Inhibition of NADPH oxidase 4 induces apoptosis in malignant mesothelioma: Role of reactive oxygen species

  • Authors:
    • Motoya Tanaka
    • Yuji Miura
    • Hiroki Numanami
    • Sivasundaram Karnan
    • Akinobu Ota
    • Hiroyuki Konishi
    • Yoshitaka Hosokawa
    • Masayuki Hanyuda
  • View Affiliations

  • Published online on: July 27, 2015     https://doi.org/10.3892/or.2015.4155
  • Pages: 1726-1732
Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Malignant pleural mesothelioma (MPM) is an aggressive tumor that is characterized by dysregulated growth and resistance to apoptosis. Reactive oxygen species (ROS)-generating NADPH oxidase (Nox) family enzymes have been suggested to be involved in neoplastic proliferation. Both the antioxidant N-acetylcysteine (NAC) and the inhibitor of flavoprotein-dependent oxidase, diphenylene iodonium (DPI), inhibited the cell viability of MPM cells in a dose-dependent manner. To examine whether Nox-mediated ROS generation confers antiapoptotic activity and thus a growth advantage to MPM cells, we analyzed the mRNA expression of Nox family members using quantitative RT-PCR in 7 MPM cell lines and a normal mesothelial cell line. Nox4 mRNA was expressed in all of the examined MPM cell lines, whereas little or no Nox2, Nox3 and Nox5 mRNA expression was detected. In 2 MPM cell lines, Nox4 mRNA expression was significantly higher than that in a normal mesothelial cell line. siRNAs targeting Nox4 suppressed ROS generation and cell viability in the MPM cell lines. In addition, DPI treatment and knockdown of Nox4 attenuated phosphorylation of AKT and ERK. Taken together, our results indicate that Nox4-mediated ROS, at least in part, transmit cell survival signals and their depletion leads to apoptosis, thus highlighting the Nox4-ROS-AKT signaling pathway as a potential therapeutic target for MPM treatment.

Introduction

Malignant pleural mesothelioma (MPM) are incurable thoracic malignancy, that has poor prognosis because it is frequently diagnosed at an advanced stage (1,2). The worldwide incidence of mesothelioma is expected to increase, particularly in Europe and Japan (35). The primary cause of MPM is often linked to asbestos exposure, and the number of patients worldwide is predicted to peak in the next 2 decades (6,7). Investigations for the molecular pathogenesis of MPM has begun (812). Recent whole-exome sequencing revealed frequent genetic alterations in BAP1, NF2, CDKN2A and CUL1 in 22 MPMs (13). The latent period between the first exposure to asbestos and the onset of this disease is ~30 years, and the first symptom is insidious and may include chest pain and breathlessness. Many clinical trials including surgery, radiotherapy, and chemotherapy were reported, but the prognosis of patients remains poor. Although there was recent progress in clinical treatment with combination chemotherapies, a curative therapy for MPM remains unknown; the median survival ranges between 9 and 17 months after diagnosis (1417). Combinations of cisplatin and pemetrexed appear to be the best chemotherapy regimen for MPM. Thus, effective clinical approaches such as molecular-targeted therapy are needed to treat MPM.

The recently-discovered epithelial NADPH oxidases (Noxs) mediate critical physiological and pathological processes including cell signaling, inflammation and mitogenesis by generating reactive oxygen species (ROS) (18). The Nox enzyme complex was first described in neutrophils, where it is normally quiescent but generates a large quantity of ROS upon activation during phagocytosis and plays a vital role in nonspecific host defense against ingested pathogens (19,20). Many non-phagocytic cells contain NADPH oxidases (20). There are 7 identified family members in the NADPH family: 5 Noxs and 2 dual oxidases (DUOXs) (20). Noxs and the mitochondria are major sources of cellular ROS (21). Cancer cells produce ROS that act as signaling molecules to promote cell survival (22,23). Nox4-mediated ROS inhibit apoptosis and promote tumor cell growth in pancreatic cancer cells (24,25). However, our understanding of the roles of the Nox family members in the development and growth of human cancer is limited (2630).

We hypothesized that intracellular ROS conferred anti-apoptotic activity and thus a growth advantage to MPM cells. In this study, we demonstrated that treatment with diphenylene iodonium (DPI), a flavoenzyme inhibitor (31) and knockdown of Nox4 suppressed ROS production in MPM cells, which induced apoptosis, suggesting that Nox4-generated ROS at least in part, transmits cell survival signals and provides a useful clinical approach for MPM treatment.

