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 drives neuronal metabolic burden to promote stress susceptibility in male mice

Abstract

Neurons are particularly susceptible to energy fluctuations in response to stress. Mitochondrial fission is highly regulated to generate ATP via oxidative phosphorylation; however, the role of a regulator of mitochondrial fission in neuronal energy metabolism and synaptic efficacy under chronic stress remains elusive. Here, we show that chronic stress promotes mitochondrial fission in the medial prefrontal cortex via activating dynamin-related protein 1 (Drp1), resulting in mitochondrial dysfunction in male mice. Both pharmacological inhibition and genetic reduction of Drp1 ameliorates the deficit of excitatory synaptic transmission and stress-related depressive-like behavior. In addition, enhancing Drp1 fission promotes stress susceptibility, which is alleviated by coenzyme Q10, which potentiates mitochondrial ATP production. Together, our findings unmask the role of Drp1-dependent mitochondrial fission in the deficits of neuronal metabolic burden and depressive-like behavior and provides medication basis for metabolism-related emotional disorders.

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: Chronic stress induces mitochondrial fission in the mPFC of mice.
Fig. 2: Chronic stress increases Drp1-dependent mitochondrial fission in the mPFC of mice.
Fig. 3: Mitochondrial Drp1 recruitment leads to the deficits of spine morphology and synaptic transmission in the mPFC of mice.
Fig. 4: Neuronal mitochondrial fission mediates depressive-like behavior.
Fig. 5: Drp1-dependent mitochondrial fission weakens synaptic transmission by disturbing the coupling of mitochondrial respiratory chain.
Fig. 6: Pharmacological inhibition of mitochondrial fission alleviates the impairment of synaptic transmission and depressive-like behavior.
Fig. 7: Mitochondrial fission regulates synaptic energy adaptation.

Similar content being viewed by others

Data availability

The data that support the finding of this study are provided within this paper and its supplementary information. Source data are provided with this paper.

References

  1. Malhi, G. S. & Mann, J. J. Depression. Lancet 392, 2299–2312 (2018).

    Article  PubMed  Google Scholar 

  2. Arias-de la Torre, J. et al. Prevalence and variability of current depressive disorder in 27 European countries: a population-based study. Lancet Public Health 6, e729–e738 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  3. Disease, G. B. D., Injury, I., Prevalence, C. & Prevalence, C. Global, regional, and national incidence, prevalence, and years lived with disability for 354 diseases and injuries for 195 countries and territories, 1990-2017: a systematic analysis for the Global Burden of Disease Study 2017. Lancet 392, 1789–1858 (2018).

    Article  Google Scholar 

  4. Wallace, D. C. A mitochondrial etiology of neuropsychiatric disorders. JAMA Psychiatry 74, 863–864 (2017).

    Article  PubMed  Google Scholar 

  5. Sobieski, C., Fitzpatrick, M. J. & Mennerick, S. J. Differential presynaptic ATP supply for basal and high-demand transmission. J. Neurosci. 37, 1888–1899 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  6. Li, S. & Sheng, Z. H. Energy matters: presynaptic metabolism and the maintenance of synaptic transmission. Nat. Rev. Neurosci. 23, 4–22 (2022).

    Article  PubMed  CAS  Google Scholar 

  7. Rangaraju, V., Calloway, N. & Ryan, T. A. Activity-driven local ATP synthesis is required for synaptic function. Cell 156, 825–835 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  8. Gardner, A. & Boles, R. G. Beyond the serotonin hypothesis: mitochondria, inflammation and neurodegeneration in major depression and affective spectrum disorders. Prog. Neuropsychopharmacol. Biol. Psychiatry 35, 730–743 (2011).

    Article  PubMed  CAS  Google Scholar 

  9. Morava, E. et al. Depressive behaviour in children diagnosed with a mitochondrial disorder. Mitochondrion 10, 528–533 (2010).

