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.

  • Review Article
  • Published:

Metabolites as signalling molecules

Abstract

Traditional views of cellular metabolism imply that it is passively adapted to meet the demands of the cell. It is becoming increasingly clear, however, that metabolites do more than simply supply the substrates for biological processes; they also provide critical signals, either through effects on metabolic pathways or via modulation of other regulatory proteins. Recent investigation has also uncovered novel roles for several metabolites that expand their signalling influence to processes outside metabolism, including nutrient sensing and storage, embryonic development, cell survival and differentiation, and immune activation and cytokine secretion. Together, these studies suggest that, in contrast to the prevailing notion, the biochemistry of a cell is frequently governed by its underlying metabolism rather than vice versa. This important shift in perspective places common metabolites as key regulators of cell phenotype and behaviour. Yet the signalling metabolites, and the cognate targets and transducers through which they signal, are only beginning to be uncovered. In this Review, we discuss the emerging links between metabolism and cellular behaviour. We hope this will inspire further dissection of the mechanisms through which metabolic pathways and intermediates modulate cell function and will suggest possible drug targets for diseases linked to metabolic deregulation.

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: Principles of metabolite signals drawn from classical experiments.
Fig. 2: Metabolite signals in the regulation of glycolysis, lactic acid fermentation and the TCA cycle.
Fig. 3: Control of mTORC1 activity by AMP and amino acids.
Fig. 4: Extracellular succinate signalling.
Fig. 5: βOHB signalling in ISCs and PEP signalling in TILs.

Similar content being viewed by others

References

  1. Humphrey, S. J., James, D. E. & Mann, M. Protein phosphorylation: a major switch mechanism for metabolic regulation. Trends Endocrinol. Metab. 26, 676–687 (2015).

    Article  CAS  PubMed  Google Scholar 

  2. Marín-Hernández, Á., Rodríguez-Zavala, J. S., Jasso-Chávez, R., Saavedra, E. & Moreno-Sánchez, R. Protein acetylation effects on enzyme activity and metabolic pathway fluxes. J. Cell. Biochem. 123, 701–718 (2022).

    Article  PubMed  Google Scholar 

  3. Sabari, B. R., Zhang, D., Allis, C. D. & Zhao, Y. Metabolic regulation of gene expression through histone acylations. Nat. Rev. Mol. Cell Biol. 18, 90–101 (2017).

    Article  CAS  PubMed  Google Scholar 

  4. Wu, Q., Schapira, M., Arrowsmith, C. H. & Barsyte-Lovejoy, D. Protein arginine methylation: from enigmatic functions to therapeutic targeting. Nat. Rev. Drug. Discov. 20, 509–530 (2021).

    Article  CAS  PubMed  Google Scholar 

  5. Murn, J. & Shi, Y. The winding path of protein methylation research: milestones and new frontiers. Nat. Rev. Mol. Cell Biol. 18, 517–527 (2017).

    Article  CAS  PubMed  Google Scholar 

  6. Husted, A. S., Trauelsen, M., Rudenko, O., Hjorth, S. A. & Schwartz, T. W. GPCR-mediated signaling of metabolites. Cell Metab. 25, 777–796 (2017).

    Article  CAS  PubMed  Google Scholar 

  7. Krebs, H. A. The history of the tricarboxylic acid cycle. Perspect. Biol. Med. 14, 154–172 (1970).

    Article  CAS  PubMed  Google Scholar 

  8. Barnett, J. A. A history of research on yeasts 5: the fermentation pathway. Yeast 20, 509–543 (2003).

    Article  CAS  PubMed  Google Scholar 

  9. Houten, S. M. & Wanders, R. J. A. A general introduction to the biochemistry of mitochondrial fatty acid β-oxidation. J. Inherit. Metab. Dis. 33, 469–477 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Lowry, O. H. & Passonneau, J. V. Kinetic evidence for multiple binding sites on phosphofructokinase. J. Biol. Chem. 241, 2268–2279 (1966).

    Article  CAS  PubMed  Google Scholar 

  11. Kemp, R. G. & Foe, L. G. Allosteric regulatory properties of muscle phosphofructokinase. Mol. Cell. Biochem. 57, 147–154 (1983).

    Article  CAS  PubMed  Google Scholar 

  12. Hue, L. & Taegtmeyer, H. The Randle cycle revisited: a new head for an old hat. Am. J. Physiol. Endocrinol. Metab. 297, E578–E591 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Icard, P. et al. Why may citrate sodium significantly increase the effectiveness of transarterial chemoembolization in hepatocellular carcinoma? Drug. Resist. Updat. 59, 100790 (2021).

    Article  CAS  PubMed  Google Scholar 

  14. Van Schaftingen, E., Hue, L. & Hers, H. G. Fructose 2,6-bisphosphate, the probably structure of the glucose- and glucagon-sensitive stimulator of phosphofructokinase. Biochem. J. 192, 897–901 (1980).

    Article  PubMed  PubMed Central  Google Scholar 

  15. Hers, H. G. & Van Schaftingen, E. Fructose 2,6-bisphosphate 2 years after its discovery. Biochem. J. 206, 1–12 (1982). An insightful history of the discovery and characterization of F2,6BP. It includes reproductions of plots from early experiments.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Uyeda, K., Furuya, E. & Luby, L. The effect of natural and synthetic d-fructose 2,6-bisphosphate on the regulatory kinetic properties of liver and muscle phosphofructokinases. J. Biol. Chem. 256, 8394–8399 (1981).

    Article  CAS  PubMed  Google Scholar 

  17. Hers, H. G. & Van Schaftingen, E. Fructose 2,6-bisphosphate 2 years after its discovery. Biochem. J. 206, 1–12 (1982).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Christophe, J. Glucagon receptors: from genetic structure and expression to effector coupling and biological responses. Biochim. Biophys. Acta 1241, 45–57 (1995).

