Generic placeholder image

Current Cancer Drug Targets

Editor-in-Chief

ISSN (Print): 1568-0096
ISSN (Online): 1873-5576

Review Article

Current Development of ROS-Modulating Agents as Novel Antitumor Therapy

Author(s): Nan Wang, Yue Wu, Jinlei Bian, Xue Qian, Hongzhi Lin, Haopeng Sun, Qidong You and Xiaojin Zhang

Volume 17, Issue 2, 2017

Page: [122 - 136] Pages: 15

DOI: 10.2174/1568009616666160216125833

Price: $65

Abstract

Compared to normal cells, usually cancer cells are under higher oxidative stress. Elevating intracellular levels of reactive oxygen species (ROS) by introducing excessive ROS or inhibiting antioxidant system may enhance selectively of cancer cell killing by ROS-modulating agents through stress sensitization or stress overload. Meanwhile due to the adaptive response, normal cells may be capable of maintaining redox homeostasis under exogenous ROS. Here we review ROS-modulating agents in different mechanisms and classify them into groups by various targets for illustrating more clearly. At last, we discuss their side effects and the potential troubles of developing these agents and argue that might be an effective strategy for further exploring to modulate the unique redox regulatory mechanisms of cancer cells.

Keywords: ROS-modulating agents, redox homeostasis, antioxidant system, exogenous generating ROS agents, mitochondrial respiratory chain, antitumor therapy.