    Article  PubMed  CAS  Google Scholar 

  10. Abdallah, C. G. et al. Glutamate metabolism in major depressive disorder. Am. J. Psychiatry 171, 1320–1327 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  11. Ashrafi, G., de Juan-Sanz, J., Farrell, R. J. & Ryan, T. A. Molecular tuning of the axonal mitochondrial Ca2+ uniporter ensures metabolic flexibility of neurotransmission. Neuron 105, 678–687 (2020).

  12. Hall, C. N., Klein-Flugge, M. C., Howarth, C. & Attwell, D. Oxidative phosphorylation, not glycolysis, powers presynaptic and postsynaptic mechanisms underlying brain information processing. J. Neurosci. 32, 8940–8951 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  13. Rangaraju, V. et al. Pleiotropic mitochondria: the influence of mitochondria on neuronal development and disease. J. Neurosci. 39, 8200–8208 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  14. Twig, G. et al. Fission and selective fusion govern mitochondrial segregation and elimination by autophagy. EMBO J. 27, 433–446 (2008).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

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

    Article  PubMed  CAS  Google Scholar 

  16. Chakrabarti, R. & Higgs, H. N. Revolutionary view of two ways to split a mitochondrion. Nature 593, 346–347 (2021).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  17. Kageyama, Y. et al. Mitochondrial division ensures the survival of postmitotic neurons by suppressing oxidative damage. J. Cell Biol. 197, 535–551 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  18. Haileselassie, B. et al. Mitochondrial dysfunction mediated through dynamin-related protein 1 (Drp1) propagates impairment in blood brain barrier in septic encephalopathy. J. Neuroinflammation 17, 36 (2020).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  19. Pfeiffer, T., Schuster, S. & Bonhoeffer, S. Cooperation and competition in the evolution of ATP-producing pathways. Science 292, 504–507 (2001).

    Article  PubMed  CAS  Google Scholar 

  20. Rangaraju, V., Lauterbach, M. & Schuman, E. M. Spatially stable mitochondrial compartments fuel local translation during plasticity. Cell 176, 73–84 (2019).

    Article  PubMed  CAS  Google Scholar 

  21. Cao, X. et al. Astrocyte-derived ATP modulates depressive-like behaviors. Nat. Med. 19, 773–777 (2013).

    Article  PubMed  CAS  Google Scholar 

  22. Bertholet, A. M. et al. Mitochondrial fusion/fission dynamics in neurodegeneration and neuronal plasticity. Neurobiol. Dis. 90, 3–19 (2016).

    Article  PubMed  CAS  Google Scholar 

  23. Fonseca, T. B., Sanchez-Guerrero, A., Milosevic, I. & Raimundo, N. Mitochondrial fission requires DRP1 but not dynamins. Nature 570, E34–E42 (2019).

    Article  PubMed  CAS  Google Scholar 

  24. Kashatus, J. A. et al. Erk2 phosphorylation of Drp1 promotes mitochondrial fission and MAPK-driven tumor growth. Mol. Cell 57, 537–551 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  25. Martorell-Riera, A. et al. Mfn2 downregulation in excitotoxicity causes mitochondrial dysfunction and delayed neuronal death. EMBO J. 33, 2388–2407 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  26. Scaini, G. et al. Dysregulation of mitochondrial dynamics, mitophagy and apoptosis in major depressive disorder: does inflammation play a role? Mol. Psychiatry 27, 1095–1102 (2022).

    Article  PubMed  CAS  Google Scholar 

  27. Simpson, C. L. et al. NIX initiates mitochondrial fragmentation via DRP1 to drive epidermal differentiation. Cell Rep. 34, 108689 (2021).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  28. Qi, X. et al. A novel Drp1 inhibitor diminishes aberrant mitochondrial fission and neurotoxicity. J. Cell Sci. 126, 789–802 (2013).

