Anticancer agent icaritin induces apoptosis through caspase-dependent pathways in human hepatocellular carcinoma cells

  • Authors:
    • Li Sun
    • Qisong Peng
    • Lili Qu
    • Lailing Gong
    • Jin Si
  • View Affiliations

  • Published online on: November 26, 2014     https://doi.org/10.3892/mmr.2014.3007
  • Pages: 3094-3100
Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Icaritin is an active ingredient derived from the plant Herba epimedium, which exhibits various pharmacological and biological activities. However, the function, and the underlying mechanisms of icaritin on the growth of SMMC‑7721 human hepatoma cells have yet to be elucidated. The present study aimed to investigate the function and underlying mechanisms of icaritin in the growth of SMMC‑7721 cells. The cells were treated with varying concentrations of icaritin for 12, 24 and 48 h, respectively, prior to cytotoxic analysis. Apoptosis of SMMC‑7721 cells following treatment with icaritin was measured using flow cytometry. The gene expression of mitochondria‑ and Fas‑mediated caspase‑dependent pathways was detected by reverse transcription‑quantitative polymerase chain reaction and western blotting. Statistical analysis was performed by Student's t‑test and one‑way analysis or variance. The present study demonstrated that treatment with icaritin significantly inhibited growth, and induced apoptosis of SMMC‑7721 cells, in a time‑ and dose‑dependent manner. In addition, icaritin triggered the mitochondrial/caspase apoptotic pathway, by decreasing the Bcl‑2/Bax protein ratio and increasing activation of caspase‑3. Icaritin also activated the Fas‑mediated apoptosis pathway, as was evident by the increased expression levels of Fas and activation of caspase‑8. These data suggest that icaritin may be a potent growth inhibitor and induce apoptosis of SMMC‑7721 cells through the mitochondria‑ and Fas‑mediated caspase‑dependent pathways. The present study may provide experimental evidence for preclinical and clinical evaluations of icaritin for HCC therapy.

Introduction

Hepatocellular carcinoma (HCC) is one of the most common types of malignant tumor worldwide, and is the third leading cause of cancer-associated mortality, after lung and colon cancer (14). It is one of the most aggressive human malignancies and previous data have shown that the five-year survival rate is poor (5). The most effective therapy for HCC is usually surgical resection or liver transplantation; however, in ~70% of patients surgery is unsuitable due to tumor metastasis or liver cirrhosis (6). The standard therapy for HCC is currently chemotherapy, particularly when surgical resection is not suitable (7). However, the current therapeutic options for HCC are not effective, since resistance and tumor relapse eventually develop (8). Novel therapeutic strategies, such as effective chemotherapy with low toxicity, are required, in order to decrease the incidence and improve the prognosis of patients with HCC.

Flavonoids are plant polyphenols that are well-known for their analgesic, physiological antipyretic and anti-inflammatory activities. Flavonoids have recently gained attention due to their antitumor activity (911). Icaritin is a prenylflavonoid and is a hydrolytic product of icaritin, derived from Herba Epimedii. Icaritin exhibits various pharmacological and biological activities, including antirheumatic and antidepressant effects, stimulation of cardiac and neuronal differentiation (1213), prevention of steroid-associated osteonecrosis (14), and inhibition of growth and induction of apoptosis of PC-3 human prostate carcinoma and MCF-7 breast cancer cells (1516).

It has recently been demonstrated that icaritin induces apoptosis of breast and endometrial cancer cells, through the mitochondrial pathway (17,18). However, whether icaritin may trigger the extrinsic pathway, which is mediated by death receptors, remains unclear. The present study aimed to investigate the anticancer activities of icaritin in SMMC-7721 cells, and to identify the underlying mechanisms.

