Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Review Article
  • Published:

Cellular mechanisms and physiological consequences of redox-dependent signalling

Key Points

  • Oxidants, such as hydrogen peroxide and superoxide anions, can damage cellular components when produced in large amounts. When reactive oxygen species (ROS) are produced at lower, regulated levels, they can function in a diverse array of signalling pathways.

  • Members of the NADPH oxidase (NOX) family are important generators of intracellular ROS and have an important role in reduction–oxidation (redox) signalling.

  • Most redox signalling occurs by the reversible oxidation and reduction of crucial reactive Cys residues. One example is the family of protein Tyr phosphatases, which can be transiently inactivated by oxidation of the catalytic Cys residue.

  • Although it is less well understood than NOX-generated ROS, mitochondrial ROS production seems to be regulated in some manner and can also modulate signalling pathways.

  • Oxidant signalling might be an ancient mode of signal transduction, as it is observed in both plants and animals. Increasing evidence indicates that redox signalling might have a crucial role in the immune response, stem cell biology, cancer and ageing.

Abstract

Reactive oxygen species (ROS), which were originally characterized in terms of their harmful effects on cells and invading microorganisms, are increasingly implicated in various cell fate decisions and signal transduction pathways. The mechanism involved in ROS-dependent signalling involves the reversible oxidation and reduction of specific amino acids, with crucial reactive Cys residues being the most frequent target. In this Review, we discuss the sources of ROS within cells and what is known regarding how intracellular oxidant levels are regulated. We further discuss the recent observations that reduction–oxidation (redox)-dependent regulation has a crucial role in an ever-widening range of biological activities — from immune function to stem cell self-renewal, and from tumorigenesis to ageing.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Sites for the generation of reactive oxygen species in mitochondria.
Figure 2: Intracellular sources of reactive oxygen species.
Figure 3: Reactive oxygen species can function as mediators of intracellular signalling.
Figure 4: The reversible modulation of reactive Cys residues underlies redox-dependent signalling.
Figure 5: The potential beneficial and harmful roles of antioxidants in tumorigenesis.
Figure 6: The multiple ways in which oxidant signalling can affect cellular function.

Similar content being viewed by others

References

  1. Harman, D. Aging: a theory based on free radical and radiation chemistry. J. Gerontol. 11, 298–300 (1956).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  3. Bae, Y. S. et al. Epidermal growth factor (EGF)-induced generation of hydrogen peroxide. Role in EGF receptor-mediated tyrosine phosphorylation. J. Biol. Chem. 272, 217–221 (1997). Shows, together with reference 2, that growth factor stimulation induces a burst of ROS that is required for subsequent signalling.

    CAS  PubMed  Google Scholar 

  4. Woo, H. A. et al. Inactivation of peroxiredoxin I by phosphorylation allows localized H2O2 accumulation for cell signaling. Cell 140, 517–528 (2010).

    CAS  PubMed  Google Scholar 

  5. Murphy, M. P. How mitochondria produce reactive oxygen species. Biochem. J. 417, 1–13 (2009).

    CAS  PubMed  Google Scholar 

  6. Jensen, P. K. Antimycin-insensitive oxidation of succinate and reduced nicotinamide-adenine dinucleotide in electron-transport particles. I. pH dependency and hydrogen peroxide formation. Biochim. Biophys. Acta 122, 157–166 (1966).

    CAS  PubMed  Google Scholar 

  7. Brand, M. D. The sites and topology of mitochondrial superoxide production. Exp. Gerontol. 45, 466–472 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Mailloux, R. J. & Harper, M. E. Mitochondrial proticity and ROS signaling: lessons from the uncoupling proteins. Trends Endocrinol. Metab. 23, 451–458 (2012).

    CAS  PubMed  Google Scholar 

  9. Sena, L. A. & Chandel, N. S. Physiological roles of mitochondrial reactive oxygen species. Mol. Cell 48, 158–167 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Chandel, N. S. et al. Mitochondrial reactive oxygen species trigger hypoxia-induced transcription. Proc. Natl Acad. Sci. USA 95, 11715–11720 (1998). One of the earliest descriptions that mitochondrial oxidants function as signalling molecules.

