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:

Acylspermidines are conserved mitochondrial sirtuin-dependent metabolites

Abstract

Sirtuins are nicotinamide adenine dinucleotide (NAD+)-dependent protein lysine deacylases regulating metabolism and stress responses; however, characterization of the removed acyl groups and their downstream metabolic fates remains incomplete. Here we employed untargeted comparative metabolomics to reinvestigate mitochondrial sirtuin biochemistry. First, we identified N-glutarylspermidines as metabolites downstream of the mitochondrial sirtuin SIR-2.3 in Caenorhabditis elegans and demonstrated that SIR-2.3 functions as a lysine deglutarylase and that N-glutarylspermidines can be derived from O-glutaryl-ADP-ribose. Subsequent targeted analysis of C. elegans, mouse and human metabolomes revealed a chemically diverse range of N-acylspermidines, and formation of N-succinylspermidines and/or N-glutarylspermidines was observed downstream of mammalian mitochondrial sirtuin SIRT5 in two cell lines, consistent with annotated functions of SIRT5. Finally, N-glutarylspermidines were found to adversely affect C. elegans lifespan and mammalian cell proliferation. Our results indicate that N-acylspermidines are conserved metabolites downstream of mitochondrial sirtuins that facilitate annotation of sirtuin enzymatic activities in vivo and may contribute to sirtuin-dependent phenotypes.

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: Comparative metabolomics uncovers decrease in N-glutarylspermidines in sir-2.3 mutants.
Fig. 2: Protein lysine deglutarylation is dependent on sir-2.3.
Fig. 3: OAADPR can serve as an acyl donor.
Fig. 4: N-acylspermidines are mammalian metabolites and dependent on SIRT5.
Fig. 5: N-glutarylspermidine production is linked to mitochondrial metabolism.
Fig. 6: Bioactivity of N-glutarylspermidines and model for N-acylspermidine production and function.

Similar content being viewed by others

Data availability

Information on newly identified metabolites has been deposited in the SMID database (https://www.smid-db.org) and the ChEBI dictionary (https://www.ebi.ac.uk/chebi/); web links and accession codes for each metabolite are listed in Supplementary Table 1. HPLC–MS data are available in the MassIVE database under accession number MSV000091755, including C18LC–MS data of WT and sir-2.3-mutant C. elegans for untargeted metabolomic analysis, HILIC–MS data showing separation of N-glutarylspermidine isomers, HILIC-targeted MS/MS data for characterization of N-acylspermidines and N-acylspermines and HILIC–MS detection of N-acylspermidines and N-acylspermines in mouse tissues. Protein sequences were retrieved from UniProt (https://www.uniprot.org), and protein 3D structure predictions were retrieved from the AlphaFold Protein Structure Database (https://alphafold.ebi.ac.uk) under accession codes SIRT4 (human), Q9Y6E7; SIRT4 (mouse), Q8R216; SIR-2.2 (C. elegans), Q20480; SIR-2.3 (C. elegans), Q20481. Source data are provided with this paper.

References

  1. Diehl, K. L. & Muir, T. W. Chromatin as a key consumer in the metabolite economy. Nat. Chem. Biol. 16, 620–629 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Ringel, A. E., Tucker, S. A. & Haigis, M. C. Chemical and physiological features of mitochondrial acylation. Mol. Cell 72, 610–624 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Jiang, H. et al. Protein lipidation: occurrence, mechanisms, biological functions, and enabling technologies. Chem. Rev. 118, 919–988 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Carrico, C., Meyer, J. G., He, W., Gibson, B. W. & Verdin, E. The mitochondrial acylome emerges: proteomics, regulation by sirtuins, and metabolic and disease implications. Cell Metab. 27, 497–512 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Walsh, C. T., Tu, B. P. & Tane, Y. Eight kinetically stable but thermodynamically activated molecules that power cell metabolism. Chem. Rev. 118, 1460–1494 (2018).

    Article  CAS  PubMed  Google Scholar 

  6. Harmel, R. & Fiedler, D. Features and regulation of non-enzymatic post-translational modifications. Nat. Chem. Biol. 14, 244–252 (2018).

