Anticancer activity of the supercritical extract of Brazilian green propolis and its active component, artepillin C: Bioinformatics and experimental analyses of its mechanisms of action

  • Authors:
    • Priyanshu Bhargava
    • Abhinav Grover
    • Nupur Nigam
    • Ashish Kaul
    • Motomichi Doi
    • Yoshiyuki Ishida
    • Hanzo Kakuta
    • Sunil C. Kaul
    • Keiji Terao
    • Renu Wadhwa
  • View Affiliations

  • Published online on: January 22, 2018     https://doi.org/10.3892/ijo.2018.4249
  • Pages: 925-932
Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Propolis, a resinous substance collected by honeybees by mixing their saliva with plant sources, including tree bark and leaves and then mixed with secreted beeswax, possesses a variety of bioactivities. Whereas caffeic acid phenethyl ester (CAPE) has been recognized as a major bioactive ingredient in New Zealand propolis, Brazilian green propolis, on the other hand, possesses artepillin C (ARC). In this study, we report that, similar to CAPE, ARC docks into and abrogates mortalin-p53 complexes, causing the activation of p53 and the growth arrest of cancer cells. Cell viability assays using ARC and green propolis-supercritical extract (GPSE) revealed higher cytotoxicity in the latter, supported by nuclear translocation and the activation of p53. Furthermore, in vivo tumor suppression assays using nude mice, we found that GPSE and its conjugate with γ cyclodextrin (γCD) possessed more potent anticancer activity than purified ARC. GPSE‑γCD may thus be recommended as a natural, effective and economic anticancer amalgam.

Introduction

Propolis is a complex mixture of resinous material, produced by bees, created by mixing their saliva with the botanical sources they live on. The color (yellowish green to dark brown) and odor (odorless to aromatic) of propolis vary and depend on its botanical source, origin of place and bee characteristics, such as strain and age (1,2). Apart from the structural and functional attributes of propolis for beehives (1), propolis has been reported to possess a variety of disease-preventive and therapeutic potentials for the human population (3). There are mainly two types of propolis known that differ in their constituents: New Zealand propolis that possesses caffeic acid phenethyl ester (CAPE) and Brazilian green propolis that possesses artepillin C (ARC) as predominant bioactive ingredients. Besides these, over 150 constituents, including flavonoids, phenolic acids, esters, terpenoids, steroids, amino acids and cinnamic acid derivatives have been identified (3), and are hence considered as popular pharmacological research material (4). A broad spectrum of biological activities identified in propolis include antitumor (512), anti-inflammatory (1315), anti-bacterial (1618), anti-viral (16,19,20) and anti-fungal (16) activities. Propolis is used in cosmetic products, such as body lotions, ointments, face creams and in functional food in various forms, such as tablets, capsules, toothpaste and mouthwash preparations (21). Molecular studies on the anticancer activity of propolis have revealed that its phenolic components cause cell cycle arrest and apoptosis (11), mitochondrial stress (22) and the inhibition of tumor growth (11,22,23). CAPE-based propolis extract (Bio-30) has been reported to block p21 (RAC1) activated kinase 1 (PAK1) signalling and to suppress tumors in neurofibromatosis (9,24,25).

ARC (3,5-diprenyl-4-hydroxycinnamic acid) is one of the active phenolic acid components of Brazilian propolis generated by bees from materials from a Brazilian plant, Baccharis dracunculifolia. It has been shown to possess various biological activities, such as anti-viral (14), anti-bacterial (14,26), antioxidant (26) and anti-carcinogenic (14,2628) activities. ARC has been reported to inhibit the growth of transplanted solid human and mouse tumors, including malignant melanoma in athymic and thymic mice, respectively (27). CAPE and ARC have been reported to differ in their bioavailability profile that in turn is determined by the stability of the compounds to the digestive enzymes and absorption through the intestinal lining. CAPE alone becomes degraded by secreted esterases (12); however, when combined with γ cyclodextrin (γCD) it is protected and has improved activity in in vitro and in vivo antitumor assays (12). On the other hand, ARC has been reported to possess extremely low absorption efficiency and bioavailability. Whereas CAPE is absorbed and distributed by the monocarboxylic acid transporter (MCT)-mediated transport system (29), ARC is mainly permeated across by transcellular passive diffusion (30). We previously performed a cDNA array of CAPE-treated normal human cells and found that the cytotoxicity of CAPE was mediated by the activation of p53-growth arrest and DNA damage-inducible 45 (GADD45). Bioinformatics and experimental analyses revealed that CAPE targeted mortalin-p53 interactions, resulting in the nuclear translocation and re-activation of p53, leading to growth arrest in cancer cells (12). In the present study, we report that ARC possesses similar capabilities; however its cytotoxic efficacy is low. In order to increase the potential of ARC, we prepared the following: i) supercritical extract [green propolis supercritical extract (GPSE)]; and ii) its complex with γCD (GPSE-γCD). We then tested their cytotoxicity in human cancer cells. We report that GPSE contains 9.6% ARC and exerted cytotoxic effects at 0.5% (16.6 µM) and was equivalent to ~500 µM of pure ARC in in vitro toxicity assays. The extract exhibited marked anti-migratory activity. Furthermore, the in vivo efficacy of GPSE was significantly enhanced by its complex with γCD.