Graphical Abstract
[1]
He, H.; Shen, Y.; Zhu, Y.; Chen, S. Prognostic analysis of chronic myeloid leukemia in Chinese population in an imatinib era. Front. Med., 2012, 6(2), 204-211.
[2]
Costa, A.; Scholer-Dahirel, A.; Mechta-Grigoriou, F. The role of reactive oxygen species and metabolism on cancer cells and their microenvironment. Semin. Cancer Biol., 2014, 25, 23-32.
[3]
Chen, F.; Zhuang, X.; Lin, L.; Yu, P.; Wang, Y.; Shi, Y.; Hu, G.; Sun, Y. New horizons in tumor microenvironment biology: challenges and opportunities. BMC Med., 2015, 13, 45.
[4]
Briehl, M.M.; Tome, M.E.; Wilkinson, S.T.; Jaramillo, M.C.; Lee, K. Mitochondria and redox homoeostasis as chemotherapeutic targets. Biochem. Soc. Trans., 2014, 42(4), 939-944.
[5]
Trachootham, D.; Alexandre, J.; Huang, P. Targeting cancer cells by ROS-mediated mechanisms: a radical therapeutic approach? Nat. Rev. Drug Discov., 2009, 8(7), 579-591.
[6]
Halliwell, B. Oxidative stress and neurodegeneration: where are we now? J. Neurochem., 2006, 97(6), 1634-1658.
[7]
Muralidhar, L.H.; Anil, K.M.; Tapas, K.H.; Kishor, K.B.; Sankar, M.; Bartosz, S. Oxidative genome damage and its repair: Implications in aging and neurodegenerative diseases. Mech. Ageing Dev., 2012, 133, 157-168.
[8]
Divya, P.; Mario, S.; Michele, P.; Vanna, S.; Francesco, B.; Angela, S.; Laleh, T.; Nouri, N. Design and discovery of novel quinazolinedione-based redox modulators as therapies for pancreatic cancer. Biochim. Biophys. Acta, 2014, 1840(1), 332-343.
[9]
Dikalov, S.; Griendling, K.K.; Harrison, D.G. Measurement of reactive oxygen species in cardiovascular studies. Hypertension, 2007, 49(4), 717-727.
[10]
Brand, M.D. The sites and topology of mitochondrial superoxide production. Exp. Gerontol., 2010, 45(7-8), 466-472.
[11]
Limon-Pacheco, J.; Gonsebatt, M.E. The role of antioxidants and antioxidant-related enzymes in protective responses to environmentally induced oxidative stress. Mutat. Res., 2009, 674(1-2), 137-147.
[12]
Schlieve, C.R.; Lieven, C.J.; Levin, L.A. Biochemical activity of reactive oxygen species scavengers do not predict retinal ganglion cell survival. Invest. Ophthalmol. Vis. Sci., 2006, 47(9), 3878-3886.
[13]
Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: the next generation. Cell, 2011, 144(5), 646-674.
[14]
Sanz, A.; Caro, P.; Gomez, J.; Barja, G. Testing the vicious cycle theory of mitochondrial ROS production: effects of H2O2 and cumene hydroperoxide treatment on heart mitochondria. J. Bioenerg. Biomembr., 2006, 38(2), 121-127.
[15]
Alberto, J.M.; Jacek, J. Cellular redox pathways as a therapeutic target in the treatment of cancer. Drugs, 2011, 71(11), 1385-1396.
[16]
Su, Z.; Chen, M.; Xiao, Y.; Sun, M.; Zong, L.; Asghar, S.; Dong, M.; Li, H.; Ping, Q.; Zhang, C. ROS-triggered and regenerating anticancer nanosystem: an effective strategy to subdue tumor’s multidrug resistance. J. Control. Release, 2014, 196, 370-383.
[17]
Glasauer, A.; Chandel, N.S. Targeting antioxidants for cancer therapy. Biochem. Pharmacol., 2014, 92(1), 90-101.
[18]
Kumar, B.; Koul, S.; Khandrika, L.; Meacham, R.B.; Koul, H.K. Oxidative stress is inherent in prostate cancer cells and is required for aggressive phenotype. Cancer Res., 2008, 68(6), 1777-1785.
[19]
Valko, M.; Leibfritz, D.; Moncol, J.; Cronin, M.T.; Mazur, M.; Telser, J. Free radicals and antioxidants in normal physiological functions and human disease. J. Biochem. Cell B., 2007, 39(1), 44-84.
[20]
Zhu, C.; Hu, W.; Wu, H.; Hu, X. No evident dose-response relationship between cellular ROS level and its cytotoxicity--a paradoxical issue in ROS-based cancer therapy. Sci. Rep., 2014, 4, 5029.
[21]
Saeidnia, S.; Abdollahi, M. Antioxidants: friends or foe in prevention or treatment of cancer: the debate of the century. Toxicol. Appl. Pharmacol., 2013, 271(1), 49-63.
[22]
Sosa, V.; Moline, T.; Somoza, R.; Paciucci, R.; Kondoh, H. ME, L.L. Oxidative stress and cancer: an overview. Ageing Res. Rev., 2013, 12(1), 376-390.
[23]
