Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

Mitochondrial TNAP controls thermogenesis by hydrolysis of phosphocreatine

Abstract

Adaptive thermogenesis has attracted much attention because of its ability to increase systemic energy expenditure and to counter obesity and diabetes1,2,3. Recent data have indicated that thermogenic fat cells use creatine to stimulate futile substrate cycling, dissipating chemical energy as heat4,5. This model was based on the super-stoichiometric relationship between the amount of creatine added to mitochondria and the quantity of oxygen consumed. Here we provide direct evidence for the molecular basis of this futile creatine cycling activity in mice. Thermogenic fat cells have robust phosphocreatine phosphatase activity, which is attributed to tissue-nonspecific alkaline phosphatase (TNAP). TNAP hydrolyses phosphocreatine to initiate a futile cycle of creatine dephosphorylation and phosphorylation. Unlike in other cells, TNAP in thermogenic fat cells is localized to the mitochondria, where futile creatine cycling occurs. TNAP expression is powerfully induced when mice are exposed to cold conditions, and its inhibition in isolated mitochondria leads to a loss of futile creatine cycling. In addition, genetic ablation of TNAP in adipocytes reduces whole-body energy expenditure and leads to rapid-onset obesity in mice, with no change in movement or feeding behaviour. These data illustrate the critical role of TNAP as a phosphocreatine phosphatase in the futile creatine cycle.

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: Isolation and identification of TNAP as a cold-inducible PCr phosphatase from mitochondria of thermogenic fat.
Fig. 2: TNAP targets mitochondria in brown adipocytes.
Fig. 3: Ablation of TNAP activity abolishes the FCC by hydrolysis of phosphocreatine.
Fig. 4: Ablation of TNAP in fat represses adaptive thermogenesis and stimulates obesity.

Similar content being viewed by others

Data availability

The mass spectrometry proteomics data have been deposited to the ProteomeXchange Consortium via the PRIDE partner repository (https://www.ebi.ac.uk/pride/) with the dataset identifier PXD025032. The published adipocyte-specific ribosomal profiling dataset can be downloaded at https://ars.els-cdn.com/content/image/1-s2.0-S1550413118301839-mmc2.xlsx. Full scans for all western blots are provided in the Supplementary Information. All other data are available from the corresponding author on reasonable request. Source data are provided with this paper.

References

  1. Rosen, E. D. & Spiegelman, B. M. What we talk about when we talk about fat. Cell 156, 20–44 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Harms, M. & Seale, P. Brown and beige fat: development, function and therapeutic potential. Nat. Med. 19, 1252–1263 (2013).

    Article  CAS  PubMed  Google Scholar 

  3. Carobbio, S., Guénantin, A. C., Samuelson, I., Bahri, M. & Vidal-Puig, A. Brown and beige fat: from molecules to physiology and pathophysiology. Biochim. Biophys. Acta Mol. Cell Biol. Lipids 1864, 37–50 (2019).

    Article  CAS  PubMed  Google Scholar 

  4. Chouchani, E. T., Kazak, L. & Spiegelman, B. M. New advances in adaptive thermogenesis: UCP1 and beyond. Cell Metab. 29, 27–37 (2019).

    Article  CAS  PubMed  Google Scholar 

  5. Kazak, L. et al. A creatine-driven substrate cycle enhances energy expenditure and thermogenesis in beige fat. Cell 163, 643–655 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Cypess, A. M. et al. Anatomical localization, gene expression profiling and functional characterization of adult human neck brown fat. Nat. Med. 19, 635–639 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Cypess, A. M. et al. Activation of human brown adipose tissue by a β3-adrenergic receptor agonist. Cell Metab. 21, 33–38 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Leitner, B. P. et al. Mapping of human brown adipose tissue in lean and obese young men. Proc. Natl Acad. Sci. USA 114, 8649–8654 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Bachman, E. S. et al. βAR signaling required for diet-induced thermogenesis and obesity resistance. Science 297, 843–845 (2002).

    Article  CAS  PubMed  ADS  Google Scholar 

  10. Zeng, X. et al. Innervation of thermogenic adipose tissue via a calsyntenin 3β–S100b axis. Nature 569, 229–235 (2019).

    Article  PubMed  PubMed Central  ADS  CAS  Google Scholar 

  11. Enerbäck, S. et al. Mice lacking mitochondrial uncoupling protein are cold-sensitive but not obese. Nature 387, 90–94 (1997).