    PubMed  PubMed Central  CAS  Google Scholar 

  29. Guo, X. et al. Inhibition of mitochondrial fragmentation diminishes Huntington′s disease-associated neurodegeneration. J. Clin. Invest. 123, 5371–5388 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  30. Duman, R. S. & Aghajanian, G. K. Synaptic dysfunction in depression: potential therapeutic targets. Science 338, 68–72 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  31. Duman, R. S., Shinohara, R., Fogaca, M. V. & Hare, B. Neurobiology of rapid-acting antidepressants: convergent effects on GluA1-synaptic function. Mol. Psychiatry 24, 1816–1832 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  32. Zhou, Z. et al. The C-terminal tails of endogenous GluA1 and GluA2 differentially contribute to hippocampal synaptic plasticity and learning. Nat. Neurosci. 21, 50–62 (2018).

    Article  PubMed  CAS  Google Scholar 

  33. Alle, H., Roth, A. & Geiger, J. R. Energy-efficient action potentials in hippocampal mossy fibers. Science 325, 1405–1408 (2009).

    Article  PubMed  CAS  Google Scholar 

  34. Harris, J. J., Jolivet, R. & Attwell, D. Synaptic energy use and supply. Neuron 75, 762–777 (2012).

    Article  PubMed  CAS  Google Scholar 

  35. Padamsey, Z., Katsanevaki, D., Dupuy, N. & Rochefort, N. L. Neocortex saves energy by reducing coding precision during food scarcity. Neuron 110, 280–296 (2022).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  36. Harris, J. J., Jolivet, R., Engl, E. & Attwell, D. Energy-efficient information transfer by visual pathway synapses. Curr. Biol. 25, 3151–3160 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  37. Cheng, X. T. et al. Programming axonal mitochondrial maintenance and bioenergetics in neurodegeneration and regeneration. Neuron 110, 1899–1923 (2022).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  38. Guo, X. et al. Mitochondrial stress is relayed to the cytosol by an OMA1-DELE1-HRI pathway. Nature 579, 427–432 (2020).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  39. Fessler, E. et al. A pathway coordinated by DELE1 relays mitochondrial stress to the cytosol. Nature 579, 433–437 (2020).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  40. Xie, T. R., Liu, C. F. & Kang, J. S. Sympathetic transmitters control thermogenic efficacy of brown adipocytes by modulating mitochondrial complex V. Signal Transduct. Target. Ther. 2, 17060 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  41. Choy, J. M. C., Agahari, F. A., Li, L. & Stricker, C. Noradrenaline increases mEPSC frequency in pyramidal cells in layer II of rat barrel cortex via calcium release from presynaptic stores. Front. Cell Neurosci. 12, 213 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  42. Bak, L. K., Schousboe, A., Sonnewald, U. & Waagepetersen, H. S. Glucose is necessary to maintain neurotransmitter homeostasis during synaptic activity in cultured glutamatergic neurons. J. Cereb. Blood Flow. Metab. 26, 1285–1297 (2006).

    Article  PubMed  CAS  Google Scholar 

  43. Chandra, R. et al. Drp1 Mitochondrial Fission in D1 Neurons Mediates Behavioral and Cellular Plasticity during Early Cocaine Abstinence. Neuron 96, 1327–1341 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  44. Ernst, J. et al. Increased pregenual anterior cingulate glucose and lactate concentrations in major depressive disorder. Mol. Psychiatry 22, 113–119 (2017).

    Article  PubMed  CAS  Google Scholar 

  45. de Kloet, E. R., Joels, M. & Holsboer, F. Stress and the brain: from adaptation to disease. Nat. Rev. Neurosci. 6, 463–475 (2005).

    Article  PubMed  Google Scholar 

  46. Wang, C. et al. Targeting PDK2 rescues stress-induced impaired brain energy metabolism. Mol. Psychiatry https://doi.org/10.1038/s41380-023-02098-9 (2023).

  47. Li, X. et al. Astrocytic ApoE reprograms neuronal cholesterol metabolism and histone-acetylation-mediated memory. Neuron 109, 957–970 (2021).

    Article  PubMed  CAS  Google Scholar 

  48. Lu, C. L. et al. Glucocorticoid receptor-dependent astrocytes mediate stress vulnerability. Biol. Psychiatry 92, 204–215 (2022).