Schumacker, P.T. Reactive oxygen species in cancer cells: live by the sword, die by the sword. Cancer Cell, 2006, 10(3), 175-176.
[24]
Catalano, V.; Turdo, A.; Di Franco, S.; Dieli, F.; Todaro, M.; Stassi, G. Tumor and its microenvironment: a synergistic interplay. Semin. Cancer Biol., 2013, 23(6 Pt B), 522-532.
[25]
Lau, A.T.; Wang, Y.; Chiu, J.F. Reactive oxygen species: current knowledge and applications in cancer research and therapeutic. J. Cell. Biochem., 2008, 104(2), 657-667.
[26]
Casteilla, L.; Rigoulet, M. Pe ́nicaud, L. Mitochondrial ROS metabolism: modulation by uncoupling proteins. IUBMB Life, 2001, 52, 181-188.
[27]
(a) Weir, I.E.; Pham, N.A.; Hedley, D.W. Oxidative stress is generated via the mitochondrial respiratory chain during plant cell apoptosis. Cytometry A, 2003, 54(2), 109-117.
(b) Andreyev, A.Y.; Kushnareva, Y.E.; Murphy, A.N.; Starkov, A.A. Mitochondrial ROS metabolism: 10 Years later. Biochemistry, 2015, 80(5), 517-531.
[28]
Halestrap, A.P.; Brennerb, C. The adenine nucleotide translocase: a central component of the mitochondrial permeability transition pore and key player in cell death. Curr. Med. Chem., 2003, 10, 1507-1525.
[29]
Cardoso, S.; Santos, M.S.; Moreno, A.; Moreira, P.I. UCP2 and ANT differently modulate proton-leak in brain mitochondria of long-term hyperglycemic and recurrent hypoglycemic rats. J. Bioenerg. Biomembr., 2013, 45(4), 397-407.
[30]
Lena, A.; Rechichi, M.; Salvetti, A.; Bartoli, B.; Vecchio, D.; Scarcelli, V.; Amoroso, R.; Benvenuti, L.; Gagliardi, R.; Gremigni, V.; Rossi, L. Drugs targeting the mitochondrial pore act as cytotoxic and cytostatic agents in temozolomide-resistant glioma cells. J. Transl. Med., 2009, 7, 13.
[31]
Fulda, S.; Galluzzi, L.; Kroemer, G. Targeting mitochondria for cancer therapy. Nat. Rev. Drug Discov., 2010, 9(6), 447-464.
[32]
Chowdhury, R.; Chatterjee, R.; Giri, A.K.; Mandal, C.; Chaudhuri, K. Arsenic-induced cell proliferation is associated with enhanced ROS generation, Erk signaling and CyclinA expression. Toxicol. Lett., 2010, 198(2), 263-271.
[33]
Wang, H.; Chen, X.Y.; Wang, B.S.; Rong, Z.X.; Qi, H.; Chen, H.Z. The efficacy and safety of arsenic trioxide with or without all-trans retinoic acid for the treatment of acute promyelocytic leukemia: a meta-analysis. Leuk. Res., 2011, 35(9), 1170-1177.
[34]
Yolanda, S.; Donna, A.; Elena, B.; Patricio, A. Arsenic trioxide as an anti-tumour agent: mechanisms of action and strategies of sensitization. J. Appl. Biomed., 2010, 8(4), 199-208.
[35]
Emadi, A.; Gore, S.D. Arsenic trioxide - An old drug rediscovered. Blood Rev., 2010, 24(4-5), 191-199.
[36]
(a) Fruehauf, J.P.; Meyskens, F.L., Jr Reactive oxygen species: a breath of life or death? Clin. Cancer Res., 2007, 13(3), 789-794.
(b) Lu, J.; Chew, E.H.; Holmgren, A. Targeting thioredoxin reductase is a basis for cancer therapy by arsenic trioxide. Proc. Natl. Acad. Sci. USA, 2007, 104(30), 12288-12293.
[37]
Filomeni, G.; Desideri, E.; Cardaci, S.; Rotilio, G.; Ciriolo, M.R. Under the ROS: Thiol network is the principal suspect for autophagy commitment. Autophagy, 2014, 6(7), 999-1005.
[38]
(a) Loor, G.; Kondapalli, J.; Schriewer, J.M.; Chandel, N.S.; Vanden Hoek, T.L.; Schumacker, P.T. Menadione triggers cell death through ROS-dependent mechanisms involving PARP activation without requiring apoptosis. Free Radic. Biol. Med., 2010, 49(12), 1925-1936.
(b) Chowdhury, R.; Chowdhury, S.; Roychoudhury, P.; Mandal, C.; Chaudhuri, K. Arsenic induced apoptosis in malignant melanoma cells is enhanced by menadione through ROS generation, p38 signaling and p53 activation. Apoptosis, 2009, 14(1), 108-123.
[39]
Cao, J.; Ying, M.; Xie, N.; Lin, G.; Dong, R.; Zhang, J.; Yan, H.; Yang, X.; He, Q.; Yang, B. The oxidation states of DJ-1 dictate the cell fate in response to oxidative stress triggered by 4-hpr: autophagy or apoptosis? Antioxid. Redox Signal., 2014, 21(10), 1443-1459.
[40]
