Extracts of Bridelia ovata and Croton oblongifolius induce apoptosis in human MDA‑MB‑231 breast cancer cells via oxidative stress and mitochondrial pathways

  • Authors:
    • Juthathip Poofery
    • Bungorn Sripanidkulchai
    • Ratana Banjerdpongchai
  • View Affiliations

  • Published online on: January 31, 2020     https://doi.org/10.3892/ijo.2020.4973
  • Pages: 969-985
Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Breast cancer is the most common type of cancer and is also the second leading cause of cancer‑associated death in women worldwide. Thus, there is an urgent requirement for the development of effective treatments for this disease. Bridelia ovata and Croton oblongifolius are herbs used in Thai traditional medicine that have been used to treat various health problems; B. ovata has traditionally been used as a purgative, an antipyretic, a leukorrhea treatment and as a birth control herb. C. oblongifolius has been used to increase breast milk production, for post‑partum care (where it is used as a hot bath herb), and as a treatment for flat worms and dysmenorrhea. However, there is little research investigating the anticancer properties of these herbs. The present study aimed to investigate the anticancer properties of crude ethyl acetate extracts of B. ovata (BEA) and C. oblongifolius (CEA) in order to explore their underlying mechanisms in breast cancer cell death. The phytoconstituents of the crude extracts of BEA and CEA were studied using gas chromatography‑mass spectrometry (GC‑MS). GC‑MS analysis showed that the primary compound in BEA is friedelan‑3‑one, and kaur‑16‑en‑18‑oic acid in CEA. Cytotoxicity was investigated using an MTT assay, both BEA and CEA showed greater toxicity against MDA‑MB‑231 breast cancer cells compared with their effect on MCF10A normal epithelial mammary cells. BEA and CEA exerted various effects, including inducing apoptotic cell death, reducing mitochondrial transmembrane potential, increasing the levels of intracellular ROS, activating caspases, upregulating pro‑apoptotic and downregulating anti‑apoptotic genes and proteins. BEA and CEA were shown to have anticancer activity against breast cancer cells and induce apoptosis in these cells via a mitochondrial pathway and oxidative stress.

Introduction

Breast cancer is the most common type of cancer and the leading cause of cancer-associated mortality and morbidity in women worldwide (1,2). The mortality rate for this disease has decreased due to early diagnosis, but treatment outcomes of patients with advanced breast cancer remain poor (3). There are currently several options for the treatment of breast cancer such as surgery, radiotherapy and chemotherapy, but their use is limited by adverse effects, drug resistance and rapid cancer cell metastasis (4). Therefore, an effective treatment for certain aggressive forms of breast cancer, such as inflammatory breast cancer and triple-negative breast cancer (TNBC), is still required. TNBC cannot be treated with hormonal therapy, as TNBC cells do not express receptors for estrogen, progesterone and epidermal growth factor (5). Despite some considerable limitations, the leading cancer treatments have generally improved over the years; though the financial cost remains high (6). Hence, cancer research should move towards developing and applying herbal treatment strategies with lower costs and fewer side effects to patients.

Bridelia ovata and Croton oblongifolius, from the Euphorbiaceae family, are traditional herbs found in Thailand; Thai traditional medicine uses the dried leaf of B. ovata as an expectorant and a laxative, whereas the bark is used as a medicinal astringent (7). Few previous phytochemical studies have demonstrated the presence of triterpenes and phytosterols in B. ovata (8,9); its crude ethanolic extract has been shown to inhibit the invasive and migratory abilities of HepG2 cells (10), thus the present study focused on the identification of these compounds in B. ovata.

C. oblongifolius has been used as a medicinal plant for the treatment of several diseases and ailments; for example, the leaf is used as a blood tonic for general weakness and poor appetite, and the flower is used to treat flatworm infections. Furthermore, the fruit is used to alleviate dysmenorrhea, the seed acts as a purgative (11) and the root is used in the treatment of dysentery (12). In addition, C. oblongifolius bark has been used for the treatment of dyspepsia, chronic liver enlargement and remittent fever (13). Moreover, in Thai traditional medicine, C. oblongifolius is utilized in combination with C. sublyratus to treat gastric ulcers and gastric cancers (14,15). Several types of phytochemical have been isolated from C. oblongifolius, including megastigmane glycosides (16), labdane diterpenoids (11,17), clerodanes (18-21), halimanes (22), cembranes (21,23,24), pimarane-types (acantoic acid) (25) and (-)-ent-kaur-16-en-19-oic acid (12). A study investigating clerodanes from C. oblongifolius showed that this phytochemical has cytotoxic effect against human cancer cell lines, including HepG2, SW620, CHAGO, KATO3 and BT474 (17). However, there are currently no reports of the activity of B. ovata and C. oblongifolius ethyl acetate extracts against TNBC cells, and to the best of our knowledge, the underlying mechanisms mediating the induction of cancer cell death remain unknown.

In the present study, apoptosis (programmed cell death) was investigated. Apoptosis is an important process for normal tissue homeostasis (26), and any disruption of the underlying mechanisms may cause the onset of a variety of diseases, including autoimmunity, neurodegenerative diseases and cancer (27). Therefore, therapeutic strategies targeting apoptosis have been applied to treat these diseases (28). The two primarily routes of apoptosis are the extrinsic and intrinsic pathways; the extrinsic, or death receptor pathway is induced by the binding of death ligands (such as FasL, TRAIL and TNF) to specific receptors (such as FAS, TRAILR1-2 and TNFR1) at the cell membrane (29). Subsequently, the intracellular death domain of the receptor recruits the adaptor FAS-associated death domain protein and the initiator caspase 8. These form the death-inducing signaling complex (DISC). DISC proteolytically auto-activates caspase 8, which then cleaves and activates the effector caspase 3 to induce apoptosis (30,31).

The intrinsic, or mitochondrial pathway, is triggered by diverse stimuli, including DNA damage, chemotherapeutic drugs and oxidative stress. These stimuli induce the loss of mitochondrial transmembrane potential, permitting cytochrome c release from the mitochondria intermembranous space into the cytosol, resulting in apoptosome formation. The apoptosome is a protein complex comprising cytochrome c, apoptosis-activating factor 1, adenosine triphosphate (ATP), and procaspase-9 (32). The apoptosome autoactivates caspase 9, which in turn cleaves and activates caspase 3 (33), resulting in apoptosis. Mitochondrial apoptosis is controlled by the Bcl-2 protein family. This comprises anti-apoptotic proteins, including Bcl-2, Bcl-xL and Mcl-1, and two other groups of pro-apoptotic proteins, including the multi-domain proteins BAX and Bcl-2 homologous antagonist/killer (BAK), and BH3-only proteins Bcl-2-like protein 11 (BIM), and phorbol-12-myristate-13-acetate-induced protein 1 (NOXA). During apoptosis, the expression of anti-apoptotic molecules is downregulated, and regulatory BH3-only proteins induce the formation of transport channels on the outer mitochondrial membrane, comprised of dimers of BAX and/or BAK (34). These channels promote cytochrome c release and the sequential activation of caspase 9 and 3 (35). Therefore, the aim of the present study was to investigate the pro-apoptotic potential of BEA and CEA against MDA-MB-231 triple-negative breast cancer cells. This cell line has the highest potential for invasiveness (36) and is commonly used as a model for late-stage breast cancer (37).

Materials and methods

Plant materials

The trunks of B. ovata and C. oblongifolius were collected from northeastern Thailand by the plant genetic conservation project under the Royal Initiative of Her Royal Highness Princess Maha Chakri Sirindhorn (RSPG). The plants were extracted using ethyl acetate as a solvent, and authenticated by Professor Dr. Bungorn Sripanidkulchai (Center for Research and Development of Herbal Health Products, Department of Pharmaceutical Chemistry, Faculty of Pharmaceutical Sciences, Khon Kaen University, Thailand); the voucher numbers of B. ovata and C. oblongifolius are TT-OC-SK-1253 and TT-OC-SK-1215, respectively. The collected plants were washed, cut and dried, and then crushed into powders. The extracts were obtained using ethyl acetate at a plant-to-solvent ratio of 1:5 at room temperature for 72 h, filtered using Whatman filter paper no. 1 and then centrifuged at 500 × g for 10 min to obtain the supernatant. The extracts were then concentrated under a rotary evaporator and the final products were weighed prior to determination of the percentage yield. For the cell treatment, the crude ethyl acetate extracts were dissolved in dimethyl sulfoxide (DMSO) to obtain an appropriate concentration of the extract stock solutions (to avoid DMSO-associated cellular toxicity). The final concentrations of DMSO under the treatment conditions were less than 0.2%, at which they were not cytotoxic (38).

