Chemotherapy: Open Access

Chemotherapy: Open Access
Open Access

ISSN: 2167-7700

+44 1223 790975

Mini Review - (2016) Volume 5, Issue 2

Targeting ROS for Cancer Therapy

Jingwu Dong1, Bin Liu2,3 and Runzhi Zhu1,2,4*
1Clinical Research Center, Xuyi People’s Hospital, Xuyi 211700, PR China
2Key Laboratory of Hepatobiliary Surgery, Affiliated Hospital of Guangdong Medical University, Zhanjiang Key Laboratory of Hepatobiliary Diseases, Zhanjiang 524001, PR China
3Institute of Traditional Chinese Medicine & Natural Products, College of Pharmacy, Jinan University, Guangzhou 510632, PR China
4Hamon Center for Therapeutic Oncology Research, UT Southwestern Medical Center at Dallas, Dallas, Texas 75390, USA
*Corresponding Author: Prof. Runzhi Zhu, Clinical Research Center of Xuyi People’s Hospital, 28 Hongwu Road Xuyi Economic Development District, Xuyi, Jiangsu 211700, PR China, Tel: +86 138-5166-1148 Email:

Abstract

Reactive oxygen species (ROS) act as a second messenger in cell signaling and participate in a variety of biological processes in normal cells. Under normal physiological conditions, ROS are produced and then eliminated by ROS scavenging system to maintain cellular redox balance. Levels of ROS could change with endogenous or exogenous reasons, leading to DNA damage, oxidative stress and abnormal cell signaling. Reduction of ROS could lead to interruption of cell signaling, thus influence the cellular homeostasis. Cancer cell metabolism that often appear in the redox imbalance and oxidative stress, in the situation, antioxidant mechanism needs to balance this pressure, which has been the symbol of tumorigenesis [

Keywords: Preventing cancer; Genomic; Cell signalling; Reactive oxygen species

Introduction

Reactive oxygen species (ROS) act as a second messenger in cell signaling and participate in a variety of biological processes in normal cells. Under normal physiological conditions, ROS are produced and then eliminated by ROS scavenging system to maintain cellular redox balance. Levels of ROS could change with endogenous or exogenous reasons, leading to DNA damage, oxidative stress and abnormal cell signaling. Reduction of ROS could lead to interruption of cell signaling, thus influence the cellular homeostasis. Cancer cell metabolism that often appear in the redox imbalance and oxidative stress, in the situation, antioxidant mechanism needs to balance this pressure, which has been the symbol of tumorigenesis [1]. The role of ROS in cancer is two-fold. On the one hand, ROS could lead to cancer development by activating signaling pathways that regulate cell proliferation, survival, angiogenesis and metastasis pathways [2-4]. Cancer cells are characterized by an increased rate of spatially localized ROS production, compared to normal cells due to a loss in proper redox control. cells adapt to oxidative stress by up-regulating antioxidant proteins. However, on the other hand, excessive ROS could induce cell death, aging and cell cycle arrest [5,6]. For example, paclitaxel could induce breast cancer cell death by increasing ROS generation [7]. Previous studies demonstrated that cancer cells make them adapt the high level of ROS and activate the antioxidant pathways [8,9] resulting in increased ROS clearance [10] in order to maintain the level of ROS. Various studies showed that the use of antioxidants inhibitors could selectively kill tumor cells, induce inhibition of tumorigenesis [11-15]. Compared with the cancer cells, the rate of producing ROS in normal cells is lower. Targeting the ROS signal to treat cancer is a very promising way. Elevating ROS and promoting ROS by using dietary antioxidants or ROS-generation drugs (paclitaxel) could selectively kill or arrest cancer cells. But most clinical trials have failed to show beneficial effects of dietary antioxidants in variety of cancer types or cause significant toxicity to normal cells [16]. The purpose of this review is to summarize the ROS key mechanisms in cancer development, as well as to target these potential redox mechanisms, present new strategy to prevent cancer.

The basic conception of ROS

The sources of ROS are mainly produced from mitochondria and NADPH oxidase (NOXs) family. Different levels of ROS could directly regulate different cell signals and even adjust the ROS production. H2O2 generated by the NOX protein and mitochondria could damage proteins and iron-sulfur clusters. However, localized in mitochondria (SOD2), cytoplasm (SOD1) and extracellular matrix (SOD3) of SODs could be quickly transformed ROS into H2O2 [17]. The intracellular concentration of peroxide is very low. H2O2 is a more stable form of diffusion and ROS could be selectively applied to reaction with cysteine residues of protein, thereby controlling cell signals. There are many ways to make intracellular H2O2 transform into H2O. During these processes, cysteine can be modified, to regulate active protein in relative signal pathways. Peroxiredoxins (PRXs) become a key regulator of quenching H2O2. PRXs regulates H2O2-mediated oxide by regulating the active site of cysteine [18]. The function of Glutathione peroxidase (GPXs) is similar with PRXs, transforming into oxidation of H2O2-mediated GSH while scavenging H2O2, and GPXs is the most abundant intracellular antioxidant [19]. Once GSH levels reduced, it could be reactivated by glutathione reductase and NADPH.

ROS and Cancer

ROS regulate cell signaling

The traditional views considered that ROS are the toxic metabolic by-products, leading to cell damage [20]. Recently, research demonstrates that low levels of ROS could maintain biological functions and cellular homeostasis [21]. ROS levels fluctuating, and then influence cell signaling and the regulation of biological functions, including the regulation of cell signaling could be directly, or through structural modification of proteins, genes and transcription factors, regulate their functions. H2O2 could be reversible with oxidation of cysteine residues of the protein. This type of redox regulation targets phosphorylation, which could produce dephosphorylated proteins in cell signaling and lead to inactivation of kinases [22]. Of course, kinases, transcription factors and antioxidant proteins such as glutathione and PRX could also be directly regulated by H2O2 [23,24]. The active site of the protein tyrosine phosphatases are susceptible to redox regulation, because they have a common complex (cysteine-X5- arginine), which presents the state of the active site cysteine thiol acid (SH) of the (S-) form [25]. H2O2 signaling pathways including the regulation of cell proliferation and PI3K/AKT, HIF and MAPK/ERK signaling cascade and death-inducing pathways similar to JNK and P38 MAPK signaling pathway [26,27]. However, the reaction of superoxide with protein thiols enough to compete with the powerful SOD enzyme, SOD enzyme rapidly degraded by peroxide, allowing the cells to maintain homeostasis [28].

