  4. Naik, R.; Harmalkar, S.D.; Xu, X.; Jang, K.; Lee, K. Bioactive benzofuran derivatives: Moracins AeZ in medicinal chemistry. Eur. J. Med. Chem. 2015, 90, 379–393. [Google Scholar] [CrossRef] [PubMed]
  5. Rangaswamy, J.; Kumar, H.V.; Harini, S.T.; Naik, N. Functionalized 3-(benzofuran-2-yl)-5-(4-methoxyphenyl)-4,5-dihydro-1H-pyrazole scaffolds: A new class of antimicrobials and antioxidants. Arab. J. Chem. 2017, 10, S2685–S2696. [Google Scholar] [CrossRef] [Green Version]
  6. Romagnoli, R.; Baraldi, G.P.; Sarkar, T.; Cara, L.C.; Lopez, C.O.; Carrion, D.M.; Preti, D.; Tolomeo, M.; Balzarini, J.; Hamel, E. Synthesis and biological evaluation of 2-aroyl-4-phenyl-5- hydroxybenzofurans as a new class of anti-tubulin agents. Med. Chem. 2008, 4, 558–564. [Google Scholar] [CrossRef] [Green Version]
  7. Zeni, G.; Ludtke, D.S.; Nogueira, C.W.; Panatieri, R.B.; Antonio, L.; Braga, A.L.; Silveira, C.C.; Stefanib, A.H.; Rochaa, J.B. New acetylenic furan derivatives: Synthesis and anti-inflammatory activity. Tetrahedron Lett. 2001, 42, 8927–8930. [Google Scholar] [CrossRef]
  8. Matsuya, Y.; Sasaki, K.; Nagaoka, M.; Kakuda, H.; Toyooka, N.; Imanishi, N.; Ochiai, H.; Nemoto, H. Synthesis of a new class of furan-fused tetracyclic compounds using o-quinodimethane chemistry and investigation of their antiviral activity. J. Org. Chem. 2004, 69, 7989–7993. [Google Scholar] [CrossRef]
  9. Singh, F.V.; Chaurasia, S.; Joshi, M.D.; Srivastava, A.K.; Goel, A. Synthesis and in vivo antihyperglycemic activity of nature-mimicking furanyl-2-pyranones in STZ-S model. Bioorg. Med. Chem. Lett. 2007, 17, 2425–2429. [Google Scholar] [CrossRef]
  10. Farag, A.A.; El Shehry, F.M.; Abbas, S.Y.; Abd-Alrahman, S.N.; Atrees, A.A.; Al-basheer, H.Z.; Ammar, Y.A. Synthesis of pyrazoles containing benzofuran and trifluoromethyl moieties as possible anti-inflammatory and analgesic agents. Z. Nat. B 2015, 70, 519–526. [Google Scholar] [CrossRef]
  11. Lu, D.; Zhou, Y.; Li, Q.; Luo, J.; Jiang, Q.; He, B.; Tang, Q. Synthesis, in vitro antitumor activity and molecular mechanism of novel furan derivatives and their precursors. Anticancer Agents Med. Chem. 2020, 20, 1475–1486. [Google Scholar] [CrossRef]
  12. Islam, K.; Pal, K.; Debnath, U.; Basha, R.S.; Khan, A.T.; Jana, K.; Misra, A.K. Anti-cancer potential of (1,2-dihydronaphtho[2,1-b]furan-2-yl)methanone derivatives. Bioorg. Med. Chem. Lett. 2020, 30, 127476. [Google Scholar] [CrossRef] [PubMed]
  13. Tighadouini, S.; Radi, S.; Benabbes, R.; Youssoufi, M.H.; Shityakov, S.; El Massaoudi, M.; Garcia, Y. Synthesis, biochemical characterization, and theoretical studies of novel β-keto-enol pyridine and furan derivatives as potent antifungal agents. ACS Omega 2020, 5, 17743–17752. [Google Scholar] [CrossRef] [PubMed]