Cuperus, R.; Leen, R.; Tytgat, G.A.; Caron, H.N.; van Kuilenburg, A.B. Fenretinide induces mitochondrial ROS and inhibits the mitochondrial respiratory chain in neuroblastoma. Cell. Mol. Life Sci., 2010, 67(5), 807-816.
[41]
Ledet, G.A.; Graves, R.A.; Glotser, E.Y.; Mandal, T.K.; Bostanian, L.A. Preparation and in vitro evaluation of hydrophilic fenretinide nanoparticles. Int. J. Pharm., 2015, 479(2), 329-337.
[42]
Mody, N.; McIlroy, G.D. The mechanisms of Fenretinide-mediated anti-cancer activity and prevention of obesity and type-2 diabetes. Biochem. Pharmacol., 2014, 91(3), 277-286.
[43]
You, B.R.; Park, W.H. Gallic acid-induced lung cancer cell death is related to glutathione depletion as well as reactive oxygen species increase. Toxicol. In Vitro, 2010, 24(5), 1356-1362.
[44]
Subramanian, A.P.; John, A.A.; Vellayappan, M.V.; Balaji, A.; Jaganathan, S.K.; Supriyantoa, E.; Yusofa, M. Gallic acid: prospects and molecular mechanisms of its anticancer activity. RSC Advances, 2015, 5, 35608-35621.
[45]
Almeida, A.S.; Figueiredo-Pereira, C.; Vieira, H.L. Carbon monoxide and mitochondria-modulation of cell metabolism, redox response and cell death. Front. Physiol., 2015, 6, 33.
[46]
Choi, Y.K.; Por, E.D.; Kwon, Y.G.; Kim, Y.M. Regulation of ROS production and vascular function by carbon monoxide. Oxid. Med. Cell. Longev., 2012. 794237
[47]
Kim, H.S.; Loughran, P.A.; Rao, J.; Billiar, T.R.; Zuckerbraun, B.S. Carbon monoxide activates NF-kappaB via ROS generation and Akt pathways to protect against cell death of hepatocytes. Am. J. Physiol. Gastrointest. Liver Physiol., 2008, 295(1), G146-G152.
[48]
Sun, J.; Xu, Y.; Dai, Z.; Sun, Y. Intermittent high glucose stimulate MCP-l, IL-18, and PAI-1, but inhibit adiponectin expression and secretion in adipocytes dependent of ROS. Cell Biochem. Biophys., 2009, 55(3), 173-180.
[49]
Slattery, K.; Bentley, D.; Coutts, A.J. The role of oxidative, inflammatory and neuroendocrinological systems during exercise stress in athletes: implications of antioxidant supplementation on physiological adaptation during intensified physical training. Sports Med., 2015, 45(4), 453-471.
[50]
Bennett, L.L.; Rojas, S.; Seefeldt, T. Role of Antioxidants in the Prevention of Cancer. J. Exp. Clin. Med., 2012, 4(4), 215-222.
[51]
Morris, G.; Anderson, G.; Dean, O.; Berk, M.; Galecki, P.; Martin-Subero, M.; Maes, M. The glutathione system: a new drug target in neuroimmune disorders. Mol. Neurobiol., 2014, 50(3), 1059-1084.
[52]
Romani, A.A.; Desenzani, S.; Soliani, P.; Borghetti, A.F. HSP27 expression is associated with increased gemcitabine resistance in cholangiocarcinoma cell lines. Biomed. Pharmacother., 2008, 62(8), 514-515.
[53]
Wang, Z.; Wang, J.; Xie, R.; Liu, R.; Lu, Y. Mitochondria-derived reactive oxygen species play an important role in Doxorubicin-induced platelet apoptosis. Int. J. Mol. Sci., 2015, 16(5), 11087-11100.
[54]
Ravi, D.; Das, K.C. Redox-cycling of anthracyclines by thioredoxin system: increased superoxide generation and DNA damage. Cancer Chemother. Pharmacol., 2004, 54(5), 449-458.
[55]
Wellington, K.W. Understanding cancer and the anticancer activities of naphthoquinones - a review. RSC Advances, 2015, 5(26), 20309-20338.
[56]
Walsh, M.E.; Shi, Y.; Van Remmen, H. The effects of dietary restriction on oxidative stress in rodents. Free Radic. Biol. Med., 2014, 66, 88-99.
[57]
Owen, W.G. Mechanism of action, metabolism, and toxicitoyf buthionine sulfoximine and its higher homologs, potent inhibitors of glutathione synthesis. J. Biol. Chem., 1982, 257, 13704-13714.
[58]
Raza, A.; Galili, N.; Callander, N.; Ochoa, L.; Piro, L.; Emanuel, P.; Williams, S.; Burris, H., III; Faderl, S.; Estrov, Z.; Curtin, P.; Larson, R.A.; Keck, J.G.; Jones, M.; Meng, L.; Brown, G.L. Phase 1-2a multicenter dose-escalation study of ezatiostat hydrochloride liposomes for injection (Telintra, TLK199), a novel glutathione analog prodrug in patients with myelodysplastic syndrome. J. Hematol. Oncol., 2009, 2, 20.
[59]