Materials and methods

Docking of mortalin and p53 with ARC

The crystal structures of human mortalin (PDB ID: 4KBO) and p53 (PDB ID: 1OLG) were obtained from Protein Data Bank (PDB) (31), while the ligand, ARC (Compound ID: 5472440) was from PubChem Compound (32) database. ArgusLab 4.0.1 (M 2004), a freely available molecular modelling, graphics and drug design package, was used for docking analyses. The structures of protein and the ligand molecules were prepared and the receptor grid around the binding site residues selected from the receptor was defined. The grid resolution was kept as 0.4 Å and an exhaustive search docking was performed. The docking process was carried out using the 'ArgusDock' docking engine, calculation type as 'Dock' and the ligand was kept in rigid mode. The scoring function used was an empirical scoring function, 'AScore', which takes into account van der Waals energy, hydrophobic component, hydrogen bond and deformation penalty. The parameter file, 'AScore.prm' was used to compute the binding energies. A total of 150 docking poses were generated, ranked according to the scoring function, and the highest scoring pose was used in further analyses.

Molecular dynamics simulations of ARC-docked mortalin and p53 structure

MD simulations were carried out using the GROMACS package (33,34). The force field Gromos43a1 (35) was used for ARC-docked mortalin structure (33,36,37). The GROMACS topology file was generated using the antechamber python parser interface (ACPYPE) script. The docked protein structure was solvated in a cubic box. Water molecules and appropriate counter-ions were added to neutralize the system. The solvated system was minimized using steepest descent and conjugate gradient methods until the force on each atom was less than 100 kJ/mol/nm. These geometry minimized systems were used for 10 nsec for carrying out isobaric (constant pressure-temperature) MD simulations. The temperature and pressure of the system was maintained at 300 K and 1 atmosphere pressure, respectively, with a time constant of 5 psec. A 2-fsec time step was used for integrating the equations of motion. Particle Mesh Ewald summation method along with periodic boundary conditions were also applied throughout to calculate the electrostatic potential between partial charges on atoms. Visual Molecular Dynamics (VMD) version 1.9.2 was used to calculate the root mean square deviations (RMSD) and hydrogen bond dynamics.

Cell culture

HT1080 (fibrosarcoma), A549 (lung carcinoma) and U2OS (osteosarcoma) human cell lines were purchased from DS Pharma Biomedical Co., Ltd., Osaka, Japan, and cultured in complete Dulbecco's modified Eagle's medium (DMEM; Life Technologies) medium using normal cell culture conditions with 10% fetal bovine serum as a supplement at 37°C, 5% CO2 and 95% air in a humidified incubator. The cultured cells were maintained for 6–10 passages. ARC (Wako Pure Chemical Industries, Ltd., Osaka, Japan), GPSE and GPSE:50%γCD complex were dissolved in dimethylsulfoxide (DMSO), and directly added to the cell culture medium to obtain the working concentrations (as indicated in the respective figures).

Preparation of the GPSE-50%γCD complex

The complex of green propolis supercritical extract (GPSE) with γCD was prepared by a conventional kneading method. GPSE and γCD were mixed in a small amount of water. The slurry was kneaded until it became a homogeneous paste. After freeze-drying of the paste, the GPSE-γCD complex thus obtained was used in the present study. The levels of artepillin in GPSE, and its complex with γCD, were determined by high-performance liquid chromatography (HPLC) using the Shimadzu HPLC system (LC-2010C; Shimadzu Corp., Kyoto, Japan) as previously described (38). A Phenomenex HPLC column [Luna 5u C18(2) 100A: 4.60 mm I.D. ×150 mm] was used and the fractionation was performed at 40°C using solution A: H2O (0.5% acetic acid) and solution B: Acetonitrile with gradient program as follows: Isocratic 70% A (0–5 min), linear gradient 70% → 0% A (5–30 min); flow rate, 1 ml/min; injection volume, 10 µl. Detection was performed at 320 nm (Shimadzu HPLC system LC-2010C; Shimadzu Corp.).

Cell proliferation assay

Cytotoxicity assay was performed using 3-(4,5-dimethylthiazol-2-yl)-2, 5-diphenyltetrazolium bromide (MTT) assay (Life Technologies/Thermo Fisher Scientific, Waltham, MA, USA) in which cell viability was observed by the conversion of yellow MTT by the mitochondrial dehydrogenases of living cells into purple formazon (39). The statistical significance of the results was determined from 3–4 independent experiments including triplet or quadruplet sets in each experiment. IC50 values were calculated by the linear regression methods in MS excel.