  14. Xie, Y.-S.; Kumar, D.; Bodduri, V.V.; Tarani, P.S.; Zhao, B.-X.; Miao, J.-Y.; Jang, K.; Shin, D.-S. Microwave-assisted parallel synthesis of benzofuran-2-carboxamide derivatives bearing anti-inflammatory, analgesic and antipyretic agents. Tetrahedron Lett. 2014, 55, 2796–2800. [Google Scholar] [CrossRef]
  15. Dong, Y.; Shi, Q.; Liu, Y.-N.; Wang, X.; Bastow, K.F.; Lee, K.-H. Antitumor Agents. 266. Design, Synthesis, and Biological Evaluation of Novel 2-(Furan-2-yl)naphthalen-1-ol Derivatives as Potent and Selective Antibreast Cancer Agents. J. Med. Chem. 2009, 52, 3586–3590. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Miao, Y.; Hu, Y.; Yang, J.; Liu, T.; Sun, J.; Wang, X. Natural source, bioactivity and synthesis of benzofuran derivatives. RSC Adv. 2019, 9, 27510–27540. [Google Scholar] [CrossRef] [Green Version]
  17. Xu, Z.; Zhao, S.J.; Liu, Y. 1,2,3-Triazole-containing hybrids as potential anticancer agents: Current developments, action mechanisms and structure-activity relationships. Eur. J. Med. Chem. 2019, 183, 111700. [Google Scholar] [CrossRef]
  18. Kapoor, G.; Bhutani, R.; Pathak, D.P.; Chauhan, G.; Kant, R.; Grover, P.; Siddiqui, S.A. Current Advancement in the Oxadiazole-Based Scaffolds as Anticancer Agents. Polycycl. Aromat. Compd. 2021, in press. [Google Scholar] [CrossRef]
  19. Irfan, A.; Ullah, S.; Anum, A.; Jabeen, N.; Zahoor, F.A.; Kanwal, H.; Kotwica-Mojzych, K.; Mojzych, M. Synthetic transformations and medicinal significance of 1,2,3-thiadiazoles derivatives: An update. Appl. Sci. 2021, 11, 5742. [Google Scholar] [CrossRef]
  20. Shahzadi, I.; Zahoor, A.F.; Rasul, A.; Mansha, A.; Ahmad, S.; Raza, Z. Synthesis, hemolytic studies, and in silico modeling of novel acefylline-1,2,4-triazole hybrids as potential anti-cancer agents against MCF-7 and A549. ACS Omega 2021, 6, 11943–11953. [Google Scholar] [CrossRef]
  21. Irfan, A.; Batool, F.; Naqvi, Z.A.S.; Islam, A.; Osman, M.S.; Nocentini, A.; Alissa, A.S.; Supuran, T.C. Benzothiazole derivatives as anticancer agents. J. Enzyme Inhib. Med. Chem. 2020, 35, 265–279. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Van Miert, S.; Van Dyck, S.; Schmidt, T.J.; Brun, R.; Vlietinck, A.; Lemiere, G.; Pieters, L. Antileishmanial activity, cytotoxicity and QSAR analysis of synthetic dihydrobenzofuran lignans and related benzofurans. Bioorg. Med. Chem. 2005, 13, 661–669. [Google Scholar] [CrossRef] [PubMed]
  23. Zhang, G.N.; Zhong, L.Y.; Bligh, S.W.A.; Guo, Y.L.; Zhang, C.F.; Zhang, M.; Wang, Z.T.; Xu, L.S. Bi-bicyclic and bi-tricyclic compounds from Dendrobiumthyrsiflorum. Phytochemistry 2005, 66, 1113–1120. [Google Scholar] [CrossRef] [PubMed]
  24. Murthy, I.S.; Sireesha, R.; Deepthi, K.; Rao, P.S.; Raju, R.R. Design, synthesis and biological evaluation of sulphonamide derivatives of benzofuran-imidazopyridines as anticancer agents. Chem. Data Collect. 2020, 31, 100608. [Google Scholar] [CrossRef]