    Article  PubMed  Google Scholar 

  19. el-Maghrabi, M. R., Claus, T. H., Pilkis, J. & Pilkis, S. J. Regulation of 6-phosphofructo-2-kinase activity by cyclic AMP-dependent phosphorylation. Proc. Natl Acad. Sci. USA 79, 315–319 (1982).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Pilkis, S. J., El-Maghrabi, M. R. & Claus, T. H. Fructose-2,6-bisphosphate in control of hepatic gluconeogenesis: from metabolites to molecular genetics. Diabetes Care 13, 582–599 (1990).

    Article  CAS  PubMed  Google Scholar 

  21. Muller, A., Unthan-Fechner, K. & Probst, I. Activation of phosphofructokinase 2 by insulin in cultured hepatocytes without accompanying changes of effector levels or cAMP-stimulated protein kinase activity ratios. Eur. J. Biochem. 176, 415–420 (1988).

    Article  CAS  PubMed  Google Scholar 

  22. Nishimura, M. & Uyeda, K. Purification and characterization of a novel xylulose 5-phosphate-activated protein phosphatase catalyzing dephosphorylation of fructose-6-phosphate,2-kinase:fructose-2,6-bisphosphatase. J. Biol. Chem. 270, 26341–26346 (1995).

    Article  CAS  PubMed  Google Scholar 

  23. Morimoto, Y. et al. Insulin pretreatment protects the liver from ischemic damage during Pringle’s maneuver. Surgery 120, 808–815 (1996).

    Article  CAS  PubMed  Google Scholar 

  24. Mokrasch, L. C. & McGilvery, R. W. Purification and properties of fructose-1,6-diphosphatase. J. Biol. Chem. 221, 909–917 (1956).

    Article  CAS  PubMed  Google Scholar 

  25. Van Schaftingen, E. & Hers, H. G. Inhibition of fructose-1,6-bisphosphatase by fructose 2,6-biphosphate. Proc. Natl Acad. Sci. USA 78, 2861–2863 (1981).

    Article  PubMed  PubMed Central  Google Scholar 

  26. Bricker Daniel, K. et al. A mitochondrial pyruvate carrier required for pyruvate uptake in yeast, Drosophila, and humans. Science 337, 96–100 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Jagannathan, V. & Schweet, R. Pyruvic oxidase of pigeon breast muscle. I. Purification and properties of the enzyme. J. Biol. Chem. 196, 551–562 (1952).

    Article  CAS  PubMed  Google Scholar 

  28. Gudi, R., Melissa, M. B.-K., Kedishvili, N. Y., Zhao, Y. & Popov, K. M. Diversity of the pyruvate dehydrogenase kinase gene family in humans. J. Biol. Chem. 270, 28989–28994 (1995).

    Article  CAS  PubMed  Google Scholar 

  29. Kerbey, A. L. et al. Regulation of pyruvate dehydrogenase in rat heart. Mechanism of regulation of proportions of dephosphorylated and phosphorylated enzyme by oxidation of fatty acids and ketone bodies and of effects of diabetes: role of coenzyme A, acetyl-coenzyme A and reduced and oxidized nicotinamide-adenine dinucleotide. Biochem. J. 154, 327–348 (1976).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Bowker-Kinley, M. M., Davis, I. W., Wu, P., Harris, A. R. & Popov, M. K. Evidence for existence of tissue-specific regulation of the mammalian pyruvate dehydrogenase complex. Biochem. J. 329, 191–196 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Sharma, P., Walsh, K. T., Kerr-Knott, K. A., Karaian, J. E. & Mongan, P. D. Pyruvate modulates hepatic mitochondrial functions and reduces apoptosis indicators during hemorrhagic shock in rats. Anesthesiology 103, 65–73 (2005).

    Article  PubMed  Google Scholar 

  32. Woods, M. & Burk, D. Inhibition of tumor cell glycolysis by DPNH2, and reversal of the inhibition by DPN, pyruvate or methylene blue. Z. f.ür. Naturforsch. B 18, 731–748 (1963).

    Article  CAS  Google Scholar 

  33. Huckabee, W. E. Relationships of pyruvate and lactate during anaerobic metabolism. I. Effects of infusion of pyruvate or glucose and of hyperventilation. J. Clin. Invest. 37, 244–254 (1958).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Baumberger, J. P., Jürgensen, J. J. & Bardwell, K. The coupled redox potential of the lactate–enzyme–pyruvate system. J. Gen. Physiol. 16, 961–976 (1933).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. O’Carra, P. & Mulcahy, P. Tissue distribution of mammalian lactate dehydrogenase isoenzymes. Biochem. Soc. Trans. 18, 272–274 (1990).

    Article  PubMed  Google Scholar 

  36. Wang, C.-S. Inhibition of human erythrocyte lactate dehydrogenase by high concentrations of pyruvate. Eur. J. Biochem. 78, 569–574 (1977). Experiments demonstrating the inhibition of high levels of pyruvate on human LDHA. The mechanism involves competition with NADH for binding within the active site.

    Article  CAS  PubMed  Google Scholar 

  37. Rao, Y. et al. Excess exogenous pyruvate inhibits lactate dehydrogenase activity in live cells in an MCT1-dependent manner. J. Biol. Chem. 297, (2021).

  38. DeBerardinis Ralph, J. et al. Beyond aerobic glycolysis: transformed cells can engage in glutamine metabolism that exceeds the requirement for protein and nucleotide synthesis. Proc. Natl Acad. Sci. USA 104, 19345–19350 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Cohen, P. F. & Colman, R. F. Diphosphopyridine nucleotide dependent isocitrate dehydrogenase from pig heart. Characterization of active substrate and modes of regulation. Biochemistry 11, 1501–1508 (1972).

    Article  CAS  PubMed  Google Scholar 

  40. Gabriel, J. L. & Plaut, G. W. E. Inhibition of bovine heart NAD-specific isocitrate dehydrogenase by reduced pyridine nucleotides: modulation of inhibition by ADP, NAD+, calcium2+, citrate, and isocitrate. Biochemistry 23, 2773–2778 (1984).

    Article  CAS  PubMed  Google Scholar 

  41. Chen, R. F. & Plaut, G. W. E. Activation and inhibition of DPN-linked isocitrate dehydrogenase of heart by certain nucleotides. Biochemistry 2, 1023–1032 (1963).

