Inhibition of NADPH oxidase 2 induces apoptosis in osteosarcoma: The role of reactive oxygen species in cell proliferation

  • Authors:
    • Kazumasa Kitamoto
    • Yuji Miura
    • Sivasundaram Karnan
    • Akinobu Ota
    • Hiroyuki Konishi
    • Yoshitaka Hosokawa
    • Keiji Sato
  • View Affiliations

  • Published online on: March 19, 2018     https://doi.org/10.3892/ol.2018.8291
  • Pages: 7955-7962
Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Osteosarcomas (OS) are aggressive tumors that are characterized by dysregulated growth and resistance to apoptosis. Reactive oxygen species (ROS) are thought to be important signal transduction molecules in the regulation of cell growth. ROS‑generating nicotinamide adenine dinucleotide phosphate oxidase (NOX) family enzymes have previously been suggested to be involved in neoplastic proliferation. To examine whether NOX‑mediated generation of intracellular ROS confers anti‑apoptotic activity, and thus a growth advantage, the current study first analyzed the mRNA expression of NOX family members by reverse transcription‑quantitative polymerase chain reaction (RT‑qPCR) in five human OS cell lines. RT‑PCR analysis revealed that NOX2 and NOX4 mRNAs were expressed in all the OS cell lines examined, whereas little or no NOX1 and NOX3 mRNAs were detected. By RT‑qPCR, NOX2 mRNA expression levels were demonstrated to be higher than NOX4 mRNA expression levels. The viability of OS cells decreased in a dose‑dependent manner with treatment of diphenylene iodonium (DPI), an inhibitor of flavoprotein‑dependent oxidase. DPI treatment was observed to reduce intracellular ROS levels by ~50%, and increase the frequency of apoptosis by 30%. Notably, small interfering RNAs (siRNAs) targeting NOX2 significantly suppressed ROS generation; ROS depletion by DPI or NOX2 siRNAs induced apoptosis in OS cells. Together, the results of the present study indicate that NOX2‑mediated ROS generation promotes cell survival and ROS depletion leads to apoptosis, thus highlighting the NOX2‑ROS signaling pathway as a potential therapeutic target for OS treatment.

Introduction

Osteosarcoma (OS) is the most common primary malignant bone tumor in children and young adults, affecting three to five people per million annually (1,2). OS typically presents in the bones around the knee (60% of cases), accounting for ~5% of the pediatric malignancies (1,2). The second most common site for OS to present is the pelvis; other initial presentation sites have also been reported, including the end of the humerus, skull and clavicle (35). OS usually occurs in teenagers (60%): The majority of OS cases are diagnosed before 20 years of age, a total of 75% (6,7).

The treatment outcome of OS has improved markedly over the course of several years. The current standard treatment for OS is a combination of surgery and chemotherapy: Preoperative (neoadjuvant) chemotherapy, limb sparing surgery and postoperative (adjuvant) chemotherapy. In the late 1970s, the 5-year survival rate was 10–20% due to apparent lung metastases post-surgery; however, 5-year survival has improved to the current rate of 50–80% (812). In addition to improvements in surgical and diagnostic techniques, the introduction of intensive chemotherapy has reduced the rate of lung metastases. The majority of patients today receive the same drugs as 25 years ago, including doxorubicin, cisplatin, high-dose methotrexate and ifosfamide in varying combinations (13,14). In order to further improve the survival rate, the development of novel anticancer agents is also necessary. Genetic studies have been performed to identify novel therapeutic targets for OS; however, the underlying molecular mechanisms have not yet been completely established (1517). Candidate gene studies and a recent genome-wide association study have identified several common single nucleotide polymorphisms associated with OS (18,19).

The recently-discovered epithelial nicotinamide adenine dinucleotide phosphate (NADPH) oxidase (NOX) family of enzymes mediate critical physiological and pathological processes, including cell signaling, inflammation and mitogenesis through the generation of reactive oxygen species (ROS) (20). The role of the ROS produced by these enzymes in specific tissue types and distinct cellular compartments has been the subject of several recent investigations (21,22). Cancer cells, like non-malignant tissues, produce ROS; in tumors, reactive oxygen metabolites may act as signaling molecules to promote cell survival over apoptosis (23,24). NOX enzymes and the mitochondria are a major source of cellular ROS (25). There are currently seven identified enzymes in the NADPH family, including five NOX enxymes (NOX1-5) (26,27). NOX enzymes have a fundamental role in numerous cell functions, including signal transduction, differentiation, proliferation and cell death (25,26). However, current understanding of the roles of the NOX family members in the development and growth of human cancers remains limited (2731).

In the present study, it was hypothesized that NOX-mediated ROS generation conferred anti-apoptotic activity, and, thus, a growth advantage to OS cells. It was demonstrated that treatment with a flavoenzyme inhibitor, diphenylene iodonium (DPI), and the knockdown of NOX2 significantly suppressed ROS generation in OS cells, which induced apoptosis, indicating that NOX2-mediated ROS may transmit cell survival signals and provide a potential clinical approach for OS treatment.

