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 adaptation to hypoxia through hypoxia inducible factors and beyond

Abstract

Molecular oxygen (O2) sustains intracellular bioenergetics and is consumed by numerous biochemical reactions, making it essential for most species on Earth. Accordingly, decreased oxygen concentration (hypoxia) is a major stressor that generally subverts life of aerobic species and is a prominent feature of pathological states encountered in bacterial infection, inflammation, wounds, cardiovascular defects and cancer. Therefore, key adaptive mechanisms to cope with hypoxia have evolved in mammals. Systemically, these adaptations include increased ventilation, cardiac output, blood vessel growth and circulating red blood cell numbers. On a cellular level, ATP-consuming reactions are suppressed, and metabolism is altered until oxygen homeostasis is restored. A critical question is how mammalian cells sense oxygen levels to coordinate diverse biological outputs during hypoxia. The best-studied mechanism of response to hypoxia involves hypoxia inducible factors (HIFs), which are stabilized by low oxygen availability and control the expression of a multitude of genes, including those involved in cell survival, angiogenesis, glycolysis and invasion/metastasis. Importantly, changes in oxygen can also be sensed via other stress pathways as well as changes in metabolite levels and the generation of reactive oxygen species by mitochondria. Collectively, this leads to cellular adaptations of protein synthesis, energy metabolism, mitochondrial respiration, lipid and carbon metabolism as well as nutrient acquisition. These mechanisms are integral inputs into fine-tuning the responses to hypoxic stress.

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: Transcription regulation induced by hypoxia.
Fig. 2: Hypoxic adaptations in proteostasis.
Fig. 3: Impact of hypoxia on mitochondrial function.
Fig. 4: Overview of sugar and lipid metabolic pathways affected by oxygen availability.
Fig. 5: Regulation of nutrient acquisition and use under hypoxia.

Similar content being viewed by others

References

  1. Pouyssegur, J. & López-Barneo, J. Hypoxia in health and disease. Mol. Asp. Med. 4748, 1–2 (2016).

    Google Scholar 

  2. Taylor, C. T., Doherty, G., Fallon, P. G. & Cummins, E. P. Hypoxia-dependent regulation of inflammatory pathways in immune cells. J. Clin. Invest. 126, 3716–3724 (2016).

    PubMed  PubMed Central  Google Scholar 

  3. Semenza, G. L. Hypoxia-inducible factors in physiology and medicine. Cell 148, 399–408 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Keith, B., Johnson, R. S. & Simon, M. C. HIF1α and HIF2α: sibling rivalry in hypoxic tumour growth and progression. Nat. Rev. Cancer 12, 9–22 (2011).

    PubMed  PubMed Central  Google Scholar 

  5. Kaelin, W. G. et al. Oxygen sensing by metazoans: the central role of the HIF hydroxylase pathway. Mol. Cell 30, 393–402 (2008). Together with Semenza (2012), this review provides an excellent outline of the regulation and importance of HIFs in physiology and diseases.

    CAS  PubMed  Google Scholar 

  6. Markolovic, S., Wilkins, S. E. & Schofield, C. J. Protein hydroxylation catalyzed by 2-oxoglutarate-dependent oxygenases. J. Biol. Chem. 290, 20712–20722 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Takeda, K. et al. Placental but not heart defects are associated with elevated hypoxia-inducible factor levels in mice lacking prolyl hydroxylase domain protein 2. Mol. Cell. Biol. 26, 8336–8346 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Lando, D., Peet, D. J., Whelan, D. A., Gorman, J. J. & Whitelaw, M. L. Asparagine hydroxylation of the HIF transactivation domain: a hypoxic switch. Science 295, 858–861 (2002).

    CAS  PubMed  Google Scholar 

  9. Zhang, N. et al. The asparaginyl hydroxylase factor inhibiting HIF-1α is an essential regulator of metabolism. Cell Metab. 11, 364–378 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Cockman, M. E., Webb, J. D., Kramer, H. B., Kessler, B. M. & Ratcliffe, P. J. Proteomics-based identification of novel factor inhibiting hypoxia-inducible factor (FIH) substrates indicates widespread asparaginyl hydroxylation of ankyrin repeat domain-containing proteins. Mol. Cell. Proteom. 8, 535–546 (2009).

    CAS  Google Scholar 

  11. Cockman, M. E. et al. Lack of activity of recombinant HIF prolyl hydroxylases (PHDs) on reported non-HIF substrates. eLife 8, e46490 (2019).

    PubMed  PubMed Central  Google Scholar 

  12. Lee, F. S. Substrates of PHD. Cell Metab. 30, 626–627 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Hirsilä, M., Koivunen, P., Günzler, V., Kivirikko, K. I. & Myllyharju, J. Characterization of the human prolyl 4-hydroxylases that modify the hypoxia-inducible factor. J. Biol. Chem. 278, 30772–30780 (2003).

    PubMed  Google Scholar 

  14. Ast, T. & Mootha, V. K. Oxygen and mammalian cell culture: are we repeating the experiment of Dr. Ox? Nat. Metab. 1, 858–860 (2019).

    CAS  PubMed  Google Scholar 

  15. Koivunen, P., Hirsilä, M., Günzler, V., Kivirikko, K. I. & Myllyharju, J. Catalytic properties of the asparaginyl hydroxylase (FIH) in the oxygen sensing pathway are distinct from those of its prolyl 4-hydroxylases. J. Biol. Chem. 279, 9899–9904 (2004).

    CAS  PubMed  Google Scholar 

  16. Jiang, B. H., Semenza, G. L., Bauer, C. & Marti, H. H. Hypoxia-inducible factor 1 levels vary exponentially over a physiologically relevant range of O2 tension. Am. J. Physiol. 271, C1172–C1180 (1996).

    CAS  PubMed  Google Scholar 

  17. Pan, Y. et al. Multiple factors affecting cellular redox status and energy metabolism modulate hypoxia-inducible factor prolyl hydroxylase activity in vivo and in vitro. Mol. Cell. Biol. 27, 912–925 (2007).

    CAS  PubMed  Google Scholar 

  18. Hagen, T., Taylor, C. T., Lam, F. & Moncada, S. Redistribution of intracellular oxygen in hypoxia by nitric oxide: effect on HIF1α. Science 302, 1975–1978 (2003).

    CAS  PubMed  Google Scholar 

  19. Yang, J. et al. Human CHCHD4 mitochondrial proteins regulate cellular oxygen consumption rate and metabolism and provide a critical role in hypoxia signaling and tumor progression. J. Clin. Invest. 122, 600–611 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. 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).

    CAS  PubMed  Google Scholar 

  21. Brunelle, J. K. et al. Oxygen sensing requires mitochondrial ROS but not oxidative phosphorylation. Cell Metab. 1, 409–414 (2005).

    CAS  PubMed  Google Scholar 

  22. Chandel, N. S. et al. Mitochondrial reactive oxygen species trigger hypoxia-induced transcription. Proc. Natl Acad. Sci. USA 95, 11715–11720 (1998). This key paper initially links mitochondria to hypoxia-induced gene transcription.