Cancer cells adapt ROS stress

ROS homeostasis is necessary for cell survival. Low ROS levels could effective regulate mitosis, cell survival, cell growth, cell proliferation and angiogenesis [20]. However, high ROS levels show toxic to cells, which could trigger a signal transduction pathway involving cell proliferation inhibition or cell death. It’s a very novel character that ROS levels of cancer cells are higher than normal cells, and it is often due to the self-needing of cancer cells [29]. P53 plays a crucial role in regulating antioxidant genes, thereby preventing damage to DNA and protein, and restore the redox balance [30]. P53 mutations or loss is about more than 50% of cancer, which is related with ROS stress [31]. In addition, the proto-oncogene increase production of ROS [32]. Downstream of RAS, PI3K/AKT/mTOR pathways are activated survival signal in most tumors [33]. AKT phosphorylation and inhibition of FOXO transcription factors, could lead to antioxidation blocked, leading to increased production of ROS.

ROS promote pro-tumorigenic cell signaling

The increased rate of ROS production in cancer cells play causal roles in acquiring hallmarks of cancer: sustained cell proliferation and mitogenic signaling, increased cell survival and disruption of cell death signaling, EMT, metastasis and angiogenesis [34-37].

ROS promote cancer cell proliferation

In various cancer cells, ROS could enhance cell proliferation by increasing pro-proliferative pathways like PI3K/AKT/mTOR and MAPK/ERK cascades. ROS-mediated cancer cell proliferation was observed in various cancer cells and could be prevented by the addition of ROS-scavenging antioxidants [38-40]. Recently, study showed that mitochondrial-derived ROS are required for cell growth in KRAS-driven lung cancer cells through the regulation of MAPK/ERKM signaling, and increasing ROS by disruption of mitochondrial function could reduce tumorigenesis.

ROS promote cancer cell survival

Aberrant ROS levels could activate PI3K/AKT pathway through inhibiting PTEN in various cancers [41,42]. Oncogenes including KRAS and AKT could also activate and stabilize the antioxidant regulator NRF2 [43], which plays a critical role in protecting against oxidative stress to promote cancer cell survival [44]. Furthermore, ROS could also promote cell survival through activating redox sensitive NF?B signaling in cancer cells [45,46]. NF?B upregulating Bcl-2, caspase inhibitors and antioxidant proteins leads to enhance cell survival in response to ROS accumulation [47,48].

ROS promote angiogenesis and metastasis

Tumor angiogenesis, invasion and metastasis are inter-related processes and the most severe inducements of tumorigenicity. Angiogenesis vascularizes solid tumors in order to provide enough nutrients and oxygen for tumor growth. The main angiogenic growth factor triggering formation of new blood vessels is VEGF, which is activated by HIFs [49]. As we know, hypoxia is an important character of cancers and could increase ROS production [50,51]. ROS play crucial roles in stabilization of HIFa, leading to VEGF activation [52]. And then induce angiogenesis and tumor progression, but the antioxidant of N-acetyl-cysteine (NAC) could prevent the effects [53]. NAC treatment also prevented HIF stabilization and diminished MYCmediated tumorigenesis [38]. Angiogenesis allows for cancer cell migration and metastasis. Metastasis requires extracellular remodeling and intracellular adaptations including EMT, reduced cell adhesion, increased migration and degradation of tissue barriers and ECM [54]. ROS have been shown to mediate metastasis in various cancer cells through regulating transcription factors and MAPK and PI3K/AKT pathways, HIFs and the EMT regulator Snail [55-57]. ROS could also modulate structural changes in tumor cells like the formation of invasive microdomains, which promote cell invasion and metastasis [58]. Furthermore, ROS promote the activation of MMPs that participate in the degradation of membranes and help detach primary tumor cells from ECM [59-61]. Taken together, these ROS-mediated events promote tumor progression.

ROS promote anti-tumorigenic cell signaling

While ROS are associated with the activation of pro-tumorigenic survival and growth pathways, oxidative stress could also lead to the induction of cell death and cell cycle arrest.

ROS promote cell death through ASK1/JNK/P38 MAPK pathways

Apoptosis signal-regulating kinase 1 (ASK1) acts as a redox sensor by mediating JNK and P38 MAPK, resulting in apoptosis upon excessive ROS [62]. P38 MAPK is also a tumor suppressor and ROS sensor in cancer cells [63]. So we could target P38 MAPK activation to induce cell death in cancer cells by identifying its high levels of ROS [64].

ROS regulate cell cycle arrest

High levels of ROS inhibit cell proliferation to prevent cell division by negatively regulating pro-proliferative kinases. For example, ROSCitation induced P38/JNK MAPK signaling could lead to down-regulate cyclins and the induction of cyclin-dependent kinase (CDK) inhibitors resulting in cell cycle arrest [65,66]. ROS could also induce cell cycle arrest directly through the oxidation of various cell cycle regulators. The H2O2-mediated oxidation and inactivation of the cell cycle phosphatase CDC25 is required for cell cycle progression from G2 to M phase [67].

In summary, ROS play causal roles in tumorigenesis but could also be toxic to the cell and could potentially induce cancer cell death, cell cycle arrest and inhibit cancer progression. Therefore cancer cells are dependent on maintaining high enough ROS levels that allow for protumorigenic cell signaling, but they still need to keep ROS at limit levels to avoid causing cell death.

