Cordycepin induces apoptosis in human liver cancer HepG2 cells through extrinsic and intrinsic signaling pathways

  • Authors:
    • Le‑Wen Shao
    • Li‑Hua Huang
    • Sheng Yan
    • Jian‑Di Jin
    • Shao‑Yan Ren
  • View Affiliations

  • Published online on: June 13, 2016     https://doi.org/10.3892/ol.2016.4706
  • Pages: 995-1000
Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Cordycepin, also termed 3'-deoxyadenosine, is a nucleoside analogue from Cordyceps sinensis and has been reported to demonstrate numerous biological and pharmacological properties. Our previous study illustrated that the anti‑tumor effect of cordycepin may be associated with apoptosis. In the present study, the apoptotic effect of cordycepin on HepG2 cells was investigated using 4',6-diamidino-2-phenylindole, tetraethylbenzimidazolylcarbocyanine iodide and propidium iodide staining analysis and flow cytometry. The results showed that cordycepin exhibited the ability to inhibit HepG2 cells in a time‑ and dose‑dependent manner when cells produced typical apoptotic morphological changes, including chromatin condensation, the accumulation of sub‑G1 cells and change mitochondrial permeability. A potential mechanism for cordycepin-induced apoptosis of human liver cancer HepG2 cells may occur through the extrinsic signaling pathway mediated by the transmembrane Fas-associated with death domain protein. Apoptosis was also associated with Bcl-2 family protein regulation, leading to altered mitochondrial membrane permeability and resulting in the release of cytochrome c into the cytosol. The activation of the caspase cascade is responsible for the execution of apoptosis. In conclusion, cordycepin‑induced apoptosis in HepG2 cells involved the extrinsic and intrinsic signaling pathway and was primarily regulated by the Bcl-2 family proteins.

Introduction

Cordyceps militaris is a fungus that parasitizes Lepidoptera larvae and has been extensively used as a folk tonic food and crude drug (1). Cordyceps militaris has long been considered to have natural medicinal properties, such as anti-angiogenic, anti-tumor, anti-viral, anti-inflammatory and hypoglycemic effects (18).

Cordycepin, also termed 3′-deoxyadenosine, is a nucleoside analogue from C. militaris (9) and has been reported to demonstrate numerous notable biological and pharmacological properties, including immunological stimulation, anti-cancer and anti-viral effects (1015), a stimulating effect on interlukin-10 production as an immune modulator (16) and preventing hyperlipidermia (17).

Apoptosis, also termed programmed cell death, is a key regulator of tissue homeostasis and is characterized by typical morphological and biochemical hallmarks, including cell shrinkage, membrane blebbing, chromatin condensation and nuclear fragmentation (18). The mechanisms for foreign chemicals disrupting cell functions and causing tissue damage have been associated with the triggering of apoptosis signal transduction pathways, consisting of the intrinsic (mitochondrial) and extrinsic (death receptor) pathways (19).

Although cordycepin has been shown to have a cytotoxic effect on HepG2 cells, the detailed molecular mechanisms have not been well elucidated (17,19). In the present study, the mechanistic understanding of how cordycepin mediates apoptosis in HepG2 cells was investigated and the intrinsic and extrinsic apoptosis pathways were the main focus.

