Skip to main content

Advertisement

Log in

Metastasis: cancer cell’s escape from oxidative stress

  • NON-THEMATIC REVIEW
  • Published:
Cancer and Metastasis Reviews Aims and scope Submit manuscript

Abstract

According to a “canonical” view, reactive oxygen species (ROS) positively contribute, in different ways, to carcinogenesis and to malignant progression of tumor cells: they drive genomic damage and genetic instability, transduce, as signaling intermediates, mitogenic and survival inputs by growth factor receptors and adhesion molecules, promote cell motility and shape the tumor microenvironment by inducing inflammation/repair and angiogenesis. Chemopreventive and tumor-inhibitory effects of endogenous, diet-derived or supplemented antioxidants largely support this notion. However, emerging lines of evidence indicates that tumor cells also need to defend themselves from oxidative damage in order to survive and successfully spread at distance. This “heresy” has recently received important impulse from studies on the role of antioxidant capacity in cancer stem cells self-renewal and resistance to therapy; additionally, the transforming activity of some oncogenes has been unexpectedly linked to their capacity to maintain elevated intracellular levels of reduced glutathione (GSH), the principal redox buffer. These studies underline the importance of cellular antioxidant capacity in metastasis, as the result of a complex cell program involving enhanced motility and a profound change in energy metabolism. The glycolytic switch (Warburg effect) observed in malignant tissues is triggered by mitochondrial oxidative damage and/or activation of redox-sensitive transcription factors, and results in an increase of cell resistance to oxidants. On the other hand, cytoskeleton rearrangement underlying cell motile and tumor-aggressive behavior use ROS as intermediates and are therefore facilitated by oxidative stress. Along this line of speculation, we suggest that metastasis represents an integrated strategy for cancer cells to avoid oxidative damage and escape excess ROS in the primary tumor site, explaning why redox signaling pathways are often up-regulated in malignancy and metastasis.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9

Similar content being viewed by others

References

  1. Steeg, P. S. (2006). Tumor metastasis: mechanistic insights and clinical challenges. Nature Medicine, 12, 895–904.

    Article  PubMed  CAS  Google Scholar 

  2. Nguyen, D. X., & Massague, J. (2007). Genetic determinants of cancer metastasis. Nature Reviews Genetics, 8, 341–352.

    Article  PubMed  CAS  Google Scholar 

  3. Chiang, A. C., & Massague, J. (2008). Molecular basis of metastasis. New England Journal of Medicine, 359, 2814–2823.

    Article  PubMed  CAS  Google Scholar 

  4. Fidler, I. J. (2003). The pathogenesis of cancer metastasis: the ‘seed and soil’ hypothesis revisited. Nature Reviews Cancer, 3, 453–458.

    Article  PubMed  CAS  Google Scholar 

  5. Klein, C. A. (2009). Parallel progression of primary tumours and metastases. Nature Reviews Cancer, 9, 302–312.

    Article  PubMed  CAS  Google Scholar 

  6. Mani, S. A., Guo, W., Liao, M. J., Eaton, E. N., Ayyanan, A., Zhou, A. Y., et al. (2008). The epithelial–mesenchymal transition generates cells with properties of stem cells. Cell, 133, 704–715.

    Article  PubMed  CAS  Google Scholar 

  7. Barnhart, B. C., & Simon, M. C. (2007). Metastasis and stem cell pathways. Cancer and Metastasis Reviews, 26, 261–271.

    Article  PubMed  CAS  Google Scholar 

  8. Bertout, J. A., Patel, S. A., & Simon, M. C. (2008). The impact of O2 availability on human cancer. Nature Reviews Cancer, 8, 967–975.

    Article  PubMed  CAS  Google Scholar 

  9. Mantovani, A. (2009). Cancer: inflaming metastasis. Nature, 457, 36–37.

    Article  PubMed  CAS  Google Scholar 

  10. Comoglio, P. M., & Trusolino, L. (2002). Invasive growth: from development to metastasis. Journal of Clinical Investigation, 109, 857–862.

    PubMed  CAS  Google Scholar 

  11. Kalluri, R., & Weinberg, R. A. (2009). The basics of epithelial–mesenchymal transition. Journal of Clinical Investigation, 119, 1420–1428.

    Article  PubMed  CAS  Google Scholar 

  12. Etienne-Manneville, S., & Hall, A. (2002). Rho GTPases in cell biology. Nature, 420, 629–635.

    Article  PubMed  CAS  Google Scholar 

  13. Hall, A. (2005). Rho GTPases and the control of cell behaviour. Biochemical Society Transactions, 33, 891–895.

    Article  PubMed  CAS  Google Scholar 

  14. Brakebusch, C., Bouvard, D., Stanchi, F., Sakai, T., & Fassler, R. (2002). Integrins in invasive growth. Journal of Clinical Investigation, 109, 999–1006.

    PubMed  CAS  Google Scholar 

  15. Mitra, S. K., & Schlaepfer, D. D. (2006). Integrin-regulated FAK-Src signaling in normal and cancer cells. Current Opinion in Cell Biology, 18, 516–523.

    Article  PubMed  CAS  Google Scholar 

  16. Raftopoulou, M., Etienne-Manneville, S., Self, A., Nicholls, S., & Hall, A. (2004). Regulation of cell migration by the C2 domain of the tumor suppressor PTEN. Science, 303, 1179–1181.

    Article  PubMed  CAS  Google Scholar 

  17. Yu, D. H., Qu, C. K., Henegariu, O., Lu, X., & Feng, G. S. (1998). Protein-tyrosine phosphatase Shp-2 regulates cell spreading, migration, and focal adhesion. Journal of Biological Chemistry, 273, 21125–21131.

    Article  PubMed  CAS  Google Scholar 

  18. Friedl, P. (2004). Prespecification and plasticity: shifting mechanisms of cell migration. Current Opinion in Cell Biology, 16, 14–23.

    Article  PubMed  CAS  Google Scholar 

  19. Sanz-Moreno, V., Gadea, G., Ahn, J., Paterson, H., Marra, P., Pinner, S., et al. (2008). Rac activation and inactivation control plasticity of tumor cell movement. Cell, 135, 510–523.

    Article  PubMed  CAS  Google Scholar 

  20. Habets, G. G., Scholtes, E. H., Zuydgeest, D., van der Kammen, R. A., Stam, J. C., Berns, A., et al. (1994). Identification of an invasion-inducing gene, Tiam-1, that encodes a protein with homology to GDP-GTP exchangers for Rho-like proteins. Cell, 77, 537–549.

    Article  PubMed  CAS  Google Scholar 

  21. Clark, E. A., Golub, T. R., Lander, E. S., & Hynes, R. O. (2000). Genomic analysis of metastasis reveals an essential role for RhoC. Nature, 406, 532–535.

    Article  PubMed  CAS  Google Scholar 

  22. Murakami, M., Meneses, P. I., Knight, J. S., Lan, K., Kaul, R., Verma, S. C., et al. (2008). Nm23-H1 modulates the activity of the guanine exchange factor Dbl-1. International Journal of Cancer, 123, 500–510.

    Article  CAS  Google Scholar 

  23. Birchmeier, C., Birchmeier, W., Gherardi, E., & Vande Woude, G. F. (2003). Met, metastasis, motility and more. Nature Reviews Molecular Cell Biology, 4, 915–925.

    Article  PubMed  CAS  Google Scholar 

  24. Douma, S., Van, L. T., Zevenhoven, J., Meuwissen, R., Van, G. E., & Peeper, D. S. (2004). Suppression of anoikis and induction of metastasis by the neurotrophic receptor TrkB. Nature, 430, 1034–1039.

    Article  PubMed  CAS  Google Scholar 

  25. Geiger, T. R., & Peeper, D. S. (2009). Metastasis mechanisms. Biochimica et Biophysica Acta, 1796, 293–308.

    PubMed  CAS  Google Scholar 

  26. Wei, L., Yang, Y., Zhang, X., & Yu, Q. (2004). Altered regulation of Src upon cell detachment protects human lung adenocarcinoma cells from anoikis. Oncogene, 23, 9052–9061.

    Article  PubMed  CAS  Google Scholar 

  27. Aguirre-Ghiso, J. A. (2007). Models, mechanisms and clinical evidence for cancer dormancy. Nature Reviews Cancer, 7, 834–846.

    Article  PubMed  CAS  Google Scholar 

  28. Kim, S., Takahashi, H., Lin, W. W., Descargues, P., Grivennikov, S., Kim, Y., et al. (2009). Carcinoma-produced factors activate myeloid cells through TLR2 to stimulate metastasis. Nature, 457, 102–106.

    Article  PubMed  CAS  Google Scholar 

  29. Kaplan, R. N., Riba, R. D., Zacharoulis, S., Bramley, A. H., Vincent, L., Costa, C., et al. (2005). VEGFR1-positive haematopoietic bone marrow progenitors initiate the pre-metastatic niche. Nature, 438, 820–827.

    Article  PubMed  CAS  Google Scholar 

  30. Bates, R. C., & Mercurio, A. M. (2003). Tumor necrosis factor-alpha stimulates the epithelial-to-mesenchymal transition of human colonic organoids. Molecular Biology of the Cell, 14, 1790–1800.