    Article  CAS  PubMed  Google Scholar 

  7. Wagner, G. R. & Payne, R. M. Widespread and enzyme-independent Nε-acetylation and Nε-succinylation of proteins in the chemical conditions of the mitochondrial matrix. J. Biol. Chem. 288, 29036–29045 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Wagner, G. R. et al. A class of reactive acyl-CoA species reveals the non-enzymatic origins of protein acylation. Cell Metab. 25, 823–837 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Dokmanovic, M., Clarke, C. & Marks, P. A. Histone deacetylase inhibitors: overview and perspectives. Mol. Cancer Res. 5, 981–989 (2007).

    Article  CAS  PubMed  Google Scholar 

  10. Michishita, E., Park, J. Y., Burneskis, J. M., Barrett, J. C. & Horikawa, I. Evolutionarily conserved and nonconserved cellular localizations and functions of human SIRT proteins. Mol. Biol. Cell 16, 4623–4635 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. van de Ven, R. A. H., Santos, D. & Haigis, M. C. Mitochondrial sirtuins and molecular mechanisms of aging. Trends Mol. Med. 23, 320–331 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  12. Chalkiadaki, A. & Guarente, L. The multifaceted functions of sirtuins in cancer. Nat. Rev. Cancer 15, 608–624 (2015).

  13. Abril, Y. L. N. et al. Pharmacological and genetic perturbation establish SIRT5 as a promising target in breast cancer. Oncogene 40, 1644–1658 (2021).

    Article  CAS  PubMed  Google Scholar 

  14. Jing, H. & Lin, H. Sirtuins in epigenetic regulation. Chem. Rev. 115, 2350–2375 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Jing, H. et al. SIRT2 and lysine fatty acylation regulate the transforming activity of K-Ras4a. eLife 6, e32436 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  16. Sadhukhan, S. et al. Metabolomics-assisted proteomics identifies succinylation and SIRT5 as important regulators of cardiac function. Proc. Natl Acad. Sci. USA 113, 4320–4325 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Tan, M. et al. Lysine glutarylation is a protein posttranslational modification regulated by SIRT5. Cell Metab. 19, 605–617 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Artyukhin, A. B. et al. Metabolomic ‘dark matter’ dependent on peroxisomal β-oxidation in Caenorhabditis elegans. J. Am. Chem. Soc. 140, 2841–2852 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Frye, R. A. Phylogenetic classification of prokaryotic and eukaryotic Sir2-like proteins. Biochem. Biophys. Res. Commun. 273, 793–798 (2000).

    Article  CAS  PubMed  Google Scholar 

  20. Wang, Y. M. & Tissenbaum, H. A. Overlapping and distinct functions for a Caenorhabditis elegans SIR2 and DAF-16/FOXO. Mech. Ageing Dev. 127, 48–56 (2006).

    Article  CAS  PubMed  Google Scholar 

  21. Jedrusik-Bode, M. et al. The sirtuin SIRT6 regulates stress granule formation in C. elegans and mammals. J. Cell Sci. 126, 5166–5167 (2013).

    CAS  PubMed  Google Scholar 

  22. Wirth, M. et al. Mitochondrial SIRT4-type proteins in Caenorhabditis elegans and mammals interact with pyruvate carboxylase and other acetylated biotin-dependent carboxylases. Mitochondrion 13, 705–720 (2013).

    Article  CAS  PubMed  Google Scholar 

  23. Helf, M. J., Fox, B. W., Artyukhin, A. B., Zhang, Y. K. & Schroeder, F. C. Comparative metabolomics with Metaboseek reveals functions of a conserved fat metabolism pathway in C. elegans. Nat. Commun. 13, 782 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Hada, K. et al. Tricarboxylic acid cycle activity suppresses acetylation of mitochondrial proteins during early embryonic development in Caenorhabditis elegans. J. Biol. Chem. 294, 3091–3099 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Dubin, D. T. & Rosenthal, S. M. The acetylation of polyamines in Escherichia coli. J. Biol. Chem. 235, 776–782 (1960).

    Article  CAS  PubMed  Google Scholar 

  26. Pannek, M. et al. Crystal structures of the mitochondrial deacylase sirtuin 4 reveal isoform-specific acyl recognition and regulation features. Nat. Commun. 8, 1513 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  27. Galligan, J. J. et al. Quantitative analysis and discovery of lysine and arginine modifications. Anal. Chem. 89, 1299–1306 (2017).

