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:

The role of ROS in tumour development and progression

Abstract

Eukaryotic cells have developed complex systems to regulate the production and response to reactive oxygen species (ROS). Different ROS control diverse aspects of cell behaviour from signalling to death, and deregulation of ROS production and ROS limitation pathways are common features of cancer cells. ROS also function to modulate the tumour environment, affecting the various stromal cells that provide metabolic support, a blood supply and immune responses to the tumour. Although it is clear that ROS play important roles during tumorigenesis, it has been difficult to reliably predict the effect of ROS modulating therapies. We now understand that the responses to ROS are highly complex and dependent on multiple factors, including the types, levels, localization and persistence of ROS, as well as the origin, environment and stage of the tumours themselves. This increasing understanding of the complexity of ROS in malignancies will be key to unlocking the potential of ROS-targeting therapies for cancer treatment.

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

Fig. 1: Generation and metabolism of ROS.
Fig. 2: NADPH and glutathione metabolism.
Fig. 3: ROS and tumour progression: tumour cell survival and invasiveness.
Fig. 4: ROS and tumour progression: interactions with stromal compartments.

Similar content being viewed by others

References

  1. Holmstrom, K. M. & Finkel, T. Cellular mechanisms and physiological consequences of redox-dependent signalling. Nat. Rev. Mol. Cell Biol. 15, 411–421 (2014).

    Article  CAS  PubMed  Google Scholar 

  2. D’Autreaux, B. & Toledano, M. B. ROS as signalling molecules: mechanisms that generate specificity in ROS homeostasis. Nat. Rev. Mol. Cell Biol. 8, 813–824 (2007).

    Article  PubMed  CAS  Google Scholar 

  3. Winterbourn, C. C. Hydrogen peroxide reactivity and specificity in thiol-based cell signalling. Biochem. Soc. Trans. 48, 745–754 (2020).

    Article  CAS  PubMed  Google Scholar 

  4. Kong, H. & Chandel, N. S. Regulation of redox balance in cancer and T cells. J. Biol. Chem. 293, 7499–7507 (2018).

    Article  CAS  PubMed  Google Scholar 

  5. Poyton, R. O., Ball, K. A. & Castello, P. R. Mitochondrial generation of free radicals and hypoxic signaling. Trends Endocrinol. Metab. 20, 332–340 (2009).

    Article  CAS  PubMed  Google Scholar 

  6. Hayes, J. D. & Pulford, D. J. The glutathione S-transferase supergene family: regulation of GST and the contribution of the isoenzymes to cancer chemoprotection and drug resistance. Crit. Rev. Biochem. Mol. Biol. 30, 445–600 (1995).

    Article  CAS  PubMed  Google Scholar 

  7. Townsend, D. M. & Tew, K. D. The role of glutathione-S-transferase in anti-cancer drug resistance. Oncogene 22, 7369–7375 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Chen, L. et al. NADPH production by the oxidative pentose-phosphate pathway supports folate metabolism. Nat. Metab. 1, 404–415 (2019). This study dissects the roles of different NADPH metabolic pathways and determines the importance of the oxPPP in maintaining the NADPH to NADP ratio and supporting folate metabolism.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Fan, J. et al. Quantitative flux analysis reveals folate-dependent NADPH production. Nature 510, 298–302 (2014). By combining deuterium tracing and carbon labelling, this paper shows that in addition to the oxPPP, one-carbon metabolism also makes a significant contribution to the maintenance of NADPH.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Ju, H. Q., Lin, J. F., Tian, T., Xie, D. & Xu, R. H. NADPH homeostasis in cancer: functions, mechanisms and therapeutic implications. Signal Transduct. Target. Ther. 5, 231 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Hoxhaj, G. et al. Direct stimulation of NADP+ synthesis through Akt-mediated phosphorylation of NAD kinase. Science 363, 1088–1092 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Schild, T. et al. NADK is activated by oncogenic signaling to sustain pancreatic ductal adenocarcinoma. Cell Rep. 35, 109238 (2021).

    Article  CAS  PubMed  Google Scholar 

  13. Pollak, N., Niere, M. & Ziegler, M. NAD kinase levels control the NADPH concentration in human cells. J. Biol. Chem. 282, 33562–33571 (2007).

    Article  CAS  PubMed  Google Scholar 

  14. Lerner, F., Niere, M., Ludwig, A. & Ziegler, M. Structural and functional characterization of human NAD kinase. Biochem. Biophys. Res. Commun. 288, 69–74 (2001).

    Article  CAS  PubMed  Google Scholar 

  15. Ding, C. C. et al. MESH1 is a cytosolic NADPH phosphatase that regulates ferroptosis. Nat. Metab. 2, 270–277 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Estrella, M. A. et al. The metabolites NADP+ and NADPH are the targets of the circadian protein Nocturnin (Curled). Nat. Commun. 10, 2367 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  17. Li, S., Yan, T., Yang, J. Q., Oberley, T. D. & Oberley, L. W. The role of cellular glutathione peroxidase redox regulation in the suppression of tumor cell growth by manganese superoxide dismutase. Cancer Res. 60, 3927–3939 (2000).

    CAS  PubMed  Google Scholar 

  18. Zhong, W. et al. Inhibition of cell growth and sensitization to oxidative damage by overexpression of manganese superoxide dismutase in rat glioma cells. Cell Growth Differ. 7, 1175–1186 (1996).

    CAS  PubMed  Google Scholar 

  19. Tran, D. H. et al. Mitochondrial NADP+ is essential for proline biosynthesis during cell growth. Nat. Metab. 3, 571–585 (2021).

    Article  CAS  PubMed  Google Scholar 

  20. Zhu, J. et al. Mitochondrial NADP(H) generation is essential for proline biosynthesis. Science 372, 968–972 (2021). Along with Tran et al. (2021), this study illustrates the importance of mitochondrial NADPH in proline biosynthesis for cell growth and proliferation.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Kuhajda, F. P. et al. Fatty acid synthesis: a potential selective target for antineoplastic therapy. Proc. Natl Acad. Sci. USA 91, 6379–6383 (1994).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Jiang, L. et al. Reductive carboxylation supports redox homeostasis during anchorage-independent growth. Nature 532, 255–258 (2016). This paper shows that IDH1 and IDH2-generated NADPH, although not critical for maintaining survival of cells in 2D cell growth conditions, is important for anchorage-independent survival in spheroids to mitigate mtROS.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. 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).

    Article  CAS  PubMed  Google Scholar 

  24. Hawk, M. A. et al. RIPK1-mediated induction of mitophagy compromises the viability of extracellular-matrix-detached cells. Nat. Cell Biol. 20, 272–284 (2018).

    Article  CAS  PubMed  Google Scholar 

  25. Lewis, C. A. et al. Tracing compartmentalized NADPH metabolism in the cytosol and mitochondria of mammalian cells. Mol. Cell 55, 253–263 (2014). This study dissects the NADPH metabolism pathways in the cytosol and mitochondria in intact cells.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Bedard, K. & Krause, K. H. The NOX family of ROS-generating NADPH oxidases: physiology and pathophysiology. Physiol. Rev. 87, 245–313 (2007).

    Article  CAS  PubMed  Google Scholar 

  27. Ohashi, K., Kawai, S. & Murata, K. Identification and characterization of a human mitochondrial NAD kinase. Nat. Commun. 3, 1248 (2012).

    Article  PubMed  CAS  Google Scholar 

  28. Alshaabi, H. et al. Miro1-mediated mitochondrial positioning supports subcellular redox status. Redox Biol. 38, 101818 (2021).

    Article  CAS  PubMed  Google Scholar 

  29. Horn, A., Raavicharla, S., Shah, S., Cox, D. & Jaiswal, J. K. Mitochondrial fragmentation enables localized signaling required for cell repair. J. Cell Biol. 219, e201909154 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Fukai, T. & Ushio-Fukai, M. Cross-talk between NADPH oxidase and mitochondria: role in ROS signaling and angiogenesis. Cells 9, 1849 (2020).

    Article  CAS  PubMed Central  Google Scholar 

  31. Gianni, D. et al. Novel p47phox-related organizers regulate localized NADPH oxidase 1 (Nox1) activity. Sci. Signal. 2, ra54 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  32. Diaz, B. et al. Tks5-dependent, nox-mediated generation of reactive oxygen species is necessary for invadopodia formation. Sci. Signal. 2, ra53 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  33. Caino, M. C. et al. PI3K therapy reprograms mitochondrial trafficking to fuel tumor cell invasion. Proc. Natl Acad. Sci. USA 112, 8638–8643 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  35. Weinberg, F. et al. Mitochondrial metabolism and ROS generation are essential for Kras-mediated tumorigenicity. Proc. Natl Acad. Sci. USA 107, 8788–8793 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Radisky, D. C. et al. Rac1b and reactive oxygen species mediate MMP-3-induced EMT and genomic instability. Nature 436, 123–127 (2005). This study shows that expression of MMP3 in immortalized breast epithelial cells results in an increase in ROS and induction of EMT.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Nieborowska-Skorska, M. et al. Rac2-MRC-cIII-generated ROS cause genomic instability in chronic myeloid leukemia stem cells and primitive progenitors. Blood 119, 4253–4263 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Tonks, N. K. Redox redux: revisiting PTPs and the control of cell signaling. Cell 121, 667–670 (2005).