    Article  PubMed  CAS  Google Scholar 

  49. Li, M. X. et al. Gene deficiency and pharmacological inhibition of caspase-1 confers resilience to chronic social defeat stress via regulating the stability of surface AMPARs. Mol. Psychiatry 23, 556–568 (2018).

    Article  PubMed  CAS  Google Scholar 

  50. Kim, J., Lei, Y., Lu, X. Y. & Kim, C. S. Glucocorticoid-glucocorticoid receptor-HCN1 channels reduce neuronal excitability in dorsal hippocampal CA1 neurons. Mol. Psychiatry 27, 4035–4049 (2022).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  51. Shields, L. Y. et al. Dynamin-related protein 1 is required for normal mitochondrial bioenergetic and synaptic function in CA1 hippocampal neurons. Cell Death Dis. 6, e1725 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  52. Gao, Q. et al. PINK1-mediated Drp1(S616) phosphorylation modulates synaptic development and plasticity via promoting mitochondrial fission. Signal Transduct. Target Ther. 7, 103 (2022).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  53. Belanger, M., Allaman, I. & Magistretti, P. J. Brain energy metabolism: focus on astrocyte-neuron metabolic cooperation. Cell Metab. 14, 724–738 (2011).

    Article  PubMed  CAS  Google Scholar 

  54. Zhang, J. M. et al. ATP released by astrocytes mediates glutamatergic activity-dependent heterosynaptic suppression. Neuron 40, 971–982 (2003).

    Article  PubMed  CAS  Google Scholar 

  55. Yang, J. et al. Astrocytes contribute to synapse elimination via type 2 inositol 1,4,5-trisphosphate receptor-dependent release of ATP. eLife 5, e15043 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  56. Bowser, D. N. & Khakh, B. S. ATP excites interneurons and astrocytes to increase synaptic inhibition in neuronal networks. J. Neurosci. 24, 8606–8620 (2004).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  57. Liddell, J. R. et al. Sustained hydrogen peroxide stress decreases lactate production by cultured astrocytes. J. Neurosci. Res. 87, 2696–2708 (2009).

    Article  PubMed  CAS  Google Scholar 

  58. Bonvento, G. & Bolanos, J. P. Astrocyte-neuron metabolic cooperation shapes brain activity. Cell Metab. 33, 1546–1564 (2021).

    Article  PubMed  CAS  Google Scholar 

  59. Baker, M. J. et al. Stress-induced OMA1 activation and autocatalytic turnover regulate OPA1-dependent mitochondrial dynamics. EMBO J. 33, 578–593 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  60. Manczak, M. & Reddy, P. H. Mitochondrial division inhibitor 1 protects against mutant huntingtin-induced abnormal mitochondrial dynamics and neuronal damage in Huntington′s disease. Hum. Mol. Genet. 24, 7308–7325 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  61. 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  PubMed  PubMed Central  CAS  Google Scholar 

  62. Zhao, R. Z., Jiang, S., Zhang, L. & Yu, Z. B. Mitochondrial electron transport chain, ROS generation and uncoupling (review). Int. J. Mol. Med. 44, 3–15 (2019).

    PubMed  PubMed Central  CAS  Google Scholar 

  63. Zhang, Y., Liu, J., Chen, X. Q. & Oliver Chen, C. Y. Ubiquinol is superior to ubiquinone to enhance coenzyme Q10 status in older men. Food Funct. 9, 5653–5659 (2018).

    Article  PubMed  CAS  Google Scholar 

  64. Sanoobar, M., Dehghan, P., Khalili, M., Azimi, A. & Seifar, F. Coenzyme Q10 as a treatment for fatigue and depression in multiple sclerosis patients: a double blind randomized clinical trial. Nutr. Neurosci. 19, 138–143 (2016).

    Article  PubMed  CAS  Google Scholar 

  65. Jones, E. et al. A threshold of transmembrane potential is required for mitochondrial dynamic balance mediated by DRP1 and OMA1. Cell. Mol. Life Sci. 74, 1347–1363 (2017).