Materials and methods

Drugs and reagents

Icaritin (purity, >98%; Fig. 1A) was purchased from Yousi Biotechnology Co., Ltd. (Shanghai, China). Dulbecco’s modified Eagle’s medium (DMEM), RPMI-1640 medium and newborn calf serum were purchased from Gibco Life Technologies (Carlsbad, CA, USA). An Annexin V-fluorscein isothiocyanate (FITC) Apoptosis Detection kit was purchased from Bipec Biopharma Inc. (Cambridge, MA, USA); and a Bicinchoninic Acid (BCA) Protein Quantitation kit was purchased from GENMED Scientifics Inc., USA (Shanghai, China). Primary antibodies targeting β-actin (rabbit polyclonal antibody, 4967S), cleaved caspase-3 (rabbit monoclonal antibody, Asp175) and cleaved caspase-8 (mouse monoclonal antibody, Asp387) were purchased from Cell Signaling Technology, Inc. (Danvers, MA, USA). Primary antibodies targeting Bax (mouse monoclonal antibody, sc-7480) and extrinsic signal-regulated Fas (rabbit polyclonal antibody, sc-7886) were purchased from Santa Cruz Biotechnology, Inc. (Dallas, TX, USA). All of the other chemicals and solvents used in the present study were of the highest commercially available grade.

Cell culture and icaritin treatment

L02 human liver cells and SMMC-7721 human HCC cells were purchased from Nanjing KeyGen Biotech., Co., Ltd. (Nanjing, China). The cells were cultured in DMEM supplemented with 10% newborn calf serum at 37°C in a humidified atmosphere containing 5% CO2.

Stock solution of icaritin, which could be further diluted with cell culture medium, was prepared in dimethyl sulfoxide (DMSO; Sigma-Aldrich, St. Louis, MO, USA) to a concentration of 10 mM, and maintained at −20°C. The final concentration of DMSO was <0.1% in culture.

Cell proliferation assay

Cell proliferation was analyzed by an 3-(4, 5-dimethylthiazol-2-yl)-2, 5-diphenyltetrazolium bromide assay (Sigma-Aldrich). The L02 and SMMC-7721 cells were seeded into 96-well plates, at a density of 5×103 cells/200 μl/well and incubated at 37°C in a humidified 5% CO2 incubator. The cells were then treated with 0, 5, 10, 20, 40 or 60 μmol/l icaritin for 12, 24 or 48 h, and the plates were returned to standard tissue incubator conditions for an additional 4 h. The media was removed and the cells were solubilized in 150 μl DMSO. The color intensity of formazan was measured at a wavelength of 490 nm, using an automated microplate spectrophotometer (iMark, Bio-Rad, Hercules, CA, USA). The survival rate of the cells was calculated using the following formula: (OD value of the treated group/OD value of untreated group) × 100%. Where OD is the optical density. The assays were performed in triplicate, in three independent experiments.

Cell apoptosis assay by flow cytometry (FCM)

The cells were stained with Annexin V-FITC/propidium iodide (PI) and measured using a FACSCalibur™ flow cytometer (BD FACSCanto™ II Flow Cytometer, BD Biosciences, Franklin Lakes, NJ USA). The samples were washed twice, and adjusted to a concentration of 1×106 cells/ml, with phosphate-buffered saline (PBS). A total of 200 μl of cell suspension was added to each tube, 5 μl Annexin V-FITC and 10 μl PI were added to the labeled tubes, and the samples were incubated for 15 min at room temperature in the dark. Subsequently, 200 μl PBS binding buffer was added to each tube and the samples were analyzed by FCM analysis (BD Biosciences) as soon as possible. The results were analyzed using FlowJo software (version 6.1.3, FlowJo, LLC, Ashland, OR, USA). The rate of apoptosis was calculated as the relative number of apoptotic cells, as compared with the total number of cells.

Reverse transcription-quantitative polymerase chain reaction (RT-qPCR)

The cells were frozen in liquid nitrogen and stored at −80°C, until further use. Bcl-2, Bax and Fas mRNA expression levels were quantified by RT-qPCR using a Reverse Transcription system (Promega Corporation, Madison, WI, USA). The reaction was conducted at 42°C for 60 min.