    Article  CAS  PubMed  Google Scholar 

  28. Baldensperger, T., Sanzo, S. D., Ori, A. & Glomb, M. A. Quantitation of reactive acyl-CoA species mediated protein acylation by HPLC–MS/MS. Anal. Chem. 91, 12336–12343 (2019).

    Article  CAS  PubMed  Google Scholar 

  29. Anderson, K. A. et al. SIRT4 is a lysine deacylase that controls leucine metabolism and insulin secretion. Cell Metab. 25, 838–855 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Toninello, A., Dalla Via, L., Siliprandi, D. & Garlid, K. D. Evidence that spermine, spermidine, and putrescine are transported electrophoretically in mitochondria by a specific polyamine uniporter. J. Biol. Chem. 267, 18393–18397 (1992).

    Article  CAS  PubMed  Google Scholar 

  31. Chen, D. et al. Identification of macrodomain proteins as novel O-acetyl-ADP-ribose deacetylases. J. Biol. Chem. 286, 13261–13271 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Du, J. et al. Sirt5 is a NAD-dependent protein lysine demalonylase and desuccinylase. Science 334, 806–809 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Nishida, Y. et al. SIRT5 regulates both cytosolic and mitochondrial protein malonylation with glycolysis as a major target. Mol. Cell 59, 321–332 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Pegg, A. E. & Michael, A. J. Spermine synthase. Cell. Mol. Life Sci. 67, 113–121 (2010).

    Article  CAS  Google Scholar 

  35. Leandro, J. & Houten, S. M. The lysine degradation pathway: subcellular compartmentalization and enzyme deficiencies. Mol. Genet. Metab. 131, 14–22 (2020).

    Article  CAS  PubMed  Google Scholar 

  36. Lee, S. H., Lee, J. H., Lee, H. Y. & Min, K. J. Sirtuin signaling in cellular senescence and aging. BMB Rep. 52, 24–34 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Tissenbaum, H. A. & Guarente, L. Increased dosage of a sir-2 gene extends lifespan in Caenorhabditis elegans. Nature 410, 227–230 (2001).

    Article  CAS  PubMed  Google Scholar 

  38. Schmeisser, K. et al. Role of sirtuins in lifespan regulation is linked to methylation of nicotinamide. Nat. Chem. Biol. 9, 693–700 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Chang, S. M., McReynolds, M. R. & Hanna-Rose, W. Mitochondrial sirtuins sir-2.2 and sir-2.3 regulate lifespan in C. elegans. Preprint at bioRxiv https://doi.org/10.1101/181727 (2017).

  40. Ludewig, A. H. et al. An excreted small molecule promotes C. elegans reproductive development and aging. Nat. Chem. Biol. 15, 838–845 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Ludewig, A. H. et al. Larval crowding accelerates C. elegans development and reduces lifespan. PLoS Genet. 13, e1006717 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  42. Pegg, A. E. Toxicity of polyamines and their metabolic products. Chem. Res. Toxicol. 26, 1782–1800 (2013).

    Article  CAS  PubMed  Google Scholar 

  43. Casero, R. A. Jr., Murray Stewart, T. & Pegg, A. E. Polyamine metabolism and cancer: treatments, challenges and opportunities. Nat. Rev. Cancer 18, 681–695 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Rafty, L. A., Schmidt, M. T., Perraud, A. L., Scharenberg, A. M. & Denu, J. M. Analysis of O-acetyl-ADP-ribose as a target for nudix ADP-ribose hydrolases. J. Biol. Chem. 277, 47114–47122 (2002).

    Article  CAS  PubMed  Google Scholar 

  45. Madeo, F., Eisenberg, T., Pietrocola, F. & Kroemer, G. Spermidine in health and disease. Science 359, eaan2788 (2018).