Morphological observation

The cells were seeded in 12-well plates and treated with various concentrations of ARC (100–400 µM, as indicated in Fig. 2C), GPSE and GPSE-50%γCD (0.5%). Morphological changes were observed under a phase contrast microscope (Nikon Eclipse TE300; Nikon, Tokyo, Japan) after 48 h.

Colony formation assay

The cells (500 cells/well) were seeded in a 6-well plate. Following 24 h of incubation, when the cells had attached to the surface of the dish they were treated with ARC, GPSE or GPSE-50%γCD (0.1%) in culture medium (DMEM). The cultures were maintained for 10–15 days (when colonies formed in the control cultures) with a regular change of medium every 3rd day. The colonies were fixed in acetone: methanol (1:1), and stained with crystal violet (Wako Pure Chemical Industries Ltd.) at the end of experiments.

Immunofluorescence staining

The cells were seeded on glass coverslips placed in 12-well plates and treated with either ARC (300 µM), GPSE or GPSE-50%γCD (0.5%) in DMEM. After 48 h of treatment, cells were fixed in 4% paraformaldehyde in phosphate-buffered saline (PBS), permeabilized with 0.1% Triton X-100 for 10 min, blocked with 0.2% bovine serum albumin (BSA)/PBS for 1 h and incubated with specific primary antibodies including anti-mortalin (clone 088: raised in our laboratory) (3–5 µg/ml) and p53 (DO-1: SC-126) (3–5 µg/ml) (Santa Cruz Biotechnology, Santa Cruz, CA, USA) at 4°C overnight. The cells were then incubated with either Alexa-488 or Alexa-594 conjugated secondary antibodies (1 µg/ml) (A11034 or A11032, Molecular Probes, Eugene, OR, USA) for 1 h. Counterstaining was performed with Hoechst 33342 (Sigma, St. Louis, MO, USA) in the dark for 10 min. The cells were examined on a Zeiss Axiovert 200 M microscope and analyzed using AxioVision 4.6 software (Carl Zeiss, Oberkochen, Germany).

Wound healing assay

The cells were plated in 6-well plates and allowed to make monolayers following which a scratch was made using a 100-ml pipette tip. The cells were then washed with PBS and cultured with the control or test compounds [GPSE or GPSE-50%γCD or γCD, (0.1%) as indicated in Fig. 4C]. The movement of the cells to the scratch area was followed for the following 24–48 h. Images were acquired under a phase contrast microscope (Nikon) with a X10 phase objective lens at the 0, 24 and 48 h time-points. The movement was quantitated by measuring the empty surface area in the control and test samples by ImageJ software. The results are expressed as the means ± standard deviation of the mean (SEM). Statistical significance of the data was determined using a Student's t-test. Values of P<0.05, P<0.01 and P<0.001 were considered to indicate significant, very significant and highly significant differences, respectively.

In vivo tumor progression

The effects of GPSE and GPSE-50%γCD on in vivo tumor progression were investigated in a nude mouse subcutaneous tumor xenograft model. BALB/c nude mice (4 weeks old, female, 3 mice in each group/average weight 21.8 g) were obtained from Nihon Clea (Tokyo, Japan). The animals were allowed to acclimatize in our laboratory for 1 week. The cells were injected subcutaneously (1×107 suspended in 0.2 ml of growth medium) into the abdomen of the nude mice and the tail vein (1×106 suspended in 0.2 ml of growth medium). GPSE and GPSE 50%γCD were administered (by oral gavage) every alternate day beginning at 1 day after the injection of the cells for 3 weeks. Tumor formation and the body weight of the mice were monitored every alternate day. The volume of the subcutaneous tumors was calculated as V = L × W2/2, where 'L' was the length and 'W' was the width of the tumor, respectively. Statistical significance of the data was calculated from 3 independent experiments (n=3 per experiment). All procedures were carried out in accordance with the Animal Experiment and Ethics Committee, Safety and Environment Management Division, National Institute of Advanced Industrial Science and Technology (AIST), Tsukuba, Japan.

Statistical analysis

All the experiments were performed in triplicate and data are expressed as the means ± SEM. An unpaired t-test (GraphPad Prism GraphPad Software, San Diego, CA, USA) and one-way ANOVA followed by Tukey's honestly significant difference (HSD) post hoc test were used to determine the degree of significance between the control and experimental samples. Statistical significance was defined as P-values as follows: P<0.05 (significant), P<0.01 (very significant) and P<0.001 (highly significant) for the t-test and P<0.05 for ANOVA followed by a post hoc test, as indicated in the figure legends.

