Skip to main content
Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Bioinform Biol Insights. 2021; 15: 11779322211010703.
Published online 2021 Apr 27. doi: 10.1177/11779322211010703
PMCID: PMC8842389
PMID: 35173424

Anticancer Potential of Moringa oleifera on BRCA-1 Gene: Systems Biology

Abstract

Breast cancer has consistently been a global challenge that is prevalent among women. There is a continuous increase in the high number of women mortality rates because of breast cancer and affecting nations at all modernization levels. Women with high-risk factors, including hereditary, obesity, and menopause, have the possibility of developing breast cancer growth. With the advent of radiotherapy, chemotherapy, hormone therapy, and surgery in breast cancer treatment, breast cancer survivors have increased. Also, the design and development of drugs targeting therapeutic enzymes effectively treat the tumour cells early. However, long-term use of anticancer drugs has been linked to severe side effects. This research aims to develop potential drug candidates from Moringa oleifera, which could serve as anticancer agents. In silico analysis using Schrödinger Molecular Drug Discovery Suite and SWISS ADME was employed to determine the therapeutic potential of phytochemicals from M oleifera against breast cancer via molecular docking, pharmacokinetic parameters, and drug-like properties. The result shows that rutin, vicenin-2, and quercetin-3-O-glucoside have the highest binding energy of −7.522, −6.808, and −6.635 kcal/mol, respectively, in the active site of BRCA-1. The essential amino acids involved in the protein-ligand interaction following active site analysis are ASN 1678, ASN 1774, GLY 1656, LEU 1657, GLN 1779, LYS 1702, SER 1655, PHE 1662, ARG 1699, GLU 1698, and VAL 1654. Thus, we propose that bioactive compounds from M oleifera may be potential novel drug candidates in the treatment of breast cancer.

Keywords: Moringa oleifera, breast cancer, in silico, BRCA-1, rutin

Introduction

Breast cancer is the leading cause of death in women around the world. Several factors contribute significantly to the increased risk of breast cancer, including oral contraceptives, obesity, menopause, and elevation in serum estradiol concentration. 1 Ductal carcinoma is the most common type of breast cancer which developed from the ducts. Cancerous cells growing from lobules are called lobular cells. 2 Breast cancers are mainly diagnosed by physical examination by a health care provider or the use of mammography. 3 High occurrence of breast cancer has been reported to be prevalent in white women within the range of 40 years and above. 4

Breast cancer gene 1 (BRCA-1), also called the caretaker gene, is a tumour suppressor gene that functions in cell cycle regulation, DNA repair mechanism, and other metabolic processes.5,6 The BRCA-1 proteins interact with other essential proteins necessary to replicate and repair double-stranded DNA breaks. 7 It contains 1863 amino acid residues and helps inhibit the proliferation of cells lining the breast’s milk ducts. Thus, BRCA-1 does not contribute to the pathogenesis of breast cancer. However, mutations in the breast cancer gene sequence can consequently increase breast cancer risk. 8 Mutations evolved when an individual’s genetic makeup becomes damaged via exposure to environmental factors, including ultraviolet light, ionizing radiation, and genotoxic chemicals. 9 When the BRCA-1 is mutated, it cannot efficiently repair the broken DNA; thereby, breast cancer prevention will be hampered. 10

Several treatment methods are available for breast cancers, but hormone-blocking agents, chemotherapy, and monoclonal antibodies are commonly used.11,12 Hormone receptors (oestrogen ER+ and progesterone PR+ receptors) are a therapeutic target in breast cancer. Drugs such as tamoxifen and anastrozole act by blocking the hormone receptors. 13 Several medicinal plants such as Camptotheca acuminate, Catharanthus roseus, Taxus brevifolia, and many others have been used as anticancer therapy. 14