  25. Mokenapelli, S.; Thalari, G.; Vadiyaala, N.; Yerrabelli, J.R.; Irlapati, V.K.; Gorityala, N.; Sagurthi, S.R.; Chitneni, P.R. Synthesis, cytotoxicity, and molecular docking of substituted 3-(2-methylbenzofuran-3-yl)-5-(phenoxymethyl)-1,2,4-oxadiazoles. Arch. Pharm. 2020, 353, e2000006. [Google Scholar] [CrossRef]
  26. Mphahlele, M.J.; Maluleka, M.M.; Aro, A.; McGaw, L.J.; Choong, Y.S. Benzofuran-appended 4-aminoquinazoline hybrids as epidermal growth factor receptor tyrosine kinase inhibitors: Synthesis, biological evaluation and molecular docking studies. J. Enzyme Inhib. Med. Chem. 2018, 33, 1516–1528. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Eldehna, W.M.; Nocentini, A.; Elsayed, Z.M.; Al-Warhi, T.; Aljaeed, N.; Alotaibi, O.J.; Al-Sanea, M.M.; Abdel-Aziz, H.A.; Supuran, C.T. Benzofuran-based carboxylic acids as carbonic anhydrase inhibitors and antiproliferative agents against breast cancer. ACS Med. Chem. Lett. 2020, 11, 1022–1027. [Google Scholar] [CrossRef]
  28. Othman, D.I.; Abdelal, A.M.; El-Sayed, M.; El Bialy, S.A.A. Novel benzofuran derivatives: Synthesis and antitumor activity. Heterocycl. Commun. 2013, 19, 29–35. [Google Scholar] [CrossRef]
  29. Faiz, S.; Zahoor, A.F.; Ajmal, M.; Kamal, S.; Ahmad, S.; Abdelgawad, A.M.; Elnaggar, E.M. Design, synthesis, antimicrobial evaluation, and laccase catalysis effect of novel benzofuran–oxadiazole and benzofuran–triazole hybrids. J. Heterocycl. Chem. 2019, 56, 2839–2852. [Google Scholar] [CrossRef]
  30. Shi, Z.; Zhao, Z.; Huang, M.; Fu, X. Ultrasound-assisted, one-pot, three-component synthesis and antibacterial activities of novel indole derivatives containing 1,3,4-oxadiazole and 1,2,4-triazole moieties. C. R. Chim. 2015, 18, 1320–1327. [Google Scholar] [CrossRef]
  31. Virk, N.A.; Rehman, A.; Abbasi, M.A.; Siddiqui, S.Z.; Rashid, U.; Iqbal, J.; Saleem, M.; Ashraf, M.; Shahid, W.; Shah, S.A.A. Conventional versus microwave assisted synthesis, molecular docking and enzyme inhibitory activities of new 3,4,5-trisubstituted-1,2,4-triazole analogues. Pak. J. Pharm. Sci. 2018, 31, 1501–1510. [Google Scholar] [PubMed]
  32. Javid, J.; Rehman, A.; Abbasi, M.A.; Siddiqui, S.Z.; Iqbal, J.; Virk, N.A.; Rasool, S.; Ali, H.A.; Ashraf, M.; Shahid, W.; et al. Comparative conventional and microwave assisted synthesis of heterocyclic oxadiazole analogues having enzymatic inhibition potential. J. Heterocycl. Chem. 2021, 58, 93–110. [Google Scholar] [CrossRef]
  33. Virk, N.A.; Rehman, A.; Abbasi, M.A.; Siddiqui, S.Z.; Iqbal, J.; Rasool, S.; Khan, S.U.; Htar, T.T.; Khalid, H.; Lauloo, S.J.; et al. Microwave-assisted synthesis of triazole derivatives conjugated with piperidine as new anti-enzymatic agents. J. Heterocycl. Chem. 2020, 57, 1387–1402. [Google Scholar] [CrossRef]
  34. Sapkal, S.B.; Shelke, K.F.; Shingate, B.B.; Shingare, M.S. An efficient synthesis of benzofuran derivatives under conventional/non-conventional method. Chin. Chem. Lett. 2010, 21, 1439–1442. [Google Scholar] [CrossRef]