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Mansfield, K. D. et al. Mitochondrial dysfunction resulting from loss of cytochrome c impairs cellular oxygen sensing and hypoxic HIF-α activation. Cell Metab. 1, 393–399 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Guzy, R. D. et al. Mitochondrial complex III is required for hypoxia-induced ROS production and cellular oxygen sensing. Cell Metab. 1, 401–408 (2005).

    CAS  PubMed  Google Scholar 

  25. Lin, X. et al. A chemical genomics screen highlights the essential role of mitochondria in HIF-1 regulation. Proc. Natl Acad. Sci. USA 105, 174–179 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Lee, G. et al. Oxidative dimerization of PHD2 is responsible for its inactivation and contributes to metabolic reprogramming via HIF-1α activation. Sci. Rep. 6, 18928 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Reczek, C. R. & Chandel, N. S. ROS-dependent signal transduction. Curr. Opin. Cell Biol. 33, 8–13 (2015).

    CAS  PubMed  Google Scholar 

  28. Briggs, K. J. et al. Paracrine induction of HIF by glutamate in breast cancer: EglN1 senses cysteine. Cell 166, 126–139 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  29. King, A., Selak, M. A. & Gottlieb, E. Succinate dehydrogenase and fumarate hydratase: linking mitochondrial dysfunction and cancer. Oncogene 25, 4675–4682 (2006).

    CAS  PubMed  Google Scholar 

  30. Selak, M. A. et al. Succinate links TCA cycle dysfunction to oncogenesis by inhibiting HIF-α prolyl hydroxylase. Cancer Cell 7, 77–85 (2005).

    CAS  PubMed  Google Scholar 

  31. Pollard, P. J. et al. Accumulation of Krebs cycle intermediates and over-expression of HIF1α in tumours which result from germline FH and SDH mutations. Hum. Mol. Genet. 14, 2231–2239 (2005). Together with Selak et al. (2005), this paper provides the initial connection of mitochondrial TCA cycle metabolites to regulation of PHDs.

    CAS  PubMed  Google Scholar 

  32. Ryan, D. G. et al. Coupling Krebs cycle metabolites to signalling in immunity and cancer. Nat. Metab. 1, 16–33 (2019).

    PubMed  PubMed Central  Google Scholar 

  33. Intlekofer, A. M. et al. l-2-Hydroxyglutarate production arises from noncanonical enzyme function at acidic pH. Nat. Chem. Biol. 13, 494–500 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Nadtochiy, S. M. et al. Acidic pH Is a metabolic switch for 2-hydroxyglutarate generation and signaling. J. Biol. Chem. 291, 20188–20197 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Sonenberg, N. & Hinnebusch, A. G. Regulation of translation initiation in eukaryotes: mechanisms and biological targets. Cell 136, 731–745 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Jackson, R. J., Hellen, C. U. T. & Pestova, T. V. The mechanism of eukaryotic translation initiation and principles of its regulation. Nat. Rev. Mol. Cell Biol. 11, 113–127 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  37. van den Beucken, T. et al. Translational control is a major contributor to hypoxia induced gene expression. Radiother. Oncol. 99, 379–384 (2011).

    PubMed  Google Scholar 

  38. Feldman, D. E., Chauhan, V. & Koong, A. C. The unfolded protein response: a novel component of the hypoxic stress response in tumors. Mol. Cancer Res. 3, 597–605 (2005).

    CAS  PubMed  Google Scholar 

  39. Brewer, J. W. & Diehl, J. A. PERK mediates cell-cycle exit during the mammalian unfolded protein response. Proc. Natl Acad. Sci. USA 97, 12625–12630 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Blais, J. D. et al. Perk-dependent translational regulation promotes tumor cell adaptation and angiogenesis in response to hypoxic stress. Mol. Cell. Biol. 26, 9517–9532 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Koumenis, C. et al. Regulation of protein synthesis by hypoxia via activation of the endoplasmic reticulum kinase PERK and phosphorylation of the translation initiation factor eIF2α. Mol. Cell. Biol. 22, 7405–7416 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Arsham, A. M., Howell, J. J. & Simon, M. C. A novel hypoxia-inducible factor-independent hypoxic response regulating mammalian target of rapamycin and its targets. J. Biol. Chem. 278, 29655–29660 (2003).

    CAS  PubMed  Google Scholar 

  43. Koritzinsky, M. et al. Gene expression during acute and prolonged hypoxia is regulated by distinct mechanisms of translational control. EMBO J. 25, 1114–1125 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Connolly, E., Braunstein, S., Formenti, S. & Schneider, R. J. Hypoxia inhibits protein synthesis through a 4E-BP1 and elongation factor 2 kinase pathway controlled by mTOR and uncoupled in breast cancer cells. Mol. Cell. Biol. 26, 3955–3965 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Kenney, J. W., Moore, C. E., Wang, X. & Proud, C. G. Eukaryotic elongation factor 2 kinase, an unusual enzyme with multiple roles. Adv. Biol. Regul. 55, 15–27 (2014).

    CAS  PubMed  Google Scholar 

  46. Moore, C. E. J. et al. Elongation factor 2 kinase is regulated by proline hydroxylation and protects cells during hypoxia. Mol. Cell. Biol. 35, 1788–1804 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Romero-Ruiz, A. et al. Prolyl hydroxylase-dependent modulation of eukaryotic elongation factor 2 activity and protein translation under acute hypoxia. J. Biol. Chem. 287, 9651–9658 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Feng, T. et al. Optimal translational termination requires C4 lysyl hydroxylation of eRF1. Mol. Cell 53, 645–654 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  49. Harding, H. P. et al. An integrated stress response regulates amino acid metabolism and resistance to oxidative stress. Mol. Cell 11, 619–633 (2003).

    CAS  PubMed  Google Scholar 

  50. Harding, H. P. et al. Regulated translation initiation controls stress-induced gene expression in mammalian cells. Mol. Cell 6, 1099–1108 (2000).

    CAS  PubMed  Google Scholar 

  51. Han, J. et al. ER-stress-induced transcriptional regulation increases protein synthesis leading to cell death. Nat. Cell Biol. 15, 481–490 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Quirós, P. M. et al. Multi-omics analysis identifies ATF4 as a key regulator of the mitochondrial stress response in mammals. J. Cell Biol. 216, 2027–2045 (2017).

    PubMed  PubMed Central  Google Scholar 

  53. Kozak, M. A second look at cellular mRNA sequences said to function as internal ribosome entry sites. Nucleic Acids Res. 33, 6593–6602 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. King, H. A., Cobbold, L. C. & Willis, A. E. The role of IRES trans-acting factors in regulating translation initiation. Biochem. Soc. Trans. 38, 1581–1586 (2010).

    CAS  PubMed  Google Scholar 

  55. Stein, I. et al. Translation of vascular endothelial growth factor mRNA by internal ribosome entry: implications for translation under hypoxia. Mol. Cell. Biol. 18, 3112–3119 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Gan, W. & Rhoads, R. E. Internal initiation of translation directed by the 5′-untranslated region of the mRNA for eIF4G, a factor involved in the picornavirus-induced switch from cap-dependent to internal initiation. J. Biol. Chem. 271, 623–626 (1996).