Moringa oleifera, which belongs to the family of Moringaceae, has been reported to possess beneficial pharmacological properties such as anticonvulsant, antimicrobial, anticancer, and antiviral. 15 The extracts (phytochemicals) from the leaves, seeds, bark, and flowers of M oleifera have been used to treat several long-term diseases, including hypercholesterolemia, high blood pressure, diabetes, insulin resistance, nonalcoholic liver disease, cancer, and inflammation. 16 Bioactive compounds of M oleifera show inhibitory potential against cancerous cell line by inhibiting proliferation of carcinoma cells and malignant astrocytoma cells.17,18 Pandey and Khan 19 reported the inhibitory potential of methanolic extract of M oleifera leaves against cervical cancer cells by the downregulation of Jun activation domain-binding protein 1 and upregulation of p27 expression. It has been reported that the phytochemicals isolated from M oleifera leaves induced apoptosis of human keratin-forming cancer cells. 20 Furthermore, cold soluble aqueous extract of M oleifera leaves had demonstrated antitumour activity in A549 lung cancer cells via the mitochondrial-mediated pathway by pro-caspase activation 3 to caspase 3. 21 Nanoparticles derived from M oleifera demonstrated biological and pharmacological activity, including antimicrobial, anticancer, and antiplatelet activity. 22 The M oleifera shows critical anticancer potential on prostate cancer-3 (PC-3) carcinoma cells of prostate cancer in a dose-dependent manner. However, its cytotoxicity impact in typical Hek293 (human embryonic kidney 293) cells was minimal. 23

In this study, in silico analysis via molecular docking and pharmacokinetic profiles were employed to screen the library of bioactive compounds from M oleifera to determine their anticancer property.

Materials and Method

Ligand preparation

The phytochemicals of M oleifera were retrieved from published literature, 15 and their crystal structures were downloaded from the PubChem database (https://pubchem.ncbi.nlm.nih.gov/). The PubChem Compound Identification Numbers (CIDs) for each ligands are rutin (CID: 5280805), vicenin-2 (CID: 5280805), quercetin-3-O-glucoside (Q3G) (CID: 5748594), chlorogenic acid (CID: 1794427), gallic acid (CID: 370), sinalbin (CID: 656568), isoquercetin (CID: 5280804), astragalin (CID: 5282102), quercetin (CID: 5280343), ferulic acid (CID: 445858), myricetin (5281672), and kaempferol (CID: 5280863). The ligands were prepared using the LigPrep module of Glide tool by using the OPLS 2005 force field. 24

Protein preparation

The crystal structures of the BRCA-1 (PDB ID: 4OFB) was retrieved from Protein Data Bank (https://www.rcsb.org/) in complex with crystallized ligands. The protein was prepared using ProteinPrep Wizard of Maestro interface (11.5) by adding missing hydrogen atoms. Furthermore, the metal ionization was corrected to ensure formal charge and force field treatment. The protein was optimized and refined for docking analysis.25,26

Molecular docking

The docking analysis was conducted using the Glide tool from Schrödinger molecular drug discovery suite (version 2017-1). The grid was generated using the receptor grid generation module of the Glide tool. The coordinate (x, y, z) of the grid was centred to −9.07, 27.02, and −0.91, respectively. The refined M oleifera ligands were docked into the active site of BRCA-1. The energy calculation was achieved using the scoring function of the Glide tool. The compounds’ drug-like properties were evaluated using the QikProp module and SWISS ADME Web tool following Lipinski five rule. 27

Results and Discussion

Molecular docking was employed to perform the library’s virtual screening of phytochemicals from M oleifera against the targeted protein (BRCA-1). 28 The phytocompounds of M oleifera were ranked according to their binding poses and energy calculations. 29 The compounds were further subjected to pharmacokinetic study to predict their drug-able properties. The molecular docking analysis, which includes binding affinity (kcal/mol) predication, the interaction of the ligands within the binding pocket of BRCA-1, and their pharmacokinetic study, was shown (Table 1). Each ligand was analysed using Lipinski rule of five (ROF). The result confirms the ligands ROF with few violations. The ligand docking shows how the phytocompounds bind effectively with BRCA-1. Visualization of the protein-ligand complex was performed using the Glide tool’s surface module (Figure 1). The interaction between the compounds and BRCA-1 identified the amino acid residues involved in the interaction and each amino acid residues’ position in their ligand-binding site. The interaction was associated with a structure-based drug design depicting protein-ligand interaction.

