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:

AMPK and PFKFB3 mediate glycolysis and survival in response to mitophagy during mitotic arrest

Abstract

Blocking mitotic progression has been proposed as an attractive therapeutic strategy to impair proliferation of tumour cells. However, how cells survive during prolonged mitotic arrest is not well understood. We show here that survival during mitotic arrest is affected by the special energetic requirements of mitotic cells. Prolonged mitotic arrest results in mitophagy-dependent loss of mitochondria, accompanied by reduced ATP levels and the activation of AMPK. Oxidative respiration is replaced by glycolysis owing to AMPK-dependent phosphorylation of PFKFB3 and increased production of this protein as a consequence of mitotic-specific translational activation of its mRNA. Induction of autophagy or inhibition of AMPK or PFKFB3 results in enhanced cell death in mitosis and improves the anti-tumoral efficiency of microtubule poisons in breast cancer cells. Thus, survival of mitotic-arrested cells is limited by their metabolic requirements, a feature with potential implications in cancer therapies aimed to impair mitosis or metabolism in tumour cells.

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

Access options

Buy this article

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

Figure 1: MCD is modulated by the intrinsic apoptotic pathway.
Figure 2: Autophagy modulates cell death during mitotic arrest.
Figure 3: Reduced mitochondrial mass as a consequence of mitophagy during mitotic arrest.
Figure 4: Metabolic switch from OXPHOS to glycolysis is mediated by AMPK.
Figure 5: Glycolysis determines survival during mitotic arrest.
Figure 6: PFKFB3 regulation during mitotic arrest.
Figure 7: Taxol cooperates with different inhibitors of survival pathways to promote MCD.
Figure 8: Anti-tumoral cooperation between microtubule poisons and PFKFB3 inhibitors in vivo.

Similar content being viewed by others

References

  1. Alvarez-Fernandez, M. & Malumbres, M. Preparing a cell for nuclear envelope breakdown: spatio-temporal control of phosphorylation during mitotic entry. Bioessays 36, 757–767 (2014).

    Article  CAS  PubMed  Google Scholar 

  2. Musacchio, A. & Salmon, E. D. The spindle-assembly checkpoint in space and time. Nat. Rev. Mol. Cell Biol. 8, 379–393 (2007).

    Article  CAS  PubMed  Google Scholar 

  3. Foley, E. A. & Kapoor, T. M. Microtubule attachment and spindle assembly checkpoint signalling at the kinetochore. Nat. Rev. Mol. Cell Biol. 14, 25–37 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Peters, J. M. The anaphase promoting complex/cyclosome: a machine designed to destroy. Nat. Rev. Mol. Cell Biol. 7, 644–656 (2006).

    Article  CAS  PubMed  Google Scholar 

  5. Jia, L., Kim, S. & Yu, H. Tracking spindle checkpoint signals from kinetochores to APC/C. Trends Biochem. Sci. 38, 302–311 (2013).

    Article  CAS  PubMed  Google Scholar 

  6. Komlodi-Pasztor, E., Sackett, D. L. & Fojo, A. T. Inhibitors targeting mitosis: tales of how great drugs against a promising target were brought down by a flawed rationale. Clin. Cancer Res. 18, 51–63 (2012).

    Article  CAS  PubMed  Google Scholar 

  7. Mitchison, T. J. The proliferation rate paradox in antimitotic chemotherapy. Mol. Biol. Cell 23, 1–6 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Manchado, E. et al. Targeting mitotic exit leads to tumor regression in vivo: modulation by Cdk1, Mastl, and the PP2A/B55α,δ phosphatase. Cancer Cell 18, 641–654 (2010).

    Article  CAS  PubMed  Google Scholar 

  9. Gascoigne, K. E. & Taylor, S. S. Cancer cells display profound intra- and interline variation following prolonged exposure to antimitotic drugs. Cancer Cell 14, 111–122 (2008).

    Article  CAS  PubMed  Google Scholar 

  10. Huang, H. C., Shi, J., Orth, J. D. & Mitchison, T. J. Evidence that mitotic exit is a better cancer therapeutic target than spindle assembly. Cancer Cell 16, 347–358 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Rieder, C. L. & Medema, R. H. No way out for tumor cells. Cancer Cell 16, 274–275 (2009).