    CAS  PubMed  Google Scholar 

  57. Young, R. M. et al. Hypoxia-mediated selective mRNA translation by an internal ribosome entry site-independent mechanism. J. Biol. Chem. 283, 16309–16319 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  58. Lang, K. J. D., Kappel, A. & Goodall, G. J. Hypoxia-inducible factor-1α mRNA contains an internal ribosome entry site that allows efficient translation during normoxia and hypoxia. Mol. Biol. Cell 13, 1792–1801 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Hinnebusch, A. G., Ivanov, I. P. & Sonenberg, N. Translational control by 5′-untranslated regions of eukaryotic mRNAs. Science 352, 1413–1416 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  60. Schwerk, J. & Savan, R. Translating the untranslated region. J. Immunol. 195, 2963–2971 (2015).

    CAS  PubMed  Google Scholar 

  61. Ho, J. J. D. et al. Systemic reprogramming of translation efficiencies on oxygen stimulus. Cell Rep. 14, 1293–1300 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  62. Uniacke, J. et al. An oxygen-regulated switch in the protein synthesis machinery. Nature 486, 126–129 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Uniacke, J., Kishan Perera, J., Lachance, G., Francisco, C. B. & Lee, S. Cancer cells exploit eIF4E2-directed synthesis of hypoxia response proteins to drive tumor progression. Cancer Res. 74, 1379–1389 (2014). Together with Uniacke et al. (2012), this work demonstrates oxygen-mediated changes to protein synthesis via HIF as a translation initiation factor.

    CAS  PubMed  Google Scholar 

  64. Sørensen, B. S., Busk, M., Overgaard, J., Horsman, M. R. & Alsner, J. Simultaneous hypoxia and low extracellular pH suppress overall metabolic rate and protein synthesis in vitro. PLoS One 10, e0134955 (2015).

    PubMed  PubMed Central  Google Scholar 

  65. Walton, Z. E. et al. Acid suspends the circadian clock in hypoxia through inhibition of mTOR. Cell 174, 72–87 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  66. Hermesh, O. & Jansen, R.-P. Take the (RN)A-train: localization of mRNA to the endoplasmic reticulum. Biochim. Biophys. Acta Mol. Cell Res. 1833, 2519–2525 (2013).

    CAS  Google Scholar 

  67. Staudacher, J. J. et al. Hypoxia-induced gene expression results from selective mRNA partitioning to the endoplasmic reticulum. Nucleic Acids Res. 43, 3219–3236 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  68. Hetz, C. The unfolded protein response: controlling cell fate decisions under ER stress and beyond. Nat. Rev. Mol. Cell Biol. 13, 89–102 (2012).

    CAS  PubMed  Google Scholar 

  69. Almanza, A. et al. Endoplasmic reticulum stress signalling — from basic mechanisms to clinical applications. FEBS J. 286, 241–278 (2019).

    CAS  PubMed  Google Scholar 

  70. Bouchecareilh, M., Higa, A., Fribourg, S., Moenner, M. & Chevet, E. Peptides derived from the bifunctional kinase/RNase enzyme IRE1α modulate IRE1α activity and protect cells from endoplasmic reticulum stress. FASEB J. 25, 3115–3129 (2011).

    CAS  PubMed  Google Scholar 

  71. Liu, C. Y., Schröder, M. & Kaufman, R. J. Ligand-independent dimerization activates the stress response kinases IRE1 and PERK in the lumen of the endoplasmic reticulum. J. Biol. Chem. 275, 24881–24885 (2000).

    CAS  PubMed  Google Scholar 

  72. Calfon, M. et al. IRE1 couples endoplasmic reticulum load to secretory capacity by processing the XBP-1 mRNA. Nature 415, 92–96 (2002).

    CAS  PubMed  Google Scholar 

  73. Lee, K. et al. IRE1-mediated unconventional mRNA splicing and S2P-mediated ATF6 cleavage merge to regulate XBP1 in signaling the unfolded protein response. Genes Dev. 16, 452–466 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  74. So, J.-S. et al. Silencing of lipid metabolism genes through IRE1α-mediated mRNA decay lowers plasma lipids in mice. Cell Metab. 16, 487–499 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  75. Xie, H. et al. IRE1α Rnase-dependent lipid homeostasis promotes survival in Myc-transformed cancers. J. Clin. Invest. 128, 1300–1316 (2018).

    PubMed  PubMed Central  Google Scholar 

  76. Zhou, Y. et al. Regulation of glucose homeostasis through a XBP-1–FoxO1 interaction. Nat. Med. 17, 356–365 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  77. Liu, J. et al. Inflammation improves glucose homeostasis through IKKβ–XBP1s interaction. Cell 167, 1052–1066 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  78. Liu, Y. et al. Preventing oxidative stress: a new role for XBP1. Cell Death Differ. 16, 847–857 (2009).

    CAS  PubMed  Google Scholar 

  79. Chen, X. et al. XBP1 promotes triple-negative breast cancer by controlling the HIF1α pathway. Nature 508, 103–107 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  80. Romero-Ramirez, L. et al. XBP1 is essential for survival under hypoxic conditions and is required for tumor growth. Cancer Res. 64, 5943–5947 (2004).

    CAS  PubMed  Google Scholar 

  81. Haze, K., Yoshida, H., Yanagi, H., Yura, T. & Mori, K. Mammalian transcription factor ATF6 is synthesized as a transmembrane protein and activated by proteolysis in response to endoplasmic reticulum stress. Mol. Biol. Cell 10, 3787–3799 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  82. Ye, J. et al. ER stress induces cleavage of membrane-bound ATF6 by the same proteases that process SREBPs. Mol. Cell 6, 1355–1364 (2000).

    CAS  PubMed  Google Scholar 

  83. Yamamoto, K. et al. Transcriptional induction of mammalian ER quality control proteins is mediated by single or combined action of ATF6α and XBP1. Dev. Cell 13, 365–376 (2007).

    CAS  PubMed  Google Scholar 

  84. Yoshida, H., Matsui, T., Yamamoto, A., Okada, T. & Mori, K. XBP1 mRNA is induced by ATF6 and spliced by IRE1 in response to ER stress to produce a highly active transcription factor. Cell 107, 881–891 (2001).

    CAS  PubMed  Google Scholar 

  85. Wu, J. et al. ATF6α optimizes long-term endoplasmic reticulum function to protect cells from chronic stress. Dev. Cell 13, 351–364 (2007).

    CAS  PubMed  Google Scholar 

  86. Shen, J., Chen, X., Hendershot, L. & Prywes, R. ER stress regulation of ATF6 localization by dissociation of BiP/GRP78 binding and unmasking of Golgi localization signals. Dev. Cell 3, 99–111 (2002).