Total cellular RNA was extracted using TRIzol® reagent (Invitrogen Life Technologies, Carlsbad, CA, USA), according to the manufacturer’s instructions. The primer sequences used were as follows: Sense: 5′-CCACCAAGAAAGCAGGAAAC-3′ and antisense: 5′-GCAGGATAGCAGCACAGGA-3′ for Bcl-2; sense: 5′-CTGAGCTGACCTTGGAGC-3′ and antisense: 5′-GACTCCAGCCACAAAGATG-3′ for Bax; sense: 5′-AGCTTGGTCTAGAGTGAAAA-3′ and antisense: 5′-GAGGCAGAATCATGAGATAT-3′ for Fas; and sense 5′-CCTCTATGCCAACACAGTGC-3′ and antisense 5′-GTACTCCTGCTTGCTGATCC-3′ for β-actin. The mRNA expression levels of Bcl-2 and Bax were analyzed using one-step RT-qPCR, with RNA-direct™ SYBR® Green Realtime PCR Master mix (Toyobo Co., Ltd., Osaka, Japan), according to the manufacturer’s instructions. RNA (2 μl) was amplified under the following conditions: 95°C for 5 min for denaturation followed by 40 cycles of 95°C for 30 sec, 57°C for 45 sec and 72°C for 1 min. The amplification was monitored on an ABI Prism 7500 Real-Time PCR system (Applied Biosystems Life Technologies, Foster City, CA, USA).

Western blot analysis

The SMMC-7721 cells were incubated with 8, 16 or 32 μM icaritin for 24 h. The cells were then lysed with ice-cold lysis buffer (50 mM Tris-HCl, pH 7.4; 150 mM NaCl, 1 mM MgCl2, 100 μg/ml PMSF and 1% Triton X-100) for 30 min on ice. The total proteins were contained within the supernatant following centrifugation at 13,225 × g for 5 min at 4°C, and the protein concentrations were measured using a BCA assay (GENMED Scientifics Inc.). Equal quantities (40 μg) of lysate protein were separated on 10% SDS-PAGE gels, and electrophoretically transferred onto polyvinylidene fluoride membranes (Pall Corporation, East Hills, NY, USA). The membranes were then blocked with 5% non-fat dry milk in TBST buffer (10 mM Tris, pH 7.5, 150 mM NaCl, and 0.05% Tween®-20) for 2 h at room temperature, and probed with 1:1,000 dilutions of the primary antibodies described above, at 4°C overnight. Subsequently, the membranes were incubated with 1:5,000 dilutions of horseradish peroxidase-conjugated secondary antibody to mouse (Sigma-Aldrich, for the detection of Bax or cleaved caspase-8) or rabbit (Santa Cruz Biotechnology, Inc., for the detection of cleaved caspase-3, Fas and β-actin). Protein bands were visualized using an Enhanced Chemiluminescence Detection system (ChemiDoc, Bio-Rad). Band intensity was quantified using BandScan 5.0 software (Glyko, Hayward, CA, USA). All of the western blots were performed at least three times.

Statistical analysis

Statistical analyses were performed using SPSS version 16.0 software (SPSS Inc., Chicago, IL, USA). Each assay was performed at least three times. Data were expressed as the mean ± standard deviation. Student’s t-test and one-way analysis of variance were used to determine statistically significant differences between the data. P<0.05 was considered to indicate a statistically significant difference.

Results

Icaritin exhibits growth inhibitory activity on hepatoma cells, whereas the same dose causes little adverse effect on normal human liver cells

Treatment with icaritin significantly inhibited the growth of SMMC-7721 cells. The half maximal inhibitory concentration (IC50) of icaritin in the SMMC-7721 cells, which underwent treatment for 24 h, was 9.6 μM (Table I). The rate of survival of the SMMC-7721 cells decreased in response to an increasing icaritin concentration (0–60 μM), and this effect was time- and dose-dependent (Fig. 1B). Treatment with the same concentration of icaritin caused little effect on the survival of the L02 normal human liver cells (Fig. 1C). These results indicate that icaritin exerts a growth inhibitory activity on hepatoma cells, whereas the same dose did not have the same effect on normal human liver cells.