    Article  CAS  PubMed  Google Scholar 

  12. Topham, C. H. & Taylor, S. S. Mitosis and apoptosis: how is the balance set? Curr. Opin. Cell Biol. 25, 780–785 (2013).

    Article  CAS  PubMed  Google Scholar 

  13. Vitale, I., Galluzzi, L., Castedo, M. & Kroemer, G. Mitotic catastrophe: a mechanism for avoiding genomic instability. Nat. Rev. Mol. Cell Biol. 12, 385–392 (2011).

    Article  CAS  PubMed  Google Scholar 

  14. Wertz, I. E. et al. Sensitivity to antitubulin chemotherapeutics is regulated by MCL1 and FBW7. Nature 471, 110–114 (2011).

    Article  CAS  PubMed  Google Scholar 

  15. Inuzuka, H. et al. SCF(FBW7) regulates cellular apoptosis by targeting MCL1 for ubiquitylation and destruction. Nature 471, 104–109 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Tse, C. et al. ABT-263: a potent and orally bioavailable Bcl-2 family inhibitor. Cancer Res. 68, 3421–3428 (2008).

    Article  CAS  PubMed  Google Scholar 

  17. Wei, M. C. et al. Proapoptotic BAX and BAK: a requisite gateway to mitochondrial dysfunction and death. Science 292, 727–730 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Zeng, X. et al. Pharmacologic inhibition of the anaphase-promoting complex induces a spindle checkpoint-dependent mitotic arrest in the absence of spindle damage. Cancer Cell 18, 382–395 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Eskelinen, E. L. et al. Inhibition of autophagy in mitotic animal cells. Traffic 3, 878–893 (2002).

    Article  CAS  PubMed  Google Scholar 

  20. Kimura, S., Noda, T. & Yoshimori, T. Dissection of the autophagosome maturation process by a novel reporter protein, tandem fluorescent-tagged LC3. Autophagy 3, 452–460 (2007).

    Article  CAS  PubMed  Google Scholar 

  21. Ashrafi, G. & Schwarz, T. L. The pathways of mitophagy for quality control and clearance of mitochondria. Cell Death Differ. 20, 31–42 (2013).

    Article  CAS  PubMed  Google Scholar 

  22. Galluzzi, L., Bravo-San Pedro, J. M. & Kroemer, G. Organelle-specific initiation of cell death. Nat. Cell Biol. 16, 728–736 (2014).

    Article  CAS  PubMed  Google Scholar 

  23. Gomes, L. C., Di Benedetto, G. & Scorrano, L. During autophagy mitochondria elongate, are spared from degradation and sustain cell viability. Nat. Cell Biol. 13, 589–598 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Rambold, A. S., Kostelecky, B., Elia, N. & Lippincott-Schwartz, J. Tubular network formation protects mitochondria from autophagosomal degradation during nutrient starvation. Proc. Natl Acad. Sci. USA 108, 10190–10195 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Gomes, L. C. & Scorrano, L. High levels of Fis1, a pro-fission mitochondrial protein, trigger autophagy. Biochim. Biophys. Acta 1777, 860–866 (2008).

    Article  CAS  PubMed  Google Scholar 

  26. Youle, R. J. & Narendra, D. P. Mechanisms of mitophagy. Nat. Rev. Mol. Cell Biol. 12, 9–14 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Taguchi, N., Ishihara, N., Jofuku, A., Oka, T. & Mihara, K. Mitotic phosphorylation of dynamin-related GTPase Drp1 participates in mitochondrial fission. J. Biol. Chem. 282, 11521–11529 (2007).

    Article  CAS  PubMed  Google Scholar 

  28. Wang, Z. et al. Cyclin B1/Cdk1 coordinates mitochondrial respiration for cell-cycle G2/M progression. Dev. Cell 29, 217–232 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  29. Tsou, P., Zheng, B., Hsu, C. H., Sasaki, A. T. & Cantley, L. C. A fluorescent reporter of AMPK activity and cellular energy stress. Cell Metab. 13, 476–486 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  31. Bando, H. et al. Phosphorylation of the 6-phosphofructo-2-kinase/fructose 2,6-bisphosphatase/PFKFB3 family of glycolytic regulators in human cancer. Clin. Cancer Res. 11, 5784–5792 (2005).