    CAS  PubMed  Google Scholar 

  87. Fawcett, T. W., Martindale, J. L., Guyton, K. Z., Hai, T. & Holbrook, N. J. Complexes containing activating transcription factor (ATF)/cAMP-responsive-element-binding protein (CREB) interact with the CCAAT/enhancer-binding protein (C/EBP)-ATF composite site to regulate Gadd153 expression during the stress response. Biochem. J. 339, 135–141 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  88. Rutkowski, D. T. et al. Adaptation to ER stress is mediated by differential stabilities of pro-survival and pro-apoptotic mRNAs and proteins. PLoS Biol. 4, e374 (2006).

    PubMed  PubMed Central  Google Scholar 

  89. Bi, M. et al. ER stress-regulated translation increases tolerance to extreme hypoxia and promotes tumor growth. EMBO J. 24, 3470–3481 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  90. Rouschop, K. M. et al. PERK/eIF2α signaling protects therapy resistant hypoxic cells through induction of glutathione synthesis and protection against ROS. Proc. Natl Acad. Sci. USA 110, 4622–4627 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  91. Harding, H. P., Zhang, Y., Bertolotti, A., Zeng, H. & Ron, D. Perk is essential for translational regulation and cell survival during the unfolded protein response. Mol. Cell 5, 897–904 (2000).

    CAS  PubMed  Google Scholar 

  92. Wilson, D. F., Rumsey, W. L., Green, T. J. & Vanderkooi, J. M. The oxygen dependence of mitochondrial oxidative phosphorylation measured by a new optical method for measuring oxygen concentration. J. Biol. Chem. 263, 2712–2718 (1988).

    CAS  PubMed  Google Scholar 

  93. Cooper, C. E. The steady-state kinetics of cytochrome c oxidation by cytochrome oxidase. Biochim. Biophys. Acta 1017, 187–203 (1990).

    CAS  PubMed  Google Scholar 

  94. Wheaton, W. W. & Chandel, N. S. Hypoxia. 2. Hypoxia regulates cellular metabolism. Am. J. Physiol. Physiol. 300, C385–C393 (2011).

    CAS  Google Scholar 

  95. Chandel, N. S., Budinger, G. R. & Schumacker, P. T. Molecular oxygen modulates cytochrome c oxidase function. J. Biol. Chem. 271, 18672–18677 (1996).

    CAS  PubMed  Google Scholar 

  96. Fukuda, R. et al. HIF-1 regulates cytochrome oxidase subunits to optimize efficiency of respiration in hypoxic cells. Cell 129, 111–122 (2007).

    CAS  PubMed  Google Scholar 

  97. Hayashi, T. et al. Higd1a is a positive regulator of cytochrome c oxidase. Proc. Natl Acad. Sci. USA 112, 1553–1558 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  98. Tello, D. et al. Induction of the mitochondrial NDUFA4L2 protein by HIF-1α decreases oxygen consumption by inhibiting complex I activity. Cell Metab. 14, 768–779 (2011).

    CAS  PubMed  Google Scholar 

  99. Chan, S. Y. et al. MicroRNA-210 controls mitochondrial metabolism during hypoxia by repressing the iron–sulfur cluster assembly proteins ISCU1/2. Cell Metab. 10, 273–284 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  100. Chen, Z., Li, Y., Zhang, H., Huang, P. & Luthra, R. Hypoxia-regulated microRNA-210 modulates mitochondrial function and decreases ISCU and COX10 expression. Oncogene 29, 4362–4368 (2010).

    CAS  PubMed  Google Scholar 

  101. Puisségur, M.-P. et al. miR-210 is overexpressed in late stages of lung cancer and mediates mitochondrial alterations associated with modulation of HIF-1 activity. Cell Death Differ. 18, 465–478 (2011).

    PubMed  Google Scholar 

  102. Iyer, N. V. et al. Cellular and developmental control of O2 homeostasis by hypoxia-inducible factor 1α. Genes Dev. 12, 149–162 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    PubMed  Google Scholar 

  104. Papandreou, I., Cairns, R. A., Fontana, L., Lim, A. L. & Denko, N. C. HIF-1 mediates adaptation to hypoxia by actively downregulating mitochondrial oxygen consumption. Cell Metab. 3, 187–197 (2006). Together with Kim et al. (2006), this work provides initial links between HIF-1 target genes and inhibition of mitochondrial metabolism.

    CAS  PubMed  Google Scholar 

  105. Garcia-Bermudez, J. et al. Aspartate is a limiting metabolite for cancer cell proliferation under hypoxia and in tumours. Nat. Cell Biol. 20, 775–781 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  106. Ameri, K. et al. HIGD1A regulates oxygen consumption, ROS production, and AMPK activity during glucose deprivation to modulate cell survival and tumor growth. Cell Rep. 10, 891–899 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  107. Thomas, L. W., Staples, O., Turmaine, M. & Ashcroft, M. CHCHD4 regulates intracellular oxygenation and perinuclear distribution of mitochondria. Front. Oncol. 7, 71 (2017).

    PubMed  PubMed Central  Google Scholar 

  108. Al-Mehdi, A.-B. et al. Perinuclear mitochondrial clustering creates an oxidant-rich nuclear domain required for hypoxia-induced transcription. Sci. Signal. 5, ra47 (2012).

    PubMed  PubMed Central  Google Scholar 

  109. Fuhrmann, D. C. & Brüne, B. Mitochondrial composition and function under the control of hypoxia. Redox Biol. 12, 208–215 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  110. Kim, H. et al. Fine-tuning of Drp1/Fis1 availability by AKAP121/Siah2 regulates mitochondrial adaptation to hypoxia. Mol. Cell 44, 532–544 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  112. Liu, L. et al. Mitochondrial outer-membrane protein FUNDC1 mediates hypoxia-induced mitophagy in mammalian cells. Nat. Cell Biol. 14, 177–185 (2012).

    PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  115. Hekimi, S., Lapointe, J. & Wen, Y. Taking a ‘good’ look at free radicals in the aging process. Trends Cell Biol. 21, 569–576 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  117. 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).

    CAS  PubMed  Google Scholar 

  118. 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 

  119. Bell, E. L. & Chandel, N. S. Mitochondrial oxygen sensing: regulation of hypoxia-inducible factor by mitochondrial generated reactive oxygen species. Essays Biochem. 43, 17–27 (2007).

    CAS  PubMed  Google Scholar 

  120. Lee, S.-J., Hwang, A. B. & Kenyon, C. Inhibition of respiration extends C. elegans life span via reactive oxygen species that increase HIF-1 activity. Curr. Biol. 20, 2131–2136 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  121. Weir, E. K., López-Barneo, J., Buckler, K. J. & Archer, S. L. Acute oxygen-sensing mechanisms. N. Engl. J. Med. 353, 2042–2055 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  122. Sommer, N. et al. Mitochondrial complex IV subunit 4 isoform 2 is essential for acute pulmonary oxygen sensing. Circ. Res. 121, 424–438 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  123. Moreno-Domínguez, A. et al. Acute O2 sensing through HIF2α-dependent expression of atypical cytochrome oxidase subunits in arterial chemoreceptors. Sci. Signal. 13, eaay9452 (2019).