    Article  CAS  PubMed  Google Scholar 

  39. Weerapana, E. et al. Quantitative reactivity profiling predicts functional cysteines in proteomes. Nature 468, 790–795 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. van der Reest, J., Lilla, S., Zheng, L., Zanivan, S. & Gottlieb, E. Proteome-wide analysis of cysteine oxidation reveals metabolic sensitivity to redox stress. Nat. Commun. 9, 1581 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  41. O’Hagan, H. M. et al. Oxidative damage targets complexes containing DNA methyltransferases, SIRT1, and polycomb members to promoter CpG Islands. Cancer Cell 20, 606–619 (2011).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  42. Ellenbroek, S. I. & Collard, J. G. Rho GTPases: functions and association with cancer. Clin. Exp. Metastasis 24, 657–672 (2007).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Heyworth, P. G., Knaus, U. G., Settleman, J., Curnutte, J. T. & Bokoch, G. M. Regulation of NADPH oxidase activity by Rac GTPase activating protein(s). Mol. Biol. Cell 4, 1217–1223 (1993).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Myant, K. B. et al. ROS production and NF-κB activation triggered by RAC1 facilitate WNT-driven intestinal stem cell proliferation and colorectal cancer initiation. Cell Stem Cell 12, 761–773 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Matos, P., Collard, J. G. & Jordan, P. Tumor-related alternatively spliced Rac1b is not regulated by Rho-GDP dissociation inhibitors and exhibits selective downstream signaling. J. Biol. Chem. 278, 50442–50448 (2003).

    Article  CAS  PubMed  Google Scholar 

  47. Hodis, E. et al. A landscape of driver mutations in melanoma. Cell 150, 251–263 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Krauthammer, M. et al. Exome sequencing identifies recurrent somatic RAC1 mutations in melanoma. Nat. Genet. 44, 1006–1014 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Zhou, C. et al. The Rac1 splice form Rac1b promotes K-ras-induced lung tumorigenesis. Oncogene 32, 903–909 (2013).

    Article  CAS  PubMed  Google Scholar 

  50. Sullivan, L. B. & Chandel, N. S. Mitochondrial reactive oxygen species and cancer. Cancer Metab. 2, 17 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  51. Pak, V. V. et al. Ultrasensitive genetically encoded indicator for hydrogen peroxide identifies roles for the oxidant in cell migration and mitochondrial function. Cell Metab. 31, 642–653.e6 (2020). Using an ultrasensitive probe for H2O2, this study reveals the subcellular localization of ROS and the importance of an intracellular ROS gradient in controlling cell protrusions for migration.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Kim, H. S. et al. SIRT3 is a mitochondria-localized tumor suppressor required for maintenance of mitochondrial integrity and metabolism during stress. Cancer Cell 17, 41–52 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Kong, H. et al. Metabolic determinants of cellular fitness dependent on mitochondrial reactive oxygen species. Sci. Adv. 6, eabb7272 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Rani, V., Neumann, C. A., Shao, C. & Tischfield, J. A. Prdx1 deficiency in mice promotes tissue specific loss of heterozygosity mediated by deficiency in DNA repair and increased oxidative stress. Mutat. Res. 735, 39–45 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Neumann, C. A. et al. Essential role for the peroxiredoxin Prdx1 in erythrocyte antioxidant defence and tumour suppression. Nature 424, 561–565 (2003).

    Article  CAS  PubMed  Google Scholar 

  56. Elchuri, S. et al. CuZnSOD deficiency leads to persistent and widespread oxidative damage and hepatocarcinogenesis later in life. Oncogene 24, 367–380 (2005).

    Article  CAS  PubMed  Google Scholar 

  57. Busuttil, R. A. et al. Organ-specific increase in mutation accumulation and apoptosis rate in CuZn-superoxide dismutase-deficient mice. Cancer Res. 65, 11271–11275 (2005).

    Article  CAS  PubMed  Google Scholar 

  58. 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. Genomics 16, 29–37 (2003).

    Article  PubMed  CAS  Google Scholar 

  59. van de Wetering, C. I., Coleman, M. C., Spitz, D. R., Smith, B. J. & Knudson, C. M. Manganese superoxide dismutase gene dosage affects chromosomal instability and tumor onset in a mouse model of T cell lymphoma. Free Radic. Biol. Med. 44, 1677–1686 (2008).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  60. Chu, F. F. et al. Bacteria-induced intestinal cancer in mice with disrupted Gpx1 and Gpx2 genes. Cancer Res. 64, 962–968 (2004).

    Article  CAS  PubMed  Google Scholar 

  61. Barrett, C. W. et al. Tumor suppressor function of the plasma glutathione peroxidase gpx3 in colitis-associated carcinoma. Cancer Res. 73, 1245–1255 (2013).

    Article  CAS  PubMed  Google Scholar 

  62. Chatterjee, A. & Gupta, S. The multifaceted role of glutathione S-transferases in cancer. Cancer Lett. 433, 33–42 (2018).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  64. Schieber, M. & Chandel, N. S. ROS function in redox signaling and oxidative stress. Curr. Biol. 24, R453–R462 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Jiang, X., Stockwell, B. R. & Conrad, M. Ferroptosis: mechanisms, biology and role in disease. Nat. Rev. Mol. Cell Biol. 22, 266–282 (2021).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  66. Hangauer, M. J. et al. Drug-tolerant persister cancer cells are vulnerable to GPX4 inhibition. Nature 551, 247–250 (2017). This study shows that drug-resistant tumour cells are vulnerable to GPX4 loss, which leads to lipid peroxidation, ferroptotic cell death and a lowered rate of tumour relapse in vivo.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Kuang, F., Liu, J., Tang, D. & Kang, R. Oxidative damage and antioxidant defense in ferroptosis. Front. Cell Dev. Biol. 8, 586578 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  68. Awasthi, Y. C. et al. Regulation of 4-hydroxynonenal-mediated signaling by glutathione S-transferases. Free Radic. Biol. Med. 37, 607–619 (2004).

    Article  CAS  PubMed  Google Scholar 

  69. Viswanathan, V. S. et al. Dependency of a therapy-resistant state of cancer cells on a lipid peroxidase pathway. Nature 547, 453–457 (2017). Along with Hangauer et al. (2017), this study shows that therapy-resistant cancers acquire a mesenchymal state and are dependent on GPX4 to prevent ferroptosis.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Dai, E. et al. Ferroptotic damage promotes pancreatic tumorigenesis through a TMEM173/STING-dependent DNA sensor pathway. Nat. Commun. 11, 6339 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Asantewaa, G. & Harris, I. S. Glutathione and its precursors in cancer. Curr. Opin. Biotechnol. 68, 292–299 (2021).

    Article  CAS  PubMed  Google Scholar 

  72. Trachootham, D. et al. Selective killing of oncogenically transformed cells through a ROS-mediated mechanism by β-phenylethyl isothiocyanate. Cancer Cell 10, 241–252 (2006).

    Article  CAS  PubMed  Google Scholar 

  73. Lien, E. C. et al. Glutathione biosynthesis is a metabolic vulnerability in PI(3)K/Akt-driven breast cancer. Nat. Cell Biol. 18, 572–578 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Harris, I. S. et al. Glutathione and thioredoxin antioxidant pathways synergize to drive cancer initiation and progression. Cancer Cell 27, 211–222 (2015).

    Article  CAS  PubMed  Google Scholar 

  75. Yoo, H. C. et al. A variant of SLC1A5 is a mitochondrial glutamine transporter for metabolic reprogramming in cancer cells. Cell Metab. 31, 267–283.e12 (2020).

    Article  CAS  PubMed  Google Scholar 

  76. Cramer, S. L. et al. Systemic depletion of l-cyst(e)ine with cyst(e)inase increases reactive oxygen species and suppresses tumor growth. Nat. Med. 23, 120–127 (2017).

