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.

  • Article
  • Published:

Mitochondrial fission links ECM mechanotransduction to metabolic redox homeostasis and metastatic chemotherapy resistance

Abstract

Metastatic breast cancer cells disseminate to organs with a soft microenvironment. Whether and how the mechanical properties of the local tissue influence their response to treatment remains unclear. Here we found that a soft extracellular matrix empowers redox homeostasis. Cells cultured on a soft extracellular matrix display increased peri-mitochondrial F-actin, promoted by Spire1C and Arp2/3 nucleation factors, and increased DRP1- and MIEF1/2-dependent mitochondrial fission. Changes in mitochondrial dynamics lead to increased production of mitochondrial reactive oxygen species and activate the NRF2 antioxidant transcriptional response, including increased cystine uptake and glutathione metabolism. This retrograde response endows cells with resistance to oxidative stress and reactive oxygen species-dependent chemotherapy drugs. This is relevant in a mouse model of metastatic breast cancer cells dormant in the lung soft tissue, where inhibition of DRP1 and NRF2 restored cisplatin sensitivity and prevented disseminated cancer-cell awakening. We propose that targeting this mitochondrial dynamics- and redox-based mechanotransduction pathway could open avenues to prevent metastatic relapse.

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: ECM stiffness regulates cystine metabolism and glutathione oxidation.
Fig. 2: ECM stiffness regulates ROS levels.
Fig. 3: NRF2 activation on soft ECM potentiates cell resistance to exogenous oxidative stress.
Fig. 4: The antioxidant response induced by ECM stiffness is initiated by mtROS.
Fig. 5: ECM stiffness regulates mitochondrial fission through DRP1.
Fig. 6: ECM stiffness regulates fission via Spire1c- and Arp2/3-dependent peri-mitochondrial F-actin.
Fig. 7: A soft ECM increases resistance to cisplatin and As2O3 chemotherapy through NRF2 and DRP1.
Fig. 8: A soft metastatic niche protects dormant breast cancer cells from chemotherapy.

Similar content being viewed by others

Data availability

Previously published metabolomics data have been deposited to the Figshare database and are available at https://doi.org/10.6084/m9.figshare.7338764. The RNA sequencing data have been deposited to Gene EXpression Omnibus database as GSE189803. All other data supporting the findings of this study are available from the corresponding author on reasonable request. Source data are provided with this paper.

References

  1. Min, E. & Schwartz, M. A. Translocating transcription factors in fluid shear stress-mediated vascular remodeling and disease. Exp. Cell. Res. 376, 92–97 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Petridou, N. I., Spiró, Z. & Heisenberg, C.-P. Multiscale force sensing in development. Nat. Cell Biol. 19, 581–588 (2017).

    Article  CAS  PubMed  Google Scholar 

  3. Tschumperlin, D. J., Ligresti, G., Hilscher, M. B. & Shah, V. H. Mechanosensing and fibrosis. J. Clin. Invest. 128, 74–84 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  4. Vining, K. H. & Mooney, D. J. Mechanical forces direct stem cell behaviour in development and regeneration. Nat. Rev. Mol. Cell Biol. 18, 728–742 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Humphrey, J. D., Dufresne, E. R. & Schwartz, M. A. Mechanotransduction and extracellular matrix homeostasis. Nat. Rev. Mol. Cell Biol. 15, 802–812 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Iskratsch, T., Wolfenson, H. & Sheetz, M. P. Appreciating force and shape—the rise of mechanotransduction in cell biology. Nat. Rev. Mol. Cell Biol. 15, 825–833 (2014).

    Article  CAS  PubMed  Google Scholar 

  7. Mohammadi, H. & Sahai, E. Mechanisms and impact of altered tumour mechanics. Nat. Cell Biol. 20, 766–774 (2018).

    Article  CAS  PubMed  Google Scholar 

  8. Gensbittel, V. et al. Mechanical adaptability of tumor cells in metastasis. Dev. Cell 56, 164–179 (2021).

    Article  CAS  PubMed  Google Scholar 

  9. Montagner, M. & Dupont, S. Mechanical Forces as Determinants of Disseminated Metastatic Cell Fate. Cells 9, 250 (2020).

  10. Romani, P., Valcarcel-Jimenez, L., Frezza, C. & Dupont, S. Crosstalk between mechanotransduction and metabolism. Nat. Rev. Mol. Cell Biol. https://doi.org/10.1038/s41580-020-00306-w (2020).

  11. Romani, P. et al. Extracellular matrix mechanical cues regulate lipid metabolism through Lipin-1 and SREBP. Nat. Cell Biol. 21, 338–347 (2019).

    Article  CAS  PubMed  Google Scholar 

  12. Gutscher, M. et al. Real-time imaging of the intracellular glutathione redox potential. Nat. Methods 5, 553–559 (2008).

    Article  CAS  PubMed  Google Scholar 

  13. Tao, R. et al. Genetically encoded fluorescent sensors reveal dynamic regulation of NADPH metabolism. Nat. Methods 14, 720–728 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. 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 

  15. Jiang, L. et al. Reductive carboxylation supports redox homeostasis during anchorage-independent growth. Nature 532, 255–258 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Schafer, Z. T. et al. Antioxidant and oncogene rescue of metabolic defects caused by loss of matrix attachment. Nature 461, 109–113 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Sasaki, H. et al. Electrophile response element-mediated induction of the cystine/glutamate exchange transporter gene expression. J. Biol. Chem. 277, 44765–44771 (2002).

    Article  CAS  PubMed  Google Scholar 

  18. 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 

  19. Sies, H. & Jones, D. P. Reactive oxygen species (ROS) as pleiotropic physiological signalling agents. Nat. Rev. Mol. Cell Biol. 21, 363–383 (2020).

    Article  CAS  PubMed  Google Scholar 

  20. Agyeman, A. S. et al. Transcriptomic and proteomic profiling of KEAP1 disrupted and sulforaphane treated human breast epithelial cells reveals common expression profiles. Breast Cancer Res. Treat. 132, 175–187 (2012).

    Article  CAS  PubMed  Google Scholar 

  21. Hayes, J. D. & Dinkova-Kostova, A. T. The Nrf2 regulatory network provides an interface between redox and intermediary metabolism. Trends Biochem. Sci. 39, 199–218 (2014).

    Article  CAS  PubMed  Google Scholar 

  22. Panieri, E., Telkoparan-Akillilar, P., Suzen, S. & Saso, L. The NRF2/KEAP1 axis in the regulation of tumor metabolism: mechanisms and therapeutic perspectives. Biomolecules 10, 791 (2020).