Regulating ROS Levels to Treat Cancer

Decrease ROS levels

Due to the high levels of ROS in promoting cancer, various antioxidant promoters like NRF2 are considered as tumor suppressors [68,69]. Antioxidant treatments include supplementation of natural ROS scavengers, treatment with antioxidants and also other strategies that decrease oxidative stress like the disruption of the ROS-producing mitochondrial electron transport chain [2,70-72]. The most effective antioxidant treatment is the regulation of HIF1 levels because that ROS could regulate hypoxic activation of HIF [52]. Antioxidants like NAC and vitamin C could prevent HIF stabilization and diminish MYC-mediated tumorigenesis38. Other studies also reported antitumorigenic results of antioxidant treatments, including overexpression of SOD3, which inhibited breast cancer metastasis indicating the potential anti-tumorigenic effect of restoring extracellular superoxide scavenging capacity [73-75]. However, most clinical trials were failed to show beneficial effects of antioxidants on a variety of cancer [76,77]. Long-term study showed that vitamin E could significantly increase the risk of prostate cancer in healthy men [78]. Supplementation with β-carotene, vitamin A or E could increase the incidence of lung cancer [79,80]. A recent study demonstrated that using NAC and vitamin E could accelerate lung cancer progression in mice by reducing ROS [81]. Long-term antioxidant supplements treatment for cancer prevention did not provide evidence that they are beneficial in primary cancer prevention [82-86]. Antioxidants are not effective in targeting the locally ROS producer requiring for tumorigenic signaling due to poor efficacy of anticancer. Mitochondrial targeted antioxidants would be a way. These targeted antioxidants have shown efficacy in diminishing tumorigenic potential [87]. Recently, study showed that reduction in reactive oxygen speciescontributes to dihydromyricetin induced apoptosis in human hepatocellular carcinoma cells [88]. However, we should pay attention to antioxidant-based therapy, because these two problems could not be solved simultaneously. (1). ROS and ROS-dependent cell signaling are essential for normal cell function; (2). Antioxidants could interfere with chemotherapy depending on ROS-induced cytotoxicity [89]. Therefore, elevating ROS levels, either by increasing ROS generation or decreasing ROS scavenging potential, could be a way to selectively kill or arrest cancer cells without causing significant toxicity to normal cells.

Increase ROS levels

Many chemotherapeutic agents could induce cancer cell death or cell growth inhibition by increasing ROS production [90,91]. Chemotherapeutic drugs such as paclitaxel, vinca alkaloids and antifolates could cause cell death through releasing cytochrome c and also disrupt the mitochondrial electron transportchain which leads to ROS increase [92]. Cisplatin, carboplatin and doxorubicin could also significantly increase ROS, which is the basis of their anti-tumor ability [93,94]. Procarbazine is approved for the treatment of lymphoma and primary brain tumors by increasing ROS [95]. Monoclonal antibodies like rituximab, an anti-CD20 antibody that is used for non-Hodgkin’s lymphoma, which is targeted ROS production [96]. Arsenic trioxide (ATO) agents could effectively treat ROS-dependent leukemia [97]. Elesclomol (STA-4783) is another ROS-generating candidate in clinical trials for malignant melanoma. Either it was given as a single agent or combination with paclitaxel seems to be effective in phase II clinical trials. But antioxidants could suppressed Elesclomol’s curative effect [98]. The same problem restricts anti-tumor drugs which increase ROS is their toxic effect on normal cells.

Target antioxidants to decrease ROS scavenging capacity

Only thinking about increasing ROS seems not very ideal for cancer therapy, because cancer cells could adjust the ROS scavenging system to adapt ROS enhancing, which may induce cancers become resistant to chemotherapeutics. Various cancer cells increase antioxidant proteins, such as activation of NRF2 to maintain ROS levels that allow pro-tumorigenic signaling pathways to be activated without inducing cell death [99,100]. GSH increasing seems to play an active role in protecting cancer cells from cell death and also from ROS-inducing therapy strategies [101]. In fact, ROS scavenging pathway blocking could more efficiently trigger ROS-mediated cell death than solely increasing ROS in various cancer cells [102-104]. Phenethyl isothiocyanate (PEITC) conjugates with GSH, depleting the GSH pool and leading to oxidative stress induced cell death. Additionally, PEITC inhibits GPX activity, which leads to cell death in ovarian cancer cells and leukemia cells. The compound also prolonged survival in an ovarian cancer cell xenograft model [102]. L-Buthioninesulfoximine (BSO) targets de novo GSH synthesis as an inhibitor of glutamylcysteine synthetase (g-GCS) [105]. BSO depletes GSH and exhibits anticancer activity through apoptosis as a single agent and in combination with ATO in solid tumors and APL cells [106,107]. BSO also increased efficacy of cisplatin in preclinical studies [108]. Similar to BSO, Imexon has GSH-depleting induced ROS accumulating, and death-inducing potential as shown in a phase I study of non-Hodgkin’s lymphoma and melanoma patients [109,110]. Recently, using a cellbased molecular screen for pro-apoptotic effects in cancer identified the natural compound piperlongumine (PL), which could induce ROS accumulating by binding and modulating the antioxidant enzyme glutathione transferase and therefore change the ROS-stress response. PL induces apoptosis in several cancer cells but has little effect on normal cells103. Sulphasalazine (SASP) could decrease cysteine and GSH levels, and then increases ROS, lead to reduce pancreatic cancer cell growth and viability [111]. Another thiol-based antioxidant is thioredoxin (TRX). TRXs are up-regulated in various cancer cells and the TRX inhibitor PX-12 showed anti-tumor activity [112]. Furthermore, motexafin gadolinium is a TRXR inhibitor which showed anti-tumor activity in phase II and III clinical trials [113]. Auranofin is a TRX inhibitor and has been shown to have sensitivity effects in several cancers by causing ROS-mediated cell death [114]. Furthermore, the antioxidant SOD1 has been proved to be a potent target to selectively kill cancer cells. The SOD1 inhibitor methoxyestradiol (2-ME) increases superoxide and is currently in phase I and II clinical trials for prostate and metastatic breast cancer [115,116]. It also induces ROS-mediated apoptosis selectively in cancer cells [117]. Furthermore, recent studies identified the SOD1 inhibitor ATN-224 could cause cancer cell death and growth impairment in various cancer cells through ROS mediated mechanisms [117-119]. ATN-224-induced cancer cell death was mediated through ROS dependent activation of P38 MAPK [104]. Interestingly, extracellular SOD3 is differentially expressed in cancers. SOD3 mRNA was shown to be decreased in some clinical mammary adenocarcinoma samples compared to normal mammary tissues and research found that overexpression of SOD3 inhibited breast cancer metastasis [120]. On the other hand, SOD3 is shown to be amplified in some other cancers making it a potential target for therapy. NRF2 inhibition has the potential to suppress antioxidant systems and induce ROS-mediated cancer cell death. Effective NRF2 inhibitors have been developed and tested for anti-cancer effects.