    Article  CAS  PubMed  Google Scholar 

  77. Ducker, G. S. et al. Human SHMT inhibitors reveal defective glycine import as a targetable metabolic vulnerability of diffuse large B-cell lymphoma. Proc. Natl Acad. Sci. USA 114, 11404–11409 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Badgley, M. A. et al. Cysteine depletion induces pancreatic tumor ferroptosis in mice. Science 368, 85–89 (2020). This study shows that pancreatic ductal adenocarcinoma cells increase cysteine uptake to manage high ROS levels, and preventing cysteine import induces ferroptosis in tumour cells and impedes tumour growth.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Lim, J. K. M. et al. Cystine/glutamate antiporter xCT (SLC7A11) facilitates oncogenic RAS transformation by preserving intracellular redox balance. Proc. Natl Acad. Sci. USA 116, 9433–9442 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Daher, B. et al. Genetic ablation of the cystine transporter xCT in PDAC cells inhibits mTORC1, growth, survival, and tumor formation via nutrient and oxidative stresses. Cancer Res. 79, 3877–3890 (2019).

    Article  CAS  PubMed  Google Scholar 

  81. Dolma, S., Lessnick, S. L., Hahn, W. C. & Stockwell, B. R. Identification of genotype-selective antitumor agents using synthetic lethal chemical screening in engineered human tumor cells. Cancer Cell 3, 285–296 (2003).

    Article  CAS  PubMed  Google Scholar 

  82. LeBoeuf, S. E. et al. Activation of oxidative stress response in cancer generates a druggable dependency on exogenous non-essential amino acids. Cell Metab. 31, 339–350.e4 (2020).

    Article  CAS  PubMed  Google Scholar 

  83. Patra, K. C. & Hay, N. The pentose phosphate pathway and cancer. Trends Biochem. Sci. 39, 347–354 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Bartrons, R. et al. Fructose 2,6-bisphosphate in cancer cell metabolism. Front. Oncol. 8, 331 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  85. Peralta, D. et al. A proton relay enhances H2O2 sensitivity of GAPDH to facilitate metabolic adaptation. Nat. Chem. Biol. 11, 156–163 (2015).

    Article  CAS  PubMed  Google Scholar 

  86. Anastasiou, D. et al. Inhibition of pyruvate kinase M2 by reactive oxygen species contributes to cellular antioxidant responses. Science 334, 1278–1283 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Yamamoto, T. et al. Reduced methylation of PFKFB3 in cancer cells shunts glucose towards the pentose phosphate pathway. Nat. Commun. 5, 3480 (2014).

    Article  PubMed  CAS  Google Scholar 

  88. Dasgupta, S. et al. Metabolic enzyme PFKFB4 activates transcriptional coactivator SRC-3 to drive breast cancer. Nature 556, 249–254 (2018). This study shows that, in addition to its enzymatic activity, PFKFB4 can also modulate metabolism by regulating gene transcription.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Goidts, V. et al. RNAi screening in glioma stem-like cells identifies PFKFB4 as a key molecule important for cancer cell survival. Oncogene 31, 3235–3243 (2012).

    Article  CAS  PubMed  Google Scholar 

  90. Ros, S. et al. 6-Phosphofructo-2-kinase/fructose-2,6-biphosphatase 4 is essential for p53-null cancer cells. Oncogene 36, 3287–3299 (2017).

    Article  CAS  PubMed  Google Scholar 

  91. Strohecker, A. M. et al. Identification of 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase as a novel autophagy regulator by high content shRNA screening. Oncogene 34, 5662–5676 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Bensaad, K. et al. TIGAR, a p53-inducible regulator of glycolysis and apoptosis. Cell 126, 107–120 (2006).

    Article  CAS  PubMed  Google Scholar 

  93. Cheung, E. C. et al. TIGAR is required for efficient intestinal regeneration and tumorigenesis. Dev. Cell 25, 463–477 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Lui, V. W. et al. Inhibition of c-Met downregulates TIGAR expression and reduces NADPH production leading to cell death. Oncogene 30, 1127–1134 (2011).

    Article  CAS  PubMed  Google Scholar 

  95. Yin, L., Kosugi, M. & Kufe, D. Inhibition of the MUC1-C oncoprotein induces multiple myeloma cell death by down-regulating TIGAR expression and depleting NADPH. Blood 119, 810–816 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Wanka, C., Steinbach, J. P. & Rieger, J. Tp53-induced glycolysis and apoptosis regulator (TIGAR) protects glioma cells from starvation-induced cell death by up-regulating respiration and improving cellular redox homeostasis. J. Biol. Chem. 287, 33436–33446 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Xiao, G. et al. B-cell-specific diversion of glucose carbon utilization reveals a unique vulnerability in B cell malignancies. Cell 173, 470–484.e18 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Li, B. et al. Fructose-1,6-bisphosphatase opposes renal carcinoma progression. Nature 513, 251–255 (2014). This study shows an additional activity of FBP1 in modulating metabolism by regulating transcription.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Ghergurovich, J. M. et al. Glucose-6-phosphate dehydrogenase is not essential for K-Ras-driven tumor growth or metastasis. Cancer Res. 80, 3820–3829 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Son, J. et al. Glutamine supports pancreatic cancer growth through a KRAS-regulated metabolic pathway. Nature 496, 101–105 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Wang, Y. P. et al. Arginine methylation of MDH1 by CARM1 inhibits glutamine metabolism and suppresses pancreatic cancer. Mol. Cell 64, 673–687 (2016).

    Article  CAS  PubMed  Google Scholar 

  102. Maddocks, O. D. K. et al. Modulating the therapeutic response of tumours to dietary serine and glycine starvation. Nature 544, 372–376 (2017).

    Article  CAS  PubMed  Google Scholar 

  103. Liu, X. et al. Cystine transporter regulation of pentose phosphate pathway dependency and disulfide stress exposes a targetable metabolic vulnerability in cancer. Nat. Cell Biol. 22, 476–486 (2020). This paper describes the importance of the PPP in supporting cystine and cysteine metabolism.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Liao, R. et al. ME1 promotes basal-like breast cancer progression and associates with poor prognosis. Sci. Rep. 8, 16743 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  105. Fernandes, L. M. et al. Malic enzyme 1 (ME1) is pro-oncogenic in ApcMin/+ mice. Sci. Rep. 8, 14268 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  106. Dey, P. et al. Genomic deletion of malic enzyme 2 confers collateral lethality in pancreatic cancer. Nature 542, 119–123 (2017). This paper shows that owing to the proximity of SMAD4 and ME2 in the genome, SMAD4-mutated pancreatic ductal adenocarcinoma cells also show loss of ME2, and are therefore more dependent on ME3 to generate NADPH. Consequently, deletion of ME3 can selectively kill these ME2-deficient tumour cells.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Lu, Y. X. et al. ME1 regulates NADPH homeostasis to promote gastric cancer growth and metastasis. Cancer Res. 78, 1972–1985 (2018).

    Article  CAS  PubMed  Google Scholar 

  108. Ye, J. et al. Serine catabolism regulates mitochondrial redox control during hypoxia. Cancer Discov. 4, 1406–1417 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Koseki, J. et al. Enzymes of the one-carbon folate metabolism as anticancer targets predicted by survival rate analysis. Sci. Rep. 8, 303 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  110. Nilsson, R. et al. Metabolic enzyme expression highlights a key role for MTHFD2 and the mitochondrial folate pathway in cancer. Nat. Commun. 5, 3128 (2014).

    Article  PubMed  CAS  Google Scholar 

  111. Pikman, Y. et al. Targeting MTHFD2 in acute myeloid leukemia. J. Exp. Med. 213, 1285–1306 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Leca, J., Fortin, J. & Mak, T. W. Illuminating the cross-talk between tumor metabolism and immunity in IDH-mutated cancers. Curr. Opin. Biotechnol. 68, 181–185 (2021).

    Article  CAS  PubMed  Google Scholar 

  113. Altenberg, B. & Greulich, K. O. Genes of glycolysis are ubiquitously overexpressed in 24 cancer classes. Genomics 84, 1014–1020 (2004).

    Article  CAS  PubMed  Google Scholar 

  114. Tan, F. et al. Identification of isocitrate dehydrogenase 1 as a potential diagnostic and prognostic biomarker for non-small cell lung cancer by proteomic analysis. Mol. Cell Proteom. 11, M111 008821 (2012).

    Article  CAS  Google Scholar 

  115. Calvert, A. E. et al. Cancer-associated IDH1 promotes growth and resistance to targeted therapies in the absence of mutation. Cell Rep. 19, 1858–1873 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Geiger, T., Madden, S. F., Gallagher, W. M., Cox, J. & Mann, M. Proteomic portrait of human breast cancer progression identifies novel prognostic markers. Cancer Res. 72, 2428–2439 (2012).

    Article  CAS  PubMed  Google Scholar 

  117. Barnabas, G. D. et al. Serine biosynthesis is a metabolic vulnerability in IDH2-driven breast cancer progression. Cancer Res. 81, 1443–1456 (2021).