Materials and methods

Cell culture and materials

Five OS cell lines (HOS, MOS, MG-63, NOS-1 and HuO 9N2) were used in this study. The MOS and NOS-1 cell lines were kindly provided by Dr Masahiko Kanamori (School of Medicine, University of Toyama, Toyama, Japan) (3236). Three cell lines (HOS, MG-63, and HuO 9N2) were obtained from the Japanese Collection of Research Bioresources cell bank (Ibaraki, Osaka, Japan). Cells were maintained at 37°C containing 5% CO2 atmospheric air in Dulbecco's modified Eagle's medium (DMEM; Sigma-Aldrich, St. Louis, MO, USA) supplemented with 10% heat-inactivated fetal bovine serum (FBS; Sigma-Aldrich), 2 mM L-glutamine, 200 U/ml penicillin and 100 µg/ml streptomycin. DPI was purchased from Calbiochem (EMD Millipore, Billerica, MA, USA).

Ethical approval for the present study was obtained from the ethical committee of Aichi Medical University (approval no. 11–039), and informed consent was obtained prior to the start of the study. Heparinized peripheral blood was collected from healthy individuals (n=3). Peripheral blood mononuclear cells (PBMCs) were separated by Ficoll-Hypaque density centrifugation at 400 × g for 30 min at room temperature.

Based on human NOX2 and NOX4 cDNA sequences, small interfering RNAs (siRNAs) were designed as follows (Integrated DNA Technologies, Coralville, IA, USA): 5′-UCAGGGUUCUUUAUUCUCUTT-3′ and 5′-AGAGAAUAAAGAACCCUGATT-3′ for NOX2 siRNA-1; 5′-GUACAAUUCGUUCAGCUCCTT-3′ and 5′-GGAGCUGAACGAAUUGUACTT-3′ for NOX2 siRNA-2; 5′-GCUGAAGUAUCAAACUAAUUUAGAT-3′ and 5′-UCUAAAUUAGUUUGAUACUUCAGCAG-3′ for NOX4 siRNA-1; 5′-GAAUUACAGUGAAGACUUUGUUGAA-3′ and5′-UUCAACAAAGUCUUCACUGUAAUUCAC-3′ for NOX4 siRNA-2. Universal scrambled siRNA sequences (Invitrogen; Thermo-Fisher Scientific, Inc., Waltham, MA, USA), which have no significant homology to mouse, rat or human genome databases, were used as controls.

Quantification of NOX2 and NOX4 mRNAs by reverse transcription-quantitative polymerase chain reaction (RT-qPCR)

The total RNA was purified with a TRIzol reagent (Thermo-Fisher Scientific, Inc.). The HOS, MOS, MG-63, NOS-1 and HuO 9N2 OS cell lines (1×105 density) were seeded into 24-well plates and cultured in the presence or absence of NOX2 and/or siRNAs for 48 h. Cell lysates were prepared in 1 ml TRIzol reagent with adequate mixing. Chloroform (200 µl) was added and the solution was mixed well and centrifuged for 15 min at 12,000 × g at 4°C. The chloroform and centrifugation steps were repeated. Then, RNA pellet was precipitated with 2-propanol and rinsed with 70% ethanol. Purified RNA was dissolved in 20 µl distilled water. RT was conducted as follows: A total of 8 µl water, containing 1 µg total RNA, was added to 50 ng random primers (Thermo-Fisher Scientific, Inc.) and incubated at 65°C for 5 min. The samples were chilled on ice and cDNA was prepared with SuperScript III First-Strand Synthesis Supermix (Invitrogen; Thermo-Fisher Scientific, Inc.) according to the manufacturer's instructions. The PCR products created with the NOXs 2 and 4 primers were identified by direct DNA sequence analysis.