    Article  CAS  PubMed  Google Scholar 

  32. Novellasdemunt, L. et al. PFKFB3 activation in cancer cells by the p38/MK2 pathway in response to stress stimuli. Biochem. J. 452, 531–543 (2013).

    Article  CAS  PubMed  Google Scholar 

  33. Pyronnet, S., Pradayrol, L. & Sonenberg, N. A cell cycle-dependent internal ribosome entry site. Mol. Cell 5, 607–616 (2000).

    Article  CAS  PubMed  Google Scholar 

  34. Novoa, I., Gallego, J., Ferreira, P. G. & Mendez, R. Mitotic cell-cycle progression is regulated by CPEB1 and CPEB4-dependent translational control. Nat. Cell Biol. 12, 447–456 (2010).

    Article  CAS  PubMed  Google Scholar 

  35. Pique, M., Lopez, J. M., Foissac, S., Guigo, R. & Mendez, R. A combinatorial code for CPE-mediated translational control. Cell 132, 434–448 (2008).

    Article  CAS  PubMed  Google Scholar 

  36. Rose, K. M., Bell, L. E. & Jacob, S. T. Specific inhibition of chromatin-associated poly(A) synthesis in vitro by cordycepin 5′-triphosphate. Nature 267, 178–180 (1977).

    Article  CAS  PubMed  Google Scholar 

  37. Lieber, J. et al. The BH3 mimetic ABT-737 increases treatment efficiency of paclitaxel against hepatoblastoma. BMC Cancer 11, 362 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Campone, M. et al. Safety and pharmacokinetics of paclitaxel and the oral mTOR inhibitor everolimus in advanced solid tumours. British J. Cancer 100, 315–321 (2009).

    Article  CAS  Google Scholar 

  39. Meier, F. et al. Significant response after treatment with the mTOR inhibitor sirolimus in combination with carboplatin and paclitaxel in metastatic melanoma patients. J. Am. Acad. Dermatol. 60, 863–868 (2009).

    Article  PubMed  Google Scholar 

  40. Brito, D. A. & Rieder, C. L. Mitotic checkpoint slippage in humans occurs via cyclin B destruction in the presence of an active checkpoint. Curr. Biol. 16, 1194–1200 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Manchado, E., Guillamot, M. & Malumbres, M. Killing cells by targeting mitosis. Cell Death Differ. 19, 369–377 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Domenech, E. & Malumbres, M. Mitosis-targeting therapies: a troubleshooting guide. Curr. Opin. Pharmacol. 13, 519–528 (2013).

    Article  CAS  PubMed  Google Scholar 

  43. Matthess, Y., Raab, M., Sanhaji, M., Lavrik, I. N. & Strebhardt, K. Cdk1/cyclin B1 controls Fas-mediated apoptosis by regulating caspase-8 activity. Mol. Cell Biol. 30, 5726–5740 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Wang, P. et al. Phosphorylation of the proapoptotic BH3-only protein bid primes mitochondria for apoptosis during mitotic arrest. Cell Rep. 7, 661–671 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Wan, L. et al. APC(Cdc20) suppresses apoptosis through targeting bim for ubiquitination and destruction. Dev. Cell 29, 377–391 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Liu, L., Xie, R., Nguyen, S., Ye, M. & McKeehan, W. L. Robust autophagy/mitophagy persists during mitosis. Cell Cycle 8, 1616–1620 (2009).

    Article  CAS  PubMed  Google Scholar 

  47. Okamoto, K. Organellophagy: eliminating cellular building blocks via selective autophagy. J. Cell Biol. 205, 435–445 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Tasdemir, E. et al. Cell cycle-dependent induction of autophagy, mitophagy and reticulophagy. Cell Cycle 6, 2263–2267 (2007).