    Google Scholar 

  124. Fernández-Agüera, M. C. et al. Oxygen sensing by arterial chemoreceptors depends on mitochondrial complex I signaling. Cell Metab. 22, 825–837 (2015).

    PubMed  Google Scholar 

  125. Waypa, G. B. et al. Superoxide generated at mitochondrial complex III triggers acute responses to hypoxia in the pulmonary circulation. Am. J. Respir. Crit. Care Med. 187, 424–432 (2013). This paper provides genetic evidence that the mitochondrial respiratory chain is necessary for organismal acute responses to hypoxia.

    CAS  PubMed  PubMed Central  Google Scholar 

  126. Schleifer, G. et al. Impaired hypoxic pulmonary vasoconstriction in a mouse model of Leigh syndrome. Am. J. Physiol. Cell. Mol. Physiol. 316, L391–L399 (2019).

    CAS  Google Scholar 

  127. Chance, B. & Williams, G. R. The respiratory chain and oxidative phosphorylation. Adv. Enzymol. Relat. Subj. Biochem. 17, 65–134 (1956).

    CAS  PubMed  Google Scholar 

  128. Milligan, L. P. & McBride, B. W. Energy costs of ion pumping by animal tissues. J. Nutr. 115, 1374–1382 (1985).

    CAS  PubMed  Google Scholar 

  129. Helenius, I. T., Dada, L. A. & Sznajder, J. I. Role of ubiquitination in Na,K-ATPase regulation during lung injury. Proc. Am. Thorac. Soc. 7, 65–70 (2010).

    PubMed  PubMed Central  Google Scholar 

  130. Emerling, B. M. et al. Hypoxic activation of AMPK is dependent on mitochondrial ROS but independent of an increase in AMP/ATP ratio. Free. Radic. Biol. Med. 46, 1386–1391 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  131. Mungai, P. T. et al. Hypoxia triggers AMPK activation through reactive oxygen species-mediated activation of calcium release-activated calcium channels. Mol. Cell. Biol. 31, 3531–3545 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  132. Gusarova, G. A. et al. Hypoxia leads to Na,K-ATPase downregulation via Ca2+ release-activated Ca2+ channels and AMPK activation. Mol. Cell. Biol. 31, 3546–3556 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  133. Laderoute, K. R. et al. 5′-AMP-activated protein kinase (AMPK) is induced by low-oxygen and glucose deprivation conditions found in solid-tumor microenvironments. Mol. Cell. Biol. 26, 5336–5347 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  134. Garcia, D. & Shaw, R. J. AMPK: mechanisms of cellular energy sensing and restoration of metabolic balance. Mol. Cell 66, 789–800 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  135. Liu, L. et al. Hypoxia-induced energy stress regulates mRNA translation and cell growth. Mol. Cell 21, 521–531 (2006).

    PubMed  PubMed Central  Google Scholar 

  136. Orr, A. L. et al. Suppressors of superoxide production from mitochondrial complex III. Nat. Chem. Biol. 11, 834–836 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  137. Muller, F. L., Liu, Y. & Van Remmen, H. Complex III releases superoxide to both sides of the inner mitochondrial membrane. J. Biol. Chem. 279, 49064–49073 (2004).

    CAS  PubMed  Google Scholar 

  138. Hernansanz-Agustín, P. et al. Mitochondrial Na+ import controls oxidative phosphorylation and hypoxic redox signalling. bioRxiv https://doi.org/10.1101/385690 (2018).

    Article  Google Scholar 

  139. Szibor, M. et al. Broad AOX expression in a genetically tractable mouse model does not disturb normal physiology. Dis. Model. Mech. 10, 163–171 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  140. Samanta, D. & Semenza, G. L. Metabolic adaptation of cancer and immune cells mediated by hypoxia-inducible factors. Biochim. Biophys. Acta Rev. Cancer 1870, 15–22 (2018). This timely review article summarizes how HIFs regulate immune cells within the tumour microenvironment.

    CAS  PubMed  Google Scholar 

  141. Nakazawa, M. S., Keith, B. & Simon, M. C. Oxygen availability and metabolic adaptations. Nat. Rev. Cancer 16, 663–673 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  142. Xie, H. & Simon, M. C. Oxygen availability and metabolic reprogramming in cancer. J. Biol. Chem. 292, 16825–16832 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  143. Faubert, B. et al. Lactate metabolism in human lung tumors. Cell 171, 358–371 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  144. Jang, C. et al. Metabolite exchange between mammalian organs quantified in pigs. Cell Metab. 30, 594–606.e3 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  145. Morotti, M. et al. Hypoxia-induced switch in SNAT2/SLC38A2 regulation generates endocrine resistance in breast cancer. Proc. Natl Acad. Sci. USA 116, 12452–12461 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  146. Sun, R. C. & Denko, N. C. Hypoxic regulation of glutamine metabolism through HIF1 and SIAH2 supports lipid synthesis that is necessary for tumor growth. Cell Metab. 19, 285–292 (2014). This important paper indicates why hypoxia promotes reductive carboxylation of α-ketoglutarate via isocitrate dehydrogenase by degrading an α-ketoglutarate dehydrogenase complex subunit, and alters use of glutamine carbons for lipid synthesis.

    CAS  PubMed  PubMed Central  Google Scholar 

  147. Huang, D. et al. HIF-1-mediated suppression of acyl-CoA dehydrogenases and fatty acid oxidation is critical for cancer progression. Cell Rep. 8, 1930–1942 (2014).

    CAS  PubMed  Google Scholar 

  148. Kamphorst, J. J. et al. Hypoxic and Ras-transformed cells support growth by scavenging unsaturated fatty acids from lysophospholipids. Proc. Natl Acad. Sci. USA 110, 8882–8887 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  149. Young, R. M. et al. Dysregulated mTORC1 renders cells critically dependent on desaturated lipids for survival under tumor-like stress. Genes Dev. 27, 1115–1131 (2013). Together with Kamphorst et al. (2013), this paper demonstrates that tumour-relevant oxygen levels can inhibit enzymatic activity of the prominent fatty acyl desaturase SCD1.

    CAS  PubMed  PubMed Central  Google Scholar 

  150. Peck, B. & Schulze, A. Lipid desaturation — the next step in targeting lipogenesis in cancer? FEBS J. 283, 2767–2778 (2016).

    CAS  PubMed  Google Scholar 

  151. Shimabukuro, M., Zhou, Y.-T., Levi, M. & Unger, R. H. Fatty acid-induced cell apoptosis: a link between obesity and diabetes. Proc. Natl Acad. Sci. USA 95, 2498–2502 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  152. Qiu, B. et al. HIF2-dependent lipid storage promotes endoplasmic reticulum homeostasis in clear-cell renal cell carcinoma. Cancer Discov. 5, 652–667 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  153. Ackerman, D. et al. Triglycerides promote lipid homeostasis during hypoxic stress by balancing fatty acid saturation. Cell Rep. 24, 2596–2605.e5 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  154. Samanta, D. & Semenza, G. L. Serine synthesis helps hypoxic cancer stem cells regulate redox. Cancer Res. 76, 6458–6462 (2016).