    Article  PubMed  ADS  Google Scholar 

  12. Hofmann, W. E., Liu, X., Bearden, C. M., Harper, M. E. & Kozak, L. P. Effects of genetic background on thermoregulation and fatty acid-induced uncoupling of mitochondria in UCP1-deficient mice. J. Biol. Chem. 276, 12460–12465 (2001).

    Article  CAS  PubMed  Google Scholar 

  13. 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 

  14. Ikeda, K. et al. UCP1-independent signaling involving SERCA2b-mediated calcium cycling regulates beige fat thermogenesis and systemic glucose homeostasis. Nat. Med. 23, 1454–1465 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Long, J. Z. et al. The secreted enzyme PM20D1 regulates lipidated amino acid uncouplers of mitochondria. Cell 166, 424–435 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Kazak, L. et al. Genetic depletion of adipocyte creatine metabolism inhibits diet-induced thermogenesis and drives obesity. Cell Metab. 26, 660–671.e3 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Kazak, L. et al. Ablation of adipocyte creatine transport impairs thermogenesis and causes diet-induced obesity. Nat. Metab. 1, 360–370 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Rahbani, J. F. et al. Creatine kinase B controls futile creatine cycling in thermogenic fat. Nature 590, 480–485 (2021).

    Article  CAS  PubMed  ADS  PubMed Central  Google Scholar 

  19. Müller, S. et al. Proteomic analysis of human brown adipose tissue reveals utilization of coupled and uncoupled energy expenditure pathways. Sci. Rep. 6, 30030 (2016).

    Article  PubMed  PubMed Central  ADS  CAS  Google Scholar 

  20. Svensson, P. A. et al. Gene expression in human brown adipose tissue. Int. J. Mol. Med. 27, 227–232 (2011).

    Article  CAS  PubMed  Google Scholar 

  21. Bessman, S. P. & Carpenter, C. L. The creatine–creatine phosphate energy shuttle. Annu. Rev. Biochem. 54, 831–862 (1985).

    Article  CAS  PubMed  Google Scholar 

  22. Pinkerton, A. B. et al. Discovery of 5-((5-chloro-2-methoxyphenyl)sulfonamido)nicotinamide (SBI-425), a potent and orally bioavailable tissue-nonspecific alkaline phosphatase (TNAP) inhibitor. Bioorg. Med. Chem. Lett. 28, 31–34 (2018).

    Article  CAS  PubMed  Google Scholar 

  23. Roh, H. C. et al. Warming induces significant reprogramming of beige, but not brown, adipocyte cellular identity. Cell Metab. 27, 1121–1137.e5 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Buchet, R., Millán, J. L. & Magne, D. Multisystemic functions of alkaline phosphatases. Methods Mol. Biol. 1053, 27–51 (2013).

    Article  CAS  PubMed  Google Scholar 

  25. Hessle, L. et al. Tissue-nonspecific alkaline phosphatase and plasma cell membrane glycoprotein-1 are central antagonistic regulators of bone mineralization. Proc. Natl Acad. Sci. USA 99, 9445–9449 (2002).

    Article  CAS  PubMed  ADS  PubMed Central  Google Scholar 

  26. Anderson, H. C. et al. Impaired calcification around matrix vesicles of growth plate and bone in alkaline phosphatase-deficient mice. Am. J. Pathol. 164, 841–847 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Lam, S. S. et al. Directed evolution of APEX2 for electron microscopy and proximity labeling. Nat. Methods 12, 51–54 (2015).

    Article  CAS  PubMed  ADS  Google Scholar 

  28. Rhee, H. W. et al. Proteomic mapping of mitochondria in living cells via spatially restricted enzymatic tagging. Science 339, 1328–1331 (2013).

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  29. Kiffer-Moreira, T. et al. Catalytic signature of a heat-stable, chimeric human alkaline phosphatase with therapeutic potential. PLoS ONE 9, e89374 (2014).

    Article  PubMed  PubMed Central  ADS  CAS  Google Scholar 

  30. Chen, W. W., Freinkman, E., Wang, T., Birsoy, K. & Sabatini, D. M. Absolute quantification of matrix metabolites reveals the dynamics of mitochondrial metabolism. Cell 166, 1324–1337.e11 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Chen, W. W., Freinkman, E. & Sabatini, D. M. Rapid immunopurification of mitochondria for metabolite profiling and absolute quantification of matrix metabolites. Nat. Protoc. 12, 2215–2231 (2017).

    PubMed  PubMed Central  Google Scholar 

  32. Haarhaus, M., Brandenburg, V., Kalantar-Zadeh, K., Stenvinkel, P. & Magnusson, P. Alkaline phosphatase: a novel treatment target for cardiovascular disease in CKD. Nat. Rev. Nephrol. 13, 429–442 (2017).