Reagents and assay kits

The chemicals and solvents used were of analytical grade, which included chloroform, ethyl acetate, ferric chloride, hydrochloric acid, sulfuric acid, acetic acid anhydride, sodium hydroxide, copper acetate, potassium iodide and mercuric chloride (all Merck KGaA). Dulbecco's modified Eagle's medium (DMEM), DMEM/Ham's F-12, fetal bovine serum (FBS), horse serum (HS), phosphate-buffered saline (PBS), trypsin-EDTA solution, penicillin and streptomycin were purchased from Gibco; Thermo Fisher Scientific, Inc. 3-(4,5-Dimethythiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT), the Annexin V-fluorescein isothiocyanate (FITC)/propidium iodide (PI) kit, 3,3′-dihexyloxacarbocyanine iodide (DiOC6), dichlorodihydrofluorescein diacetate (DCFH-DA), dihydroethidium (DHE), DMSO, epidermal growth factor (EGF), cholera toxin, hydrocortisone and insulin were purchased from Sigma-Aldrich; Merck KGaA. Caspase 9 (LEHD-paranitroaniline; LEHD-p-NA), caspase 8 (IETD-p-NA) and caspase 3 (DEVD-p-NA) activity colorimetric assay kits were obtained from Invitrogen; Thermo Fisher Scientific, Inc. The primers and RevertAid First Strand cDNA Synthesis kit were obtained from Thermo Fisher Scientific, Inc. The RNA Prep Pure kit was purchased from Tiangen Biotech Co., Ltd and the SensiFAST SYBR Lo-ROX kit was purchased from Bioline; Meridian Biosciences. The mitochondria/cytosol fractionation kit, primary antibodies against NOXA (ab13654), BAX (ab32503), Bcl-xL (ab2568), cytochrome c (ab6400), β-actin (ab8227), COX-4 (abl6056) and Rip3 (ab56164), rabbit HRP-conjugated secondary antibody (ab97051), and mouse HRP-conjugated secondary antibody (ab97046) were purchased from Abcam. The ECL Prime Western Blotting System was purchased from GE Healthcare.

Phytochemical screening

The ethyl acetate crude extracts were screened for the presence of bioactive chemical constituents, including alkaloids, flavonoids, phytosterols, terpenoids, saponins, phenols and tannins using the standard methods outlined below (39-41). The results were reported as -, absent; +, present; ++ abundant.

Detection of alkaloids

Dry extracts (50 mg each) were dissolved in 1 ml of 2% hydrochloric acid and filtered using Whatman filter paper no. 1. The filtrates were treated with 100 μ1 potassium mercuric iodide, and a yellow-colored precipitate indicated the presence of alkaloids.

Detection of flavonoids

Dry extracts (100 mg each) were dissolved in 4 ml ethyl acetate and filtered using Whatman filter paper no. 1. The filtrates were treated with 200 μl sodium hydroxide solution followed by 1% sulfuric acid. The formation of an intense yellow color became colorless when 1% sulfuric acid was added, which indicated the presence of flavonoids.

Detection of phytosterols

Dry extracts (50 mg each) were dissolved in 1 ml chloroform and filtered using Whatman filter paper no. 1. The filtrates were treated with 100 μl acetic anhydride, boiled at 90-100°C and then subsequently cooled at room temperature. Then, 1 ml 95% sulfuric acid was added and the formation of a brown ring at the junction indicated the presence of phytosterols.

Detection of triterpenoids

Dry extracts (50 mg each) were dissolved in 1 ml chloroform and filtered using Whatman filter paper no. 1. The filtrates were treated with 100 μ1 95% sulfuric acid, gently shaken and then allowed to stand for 5 min. Formation of a golden-yellow color indicated the presence of triterpenoids.

Detection of diterpenes

Dry extracts (50 mg each) were dissolved in 1 ml distilled water and treated with 200 μ1 copper acetate solution. The Appearance of an emerald green color indicated the presence of diterpenes.

Detection of saponins

Dry extracts (50 mg each) were dissolved in 20 ml distilled water and shaken in a graduated cylinder for 15 min. The formation of a 1-cm layer of foam indicated the presence of saponins.

Detection of phenols

Dry extracts (50 mg each) were dissolved in 1 ml distilled water and treated with 100 μl 2% ferric chloride. The formation of a bluish-green or black coloration confirmed the presence of phenols.

Detection of tannins

Dry extracts (50 mg each) were boiled at 90-100°C in 1 ml distilled water for 5 min and then filtered using Whatman filter paper no. 1; 100 μ1 0.1% ferric chloride was added to the filtrates, and the appearance of a brownish-green or a blue-black color indicated the presence of tannins.

Gas chromatography-mass spectrometry (GC-MS) analysis

BEA and CEA were subjected to GC-MS detection, which was performed using a GC 7890A attached to a MSD 5975C (Agilent Technologies, Inc.). The chromatograph was fitted with a HP5-MS capillary column (30 m × 0.25 mm, 0.25-μm film thickness) using a temperature program initially set to 100-150°C, at a rate of 3°C/min for 16.67 min. Then, the temperature was raised to 300°C at a rate of 5°C/min, for 30 min, and then held at 300°C for 10 min; the total run time was 56.67 min. The injector was set at 280°C and the injection volume was at 1.0 μ1 in the split mode (2:1). The flow speed of helium, the carrier gas, was 1 ml/min. The mass spectra were scanned from m/z 50-550, and the electron impact ionization energy was set at 70 eV. Identification of the components was based on the comparison of their mass spectra and retention times with those of the standard authentic compounds. This was accomplished using Agilent ChemStation Integrator (Agilent Technologies, Inc.) software computer matching using the National Institute Standard and Technology Library (nist.gov/nist-research-library).

Cell culture

The human MDA-MB-231 breast cancer cell line was obtained from the American Type Culture Collection (ATCC). The cells were cultured in DMEM and supplemented with 10% FBS, 100 U/ml penicillin and 100 μg/ml streptomycin. The human MCF10A mammary epithelial cell line was obtained from the ATCC and cultured in DMEM/Ham's F-12 supplemented with 5% HS, 20 ng/ml EGF, 0.5 mg/ml hydrocortisone, 100 ng/ml cholera toxin, 10 μg/ml insulin, 50 U/ml penicillin and 50 μg/ml streptomycin. Both cell lines were cultured at 37°C in a 5% CO2 atmosphere.

Cytotoxicity determination by MTT assay

MDA-MB-231 and MCF10A cells were seeded into 96-well plates at 5×104 cells/ml and treated with BEA, CEA or Doxorubicin at various concentrations. The cells were then incubated for 24 h at 37°C before detecting cell viability using the MTT method, as previously described (42). Briefly, MTT reagent was added to the cells, which were then incubated for 4 h at 37°C. The formazan crystals were dissolved using DMSO and the optical density was measured at 540 nm (with a reference wavelength of 630 nm). The results were expressed as a percentage of cell viability and were compared with untreated control cells. The inhibitory concentration (IC)10, IC20 and IC50 values were calculated in the presence of each of the extracts.

Apoptosis investigation using annexin V-FITC and PI staining

MDA-MB-231 cells were treated with BEA and CEA at IC10, IC20 and IC50 for 24 h, at 37°C, and then stained with annexin V-FITC and PI for 15 min. The percentages of apoptotic cells were quantified by annexin V-FITC positive staining (43), detected using the BD FACSAria™ III Cell Sorter flow cytometer (BD Biosciences) and then analyzed with BD FACSDiva software version 7.0 (BD Biosciences). The treated cell results were compared with untreated control cells.

Assessment of mitochondrial transmembrane depolarization

To detect changes in mitochondrial transmembrane potential (A1 I'm), MDA-MB-231 cells were treated with BEA and CEA for 24 h at 37°C. BEA and CEA-treated and untreated control cells were stained using DiOC6 at a final concentration of 40 nM for 15 min at 37°C. Then, the cells were washed and analyzed using a FACSCAtia™ III Cell Sorter flow cytometer (BD Biosciences) (44).