  23. Romero, R. et al. Keap1 loss promotes Kras-driven lung cancer and results in dependence on glutaminolysis. Nat. Med. 23, 1362–1368 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Rodriguez-Barrueco, R. et al. Inhibition of the autocrine IL-6–JAK2–STAT3–calprotectin axis as targeted therapy for HR/HER2+ breast cancers. Genes Dev. 29, 1631–1648 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Vera-Ramirez, L., Vodnala, S. K., Nini, R., Hunter, K. W. & Green, J. E. Autophagy promotes the survival of dormant breast cancer cells and metastatic tumour recurrence. Nat. Commun. 9, 1944 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  26. Medina, S. H. et al. Identification of a mechanogenetic link between substrate stiffness and chemotherapeutic response in breast cancer. Biomaterials 202, 1–11 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Havas, K. M. et al. Metabolic shifts in residual breast cancer drive tumor recurrence. J. Clin. Invest. 127, 2091–2105 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  28. Jing, H. et al. Early evaluation of relative changes in tumor stiffness by shear wave elastography predicts the response to neoadjuvant chemotherapy in patients with breast cancer. J. Ultrasound Med. 35, 1619–1627 (2016).

    Article  PubMed  Google Scholar 

  29. Hsu, C.-K. et al. Caveolin-1 controls hyperresponsiveness to mechanical stimuli and fibrogenesis-associated RUNX2 activation in keloid fibroblasts. J. Invest. Dermatol. 138, 208–218 (2018).

    Article  CAS  PubMed  Google Scholar 

  30. Huang, C. & Ogawa, R. Fibroproliferative disorders and their mechanobiology. Connect. Tissue Res. 53, 187–196 (2012).

    Article  CAS  PubMed  Google Scholar 

  31. Dupont, S. et al. Role of YAP/TAZ in mechanotransduction. Nature 474, 179–183 (2011).

    Article  CAS  PubMed  Google Scholar 

  32. Ahn, S.-G. & Thiele, D. J. Redox regulation of mammalian heat shock factor 1 is essential for Hsp gene activation and protection from stress. Genes Dev. 17, 516–528 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Mendillo, M. L. et al. HSF1 drives a transcriptional program distinct from heat shock to support highly malignant human cancers. Cell 150, 549–562 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Tharp, K. M. et al. Adhesion-mediated mechanosignaling forces mitohormesis. Cell Metab. 33, 1322–1341 (2021).

    Article  CAS  PubMed  Google Scholar 

  35. Luo, M. et al. Heat stress activates YAP/TAZ to induce the heat shock transcriptome. Nat. Cell Biol. 22, 1447–1459 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Chandel, N. S. Evolution of mitochondria as signaling organelles. Cell Metab. 22, 204–206 (2015).

    Article  CAS  PubMed  Google Scholar 

  37. Shadel, G. S. & Horvath, T. L. Mitochondrial ROS signaling in organismal homeostasis. Cell 163, 560–569 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Yun, J. & Finkel, T. Mitohormesis. Cell Metab. 19, 757–766 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Dixon, S. J. & Stockwell, B. R. The hallmarks of ferroptosis. Ann. Rev. Cancer Biol. 3, 35–54 (2019).

    Article  Google Scholar 

  40. Ursini, F. & Maiorino, M. Lipid peroxidation and ferroptosis: the role of GSH and GPx4. Free Radic. Biol. Med. 152, 175–185 (2020).

    Article  CAS  PubMed  Google Scholar 

  41. Bertero, T. et al. Tumor-stroma mechanics coordinate amino acid availability to sustain tumor growth and malignancy. Cell Metab. 29, 124–140 (2019).

    Article  CAS  PubMed  Google Scholar 

  42. Chakraborty, M. et al. Mechanical stiffness controls dendritic cell metabolism and function. Cell Rep. 34, 108609 (2021).

    Article  CAS  PubMed  Google Scholar 

  43. Khan, A. U. H. et al. Mitochondrial complex I activity signals antioxidant response through ERK5. Sci. Rep. 8, 7420 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  44. Santacatterina, F. et al. Down-regulation of oxidative phosphorylation in the liver by expression of the ATPase inhibitory factor 1 induces a tumor-promoter metabolic state. Oncotarget 7, 490–508 (2015).

    Article  PubMed Central  Google Scholar 

  45. Wu, D. et al. Identification of novel dynamin-related protein 1 (Drp1) GTPase inhibitors: therapeutic potential of Drpitor1 and Drpitor1a in cancer and cardiac ischemia-reperfusion injury. FASEB J. 34, 1447–1464 (2020).

    Article  CAS  PubMed  Google Scholar 

  46. Giacomello, M., Pyakurel, A., Glytsou, C. & Scorrano, L. The cell biology of mitochondrial membrane dynamics. Nat. Rev. Mol. Cell Biol. 21, 204–224 (2020).

    Article  CAS  PubMed  Google Scholar 

  47. Kraus, F., Roy, K., Pucadyil, T. J. & Ryan, M. T. Function and regulation of the divisome for mitochondrial fission. Nature 590, 57–66 (2021).

    Article  CAS  PubMed  Google Scholar 

  48. Cieri, D. et al. SPLICS: a split green fluorescent protein-based contact site sensor for narrow and wide heterotypic organelle juxtaposition. Cell Death Differ. 25, 1131–1145 (2018).

    Article  CAS  PubMed  Google Scholar 

  49. Vallese, F. et al. An expanded palette of improved SPLICS reporters detects multiple organelle contacts in vitro and in vivo. Nat. Commun. 11, 6069 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Bordt, E. A. et al. The putative Drp1 inhibitor mdivi-1 is a reversible mitochondrial complex I inhibitor that modulates reactive oxygen species. Dev. Cell 40, 583–594 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Cassidy-Stone, A. et al. Chemical inhibition of the mitochondrial division dynamin reveals its role in Bax/Bak-dependent mitochondrial outer membrane permeabilization. Dev. Cell 14, 193–204 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Atkins, K., Dasgupta, A., Chen, K.-H., Mewburn, J. & Archer, S. L. The role of Drp1 adaptor proteins MiD49 and MiD51 in mitochondrial fission: implications for human disease. Clin. Sci. 130, 1861–1874 (2016).

    Article  CAS  Google Scholar 

  53. Dasgupta, A. et al. An epigenetic increase in mitochondrial fission by MiD49 and MiD51 regulates the cell cycle in cancer: diagnostic and therapeutic implications. FASEB J. 34, 5106–5127 (2020).

    Article  CAS  PubMed  Google Scholar 

  54. Koirala, S. et al. Interchangeable adaptors regulate mitochondrial dynamin assembly for membrane scission. Proc. Natl Acad. Sci. USA 110, E1342–E1351 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Osellame, L. D. et al. Cooperative and independent roles of the Drp1 adaptors Mff, MiD49 and MiD51 in mitochondrial fission. J. Cell Sci. 129, 2170–2181 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Lomakin, A. J. et al. Competition for actin between two distinct F-actin networks defines a bistable switch for cell polarization. Nat. Cell Biol. 17, 1435–1445 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Pocaterra, A. et al. Fascin1 empowers YAP mechanotransduction and promotes cholangiocarcinoma development. Commun. Biol. 4, 1–13 (2021).