RT-qPCR was performed with SYBR Premix Ex Taq II (Takara Bio, Inc., Otsu, Shiga, Japan) in an ABI PRISM 7500 Sequence Detection system (Applied Biosystems; Thermo Fisher Scientific, Inc.). Briefly, a solution of SYBR Premix Ex Taq II (10 µl) containing sense and antisense primers (10 µM each) was prepared and aliquoted into individual wells of a MicroAmp Optical Plate (ABI-PE; Applied Biosystems; Thermo-Fisher Scientific, Inc.): 2 µl cDNA was added to give a final volume of 20 µl. The cycling conditions for the PCR were 42°C for 5 min, 95°C for 10 sec and 40 cycles of 95°C for 5 sec (denaturation) and 60°C for 34 sec (annealing/extension). The data were analyzed with Sequence Detector software (version 1.6; ABI-PE; Applied Biosystems; Thermo-Fisher Scientific, Inc.). The quantitative cycle (Cq) during the exponential phase of amplification was determined by real-time monitoring of fluorescent emission by the nuclease activity of Taq polymerase. β-actin was used as an internal control gene for mRNA expression. Relative transcripts were determined by the formula: 1/2(Cq target - Cq control) (36). NOX1, NOX2, NOX3, NOX4, NOX5 and β-actin genes were amplified with specific primer sequences (Star Oligo, Rikaken, Nagoya, Japan) according to the NCBI reference sequences (http://www.ensembl.org/Homo_sapiens/index.html). The PCR primer pairs and probes used, were as follows: NOX1, 5′-AGCGTCTGCTCTCTGCTTGAA-3′ and 5′-GGCTGCAAAATGAGCAGGT-3′ (junction between exons 3 and 4); NOX2, 5′-TGCCTTTGAGTGGTTTGCAGAT-3′ and 5′-ATTGGCCTGAGACTCATCCCA-3′ (junction between exons 11 and 12); NOX3, 5′-GAACCCTCGGCTTGGAAAT-3′ (junction between exons 7 and 8) and 5′-TGGCTTACCACCTTGGTAATGA-3′ (junction between exons 8 and 9); NOX4, 5′-CCCTCACAATGTGTCCAACTGA-3′ (junction between exons 11 and 12) and 5′-GGCAGAATTTCGGAGTCTTGAC-3′; NOX5, 5′-AAGAGTCAAAGGTCGTCCAAGG-3′ and 5′-GCTTTCTTTTCTGGTGCCTGT-3′ (junction between exons 13 and 14); β-actin, 5′-GATGACCCAGATCATGTTTGAGACC-3′ (junction between exons 2 and 3) and 5′-CGGTGAGGATCTTCATGAGGTAGT-3′. Semi-quantitative RT-PCR was performed with 30 cycles (94°C for 1 min, 55°C for 1 min, 72°C for 1 min).

Cell transfection

The cells were transfected with NOX2- and/or NOX4-specific siRNAs, or scramble RNAs, using Lipofectamine® 2000 (Invitrogen; Thermo-Fisher Scientific, Inc.) according to the manufacturer's instructions.

Assessment of intracellular ROS production

HOS, MOS, MG-63, NOS-1 and HuO 9N2 OS cells (2×105) were seeded in 6-well plates and treated with 10 µM DPI for 48 h. Following this, the cells were transfected with NOX2- and NOX4-specific siRNAs and cultured for 48 h. The cells were then incubated with 2.5 µM dihydroethidium (DCFH-DA; Molecular Probes, Thermo-Fisher Scientific, Inc.) for 30 min at 37°C in the dark. Subsequently, the cells were washed with Hank's buffer and fixed in 1% paraformaldehyde. Fluorescence-activated cell sorting (FACS) was used to measure the fluorescence emission intensities at 488 nm for excitation and at 580 nm for detection. The histograms were analyzed with the BD FACStation System Data Management system (BD Biosciences; Franklin Lakes, NJ, USA). Background fluorescence from a blank sample was subtracted from each reading to normalize the results.

In vitro 3-(4, 5-dimethyl thiazol-2-yl)-2, 5-diphenyl tetrazolium bromide (MTT) assay

An MTT assay was used to evaluate cell viability after 48 h incubation at 37°C. OS cells were incubated (in triplicate) in 96-well culture plates at 37°C in humidified air with 5% CO2. Three wells that contained OS cells in drug-free DMEM were included to determine the control cell survival rate; another three wells contained only DMEM to calibrate the spectrophotometer. After two days, 10 µl (5 mg/ml) MTT salt (Sigma-Aldrich: St Louis, MO, USA) was added to each well in 96-well culture plates for 6 h. The MTT compound was reduced to colored formazan crystals by the living cells alone. The crystals were dissolved with 100 µl of acidified isopropanol and the formazan crystal production was quantified using a spectrophotometer (562 nm). The optical density (OD) is linearly related to the cell number. Cell survival (CS) was calculated for each drug concentration using the following equation: CS=(ODtreated well/ODcontrol well mean) ×100%.

Measuring apoptosis using flow cytometry

The externalization of phosphatidylserine was measured by flow cytometry with fluorescein isothiocyanate-conjugated Annexin V (BD Pharmingen, San Diego, CA, USA) (37). Flow cytometry analyses were performed using a FACSCalibur flow cytometer (BD Biosciences) and CellQuest Pro Version 4.0.2 (BD Biosciences) software. Cells (2×105) seeded in 6-well plates were cultured for 48 h following transfection of NOX siRNAs or scramble RNAs, washed and resuspended in 100 µl Annexin-binding buffer [10 mM HEPES (N-2-hydroxyethylpiperazine-N'-2-ethanesulfonic acid), 140 mM NaCl and 2.5 mM CaCl2 in PBS], stained with 5 µl Annexin V-Alexa Fluor 488 conjugate for 20 min then analyzed by flow cytometry (FACSCalibur). The cells that stained positive with Annexin V were counted as apoptotic populations.