    Article  CAS  PubMed  Google Scholar 

  49. Tasdemir, E. et al. p53 represses autophagy in a cell cycle-dependent fashion. Cell Cycle 7, 3006–3011 (2008).

    Article  CAS  PubMed  Google Scholar 

  50. Diaz-Martinez, L. A. et al. Genome-wide siRNA screen reveals coupling between mitotic apoptosis and adaptation. EMBO J. 33, 1960–1976 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Hardie, D. G., Ross, F. A. & Hawley, S. A. AMP-activated protein kinase: a target for drugs both ancient and modern. Chem. Biol. 19, 1222–1236 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Hardie, D. G., Ross, F. A. & Hawley, S. A. AMPK: a nutrient and energy sensor that maintains energy homeostasis. Nat. Rev. Mol. Cell Biol. 13, 251–262 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Kim, J., Kundu, M., Viollet, B. & Guan, K. L. AMPK and mTOR regulate autophagy through direct phosphorylation of Ulk1. Nat. Cell Biol. 13, 132–141 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Banko, M. R. et al. Chemical genetic screen for AMPKα2 substrates uncovers a network of proteins involved in mitosis. Mol. Cell 44, 878–892 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Davila, D. et al. Two-step activation of FOXO3 by AMPK generates a coherent feed-forward loop determining excitotoxic cell fate. Cell Death Differ. 19, 1677–1688 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Concannon, C. G. et al. AMP kinase-mediated activation of the BH3-only protein Bim couples energy depletion to stress-induced apoptosis. J. Cell Biol. 189, 83–94 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Kilbride, S. M. et al. AMP-activated protein kinase mediates apoptosis in response to bioenergetic stress through activation of the pro-apoptotic Bcl-2 homology domain-3-only protein BMF. J. Biol. Chem. 285, 36199–36206 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Marsin, A. S. et al. Phosphorylation and activation of heart PFK-2 by AMPK has a role in the stimulation of glycolysis during ischaemia. Curr. Biol. 10, 1247–1255 (2000).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Herrero-Mendez, A. et al. The bioenergetic and antioxidant status of neurons is controlled by continuous degradation of a key glycolytic enzyme by APC/C-Cdh1. Nat. Cell Biol. 11, 747–752 (2009).

    Article  CAS  PubMed  Google Scholar 

  61. Oltersdorf, T. et al. An inhibitor of Bcl-2 family proteins induces regression of solid tumours. Nature 435, 677–681 (2005).

    Article  CAS  PubMed  Google Scholar 

  62. Salazar, M. et al. Cannabinoid action induces autophagy-mediated cell death through stimulation of ER stress in human glioma cells. J. Clin. Invest. 119, 1359–1372 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Tsukamoto, S. et al. Autophagy is essential for preimplantation development of mouse embryos. Science 321, 117–120 (2008).

    Article  CAS  PubMed  Google Scholar 

  64. Buytaert, E., Callewaert, G., Vandenheede, J. R. & Agostinis, P. Deficiency in apoptotic effectors Bax and Bak reveals an autophagic cell death pathway initiated by photodamage to the endoplasmic reticulum. Autophagy 2, 238–240 (2006).

    Article  CAS  PubMed  Google Scholar 

  65. Ortiz-Zapater, E. et al. Key contribution of CPEB4-mediated translational control to cancer progression. Nat. Med. 18, 83–90 (2012).

    Article  CAS  Google Scholar 

  66. Esteban-Martinez, L. & Boya, P. Autophagic flux determination in vivo and ex vivo. Methods 75C, 79–86 (2015).

    Article  Google Scholar 

  67. Mauro-Lizano, M. et al. New method to assess mitophagy flux by flow cytometry. Autophagy 11, 833–843 (2015).

    Article  Google Scholar 

  68. Saha, A. K. et al. Pioglitazone treatment activates AMP-activated protein kinase in rat liver and adipose tissue in vivo. Biochem. Biophys. Res. Commun. 314, 580–585 (2004).