Materials and methods

Cell culture and materials

Seven MPM cell lines (ACC-MESO-1, ACC-MESO4, Y-MESO-8A, MSTO-211H, NCI-H28, NCI-H290 and NCI-H2052) and a normal mesothelial cell line (Met-5A) were kindly provided by Dr Y. Sekido, Division of Molecular oncology, Aichi Cancer Center Research Institute. Cells were maintained at 37°C under 5% CO2 air atmosphere in DMEM culture medium (Sigma, St. Louis, MO, USA) supplemented with 10% heat-inactivated FBS, 2 mM L-glutamine, 200 U/ml penicillin and 100 μg/ml streptomycin. Heparinized peripheral blood was collected from normal individuals after informed consent was obtained, and PBMCs were separated using density-gradient centrifugation.

Analysis and quantification of Nox4 mRNA levels by RT-PCR and real-time PCR

Reverse transcription (RT) was conducted as follows: 8 μl water containing 1 μg total RNA was added to 50 ng random primers (Life Technologies) and incubated at 65°C for 5 min. cDNA was prepared with SuperScript III First-Strand Synthesis Supermix (Invitrogen, Carsbad, CA, USA) according to the manufacturer's protocol.

Real-time PCR was performed using SYBR Premix Ex Taq II (Takara Bio, Otsu, Shiga, Japan), and PCR amplifications were performed in an ABI PRISM 7500 Sequence Detection System (Applied Biosystems, Foster City, CA, USA). Briefly, a solution of SYBR Premix Ex Taq II (10 μl; Takara Bio) containing sense and antisense primers (10 μM each) was prepared and 2 μl cDNA was added to a final volume of 20 μl. Conditions for PCR included 42°C for 5 min, 95°C for 10 sec, and 40 cycles of 95°C for 5 sec and 60°C for 34 sec. Data were analyzed with Sequencer Detector version 1.6 software (ABI-PE). The threshold cycle (CT) during the exponential phase of amplification was determined by real-time monitoring of fluorescent emission by nuclease activity of Taq polymerase. β-actin was used as an internal control. Relative transcripts were determined by the following formula: 1/2(CTtarget − CTcontrol) (32). Specific primers for Noxs 1-5 and β-actin were synthesized (Star Oligo Rikaken, Nagoya, Japan). PCR primer pairs were as follows: Nox1, sense 5′-AGCGTCTGCTCTCTGCTTGAA-3′ and antisense 5′-GGCTGCAAAATGAGCAGGT-3′; Nox2, sense 5′-TGCCTTTGAGTGGTTTGCAGAT-3′ and antisense 5′-ATTGGCCTGAGACTCATCCCA-3′; Nox3, sense 5′-GAACCCTCGGCTTGGAAAT-3′ and antisense 5′-TGGCTTACCACCTTGGTAATGA-3′; Nox4, sense 5′-CCCTCACAATGTGTCCAACTGA-3′ and antisense 5′-GGCAGAATTTCGGAGTCTTGAC-3′; Nox5, sense 5′-AAGAGTCAAAGGTCGTCCAAGG-3′ and antisense 5′-GCTTTCTTTTCTGGTGCCTGT-3′; β-actin, sense 5′-GATGACCCAGATCATGTTTGAGACC-3′ and antisense 5′-CGGTGAGGATCTTCATGAGGTAGT-3′.

Cell viability assay

The viability of the cells transfected with Nox4 siRNAs or treated with NAC, DPI, or specific inhibitors for protein kinases was determined using the MTT assay. MPM cells (1×103) were incubated with each reagent at each concentration in triplicate in 96-well culture plates at 37°C in humidified air with 5% CO2. Three wells contained MPM cells in drug-free medium to determine the control cell survival and the percentage of cells after culture. Three wells contained medium only to blank the spectrophotometer. After 2 days, 10 μl (5 mg/ml) MTT salt was added for 6 h. Formazan production was quantitated using a spectrophotometer at 562 nm. The optical density (OD) is linearly related to the cell number. Cell survival (CS) was calculated at each drug concentration by the equation CS = (OD treated well/mean OD control wells) × 100%.

Flow cytometric analysis of apoptosis

To analyze apoptosis, the externalization of phosphatidylserine was measured by flow cytometric staining with FITC-conjugated Annexin V (BD Pharmingen). Cells in 6-well plates (2×105 cells per well) were treated for 48 h, washed, resuspended in 100 μl Annexin-binding buffer, and stained with 5 μl Annexin V-FITC and propidium iodide for 20 min. Flow cytometric analysis was performed using FACSCalibur (BD Biosciences) and Cell Quest Pro Version 4.0.2 (BD Biosciences) software. Cells that were positively stained with Annexin V were counted as apoptotic populations.