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Nemoto, S., Takeda, K., Yu, Z. X., Ferrans, V. J. & Finkel, T. Role for mitochondrial oxidants as regulators of cellular metabolism. Mol. Cell. Biol. 20, 7311–7318 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Hampton, M. B., Kettle, A. J. & Winterbourn, C. C. Inside the neutrophil phagosome: oxidants, myeloperoxidase, and bacterial killing. Blood 92, 3007–3017 (1998).

    CAS  PubMed  Google Scholar 

  13. Lo, Y. Y. & Cruz, T. F. Involvement of reactive oxygen species in cytokine and growth factor induction of c-fos expression in chondrocytes. J. Biol. Chem. 270, 11727–11730 (1995).

    CAS  PubMed  Google Scholar 

  14. Rajagopalan, S. et al. Angiotensin II-mediated hypertension in the rat increases vascular superoxide production via membrane NADH/NADPH oxidase activation. Contribution to alterations of vasomotor tone. J. Clin. Invest. 97, 1916–1923 (1996).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Sundaresan, M. et al. Regulation of reactive-oxygen-species generation in fibroblasts by Rac1. Biochem. J. 318, 379–382 (1996).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  17. Suh, Y. A. et al. Cell transformation by the superoxide-generating oxidase Mox1. Nature 401, 79–82 (1999). Describes the cloning of the first non-phagocytic cell member of the NOX superfamily of NADPH oxidases.

    CAS  PubMed  Google Scholar 

  18. Aguirre, J. & Lambeth, J. D. Nox enzymes from fungus to fly to fish and what they tell us about Nox function in mammals. Free Radic. Biol. Med. 49, 1342–1353 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Nakano, Y. et al. Mutation of the Cyba gene encoding p22phox causes vestibular and immune defects in mice. J. Clin. Invest. 118, 1176–1185 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Dixon, S. J. & Stockwell, B. R. The role of iron and reactive oxygen species in cell death. Nature Chem. Biol. 10, 9–17 (2014).

    CAS  Google Scholar 

  21. Kil, I. S. et al. Feedback control of adrenal steroidogenesis via H2O2-dependent, reversible inactivation of peroxiredoxin III in mitochondria. Mol. Cell 46, 584–594 (2012).

    CAS  PubMed  Google Scholar 

  22. Truong, T. H. & Carroll, K. S. Redox regulation of epidermal growth factor receptor signaling through cysteine oxidation. Biochemistry 51, 9954–9965 (2012).

    CAS  PubMed  Google Scholar 

  23. Meng, T. C., Fukada, T. & Tonks, N. K. Reversible oxidation and inactivation of protein tyrosine phosphatases in vivo. Mol. Cell 9, 387–399 (2002). Demonstrates the reversible oxidation and inactivation of an intracellular target by physiological levels of ROS.

    CAS  PubMed  Google Scholar 

  24. Denu, J. M. & Tanner, K. G. Specific and reversible inactivation of protein tyrosine phosphatases by hydrogen peroxide: evidence for a sulfenic acid intermediate and implications for redox regulation. Biochemistry 37, 5633–5642 (1998).

    CAS  PubMed  Google Scholar 

  25. Paulsen, C. E. et al. Peroxide-dependent sulfenylation of the EGFR catalytic site enhances kinase activity. Nature Chem. Biol. 8, 57–64 (2012).

    CAS  Google Scholar 

  26. Frijhoff, J., Dagnell, M., Godfrey, R. & Ostman, A. Regulation of protein tyrosine phosphatase oxidation in cell adhesion and migration. Antioxid. Redox Signal 20, 1994–2010 (2014).

    CAS  PubMed  Google Scholar 

  27. Leslie, N. R. et al. Redox regulation of PI 3-kinase signalling via inactivation of PTEN. EMBO J. 22, 5501–5510 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Kwon, J. et al. Reversible oxidation and inactivation of the tumor suppressor PTEN in cells stimulated with peptide growth factors. Proc. Natl Acad. Sci. USA 101, 16419–16424 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Savitsky, P. A. & Finkel, T. Redox regulation of Cdc25C. J. Biol. Chem. 277, 20535–20540 (2002).