    Article  CAS  PubMed  Google Scholar 

  42. Roche, T. E. & Lawlis, V. B. Structure and regulation of α-ketoglutarate dehydrogenase of bovine kidney. Ann. N. Y. Acad. Sci. 378, 236–249 (1982).

    Article  CAS  PubMed  Google Scholar 

  43. Smith, C., Bryla, J. & Williamson, J. Regulation of mitochondrial α-ketoglutarate metabolism by product inhibition at α-ketoglutarate dehydrogenase. J. Biol. Chem. 249, 1497–1505 (1974).

    Article  CAS  PubMed  Google Scholar 

  44. Fang, J. et al. Expression, purification and characterization of human glutamate dehydrogenase (GDH) allosteric regulatory mutations. Biochem. J. 363, 81–87 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Smith, T. J., Peterson, P. E., Schmidt, T., Fang, J. & Stanley, C. A. Structures of bovine glutamate dehydrogenase complexes elucidate the mechanism of purine regulation. J. Mol. Biol. 307, 707–720 (2001).

    Article  CAS  PubMed  Google Scholar 

  46. Plaitakis, A., Metaxari, M. & Shashidharan, P. Nerve tissue-specific (GLUD2) and housekeeping (GLUD1) human glutamate dehydrogenases are regulated by distinct allosteric mechanisms. J. Neurochem. 75, 1862–1869 (2000).

    Article  CAS  PubMed  Google Scholar 

  47. Jitrapakdee, S. et al. Structure, mechanism and regulation of pyruvate carboxylase. Biochem. J. 413, 369–387 (2008).

    Article  CAS  PubMed  Google Scholar 

  48. Yang, R.-Z., Blaileanu, G., Hansen, B. C., Shuldiner, A. R. & Gong, D.-W. cDNA cloning, genomic structure, chromosomal mapping, and functional expression of a novel human alanine aminotransferase. Genomics 79, 445–450 (2002).

    Article  CAS  PubMed  Google Scholar 

  49. Ouyang, Q. et al. Mutations in mitochondrial enzyme GPT2 cause metabolic dysfunction and neurological disease with developmental and progressive features. Proc. Natl Acad. Sci. USA 113, E5598–E5607 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Scrutton, M. & Utter, M. Pyruvate carboxylase. 3. Some physical and chemical properties of the highly purified enzyme. J. Biol. Chem. 240, 1–9 (1965). The detailed measurement of the effects of acetyl-CoA and ADP on pyruvate carboxylase. Although it later turned out to not be entirely essential, acetyl-CoA appeared to be required for pyruvate carboxylase activity.

    Article  CAS  PubMed  Google Scholar 

  51. Utter, M. & Kurahashi, K. Mechanism of action of oxalacetic carboxylase. J. Biol. Chem. 207, 821–841 (1954).

    Article  CAS  PubMed  Google Scholar 

  52. Freedman, A. D. & Kohn, L. Pyruvate metabolism and control: factors affecting pyruvic carboxylase activity. Science 145, 58–60 (1964).

    Article  CAS  PubMed  Google Scholar 

  53. Keech, D. & Utter, M. Pyruvate carboxylase. II. Properties. J. Biol. Chem. 238, 2609–2614 (1963).

    Article  CAS  PubMed  Google Scholar 

  54. Krebs, H. A., Speake, R. & Hems, R. Acceleration of renal gluconeogenesis by ketone bodies and fatty acids. Biochem. J. 94, 712–720 (1965). A cross-species study showing that in acute kidney slices, acetyl-CoA levels could be increased by treating with acetoacetate or short-chain fatty acids. This stimulated gluconeogenesis beyond a small “sparing action” effect attributed to pyruvate carboxylase activation.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Seegmiller, J. & Horecker, B. Metabolism of 6-phosphogluconic acid in liver and bone marrow. J. Biol. Chem. 194, 261–268 (1952).

    Article  CAS  PubMed  Google Scholar 

  56. Kramer, L. et al. Partial pressure of ammonia versus ammonia in hepatic encephalopathy. Hepatology 31, 30–34 (2000).

    Article  CAS  PubMed  Google Scholar 

  57. Nowinski, S. M., Van Vranken, J. G., Dove, K. K. & Rutter, J. Impact of mitochondrial fatty acid synthesis on mitochondrial biogenesis. Curr. Biol. 28, R1212–R1219 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Srere, P. A. & Bhaduri, A. Incorporation of radioactive citrate into fatty acids. Biochim. Biophys. Acta 59, 487–489 (1962).

    Article  CAS  PubMed  Google Scholar 

  59. Wakil, S. J. A malonic acid derivitive as an intermediate in fatty acid synthesis. J. Am. Chem. Soc. 80, 6465–6465 (1958).

    Article  CAS  Google Scholar 

  60. Martin, D. & Vagelos, P. The mechanism of tricarboxylic acid cycle regulation of fatty acid synthesis. J. Biol. Chem. 237, 1787–1792 (1962).

    Article  CAS  PubMed  Google Scholar 

  61. Bortz, W. & Lynen, F. The inhibition of acetyl CoA carboxylase by long chain acyl CoA derivatives. Biochem. Z. 337, 505–509 (1963).

    CAS  PubMed  Google Scholar 

  62. Williamson, J. R., Browning, E. T., Scholz, R., Kreisberg, R. A. & Fritz, I. B. Inhibition of fatty acid stimulation of gluconeogenesis by (+)-decanoylcarnitine in perfused rat liver. Diabetes 17, 194–208 (1968).

    Article  CAS  PubMed  Google Scholar 

  63. Van Tol, A. & Hülsmann, W. C. The localization of palmitoyl-CoA: carnitine palmitoyltransferase in rat liver. Biochim. Biophys. Acta Bioenerg. 189, 342–353 (1969).

    Article  Google Scholar 

  64. McGarry, J., Leatherman, G. & Foster, D. Carnitine palmitoyltransferase I. The site of inhibition of hepatic fatty acid oxidation by malonyl-CoA. J. Biol. Chem. 253, 4128–4136 (1978).