Results

ARC docks into the mortalin-p53 complexes and induces the activation of p53

The interaction of mortalin and p53 occurs at N-terminal amino acids residues (253–282) of mortalin and C-terminal residues (323–337) of p53 (12,4042). Based on this information, we docked ARC with mortalin or p53 or at mortalin-p53 complex interface. ARC interacted with the crystal structure of mortalin and p53 at the binding regions. It showed a docking score of −6.99 kcal/mol with mortalin. The docked complex exhibited multiple hydrogen bonds and hydrophobic interactions (Fig. 1A–a). The ligand formed 3 hydrogen bonds, 2 with Arg 126 and 1 with Thr 271 of mortalin, while residues Glu 132, Ala 195, Tyr 196, Thr 267, Asn 268 and Gly 269 played a role in hydrophobic interactions. The second oxygen atom of ARC formed 2 hydrogen bonds with nitrogen atoms (NH1 and NH2) of Arg 126. The third hydrogen bond was formed between the third oxygen atom of ARC and the third oxygen atom of Thr 271. Together the hydrogen bonds and hydrophobic interactions formed a stable protein-ligand complex.

The docking of ARC with p53 exhibited a score of −10.89 kcal/mol. The ligand was observed to form 1 hydrogen bond and various hydrophobic interactions (Fig. 1B–a). The nitrogen atom of Arg 333 was observed to form a hydrogen bond with the third oxygen atom of ARC. Residues Thr 329, Leu 330, Gln 331, Ile 332, Phe 338, Phe 341 and Arg 342 were involved in the hydrophobic interaction with p53. ARC docked in the cavity of p53 forming a stable complex.

Molecular dynamics simulations and energy stabilization of ARC in the mortalin-p53 complex

The stability of the ARC-docked mortalin structure was further analyzed by simulating the complex for 10 nsec under conditions mimicking the bodily environment. The complex was observed to be stable for a period of 5 nsec from 5 to 10 nsec, as seen in the RMSD trajectory. Changes observed in the interacting residues of mortalin are shown in Fig. 1A–b. One hydrogen bond was observed between the oxygen atom of Arg 127 and the third oxygen atom of ARC, while residues Tyr 196, Ser 266, Thr 267 and Thr 271 were involved in hydrophobic interaction following simulations. Despite the changes in the interaction pattern of ligand-docked protein complex, the structure was more stable following simulations. The RMSD trajectory is shown Fig. 1A–c. A change in the interaction pattern of ARC docked p53 complex was observed (Fig. 1B–b). The number of hydrogen bonds between ARC and p53 increased to 2 post-simulations. One bond was formed between the second oxygen atom of ARC and nitrogen atom (NH1) of Arg 333, while the second was observed to be formed between nitrogen atom (NE) of Arg 337. The number of residues involved in hydrophobic interactions were the same, with 2 residues (Leu 330 and Ile 332) replaced with Phe 328 and Asn 345. The complex was more stable post-simulations. The ARC-docked p53 structure was observed to be stable from 5 to 10 nsec that can be seen in the RMSD trajectory (Fig. 1B–c). Bioinformatics analyses revealed that the ARC docked into the mortalin-p53 complexes (Fig. 1), similar to CAPE, and hence may cause abrogation of these complexes, and the re-activation of p53 function in cancer cells.

Anticancer activity of ARC, GPSE and the GPSE-γCD complex

We then examined the cytotoxicity of ARC in HT1080 cells and found that it induced a considerable (>20%) decrease in cell viability at concentrations of 250 µM (IC50, 275 µM) (Fig. 2A). Based on such high IC50 values, we considered whether the extract of green propolis could be more effective than purified ARC for the reasons including that: i) it may provide a complex mixture of bioactives with multi-module action; and ii) it may protect the bioactive components from degradation. We prepared the GPSE and its conjugate with γCD (GPSE-γCD) and HPLC analyses revealed that they contained 9.6 and 3.0% ARC, respectively. The IC50 values were 0.2–0.5% for GPSE and GPSE-γCD, whereas they were 250–300 µM for ARC (Fig. 2B). Similar results were obtained with the A549 and U2OS cells (data not shown). Cell morphological analysis revealed that the cytotoxicity induced by 0.5% GPSE and GPSE-γCD was comparable to that induced by 300 µM ARC (Fig. 2C). Long-term colony forming assays revealed that ARC (100 µM) induced a marked decrease in the number and size of the colonies; a more potent decrease was observed by treatment with 0.1% GPSE and GPSE-γCD (Fig. 2D).