Table I

Half maximal inhibitory concentration (IC50) of icaritin in different cell lines.

Table I

Half maximal inhibitory concentration (IC50) of icaritin in different cell lines.

Cell lineIC50 of icaritin (24 h)
LO271.6±4.4
SMMC-77219.6±0.8
Icaritin induces significant apoptosis of SMMC-7721 cells in a time- and dose-dependent manner

Treatment with icaritin markedly induced apoptosis in the SMMC-7721 human hepatoma cells (Fig. 2A). The rate of apoptosis of the SMMC-7721 cells was 3.0±0.3, 7.8±0.9 and 56.5±6.5%, in response to treatment with 8, 16 and 32 μM icaritin, respectively. However, 0–32 μM icaritin did not trigger significant levels of apoptosis in the L02 human normal hepatocyte cells (Fig. 2B and C). In addition, icaritin induced an increased rate of apoptosis in a dose-dependent manner, when the time of treatment was fixed (Fig. 2D). These results suggest that treatment with icaritin markedly induced apoptosis of SMMC-7721 HCC cells in a time- and dose-dependent manner.

Icaritin upregulates the protein expression levels of Bax and cleaved caspase-3 and effects gene expression in SMMC-7721 cells

Treatment of the SMMC-7721 cells with various concentrations of icaritin for 24 h significantly increased the protein expression levels of Bax and cleaved caspase-3 (Fig. 3A). The protein expression levels of Bax and cleaved caspase-3, at the mitochondrial membrane, were significantly increased in response to treatment with icaritin, in a dose-dependent manner (Fig. 3B). Furthermore, treatment with icaritin decreased Bcl-2 and increased Bax mRNA expression levels, resulting in a decreased Bcl-2/Bax ratio (Fig. 3C).

Effects of icaritin on SMMC-7721 apoptosis, by the extrinsic pathway

The results of the present study indicate that icaritin may trigger the mitochondrial pathway of apoptosis in SMMC-7721 cells (Fig. 3). The present study also aimed to determine whether icaritin could induce apoptosis through the death receptor pathway. The protein expression levels of Fas were increased following treatment with icaritin (Fig. 4A). Treatment of the SMMC-7721 cells with 32 μM icaritin for 24 h also induced cleaved caspase-8 protein expression. Icaritin also upregulated the mRNA expression levels of Fas in the SMMC-7721 cells (Fig. 4B).

Discussion

HCC is the third most common cause of cancer-associated mortality, with ~600,000 fatalities worldwide (19). The most effective therapy for HCC is surgical resection; however, <15% of patients can benefit from this treatment, due to the presence of numerous tumor nodules. Until recently, chemotherapy for patients with HCC was toxic and not particularly effective. Thus emphasizing the requirement for further research into the molecular mechanisms of HCC development, and identification of non-cytotoxic and effective antineoplastic compounds for chemoprevention and treatment. Recently, more traditional herbal medicines are being considered for use as anticancer treatments (2022). However, few anticancer agents with low toxicity have been identified to be safe and effective for the treatment of HCC. It has previously been demonstrated that certain Chinese natural ingredients, or herbal formulas, have preventive and therapeutic effects against cancer. In particular, the development of icaritin has gained attention as an antifibrotic and antineoplastic monomer, purified from the herb Epimedium (2325). The present study demonstrated that icaritin may inhibit growth and significantly induce apoptosis in SMMC-7721 cells, with an effective concentration range between 8 and 32 μM. The icaritin-induced growth inhibition and apoptosis were time- and dose-dependent. Furthermore, icaritin at the concentrations mentioned above, hardly affected the growth and apoptosis of L02 non-tumor human hepatocyte cells. These results suggest that icaritin may possess a selective antitumor action.