    Article  Google Scholar 

  58. Suarez, C. & Kovar, D. R. Internetwork competition for monomers governs actin cytoskeleton organization. Nat. Rev. Mol. Cell Biol. 17, 799–810 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Yang, Q., Zhang, X.-F., Pollard, T. D. & Forscher, P. Arp2/3 complex-dependent actin networks constrain myosin II function in driving retrograde actin flow. J. Cell Biol. 197, 939–956 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Carlier, M.-F. & Shekhar, S. Global treadmilling coordinates actin turnover and controls the size of actin networks. Nat. Rev. Mol. Cell Biol. 18, 389–401 (2017).

    Article  CAS  PubMed  Google Scholar 

  61. Korobova, F., Ramabhadran, V. & Higgs, H. N. An actin-dependent step in mitochondrial fission mediated by the ER-associated formin INF2. Science 339, 464–467 (2013).

    Article  CAS  PubMed  Google Scholar 

  62. Li, S. et al. Transient assembly of F-actin on the outer mitochondrial membrane contributes to mitochondrial fission. J. Cell Biol. 208, 109–123 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Manor, U. et al. A mitochondria-anchored isoform of the actin-nucleating spire protein regulates mitochondrial division. eLife 4, e08828 (2015).

  64. Posern, G., Sotiropoulos, A. & Treisman, R. Mutant actins demonstrate a role for unpolymerized actin in control of transcription by serum response factor. Mol. Biol. Cell 13, 4167–4178 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Helle, S. C. J. et al. Mechanical force induces mitochondrial fission. eLife 6, e30292 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  66. Kleele, T. et al. Distinct fission signatures predict mitochondrial degradation or biogenesis. Nature 593, 435–439 (2021).

    Article  CAS  PubMed  Google Scholar 

  67. Moore, A. S. et al. Actin cables and comet tails organize mitochondrial networks in mitosis. Nature 591, 659–664 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Barkan, D. & Chambers, A. F. β1-Integrin: a potential therapeutic target in the battle against cancer recurrence. Clin. Cancer Res. 17, 7219–7223 (2011).

    Article  CAS  PubMed  Google Scholar 

  69. Montagner, M. & Sahai, E. In vitro models of breast cancer metastatic dormancy. Front. Cell Dev. Biol. 8, 37 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  70. Klein, C. A. Cancer progression and the invisible phase of metastatic colonization. Nat. Rev. Cancer 20, 681–694 (2020).

    Article  CAS  PubMed  Google Scholar 

  71. Barkan, D. et al. Inhibition of metastatic outgrowth from single dormant tumor cells by targeting the cytoskeleton. Cancer Res. 68, 6241–6250 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Barkan, D. et al. Metastatic growth from dormant cells induced by a Col-I-enriched fibrotic environment. Cancer Res. 70, 5706–5716 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Alsafadi, H. N. et al. Applications and approaches for three-dimensional precision-cut lung slices. Disease modeling and drug discovery. Am. J. Respir. Cell Mol. Biol. 62, 681–691 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Giobbe, G. G. et al. Extracellular matrix hydrogel derived from decellularized tissues enables endodermal organoid culture. Nat. Commun. 10, 5658 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  75. Maghsoudlou, P. et al. Preservation of micro-architecture and angiogenic potential in a pulmonary acellular matrix obtained using intermittent intra-tracheal flow of detergent enzymatic treatment. Biomaterials 34, 6638–6648 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Liu, F. et al. Feedback amplification of fibrosis through matrix stiffening and COX-2 suppression. J. Cell Biol. 190, 693–706 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Girton, T. S., Oegema, T. R. & Tranquillo, R. T. Exploiting glycation to stiffen and strengthen tissue equivalents for tissue engineering. J. Biomed. Mater. Res. 46, 87–92 (1999).

    Article  CAS  PubMed  Google Scholar 

  78. Roy, R., Boskey, A. & Bonassar, L. J. Processing of type I collagen gels using non-enzymatic glycation. J. Biomed. Mater. Res. A 93, 843–851 (2010).

    PubMed  PubMed Central  Google Scholar 

  79. Goddard, E. T., Bozic, I., Riddell, S. R. & Ghajar, C. M. Dormant tumour cells, their niches and the influence of immunity. Nat. Cell Biol. 20, 1240–1249 (2018).

    Article  CAS  PubMed  Google Scholar 

  80. Rehman, J. et al. Inhibition of mitochondrial fission prevents cell cycle progression in lung cancer. FASEB J. 26, 2175–2186 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Takahashi, N. et al. 3D culture models with CRISPR screens reveal hyperactive NRF2 as a prerequisite for spheroid formation via regulation of proliferation and ferroptosis. Mol. Cell https://doi.org/10.1016/j.molcel.2020.10.010 (2020).

  82. Chen, K., Wang, Y., Deng, X., Guo, L. & Wu, C. Extracellular matrix stiffness regulates mitochondrial dynamics through PINCH-1- and kindlin-2-mediated signalling. Curr. Res. Cell Biol. 2, 100008 (2021).

    Article  Google Scholar 

  83. Yang, H. et al. Materials stiffness-dependent redox metabolic reprogramming of mesenchymal stem cells for secretome-based therapeutic angiogenesis. Adv. Healthcare Mater. 8, 1900929 (2019).

    Article  CAS  Google Scholar 

  84. Khacho, M. et al. Mitochondrial dynamics impacts stem cell identity and fate decisions by regulating a nuclear transcriptional program. Cell Stem Cell 19, 232–247 (2016).

    Article  CAS  PubMed  Google Scholar 

  85. Hatch, A. L., Ji, W.-K., Merrill, R. A., Strack, S. & Higgs, H. N. Actin filaments as dynamic reservoirs for Drp1 recruitment. Mol. Biol. Cell 27, 3109–3121 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Palmer, C. S. et al. MiD49 and MiD51, new components of the mitochondrial fission machinery. EMBO Rep. 12, 565–573 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Moore, A. S., Wong, Y. C., Simpson, C. L. & Holzbaur, E. L. F. Dynamic actin cycling through mitochondrial subpopulations locally regulates the fission-fusion balance within mitochondrial networks. Nat. Commun. 7, 12886 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Liu, X. & Hajnóczky, G. Altered fusion dynamics underlie unique morphological changes in mitochondria during hypoxia–reoxygenation stress. Cell Death Differ. 18, 1561–1572 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Miyazono, Y. et al. Uncoupled mitochondria quickly shorten along their long axis to form indented spheroids, instead of rings, in a fission-independent manner. Sci. Rep. 8, 350 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  90. Maguire, S. L. et al. Three-dimensional modelling identifies novel genetic dependencies associated with breast cancer progression in the isogenic MCF10 model. J. Pathol. 240, 315–328 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Pocaterra, A. et al. F-actin dynamics regulates mammalian organ growth and cell fate maintenance. J. Hepatol. 71, 130–142 (2019).