  35. Vani, I.; Sireesha, R.; Mak, K.; Rao, P.M.; Prasad, K.R.S.; Rao, M.V.B. Microwave assisted synthesis and antimicrobial and antioxidant activities of dimers of 1,2,3-triazole-benzofuran bearing alkyl spacer derivatives. Chem. Data Collect. 2020, 31, 100605. [Google Scholar] [CrossRef]
  36. Reddy, E.R.; Ramesh, S.; Anitha, K.; Reddy, A.P.; Reddy, V.P. Microwave-assisted synthesis and antibacterial activity of 1-(5-((2-(4-bromobenzoyl)-3-methylbenzofuran-5-yl)methyl)-2-((1-aryl-1H-1,2,3-triazol-4-yl)methoxy)phenyl)ethanones. Chem. Data Collect. 2021, 34, 100730. [Google Scholar] [CrossRef]
  37. Ashok, D.; Gandhi, D.M.; Srinivas, G.; Kumar, A.V. Microwave-assisted synthesis of novel 1,2,3-triazole derivatives and their antimicrobial activity. Med. Chem. Res. 2014, 23, 3005–3018. [Google Scholar] [CrossRef]
  38. Farshori, N.N.; Rauf, A.; Siddiqui, M.A.; Al-Sheddi, E.S.; Al-Oqail, M.M. A facile one-pot synthesis of novel 2,5-disubstituted-1,3,4-oxadiazoles under conventional and microwave conditions and evaluation of their in vitro antimicrobial activities. Arab. J. Chem. 2017, 10, S2853–S2861. [Google Scholar] [CrossRef] [Green Version]
  39. Jaisankar, K.R.; Kumaran, K.; Kamil, S.R.M.; Srinivasan, T. Microwave-assisted synthesis of 1,2,4-triazole-3-carboxamides from esters and amines under neutral conditions. Res. Chem. Intermed. 2015, 41, 1975–1984. [Google Scholar] [CrossRef] [Green Version]
  40. Riaz, M.; Rasool, N.; Bukhari, I.; Shahid, M.; Zubair, M.; Rizwan, K.; Rashid, U. In vitro antimicrobial, antioxidant, cytotoxicity and GC-MS analysis of Mazus goodenifolius. Molecules 2012, 17, 14275–14287. [Google Scholar] [CrossRef] [Green Version]
  41. Batool, M.; Tajammal, A.; Farhat, F.; Verpoort, F.; Khattak, Z.A.K.; Mehr-un-Nisa, M.S.; Ahmad, H.A.; Munawar, M.A.; Zia-ur-Rehman, M.; Basra, M.A.R. Molecular Docking, Computational, and Antithrombotic Studies of Novel 1, 3, 4-Oxadiazole Derivatives. Int. J. Mol. Sci. 2018, 19, 3606. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Rasul, A.; Di, J.; Millimouno, F.; Malhi, M.; Tsuji, I.; Ali, M.; Li, J.; Li, X. Reactive oxygen species mediate isoalantolactone-induced apoptosis in human prostate cancer cells. Molecules 2013, 18, 9382–9396. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Arun, Y.; Saranraj, K.; Balachandran, C.; Perumal, P.T. Novel spirooxindole-pyrrolidine compounds: Synthesis, anticancer and molecular docking studies. Eur. J. Med. Chem. 2014, 74, 50–64. [Google Scholar] [CrossRef] [PubMed]
  44. Mustafa, M.; Abdelhamid, D.; Abdelhafez, E.M.; Ibrahim, M.A.; Gamal-Eldeen, A.M.; Aly, O.M. Synthesis, antiproliferative, anti-tubulin activity, and docking study of new 1,2,4-triazoles as potential combretastatin analogues. Eur. J. Med. Chem. 2017, 141, 293–305. [Google Scholar] [CrossRef] [PubMed]