    Article  PubMed  Google Scholar 

  46. Hai, Y., Shinsky, S. A., Porter, N. J. & Christianson, D. W. Histone deacetylase 10 structure and molecular function as a polyamine deacetylase. Nat. Commun. 8, 15368 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Stewart, T. M. et al. Histone deacetylase-10 liberates spermidine to support polyamine homeostasis and tumor cell growth. J. Biol. Chem. 298, 102407 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Wagner, M., Zhang, B., Tauffenberger, A., Schroeder, F. C. & Skirycz, A. Experimental methods for dissecting the terraincognita of protein–metabolite interactomes. Curr. Opin. Syst. Biol. 28, 100403 (2021).

    Article  CAS  Google Scholar 

  49. O’Neill, E. M. et al. Phevamine A, a small molecule that suppresses plant immune responses. Proc. Natl Acad. Sci. USA 115, E9514–E9522 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  50. Edreva, A. M., Velikova, V. B. & Tsonev, T. D. Phenylamides in plants. Russ. J. Plant Physiol. 54, 287–301 (2007).

    Article  CAS  Google Scholar 

  51. Wang, Q. P. et al. Kukoamine A inhibits human glioblastoma cell growth and migration through apoptosis induction and epithelial–mesenchymal transition attenuation. Sci. Rep. 6, 36543 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Laurent, G. et al. SIRT4 coordinates the balance between lipid synthesis and catabolism by repressing malonyl CoA decarboxylase. Mol. Cell 50, 686–698 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Haigis, M. C. et al. SIRT4 inhibits glutamate dehydrogenase and opposes the effects of calorie restriction in pancreatic beta cells. Cell 126, 941–954 (2006).

    Article  CAS  PubMed  Google Scholar 

  54. Mathias, R. A. et al. Sirtuin 4 is a lipoamidase regulating pyruvate dehydrogenase complex activity. Cell 159, 1615–1625 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. King, J. L. & Jukes, T. H. Non-Darwinian evolution. Science 164, 788–798 (1969).

    Article  CAS  PubMed  Google Scholar 

  56. Keszthelyi, D., Troost, F. J. & Masclee, A. A. M. Understanding the role of tryptophan and serotonin metabolism in gastrointestinal function. Neurogastroenterol. Motil. 21, 1239–1249 (2009).

    Article  CAS  PubMed  Google Scholar 

  57. Bhatt, D. P. et al. Deglutarylation of glutaryl-CoA dehydrogenase by deacylating enzyme SIRT5 promotes lysine oxidation in mice. J. Biol. Chem. 298, 101723 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Minogue, E. et al. Glutarate regulates T cell metabolism and anti-tumour immunity. Nat. Metab. 5, 1747–1764 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Kraemer, B. C., Burgess, J. K., Chen, J. H., Thomas, J. H. & Schellenberg, G. D. Molecular pathways that influence human tau-induced pathology in Caenorhabditis elegans. Hum. Mol. Genet. 15, 1483–1496 (2006).

    Article  CAS  PubMed  Google Scholar 

  60. Kontaxi, C., Piccardo, P. & Gill, A. C. Lysine-directed post-translational modifications of tau protein in Alzheimer’s disease and related tauopathies. Front. Mol. Biosci. 4, 56 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  61. Stewart, T. J. & Abrams, S. I. Altered immune function during long-term host–tumor interactions can be modulated to retard autochthonous neoplastic growth. J. Immunol. 179, 2851–2859 (2007).

    Article  CAS  PubMed  Google Scholar 

  62. Yan, D. Q. et al. SIRT5 is a druggable metabolic vulnerability in acute myeloid. Leuk. Blood Cancer Discov. 2, 266–287 (2021).

    CAS  Google Scholar 

  63. Yu, J. J. et al. Metabolic characterization of a Sirt5 deficient mouse model. Sci. Rep. 3, 2806 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  64. Artyukhin, A. B., Schroeder, F. C. & Avery, L. Density dependence in Caenorhabditis larval starvation. Sci. Rep. 3, 2777 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  65. Karpievitch, Y. V., Dabney, A. R. & Smith, R. D. Normalization and missing value imputation for label-free LC–MS analysis. BMC Bioinformatics 13, S5 (2012).

    CAS  Google Scholar 

  66. Sievers, F. et al. Fast, scalable generation of high-quality protein multiple sequence alignments using Clustal Omega. Mol. Syst. Biol. 7, 539 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  67. Varadi, M. et al. AlphaFold Protein Structure Database: massively expanding the structural coverage of protein-sequence space with high-accuracy models. Nucleic Acids Res. 50, D439–D444 (2022).