    Article  CAS  PubMed  Google Scholar 

  92. Ignesti, M. et al. A polydnavirus-encoded ANK protein has a negative impact on steroidogenesis and development. Insect Biochem. Mol. Biol. 95, 26–32 (2018).

    Article  CAS  PubMed  Google Scholar 

  93. Romani, P. et al. Dynamin controls extracellular level of Awd/Nme1 metastasis suppressor protein. Naunyn Schmiedebergs Arch. Pharmacol. 389, 1171–1182 (2016).

    Article  CAS  PubMed  Google Scholar 

  94. Romani, P., Duchi, S., Gargiulo, G. & Cavaliere, V. Evidence for a novel function of Awd in maintenance of genomic stability. Sci. Rep. 7, 16820 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  95. Chen, K.-H. et al. Epigenetic dysregulation of the dynamin-related protein 1 binding partners MiD49 and MiD51 increases mitotic mitochondrial fission and promotes pulmonary arterial hypertension: mechanistic and therapeutic implications. Circulation 138, 287–304 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Enzo, E. et al. Aerobic glycolysis tunes YAP/TAZ transcriptional activity. EMBO J. 34, 1349–1370 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Hagen, C. K. et al. High contrast microstructural visualization of natural acellular matrices by means of phase-based x-ray tomography. Sci. Rep. 5, 18156 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Urciuolo, A. et al. Intravital three-dimensional bioprinting. Nat. Biomed. Eng. 4, 901–915 (2020).

    Article  CAS  PubMed  Google Scholar 

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

  100. Montagner, M. et al. Crosstalk with lung epithelial cells regulates Sfrp2-mediated latency in breast cancer dissemination. Nat. Cell Biol. 22, 289–296 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Shibue, T., Brooks, M. W. & Weinberg, R. A. An integrin-linked machinery of cytoskeletal regulation that enables experimental tumor initiation and metastatic colonization. Cancer Cell 24, 481–498 (2013).

    Article  CAS  PubMed  Google Scholar 

  102. Aston, W. J. et al. A systematic investigation of the maximum tolerated dose of cytotoxic chemotherapy with and without supportive care in mice. BMC Cancer 17, 684 (2017).

Download references

Acknowledgements

We thank D. De Stefani for help with plasmids and advice on toroidal mitochondria, L. Scorrano and M. Zamberlan for protocols to measure mitochondrial morphology and reagents to study DRP1, C. Laterza for help with tissue sectioning, J. Weedon who contributed to the set-up of the bleomycin model, and I. Szabo for help with the transmission electron microscopy and critically reading the manuscript. This study was supported by: Worldwide Cancer Research grant no. 21-0156, AIRC Foundation Investigator grant no. 21392 and a CARIPARO Excellence Grant to S.D.; P.R. is supported by a Veronesi Foundation Postdoctoral Fellowship; AIRC Foundation Investigator grant no. 2018-ID 2135, AIRC Foundation Investigator Grant 5 per mille grant no. 22759, Ministero della Salute grant no. RCR-2019-23669115, Ministero della Salute grant no. NET-2016-02361632 and Istituto Oncologico Veneto to A.R.; Giovanni Armenise–Harvard Foundation and ERC Starting Grant (MetEpiStem) to G.M.; Canadian Institutes for Health Research (CIHR) Foundation Grant, National Institutes of Health grant nos R01-HL071115 and 1RC1HL099462, a Tier I Canada Research Chair and the William J. Henderson Foundation to S.L.A.; Longfonds (Voorhen Astma Fonds) BREATH Consortium, NIHR GOSH Biomedical Research Centre, the National Institute for Health Research grant no. NIHR-RP-2014-04-046 to P.D.C.; F.M. is supported by a GOSH BRC Catalyst Fellowship; University of Padua STARS Consolidator Grant 2019 to T.C.; University of Padua TWINING-VELT 2017 Grant to N.E.; and Fondazione IRP Città della Speranza to A.U.

Author information

Authors and Affiliations

Authors

Contributions

P.R. carried out experiments and analysed data with help from N.N. M.A. retrieved raw data for gene-set enrichment analysis and performed RNA sequencing analysis, with support from G.M. A.T. and V.B. performed the immunohistochemistry and helped with the mouse experiments, with support from A.R. D.W. performed the analysis of mitochondrial morphology; C.C.T.H. and S.L.A. provided Drpitor1a and support. F.M., S.S. and T.B. prepared the mouse lung decellularized ECM slices and performed the lung infiltration experiments, with support from P.D.C. M.G. and A.U. performed the AFM measurements, with support from N.E. F.G. helped with the SPLICS experiments, with support from T.C. A.R. performed the GSH and GSSG quantifications. P.C. and M.M. helped with the gene-set enrichment analysis. S.D. conceptualized, coordinated and supervised the project, and wrote the manuscript.

Corresponding author

Correspondence to Sirio Dupont.

Ethics declarations

Competing interests

The invention of Drpitor1a is the subject of the US Patent Application 20200323829 by D.W. and S.L.A. The remaining authors declare no competing interests.

Peer review

Peer review information

Nature Cell Biology thanks the anonymous reviewers for their contribution to the peer review of this work. Peer reviewer reports are available.

Additional information

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

Extended data

Extended Data Fig. 1 ECM stiffness regulates cystine metabolism and glutathione oxidation.

a, A simplified scheme depicting the major metabolic intermediates of the glutathione synthesis pathway, together with the enzymes mediating each reaction. In blue, the intermediates increased in MCF10A-RAS cells cultured in conditions of decreased actomyosin contractility. b, Representative gating scheme for quantification of FITC-labelled cystine uptake by flow cytometry. Cells without stain were used as negative control (outlined white area). Cells treated with Y27632 and ML7 (YM - orange) display increased uptake compared to DMSO (grey), while treatment with the sulfasalazine (SAS - magenta) inhibitor of the cystine transporter inhibits uptake. Scale bar, 50 μm. On the right: quantification of FITC-labelled cystine uptake. Data are mean and single points relative to DMSO (black bar). P-values by Dunnet’s test from a sample size of n = 4 biologically independent samples pooled across two independent experiments for each bar. c,d, Glutathione redox analysis using the mitochondrial mito-Grx1-roGFP2 sensor in MCF10A-RAS cells treated with YM (c), or cultured on stiff or soft Matrigel substrata (d). Data are mean and single cells. P-values by Dunnet’s test or Student’s t-test from a sample size of n = 21 in c and n = 31 in d cells pooled across two independent experiments for each bar. e-g, Direct quantification of reduced (GSH, e) and oxidized (GSSH, f) glutathione in extracts from MCF10A-RAS cells treated with YM for 24 h. The ratio of reduced to oxidized glutathione was calculated in g. Data are mean and single points. n = 2 independent experiments. Images in b,c are representative of two independent experiments with similar results. See Source Data Extended Data Fig. 1.