We then examined the above-mentioned concentrations of ARC, GPSE and GPSE-γCD, and determined their effects on mortalin-p53 interactions as predicted by bioinformatics analysis (Fig. 1). In several independent experiments, we found that treatment with GPSE and ARC led to the nuclear trans-location of p53 (Fig. 3A), accompanied by an increase in p53 expression (Fig. 3B). We performed antitumor assays using a subcutaneous tumor xenograft model with nude mice. As shown in Fig. 4, mice fed the GPSE or GPSE-γCD exhibited no change in body weight (Fig. 4A) or physical activity (data not shown) with respect to the control group. A reduction in tumour growth was observed in the mice fed GPSE or GPSE-γCD as compared to the control group over the 3 weeks of treatment (Fig. 4B). The tumor growth data were strongly suggestive that the GPSE-γCD induced a more marked suppressive effect on tumor growth, endorsing its potential as an antitumor natural compound. We also performed in vitro wound healing assays to examine whether the GPSE and GPSE-γCD extracts possess anti-migratory activity and to determine whether they can be used to inhibit cancer metastasis. As shown in Fig. 4C and D, a considerable inhibition of cell migration was detected in the cells cultured in the presence of non-toxic doses of GPSE and GPSE-γCD.

Discussion

A number of different methods have been used to extract active components from propolis, including solvent extraction, e.g., ethanol (3), water or diluted aqueous ethanol (43) and supercritical carbon dioxide extracts (8,44,45). Apart from differences in the yield due to the function of temperature, pressure and solvent concentration, the different extraction methods generate mixtures with different components and also show different activities. Supercritical extracts have non-polar characteristics that are tolerant to oxidation and provide high yields (8,44,46). However, due to their poor solubility in aqueous solvents, their use is limited in in vitro and in vivo studies. Cyclodextrins (CDs), with inner lipophilic cavities and hydrophilic outer surfaces, are capable of encapsulating molecules by non-covalent inclusion (47,48). We previously used γCD to conjugate with CAPE to generate a CAPE-γCD complex that showed significant anticancer activity (12). In the present study, we demonstrate that ARC, an active component of Brazilian green propolis, activates p53 tumor suppressor protein by abrogating its complex with mortalin (p53 inactivating protein), and thus possesses anticancer activity. However, it turned out to be a low efficacy compound (IC50, 275 µM). We hypothesized that crude extracts of green propolis may possess better efficacy. Hence, we generated its supercritical extract (GPSE) and its complex with γCD (GPSE-γCD). We herein report that the GPSE containing 9.6% ARC possesses high cytotoxicity to cancer cells, and was effective at concentrations of 0.25 to 0.5% (8.3–16.6 µM ARC). Furthermore, the GPSE-γCD complex (containing 3% ARC) exhibited greater cytotoxicity in vitro (0.25 to 0.5% = 2.7–5.5 µM ARC) and greater antitumor activity in vivo. Anti-migration assays revealed that GPSE, as well as the GPSE-50%γCD complex, at non-toxic concentrations, caused the delayed migration of cells to the scratch in wound healing assays, suggesting its potential for the treatment of metastatic cancers. Based on these data, GPSE-γCD is proposed as a Natural Efficient and Welfare (NEW) antitumor composite. Further studies are warranted to elucidate the molecular mechanism(s) of the potent anticancer activity of GPSE and its complex with γCD.

Abbreviations:

ARC

artepillin C

GPSE

green propolis supercritical extract

γCD

γ cyclodextrin

DMEM

Dulbecco's modified Eagle's medium

CAPE

caffeic acid phenethyl ester

GADD45

growth arrest and DNA damage-inducible 45

MTT

3-(4,5-dimethylthiazol-2-yl)-2, 5-diphenyltetrazolium bromide

BSA

bovine serum albumin

PBS

phosphate-buffered saline

Acknowledgments

Priyanshu Bhargava is a recipient of the Ministry of Education, Culture, Sports, Science and Technology (MEXT) Scholarship, Japan. This study was supported by AIST (Japan) and DBT (Government of India) funds.

Notes

[1] Competing interests

The authors declare that they have no competing interests.

References

1 

Bankova V: Chemical diversity of propolis and the problem of standardization. J Ethnopharmacol. 100:114–117. 2005. View Article : Google Scholar : PubMed/NCBI

2 

Kuropatnicki AK, Szliszka E and Krol W: Historical aspects of propolis research in modern times. Evid Based Complement Alternat Med. 2013:9641492013. View Article : Google Scholar : PubMed/NCBI

3 

Burdock GA: Review of the biological properties and toxicity of bee propolis (propolis). Food Chem Toxicol. 36:347–363. 1998. View Article : Google Scholar : PubMed/NCBI

4 

Bankova V, Popova M and Trusheva B: Propolis volatile compounds: chemical diversity and biological activity: A review. Chem Cent J. 8:282014. View Article : Google Scholar : PubMed/NCBI

5 

Fauzi AN, Norazmi MN and Yaacob NS: Tualang honey induces apoptosis and disrupts the mitochondrial membrane potential of human breast and cervical cancer cell lines. Food Chem Toxicol. 49:871–878. 2011. View Article : Google Scholar

6 

Ghashm AA, Othman NH, Khattak MN, Ismail NM and Saini R: Antiproliferative effect of Tualang honey on oral squamous cell carcinoma and osteosarcoma cell lines. BMC Complement Altern Med. 10:492010. View Article : Google Scholar : PubMed/NCBI