    CAS  PubMed  Google Scholar 

  30. Jeong, W., Bae, S. H., Toledano, M. B. & Rhee, S. G. Role of sulfiredoxin as a regulator of peroxiredoxin function and regulation of its expression. Free Radic. Biol. Med. 53, 447–456 (2012).

    CAS  PubMed  Google Scholar 

  31. Wood, Z. A., Poole, L. B. & Karplus, P. A. Peroxiredoxin evolution and the regulation of hydrogen peroxide signaling. Science 300, 650–653 (2003).

    CAS  PubMed  Google Scholar 

  32. Edgar, R. S. et al. Peroxiredoxins are conserved markers of circadian rhythms. Nature 485, 459–464 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Chen, K., Kirber, M. T., Xiao, H., Yang, Y. & Keaney, J. F. Jr. Regulation of ROS signal transduction by NADPH oxidase 4 localization. J. Cell Biol. 181, 1129–1139 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Drazic, A. & Winter, J. The physiological role of reversible methionine oxidation. Biochim. Biophys. Acta http://dx.doi.org/10.1016/j.bbapap.2014.01.001 (2014).

  35. Erickson, J. R. et al. A dynamic pathway for calcium-independent activation of CaMKII by methionine oxidation. Cell 133, 462–474 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Hung, R. J., Spaeth, C. S., Yesilyurt, H. G. & Terman, J. R. SelR reverses Mical-mediated oxidation of actin to regulate F-actin dynamics. Nature Cell Biol. 15, 1445–1454 (2013).

    CAS  PubMed  Google Scholar 

  37. Lee, B. C. et al. MsrB1 and MICALs regulate actin assembly and macrophage function via reversible stereoselective methionine oxidation. Mol. Cell 51, 397–404 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Lee, C. et al. Redox regulation of OxyR requires specific disulfide bond formation involving a rapid kinetic reaction path. Nature Struct. Mol. Biol. 11, 1179–1185 (2004).

    CAS  Google Scholar 

  40. Putker, M. et al. Redox-dependent control of FOXO/DAF-16 by transportin-1. Mol. Cell 49, 730–742 (2013).

    CAS  PubMed  Google Scholar 

  41. Taguchi, K., Motohashi, H. & Yamamoto, M. Molecular mechanisms of the Keap1–Nrf2 pathway in stress response and cancer evolution. Genes Cells 16, 123–140 (2011).

    CAS  PubMed  Google Scholar 

  42. Scherz-Shouval, R. & Elazar, Z. Monitoring starvation-induced reactive oxygen species formation. Methods Enzymol. 452, 119–130 (2009).

    CAS  PubMed  Google Scholar 

  43. Scherz-Shouval, R. et al. Reactive oxygen species are essential for autophagy and specifically regulate the activity of Atg4. EMBO J. 26, 1749–1760 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Zhang, J. et al. A tuberous sclerosis complex signalling node at the peroxisome regulates mTORC1 and autophagy in response to ROS. Nature Cell Biol. 15, 1186–1196 (2013).

    CAS  PubMed  Google Scholar 

  45. Ha, E. M., Oh, C. T., Bae, Y. S. & Lee, W. J. A direct role for dual oxidase in Drosophila gut immunity. Science 310, 847–850 (2005).

    CAS  PubMed  Google Scholar 

  46. Kumar, A. et al. Commensal bacteria modulate cullin-dependent signaling via generation of reactive oxygen species. EMBO J. 26, 4457–4466 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Jones, R. M. et al. Symbiotic lactobacilli stimulate gut epithelial proliferation via Nox-mediated generation of reactive oxygen species. EMBO J. 32, 3017–3028 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Neish, A. S. et al. Prokaryotic regulation of epithelial responses by inhibition of IκB-α ubiquitination. Science 289, 1560–1563 (2000).

    CAS  PubMed  Google Scholar 

  49. Lee, W. J. Bacterial-modulated host immunity and stem cell activation for gut homeostasis. Genes Dev. 23, 2260–2265 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. West, A. P. et al. TLR signalling augments macrophage bactericidal activity through mitochondrial ROS. Nature 472, 476–480 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Bulua, A. C. et al. Mitochondrial reactive oxygen species promote production of proinflammatory cytokines and are elevated in TNFR1-associated periodic syndrome (TRAPS). J. Exp. Med. 208, 519–533 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Tal, M. C. et al. Absence of autophagy results in reactive oxygen species-dependent amplification of RLR signaling. Proc. Natl Acad. Sci. USA 106, 2770–2775 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Zhou, R., Yazdi, A. S., Menu, P. & Tschopp, J. A role for mitochondria in NLRP3 inflammasome activation. Nature 469, 221–225 (2011).