Measurement of intracellular ROS production

Cells (2×105 per well) were seeded in 6-well plates and treated with 10 μM DPI for 48 h or transfected Nox4 siRNAs. Then, cells were incubated with 2.5 μM of 2′,7′-dichlorodihydrofluorescein diacetate (DCFH-DA; Molecular Probes, Eugene, OR, USA) for 30 min at 37°C in the dark, washed with Hank's buffer, and fixed in 1% paraformaldehyde. The fluorescence intensity was measured using FACS, with the excitation source at 488 nm and an emission wavelength of 580 nm. An analysis was performed with the software program BD FACStationt System Data Management System (Becton-Dickinson). Background fluorescence from the blank was subtracted from each reading.

Transfection and immunoblotting

Cells were transfected with Nox4 siRNAs or scramble siRNAs utilizing Lipofectamine 2000 (Invitrogen) according to the manufacturer's protocol. siRNAs were designed from the human Nox4 cDNA sequences as follows (Integrated DNA Technologies, Coralville, IA, USA): 5′-GCUGAAGUAUCAAACUAUUUAGAT-3′ and 5′-AUCUAAAUUAGUUUGAUACUUCAGCAG-3′ for Nox4RNAi-1, and 5′-GAAUUACAGUGAAGACUUUGUUGAA-3′ and 5′-UUCAACAAAGUCUUCACUGUAAUUCAC-3′ for Nox4RNAi-2. Universal scrambled siRNA sequences, which have no significant homology to mouse, rat, or human genome databases, were used as controls (Invitrogen).

For western blot analysis, equal amounts of reduced proteins (20 μg) were loaded on 10% Bis-Tris-buffered polyacrylamide gels. After gel electrophoresis, proteins were transferred to PVDF membranes (Invitrogen) by electroblotting. The membranes were preincubated for 1 h in 5% low-fat dried milk in TBS and 0.1% Tween-20 (TBS-T) to block non-specific binding sites. After washing with TBS-T, membranes were incubated overnight with a 1:1,000 dilution of primary antibody in TBS-T containing 5% BSA at 4°C (Cell Signaling Technology), and probed with horseradish peroxidase-conjugated secondary antibody (1:2,000 dilution) for 1 h at room temperature. The bound antibodies were visualized with the ECL reaction (GE Healthcare).

Statistical analysis

Data were analyzed by the Welch t-test, Fisher's exact-test, or ANOVA using Statview software (SAS, Cary, NC, USA), and P-values at <0.05 were considered to be statistically significant.

Results

Inhibition of cell growth, suppression of ROS generation, and induction of apoptosis by antioxidants

The flavoenzyme inhibitor, DPI inhibits membrane-bound, flavoprotein-containing Noxs. We examined whether the antioxidant NAC and the flavoenzyme inhibitor DPI, affected the cell viability of mesothelioma cell lines. Five MPM cell lines (ACC1-MESO, ACC4-MESO, Y8-MESO, MSTO-211H and H290) were treated with various concentrations of NAC or DPI for 48 h. Both NAC and DPI treatments inhibited the cell viability in a dose-dependent manner (Fig. 1). The IC50 values for DPI were as follows: 2.1 μM (ACC1-MESO), 0.8 μM (ACC4-MESO), 2.5 μM (Y8-MESO), 2.2 μM (MSTO-211H) and 2.2 μM (H290).

To verify that antioxidants affect ROS generation, we evaluated ROS production using flow cytometry. With vitamin E and DPI treatment, DCF fluorescence intensity, 2.1×104 in untreated cells was reduced to 0.7×104 and 0.1×104, respectively (Fig. 2). Thus, MPM cells regularly generated ROS, and both vitamin E and DPI treatment suppressed ROS generation.

We further examined the effect of DPI on the induction of apoptosis of MPM cells using Annexin V assay. DPI treatment significantly induced apoptosis (27 and 26%) in ACC1 and MSTO-221H cells, respectively (Fig. 3). Our results strongly suggest that depletion of ROS leads to the apoptosis of MPM cells.