Measurement of reactive oxygen species (ROS) generation

The levels of intracellular ROS were measured, and included those of superoxide anion, hydrogen peroxide and peroxide radicals; two intracellular redox-sensitive fluorescent dyes, DHE and DCFH-DA, were used according to previously described methods (45). Briefly, after the MDA-MB-231 cells were treated with BEA and CEA at IC10, IC20 and IC50 for 12 h, the cells were stained with DHE and DCFH-DA at a final concentration of 2 for 30 min at 37°C. Fluorescence intensity was measured using a fluorescence microplate reader (BioTek Instruments, Inc.) at 535 nm excitation and 635 nm emission wavelengths for DHE, and 485 nm excitation and 525 nm emission wavelengths for DCFH-DA (46,47). The results were compared between treated and untreated cells. 3% H2O2 was used as the positive control for DCFH-DA staining, whereas FeSO4 was used as the positive control for DHE staining.

Determination of caspase activities

Caspase 3, 8 and 9 activities were measured in cellular protein extracts using colorimetric assay kits. Following treatment with BEA and CEA at IC10, IC20 and IC50, the MDA-MB-231 cells were lysed and the total protein was quantified using a Bradford protein assay. Caspase activity was detected using the specific substrates DEVD-p-NA, IETD-p-NA and LEHD-p-NA for caspase 3, 8 and 9, respectively. The absorbance was then determined at 405 nm using a microplate reader (BioTek Instruments, Inc.) (44). The results were compared between treated and untreated cells.

Analysis of gene expression levels by reverse transcription-quantitative (RT-qPCR)

After treatment of the MDA-MB-231 cells with BEA and CEA at IC10, IC20 and IC50 (for 24 h at 37°C), the total RNA was extracted using the Tiangen RNA Prep Pure kit, and then reverse transcribed into cDNA using the RevertAid First Strand cDNA Synthesis kit. mRNA expression levels of NOXA, BAX and Bcl-xL were quantified using the SensiFAST SYBR Lo-ROX kit (Bioline; Meridian Biosciences) and the 7500 Fast Real-time PCR system (Applied Biosystems; Thermo Fisher Scientific, Inc.). The relative gene expression levels were analyzed using the 2-ΔΔCq method (48) with GAPDH as the housekeeping gene. The primers for NOXA, BAX, Bcl-xL and GAPDH were as follows: NOXA, forward 5′GCT GGA AGT CGAGTG TGC TA3′ and reverse 5′CCTGAGCAGAAGAGTTTG GA3′; BAX, forward 5′AGC AGATCATGAAGACAGGGG3′ and reverse 5′ACACTCGCT CAGCTTCTTGG3′; Bcl-xL, forward 5′GCTGGGACACTTT TGT GGAT3′ and reverse 5′TGT CTG GTCACT TCC GAC TG3′; and GAPDH, forward 5′TGCACCACCAACTGCTTAGC3′ and reverse 5′GGCATGGACTGTGGTCATGAG3′. The primers were designed and selected using the Primer-BLAST tool (ncbi.nlm.nih.gov/tools/primer-blast/) (48,49).

Western blotting

To determine the protein expression levels in MDA-MB-231 cells, the cells were seeded into culture dishes and treated with BEA or CEA for 24 h at IC10, IC20 and IC50; untreated MDA-MB-231 cells were used as the control. Following treatment, the cells were harvested and lysed using RIPA lysis buffer supplemented with a protease inhibitor. The mitochondria/cytosol fractionation kit was used to extract cytosolic and mitochondrial proteins per the manufacturer's protocol. Protein concentrations were measured using a Bradford protein assay (50), and protein expression levels were assessed using western blot analysis (51). A total of 20 μg (for NOXA, BAX, Bcl-xL, Rip3, COX-IV and p-actin) and 80 μg (for cytochrome c) total protein per lane was loaded onto a 12% gel, resolved using SDS-PAGE and subsequently transferred onto nitrocellulose membranes. The membranes were blocked with 5% skim milk at room temperature for 1 h, and then incubated with primary antibodies against NOXA, BAX, Bcl-xL, cytochrome c, Rip3 and COX-IV (1:1,000) at 4°C overnight (B-actin was used at 1:20,000). Next, the membranes were washed and incubated with HRP-conjugated secondary antibodies (1:5,000) for 2 h at room temperature. For visualization, an ECL Prime Western Blotting System was used and antibody-specific chemiluminescent bands were detected using an Omega Lum W Imager (Gel Company, Inc.). Quantification of the band density was conducted using ImageJ 1.52p (National Institutes of Health). B-actin and COX-IV were used as the loading controls for whole cell/cytosolic and mitochondrial proteins, respectively.

Statistical analysis

The data are presented as the mean ± standard deviation (SD), calculated from three independent experiments (unless otherwise stated). The statistical differences between multiple-group comparisons were determined using the Kruskal-Wallis test and Dunn's multiple comparisons post-hoc test, and P<0.05 was considered to indicate a statistically significant difference. All statistical analyses were performed using SPSS version 22.0. (IBM Corp.).

Results

Percentage yield and phytochemical screening of BEA and CEA

The percentage yields of BEA and CEA were 0.31 and 0.95%, respectively. For phytochemical screening of the extracts, standard methods were applied (41). BEA contained alkaloids, flavonoids, phytosterols, triterpenoids, saponins, phenols and tannins with an absence of diterpenoids. The CEA results revealed the presence of alkaloids, flavonoids, phytosterols, triterpenoids, diterpenoids and tannins with an absence of saponins and phenols. The results of these qualitative phytochemical analyses are presented in Table I.

Table I

Phytochemical screening test of BEA and CEA.

Table I

Phytochemical screening test of BEA and CEA.

PhytoconstituentBEACEA
Alkaloids++
Flavonoids++++
Phytosterols+++
Triterpenoids+++
Diterpenoids++
Saponins+
Phenols++
Tannins+++

[i] BEA, Bridelia ovata; CEA, Croton oblongifolius; −, absent; +, present; ++ abundant.

Identification of BEA and CEA compounds using GC-MS

The characterization of BEA and CEA extracts are presented in Table II. A total of 24 BEA compounds and 12 CEA compounds were identified. Their respective retention times and percentages of the total are stated. The GC-MS chromatogram and structures of the 3 major identified compounds of each extract are presented in Fig. 1. The GC-MS results show that the primary phytoconstituent of BEA is friedelan-3-one (20.1%) and that the primary active compound of CEA is kaur-16-en-18-oic acid (36.1%).

Table II

Phytochemical analyses and identified compound profiles with retention time, name and percentage of total in BEA and CEA treatment.

Table II

Phytochemical analyses and identified compound profiles with retention time, name and percentage of total in BEA and CEA treatment.

A, BEA
Retention time (min)Compounds% of total
8.948 2-methoxy-4-vinylphenol1.62
10.960 1,2,3-benzenetriol7.70
11.618Benzaldehyde, 4-hydroxy-3-methoxy-1.15
16.653Benzene, 1,2,3-trimethoxy-5 (2-propenyl)0.43
20.070Benzaldehyde, 4-hydroxy-3,5-dimethoxy-0.43
23.150Tetradecanoic acid0.88
28.006Palmitic acid10.28
30.986Linoleic acid0.25
31.429Oleic acid7.92
31.733Stearic acid0.63
37.568 Di-n-octylphthalate0.72
42.9571-Nonadecene0.69
45.6741-Hexacosanol2.59
45.862α-Tocopherol1.03
46.593Cirsilineol0.83
47.059Campesterol4.18
47.477Stigmasterol7.74
48.060 24-n-propylidenecholesterol1.06
48.433 Stigmast-5-en-3-ol19.55
48.558Fucosterol1.09
49.689 Stigmasta-3,5-dien-7-one1.18
50.367 Stigmast-4-en-3-one0.92
50.522 9,19-Cyclolanostan-3-ol, 24-methylene-, (3.beta.)-0.80
52.247 Friedelan-3-one20.09

B, CEA
Retention time (min)Compounds% of total
5.589Benzoic acid4.54
27.688Palmitic acid2.38
29.082Biformen1.15
29.210 (-)-Kaur-16-ene2.00
33.331 4-Hydroxy-3,5,4′-trimethoxystilbene3.25
36.634Kaur-16-en-18-oic acid36.06
36.794 9-Isopropylidene-5,6,7,8-tetramethyl-1,4-dihydro-1,4-methano-naphthalene2.66
43.416Aflatoxin G12.54
47.055Campesterol1.25
47.413Stigmasterol2.52
48.286 Stigmast-5-en-3-ol4.70
49.673 Stigmasta-3,5-dien-7-one3.06

[i] BEA, Bridelia ovata; CEA, Croton oblongifolius.