    Article  CAS  PubMed  Google Scholar 

  65. Goodridge, A. G. Regulation of fatty acid synthesis in isolated hepatocytes: evidence for a physiological role for long chain fatty acyl coenzyme A and citrate. J. Biol. Chem. 248, 4318–4326 (1973).

    Article  CAS  PubMed  Google Scholar 

  66. Heesom, K. J., Moule, S. K. & Denton, R. M. Purification and characterization of an insulin-stimulated protein-serine kinase which phosphorylates acetyl-CoA carboxylase. FEBS Lett. 422, 43–46 (1998).

    Article  CAS  PubMed  Google Scholar 

  67. Brownsey, R. W., Boone, A. N., Elliott, J. E., Kulpa, J. E. & Lee, W. M. Regulation of acetyl-CoA carboxylase. Biochem. Soc. Trans. 34, 223–227 (2006).

    Article  CAS  PubMed  Google Scholar 

  68. Lennox, W. G. X. I. V. The respiratory quotient of the brain and of the extremities in man. Arch. Neurol. Psychiatry 26, 719–724 (1931). An early study of the respiratory quotient, which is the ratio of carbon dioxide produced per molecular oxygen consumed, performed in a large number of human subjects. This work established that the brain relies primarily on carbohydrate oxidation in the fed state.

    Article  CAS  Google Scholar 

  69. Owen, O. E. et al. Brain metabolism during fasting. J. Clin. Invest. 46, 1589–1595 (1967).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Zhang, W.-W. & Churchill, P. Purification of d-β-hydroxybutyrate dehydrogenase from rat brain. Biochem. Cell Biol. 68, 980–983 (1990).

    Article  CAS  PubMed  Google Scholar 

  71. Marks, A. R. et al. Molecular cloning and characterization of (R)-3-hydroxybutyrate dehydrogenase from human heart. J. Biol. Chem. 267, 15459–15463 (1992).

    Article  CAS  PubMed  Google Scholar 

  72. Shepher, D., Yates, D. W. & Garland, P. B. The relationship between the rates of conversion of palmitate into citrate or acetoacetate and the acetyl-coenzyme A content of rat-liver mitochondria. Biochem. J. 97, 38C–40C (1965).

    Article  Google Scholar 

  73. Fukao, T. et al. Molecular cloning and sequence of the complementary DNA encoding human mitochondrial acetoacetyl-coenzyme A thiolase and study of the variant enzymes in cultured fibroblasts from patients with 3-ketothiolase deficiency. J. Clin. Invest. 86, 2086–2092 (1990).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Robinson, A. M. & Williamson, D. H. Physiological roles of ketone bodies as substrates and signals in mammalian tissues. Physiol. Rev. 60, 143–187 (1980).

    Article  CAS  PubMed  Google Scholar 

  75. Sauer, F. & Erfle, J. On the mechanism of acetoacetate synthesis by guinea pig liver fractions. J. Biol. Chem. 241, 30–37 (1966).

    Article  CAS  PubMed  Google Scholar 

  76. Kapahi, P. in Protein Metabolism and Homeostasis in Aging (ed. Tavernarakis, N.) 30–37 (Springer US, 2010). https://doi.org/10.1007/978-1-4419-7002-2_3.

  77. Austad, S. N. & Hoffman, J. M. Is antagonistic pleiotropy ubiquitous in aging biology? Evol. Med. Public Health 2018, 287–294 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  78. Saxton, R. A. & Sabatini, D. M. mTOR signaling in growth, metabolism, and disease. Cell 168, 960–976 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Gwinn, D. M. et al. AMPK phosphorylation of raptor mediates a metabolic checkpoint. Mol. Cell 30, 214–226 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Inoki, K., Zhu, T. & Guan, K.-L. TSC2 mediates cellular energy response to control cell growth and survival. Cell 115, 577–590 (2003).

    Article  CAS  PubMed  Google Scholar 

  81. Carling, D., Zammit, V. A. & Hardie, D. G. A common bicyclic protein kinase cascade inactivates the regulatory enzymes of fatty acid and cholesterol biosynthesis. FEBS Lett. 223, 217–222 (1987).

    Article  CAS  PubMed  Google Scholar 

  82. Harrison, D. E. et al. Rapamycin fed late in life extends lifespan in genetically heterogeneous mice. Nature 460, 392–395 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Martin-Montalvo, A. et al. Metformin improves healthspan and lifespan in mice. Nat. Commun. 4, 2192 (2013).

    Article  PubMed  Google Scholar 

  84. Sancak, Y. et al. The Rag GTPases bind raptor and mediate amino acid signaling to mTORC1. Science 320, 1496–1501 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Chantranupong, L. et al. The CASTOR proteins are arginine sensors for the mTORC1 pathway. Cell 165, 153–164 (2016). Demonstration that arginine binds to CASTOR1 to disrupt its interaction with the GATOR2 complex. This mechanism activates mTORC1 in response to arginine in addition to potential sensing by SLC38A9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Chen, J. et al. SAR1B senses leucine levels to regulate mTORC1 signalling. Nature 596, 281–284 (2021). This work complemented the study by Wolfson et al. (ref. 87) revealing that leucine is sensed in a biphasic manner activating mTORC1. The higher affinity sensor SAR1B interacts with leucine at low concentrations and the lower affinity protein sestrin2 interacts with leucine at higher abundance.

    Article  CAS  PubMed  Google Scholar 

  87. Wolfson Rachel, L. et al. Sestrin2 is a leucine sensor for the mTORC1 pathway. Science 351, 43–48 (2016).

    Article  CAS  PubMed  Google Scholar 

  88. Gu, X. et al. SAMTOR is an S-adenosylmethionine sensor for the mTORC1 pathway. Science 358, 813–818 (2017). This study showed that SAMTOR interacts with SAM to activate mTORC1. This sensor mediates methionine sensing by this pathway through the conversion of methionine into SAM.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Covarrubias, A. J., Perrone, R., Grozio, A. & Verdin, E. NAD+ metabolism and its roles in cellular processes during ageing. Nat. Rev. Mol. Cell Biol. 22, 119–141 (2021).