Statistical analysis

The results of MTT assay, apoptosis assay with siRNA transfection, and RT-qPCR assay were analyzed using one-way analysis of variance with a Turkey Kramer post hoc test using the Statview software (version 5; SAS Institute, Inc., Cary, NC, USA). Results are expressed as the mean ± standard deviation. P<0.05 was considered to indicate a statistically significant difference.

Results

Inhibition of cell growth and ROS generation by the flavoenzyme inhibitor DPI

Flavoprotein-dependent ROS play a critical role in cytokine-mediated signal transduction in normal tissues and tumor cells. DPI, a flavoenzyme inhibitor, inhibits the membrane-bound, flavoprotein-containing NOX enzymes (25,30). The present study examined whether DPI affected cell viability in OS cell lines using an MTT assay. A total of five OS cell lines were treated with various DPI concentrations for 24 h and the representative results for HOS and MOS cells are shown in Fig. 1. As hypothesized, DPI treatment decreased the viability of HOS and MOS cells in a dose-dependent manner (P<0.0001 for both cell lines); the IC50 values of HOS and MOS cells were 0.5 and 0.8 µM, respectively. The IC50 values for the other three cell lines were as follows: 0.9 µM (MG-63), 1.2 µM (NOS-1) and 0.7 µM (HuO 9N2), respectively.

To determine whether DPI affects ROS generation, intracellular ROS levels were evaluated by using flow cytometry. It was observed that untreated HOS cells generated ROS, and DPI treatment eliminated ROS generation in HOS cells (Fig. 2). Therefore, dichlorofluorescein (DCF) fluorescence intensity was reduced from 2.8×103 in untreated cells, to 0.9×103 in DPI-treated cells.

Induction of apoptosis by DPI

The present study examined the effect of DPI on apoptosis in OS cells using Annexin V, and observed that DPI treatment markedly increased apoptosis in HOS (54%), HuO 9N2 (43%) and MG63 (28%) cells (Fig. 3). The results suggest that the depletion of ROS, generated by the NOX-like enzymes, triggered apoptosis in OS cells.

Expression of NOX1-5 mRNAs in human OS cell lines

NOX family members produce ROS that are pivotal for cell proliferation. To examine the role of the NOX family in the proliferation of OS cells, mRNA expression of the NOX family members was measured in 5 human OS cell lines by semi-quantitative RT-PCR. NOX2 mRNA was highly expressed in all of the examined OS cell lines, whereas little or no NOX1, NOX3 and NOX5 mRNA was detected (Fig. 4). The OS cell lines also expressed low-moderate levels of NOX4 mRNA (Fig. 4). To provide a comparison, high levels of NOX2 mRNA were also detected in human PBMCs (Fig. 4).

NOX2 and NOX4 expression in human OS cell lines

The present study measured the expression levels of NOX family mRNAs relative to β-actin by RT-qPCR (Fig. 5). The expression was graded as low (NOX/β-actin ratio <2×10−5), moderate (ratio >2×10−5 and <50×10−5) or high (ratio >50×10−5). Relative transcripts were determined by the formula: 1/2(Cqtarget - Cqcontrol) (36). High-level NOX2 mRNA expression was observed in all of the examined OS cell lines, with the highest expression detected in MOS cells. Moderate-level NOX4 mRNA expression was detected in all of the examined OS cell lines. Therefore, NOX2 mRNA expression was higher than NOX4 mRNA expression in the included OS cell lines. For comparison, high-level expression of NOX2 and low-level expression of NOX4 was detected in human peripheral blood mononuclear cells (Fig. 5).

NOX2 and NOX4 mediate ROS production in OS cells

The present study utilized RNA interference to determine whether NOX2/4 mediates ROS generation in OS cells. Primarily, the effects of NOX2 or NOX4-specific siRNAs on the endogenous expression of their mRNAs in HOS cells were evaluated (Fig. 6), revealing that their siRNAs functioned effectively (Fig. 6). In order to elucidate how NOX2 and NOX4 mRNA expression affects ROS generation, siRNAs targeting NOX2 and NOX4 were transiently transfected into OS cells and intracellular ROS levels were evaluated using flow cytometry. Double transfection of NOX2 and NOX4 siRNAs reduced DCF fluorescence intensity to 1.1×103 from the untreated control intensity of 2.8×103 (Fig. 2). Thus, the double transfection of OS cells with NOX2 and NOX4 siRNAs suppressed intracellular ROS levels (39%) compared with the controls (Fig. 2). The results indicate that NOX2 and/or NOX4, at least in part, are responsible for intracellular ROS generation in OS cells. However, ROS generation was not completely inhibited by NOX2 and NOX4 knockdown, suggesting that other NOX proteins and mitochondrial components may also contribute to ROS generation.