    Article  PubMed  CAS  Google Scholar 

  66. Pantoja-Urban, A. H. et al. Gains and losses: resilience to social defeat stress in adolescent female mice. Biol. Psychiatry https://doi.org/10.1016/j.biopsych.2023.06.014 (2023).

  67. Krishnan, V. et al. Molecular adaptations underlying susceptibility and resistance to social defeat in brain reward regions. Cell 131, 391–404 (2007).

    Article  PubMed  CAS  Google Scholar 

  68. Cui, Q. Q. et al. Hippocampal CD39/ENTPD1 promotes mouse depression-like behavior through hydrolyzing extracellular ATP. EMBO Rep. 21, e47857 (2020).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  69. Acin-Perez, R. et al. A novel approach to measure mitochondrial respiration in frozen biological samples. EMBO J. 39, e104073 (2020).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  70. Ji, W. K. et al. Receptor-mediated Drp1 oligomerization on endoplasmic reticulum. J. Cell Biol. 216, 4123–4139 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

Download references

Acknowledgements

This work was supported by the Foundation for National Key R&D Program of China (STI2030-Major Projects, grant no. 2021ZD0202900 to J.-G.C.), National Natural Science Foundation of China (grant no. 82130110 to J.-G.C. and grant no. U21A20363 to F.W.), National Natural Science Foundation of China (no. 81872849 to L.-H.L., no. 81971279 to F.W. and no. 81973310 to J.-G.C.) and the Innovative Research Groups of National Natural Science Foundation of China (grant no. 81721005 to J.-G.C. and F.W.). The authors are grateful to W.-K. Ji for the plasmid mCherry-mito-7.

Author information

Authors and Affiliations

Authors

Contributions

J.-G.C., F.W. and L.-H.L. designed the experiment and directed the study. W.-T.D. performed the experiments. W.-T.D., L.-H.L. and Q.D. analyzed data and W.-T.D. and L.-H.L. wrote the manuscript. D.L. and J.-L.W. made CSDS models and performed the behavioral tests.

Corresponding authors

Correspondence to Fang Wang or Jian-Guo Chen.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Metabolism thanks Patricia Jensen and the other, anonymous, reviewers for their contribution to the peer review of this work. Primary Handling Editor: Ashley Castellanos-Jankiewicz, in collaboration with the Nature Metabolism team.

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 Chronic stress has no significant effect on mitochondrial function of resilient mice.

a-c, Typical heat plot of social interaction test (a), social interaction ratio (b), time in interaction zone (c), time in corner zone (d) and sucrose preference (e). N = 10 mice for control and N = 12 mice for resilient. f, Pie chart showing the percentage of eligible and ineligible control mice. N = 519 mice for control and N = 893 mice for CSDS. g, h, Spontaneous activity of mice in open-field test (OFT). N = 54 mice for control and N = 73 mice for CSDS. i, j, Pie chart showing the percentage of mice with immobility time more than 90 s in the forced swim test (FST). N = 54 mice for control and N = 73 mice for CSDS. k, l, Pie chart showing the percentage of mice with immobility time more than 90 s in the tail suspension test (TST). N = 54 mice for control and N = 73 mice for CSDS. m, Timelines of experimental procedure, selection of resilient mice after exposure to CSDS, extraction of mitochondria to detect the mitochondrial membrane potential and ATP. n, JC-1 detection of mitochondrial membrane potential in control and resilient mice. N = 6 mice for each group. o, Bioluminescence kit to observe mitochondrial ATP level. N = 5 mice for control and N = 7 mice for resilient. p, Representative electron microscope graphs indicating the mitochondrial morphology and quantitative analysis of the relative length from fission site to the tip of a mitochondrion. N = 6 mice for each group, white line depicts the mitochondrial edges and white arrows represent fission site. Scale bar represents 100 nm. q, Representative electron microscope graphs and quantitative analysis of tiny mitochondria. N = 6 mice for control and N = 6 mice for resilient, white line depicts the mitochondrial edges and white arrows represent tiny mitochondria. Scale bar represents 500 nm. Data are represented as means ± s.e.m. Two-tailed Student’s t-tests for b, e, h, j, l and n-q. Two-way ANOVA followed by Bonferroni’s post hoc tests for c and d.