    Article  CAS  PubMed  Google Scholar 

  118. Tsang, Y. H. et al. Functional annotation of rare gene aberration drivers of pancreatic cancer. Nat. Commun. 7, 10500 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Yau, E. H. et al. Genome-wide CRISPR screen for essential cell growth mediators in mutant KRAS colorectal cancers. Cancer Res. 77, 6330–6339 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  120. Wild, A. C., Moinova, H. R. & Mulcahy, R. T. Regulation of γ-glutamylcysteine synthetase subunit gene expression by the transcription factor Nrf2. J. Biol. Chem. 274, 33627–33636 (1999).

    Article  CAS  PubMed  Google Scholar 

  121. Mitsuishi, Y. et al. Nrf2 redirects glucose and glutamine into anabolic pathways in metabolic reprogramming. Cancer Cell 22, 66–79 (2012).

    Article  CAS  PubMed  Google Scholar 

  122. Thimmulappa, R. K. et al. Identification of Nrf2-regulated genes induced by the chemopreventive agent sulforaphane by oligonucleotide microarray. Cancer Res. 62, 5196–5203 (2002).

    CAS  PubMed  Google Scholar 

  123. Rojo de la Vega, M., Chapman, E. & Zhang, D. D. NRF2 and the hallmarks of cancer. Cancer Cell 34, 21–43 (2018).

    Article  CAS  PubMed  Google Scholar 

  124. Marhenke, S. et al. Activation of nuclear factor E2-related factor 2 in hereditary tyrosinemia type 1 and its role in survival and tumor development. Hepatology 48, 487–496 (2008).

    Article  CAS  PubMed  Google Scholar 

  125. Ramos-Gomez, M. et al. Sensitivity to carcinogenesis is increased and chemoprotective efficacy of enzyme inducers is lost in nrf2 transcription factor-deficient mice. Proc. Natl Acad. Sci. USA 98, 3410–3415 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Iida, K. et al. Nrf2 and p53 cooperatively protect against BBN-induced urinary bladder carcinogenesis. Carcinogenesis 28, 2398–2403 (2007).

    Article  CAS  PubMed  Google Scholar 

  127. DeNicola, G. M. et al. Oncogene-induced Nrf2 transcription promotes ROS detoxification and tumorigenesis. Nature 475, 106–109 (2011). This study shows that various oncogenes, such as mutant KRAS and BRAF and activated MYC, increase NRF2 expression to induce antioxidant pathways, and deletion of NRF2 impairs tumorigenesis in a mouse model of pancreatic ductal adenocarcinoma.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Romero, R. et al. Keap1 loss promotes Kras-driven lung cancer and results in dependence on glutaminolysis. Nat. Med. 23, 1362–1368 (2017). This study shows that tumours that have increased NRF2 expression are dependent on increased glutaminolysis for antioxidant defence.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Cancer Genome Atlas Research Network. Comprehensive genomic characterization of squamous cell lung cancers. Nature 489, 519–525 (2012).

    Article  CAS  Google Scholar 

  130. Lim, J. K. M. & Leprivier, G. The impact of oncogenic RAS on redox balance and implications for cancer development. Cell Death Dis. 10, 955 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Vafa, O. et al. c-Myc can induce DNA damage, increase reactive oxygen species, and mitigate p53 function: a mechanism for oncogene-induced genetic instability. Mol. Cell 9, 1031–1044 (2002).

    Article  CAS  PubMed  Google Scholar 

  132. Fox, D. B. et al. NRF2 activation promotes the recurrence of dormant tumour cells through regulation of redox and nucleotide metabolism. Nat. Metab. 2, 318–334 (2020). This study shows that the survival of dormant tumour cells after therapy depends on NRF2 activity to support redox homeostasis and nucleotide metabolism, and higher NRF2 expression can accelerate the recurrence of these dormant cells.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Humpton, T. J. & Vousden, K. H. Regulation of cellular metabolism and hypoxia by p53. Cold Spring Harb. Perspect. Med. 6, a026146 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  134. Sablina, A. A. et al. The antioxidant function of the p53 tumor suppressor. Nat. Med. 11, 1306–1313 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Kang, M. Y. et al. The critical role of catalase in prooxidant and antioxidant function of p53. Cell Death Differ. 20, 117–129 (2013).

    Article  CAS  PubMed  Google Scholar 

  136. Jiang, L. et al. Ferroptosis as a p53-mediated activity during tumour suppression. Nature 520, 57–62 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Liu, G. & Chen, X. The ferredoxin reductase gene is regulated by the p53 family and sensitizes cells to oxidative stress-induced apoptosis. Oncogene 21, 7195–7204 (2002).

    Article  CAS  PubMed  Google Scholar 

  138. Li, P. F., Dietz, R. & von Harsdorf, R. p53 regulates mitochondrial membrane potential through reactive oxygen species and induces cytochrome c-independent apoptosis blocked by Bcl-2. EMBO J. 18, 6027–6036 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  139. Maxwell, S. A. & Rivera, A. Proline oxidase induces apoptosis in tumor cells, and its expression is frequently absent or reduced in renal carcinomas. J. Biol. Chem. 278, 9784–9789 (2003).

    Article  CAS  PubMed  Google Scholar 

  140. Humpton, T. J. et al. p53-mediated redox control promotes liver regeneration and maintains liver function in response to CCl4. Cell Death Differ. https://doi.org/10.1038/s41418-021-00871-3 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  141. Huo, Y. et al. Protective role of p53 in acetaminophen hepatotoxicity. Free Radic. Biol. Med. 106, 111–117 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  142. Humpton, T. J., Hock, A. K., Maddocks, O. D. K. & Vousden, K. H. p53-mediated adaptation to serine starvation is retained by a common tumour-derived mutant. Cancer Metab. 6, 18 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  143. Chaffer, C. L. & Weinberg, R. A. A perspective on cancer cell metastasis. Science 331, 1559–1564 (2011).

    Article  CAS  PubMed  Google Scholar 

  144. Gianni, D., Taulet, N., DerMardirossian, C. & Bokoch, G. M. c-Src-mediated phosphorylation of NoxA1 and Tks4 induces the reactive oxygen species (ROS)-dependent formation of functional invadopodia in human colon cancer cells. Mol. Biol. Cell 21, 4287–4298 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Sadok, A. et al. NADPH oxidase 1 controls the persistence of directed cell migration by a Rho-dependent switch of α2/α3 integrins. Mol. Cell Biol. 29, 3915–3928 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Park, M. T. et al. Novel signaling axis for ROS generation during K-Ras-induced cellular transformation. Cell Death Differ. 21, 1185–1197 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  147. Satooka, H. & Hara-Chikuma, M. Aquaporin-3 controls breast cancer cell migration by regulating hydrogen peroxide transport and its downstream cell signaling. Mol. Cell Biol. 36, 1206–1218 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  148. 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).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  149. Chow, P. H., Bowen, J. & Yool, A. J. Combined systematic review and transcriptomic analyses of mammalian aquaporin classes 1 to 10 as biomarkers and prognostic indicators in diverse cancers. Cancers 12, 1911 (2020).

    Article  CAS  PubMed Central  Google Scholar 

  150. Nelson, K. K. & Melendez, J. A. Mitochondrial redox control of matrix metalloproteinases. Free Radic. Biol. Med. 37, 768–784 (2004).

    Article  CAS  PubMed  Google Scholar 

  151. Shinohara, M. et al. Reactive oxygen generated by NADPH oxidase 1 (Nox1) contributes to cell invasion by regulating matrix metalloprotease-9 production and cell migration. J. Biol. Chem. 285, 4481–4488 (2010).

    Article  CAS  PubMed  Google Scholar 

  152. Mori, K. et al. A mitochondrial ROS pathway controls matrix metalloproteinase 9 levels and invasive properties in RAS-activated cancer cells. FEBS J. 286, 459–478 (2019).

    Article  CAS  PubMed  Google Scholar 

  153. Connor, K. M. et al. Manganese superoxide dismutase enhances the invasive and migratory activity of tumor cells. Cancer Res. 67, 10260–10267 (2007).

    Article  CAS  PubMed  Google Scholar 

  154. Zhang, H. J. et al. Activation of matrix metalloproteinase-2 by overexpression of manganese superoxide dismutase in human breast cancer MCF-7 cells involves reactive oxygen species. J. Biol. Chem. 277, 20919–20926 (2002).

    Article  CAS  PubMed  Google Scholar 

  155. Hemachandra, L. P. et al. Mitochondrial superoxide dismutase has a protumorigenic role in ovarian clear cell carcinoma. Cancer Res. 75, 4973–4984 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Malafa, M., Margenthaler, J., Webb, B., Neitzel, L. & Christophersen, M. MnSOD expression is increased in metastatic gastric cancer. J. Surg. Res. 88, 130–134 (2000).