    Article  PubMed  CAS  Google Scholar 

  31. Luo, J. L., Tan, W., Ricono, J. M., Korchynskyi, O., Zhang, M., Gonias, S. L., et al. (2007). Nuclear cytokine-activated IKKalpha controls prostate cancer metastasis by repressing Maspin. Nature, 446, 690–694.

    Article  PubMed  CAS  Google Scholar 

  32. Kaelin, W. G., Jr., & Ratcliffe, P. J. (2008). Oxygen sensing by metazoans: the central role of the HIF hydroxylase pathway. Molecular Cell, 30, 393–402.

    Article  PubMed  CAS  Google Scholar 

  33. Pouyssegur, J., Dayan, F., & Mazure, N. M. (2006). Hypoxia signalling in cancer and approaches to enforce tumour regression. Nature, 441, 437–443.

    Article  PubMed  CAS  Google Scholar 

  34. Kim, J. W., Tchernyshyov, I., Semenza, G. L., & Dang, C. V. (2006). HIF-1-mediated expression of pyruvate dehydrogenase kinase: a metabolic switch required for cellular adaptation to hypoxia. Cell Metab, 3, 177–185.

    Article  PubMed  CAS  Google Scholar 

  35. Fukuda, R., Zhang, H., Kim, J. W., Shimoda, L., Dang, C. V., & Semenza, G. L. (2007). HIF-1 regulates cytochrome oxidase subunits to optimize efficiency of respiration in hypoxic cells. Cell, 129, 111–122.

    Article  PubMed  CAS  Google Scholar 

  36. Dewhirst, M. W., Cao, Y., & Moeller, B. (2008). Cycling hypoxia and free radicals regulate angiogenesis and radiotherapy response. Nature Reviews Cancer, 8, 425–437.

    Article  PubMed  CAS  Google Scholar 

  37. Pennacchietti, S., Michieli, P., Galluzzo, M., Mazzone, M., Giordano, S., & Comoglio, P. M. (2003). Hypoxia promotes invasive growth by transcriptional activation of the met protooncogene. Cancer Cell, 3, 347–361.

    Article  PubMed  Google Scholar 

  38. Erler, J. T., Bennewith, K. L., Nicolau, M., Dornhofer, N., Kong, C., Le, Q. T., et al. (2006). Lysyl oxidase is essential for hypoxia-induced metastasis. Nature, 440, 1222–1226.

    Article  PubMed  CAS  Google Scholar 

  39. Denko, N. C. (2008). Hypoxia, HIF1 and glucose metabolism in the solid tumour. Nature Reviews Cancer., 8, 705–13.

    Article  PubMed  CAS  Google Scholar 

  40. Vander Heiden, M. G., Cantley, L. C., & Thompson, C. B. (2009). Understanding the Warburg effect: the metabolic requirements of cell proliferation. Science, 324, 1029–1033.

    Article  PubMed  CAS  Google Scholar 

  41. Hsu, P. P., & Sabatini, D. M. (2008). Cancer cell metabolism: Warburg and beyond. Cell, 134, 703–707.

    Article  PubMed  CAS  Google Scholar 

  42. Gatenby, R. A., & Gawlinski, E. T. (2001). Mathematical models of tumour invasion mediated by transformation-induced alteration of microenvironmental pH. Novartis Foundation Symposium, 240, 85–96.

    Article  PubMed  CAS  Google Scholar 

  43. Rofstad, E. K., Mathiesen, B., Kindem, K., & Galappathi, K. (2006). Acidic extracellular pH promotes experimental metastasis of human melanoma cells in athymic nude mice. Cancer Research, 66, 6699–6707.

    Article  PubMed  CAS  Google Scholar 

  44. Clarke, M. F., Dick, J. E., Dirks, P. B., Eaves, C. J., Jamieson, C. H., Jones, D. L., et al. (2006). Cancer stem cells—perspectives on current status and future directions: AACR Workshop on cancer stem cells. Cancer Research, 66, 9339–9344.

    Article  PubMed  CAS  Google Scholar 

  45. Cipolleschi, M. G., Rovida, E., Ivanovic, Z., Praloran, V., Olivotto, M., & Dello, S. P. (2000). The expansion of murine bone marrow cells preincubated in hypoxia as an in vitro indicator of their marrow-repopulating ability. Leukemia, 14, 735–739.

    Article  PubMed  CAS  Google Scholar 

  46. Simon, M. C., & Keith, B. (2008). The role of oxygen availability in embryonic development and stem cell function. Nature Reviews Molecular Cell Biology, 9, 285–296.

    Article  PubMed  CAS  Google Scholar 

  47. Ezashi, T., Das, P., & Roberts, R. M. (2005). Low O2 tensions and the prevention of differentiation of hES cells. Proceedings of the National Academy of Sciences of the United States of America, 102, 4783–4788.

    Article  PubMed  CAS  Google Scholar 

  48. Kondoh, H., Lleonart, M. E., Gil, J., Wang, J., Degan, P., Peters, G., et al. (2005). Glycolytic enzymes can modulate cellular life span. Cancer Research, 65, 177–185.

    PubMed  CAS  Google Scholar 

  49. Giuntoli, S., Rovida, E., Gozzini, A., Barbetti, V., Cipolleschi, M. G., Olivotto, M., et al. (2007). Severe hypoxia defines heterogeneity and selects highly immature progenitors within clonal erythroleukemia cells. Stem Cells, 25, 1119–1125.

    Article  PubMed  CAS  Google Scholar 

  50. Das, B., Tsuchida, R., Malkin, D., Koren, G., Baruchel, S., & Yeger, H. (2008). Hypoxia enhances tumor stemness by increasing the invasive and tumorigenic side population fraction. Stem Cells, 26, 1818–1830.

    Article  PubMed  Google Scholar 

  51. Bjerkvig, R., Johansson, M., Miletic, H., & Niclou, S. P. (2009). Cancer stem cells and angiogenesis. Seminars in Cancer Biology, 19, 279–284.

    Article  PubMed  CAS  Google Scholar 

  52. Janssen-Heininger, Y. M., Mossman, B. T., Heintz, N. H., Forman, H. J., Kalyanaraman, B., Finkel, T., et al. (2008). Redox-based regulation of signal transduction: principles, pitfalls, and promises. Free Radical Biology and Medicine, 45, 1–17.

    Article  PubMed  CAS  Google Scholar 

  53. Murphy, M. P. (2009). How mitochondria produce reactive oxygen species. Biochemical Journal, 417, 1–13.

    Article  PubMed  CAS  Google Scholar 

  54. Paget, M. S., & Buttner, M. J. (2003). Thiol-based regulatory switches. Annual Review of Genetics, 37, 91–121.

    Article  PubMed  CAS  Google Scholar 

  55. Lambeth, J. D. (2004). NOX enzymes and the biology of reactive oxygen. Nature Reviews Cancer, 4, 181–189.

    CAS  Google Scholar 

  56. Sundaresan, M., Yu, Z. X., Ferrans, V. J., Irani, K., & Finkel, T. (1995). Requirement for generation of H2O2 for platelet-derived growth factor signal transduction. Science, 270, 296–299.

    Article  PubMed  CAS  Google Scholar 

  57. Baumer, A. T., Ten, F. H., Sauer, H., Wartenberg, M., Kappert, K., Schnabel, P., et al. (2008). Phosphatidylinositol 3-kinase-dependent membrane recruitment of Rac-1 and p47phox is critical for alpha-platelet-derived growth factor receptor-induced production of reactive oxygen species. Journal of Biological Chemistry, 283, 7864–7876.

    Article  PubMed  CAS  Google Scholar 

  58. Lo, Y. Y., & Cruz, T. F. (1995). Involvement of reactive oxygen species in cytokine and growth factor induction of c-fos expression in chondrocytes. Journal of Biological Chemistry, 270, 11727–11730.

    Article  PubMed  CAS  Google Scholar 

  59. Ushio-Fukai, M., Zafari, A. M., Fukui, T., Ishizaka, N., & Griendling, K. K. (1996). p22phox is a critical component of the superoxide-generating NADH/NADPH oxidase system and regulates angiotensin II-induced hypertrophy in vascular smooth muscle cells. Journal of Biological Chemistry, 271, 23317–23321.

    Article  PubMed  CAS  Google Scholar 

  60. Chiarugi, P., Pani, G., Giannoni, E., Taddei, L., Colavitti, R., Raugei, G., et al. (2003). Reactive oxygen species as essential mediators of cell adhesion: the oxidative inhibition of a FAK tyrosine phosphatase is required for cell adhesion. Journal of Biological Chemistry, 161, 933–944.

    CAS  Google Scholar 

  61. Inoue, M., Sato, E. F., Nishikawa, M., Park, A. M., Kira, Y., Imada, I., et al. (2003). Mitochondrial generation of reactive oxygen species and its role in aerobic life. Current Medicinal Chemistry, 10, 2495–2505.