Table 1.

Docking results of phytochemicals from Moringa oleifera in terms of binding affinity (kcal/mol), the interaction of the compounds with BRCA-1, and the drug-like properties.

PhytochemicalsAffinity (kcal/mol)Structure of the compounds and their interaction with BRCA-1Drug-like properties (Lipinski rule of five)
Rutin−7.522 An external file that holds a picture, illustration, etc.
Object name is 10.1177_11779322211010703-img1.jpg Molecular weight (<500 Da): 610.52
Log P (<5): 2.43
H-bond donor (5): 10
H-bond acceptor (<10): 16
MlogP (<4.15): 3.89
Violations: 3
Vicenin-2−6.808 An external file that holds a picture, illustration, etc.
Object name is 10.1177_11779322211010703-img2.jpg Molecular weight (<500 Da): 594.52
Log P (<5): 1.27
H-bond donor (5): 8
H-bond acceptor (<10): 12
MlogP (<4.15): 2.59
Violations: 3
Quercetin-3-O-glucoside−6.635 An external file that holds a picture, illustration, etc.
Object name is 10.1177_11779322211010703-img3.jpg Molecular weight (<500 Da): 464.38
Log P (<5): 2.02
H-bond donor (5): 11
H-bond acceptor (<10): 15
MlogP (<4.15): 2.59
Violations: 2
Chlorogenic acid−6.181 An external file that holds a picture, illustration, etc.
Object name is 10.1177_11779322211010703-img4.jpg Molecular weight (<500 Da): 354.31
Log P (<5): 0.96
H-bond donor (5): 6
H-bond acceptor (<10): 9
MlogP (<4.15): 1.05
Violations: 1
Gallic acid−5.771 An external file that holds a picture, illustration, etc.
Object name is 10.1177_11779322211010703-img5.jpg Molecular weight (<500 Da): 170.12
Log P (<5): 0.21
H-bond donor (5): 4
H-bond acceptor (<10): 5
MlogP (<4.15): 0.16
Violations: 0
Sinalbin−4.893 An external file that holds a picture, illustration, etc.
Object name is 10.1177_11779322211010703-img6.jpg Molecular weight (<500 Da): 425.43
Log P (<5): 0.49
H-bond donor (5): 6
H-bond acceptor (<10): 11
MlogP (<4.15): 2.14
Violations: 2
Isoquercetin−4.766 An external file that holds a picture, illustration, etc.
Object name is 10.1177_11779322211010703-img7.jpg Molecular weight (<500 Da): 464.38
Log P (<5): 2.11
H-bond donor (5): 8
H-bond acceptor (<10): 12
MlogP (<4.15): 2.59
Violations: 2
Astragalin−4.492 An external file that holds a picture, illustration, etc.
Object name is 10.1177_11779322211010703-img8.jpg Molecular weight (<500 Da): 448.38
Log P (<5): 0.53
H-bond donor (5): 7
H-bond acceptor (<10): 11
MlogP (<4.15): 2.10
Violations: 2
Quercetin4.415 An external file that holds a picture, illustration, etc.
Object name is 10.1177_11779322211010703-img9.jpg Molecular weight (<500 Da): 448.38
Log P (<5): 0.53
H-bond donor (5): 7
H-bond acceptor (<10): 11
MlogP (<4.15): 2.10
Violations: 2
Ferulic acid−4.090 An external file that holds a picture, illustration, etc.
Object name is 10.1177_11779322211010703-img10.jpg Molecular weight (<500 Da): 194.18
Log P (<5): 1.62
H-bond donor (5): 2
H-bond acceptor (<10): 4
MlogP (<4.15): 1.00
Violations: 0
Myricetin−3.819 An external file that holds a picture, illustration, etc.
Object name is 10.1177_11779322211010703-img11.jpg Molecular weight (<500 Da): 318.24
Log P (<5): 1.08
H-bond donor (5): 6
H-bond acceptor (<10): 8
MlogP (<4.15): 1.08
Violations: 1
Kaempferol−3.666 An external file that holds a picture, illustration, etc.
Object name is 10.1177_11779322211010703-img12.jpg Molecular weight (<500 Da): 286.24
Log P (<5): 1.70
H-bond donor (5): 4
H-bond acceptor (<10): 6
MlogP (<4.15): 0.03
Violations: 0

Abbreviation: BRCA-1, breast cancer gene 1.