    Article  CAS  PubMed  Google Scholar 

  68. Jurrus, E. et al. Improvements to the APBS biomolecular solvation software suite. Protein Sci. 27, 112–128 (2018).

    Article  CAS  PubMed  Google Scholar 

  69. Guex, N. & Peitsch, M. C. SWISS-MODEL and the Swiss-PdbViewer: an environment for comparative protein modeling. Electrophoresis 18, 2714–2723 (1997).

    Article  CAS  PubMed  Google Scholar 

  70. Jonassen, T., Marbois, B. N., Faull, K. F., Clarke, C. F. & Larsen, P. L. Development and fertility in Caenorhabditis elegans clk-1 mutants depend upon transport of dietary coenzyme Q8 to mitochondria. J. Biol. Chem. 277, 45020–45027 (2002).

    Article  CAS  PubMed  Google Scholar 

  71. Byun, J. A. et al. Analysis of polyamines as carbamoyl derivatives in urine and serum by liquid chromatography–tandem mass spectrometry. Biomed. Chromatogr. 22, 73–80 (2008).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank the CGC for C. elegans strains, N. Burgess-Brown for the MACROD1 plasmid, S. Abrams for AT-3 cells, J. Auwerx for Sirt5-KO mice, J. Blum for mouse Sirt5 sgRNA constructs, M. Deininger for the plasmid with SIRT5-targeted shRNA, D. Fajardo Palomino for maintaining and genotyping C. elegans strains, the Cornell Institute of Biotechnology for DNA sequencing, T. H. Won for advice on mouse metabolite extraction, H. Lin and the laboratory of F.C.S. for advice and the laboratory of F.C.S. for comments on the manuscript. This work was partly supported by the NIH (R35 GM131877 to F.C.S., R01 CA223534 to R.S.W., P40 OD010440 to the CGC, GM69702 for APBS-PDB2PQR software). I.R.F. was supported by HHMI Gilliam Fellowship GT11525, and J.M. was supported by NIH grant F30 CA250451.

Author information

Authors and Affiliations

Authors

Contributions

R.S.W. and F.C.S. supervised the study. B.Z., R.S.W. and F.C.S. designed experiments. J.M. and I.R.F. prepared mammalian samples. A.H.L. generated the sir-2.2;sir-2.3 double mutant strain and performed C. elegans-aging assays. J.M. performed compound testing in mammalian cells. B.Z. and T.R.B. performed all other chemical and biological experiments. B.Z. and F.C.S. wrote the manuscript with input from the other authors.

Corresponding author

Correspondence to Frank C. Schroeder.

Ethics declarations

Competing interests

F.C.S. is a cofounder and stockholder of Holoclara and Ascribe Bioscience, which develop nematode-derived small molecules for biomedical and agricultural applications. The remaining authors declare no competing interests.

Peer review

Peer review information

Nature Chemical Biology thanks Matthew Hirschey, and the other, anonymous reviewers for their contribution to the peer review of this work.

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 Model for the regulation of mitochondrial protein lysine acylation by sirtuins.

a, Non-enzymatic protein lysine acylation occurs in the high-pH environment of mitochondria. Four or five-carbon (main chain) dicarboxyl-CoAs can form reactive anhydride intermediates and can be further captured by lysine residues. Sirtuin-mediated deacylation requires NAD+ as a cofactor: following formation of a complex of acylated substrate-ADP-ribose and sirtuin, a conserved histidine residue of the sirtuins deprotonates the 2′-OH of ribose that then attacks the carbonyl carbon of the alkylimidate intermediate to form a 1′,2′-cyclic intermediate, which is finally hydrolyzed to 2′/3′-O-acyl-ADP-ribose (OAADPR). b, Table summarizing the localization and in vitro deacylation activity of seven mammalian sirtuins, with their C. elegans orthologs. C. elegans has two mitochondrial sirtuins, sir-2.2 and sir-2.3; both are orthologs of mammalian SIRT4 that robustly removes five-carbon (main chain) dicarboxyl modifications.