The present study also investigated the underlying mechanisms of icaritin-induced apoptosis of HCC cells. The expression of a number of genes, including Bcl-2, Bax and Caspase-3 have been revealed to be altered following treatment with icaritin, but the study, which demonstrated this was focussed on the mitochondrial pathway (26). Bax and Bcl-2 family members are critical regulators of mitochondrial function (2730). They may activate or inhibit the release of cytochrome c, which leads to the activation of caspase-3 in the process of apoptosis. Bax exerts pro-apoptotic activity by translocation from the cytosol to the mitochondria, and induces the release of cytochrome c; whereas Bcl-2 exerts anti-apoptotic activity by inhibiting the translocation of Bax to the mitochondria (31). The Bcl-2/Bax ratio is correlated with the extent of apoptosis, and is a crucial factor in the induction of apoptosis (32). In the present study, treatment with icaritin resulted in increased Bax and decreased Bcl-2 expression levels, leading to a decreased Bcl-2/Bax ratio. Caspase-3 protein was cleaved and cleaved caspase-3 expression levels were shown to be increased. The alterations to the Bcl-2/Bax ratio and the expression of cleaved caspase-3, in response to treatment with icaritin, were dose-dependent.

Mitochondria have a pivotal role in the signal transduction of apoptosis (33); however, activation of the mitochondrial pathway is not the only mechanism by which icaritin may induce the apoptosis of HCC cells. The extrinsic pathway is mediated by death receptors, and involves Fas and the binding and activation of caspase-8 (3436). It has previously been reported that the engagement of Fas may promote the formation of the death-inducing signaling complex, resulting in activated T cell apoptosis (37). The Fas ligand activates caspase-8, following interaction with the Fas receptor (38). Activation of caspase-8 may lead to the activation of downstream effector proteases, such as caspase-3, resulting in apoptosis (3940). The present study showed that the expression levels of Fas were significantly increased following treatment with icaritin and the protein expression levels of cleaved caspase-8 were detected following treatment with 32 μM icaritin. These results demonstrate that increased apoptosis may be achieved through the extrinsic pathway in SMMC-7721 cells treated with icaritin.

The results of the present study indicate that icaritin may inhibit growth and induce apoptosis in SMMC-7721 cells. These effects were dose-dependent and not observed in L02 normal hepatocyte cells. The underlying mechanisms of icaritin-induced apoptosis were also examined. The results of the present study indicate that icaritin-induced apoptosis involves the intrinsic mitochondria and extrinsic death receptor pathways. Fas activation was shown to be required for icaritin-induced apoptosis. These data suggest that icaritin may be developed as a promising anticancer agent against human HCC.

Acknowledgements

This study was supported by the National Natural Science Foundation of China (grant no. 81472831), the Medical Key Talent Foundation of Jiangsu Province (grant no. RC2011081), the Medical Key Science and Technology Development Projects of Nanjing (grant no. ZKX11176) and the Science and Technology Development Foundation of Nanjing Medical University of China (grant no. 2011NJMU150).

References

1 

McKillop IH, Moran DM, Jin X and Koniaris LG: Molecular pathogenesis of hepatocellular carcinoma. J Surg Res. 136:125–135. 2006. View Article : Google Scholar : PubMed/NCBI

2 

Bruix J, Boix L, Sala M and Llovet JM: Focus on hepatocellular carcinoma. Cancer Cell. 5:215–219. 2004. View Article : Google Scholar : PubMed/NCBI

3 

Trevisani F, Cantarini MC, Wands JR and Bernardi M: Recent advances in the natural history of hepatocellular carcinoma. Carcinogenesis. 29:1299–1305. 2008. View Article : Google Scholar : PubMed/NCBI

4 

Kusakabe A, Tanaka Y, Orito E, et al: A weak association between occult HBV infection and non-B non-c hepatocellular carcinoma in Japan. J Gastroenterol. 42:298–305. 2007. View Article : Google Scholar : PubMed/NCBI