Source Data

Extended Data Fig. 2 ECM stiffness regulates ROS levels.

a, NADPH redox analysis using the iNap1 sensor in MCF10A-RAS cells treated with YM. Co-treatment with Diamide (DIA) and the G6PD inhibitor DHEA serves as positive control for NADPH depletion. iNapC is a NADP-insensitive and pH-sensitive control sensor (n = 35 cells pooled across two independent experiments for each bar; unpaired two-tailed Student’s t-test). b,c, Representative gating schemes and quantifications of DCFDA (b) and MitoSOX (c) oxidation by flow cytometry. Cells without stain were used as negative control (outlined white area). Cells treated with the positive control cumene hydroperoxide (CUM - purple) display increased uptake compared to DMSO (grey). Median intensity in the controls were set to 1, and other samples are relative to these (n = 6 biologically independent samples pooled across three independent experiments for each bar; unpaired two-tailed Student’s t-test). d-i, Quantification of reactive-oxygen species by the DCFDA and MitoSOX reagents in MCF10A, MCF10A-RAS and D2.0R cells. Median intensity in the controls were set to 1, and other samples are relative to these (in d n = 4 (STIFF) and n = 5 (all other conditions); in e n = 11 (DMSO) n = 6 (YM and SOFT) n = 5 (BLEBBI) n = 8 (STIFF); in f n = 4 (STIFF) n = 5 (all other condotions); in g n = 4 (DMSO and YM) n = 6 (STIFF and SOFT); in h n = 6 (DMSO, YM6h and YM 24 h) n = 4 (YM1h and YM 3 h); in i n = 5 (YM 1 h) and n = 6 (all other conditions) biologically independent samples pooled across two independent experiments; unpaired two-tailed Student’s t-test or Dunnet’s test). j, Representative gating scheme and quantification of C11-Bodipy-581/591 lipid peroxidation by flow cytometry. Cells without C11-Bodipy-581/591 were used as negative control (outlined white area). Cells treated with cumene hydroperoxide (CUM - purple) or with the GPX4 lipid hydroperoxide glutathione peroxidase inhibitor RSL3 (blue) for 3 h were used as positive controls. Mean intensity in the controls were set to 1, and other samples are relative to these (n = 4 biologically independent samples pooled across two independent experiments for each bar; Dunnet’s test). k,l, Quantification of C11-Bodipy-581/591 lipid peroxidation in MCF10A-RAS and D2.0R cells treated with YM for 3 h or cultured on Fibronectin-coated stiff or soft acrylamide hydrogels. Median intensity in the controls were set to 1, and other samples are relative to these (n = 4 biologically independent samples pooled across two independent experiments for each bar; unpaired two-tailed Student’s t-test). m, Immunofluorescence of 4-hydroxy-2-nonenal (4HNE) lipid peroxidation adducts in MCF10A-RAS cells cultured on stiff or soft Matrigel substrata for 24 h. Images are representative of at two experiments with similar results. Scale bar, 5 μm. Mean intensity in the controls were set to 1, and other samples are relative to these (n = 36 (STIFF) n = 34 (SOFT) cells pooled across two independent experiments; unpaired two-tailed Student’s t-test). Data are mean and single points. See Source Data Extended Data Fig. 2.

Source Data

Extended Data Fig. 3 ECM stiffness regulates NRF2 activity.

a,b, qPCR for established NRF2 target genes in D2.0R and MCF10A cells treated with YM for 6 h (in a FTH1 n = 2 (DMSO) n = 4 (YM); HMOX1 n = 2 (DMSO) n = 4 (YM); NQO1 n = 2 (DMSO) n = 4 (YM); GCLC n = 2 (DMSO) n = 4 (YM); SLC7A11 n = 7 (DMSO) n = 5 (YM) independent samples pooled across two independent experiments; in b HMOX1 n = 8; FTH1 n = 8 (DMSO) n = 7 (YM); GCLC n = 8 (DMSO) n = 6 (YM); GCLM n = 8 (DMSO) n = 6 (YM); SLC7A11 n = 8 (DMSO) n = 6 (YM) independent samples pooled across two independent experiments; unpaired two-tailed Student’s t-tests. c qPCR for established NRF2 target genes in D2.0R cells cultured on Matrigel substrata of different stiffness. (Fth1 n = 7(STIFF) n = 5 (SOFT); Hmox1 n = 4; Nqo1 n = 7; Gclm n = 4; SLC7A11 n = 6 (STIFF) n = 5 (SOFT) independent samples pooled across two independent experiments; unpaired two-tailed Student’s t-tests. d qPCR for established NRF2 target genes in MCF10A-RAS (FTH1 n = 4; HMOX1 n = 6 (DMSO and YM) n = 4 (Ki696); NQO1 n = 6 (DMSO and YM) n = 4 (ki696); GCLC n = 6 (DMSO and YM) n = 4 (ki696) independent samples pooled across two independent experiments; unpaired two-tailed Student’s t-tests. e-g, qPCR for established YAP/TAZ targets in D2.0R (e) and NRF2 and YAP/TAZ target genes in MCF10A-RAS (f,g). YAP/TAZ targets serve as internal positive controls.(in e n = 4, in f n = 4, in g n = 10 samples pooled across two independent experiments for each bars; unpaired two-tailed Student’s t-tests). h, qPCR for established NRF2 target genes in MCF10A cells cultured on soft (E ≈ 0.2kPa) or stiff (E ≈ 50kPa) CollagenI-coated hydrogels of different stiffness (n = 4 samples pooled across two independent experiments for each bars; unpaired two-tailed Student’s t-tests). i, Quantification of active nuclear S40-phosphorylated NRF2 (pNRF2) intensity from immunofluorescence stainings of MCF10A-RAS and D2.0R cells cultured on stiff or soft Matrigel substrata. KI696 is a KEAP1 inhibitor used as a positive control. Mean levels in the controls were set to 1, and other samples are relative to these (n = 59 (DMSO) n = 60 (YM and ki696) cells were imaged for each condition, pooled across two independent experiments; unpaired two-tailed Student’s t-tests). j, qPCR to control for efficient knockdown of NEF2L2 (encoding for NRF2) in MCF10A-RAS (n = 10 (siCO) n = 8 (siNRF2a and siNRF2b) samples pooled across two independent experiments for each bars; unpaired two-tailed Student’s t-tests). k, Immunoblotting for endogenous NRF2 from extracts of MCF10A-RAS cells transfected with the indicated siRNAs. Equal total proteins were loaded in each lane, and GAPDH was used as loading control. Images are representative of two independent experiments with similar results. Unprocessed blots in Source Data Extended Data Fig. 3. j, qPCR to control for efficient knockdown of Nfe2L2 in D2.0R cells (n = 5 (shCO) n = 6 (shNrf2a and shNrf2b) samples pooled across two independent experiments for each bars; unpaired two-tailed Student’s t-tests). Data are mean and single points. mRNA expression data are relative to GAPDH levels; mean level in the control was set to 1, and other samples are relative to this. See Source Data Extended Data Fig. 3.