Expression and quantification of Nox 1-5 mRNAs in MPM cell lines

The Nox family members produce ROS that are though to be pivotal for cell proliferative signaling. To examine the role of the Nox family in proliferation of MPM cells, we analyzed the mRNA expression of Nox family members in 7 MPM cell lines and a non-malignant mesothelial cell line (Met-5A) (Fig. 4). Nox4 mRNA was expressed in all of the examined MPM cell lines, whereas little or no Nox2, Nox3 and Nox5 mRNAs were detected (Fig. 4). In the ACC-MESO4 cell line, Nox1 mRNA expression was readily detected.

Subsequently, the expression levels of Noxs 1-5 relative to β-actin were measured by quantitative real-time RT-PCR (Fig. 5). Expression was arbitrarily graded as low (Nox copy number/β-actin copy number <500×10−8), intermediate (ratio >500 but <2,000×10−8) or high (ratio >2,000×10−8). Nox genes with expression ratios >500×10−8 were routinely visible by RT-PCR analysis using ≥40 μg total RNA. High- or intermediate-level Nox4 mRNA expression was observed in all examined MPM cell lines. Especially, 2 ACC-MESO4 and MSTO-211H cell lines expressed Nox4 mRNA at a high level. For comparison, we examined expression of the Nox family in several lymphoma/leukemia cell lines; mantle lymphoma cells and Jurkat leukemia cells expressed low levels of Nox2 mRNA (Fig. 5). High levels of Nox2 were detected in human PBMCs (Fig. 5) and the Jurkat cell line (data not shown).

Nox4 mediates ROS production in MPM cells

We utilized an RNA interference approach to verify whether Nox4 mediates ROS production in MPM cells. The expression of endogenous Nox4 mRNAs in ACC-MESO4 and MSTO-211H cells was significantly suppressed upon transfection of Nox4 siRNAs, (Fig. 6A). Intracellular ROS production was evaluated by flow cytometry. The transfection of Nox4 siRNAs significantly suppressed ROS levels compared to controls (Fig. 6B), indicating that Nox4, at least in part, is responsible for intracellular ROS generation. However, the ROS generation was not completely inhibited by transfection of Nox4 siRNAs, suggesting the possibility that other Nox members and mitochondria sources may also contribute to ROS generation.

Suppression of ROS generation by Nox4 siRNAs induces apoptosis

To explore whether Nox4-generated ROS regulate cell survival, we examined the effect of Nox4 siRNAs on cell viability and apoptosis. The knockdown of Nox4 significantly reduced the cell viability of ACC-MESO4 and MSTO-211H cells by 30% (Fig. 6C). To verify whether the inhibitory effect of Nox4 siRNAs is associated with apoptosis, an Annexin V assay was performed. The transfection of Nox4 siRNAs induced apoptosis by 13% in MPM cells (Fig. 6D). Thus, Nox4 siRNAs suppressed ROS production, and the depletion of ROS by Nox4 siRNAs and DPI treatment induced apoptosis in MPM cells.

The role of protein kinases in cell survival signaling in MPM cells

Both PI3K/AKT and MEK/ERK1/2 signaling cascades have important roles in cell proliferation, but they also mediate apoptosis (33). Western blot analysis showed that ACC1, ACC4, and MSTO-211H cell lines expressed phospho-AKT and phospho-ERK (Fig. 7). Nox4 siRNAs transfection and DPI treatment attenuated the phosphorylation of AKT and ERK. Given that transfection of Nox4 siRNAs and DPI treatment induced apoptosis in MPM cells, the inactivation of PI3K/AKT and MEK/ERK1/2 signaling cascades likely plays an important role in the induction of apoptosis.

Discussion

It is well established that the development of MPM is associated with asbestos exposure (34,35). Chronic inflammation accelerates the development and progression of malignant mesothelioma, possibly because of cytokine release and ROS generation. The inflammation that infiltrate into tissue areas containing asbestos deposits consists largely of phagocytic macrophages that internalize asbestos and release numerous cytokines and mutagenic ROS.

In this study, we first examined the role of ROS in cell proliferation. Although ROS are thought to cause stress-induced apoptosis, ROS often provide cancer cells with a survival advantage. In fact, we showed that suppression of ROS levels by NAC and DPI treatment reduced the viability of MPM cells. Similar results were observed in other cancer cells including pancreatic cells. Second, we examined whether the Nox4-mediated generation of intracellular ROS conferred anti-apoptotic activity and thus a growth advantage to MPM cells. To this end, we analyzed the expression levels of Nox genes in 7 MPM cell lines, and a normal mesothelial cell line. RT-PCR analysis revealed that Nox4 mRNA was expressed in all MPM cell lines, whereas little or no Nox2, Nox3 and Nox5 mRNAs were detected. Quantitative real-time RT-PCR also revealed that a high or intermediate level of Nox4 mRNA expression was observed in all MPM cell lines; high-level expression was detected in ACC-MESO4 and MSTO-211H cells compared to the normal mesothelial cell line Met-5A (P<0.01).