Combination therapy

The difference between normal cells and cancer cells is the rate of ROS production and results in redox regulation. Combination therapy need to solve these two problems we would meet: (1) decrease the toxicity of chemotherapeutics to normal cells, (2) augment efficacy of anti-tumor. Upon the previous states, metabolic modulation is a suitable candidate for combination therapy with ROS-generators or ROS-scavenging inhibitors. We target the deference of ROS regulation between cancer cell and normal cell to design combination therapy. For example, regulating redox balance particularly through the production of reducing equivalents like NADH and NADPH, which are required for the function of various antioxidant proteins inducing cancer-killing ROS. The metabolic pathway predominantly involved in redox modulation is glutamine metabolism. It plays a central role in redox regulation and antioxidant response. Glutamine is the precursor of glutamate, which, as previously mentioned, is required for GSH synthesis and therefore antioxidant response. Glutamine metabolism has been shown to be required for cancer cell survival leading to the notion that some cancers are glutamine addicted [121]. Inhibition of glutaminase 1 (GLS1), the enzyme that converts glutamine to glutamate for entry into the TCA cycle, inhibits oncogenic transformation [122,123]. Furthermore, an approved agent for the treatment of leukemia has been shown to deplete glutamine levels and therefore GSH synthesis. The drug L-asparaginase was thought to function through its role in limiting asparagine levels, however recent data has shown that the anticancer effect of the drug could be attributed to its effect on glutamine levels [124]. Finally, the alternative glutamine pathway, mediated by the aspartate transaminase GOT1 has been shown to be required for KRAS-driven pancreatic ductal adenocarcinoma (PDAC) growth in vitro and in vivo [125]. GOT1 is a key enzyme in the aspartate-malate shuttle, producing pyruvate and increasing the NADPH/NADP+ ratio, which in turn maintains reduced GSH levels and therefore redox homeostasis. Indeed, GOT1 inhibition led to a decrease in the ratio of reduced-to-oxidized GSH, an increase in ROS levels and suppression of PDAC growth. The role of glutamine metabolism in NADPH production and GSH synthesis makes glutamine pathway inhibitors potentially suitable partners for pro-oxidants in anti-cancer therapy [126].

Conclusion

Studies have established a causal role of ROS in maintaining cellular homeostasis and in triggering cell signaling events. Cancer cells, compared to normal cells, have an increased rate of ROS production and have aberrant ROS regulation mechanisms to adapt their unique redox status. ROS play roles in promoting and maintaining tumorigenicity, and antioxidants could prevent or reduce tumorigenesis by targeting ROS regulation. Some clinical trials showed that dietary antioxidants could not affect cancer development. In fact, dietary antioxidants contribute to tumorigenesis by reducing the potentially death-inducing oxidative stress in cancer cells. This failure of therapeutic antioxidants might be due to the fact that they are unlikely to diminish the local ROS required for cancer survival. Thus, recent studies have focused on the aberrant redox status of cancer cells by exploiting pro-oxidant approaches to cancer therapy16. Chemotherapy and radiation could induce cancer cell death by increasing intracellular ROS. However, cancer cells are master of adaptation to oncogene mutations, reprogrammed metabolism, extreme microenvironments and nutrient starvation induce highly stressful conditions. Hence, cancer cells have the ability to develop resistance to therapeutics that exogenously raise ROS by increasing their antioxidant mechanisms. With that, tumor cells can maintain ROS levels that allow pro-tumorigenic signaling without inducing cell death, but at the same time they also rely on ROS detoxification. Redox scavenging and antioxidant system provide new strategy to selectively induce cancer cell death via oxidative stress, while sparing normal cells. In fact, studies have shown that disabling antioxidant mechanisms trigger ROS-mediated cell death in a variety of cancer cell types. Hence, the design of dual pro-oxidant therapies has the potential to be efficacious in selectively killing cancer cells. Combining ROSgenerating agents with ROS-scavenging inhibitors like GSH, TRX or SOD inhibitors could diminish the ability of cancer cells to adapt to sole chemotherapeutics.

Various antioxidant proteins and regulators are upregulated in cancers, but cancers have very complicated systems and strong adaption ability to cope with ectogenic toxicity. To targeting ROS signals for cancer therapy, we still need to study the mechanism of the endogenic ROS regulation in cancer cells comprehensively and systematicly.

Acknowledgements

This study was supported by grants from the Yangfan Plan of Talents Recruitment Grant, Guangdong, China (grant no. Yue Ren Cai Ban [2016].