Lyons, R.M.; Wilks, S.T.; Young, S.; Brown, G.L. Oral ezatiostat HCl (Telintra (R), TLK199) and idiopathic chronic neutropenia(ICN): a case report of complete response of a patient with G-CSF resistant ICN following treatment with ezatiostat, a glutathione S-transferase P1-1 (GSTP1-1) inhibitor. J. Hematol. Oncol., 2011, 4, 43.
[60]
Tai, K.W.; Lii, C.K.; Chou, M.Y.; Chang, Y.C. Relationship between intracellular glutathione level and the mode of cell death induced by pingyangmycin. Oral Oncol., 2003, 39, 13-18.
[61]
Yoon, J.Y.; Ishdorj, G.; Graham, B.A.; Johnston, J.B.; Gibson, S.B. Valproic acid enhances fludarabine-induced apoptosis mediated by ROS and involving decreased AKT and ATM activation in B-cell-lymphoid neoplastic cells. Apoptosis, 2014, 19(1), 191-200.
[62]
Park, S.; Park, J.A.; Kim, Y.E.; Song, S.; Kwon, H.J.; Lee, Y. Suberoylanilide hydroxamic acid induces ROS-mediated cleavage of HSP90 in leukemia cells. Cell Stress Chaperones, 2015, 20(1), 149-157.
[63]
Park, J.M.; Kim, A.; Oh, J.H.; Chung, A.S. Methylseleninic acid inhibits PMA-stimulated pro-MMP-2 activation mediated by MT1-MMP expression and further tumor invasion through suppression of NF-kappaB activation. Carcinogenesis, 2007, 28(4), 837-847.
[64]
Shi, J.; Zuo, H.; Ni, L.; Xia, L.; Zhao, L.; Gong, M.; Nie, D.; Gong, P.; Cui, D.; Shi, W.; Chen, J. An IDH1 mutation inhibits growth of glioma cells via GSH depletion and ROS generation. Neurol. Sci., 2014, 35(6), 839-845.
[65]
Sahu, R.P.; Zhang, R.; Batra, S.; Shi, Y.; Srivastava, S.K. Benzyl isothiocyanate-mediated generation of reactive oxygen species causes cell cycle arrest and induces apoptosis via activation of MAPK in human pancreatic cancer cells. Carcinogenesis, 2009, 30(10), 1744-1753.
[66]
Khan, M.; Yi, F.; Rasul, A.; Li, T.; Wang, N.; Gao, H.; Gao, R.; Ma, T. Alantolactone induces apoptosis in glioblastoma cells via GSH depletion, ROS generation, and mitochondrial dysfunction. IUBMB Life, 2012, 64(9), 783-794.
[67]
Tusskorn, O.; Prawan, A.; Senggunprai, L.; Kukongviriyapan, U.; Kukongviriyapan, V. Phenethyl isothiocyanate induces apoptosis of cholangiocarcinoma cells through interruption of glutathione and mitochondrial pathway. Naunyn Schmiedebergs Arch. Pharmacol., 2013, 386(11), 1009-1016.
[68]
Indira, J.; Aaron, S.G.; Gayathri, C.; Satya, P. KyoungHyun, K.; Robert, B.; Un-Ho, J.; Stephen, S. Mechanism of action of phenethylisothiocyanate and other reactive oxygen species-inducing anticancer agents. Mol. Cell. Biol., 2014, 34(13), 2382-2395.
[69]
Katerina, D.; Claire, M.P.; Margaret, E.T.; Margaret, M.B.; Thomas, M.; Robert, T.D. Induction of oxidative stress and apoptosis in myeloma cells by the aziridine-containing agent imexon. Biochem. Pharmacol., 2000, 60, 749-758.
[70]
Sheveleva, E.V.; Landowski, T.H.; Samulitis, B.K.; Bartholomeusz, G.; Powis, G.; Dorr, R.T. Imexon induces an oxidative endoplasmic reticulum stress response in pancreatic cancer cells. Mol. Cancer Res., 2012, 10(3), 392-400.
[71]
Mittal, M.; Khan, K.; Pal, S.; Porwal, K.; China, S.P.; Barbhuyan, T.K.; Baghel, K.S.; Rawat, T.; Sanyal, S.; Bhadauria, S.; Sharma, V.L.; Chattopadhyay, N. The thiocarbamate disulphide drug, disulfiram induces osteopenia in rats by inhibition of osteoblast function due to suppression of acetaldehyde dehydrogenase activity. Toxicol. Sci., 2014, 139(1), 257-270.
[72]
Witaicenis, A.; Luchini, A.C.; Hiruma-Lima, C.A.; Felisbino, S.L.; Garrido-Mesa, N.; Utrilla, P.; Galvez, J.; Di Stasi, L.C. Suppression of TNBS-induced colitis in rats by 4-methylesculetin, a natural coumarin: comparison with prednisolone and sulphasalazine. Chem. Biol. Interact., 2012, 195(1), 76-85.
[73]
Park, W.H.; Kim, S.H. MG132, a proteasome inhibitor, induces human pulmonary fibroblast cell death via increasing ROS levels and GSH depletion. Oncol. Rep., 2012, 27(4), 1284-1291.
[74]
Nam, Y.J. Lee da, H.; Shin, Y.K.; Sohn, D.S.; Lee, C.S. Flavanonol taxifolin attenuates proteasome inhibition-induced apoptosis in differentiated PC12 cells by suppressing cell death process. Neurochem. Res., 2015, 40(3), 480-491.