Source Data

Extended Data Fig. 4 Activation of NRF2 gene signatures in response to a soft ECM in published datasets.

a, Gene-Set Enrichment Analysis (GSEA) was performed on genes upregulated by culturing MCF10A cells on stiff 2D Matrigel-coated plates (red) or soft 3D Matrigel gels (blue). NRF2 signatures: KEAP/SFN (normalized enrichment score NES = −1.53 P < 0.0001 false discovery rate FDR = 0.038), LIST (NES = −1.91 P < 0.0001 FDR < 0.0001), PANIERI (NES = −1.75 P < 0.0001 FDR = 0.0006), HER2 (NES = −1.80 P < 0.001 FDR = 0.0005). YAP/TAZ (NES = 2.01 P < 0.0001 FDR < 0.0001) and SREBP (NES = −1.47 P < 0.017 FDR 0.06) serve as positive controls for gene signatures regulated by ECM stiffness. b, GSEA was performed on genes upregulated by culturing D2.0R cells on stiff Matrigel+Collagen-I gels (red) or soft Matrigel gels (blue). NRF2 signatures: KEAP/SFN (NES = −2.11 P < 0.0001 FDR < 0.0001), LIST (NES = −2.11 P < 0.0001 FDR < 0.0001), HER2 (NES = −1.86 P < 0.001 FDR = 0.0008). YAP/TAZ (NES = 1.64 P = 0.004 FDR = 0.009) and SREBP (NES = −1.77 P < 0.0001 FDR = 0.0026) serve as positive controls for gene signatures regulated by ECM stiffness. c, GSEA was performed on genes upregulated by culturing human MDA-MB-453 breast cancer cells on stiff (red) Fibronectin-coated plates or soft (blue) acrylamide hydrogels. NRF2 signatures: KEAP/SFN (NES = −3.25 P < 0.001 FDR < 0.001), LIST (NES = −3.60 P < 0.001 FDR < 0.001), HER2 (NES = −4.50 P < 0.001 FDR < 0.001). The YAP/TAZ signature (NES = 1.71 P = 0.025 FDR = 0.048 Zhao NES = 2.43 P < 0.001 FDR < 0.001) serves as positive control for genes regulated by ECM stiffness. d, Heatmap of NRF2 target genes in RNAseq data of mouse MMTV-PyMT breast cancer cells cultured ex vivo on Fibronectin-coated plates, stiff or soft acrylamide hydrogels. Each column is an independent biological sample (n = 4 for each condition); each line is a single gene. Expression levels for each gene were normalized relative to the mean expression on plates, which was set to 1 (white). Black and blue indicate downregulation and upregulation (P < 0.05), respectively. YAP/TAZ target genes serve as positive controls. e, GSEA was performed on genes regulated in matched cases of pre-treatment fine-needle aspiration (PRE, red) and the respective post neoadjuvant treatment operative sample (POST, blue) of primary breast cancer patients. Neoadjuvant treatment is known to reduce tumour stiffness. NRF2 signatures: KEAP/SFN (NES = −1.32 P = 0.012 FDR = 0.096), LIST (NES = −1.34 P = 0.11 FDR = 0.088), PANIERI (NES = −1.34 P = 0.093 FDR = 0.087). SREBP (NES = −1.94 P < 0.001 FDR < 0.001) and YAP/TAZ (NES = 2.56 P < 0.001 FDR < 0.001) serve as positive controls for gene signatures regulated by ECM stiffness. f, Heatmap of NRF2 and YAP target levels in n = 7 patient-matched stiff keloid tissue or soft normal skin. Each column represents -log2(keloid/skin) values for a single patient; each line is a single gene probe; genes ranked according to expression in patient #1. Black and blue indicate downregulation and upregulation (P < 0.05), respectively. YAP target genes serve as positive controls. g,h, qPCR for established HSF1 target genes in MCF10A cells cultured on Fibronectin-coated stiff or soft hydrogels (g), or transfected on plastics with control (siCO) or with YAP/TAZ siRNAs (h). Data are mean and single points. mRNA expression data are relative to GAPDH levels; mean level in the control was set to 1, and other samples are relative to this (n = 2 biologically independent samples from one experiment). See Source Data Extended Data Fig. 4.

Source Data

Extended Data Fig. 5 ECM stiffness regulates resistance to oxidative stress.

a, Representative gating scheme for quantification of CytoGrx1-roGFP2 by flow cytometry. Cells without CytoGrx1-roGFP2 expression were used as negative control (outlined white area). Cells treated with cumene hydroperoxide (CUM, purple) for 10 min. to induce glutathione oxidation were used as positive control (n = 4 biologically independent samples pooled across two independent experiments for each bar; unpaired two-tailed Student’s t-test). b, Cell survival by the resazurin assay in MCF10A cells with the indicated knockdowns, plated on stiff or soft Matrigel substrata, and treated for 48 h with Cumene hydroperoxide (CUM). Mean cell number in controls was set to 100%, and all other samples are relative to this (n = 12 biologically independent samples pooled across two independent experiments for each bar; Dunnet’s tests). c, Cell survival of MCF10A-RAS cells transfected with the indicated KEAP1 siRNAs and treated for 48 h with Cumene hydroperoxide (CUM). Mean cell number in Stiff controls was set to 100%, and all other samples are relative to this (n = 12 biologically independent samples pooled across two independent experiments for each bar; Dunnet’s test). d, qPCR to control for efficient knockdown of KEAP1 in MCF10A-RAS. mRNA expression data are relative to GAPDH levels; mean level in the control was set to 1, and other samples are relative to this (n = 4 biologically independent samples from two independent experiments; Dunnet’s test). e, Immunoblotting for endogenous KEAP1 from extracts of MCF10A-RAS cells transfected with the indicated siRNAs. Equal total proteins were loaded in each lane, and GAPDH was used as loading control. Images are representative of two independent experiments with similar results. Unprocessed blots in Source Data Extended Data Fig. 5. f, Cell survival of MCF10A-RAS cells transfected with the indicated NRF1 siRNA, plated on stiff or soft Matrigel substrata, and treated for 48 h with Cumene hydroperoxide (CUM). Mean cell number in Stiff controls was set to 100%, and all other samples are relative to this (n = 12 biologically independent samples pooled across two independent experiments for each bar; unpaired two-tailed Student’s t-tests). g, qPCR for NFE2L1 (encoding for NRF1) and NFE2L2 (encoding for NRF2) expression levels in MCF10A-RAS transfected with NRF1 siRNAs. mRNA expression data are relative to GAPDH levels; mean level in the control was set to 1, and other samples are relative to this (n = 6 biologically independent samples from two independent experiments; Dunnet’s test). h,i, Cell survival of MCF10A-RAS cells treated for 48 h with Erastin (ERA, g) or RSL3 (h) alone and together with the anti-ferroptosis compounds Liproxstatin-1 (LIP) or Ferrostatin (FER). Mean cell number in Stiff controls was set to 100%, and all other samples are relative to this (in h n = 7 (VEHICLE DMSO, ERASTIN 10μM FER, ERASTIN 10μM LIP, ERASTIN 30μM DMSO) n = 6 (ERASTIN 10μM DMSO) n = 8 (all other conditions) biologically independent samples pooled from a single experiment; in i n = 8 (VEHICLE) n = 6 (RSL3 5μM LIP) n = 7 (all other conditions) biologically independent samples pooled from a single experiment; Dunnet’s tests). j, qPCR for YAP/TAZ targets in MCF10A-RAS cells used as a control for differential stiffness of the Matrigel substrata within the extracellular flux analyser plates (see Fig. 4f). mRNA expression data are relative to GAPDH levels; mean level in the control was set to 1, and other samples are relative to this (n = 6 biologically independent samples from three independent experiments; unpaired two-tailed Student’s t-tests). k, qPCR for YAP/TAZ targets in D2.0R cells used as a control for differential stiffness of the Matrigel substrata within the extracellular flux analyser plates (see Fig. 4g). mRNA expression data are relative to GAPDH levels; mean level in the control was set to 1, and other samples are relative to this (n = 4 biologically independent samples from two independent experiments; unpaired two-tailed Student’s t-tests). Data are mean and single points. See Source Data Extended Data Fig. 5.