References

  1. Cairns RA, Harris IS, Mak TW (2011) Regulation of cancer cell metabolism. Nat Rev Cancer 11: 85-95.
  2. Weinberg F, Hamanaka R, Wheaton WW, Weinberg S, Joseph J, et al. (2010) Mitochondrial metabolism and ROS generation are essential for Kras-mediated tumorigenicity. Proceedings of the National Academy of Sciences 107: 8788-8793.
  3. Glasauer A, Chandel NS (2014) Targeting antioxidants for cancer therapy. Biochem Pharmacol 92: 90-101.
  4. Ichijo H, Nishida E, Irie K, ten Dijke P, Saitoh M, et al. (1997) Induction of apoptosis by ASK1, a mammalian MAPKKK that activates SAPK/JNK and p38 signaling pathways. Science 275: 90-94.
  5. Moon DO, Kim MO, Choi YH, Hyun JW, Chang WY, et al. (2010) Butein induces G(2)/M phase arrest and apoptosis in human hepatoma cancer cells through ROS generation. Cancer Lett 288: 204-213.
  6. Alexandre J, Hu Y, Lu W, Pelicano H, Huang P (2007) Novel action of paclitaxel against cancer cells: bystander effect mediated by reactive oxygen species. Cancer research 67: 3512-3517.
  7. Schafer ZT, Grassian AR, Song L, Jiang Z, Gerhart-Hines Z, et al. (2009) Antioxidant and oncogene rescue of metabolic defects caused by loss of matrix attachment. Nature 461: 109-113.
  8. Tiligada E (2006) Chemotherapy: induction of stress responses. Endocr Relat Cancer 13 Suppl 1: S115-124.
  9. Tai DJ, Jin WS, Wu CS, Si HW, Cao XD, et al. (2012) Changes in intracellular redox status influence multidrug resistance in gastric adenocarcinoma cells. Experimental and therapeutic medicine 4: 291-296.
  10. Trachootham D, Zhou Y, Zhang H, Demizu Y, Chen Z, et al. (2006) Selective killing of oncogenically transformed cells through a ROS-mediated mechanism by beta-phenylethyl isothiocyanate. Cancer Cell 10: 241-252.
  11. Shaw AT, Winslow MM, Magendantz M, Ouyang C, Dowdle J, et al. (2011) Selective killing of K-ras mutant cancer cells by small molecule inducers of oxidative stress.&nbasp; Proc Natl Acad Sci U S A 108: 8773-8778.
  12. Raj L, Ide T, Gurkar AU, Foley M, Schenone M, et al. (2015) Corrigendum: Selective killing of cancer cells by a small molecule targeting the stress response to ROS. Nature 526: 596.
  13. Adhikary A, Mohanty S, Lahiry L, Hossain DM, Chakraborty S, et al. (2010) Theaflavins retard human breast cancer cell migration by inhibiting NF-kappaB via p53-ROS cross-talk. FEBS Lett 584: 7-14.
  14. Glasauer A, Sena LA, Diebold LP, Mazar AP, Chandel NS (2014) Targeting SOD1 reduces experimental non–small-cell lung cancer. J Clin Invest 124: 117-128.
  15. Gorrini C, Harris IS, Mak TW (2013) Modulation of oxidative stress as an anticancer strategy. Nat Rev Drug Discov 12: 931-947.
  16. McCord JM, Fridovich I (1969) Superoxide dismutase. An enzymic function for erythrocuprein (hemocuprein). J Biol Chem 244: 6049-6055.
  17. Berndt C, Lillig CH, Holmgren A (2007) Thiol-based mechanisms of the thioredoxin and glutaredoxin systems: implications for diseases in the cardiovascular system. Am J Physiol Heart Circ Physiol 292: H1227-36.
  18. Meister A (1983) Selective modification of glutathione metabolism. Science 220: 472-477.
  19. Weinberg SE, Chandel NS (2015) Targeting mitochondria metabolism for cancer therapy. Nat Chem Biol 11: 9-15.
  20. Finkel T (2012) Signal transduction by mitochondrial oxidants. J Biol Chem 287: 4434-4440.
  21. Meng TC, Fukada T, Tonks NK (2002) Reversible oxidation and inactivation of protein tyrosine phosphatases in vivo. Mol Cell 9: 387-399.
  22. Wani R, Qian J, Yin L, Bechtold E, King SB, et al. (2011) Isoform-specific regulation of Akt by PDGF-induced reactive oxygen species. Proc Natl Acad Sci U S A 108: 10550-10555.
  23. Kemble DJ, Sun G (2009) Direct and specific inactivation of protein tyrosine kinases in the Src and FGFR families by reversible cysteine oxidation. Proc Natl Acad Sci U S A 106: 5070-5075.
  24. Hamanaka RB, Chandel NS (2010) Mitochondrial reactive oxygen species regulate cellular signaling and dictate biological outcomes. Trends Biochem Sci 35: 505-513.
  25. Weinberg F, Chandel NS (2009) Reactive oxygen species-dependent signaling regulates cancer. Cell Mol Life Sci 66: 3663-3673.
  26. Kamata H, Honda S, Maeda S, Chang L, Hirata H, et al. (2005) Reactive oxygen species promote TNFalpha-induced death and sustained JNK activation by inhibiting MAP kinase phosphatases. Cell 120: 649-661.
  27. Winterbourn CC, Hampton MB (2008) Thiol chemistry and specificity in redox signaling. Free Radic Biol Med 45: 549-561.
  28. Cairns RA, Harris IS, Mak TW (2011) Regulation of cancer cell metabolism. Nat Rev Cancer 11: 85-95.
  29. Sablina AA, Budanov AV, Ilyinskaya GV, Agapova LS, Kravchenko JE, et al. (2005) The antioxidant function of the p53 tumor suppressor. Nat Med 11: 1306-1313.
  30. Attardi LD, Donehower LA (2005) Probing p53 biological functions through the use of genetically engineered mouse models. Mutation Research/Fundamental and Molecular Mechanisms of Mutagenesis 576: 4-21.
  31. Behrend L, Henderson G, Zwacka RM (2003) Reactive oxygen species in oncogenic transformation. Biochem Soc Trans 31: 1441-1444.
  32. Cantley LC (2002) The phosphoinositide 3-kinase pathway. Science 296: 1655-1657.
  33. Trachootham D, Zhang H, Zhang W, Feng L, Du M, et al. (2008) Effective elimination of fludarabine-resistant CLL cells by PEITC through a redox-mediated mechanism. Blood 112: 1912-1922.
  34. Clerkin JS, Naughton R, Quiney C, Cotter TG (2008) Mechanisms of ROS modulated cell survival during carcinogenesis. Cancer Lett 266: 30-36.
  35. Nishikawa M (2008) Reactive oxygen species in tumor metastasis. Cancer Lett 266: 53-59.
  36. Ushio-Fukai M, Nakamura Y (2008) Reactive oxygen species and angiogenesis: NADPH oxidase as target for cancer therapy. Cancer Lett 266: 37-52.