Materials and methods

Materials

Dried Cordyceps militaris was purchased from Cordyceps Garden Biotechnology Company (Fuzhou, China). The macroprous adsorption resin NKA-II was purchased from the Chemical Plant of Nankai University (Tianjin, China). Sulforhodamine B (SRB), 4′,6-diamidino-2-phenylindole (DAPI), propidium iodide (PI) and tetraethylbenzimidazolylcarbocyanine iodide (JC-1) were purchased from Sigma-Aldrich (St. Louis, MO, USA). The following antibodies were purchased from Sangon Biotech Co., Ltd. (Shanghai, China): Mouse monoclonal anti-Fas (1:3,000; catalog no. D198888); rabbit polyclonal anti-Fas ligand (1:3,000; catalog no. D262701); mouse monoclonal anti-Fas associated with death domain protein (FADD; 1:3,000; catalog no. D199671); rabbit polyclonal anti-caspase-8 (1:2,000; catalog no. D155240); rabbit polyclonal anti-caspase-9 (1:2,000; catalog no. D220078); rabbit ployclonal anti-caspase-10 (1:2,000; catalog no. D260010); rabbit monoclonal anti-caspase-3 (1:3,000; catalog no. D120074); mouse monoclonal anti-B-cell lymphoma-2 (Bcl-2; 1:3,000; catalog no. D198628); rabbit polyclonal anti-Bcl-2 associated X protein (Bax; 1:3,000; catalog no. D120073); mouse monoclonal anti-BH3 interacting domain death agonist (Bid; 1:3,000; catalog no. D198911); and rabbit polyclonal anti-cytochrome c (1:2,000; catalog no. D110006). Rabbit polyclonal anti-mouse IgG (1:3,000; catalog no. sc-358920) and goat anti-rabbit IgG (1:3,000; catalog no. sc-2768) antibodies were purchased from Santa Cruz Biotechnology, Inc. (Dallas, TX, USA). The Mitochondrial Membrane Potential Assay kit with JC-1, radioimmunoprecipitation assay (RIPA) lysis buffer and Cell Mitochondria Isolation kit were purchased from Beyotime Institute of Biotechnology (Haimen, China).

Cell lines and culture

HepG2 cell lines were purchased from the Culture Center of the Institute of Basic Medical Sciences of the Chinese Academy of Medical Sciences (Beijing, China). The cells were maintained in Dulbecco's modified Eagle's medium (DMEM; Gibco; Thermo Fisher Scientific, Inc., Waltham, MA, USA) supplemented with 10% fetal bovine serum (Invitrogen; Thermo Fisher Scientific, Inc.), 100 IU/ml penicillin and 100 µg/ml streptomycin (Sigma-Aldrich), and cultured at 37°C for 48 h in a humidified atmosphere consisting of 95% air and 5% CO2.

Isolation and purification of cordycepin

A total of 150 g dried fruiting bodies of C. militaris were cleaned with distilled water, dried in a dark and ventilated place and crushed prior to soaking in 500 ml distilled water overnight at room temperature, and then boiled in distilled water for 3 h. Subsequent to filtration to remove debris, the filtrate (crude aqueous extract) was concentrated in a rotary evaporator (model no. 0010001820; IKA® Works Guangzhou, Guangzhou, China). Ethanol was added and the mixture was left at room temperature overnight. The precipitate obtained subsequent to filtration was discarded and the supernatant was lyophilized to yield the crude sample. The crude extract was chromatographed on a NKA-II column (Sigma-Aldrich) and eluted with 50% (v/v) ethanol to collect the fractions, which were then further purified by high performance liquid chromatography (HPLC). The Lab Alliance HPLC system (model no. 2690/2695; Waters Technologies Ltd., Shanghai, China), including two Series III pumps (model no. 90–2489 rev M; Scientific Systems, Inc. State College, PA, USA) and anabsorbance detector (model no. 500; Waters Technologies Ltd.) set at 254 nm, connected to a Cs420 Hardware integrator (model no. Cs420; Waters Technologies Ltd.). The column was a 250×10 mm reversed-phase Merck Millipore C18 ODS column (Merck Millipore, Darmstadt, Germany) with an internal diameter of 12 µm. The temperature was set at 30°C and the elution conditions were as follows: Flow rate, 0.8 ml/min; and solvent, 15% methanol.

Antiproliferative activity assay of cordycepin

Various concentrations of cordycepin (0, 125, 250 and 500 µM) were inoculated into 96-well microplates when the cell concentration was adjusted to 1×105 cells/ml. Cultures in triplicate were treated with cordycepin for 48 h and the control wells received only maintenance medium (DMEM). Cellular responses were colorimetrically evaluated by an SRB assay. Briefly, the cells were fixed with 25% trichloroacetic acid (Sigma-Aldrich) and washed and stained with 0.4% SRB. Subsequent to the cell-bound SRB being solubilized by the addition of 10 mM Tris-HCl, bound SRB was colorimetrically assessed using an ELISA microplate reader (MK3; Thermo Fisher Scientific, Inc.) at 490 nm. Cell growth inhibition was expressed as a percentage of the untreated control absorbance following the subtraction of the mean background absorbance. Compounds were considered to have potent growth inhibitory activity when the reduction in SRB absorbance was >25% compared with the untreated control cells. The half maximal inhibitory concentration (IC50) values were calculated from the dose-response curves.