Extended Data Fig. 2 Decreased abundance of N-glutarylspermidine in sir-2.3 null mutants.

a, Untargeted comparative metabolomics of the exo-metabolome from WT and sir-2.2(n5134) mutant C. elegans. Bubble sizes reflect peak areas. n = 4. b, Extracted ion chromatograms (EICs) for m/z 302.2074 ± 5 ppm (C14H28N3O4+) showing that this feature is absent in bacterial food (OP50) and downregulated in the sir-2.3 mutant compared to WT. Representative of n = 5. c, daspid#4 (N1-glutaryl isomer) is more abundant than daspid#3 (N8-glutaryl isomer) quantified by HILIC-MS1. N2 n = 6, sir-2.3 n = 4, sir-2.2;sir-2.3 n = 2. d, Numbering scheme of spermidine. e, f, N-acetyl (e), N-succinyl (f), and N-malonyl (f) spermidines remained unchanged upon sir-2.3 deletion. n = 4. g, Major fragmentation reactions and resulting fragment ions of maspid#3 and maspid#4 in tandem mass spectrometry. Comparison of levels of maspid#3 and maspid#4 in different biological conditions were based on HILIC-PRM. EIC is ESI+ of indicated m/z ± 5 ppm. h, daspid#3 and daspid#4 are mainly secreted, and maspid#4 are mainly retained in worm bodies. daspid#3 and daspid#4: WT, n = 8; sir-2.3, n = 6; sir-2.2;sir-2.3, n = 3. maspid#3 and maspid#4: WT, n = 5; sir-2.3, n = 3; sir-2.2;sir-2.3, n = 2. i, HPLC-MS-based quantification of daspid#3 and daspid#4 in exo-metabolomes collected at different time intervals, corresponding to different developmental stages. daspid#3 and daspid#4 are mainly produced starting in young adulthood. Data were normalized to ascr#3. n = 2. j, Only trace amounts of N-glutarylspermidine derivatives are present in the metabolome of bacterial food (OP50), compared to the C. elegans metabolome. Representative of n > 3. n represents the number of independent experiments in a, e, f, and h, and the number of biologically distinct samples in c and i. P values were calculated by unpaired, two-tailed t-test with Welch correction in a. Data are mean (±s.d.) in c, e, f, and h.

Source data

Extended Data Fig. 3 MS/MS analyses of newly characterized dicarboxylic acid-derived N-acylspermidines.

al, Major fragmentation reactions and resulting fragment ions are indicated with MS/MS spectra of synthetic standards of the glutaryl, succinyl, or malonyl spermidine derivatives, all obtained in ESI+ mode.

Extended Data Fig. 4 O-acyl-ADP-ribose can serve as an acyl donor.

a, Scheme for comparing acyl transfer from acetyl-CoA and O-acetyl-ADPR to spermidine at neutral or slightly basic pH. Products of acyl transfer were derivatized with isobutyl chloroformate and analyzed by HPLC-MS. b, c, Estimation of the second-order rate constants for acyl transfer to form N1-acetylspermidine from AcCoA (b) or OAcADPR (c) at pH 8.0. Number of independent assays, n = 2. d, daspid#3 and daspid#4 are also produced from supplemented D4-glutaric acid. Representative of three experiments. e, Amounts of daspid#3 and daspid#4 produced from supplemented D4-glutaric acid (detected at m/z 306.2325) do not differ between WT and sir-2.3. Number of independent experiments, n = 3. f, Cartoon illustrating the production of N-glutarylspermidine from extracellular lysine and glutaric acid. g, Expression of human MacroD1 (ΔN58aa truncated) with an infusion thioredoxin and His6 tag in Rosetta(DE3) cells grown in LB induced by isopropyl β-D-1-thiogalactopyranoside. Elutions from a Ni-NTA column for recombinant protein purification are shown. h, TEV sequence was cleaved by TEV protease to release the MacroD1 for assays. Shown is the input and column-purified product. i, Steady-state kinetic analysis of OAcADPR and OGADPR hydrolysis catalyzed by human MacroD1 (ΔN58aa truncated) reveals both deacylation reactions follow saturation kinetics. The steady-state kinetic parameters Km and Vmax are determined by HPLC-MS for ADP-ribose formation to be 427 ± 74 µM and 0.17 ± 0.01 s−1 for OAcADPR, and 1020 ± 283 µM and 0.46 ± 0.08 s−1 for OGADPR. The reaction mixtures contain 0.5 µM MacroD1. Number of independent assays using the same batch of purified enzyme, n = 3. Data are mean ± s.d. in e and i. P values were calculated using unpaired, two-tailed t-tests with Welch correction in e. g and h are results from one experiment without attempts of replication.