Cytotoxicity of BEA and CEA in MDA-MB-231 and MCF10A cells

BEA and CEA exhibited a cytotoxic effect on MDA-MB-231 and MCF10A cells. The percentage of cell viability was calculated compared with the untreated control cells (100% cell viability). The IC10, IC20, IC50 and selectivity index (SI) of BEA, CEA and doxorubicin for both cell types were calculated (Table III). BEA and CEA significantly decreased MDA-MB-231 cell viability in a dose-dependent manner (Fig. 2A and B). The IC50 of BEA and CEA were 40.0±3.6 and 29.7±3.4 μg/ml, respectively. However, BEA and CEA also decreased the viability of MCF10A cells (Fig. 2D and E) but at higher concentrations compared with their effect on MDA-MB-231 cells. The results are in accordance with the selectivity index criteria: SI>3=high selectivity and SI<3=less selectivity (52,53). Therefore, BEA possessed a high SI (9.84), whereas CEA had less selectivity (2.76). Doxorubicin, the conventional chemotherapeutic drug used for breast cancer (54), was used to treat both cell types (Fig. 2C and F). The IC50 of doxorubicin in MCF10A cells was higher than that in MDA-MB-231 cells and the SI of both BEA and CEA were higher than doxorubicin (2.13). This suggests that these two extracts have more selectivity towards breast cancer cells than noncancerous cells, which may relate to a reduction in adverse effects.

Table III

IC10, IC20 and IC50 values in MDA MB 231 cells and IC50 in MCF10A cells treated with BEA, CEA and doxorubicin.

Table III

IC10, IC20 and IC50 values in MDA MB 231 cells and IC50 in MCF10A cells treated with BEA, CEA and doxorubicin.

A, MDA MA 231
ConcentrationBEA ± SD (µg/ml)CEA ± SD (µg/ml)Doxorubicin ± SD (µM)
IC107.6±1.93.9±0.7-
IC2014.4±2.48.5±1.3-
IC5040.0±3.629.7±3.44.1±0.8

B, MCF10A

ConcentrationBEA ± SD, µg/mlCEA ± SD, µg/mlDoxorubicin ± SD (µM)

IC50393.8±4.382.0±4.38.7±0.8
SIa9.82.82.1

a SI values of each agent, which are ratio of IC50 of MCF10A/MDA MB-231 cells. IC, inhibitory concentration; SI, selectivity index; BEA, Bridelia ovata; CEA, Croton oblongifolius; SD, standard deviation.

Mode of cell death induced by BEA and CEA in MDA-MB-231 cells

Annexin V-FITC/PI staining was employed to determine the mode of cell death in MDA-MB-231 cells after treatment with both extracts. Dot plot analyses show the percentage of cells in each quadrant (Fig. 3A). The percentages of cells in early (Q4) and late apoptosis (Q2) were significantly increased compared with the control, whereas the percentage of living cells (Q3) was significantly decreased after BEA treatment at IC50 concentration compared with the control; similar results were observed following CEA treatment (Fig. 3B and C). However, the number of necrotic cells in Q1 was increased, but not significantly compared with the control cells; supplementary data demonstrates that necroptosis marker protein RIP3 did not change the expression levels following BEA and BEA treatment (Fig. S1). Therefore, BEA and CEA induced MDA-MB-231 cell death via the apoptosis pathway.

Reduction of Δψm and intracellular ROS generation

Mitochondrial dysfunction is a key event in the intrinsic apoptosis pathway, and depolarization of Δψm, which occurs during apoptosis (55), was measured to determine the level of mitochondrial disruption induced by BEA and CEA. The percentage of MDA-MB-231 cells with Δψm loss was significantly increased after treatment with BEA and CEA at IC50 (Fig. 4B). The reduction of Δψm indicates mitochondrial dysfunction. The intracellular ROS were measured using two fluorescent dyes. DCFH-DA is widely used probe for detecting H2O2 and oxidative stress and DHE is another used probe for detecting intracellular O•-2 (56). The fluorescent intensity of both DCFH-DA and DHE probes were increased following BEA and CEA treatment, compared with the untreated control (Fig. 5), thus H2O2 and O•-2 were generated. However, the level of ROS generation did not increase in a dose-dependent manner (though they were significantly increased). This indicates that BEA and CEA induced oxidative stress and mitochondrial dysfunction during MDA-MB-231 cell apoptosis.

Effect of BEA and CEA on caspase activity

In MDA-MB-231 cells, caspase 3 and 9 activities were significantly increased following treatment with BEA and CEA (Fig. 6A and C), while caspase 8 activity was not significantly altered, compared with untreated control cells (Fig. 6B). These results indicate that BEA and CEA induce MDA-MB-231 apoptosis through the intrinsic pathway.

Effects of BEA and CEA treatment on Bcl-2 family mRNA expression levels

Bcl-2 family proteins play an important role in the mitochondrial pathway (57). Expression levels of NOXA and BAX were significantly increased in cells treated with BEA and CEA, whereas Bcl-xL mRNA expression levels were significantly attenuated compared with those of the control cells (Fig. 7). These results indicate that BEA and CEA induce NOXA expression, which subsequently downregulates Bcl-xL and upregulates BAX expression, inducing apoptosis.

Effects of BEA and CEA treatment on apoptotic-related protein expression levels

The expression levels of NOXA, BAX, Bcl-xL and cytochrome c proteins were investigated in MDA-MB-231 cells following BEA and CEA treatment. Protein expression levels of NOXA and BAX were significantly increased, whereas Bcl-xL expression levels were significantly decreased following BEA and CEA treatment (Fig. 8). In addition, cytochrome c expression levels were measured in the mitochondria and cytosol. Following treatment with BEA and CEA, cytochrome c expression levels in the mitochondria decreased in a dose-dependent manner, while those in the cytosol were increased (Fig. 9). These results suggest that BEA and CEA induce the mitochondrial apoptosis pathway in MDA-MB-231 cells.

Discussion

B. ovata and C. oblongifolius are medicinal plants used in Thai traditional medicine. These herbs are used in the treatment of several diseases and symptoms, including as an expectorant (7), and to treat flatworm infections and dysmenorrhea (11). However, previous studies have also reported the cytotoxicity of these herbs against some cancer cell lines (10,17). However, there are no reports investigating the underlying mechanisms of action of these two herbs for potential use in breast cancer treatment. In the present study, BEA and CEA were screened for the presence of phytoconstituent groups using standard methods. The phytochemical screening results showed several groups of phytochemical compounds, such as alkaloids, flavonoids, phytosterols and tannins. However, previous phytochemical reports of B. ovata showed the presence of triterpenoids and phytosterols, whereas those of C. oblongifolius revealed diterpenoids and megastigmane glycosides. Therefore, GC-MS analysis in the present study primarily aimed to identify triterpenoids, phytosterols and diterpenoids in BEA and CEA. To the best of our knowledge, this is first report of the phytoconstituents of crude ethyl acetate extracts of BEA and CEA by GC-MS analysis.

Friedelan-3-one is a friedelane-type triterpenoid, which was a primary phytoconstituent of BEA in the present study. A previous study reported the cytotoxic effect of friedelan-3-one against mouse 4T1 breast cancer cells (58). Friedelan-3-one also inhibits VEGF-induced Kaposi's sarcoma cell proliferation and induces apoptotic cell death (59). These findings indicate the cytotoxic effect and apoptotic induction of BEA in MDA-MB-231 cells, which may be associated with friedelan-3 -one.

Kaur-16-en-18-oic acid (or kaurenoic acid) is a kaurane-type diterpenoid, which is present in some medicinal plants that possess potential anti-cancer properties (60). Kaurane diterpenes have been considered for the development of novel effective anti-cancer chemotherapeutic agents (61). Kaurenoic acid is also reported to possess apoptosis-inducing properties against various cell lines, including the HL-60 human leukemia cell line (62), the PC-3 human prostate cancer cell line and the A-549 human lung cancer cell line (63). Furthermore, the genotoxicity of kaurenoic acid has been demonstrated in gastric cancer cell lines, including ACP02, ACP03, AGP01 and PG100 (64), and cervical cancer cell lines such as HeLa, CaSki and C33A (65). Moreover, the epimer of kaurenoic acid isolated from Croton antisyphiliticus exhibits moderate cytotoxic activity against HeLa cervical cancer cells and B-16 murine melanoma cells, by inducing apoptosis (66).