    Article  CAS  PubMed  Google Scholar 

  90. 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 

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

    Article  CAS  PubMed  Google Scholar 

  92. Luo, J. et al. Negative control of p53 by Sir2α promotes cell survival under stress. Cell 107, 137–148 (2001).

    Article  CAS  PubMed  Google Scholar 

  93. Mouchiroud, L. et al. The NAD+/Sirtuin pathway modulates longevity through activation of mitochondrial UPR and FOXO signaling. Cell 154, 430–441 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Libert, S. et al. SIRT1 activates MAO-A in the brain to mediate anxiety and exploratory drive. Cell 147, 1459–1472 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Nakagawa, T., Lomb, D. J., Haigis, M. C. & Guarente, L. SIRT5 deacetylates carbamoyl phosphate synthetase 1 and regulates the urea cycle. Cell 137, 560–570 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Nakahata, Y. et al. The NAD+-dependent deacetylase SIRT1 modulates CLOCK-mediated chromatin remodeling and circadian control. Cell 134, 329–340 (2008). This study revealed that SIRT1 activity is controlled in a circadian manner. SIRT1 was shown to deacetylate BMAL1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. de Murcia, J. M. et al. Requirement of poly(ADP-ribose) polymerase in recovery from DNA damage in mice and in cells. Proc. Natl Acad. Sci. USA 94, 7303–7307 (1997).

    Article  PubMed  PubMed Central  Google Scholar 

  98. Caldecott, K. W. XRCC1 protein; form and function. DNA Repair. 81, 102664 (2019).

    Article  CAS  PubMed  Google Scholar 

  99. Ruscetti, T. et al. Stimulation of the DNA-dependent protein kinase by poly(ADP-ribose) polymerase. J. Biol. Chem. 273, 14461–14467 (1998).

    Article  CAS  PubMed  Google Scholar 

  100. Gomes, A. P. et al. Declining NAD+ induces a pseudohypoxic state disrupting nuclear–mitochondrial communication during aging. Cell 155, 1624–1638 (2013). This study revealed that NAD+ levels decline with age in mice. Restoring NAD+ levels ameliorates age-related phenotypes in a SIRT1 dependent manner.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Camacho-Pereira, J. et al. CD38 dictates age-related NAD decline and mitochondrial dysfunction through an SIRT3-dependent mechanism. Cell Metab. 23, 1127–1139 (2016). By knocking out Cd38 in mice, this work revealed that increased expression of CD38 with age contributes strongly to the age-related decline in NAD+ levels.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. He, S. & Sharpless, N. E. Senescence in health and disease. Cell 169, 1000–1011 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Covarrubias, A. J. et al. Senescent cells promote tissue NAD+ decline during ageing via the activation of CD38+ macrophages. Nat. Metab. 2, 1265–1283 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Chini, C. C. S. et al. CD38 ecto-enzyme in immune cells is induced during aging and regulates NAD+ and NMN levels. Nat. Metab. 2, 1284–1304 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Tarragó, M. G. et al. A potent and specific CD38 inhibitor ameliorates age-related metabolic dysfunction by reversing tissue NAD+ decline. Cell Metab. 27, 1081–1095.e10 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  106. Peclat, T. R. et al. CD38 inhibitor 78c increases mice lifespan and healthspan in a model of chronological aging. Aging Cell 21, e13589 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Kazak, L. et al. UCP1 deficiency causes brown fat respiratory chain depletion and sensitizes mitochondria to calcium overload-induced dysfunction. Proc. Natl Acad. Sci. USA 114, 7981–7986 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Mills, E. L. et al. Accumulation of succinate controls activation of adipose tissue thermogenesis. Nature 560, 102–106 (2018). Succinate was found to be enriched in BAT and increased there upon cold exposure in mice. This TCA metabolite was taken up from the circulation into these cells where it is oxidized.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Mills, E. L. et al. UCP1 governs liver extracellular succinate and inflammatory pathogenesis. Nat. Metab. 3, 604–617 (2021). Ucp1 knockout mice have increased levels of circulating succinate which signals through SUCNR1 in the liver to mediate inflammation.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Tong, W. et al. The intestine is a major contributor to circulating succinate in mice. FASEB J. 36, e22546 (2022).

    Article  CAS  PubMed  Google Scholar 

  111. Reddy, A. et al. pH-gated succinate secretion regulates muscle remodeling in response to exercise. Cell 183, 62–75.e17 (2020). This work demonstrated that succinate is released from exercising muscle to activate SUCNR1 in stromal cells surrounding myofibres. This signalling induces physiological changes associated with skeletal muscle exercise.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. He, W. et al. Citric acid cycle intermediates as ligands for orphan G-protein-coupled receptors. Nature 429, 188–193 (2004).

    Article  CAS  PubMed  Google Scholar 

  113. Cao, W., Medvedev, A. V., Daniel, K. W. & Collins, S. β-Adrenergic activation of p38 MAP kinase in adipocytes: cAMP induction of the uncoupling protein 1 (UCP1) gene requires p38 MAP kinase. J. Biol. Chem. 276, 27077–27082 (2001).

    Article  CAS  PubMed  Google Scholar 

  114. Wang, Y.-X. et al. Peroxisome-proliferator-activated receptor δ activates fat metabolism to prevent obesity. Cell 113, 159–170 (2003).

    Article  CAS  PubMed  Google Scholar 

  115. Whitehead, A. et al. Brown and beige adipose tissue regulate systemic metabolism through a metabolite interorgan signaling axis. Nat. Commun. 12, 1905 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  116. Pinkosky, S. L. et al. Long-chain fatty acyl-CoA esters regulate metabolism via allosteric control of AMPK β1 isoforms. Nat. Metab. 2, 873–881 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Haczeyni, F., Bell-Anderson, K. S. & Farrell, G. C. Causes and mechanisms of adipocyte enlargement and adipose expansion. Obes. Rev. 19, 406–420 (2018).