7 

Khacha-ananda S, Tragoolpua K, Chantawannakul P and Tragoolpua Y: Antioxidant and anti-cancer cell proliferation activity of propolis extracts from two extraction methods. Asian Pac J Cancer Prev. 14:6991–6995. 2013. View Article : Google Scholar

8 

Machado BA, Silva RP, Barreto GA, Costa SS, Silva DF, Brandão HN, Rocha JL, Dellagostin OA, Henriques JA, Umsza-Guez MA, et al: Chemical composition and biological activity of extracts obtained by supercritical extraction and ethanolic extraction of brown, green and red propolis derived from different geographic regions in Brazil. PLoS One. 11:e01459542016. View Article : Google Scholar : PubMed/NCBI

9 

Messerli SM, Ahn MR, Kunimasa K, Yanagihara M, Tatefuji T, Hashimoto K, Mautner V, Uto Y, Hori H, Kumazawa S, et al: Artepillin C (ARC) in Brazilian green propolis selectively blocks oncogenic PAK1 signaling and suppresses the growth of NF tumors in mice. Phytother Res. 23:423–427. 2009. View Article : Google Scholar

10 

Rao CV, Desai D, Simi B, Kulkarni N, Amin S and Reddy BS: Inhibitory effect of caffeic acid esters on azoxymethane-induced biochemical changes and aberrant crypt foci formation in rat colon. Cancer Res. 53:4182–4188. 1993.PubMed/NCBI

11 

Sawicka D, Car H, Borawska MH and Nikliński J: The anticancer activity of propolis. Folia Histochem Cytobiol. 50:25–37. 2012. View Article : Google Scholar : PubMed/NCBI

12 

Wadhwa R, Nigam N, Bhargava P, Dhanjal JK, Goyal S, Grover A, Sundar D, Ishida Y, Terao K and Kaul SC: Molecular characterization and enhancement of anticancer activity of caffeic acid phenethyl ester by γ cyclodextrin. J Cancer. 7:1755–1771. 2016. View Article : Google Scholar :

13 

Gao W, Wu J, Wei J, Pu L, Guo C, Yang J, Yang M and Luo H: Brazilian green propolis improves immune function in aged mice. J Clin Biochem Nutr. 55:7–10. 2014. View Article : Google Scholar : PubMed/NCBI

14 

Khayyal MT, el-Ghazaly MA and el-Khatib AS: Mechanisms involved in the antiinflammatory effect of propolis extract. Drugs Exp Clin Res. 19:197–203. 1993.PubMed/NCBI

15 

Wang L, Yang L, Debidda M, Witte D and Zheng Y: Cdc42 GTPase-activating protein deficiency promotes genomic instability and premature aging-like phenotypes. Proc Natl Acad Sci USA. 104:1248–1253. 2007. View Article : Google Scholar : PubMed/NCBI

16 

Kujumgiev A, Tsvetkova I, Serkedjieva Y, Bankova V, Christov R and Popov S: Antibacterial, antifungal and antiviral activity of propolis of different geographic origin. J Ethnopharmacol. 64:235–240. 1999. View Article : Google Scholar : PubMed/NCBI

17 

Pepeljnjak S, Jalsenjak I and Maysinger D: Flavonoid content in propolis extracts and growth inhibition of Bacillus subtilis. Pharmazie. 40:122–123. 1985.PubMed/NCBI

18 

Velikova M, Bankova V, Tsvetkova I, Kujumgiev A and Marcucci MC: Antibacterial ent-kaurene from Brazilian propolis of native stingless bees. Fitoterapia. 71:693–696. 2000. View Article : Google Scholar : PubMed/NCBI

19 

Jafarzadeh Kashi TS, Kasra Kermanshahi R, Erfan M, Vahid Dastjerdi E, Rezaei Y and Tabatabaei FS: Evaluating the in vitro antibacterial effect of Iranian propolis on oral microorganisms. Iran J Pharm Res. 10:363–368. 2011.PubMed/NCBI

20 

Sartori G, Pesarico AP, Pinton S, Dobrachinski F, Roman SS, Pauletto F, Rodrigues LC Jr and Prigol M: Protective effect of brown Brazilian propolis against acute vaginal lesions caused by herpes simplex virus type 2 in mice: Involvement of antioxidant and anti-inflammatory mechanisms. Cell Biochem Funct. 30:1–10. 2012. View Article : Google Scholar

21 

Choudhari MK, Haghniaz R, Rajwade JM and Paknikar KM: Anticancer activity of Indian stingless bee propolis: An in vitro study. Evid Based Complement Alternat Med. 2013.928280:2013.