    Article  PubMed  CAS  Google Scholar 

  62. Nemoto, S., Takeda, K., Yu, Z. X., Ferrans, V. J., & Finkel, T. (2000). Role for mitochondrial oxidants as regulators of cellular metabolism. Molecular and Cellular Biology, 20, 7311–7318.

    Article  PubMed  CAS  Google Scholar 

  63. Guzy, R. D., Hoyos, B., Robin, E., Chen, H., Liu, L., Mansfield, K. D., et al. (2005). Mitochondrial complex III is required for hypoxia-induced ROS production and cellular oxygen sensing. Cell Metab, 1, 401–408.

    Article  PubMed  CAS  Google Scholar 

  64. Cai, J., & Jones, D. P. (1998). Superoxide in apoptosis. Mitochondrial generation triggered by cytochrome c loss. Journal of Biological Chemistry, 273, 11401–11404.

    Article  PubMed  CAS  Google Scholar 

  65. Giorgio, M., Migliaccio, E., Orsini, F., Paolucci, D., Moroni, M., Contursi, C., et al. (2005). Electron transfer between cytochrome c and p66Shc generates reactive oxygen species that trigger mitochondrial apoptosis. Cell, 122, 221–233.

    Article  PubMed  CAS  Google Scholar 

  66. Pinton, P., Rimessi, A., Marchi, S., Orsini, F., Migliaccio, E., Giorgio, M., et al. (2007). Protein kinase C beta and prolyl isomerase 1 regulate mitochondrial effects of the life-span determinant p66Shc. Science, 315, 659–663.

    Article  PubMed  CAS  Google Scholar 

  67. Pani, G., Koch, O. R., & Galeotti, T. (2009). The p53-p66shc-Manganese Superoxide Dismutase (MnSOD) network: a mitochondrial intrigue to generate reactive oxygen species. International Journal of Biochemistry and Cell Biology, 41, 1002–1005.

    Article  PubMed  CAS  Google Scholar 

  68. Rhee, S. G., Yang, K. S., Kang, S. W., Woo, H. A., & Chang, T. S. (2005). Controlled elimination of intracellular H(2)O(2): regulation of peroxiredoxin, catalase, and glutathione peroxidase via post-translational modification. Antioxidants Redox Signaling, 7, 619–626.

    Article  PubMed  CAS  Google Scholar 

  69. Veal, E. A., Day, A. M., & Morgan, B. A. (2007). Hydrogen peroxide sensing and signaling. Molecular Cell, 26, 1–14.

    Article  PubMed  CAS  Google Scholar 

  70. Chiarugi, P., & Cirri, P. (2003). Redox regulation of protein tyrosine phosphatases during receptor tyrosine kinase signal transduction. Trends in Biochemical Sciences, 28, 509–514.

    Article  PubMed  CAS  Google Scholar 

  71. Pani, G., Colavitti, R., Bedogni, B., Anzevino, R., Borrello, S., & Galeotti, T. (2000). A redox signaling mechanism for density-dependent inhibition of cell growth. Journal of Biological Chemistry, 275, 38891–38899.

    Article  PubMed  CAS  Google Scholar 

  72. Mannick, J. B., Hausladen, A., Liu, L., Hess, D. T., Zeng, M., Miao, Q. X., et al. (1999). Fas-induced caspase denitrosylation. Science, 284, 651–654.

    Article  PubMed  CAS  Google Scholar 

  73. Gu, Z., Kaul, M., Yan, B., Kridel, S. J., Cui, J., Strongin, A., et al. (2002). S-nitrosylation of matrix metalloproteinases: signaling pathway to neuronal cell death. Science, 297, 1186–1190.

    Article  PubMed  CAS  Google Scholar 

  74. Saitoh, M., Nishitoh, H., Fujii, M., Takeda, K., Tobiume, K., Sawada, Y., et al. (1998). Mammalian thioredoxin is a direct inhibitor of apoptosis signal-regulating kinase (ASK) 1. EMBO Journal, 17, 2596–2606.

    Article  PubMed  CAS  Google Scholar 

  75. Dinkova-Kostova, A. T., Holtzclaw, W. D., Cole, R. N., Itoh, K., Wakabayashi, N., Katoh, Y., et al. (2002). Direct evidence that sulfhydryl groups of Keap1 are the sensors regulating induction of phase 2 enzymes that protect against carcinogens and oxidants. Proceedings of the National Academy of Sciences of the United States of America, 99, 11908–11913.

    Article  PubMed  CAS  Google Scholar 

  76. Liu, H., Colavitti, R., Rovira, I. I., & Finkel, T. (2005). Redox-dependent transcriptional regulation. Circulation Research, 97, 967–974.

    Article  PubMed  CAS  Google Scholar 

  77. Xanthoudakis, S., & Curran, T. (1992). Identification and characterization of Ref-1, a nuclear protein that facilitates AP-1 DNA-binding activity. EMBO Journal, 11, 653–665.

    PubMed  CAS  Google Scholar 

  78. Dansen, T. B., Smits, L. M., van Triest, M. H., de Keizer, P. L., van Leenen, D., Koerkamp, M. G., et al. (2009). Redox-sensitive cysteines bridge p300/CBP-mediated acetylation and FoxO4 activity. Nat Chem Biol, 5, 664–672.

    Article  PubMed  CAS  Google Scholar 

  79. van der Horst, A., Tertoolen, L. G., de Vries-Smits, L. M., Frye, R. A., Medema, R. H., & Burgering, B. M. (2004). FOXO4 is acetylated upon peroxide stress and deacetylated by the longevity protein hSir2(SIRT1). Journal of Biological Chemistry, 279, 28873–28879.

    Article  PubMed  CAS  Google Scholar 

  80. Brunet, A., Bonni, A., Zigmond, M. J., Lin, M. Z., Juo, P., Hu, L. S., et al. (1999). Akt promotes cell survival by phosphorylating and inhibiting a Forkhead transcription factor. Cell, 96, 857–868.

    Article  PubMed  CAS  Google Scholar 

  81. Nemoto, S., & Finkel, T. (2002). Redox regulation of forkhead proteins through a p66shc-dependent signaling pathway. Science, 295, 2450–2452.

    Article  PubMed  CAS  Google Scholar 

  82. Yasinska, I. M., & Sumbayev, V. V. (2003). S-nitrosation of Cys-800 of HIF-1alpha protein activates its interaction with p300 and stimulates its transcriptional activity. FEBS Letters, 549, 105–109.

    Article  PubMed  CAS  Google Scholar 

  83. Gu, J., Milligan, J., & Huang, L. E. (2001). Molecular mechanism of hypoxia-inducible factor 1alpha -p300 interaction. A leucine-rich interface regulated by a single cysteine. Journal of Biological Chemistry, 276, 3550–3554.

    Article  PubMed  CAS  Google Scholar 

  84. Gerald, D., Berra, E., Frapart, Y. M., Chan, D. A., Giaccia, A. J., Mansuy, D., et al. (2004). JunD reduces tumor angiogenesis by protecting cells from oxidative stress. Cell, 118, 781–794.

    Article  PubMed  CAS  Google Scholar 

  85. Schroedl, C., McClintock, D. S., Budinger, G. R., & Chandel, N. S. (2002). Hypoxic but not anoxic stabilization of HIF-1alpha requires mitochondrial reactive oxygen species. American Journal of Physiology Lung Cellular and Molecular Physiology, 283, L922–L931.

    PubMed  CAS  Google Scholar 

  86. Lu, H., Dalgard, C. L., Mohyeldin, A., Mcfate, T., Tait, A. S., & Verma, A. (2005). Reversible inactivation of HIF-1 prolyl hydroxylases allows cell metabolism to control basal HIF-1. Journal of Biological Chemistry, 280, 41928–41939.

    Article  PubMed  CAS  Google Scholar 

  87. Arbiser, J. L., Petros, J., Klafter, R., Govindajaran, B., McLaughlin, E. R., Brown, L. F., et al. (2002). Reactive oxygen generated by Nox1 triggers the angiogenic switch. Proceedings of the National Academy of Sciences of the United States of America, 99, 715–720.

    Article  PubMed  CAS  Google Scholar 

  88. Wang, F. S., Wang, C. J., Chen, Y. J., Chang, P. R., Huang, Y. T., Sun, Y. C., et al. (2004). Ras induction of superoxide activates ERK-dependent angiogenic transcription factor HIF-1alpha and VEGF-A expression in shock wave-stimulated osteoblasts. Journal of Biological Chemistry, 279, 10331–10337.

    Article  PubMed  CAS  Google Scholar 

  89. Gao, P., Zhang, H., Dinavahi, R., Li, F., Xiang, Y., Raman, V., et al. (2007). HIF-dependent antitumorigenic effect of antioxidants in vivo. Cancer Cell, 12, 230–238.

    Article  PubMed  CAS  Google Scholar 

  90. Nott, A., Watson, P. M., Robinson, J. D., Crepaldi, L., & Riccio, A. (2008). S-Nitrosylation of histone deacetylase 2 induces chromatin remodelling in neurons. Nature, 455, 411–415.