In the present study, the second most abundant compound in both BEA and CEA was the phytosterol stigmast-5-en-3-ol. The anti-proliferative and pro-apoptotic effects of stigmast-5-en-3-ol (from Dendronephthya gigantean) have been demonstrated in HL-60 and MCF-7 cells (67). Diterpenes, triterpenes and phytosterols have shown anti-tumor activity in several models (68-71), hence they may have value as novel anti-tumor agents. In the present study, high contents of friedelan-3-one and stigmast-5-en-3-ol in BEA, and kaur-16-en-18-oic acid in CEA, were observed. Therefore, these two herbs show potential as anti-breast cancer agents.

However, in the present study, crude ethyl acetate extracts of B. ovata and C. oblongifolius were investigated for their anti-cancer activity as an alternative to purified forms of the constituents identified. The present study aimed to investigate all of the constituents in these herbal extracts on the basis of possible synergy required to generate the anti-cancer effects. A previous study demonstrated that the IC50 values of crude ethanolic extracts of B. ovata and C. oblongifolius in the HepG2 cell line (72) were higher than the crude ethyl acetate extracts in the MDA-MB-231 cell line in the present study. These differences suggest that these two herbs have more potential as ethyl acetate extracts against breast cancer cells than ethanolic extracts against liver cancer; however, further study of these extracts against liver cancer is required. Comparing the IC50 and SI values of BEA and CEA suggests that CEA exerts greater toxic against MDA-MB-213 cells than BEA, whereas BEA acts more selectively towards breast cancer cells than CEA.

Annexin V-FITC and PI staining showed that MDA-MB-231 cells had undergone apoptosis and necrosis. However, the number of necrotic cells was not significantly different compared with that of the control cells, and Rip3 protein expression levels did not change following BEA and CEA treatment. The percentage of apoptotic cells was significantly increased following treatment with both extracts in a dose-dependent manner, compared with that of the control cells.

The present study further investigated the mechanism of apoptosis, which revealed that the percentage of cells with mitochondrial disruption increased following treatment with both extracts. The mitochondria are the primary source of intracellular ROS production. High generation of ROS induces mitochondrial dysfunction, cell damage and apoptosis (73). This Δψm loss was associated with ROS generation, as the high energy source equates to the mitochondrial electron transport chain (74). Intracellular ROS induce cellular stress and consequently induce apoptosis (75). During apoptosis, the loss of mitochondrial membrane potential leads to the inhibition of ATP synthase, subsequently resulting in ROS production (76). However, a number of drugs and natural compounds can directly induce ROS generation; this includes doxorubicin, which also induces mitochondrial dysfunction and apoptosis (77,78). In the present study, the level of ROS increased ~1.5-fold in MDA-MB-213 cells treated with BEA and CEA, compared with the control group, demonstrating that BEA and CEA induce oxidative stress, and subsequent apoptotic cell death. ROS level increased in a dose-independent manner, which may be due to the presence of antioxidants in the cell or the cellular microenvironment (79). Alternatively, this may be due to the oxidant content of the microenvironment, which contains superoxide dismutase (80), reduced glutathione (81), catalase, bilirubin, NADPH oxidase and lipoxygenase and toxins (82,83).

Cellular stress can lead to DNA damage that activates the p53 tumor suppressor gene; this can induce apoptotic cell death and cell cycle arrest. However, the downstream effects of p53 activation depend on cell type, differentiation state and the local microenvironment (84). It has been reported that p53-mediated apoptosis is favorably selected in cells expressing oncoproteins (85); for example, mouse embryonic fibroblasts (MEFs) normally undergo p53-mediated cell cycle arrest, whereas MEFs expressing E1A oncoprotein undergo p53-mediated apoptosis (86). p53-mediated apoptosis involves the transcriptional targets PUMA and NOXA, and the expression of other BH-3 only proteins (85,87). Several studies have reported that NOXA functions as an integral mediator of p53-dependent apoptosis (88). Moreover, PUMA and NOXA indirectly induce mitochondrial outer membrane permeabilization (MOMP) by direct binding to members of the anti-apoptotic Bcl-2 family, such as Bcl-2 and Bcl-xL, which interferes with pro-apoptotic BAX and BAK and anti-apoptotic interactions. Consequently, BAX serves a role in MOMP by homo- or heterodimerizing with BAK, forming pores in the mitochondrial outer membrane and permitting the release of cytochrome c from the mitochondrial intermembranous space into the cytosol (89). This subsequently leads to the activation of initiator caspase 9, resulting in downstream activation of executioner caspase 3, which proteolyzes key proteins, including those of cytoskeletal proteins. Also, caspase 3 inactivates other enzymes such as poly(ADP) ribose polymerase (PARP), a DNA repair enzyme (90). Finally, the cells undergo apoptotic cell death and exhibit unique morphological changes, including DNA fragmentation and membrane blebbing (91).

The present study investigated the effects of BEA and CEA on MDA-MB-231 human breast cancer cells, showing that treatment with these herbs increased the mRNA and protein expression levels of the pro-apoptotic proteins NOXA and BAX. By contrast, BEA and CEA treatment led to the downregulation of the anti-apoptotic protein Bcl-xL. Additionally, cytochrome c, which is essential for apoptosome formation and the progression of apoptosis (92), was released from the mitochondria into the cytosol. These data show that BEA and CEA induced breast cancer cell death via the mitochondria-mediated pathway. Caspases are a group of cysteine proteases which play a pivotal role in the apoptosis pathway (93). The mitochondrial dysfunction observed following BEA and CEA treatment demonstrated that only the intrinsic apoptotic pathway was activated by these herbs, and that there was no evidence of a significant increase in caspase 8 activity, which would indicate the activation of the extrinsic pathway. By contrast, significantly enhanced caspase 9 activity was observed. Caspase 3 activity was activated by active caspase 9, and this resulted in MDA-MB-231 cell apoptosis. Therefore, BEA and CEA induced apoptotic cell death via the mitochondrial apoptosis pathway.

To conclude, crude BEA and CEA extracts induced MDA-MB-231 human breast cancer cell apoptosis via oxidative stress and mitochondria-mediated pathways. The altered expression levels of NOXA, BAX and Bcl-xL resulted in disruption of the mitochondrial membrane potential and the release of cytochrome c into the cytosol. This led to caspase 9 and 3 activation and induced apoptotic cell death. The pathway by which BEA and CEA induce MDA-MB-231 cell apoptosis is illustrated in Fig. 10. The lack of in vivo experiments using animal models and patient studies are the primary limitations of the present study. Therefore, further investigation into the associated active compounds in these extracts is required, and in vivo experiments need to be performed before these herbs can be developed as therapeutic agents for human breast cancer treatment.

Supplementary Data

Funding

The present study was supported by The Research and Researchers for Industries (grant no. PHD58I0008), and the Research Fund for Graduate Student of Faculty of Medicine, Chiang Mai University (grant no. 038/2018).

Availability of data and materials

The datasets used and/or analyzed during the present study are available from the corresponding author on reasonable request.

Authors' contributions

JP and RB designed the study. BS collected the herbs and took responsibility for the extraction. JP performed the biochemical experiments and wrote the manuscript. All authors read and approved the final manuscript.

Ethical approval and consent to participate

Not applicable.

Patient consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Abbreviations:

BEA

Bridelia ovata ethyl acetate extract

CEA

Croton oblongifolius ethyl acetate extract

ΔΨm

mitochondrial transmembrane potential

NOXA

phorbol-12-myristate-13-acetate-induced protein 1 or PMAIP1

BAX

Bcl2-associated X protein

Bcl-xL

B-cell lymphoma-extra large

Acknowledgments

The group would like to thank The Medical Science Research Equipment Center, Faculty of Medicine, Chiang Mai University.