    Article  CAS  PubMed  Google Scholar 

  54. Sena, L. A. et al. Mitochondria are required for antigen-specific T cell activation through reactive oxygen species signaling. Immunity 38, 225–236 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Zhang, Y. et al. ROS play a critical role in the differentiation of alternatively activated macrophages and the occurrence of tumor-associated macrophages. Cell Res. 23, 898–914 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Kobayashi, C. I. & Suda, T. Regulation of reactive oxygen species in stem cells and cancer stem cells. J. Cell. Physiol. 227, 421–430 (2012).

    CAS  PubMed  Google Scholar 

  57. Urao, N. & Ushio-Fukai, M. Redox regulation of stem/progenitor cells and bone marrow niche. Free Radic. Biol. Med. 54, 26–39 (2013).

    CAS  PubMed  Google Scholar 

  58. Ito, K. et al. Regulation of oxidative stress by ATM is required for self-renewal of haematopoietic stem cells. Nature 431, 997–1002 (2004).

    CAS  PubMed  Google Scholar 

  59. Tothova, Z. et al. FoxOs are critical mediators of hematopoietic stem cell resistance to physiologic oxidative stress. Cell 128, 325–339 (2007).

    CAS  PubMed  Google Scholar 

  60. Miyamoto, K. et al. Foxo3a is essential for maintenance of the hematopoietic stem cell pool. Cell Stem Cell 1, 101–112 (2007). References 58–60 provide convincing evidence for the relationship between redox homeostasis and stem cell self-renewal.

    CAS  PubMed  Google Scholar 

  61. Kops, G. J. et al. Forkhead transcription factor FOXO3a protects quiescent cells from oxidative stress. Nature 419, 316–321 (2002).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  63. Liu, J. et al. Bmi1 regulates mitochondrial function and the DNA damage response pathway. Nature 459, 387–392 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  64. Chatoo, W. et al. The polycomb group gene Bmi1 regulates antioxidant defenses in neurons by repressing p53 pro-oxidant activity. J. Neurosci. 29, 529–542 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  65. Chuikov, S., Levi, B. P., Smith, M. L. & Morrison, S. J. Prdm16 promotes stem cell maintenance in multiple tissues, partly by regulating oxidative stress. Nature Cell Biol. 12, 999–1006 (2010).

    CAS  PubMed  Google Scholar 

  66. Abbas, H. A. et al. Mdm2 is required for survival of hematopoietic stem cells/progenitors via dampening of ROS-induced p53 activity. Cell Stem Cell 7, 606–617 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  67. Le Belle, J. E. et al. Proliferative neural stem cells have high endogenous ROS levels that regulate self-renewal and neurogenesis in a PI3K/Akt-dependant manner. Cell Stem Cell 8, 59–71 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  68. Dickinson, B. C., Peltier, J., Stone, D., Schaffer, D. V. & Chang, C. J. Nox2 redox signaling maintains essential cell populations in the brain. Nature Chem. Biol. 7, 106–112 (2011).

    CAS  Google Scholar 

  69. Morimoto, H. et al. ROS are required for mouse spermatogonial stem cell self-renewal. Cell Stem Cell 12, 774–786 (2013).

    CAS  PubMed  Google Scholar 

  70. Owusu-Ansah, E. & Banerjee, U. Reactive oxygen species prime Drosophila haematopoietic progenitors for differentiation. Nature 461, 537–541 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  71. Li, T. S. & Marban, E. Physiological levels of reactive oxygen species are required to maintain genomic stability in stem cells. Stem Cells 28, 1178–1185 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  72. Funato, Y., Michiue, T., Asashima, M. & Miki, H. The thioredoxin-related redox-regulating protein nucleoredoxin inhibits Wnt-β-catenin signalling through dishevelled. Nature Cell Biol. 8, 501–508 (2006).