Assessment of apoptosis by DAPI staining

HepG2 cells were seeded into a 50 ml culture flask at a density of 1×105 cells/ml for 24 h. Subsequent to treatment with or without cordycepin (0, 125, 250 and 500 µM) at 37°C for 48 h, the cells were collected and washed with phosphate-buffered saline (PBS). Cell pellets were fixed with 4% paraformaldehyde for 10 min and washed three times with PBS. The cells were then incubated with 5 µg/ml DAPI for 20 min. Subsequent to washing with PBS, the cells were observed under a fluorescence microscope (model no. Bx51; Olympus Corporation, Tokyo, Japan) and images were captured.

Assessment of apoptosis by flow cytometry

HepG2 cells cultured with or without cordycepin at 37°C for 48 h were harvested, washed with PBS and fixed with 70% cold ethanol at 4°C for 4 h. The fixed cells were washed and stained with a PI solution containing 20 µg/ml PI and 10 mg/ml RNase (Takara Biotech, Inc., Dalian, China) in PBS for 20 min in the dark. The stained cells were then analyzed by flow cytometry using fluorescence-activated cell sorting (FACS) and the FACSCalibur system (BD Biosciences, Franklin Lakes, NJ, USA). In addition, subsequent to being treated with various concentrations of cordycepin (0, 125, 250 and 500 µM) for 48 h, the harvested HepG2 cells were washed in PBS and incubated in a freshly prepared JC-1 solution at 37°C for 20 min. Cell-associated fluorescence was also measured by FACS.

Western blot analysis

HepG2 cells were treated for 48 h with various concentrations of cordycepin (0, 125, 250 and 500 µM) and harvested and lysed in RIPA lysis buffer that contained 1% NP-40, 0.5% sodium deoxycholate and 0.1% SDS. The cell lysates were resolved by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and electrophoretically transferred to a polyvinylidene fluoride membrane, and then blocked for 1 h with 5% fat-free milk in PBS with Tween-20 (PBST) at room temperature. The membrane was immunoblotted overnight with antibodies in PBST at 4°C. Subsequent to being washed with PBST, the membrane was detected by enhanced chemiluminescence (Kodak, Rochester, NY, USA). Equal loading was confirmed by probing with an antibody against β-actin (1:3000; catalog no. sc-8432; Santa Cruz Biotechnology, Inc.). Mitochondrial and cytosolic fractions were prepared in order to detect cytochrome c. Following treatment with the indicated concentrations of cordycepin for 48 h, HepG2 cells were harvested and disposed using the Cell Mitochondria Isolation kit, according to the manufacturer's protocol. Finally, the mitochondrial pellet and cytosolic supernatant were separated using a centrifuge (model no. 5415C; Eppendorf, Hamburg, Germany).

Results

Isolation and purification of cordycepin

Two fractions from crude extract, termed A and B, were eluted using NKA macroporous resin (data not shown). Fraction A was then purified on a HPLC preparative column followed by an analytical column using CH3OH:H2O = 15:85 (v/v) as the mobile phase. The retention time of fraction A was ~10.5 min, which is consistent with that of standard cordycepin (Fig. 1A). In addition, mass spectrometric analysis revealed that the m/z of fraction A was 252.1096 Da, which corresponds with the m/z of standard cordycepin (Fig. 1B).

Effect of cordycepin on HepG2 cell viability

Cordycepin exhibited the ability to inhibit the proliferation of HepG2 cells in a time- and dose-dependent manner (Fig. 2). The IC50 values for HepG2 cells treated with cordycepin for 24, 48, 72 and 96 h were 735±3.67, 497.5±2.49, 385±1.93 and 307.5±1.54 µM, respectively.