  37. Gao P, Zhang H, Dinavahi R, Li F, Xiang Y, et al. (2007) HIF-dependent antitumorigenic effect of antioxidants in vivo. Cancer Cell 12: 230-238.
  38. Policastro L, Molinari B, Larcher F, Blanco P, Podhajcer OL, et al. (2004) Imbalance of antioxidant enzymes in tumor cells and inhibition of proliferation and malignant features by scavenging hydrogen peroxide. Mol Carcinog 39: 103-113.
  39. Qin Y, Pan X, Tang TT, Zhou L, Gong XG (2011) Anti-proliferative effects of the novel squamosamide derivative (FLZ) on HepG2 human hepatoma cells by regulating the cell cycle-related proteins are associated with decreased Ca 2+/ROS levels. Chemico-biological interactions 193: 246-253.
  40. Chen Z, Trotman LC, Shaffer D, Lin HK, Dotan ZA, et al. (2005) Crucial role of p53-dependent cellular senescence in suppression of Pten-deficient tumorigenesis. Nature 436: 725-730.
  41. Nitsche C, Edderkaoui M, Moore RM, Eibl G, Kasahara N, et al. (2012) The phosphatase PHLPP1 regulates Akt2, promotes pancreatic cancer cell death, and inhibits tumor formation. Gastroenterology 142: 377-387.
  42. DeNicola GM, Karreth FA, Humpton TJ, Gopinathan A, Wei C, et al. (2011) Oncogene-induced Nrf2 transcription promotes ROS detoxification and tumorigenesis. Nature 475: 106-109.
  43. Kotlo KU, Yehiely F, Efimova E, Harasty H, Hesabi B, et al. (2003) Nrf2 is an inhibitor of the Fas pathway as identified by Achilles' Heel Method, a new function-based approach to gene identification in human cells. Oncogene 22: 797-806.
  44. Kobayashi M, Yamamoto M (2006) Nrf2-Keap1 regulation of cellular defense mechanisms against electrophiles and reactive oxygen species. Adv Enzyme Regul 46: 113-140.
  45. Rahman I, Gilmour PS, Jimenez LA, MacNee W (2002) Oxidative stress and TNF-alpha induce histone acetylation and NF-kappaB/AP-1 activation in alveolar epithelial cells: potential mechanism in gene transcription in lung inflammation. Mol Cell Biochem 235: 239-248.
  46. Karin M, Lin A (2002) NF-kappaB at the crossroads of life and death. Nat Immunol 3: 221-227.
  47. Chandel NS, Schumacker PT, Arch RH (2001) Reactive oxygen species are downstream products of TRAF-mediated signal transduction. J Biol Chem 276: 42728-42736.
  48. Forsythe JA, Jiang BH, Iyer NV, Agani F, Leung SW, et al. (1996) Activation of vascular endothelial growth factor gene transcription by hypoxia-inducible factor 1. Mol Cell Biol 16: 4604-4613.
  49. Jain RK (2005) Normalization of tumor vasculature: an emerging concept in antiangiogenic therapy. Science 307: 58-62.
  50. Bell EL, Emerling BM, Chandel NS (2005) Mitochondrial regulation of oxygen sensing. Mitochondrion 5: 322-332.
  51. Klimova T, Chandel NS (2008) Mitochondrial complex III regulates hypoxic activation of HIF. Cell Death Differ 15: 660-666.
  52. Ma Q, Cavallin LE, Yan B, Zhu S, Duran EM, et al. (2009) Antitumorigenesis of antioxidants in a transgenic Rac1 model of Kaposi's sarcoma. Proc Natl Acad Sci U S A 106: 8683-8688.
  53. Chiang AC, Massagué J (2008) Molecular basis of metastasis. N Engl J Med 359: 2814-2823.
  54. Ho BY, Wu YM, Chang KJ, Pan TM (2011) Dimerumic acid inhibits SW620 cell invasion by attenuating Hâ‚‚Oâ‚‚-mediated MMP-7 expression via JNK/C-Jun and ERK/C-Fos activation in an AP-1-dependent manner. Int J Biol Sci 7: 869-880.
  55. Radisky DC, Levy DD, Littlepage LE, Liu H, Nelson CM, et al. (2005) Rac1b and reactive oxygen species mediate MMP-3-induced EMT and genomic instability. Nature 436: 123-127.
  56. Gort EH, Groot AJ, van der Wall E, van Diest PJ, Vooijs MA (2008) Hypoxic regulation of metastasis via hypoxia-inducible factors. Curr Mol Med 8: 60-67.
  57. Courtneidge SA (2012) Cell migration and invasion in human disease: the Tks adaptor proteins. Biochem Soc Trans 40: 129-132.
  58. Binker MG, Binker-Cosen AA, Richards D, Oliver B, Cosen-Binker LI (2009) EGF promotes invasion by PANC-1 cells through Rac1/ROS-dependent secretion and activation of MMP-2. Biochem Biophys Res Commun 379: 445-450.
  59. Seo JM, Park S, Kim JH (2012) Leukotriene B4 receptor-2 promotes invasiveness and metastasis of ovarian cancer cells through signal transducer and activator of transcription 3 (STAT3)-dependent up-regulation of matrix metalloproteinase 2. J Biol Chem 287: 13840-13849.
  60. Kleiner DE Jr, Stetler-Stevenson WG (1993) Structural biochemistry and activation of matrix metalloproteases. Curr Opin Cell Biol 5: 891-897.
  61. Dolado I, Swat A, Ajenjo N, De Vita G, Cuadrado A, et al. (2007) p38alpha MAP kinase as a sensor of reactive oxygen species in tumorigenesis. Cancer Cell 11: 191-205.
  62. Han J, Sun P (2007) The pathways to tumor suppression via route p38. Trends Biochem Sci 32: 364-371.
  63. Ambrosino C, Nebreda AR (2001) Cell cycle regulation by p38 MAP kinases. Biol Cell 93: 47-51.
  64. Thornton TM, Rincon M (2009) Non-classical p38 map kinase functions: cell cycle checkpoints and survival. Int J Biol Sci 5: 44-51.
  65. Rudolph J (2005) Redox regulation of the Cdc25 phosphatases. Antioxid Redox Signal 7: 761-767.
  66. Siddiqui IA, Adhami VM, Saleem M, Mukhtar H (2006) Beneficial effects of tea and its polyphenols against prostate cancer. Mol Nutr Food Res 50: 130-143.
  67. Bianchini F, Vainio H (2003) Wine and resveratrol: mechanisms of cancer prevention? Eur J Cancer Prev 12: 417-425.
  68. Pantuck AJ, Zomorodian N, Belldegrun AS (2006) Phase-II Study of pomegranate juice for men with prostate cancer and increasing PSA. Curr Urol Rep 7: 7.
  69. Kim HS, Bowen P, Chen L, Duncan C, Ghosh L, et al. (2003) Effects of tomato sauce consumption on apoptotic cell death in prostate benign hyperplasia and carcinoma. Nutr Cancer 47: 40-47.