    CAS  PubMed  Google Scholar 

  73. Kajla, S. et al. A crucial role for Nox 1 in redox-dependent regulation of Wnt-β-catenin signaling. FASEB J. 26, 2049–2059 (2012).

    CAS  PubMed  Google Scholar 

  74. Hamanaka, R. B. et al. Mitochondrial reactive oxygen species promote epidermal differentiation and hair follicle development. Sci. Signal 6, ra8 (2013).

    PubMed  PubMed Central  Google Scholar 

  75. Duncan, A. W. et al. Integration of Notch and Wnt signaling in hematopoietic stem cell maintenance. Nature Immunol. 6, 314–322 (2005).

    CAS  Google Scholar 

  76. de Keizer, P. L., Burgering, B. M. & Dansen, T. B. Forkhead box o as a sensor, mediator, and regulator of redox signaling. Antioxid. Redox Signal 14, 1093–1106 (2011).

    CAS  PubMed  Google Scholar 

  77. Coant, N. et al. NADPH oxidase 1 modulates WNT and NOTCH1 signaling to control the fate of proliferative progenitor cells in the colon. Mol. Cell. Biol. 30, 2636–2650 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  78. Schroeder, E. A., Raimundo, N. & Shadel, G. S. Epigenetic silencing mediates mitochondria stress-induced longevity. Cell. Metab. 17, 954–964 (2013). The first study to show a link between mROS and epigenetics.

    CAS  PubMed  PubMed Central  Google Scholar 

  79. Lee, J. G., Baek, K., Soetandyo, N. & Ye, Y. Reversible inactivation of deubiquitinases by reactive oxygen species in vitro and in cells. Nature Commun. 4, 1568 (2013).

    Google Scholar 

  80. Cotto-Rios, X. M., Bekes, M., Chapman, J., Ueberheide, B. & Huang, T. T. Deubiquitinases as a signaling target of oxidative stress. Cell Rep. 2, 1475–1484 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  81. Kulathu, Y. et al. Regulation of A20 and other OTU deubiquitinases by reversible oxidation. Nature Commun. 4, 1569 (2013).

    Google Scholar 

  82. The Alpha-Tocopherol Beta Carotene Cancer Prevention Study Group. The effect of vitamin E and β carotene on the incidence of lung cancer and other cancers in male smokers. N. Engl. J. Med. 330, 1029–1035 (1994).

  83. Watson, J. Oxidants, antioxidants and the current incurability of metastatic cancers. Open Biol. 3, 120144 (2013).

    PubMed  PubMed Central  Google Scholar 

  84. Christofk, H. R. et al. The M2 splice isoform of pyruvate kinase is important for cancer metabolism and tumour growth. Nature 452, 230–233 (2008).

    CAS  PubMed  Google Scholar 

  85. Anastasiou, D. et al. Inhibition of pyruvate kinase M2 by reactive oxygen species contributes to cellular antioxidant responses. Science 334, 1278–1283 (2011). An interesting link between cancer metabolism and redox signalling.

    CAS  PubMed  PubMed Central  Google Scholar 

  86. Mathew, R. & White, E. Autophagy, stress, and cancer metabolism: what doesn't kill you makes you stronger. Cold Spring Harb. Symp. Quant. Biol. 76, 389–396 (2011).

    CAS  PubMed  Google Scholar 

  87. Bensaad, K., Cheung, E. C. & Vousden, K. H. Modulation of intracellular ROS levels by TIGAR controls autophagy. EMBO J. 28, 3015–3026 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  88. Mathew, R. et al. Autophagy suppresses tumorigenesis through elimination of p62. Cell 137, 1062–1075 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  89. Lee, I. H. et al. Atg7 modulates p53 activity to regulate cell cycle and survival during metabolic stress. Science 336, 225–228 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  90. Ishikawa, K. et al. ROS-generating mitochondrial DNA mutations can regulate tumor cell metastasis. Science 320, 661–664 (2008).

    CAS  PubMed  Google Scholar 

  91. Irani, K. et al. Mitogenic signaling mediated by oxidants in Ras-transformed fibroblasts. Science 275, 1649–1652 (1997).