    Article  CAS  PubMed  Google Scholar 

  157. Thomas, P. A. et al. Immunohistochemical characterization of antioxidant enzymes in human breast cancer. Pathol. Oncol. Res. 3, 278–286 (1997).

    Article  CAS  PubMed  Google Scholar 

  158. Arnandis, T. et al. Oxidative stress in cells with extra centrosomes drives non-cell-autonomous invasion. Dev. Cell 47, 409–424.e409 (2018). This study shows that extra centrosomes in mammary cancer cells induce an early NOX1-mediated ROS increase, which drives secretion of pro-metastatic factors.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  159. Lignitto, L. et al. Nrf2 activation promotes lung cancer metastasis by inhibiting the degradation of bach1. Cell 178, 316–329.e18 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  160. Wiel, C. et al. BACH1 stabilization by antioxidants stimulates lung cancer metastasis. Cell 178, 330–345.e22 (2019). Along with Lignitto et al. (2019), this paper shows that in non-small-cell lung cancer, addition of an antioxidant or increased NRF2 activity stabilizes BACH1, increasing BACH-regulated glycolysis and pro-metastatic gene expression and, thereby, promoting metastasis.

    Article  CAS  PubMed  Google Scholar 

  161. Brabletz, T. To differentiate or not — routes towards metastasis. Nat. Rev. Cancer 12, 425–436 (2012).

    Article  CAS  PubMed  Google Scholar 

  162. Lambert, A. W. & Weinberg, R. A. Linking EMT programmes to normal and neoplastic epithelial stem cells. Nat. Rev. Cancer 21, 325–338 (2021).

    Article  CAS  PubMed  Google Scholar 

  163. Chaffer, C. L., San Juan, B. P., Lim, E. & Weinberg, R. A. EMT, cell plasticity and metastasis. Cancer Metastasis Rev. 35, 645–654 (2016).

    Article  PubMed  Google Scholar 

  164. Pastushenko, I. et al. Fat1 deletion promotes hybrid EMT state, tumour stemness and metastasis. Nature 589, 448–455 (2021).

    Article  CAS  PubMed  Google Scholar 

  165. Wu, J. B. et al. Monoamine oxidase A mediates prostate tumorigenesis and cancer metastasis. J. Clin. Invest. 124, 2891–2908 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  166. Hudson, J. et al. p66ShcA promotes breast cancer plasticity by inducing an epithelial-to-mesenchymal transition. Mol. Cell Biol. 34, 3689–3701 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  167. Kesanakurti, D. et al. A novel interaction of PAK4 with PPARγ to regulate Nox1 and radiation-induced epithelial-to-mesenchymal transition in glioma. Oncogene 36, 5309–5320 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  168. Cheung, E. C. et al. Dynamic ROS control by TIGAR regulates the initiation and progression of pancreatic cancer. Cancer Cell 37, 168–182.e4 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  169. Li, J. et al. Combined treatment with N-acetylcysteine and gefitinib overcomes drug resistance to gefitinib in NSCLC cell line. Cancer Med. 9, 1495–1502 (2020).

    Article  CAS  PubMed  Google Scholar 

  170. Mani, S. A. et al. The epithelial–mesenchymal transition generates cells with properties of stem cells. Cell 133, 704–715 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  171. Diehn, M. et al. Association of reactive oxygen species levels and radioresistance in cancer stem cells. Nature 458, 780–783 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  172. Oren, Y. et al. Cycling cancer persister cells arise from lineages with distinct programs. Nature 596, 576–582 (2021). Along with Diehn et al. (2009), this study examines the importance of ROS regulation in supporting cancer cells that can evade therapy, as CSCs or cycling persister cells.

    Article  CAS  PubMed  Google Scholar 

  173. Schafer, Z. T. et al. Antioxidant and oncogene rescue of metabolic defects caused by loss of matrix attachment. Nature 461, 109–113 (2009). This paper shows that matrix detachment of mammary cells decreases glucose uptake and lowers flux into the oxPPP, leading to increased ROS and cell death by anoikis. This is rescued by overexpression of the oncogene ERBB2.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  174. Zhang, Y. et al. Upregulation of antioxidant capacity and nucleotide precursor availability suffices for oncogenic transformation. Cell Metab. 33, 94–109.e8 (2021).

    Article  CAS  PubMed  Google Scholar 

  175. Bueno, M. J. et al. Essentiality of fatty acid synthase in the 2D to anchorage-independent growth transition in transforming cells. Nat. Commun. 10, 5011 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  176. Labuschagne, C. F., Cheung, E. C., Blagih, J., Domart, M. C. & Vousden, K. H. Cell clustering promotes a metabolic switch that supports metastatic colonization. Cell Metab. 30, 720–734.e5 (2019). This study shows that cell clustering following detachment maintains cancer cell survival by promoting mitophagy to clear damaged mitochondria and limit ROS production.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  177. Piskounova, E. et al. Oxidative stress inhibits distant metastasis by human melanoma cells. Nature 527, 186–191 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  178. Le Gal, K. et al. Antioxidants can increase melanoma metastasis in mice. Sci. Transl Med. 7, 308re308 (2015).

    Google Scholar 

  179. Sayin, V. I. et al. Antioxidants accelerate lung cancer progression in mice. Sci. Transl Med. 6, 221ra215 (2014). Along with Piskounova et al. (2015) and Le Gal et al. (2015), this paper shows that systemic treatment with antioxidants increases survival of CTCs in the blood and promotes distant metastasis in melanoma and lung cancer mouse models.

    Article  CAS  Google Scholar 

  180. Tasdogan, A. et al. Metabolic heterogeneity confers differences in melanoma metastatic potential. Nature 577, 115–120 (2020).

    Article  CAS  PubMed  Google Scholar 

  181. Zheng, Y. et al. Expression of β-globin by cancer cells promotes cell survival during blood-borne dissemination. Nat. Commun. 8, 14344 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  182. Ubellacker, J. M. et al. Lymph protects metastasizing melanoma cells from ferroptosis. Nature 585, 113–118 (2020). This paper shows that compared with blood, melanoma cells in the lymph display lower ROS levels and less ferroptosis, reflective of higher oleic acid and lower iron levels in the lymph. This leads to increased tumour cell survival and enhanced metastatic capacity.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  183. Goh, J. et al. Mitochondrial targeted catalase suppresses invasive breast cancer in mice. BMC Cancer 11, 191 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  184. van Rossen, M. E. et al. Scavenging of reactive oxygen species leads to diminished peritoneal tumor recurrence. Cancer Res. 60, 5625–5629 (2000).

    PubMed  Google Scholar 

  185. van Rossen, M. E. et al. Red blood cells inhibit tumour cell adhesion to the peritoneum. Br. J. Surg. 86, 509–513 (1999).

    Article  PubMed  Google Scholar 

  186. van den Tol, P. M. et al. Reduction of peritoneal trauma by using nonsurgical gauze leads to less implantation metastasis of spilled tumor cells. Ann. Surg. 227, 242–248 (1998).

    Article  PubMed  PubMed Central  Google Scholar 

  187. Porporato, P. E. et al. A mitochondrial switch promotes tumor metastasis. Cell Rep. 8, 754–766 (2014). This study shows that an increase in mtROS, either by partial electron transport chain inhibition or increased mitochondrial activity, facilitates distant metastasis.

    Article  CAS  PubMed  Google Scholar 

  188. Ishikawa, K. et al. ROS-generating mitochondrial DNA mutations can regulate tumor cell metastasis. Science 320, 661–664 (2008). This paper shows that mitochondrial DNA mutations in complex I lead to an increase in mtROS and metastasis of tumour cells, whereas treatment with ROS scavengers reduces the metastatic potential of these cells.

    Article  CAS  PubMed  Google Scholar 

  189. Chourasia, A. H. et al. Mitophagy defects arising from BNip3 loss promote mammary tumor progression to metastasis. EMBO Rep. 16, 1145–1163 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  190. Kenny, T. C., Craig, A. J., Villanueva, A. & Germain, D. Mitohormesis primes tumor invasion and metastasis. Cell Rep. 27, 2292–2303.e6 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  191. Joyce, J. A. & Pollard, J. W. Microenvironmental regulation of metastasis. Nat. Rev. Cancer 9, 239–252 (2009).

    Article  CAS  PubMed  Google Scholar 

  192. Gerald, D. et al. JunD reduces tumor angiogenesis by protecting cells from oxidative stress. Cell 118, 781–794 (2004).