An external file that holds a picture, illustration, etc.
Object name is 10.1177_11779322211010703-fig1.jpg

Visualization of docking results showing binding of (A) rutin, (B) vicenin-2, (C) quercetin-3-O-glucoside, (D) chlorogenic acid, (E) gallic acid, (F) sinalbin, (G) isoquercetin, (H) astragalin, (I) quercetin, (J) ferulic acid, (K) myricetin, and (L) kaempferol with BRCA-1. BRCA-1 indicates breast cancer gene 1.

The molecular docking demonstrates hydrophobic, pi-pi stacking, hydrogen bonding, and many others between the protein and the ligands. 26 Breast cancer gene 1 was cocrystallized with a natural inhibitor that defines its active site. This allows the binding of the ligands into the protein-binding domain. The primary amino acids involved in the protein-ligand interaction following active site analysis are ASN 1678, ASN 1774, GLY 1656, LEU 1657, GLN 1779, LYS 1702, SER 1655, PHE 1662, ARG 1699, GLU 1698, and VAL 1654. The phytoconstituents show a favourable interaction with BRCA-1. Rutin is a flavonoid found in natural products and has shown antitumour potential against various cancer cells. 30 Rutin’s inhibitory potential against cervical cancer cells, majorly the human papillomavirus–negative C33A cell line, has been elucidated. 31 Following rutin’s extra precision docking against BRCA-1, it shows hydrogen bonding interaction and pi-pi stalking with amino acid residues LEU 1701, ASN 1774, ARG 1699, GLU 1698, ASN 1678, LEU 1657, and SER 1655 and a binding affinity of −11.769 kcal/mol. Rutin’s toxicity study confirms that it has low bioavailability and solubility, which has affected its application in the delivery system. It binds firmly to the human serum albumin, shows a high metabolic rate, and can be easily excreted.32,33

Vicenin-2 is a nontoxic flavonoid with pharmacological properties including antioxidant, hepatoprotective, anti-inflammatory, and anticancer. 34 It significantly inhibits vascular endothelial growth factor receptor tyrosine kinases in prostate cancer cells. 35 It shows inhibitory potential against hepatocarcinoma cell proliferation via the signal transduction pathway involved in activating the signal transducer and activator of transcription 3 gene in a dose-dependent manner. 36 In addition, it reduced the growth of cancer cells by targeting the β-catenin pathway. 37 Vicenin-2 exhibited promising ligand interaction when complexed with BRCA-1. It binds with an energy of −6.808 kcal/mol by hydrophobic interaction with VAL 1654.

Aqueous Q3G extract isolated from medicinal plants has exhibited antiproliferative against different cancer cell lines. 38 The Q3G triggers apoptotic cell death through inhibition of the extrinsic pathway in caspase-3. Also, it is a potent inhibitor of DNA topoisomerase II involved in carcinogenesis. 39 The docking of Q3G with BRCA-1 shows a glide score of −6.635 kcal/mol by forming 5 hydrogen bonds with ASN 1774, GLY 1779, ASN 1678, GLY 1656, and Ser 1655 accompanied with pi-pi stacking at amino acid residue LYS 1702. A chlorogenic acid is an esterified form of caffeic acid via the shikimate pathway found naturally in plants. The molecular mechanism underlying the anticancer potential of chlorogenic acid involved inhibition of signalling transduction pathways (NF-κB, c-Jun NH2-terminal kinase, p38 kinase) to reduce the growth of cancer cells. 40 It binds well with the targeted protein with an affinity of −6.181 kcal/mol.