    Article  CAS  PubMed  Google Scholar 

  69. Maizels, E. Z., Ruderman, N. B., Goodman, M. N. & Lau, D. Effect of acetoacetate on glucose metabolism in the soleus and extensor digitorum longus muscles of the rat. Biochem. J. 162, 557–568 (1977).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Bradley, S. A. et al. Fermentanomics: monitoring mammalian cell cultures with NMR spectroscopy. J. Am. Chem. Soc. 132, 9531–9533 (2010).

    Article  CAS  PubMed  Google Scholar 

  71. Van Schaftingen, E., Lederer, B., Bartrons, R. & Hers, H. G. A kinetic study of pyrophosphate: fructose-6-phosphate phosphotransferase from potato tubers. Application to a microassay of fructose 2,6-bisphosphate. Eur. J. Biochem. 129, 191–195 (1982).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank M. Soengas (Spanish National Cancer Research Centre, Spain), J. P. Bolaños (Institute of Functional Biology and Genomics, Spain) and N. Mizushima (The University of Tokyo, Japan) for reagents. We also thank S. Velasco for help with the Seahorse apparatus, J. Muñoz and M. P. Ximénez for their support in phospho-proteomics experiments, and F. Escobar and B. Pou for help with electron microscopy. E.D., C.M. and M.S.-R. were supported by the Spanish Fondo de Investigaciones Sanitarias (Madrid), MINECO (Juan de la Cierva programme) and Asociación Española contra el Cáncer (AECC), respectively. L.E.-M. is a recipient of a JAE predoctoral fellowship from the CSIC. A.K.S. was supported by USPHS grants RO1DK19514, RO1DK67509. G.V. was supported by grants from the Spanish Ministry of Economy and Competitiveness (MINECO) and Fondo Europeo de Desarrollo Regional (FEDER) (PI12/02248), Fundació La Marató de TV3 (m12 20134031), and Fundación Mutua Madrileña (AP101042012). M.L. was supported by the European Community’s Seventh Framework Programme under grant agreement no. 281854—the ObERStress (European Research Council project). E.R. was financially supported by a MINECO grant (SAF 2010-20256). Work in the R.M. laboratory was supported by the Fundación Botín, Banco Santander and MINECO (BFU2011-30121, BFU2014-52125-REDT and Consolider RNAREG CSD2009-00080). Work in the P.B. laboratory is supported by a grant from the Spanish Ministry for Economy and Competitiveness (MINECO; SAF2012-36079). Work in the M.M. laboratory was supported by grants from the MINECO (SAF2012-38215), Consolider-Ingenio 2010 Programme (SAF2014-57791-REDC), Excellence Network CellSYS (BFU2014-52125-REDT), the OncoCycle Programme (S2010/BMD-2470) from the Comunidad de Madrid, Worldwide Cancer Research (WCR no. 15-0278), and the European Union Seventh Framework Programme (MitoSys project; HEALTH-F5-2010-241548).

Author information

Authors and Affiliations

Authors

Contributions

E.D. and M.S.-R. performed most of the cellular and biochemical assays with the help of C.M. and D.P. L.E.-M., E.S. and P.B. collaborated in the analysis of mitophagy, and G.F.-M., R.P. and R.M. collaborated in the analysis of translation. R.C.-O. generated the NMR data. M.P. and D.M. helped with microscopy analysis. A.K.S. performed the AMP/ATP measurements and M.L. contributed to the analysis of AMPK. E.R. helped with metabolic measurements. E.D., C.M., G.V., E.R., P.B., M.S.-R. and M.M. analysed the data. M.M. conceived the project, M.S.R. and M.M., supervised the experiments, and M.M. wrote the manuscript with the help of all co-authors.

Corresponding authors

Correspondence to María Salazar-Roa or Marcos Malumbres.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 4 Analysis of apoptosis, necrosis and necroptosis.