The siRNAs targeting Nox4 in 2 MPM cell lines reduced intracellular ROS generation by 50%, and cell viability by 30%. The depletion of ROS by DPI treatment or knockdown of Nox4 induced apoptosis. Collectively, our findings suggest that ROS generated by Nox4, at least in part, transmit cell survival signals and their depletion leads to apoptosis. High-level Nox1 mRNA expression was observed in colorectal cancer cell lines. Growth inhibition profiling of DPI revealed a modest positive correlation with Nox1 levels (31). Exposure of HT-29 colon cancer cells, which expresses Nox1, to DPI had inhibitory effects on the steady-state ROS levels, and decreased STAT, ERK1/2, and AKT signaling activity (31). Nox4 overexpression is also reported in primary breast, ovarian, prostate, melanoma, and glioblastoma cancer cell lines (3638). Nox4 expression was intermediate to high in 2 of the 4 tested ovarian cancer cell lines. Notably, in A2780/DDP cells, high level acquired resistance to cisplatin was associated with a marked decrease in the expression level of Nox4. Recently, Nox4 was shown to be an oncoprotein localized to mitochondria (39). Together with these studies, our study suggests that Nox4 may act as an oncogene, and Nox4-related signaling molecules may be good candidates for molecular-targeted therapy for MPM. Given a possible role of Nox4 in tumorigenesis, it will be of particular interest to investigate the correlation of Nox4 expression in primary mesotheliomas and prognosis in the patients.

AKT (protein kinase B) is a regulator of cell survival in response to a growth factor. AKT is activated through its phosphorylation, and it inhibits apoptosis-inducing proteins, thereby promoting cell survival. The ERK pathway is mostly activated by growth factors and mediates cell proliferation, but it also mediates apoptosis by stress stimuli (33). To examine the role of AKT and ERK in the apoptosis of MPM cells, we evaluated the phosphorylation state of AKT and ERK. Both DPI treatment and knockdown of Nox4 attenuated their phosphorylation levels, suggesting that AKT and ERK play a role in cell survival signals in MPM cells. Consistent with our results, the PI3K-AKT pathway was reported to be activated in human malignant mesothelioma (40). In addition, both selective inhibitors for PI3K/AKT and MEK/ERK1/2 were effective in downregulating the expression of prometastasis phenotypes of MPM cells (41).

In conclusion, we demonstrated that Nox4-mediated ROS generation, at least in part, transmits cell survival signals, and ROS depletion by the knockdown of Nox4 and DPI treatment leads to apoptosis in a subset of MPM cells. Our study raises the possibility of the Nox4-ROS-AKT signaling pathway as a novel therapeutic target for MPM. Antioxidant treatment targeted to this signaling pathway has the potential to enhance the therapeutic index of cisplatin-based therapy. Further studies are warranted to contribute to the knowledge required to ultimately develop targeted therapies for MPM.

Abbreviations:

DCF

dichlorofluorescin

Duox

dual oxidase

DCFH-DA

2′,7′-dichlorodihydrofluorescein diacetate

DPI

diphenylene iodonium

MPM

malignant pleural mesothelioma

NAC

N-acetylcysteine

Nox

NADPH oxidase

PBMC

peripheral blood mononuclear cell

ROS

reactive oxygen species

Acknowledgments

This study was supported by grant AI 52213 from the Aichi Cancer Center to Y.M. and a grant of Strategic Research Foundation Grant-Aided Project for Private Universities from the Ministry of Education, Culture, Sports, Science and Technology, Japan (MEXT) to Y.H. The authors thank Dr Yoshitaka Sekido (Division of Molecular Oncology Aichi Cancer Center Research Institute) for providing MPM cell lines.