    Article  PubMed  CAS  Google Scholar 

  91. Bedogni, B., Pani, G., Colavitti, R., Riccio, A., Borrello, S., Murphy, M., et al. (2003). Redox regulation of cAMP-responsive element-binding protein and induction of manganous superoxide dismutase in nerve growth factor-dependent cell survival. Journal of Biological Chemistry, 278, 16510–16519.

    Article  PubMed  CAS  Google Scholar 

  92. Riccio, A., Alvania, R. S., Lonze, B. E., Ramanan, N., Kim, T., Huang, Y., et al. (2006). A nitric oxide signaling pathway controls CREB-mediated gene expression in neurons. Molecular Cell, 21, 283–294.

    Article  PubMed  CAS  Google Scholar 

  93. Ito, K., Hanazawa, T., Tomita, K., Barnes, P. J., & Adcock, I. M. (2004). Oxidative stress reduces histone deacetylase 2 activity and enhances IL-8 gene expression: role of tyrosine nitration. Biochemical and Biophysical Research Communications, 315, 240–245.

    Article  PubMed  CAS  Google Scholar 

  94. Moldovan, L., Irani, K., Moldovan, N. I., Finkel, T., & Goldschmidt-Clermont, P. J. (1999). The actin cytoskeleton reorganization induced by Rac1 requires the production of superoxide. Antioxidants Redox Signaling, 1, 29–43.

    Article  PubMed  CAS  Google Scholar 

  95. Joneson, T., & Bar-Sagi, D. (1998). A Rac1 effector site controlling mitogenesis through superoxide production. Journal of Biological Chemistry, 273, 17991–17994.

    Article  PubMed  CAS  Google Scholar 

  96. Kheradmand, F., Werner, E., Tremble, P., Symons, M., & Werb, Z. (1998). Role of Rac1 and oxygen radicals in collagenase-1 expression induced by cell shape change. Science, 280, 898–902.

    Article  PubMed  CAS  Google Scholar 

  97. Werner, E., & Werb, Z. (2002). Integrins engage mitochondrial function for signal transduction by a mechanism dependent on Rho GTPases. Journal of Cell Biology, 158, 357–368.

    Article  PubMed  CAS  Google Scholar 

  98. Zhou, C., Ziegler, C., Birder, L. A., Stewart, A. F., & Levitan, E. S. (2006). Angiotensin II and stretch activate NADPH oxidase to destabilize cardiac Kv4.3 channel mRNA. Circulation Research, 98, 1040–1047.

    Article  PubMed  CAS  Google Scholar 

  99. Taddei, M. L., Parri, M., Mello, T., Catalano, A., Levine, A. D., Raugei, G., et al. (2007). Integrin-mediated cell adhesion and spreading engage different sources of reactive oxygen species. Antioxidants Redox Signaling, 9, 469–481.

    Article  PubMed  CAS  Google Scholar 

  100. Nimnual, A. S., Taylor, L. J., & Bar-Sagi, D. (2003). Redox-dependent downregulation of Rho by Rac. Nature Cell Biology, 5, 236–241.

    Article  PubMed  CAS  Google Scholar 

  101. Ng, J., & Luo, L. (2004). Rho GTPases regulate axon growth through convergent and divergent signaling pathways. Neuron, 44, 779–793.

    Article  PubMed  CAS  Google Scholar 

  102. Zondag, G. C., Evers, E. E., ten Klooster, J. P., Janssen, L., van der Kammen, R. A., & Collard, J. G. (2000). Oncogenic Ras downregulates Rac activity, which leads to increased Rho activity and epithelial–mesenchymal transition. Journal of Cell Biology, 149, 775–782.

    Article  PubMed  CAS  Google Scholar 

  103. Buricchi, F., Giannoni, E., Grimaldi, G., Parri, M., Raugei, G., Ramponi, G., et al. (2007). Redox regulation of ephrin/integrin cross-talk. Cell Adh Migr, 1, 33–42.

    PubMed  Google Scholar 

  104. Taddei, M. L., Parri, M., Angelucci, A., Onnis, B., Bianchini, F., Giannoni, E., et al. (2009). Kinase-dependent and -independent roles of EphA2 in the regulation of prostate cancer invasion and metastasis. American Journal of Pathology, 174, 1492–1503.

    Article  PubMed  CAS  Google Scholar 

  105. Giannoni, E., Buricchi, F., Raugei, G., Ramponi, G., & Chiarugi, P. (2005). Intracellular reactive oxygen species activate Src tyrosine kinase during cell adhesion and anchorage-dependent cell growth. Molecular and Cellular Biology, 25, 6391–6403.

    Article  PubMed  CAS  Google Scholar 

  106. Fratelli, M., Demol, H., Puype, M., Casagrande, S., Eberini, I., Salmona, M., et al. (2002). Identification by redox proteomics of glutathionylated proteins in oxidatively stressed human T lymphocytes. Proceedings of the National Academy of Sciences of the United States of America, 99, 3505–3510.

    Article  PubMed  CAS  Google Scholar 

  107. Fiaschi, T., Cozzi, G., Raugei, G., Formigli, L., Ramponi, G., & Chiarugi, P. (2006). Redox regulation of beta-actin during integrin-mediated cell adhesion. Journal of Biological Chemistry, 281, 22983–22991.

    Article  PubMed  CAS  Google Scholar 

  108. Terman, J. R., Mao, T., Pasterkamp, R. J., Yu, H. H., & Kolodkin, A. L. (2002). MICALs, a family of conserved flavoprotein oxidoreductases, function in plexin-mediated axonal repulsion. Cell, 109, 887–900.

    Article  PubMed  CAS  Google Scholar 

  109. Ventura, A., & Pelicci, P. G. (2002). Semaphorins: green light for redox signaling? Sci STKE, 2002, e44.

    Article  Google Scholar 

  110. Suzuki, T., Nakamoto, T., Ogawa, S., Seo, S., Matsumura, T., Tachibana, K., et al. (2002). MICAL, a novel CasL interacting molecule, associates with vimentin. Journal of Biological Chemistry, 277, 14933–14941.

    PubMed  CAS  Google Scholar 

  111. Nadella, M., Bianchet, M. A., Gabelli, S. B., Barrila, J., & Amzel, L. M. (2005). Structure and activity of the axon guidance protein MICAL. Proceedings of the National Academy of Sciences of the United States of America, 102, 16830–16835.

    Article  PubMed  CAS  Google Scholar 

  112. Fischer, J., Weide, T., & Barnekow, A. (2005). The MICAL proteins and rab1: a possible link to the cytoskeleton? Biochemical and Biophysical Research Communications, 328, 415–423.

    Article  PubMed  CAS  Google Scholar 

  113. Kanda, I., Nishimura, N., Nakatsuji, H., Yamamura, R., Nakanishi, H., & Sasaki, T. (2008). Involvement of Rab13 and JRAB/MICAL-L2 in epithelial cell scattering. Oncogene, 27, 1687–1695.

    Article  PubMed  CAS  Google Scholar 

  114. Ferraro, D., Corso, S., Fasano, E., Panieri, E., Santangelo, R., Borrello, S., et al. (2006). Pro-metastatic signaling by c-Met through RAC-1 and reactive oxygen species (ROS). Oncogene, 25, 3689–3698.

    Article  PubMed  CAS  Google Scholar 

  115. Jagadeeswaran, R., Jagadeeswaran, S., Bindokas, V. P., & Salgia, R. (2007). Activation of HGF/c-Met pathway contributes to the reactive oxygen species generation and motility of small cell lung cancer cells. American Journal of Physiology Lung Cellular and Molecular Physiology, 292, L1488–L1494.

    Article  PubMed  CAS  Google Scholar 

  116. Li, W., Liu, G., Chou, I. N., & Kagan, H. M. (2000). Hydrogen peroxide-mediated, lysyl oxidase-dependent chemotaxis of vascular smooth muscle cells. Journal of Cellular Biochemistry, 78, 550–557.

    Article  PubMed  CAS  Google Scholar 

  117. Ushio-Fukai, M. (2009). Compartmentalization of redox signaling through NADPH oxidase-derived ROS. Antioxidants Redox Signaling, 11, 1289–1299.

    Article  PubMed  CAS  Google Scholar 

  118. Wu, R. F., Xu, Y. C., Ma, Z., Nwariaku, F. E., Sarosi, G. A., Jr., & Terada, L. S. (2005). Subcellular targeting of oxidants during endothelial cell migration. Journal of Cell Biology, 171, 893–904.

    Article  PubMed  CAS  Google Scholar 

  119. Diaz, B., Shani, G., Pass, I., Anderson, D., Quintavalle, M., & Courtneidge, S. A. (2009). Tks5-dependent, nox-mediated generation of reactive oxygen species is necessary for invadopodia formation. Sci Signal, 2, ra53.

    Article  PubMed  Google Scholar 

  120. Gianni, D., Diaz, B., Taulet, N., Fowler, B., Courtneidge, S. A., & Bokoch, G. M. (2009). Novel p47(phox)-related organizers regulate localized NADPH oxidase 1 (Nox1) activity. Science Signal, 2, ra54.