22 

Benguedouar L, Boussenane HN, Wided K, Alyane M, Rouibah H and Lahouel M: Efficiency of propolis extract against mitochondrial stress induced by antineoplasic agents (doxorubicin and vinblastin) in rats. Indian J Exp Biol. 46:112–119. 2008.PubMed/NCBI

23 

Chen CN, Weng MS, Wu CL and Lin JK: Comparison of radical scavenging activity, cytotoxic effects and apoptosis induction in human melanoma cells by Taiwanese propolis from different sources. Evid Based Complement. Alternat Med. 1:175–185. 2004.

24 

Hirokawa Y, Levitzki A, Lessene G, Baell J, Xiao Y, Zhu H and Maruta H: Signal therapy of human pancreatic cancer and NF1-deficient breast cancer xenograft in mice by a combination of PP1 and GL-2003, anti-PAK1 drugs (Tyr-kinase inhibitors). Cancer Lett. 245:242–251. 2007. View Article : Google Scholar

25 

Maruta H: Effective neurofibromatosis therapeutics blocking the oncogenic kinase PAK1. Drug Discov Ther. 5:266–278. 2011. View Article : Google Scholar

26 

Veiga RS, De Mendonça S, Mendes PB, Paulino N, Mimica MJ, Lagareiro Netto AA, Lira IS, López BG, Negrão V and Marcucci MC: Artepillin C and phenolic compounds responsible for antimicrobial and antioxidant activity of green propolis and Baccharis dracunculifolia DC. J Appl Microbiol. 122:911–920. 2017. View Article : Google Scholar : PubMed/NCBI

27 

Kimoto T, Arai S, Kohguchi M, Aga M, Nomura Y, Micallef MJ, Kurimoto M and Mito K: Apoptosis and suppression of tumor growth by artepillin C extracted from Brazilian propolis. Cancer Detect Prev. 22:506–515. 1998. View Article : Google Scholar : PubMed/NCBI

28 

Matsuno T, Jung SK, Matsumoto Y, Saito M and Morikawa J: Preferential cytotoxicity to tumor cells of 3,5-diprenyl-4-hydroxycinnamic acid (artepillin C) isolated from propolis. Anticancer Res. 17A:3565–3568. 1997.

29 

Konishi Y, Hitomi Y, Yoshida M and Yoshioka E: Absorption and bioavailability of artepillin C in rats after oral administration. J Agric Food Chem. 53:9928–9933. 2005. View Article : Google Scholar : PubMed/NCBI

30 

Konishi Y: Transepithelial transport of artepillin C in intestinal Caco-2 cell monolayers. Biochim Biophys Acta. 1713:138–144. 2005. View Article : Google Scholar : PubMed/NCBI

31 

Berman HM, Battistuz T, Bhat TN, Bluhm WF, Bourne PE, Burkhardt K, Feng Z, Gilliland GL, Iype L, Jain S, et al: The Protein Data Bank. Acta Crystallogr D Biol Crystallogr. 58:899–907. 2002. View Article : Google Scholar : PubMed/NCBI

32 

Bolton EE, Wang Y, Thiessen PA and Bryant SH: Chapter 12-PubChem: Integrated platform of small molecules and biological activities. Annu Rep Comput Chem. 4:217–241. 2008. View Article : Google Scholar

33 

Van Der Spoel D, Lindahl E, Hess B, Groenhof G, Mark AE and Berendsen HJ: GROMACS: Fast, flexible, and free. J Comput Chem. 26:1701–1718. 2005. View Article : Google Scholar : PubMed/NCBI

34 

van der Spoel D, van Maaren PJ and Caleman C: GROMACS molecule & liquid database. Bioinformatics. 28:752–753. 2012. View Article : Google Scholar : PubMed/NCBI

35 

Lindorff-Larsen K, Piana S, Palmo K, Maragakis P, Klepeis JL, Dror RO and Shaw DE: Improved side-chain torsion potentials for the Amber ff99SB protein force field. Proteins. 78:1950–1958. 2010.PubMed/NCBI

36 

van der Spoel D, Lindahl E, Hess B, Van Buuren AR, Apol E, Meulenhoff PJ, Tieleman DP, Sijbers ALTM, Feenstra KA, van Drunen R, et al: GROMACS user manual version 3.3. 2008, http://ftp.gromacs.org/pub/manual/manual-3.3.pdfurisimpleftp://ftp.gromacs.org/pub/manual/manual-3.3.pdf.

37 

van Gunsteren W, Billeter S, Eising A, Hünenberger P, Krüger P, Mark A, Scott W and Tironi I: Gromos43a1. Hochschulverlag AG an der ETH Zürich; Zürich: 1996

38 

Nobushi Y, Oikawa N, Okazaki Y, Tsutsumi S, Park YK, Kurokawa M and Yasukawa K: Determination of Artepillin-C in Brazilian propolis by HPLC with photodiode array detector. J Pharm Nutr Sci. 2:127–131. 2012.