    CAS  PubMed  Google Scholar 

  155. Levine, B. & Klionsky, D. J. Development by self-digestion: Molecular mechanisms and biological functions of autophagy. Dev. Cell 6, 463–477 (2004).

    CAS  PubMed  Google Scholar 

  156. Feng, Y., He, D., Yao, Z. & Klionsky, D. J. The machinery of macroautophagy. Cell Res. 24, 24–41 (2014).

    CAS  PubMed  Google Scholar 

  157. Kimmelman, A. C. & White, E. Autophagy and tumor metabolism. Cell Metab. 25, 1037–1043 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  158. Mazure, N. M. & Pouysségur, J. Hypoxia-induced autophagy: cell death or cell survival? Curr. Opin. Cell Biol. 22, 177–180 (2010).

    CAS  PubMed  Google Scholar 

  159. Choi, A. M. K., Ryter, S. W. & Levine, B. Autophagy in human health and disease. N. Engl. J. Med. 368, 651–662 (2013).

    CAS  PubMed  Google Scholar 

  160. Poillet-Perez, L. & White, E. Role of tumor and host autophagy in cancer metabolism. Genes Dev. 33, 610–619 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  161. Zhang, H. et al. Mitochondrial autophagy is an HIF-1-dependent adaptive metabolic response to Hypoxia. J. Biol. Chem. 283, 10892–10903 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  162. Bellot, G. et al. Hypoxia-induced autophagy is mediated through hypoxia-inducible factor induction of BNIP3 and BNIP3L via their BH3 domains. Mol. Cell. Biol. 29, 2570–2581 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    PubMed  Google Scholar 

  164. Azad, M. B. et al. Hypoxia induces autophagic cell death in apoptosis-competent cells through a mechanism involving BNIP3. Autophagy 4, 195–204 (2008).

    CAS  PubMed  Google Scholar 

  165. Fei, P. et al. Bnip3L is induced by p53 under hypoxia, and its knockdown promotes tumor growth. Cancer Cell 6, 597–609 (2004).

    CAS  PubMed  Google Scholar 

  166. Mazure, N. M. & Pouysségur, J. Atypical BH3-domains of BNIP3 and BNIP3L lead to autophagy in hypoxia. Autophagy 5, 868–869 (2009).

    PubMed  Google Scholar 

  167. Band, M., Joel, A., Hernandez, A. & Avivi, A. Hypoxia-induced BNIP3 expression and mitophagy: in vivo comparison of the rat and the hypoxia-tolerant mole rat, Spalax ehrenbergi. FASEB J. 23, 2327–2335 (2009).

    CAS  PubMed  Google Scholar 

  168. Chourasia, A. H. & Macleod, K. F. Tumor suppressor functions of BNIP3 and mitophagy. Autophagy 11, 1937–1938 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  169. Rozpedek, W. et al. The role of the PERK/eIF2α/ATF4/CHOP signaling pathway in tumor progression during endoplasmic reticulum stress. Curr. Mol. Med. 16, 533–544 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  170. Mihaylova, M. M. & Shaw, R. J. The AMPK signalling pathway coordinates cell growth, autophagy and metabolism. Nat. Cell Biol. 13, 1016–1023 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  171. Hardie, D. G. AMPK and autophagy get connected. EMBO J. 30, 634–635 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  172. Wang, C. et al. Phosphorylation of ULK1 affects autophagosome fusion and links chaperone-mediated autophagy to macroautophagy. Nat. Commun. 9, 3492 (2018).

    PubMed  PubMed Central  Google Scholar 

  173. Russell, R. C. et al. ULK1 induces autophagy by phosphorylating Beclin-1 and activating VPS34 lipid kinase. Nat. Cell Biol. 15, 741–750 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  174. Egan, D. F. et al. Phosphorylation of ULK1 (hATG1) by AMP-activated protein kinase connects energy sensing to mitophagy. Science 331, 456–461 (2011).

    CAS  PubMed  Google Scholar 

  175. Frezza, C. et al. Metabolic profiling of hypoxic cells revealed a catabolic signature required for cell survival. PLoS One 6, e24411 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  176. Antonescu, C. N., McGraw, T. E. & Klip, A. Reciprocal regulation of endocytosis and metabolism. Cold Spring Harb. Perspect. Biol. 6, a016964 (2014).

    PubMed  PubMed Central  Google Scholar 

  177. Huang, S. & Czech, M. P. The GLUT4 glucose transporter. Cell Metab. 5, 237–252 (2007).

    CAS  PubMed  Google Scholar 

  178. Herman, M. A. & Kahn, B. B. Glucose transport and sensing in the maintenance of glucose homeostasis and metabolic harmony. J. Clin. Invest. 116, 1767–1775 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  179. Foley, K., Boguslavsky, S. & Klip, A. Endocytosis, recycling, and regulated exocytosis of glucose transporter 4. Biochemistry 50, 3048–3061 (2011).

    CAS  PubMed  Google Scholar 

  180. Fazakerley, D. J. et al. Kinetic evidence for unique regulation of GLUT4 trafficking by insulin and AMP-activated protein kinase activators in L6 myotubes. J. Biol. Chem. 285, 1653–1660 (2010).

    CAS  PubMed  Google Scholar 

  181. Mu, J., Brozinick, J. T., Valladares, O., Bucan, M. & Birnbaum, M. J. A role for AMP-activated protein kinase in contraction- and hypoxia-regulated glucose transport in skeletal muscle. Mol. Cell 7, 1085–1094 (2001).

    CAS  PubMed  Google Scholar 

  182. Sakagami, H. et al. Loss of HIF-1α impairs GLUT4 translocation and glucose uptake by the skeletal muscle cells. Am. J. Physiol. Metab. 306, E1065–E1076 (2014).

    CAS  Google Scholar 

  183. Görgens, S. W. et al. Hypoxia in combination with muscle contraction improves insulin action and glucose metabolism in human skeletal muscle via the HIF-1α pathway. Diabetes 66, 2800–2807 (2017).

    PubMed  Google Scholar 

  184. Li, G., Wang, J., Ye, J., Zhang, Y. & Zhang, Y. PPARα protein expression was increased by four weeks of intermittent hypoxic training via ampkα2-dependent manner in mouse skeletal muscle. PLoS One 10, e0122593 (2015).

    PubMed  PubMed Central  Google Scholar 

  185. Siques, P. et al. Long-term chronic intermittent hypobaric hypoxia induces glucose transporter (GLUT4) translocation through AMP-activated protein kinase (AMPK) in the soleus muscle in lean rats. Front. Physiol. 9, 799 (2018).

    PubMed  PubMed Central  Google Scholar 

  186. Wang, Y. et al. Effects of four weeks intermittent hypoxia intervention on glucose homeostasis, insulin sensitivity, GLUT4 translocation, insulin receptor phosphorylation, and Akt activity in skeletal muscle of obese mice with type 2 diabetes. PLoS One 13, e0203551 (2018).