Assessment of apoptosis by DAPI staining

The fluorescence microscopic examination with DAPI staining revealed that morphological changes, such as chromatin condensation, nuclear shrinkage or fragmentation and apoptotic body formation, occurred in HepG2 cells subsequent to being treated with cordycepin for 48 h (Fig. 3A).

Assessment of apoptosis by flow cytometry

The flow cytometric analysis with PI staining revealed that the cells treated with various concentrations of cordycepin for 48 h exhibited an apoptotic hypodiploid sub-G1 peak in a concentration-dependent manner (Fig. 3B). The flow cytometric analysis with JC-1 staining showed the respective green fluorescence intensity ratios, which indicated the apoptotic HepG2 cells, were 5.72, 18.58, 32.12 and 36.65%, respectively (Fig. 4).

Cordycepin-induced apoptosis signaling pathway

Cordycepin increased the expression levels of Fas and FADD, but had no effect on FasL expression. Additionally, the level of procaspase-8 decreased, while the level of cleaved caspase-8 was elevated (Fig. 5A). The expression of the Bcl-2 family proteins was also detected and the results showed that cordycepin enhanced the expression level of truncated Bid (tBid) and decreased the level of Bid, but had little effect on the levels of Bax and Bcl-2 (Fig. 5B). A decreased level of cytochrome c was detected in the mitochondrial fraction and an increased level was detected in the cytosolic fraction (Fig. 5C). Furthermore, the results also showed that HepG2 cells treated with cordycepin for 48 h underwent significant cleavage of precaspase-9 and precaspase-3 (Fig. 5C).

Discussion

As a natural compound, cordycepin plays a key role in cancer therapy and induces apoptosis in numerous cells by targeting molecules and pathways (2022). However, studies investigating the induction of apoptosis by cordycepin through the extrinsic and intrinsic signaling pathways in HepG2 cell are currently available (21,22). The present study investigated the effect of cordycepin on HepG2 cells and the molecular pathway by which cordycepin induces apoptosis. Subsequently, it was demonstrated that cordycepin induced apoptosis by the activation of caspase, interaction between Fas and FADD, and modulation of the protein levels of Bid and tBid.

It is well known that mitochondria play a major role in several stress-induced cell death pathways, and damage to mitochondria with subsequent loss of mitochondrial membrane potential has been known to be the point of no return in apoptotic cascades (20,23). In the present study, it was found that cordycepin had an inhibitory effect on the viability of HepG2 cells by DAPI staining and could increase the fraction of sub-G1 cells by PI staining. The JC-1 assay also revealed that cordycepin-induced collapse of the mitochondrial membrane occurred in a concentration-dependent manner. All these findings demonstrated that cordycepin may induce apoptosis in HepG2 cells.

In mammalian cells, apoptosis occurs via either the extrinsic (receptor-dependent) or the intrinsic (mitochondria-dependent) pathway, with each involving caspase activation. Fas is a TNF-receptor that transduces the apoptotic signal into cells (23,24). FADD is an apoptotic adaptor molecule that recruits caspase-8 or caspase-10 to activated Fas (25). FasL is a cytokine that binds to Fas, and it has been proposed that Fas-mediated apoptosis involves the interaction between Fas and FasL (26). In the present study, cordycepin increased the expression levels of Fas and FADD, without changing the level of FasL, and the expression of cleaved caspase-8 was elevated. Therefore, the present study hypothesizes that cordycepin may stimulate Fas/FADD signaling independently of FasL and subsequently activate caspase-8 in HepG2 cells.

Fas may induce caspase-mediated cleavage of p22 Bid into a major p15, and minor p13 and p11 product; the major proteolytic product p15 tBid allows the release of cytochrome c (2729). The mitochondrial membrane potential was shown to be decreased by the JC-1 staining analysis and confirmed the increased cytosolic cytochrome c by western blotting. The activation cascade of caspase-9 and caspase-3, which are responsible for apoptosis execution, was detected. In particular, caspase-3 plays a central role in the execution of apoptosis and its activation requires the activation of initiator caspases, such as caspase-8 or caspase-9, in response to pro-apoptotic stimuli (30). Caspase-8 is crucial for triggering apoptosis via the death receptor pathway and its activation requires the binding of death ligands to death receptors (31). In addition, in the prior time course assay, the active forms of caspase-8 was detected notably earlier than that of caspase-3 and −9, which clearly demonstrated that caspase-8 may act as an upstream initiator and the extrinsic pathway was involved in cordycepin-induced apoptosis.