39 

Ryu J, Kaul Z, Yoon AR, Liu Y, Yaguchi T, Na Y, Ahn HM, Gao R, Choi IK, Yun CO, et al: Identification and functional characterization of nuclear mortalin in human carcinogenesis. J Biol Chem. 289:24832–24844. 2014. View Article : Google Scholar : PubMed/NCBI

40 

Grover A, Priyandoko D, Gao R, Shandilya A, Widodo N, Bisaria VS, Kaul SC, Wadhwa R and Sundar D: Withanone binds to mortalin and abrogates mortalin-p53 complex: Computational and experimental evidence. Int J Biochem Cell Biol. 44:496–504. 2012. View Article : Google Scholar

41 

Lu WJ, Lee NP, Kaul SC, Lan F, Poon RT, Wadhwa R and Luk JM: Mortalin-p53 interaction in cancer cells is stress dependent and constitutes a selective target for cancer therapy. Cell Death Differ. 18:1046–1056. 2011. View Article : Google Scholar : PubMed/NCBI

42 

Nagpal N, Goyal S, Dhanjal JK, Ye L, Kaul SC, Wadhwa R, Chaturvedi R and Grover A: Molecular dynamics-based identification of novel natural mortalin-p53 abrogators as anticancer agents. J Recept Signal Transduct Res. 37:8–16. 2017. View Article : Google Scholar

43 

Markiewicz-Żukowska R, Car H, Naliwajko SK, Sawicka D, Szynaka B, Chyczewski L, Isidorov V and Borawska MH: Ethanolic extract of propolis, chrysin, CAPE inhibit human astroglia cells. Adv Med Sci. 57:208–216. 2012. View Article : Google Scholar

44 

Machado BA, Barreto GA, Costa AS, Costa SS, Silva RP, da Silva DF, Brandão HN, da Rocha JL, Nunes SB, Umsza-Guez MA, et al: Determination of parameters for the supercritical extraction of antioxidant compounds from green propolis using carbon dioxide and ethanol as co-solvent. PLoS One. 10:e01344892015. View Article : Google Scholar : PubMed/NCBI

45 

Takara K, Fujita M, Matsubara M, Minegaki T, Kitada N, Ohnishi N and Yokoyama T: Effects of propolis extract on sensitivity to chemotherapeutic agents in HeLa and resistant sublines. Phytother Res. 21:841–846. 2007. View Article : Google Scholar : PubMed/NCBI

46 

Akanda MJ, Sarker MZ, Ferdosh S, Manap MY, Ab Rahman NN and Ab Kadir MO: Applications of supercritical fluid extraction (SFE) of palm oil and oil from natural sources. Molecules. 17:1764–1794. 2012. View Article : Google Scholar : PubMed/NCBI

47 

Szente L and Szejtli J: Highly soluble cyclodextrin derivatives: Chemistry, properties, and trends in development. Adv Drug Deliv Rev. 36:17–28. 1999. View Article : Google Scholar

48 

Charumanee S, Okonogi S, Sirithunyalug J, Wolschann P and Viernstein H: Effect of cyclodextrin types and co-solvent on solubility of a poorly water soluble drug. Sci Pharm. 84:694–704. 2016. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

March-2018
Volume 52 Issue 3

Print ISSN: 1019-6439
Online ISSN:1791-2423

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Bhargava P, Grover A, Nigam N, Kaul A, Doi M, Ishida Y, Kakuta H, Kaul SC, Terao K, Wadhwa R, Wadhwa R, et al: Anticancer activity of the supercritical extract of Brazilian green propolis and its active component, artepillin C: Bioinformatics and experimental analyses of its mechanisms of action. Int J Oncol 52: 925-932, 2018
APA
Bhargava, P., Grover, A., Nigam, N., Kaul, A., Doi, M., Ishida, Y. ... Wadhwa, R. (2018). Anticancer activity of the supercritical extract of Brazilian green propolis and its active component, artepillin C: Bioinformatics and experimental analyses of its mechanisms of action. International Journal of Oncology, 52, 925-932. https://doi.org/10.3892/ijo.2018.4249
MLA
Bhargava, P., Grover, A., Nigam, N., Kaul, A., Doi, M., Ishida, Y., Kakuta, H., Kaul, S. C., Terao, K., Wadhwa, R."Anticancer activity of the supercritical extract of Brazilian green propolis and its active component, artepillin C: Bioinformatics and experimental analyses of its mechanisms of action". International Journal of Oncology 52.3 (2018): 925-932.
Chicago
Bhargava, P., Grover, A., Nigam, N., Kaul, A., Doi, M., Ishida, Y., Kakuta, H., Kaul, S. C., Terao, K., Wadhwa, R."Anticancer activity of the supercritical extract of Brazilian green propolis and its active component, artepillin C: Bioinformatics and experimental analyses of its mechanisms of action". International Journal of Oncology 52, no. 3 (2018): 925-932. https://doi.org/10.3892/ijo.2018.4249