  45. Saidi, I.; Nimbarte, V.D.; Schwalbe, H.; Waffo-Téguo, P.; Harrath, A.H.; Mansour, L.; Jannet, H.B. Anti-tyrosinase, anti-cholinesterase and cytotoxic activities of extracts and phytochemicals from the Tunisian Citharexylum spinosum L.: Molecular docking and SAR analysis. Bioorg.Chem. 2020, 102, 104093. [Google Scholar] [CrossRef]
  46. Lakshmanan, S.; Govindaraj, D.; Ramalakshmi, N.; Antony, S.A. Synthesis, molecular docking, DFT calculations and cytotoxicity activity of benzo [g] quinazoline derivatives in choline chloride-urea. J. Mol. Struct. 2017, 1150, 88–95. [Google Scholar] [CrossRef]
  47. Bang, Y.-J. The potential for crizotinib in non-small cell lung cancer: A perspective review. Ther. Adv. Med. Oncol. 2011, 3, 279–291. [Google Scholar] [CrossRef] [Green Version]
  48. Dai, X.; Guo, G.; Zou, P.; Cui, R.; Chen, W.; Chen, X.; Yin, C.; He, W.; Vinothkumar, R.; Yang, F.; et al. (S)-crizotinib induces apoptosis in human non-small cell lung cancer cells by activating ROS independent of MTH1. J. Exp. Clin. Cancer Res. 2017, 36, 120. [Google Scholar] [CrossRef] [PubMed]
  49. Chuang, J.C.; Neal, J.W. Crizotinib as first line therapy for advanced ALK-positive non-small cell lung cancers. Transl. Lung Cancer Res. 2015, 4, 639–641. [Google Scholar] [CrossRef]
  50. Chen, R.L.; Zhao, J.; Zhang, X.C.; Lou, N.N.; Chen, H.J.; Yang, X.; Su, J.; Xei, Z.; Zhou, O.; Tu, H.Y.; et al. Crizotinib in advanced non-small-cell lung cancer with concomitant ALK rearrangement and c-Met overexpression. BMC Cancer 2018, 18, 1171. [Google Scholar] [CrossRef] [Green Version]
  51. Roberts, P.J. Clinical use of crizotinib for the treatment of non-small cell lung cancer. Biol. Targets Ther. 2013, 7, 91–101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Liu, C.; Yu, H.; Long, Q.; Chen, H.; Li, Y.; Zhao, W.; Zhao, K.; Zhu, Z.; Sun, S.; Fan, M.; et al. Real world experience of crizotinib in 104 patients with alk rearrangement non-small-cell lung cancer in a single chinese cancer center. Front. Oncol. 2019, 9, 1116. [Google Scholar] [CrossRef] [PubMed]
  53. Wang, Q.; He, Y.; Yang, X.; Wang, Y.; Xiao, H. Extraordinary response to crizotinib in a woman with squamous cell lung cancer after two courses of failed chemotherapy. BMC Pulm. Med. 2014, 14, 83. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Drilon, A.; Clark, J.W.; Weiss, J.; Ou, S.-H.I.; Camidge, D.R.; Solomon, B.J.; Otterson, G.A.; Villaruz, L.C.; Riely, G.J.; Heist, R.S.; et al. Antitumor activity of crizotinib in lung cancers harboring a MET exon 14 alteration. Nat. Med. 2020, 26, 47–51. [Google Scholar] [CrossRef] [PubMed]
  55. Ma, Y.; Zheng, X.; Gao, H.; Wan, C.; Rao, G.; Mao, Z. Design, synthesis, and biological evaluation of novel benzofuran derivatives bearing n-aryl piperazine moiety. Molecules 2016, 21, 1684. [Google Scholar] [CrossRef]