Source data

Extended Data Fig. 5 Characterization of acylated polyamines in mouse.

a, b, Representative EICs of characteristic fragment ions of different N-acylspermidines using targeted MS/MS across different mouse tissues, demonstrating that dicarboxylic acid-derived spermidines are mammalian metabolites. Gaussian smoothing of the HPLC-MS raw data was enabled for the following metabolites: daspid#3, daspid#7, daspid#8, maspid#7 and #8 (using 5 raw data points); daspid#5, maspid#3 and #4 (using 7 data points). c, d, MS/MS analyses of N1-glutarylspermine (c) and N1-succinylspermine (d). Major fragmentation reactions and resulting fragment ions are indicated with MS/MS spectra of synthetic standards obtained in ESI+ mode. e, Representative EICs of N1-acylspermines across different mouse tissues. The dicarboxylic acid-derived spermines were below detection limit.

Extended Data Fig. 6 Relative abundances of N-acylspermidines detected in mouse liver, brain, and heart samples.

N-acylspermidine levels did not significantly differ between WT and Sirt5 KO samples, except for a decrease of maspid#3 (N8-glutaryl) and maspid#5 (N8-succinyl) in the Sirt5 KO brain samples. Data represent total peak intensity in each sample, based on detected peak area and sample volume, per tissue weight. Mice were all female. n = 5, except for WT mouse brain, n = 4. Data are mean ± s.d. and P values were calculated using unpaired, two-tailed t-tests with Welch correction (P ≥ 0.1 are not shown).

Source data

Extended Data Fig. 7 N-acylspermidines levels in WT and SIRT5 KO cell cultures.

a-g. Comparison of N-acylspermidine levels in WT and Sirt5 KO AT-3 cell cultures, corresponding to Fig. 4b. h-o. Comparison of N-acylspermidine levels in WT and SIRT5 KO HCT116 cell cultures, corresponding to Fig. 4c. Number of biologically distinct samples in each of the three independent experiments, n = 3. Data are mean ± s.d., and P values were calculated by unpaired, two-tailed t-tests with Welch correction (P ≥ 0.15 are not shown).

Source data

Extended Data Fig. 8 N-glutarylspermidine production is linked to mitochondrial metabolism.

a, The lysine origin of the glutaryl moiety in daspid#3 and daspid#4 can be inferred from robust incorporation of five 13C atoms in the molecular ion (a) as well as in the glutaryl-containing fragment (Fig. 5c) in isotope tracing experiments with 13C6,15N2-L-lysine. b, The ratio of labelled daspid#3 and daspid#4 closely reflected that of the corresponding unlabeled compounds in the lysine tracing experiment and was similar to the ratio of products obtained from reaction of O-glutaryl-ADP-ribose with spermidine (Fig. 3d), but differed from that of in the D4-glutaric acid labelling experiments (Fig. 3f), consistent with sir-2.3-dependent formation of daspid#3 and daspid#4 via mitochondrial lysine catabolism. Number of biologically distinct samples, n = 3. c, Representative images for delayed/aborted development of lysine-supplemented sir-2.3 mutant animals. Developmentally delayed animals are indicated by arrows. Scale bar: 1 mm. Spectra in a are representative of three independent experiments. Images in c are representative of four experiments.