[75]
Boldogh, I.; Liebenthal, D.; Hughes, T.K.; Juelich, T.L.; Georgiades, J.A.; Kruzel, M.L.; Stanton, G.J. Modulation of 4HNE-Mediated signaling by proline-rich peptides from ovine colostrum. J. Mol. Neurosci., 2002, 20, 125-134.
[76]
You, B.R.; Kim, S.H.; Park, W.H. Reactive oxygen species, glutathione, and thioredoxin influence suberoyl bishydroxamic acid-induced apoptosis in A549 lung cancer cells. Tumour Biol., 2015, 36(5), 3429-3439.
[77]
Wangpaichitr, M.; Sullivan, E.J.; Theodoropoulos, G.; Wu, C.; You, M.; Feun, L.G.; Lampidis, T.J.; Kuo, M.T.; Savaraj, N. The relationship of thioredoxin-1 and cisplatin resistance: its impact on ROS and oxidative metabolism in lung cancer cells. Mol. Cancer Ther., 2012, 11(3), 604-615.
[78]
Baker, A.F.; Dragovich, T.; Tate, W.R.; Ramanathan, R.K.; Roe, D.; Hsu, C.H.; Kirkpatrick, D.L.; Powis, G. The antitumor thioredoxin-1 inhibitor PX-12 (1-methylpropyl 2-imidazolyl disulfide) decreases thioredoxin-1 and VEGF levels in cancer patient plasma. J. Lab. Clin. Med., 2006, 147(2), 83-90.
[79]
Magda, D.; Miller, R.A. Motexafin gadolinium: a novel redox active drug for cancer therapy. Semin. Cancer Biol., 2006, 16(6), 466-476.
[80]
Luo, Z.; Yu, L.; Yang, F.; Zhao, Z.; Yu, B.; Lai, H.; Wong, K.H.; Ngai, S.M.; Zheng, W.; Chen, T. Ruthenium polypyridyl complexes as inducer of ROS-mediated apoptosis in cancer cells by targeting thioredoxin reductase. Metallomics, 2014, 6(8), 1480-1490.
[81]
Lin, T.S.; Naumovski, L.; Lecane, P.S.; Lucas, M.S.; Moran, M.E.; Cheney, C.; Lucas, D.M.; Phan, S.C.; Miller, R.A.; Byrd, J.C. Effects of motexafin gadolinium in a phase II trial in refractory chronic lymphocytic leukemia. Leuk. Lymphoma, 2009, 50(12), 1977-1982.
[82]
Satoshi, T.; Nobuo, S.; Mikio, Y.; Tohru, K. Inactivation of acrolein by sodium 2-mercaptoethanesulfonate using headspace-solid-phase microextraction gas chromatography and mass spectrometry. J. Chromatogr. B Analyt., 2003, 791(1-2), 365-369.
[83]
Sun, Y.; Rigas, B. The thioredoxin system mediates redox-induced cell death in human colon cancer cells: implications for the mechanism of action of anticancer agents. Cancer Res., 2008, 68(20), 8269-8277.
[84]
Rigas, B.; Sun, Y. Induction of oxidative stress as a mechanism of action of chemopreventive agents against cancer. Br. J. Cancer, 2008, 98(7), 1157-1160.
[85]
Sun, Y.; Chen, J.; Rigas, B. Chemopreventive agents induce oxidative stress in cancer cells leading to COX-2 overexpression and COX-2-independent cell death. Carcinogenesis, 2009, 30(1), 93-100.
[86]
Zhou, H.; Huang, L.; Sun, Y.; Rigas, B. Nitric oxide-donating aspirin inhibits the growth of pancreatic cancer cells through redox-dependent signaling. Cancer Lett., 2009, 273(2), 292-299.
[87]
Karin, B.; Manuel, L.; Christoph, R.; Renate, G.; Patrick, B. Correlations between the activities of 19 standard anticancer agents, antioxidative enzyme activities and the expression of ATP-binding cassette transporters: comparison with the National Cancer Institute data. Anticancer Drugs, 2007, 18, 389-404.
[88]
Visagie, M.H.; Joubert, A.M. In vitro effects of 2-methoxyestradiol-bis-sulphamate on reactive oxygen species and possible apoptosis induction in a breast adenocarcinoma cell line. Cancer Cell Int., 2011, 11(1), 43.
[89]
(a) Zhang, Q.; Ma, Y.; Cheng, Y.F.; Li, W.J.; Zhang, Z.; Chen, S.Y. Involvement of reactive oxygen species in 2-methoxyestradiol-induced apoptosis in human neuroblastoma cells. Cancer Lett., 2011, 313(2), 201-210.
(b) Xin, W.; Jinming, W.; Fengtian, W.; Bo, L.; Canhua, H.; Yuquan, W. Deconvoluting the role of reactive oxygen species and autophagy in human diseases. Free Radic. Biol. Med., 2013, 65, 402-410.
[90]
Kim, K.K.; Kawar, N.M.; Singh, R.K.; Lange, T.S.; Brard, L.; Moore, R.G. Tetrathiomolybdate induces doxorubicin sensitivity in resistant tumor cell lines. Gynecol. Oncol., 2011, 122(1), 183-189.
[91]