  56. Tseng, C.-Y.; Lin, C.-H.; Wu, L.-Y.; Wang, J.-S.; Chung, M.-C.; Chang, J.-F.; Chao, M.-W. Potential combinational anti-cancer therapy in non-small cell lung cancer with traditional chinese medicine sun-bai-pi extract and cisplatin. PLoS ONE 2016, 11, e0155469. [Google Scholar] [CrossRef]
  57. Ozturk, O.H.; Bozcuk, H.; Burgucu, D.; Ekinci, D.; Ozdogan, M.; Akca, S.; Yildiz, M. Cisplatin cytotoxicity is enhanced with zoledronic acid in A549 lung cancer cell line: Preliminary results of an in vitro study. Cell Biol. Int. 2007, 31, 1069–1071. [Google Scholar] [CrossRef]
  58. .Ahmed, F.F.; Abd El-Hafeez, A.A.; Abbas, S.H.; Abdelhamid, D.; Abdel-Aziz, M. New 1,2,4-triazole-Chalcone hybrids induce Caspase-3 dependent apoptosis in A549 human lung adenocarcinoma cells. Eur. J. Med. Chem. 2018, 151, 705–722. [Google Scholar] [CrossRef]
  59. Özdemir, A.; Sever, B.; Altıntop, M.D.; Temel, H.E.; Atlı, Ö.; Baysal, M.; Demirci, F. Synthesis and Evaluation of New Oxadiazole, Thiadiazole, and Triazole Derivatives as Potential Anticancer Agents Targeting MMP-9. Molecules 2017, 22, 1109. [Google Scholar] [CrossRef] [Green Version]
  60. Wagner, E.; Wietrzyk, J.; Psurski, M.; Becan, L.; Turlej, E. Synthesis and Anticancer Evaluation of Novel Derivatives of Isoxazolo[4,5-e][1,2,4]triazepine Derivatives and Potential Inhibitors of Protein Kinase C. ACS Omega 2020, 6, 119–134. [Google Scholar] [CrossRef]
Figure 1. Structures of natural and synthetic furan derivatives.
Figure 1. Structures of natural and synthetic furan derivatives.
Molecules 27 01023 g001
Figure 2. Structures of some synthetic anticancer benzofuran derivatives.
Figure 2. Structures of some synthetic anticancer benzofuran derivatives.
Molecules 27 01023 g002
Scheme 1. Synthetic pathway for microwave/ultrasound-assisted synthesis of S-alkylated oxadiazole/triazole-based benzofuran derivatives.
Scheme 1. Synthetic pathway for microwave/ultrasound-assisted synthesis of S-alkylated oxadiazole/triazole-based benzofuran derivatives.
Molecules 27 01023 sch001
Figure 3. (A) 3D model of crizotinib with carbons (gray), oxygen (red), nitrogen (blue), fluorine (sky blue) and chlorine (green) shown; (B) 3D model of compound 5d with carbons (gray), oxygen (red), sulfur (yellow), nitrogen (blue) shown.
Figure 3. (A) 3D model of crizotinib with carbons (gray), oxygen (red), nitrogen (blue), fluorine (sky blue) and chlorine (green) shown; (B) 3D model of compound 5d with carbons (gray), oxygen (red), sulfur (yellow), nitrogen (blue) shown.
Molecules 27 01023 g003
Figure 4. (a) Structure–activity relationship of hemolytic compounds 5b, 5e, 7g and 7b. (b) Structure–activity relationship of thrombolytic compounds 5a, 5g, 7f and 7h. (c) Structure–activity relationship of anticancer 5d, 5e, 7h and 7f derivatives.
Figure 4. (a) Structure–activity relationship of hemolytic compounds 5b, 5e, 7g and 7b. (b) Structure–activity relationship of thrombolytic compounds 5a, 5g, 7f and 7h. (c) Structure–activity relationship of anticancer 5d, 5e, 7h and 7f derivatives.