Source Data

Extended Data Fig. 6 ECM stiffness regulates mitochondrial fission through DRP1.

a, Quantification of mitochondria length in D2.0R cells transfected with mitoRFP and treated with YM (n = 128 (DMSO) n = 117(YM) mitochondria pooled across 10 cells per condition in two independent experiments for each bar; unpaired two-tailed Student’s t-tests). The same set of images was analysed by automated classification of mitochondrial length (centre) and by MFC analysis (right). Data are mean and single mitochondria (left), mean and s.d. (centre) and mean and single pictures (right). b, Quantification of mitochondria content by MitoTracker staining in MCF10A-RAS and D2.0R cells treated with YM. Median level in the control was set to 1, and other samples are relative to this (n = 7 biologically independent samples pooled across three independent experiments for MCF10A-RAS; n = 4 biologically independent samples pooled across two independent experiments for D2.0R). c, Quantification of mitochondrial DNA content in MCF10A-RAS cells treated with YM, based on qPCR for three different mtDNA loci. mtDNA levels are relative to genomic DNA (gDNA) levels; mean level in the control was set to 1, and other samples are relative to this (n = 6 biologically independent samples from two independent experiments). d, qPCR for DRP1 in MCF10A cells treated with YM. mRNA expression data are relative to GAPDH levels; mean level in the control was set to 1, and other samples are relative to this (n = 6 biologically independent samples from three independent experiments). e, Immunoblotting for endogenous total and phosphorylated DRP1 (p616, p637) from extracts of MCF10A-RAS cells treated with ROCK/MLCK inhibitors. Equal total proteins were loaded in each lane, and GAPDH was used as loading control. f,g. qPCR to control for efficient knockdown of DRP1 in MCF10A-RAS (f) or in D2.0R (g). mRNA expression data are relative to GAPDH levels; mean level in the control was set to 1, and other samples are relative to this (n = 4 biologically independent samples from two independent experiments for each bar; unpaired two-tailed Student’s t-test in MCF10A-RAS, Dunnet’s test in D2.0R). h. Immunoblotting for endogenous DRP1 from extracts of MCF10A-RAS cells transfected with the indicated siRNAs. Equal total proteins were loaded in each lane, and GAPDH was used as loading control. i. Quantification of mtROS in MCF10A-RAS cells with knockdown of the indicated fission factors, and treated with YM. Median intensity in the control were set to 1, and other samples are relative to these (n = 8 biologically independent samples pooled across four independent experiments for siCO bars; n = 4 biologically independent samples pooled across two independent experiments for other bars; unpaired two-tailed Student’s t-tests). j-m. qPCR in MCF10A-RAS to control for efficient knockdown of the FIS1, MFF, MIEF1 and MIEF2 fission factors. mRNA expression data are relative to GAPDH levels; mean level in the control was set to 1, and other samples are relative to this (n = 4 biologically independent samples from two independent experiments for each bar; Dunnet’s tests). n. Oxygen Consumption Rate (OCR) analysis performed on monolayers of MCF10A-RAS cells transfected with control (siCO) or DRP1 siRNA and plated on stiff or soft Matrigel substrata (n = 20 biologically independent samples pooled across two independent experiments). o. OCR analysis on cells were treated with Drpitor1a (n = 20 biologically independent samples pooled across two independent experiments). p. OCR analysis performed on monolayers of D2.0R cells stably expressing control (shCo.) or DRP1a shRNA and cultured on stiff or soft Matrigel substrata (n = 20 biologically independent samples pooled across two independent experiments). q. OCR analysis on cells were treated with Drpitor1a (n = 20 biologically independent samples pooled across two independent experiments). Images in e,h are representative of two independent experiments with similar results. Unprocessed blots in Source Data Extended Data Fig. 6. Data are mean and single points, except n-q (mean and s.d. - shaded areas). See Source Data Extended Data Fig. 6.

Source Data

Extended Data Fig. 7 A soft ECM increases resistance to Cisplatin and As2O3 chemotherapy through NRF2 and DRP1 in MCF10A-RAS cells.

a,b, Cell survival of MCF10A-RAS cells treated for 48 h with the indicated concentrations of Cisplatin (a) or As2O3 (b), in the absence or presence of the antioxidant N-Acetyl-L-cysteine (NAC, 5 mM) (in a n = 14 (VEHICLE) n = 13 (VEHICLE NAC) n = 15 (CISPLATIN 5μM) n = 16 (CISPLATIN 20 μM) biologically independent samples pooled across two independent experiments; in b n = 22 (VEHICLE and As2O3 20μM) n = 19 (VEHICLE NAC) n = 21 (As2O3 5μM) biologically independent samples pooled across two independent experiments; Dunnet’s tests). c,d, Cell survival assay in MCF10A-RAS cells cultured on stiff Matrigel-coated plates or soft Matrigel gels, and treated for 48 h with the indicated concentrations of Cisplatin (c) or As2O3 (d) (n = 16 biologically independent samples pooled across two independent experiments for each bar; Dunnet’s test). e, Cell survival assay in MCF10A-RAS cells treated for 48 h with the indicated concentrations of Doxorubicin (DOXO), in the absence or presence of N-Acetyl-L-cysteine (NAC, 5 mM) (n = 16 biologically independent samples pooled across two independent experiments for each bar). f, Cell survival assay in MCF10A-RAS cells cultured on stiff Matrigel-coated plates (black) or soft Matrigel gels (green), and treated for 48 h with the indicated concentrations of Doxorubicin (DOXO) (n = 14 biologically independent samples pooled across two independent experiments for each bar). g, Cell survival of MCF10A-RAS cells transiently transfected as indicated and cultured on stiff Matrigel-coated plates or soft Matrigel gels, and treated for 48 h with the indicated concentration of Cisplatin. (n = 8 (siCO and siNRF2b CISPLATIN SOFT) n = 6 (all other conditions) biologically independent samples pooled across two independent experiments; Dunnet’s test). h. Cell survival of MCF10A-RAS cells transiently transfected as indicated, cultured on stiff Matrigel-coated plates or soft Matrigel gels, and treated for 48 h with the indicated concentration of Cisplatin. Where indicated cells were also treated with the DRP1 inhibitor Drpitor1a. (n = 14 (siCO) n = 12 (siDRP1) n = 16 (DRPITOR1A) biologically independent samples pooled across two independent experiments; Dunnet’s test). Data are mean and single points. Mean expression in untreated controls were set to 100%, and all other samples are relative to this. See Source Data Extended Data Fig. 7.