    Article  Google Scholar 

  121. Patel, H. H., & Insel, P. A. (2009). Lipid rafts and caveolae and their role in compartmentation of redox signaling. Antioxidants Redox Signaling, 11, 1357–1372.

    Article  PubMed  CAS  Google Scholar 

  122. Gaus, K., Le, L. S., Balasubramanian, N., & Schwartz, M. A. (2006). Integrin-mediated adhesion regulates membrane order. Journal of Cell Biology, 174, 725–734.

    Article  PubMed  CAS  Google Scholar 

  123. Meng, T. C., Fukada, T., & Tonks, N. K. (2002). Reversible oxidation and inactivation of protein tyrosine phosphatases in vivo. Molecular Cell, 9, 387–399.

    Article  PubMed  CAS  Google Scholar 

  124. Campbell, M., Anderson, P., & Trimble, E. R. (2008). Glucose lowers the threshold for human aortic vascular smooth muscle cell migration: inhibition by protein phosphatase-2A. Diabetologia, 51, 1068–1080.

    Article  PubMed  CAS  Google Scholar 

  125. Leopold, J. A., Walker, J., Scribner, A. W., Voetsch, B., Zhang, Y. Y., Loscalzo, A. J., et al. (2003). Glucose-6-phosphate dehydrogenase modulates vascular endothelial growth factor-mediated angiogenesis. Journal of Biological Chemistry, 278, 32100–32106.

    Article  PubMed  CAS  Google Scholar 

  126. Ames, B. N., Gold, L. S., & Willett, W. C. (1995). The causes and prevention of cancer. Proceedings of the National Academy of Sciences of the United States of America, 92, 5258–5265.

    Article  PubMed  CAS  Google Scholar 

  127. Ambrosone, C. B., Freudenheim, J. L., Thompson, P. A., Bowman, E., Vena, J. E., Marshall, J. R., et al. (1999). Manganese superoxide dismutase (MnSOD) genetic polymorphisms, dietary antioxidants, and risk of breast cancer. Cancer Research, 59, 602–606.

    PubMed  CAS  Google Scholar 

  128. Ratnasinghe, D., Tangrea, J. A., Andersen, M. R., Barrett, M. J., Virtamo, J., Taylor, P. R., et al. (2000). Glutathione peroxidase codon 198 polymorphism variant increases lung cancer risk. Cancer Research, 60, 6381–6383.

    PubMed  CAS  Google Scholar 

  129. Zhao, Y., Oberley, T. D., Chaiswing, L., Lin, S. M., Epstein, C. J., Huang, T. T., et al. (2002). Manganese superoxide dismutase deficiency enhances cell turnover via tumor promoter-induced alterations in AP-1 and p53-mediated pathways in a skin cancer model. Oncogene, 21, 3836–3846.

    Article  PubMed  CAS  Google Scholar 

  130. Marnett, L. J. (2000). Oxyradicals and DNA damage. Carcinogenesis, 21, 361–370.

    Article  PubMed  CAS  Google Scholar 

  131. Hussain, S. P., Hofseth, L. J., & Harris, C. C. (2003). Radical causes of cancer. Nature Reviews Cancer, 3, 276–285.

    Article  PubMed  CAS  Google Scholar 

  132. Cerutti, P. A. (1985). Prooxidant states and tumor promotion. Science, 227, 375–381.

    Article  PubMed  CAS  Google Scholar 

  133. Irani, K., Xia, Y., Zweier, J. L., Sollott, S. J., Der, C. J., Fearon, E. R., et al. (1997). Mitogenic signaling mediated by oxidants in Ras-transformed fibroblasts. Science, 275, 1649–1652.

    Article  PubMed  CAS  Google Scholar 

  134. Hanahan, D., & Weinberg, R. A. (2000). The hallmarks of cancer. Cell, 100, 57–70.

    Article  PubMed  CAS  Google Scholar 

  135. Szatrowski, T. P., & Nathan, C. F. (1991). Production of large amounts of hydrogen peroxide by human tumor cells. Cancer Research, 51, 794–798.

    PubMed  CAS  Google Scholar 

  136. Lim, S. D., Sun, C., Lambeth, J. D., Marshall, F., Amin, M., Chung, L., et al. (2005). Increased Nox1 and hydrogen peroxide in prostate cancer. Prostate, 62, 200–207.

    Article  PubMed  CAS  Google Scholar 

  137. Mitsushita, J., Lambeth, J. D., & Kamata, T. (2004). The superoxide-generating oxidase Nox1 is functionally required for Ras oncogene transformation. Cancer Research, 64, 3580–3585.

    Article  PubMed  CAS  Google Scholar 

  138. Shinohara, M., Shang, W. H., Kubodera, M., Harada, S., Mitsushita, J., Kato, M., et al. (2007). Nox1 redox signaling mediates oncogenic Ras-induced disruption of stress fibers and focal adhesions by down-regulating Rho. Journal of Biological Chemistry, 282, 17640–17648.

    Article  PubMed  CAS  Google Scholar 

  139. Lander, H. M., Milbank, A. J., Tauras, J. M., Hajjar, D. P., Hempstead, B. L., Schwartz, G. D., et al. (1996). Redox regulation of cell signalling. Nature, 381, 380–381.

    Article  PubMed  CAS  Google Scholar 

  140. Benhar, M., Dalyot, I., Engelberg, D., & Levitzki, A. (2001). Enhanced ROS production in oncogenically transformed cells potentiates c-Jun N-terminal kinase and p38 mitogen-activated protein kinase activation and sensitization to genotoxic stress. Molecular and Cellular Biology, 21, 6913–6926.

    Article  PubMed  CAS  Google Scholar 

  141. Rhee, S. G., Bae, Y. S., Lee, S. R., & Kwon, J. (2000). Hydrogen peroxide: a key messenger that modulates protein phosphorylation through cysteine oxidation. Sci STKE, 2000, e1.

    Article  Google Scholar 

  142. Konishi, H., Tanaka, M., Takemura, Y., Matsuzaki, H., Ono, Y., Kikkawa, U., et al. (1997). Activation of protein kinase C by tyrosine phosphorylation in response to H2O2. Proceedings of the National Academy of Sciences of the United States of America, 94, 11233–11237.

    Article  PubMed  CAS  Google Scholar 

  143. Pani, G., Bedogni, B., Colavitti, R., Anzevino, R., Borrello, S., & Galeotti, T. (2001). Cell compartmentalization in redox signaling. IUBMB Life, 52, 7–16.

    Article  PubMed  CAS  Google Scholar 

  144. Coats, S., Flanagan, W. M., Nourse, J., & Roberts, J. M. (1996). Requirement of p27Kip1 for restriction point control of the fibroblast cell cycle. Science, 272, 877–880.

    Article  PubMed  CAS  Google Scholar 

  145. Medema, R. H., Kops, G. J., Bos, J. L., & Burgering, B. M. (2000). AFX-like Forkhead transcription factors mediate cell-cycle regulation by Ras and PKB through p27kip1. Nature, 404, 782–787.

    Article  PubMed  CAS  Google Scholar 

  146. Nogueira, V., Park, Y., Chen, C. C., Xu, P. Z., Chen, M. L., Tonic, I., et al. (2008). Akt determines replicative senescence and oxidative or oncogenic premature senescence and sensitizes cells to oxidative apoptosis. Cancer Cell, 14, 458–470.

    Article  PubMed  CAS  Google Scholar 

  147. Itahana, K., Campisi, J., & Dimri, G. P. (2004). Mechanisms of cellular senescence in human and mouse cells. Biogerontology, 5, 1–10.

    Article  PubMed  CAS  Google Scholar 

  148. Parrinello, S., Samper, E., Krtolica, A., Goldstein, J., Melov, S., & Campisi, J. (2003). Oxygen sensitivity severely limits the replicative lifespan of murine fibroblasts. Nature Cell Biology, 5, 741–747.

    Article  PubMed  CAS  Google Scholar 

  149. Packer, L., & Fuehr, K. (1977). Low oxygen concentration extends the lifespan of cultured human diploid cells. Nature, 267, 423–425.

    Article  PubMed  CAS  Google Scholar 

  150. Bell, E. L., Klimova, T. A., Eisenbart, J., Schumacker, P. T., & Chandel, N. S. (2007). Mitochondrial reactive oxygen species trigger hypoxia-inducible factor-dependent extension of the replicative life span during hypoxia. Molecular and Cellular Biology, 27, 5737–5745.

    Article  PubMed  CAS  Google Scholar 

  151. Bermudez, Y., Ahmadi, S., Lowell, N. E., & Kruk, P. A. (2007). Vitamin E suppresses telomerase activity in ovarian cancer cells. Cancer Detection and Prevention, 31, 119–128.

    Article  PubMed  CAS  Google Scholar 

  152. Kwon, J., Lee, S. R., Yang, K. S., Ahn, Y., Kim, Y. J., Stadtman, E. R., et al. (2004). Reversible oxidation and inactivation of the tumor suppressor PTEN in cells stimulated with peptide growth factors. Proceedings of the National Academy of Sciences of the United States of America, 101, 16419–16424.