  70. Jian L, Xie LP, Lee AH, Binns CW (2004) Protective effect of green tea against prostate cancer: a case-control study in southeast China. Int J Cancer 108: 130-135.
  71. Blot WJ, Li JY, Taylor PR, Guo W, Dawsey S, et al. (1993) Nutrition intervention trials in Linxian, China: supplementation with specific vitamin/mineral combinations, cancer incidence, and disease-specific mortality in the general population. J Natl Cancer Inst 85: 1483-1492.
  72. Qiao YL, Dawsey SM, Kamangar F, Fan JH, Abnet CC, et al. (2009) Total and cancer mortality after supplementation with vitamins and minerals: follow-up of the Linxian General Population Nutrition Intervention Trial. J Natl Cancer Inst 101: 507-518.
  73. Teoh-Fitzgerald M, Fitzgerald M, Zhong W, Askeland R, Domann F (2014) Epigenetic reprogramming governs EcSOD expression during human mammary epithelial cell differentiation, tumorigenesis and metastasis. Oncogene 33: 358-368.
  74. Khaw KT, Bingham S, Welch A, Luben R, Wareham N, et al. (2001) Relation between plasma ascorbic acid and mortality in men and women in EPIC-Norfolk prospective study: a prospective population study. Lancet 357: 657-663.
  75. Burr M, Appleby P, Key T, Thorogood M (2001) Plasma ascorbic acid and risk of heart disease and cancer. Lancet 357: 2135-2136.
  76. Klein EA, Thompson IM Jr, Tangen CM, Crowley JJ, Lucia MS, et al. (2011) Vitamin E and the risk of prostate cancer: the Selenium and Vitamin E Cancer Prevention Trial (SELECT). JAMA 306: 1549-1556.
  77. [No authors listed] (1994) The effect of vitamin E and beta carotene on the incidence of lung cancer and other cancers in male smokers. The Alpha-Tocopherol, Beta Carotene Cancer Prevention Study Group. N Engl J Med 330: 1029-1035.
  78. Omenn GS, Goodman GE, Thornquist MD, Balmes J, Cullen MR, et al. (1996) Effects of a combination of beta carotene and vitamin A on lung cancer and cardiovascular disease. N Engl J Med 334: 1150-1155.
  79. Sayin VI, Ibrahim MX, Larsson E, Nilsson JA, Lindahl P, et al. (2014) Antioxidants accelerate lung cancer progression in mice. Sci Transl Med 6: 221ra15.
  80. Hennekens CH, Buring JE, Manson JE, Stampfer M, Rosner B, et al. (1996) Lack of effect of long-term supplementation with beta carotene on the incidence of malignant neoplasms and cardiovascular disease. N Engl J Med 334: 1145-1149.
  81. Lonn E, Bosch J, Yusuf S, Sheridan P, Pogue J, et al. (2005) Effects of long-term vitamin E supplementation on cardiovascular events and cancer: a randomized controlled trial. JAMA 293: 1338-1347.
  82. Gaziano JM, Glynn RJ, Christen WG, Kurth T, Belanger C, MacFadyen J, et al. (2009) Vitamins E and C in the Prevention of Prostate and Total Cancer in Men The Physicians' Health Study II Randomized Controlled Trial. Jama-J Am Med Assoc 301: 52-62.
  83. Nazarewicz RR, Dikalova A, Bikineyeva A, Ivanov S, Kirilyuk IA, et al. (2013) Does Scavenging of Mitochondrial Superoxide Attenuate Cancer Prosurvival Signaling Pathways? Antioxid Redox Sign 19: 344-349.
  84. Liu B, Tan X, Liang J, Wu S, Liu J, et al. (2014) A reduction in reactive oxygen species contributes to dihydromyricetin-induced apoptosis in human hepatocellular carcinoma cells. Sci Rep 4: 7041.
  85. Conklin KA (2000) Dietary antioxidants during cancer chemotherapy: Impact on chemotherapeutic effectiveness and development of side effects. Nutrition and Cancer-an International Journal 37: 1-18.
  86. Kong Q, Lillehei KO (1998) Antioxidant inhibitors for cancer therapy. Med Hypotheses 51: 405-409.
  87. Wondrak GT (2009) Redox-directed cancer therapeutics: molecular mechanisms and opportunities. Antioxid Redox Signal 11: 3013-3069.
  88. Kaufmann SH, Earnshaw WC (2000) Induction of apoptosis by cancer chemotherapy. Exp Cell Res 256: 42-49.
  89. Conklin KA (2004) Chemotherapy-associated oxidative stress: impact on chemotherapeutic effectiveness. Integr Cancer Ther 3: 294-300.
  90. Kotamraju S, Chitambar CR, Kalivendi SV, Joseph J, Kalyanaraman B (2002) Transferrin receptor-dependent iron uptake is responsible for doxorubicin-mediated apoptosis in endothelial cells: role of oxidant-induced iron signaling in apoptosis. J Biol Chem 277: 17179-17187.
  91. Berneis K, Bollag W, Kofler M, Lüthy H (2004) The enhancement of the after effect of ionizing radiation by a cytotoxic methylhydrazine derivative. 1966. Eur J Cancer 40: 1928-1933.
  92. Renschler MF (2004) The emerging role of reactive oxygen species in cancer therapy. Eur J Cancer 40: 1934-1940.
  93. Chou WC, Jie C, Kenedy AA, Jones RJ, Trush MA, et al. (2004) Role of NADPH oxidase in arsenic-induced reactive oxygen species formation and cytotoxicity in myeloid leukemia cells. Proc Natl Acad Sci U S A 101: 4578-4583.
  94. Kirshner JR, He S, Balasubramanyam V, Kepros J, Yang CY, et al. (2008) Elesclomol induces cancer cell apoptosis through oxidative stress. Mol Cancer Ther 7: 2319-2327.
  95. DeNicola GM, Karreth FA, Humpton TJ, Gopinathan A, Wei C, et al. (2011) Oncogene-induced Nrf2 transcription promotes ROS detoxification and tumorigenesis. Nature 475: 106-109.
  96. Hayes JD, McMahon M, Chowdhry S, Dinkova-Kostova AT (2010) Cancer Chemoprevention Mechanisms Mediated Through the Keap1-Nrf2 Pathway. Antioxid Redox Sign 13: 1713-1748.
  97. Raj L, Ide T, Gurkar AU, Foley M, Schenone M, et al. (2011) Selective killing of cancer cells by a small molecule targeting the stress response to ROS 526: 596.
  98. Glasauer A, Sena LA, Diebold LP, Mazar AP, Chandel NS (2014) Targeting SOD1 reduces experimental non–small-cell lung cancer. J Clin Invest 124: 117-128.
  99. Griffith OW, Meister A (1979) Potent and Specific-Inhibition of Glutathione Synthesis by Buthionine Sulfoximine (S-Normal-Butyl Homocysteine Sulfoximine). Journal of Biological Chemistry 254: 7558-7560.