    Article  CAS  PubMed  Google Scholar 

  33. Liu, X. et al. Paradoxical resistance to diet-induced obesity in UCP1-deficient mice. J. Clin. Invest. 111, 399–407 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Feldmann, H. M., Golozoubova, V., Cannon, B. & Nedergaard, J. UCP1 ablation induces obesity and abolishes diet-induced thermogenesis in mice exempt from thermal stress by living at thermoneutrality. Cell Metab. 9, 203–209 (2009).

    Article  CAS  PubMed  Google Scholar 

  35. Vendelin, M., Lemba, M. & Saks, V. A. Analysis of functional coupling: mitochondrial creatine kinase and adenine nucleotide translocase. Biophys. J. 87, 696–713 (2004).

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  36. Shulman, G. I., Ladenson, P. W., Wolfe, M. H., Ridgway, E. C. & Wolfe, R. R. Substrate cycling between gluconeogenesis and glycolysis in euthyroid, hypothyroid, and hyperthyroid man. J. Clin. Invest. 76, 757–764 (1985).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Wolfe, R. R., Herndon, D. N., Jahoor, F., Miyoshi, H. & Wolfe, M. Effect of severe burn injury on substrate cycling by glucose and fatty acids. N. Engl. J. Med. 317, 403–408 (1987).

    Article  CAS  PubMed  Google Scholar 

  38. Clark, M. G. et al. Accelerated substrate cycling of fructose-6-phosphate in the muscle of malignant hyperthermic pigs. Nature 245, 99–101 (1973).

    Article  CAS  PubMed  ADS  Google Scholar 

  39. Newsholme, E. A. & Crabtree, B. Substrate cycles in metabolic regulation and in heat generation. Biochem. Soc. Symp. 61–109 (1976).

  40. Klein, S. & Wolfe, R. R. Whole-body lipolysis and triglyceride-fatty acid cycling in cachectic patients with esophageal cancer. J. Clin. Invest. 86, 1403–1408 (1990).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Love, M. I., Huber, W. & Anders, S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 15, 550 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  42. Klein, J. et al. β3-adrenergic stimulation differentially inhibits insulin signaling and decreases insulin-induced glucose uptake in brown adipocytes. J. Biol. Chem. 274, 34795–34802 (1999).

    Article  CAS  PubMed  Google Scholar 

  43. Wu, J. et al. Beige adipocytes are a distinct type of thermogenic fat cell in mouse and human. Cell 150, 366–376 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Sharabi, K. et al. Selective chemical inhibition of PGC-1α gluconeogenic activity ameliorates type 2 diabetes. Cell 169, 148–160.e15 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Springer, M. L., Rando, T. A. & Blau, H. M. Gene delivery to muscle. Curr. Protoc. Hum. Genet. 31, 13.4.1–13.4.19 (2002).

    Google Scholar 

  46. Allan, C. et al. OMERO: flexible, model-driven data management for experimental biology. Nat. Methods 9, 245–253 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Schindelin, J. et al. Fiji: an open-source platform for biological-image analysis. Nat. Methods 9, 676–682 (2012).

    Article  CAS  PubMed  Google Scholar 

  48. Yuan, M., Breitkopf, S. B., Yang, X. & Asara, J. M. A positive/negative ion-switching, targeted mass spectrometry-based metabolomics platform for bodily fluids, cells, and fresh and fixed tissue. Nat. Protoc. 7, 872–881 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Foster, B. L. et al. Conditional Alpl ablation phenocopies dental defects of hypophosphatasia. J. Dent. Res. 96, 81–91 (2017).

    Article  CAS  PubMed  Google Scholar 

  50. Gettins, P., Metzler, M. & Coleman, J. E. Alkaline phosphatase. 31P NMR probes of the mechanism. J. Biol. Chem. 260, 2875–2883 (1985).

    Article  CAS  PubMed  Google Scholar 

  51. Subramanian, A. et al. Gene set enrichment analysis: a knowledge-based approach for interpreting genome-wide expression profiles. Proc. Natl Acad. Sci. USA 102, 15545–15550 (2005).