    Article  PubMed  CAS  Google Scholar 

  153. Romashkova, J. A., & Makarov, S. S. (1999). NF-kappaB is a target of AKT in anti-apoptotic PDGF signalling. Nature, 401, 86–90.

    Article  PubMed  CAS  Google Scholar 

  154. Ozes, O. N., Mayo, L. D., Gustin, J. A., Pfeffer, S. R., Pfeffer, L. M., & Donner, D. B. (1999). NF-kappaB activation by tumour necrosis factor requires the Akt serine-threonine kinase. Nature, 401, 82–85.

    Article  PubMed  CAS  Google Scholar 

  155. Anderson, M. T., Staal, F. J., Gitler, C., Herzenberg, L. A., & Herzenberg, L. A. (1994). Separation of oxidant-initiated and redox-regulated steps in the NF-kappa B signal transduction pathway. Proceedings of the National Academy of Sciences of the United States of America, 91, 11527–11531.

    Article  PubMed  CAS  Google Scholar 

  156. Sulciner, D. J., Irani, K., Yu, Z. X., Ferrans, V. J., Goldschmidt-Clermont, P., & Finkel, T. (1996). rac1 regulates a cytokine-stimulated, redox-dependent pathway necessary for NF-kappaB activation. Molecular and Cellular Biology, 16, 7115–7121.

    PubMed  CAS  Google Scholar 

  157. Yang, J. Q., Zhao, W., Duan, H., Robbins, M. E., Buettner, G. R., Oberley, L. W., et al. (2001). v-Ha-RaS oncogene upregulates the 92-kDa type IV collagenase (MMP-9) gene by increasing cellular superoxide production and activating NF-kappaB. Free Radical Biology and Medicine, 31, 520–529.

    Article  PubMed  CAS  Google Scholar 

  158. Joneson, T., & Bar-Sagi, D. (1999). Suppression of Ras-induced apoptosis by the Rac GTPase. Molecular and Cellular Biology, 19, 5892–5901.

    PubMed  CAS  Google Scholar 

  159. Lluis, J. M., Buricchi, F., Chiarugi, P., Morales, A., & Fernandez-Checa, J. C. (2007). Dual role of mitochondrial reactive oxygen species in hypoxia signaling: activation of nuclear factor-{kappa}B via c-SRC and oxidant-dependent cell death. Cancer Research, 67, 7368–7377.

    Article  PubMed  CAS  Google Scholar 

  160. Barbie, D. A., Tamayo, P., Boehm, J. S., Kim, S. Y., Moody, S. E., Dunn, I. F., et al. (2009). Systematic RNA interference reveals that oncogenic KRAS-driven cancers require TBK1. Nature, 462, 108–112.

    Article  PubMed  CAS  Google Scholar 

  161. Mayo, M. W., Wang, C. Y., Cogswell, P. C., Rogers-Graham, K. S., Lowe, S. W., Der, C. J., et al. (1997). Requirement of NF-kappaB activation to suppress p53-independent apoptosis induced by oncogenic Ras. Science, 278, 1812–1815.

    Article  PubMed  CAS  Google Scholar 

  162. Ciani, E., Guidi, S., Della, V. G., Perini, G., Bartesaghi, R., & Contestabile, A. (2002). Nitric oxide protects neuroblastoma cells from apoptosis induced by serum deprivation through cAMP-response element-binding protein (CREB) activation. Journal of Biological Chemistry, 277, 49896–49902.

    Article  PubMed  CAS  Google Scholar 

  163. Hess, D. T., Matsumoto, A., Kim, S. O., Marshall, H. E., & Stamler, J. S. (2005). Protein S-nitrosylation: purview and parameters. Nature Reviews Molecular Cell Biology, 6, 150–166.

    Article  PubMed  CAS  Google Scholar 

  164. Pan, S., & Berk, B. C. (2007). Glutathiolation regulates tumor necrosis factor-alpha-induced caspase-3 cleavage and apoptosis: key role for glutaredoxin in the death pathway. Circulation Research, 100, 213–219.

    Article  PubMed  CAS  Google Scholar 

  165. Giannoni, E., Fiaschi, T., Ramponi, G., & Chiarugi, P. (2009). Redox regulation of anoikis resistance of metastatic prostate cancer cells: key role for Src and EGFR-mediated pro-survival signals. Oncogene, 28, 2074–2086.

    Article  PubMed  CAS  Google Scholar 

  166. Pani, G., Giannoni, E., Galeotti, T., & Chiarugi, P. (2009). Redox-based escape mechanism from death: the cancer lesson. Antioxidants Redox Signaling, 11, 2791–2806.

    Article  PubMed  CAS  Google Scholar 

  167. Maulik, N. (2002). Redox signaling of angiogenesis. Antioxidants Redox Signaling, 4, 805–815.

    Article  PubMed  CAS  Google Scholar 

  168. Komatsu, D., Kato, M., Nakayama, J., Miyagawa, S., & Kamata, T. (2008). NADPH oxidase 1 plays a critical mediating role in oncogenic Ras-induced vascular endothelial growth factor expression. Oncogene, 27, 4724–4732.

    Article  PubMed  CAS  Google Scholar 

  169. Zundel, W., Schindler, C., Haas-Kogan, D., Koong, A., Kaper, F., Chen, E., et al. (2000). Loss of PTEN facilitates HIF-1-mediated gene expression. Genes and Development, 14, 391–396.

    PubMed  CAS  Google Scholar 

  170. Hudson, C. C., Liu, M., Chiang, G. G., Otterness, D. M., Loomis, D. C., Kaper, F., et al. (2002). Regulation of hypoxia-inducible factor 1alpha expression and function by the mammalian target of rapamycin. Molecular and Cellular Biology, 22, 7004–7014.

    Article  PubMed  CAS  Google Scholar 

  171. Sarbassov, D. D., & Sabatini, D. M. (2005). Redox regulation of the nutrient-sensitive raptor-mTOR pathway and complex. Journal of Biological Chemistry, 280, 39505–39509.

    Article  PubMed  CAS  Google Scholar 

  172. Metzen, E., Zhou, J., Jelkmann, W., Fandrey, J., & Brune, B. (2003). Nitric oxide impairs normoxic degradation of HIF-1alpha by inhibition of prolyl hydroxylases. Molecular Biology of the Cell, 14, 3470–3481.

    Article  PubMed  CAS  Google Scholar 

  173. Kasuno, K., Takabuchi, S., Fukuda, K., Kizaka-Kondoh, S., Yodoi, J., Adachi, T., et al. (2004). Nitric oxide induces hypoxia-inducible factor 1 activation that is dependent on MAPK and phosphatidylinositol 3-kinase signaling. Journal of Biological Chemistry, 279, 2550–2558.

    Article  PubMed  CAS  Google Scholar 

  174. Parenti, A., Morbidelli, L., Cui, X. L., Douglas, J. G., Hood, J. D., Granger, H. J., et al. (1998). Nitric oxide is an upstream signal of vascular endothelial growth factor-induced extracellular signal-regulated kinase1/2 activation in postcapillary endothelium. Journal of Biological Chemistry, 273, 4220–4226.

    Article  PubMed  CAS  Google Scholar 

  175. Colavitti, R., Pani, G., Bedogni, B., Anzevino, R., Borrello, S., Waltenberger, J., et al. (2002). Reactive oxygen species as downstream mediators of angiogenic signaling by vascular endothelial growth factor receptor-2/KDR. Journal of Biological Chemistry, 277, 3101–3108.

    Article  PubMed  CAS  Google Scholar 

  176. Harfouche, R., Malak, N. A., Brandes, R. P., Karsan, A., Irani, K., & Hussain, S. N. (2005). Roles of reactive oxygen species in angiopoietin-1/tie-2 receptor signaling. FASEB Journal, 19, 1728–1730.

    PubMed  CAS  Google Scholar 

  177. Ushio-Fukai, M. (2006). Redox signaling in angiogenesis: role of NADPH oxidase. Cardiovascular Research, 71, 226–235.

    Article  PubMed  CAS  Google Scholar 

  178. Urao, N., Inomata, H., Razvi, M., Kim, H. W., Wary, K., McKinney, R., et al. (2008). Role of nox2-based NADPH oxidase in bone marrow and progenitor cell function involved in neovascularization induced by hindlimb ischemia. Circulation Research, 103, 212–220.

    Article  PubMed  CAS  Google Scholar 

  179. Lelkes, P. I., Hahn, K. L., Sukovich, D. A., Karmiol, S., & Schmidt, D. H. (1998). On the possible role of reactive oxygen species in angiogenesis. Advances in Experimental Medicine and Biology, 454, 295–310.

    PubMed  CAS  Google Scholar 

  180. Fukumura, D., Kashiwagi, S., & Jain, R. K. (2006). The role of nitric oxide in tumour progression. Nature Reviews Cancer, 6, 521–534.

    Article  PubMed  CAS  Google Scholar 

  181. Radisky, D. C., Levy, D. D., Littlepage, L. E., Liu, H., Nelson, C. M., Fata, J. E., et al. (2005). Rac1b and reactive oxygen species mediate MMP-3-induced EMT and genomic instability. Nature, 436, 123–127.