Source Data

Extended Data Fig. 8 Controls to experiments shown in main Fig. 8.

a, A simplified model of the mechanical conditions used throughout Fig. 8. Cells are seeded in a soft microenvironment (wavy lines: in vitro BM ECM, ex vivo decellularized lung ECM, in vivo lung tissue) or directly in stiff conditions (straight bold lines: lung ECM slices treated with Ribose or derived from fibrotic mice). Cancer cells remodel the ECM, such that they stiffen their microenvironment at late time points (LATE: 14 days in vitro, 1.5 months in vivo). b, Quantification of tissue stiffness by atomic force microscopy on mouse mammary glands and normal lungs, using frozen tissue microtome sections. Data are mean and single points (n = 4 and n = 3 tissue sections from two mice). c, YAP/TAZ immunofluorescence in D2.0R cells cultured on normal and Ribose-stiffened (RIB) lung ECM slices. Images are representative of two independent experiments with similar results. Scale bars, 25 μm. d, Collagen-I and YAP/TAZ immunofluorescence in D2.0R cells cultured on normal and fibrosis-stiffened (FIB) lung ECM slices. Images are representative of two independent experiments with similar results. Scale bars, 25 μm. e, qPCR to control for efficient regulation of YAP/TAZ in D2.0R cells infiltrated into normal or fibrotic (FIB) lung ECM scaffolds. Data are mean and single points; mRNA expression data are relative to GAPDH levels; mean level in the control was set to 1, and other samples are relative to this (n = 7 biologically independent samples from two independent experiments for bars 1 and 2; n = 2 biologically independent samples from one single experiment for bars 3). See Source Data Extended Data Fig. 8.

Source Data

Extended Data Fig. 9 Controls to in vivo experiments shown in main Fig. 8.

a, Experimental set-up to study D2.0R metastatic cell behaviour in mice. D2.0R cells expressing EGFP and Firefly-luciferase (EGFP/Luc) were injected via the tail vein to induce metastatic dissemination into the lungs and their initial cell number was quantified by intravital bioluminescence imaging (BLI). After letting cells settle for one week, mice were injected i.p. with Cisplatin or with the equivalent volume of vehicle (1xPBS) for four consecutive rounds. Cell growth was monitored after every injection and then every 11 days. b, EGFP/Luc D2.0R cells expressing the indicated shRNAs were injected via the tail vein mixed at a 1:1 ratio together with mCherry D2.0R cells to induce metastatic spread into the lungs. After three days, the ratio of red/green cells in the lung parenchyma was quantified to measure extravasation efficiency. Scale bars, 20 μm. Images are representative of three mice with similar results. Data are mean and single points (n = 3 mice for each condition). c, Intravital imaging of mice with lung metastases from EGFP/Luc D2.0R cells quantified in Fig. 8l, taken at day 90. The rainbow LUT was used to visualize the radiance. Images are representative of six mice with similar results per condition. d, Quantification of activated Caspase-3 in D2.0R cells disseminated to the mouse lung after treatment with two doses of Cisplatin (see scheme above) (n = 2 mice). On the right, an immunofluorescent picture with spectral unmixing showing GFP-positive D2.0R cells (cyan) disseminated into the lung parenchyma and one example of a double-positive Caspase-3/GFP cell (yellow). Tissue sections were also stained with CD31 (magenta) to visualize blood vessels. Scale bar, 20 μm. e, Intravital imaging of mice with lung metastases from EGFP/Luc D2.0R cells quantified in Fig. 8m, taken at day 75. The rainbow LUT was used to visualize the radiance. Images are representative of six mice with similar results per condition. Data are mean and single points. See Source Data Extended Data Fig. 9.

Source Data

Extended Data Fig. 10 Proposed model.

A simplified model that recapitulates the main findings of the manuscript, ordered in time after shifting cells to a soft microenvironment. On a soft ECM, cells develop reduced actomyosin tension, which is associated with increased formation of peri-mitochondrial F-actin, increased DRP1-dependent mitochondrial fission and enhanced mtROS production. mtROS activate a NRF2-dependent antioxidant metabolic response which includes increased cysteine uptake and glutathione synthesis. Functionally, this response has two main outcomes: it keeps redox balance in the face of increased ROS and glutathione oxidation (upward facing arrow), and it results in a better ability of cells to resist exogenous oxidative stress and ROS-dependent chemotherapy (downward facing arrow), compared to cells in a stiff microenvironment. mt is an abbreviation for mitochondria.

Supplementary information

Source data

Source Data Fig. 1

Statistical source data.

Source Data Fig. 2

Statistical source data.

Source Data Fig. 3

Statistical source data.

Source Data Fig. 4

Statistical source data.

Source Data Fig. 5

Statistical source data.

Source Data Fig. 6

Statistical source data.

Source Data Fig. 7

Statistical source data.

Source Data Fig. 8

Statistical source data.

Source Data Extended Data Fig. 1

Statistical source data.

Source Data Extended Data Fig. 2

Statistical source data.

Source Data Extended Data Fig. 3

Statistical source data.

Source Data Extended Data Fig. 3

Unprocessed western blots.

Source Data Extended Data Fig. 4

Statistical source data.

Source Data Extended Data Fig. 5

Statistical source data.

Source Data Extended Data Fig. 5

Unprocessed western blots.

Source Data Extended Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 6

Unprocessed western blots.

Source Data Extended Data Fig. 7

Statistical source data.

Source Data Extended Data Fig. 8

Statistical source data.

Source Data Extended Data Fig. 9

Statistical source data.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Romani, P., Nirchio, N., Arboit, M. et al. Mitochondrial fission links ECM mechanotransduction to metabolic redox homeostasis and metastatic chemotherapy resistance. Nat Cell Biol 24, 168–180 (2022). https://doi.org/10.1038/s41556-022-00843-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41556-022-00843-w

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