Molecules 27 01023 g004
Figure 5. (A) 3D Model of crizotinib (brown) inside the bonding pocket of protein with interactions shown; (B) 3D model of compound 5d (pink) with all atoms shown, collaborating with amino acid residues with bond distances displayed.
Figure 5. (A) 3D Model of crizotinib (brown) inside the bonding pocket of protein with interactions shown; (B) 3D model of compound 5d (pink) with all atoms shown, collaborating with amino acid residues with bond distances displayed.
Molecules 27 01023 g005
Figure 6. (A) 2D Model of crizotinib inside the bonding pocket of protein with interactions shown; (B) 2D model of compound 5d (B) with all atoms shown, collaborating with amino acid residues with bond distances displayed.
Figure 6. (A) 2D Model of crizotinib inside the bonding pocket of protein with interactions shown; (B) 2D model of compound 5d (B) with all atoms shown, collaborating with amino acid residues with bond distances displayed.
Molecules 27 01023 g006
Table 1. Yields of benzofuran–oxadiazole derivatives via conventional, ultrasound- and microwave-assisted synthetic approaches.
Table 1. Yields of benzofuran–oxadiazole derivatives via conventional, ultrasound- and microwave-assisted synthetic approaches.
CompoundsR-NHProducts Percentage Yieldsm.p. (°C)
Conventional Approach
(Faiz S. et al.) [29]
Ultrasound-Assisted Approach
(This Work)
Microwave-Assisted Approach
(This Work)
FoundReported
(Faiz S. et al.) [29]
5a Molecules 27 01023 i001637075212212–214
5b Molecules 27 01023 i002536670181–182180–182
5c Molecules 27 01023 i003596875171–173171–173
5d Molecules 27 01023 i004808694177176–178
5e Molecules 27 01023 i005677382154–156155–157
5f Molecules 27 01023 i006366069205204–206
5g Molecules 27 01023 i00749667496–9795–97
Table 2. Yields of benzofuran–triazole derivatives via conventional, ultrasoundand microwave-assisted synthetic approaches.
Table 2. Yields of benzofuran–triazole derivatives via conventional, ultrasoundand microwave-assisted synthetic approaches.
CompoundsR-NHProducts Percentage Yieldsm.p. (°C)
Conventional Approach
(Faiz S. et al.) [29]
Ultrasound-Assisted Approach
(This Work)
Microwave-Assisted Approach
(This Work)
FoundReported
(Faiz S. et al.) [29]
7a Molecules 27 01023 i008738090220–221220–222
7b Molecules 27 01023 i009426473173–175173–175
7c Molecules 27 01023 i010386068189–190188–190
7d Molecules 27 01023 i011426374218–219218–220
7e Molecules 27 01023 i012466977249–251249–251
7f Molecules 27 01023 i013396170226–227226–228
7g Molecules 27 01023 i014647789229–230228–230
7h Molecules 27 01023 i015799096217–219216–218
Table 3. Hemolytic, thrombolytic and anticancer activities of benzofuran–oxadiazole/triazole derivatives.
Table 3. Hemolytic, thrombolytic and anticancer activities of benzofuran–oxadiazole/triazole derivatives.