    Article  CAS  PubMed  Google Scholar 

  193. Arbiser, J. L. et al. Reactive oxygen generated by Nox1 triggers the angiogenic switch. Proc. Natl Acad. Sci. USA 99, 715–720 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  194. Xia, C. et al. Reactive oxygen species regulate angiogenesis and tumor growth through vascular endothelial growth factor. Cancer Res. 67, 10823–10830 (2007).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  196. Chandel, N. S. et al. Reactive oxygen species generated at mitochondrial complex III stabilize hypoxia-inducible factor-1α during hypoxia: a mechanism of O2 sensing. J. Biol. Chem. 275, 25130–25138 (2000).

    Article  CAS  PubMed  Google Scholar 

  197. Connor, K. M. et al. Mitochondrial H2O2 regulates the angiogenic phenotype via PTEN oxidation. J. Biol. Chem. 280, 16916–16924 (2005).

    Article  CAS  PubMed  Google Scholar 

  198. Kim, Y. M., Kim, K. E., Koh, G. Y., Ho, Y. S. & Lee, K. J. Hydrogen peroxide produced by angiopoietin-1 mediates angiogenesis. Cancer Res. 66, 6167–6174 (2006).

    Article  CAS  PubMed  Google Scholar 

  199. West, X. Z. et al. Oxidative stress induces angiogenesis by activating TLR2 with novel endogenous ligands. Nature 467, 972–976 (2010). This study shows that inflammation leads to the generation of lipid oxidation products that promote angiogenesis through VEGF-independent but TLR2-dependent mechanisms, distinguishing this response from hypoxia-induced angiogenesis.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  200. ten Kate, M. et al. The role of superoxide anions in the development of distant tumour recurrence. Br. J. Cancer 95, 1497–1503 (2006).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  201. Carretero, J. et al. Tumoricidal activity of endothelial cells. Inhibition of endothelial nitric oxide production abrogates tumor cytotoxicity induced by hepatic sinusoidal endothelium in response to B16 melanoma adhesion in vitro. J. Biol. Chem. 276, 25775–25782 (2001).

    Article  CAS  PubMed  Google Scholar 

  202. Wang, H. H. et al. B16 melanoma cell arrest in the mouse liver induces nitric oxide release and sinusoidal cytotoxicity: a natural hepatic defense against metastasis. Cancer Res. 60, 5862–5869 (2000).

    CAS  PubMed  Google Scholar 

  203. van Wetering, S. et al. Reactive oxygen species mediate Rac-induced loss of cell–cell adhesion in primary human endothelial cells. J. Cell Sci. 115, 1837–1846 (2002).

    Article  PubMed  Google Scholar 

  204. McDowell, S. A. C. et al. Neutrophil oxidative stress mediates obesity-associated vascular dysfunction and metastatic transmigration. Nat. Cancer 2, 545–562 (2021). This study shows that obesity leads to an increase in ROS produced by neutrophils, which leads to NET production, a weakening of endothelial integrity and, subsequently, increased tumour extravasation and metastasis.

    Article  CAS  PubMed  Google Scholar 

  205. Carmeliet, P. & Jain, R. K. Principles and mechanisms of vessel normalization for cancer and other angiogenic diseases. Nat. Rev. Drug Discov. 10, 417–427 (2011).

    Article  CAS  PubMed  Google Scholar 

  206. Ippolito, L. et al. Cancer-associated fibroblasts promote prostate cancer malignancy via metabolic rewiring and mitochondrial transfer. Oncogene 38, 5339–5355 (2019).

    Article  CAS  PubMed  Google Scholar 

  207. Toullec, A. et al. Oxidative stress promotes myofibroblast differentiation and tumour spreading. EMBO Mol. Med. 2, 211–230 (2010). This study shows that JUND deletion in the stroma, which increases ROS, can facilitate metastasis by increasing the conversion of fibroblasts to activated myofibroblasts.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  208. Jezierska-Drutel, A. et al. The peroxidase PRDX1 inhibits the activated phenotype in mammary fibroblasts through regulating c-Jun N-terminal kinases. BMC Cancer 19, 812 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  209. Tobar, N. et al. c-Jun N terminal kinase modulates NOX-4 derived ROS production and myofibroblasts differentiation in human breast stromal cells. BMC Cancer 14, 640 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  210. Cat, B. et al. Enhancement of tumor invasion depends on transdifferentiation of skin fibroblasts mediated by reactive oxygen species. J. Cell Sci. 119, 2727–2738 (2006).

    Article  CAS  PubMed  Google Scholar 

  211. Giannoni, E., Bianchini, F., Calorini, L. & Chiarugi, P. Cancer associated fibroblasts exploit reactive oxygen species through a proinflammatory signature leading to epithelial mesenchymal transition and stemness. Antioxid. Redox Signal. 14, 2361–2371 (2011).

    Article  CAS  PubMed  Google Scholar 

  212. Xiang, H. et al. Cancer-associated fibroblasts promote immunosuppression by inducing ROS-generating monocytic MDSCs in lung squamous cell carcinoma. Cancer Immunol. Res. 8, 436–450 (2020).

    Article  CAS  PubMed  Google Scholar 

  213. Ford, K. et al. NOX4 inhibition potentiates immunotherapy by overcoming cancer-associated fibroblast-mediated CD8 T-cell exclusion from tumors. Cancer Res. 80, 1846–1860 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  214. Lood, C. et al. Neutrophil extracellular traps enriched in oxidized mitochondrial DNA are interferogenic and contribute to lupus-like disease. Nat. Med. 22, 146–153 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  215. Papayannopoulos, V. Neutrophil extracellular traps in immunity and disease. Nat. Rev. Immunol. 18, 134–147 (2018).

    Article  CAS  PubMed  Google Scholar 

  216. Fridlender, Z. G. et al. Polarization of tumor-associated neutrophil phenotype by TGF-β: “N1” versus “N2” TAN. Cancer Cell 16, 183–194 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  217. Albrengues, J. et al. Neutrophil extracellular traps produced during inflammation awaken dormant cancer cells in mice. Science 361, eaao4227 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  218. Park, J. et al. Cancer cells induce metastasis-supporting neutrophil extracellular DNA traps. Sci. Transl Med. 8, 361ra138 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  219. Wculek, S. K., Bridgeman, V. L., Peakman, F. & Malanchi, I. Early neutrophil responses to chemical carcinogenesis shape long-term lung cancer susceptibility. iScience 23, 101277 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  220. Granot, Z. et al. Tumor entrained neutrophils inhibit seeding in the premetastatic lung. Cancer Cell 20, 300–314 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  221. Wynn, T. A., Chawla, A. & Pollard, J. W. Macrophage biology in development, homeostasis and disease. Nature 496, 445–455 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  222. Noy, R. & Pollard, J. W. Tumor-associated macrophages: from mechanisms to therapy. Immunity 41, 49–61 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  223. Kennel, K. B. & Greten, F. R. Immune cell-produced ROS and their impact on tumor growth and metastasis. Redox Biol. 42, 101891 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  224. Zhang, J. et al. Tumoral NOX4 recruits M2 tumor-associated macrophages via ROS/PI3K signaling-dependent various cytokine production to promote NSCLC growth. Redox Biol. 22, 101116 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  225. Kuo, C. L. et al. Mitochondrial oxidative stress by Lon-PYCR1 maintains an immunosuppressive tumor microenvironment that promotes cancer progression and metastasis. Cancer Lett. 474, 138–150 (2020).

    Article  CAS  PubMed  Google Scholar 

  226. Molon, B. et al. Chemokine nitration prevents intratumoral infiltration of antigen-specific T cells. J. Exp. Med. 208, 1949–1962 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  227. Diwanji, N. & Bergmann, A. Basement membrane damage by ROS- and JNK-mediated Mmp2 activation drives macrophage recruitment to overgrown tissue. Nat. Commun. 11, 3631 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  228. Fogarty, C. E. et al. Extracellular reactive oxygen species drive apoptosis-induced proliferation via Drosophila macrophages. Curr. Biol. 26, 575–584 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  229. Kono, K. et al. Hydrogen peroxide secreted by tumor-derived macrophages down-modulates signal-transducing ζ molecules and inhibits tumor-specific T cell-and natural killer cell-mediated cytotoxicity. Eur. J. Immunol. 26, 1308–1313 (1996).

    Article  CAS  PubMed  Google Scholar 

  230. Corzo, C. A. et al. Mechanism regulating reactive oxygen species in tumor-induced myeloid-derived suppressor cells. J. Immunol. 182, 5693–5701 (2009).

    Article  CAS  PubMed  Google Scholar 

  231. Beury, D. W. et al. Myeloid-derived suppressor cell survival and function are regulated by the transcription factor Nrf2. J. Immunol. 196, 3470–3478 (2016).

    Article  CAS  PubMed  Google Scholar 

  232. Xia, H. et al. Autophagic adaptation to oxidative stress alters peritoneal residential macrophage survival and ovarian cancer metastasis. JCI Insight 5, e141115 (2020).