5 

El-Serag HB, Siegel AB, Davila JA, Shaib YH, Cayton-Woody M, McBride R and McGlynn KA: Treatment and outcomes of treating of hepatocellular carcinoma among Medicare recipients in the United States: a population-based study. J Hepatol. 44:158–166. 2006. View Article : Google Scholar

6 

Llovet JM, Burroughs A and Bruix J: Hepatocellular carcinoma. Lancet. 362:1907–1917. 2003. View Article : Google Scholar : PubMed/NCBI

7 

Plataniotis G and Castiglione M; ESMO Guidelines Working Group. Endometrial cancer : ESMO Clinical Practice Guidelines for diagnosis, treatment and follow-up. Ann Oncol. 2:v41–v45. 2010. View Article : Google Scholar

8 

Gehrig PA and Bae-Jump VL: Promising novel therapies for the treatment of endometrial cancer. Gynecol Oncol. 116:187–194. 2010. View Article : Google Scholar

9 

Middleton E Jr, Kandaswami C and Theoharides TC: The effects of plant flavonoids on mammalian cells: implications for inflammation, heart disease, and cancer. Pharmacol Rev. 52:673–751. 2000.PubMed/NCBI

10 

Li YL, Gan GP, Zhang HZ, Wu HZ, Li CL, Huang YP, Liu YW and Liu JW: A flavonoid glycoside isolated from Smilax china L. rhizome in vitro anticancer effects on human cancer cell lines. J Ethnopharmacol. 113:115–124. 2007. View Article : Google Scholar : PubMed/NCBI

11 

Díaz JG, Carmona AJ, Torres F, Quintana J, Estévez F and Herz W: Cytotoxic activities of flavonoid glycoside acetates from Consolida oliveriana. Planta Med. 74:171–174. 2008. View Article : Google Scholar : PubMed/NCBI

12 

Wang Z, Wang H, Wu J, Zhu D, Zhang X, Ou L, Yu Y and Lou Y: Enhanced co-expression of beta-tubulin III and choline acetyltransferase in neurons from mouse embryonic stem cells promoted by icaritin in an estrogen receptor-independent manner. Chem Biol Interact. 179:375–385. 2009. View Article : Google Scholar : PubMed/NCBI

13 

Wo YB, Zhu DY, Hu Y, Wang ZQ, Liu J and Lou YJ: Reactive oxygen species involved in prenylflavonoids, icariin and icaritin, initiating cardiac differentiation of mouse embryonic stem cells. J Cell Biochem. 103:1536–1550. 2008. View Article : Google Scholar

14 

Zhang G, Qin L, Sheng H, Wang XL, Wang YX, Yeung DK, Griffith JF, Yao XS, Xie XH, Li ZR, Lee KM and Leung KS: A novel semisynthesized small molecule icaritin reduces incidence of steroid-associated osteonecrosis with inhibition of both thrombosis and lipid-deposition in a dose-dependent manner. Bone. 44:345–356. 2009. View Article : Google Scholar

15 

Huang X, Zhu D and Lou Y: A novel anticancer agent, icaritin, induced cell growth inhibition, Gl arrest and mitochondrial transmenbrane potential drop in human prostate carcinoma PC-3 cells. Eur J Pharmacol. 564:26–36. 2007. View Article : Google Scholar : PubMed/NCBI

16 

Wang ZQ and Lou YJ: Proliferation-stimulating effects of icaritin and desmethylicaritin in MCF-7 cells. Eur J Pharmacol. 504:147–153. 2004. View Article : Google Scholar : PubMed/NCBI

17 

Guo Y, Zhang X, Meng J and Wang ZY: An anticancer agent icaritin induces sustained activation of the extracellular signal-regulated kinase (ERK) pathway and inhibits growth of breast cancer cells. Eur J Pharmacol. 658:114–122. 2011. View Article : Google Scholar : PubMed/NCBI