    PubMed  PubMed Central  Google Scholar 

  187. Lee, K. Y., Gesta, S., Boucher, J., Wang, X. L. & Kahn, C. R. The differential role of Hif1β/Arnt and the hypoxic response in adipose function, fibrosis, and inflammation. Cell Metab. 14, 491–503 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  188. Bloomfield, G. & Kay, R. R. Uses and abuses of macropinocytosis. J. Cell Sci. 129, 2697–2705 (2016).

    CAS  PubMed  Google Scholar 

  189. Recouvreux, M. V. & Commisso, C. Macropinocytosis: a metabolic adaptation to nutrient stress in cancer. Front. Endocrinol. 8, 261 (2017).

    Google Scholar 

  190. Commisso, C. et al. Macropinocytosis of protein is an amino acid supply route in Ras-transformed cells. Nature 497, 633–637 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  191. Palm, W. et al. The utilization of extracellular proteins as nutrients is suppressed by mTORC1. Cell 162, 259–270 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  192. Kamphorst, J. J. et al. Human pancreatic cancer tumors are nutrient poor and tumor cells actively scavenge extracellular protein. Cancer Res. 75, 544–553 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  193. Li, X., Egervari, G., Wang, Y., Berger, S. L. & Lu, Z. Regulation of chromatin and gene expression by metabolic enzymes and metabolites. Nat. Rev. Mol. Cell Biol. 19, 563–578 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  194. Chen, W. et al. Targeting renal cell carcinoma with a HIF-2 antagonist. Nature 539, 112–117 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  195. Cho, H. et al. On-target efficacy of a HIF-2α antagonist in preclinical kidney cancer models. Nature 539, 107–111 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  196. Courtney, K. D. et al. HIF-2 complex dissociation, target inhibition, and acquired resistance with PT2385, a first-in-class HIF-2 inhibitor, in patients with clear cell renal cell carcinoma. Clin. Cancer Res. 26, 793–803 (2020).

    PubMed  Google Scholar 

  197. Maxwell, P. H. & Eckardt, K. U. HIF prolyl hydroxylase inhibitors for the treatment of renal anaemia and beyond. Nat. Rev. Nephrol. 12, 157–168 (2016).

    CAS  PubMed  Google Scholar 

  198. Jain, I. H. et al. Leigh syndrome mouse model can be rescued by interventions that normalize brain hyperoxia, but not HIF activation. Cell Metab. 30, 824–832 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  199. Ast, T. et al. Hypoxia rescues frataxin loss by restoring iron sulfur cluster biogenesis. Cell 177, 1507–1521 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  200. Krock, B. L., Skuli, N. & Simon, M. C. Hypoxia-induced angiogenesis: good and evil. Genes Cancer 2, 1117–1133 (2011).

    PubMed  PubMed Central  Google Scholar 

  201. Wong, B. W., Marsch, E., Treps, L., Baes, M. & Carmeliet, P. Endothelial cell metabolism in health and disease: impact of hypoxia. EMBO J. 36, 2187–2203 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  202. Eales, K. L., Hollinshead, K. E. R. & Tennant, D. A. Hypoxia and metabolic adaptation of cancer cells. Oncogenesis 5, e190 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  203. Höckel, M. et al. Intratumoral pO2 predicts survival in advanced cancer of the uterine cervix. Radiother. Oncol. 26, 45–50 (1993).

    PubMed  Google Scholar 

  204. Brizel, D. M. et al. Tumor oxygenation predicts for the likelihood of distant metastases in human soft tissue sarcoma. Cancer Res. 56, 941–943 (1996).

    CAS  PubMed  Google Scholar 

  205. Vaupel, P., Kelleher, D. K. & Höckel, M. Oxygenation status of malignant tumors: pathogenesis of hypoxia and significance for tumor therapy. Semin. Oncol. 28, 29–35 (2001).

    CAS  PubMed  Google Scholar 

  206. Eisinger-Mathason, T. S. K. et al. Hypoxia-dependent modification of collagen networks promotes sarcoma metastasis. Cancer Discov. 3, 1190–1205 (2013).

    CAS  PubMed  Google Scholar 

  207. Sun, J. D. et al. Selective tumor hypoxia targeting by hypoxia-activated prodrug TH-302 inhibits tumor growth in preclinical models of cancer. Clin. Cancer Res. 18, 758–770 (2012).

    CAS  PubMed  Google Scholar 

  208. Denny, W. A. The role of hypoxia-activated prodrugs in cancer therapy. Lancet Oncol. 1, 25–29 (2000).

    CAS  PubMed  Google Scholar 

  209. Baran, N. & Konopleva, M. Molecular pathways: hypoxia-activated prodrugs in cancer therapy. Clin. Cancer Res. 23, 2382–2390 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  210. Shmakova, A., Batie, M., Druker, J. & Rocha, S. Chromatin and oxygen sensing in the context of JmjC histone demethylases. Biochem. J. 462, 385–395 (2014).

    CAS  PubMed  Google Scholar 

  211. Chakraborty, A. A. et al. Histone demethylase KDM6A directly senses oxygen to control chromatin and cell fate. Science 363, 1217–1222 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  212. Batie, M. et al. Hypoxia induces rapid changes to histone methylation and reprograms chromatin. Science 363, 1222–1226 (2019).

    CAS  PubMed  Google Scholar 

  213. Masson, N. et al. Conserved N-terminal cysteine dioxygenases transduce responses to hypoxia in animals and plants. Science 365, 65–69 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  214. Licausi, F. et al. Oxygen sensing in plants is mediated by an N-end rule pathway for protein destabilization. Nature 479, 419–422 (2011).

    CAS  PubMed  Google Scholar 

  215. Gibbs, D. J. et al. Homeostatic response to hypoxia is regulated by the N-end rule pathway in plants. Nature 479, 415–418 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  216. Weits, D. A. et al. Plant cysteine oxidases control the oxygen-dependent branch of the N-end-rule pathway. Nat. Commun. 5, 3425 (2014).

    PubMed  Google Scholar 

  217. Brugarolas, J. et al. Regulation of mTOR function in response to hypoxia by REDD1 and the TSC1/TSC2 tumor suppressor complex. Genes Dev. 18, 2893–2904 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  218. Sofer, A., Lei, K., Johannessen, C. M. & Ellisen, L. W. Regulation of mTOR and cell growth in response to energy stress by REDD1. Mol. Cell. Biol. 25, 5834–5845 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  219. Ding, M. et al. The mTOR targets 4E-BP1/2 restrain tumor growth and promote hypoxia tolerance in PTEN-driven prostate cancer. Mol. Cancer Res. 16, 682–695 (2018).

    CAS  PubMed  Google Scholar 

  220. Seong, M., Lee, J. & Kang, H. Hypoxia−induced regulation of mTOR signaling by miR-7 targeting REDD1. J. Cell. Biochem. 120, 4523–4532 (2019).

    CAS  PubMed  Google Scholar 

  221. Bernardi, R. et al. PML inhibits HIF-1α translation and neoangiogenesis through repression of mTOR. Nature 442, 779–785 (2006).