    Article  PubMed Central  Google Scholar 

  233. Jackson, S. H., Devadas, S., Kwon, J., Pinto, L. A. & Williams, M. S. T cells express a phagocyte-type NADPH oxidase that is activated after T cell receptor stimulation. Nat. Immunol. 5, 818–827 (2004).

    Article  CAS  PubMed  Google Scholar 

  234. Kwon, J. et al. The nonphagocytic NADPH oxidase Duox1 mediates a positive feedback loop during T cell receptor signaling. Sci. Signal. 3, ra59 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  235. Devadas, S., Zaritskaya, L., Rhee, S. G., Oberley, L. & Williams, M. S. Discrete generation of superoxide and hydrogen peroxide by T cell receptor stimulation: selective regulation of mitogen-activated protein kinase activation and fas ligand expression. J. Exp. Med. 195, 59–70 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  236. Kaminski, M. M. et al. Mitochondrial reactive oxygen species control T cell activation by regulating IL-2 and IL-4 expression: mechanism of ciprofloxacin-mediated immunosuppression. J. Immunol. 184, 4827–4841 (2010).

    Article  CAS  PubMed  Google Scholar 

  237. Yi, J. S., Holbrook, B. C., Michalek, R. D., Laniewski, N. G. & Grayson, J. M. Electron transport complex I is required for CD8+T cell function. J. Immunol. 177, 852–862 (2006).

    Article  CAS  PubMed  Google Scholar 

  238. Reth, M. Hydrogen peroxide as second messenger in lymphocyte activation. Nat. Immunol. 3, 1129–1134 (2002).

    Article  CAS  PubMed  Google Scholar 

  239. Siska, P. J. et al. Mitochondrial dysregulation and glycolytic insufficiency functionally impair CD8 T cells infiltrating human renal cell carcinoma. JCI Insight 2, e93411 (2017).

    Article  PubMed Central  Google Scholar 

  240. Ligtenberg, M. A. et al. Coexpressed catalase protects chimeric antigen receptor-redirected T cells as well as bystander cells from oxidative stress-induced loss of antitumor activity. J. Immunol. 196, 759–766 (2016). This paper shows that an enhanced antioxidant capacity in CAR T cells can increase not only their own antitumour activity but also that of the neighbouring immune cells, such as NK cells.

    Article  CAS  PubMed  Google Scholar 

  241. Blank, C. U. et al. Defining ‘T cell exhaustion’. Nat. Rev. Immunol. 19, 665–674 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  242. Scharping, N. E. et al. Mitochondrial stress induced by continuous stimulation under hypoxia rapidly drives T cell exhaustion. Nat. Immunol. 22, 205–215 (2021). This study shows that increased mtROS in intratumoural T cells owing to hypoxia and persistent antigen stimulation leads to T cell dysfunction and exhaustion. Reducing T cell intrinsic ROS damage can synergize with tumour immunotherapy.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  243. Gu, M. et al. NF-κB-inducing kinase maintains T cell metabolic fitness in antitumor immunity. Nat. Immunol. 22, 193–204 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  244. Tkachev, V. et al. Programmed death-1 controls T cell survival by regulating oxidative metabolism. J. Immunol. 194, 5789–5800 (2015).

    Article  CAS  PubMed  Google Scholar 

  245. Togashi, Y., Shitara, K. & Nishikawa, H. Regulatory T cells in cancer immunosuppression — implications for anticancer therapy. Nat. Rev. Clin. Oncol. 16, 356–371 (2019).

    Article  CAS  PubMed  Google Scholar 

  246. Josefowicz, S. Z., Lu, L. F. & Rudensky, A. Y. Regulatory T cells: mechanisms of differentiation and function. Annu. Rev. Immunol. 30, 531–564 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  247. Kraaij, M. D. et al. Induction of regulatory T cells by macrophages is dependent on production of reactive oxygen species. Proc. Natl Acad. Sci. USA 107, 17686–17691 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  248. Wen, Z. et al. NADPH oxidase deficiency underlies dysfunction of aged CD8+ Tregs. J. Clin. Invest. 126, 1953–1967 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  249. Maj, T. et al. Oxidative stress controls regulatory T cell apoptosis and suppressor activity and PD-L1-blockade resistance in tumor. Nat. Immunol. 18, 1332–1341 (2017). This study shows that ROS stress in tumours induces Treg cell apoptosis, which releases adenosine and mediates immunosuppression.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  250. Guo, Z. et al. DCAF1 regulates Treg senescence via the ROS axis during immunological aging. J. Clin. Invest. 130, 5893–5908 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  251. Sies, H. Hydrogen peroxide as a central redox signaling molecule in physiological oxidative stress: oxidative eustress. Redox Biol. 11, 613–619 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  252. Cheung, E. C. et al. Opposing effects of TIGAR- and RAC1-derived ROS on Wnt-driven proliferation in the mouse intestine. Genes Dev. 30, 52–63 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  253. Bersuker, K. et al. The CoQ oxidoreductase FSP1 acts parallel to GPX4 to inhibit ferroptosis. Nature 575, 688–692 (2019). This paper identifies another pathway inhibiting ferroptosis, mediated by FSP1, that acts in parallel to the GPX4 system.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  254. Takahashi, N. et al. Cancer cells co-opt the neuronal redox-sensing channel TRPA1 to promote oxidative-stress tolerance. Cancer Cell 33, 985–1003.e7 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  255. Kang, Y. P. et al. Non-canonical glutamate-cysteine ligase activity protects against ferroptosis. Cell Metab. 33, 174–189.e7 (2021).

    Article  CAS  PubMed  Google Scholar 

  256. Soula, M. et al. Metabolic determinants of cancer cell sensitivity to canonical ferroptosis inducers. Nat. Chem. Biol. 16, 1351–1360 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  257. Arora, S. et al. An undesired effect of chemotherapy: gemcitabine promotes pancreatic cancer cell invasiveness through reactive oxygen species-dependent, nuclear factor κB- and hypoxia-inducible factor 1α-mediated up-regulation of CXCR4. J. Biol. Chem. 288, 21197–21207 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  258. Mayers, J. R. et al. Tissue of origin dictates branched-chain amino acid metabolism in mutant Kras-driven cancers. Science 353, 1161–1165 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  259. Ward, M. P. et al. Platelets, immune cells and the coagulation cascade; friend or foe of the circulating tumour cell? Mol. Cancer 20, 59 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  260. Liou, G. Y. et al. Mutant KRas-induced mitochondrial oxidative stress in acinar cells upregulates EGFR signaling to drive formation of pancreatic precancerous lesions. Cell Rep. 14, 2325–2336 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  261. Humpton, T. J. et al. Oncogenic KRAS induces NIX-mediated mitophagy to promote pancreatic cancer. Cancer Discov. 9, 1268–1287 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  262. 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).

    Article  Google Scholar 

  263. Klein, E. A. et al. Vitamin E and the risk of prostate cancer: the Selenium and Vitamin E Cancer Prevention Trial (SELECT). JAMA 306, 1549–1556 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  264. Chandel, N. S. & Tuveson, D. A. The promise and perils of antioxidants for cancer patients. N. Engl. J. Med. 371, 177–178 (2014).

    Article  CAS  PubMed  Google Scholar 

  265. Shimura, T. et al. Radiation-induced myofibroblasts promote tumor growth via mitochondrial ROS-activated TGFβ signaling. Mol. Cancer Res. 16, 1676–1686 (2018).

    Article  CAS  PubMed  Google Scholar 

  266. Roux, C. et al. Reactive oxygen species modulate macrophage immunosuppressive phenotype through the up-regulation of PD-L1. Proc. Natl Acad. Sci. USA 116, 4326–4335 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  267. Yun, J. et al. Vitamin C selectively kills KRAS and BRAF mutant colorectal cancer cells by targeting GAPDH. Science 350, 1391–1396 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  268. Ezerina, D., Takano, Y., Hanaoka, K., Urano, Y. & Dick, T. P. N-acetyl cysteine functions as a fast-acting antioxidant by triggering intracellular H2S and sulfane sulfur production. Cell Chem. Biol. 25, 447–459.e4 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  271. Edmondson, D. E. & Binda, C. Monoamine oxidases. Subcell. Biochem. 87, 117–139 (2018).

    Article  CAS  PubMed  Google Scholar 

  272. Takac, I. et al. The E-loop is involved in hydrogen peroxide formation by the NADPH oxidase Nox4. J. Biol. Chem. 286, 13304–13313 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  273. Tu, B. P. & Weissman, J. S. The FAD- and O2-dependent reaction cycle of Ero1-mediated oxidative protein folding in the endoplasmic reticulum. Mol. Cell 10, 983–994 (2002).