    Article  PubMed  CAS  Google Scholar 

  182. Nelson, C. M., Khauv, D., Bissell, M. J., & Radisky, D. C. (2008). Change in cell shape is required for matrix metalloproteinase-induced epithelial-mesenchymal transition of mammary epithelial cells. Journal of Cellular Biochemistry, 105, 25–33.

    Article  PubMed  CAS  Google Scholar 

  183. Cannito, S., Novo, E., Compagnone, A., di Valfre, B. L., Busletta, C., Zamara, E., et al. (2008). Redox mechanisms switch on hypoxia-dependent epithelial–mesenchymal transition in cancer cells. Carcinogenesis, 29, 2267–2278.

    Article  PubMed  CAS  Google Scholar 

  184. Ishikawa, K., Takenaga, K., Akimoto, M., Koshikawa, N., Yamaguchi, A., Imanishi, H., et al. (2008). ROS-generating mitochondrial DNA mutations can regulate tumor cell metastasis. Science, 320, 661–664.

    Article  PubMed  CAS  Google Scholar 

  185. Binker, M. G., Binker-Cosen, A. A., Richards, D., Oliver, B., & Cosen-Binker, L. I. (2009). EGF promotes invasion by PANC-1 cells through Rac1/ROS-dependent secretion and activation of MMP-2. Biochemical and Biophysical Research Communications, 379, 445–450.

    Article  PubMed  CAS  Google Scholar 

  186. Oberley, T. D., Schultz, J. L., Li, N., & Oberley, L. W. (1995). Antioxidant enzyme levels as a function of growth state in cell culture. Free Radical Biology and Medicine, 19, 53–65.

    Article  PubMed  CAS  Google Scholar 

  187. 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.

    Article  PubMed  CAS  Google Scholar 

  188. Lee, A. C., Fenster, B. E., Ito, H., Takeda, K., Bae, N. S., Hirai, T., et al. (1999). Ras proteins induce senescence by altering the intracellular levels of reactive oxygen species. Journal of Biological Chemistry, 274, 7936–7940.

    Article  PubMed  CAS  Google Scholar 

  189. Serrano, M., Lin, A. W., McCurrach, M. E., Beach, D., & Lowe, S. W. (1997). Oncogenic ras provokes premature cell senescence associated with accumulation of p53 and p16INK4a. Cell, 88, 593–602.

    Article  PubMed  CAS  Google Scholar 

  190. Takahashi, A., Ohtani, N., Yamakoshi, K., Iida, S., Tahara, H., Nakayama, K., et al. (2006). Mitogenic signalling and the p16INK4a-Rb pathway cooperate to enforce irreversible cellular senescence. Nature Cell Biology, 8, 1291–1297.

    Article  PubMed  CAS  Google Scholar 

  191. Lowe, S. W., Jacks, T., Housman, D. E., & Ruley, H. E. (1994). Abrogation of oncogene-associated apoptosis allows transformation of p53-deficient cells. Proceedings of the National Academy of Sciences of the United States of America, 91, 2026–2030.

    Article  PubMed  CAS  Google Scholar 

  192. Pani, G., Bedogni, B., Anzevino, R., Colavitti, R., Palazzotti, B., Borrello, S., et al. (2000). Deregulated manganese superoxide dismutase expression and resistance to oxidative injury in p53-deficient cells. Cancer Research, 60, 4654–4660.

    PubMed  CAS  Google Scholar 

  193. Forrester, K., Ambs, S., Lupold, S. E., Kapust, R. B., Spillare, E. A., Weinberg, W. C., et al. (1996). Nitric oxide-induced p53 accumulation and regulation of inducible nitric oxide synthase expression by wild-type p53. Proceedings of the National Academy of Sciences of the United States of America, 93, 2442–2447.

    Article  PubMed  CAS  Google Scholar 

  194. Ambs, S., Merriam, W. G., Ogunfusika, M. O., Bennett, W. P., Ishibe, N., Hussain, S. P., et al. (1998). p53 and vascular endothelial growth factor regulate tumor growth of NOS2-expressing human carcinoma cells. Nature Medicine, 4, 1371–1376.

    Article  PubMed  CAS  Google Scholar 

  195. Ambs, S., Bennett, W. P., Merriam, W. G., Ogunfusika, M. O., Oser, S. M., Harrington, A. M., et al. (1999). Relationship between p53 mutations and inducible nitric oxide synthase expression in human colorectal cancer. Journal of the National Cancer Institute, 91, 86–88.

    Article  PubMed  CAS  Google Scholar 

  196. Kops, G. J., Dansen, T. B., Polderman, P. E., Saarloos, I., Wirtz, K. W., Coffer, P. J., et al. (2002). Forkhead transcription factor FOXO3a protects quiescent cells from oxidative stress. Nature, 419, 316–321.

    Article  PubMed  CAS  Google Scholar 

  197. Park, H. J., Carr, J. R., Wang, Z., Nogueira, V., Hay, N., Tyner, A. L., et al. (2009). FoxM1, a critical regulator of oxidative stress during oncogenesis. EMBO Journal, 28, 2908–2918.

    Article  PubMed  CAS  Google Scholar 

  198. Tanaka, H., Matsumura, I., Ezoe, S., Satoh, Y., Sakamaki, T., Albanese, C., et al. (2002). E2F1 and c-Myc potentiate apoptosis through inhibition of NF-kappaB activity that facilitates MnSOD-mediated ROS elimination. Molecular Cell, 9, 1017–1029.

    Article  PubMed  CAS  Google Scholar 

  199. Dolado, I., Swat, A., Ajenjo, N., De, V. G., Cuadrado, A., & Nebreda, A. R. (2007). p38alpha MAP kinase as a sensor of reactive oxygen species in tumorigenesis. Cancer Cell, 11, 191–205.

    Article  PubMed  CAS  Google Scholar 

  200. Kennedy, N. J., Cellurale, C., & Davis, R. J. (2007). A radical role for p38 MAPK in tumor initiation. Cancer Cell, 11, 101–103.

    Article  PubMed  CAS  Google Scholar 

  201. Bulavin, D. V., Phillips, C., Nannenga, B., Timofeev, O., Donehower, L. A., Anderson, C. W., et al. (2004). Inactivation of the Wip1 phosphatase inhibits mammary tumorigenesis through p38 MAPK-mediated activation of the p16(Ink4a)-p19(Arf) pathway. Nature Genetics, 36, 343–350.

    Article  PubMed  CAS  Google Scholar 

  202. Helzlsouer, K. J., Selmin, O., Huang, H. Y., Strickland, P. T., Hoffman, S., Alberg, A. J., et al. (1998). Association between glutathione S-transferase M1, P1, and T1 genetic polymorphisms and development of breast cancer. Journal of the National Cancer Institute, 90, 512–518.

    Article  PubMed  CAS  Google Scholar 

  203. Havre, P. A., O'Reilly, S., McCormick, J. J., & Brash, D. E. (2002). Transformed and tumor-derived human cells exhibit preferential sensitivity to the thiol antioxidants. N-acetyl cysteine and penicillamine. Cancer Research, 62, 1443–1449.

    PubMed  CAS  Google Scholar 

  204. Polyak, K., Xia, Y., Zweier, J. L., Kinzler, K. W., & Vogelstein, B. (1997). A model for p53-induced apoptosis. Nature, 389, 300–305.

    Article  PubMed  CAS  Google Scholar 

  205. Johnson, T. M., Yu, Z. X., Ferrans, V. J., Lowenstein, R. A., & Finkel, T. (1996). Reactive oxygen species are downstream mediators of p53-dependent apoptosis. Proceedings of the National Academy of Sciences of the United States of America, 93, 11848–11852.

    Article  PubMed  CAS  Google Scholar 

  206. Drane, P., Bravard, A., Bouvard, V., & May, E. (2001). Reciprocal down-regulation of p53 and SOD2 gene expression-implication in p53 mediated apoptosis. Oncogene, 20, 430–439.

    Article  PubMed  CAS  Google Scholar 

  207. Dhar, S. K., Xu, Y., Chen, Y., & St Clair, D. K. (2006). Specificity protein 1-dependent p53-mediated suppression of human manganese superoxide dismutase gene expression. Journal of Biological Chemistry, 281, 21698–21709.

    Article  PubMed  CAS  Google Scholar 

  208. Pani, G., Colavitti, R., Bedogni, B., Fusco, S., Ferraro, D., Borrello, S., et al. (2004). Mitochondrial superoxide dismutase: a promising target for new anticancer therapies. Current Medicinal Chemistry, 11, 1299–1308.

    PubMed  CAS  Google Scholar 

  209. Ito, K., Hirao, A., Arai, F., Matsuoka, S., Takubo, K., Hamaguchi, I., et al. (2004). Regulation of oxidative stress by ATM is required for self-renewal of haematopoietic stem cells. Nature, 431, 997–1002.

    Article  PubMed  CAS  Google Scholar 

  210. Diehn, M., Cho, R. W., Lobo, N. A., Kalisky, T., Dorie, M. J., Kulp, A. N., et al. (2009). Association of reactive oxygen species levels and radioresistance in cancer stem cells. Nature, 458, 780–783.