Gallic acid is a polyphenol known as 3,4,5-trihydroxy benzoic acid, with a significant anticancer property. 41 When A375 melanoma cancer cells were exposed to gallic acid in vitro, it stops cancer cells’ growth. Synchronous treatment with low-level laser and subsequently gallic acid increases reactive oxygen species’ production in both breast and melanoma cancer growth cells compared with gallic acid alone, which causes more apoptotic cells death in the tumour cells. 42 Sinalbin (a glucosinolate), quercetin, and isoquercetin are phenolic compounds that exhibit various biological functions, such as antioxidant, radical-scavenging, anti-inflammatory, antibacterial, antiviral, and anticancer. Khan et al 43 reported that quercetin and its derivatives inhibit cell proliferation in the colon (HCT-116) cancer cells. There was a favourable interaction of gallic acid, sinalbin, and isoquercetin against BRCA-1 with a binding energy of −5.771, −4.893, and 4.766 kcal/mol, respectively. The drug-like properties of gallic acid demonstrated that it does not violate Lipinski 5-year rule with a promising therapeutic potential. Isoquercetin interacts with an amino acid at GLY 1656. Astragalin and quercetin’s pharmacokinetic profiles adhere to the ROF with only 2 violations and docking scores of 4.415 and −4.090 kcal/mol, respectively.

Ferulic acid (4-hydroxy-3-methoxy cinnamic acid) is a widely distributed phenolic compound in plants. 44 Furthermore, ferulic acid causes cell death by the downregulation of cyclin-dependent kinases and inhibiting the activation of the PI3K/Akt signalling pathway in proliferative cells. 45 Myricetin demonstrates anticancer activity against human acute leukaemia HL-60 cells in a dose-dependent manner while inhibiting tumour promoter–induced neoplastic cells in skin cancer.46,47 Ferulic acid and myricetin have a binding energy of −4.090 and −3.819 kcal/mol, respectively, when complexed with the targeted protein. Kaempferol is an aglycone flavonoid that possesses anticancer activity against glioblastoma, breast cancer, hepatocellular carcinoma, colorectal cancer, and acute promyelocytic leukaemia pancreatic cancer, prostate cancer, and renal cell carcinoma. 48 Kaempferol has drug-like properties without violating Lipinski ROF and binding energy of −3.666 kcal/mol.

Conclusions

Several anticancer drugs, such as tamoxifen, anastrozole, and exemestane have been developed and are effective but posed severe side effects, including liver toxicity, cardiovascular diseases, and many others, following long-time use. In this study, we used computational modelling techniques to predict the inhibitory potential of M oleifera against BRCA-1. The binding of the compounds with BRCA-1, toxicity, and drug-like property as confirmed by docking analysis show that the M oleifera ligands are promising anticancer agents. Following the phytocompound screening from M oleifera by docking technique, rutin was found to exhibit the highest degree of interaction and binding affinity with BRCA-1 accompanied by favourable drug-like properties. Thus, we proposed that the phytochemicals from M oleifera may be potential BRCA-1 inhibitors. Further biochemical analysis such as in vitro and in vivo study is required to establish the compounds’ pharmacological properties.

Footnotes

Funding: The author(s) received no financial support for the research, authorship, and/or publication of this article.

Declaration of conflicting interests: The author(s) declared no potential conflicts of interest with respect to the research, authorship, and/or publication of this article.

Author Contributions: Toheeb A Balogun conceptualized and designed the study, performed the analysis and wrote the manuscript. Kaosarat D Buliaminu contributed to the drafting of the manuscript. Onyeka S Chukwudozie critically review the manuscript. Zainab A Tiamiyu and Taiwo J Idowu assisted with the editorial works.