Lin, J.; Zahurak, M.; Beer, T.M.; Ryan, C.J.; Wilding, G.; Mathew, P.; Morris, M.; Callahan, J.A.; Gordon, G.; Reich, S.D.; Carducci, M.A.; Antonarakis, E.S. A non-comparative randomized phase II study of 2 doses of ATN-224, a copper/zinc superoxide dismutase inhibitor, in patients with biochemically recurrent hormone-naive prostate cancer. Urol. Oncol., 2013, 31(5), 581-588.
[92]
Alexandre, J.; Nicco, C.; Chereau, C.; Laurent, A.; Weill, B.; Goldwasser, F.; Batteux, F. Improvement of the therapeutic index of anticancer drugs by the superoxide dismutase mimic mangafodipir. J. Natl. Cancer Inst., 2006, 98(4), 236-244.
[93]
Liu, J.; Yin, Y.; Song, Z.; Li, Y.; Jiang, S.; Shao, C.; Wang, Z. NADH: flavin oxidoreductase/NADH oxidase and ROS regulate microsclerotium development in Nomuraea rileyi. World J. Microbiol. Biotechnol., 2014, 30(7), 1927-1935.
[94]
Zhang, S.; Ong, C.N.; Shen, H.M. Critical roles of intracellular thiols and calcium in parthenolide-induced apoptosis in human colorectal cancer cells. Cancer Lett., 2004, 208(2), 143-153.
[95]
Sun, Y.; St Clair, D.K.; Xu, Y.; Crooks, P.A.; St Clair, W.H. A NADPH oxidase-dependent redox signaling pathway mediates the selective radiosensitization effect of parthenolide in prostate cancer cells. Cancer Res., 2010, 70(7), 2880-2890.
[96]
Lan, B.; Wan, Y.J.; Pan, S.; Wang, Y.; Yang, Y.; Leng, Q.L.; Jia, H.; Liu, Y.H.; Zhang, C.Z.; Cao, Y. Parthenolide induces autophagy via the depletion of 4E-BP1. Biochem. Biophys. Res. Commun., 2015, 456(1), 434-439.
[97]
(a) Karmakar, A.; Xu, Y.; Mustafa, T.; Kannarpady, G.; Bratton, S.M.; Radominska-Pandya, A.; Crooks, P.A.; Biris, A.S. Nanodelivery of parthenolide using functionalized nanographene enhances its anticancer activity. RSC Advances, 2015, 5(4), 2411-2420.
(b) Rajasubramaniam, S.; Praveen, K.; Hitesh, A.; Liang, C.; Peter, C.; Sundar, N.; Tyler, P.; James, K.; William, M.; Harikrishna, N.; Christopher, J.S. A water soluble parthenolide analog suppresses in vivo tumer growth of two tobacco-associated cancer, lung and bladder cancer, by targeting NF-kB and generating reactive oxygen species. Int. J. Cancer, 2011, 128, 14.
[98]
Butterfield, D.A. The 2013 SFRBM discovery award: selected discoveries from the butterfield laboratory of oxidative stress and its sequela in brain in cognitive disorders exemplified by Alzheimer disease and chemotherapy induced cognitive impairment. Free Radic. Biol. Med., 2014, 74, 157-174.
[99]
(a) Li, J.Z.; Ke, Y.; Misra, H.P.; Trush, M.A.; Li, Y.R.; Zhu, H.; Jia, Z. Mechanistic studies of cancer cell mitochondria- and NQO1-mediated redox activation of beta-lapachone, a potentially novel anticancer agent. Toxicol. Appl. Pharmacol., 2014, 281(3), 285-293.
(b) Park, E.J.; Choi, K.S.; Kwon, T.K. beta-Lapachone-induced reactive oxygen species (ROS) generation mediates autophagic cell death in glioma U87 MG cells. Chem. Biol. Interact., 2011, 189(1-2), 37-44.
[100]
Bair, J.S.; Palchaudhuri, R.; Hergenrother, P.J. Chemistry and biology of deoxynyboquinone, a potent inducer of cancer cell death. J. Am. Chem. Soc., 2010, 132(15), 5469-5478.
[101]
Huang, X.; Dong, Y.; Bey, E.A.; Kilgore, J.A.; Bair, J.S.; Li, L.S.; Patel, M.; Parkinson, E.I.; Wang, Y.; Williams, N.S.; Gao, J.; Hergenrother, P.J.; Boothman, D.A. An NQO1 substrate with potent antitumor activity that selectively kills by PARP1-induced programmed necrosis. Cancer Res., 2012, 72(12), 3038-3047.
[102]
Bian, J.; Xu, L.; Deng, B.; Qian, X.; Fan, J.; Yang, X.; Liu, F.; Xu, X.; Guo, X.; Li, X.; Sun, H.; You, Q.; Zhang, X. Synthesis and evaluation of (+/-)-dunnione and its ortho-quinone analogues as substrates for NAD(P)H:quinone oxidoreductase 1 (NQO1). Med. Chem. Lett., 2015, 25(6), 1244-1248.
[103]
Bian, J.; Deng, B.; Xu, L.; Xu, X.; Wang, N.; Hu, T.; Yao, Z.; Du, J.; Yang, L.; Lei, Y.; Li, X.; Sun, H.; Zhang, X.; You, Q. 2-Substituted 3-methylnaphtho[1,2-b]furan-4,5-diones as novel L-shaped ortho-quinone substrates for NAD(P)H:quinone oxidoreductase (NQO1). Eur. J. Med. Chem., 2014, 82, 56-67.
[104]