References

1 

Robinson BW and Lake RA: Advances in malignant mesothelioma. N Engl J Med. 353:1591–1603. 2005. View Article : Google Scholar : PubMed/NCBI

2 

Campbell NP and Kindler HL: Update on malignant pleural mesothelioma. Semin Respir Crit Care Med. 32:102–110. 2011. View Article : Google Scholar : PubMed/NCBI

3 

Peto J, Decarli A, La Vecchia C, Levi F and Negri E: The European mesothelioma epidemic. Br J Cancer. 79:666–672. 1999. View Article : Google Scholar : PubMed/NCBI

4 

Murayama T: Epidemic of asbestos related diseases. Proceedings of the global Asbestos Congress (Tokyo); pp. 172004

5 

Stewart DJ, Martin-Ucar A, Pilling JE, Edwards JG, O'Byrne KJ and Waller DA: The effect of extent of local resection on patterns of disease progression in malignant pleural mesothelioma. Ann Thorac Surg. 78:245–252. 2004. View Article : Google Scholar : PubMed/NCBI

6 

Takahashi K: Emerging health effects of asbestos in Asia. Proceedings of the Global Asbestos Congress (Tokyo); pp. 22004

7 

Murayama T, Takahashi K, Natori Y and Kurumatani N: Estimation of future mortality from pleural malignant mesothelioma in Japan based on an age-cohort model. Am J Ind Med. 49:1–7. 2006. View Article : Google Scholar

8 

Rascoe PA, Jupiter D, Cao X, Littlejohn JE and Smythe WR: Molecular pathogenesis of malignant mesothelioma. Expert Rev Mol Med. 14:e122012. View Article : Google Scholar : PubMed/NCBI

9 

Murakami H, Mizuno T, Taniguchi T, Fujii M, Ishiguro F, Fukui T, Akatsuka S, Horio Y, Hida T, Kondo Y, et al: LATS2 is a tumor suppressor gene of malignant mesothelioma. Cancer Res. 71:873–883. 2011. View Article : Google Scholar : PubMed/NCBI

10 

Fujii M, Toyoda T, Nakanishi H, Yatabe Y, Sato A, Matsudaira Y, Ito H, Murakami H, Kondo Y, Kondo E, et al: TGF-β synergizes with defects in the Hippo pathway to stimulate human malignant mesothelioma growth. J Exp Med. 209:479–494. 2012. View Article : Google Scholar : PubMed/NCBI

11 

Shi Y, Moura U, Opitz I, Soltermann A, Rehrauer H, Thies S, Weder W, Stahel RA and Felley-Bosco E: Role of hedgehog signaling in malignant pleural mesothelioma. Clin Cancer Res. 18:4646–4656. 2012. View Article : Google Scholar : PubMed/NCBI

12 

Testa JR, Cheung M, Pei J, Below JE, Tan Y, Sementino E, Cox NJ, Dogan AU, Pass HI, Trusa S, et al: Germline BAP1 mutations predispose to malignant mesothelioma. Nat Genet. 43:1022–1025. 2011. View Article : Google Scholar : PubMed/NCBI

13 

Guo G, Chmielecki J, Goparaju C, Heguy A, Dolgalev I, Carbone M, Seepo S, Meyerson M and Pass HI: Whole-exome sequencing reveals frequent genetic alterations in BAP1, NF2, CDKN2A, and CUL1 in malignant pleural mesothelioma. Cancer Res. 75:264–269. 2015. View Article : Google Scholar

14 

Vogelzang NJ, Rusthoven JJ, Symanowski J, Denham C, Kaukel E, Ruffie P, Gatzemeier U, Boyer M, Emri S, Manegold C, et al: Phase III study of pemetrexed in combination with cisplatin versus cisplatin alone in patients with malignant pleural mesothelioma. J Clin Oncol. 21:2636–2644. 2003. View Article : Google Scholar : PubMed/NCBI

15 

Tsao AS, Wistuba I, Roth JA and Kindler HL: Malignant pleural mesothelioma. J Clin Oncol. 27:2081–2090. 2009. View Article : Google Scholar : PubMed/NCBI

16 

Stahel RA and Weder W: Improving the outcome in malignant pleural mesothelioma: Nonaggressive or aggressive approach? Curr Opin Oncol. 21:124–130. 2009. View Article : Google Scholar : PubMed/NCBI

17 

Kindler HL, Karrison TG, Gandara DR, Lu C, Krug LM, Stevenson JP, Jänne PA, Quinn DI, Koczywas MN, Brahmer JR, et al: Multicenter, double-blind, placebo-controlled, randomized phase II trial of gemcitabine/cisplatin plus bevacizumab or placebo in patients with malignant mesothelioma. J Clin Oncol. 30:2509–2515. 2012. View Article : Google Scholar : PubMed/NCBI