    Article  CAS  PubMed  Google Scholar 

  118. Hilgendorf, K. I. et al. Omega-3 fatty acids activate ciliary FFAR4 to control adipogenesis. Cell 179, 1289–1305.e21 (2019). This work demonstrated that omega-3 fatty acids from the diet signal to pre-adipocytes using the cell surface receptor FFAR4. This receptor localizes to the primary cilium of these cells and its activation induces WAT differentiation.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Stiles, J. & Jernigan, T. L. The basics of brain development. Neuropsychol. Rev. 20, 327–348 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  120. Imayoshi, I., Sakamoto, M., Yamaguchi, M., Mori, K. & Kageyama, R. Essential roles of notch signaling in maintenance of neural stem cells in developing and adult brains. J. Neurosci. 30, 3489 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Zheng, X. et al. Metabolic reprogramming during neuronal differentiation from aerobic glycolysis to neuronal oxidative phosphorylation. eLife 5, e13374 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  122. Richard, J. P. Mechanism for the formation of methylglyoxal from triosephosphates. Biochem. Soc. Trans. 21, 549–553 (1993).

    Article  CAS  PubMed  Google Scholar 

  123. Yang, G. et al. A Glo1–methylglyoxal pathway that is perturbed in maternal diabetes regulates embryonic and adult neural stem cell pools in murine offspring. Cell Rep. 17, 1022–1036 (2016).

    Article  CAS  PubMed  Google Scholar 

  124. Rodrigues, D. C. et al. Methylglyoxal couples metabolic and translational control of Notch signalling in mammalian neural stem cells. Nat. Commun. 11, 2018 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Nakadate, Y. et al. The formation of argpyrimidine, a methylglyoxal–arginine adduct, in the nucleus of neural cells. Biochem. Biophys. Res. Commun. 378, 209–212 (2009).

    Article  CAS  PubMed  Google Scholar 

  126. Horner, P. J. & Gage, F. H. Regenerating the damaged central nervous system. Nature 407, 963–970 (2000).

    Article  CAS  PubMed  Google Scholar 

  127. Liuzzi Francis, J. & Lasek Raymond, J. Astrocytes block axonal regeneration in mammals by activating the physiological stop pathway. Science 237, 642–645 (1987).

    Article  Google Scholar 

  128. Li, F. et al. Glial metabolic rewiring promotes axon regeneration and functional recovery in the central nervous system. Cell Metab. 32, 767–785.e7 (2020). This work showed that increasing glycolysis in glial cells induces the secretion of lactate and L-2HG. These molecules signal through inverse agonism of GABAB receptors to increase neuronal cAMP concentration. This led to axonal growth ameliorating damage from spinal cord injury.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Meacham, C. E., DeVilbiss, A. W. & Morrison, S. J. Metabolic regulation of somatic stem cells in vivo. Nat. Rev. Mol. Cell Biol. 23, 428–443 (2022).

    Article  CAS  PubMed  Google Scholar 

  130. Faubert, B., Solmonson, A. & DeBerardinis, R. J. Metabolic reprogramming and cancer progression. Science 368, eaaw5473 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Tyrakis, P. A. et al. S-2-Hydroxyglutarate regulates CD8+ T-lymphocyte fate. Nature 540, 236–241 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. Du, X. & Hu, H. The roles of 2-hydroxyglutarate. Front. Cell Dev. Biol. 9, 651317 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  133. Barker, N. et al. Identification of stem cells in small intestine and colon by marker gene Lgr5. Nature 449, 1003–1007 (2007).

    Article  CAS  PubMed  Google Scholar 

  134. Sato, T. et al. Single Lgr5 stem cells build crypt–villus structures in vitro without a mesenchymal niche. Nature 459, 262–265 (2009).

    Article  CAS  PubMed  Google Scholar 

  135. Cheng, C.-W. et al. Ketone body signaling mediates intestinal stem cell homeostasis and adaptation to diet. Cell 178, 1115–1131.e15 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. Shimazu, T. et al. Suppression of oxidative stress by β-hydroxybutyrate, an endogenous histone deacetylase inhibitor. Science 339, 211–214 (2013).

    Article  CAS  PubMed  Google Scholar 

  137. Chriett, S. et al. Prominent action of butyrate over β-hydroxybutyrate as histone deacetylase inhibitor, transcriptional modulator and anti-inflammatory molecule. Sci. Rep. 9, 742 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  138. Dmitrieva-Posocco, O. et al. β-Hydroxybutyrate suppresses colorectal cancer. Nature 605, 160–165 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  139. Kim, D. & Scherer, P. Obesity, diabetes, and increased cancer progression. Diabetes Metab. J. 45, 799–812 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  140. Bella, F., Godos, J., Ippolito, A., Di Prima, A. & Sciacca, S. Differences in the association between empirically derived dietary patterns and cancer: a meta-analysis. Int. J. Food Sci. Nutr. 68, 402–410 (2017).

    Article  PubMed  Google Scholar 

  141. Liese, A. D. et al. The dietary patterns methods project: synthesis of findings across cohorts and relevance to dietary guidance. J. Nutr. 145, 393–402 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  142. Joh, H.-K. et al. Simple sugar and sugar-sweetened beverage intake during adolescence and risk of colorectal cancer precursors. Gastroenterology 161, 128–142.e20 (2021).

    Article  CAS  PubMed  Google Scholar 

  143. Meyerhardt, J. A. et al. Dietary glycemic load and cancer recurrence and survival in patients with stage III colon cancer: findings from CALGB 89803. JNCI: J. Natl Cancer Inst. 104, 1702–1711 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. Hageman, J. H. et al. Intestinal regeneration: regulation by the microenvironment. Dev. Cell 54, 435–446 (2020).

    Article  CAS  PubMed  Google Scholar 

  145. Taylor, S. R. et al. Dietary fructose improves intestinal cell survival and nutrient absorption. Nature 597, 263–267 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Hall, P. A., Coates, P. J., Ansari, B. & Hopwood, D. Regulation of cell number in the mammalian gastrointestinal tract: the importance of apoptosis. J. Cell Sci. 107, 3569–3577 (1994).