References

1 

DeSantis CE, Ma J, Goding Sauer A, Newman LA and Jemal A: Breast cancer statistics, 2017, racial disparity in mortality by state. CA Cancer J Clin. 67:439–448. 2017. View Article : Google Scholar : PubMed/NCBI

2 

Jemal A, Bray F, Center MM, Ferlay J, Ward E and Forman D: Global cancer statistics. CA Cancer J Clin. 61:69–90. 2011. View Article : Google Scholar : PubMed/NCBI

3 

Sledge GW, Mamounas EP, Hortobagyi GN, Burstein HJ, Goodwin PJ and Wolff AC: Past, present, and future challenges in breast cancer treatment. J Clin Oncol. 32:1979–1986. 2014. View Article : Google Scholar : PubMed/NCBI

4 

Moiseenko F, Volkov N, Bogdanov A, Dubina M and Moiseyenko V: Resistance mechanisms to drug therapy in breast cancer and other solid tumors: An opinion. F1000Res. 6:2882017. View Article : Google Scholar : PubMed/NCBI

5 

Liu W, Fu X, Yang Z, Li S, Cao Y, Li Q and Luan J: Moderate intermittent negative pressure increases invasiveness of MDA-MB-231 triple negative breast cancer cells. Breast. 38:14–21. 2018. View Article : Google Scholar

6 

Sun L, Legood R, Dos-Santos-Silva I, Gaiha SM and Sadique Z: Global treatment costs of breast cancer by stage: A systematic review. PLoS One. 13:e02079932018. View Article : Google Scholar : PubMed/NCBI

7 

Chotchoungchatchai S, Saralamp P, Jenjittikul T, Pornsiripongse S and Prathanturarug S: Medicinal plants used with thai traditional medicine in modern healthcare services: A case study in Kabchoeng Hospital, Surin Province, Thailand. J Ethnopharmacol. 141:193–205. 2012. View Article : Google Scholar : PubMed/NCBI

8 

Boonyaratavej S, Tantayanontha S, Kitchanachai P, Chaichantipyuth C, Chittawong V and Miles DH: Trans-triacontyl-4-hydroxy-3-methoxycinnamate, a new compound from the thai plant bridelia ovata. J Nat Prod. 55:1761–1763. 1992. View Article : Google Scholar

9 

Thongkorn N: Chemical constituents of the leaves of bridelia ovata decne (unpublished PhD thesis). Chulalongkorn University; 1995

10 

Baig H, Diskul-Na-Ayudthaya P, Weeraphan C, Paricharttanakul M, Svasti J and Srisomsap C: Inhibitory effect of bridelia ovata decne extract on HepG2 cell migration and invasion stimulated by fibroblast-conditioned media. Naresuan Phayao J. 1:6–10. 2015.

11 

Sommit D, Petsom A, Ishikawa T and Roengsumran S: Cytotoxic activity of natural labdanes and their semi-synthetic modified derivatives from Croton oblongifolius. Planta Med. 69:167–170. 2003. View Article : Google Scholar : PubMed/NCBI

12 

Ngamrojnavanich N, Sirimongkon S, Roengsumran S, Petsom A and Kamimura H: Inhibition of Na+,K+-ATPase activity by (-)-ent-Kaur-16-en-19-oic acid and its derivatives. Planta Med. 69:555–556. 2003. View Article : Google Scholar : PubMed/NCBI

13 

Ahmed B, Alam T, Varshney M and Khan SA: Hepatoprotective activity of two plants belonging to the Apiaceae and the euphor-biaceae family. J Ethnopharmacol. 79:313–316. 2002. View Article : Google Scholar : PubMed/NCBI

14 

Salatino A, Salatino MLF and Negri G: Traditional uses, chemistry and pharmacology of Croton species (Euphorbiaceae). J Braz Chem Soc. 18:11–33. 2007. View Article : Google Scholar

15 

Singh M, Pal M and Sharma RP: Biological activity of the labdane diterpenes. Planta Med. 65:2–8. 1999. View Article : Google Scholar : PubMed/NCBI

16 

Takeshige Y, Kawakami S, Matsunami K, Otsuka H, Lhieochaiphant D and Lhieochaiphant S: Oblongionosides A-F, megastigmane glycosides from the leaves of Croton oblongifolius Roxburgh. Phytochemistry. 80:132–136. 2012. View Article : Google Scholar : PubMed/NCBI

17 

Roengsumran S, Petsom A, Kuptiyanuwat N, Vilaivan T, Ngamrojnavanich N, Chaichantipyuth C and Phuthong S: Cytotoxic labdane diterpenoids from Croton oblongifolius. Phytochemistry. 56:103–107. 2001. View Article : Google Scholar : PubMed/NCBI

18 

Pudhom K and Sommit D: Clerodane diterpenoids and a trisubstituted furan from Croton oblongifolius. Phytochem Lett. 4:147–150. 2011. View Article : Google Scholar

19 

Roengsumran S, Musikul K, Petsom A, Vilaivan T, Sangvanich P, Pornpakakul S, Puthong S, Chaichantipyuth C, Jaiboon N and Chaichit N: Croblongifolin, a new anticancer clerodane from Croton oblongifolius. Planta Med. 68:274–277. 2002. View Article : Google Scholar : PubMed/NCBI

20 

Youngsa-ad W, Ngamrojanavanich N, Mahidol C, Ruchirawat S, Prawat H and Kittakoop P: Diterpenoids from the roots of Croton oblongifolius. Planta Med. 73:1491–1494. 2007. View Article : Google Scholar : PubMed/NCBI

21 

Pudhom K, Vilaivan T, Ngamrojanavanich N, Dechangvipart S, Sommit D, Petsom A and Roengsumran S: Furanocembranoids from the stem bark of Croton oblongifolius. J Nat Prod. 70:659–661. 2007. View Article : Google Scholar : PubMed/NCBI

22 

Roengsumran S, Pornpakakul S, Muangsin N, Sangvanich P, Nhujak T, Singtothong P, Chaichit N, Puthong S and Petsom A: New halimane diterpenoids from Croton oblongifolius. Planta Med. 70:87–89. 2004. View Article : Google Scholar : PubMed/NCBI

23 

Roengsumran S, Achayindee S, Petsom A, Pudhom K, Singtothong P, Surachetapan C and Vilaivan T: Two new cembranoids from Croton oblongifolius. J Nat Prod. 61:652–654. 1998. View Article : Google Scholar : PubMed/NCBI

24 

Roengsumran S, Singtothong P, Pudhom K, Ngamrochanavanich N, Petsom A and Chaichantipyuth C: Neocrotocembranal from Croton oblongifolius. J Nat Prod. 62:1163–1164. 1999. View Article : Google Scholar : PubMed/NCBI

25 

Suwancharoen S, Tommeurd W, Phurat C, Muangsin N and Pornpakakul S: Acanthoic acid. Acta Crystallogr Sect E Struct Rep Online. 66:o15312010. View Article : Google Scholar : PubMed/NCBI

26 

Renehan AG, Booth C and Potten CS: What is apoptosis, and why is it important? BMJ. 322:1536–1538. 2001. View Article : Google Scholar : PubMed/NCBI

27 

Favaloro B, Allocati N, Graziano V, Di Ilio C and De Laurenzi V: Role of apoptosis in disease. Aging (Albany N Y). 4:330–349. 2012.

28 

Fathi N, Rashidi G, Khodadadi A, Shahi S and Sharifi S: STAT3 and apoptosis challenges in cancer. Int J Biol Macromol. 117:993–1001. 2018. View Article : Google Scholar : PubMed/NCBI

29 

Kumar R, Herbert PE and Warrens AN: An introduction to death receptors in apoptosis. Int J Surg. 3:268–277. 2005. View Article : Google Scholar

30 

O'Brien MA and Kirby R: Apoptosis: A review of pro-apoptotic and anti-apoptotic pathways and dysregulation in disease. J Veterinary Emergency Crit Care. 18:572–585. 2008. View Article : Google Scholar

31 

Schneider P and Tschopp J: Apoptosis induced by death receptors. Pharm Acta Helv. 74:281–286. 2000. View Article : Google Scholar : PubMed/NCBI

32 

Chinnaiyan AM: The apoptosome: Heart and soul of the cell death machine. Neoplasia. 1:5–15. 1999. View Article : Google Scholar

33 

Tait SW and Green DR: Mitochondria and cell death: Outer membrane permeabilization and beyond. Nat Rev Mol Cell Biol. 11:621–632. 2010. View Article : Google Scholar : PubMed/NCBI

34 

Cao K and Tait SWG: Apoptosis and cancer: Force awakens, phantom menace, or both? Int Rev Cell Mol Biol. 337:135–152. 2018. View Article : Google Scholar : PubMed/NCBI

35 

Dong N, Liu X, Zhao T, Wang L, Li H, Zhang S, Li X, Bai X, Zhang Y and Yang B: Apoptosis-inducing effects and growth inhibitory of a novel chalcone, in human hepatic cancer cells and lung cancer cells. Biomed Pharmacother. 105:195–203. 2018. View Article : Google Scholar : PubMed/NCBI

36 

Neve RM, Chin K, Fridlyand J, Yeh J, Baehner FL, Fevr T, Clark L, Bayani N, Coppe JP, Tong F, et al: A collection of breast cancer cell lines for the study of functionally distinct cancer subtypes. Cancer Cell. 10:515–527. 2006. View Article : Google Scholar : PubMed/NCBI

37 

Dai X, Cheng H, Bai Z and Li J: Breast cancer cell line classification and its relevance with breast tumor subtyping. J Cancer. 8:3131–3141. 2017. View Article : Google Scholar : PubMed/NCBI

38 

Chen X and Thibeault S: Effect of DMSO concentration, cell density and needle gauge on the viability of cryopreserved cells in three dimensional hyaluronan hydrogel. Conf Proc IEEE Eng Med Biol Soc. 2013:6228–6231. 2013.PubMed/NCBI

39 

Arega ED: Phytochemical studies of the ethyl acetate extract of the fruit of piper capense. J Pharm Nat Products. 4:148–152. 2018.