18 

Brar SS, Kennedy TP, Sturrock AB, Huecksteadt TP, Quinn MT, Whorton AR and Hoidal JR: An NAD(P)H oxidase regulates growth and transcription in melanoma cells. Am J Physiol Cell Physiol. 282:C1212–C1224. 2002. View Article : Google Scholar : PubMed/NCBI

19 

Lambeth JD: NOX enzymes and the biology of reactive oxygen. Nat Rev Immunol. 4:181–189. 2004. View Article : Google Scholar : PubMed/NCBI

20 

Dworakowski R, Anilkumar N, Zhang M and Shah AM: Redox signalling involving NADPH oxidase-derived reactive oxygen species. Biochem Soc Trans. 34:960–964. 2006. View Article : Google Scholar : PubMed/NCBI

21 

Chen K, Craige SE and Keaney JF Jr: Downstream targets and intracellular compartmentalization in Nox signaling. Antioxid Redox Signal. 11:2467–2480. 2009. View Article : Google Scholar : PubMed/NCBI

22 

Storz P: Reactive oxygen species in tumor progression. Front Biosci. 10:1881–1896. 2005. View Article : Google Scholar : PubMed/NCBI

23 

Szatrowski TP and Nathan CF: Production of large amounts of hydrogen peroxide by human tumor cells. Cancer Res. 51:794–798. 1991.PubMed/NCBI

24 

Mochizuki T, Furuta S, Mitsushita J, Shang WH, Ito M, Yokoo Y, Yamaura M, Ishizone S, Nakayama J, Konagai A, et al: Inhibition of NADPH oxidase 4 activates apoptosis via the AKT/apoptosis signal-regulating kinase 1 pathway in pancreatic cancer PANC-1 cells. Oncogene. 25:3699–3707. 2006. View Article : Google Scholar : PubMed/NCBI

25 

Vaquero EC, Edderkaoui M, Pandol SJ, Gukovsky I and Gukovskaya AS: Reactive oxygen species produced by NAD(P)H oxidase inhibit apoptosis in pancreatic cancer cells. J Biol Chem. 279:34643–34654. 2004. View Article : Google Scholar : PubMed/NCBI

26 

Cheng G, Cao Z, Xu X, van Meir EG and Lambeth JD: Homologs of gp91phox: Cloning and tissue expression of Nox3, Nox4, and Nox5. Gene. 269:131–140. 2001. View Article : Google Scholar : PubMed/NCBI

27 

Donkó A, Péterfi Z, Sum A, Leto T and Geiszt M: Dual oxidases. Philos Trans R Soc Lond B Biol Sci. 360:2301–2308. 2005. View Article : Google Scholar : PubMed/NCBI

28 

Geiszt M, Witta J, Baffi J, Lekstrom K and Leto TL: Dual oxidases represent novel hydrogen peroxide sources supporting mucosal surface host defense. FASEB J. 17:1502–1504. 2003.PubMed/NCBI

29 

Juhasz A, Ge Y, Markel S, Chiu A, Matsumoto L, van Balgooy J, Roy K and Doroshow JH: Expression of NADPH oxidase homologues and accessory genes in human cancer cell lines, tumours and adjacent normal tissues. Free Radic Res. 43:523–532. 2009. View Article : Google Scholar : PubMed/NCBI

30 

Wu Y, Antony S, Juhasz A, Lu J, Ge Y, Jiang G, Roy K and Doroshow JH: Up-regulation and sustained activation of Stat1 are essential for interferon-gamma (IFN-gamma)-induced dual oxidase 2 (Duox2) and dual oxidase A2 (DuoxA2) expression in human pancreatic cancer cell lines. J Biol Chem. 286:12245–12256. 2011. View Article : Google Scholar : PubMed/NCBI

31 

Doroshow JH, Juhasz A, Ge Y, Holbeck S, Lu J, Antony S, Wu Y, Jiang G and Roy K: Antiproliferative mechanisms of action of the flavin dehydrogenase inhibitors diphenylene iodonium and di-2-thienyliodonium based on molecular profiling of the NCI-60 human tumor cell panel. Biochem Pharmacol. 83:1195–1207. 2012. View Article : Google Scholar : PubMed/NCBI

32 

Miura Y, Thoburn CJ, Bright EC, Phelps ML, Shin T, Matsui EC, Matsui WH, Arai S, Fuchs EJ, Vogelsang GB, et al: Association of Foxp3 regulatory gene expression with graft-versus-host disease. Blood. 104:2187–2193. 2004. View Article : Google Scholar : PubMed/NCBI