ORCID iD: Toheeb A Balogun An external file that holds a picture, illustration, etc.
Object name is 10.1177_11779322211010703-img13.jpg https://orcid.org/0000-0002-6267-425X

References

1. Key T, Verkasalo P, Banks E. Epidemiology of breast cancer. Lancet Oncol. 2001;2:133-140. [PubMed] [Google Scholar]
2. National Cancer Institute. Breast Cancer Treatment (Adult). Health professional version. https://www.cancer.gov/types/breast/hp/breast-treatment-pdq. Published 2020.
3. Saslow D, Hannan J, Osuch J, et al. Clinical breast examination: practical recommendations for optimizing performance and reporting. CA Cancer J Clin. 2004;54:327-344. [PubMed] [Google Scholar]
4. Anders CK, Johnson R, Litton J, Phillips M, Bleyer A. Breast cancer before age 40 years. Semin Oncol. 2009;36(3):237-249. DOI: 10.1053/j.seminoncol.2009.03.001. [PMC free article] [PubMed] [Google Scholar]
5. Duncan JA, Reeves JR, Cooke TG. BRCA1 and BRCA2 proteins: roles in health and disease. Mol Pathol. 1998;51:237-247. [PMC free article] [PubMed] [Google Scholar]
6. Yoshida K, Miki Y. Role of BRCA1 and BRCA2 as regulators of DNA repair, transcription, and cell cycle in response to DNA damage. Cancer Sci. 2004;95:866-871. [PMC free article] [PubMed] [Google Scholar]
7. Irminger-Finger I, Ratajska M, Pilyugin M. New concepts on BARD1: regulator of BRCA pathways and beyond. Int J Biochem Cell Biol. 2016;72:1-17. [PubMed] [Google Scholar]
8. Friedenson B. The BRCA1/2 pathway prevents hematologic cancers in addition to breast and ovarian cancers. BMC Cancer. 2007;7:152. [PMC free article] [PubMed] [Google Scholar]
9. Dizdaroglu M, Coskun E, Jaruga P. Measurement of oxidatively induced DNA damage and its repair, by mass spectrometric techniques. Free Radic Res. 2015;49:525-548. [PubMed] [Google Scholar]
10. Boulton SJ, Jackson SP. Saccharomyces cerevisiae Ku70 potentiates illegitimate DNA double-strand break repair and serves as a barrier to error-prone DNA repair pathways. EMBO J. 1996;15:5093-5103. [PMC free article] [PubMed] [Google Scholar]
11. Leite AM, Macedo AV, Jorge AJ, Martins WA. Antiplatelet therapy in breast cancer patients using hormonal therapy: myths, evidence, and potentialities – systematic review. Arq Bras Cardiol. 2018;111:205-212. [PMC free article] [PubMed] [Google Scholar]
12. Holmes MD, Chen WY, Li L, Hertzmark E, Spiegelman D, Hankinson SE. Aspirin intake and survival after breast cancer. J Clin Oncol. 2010;28:1467-1472. [PMC free article] [PubMed] [Google Scholar]
13. Bao T, Rudek MA. The clinical pharmacology of anastrozole. Eur Oncol Haematol. 2011;7:106-108. [Google Scholar]
14. Cragg GM, Newman DJ. Plants as a source of anti-cancer agents. J Ethnopharmacol. 2005;100:72-79. [PubMed] [Google Scholar]
15. Abd Rani N, Husain K, Kumolosasi E. Moringa genus: a review of phytochemistry and pharmacology. Front Pharmacol. 2018;16:108. [PMC free article] [PubMed] [Google Scholar]
16. Marcela Vergara J, Manal Mused A, Maria Luz F. Bioactive components in Moringa oleifera leaves protect against chronic disease. Antioxidants (Basel). 2017;6:91. [PMC free article] [PubMed] [Google Scholar]
17. Rajan TS, De Nicola GR, Iori R, Rollin P, Bramanti P, Mazzon E. Anticancer activity of glucomoringin isothiocyanate in human malignant astrocytoma cells. Fitoterapia. 2016;110:1-7. [PubMed] [Google Scholar]