40 

Obasi NL, Egbuonu ACC, Ukoha PO and Ejikeme PM: Comparative phytochemical and antimicrobial screening of some solvent extracts of Samanea saman (fabaceae or mimosaceae) pods. Afr J Pure Appl Chem. 4:206–212. 2010.

41 

Tiwari P, Kumar B, Kaur M, Kaur G and Kaur H: Phytochemical screening and extraction: A review. Int Pharm Sci. 1:98–106. 2011.

42 

Banjerdpongchai R, Yingyurn S and Kongtawelert P: Sesamin induces human leukemic cell apoptosis via mitochondrial and endoplasmic reticulum stress pathways. World J Oncol. 1:78–86. 2010.PubMed/NCBI

43 

Wudtiwai B, Sripanidkulchai B, Kongtawelert P and Banjerdpongchai R: Methoxyflavone derivatives modulate the effect of TRAIL-induced apoptosis in human leukemic cell lines. J Hematol Oncol. 4:522011. View Article : Google Scholar : PubMed/NCBI

44 

Khaw-on P and Banjerdpongchai R: Induction of intrinsic and extrinsic apoptosis pathways in the human leukemic MOLT-4 cell line by terpinen-4-ol. Asian Pac J Cancer Prev. 13:3073–3076. 2012. View Article : Google Scholar : PubMed/NCBI

45 

Khaw-On P, Pompimon W and Banjerdpongchai R: Goniothalamin induces necroptosis and anoikis in human invasive breast cancer MDA-MB-231 cells. Int J Mol Sci. 20:E39532019. View Article : Google Scholar : PubMed/NCBI

46 

Dikalov S, Griendling KK and Harrison DG: Measurement of reactive oxygen species in cardiovascular studies. Hypertension. 49:717–727. 2007. View Article : Google Scholar : PubMed/NCBI

47 

Zhao H, Kalivendi S, Zhang H, Joseph J, Nithipatikom K, Vasquez-Vivar J and Kalyanaraman B: Superoxide reacts with hydroethidine but forms a fluorescent product that is distinctly different from ethidium: Potential implications in intracellular fluorescence detection of superoxide. Free Radic Biol Med. 34:1359–1368. 2003. View Article : Google Scholar : PubMed/NCBI

48 

Livak KJ and Schmittgen TD: Analysis of relative gene expression data using real-time quantitative PCR and the 2(-Delta Delta C(T)) method. Methods. 25:402–408. 2001. View Article : Google Scholar

49 

Mitas M, Mikhitarian K, Walters C, Baron PL, Elliott BM, Brothers TE, Robison JG, Metcalf JS, Palesch YY, Zhang Z, et al: Quantitative real-time RT-PCR detection of breast cancer micrometastasis using a multigene marker panel. Int J Cancer. 93:162–171. 2001. View Article : Google Scholar : PubMed/NCBI

50 

Hanf A, Oelze M, Manea A, Li H, Munzel T and Daiber A: The anti-cancer drug doxorubicin induces substantial epigenetic changes in cultured cardiomyocytes. Chem Biol Interact. 313:1088342019. View Article : Google Scholar : PubMed/NCBI

51 

Wudtiwai B, Pitchakarn P and Banjerdpongchai R: Alpha-mangostin, an active compound in Garcinia mangostana, abrogates anoikis-resistance in human hepatocellular carcinoma cells. Toxicol In Vitro. 53:222–232. 2018. View Article : Google Scholar : PubMed/NCBI

52 

Badisa RB, Darling-Reed SF, Joseph P, Cooperwood JS, Latinwo LM and Goodman CB: Selective cytotoxic activities of two novel synthetic drugs on human breast carcinoma MCF-7 cells. Anticancer Res. 29:2993–2996. 2009.PubMed/NCBI

53 

Prayong P, Barusrux S and Weerapreeyakul N: Cytotoxic activity screening of some indigenous Thai plants. Fitoterapia. 79:598–601. 2008. View Article : Google Scholar : PubMed/NCBI

54 

Rivankar S: An overview of doxorubicin formulations in cancer therapy. J Cancer Res Ther. 10:853–858. 2014. View Article : Google Scholar

55 

Kalogeris T, Bao Y and Korthuis RJ: Mitochondrial reactive oxygen species: A double edged sword in ischemia/reperfusion vs. preconditioning Redox Biol. 2:702–714. 2014. View Article : Google Scholar

56 

Kalyanaraman B, Darley-Usmar V, Davies KJ, Dennery PA, Forman HJ, Grisham MB, Mann GE, Moore K, Roberts LJ II and Ischiropoulos H: Measuring reactive oxygen and nitrogen species with fluorescent probes: Challenges and limitations. Free Radic Biol Med. 52:1–6. 2012. View Article : Google Scholar

57 

Tsujimoto Y: Role of Bcl-2 family proteins in apoptosis: Apoptosomes or mitochondria? Genes Cells. 3:697–707. 1998. View Article : Google Scholar

58 

Sousa GFd, Soares DCF, Mussel WdN, Pompeu NFE, Silva GDdF, Vieira Filho SA and Duarte LP: Pentacyclic triterpenes from branches of Maytenus robusta and in vitro cytotoxic property against. J Braz Chem Soc. 25:1338–1345. 2014.

59 

Martucciello S, Balestrieri ML, Felice F, Estevam Cdos S, Sant'Ana AE, Pizza C and Piacente S: Effects of triterpene derivatives from Maytenus rigida on VEGF-induced Kaposi's sarcoma cell proliferation. Chem Biol Interact. 183:450–454. 2010. View Article : Google Scholar

60 

Balbinot RB, de Oliveira JAM, Bernardi DI, Melo UZ, Zanqueta EB, Endo EH, Ribeiro FM, Volpato H, Figueiredo MC, Back DF, et al: Structural characterization and biological evaluation of 18-Nor-ent-labdane diterpenoids from grazielia gaudichaudeana. Chem Biodivers. 16:e18006442019. View Article : Google Scholar : PubMed/NCBI

61 

Cavalcanti BC, Ferreira JR, Moura DJ, Rosa RM, Furtado GV, Burbano RR, Silveira ER, Lima MA, Camara CA, Saffi J, et al: Structure-mutagenicity relationship of kaurenoic acid from Xylopia sericeae (Annonaceae). Mutat Res. 701:153–163. 2010. View Article : Google Scholar : PubMed/NCBI

62 

Cavalcanti BC, Bezerra Dp, Magalhaes HI, Moraes MO, Lima MA, Silveira ER, Camara CA, Rao VS, Pessoa C and Costa-Lotufo LV: Kauren-19-oic acid induces DNA damage followed by apoptosis in human leukemia cells. J Appl Toxicol. 29:560–568. 2009. View Article : Google Scholar : PubMed/NCBI

63 

Cuca LE, Coy ED, Alarcon MA, Fernandez A and Aristizabal FA: Cytotoxic effect of some natural compounds isolated from Lauraceae plants and synthetic derivatives. Biomedica. 31:335–343. 2011. View Article : Google Scholar

64 

Cardoso PCDS, Rocha CAMD, Leal MF, Bahia MO, Alcantara DDFA, Santos RAD, Gongalves NDS, Ambrosio SR, Cavalcanti BC, Moreira-Nunes CA, et al: Effect of diterpenoid kaurenoic acid on genotoxicity and cell cycle progression in gastric cancer cell lines. Biomed Pharmacother. 89:772–780. 2017. View Article : Google Scholar : PubMed/NCBI

65 

Rocha SMMD, Cardoso PCDS, Bahia MO, Pessoa CDO, Soares PC, Rocha SMD, Burbano RMR and Rocha CAMD: Effect of the kaurenoic acid on genotoxicity and cell cycle progression in cervical cancer cells lines. Toxicol In Vitro. 57:126–131. 2019. View Article : Google Scholar : PubMed/NCBI