Source data

Extended Data Fig. 2 Chronic stress increases the level of phosphorylated Drp1 (p-Drp1) in the mPFC.

a, Double immunostaining for Tomm20 (Green) with AAV-mito-7-mCherry (AAV-mito, Red) (from 3 independent experiments). Scale bar represents 5 µm. b, Representative micrographs showing the mitochondrial morphology (from 3 independent experiments). Scale bar represents 5 µm (left) and 2 µm (top right). c, d, Double immunostaining (c) and quantitative analysis (d) for the NeuN (Red), Iba1 (Red) and GFAP (Red) with Drp1 (Green) in the mPFC. N = 7 brain slices from 4 mice. Scale bar represents 20 µm. e-g, The representative protein bands (e) and quantitative analysis of p-Drp1 (f) and Drp1 (g) in the mPFC. N = 23 mice for control, N = 24 mice for SUS and N = 11 mice for resilient. h,i, Mitochondrial membrane potential (h) and ATP level (i) in the nucleus accumbens (NAc). N = 6 mice for each group. j-l, The representative protein bands (j) and quantitative analysis of p-Drp1 (k) and Drp1 (l) in the NAc. N = 7 mice for control, N = 6 mice for SUS and resilient. m,n, Mitochondrial membrane potential (m) and ATP level (n) in the hippocampus. N = 6 mice for each group. o-q, The representative protein bands (o) and quantitative analysis of p-Drp1 (p) and Drp1 (q) in the Hippocampus. N = 7 mice for control, N = 6 mice for SUS and resilient. r,s, Mitochondrial membrane potential (r) and ATP level (s) in the lateral habenula (LHb). N = 6 mice for each group. t-v, The representative protein bands (t) and quantitative analysis of p-Drp1 (u) and Drp1 (v) in LHb. N = 7 mice for control, N = 6 mice for SUS and resilient. w, The glutamate concentration in the mPFC. N = 5 mice for each group. x, y, The representative protein bands (x) and quantitative analysis (y) of mitofusin 2 (Mfn2) in the mPFC. N = 6 mice for control, N = 7 mice for SUS. Data are represented as means ± s.e.m. Two-tailed Student’s t-tests for h, i, m, n, r, s, w and y. One-way ANOVA followed by Bonferroni’s post hoc tests for f, g, k, l, p, q, u and v.

Source data

Extended Data Fig. 3 Binding of Drp1 to Fis1 promotes stress susceptibility.

a, b, mRNA expression of mitochondrial fission protein 1 (Fis1) (a) and mitochondrial fission factor (Mff) (b). N = 8 mice for each group. c, Experimental timelines for P110 administration. d, The representative protein bands and quantitative analysis of mitochondrial Drp1 in the mPFC. N = 6 mice for TAT and N = 7 mice for P110. e,f, Mitochondrial membrane potential (e) and mitochondrial ATP level (f). N = 9 mice for each group. g, Experimental timelines for P110 administration, behavioral study and Co-immunoprecipitation (Co-IP) analysis. h,i, Co-IP analysis showing the interaction between Drp1 and Fis1 in the mPFC. N = 4 mixed samples from 12 mice for each group. j, Spontaneous activity of mice. N = 12 mice for each group. k-m, Typical heat plot (k) of social interaction test, social interaction ratio, time spent in interaction zone and corner zone (l), and sucrose preference (m). N = 12 mice for each group. n, Experiment timeline for P110 pre-administration and behavioral study. o, Spontaneous activity of mice. N = 8 mice for control + TAT and control + P110, N = 9 mice for CSDS + TAT and CSDS + P110. p, q, Typical heat plot (p) of social interaction test, social interaction ratio, time spent in interaction zone and time spent in corner zone (q). N = 8 mice for control + TAT and control + P110, N = 9 mice for CSDS + TAT and CSDS + P110. r, The proportion of social interaction deficient mice. s, Sucrose preference test. N = 8 mice for control + TAT and control + P110, N = 9 mice for CSDS + TAT and CSDS + P110. t, Experiment timeline for behavioral test. u, Spontaneous activity of mice. N = 10 mice for each group. v-x, The depressive-like behavior was evaluated by social interaction test (v, w), and sucrose preference test (x). N = 10 mice for each group. Data are represented as means ± s.e.m. Two-tailed Student’s t-tests for a, b and d-f. Two-way ANOVA followed by Bonferroni’s post hoc tests for j, l, m, o, q, s, u, w, x and Tukey’s post hoc tests for i.

Source data

Extended Data Fig. 4 Drp1 knockdown rescues the inhibition of excitatory synaptic transmission.

a, b, The depressive-like behavior was evaluated by social interaction test (a) and sucrose preference test (b) after LV-Dnm1l-RNAi [LV-Dnm1l (Ri)] infusion. N = 33 mice for control + LV-GFP (GFP), N = 35 mice for susceptible (SUS) + GFP, N = 31 mice for control + Dnm1l (Ri) and N = 33 mice for SUS + Dnm1l (Ri). c-e, Representative traces of AMPAR-mediated miniature excitatory postsynaptic currents (mEPSCs) recordings after Dnm1l (Ri) injection (c). The mEPSCs amplitude scatter diagram (upper panel) and cumulative plots (bottom panel, d), and mEPSCs frequency scatter diagram (upper panel) and cumulative plots (bottom panel, e) for representative cells. N = 9 cells from 5 mice for control + GFP, N = 11 cells from 5 mice for SUS + GFP, N = 9 cells from 5 mice for control + Dnm1l (Ri) and N = 14 cells from 6 mice for SUS + Dnm1l (Ri). f-h, Representative traces of miniature inhibitory postsynaptic currents (mIPSCs) recordings after Dnm1l (Ri) injection (f). The mIPSCs amplitude scatter diagram (upper panel) and cumulative plots (bottom panel, g), and mIPSCs frequency scatter diagram (upper panel) and cumulative plots (bottom panel, h) for representative cells. N = 7 cells from 4 mice for control + GFP, N = 6 cells from 3 mice for SUS + GFP, N = 6 cells from 3 mice for control + Dnm1l (Ri) and N = 11 cells from 5 mice for SUS + Dnm1l (Ri). Data are represented as means ± s.e.m. Two-way ANOVA followed by Bonferroni’s post hoc tests for a, b, d, e, g and h.

Source data

Extended Data Fig. 5 Drp1 knockdown reverses the mRNA expression of OMA1/DELE1/HRI in susceptible mice.

a, b, The depressive-like behavior was evaluated by social interaction test (a) and sucrose preference test (b) after LV-Dnm1l-RNAi [LV-Dnm1l (Ri)] infusion. N = 18 mice for control + LV-GFP (GFP), N = 19 mice for susceptible (SUS) + GFP, N = 21 mice for control + Dnm1l (Ri) and N = 22 mice for SUS + Dnm1l (Ri). c-e, The mRNA level of mitochondrial protease OMA1 (c), DAP3 binding cell death enhancer 1 (DELE1) (d) and eukaryotic translation initiation factor 2α kinase (HRI) (e) after Dnm1l (Ri) injection. N = 6 mice for control + GFP, N = 6 mice for SUS + GFP, N = 7 mice for control + Dnm1l (Ri) and N = 7 mice for SUS + Dnm1l (Ri). f, g, The time in the interaction zone (f) and corner zone (g) during the no target phase of social interaction test of mice. N = 8 mice for control + GFP, N = 10 mice for SUS + GFP, N = 7 mice for control + Dnm1l (Ri) and N = 10 mice for SUS + Dnm1l (Ri). Data are represented as means ± s.e.m. Two-way ANOVA followed by Bonferroni’s post hoc tests for a-g.

Source data

Extended Data Fig. 6 Norepinephrine and oligomycin A differentially decreases mitochondrial ATP.

a, Norepinephrine moderately dissipates mitochondrial ATP in N2A cell. N = 6 well for each group. b, c, the time in interaction zone (b) and corner zone of mice (c) during the no target phase of social interaction test. N = 12 mice for each group. d, Oligomycin A continuously inhibits mitochondrial ATP in N2A cell. N = 6 well for each group. e-g, The time in interaction zone during the target phase of social interaction test (e), and time in interaction zone (f) and corner zone (g) during the no target phase of social interaction test of mice. N = 14 mice for control + LV-GFP (GFP), N = 12 mice for susceptible (SUS) + GFP, N = 13 mice for control + LV-Dnm1l-RNAi [Dnm1l (Ri)] and N = 12 mice for SUS + Dnm1l (Ri). Data are represented as means ± s.e.m. One-way ANOVA followed by Bonferroni’s post hoc tests for a and d. Two-way ANOVA followed by Bonferroni’s post hoc tests for b, c and e-g.

Source data

Extended Data Fig. 7 Mdivi1 reduces ROS release and rescues Drp1 recruitment to mitochondria in susceptible mice.

a, Experimental timelines for glucose administration and behavioral study, susceptible (SUS) mice were selected for implanting cannula into the mPFC, and behavioral test was performed after glucose administration. b-d, Schematic and typical heat plot of mice in social interaction test (b), social interaction ratio, time in interaction zone, time in corner zone (c) and sucrose preference test (d). N = 7 mice for control + vehicle, N = 6 mice for SUS + vehicle, N = 7 mice for control + glucose and N = 6 mice for SUS + glucose. e, Spontaneous activity of mice in open-field test (OFT). N = 7 mice for control + vehicle, N = 6 mice for SUS + vehicle, N = 7 mice for control + glucose and N = 6 mice for SUS + glucose. f, Representative micrographs showing immunostaining for dihydropyridine (DHE) (Red, representing mitochondrial ROS), NeuN (Green) and DAPI (Blue), and the quantification of mitochondrial ROS in control and SUS mice. N = 3 mice each group. Scale bar represents 20 µm. g, Representative micrographs showing the recruitment of Drp1 to mitochondria in the mPFC of SUS + vehicle and SUS + Mdivi1 mice (from 3 independent experiments). Scale bar represents 5 µm. h-j, mRNA expression of mitochondrial protease OMA1 (h), DAP3 binding cell death enhancer 1 (DELE1) (i) and eukaryotic translation initiation factor 2α kinase (HRI) (j). N = 10 mice for each group. Data are represented as means ± s.e.m. Two-tailed Student’s t-tests for f. Two-way ANOVA followed by Bonferroni’s post hoc tests for c-e and h-j.

Source data

Supplementary information

Supplementary Information

Supplementary Tables 1–5 and uncropped scans of all blots.

Reporting Summary

Source data

Source Data Fig. 1

Statistical source data, source cartoon schematic and image source data.

Source Data Fig. 2

Statistical source data, source cartoon schematic and image source data.

Source Data Fig. 3

Statistical source data, cartoon pattern and image source data.

Source Data Fig. 4

Statistical Source Data and image source data.

Source Data Fig. 5

Statistical Source Data and source cartoon schematic.

Source Data Fig. 6

Statistical Source Data and source cartoon schematic.

Source Data Extended Data Fig. 1

Statistical Source Data and image source data.

Source Data Extended Data Fig. 2

Statistical Source Data and image source data.

Source Data Extended Data Fig. 3

Statistical Source Data.

Source Data Extended Data Fig. 4

Statistical Source Data.

Source Data Extended Data Fig. 5

Statistical Source Data.

Source Data Extended Data Fig. 6

Statistical Source Data.

Source Data Extended Data Fig. 7

Statistical Source Data and image source data.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Dong, WT., Long, LH., Deng, Q. et al. Mitochondrial fission drives neuronal metabolic burden to promote stress susceptibility in male mice. Nat Metab 5, 2220–2236 (2023). https://doi.org/10.1038/s42255-023-00924-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s42255-023-00924-6

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