18 

Tong JS, Zhang QH, Huang X, Fu XQ, Qi ST, Wang YP, Hou Y, Sheng J and Sun QY: Icaritin causes sustained ERK1/2 activation and induces apoptosis in human endometrial cancer cells. PLoS One. 6:e167812011. View Article : Google Scholar : PubMed/NCBI

19 

Parkin DM, Bray F, Ferlay J and Pisani P: Global cancer statistics, 2002. CA Cancer J Clin. 55:74–108. 2005. View Article : Google Scholar : PubMed/NCBI

20 

Zhou NN, Tang J, Chen WD, Feng GK, Xie BF, Liu ZC, Yang D and Zhu XF: Houttuyninum, an active constituent of Chinese herbal medicine, inhibits phosphorylation of HER2/neu receptor tyrosine kinase and the tumor growth of HER2/neu-overexpressing cancer cells. Life Sci. 90:770–775. 2012. View Article : Google Scholar : PubMed/NCBI

21 

Wang CZ, Calway T and Yuan CS: Herbal medicines as adjuvants for cancer therapeutics. Am J Chin Med. 40:657–669. 2012. View Article : Google Scholar : PubMed/NCBI

22 

Ben-Arye E, Schiff E, Hassan E, Mutafoglu K, Lev-Ari S, Steiner M, Lavie O, Polliack A, Silbermann M and Lev E: Integrative oncology in the Middle East: from traditional herbal knowledge to contemporary cancer care. Ann Oncol. 23:211–221. 2012. View Article : Google Scholar

23 

Li J, Liu P, Zhang R, Cao L, Qian H, Liao J, Xu W, Wu M and Yin Z: Icaritin induces cell death in activated hepatic stellate cells through mitochondrial activated apoptosis and ameliorates the development of liver fibrosis in rats. J Ethnopharmacol. 137:714–723. 2011. View Article : Google Scholar : PubMed/NCBI

24 

Zhou J, Wu J, Chen X, Fortenbery N, Eksioglu E, Kodumudi KN, Pk EB, Dong J, Djeu JY and Wei S: Icariin and its derivative, ICT, exert anti-inflammatory, anti-tumor effects, and modulate myeloid derived suppressive cells (MDSCs) functions. Int Immunopharmacol. 11:890–898. 2011. View Article : Google Scholar : PubMed/NCBI

25 

Huang J, Yuan L, Wang X, Zhang TL and Wang K: Icaritin and its glycosides enhance osteoblastic, but suppress osteoclastic, differentiation and activity in vitro. Life Sci. 81:832–840. 2007. View Article : Google Scholar : PubMed/NCBI

26 

He J, Wang Y, Duan F, Jiang H, Chen MF and Tang SY: Icaritin induces apoptosis of HepG2 cells via the JNK1 signaling pathway independent of the estrogen receptor. Planta Med. 76:1834–1839. 2010. View Article : Google Scholar : PubMed/NCBI

27 

Golbano JM, Lóppez-Aparicio P, Recio MN and Pérez-Albarsanz MA: Finasteride induces apoptosis via Bcl-2, Bcl-xL, Bax and caspase-3 proteins in LNCaP human prostate cancer cell line. Int J Oncol. 32:919–924. 2008.PubMed/NCBI

28 

Kroemer G and Reed JC: Mitochondrial control of cell death. Nat Med. 6:513–519. 2000. View Article : Google Scholar : PubMed/NCBI

29 

van Loo G, Saelens X, van Gurp M, MacFarlane M, Martin SJ and Vandenabeele P: The role of mitochondrial factors in apoptosis: a Russian roulette with more than one bullet. Cell Death Differ. 9:1031–1042. 2002. View Article : Google Scholar : PubMed/NCBI

30 

Llambi F, Moldoveanu T, Tait SW, Bouchier-Hayes L, Temirov J, McCormick LL, Dillon CP and Green DR: A unified model of mammalian BCL-2 protein family interactions at the mitochondria. Mol Cell. 44:517–531. 2011. View Article : Google Scholar : PubMed/NCBI