Source data

Extended Data Fig. 9 Bioactivity of N-glutarylspermidines.

a, Abundance of N-glutarylspermidines in worm bodies after supplementation of each compound at 5 µM in the growth media. The endogenous levels were first shown in Fig. 3g. Number of biologically distinct samples, n = 3. b, maspid#3 inhibited the clonogenic growth of AT-3 cells. Colony formation assay with AT-3 cells over 12 days after treatment with maspid#3 or maspid#4 in the first 3 days. See Supplementary Fig. 6 for additional images. c, maspid#3 inhibited the clonogenic growth of AT-3 cells culturing in both regular and heat-inactivated bovine serum. d, maspid#3 inhibited the clonogenic growth of MDA-MB-231 cells. +Dox, doxycycline-induced knockdown of SIRT5. e, Cells took up maspid#3 and maspid#4 to similar extents. Cells were treated with 10 µM of compound for 48 h, and the abundance of compounds in cell pellets was measured by HILIC-HRMS. Number of biologically distinct samples n = 4. f, Known mammalian oxidation reactions of polyamines and polyamine derivatives produce reactive nucleophiles (aldehydes) and hydrogen peroxide. Polyamine oxidase (PAOX) oxidizes N1-acetylated spermidine or spermine. Spermine oxidase (SMOX) oxidizes spermine. g, Hypothetical oxidation intermediates and products of N-glutarylspermidines that contain isotope labels in the glutaryl moiety. Oxidation sites are shown in colored bonds (top). One of the predicted oxidation products of maspid#3 was observed (see h for characterization) when supplementing AT-3 cells. h, Supplementation with maspid#3 resulted in production of N-glutarylputrescine, consistent with polyamine oxidase acting on maspid#3, in AT-3 cells. The glutaryl-containing fragments of N-glutarylputrescine contained the D4-label derived from N8-glutaryl(D4)-spermidine, confirming that N-glutarylputrescine originated from supplemented maspid#3. Representative of three experiments. i, Oxidase(s) in the bovine serum can also oxidize maspid#3 to form N-glutarylputrescine, but this activity can be largely inhibited by heat inactivation before the assay. The growth-inhibition effect of maspid#3 reproduced when culturing cells using heat-inactivated serum (c), indicating that oxidase activity in the culture media was not responsible for the observed effects of maspid#3 on cell proliferation. Number of biologically distinct samples, n = 3. Data are mean ± s.d. in a, e and i.

Source data

Extended Data Fig. 10 Examination of effects of N-glutarylspermidines on SIRT5 in vitro.

a, maspid#3 and maspid#4 were not SIRT5 substrates. O-glutaryl-ADP-ribose was not produced when incubating 100 µM maspid#3 or maspid#4 with 1 µM recombinant SIRT5 and 1 mM NAD+ in Tris buffer (pH 7.5) for 2 hr at 37 °C. A peptide with succinylated lysine was used as positive control. n = 2. b, N-glutarylspermidines had no inhibitory effects on SIRT5. maspid#3 to maspid#4 ratio was 1:1. Nicotinamide was used as a positive control, IC50 = 22 µM. n = 2. n, number of assays using the same commercial enzyme sample.

Source data

Supplementary information

Supplementary Information

Supplementary Figs. 1–7, Notes 1 and 2 and Table 1.

Reporting Summary

Supplementary Data 1

Supporting data for Supplementary Fig. 4.

Source data

Source Data Fig. 1

Source Data for Fig. 1.

Source Data Fig. 2

Source Data for Fig. 2.

Source Data Fig. 3

Source Data for Fig. 3.

Source Data Fig. 4

Source Data for Fig. 4.

Source Data Fig. 5

Source Data for Fig. 5.

Source Data Fig. 6

Source Data for Fig. 6.

Source Data Extended Data Fig. 2

Source Data for Extended Data Fig. 2.

Source Data Extended Data Fig. 4

Source Data for Extended Data Fig. 4.

Source Data Extended Data Fig. 6

Source Data for Extended Data Fig. 6.

Source Data Extended Data Fig. 7

Source Data for Extended Data Fig. 7.

Source Data Extended Data Fig. 8

Source Data for Extended Data Fig. 8.

Source Data Extended Data Fig. 9

Source Data for Extended Data Fig. 9.

Source Data Extended Data Fig. 10

Source Data for Extended Data Fig. 10.

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

Zhang, B., Mullmann, J., Ludewig, A.H. et al. Acylspermidines are conserved mitochondrial sirtuin-dependent metabolites. Nat Chem Biol (2024). https://doi.org/10.1038/s41589-023-01511-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1038/s41589-023-01511-2

This article is cited by

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