EntryPercentage
Hemolysis ± SD
Percentage
Thrombolysis ± SD
a Percentage Cell viability A549
(Lung Cancer) ± SD
IC50 (μM)
A549
(Lung Cancer)
5a3.7 ± 0.00856.8 ± 0.08164.1 ± 1.72-
5b22.12 ± 0.00850.7 ± 0.08145.99 ± 4.22-
5c1.3 ± 0.00852.8 ± 0.08143.7 ± 0.94-
5d5.02 ± 0.00853.5 ± 0.08127.49 ± 1.906.3 ± 0.7
5e0.5 ± 0.00852.4 ± 0.08134.47 ± 2.1917.9 ± 0.46
5f0.74 ± 0.00856.5 ± 0.8143.67 ± 4.43-
5g4.86 ± 0.04748.3 ± 0.08141.45 ± 4.10-
7a9.6 ± 0.08152.2 ± 0.08149.8 ± 1.06-
7b0.1 ± 0.00452.5 ± 0.08157.62 ± 4.94-
7c2.15 ± 0.00854 ± 0.08144.52 ± 5.01-
7d6.13 ± 0.04756.2 ± 0.08139.12 ± 2.21-
7e3.11 ± 0.00859.1 ± 0.00844.72 ± 0.84-
7f15.7 ± 0.08161.4 ± 0.08199.1 ± 5.04-
7g23.4 ± 0.08149.06 ± 0.04736.26 ± 0.4119.8 ± 0.54
7h14.8 ± 0.08148.1 ± 0.08129.29 ± 3.9810.9 ± 0.94
ABTS
(+ve control)
95.986--
DMSO
(-ve control)
--100 ± 0-
Crizotinib [41,42,43,44,45,46,47,48] 28.22 ± 3.888.54 ± 0.84
Cisplatin [49,50,51,52,53,54] 15.34 ± 2.983.88 ± 0.76
a Cell viability, IC50: (Mean ± SD) in triplicate.
Table 4. Molecular docking parameters analysis against ALK receptors.
Table 4. Molecular docking parameters analysis against ALK receptors.
TargetLigandBinding Energies (Kcal/mol)Binding ResiduesType of InteractionBond Distance Range (Å)
Anaplastic lymphoma kinase (ALK) receptorsCrizotinib−8.985LEU A: 1122, ALA A: 1148, MET A: 1199, GLU A: 1197, ARG A: 1253, LEU A: 1256Conventional hydrogen bond, carbon hydrogen bond, pi-sigma, alkyl interaction, pi-alkyl, van der Waals interactions2.98–4.48
Compound 5d−9.925VAL A: 1130, ALA A: 1148, GLY A: 1201, ASP A: 1203, GLU A: 1210, LEU A: 1256, PRO A: 1260Conventional hydrogen bond, pi-anion, pi-donor hydrogen bond, pi-sigma, pi-alkyl, van der Waals interactions2.68–4.57
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Irfan, A.; Faiz, S.; Rasul, A.; Zafar, R.; Zahoor, A.F.; Kotwica-Mojzych, K.; Mojzych, M. Exploring the Synergistic Anticancer Potential of Benzofuran–Oxadiazoles and Triazoles: Improved Ultrasound- and Microwave-Assisted Synthesis, Molecular Docking, Hemolytic, Thrombolytic and Anticancer Evaluation of Furan-Based Molecules. Molecules 2022, 27, 1023. https://doi.org/10.3390/molecules27031023

AMA Style

Irfan A, Faiz S, Rasul A, Zafar R, Zahoor AF, Kotwica-Mojzych K, Mojzych M. Exploring the Synergistic Anticancer Potential of Benzofuran–Oxadiazoles and Triazoles: Improved Ultrasound- and Microwave-Assisted Synthesis, Molecular Docking, Hemolytic, Thrombolytic and Anticancer Evaluation of Furan-Based Molecules. Molecules. 2022; 27(3):1023. https://doi.org/10.3390/molecules27031023

Chicago/Turabian Style

Irfan, Ali, Sadia Faiz, Azhar Rasul, Rehman Zafar, Ameer Fawad Zahoor, Katarzyna Kotwica-Mojzych, and Mariusz Mojzych. 2022. "Exploring the Synergistic Anticancer Potential of Benzofuran–Oxadiazoles and Triazoles: Improved Ultrasound- and Microwave-Assisted Synthesis, Molecular Docking, Hemolytic, Thrombolytic and Anticancer Evaluation of Furan-Based Molecules" Molecules 27, no. 3: 1023. https://doi.org/10.3390/molecules27031023

Article Metrics

Back to TopTop