NOX2 and NOX4 siRNAs reduce cell viability and induce apoptosis

To explore whether NOX2 and NOX4-mediated ROS affect cell survival, the effect of their knockdown on cell viability and apoptosis was examined. NOX2 and NOX4 knockdown significantly reduced HOS cell viability, by 74 and 65%, respectively, relative to that of untreated cells (P<0.0001 for NOX2 knockdown; P=0.0033 for NOX4 knockdown; Fig. 7A). NOX2 knockdown in HuO 9N2 cells reduced viability by 61% (P=0.0003; Fig. 7B), while NOX4 knockdown in the same cell line reduced viability by 11%, which was not statistically significant. Collectively, these results suggest that among NOX family members, NOX2 has a major role in survival of HOS and HuO 9N2 cells.

To verify whether this reduced cell viability was associated with apoptosis, an Annexin V assay was performed and NOX2 knockdown was observed to markedly induce apoptosis in HOS and HuO 9N2 cells (P=0.0003 for HOS cells; P<0.0001 for HuO 9N2 cells; Fig. 8A and B). Thus, the reduction in cell viability observed with siRNA knockdown was associated with the induction of apoptosis. Therefore, NOX2 and NOX4 siRNAs suppressed ROS generation, and the depletion of ROS by NOX2 knockdown, and DPI treatment induced apoptosis in HOS and HuO 9N2 cells.

Discussion

Cancer cells produce ROS that may act as signaling molecules to promote cell survival and cell growth (23,24,38). It is possible that chronic inflammation may accelerate the development and progression of malignant OS, due to cytokine release and ROS generation.

The present study examined the role of ROS in the viability and apoptosis of OS cell lines. Although ROS are considered to cause stress-induced apoptosis, ROS often confer a survival advantage on cancer cells. The results of the current study demonstrated that suppressing ROS levels by DPI treatment reduced the viability of OS cells. Similar results have been observed in other types of cancer cells, including pancreatic tumor cells (39). The current study also investigated whether the NOX2 and 4-mediated ROS generation conferred anti-apoptotic activity and, thus, a growth advantage to OS cells. As such, the expression levels of NOX genes in five human OS cell lines were examined. RT-PCR analysis revealed that NOX2 and NOX4 mRNAs were expressed in the OS cell lines; however little or no NOX2, NOX3 and NOX5 mRNAs were detected. RT-qPCR revealed that NOX2 and NOX4 mRNAs were expressed in OS cells at high and moderate levels, respectively. In all the examined OS cells, NOX2 mRNA exhibited the highest expression levels. NOX2 siRNAs significantly reduced intracellular ROS generation and OS cell viability. Concordantly, ROS depletion by DPI treatment or NOX2 knockdown induced apoptosis. The results of the present study suggested that NOX2-mediated ROS generation promotes the production of cell survival signals and that ROS depletion induces apoptosis in OS cells.

NOX4 mRNA overexpression has previously been reported in primary breast, ovarian, prostate, melanoma and glioblastoma cancer cell lines (4042). NOX4 expression was moderate-high in two of the four tested ovarian cancer cell lines. Notably, high-level acquired resistance to cisplatin was associated with a marked decrease in the NOX4 mRNA levels in A2780/DDP cells (30). Previously, NOX4 was demonstrated to be an oncoprotein localized in mitochondria (43). In the current study, NOX4 knockdown induced apoptosis in a subset of OS cells. Considering that NOX2 knockdown significantly reduced cell viability and induced apoptosis in all the examined OS cell lines, it is possible that NOX2-mediated ROS generation has a major role in promoting the survival of OS cells. The present study raises the possibility that NOX2 may act as an oncoprotein in the pathogenesis of OS, similar to NOX4 in other types of cancer. Therefore, NOX2 and NOX2-associated signaling molecules may be good candidates for the targeted therapy of OS. NOX2 has been previously reported to be involved in a variety of physiological and pathological conditions, including prion disease (4447). NOX2 has also been implicated in cancer biology (4850) as NOX2 was established to serve a pro-survival role in human leukemia and be involved to promote apoptosis in human glioma (48,50). Considering the possible role of NOX2 in cancer development, it may be of interest to investigate the correlation between NOX2 expression and OS prognosis in further studies.

In conclusion, the present study demonstrated that NOX2-mediated ROS generation promotes the survival of OS cells and that ROS depletion through NOX2 knockdown and DPI treatment leads to apoptosis. The current study raises the possibility of the NOX2-ROS signaling pathway being used as a novel therapeutic target for OS. For example, antioxidant treatments targeted to this signaling pathway have the potential to enhance the therapeutic index of cisplatin-based therapies. Further studies are required to contribute to the development of targeted therapies for OS.