    Article  CAS  PubMed  Google Scholar 

  147. Luo, W. et al. Pyruvate kinase M2 is a PHD3-stimulated coactivator for hypoxia-inducible factor 1. Cell 145, 732–744 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  148. Chang, C.-H. et al. Metabolic competition in the tumor microenvironment is a driver of cancer progression. Cell 162, 1229–1241 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  149. Trebak, M. & Kinet, J.-P. Calcium signalling in T cells. Nat. Rev. Immunol. 19, 154–169 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  150. Ho, P.-C. et al. Phosphoenolpyruvate is a metabolic checkpoint of anti-tumor T cell responses. Cell 162, 1217–1228 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  151. Imamura, K. & Tanaka, T. in Methods in Enzymology Vol. 90 (ed. Wood, W. A.) 150–165 (Academic, 1982).

  152. Xu, K. et al. Glycolysis fuels phosphoinositide 3-kinase signaling to bolster T cell immunity. Science 371, 405–410 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Cham, C. M., Driessens, G., O’Keefe, J. P. & Gajewski, T. F. Glucose deprivation inhibits multiple key gene expression events and effector functions in CD8+ T cells. Eur. J. Immunol. 38, 2438–2450 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  154. Somoza, J. R. et al. Structural, biochemical, and biophysical characterization of idelalisib binding to phosphoinositide 3-kinase δ. J. Biol. Chem. 290, 8439–8446 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  155. Furze, R. C. & Rankin, S. M. Neutrophil mobilization and clearance in the bone marrow. Immunology 125, 281–288 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Khatib-Massalha, E. et al. Lactate released by inflammatory bone marrow neutrophils induces their mobilization via endothelial GPR81 signaling. Nat. Commun. 11, 3547 (2020). These experiments demonstrated that bone marrow neutrophils increase lactate production upon infection with a bacterial pathogen. The secreted lactate signals to endothelial cells via HCAR1 to promote permeability of the endothelial barrier and entry of neutrophils into the circulation.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  157. Zhang, W. et al. Lactate is a natural suppressor of RLR signaling by targeting MAVS. Cell 178, 176–189.e15 (2019). This work revealed that lactate can interact with the MAVS transmembrane domain, preventing its aggregation. Notably, inhibition of LDHA increases type I interferon signalling by limiting lactate inhibition of MAVS, providing an enhanced anti-viral immune response.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  158. Franz, K. M. & Kagan, J. C. Innate immune receptors as competitive determinants of cell fate. Mol. Cell 66, 750–760 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  159. Seth, R. B., Sun, L., Ea, C.-K. & Chen, Z. J. Identification and characterization of MAVS, a mitochondrial antiviral signaling protein that activates NF-κB and IRF3. Cell 122, 669–682 (2005).

    Article  CAS  PubMed  Google Scholar 

  160. Peisley, A., Wu, B., Xu, H., Chen, Z. J. & Hur, S. Structural basis for ubiquitin-mediated antiviral signal activation by RIG-I. Nature 509, 110–114 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  161. Levy, M. M. et al. 2001 SCCM/ESICM/ACCP/ATS/SIS International Sepsis Definitions Conference. Crit. Care Med. 31, 1250–1256 (2003).

    Article  PubMed  Google Scholar 

  162. Ljungström, L. et al. Diagnostic accuracy of procalcitonin, neutrophil–lymphocyte count ratio, C-reactive protein, and lactate in patients with suspected bacterial sepsis. PLOS One 12, e0181704 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  163. Apriletti, J. W., Eberhardt, N. L., Latham, K. R. & Baxter, J. D. Affinity chromatography of thyroid hormone receptors. Biospecific elution from support matrices, characterization of the partially purified receptor. J. Biol. Chem. 256, 12094–12101 (1981).

    Article  CAS  PubMed  Google Scholar 

  164. Venkata Subbaiah, K. C., Hedaya, O., Wu, J., Jiang, F. & Yao, P. Mammalian RNA switches: molecular rheostats in gene regulation, disease, and medicine. Comput. Struct. Biotechnol. J. 17, 1326–1338 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  165. Luo, X. & Kraus, W. L. On PAR with PARP: cellular stress signaling through poly(ADP-ribose) and PARP-1. Genes. Dev. 26, 417–432 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  166. Orsak, T. et al. Revealing the allosterome: systematic identification of metabolite–protein interactions. Biochemistry 51, 225–232 (2012).

    Article  CAS  PubMed  Google Scholar 

  167. Lim, Y. T. et al. An efficient proteome-wide strategy for discovery and characterization of cellular nucleotide–protein interactions. PLOS One 13, e0208273 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  168. Gaetani, M. et al. Proteome integral solubility alteration: a high-throughput proteomics assay for target deconvolution. J. Proteome Res. 18, 4027–4037 (2019).

    Article  CAS  PubMed  Google Scholar 

  169. Bakker, S. J. L., Maaten, J. C. T., Hoorntje, S. J. & Gans, R. O. B. Protection against cardiovascular collapse in an alcoholic patient with thiamine deficiency by concomitant alcoholic ketoacidosis. J. Intern. Med. 242, 179–183 (1997).

    Article  CAS  PubMed  Google Scholar 

  170. Zhang, B. et al. B cell-derived GABA elicits IL-10+ macrophages to limit anti-tumour immunity. Nature 599, 471–476 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  171. Lee, S.-E., Lee, Y. & Lee, G. H. The regulation of glutamic acid decarboxylases in GABA neurotransmission in the brain. Arch. Pharmacal Res. 42, 1031–1039 (2019).

    Article  CAS  Google Scholar 

  172. Ferkany, J. W., Smith, L. A., Seifert, W. E., Caprioli, R. M. & Enna, S. J. Measurement of gamma-aminobutyric acid (GABA) in blood. Life Sci. 22, 2121–2128 (1978).

    Article  CAS  PubMed  Google Scholar 

  173. Song, Y., Shenwu, M., Dhossche, D. M. & Liu, Y.-M. A capillary liquid chromatographic/tandem mass spectrometric method for the quantification of γ-aminobutyric acid in human plasma and cerebrospinal fluid. J. Chromatogr. B 814, 295–302 (2005).