The Bcl-2 family proteins play an essential role in apoptotic progress, such as Bcl-2 protein (anti-apoptosis) and Bax protein (pro-apoptosis). These proteins are regulators of mitochondrial membrane permeability and intermembrane space protein efflux, according to the opposing fractions of the anti-apoptosis members and pro-apoptosis members (32,33). Bax accelerates programmed cell death and undergoes a conformational change that causes translocation to the mitochondrial membrane, leading to the release of cytochrome c, which then triggers apoptosis (34). Cordycepin enhanced the expression level of Bax and decreased the expression level of Bid in the present study.

Cleaved Bid causes cytochrome c efflux from mitochondria, which in turn leads to the activation of caspase-9 and caspase-3 (35). In addition, other Bcl-2 family molecules, such as the proapoptotic protein Bax and anti-apoptotic protein Bcl-2, are also key regulators of apoptosis, which control the release of mitochondrial cytochrome c by modulating the permeability of the outer mitochondrial membrane (36). The upregulation of the ratio of Bax/Bcl-2 caused by cordycepin was confirmed by immunoblotting. The release of cytochrome c from mitochondria was also relevant to the dissipation of the mitochondrial membrane potential (ΔΨm) (37). In the present study, cordycepin treatment caused a significant decrease of ΔΨm in HepG2 cells and increased cytosolic cytochrome c. These findings indicated that activation of caspase-3 in cordycepin-induced apoptosis may result from direct caspase-8 cleavage and caspase-8-mediated caspase-3 activation through the intrinsic pathway, which is regulated by the Bcl-2 family proteins.

It has been shown that cordycepin is an analogue of adenosine (22). The possibility that the adenosine receptor is involved in the apoptosis of certain types of cancer cells, including ovarian (38), breast (39) and renal (40) cancer cells, has also been proposed. Thus, additional studies investigating whether cordycepin induces apoptosis through adenosine receptors in HepG2 cells may be valuable.

In conclusion, cordycepin-induced apoptosis in HepG2 cells may be initiated by the FADD mediated signal pathway and regulated by the Bcl-2 family proteins that can change the alternation of mitochondrial membrane permeability and cause the mitochondria mediated apoptosis signal pathway (Fig. 6). To the best of our knowledge, the present study is the first to report apoptosis induced by cordycepin in human liver cancer HepG2 cells.

References

1 

Won SY and Park EH: Anti-inflammatory and related pharmacological activities of cultured mycelia and fruiting bodies of Cordyceps militaris. J Ethnopharmacol. 96:555–561. 2005. View Article : Google Scholar : PubMed/NCBI

2 

Yoo HS, Shin JW, Cho JH, Son CG, Lee YW, Park SY and Cho CK: Effects of Cordyceps militaris extract on angiogenesis and tumor growth. Acta Pharmacol Sin. 25:657–665. 2004.PubMed/NCBI

3 

Hattori M, Isomura S, Yokoyama E, Ujita M and Hara A: Extracellular trypsin-like proteases produced by Cordyceps militaris. J Biosci Bioeng. 100:631–636. 2005. View Article : Google Scholar : PubMed/NCBI

4 

Park C, Hong SH, Lee JY, Kim GY, Choi BT, Lee YT, Park DI, Park YM, Jeong YK and Choi YH: Growth inhibition of U937 leukemia cells by aqueous extract of Cordyceps militaris through induction of apoptosis. Oncol Rep. 13:1211–1216. 2005.PubMed/NCBI

5 

Zhang G, Huang Y, Bian Y, Wong JH, Ng TB and Wagn H: Hypoglycemic activity of the fungi Cordyceps militaris, Cordyceps sinensis, Tricholoma mongolicum and Omphalia lapidescens in streptozotocin-induced diabetic rats. Appl Microbiol Biotechnol. 72:1152–1156. 2006. View Article : Google Scholar : PubMed/NCBI