Acknowledgements

The authors would like to thank Dr Masahiko Kanamori and Dr Taketoshi Yasuda (Department of Orthopedic Surgery, School of Medicine, University of Toyama) for providing the OS cell lines.

Funding

The present study was supported by the Aichi Cancer Center to Yuji Miura and a grant of Strategic Research Foundation Grant-Aided Project for Private Universities from the Ministry of Education, Culture, Sports, Science and Technology, Japan to Yoshitaka Hosokawa (grant no. AI 52213).

Availability of data and materials

All data generated or analyzed during this study are available from the corresponding author on reasonable request.

Authors' contributions

Conception and design, KK, YM, YH and KS; development of methodology, KK, YM, SK, AO and YH; acquisition of data (such as provided cells, provided facilities), KK, YM, HK, YH and SK; analysis and interpretation of data, KK, YM, AO, HK, YH and SK; writing, review, and/or revision, KK, YM, YH and SK; administrative, technical or material support (including reporting or organizing data, preparing vectors), KK, YM, SK, AO, HK, YH and KS; study supervision, AO, HK, YH and SK.

Ethics approval and consent to participate

Ethical approval for the present study was obtained from the ethical committee of Aichi Medical University (approval no. 11-039), and informed consent was obtained prior to the start of the present study.

Consent for publication

Informed consent for publication was obtained prior to the start of the present study.

Competing interests

The authors have no competing interests to declare.

Glossary

Abbreviations

Abbreviations:

DCF

dichlorofluorescin

Duox

dual oxidase

DCFH-DA

2′, 7′-dichlorodihydrofluorescein diacetate

DPI

diphenylene iodonium

NADPH

nicotinamide adenine dinucleotide phosphate

NOX

NADPH oxidase

OS

osteosarcoma

PBMC

peripheral blood mononuclear cell

PCR

polymerase chain reaction

ROS

reactive oxygen species

RT-PCR

reverse transcription quantitative-polymerase chain reaction

References

1 

Raymond AK, Ayala AG and Knuutila K: Conventional osteosarcomaFletcher CDM, Unni KK and Mertens F: World Health Organization Classification of Tumours. Pathology and Genetics of Tumours of Soft Tissue and Bone. Lyon: IARC Press; pp. 264–270. 2002

2 

Arndt CA and Crist WM: Common musculoskeletal tumors of childhood and adolescence. N Engl J Med. 341:342–352. 1999. View Article : Google Scholar : PubMed/NCBI

3 

Mirabello L, Troisi RJ and Savage SA: International osteosarcoma incidence patterns in children and adolescents, middle ages and elderly persons. Int J Cancer. 125:229–234. 2009. View Article : Google Scholar : PubMed/NCBI

4 

Mirabello L, Troisi RJ and Savage SA: Osteosarcoma incidence and survival rates from 1973 to 2004: Data from the surveillance, epidemiology, and end results program. Cancer. 115:1531–1543. 2009. View Article : Google Scholar : PubMed/NCBI

5 

Anninga JK, Gelderblom H, Fiocco M, Kroep JR, Taminiau AH, Hogendoorn PC and Egeler RM: Chemotherapeutic adjuvant treatment for osteosarcoma: Where do we stand? Eur J Cancer. 47:2431–2445. 2011. View Article : Google Scholar : PubMed/NCBI

6 

Uribe-Botero G, Russell WO, Sutow WW and Martin RG: Primary osteosarcoma of bone. Clinicopathologic investigation of 243 cases, with necropsy studies in 54. Am J Clin Pathol. 67:427–435. 1977. View Article : Google Scholar : PubMed/NCBI

7 

Bielack SS, Kempf-Bielack B, Delling G, Exner GU, Flege S, Helmke K, Kotz R, Salzer-Kuntschik M, Werner M, Winkelmann W, et al: Prognostic factors in high-grade osteosarcoma of the extremities or trunk: an analysis of 1,702 patients treated on neoadjuvant cooperative osteosarcoma study group protocols. J Clin Oncol. 20:776–790. 2002. View Article : Google Scholar : PubMed/NCBI

8 

Link MP, Goorin AM, Miser AW, Green AA, Pratt CB, Belasco JB, Pritchard J, Malpas JS, Baker AR, Kirkpatrick JA, et al: The effect of adjuvant chemotherapy on relapse-free survival in patients with osteosarcoma of the extremity. N Engl J Med. 314:1600–1606. 1986. View Article : Google Scholar : PubMed/NCBI

9 

Anderson P: Chemotherapy for osteosarcoma with high-dose methotrexate is effective and outpatient therapy is now possible. Nat Clin Pract Oncol. 4:624–625. 2007.PubMed/NCBI

10 

Ritter J and Bielack SS: Osteosarcoma. Ann Oncol. 21 Suppl 7:vii320–vii325. 2010. View Article : Google Scholar : PubMed/NCBI