    CAS  PubMed  Google Scholar 

  222. Sancak, Y. et al. The rag GTPases bind raptor and mediate amino acid signaling to mTORC1. Science 320, 1496–1501 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  223. Li, Y. et al. Bnip3 mediates the hypoxia-induced inhibition on mammalian target of rapamycin by interacting with Rheb. J. Biol. Chem. 282, 35803–35813 (2007).

    CAS  PubMed  Google Scholar 

  224. Cam, H., Easton, J. B., High, A. & Houghton, P. J. mTORC1 signaling under hypoxic conditions is controlled by ATM-dependent phosphorylation of HIF-1α. Mol. Cell 40, 509–520 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  225. Brugarolas, J. B., Vazquez, F., Reddy, A., Sellers, W. R. & Kaelin, W. G. TSC2 regulates VEGF through mTOR-dependent and -independent pathways. Cancer Cell 4, 147–158 (2003).

    CAS  PubMed  Google Scholar 

  226. Hudson, C. C. et al. Regulation of hypoxia-inducible factor 1α expression and function by the mammalian target of rapamycin. Mol. Cell. Biol. 22, 7004–7014 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  227. Thomas, G. V. et al. Hypoxia-inducible factor determines sensitivity to inhibitors of mTOR in kidney cancer. Nat. Med. 12, 122–127 (2006).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

N.S.C. acknowledges support by R35CA197532 from the National Institutes of Health (NIH). M.S.C. is supported by P01CA104838 and R35CA220483 from the NIH. P.L. is supported by T32AR53461-10 and F32CA217185-02 from the NIH. The authors have no financial interests to disclose. The authors apologize to those authors whose research could not be directly cited due to space limitations.

Author information

Authors and Affiliations

Authors

Contributions

The authors contributed equally to all aspects of the article.

Corresponding authors

Correspondence to Pearl Lee, Navdeep S. Chandel or M. Celeste Simon.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information

Nature Reviews Molecular Cell Biology thanks C. Taylor, M. Ashcroft and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Publisher’s note

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

Supplementary information

Glossary

Dioxygenases

A group of enzymes that reduce molecular oxygen by incorporating both oxygen atoms into their substrates.

von Hippel–Lindau (VHL) tumour suppressor protein

(pVHL). A protein named after the physicians von Hippel and Lindau, who characterized patients with highly vascular neoplasia of the kidney, eye and central nervous system who carried mutations in the VHL gene. pVHL is required for the ubiquitylation of hypoxia inducible factor-α and its degradation.

Carotid body

A cluster of peripheral chemoreceptor cells (glomus type I and glomus type II), which sense oxygen, carbon dioxide and pH levels of blood.

Erythropoietin

A glycoprotein cytokine secreted by the kidney in response to hypoxia to stimulate erythropoiesis.

Elongin BC–CUL2

Additional complexes that interact with von Hippel–Lindau (VHL) tumour suppressor protein (pVHL). The Elongin BC complex acts as an adaptor connecting Cullin (Cul) proteins.

Michaelis constant

(Km). The substrate concentration at half of the maximum reaction velocity.

Acidosis

The process or condition where there is increased acidity in the blood and other body tissues.

Cap-dependent protein synthesis

Eukaryotic mRNAs contain a modified guanosine (the cap) at their 5′ ends. Cap-dependent translation requires the binding of an initiation factor, eukaryotic translation initiation factor 4E (eIF4E), to the cap structure.

PERK

(Protein kinase RNA (PKR)-like endoplasmic reticulum (ER) kinase). A sensor of ER stress and stress-induced protein misfolding.

mTOR

(Mechanistic target of rapamycin). A protein kinase that regulates protein synthesis and cell growth in response to growth factors, nutrients, energy levels and stress.

Eukaryotic initiation factor 2

(eIF2). A complex comprising α, β and γ components that integrates a diverse array of stress-related signals to regulate both global and specific mRNA translation.

ER stress

A condition when the capacity of the endoplasmic reticulum (ER) to fold proteins becomes saturated due to impaired protein glycosylation or disulfide bond formation, or by overexpression of or mutations in proteins entering the secretory pathway.

eIF4F

The cap-binding eukaryotic translation initiation factor 4F (eIF4F) complex consists of three subunits, eIF4A, eIF4E and eIF4G. eIF4G strongly associates with eIF4E, the protein that directly binds the mRNA cap.

ATF4

(Activating transcription factor 4). A cAMP-response element binding protein that belongs to the cAMP response element-binding protein 2 (CREB2) family of transcription factors.

Integrated stress response

An adaptive pathway to restore cellular homeostasis by optimizing the cellular response to stress. Its activity is dependent on the cellular context and the type as well as intensity of the stress stimuli.

Signal recognition particles

Universally conserved ribonucleoproteins that recognize and target specific proteins to the endoplasmic reticulum.

ER-associated degradation

The cellular pathway that targets misfolded proteins of the endoplasmic reticulum (ER) for ubiquitylation and subsequent degradation by the proteasome.

Electron transport chain

(ETC). A series of complexes within the inner mitochondrial membrane that shuttle electrons from NADH and FADH2 to molecular oxygen.

Enzymatic maximal velocity

(Vmax). The reaction rate when an enzyme is fully saturated by substrate.

Iron–sulphur (Fe–S) clusters

Molecular ensembles of iron and sulfide, often found as components of electron transfer proteins. The ferredoxin proteins are the most common iron–sulfide clusters in nature.

Pyruvate

The conjugate base, CH3COCOO, of pyruvic acid is a key intermediate in several metabolic pathways throughout the cell.

Superoxide dismutases

A family of enzymes that catalyse the dismutation of the superoxide (O2) radical into molecular oxygen or hydrogen peroxide.

Telomerase

A ribonucleoprotein that adds a species-dependent telomere repeat sequence to the 3′ end of a region of repetitive sequences at each end of chromosomes (telomeres).

Essential electrochemical gradient

A gradient of electrochemical potential, usually for an ion that can move across membranes.

Ubisemiquinone

Ubiquinol is the reduced form of coenzyme Q10 that is oxidized by mitochondrial complex III to the partially reduced form ubisemiquinone and subsequently to the fully oxidized ubiquinone. Ubisemiquionone is a highly unstable free radical that can donate electrons to molecular oxygen to generate superoxide.

Lipid droplets

Lipid-rich storage organelles that regulate the storage and hydrolysis of neutral lipids. They also serve as a reservoir for cholesterol and acyl-glycerols for membrane formation and maintenance.

Pentose phosphate pathway

A predominantly anabolic pathway parallel to glycolysis that generates NADPH and precursors for nucleotide synthesis.

Leigh syndrome

A neurological disorder characterized by progressive loss of mental and motor abilities.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lee, P., Chandel, N.S. & Simon, M.C. Cellular adaptation to hypoxia through hypoxia inducible factors and beyond. Nat Rev Mol Cell Biol 21, 268–283 (2020). https://doi.org/10.1038/s41580-020-0227-y

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41580-020-0227-y

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