6 

Ohta Y, Lee JB, Hayashi K, Fujita A, Park DK and Hayashi T: In vivo anti-influenza virus activity of an immunomodulatory acidic polysaccharide isolated from Cordyceps militaris grown on germinated soybeans. J Agric Food Chem. 55:10194–10199. 2007. View Article : Google Scholar : PubMed/NCBI

7 

Lee H, Kim YJ, Kim HW, Lee DH, Sung MK and Park T: Induction of apoptosis by Cordyceps militaris through activation of caspase-3 in leukemia HL-60 cells. Biol Pharm Bull. 29:670–674. 2006. View Article : Google Scholar : PubMed/NCBI

8 

Xu FL, Lee YL, Tsai WY, Lin SJ, Yang ZQ, Yang CC, Liu HY, Cheng L, Xiao H and Wen L: Effect of cordycepin on Hantaan virus 76–118 infection of primary human embryonic pulmonary fibroblasts-characterization of apoptotic effects. Acta Virol. 49:183–193. 2005.PubMed/NCBI

9 

Nakamura K, Konoha K, Yoshikawa N, Yamaguchi Y, Kagota S, Shinozuka K and Kunitomo M: Effect of cordycepin (3′-deoxyadenosine) on hematogenic lung metastatic model mice. In Vivo. 19:137–141. 2005.PubMed/NCBI

10 

Nakamura K, Yoshikawa N, Yamaguchi Y, Kagota S, Shinozuka K and Kunitomo M: Antitumor effect of cordycepin (3′-deoxyadenosine) on mouse melanoma and lung carcinoma cells involves adenosine A3 receptor stimulation. Anticancer Res. 26:43–47. 2006.PubMed/NCBI

11 

Yoshikawa N, Nakamura K, Yamaguchi Y, Kagota S, Shinozuka K and Kunitomo M: Antitumour activity of cordycepin in mice. Clin Exp Pharmacol Physiol. 31(Suppl 2): S51–S53. 2004. View Article : Google Scholar : PubMed/NCBI

12 

Lee SJ, Kim SK, Choi WS, Kim WJ and Moon SK: Cordycepin causes p21WAF1-mediated G2/M cell-cycle arrest by regulating c-Jun N-terminal kinase activation in human bladder cancer cells. Arch Biochem Biophys. 490:103–109. 2009. View Article : Google Scholar : PubMed/NCBI

13 

Zhou X, Luo L, Dressel W, Shadier G, Krumbiegel D, Schmidtke P, Zepp F and Meyer CU: Cordycepin is an immunoregulatory active ingredient of Cordyceps sinensis. Am J Chin Med. 36:967–980. 2008. View Article : Google Scholar : PubMed/NCBI

14 

Shin S, Lee S, Kwon J, Moon S, Lee S, Lee CK, Cho K, Ha NJ and Kim K: Cordycepin suppresses expression of diabetes regulating genes by inhibition of lipopolysaccharide-induced inflammation in macrophages. Immune Netw. 9:98–105. 2009. View Article : Google Scholar : PubMed/NCBI

15 

Thomadaki H, Scorilas A, Tsiapalis CM and Havredaki M: The role of cordycepin in cancer treatment via induction or inhibition of apoptosis: Implication of polyadenylation in a cell type specific manner. Cancer Chemother Pharmacol. 61:251–265. 2008. View Article : Google Scholar : PubMed/NCBI

16 

Zhou X, Meyer CU, Schmidtke P and Zepp F: Effect of cordycepin on interleukin-10 production of human peripheral blood mononuclear cells. Eur J Pharmacol. 453:309–317. 2002. View Article : Google Scholar : PubMed/NCBI

17 

Guo P, Kai Q, Gao J, Lian ZQ, Wu CM, Wu CA and Zhu HB: Cordycepin prevents hyperlipidemia in hamsters fed a high-fat diet via activation of AMP-activated protein kinase. J Pharmacol Sci. 113:395–403. 2010. View Article : Google Scholar : PubMed/NCBI

18 

Lomonosova E and Chinnadurai G: BH3-only proteins in apoptosis and beyond: An overview. Oncogene. 27(Suppl 1): S2–S19. 2008. View Article : Google Scholar : PubMed/NCBI