18. Vijayarajan M, Pandian MR. Cytotoxicity of methanol and acetone root bark extracts of Moringa concanensis against A549, Hep-G2, and HT-29 cell lines. J Acad Ind Res. 2016;5:45-49. [Google Scholar]
19. Pandey P, Khan F. Jab1 inhibition by methanolic extract of Moringa Oleifera leaves in cervical cancer cells: a potent targeted therapeutic approach [published online ahead of print September 30, 2020]. Nutr Cancer. DOI: 10.1080/01635581.2020.1826989. [PubMed] [CrossRef] [Google Scholar]
20. Sreelatha S, Jeyachitra A, Padma PR. Antiproliferation and induction of apoptosis by Moringa oleifera leaf extract on human cancer cells. Food Chem Toxicol. 2011;49:1270-1275. [PubMed] [Google Scholar]
21. Ismail AA, Wasiu GB, Hasni A. Moringa oleifera: an apoptosis inducer in cancer cells. Trop J Pharm Res. 2017;16:2289-2296. [Google Scholar]
22. Ashish KS, Ajit KS, Dinesh J, et al. An eco-friendly green synthesis of tungsten nanoparticles from Moringa oleifera Lam. and their pharmacological studies. Gazi Med J. 2020;31:719-725. [Google Scholar]
23. Khan F, Pandey P, Ahmad V, Upadhyay TK. Moringa oleifera methanolic leaves extract induces apoptosis and G0/G1 cell cycle arrest via downregulation of Hedgehog Signaling Pathway in human prostate PC-3 cancer cells. J Food Biochem. 2020;44:e13338. [PubMed] [Google Scholar]
24. Balogun TA, Omoboyowa DA, Saibu OA. In silico anti-malaria activity of quinolone compounds against Plasmodium falciparum dihydrofolate reductase (pfDHFR). Int J Biochem Res Rev. 2020;29:10-17. [Google Scholar]
25. Harder E, Damm W, Maple J, et al. OPLS3: a force field providing broad coverage of drug-like small molecules and proteins. J Chem Theory Comput 2016;12:281-296. [PubMed] [Google Scholar]
26. Friesner RA, Banks JL, Murphy RB, et al. Glide: a new approach for rapid, accurate docking and scoring. 1. Method and assessment of docking accuracy. J Med Chem. 2004;47:1739-1749. [PubMed] [Google Scholar]
27. Narkhede RR, Pise AV, Cheke RS, Shinde SD. Recognition of natural products as potential inhibitors of Covid-19 main protease (Mpro): in-silico evidences. Nat Prod Bioprospect. 2020;10:297-306. [PMC free article] [PubMed] [Google Scholar]
28. Lengauer T, Rarey M. Computational methods for biomolecular docking. Curr Opin Struct Biol. 1996;6:402-406. [PubMed] [Google Scholar]
29. Feig M, Onufriev A, Lee MS, Im W, Case DA, Brooks CL. Performance comparison of generalized born and Poisson methods in the calculation of electrostatic solvation energies for protein structures. J Comput Chem. 2004;25:265-284. [PubMed] [Google Scholar]
30. Perk AA, Shatynska-Mytsyk I, Gerçek YC, et al. Rutin mediated targeting of signaling machinery in cancer cells. Cancer Cell Int. 2014;14:124. [PMC free article] [PubMed] [Google Scholar]
31. Khan F, Pandey P, Upadhyay TK, et al. Anti-cancerous effect of Rutin against HPV-C33A cervical cancer cells via G0/G1 cell cycle arrest and apoptotic induction. Endocr Metab Immune Disord Drug Targets. 2020;20:409-418. [PubMed] [Google Scholar]
32. Sharma S, Ali A, Ali J, Sahni JK, Baboota S. Rutin: therapeutic potential and recent advances in drug delivery. Expert Opin Investig Drugs. 2013;22:1063-1079. [PubMed] [Google Scholar]
33. Yang CY, Hsiu SL, Wen KC, et al. Bioavailability and metabolic pharmacokinetics of rutin and quercetin in rats. J Food Drug Anal. 2005;13:244-250. [Google Scholar]