11 

Ellegast J, Barth TF, Schulte M, Bielack SS, Schmid M and Mayer-Steinacker R: Metastasis of osteosarcoma after 16 years. J Clin Oncol. 29:e62–e66. 2011. View Article : Google Scholar : PubMed/NCBI

12 

Andreou D, Bielack SS, Carrle D, Kevric M, Kotz R, Winkelmann W, Jundt G, Werner M, Fehlberg S, Kager L, et al: The influence of tumor- and treatment-related factors on the development of local recurrence in osteosarcoma after adequate surgery. An analysis of 1355 patients treated on neoadjuvant cooperative osteosarcoma study group protocols. Ann Oncol. 22:1228–1235. 2011. View Article : Google Scholar : PubMed/NCBI

13 

Bielack SS: Osteosarcoma: Time to move on? Eur J Cancer. 46:1942–1945. 2010. View Article : Google Scholar : PubMed/NCBI

14 

Federman N, Bernthal N, Eilber FC and Tap WD: The multidisciplinary management of osteosarcoma. Curr Treat Options Oncol. 10:82–93. 2009. View Article : Google Scholar : PubMed/NCBI

15 

Ito M, Barys L, O'Reilly T, Young S, Gorbatcheva B, Monahan J, Zumstein-Mecker S, Choong PF, Dickinson I, Crowe P, et al: Comprehensive mapping of p53 pathway alterations reveals an apparent role for both SNP309 and MDM2 amplification in sarcomagenesis. Clin Cancer Res. 17:416–426. 2011. View Article : Google Scholar : PubMed/NCBI

16 

Savage SA, Stewart BJ, Liao JS, Helman LJ and Chanock SJ: Telomere stability genes are not mutated in osteosarcoma cell lines. Cancer Genet Cytogenet. 160:79–81. 2005. View Article : Google Scholar : PubMed/NCBI

17 

Kansara M and Thomas DM: Molecular pathogenesis of osteosarcoma. DNA Cell Biol. 26:1–18. 2007. View Article : Google Scholar : PubMed/NCBI

18 

Savage SA, Mirabello L, Wang Z, Gastier-Foster JM, Gorlick R, Khanna C, Flanagan AM, Tirabosco R, Andrulis IL, Wunder JS, et al: Genome-wide association study identifies two susceptibility loci for osteosarcoma. Nat Genet. 45:799–803. 2013. View Article : Google Scholar : PubMed/NCBI

19 

Mirabello L, Koster R, Moriarity BS, Spector LG, Meltzer PS, Gary J, Machiela MJ, Pankratz N, Panagiotou OA, Largaespada D, et al: A genome-wide scan identifies variants in NFIB associated with metastasis in patients with osteosarcoma. Cancer Discov. 5:920–931. 2015. View Article : Google Scholar : PubMed/NCBI

20 

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

21 

Valko M, Rhodes CJ, Moncol J, Izakovic M and Mazur M: Free radicals, metals and antioxidants in oxidative stress-induced cancer. Chem Biol Interact. 160:1–40. 2006. View Article : Google Scholar : PubMed/NCBI

22 

Rhee SG: Cell signaling. H2O2, a necessary evil for cell signaling. Science. 312:1882–1883. 2006. View Article : Google Scholar : PubMed/NCBI

23 

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

24 

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

25 

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

26 

Bedard K and Krause KH: The NOX family of ROS-generating NADPH oxidases: physiology and pathophysiology. Physiol Rev. 87:245–313. 2007. View Article : Google Scholar : PubMed/NCBI

27 

Cheng G, Cao Z, Xu X, 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

28 

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

29 

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. View Article : Google Scholar : PubMed/NCBI

30 

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

31 

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

32 

Kanamori M, Ohmori K, Yasuda T and Yudoh K: Effects of hyperthermia and differentiation on cultured Dunn osteosarcoma cells. Cancer Detect Prev. 27:76–81. 2003. View Article : Google Scholar : PubMed/NCBI

33 

Yasuda T, Kanamori M, Nogami S, Hori T, Oya T, Suzuki K and Kimura T: Establishment of a new human osteosarcoma cell line, UTOS-1: Cytogenetic characterization by array comparative genomic hybridization. J Exp Clin Cancer Res. 28:262009. View Article : Google Scholar : PubMed/NCBI

34 

Hori T, Kondo T, Kanamori M, Tabuchi Y, Ogawa R, Zhao QL, Ahmed K, Yasuda T, Seki S, Suzuki K and Kimura T: Ionizing radiation enhances tumor necrosis factor-related apoptosis-inducing ligand (TRAIL)-induced apoptosis through up-regulations of death receptor 4 (DR4) and death receptor 5 (DR5) in human osteosarcoma cells. J Orthop Res. 28:739–745. 2010.PubMed/NCBI