33 

Lee YJ, Cho HN, Soh JW, Jhon GJ, Cho CK, Chung HY, Bae S, Lee SJ and Lee YS: Oxidative stress-induced apoptosis is mediated by ERK1/2 phosphorylation. Exp Cell Res. 291:251–266. 2003. View Article : Google Scholar : PubMed/NCBI

34 

Carbone M and Yang H: Molecular pathways: Targeting mechanisms of asbestos and erionite carcinogenesis in mesothelioma. Clin Cancer Res. 18:598–604. 2012. View Article : Google Scholar :

35 

Carbone M, Ly BH, Dodson RF, Pagano I, Morris PT, Dogan UA, Gazdar AF, Pass HI and Yang H: Malignant mesothelioma: Facts, myths, and hypotheses. J Cell Physiol. 227:44–58. 2012. View Article : Google Scholar

36 

Ushio-Fukai M and Nakamura Y: Reactive oxygen species and angiogenesis: NADPH oxidase as target for cancer therapy. Cancer Lett. 266:37–52. 2008. View Article : Google Scholar : PubMed/NCBI

37 

Shono T, Yokoyama N, Uesaka T, Kuroda J, Takeya R, Yamasaki T, Amano T, Mizoguchi M, Suzuki SO, Niiro H, et al: Enhanced expression of NADPH oxidase Nox4 in human gliomas and its roles in cell proliferation and survival. Int J Cancer. 123:787–792. 2008. View Article : Google Scholar : PubMed/NCBI

38 

Yamaura M, Mitsushita J, Furuta S, Kiniwa Y, Ashida A, Goto Y, Shang WH, Kubodera M, Kato M, Takata M, et al: NADPH oxidase 4 contributes to transformation phenotype of melanoma cells by regulating G2-M cell cycle progression. Cancer Res. 69:2647–2654. 2009. View Article : Google Scholar : PubMed/NCBI

39 

Graham KA, Kulawiec M, Owens KM, Li X, Desouki MM, Chandra D and Singh KK: NADPH oxidase 4 is an oncoprotein localized to mitochondria. Cancer Biol Ther. 10:223–231. 2010. View Article : Google Scholar : PubMed/NCBI

40 

Suzuki Y, Murakami H, Kawaguchi K, Tanigushi T, Fujii M, Shinjo K, Kondo Y, Osada H, Shimokata K, Horio Y, et al: Activation of the PI3K-AKT pathway in human malignant mesothelioma cells. Mol Med Rep. 2:181–188. 2009.PubMed/NCBI

41 

Cole GW Jr, Alleva AM, Zuo JT, Sehgal SS, Yeow WS, Schrump DS and Nguyen DM: Suppression of pro-metastasis phenotypes expression in malignant pleural mesothelioma by the PI3K inhibitor LY294002 or the MEK inhibitor UO126. Anticancer Res. 26A:809–821. 2006.

Related Articles

Journal Cover

October-2015
Volume 34 Issue 4

Print ISSN: 1021-335X
Online ISSN:1791-2431

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Tanaka M, Miura Y, Numanami H, Karnan S, Ota A, Konishi H, Hosokawa Y and Hanyuda M: Inhibition of NADPH oxidase 4 induces apoptosis in malignant mesothelioma: Role of reactive oxygen species. Oncol Rep 34: 1726-1732, 2015
APA
Tanaka, M., Miura, Y., Numanami, H., Karnan, S., Ota, A., Konishi, H. ... Hanyuda, M. (2015). Inhibition of NADPH oxidase 4 induces apoptosis in malignant mesothelioma: Role of reactive oxygen species. Oncology Reports, 34, 1726-1732. https://doi.org/10.3892/or.2015.4155
MLA
Tanaka, M., Miura, Y., Numanami, H., Karnan, S., Ota, A., Konishi, H., Hosokawa, Y., Hanyuda, M."Inhibition of NADPH oxidase 4 induces apoptosis in malignant mesothelioma: Role of reactive oxygen species". Oncology Reports 34.4 (2015): 1726-1732.
Chicago
Tanaka, M., Miura, Y., Numanami, H., Karnan, S., Ota, A., Konishi, H., Hosokawa, Y., Hanyuda, M."Inhibition of NADPH oxidase 4 induces apoptosis in malignant mesothelioma: Role of reactive oxygen species". Oncology Reports 34, no. 4 (2015): 1726-1732. https://doi.org/10.3892/or.2015.4155