66 

Fernandes VC, Pereira SI, Coppede J, Martins JS, Rizo WF, Beleboni RO, Marins M, Pereira PS, Pereira AM and Fachin AL: The epimer of kaurenoic acid from Croton antisyphiliticus is cytotoxic toward B-16 and HeLa tumor cells through apoptosis induction. Genet Mol Res. 12:1005–1011. 2013. View Article : Google Scholar : PubMed/NCBI

67 

Fernando IPS, Sanjeewa KKA, Ann YS, Ko CI, Lee SH, Lee WW and Jeon YJ: Apoptotic and antiproliferative effects of Stigmast-5-en-3-ol from Dendronephthya gigantea on human leukemia HL-60 and human breast cancer MCF-7 cells. Toxicol In Vitro. 52:297–305. 2018. View Article : Google Scholar : PubMed/NCBI

68 

Jian B, Zhang H, Han C and Liu J: Anti-cancer activities of diterpenoids derived from Euphorbia fischeriana steud. Molecules. 23:E3872018. View Article : Google Scholar : PubMed/NCBI

69 

Petiwala SM and Johnson JJ: Diterpenes from rosemary (Rosmarinus officinalis): Defining their potential for anti-cancer activity. Cancer Lett. 367:93–102. 2015. View Article : Google Scholar : PubMed/NCBI

70 

Suttiarporn P, Chumpolsri W, Mahatheeranont S, Luangkamin S, Teepsawang S and Leardkamolkarn V: Structures of phytosterols and triterpenoids with potential anti-cancer activity in bran of black non-glutinous rice. Nutrients. 7:1672–1687. 2015. View Article : Google Scholar : PubMed/NCBI

71 

Saleem M: Lupeol, a novel anti-inflammatory and anti-cancer dietary triterpene. Cancer Lett. 285:109–115. 2009. View Article : Google Scholar : PubMed/NCBI

72 

Weerapreeyakul N, Nonpunya A, Barusrux S, Thitimetharoch T and Sripanidkulchai B: Evaluation of the anticancer potential of six herbs against a hepatoma cell line. Chin Med. 7:152012. View Article : Google Scholar : PubMed/NCBI

73 

Sinha K, Das J, Pal PB and Sil PC: Oxidative stress: The mitochondria-dependent and mitochondria-independent pathways of apoptosis. Arch Toxicol. 87:1157–1180. 2013. View Article : Google Scholar : PubMed/NCBI

74 

Cairns RA, Harris IS and Mak TW: Regulation of cancer cell metabolism. Nat Rev Cancer. 11:85–95. 2011. View Article : Google Scholar : PubMed/NCBI

75 

Circu ML and Aw TY: Reactive oxygen species, cellular redox systems, and apoptosis. Free Radic Biol Med. 48:749–762. 2010. View Article : Google Scholar : PubMed/NCBI

76 

Redza-Dutordoir M and Averill-Bates DA: Activation of apoptosis signalling pathways by reactive oxygen species. Biochim Biophys Acta. 1863:2977–2992. 2016. View Article : Google Scholar : PubMed/NCBI

77 

He H, Zang LH, Feng YS, Chen LX, Kang N, Tashiro S, Onodera S, Qiu F and Ikejima T: Physalin A induces apoptosis via p53-Noxa-mediated ROS generation, and autophagy plays a protective role against apoptosis through p38-NF-kB survival pathway in A375-S2 cells. J Ethnopharmacol. 148:544–555. 2013. View Article : Google Scholar : PubMed/NCBI

78 

Yu CC, Ko FY, Yu CS, Lin CC, Huang YP, Yang JS, Lin JP and Chung JG: Norcantharidin triggers cell death and DNA damage through S-phase arrest and ROS-modulated apoptotic pathways in TSGH 8301 human urinary bladder carcinoma cells. Int J Oncol. 41:1050–1060. 2012. View Article : Google Scholar : PubMed/NCBI

79 

Weinberg F, Ramnath N and Nagrath D: Reactive oxygen species in the tumor microenvironment: An overview. Cancers (Basel). 11:E11912019. View Article : Google Scholar

80 

Crompton M: The mitochondrial permeability transition pore and its role in cell death. Biochem J. 341:233–249. 1999. View Article : Google Scholar : PubMed/NCBI

81 

Griffith OW: Biologic and pharmacologic regulation of mammalian glutathione synthesis. Free Radic Biol Med. 27:922–935. 1999. View Article : Google Scholar : PubMed/NCBI

82 

Komonrit P and Banjerdpongchai R: Effect of Pseuderanthemum palatiferum (Nees) radlk fresh leaf ethanolic extract on human breast cancer MDA-MB-231 regulated cell death. Tumour Biol. 40:10104283188001822018. View Article : Google Scholar : PubMed/NCBI

83 

Kumari S, Badana AK, GMM GS and Malla R: Reactive oxygen species: A key constituent in cancer survival. Biomark Insights. 13:11772719187553912018. View Article : Google Scholar : PubMed/NCBI

84 

Kastenhuber ER and Lowe SW: Putting p53 in context. Cell. 170:1062–1078. 2017. View Article : Google Scholar : PubMed/NCBI

85 

Vousden KH and Lu X: Live or let die: The cell's response to p53. Nat Rev Cancer. 2:594–604. 2002. View Article : Google Scholar : PubMed/NCBI

86 

Lowe SW, Ruley HE, Jacks T and Housman DE: p53-dependent apoptosis modulates the cytotoxicity of anticancer agents. Cell. 74:957–967. 1993. View Article : Google Scholar : PubMed/NCBI

87 

Shibue T, Suzuki S, Okamoto H, Yoshida H, Ohba Y, Takaoka A and Taniguchi T: Differential contribution of Puma and Noxa in dual regulation of p53-mediated apoptotic pathways. EMBO J. 25:4952–4962. 2006. View Article : Google Scholar : PubMed/NCBI

88 

Shibue T, Takeda K, Oda E, Tanaka H, Murasawa H, Takaoka A, Morishita Y, Akira S, Taniguchi T and Tanaka N: Integral role of Noxa in p53-mediated apoptotic response. Genes Dev. 17:2233–2238. 2003. View Article : Google Scholar : PubMed/NCBI

89 

Zhang LN, Li JY and Xu W: A review of the role of Puma, Noxa and Bim in the tumorigenesis, therapy and drug resistance of chronic lymphocytic leukemia. Cancer Gene Ther. 20:1–7. 2013. View Article : Google Scholar

90 

McIlwain DR, Berger T and Mak TW: Caspase functions in cell death and disease. Cold Spring Harb Perspect Biol. 5:a0086562013. View Article : Google Scholar : PubMed/NCBI

91 

Saraste A and Pulkki K: Morphologic and biochemical hallmarks of apoptosis. Cardiovasc Res. 45:528–537. 2000. View Article : Google Scholar : PubMed/NCBI

92 

Huttemann M, Pecina P, Rainbolt M, Sanderson TH, Kagan VE, Samavati L, Doan JW and Lee I: The multiple functions of cytochrome c and their regulation in life and death decisions of the mammalian cell: From respiration to apoptosis. Mitochondrion. 11:369–381. 2011. View Article : Google Scholar : PubMed/NCBI

93 

Salvesen GS: Caspases and apoptosis. Essays Biochem. 38:9–19. 2002. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

April-2020
Volume 56 Issue 4

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
Poofery J, Sripanidkulchai B and Banjerdpongchai R: Extracts of Bridelia ovata and Croton oblongifolius induce apoptosis in human MDA‑MB‑231 breast cancer cells via oxidative stress and mitochondrial pathways. Int J Oncol 56: 969-985, 2020
APA
Poofery, J., Sripanidkulchai, B., & Banjerdpongchai, R. (2020). Extracts of Bridelia ovata and Croton oblongifolius induce apoptosis in human MDA‑MB‑231 breast cancer cells via oxidative stress and mitochondrial pathways. International Journal of Oncology, 56, 969-985. https://doi.org/10.3892/ijo.2020.4973
MLA
Poofery, J., Sripanidkulchai, B., Banjerdpongchai, R."Extracts of Bridelia ovata and Croton oblongifolius induce apoptosis in human MDA‑MB‑231 breast cancer cells via oxidative stress and mitochondrial pathways". International Journal of Oncology 56.4 (2020): 969-985.
Chicago
Poofery, J., Sripanidkulchai, B., Banjerdpongchai, R."Extracts of Bridelia ovata and Croton oblongifolius induce apoptosis in human MDA‑MB‑231 breast cancer cells via oxidative stress and mitochondrial pathways". International Journal of Oncology 56, no. 4 (2020): 969-985. https://doi.org/10.3892/ijo.2020.4973