    Article  PubMed  CAS  Google Scholar 

  211. Smith, J., Ladi, E., Mayer-Proschel, M., & Noble, M. (2000). Redox state is a central modulator of the balance between self-renewal and differentiation in a dividing glial precursor cell. Proceedings of the National Academy of Sciences of the United States of America, 97, 10032–10037.

    Article  PubMed  CAS  Google Scholar 

  212. Jang, Y. Y., & Sharkis, S. J. (2007). A low level of reactive oxygen species selects for primitive hematopoietic stem cells that may reside in the low-oxygenic niche. Blood, 110, 3056–3063.

    Article  PubMed  CAS  Google Scholar 

  213. Ito, K., Hirao, A., Arai, F., Takubo, K., Matsuoka, S., Miyamoto, K., et al. (2006). Reactive oxygen species act through p38 MAPK to limit the lifespan of hematopoietic stem cells. Nature Medicine, 12, 446–451.

    Article  PubMed  CAS  Google Scholar 

  214. Tothova, Z., Kollipara, R., Huntly, B. J., Lee, B. H., Castrillon, D. H., Cullen, D. E., et al. (2007). FoxOs are critical mediators of hematopoietic stem cell resistance to physiologic oxidative stress. Cell, 128, 325–339.

    Article  PubMed  CAS  Google Scholar 

  215. Chen, C., Liu, Y., Liu, R., Ikenoue, T., Guan, K. L., Liu, Y., et al. (2008). TSC-mTOR maintains quiescence and function of hematopoietic stem cells by repressing mitochondrial biogenesis and reactive oxygen species. Journal of Experimental Medicine, 205, 2397–2408.

    Article  PubMed  CAS  Google Scholar 

  216. Ranganathan, A. C., Adam, A. P., & Aguirre-Ghiso, J. A. (2006). Opposing roles of mitogenic and stress signaling pathways in the induction of cancer dormancy. Cell Cycle, 5, 1799–1807.

    PubMed  CAS  Google Scholar 

  217. Adam, A. P., George, A., Schewe, D., Bragado, P., Iglesias, B. V., Ranganathan, A. C., et al. (2009). Computational identification of a p38SAPK-regulated transcription factor network required for tumor cell quiescence. Cancer Research, 69, 5664–5672.

    Article  PubMed  CAS  Google Scholar 

  218. Schewe, D. M., & Aguirre-Ghiso, J. A. (2008). ATF6alpha-Rheb-mTOR signaling promotes survival of dormant tumor cells in vivo. Proceedings of the National Academy of Sciences of the United States of America, 105, 10519–10524.

    Article  PubMed  Google Scholar 

  219. Schafer, Z. T., Grassian, A. R., Song, L., Jiang, Z., Gerhart-Hines, Z., Irie, H. Y., et al. (2009). Antioxidant and oncogene rescue of metabolic defects caused by loss of matrix attachment. Nature, 461, 109–113.

    Article  PubMed  CAS  Google Scholar 

  220. Kuo, W., Lin, J., & Tang, T. K. (2000). Human glucose-6-phosphate dehydrogenase (G6PD) gene transforms NIH 3T3 cells and induces tumors in nude mice. International Journal of Cancer, 85, 857–864.

    Article  CAS  Google Scholar 

  221. The Alpha-Tocopherol, Beta Carotene Cancer Prevention Study Group. (1994). The effect of vitamin E and beta carotene on the incidence of lung cancer and other cancers in male smokers. New England Journal of Medicine, 330, 1029–1035.

    Article  Google Scholar 

  222. Omenn, G. S., Goodman, G. E., Thornquist, M. D., Balmes, J., Cullen, M. R., Glass, A., et al. (1996). Effects of a combination of beta carotene and vitamin A on lung cancer and cardiovascular disease. New England Journal of Medicine, 334, 1150–1155.

    Article  PubMed  CAS  Google Scholar 

  223. Shackelford, R. E., Heinloth, A. N., Heard, S. C., & Paules, R. S. (2005). Cellular and molecular targets of protein S-glutathiolation. Antioxidants Redox Signaling, 7, 940–950.

    Article  PubMed  CAS  Google Scholar 

  224. Coussens, L. M., & Werb, Z. (2002). Inflammation and cancer. Nature, 420, 860–867.

    Article  PubMed  CAS  Google Scholar 

  225. Schafer, M., & Werner, S. (2008). Cancer as an overhealing wound: an old hypothesis revisited. Nature Reviews Molecular Cell Biology, 9, 628–638.

    Article  PubMed  CAS  Google Scholar 

  226. Phinney, D. G., & Prockop, D. J. (2007). Concise review: mesenchymal stem/multipotent stromal cells: the state of transdifferentiation and modes of tissue repair—current views. Stem Cells, 25, 2896–2902.

    Article  PubMed  Google Scholar 

  227. Ishii, T., Itoh, K., Takahashi, S., Sato, H., Yanagawa, T., Katoh, Y., et al. (2000). Transcription factor Nrf2 coordinately regulates a group of oxidative stress-inducible genes in macrophages. Journal of Biological Chemistry, 275, 16023–16029.

    Article  PubMed  CAS  Google Scholar 

  228. Huang, C., Han, Y., Wang, Y., Sun, X., Yan, S., Yeh, E. T., et al. (2009). SENP3 is responsible for HIF-1 transactivation under mild oxidative stress via p300 de-SUMOylation. EMBO Journal, 28, 2748–2762.

    Article  PubMed  CAS  Google Scholar 

  229. Scortegagna, M., Ding, K., Oktay, Y., Gaur, A., Thurmond, F., Yan, L. J., et al. (2003). Multiple organ pathology, metabolic abnormalities and impaired homeostasis of reactive oxygen species in Epas1−/− mice. Nature Genetics, 35, 331–340.

    Article  PubMed  CAS  Google Scholar 

  230. Conrotto, P., Corso, S., Gamberini, S., Comoglio, P. M., & Giordano, S. (2004). Interplay between scatter factor receptors and B plexins controls invasive growth. Oncogene, 23, 5131–5137.

    Article  PubMed  CAS  Google Scholar 

  231. Fischer, O. M., Giordano, S., Comoglio, P. M., & Ullrich, A. (2004). Reactive oxygen species mediate Met receptor transactivation by G protein-coupled receptors and the epidermal growth factor receptor in human carcinoma cells. Journal of Biological Chemistry, 279, 28970–28978.

    Article  PubMed  CAS  Google Scholar 

  232. Klarmann, G. J., Hurt, E. M., Mathews, L. A., Zhang, X., Duhagon, M. A., Mistree, T., et al. (2009). Invasive prostate cancer cells are tumor initiating cells that have a stem cell-like genomic signature. Clinical & Experimental Metastasis, 26, 433–446.

    Article  CAS  Google Scholar 

  233. O'Dell, T. J., Hawkins, R. D., Kandel, E. R., & Arancio, O. (1991). Tests of the roles of two diffusible substances in long-term potentiation: evidence for nitric oxide as a possible early retrograde messenger. Proceedings of the National Academy of Sciences of the United States of America, 88, 11285–11289.

    Article  PubMed  Google Scholar 

  234. Niethammer, P., Grabher, C., Look, A. T., & Mitchison, T. J. (2009). A tissue-scale gradient of hydrogen peroxide mediates rapid wound detection in zebrafish. Nature, 459, 996–999.

    Article  PubMed  CAS  Google Scholar 

  235. Taylor, B. L., Rebbapragada, A., & Johnson, M. S. (2001). The FAD-PAS domain as a sensor for behavioral responses in Escherichia coli. Antioxidants Redox Signaling, 3, 867–879.

    Article  PubMed  CAS  Google Scholar 

  236. Graeber, T. G., Osmanian, C., Jacks, T., Housman, D. E., Koch, C. J., Lowe, S. W., et al. (1996). Hypoxia-mediated selection of cells with diminished apoptotic potential in solid tumours. Nature, 379, 88–91.

    Article  PubMed  CAS  Google Scholar 

  237. Hentze, M. W., & Kuhn, L. C. (1996). Molecular control of vertebrate iron metabolism: mRNA-based regulatory circuits operated by iron, nitric oxide, and oxidative stress. Proceedings of the National Academy of Sciences of the United States of America, 93, 8175–8182.

    Article  PubMed  CAS  Google Scholar 

Download references

Acknowledgments

The authors wish to thank Prof. Lido Calorini, as like as members of their laboratories for suggestions and insightful discussions. Authors are supported by the Italian Association for Cancer Research (AIRC) (grant# IG 8634 to G.P. and grant # IG8797 to P.C.), by Istituto Toscano Tumori and Regione Toscana (TRESOR grants and PorCreo grant to P. C.).

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Giovambattista Pani or Paola Chiarugi.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Pani, G., Galeotti, T. & Chiarugi, P. Metastasis: cancer cell’s escape from oxidative stress. Cancer Metastasis Rev 29, 351–378 (2010). https://doi.org/10.1007/s10555-010-9225-4

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10555-010-9225-4

Keywords

Navigation