    Article  CAS  Google Scholar 

  174. Pearl, P. L., Hartka, T. R., Cabalza, J. L., Taylor, J. & Gibson, M. K. Inherited disorders of GABA metabolism. Future Neurol. 1, 631–636 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  175. Carvalho, D. P. & Dupuy, C. Thyroid hormone biosynthesis and release. Mol. Cell. Endocrinol. 458, 6–15 (2017).

    Article  CAS  PubMed  Google Scholar 

  176. Ashizawa, K., McPhie, P., Lin, K. H. & Cheng, S. Y. An in vitro novel mechanism of regulating the activity of pyruvate kinase M2 by thyroid hormone and fructose 1,6-bisphosphate. Biochemistry 30, 7105–7111 (1991).

    Article  CAS  PubMed  Google Scholar 

  177. Healy, M. E. et al. Dietary sugar intake increases liver tumor incidence in female mice. Sci. Rep. 6, 22292 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  178. Reinfeld, B. I. et al. Cell-programmed nutrient partitioning in the tumour microenvironment. Nature 593, 282–288 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  179. Krautkramer, K. A., Fan, J. & Bäckhed, F. Gut microbial metabolites as multi-kingdom intermediates. Nat. Rev. Microbiol. 19, 77–94 (2021).

    Article  CAS  PubMed  Google Scholar 

  180. Sorbara, M. T. & Pamer, E. G. Microbiome-based therapeutics. Nat. Rev. Microbiol. 20, 365–380 (2022).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

The authors thank several members of the Rutter laboratory for their careful reading and helpful comments on the manuscript, specifically A. Cluntun, C. Cunningham, A. Leifer, K. Hicks, J. Morgan and S. Johnson. They also thank E. Taylor and T. Ryan for insightful discussion of particular topics.

Author information

Authors and Affiliations

Authors

Contributions

The authors contributed equally to all aspects of the article.

Corresponding authors

Correspondence to Steven Andrew Baker or Jared Rutter.

Ethics declarations

Competing interests

J.R. is a founder and consultant to Atavistik Bio. S.A.B. has no competing interests to disclose.

Peer review

Peer review information

Nature Reviews Molecular Cell Biology thanks the 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.

Supplementary information

Glossary

Carnitine

This common metabolite, named for its abundance in meat, is a critical shuttling molecule when conjugated to long-chain fatty acids. Acyl-carnitines are transported by a dedicated channel (SLC25A20) into the mitochondrial matrix for catabolism. Structurally, carnitine resembles a quaternary ammonium attached to the γ-carbon of β-hydroxybutyrate (βOHB), but in mammals this metabolite is synthesized from lysine via methylation.

Fermentation

The extraction of energy from carbohydrates without the requirement for oxygen. An example in vertebrates would be the generation of ATP from glycolysis where the final product pyruvate is reduced to lactate, thereby recycling NADH to NAD+.

Gluconeogenesis

Literally the synthesis of new glucose. During times of fasting in multicellular organisms, some tissues supply glucose to others by converting smaller carbon-containing metabolites into the six-carbon molecule d-glucose. In vertebrates there are several tissues that contribute to this process including the liver, kidney, small intestine and pancreas. Sources of carbon include lactate, pyruvate, glycerol and several amino acids.

Ketone bodies

Small, water-soluble energy carrrying metabolites derived from fat. These are generally synthesized in the liver and consist of acetone, acetoacetate and β-hydroxybutyrate (βOHB). They are characterized by a ketone functional group. 

Mass action

In chemistry, reactions typically settle on an equilibrium determined by the steady-state level of substrates and products that are achieved under a particular set of conditions. In this state, the forward and reverse reactions still occur but are equivalent in their rate. This steady state can be perturbed to drive the reaction in either the forward or reverse direction by increasing the concentration of one or more reactants such that equilibrium is re-established. Driving the reaction by the addition of reactants to restore equilibrium is an example of mass action.

One-carbon metabolism

The central metabolic pathways generating one-carbon groups for synthesis of nucleotides, amines, phospholipids and creatine. Folate serves as a central cofactor for these pathways and functions as a methyl carrier transferring one-carbon groups directly needed for nucleotide synthesis or transferring that methyl group to the methionine cycle to generate S-adenosyl methionine (SAM).

Pattern recognition receptors

The innate immune system uses certain molecular patterns that are uncommon in healthy host physiology to detect pathogens and cellular damage. These pattern antigens include lipopolysaccharide (LPS), mannose, double-stranded RNA, cytoplasmic DNA, certain bacterial proteins such as flagellin, peptidoglycan and fungal glucans. These molecules are recognized by this class of receptors and help stimulate the immune system.

Peroxisome proliferator-activated receptor-δ

(PPARδ). This nuclear hormone receptor binds to certain fatty acids, such as arachidonic acid, and mediates transcriptional regulation in the nucleus. Expression of its encoding gene (PPARD) is upregulated during adipose catabolism and regulates lipid metabolism.

Phosphoinositide 3-kinase

(PI3K). This family of kinases phosphorylates phosphatidylinositol (PI) at the 3-hydroxy position of the inositol ring. PI is a common constituent of the inner leaflet of the plasma membrane. Although this family consists of numerous kinases with diverse functions, a major class (class I) of this family converts phosphatidylinositol (4,5)-bisphosphate (PI(4,5)P2) into phosphatidylinositol (3,4,5)-trisphosphate (PI(3,4,5)P3), activating downstream Akt protein kinases to regulate cell growth and metabolism.

RNA riboswitches

Structural RNA motifs that interact with a small molecule to alter the function of the containing mRNA. These RNA segments have regulatory function that changes the stability, modification or translation of the mRNA.

S-Adenosyl methionine

(SAM). A metabolite that serves as a universal methyl donor and is synthesized from ATP and methionine by the enzyme S-adenosyl methionine synthetase.

Warburg effect

The observation made by Otto Warburg that tumour cells often consume high levels of glucose and generate lactate even in the presence of available oxygen. Also referred to as aerobic glycolysis.

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

Baker, S.A., Rutter, J. Metabolites as signalling molecules. Nat Rev Mol Cell Biol 24, 355–374 (2023). https://doi.org/10.1038/s41580-022-00572-w

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41580-022-00572-w

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