34. Ku SK, Bae JS. Vicenin-2 and scolymoside inhibit high-glucose-induced vascular inflammation in vitro and in vivo. Can J Physiol Pharmacol. 2016;94:287-295. [PubMed] [Google Scholar]
35. Singhal SS, Jain D, Singhal P, Awasthi S, Singhal J, Horne D. Targeting the mercapturic acid pathway and vicenin-2 for prevention of prostate cancer. Biochim Biophys Acta Rev Cancer. 2017;1868:167-175. [PMC free article] [PubMed] [Google Scholar]
36. Huang G, Li S, Zhang Y, Zhou X, Chen W. Vicenin-2 is a novel inhibitor of STAT3 signaling pathway in human hepatocellular carcinoma. J Funct Food. 2020;69:103921. [Google Scholar]
37. Egashira I, Takahashi-Yanaga F, Nishida R, et al. Celecoxib and 2,5-dimethylcelecoxib inhibit intestinal cancer growth by suppressing the Wnt/beta-catenin signaling pathway. Cancer Sci. 2017;108:108-115. [PMC free article] [PubMed] [Google Scholar]
38. Yoon H, Liu RH. Effect of selected phytochemicals and apple extracts on NF-κB activation in human breast cancer MCF-7 cells. J Agric Food Chem. 2007;55:3167-3173. [PubMed] [Google Scholar]
39. Sudan S, Rupasinghe HP. Quercetin-3-O-glucoside induces human DNA topoisomerase II inhibition, cell cycle arrest and apoptosis in hepatocellular carcinoma cells. Anticancer Res. 2014;34:1691-1699. [PubMed] [Google Scholar]
40. Xu R, Kang Q, Ren J, Li Z, Xu X. Antitumor molecular mechanism of chlorogenic acid on inducting genes GSK-3β and APC and inhibiting gene β-catenin. J Anal Methods Chem. 2013;2013:951319. [PMC free article] [PubMed] [Google Scholar] Retracted
41. Zhao B, Hu M. Gallic acid reduces cell viability, proliferation, invasion and angiogenesis in human cervical cancer cells. Oncol Lett. 2013;6:1749-1755. [PMC free article] [PubMed] [Google Scholar]
42. Khorsandi K, Kianmehr Z, Hosseinmardi Z, Hosseinzadeh R. Anti-cancer effect of gallic acid in presence of low level laser irradiation: ROS production and induction of apoptosis and ferroptosis. Cancer Cell Int. 2020;20:18. [PMC free article] [PubMed] [Google Scholar]
43. Khan I, Paul S, Jakhar R, Bhardwaj M, Han J, Kang SC. Novel quercetin derivative TEF induces ER stress and mitochondria-mediated apoptosis in human colon cancer HCT-116 cells. Biomed Pharmacother. 2016;84:789-799. [PubMed] [Google Scholar]
44. Ls R, Nja S. Anticancer properties of phenolic acids in colon cancer – a review. J Nutr Food Sci. 2016;6:1-7. [Google Scholar]
45. Abotaleb M, Liskova A, Kubatka P, Büsselberg D. Therapeutic potential of plant phenolic acids in the treatment of cancer. Biomolecules. 2020;10:221. [PMC free article] [PubMed] [Google Scholar]
46. Chang H, Mi MT, Gu YY, Yuan JL, Ling WH, Lin H. Effects of flavonoids with different structures on proliferation of leukemia cell line HL-60. Chin J Cancer. 2007;26:1309-1314. [PubMed] [Google Scholar]
47. Kang NJ, Jung SK, Lee KW, Lee HJ. Myricetin is a potent chemopreventive phytochemical in skin carcinogenesis. Ann N Y Acad Sci. 2011;1229:124-132. [PubMed] [Google Scholar]
48. Imran M, Salehi B, Sharifi-Rad J, et al. Kaempferol: a key emphasis to its anticancer potential. Molecules (Basel, Switzerland). 2019;24:2277. [PMC free article] [PubMed] [Google Scholar]

Articles from Bioinformatics and Biology Insights are provided here courtesy of SAGE Publications

-