31 

Wang X: The expanding role of mitochondria in apoptosis. Genes Dev. 15:2922–2933. 2001.PubMed/NCBI

32 

Susnow N, Zeng L, Margineantu D and Hockenbery DM: Bcl-2 family proteins as regulators of oxidative stress. Semin Cancer Biol. 19:42–49. 2009. View Article : Google Scholar : PubMed/NCBI

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 

Fulda S and Debatin KM: Extrinsic versus intrinsic apoptosis pathways in anticancer chemotherapy. Oncogene. 25:4798–4811. 2006. View Article : Google Scholar : PubMed/NCBI

35 

Yamaguchi Y, Shiraki K, Fuke H, Inoue T, Miyashita K, Yamanaka Y and Nakano T: Adenovirus-mediated transfection of caspase-8 sensitizes hepatocellular carcinoma to TRAIL- and chemotherapeutic agent-induced cell death. Biochim Biophys Acta. 1763:844–853. 2006. View Article : Google Scholar : PubMed/NCBI

36 

Wang SH, Chen LM, Yang WK and Lee JD: Increased extrinsic apoptotic pathway activity in patients with hepatocellular carcinoma following transarterial embolization. World J Gastroenterol. 17:4675–4681. 2011. View Article : Google Scholar : PubMed/NCBI

37 

Muppidi JR and Siegel RM: Ligand-independent redistribution of Fas (CD95) into lipid rafts mediates clonotypic T cell death. Nat Immunol. 5:182–189. 2004. View Article : Google Scholar : PubMed/NCBI

38 

Abd El-Ghany RM, Sharaf NM, Kassem LA, Mahran LG and Heikal OA: Thymoquinone triggers anti-apoptotic signaling targeting death ligand and apoptotic regulators in a model of hepatic ischemia reperfusion injury. Drug Discov Ther. 3:296–306. 2009.PubMed/NCBI

39 

Hyer ML, Shi R, Krajewska M, Meyer C, Lebedeva IV, Fisher PB and Reed JC: Apoptotic activity and mechanism of 2-cyano-3,12-dioxoolean-1,9-dien-28-oic-acid and related synthetic triterpenoids in prostate cancer. Cancer Res. 68:2927–2933. 2008. View Article : Google Scholar : PubMed/NCBI

40 

Zhang L, Zhang Y, Zhang L, Yang X and Lv Z: Lupeol, a dietary triterpene, inhibited growth, and induced apoptosis through down-regulation of DR3 in SMMC7721 cells. Cancer Invest. 27:163–170. 2009. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

April-2015
Volume 11 Issue 4

Print ISSN: 1791-2997
Online ISSN:1791-3004

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Sun L, Peng Q, Qu L, Gong L and Si J: Anticancer agent icaritin induces apoptosis through caspase-dependent pathways in human hepatocellular carcinoma cells. Mol Med Rep 11: 3094-3100, 2015
APA
Sun, L., Peng, Q., Qu, L., Gong, L., & Si, J. (2015). Anticancer agent icaritin induces apoptosis through caspase-dependent pathways in human hepatocellular carcinoma cells. Molecular Medicine Reports, 11, 3094-3100. https://doi.org/10.3892/mmr.2014.3007
MLA
Sun, L., Peng, Q., Qu, L., Gong, L., Si, J."Anticancer agent icaritin induces apoptosis through caspase-dependent pathways in human hepatocellular carcinoma cells". Molecular Medicine Reports 11.4 (2015): 3094-3100.
Chicago
Sun, L., Peng, Q., Qu, L., Gong, L., Si, J."Anticancer agent icaritin induces apoptosis through caspase-dependent pathways in human hepatocellular carcinoma cells". Molecular Medicine Reports 11, no. 4 (2015): 3094-3100. https://doi.org/10.3892/mmr.2014.3007