19 

Seitz SJ, Schleithoff ES, Koch A, Schuster A, Teufel A, Staib F, Stremmel W, Melino G, Krammer PH, Schilling T and Müller M: Chemotherapy-induced apoptosis in hepatocellular carcinoma involves the p53 family and is mediated via the extrinsic and the intrinsic pathway. Int J Cancer. 126:2049–2066. 2010.PubMed/NCBI

20 

Lee SY, Debnath T, Kim SK and Lim BO: Anti-cancer effect and apoptosis induction of cordycepin through DR3 pathway in the human colonic cancer cell HT-29. Food Chem Toxicol. 60:439–447. 2013. View Article : Google Scholar : PubMed/NCBI

21 

Lee HH, Park C, Jeong JW, Kim MJ, Seo MJ, Kang BW, Park JU, Kim GY, Choi BT, Choi YH and Jeong YK: Apoptosis induction of human prostate carcinoma cells by cordycepin through reactive oxygen species-mediated mitochondrial death pathway. Int J Oncol. 42:1036–1044. 2013.PubMed/NCBI

22 

Chen YH, Wang JY, Pan BS, Mu YF, Lai MS, So EC, Wong TS and Huang BM: Cordycepin enhances cisplatin apoptotic effect through caspase/MAPK pathways in human head and neck tumor cells. Onco Targets Ther. 6:983–998. 2013.PubMed/NCBI

23 

Sytwu HK, Liblau RS and McDevitt HO: The roles of Fas/APO-1 (CD95) and TNF in antigen-induced programmed cell death in T cell receptor transgenic mice. Immunity. 5:17–30. 1996. View Article : Google Scholar : PubMed/NCBI

24 

Lien YC, Daosukho C and St Clair DK: TNF receptor deficiency reveals a translational control mechanism for adriamycin-induced Fas expression in cardiac tissues. Cytokine. 33:226–230. 2006. View Article : Google Scholar : PubMed/NCBI

25 

Grunert M, Gottschalk K, Kapahnke J, Gündisch S, Kieser A and Jeremias I: The adaptor protein FADD and the initiator caspase-8 mediate activation of NF-κB by TRAIL. Cell Death Dis. 3:e4142012. View Article : Google Scholar : PubMed/NCBI

26 

Saggioro FP, Neder L, Stavale JN, Paixão-Becker AN, Malheiros SM, Soares FA, Pittella JE, Matias CC, Colli BO, Carlotti CG and Franco M: Fas, FasL and cleaved caspases 8 and 3 in glioblastomas: A tissue microarray-based study. Pathol Res Pract. 210:267–273. 2014. View Article : Google Scholar : PubMed/NCBI

27 

Kaufmann T, Strasser A and Jost PJ: Fas death receptor signalling: Roles of Bid and XIAP. Cell Death Differ. 19:42–50. 2012. View Article : Google Scholar : PubMed/NCBI

28 

Vogel A, Aslan JE, Willenbring H, Klein C, Finegold M, Mount H, Thomas G and Grompe M: Sustained phosphorylation of Bid is a marker for resistance to Fas-induced apoptosis during chronic liver diseases. Gastroenterology. 130:104–119. 2006. View Article : Google Scholar : PubMed/NCBI

29 

Schmich K, Schlatter R, Corazza N, Ferreira Sá K, Ederer M, Brunner T, Borner C and Mergort I: Tumor necrosis factor α sensitizes primary murine hepatocytes to Fas/CD95-induced apoptosis in a Bim- and Bid-dependent manner. Hepatology. 53:282–292. 2011. View Article : Google Scholar : PubMed/NCBI

30 

Liu G, Wang T, Wang T, Song J and Zhou Z: Effects of apoptosis-related proteins caspase-3, Bax and Bcl-2 on cerebral ischemia rats. Biomed Rep. 1:861–867. 2013.PubMed/NCBI

31 

Hu Q, Cui X, Tao L, Xiu L, Wang T and Wang X: Staphylococcus aureus induces apoptosis in primary bovine mammary epithelial cells through Fas-FADD death receptor-linked caspase-8 signaling. DNA Cell Biol. 33:388–397. 2014. View Article : Google Scholar : PubMed/NCBI