    Article  CAS  PubMed  Google Scholar 

  274. Haynes, C. M., Titus, E. A. & Cooper, A. A. Degradation of misfolded proteins prevents ER-derived oxidative stress and cell death. Mol. Cell 15, 767–776 (2004).

    Article  CAS  PubMed  Google Scholar 

  275. Gilbert, J. A. et al. Current understanding of the human microbiome. Nat. Med. 24, 392–400 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  276. Human Microbiome Project Consortium. Structure, function and diversity of the healthy human microbiome. Nature 486, 207–214 (2012).

    Article  CAS  Google Scholar 

  277. Gonzalez-Sanchez, P. & DeNicola, G. M. The microbiome(s) and cancer: know thy neighbor(s). J. Pathol. 254, 332–343 (2021).

    Article  PubMed  Google Scholar 

  278. Sepich-Poore, G. D. et al. The microbiome and human cancer. Science 371, eabc4552 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  279. Garrett, W. S. Cancer and the microbiota. Science 348, 80–86 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  280. Xavier, J. B. et al. The cancer microbiome: distinguishing direct and indirect effects requires a systemic view. Trends Cancer 6, 192–204 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  281. Fang, Y. et al. The roles of microbial products in the development of colorectal cancer: a review. Bioengineered 12, 720–735 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  282. Iida, N. et al. Commensal bacteria control cancer response to therapy by modulating the tumor microenvironment. Science 342, 967–970 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  283. Cattaneo, F., Iaccio, A., Guerra, G., Montagnani, S. & Ammendola, R. NADPH-oxidase-dependent reactive oxygen species mediate EGFR transactivation by FPRL1 in WKYMVm-stimulated human lung cancer cells. Free Radic. Biol. Med. 51, 1126–1136 (2011).

    Article  CAS  PubMed  Google Scholar 

  284. van der Post, S., Birchenough, G. M. H. & Held, J. M. NOX1-dependent redox signaling potentiates colonic stem cell proliferation to adapt to the intestinal microbiota by linking EGFR and TLR activation. Cell Rep. 35, 108949 (2021).

    Article  PubMed  CAS  Google Scholar 

Download references

Acknowledgements

This work was funded by Cancer Research UK grant C596/A26855 and supported by the Francis Crick Institute, which receives its core funding from Cancer Research UK (FC001557 and FC001029), the UK Medical Research Council (FC001557 and FC001029) and the Wellcome Trust (FC001557 and FC001029). For the purpose of Open Access, the author has applied a CC BY public copyright licence to any author accepted manuscript version arising from this submission. The authors thank C. Labuschagne for his advice.

Author information

Authors and Affiliations

Authors

Contributions

The authors contributed equally to all aspects of the article.

Corresponding author

Correspondence to Karen H. Vousden.

Ethics declarations

Competing interests

K.H.V. is on the Board of Directors and a shareholder of Bristol Myers Squibb, a shareholder of Illumina, Inc., on the Science Advisory Board (with stock or stock options) of PMV Pharma, RAZE Therapeutics, Kovina Therapeutics and Volastra Therapeutics, and a co-founder and consultant of Faeth Therapeutics. She has been in receipt of research funding from Astex Pharmaceuticals and AstraZeneca, and contributed to CRUK Cancer Research Technology filing of patent application WO/2017/144877. E.C.C. declares no competing interests.

Peer review

Peer review information

Nature Reviews Cancer thanks Sarah-Maria Fendt and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Additional information

Publisher’s note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Glossary

Reactive oxygen species

(ROS). Unstable and reactive molecules that originate from oxygen during cellular metabolism.

NADPH

A reduced form of NADP+, the reducing agent for many anabolic reactions and regenerating antioxidants. Also used for generating reactive oxygen species (ROS) via NADPH oxidase (NOX).

Fenton reaction

The formation of hydroxyl radicals (·OH) and hydroperoxyl radicals from hydrogen peroxide (H2O2) and Fe2+ ions.

Mitophagy

The selective removal of dysfunctional mitochondria. Mitophagy is a type of autophagy, which is a cell intrinsic mechanism that removes and recycles unnecessary or dysfunctional cellular components and promotes short-term survival during starvation or repair during stress.

Peroxisomes

Small membrane-bound organelles that contain several reactive oxygen species (ROS)-producing and ROS-degrading enzymes for various oxidation and lipid biosynthesis reactions.

Oxidative pentose phosphate pathway

(oxPPP). An arm of the metabolic pathway that branches from glycolysis, generating NADPH and nucleotides.

Genomic instability

A high frequency of DNA mutations, chromosomal rearrangements or aneuploidy frequently seen during tumorigenesis.

Ferroptosis

A type of cell death that is dependent on iron and reactive oxygen species (ROS), promoting a toxic accumulation of oxidized lipids.

Lipid peroxidation

A chain reaction leading to the oxidation of lipids by oxidants that reacts with carbon–carbon double bonds, resulting in damaged membranes and cell death.

Cystine

An amino acid produced by the oxidation of two cysteine molecules, which are connected through a disulfide bond, and the predominant form of cysteine in cell culture media and the extracellular space.

Invadopodia

Protrusions from the cell membrane that induce degradation of the extracellular matrix and contribute to the ability of cancers to invade and metastasize.

Epithelial-to-mesenchymal transition

(EMT). A cellular process that involves losing epithelial characteristics such as intercellular contacts and cell polarity and gaining mesenchymal characteristics such as migratory and invasive properties. Essential in normal development, wound healing and fibrosis. EMT also contributes to tumour metastasis.

N-acetyl-l-cysteine

(NAC). An exogenous antioxidant, providing cysteine for glutathione synthesis and reducing free radicals via the generation of persulfides.

Cancer stem cells

(CSCs). Cancer cells that acquire some stem-like cell properties to enable them to self-renew to increase the capacity for metastasis, relapse and drug resistance.

Reductive carboxylation

A metabolic pathway that uses glutamine to produce citrate from α-ketoglutarate (α-KG) via the enzymes isocitrate dehydrogenase 1 (IDH1) and IDH2, rather than the normal tricarboxylic acid (TCA) cycle in which citrate is generated from acetyl-coenzyme A (acetyl-CoA) and oxaloacetate.

Circulating tumour cells

(CTCs). Cancer cells released into the lymphatic system or vasculature and carried around the body in blood circulation. These cells act as a precursor or seed for distant metastasis.

Electron transport chain

(Also known as the respiratory chain). Mitochondrial protein complexes that transfer electrons to create a proton gradient across the membrane, which is coupled to generate ATP via oxidative phosphorylation through ATP synthase.

Mitohormesis

Mild mitochondrial stress that increases antioxidative defences against later, higher reactive oxygen species (ROS) insults.

Angiogenesis

The growth of new blood vessels; in cancer, this refers to the development of abnormal vascularization during tumorigenesis.

Endothelial cells

Cells in a single layer that line the blood vessel.

Extravasation

The movement of cells from the blood vessels to the surrounding tissues.

Cancer-associated fibroblasts

(CAFs). Cells derived from resident normal fibroblasts or from other sites such as the endothelium or bone marrow (mesenchymal stem cells). These cells can remodel extracellular matrix or secrete factors that affect the cancer and other cells in the tumour microenvironment.

Monocytes

A type of white blood cell that gives rise to macrophages and dendritic cells.

Myeloid-derived suppressor cells

(MDSCs). Immature myeloid cells that are immunosuppressive during chronic infection and in the tumour microenvironment.

CD8+ T cell

A cytotoxic T cell that is a key component of the adaptive immune response, recognizing peptides through the T cell receptor (TCR), and able to clear infected, damaged cells and cancer cells.

Neutrophil extracellular traps

(NETs). Extracellular fibres consisting of DNA–histone complexes and proteins such as proteases and myeloperoxidase produced by neutrophils that bind pathogens extracellularly.

Tumour-associated macrophages

(TAMs). Monocyte-derived or tissue-resident macrophages (important components of the innate immune response) present in the tumour microenvironment that modulate immune suppression and angiogenesis.

CD4+ T cells

T helper cells that support the activity of other immune cells by releasing cytokines and through cell–cell interactions.

Chimeric antigen receptor (CAR) T cells

Genetically engineered T cells that produce designed T cell receptors (TCRs) against tumour cells for immunotherapy.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Cheung, E.C., Vousden, K.H. The role of ROS in tumour development and progression. Nat Rev Cancer 22, 280–297 (2022). https://doi.org/10.1038/s41568-021-00435-0

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41568-021-00435-0

This article is cited by

Search

Quick links

Nature Briefing: Cancer

Sign up for the Nature Briefing: Cancer newsletter — what matters in cancer research, free to your inbox weekly.

Get what matters in cancer research, free to your inbox weekly. Sign up for Nature Briefing: Cancer