Nagahiko, Y.; Ryuichiro, N.; Koshiro, U.; Shin-ichiro, U. Hematoporphyrin as a sensitizer of cell-damaging effect of ultrasound. Jpn. J. Cancer Res., 1989, 80, 219-222.
[105]
Wang, J.; Guo, Y.; Gao, J.; Jin, X.; Wang, Z.; Wang, B.; Li, K.; Li, Y. Detection and comparison of reactive oxygen species (ROS) generated by chlorophyllin metal(Fe, Mg and Cu) complexes under ultrasonic and visible-light irradiation. Ultrason. Sonochem., 2011, 18(5), 1028-1034.
[106]
(a) Valko, M.; Rhodes, C.J.; Moncol, J.; Izakovic, M.; Mazur, M. Free radicals, metals and antioxidants in oxidative stress-induced cancer. Chem. Biol. Interact., 2006, 160(1), 1-40.
(b) Joseph, S.B.; Rahul, P.; Paul, J.H. Chemistry and biology of deoxynyboquinone, a potent inducer of cancer cell death. J. Am. Chem. Soc., 2010, 132, 5469-5478.
[107]
Chow, M.S.; Liu, L.V.; Solomon, E.I. Further insights into the mechanism of the reaction of activated bleomycin with DNA. Proc. Natl. Acad. Sci. USA, 2008, 105(36), 13241-13245.
[108]
Guo, Y.; Cheng, C.; Wang, J.; Jin, X.; Liu, B.; Wang, Z.; Gao, J.; Kang, P. Oxidation-extraction spectrometry of reactive oxygen species (ROS) generated by chlorophyllin magnesium (Chl-Mg) under ultrasonic irradiation. Spectrochimica acta. Part A. Mol. Biomol. Spectrosc., 2011, 79(5), 1099-1104.
[109]
(a) Wang, J.; Guo, Y.; Liu, B.; Jin, X.; Liu, L.; Xu, R.; Kong, Y.; Wang, B. Detection and analysis of reactive oxygen species (ROS) generated by nano-sized TiO2 powder under ultrasonic irradiation and application in sonocatalytic degradation of organic dyes. Ultrason. Sonochem., 2011, 18(1), 177-183.
(b) Guo, Y.; Cheng, C.; Wang, J.; Wang, Z.; Jin, X.; Li, K.; Kang, P.; Gao, J. Detection of reactive oxygen species (ROS) generated by TiO2(R), TiO2(R/A) and TiO2(A) under ultrasonic and solar light irradiation and application in degradation of organic dyes. J. Hazard. Mater., 2011, 192(2), 786-793.
[110]
Barbi de Moura, M.; Vincent, G.; Fayewicz, S.L.; Bateman, N.W.; Hood, B.L.; Sun, M.; Suhan, J.; Duensing, S.; Yin, Y.; Sander, C.; Kirkwood, J.M.; Becker, D.; Conrads, T.P.; Van Houten, B.; Moschos, S.J. Mitochondrial respiration--an important therapeutic target in melanoma. PLoS One, 2012, 7(8), e40690.
[111]
Blackman, R.K.; Cheung-Ong, K.; Gebbia, M.; Proia, D.A.; He, S.; Kepros, J.; Jonneaux, A.; Marchetti, P.; Kluza, J.; Rao, P.E.; Wada, Y.; Giaever, G.; Nislow, C. Mitochondrial electron transport is the cellular target of the oncology drug elesclomol. PLoS One, 2012, 7(1), e29798.
[112]
(a) Poschmann, G.; Grzendowski, M.; Stefanski, A.; Bruns, E.; Meyer, H.E.; Stuhler, K. Redox proteomics reveal stress responsive proteins linking peroxiredoxin-1 status in glioma to chemosensitivity and oxidative stress. Biochim. Biophys. Acta, 2015, 1854(6), 624-631.
(b) Chen, L.; Li, X.; Liu, L.; Yu, B.; Xue, Y.; Liu, Y. Erastin sensitizes glioblastoma cells to temozolomide by restraining xCT and cystathionine-gamma-lyase function. Oncol. Rep., 2015, 33(3), 1465-1474.
[113]
Voborilova, J.; Nemcova-Furstova, V.; Neubauerova, J.; Ojima, I.; Zanardi, I.; Gut, I.; Kovar, J. Cell death induced by novel fluorinated taxanes in drug-sensitive and drug-resistant cancer cells. Invest. New Drugs, 2011, 29(3), 411-423.
[114]
Hamilton, G. Cytotoxic effects of fascaplysin against small cell lung cancer cell lines. Mar. Drugs, 2014, 12(3), 1377-1389.
[115]
Kawata, K.; Shimazaki, R.; Okabe, S. Comparison of gene expression profiles in HepG2 cells exposed to arsenic, cadmium, nickel, and three model carcinogens for investigating the mechanisms of metal carcinogenesis. Environ. Mol. Mutagen., 2009, 50(1), 46-59.
[116]
Bjelakovic, G.; Nikolova, D.; Gluud, L.L.; Simonetti, R.G.; Gluud, C. Mortality in randomized trials of antioxidant supplements for primary and secondary prevention: systematic review and meta-analysis. JAMA, 2007, 297, 842-857.
[117]
Halliwell, B. The antioxidant paradox. Lancet, 2000, 355, 1179-1180.

Rights & Permissions Print Cite
© 2024 Bentham Science Publishers | Privacy Policy