  100. Davison K, Cote S, Mader S, Miller WH (2003) Glutathione depletion overcomes resistance to arsenic trioxide in arsenic-resistant cell lines. Leukemia 17: 931-940.
  101. Maeda H, Hori S, Ohizumi H, Segawa T, Kakehi Y, et al. (2004) Effective treatment of advanced solid tumors by the combination of arsenic trioxide and L-buthionine-sulfoximine. Cell Death Differ 11: 737-746.
  102. Engel RH, Evens AM (2006) Oxidative stress and apoptosis: a new treatment paradigm in cancer. Front Biosci 11: 300-312.
  103. Dragovich T, Gordon M, Mendelson D, Wong L, Modiano M, et al. (2007) Phase I trial of imexon in patients with advanced malignancy. J Clin Oncol 25: 1779-1784.
  104. Weber JS, Samlowski WE, Gonzalez R, Ribas A, Stephenson J, et al. (2010) A phase 1-2 study of imexon plus dacarbazine in patients with unresectable metastatic melanoma. Cancer 116: 3683-3691.
  105. Guan J, Lo M, Dockery P, Mahon S, Karp CM, et al. (2009) The xc- cystine/glutamate antiporter as a potential therapeutic target for small-cell lung cancer: use of sulfasalazine. Cancer Chemother Pharmacol 64: 463-472.
  106. Welsh SJ, Williams RR, Birmingham A, Newman DJ, Kirkpatrick DL, et al. (2003) The thioredoxin redox inhibitors 1-methylpropyl 2-imidazolyl disulfide and pleurotin inhibit hypoxia-induced factor 1 alpha and vascular endothelial growth factor formation. Molecular Cancer Therapeutics 2: 235-243.
  107. Mehta MP, Shapiro WR, Phan SC, Gervais R, Carrie C, et al. (2009) Motexafin Gadolinium Combined with Prompt Whole Brain Radiotherapy Prolongs Time to Neurologic Progression in Non-Small-Cell Lung Cancer Patients with Brain Metastases: Results of a Phase Iii Trial. Int J Radiat Oncol 73: 1069-1076.
  108. Marzano C, Gandin V, Folda A, Scutari G, Bindoli A, et al. (2007) Inhibition of thioredoxin reductase by auranofin induces apoptosis in cisplatin-resistant human ovarian cancer cells. Free Radical Bio Med 42: 872-881.
  109. James J, Murry DJ, Treston AM, Storniolo AM, Sledge GW, et al. (2007) Phase I safety, pharmacokinetic and pharmacodynamic studies of 2-methoxyestradiol alone or in combination with docetaxel in patients with locally recurrent or metastatic breast cancer. Invest New Drugs 25: 41-48.
  110. Sweeney C, Liu G, Yiannoutsos C, Kolesar J, Horvath D, et al. (2005) A phase II multicenter, randomized, double-blind, safety trial assessing the pharmacokinetics, pharmacodynamics, and efficacy of oral 2-methoxyestradiol capsules in hormone-refractory prostate cancer. Clinical Cancer Research 11: 6625-6633.
  111. Lee K, Briehl MM, Mazar AP, Batinic-Haberle I, Reboucas JS, et al. (2013) The copper chelator ATN-224 induces peroxynitrite-dependent cell death in hematological malignancies. Free Radical Bio Med 60:157-167.
  112. Donate F, Juarez JC, Burnett ME, Manuia MM, Guan X, et al. (2008) Identification of biomarkers for the antiangiogenic and antitumour activity of the superoxide dismutase 1 (SOD1) inhibitor tetrathiomolybdate (ATN-224). Brit J Cancer 98: 776-783.
  113. Juarez JC, Manuia M, Burnett ME, Betancourt O, Boivin B, et al. (2008) Superoxide dismutase 1 (SOD1) is essential for H2O2-mediated oxidation and inactivation of phosphatases in growth factor signaling. Proc Natl Acad Sci U S A 105: 7147-7152.
  114. Teoh-Fitzgerald ML, Fitzgerald MP, Zhong W, Askeland RW, Domann FE (2014) Epigenetic reprogramming governs EcSOD expression during human mammary epithelial cell differentiation, tumorigenesis and metastasis. Oncogene33: 358-368.
  115. Galluzzi L, Kepp O, Vander Heiden MG, Kroemer G (2013) Metabolic targets for cancer therapy. Nat Rev Drug Discov 12: 829-846.
  116. Vander Heiden MG, Cantley LC, Thompson CB (2009) Understanding the Warburg effect: the metabolic requirements of cell proliferation. Science 324: 1029-1033.
  117. Reinert RB, Oberle LM, Wek SA, Bunpo P, Wang XP, et al. (2006) Role of glutamine depletion in directing tissue-specific nutrient stress responses to L-asparaginase. J Biol Chem 281: 31222-31233.
  118. Son J, Lyssiotis CA, Ying H, Wang X, Hua S, et al. (2013) Glutamine supports pancreatic cancer growth through a KRAS-regulated metabolic pathway. Nature 496: 101-105.
  119. Sena LA, Chandel NS (2012) Physiological roles of mitochondrial reactive oxygen species. Mol Cell 48: 158-167.
  120. Bandy B, Davison AJ (1990) Mitochondrial mutations may increase oxidative stress: implications for carcinogenesis and aging? Free Radic Biol Med 8: 523-539.
  121. Trachootham D, Alexandre J, Huang P (2009) Targeting cancer cells by ROS-mediated mechanisms: a radical therapeutic approach? Nat Rev Drug Discov 8: 579-591.
  122. Hayes JD, McMahon M (2009) NRF2 and KEAP1 mutations: permanent activation of an adaptive response in cancer. Trends Biochem Sci 34: 176-188.
  123. Mitsuishi Y, Taguchi K, Kawatani Y, Shibata T, Nukiwa T, et al. (2012) Nrf2 redirects glucose and glutamine into anabolic pathways in metabolic reprogramming. Cancer Cell 22: 66-79.
  124. Lenehan PF, Gutiérrez PL, Wagner JL, Milak N, Fisher GR, et al. (1995) Resistance to oxidants associated with elevated catalase activity in HL-60 leukemia cells that overexpress multidrug-resistance protein does not contribute to the resistance to daunorubicin manifested by these cells. Cancer chemotherapy and pharmacology 35: 377-386.
  125. Ren D, Villeneuve NF, Jiang T, Wu T, Lau A, et al. (2011) Brusatol enhances the efficacy of chemotherapy by inhibiting the Nrf2-mediated defense mechanism. Proc Natl Acad Sci U S A 108: 1433-1438.
Citation: Dong J, Liu, Zhu R (2016) Targeting ROS for Cancer Therapy. Chemo Open Access 5: 199.

Copyright: © 2016 Dong J, et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.
Top