(a) Depletion of Cdc20 after treatment of Cdc20(lox/lox); CreERT2 cells with 4-hydroxytamoxifen (4-OHT) (left top panel). Immunodetection of the indicated proteins after treatment with the indicated siRNAs (rest of panels). Vinculin or α-tubulin were used as a loading control. (b) Efficiency of the indicated siRNAs by quantification of mRNA levels 48 h after siRNA nucleofection. Data were normalized against the levels of β-actin transcripts. Data are mean ± SD and represent one out of at three independent experiments. (c) Immunodetection of caspase 3 cleavage in Cdc20-null cells treated with either ZVAD, Okadaic Acid (OA) or both, at the time points indicated. α-tubulin was used as loading control. (d) Representative electron microscopy images of Bak/Bax-null and wild-type cells arrested in mitosis by the nocodazole-MG132 treatment. Scale bars, 2 μm. (e) Percentage of cells positive for Annexin V or DAPI in the indicated cultures. The left panel represents Cdc20-null or control cells expressing CreERT2 in the presence of 4-hydroxytamoxifen (4OHT). Right panel represents human cells arrested in mitosis with Taxol and proTAME (Tx + pT). Values are represented as mean ± SD (n = 3 independent experiments). For a,c unprocessed original scans of blots are shown in Supplementary Fig. 7. Source data can be found in Supplementary Table 4.

Supplementary Figure 5 Induction of autophagy during mitotic arrest.

(a) Mitotic entry in Cdc20-null cells after release from RO-3306. Cdc20(lox/lox) cells were treated with 4-OHT to eliminate Cdc20 12 h before the addition of the Cdk1 inhibitor RO-3306 to trigger G2 arrest. This compound was washed-out 18 h later allowing the accumulation of mitotic cells over time (left panel). Representative images are shown in the right panel. Scale bars, 50 μm. (b) LC3 lipidation was analyzed in MDA-MB 231 cells arrested by treatment with taxol + pro-TAME, in the presence or absence of E64D/PA. Two different time exposures for LC3 are shown. (c) Colocalization of LC3 and Annexin V (mean ± SD; n = 3 independent experiments) in Cdc20-null cells at the time points of mitotic arrest indicated in the graph. Images are representative for the double staining at 24 h. Scale bar, 10 μm. (d) mRNA levels (mean ± SD; n = 3 independent experiments) of the indicated transcripts 48 h after siRNA nucleofection. Data were normalized against the levels of β-actin transcripts. (e) Duration of mitosis (from metaphase until cell death) in Cdc20-null cells treated with ZVAD, the indicated inducers of autophagy (mTOR inhibitors PP242, temsirolimus, BEZ235), or transfected with siRNAs against Raptor (Rap.). (f) Duration of mitosis in Hela and MDA-MB 231 cells synchronized by the taxol + pro-TAME protocol and treated with the indicated inhibitors. (g) Duration of mitosis in Cdc20-null cells treated with the indicated PI3K/Akt inhibitors. (h) Immunoblot of LC3 after mTOR inhibition with PP242 or BEZ235 in Cdc20-null cells. Cdc27 band-shift (indicating phosphorylation) was used to confirm mitotic arrest. The samples were processed simultaneously and run in the same blot. (i) Immunoblot of LC3 after mTOR inhibition with PP242 in MDA-MB-231 cells synchronized by the taxol + pro-TAME protocol. (j) Duration of mitosis in Atg5-null cells treated or not with the caspase inhibitor ZVAD (left) and Bax/Bak double mutant cells treated or not with the autophagy inhibitor 3MA (right), in both cases using the taxol + pro-TAME protocol. (k) LC3 lipidation and PARP cleavage after treatment with caspase (ZVAD) and autophagy (E64d/PA) inhibitors in Cdc20-null cells. Cdc27 phosphorylation (retarded motility) indicates mitotic cells. In b,h,i,k, α-tubulin is included as a loading control. In eg; j dots represent individual cells and the mean are indicated by red lines. The number of cells analyzed (n) is indicated in each condition, and data are representative of three (e), two (f), six (g) or one (j) separate experiments. Green or red backgrounds indicate a significant delay or premature cell death in mitosis, respectively. , P < 0.05; , P < 0.01; , P < 0.001; Student’s t-test. For b,h,i,k Unprocessed original scans of blots are shown in Supplementary Fig. 7, and α-tubulin is included as a loading control. Source data can be found in Supplementary Table 4.