    CAS  PubMed  Google Scholar 

  92. Johnson, T. M., Yu, Z. X., Ferrans, V. J., Lowenstein, R. A. & Finkel, T. Reactive oxygen species are downstream mediators of p53-dependent apoptosis. Proc. Natl Acad. Sci. USA 93, 11848–11852 (1996).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  94. Vigneron, A. & Vousden, K. H. p53, ROS and senescence in the control of aging. Aging 2, 471–474 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  95. Lee, A. C. et al. Ras proteins induce senescence by altering the intracellular levels of reactive oxygen species. J. Biol. Chem. 274, 7936–7940 (1999).

    CAS  PubMed  Google Scholar 

  96. DeNicola, G. M. et al. Oncogene-induced Nrf2 transcription promotes ROS detoxification and tumorigenesis. Nature 475, 106–109 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  97. Gorrini, C., Harris, I. S. & Mak, T. W. Modulation of oxidative stress as an anticancer strategy. Nature Rev. Drug Discov. 12, 931–947 (2013).

    CAS  Google Scholar 

  98. Balaban, R. S., Nemoto, S. & Finkel, T. Mitochondria, oxidants, and aging. Cell 120, 483–495 (2005).

    CAS  PubMed  Google Scholar 

  99. Van Remmen, H. et al. Life-long reduction in MnSOD activity results in increased DNA damage and higher incidence of cancer but does not accelerate aging. Physiol. Genom. 16, 29–37 (2003).

    CAS  Google Scholar 

  100. Yang, W., Li, J. & Hekimi, S. A. Measurable increase in oxidative damage due to reduction in superoxide detoxification fails to shorten the life span of long-lived mitochondrial mutants of Caenorhabditis elegans. Genetics 177, 2063–2074 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  101. Schriner, S. E. et al. Extension of murine life span by overexpression of catalase targeted to mitochondria. Science 308, 1909–1911 (2005).

    CAS  PubMed  Google Scholar 

  102. Schulz, T. J. et al. Glucose restriction extends Caenorhabditis elegans life span by inducing mitochondrial respiration and increasing oxidative stress. Cell. Metab. 6, 280–293 (2007). One of the first demonstrations of the concept of hormesis with regard to ageing by showing that in the worm, oxidative stress can extend lifespan.

    CAS  PubMed  Google Scholar 

  103. Zarse, K. et al. Impaired insulin/IGF1 signaling extends life span by promoting mitochondrial L-proline catabolism to induce a transient ROS signal. Cell. Metab. 15, 451–465 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  104. Lapointe, J. & Hekimi, S. Early mitochondrial dysfunction in long-lived Mclk1+/− mice. J. Biol. Chem. 283, 26217–26227 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  105. Liu, X. et al. Evolutionary conservation of the clk-1-dependent mechanism of longevity: loss of mclk1 increases cellular fitness and lifespan in mice. Genes Dev. 19, 2424–2434 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  106. Ishii, N. et al. A mutation in succinate dehydrogenase cytochrome b causes oxidative stress and ageing in nematodes. Nature 394, 694–697 (1998).

    CAS  PubMed  Google Scholar 

  107. Baker, B. M., Nargund, A. M., Sun, T. & Haynes, C. M. Protective coupling of mitochondrial function and protein synthesis via the eIF2α kinase GCN-2. PLoS Genet. 8, e1002760 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  108. Pan, Y., Schroeder, E. A., Ocampo, A., Barrientos, A. & Shadel, G. S. Regulation of yeast chronological life span by TORC1 via adaptive mitochondrial ROS signaling. Cell. Metab. 13, 668–678 (2011). Demonstrates the concept of hormesis in a model organism (yeast), in which the release of mROS might extend rather than shorten lifespan.

    CAS  PubMed  PubMed Central  Google Scholar 

  109. Ristow, M. et al. Antioxidants prevent health-promoting effects of physical exercise in humans. Proc. Natl Acad. Sci. USA 106, 8665–8670 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  110. Cocheme, H. M. et al. Measurement of H2O2 within living Drosophila during aging using a ratiometric mass spectrometry probe targeted to the mitochondrial matrix. Cell. Metab. 13, 340–350 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  111. Dikalov, S. I. & Harrison, D. G. Methods for detection of mitochondrial and cellular reactive oxygen species. Antioxid. Redox Signal 20, 372–382 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  112. Miller, E. W., Albers, A. E., Pralle, A., Isacoff, E. Y. & Chang, C. J. Boronate-based fluorescent probes for imaging cellular hydrogen peroxide. J. Am. Chem. Soc. 127, 16652–16659 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  113. Miller, E. W., Tulyathan, O., Isacoff, E. Y. & Chang, C. J. Molecular imaging of hydrogen peroxide produced for cell signaling. Nature Chem. Biol. 3, 263–267 (2007).