35 

Kanamori M, Sano A, Yasuda T, Hori T and Suzuki K: Array-based comparative genomic hybridization for genomic-wide screening of DNA copy number alterations in aggressive bone tumors. J Exp Clin Cancer Res. 31:1002012. View Article : Google Scholar : PubMed/NCBI

36 

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

37 

Kumar S, Kain V and Sitasawad SL: Cardiotoxicity of calmidazolium chloride is attributed to calcium aggravation, oxidative and nitrosative stress, and apoptosis. Free Radic Biol Med. 47:699–709. 2009. View Article : Google Scholar : PubMed/NCBI

38 

Diehn M, Cho RW, Lobo NA, Kalisky T, Dorie MJ, Kulp AN, Qian D, Lam JS, Ailles LE, Wong M, et al: Association of reactive oxygen species levels and radioresistance in cancer stem cells. Nature. 458:780–783. 2009. View Article : Google Scholar : PubMed/NCBI

39 

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

40 

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

41 

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

42 

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

43 

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

44 

McCann SK, Dusting GJ and Roulston CL: Nox2 knockout delays infarct progression and increases vascular recovery through angiogenesis in mice following ischaemic stroke with reperfusion. PLoS One. 9:e1106022014. View Article : Google Scholar : PubMed/NCBI

45 

Li ZY, Jiang WY and Cui ZJ: An essential role of NAD(P)H oxidase 2 in UVA-induced calcium oscillations in mast cells. Photochem Photobiol Sci. 14:414–428. 2015. View Article : Google Scholar : PubMed/NCBI

46 

Sorce S, Nuvolone M, Keller A, Falsig J, Varol A, Schwarz P, Bieri M, Budka H and Aguzzi A: The role of the NADPH oxidase NOX2 in prion pathogenesis. PLoS Pathog. 10:e10045312014. View Article : Google Scholar : PubMed/NCBI

47 

Menden H, Welak S, Cossette S, Ramchandran R and Sampath V: Lipopolysaccharide (LPS)-mediated angiopoietin-2-dependent autocrine angiogenesis is regulated by NADPH oxidase 2 (Nox2) in human pulmonary microvascular endothelial cells. J Biol Chem. 290:5449–5461. 2015. View Article : Google Scholar : PubMed/NCBI

48 

Maraldi T, Prata C, Dalla Sega Vieceli F, Caliceti C, Zambonin L, Fiorentini D and Hakim G: NAD(P)H oxidase isoform Nox2 plays a prosurvival role in human leukaemia cells. Free Radic Res. 43:1111–1121. 2009. View Article : Google Scholar : PubMed/NCBI

49 

Jones KJ, Chetram MA, Bethea DA, Bryant LK, Odero-Marah V and Hinton CV: Cysteine (C)-X-C Receptor 4 Regulates NADPH Oxidase-2 during oxidative stress in prostate cancer cells. Cancer Microenviron. Sep 28–2013.(Epub ahead of print). View Article : Google Scholar : PubMed/NCBI

50 

Li SZ, Hu YY, Zhao J, Zhao YB, Sun JD, Yang YF, Ji CC, Liu ZB, Cao WD, Qu Y, et al: MicroRNA-34a induces apoptosis in the human glioma cell line, A172, through enhanced ROS production and NOX2 expression. Biochem Biophys Res Commun. 444:6–12. 2014. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

May-2018
Volume 15 Issue 5

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
Kitamoto K, Miura Y, Karnan S, Ota A, Konishi H, Hosokawa Y and Sato K: Inhibition of NADPH oxidase 2 induces apoptosis in osteosarcoma: The role of reactive oxygen species in cell proliferation. Oncol Lett 15: 7955-7962, 2018
APA
Kitamoto, K., Miura, Y., Karnan, S., Ota, A., Konishi, H., Hosokawa, Y., & Sato, K. (2018). Inhibition of NADPH oxidase 2 induces apoptosis in osteosarcoma: The role of reactive oxygen species in cell proliferation. Oncology Letters, 15, 7955-7962. https://doi.org/10.3892/ol.2018.8291
MLA
Kitamoto, K., Miura, Y., Karnan, S., Ota, A., Konishi, H., Hosokawa, Y., Sato, K."Inhibition of NADPH oxidase 2 induces apoptosis in osteosarcoma: The role of reactive oxygen species in cell proliferation". Oncology Letters 15.5 (2018): 7955-7962.
Chicago
Kitamoto, K., Miura, Y., Karnan, S., Ota, A., Konishi, H., Hosokawa, Y., Sato, K."Inhibition of NADPH oxidase 2 induces apoptosis in osteosarcoma: The role of reactive oxygen species in cell proliferation". Oncology Letters 15, no. 5 (2018): 7955-7962. https://doi.org/10.3892/ol.2018.8291