Supplementary Figure 6 Autophagosomes and ROS accumulate during mitotic arrest.

(a) Representative profiles of mitotracker (left) and Dihydroethidium (DHE; right) in Cdc20-null (green) or control (grey) cells expressing CreERT, at the indicated time points after the release from RO-3306 (all the cells were previously treated 18 h with 4OHT). (b) Immunodetection of the indicated mitochondrial proteins at 24 h after the release from RO-3306 (all the cells were previously treated 18 h with 4OHT) in control-CreERT2 (left) and Cdc20-null cells treated with vehicle or the autophagy inhibitor 3MA. Unprocessed original scans of blots are shown in Supplementary Fig. 7. β-actin was used as a loading control. (c) Effect of 3MA and Cyclosporin A (CsA) on mitotracker levels in control-CreERT2 (left panels) and Cdc20-null cells 24 h after the release from RO-3306 (all the cells were previously treated 18 h with vehicle or 4OHT). Representative FACS profiles are shown and quantification of mitotracker signal is represented in the bottom graphs. Data are mean ± SEM (n = 3 independent experiments). (d) Quantification of ROS (mean ± SEM; n = 3 independent experiments) in Cdc20-null cells as determined by the OxiSelect TM Intracellular ROS assay kit. The effect of NAC treatment in the accumulation of ROS (mean ± SEM; n = 3 independent experiments) is also shown at the indicated times. (e) Quantification of ROS (mean ± SEM; n = 3 independent experiments) in Cdc20-null cells or Cdc20(lox/lox) cells arrested in mitosis after 24 h with taxol and pro-TAME, as determined by the OxiSelect TM Intracellular ROS assay kit. (f) Duration of mitosis until cell death in Cdc20-null and HeLa cells treated with the antioxidant NAC or the caspase inhibitor ZVAD after synchronization with nocodazole and release in MG132. The mean is indicated by red bars and data represent two separate experiments. In d,e,f the number of cells analyzed (n) is indicated in each condition. , P < 0.05; , P < 0.01; , P < 0.001 (Student’s t-test). Source data can be found in Supplementary Table 4.

Supplementary Figure 7 AMPK is reactivated during mitotic arrest.

(a) Schematic representation of the protocol followed for synchronization in human cells. Cells were treated with RO-3306 and this compound was washed-out 18 h later allowing mitotic entry in the presence of taxol and pro-TAME (an APC/C inhibitor) which mimic the arrest in metaphase imposed by Cdc20 ablation. (b) The levels of the indicated antigens were analyzed by immunoblot in HeLa or MDA-MB-231 cells. α-tubulin was included as a loading control and unprocessed original scans of blots are shown in Supplementary Fig. 7. (c) Changes in AMPK activity were monitored by using a FRET biosensor in Cdc20(lox/lox) cells in the absence of 4-OHT. Upon treatment of these control cells with 2-DG, a glucose analog, AMPK is activated giving an increase of about 2-fold in biosensor signal (indicated in red in the images after 2DG addition; bottom panel). This increase is not observed in cells treated with siRNAs against AMPK. Data are mean ± SD (n = 4 independent experiments). (d) AMPK activity was also monitored by the FRET biosensor in mitotic cells treated with vehicle (blue), 2DG (green) or Glucose (red). Data are represented as mean ± SD (n = 3 independent experiments) after scoring the indicated number of mitotic cells in each condition. (e) mRNA levels of APMKα1 and AMPKα2 48 h after nucleofection of specific siRNAs. Data are mean ± SD (n = 3 independent experiments) normalized against β-actin mRNA levels. (f) Levels of OCR (left) and ECAR (right) in Cdc20-null cells arrested in mitosis for 24 h and transfected with siRNAs against siAMPKα1 (siα1), siAMPKα2 (siα2) or the combination of both relative to Scrambled (Scr.) sequences. Data are mean ± SD (one experiment with 12 replicates). (g) Effect of siRNAs against AMPKα1 and/or AMPKα2 on mitotracker levels in Cdc20-null cells 24 h after the release from RO-3306. Representative FACS profiles are shown on the left, and the quantification of mitotracker signal mean ± SEM (n = 3 independent experiments) is represented in the histogram. In g, NS, no significant; , P < 0.05 (Student’s t-test). Source data for all the figures can be found in Supplementary Table 4.