    CAS  Google Scholar 

  114. Belousov, V. V. et al. Genetically encoded fluorescent indicator for intracellular hydrogen peroxide. Nature Methods 3, 281–286 (2006).

    CAS  PubMed  Google Scholar 

  115. Morgan, B., Sobotta, M. C. & Dick, T. P. Measuring EGSH and H2O2 with roGFP2-based redox probes. Free Radic. Biol. Med. 51, 1943–1951 (2011).

    CAS  PubMed  Google Scholar 

  116. Rhee, S. G., Woo, H. A., Kil, I. S. & Bae, S. H. Peroxiredoxin functions as a peroxidase and a regulator and sensor of local peroxides. J. Biol. Chem. 287, 4403–4410 (2012).

    CAS  PubMed  Google Scholar 

  117. Reddi, A. R. & Culotta, V. C. SOD1 integrates signals from oxygen and glucose to repress respiration. Cell 152, 224–235 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  118. Bause, A. S. & Haigis, M. C. SIRT3 regulation of mitochondrial oxidative stress. Exp. Gerontol. 48, 634–639 (2013).

    CAS  PubMed  Google Scholar 

  119. Miller, E. W., Dickinson, B. C. & Chang, C. J. Aquaporin-3 mediates hydrogen peroxide uptake to regulate downstream intracellular signaling. Proc. Natl Acad. Sci. USA 107, 15681–15686 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The authors are grateful to members of the Finkel laboratory for helpful comments and to I. Rovira for help in preparing the manuscript. This work was supported by US National Institutes of Health (NIH) Intramural Funds and The Leducq Foundation.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Toren Finkel.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

PowerPoint slides

Glossary

Complex IV

Together with complexes I–III, a series of large multisubunit protein complexes that are located on the inner mitochondrial membrane. They facilitate the flow of electrons that are initially extracted from substrates such as NADH. The energy that is obtained from the flow of these electrons is used to produce a proton gradient across the inner mitochondrial membrane, which facilitates ATP production.

Proton-motive force

The combination of the electrical and chemical pH gradients across the inner mitochondrial membrane.

Neutrophils

The most abundant type of circulating white blood cell. This phagocytic cell type can engulf and kill a bacterium through the generation of reactive oxygen species.

Chronic granulomatous disease

A rare genetic disease that is characterized by recurring infections. The molecular defect is the inability to generate reactive oxygen species owing to inherited mutations in the NADPH oxidase that are found in immune cells.

Respiratory burst

The rapid, high level release of reactive oxygen species (ROS) following the activation of an immune cell.

Cytochrome P450 enzymes

A large family of related proteins that function as monooxygenases and metabolize a wide range of substances, including steroids, lipids, xenobiotics and toxins.

Peroxisome

A subcellular organelle that is involved in the metabolism of very long-chain fatty acids and branched-chain fatty acids. Shorter fatty acids are metabolized in mitochondria.

Toll-like receptors

(TLRs). Single-pass transmembrane receptors that have a crucial role in innate immunity. Lipopolysaccharide (LPS) is one of many ligands for members of this receptor family.

Retinoic acid-inducible gene I signalling

(RIG-I signalling). A pathway that regulates the cellular response to viral infection on the basis of the ability of RIG-I to directly bind double-stranded RNA.

Inflammasome

A large intracellular, multiprotein complex that has a crucial role in innate immunity. Once activated, the inflammasome leads to caspase 1 activation and the subsequent maturation of various inflammatory cytokines.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Holmström, K., Finkel, T. Cellular mechanisms and physiological consequences of redox-dependent signalling. Nat Rev Mol Cell Biol 15, 411–421 (2014). https://doi.org/10.1038/nrm3801

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrm3801

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing