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.

Ischaemic accumulation of succinate controls reperfusion injury through mitochondrial ROS

Subjects

Abstract

Ischaemia-reperfusion injury occurs when the blood supply to an organ is disrupted and then restored, and underlies many disorders, notably heart attack and stroke. While reperfusion of ischaemic tissue is essential for survival, it also initiates oxidative damage, cell death and aberrant immune responses through the generation of mitochondrial reactive oxygen species (ROS)1,2,3,4,5. Although mitochondrial ROS production in ischaemia reperfusion is established, it has generally been considered a nonspecific response to reperfusion1,3. Here we develop a comparative in vivo metabolomic analysis, and unexpectedly identify widely conserved metabolic pathways responsible for mitochondrial ROS production during ischaemia reperfusion. We show that selective accumulation of the citric acid cycle intermediate succinate is a universal metabolic signature of ischaemia in a range of tissues and is responsible for mitochondrial ROS production during reperfusion. Ischaemic succinate accumulation arises from reversal of succinate dehydrogenase, which in turn is driven by fumarate overflow from purine nucleotide breakdown and partial reversal of the malate/aspartate shuttle. After reperfusion, the accumulated succinate is rapidly re-oxidized by succinate dehydrogenase, driving extensive ROS generation by reverse electron transport at mitochondrial complex I. Decreasing ischaemic succinate accumulation by pharmacological inhibition is sufficient to ameliorate in vivo ischaemia-reperfusion injury in murine models of heart attack and stroke. Thus, we have identified a conserved metabolic response of tissues to ischaemia and reperfusion that unifies many hitherto unconnected aspects of ischaemia-reperfusion injury. Furthermore, these findings reveal a new pathway for metabolic control of ROS production in vivo, while demonstrating that inhibition of ischaemic succinate accumulation and its oxidation after subsequent reperfusion is a potential therapeutic target to decrease ischaemia-reperfusion injury in a range of pathologies.

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

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

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

Figure 1: Comparative metabolomics identifies succinate as a potential mitochondrial metabolite that drives reperfusion ROS production.
Figure 2: Reverse SDH activity drives ischaemic succinate accumulation by the reduction of fumarate.
Figure 3: Ischaemic succinate levels control ROS production in adult primary cardiomyocytes and in the heart in vivo.
Figure 4: NADH and AMP sensing pathways drive ischaemic succinate accumulation to control reperfusion pathologies in vivo through mitochondrial ROS production.

References

  1. Murphy, E. & Steenbergen, C. Mechanisms underlying acute protection from cardiac ischemia-reperfusion injury. Physiol. Rev. 88, 581–609 (2008)

    CAS  PubMed  Google Scholar 

  2. Yellon, D. M. & Hausenloy, D. J. Myocardial reperfusion injury. N. Engl. J. Med. 357, 1121–1135 (2007)

    CAS  PubMed  Google Scholar 

  3. Burwell, L. S., Nadtochiy, S. M. & Brookes, P. S. Cardioprotection by metabolic shut-down and gradual wake-up. J. Mol. Cell. Cardiol. 46, 804–810 (2009)

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Eltzschig, H. K. & Eckle, T. Ischemia and reperfusion–from mechanism to translation. Nature Med. 17, 1391–1401 (2011)

    CAS  PubMed  Google Scholar 

  5. Timmers, L. et al. The innate immune response in reperfused myocardium. Cardiovasc. Res. 94, 276–283 (2012)

    CAS  PubMed  Google Scholar 

  6. Harmsen, E., de Jong, J. W. & Serruys, P. W. Hypoxanthine production by ischemic heart demonstrated by high pressure liquid chromatography of blood purine nucleosides and oxypurines. Clin. Chim. Acta 115, 73–84 (1981)

    CAS  PubMed  Google Scholar 

  7. Pacher, P., Nivorozhkin, A. & Szabo, C. Therapeutic effects of xanthine oxidase inhibitors: renaissance half a century after the discovery of allopurinol. Pharmacol. Rev. 58, 87–114 (2006)

    CAS  PubMed  Google Scholar 

  8. Chouchani, E. T. et al. Cardioprotection by S-nitrosation of a cysteine switch on mitochondrial complex I. Nature Med. 19, 753–759 (2013)

    CAS  PubMed  Google Scholar 

  9. Zweier, J. L., Flaherty, J. T. & Weisfeldt, M. L. Direct measurement of free radical generation following reperfusion of ischemic myocardium. Proc. Natl Acad. Sci. USA 84, 1404–1407 (1987)

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  10. Tannahill, G. M. et al. Succinate is an inflammatory signal that induces IL-1beta through HIF-1alpha. Nature 496, 238–242 (2013)

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  11. Smith, A. C. & Robinson, A. J. A metabolic model of the mitochondrion and its use in modelling diseases of the tricarboxylic acid cycle. BMC Syst. Biol. 5, 102 (2011)

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Niatsetskaya, Z. V. et al. The oxygen free radicals originating from mitochondrial complex I contribute to oxidative brain injury following hypoxia–ischemia in neonatal mice. J. Neurosci. 32, 3235–3244 (2012)

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Taegtmeyer, H. Metabolic responses to cardiac hypoxia. Increased production of succinate by rabbit papillary muscles. Circ. Res. 43, 808–815 (1978)

    CAS  PubMed  Google Scholar 

  14. Hochachka, P. W. & Storey, K. B. Metabolic consequences of diving in animals and man. Science 187, 613–621 (1975)

    ADS  CAS  PubMed  Google Scholar 

  15. Easlon, E., Tsang, F., Skinner, C., Wang, C. & Lin, S. J. The malate-aspartate NADH shuttle components are novel metabolic longevity regulators required for calorie restriction-mediated life span extension in yeast. Genes Dev. 22, 931–944 (2008)

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Barron, J. T., Gu, L. & Parrillo, J. E. Malate-aspartate shuttle, cytoplasmic NADH redox potential, and energetics in vascular smooth muscle. J. Mol. Cell. Cardiol. 30, 1571–1579 (1998)

    CAS  PubMed  Google Scholar 

  17. Van den Berghe, G., Vincent, M. F. & Jaeken, J. Inborn errors of the purine nucleotide cycle: adenylosuccinase deficiency. J. Inherit. Metab. Dis. 20, 193–202 (1997)

    CAS  PubMed  Google Scholar 

  18. Sridharan, V. et al. O2-sensing signal cascade: clamping of O2 respiration, reduced ATP utilization, and inducible fumarate respiration. Am. J. Physiol. 295, C29–C37 (2008)

    CAS  Google Scholar 

  19. Dervartanian, D. V. & Veeger, C. Studies on succinate dehydrogenase. I. spectral properties of the purified enzyme and formation of enzyme-competitive inhibitor complexes. Biochim. Biophys. Acta 92, 233–247 (1964)

    CAS  PubMed  Google Scholar 

  20. Gutman, M. Modulation of mitochondrial succinate dehydrogenase activity, mechanism and function. Mol. Cell. Biochem. 20, 41–60 (1978)

    CAS  PubMed  Google Scholar 

  21. Bünger, R., Glanert, S., Sommer, O. & Gerlach, E. Inhibition by (aminooxy)acetate of the malate-aspartate cycle in the isolated working guinea pig heart. Hoppe-Seyler's Z. Physiol. Chem. 361, 907–914 (1980)

    Google Scholar 

  22. Swain, J. L., Hines, J. J., Sabina, R. L., Harbury, O. L. & Holmes, E. W. Disruption of the purine nucleotide cycle by inhibition of adenylosuccinate lyase produces skeletal muscle dysfunction. J. Clin. Invest. 74, 1422–1427 (1984)

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Hirst, J., King, M. S. & Pryde, K. R. The production of reactive oxygen species by complex I. Biochem. Soc. Trans. 36, 976–980 (2008)

    CAS  PubMed  Google Scholar 

  24. Kussmaul, L. & Hirst, J. The mechanism of superoxide production by NADH:ubiquinone oxidoreductase (complex I) from bovine heart mitochondria. Proc. Natl Acad. Sci. USA 103, 7607–7612 (2006)

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  25. Pryde, K. R. & Hirst, J. Superoxide is produced by the reduced flavin in mitochondrial complex I: a single, unified mechanism that applies during both forward and reverse electron transfer. J. Biol. Chem. 286, 18056–18065 (2011)

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  27. Davidson, S. M., Yellon, D. & Duchen, M. R. Assessing mitochondrial potential, calcium, and redox state in isolated mammalian cells using confocal microscopy. Methods Mol. Biol. 372, 421–430 (2007)

    CAS  PubMed  Google Scholar 

  28. Wojtovich, A. P., Smith, C. O., Haynes, C. M., Nehrke, K. W. & Brookes, P. S. Physiological consequences of complex II inhibition for aging, disease, and the mKATP channel. Biochim. Biophys. Acta 1827, 598–611 (2013)

    CAS  PubMed  Google Scholar 

  29. Brennan, J. P. et al. Mitochondrial uncoupling, with low concentration FCCP, induces ROS-dependent cardioprotection independent of KATP channel activation. Cardiovasc. Res. 72, 313–321 (2006)

    CAS  PubMed  Google Scholar 

  30. Hamel, D. et al. G-Protein-coupled receptor 91 and succinate are key contributors in neonatal postcerebral hypoxia-ischemia recovery. Arterioscler. Thromb. Vasc. Biol. 34, 285–293 (2014)

    CAS  PubMed  Google Scholar 

  31. Schmidt, K. et al. Cardioprotective effects of mineralocorticoid receptor antagonists at reperfusion. Eur. Heart J. 31, 1655–1662 (2010)

    CAS  PubMed  Google Scholar 

  32. Methner, C. et al. Protection through postconditioning or a mitochondria-targeted S-nitrosothiol is unaffected by cardiomyocyte-selective ablation of protein kinase G. Basic Res. Cardiol. 108, 337 (2013)

    PubMed  Google Scholar 

  33. Nadtochiy, S. M., Redman, E., Rahman, I. & Brookes, P. S. Lysine deacetylation in ischaemic preconditioning: the role of SIRT1. Cardiovasc. Res. 89, 643–649 (2011)

    CAS  PubMed  Google Scholar 

  34. Aksentijevic, D. et al. Myocardial creatine levels do not influence response to acute oxidative stress in isolated perfused heart. PLoS One. 9, e109021 (2014)

    ADS  PubMed  PubMed Central  Google Scholar 

  35. Ord, E. N. J. et al. Positive impact of pre-stroke surgery on survival following transient focal ischemia in hypertensive rats. J. Neurosci. Methods 211, 305–308 (2012)

    PubMed  PubMed Central  Google Scholar 

  36. Koizumi, J., Yoshida, Y., Nakazawa, T. & Ooneda, G. Experimental studies of ischemic brain edema: a new experimental model of cerebral embolism in rats in which recirculation can be introduced in the ischemic area. Jpn. J. Stroke 8, 1–8 (1986)

    Google Scholar 

  37. Hunter, A. J. et al. Functional assessments in mice and rats after focal stroke. Neuropharmacology 39, 806–816 (2000)

    CAS  PubMed  Google Scholar 

  38. Ord, E. N. J. et al. Combined antiapoptotic and antioxidant approach to acute neuroprotection for stroke in hypertensive rats. J. Cereb. Blood Flow Metab. 33, 1215–1224 (2013)

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Osborne, K. A. et al. Quantitative assessment of early brain damage in a rat model of focal cerebral ischaemia. J. Neurol. Neurosurg. Psychiatry 50, 402–410 (1987)

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Gaude, E. et al. muma, An R Package for metabolomics univariate and multivariate statistical analysis. Curr. Metabol. 1, 180–189 (2013)

    CAS  Google Scholar 

  41. Goodpaster, A. M., Romick-Rosendale, L. E. & Kennedy, M. A. Statistical significance analysis of nuclear magnetic resonance-based metabonomics data. Anal. Biochem. 401, 134–143 (2010)

    CAS  PubMed  Google Scholar 

  42. Davidson, S. M. & Duchen, M. R. Effects of NO on mitochondrial function in cardiomyocytes: Pathophysiological relevance. Cardiovasc. Res. 71, 10–21 (2006)

    CAS  PubMed  Google Scholar 

  43. Smith, A. C., Blackshaw, J. A. & Robinson, A. J. MitoMiner: a data warehouse for mitochondrial proteomics data. Nucleic Acids Res. 40, D1160–D1167 (2012)

    CAS  PubMed  Google Scholar 

  44. Chang, A., Scheer, M., Grote, A., Schomburg, I. & Schomburg, D. BRENDA, AMENDA and FRENDA the enzyme information system: new content and tools in 2009. Nucleic Acids Res. 37, D588–D592 (2009)

    CAS  PubMed  Google Scholar 

  45. Romero, P. et al. Computational prediction of human metabolic pathways from the complete human genome. Genome Biol. 6, R2 (2004)

    PubMed  PubMed Central  Google Scholar 

  46. Abad, M. F., Di Benedetto, G., Magalhaes, P. J., Filippin, L. & Pozzan, T. Mitochondrial pH monitored by a new engineered green fluorescent protein mutant. J. Biol. Chem. 279, 11521–11529 (2004)

    CAS  PubMed  Google Scholar 

  47. Orth, J. D., Thiele, I. & Palsson, B. O. What is flux balance analysis? Nature Biotechnol. 28, 245–248 (2010)

    CAS  Google Scholar 

  48. Becker, S. A. et al. Quantitative prediction of cellular metabolism with constraint-based models: the COBRA Toolbox. Nature Protocols 2, 727–738 (2007)

    ADS  CAS  PubMed  Google Scholar 

  49. Cochemé, H. M. & Murphy, M. P. Complex I is the major site of mitochondrial superoxide production by paraquat. J. Biol. Chem. 283, 1786–1798 (2008)

    PubMed  Google Scholar 

  50. Srere, P. A. Citrate synthase. Methods Enzymol. 13, 3–11 (1969)

    CAS  Google Scholar 

Download references

Acknowledgements

Supported by the Medical Research Council (UK) and by grants from Canadian Institutes of Health Research and the Gates Cambridge Trust (E.T.C.) and the British Heart Foundation (T.K., V.R.P., L.M.W.). We thank J. Hirst and G. C. Brown for discussions.

Author information

Authors and Affiliations

Authors

Contributions

E.T.C. designed research, carried out biochemical experiments, analysed data from in vivo experiments and co-wrote the paper. T.K., V.R.P. and C.-H.H. designed and carried out the ex vivo and in vivo experiments. C.F. and E.G. designed and carried out mass spectrometry and metabolomics analyses, with A.S.H.C. assisting. D.A. and M.J.S. designed and carried out ex vivo perfused heart experiments. S.Y.S., S.M.D., M.R.D., S.M.N., E.L.R. and P.S.B. designed and carried out cell experiments. L.M.W., E.N.J.O. and R.S. designed and carried out brain experiments. A.J.D., S.R. and K.S.-P. designed and carried out kidney experiments. A.L. and R.C.H. carried out ROS analyses. S.E. carried out analyses. A.M.J. helped with data interpretation. A.C.S., A.J.R. and F.E. designed and performed bioinformatic analyses. E.T.C., T.K., C.F. and M.P.M. directed the research and co-wrote the paper, with assistance from all other authors.

Corresponding authors

Correspondence to Christian Frezza, Thomas Krieg or Michael P. Murphy.

Ethics declarations

Competing interests

M.P.M., E.T.C., L.M.W., C.F. and T.K. have applied for a patent on some of the work described here.

Extended data figures and tables

Extended Data Figure 1 Comparative analysis of metabolites significantly accumulated in ischaemic conditions.

a, Various rat and mouse tissues exposed to sufficient periods of ischaemia to prime for reperfusion ROS production were subjected to targeted LC–MS metabolomics analysis and comparison of metabolites that accumulated significantly when compared to normoxic levels. After this, metabolites were scored according to the prevalence of their accumulation across five ischaemic tissue conditions. B, brain; H, whole heart ischaemia ex vivo; HL, left anterior descending coronary artery ischaemia in vivo; K, kidney, L, liver. b, Determination of linearity of the relationship between LC–MS metabolite peak intensity and concentration for CAC and related metabolites. c, Quality control determination of coefficient of variation for LC–MS quantification of CAC and related metabolites.

Extended Data Figure 2 Time course of succinate levels in the in vivo heart during ischaemia and reperfusion and potential metabolic inputs for succinate.

a, Time course of succinate levels during myocardial ischaemia and reperfusion for the in vivo heart (5 min and 15 min ischaemia n = 4; 30 min ischaemia n = 9; 5 min reperfused n = 5). b, Summary of the three potential metabolic inputs for succinate-directed ischaemic flux. To understand the metabolic pathways that could contribute to succinate production under ischaemia, an updated version of the iAS253 model of cardiac metabolism11 was used to simulate ischaemia using flux balance analysis. The model showed three possible mechanisms for producing succinate: from α-ketoglutarate produced by the CAC, derived from glycolysis, fatty acid oxidation, and glutaminolysis (grey box), from succinic semialdehyde produced from the GABA shunt (blue box), and from fumarate produced from the malate-aspartate shuttle and purine nucleotide cycle (red box) via the reversal of SDH. Data are mean ± s.e.m. of at least four biological replicates.

Extended Data Figure 3 Metabolic labelling of CAC and proximal metabolites by 13C-glucose in the ischaemic and normoxic myocardium.

Proportional isotopic labelling profile of CAC and proximal metabolites during normoxic and ischaemic myocardial perfusion. Mouse hearts were perfused with 11 mM [U-13C]glucose (+6 labelled) for 10 min followed by either 30 min no flow ischaemia or 30 min normoxic perfusion followed by snap-freezing and LC–MS metabolomic analysis (n = 4). *P < 0.05, **P < 0.01, ***P < 0.001 (two-tailed Student’s t-test). Data are mean ± s.e.m. of at least four biological replicates.

Extended Data Figure 4 Metabolic labelling of CAC and proximal metabolites by 13C-palmitate in the ischaemic and normoxic myocardium.

a, Mouse hearts were perfused with 0.3 mM [U-13C]palmitate (+16 labelled) for 10 min resulting in a significant proportion of the endogenous palmitate pool being +16 labelled. Following this, hearts were subjected to either 30 min ischaemia or continued normoxic respiration with 13C-palmitate followed by snap-freezing and metabolomic analysis. b, Isotopic flux from palmitate to CAC and proximal metabolites following normoxic and ischaemic myocardial respiration. The isotopic profile for each metabolite is expressed as a proportion of the total pool (n = 4). *P < 0.05, **P < 0.01, ***P < 0.001 (two-tailed Student’s t-test). Data are shown as the mean ± s.e.m. of at least four biological replicates.

Extended Data Figure 5 Metabolic labelling of CAC and proximal metabolites by 13C-glutamine in the ischaemic and normoxic myocardium, and measurement of the effect of inhibition of GABA transaminase on succinate accumulation in the ischaemic myocardium.

a, Mouse hearts were perfused with 4 mM [U-13C]glutamine (+5 labelled) for 10 min followed by either 30 min no flow ischaemia or 30 min normoxic respiration followed by snap freezing and metabolomic analysis. The isotopic profile for each metabolite is expressed as a proportion of the total pool (n = 4). b, Furthermore, flux to α-ketoglutarate was determined relative to the proportion of the +5 glutamine pool in the heart (n = 4). c, d, Perfused mouse hearts were subjected to 30 min no flow ischaemia ± continuous infusion of vigabatrin (vig; 100, 300 and 700 μM) 10 min before ischaemia. Heart tissue was snap frozen and GABA (c) and succinate (d) abundance quantified relative to normoxic levels by LC–MS (n = 4; ischaemia n = 5). *P < 0.05 (two-tailed Student’s t-test). Data are mean ± s.e.m. of at least four biological replicates.

Extended Data Figure 6 Unabridged metabolic model identifying pathways that can become activated by tissue ischaemia to drive succinate accumulation.

To identify the metabolic pathways that could contribute to succinate production under ischaemia, we simulated these conditions using flux balance analysis in conjunction with an expanded version of the iAS253 mitochondrial model of central cardiac metabolism. The major pathways contributing to succinate accumulation (bold red lines) were via fumarate feeding into the reverse activity of SDH. This was produced by the PNC and the MAS, which consumed glucose and aspartate, and also led to significant production of lactate and alanine. Lesser sources of succinate (thin red lines) included glycolysis and glutaminolysis but this was relatively minor as this route was constrained by the overproduction of NADH. In addition, a small amount of fumarate was generated by pyruvate carboxylase activity. The GABA shunt did not contribute (black dashed line).

Extended Data Figure 7 Effects of dimethyl malonate and dimethyl succinate treatment of cells and in vivo on intracellular accumulation of malonate and succinate, and respiration and comparison of 13C-labelled ischaemic metabolite fluxes to succinate relative to isotopic donor pools.

a, Intravenous infusion of dimethyl malonate in vivo results in accumulation of malonate in the ischaemic myocardium (n = 4). b, c, C2C12 cells were incubated with: no additions, glucose, 5 mM dimethyl succinate, 5 mM dimethyl malonate, or 5 mM dimethyl malonate and 5 mM dimethyl succinate. Cellular oxygen consumption rate due to ATP synthesis (b) and maximal rates (c) in the presence of p-triflouromethoxyphenylhydrazone (FCCP) were determined using a Seahorse XF96 analyser (n = 4). d, Mouse hearts were perfused with 13C-glucose (+6 labelled), 13C-glutamine (+5 labelled), 13C-aspartate (+1 labelled), or 13C-palmitate (+16 labelled) for 10 min followed by 30 min no-flow ischaemia or 30 min normoxic respiration, followed by snap-freezing and metabolomic analysis. To compare the relative magnitude of metabolite flux from each carbon source, 13C incorporation to succinate during normoxia and ischaemia was determined relative to the proportion of the total pool of the relevant infused 13C donor. 13C incorporation into succinate was considered in terms of the proportion of the +4 isotope in the entire succinate pool for 13C-glucose, 13C-glutamine and 13C-palmitate infusions; and the proportion of the +1 isotope in the entire pool for the 13C-aspartate infusion (n = 4). *P < 0.05, ***P < 0.001 (two-tailed Student’s t-test for pairwise comparisons, and one-way ANOVA followed by Bonferroni’s test for multiple comparisons). Data are mean ± s.e.m. of at least four biological replicates.

Extended Data Figure 8 Predicted changes in pathways of succinate and oxidative phosphorylation metabolism during ischaemia and following reperfusion.

To determine possible changes in succinate metabolism during ischaemia, reperfusion and normoxia, cardiac metabolism was simulated in these conditions using an expanded version of the iAS253 model with flux balance analysis. a, The simulations predicted that under ischaemia, SDH ran in reverse by using ubiquinol produced by complex I to reduce fumarate to succinate, thereby acting as a terminal electron acceptor instead of oxygen. Fumarate was produced from the PNC and reversal of the CAC. Flux through the rest of the respiratory chain was diminished and AMP was produced from ADP owing to insufficient ATP production. b, With oxygen restored SDH metabolised excess succinate. A delay in regenerating AMP to ADP, as typified in the first minute of reperfusion, limited the flux through ATP-synthase. This in turn prevented complex III consuming all the ubiquinol generated by SDH, as the membrane became hyperpolarized. The excess flux of ubiquinol and protons forced complex I to run in reverse, which would generate ROS by RET. c, Once the flux of succinate was reduced to normal levels, as in the transition from late reperfusion to normoxia, the fluxes through the respiratory chain and citric acid cycle returned to normal.

Extended Data Figure 9 Tracking DHE oxidation, NAD(P)H reduction state, and mitochondrial membrane potential in primary cardiomyocytes during in situ IR.

a, Inhibition of mitochondrial complex I RET reduces DHE oxidation on reperfusion (n = 6; rotenone n = 4). b, c, Effect of manipulation of ischaemic succinate levels on NAD(P)H oxidation during early reperfusion (n = 3). Primary rat cardiomyocytes were subjected to 40 min ischaemia followed by reoxygenation and NAD(P)H reduction state was tracked throughout the experiment by measurement of NAD(P)H autofluorescence. Ischaemic buffer contained no additions, 4 mM dimethyl malonate, or 4 mM dimethyl succinate. Average (b) and representative (c) traces from each condition are shown. The highlighted window in c indicates the period of the experiment expanded in detail in b. d, Effect of inhibition of ischaemic succinate accumulation on mitochondrial membrane potential following late ischaemia (left) and early reperfusion (right). e, f, Primary rat cardiomyocytes were subjected to 40 min ischaemia and reoxygenation and mitochondrial membrane potential was tracked throughout the experiment by measurement of TMRM fluorescence. Ischaemic buffer contained either no additions or 4 mM dimethyl malonate. e, TMRM signal throughout the entire experiment. f, TMRM signal during the transition from ischaemia to reoxygenation (n = 3). *P < 0.05 (two-tailed Student's t-test and one-way ANOVA). Data are mean ± s.e.m. of at least three biological replicates. Replicates represent separate experiments on independent cell preparations.

Extended Data Figure 10 Quantification of CAC intermediates in the heart following infusion of dimethyl succinate and in the brain after infusion of dimethyl malonate, and extended summary cytoprotection and neurological scores of rats subjected to tMCAO IR in vivo ± dimethyl malonate infusion.

a, Effect of intravenous infusion of dimethyl succinate on CAC metabolite abundance in the ischaemic and non-ischaemic myocardium (normoxia and peripheral heart tissue plus dimethyl succinate n = 3; ischaemia plus dimethyl succinate n = 4; α- ketoglutarate and aconitate in peripheral heart tissue n = 2). b, Profile of mitochondrial CAC metabolite levels after tMCAO ischaemia ± dimethyl malonate (n = 4). c, Representative images of cross-sections from rat brains after undergoing tMCAO in vivo ± treatment with dimethyl malonate. Brains were treated with haematoxylin and eosin to delineate infarcted tissue. d, Locomotor and sensorimotor assessment of rats by quantification of average number of footfalls after tCMAO ± dimethyl malonate (control n = 6; dimethyl malonate n = 4). *P < 0.05, **P < 0.01 (two-tailed Student’s t-test for pairwise comparisons, and one-way (a, b) or two-way (d) ANOVA for multiple comparisons). Data are mean ± s.e.m. of at least three biological replicates, unless otherwise stated.

Supplementary information

Supplementary Data

This file contains Supplementary Table 1. (XLSX 82 kb)

Supplementary Data

This file contains Supplementary Table 2. (XLSX 29 kb)

Supplementary Data

This file contains Supplementary Table 3. (XLSX 43 kb)

Supplementary data

This file contains Supplementary Table 4. (XLSX 13 kb)

Supplementary Data

This file contains Supplementary Table 5. (XLSX 122 kb)

Supplementary Data

This file contains Supplementary Table 6. (XLSX 14 kb)

PowerPoint slides

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chouchani, E., Pell, V., Gaude, E. et al. Ischaemic accumulation of succinate controls reperfusion injury through mitochondrial ROS. Nature 515, 431–435 (2014). https://doi.org/10.1038/nature13909

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nature13909

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Search

Quick links

Nature Briefing: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research