Supplementary Figure 8 Metabolic profile of mitotic arrested cells.

(a) Protocol for synchronized mitotic entry. Cdc20(lox/lox) cells were untreated or treated with 4-OHT to eliminate Cdc20 expression, 12 h prior the addition of the Cdk1 inhibitor RO-3306 to trigger G2 arrest. This compound was washed-out 18 h later, allowing cells to progress through the cell cycle (Cdc20(lox/lox); blue) or arrest in mitosis in the absence of Cdc20 (Cdc20(Δ/Δ); red). Green lines indicate asynchronous cultures and these colour codes are maintained throughout the figure to indicate the three different cell cultures. (b) Similar protocol followed for synchronization in human MDA-MB-231 cells. (c) Cell viability (TO-PRO3 staining) assessed by flow cytometry in the different cultures represented in a. (d) Differential extracellular concentration of the indicated metabolites after 24 h of adding new medium, as monitored by NMR. 34 metabolites were analyzed and only those exhibiting significant differences are represented (23 in this figure + glucose and lactate). No differences were found in carnosine, adenine, glyoxylate, cholesterol, EtOH, niacinamide, acetate, asparagine and aspartic acid. (e) Cell viability (TO-PRO3 staining) assessed by flow cytometry in the cultures represented in b. (f) Glucose uptake and concentration of extracellular lactate in asynchronous (green), normally release from G2 arrest (blue) or mitotic arrested (red) MDA-MB-231 cells, in the presence of 13C-labelled glucose. Data in d and f are not normalized for death cells and the differences in mitotic-arrested cells are likely under-estimated. (g) PFKFB1/4 protein levels in asynchronous (As.), G2-arrested or mitotic Cdc20-deficient cells in the presence or absence or cordycepin (Cord.) or cycloheximide (Chx.). GAPDH was used as a loading control. Images are representative of 2 independent experiments. (h) Immunodetection of PFKFB1 and PFKFB3 48 h after transfection. α-tubulin was included as a loading control. For panels (g,h), Unprocessed original scans of blots are shown in Supplementary Fig. 7. (i) Duration of mitosis (from mitotic entry until cell death) in Cdc20-null cells transfected with siRNAs against PFKFB1 or Scrambled sequences. Each dot indicates a single cells and red lines indicate mean (n = 50 cells per condition; one independent experiment). NS, not significant differences. In c,-f) data are mean ± SD (n = 3 independent experiments. Source data can be found in Supplementary Table 4.

Supplementary Figure 9 Cooperation between taxol and inhibitors of mitotic survival pathways.

Cell death was quantified by scoring To-Pro3 levels by high-throughput microscopy after the addition of the indicated compounds in the presence or absence of taxol in MDA-MB-468, EVSA-T or MCF7 breast cancer cells (left panels) or BJ-TERT-2 and RPE non-transformed cells (right panels). Pink columns indicate significant cooperation when compared to both single inhibitors (black dotted comparisons) and taxol (red dotted comparisons). The number of cells analyzed (n) is indicated in each condition. Data are mean ± SEM (n = 6 independent experiments). , P < 0.001; Student’s t-test, and represent one out of 6 independent experiments. Source data can be found in Supplementary Table 4.

Supplementary Table 1 Small-molecule inhibitors and other chemicals used in this work.
Supplementary Table 2 Oligonucleotides used in this work.
Supplementary Table 3 Antibodies used in this work for immunodetection of proteins.

Supplementary information

Supplementary Information

Supplementary Information (PDF 6202 kb)

Supplementary Table 4

Supplementary Information (XLSX 5405 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Doménech, E., Maestre, C., Esteban-Martínez, L. et al. AMPK and PFKFB3 mediate glycolysis and survival in response to mitophagy during mitotic arrest. Nat Cell Biol 17, 1304–1316 (2015). https://doi.org/10.1038/ncb3231

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/ncb3231

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