32 

Jung EB and Lee CS: Baicalein attenuates proteasome inhibition-induced apoptosis by suppressing the activation of the mitochondrial pathway and the caspase-8- and Bid-dependent pathways. Eur J Pharmacol. 730:116–124. 2014. View Article : Google Scholar : PubMed/NCBI

33 

Wang J and Li W: Discovery of novel second mitochondria-derived activator of caspase mimetics as selective inhibitor of apoptosis protein inhibitors. J Pharmacol Exp Ther. 349:319–329. 2014. View Article : Google Scholar : PubMed/NCBI

34 

Gómez-Crisóstomo NP, López-Marure R, Zapata E, Zazueta C and Martínez-Abundis E: Bax induces cytochrome c release by multiple mechanisms in mitochondria from MCF7 cells. J Bioenerg Biomembr. 45:441–448. 2013. View Article : Google Scholar : PubMed/NCBI

35 

De Martino L, Marfé G, Longo M, Fiorito F, Montagnaro S, Iovane V, Decaro N and Pagnini U: Bid cleavage, cytochrome c release and caspase activation in canine coronavirus-induced apoptosis. Vet Microbiol. 141:36–45. 2010. View Article : Google Scholar : PubMed/NCBI

36 

Kumar P, Coltas IK, Kumar B, Chepeha DB, Bradford CR and Polverini P: Bcl-2 protects endothelial cells against gamma-radiation via a Raf-MEK-ERK-survivin signaling pathway that is independent of cytochrome c release. Cancer Res. 67:1193–1202. 2007. View Article : Google Scholar : PubMed/NCBI

37 

Lei X, Chen Y, Du G, Yu W, Wang X, Qu H, Xia B, He H, Mao J, Zong W, et al: Gossypol induces Bax/Bak-independent activation of apoptosis and cytochrome c release via a conformational change in Bcl-2. Faseb J. 20:2147–2149. 2006. View Article : Google Scholar : PubMed/NCBI

38 

Hajiahmadi S, Panjehpour M, Aghaei M and Shabani M: Activation of A2b adenosine receptor regulates ovarian cancer cell growth: involvement of Bax/Bcl-2 and caspase-3. Biochem Cell Biol. 93:321–329. 2015. View Article : Google Scholar : PubMed/NCBI

39 

Dastjerdi MN, Valiani A, Mardani M and Ra MZ: Adenosine A1 receptor modifies P53 expression and apoptosis in breast cancer cell Mcf-7. Bratisl Lek Listy. 117:242–246. 2016.PubMed/NCBI

40 

Nagaya H, Gotoh A, Kanno T and Nishizaki T: A3 adenosine receptor mediates apoptosis in invitro RCC4-VHL human renal cancer cells by up-regulation AMID expression. J Urol. 189:321–328. 2013. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

August-2016
Volume 12 Issue 2

Print ISSN: 1792-1074
Online ISSN:1792-1082

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Shao LW, Huang LH, Yan S, Jin JD and Ren SY: Cordycepin induces apoptosis in human liver cancer HepG2 cells through extrinsic and intrinsic signaling pathways. Oncol Lett 12: 995-1000, 2016
APA
Shao, L., Huang, L., Yan, S., Jin, J., & Ren, S. (2016). Cordycepin induces apoptosis in human liver cancer HepG2 cells through extrinsic and intrinsic signaling pathways. Oncology Letters, 12, 995-1000. https://doi.org/10.3892/ol.2016.4706
MLA
Shao, L., Huang, L., Yan, S., Jin, J., Ren, S."Cordycepin induces apoptosis in human liver cancer HepG2 cells through extrinsic and intrinsic signaling pathways". Oncology Letters 12.2 (2016): 995-1000.
Chicago
Shao, L., Huang, L., Yan, S., Jin, J., Ren, S."Cordycepin induces apoptosis in human liver cancer HepG2 cells through extrinsic and intrinsic signaling pathways". Oncology Letters 12, no. 2 (2016): 995-1000. https://doi.org/10.3892/ol.2016.4706