    CAS  PubMed  PubMed Central  ADS  Google Scholar 

  52. Mootha, V. K. et al. PGC-1α-responsive genes involved in oxidative phosphorylation are coordinately downregulated in human diabetes. Nat. Genet. 34, 267–273 (2003).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank C. J. Rosen for sharing the Alplfl/fl mouse strain with permission of J.L.M.; R. Garrity for help with CLAMS studies; the NMR Core jointly operated by Harvard Medical School and Dana-Farber Cancer Institute for help with NMR data acquisition; Nikon Imaging Center at Harvard Medical School for help with fluorescence imaging studies; the EM Core at Harvard Medical School for APEX2/EM imaging studies; and the Mass Spectrometry Facility at Beth Israel Deaconess Medical Center for targeted metabolomics studies. Y.S. and C.L.R. are supported by the American Heart Association postdoctoral fellowship. J.F.R. is supported by the Charlotte and Leo Karassik Fellowship. B.H. was a Cancer Research Institute/Leonard Kahn Foundation Fellow. P.A.D. is supported by a Damon Runyon Cancer Research Foundation Fellowship. This study was supported by NIDCR grant DE12889 to J.L.M., NIH grant DK 123095 to E.T.C., Canadian Institutes of Health Research (CIHR) grant PJT-159529 to L.K. and JPB Foundation 6293803 and NIH grant DK123228 to B.M.S.

Author information

Authors and Affiliations

Authors

Contributions

Y.S. and B.M.S. conceived the study and designed the experiments. Y.S. performed experiments and analysed the data. Y.S. and J.F.R. performed cellular and mitochondrial respiration experiments and analysed the data. M.P.J. performed mass spectrometry analysis. C.L.R. and S.V. helped with fluorescence imaging studies. S.V. assisted with protease protection assays. D.B. assisted with all experiments. B.H. assisted with animal experiments. C.L.R., S.V., P.A.D., X.Z., A.B.W. and N.H.K. helped with animal and/or cellular experiments and data analyses. C.R.K. assisted with CLAMS studies. A.M. assisted with activity measurements and animal experiments. J.L.M. contributed essential genetic and pharmacological reagents and discussed data. L.K. and E.T.C. contributed to experimental discussions. Y.S. and B.M.S. wrote the manuscript with comments from all authors. All authors provided input and reviewed the manuscript.

Corresponding author

Correspondence to Bruce M. Spiegelman.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature thanks the anonymous reviewers for their contribution to the peer review of this work.

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data figures and tables

Extended Data Fig. 1 PCr phosphatase activities of thermogenic fat and TNAP.

a, PCr phosphatase activities of total mitochondrial protein extracts from different tissues of cold-acclimated mice. Mitochondrial protein extract was prepared from tissues excised from 10 mice for BAT or 20 mice for iWAT. Each reaction contains 10 mM of PCr and 0.4 mg ml−1 of mitochondrial protein extract, except the buffer control. Data are presented as the estimated parameters ± uncertainties. Uncertainties are represented by the standard errors of nonlinear regression that fits a straight-line model to the initial linear phase of PCr hydrolysis kinetics measured by 31P NMR over 11 time points for BAT and iWAT and 6 time points for the buffer control (shown in Source Data). b, Ion-exchange chromatography of the active fraction of SEC. The PCr phosphatase activity of each fraction was measured by enzyme-coupled assay. The activity of the most active fraction was also verified by 31P NMR. Red and blue bars denote the fractions used for isobaric labelling (TMT) and quantitative mass spectrometric analysis. c, Western-blot analysis of the active SEC fraction prepared from cold-acclimated mice (cold), compared with the equivalent fraction prepared from room-temperature housed mice (RT). d, PCr phosphatase activities of total mitochondrial protein extracts from BAT of cold-acclimated mice treated with vehicle or SBI-425 (10 μM), measured by 31P NMR. n = 2 technical replicates per group. Data are presented as mean ± s.e.m. e, Stacked traces of 31P NMR spectra recorded at indicated time points, demonstrating the kinetics of PCr hydrolysis catalysed by recombinant TNAP. The minor peak marked with an asterisk on top is from glycerol-3-phosphate, a side-product of the phospho-transferase activity of TNAP50, that transfers the phosphoryl-group from PCr to glycerol present in the reaction buffer. f, Michaelis constant (Km) curves of hydrolysis of PCr (left) and PPi (right) catalysed by recombinant TNAP. Activities were measured by the enzyme-coupled assay; n = 2 technical replicates. Data are presented as mean ± s.e.m. g, Comparison of the Michaelis–Menten parameters extrapolated from f. Data are presented as the estimated parameters ± uncertainties. Uncertainties are represented by standard errors derived from the nonlinear regression fit of Michaelis–Menten model to the data in f.

Source data

Extended Data Fig. 2 Mitochondrial localization of ectopically expressed TNAP in non-thermogenic fat cell types.

Confocal fluorescence microscopic images showing subcellular localization of ectopically expressed TNAP in different cell types. PTEC, kidney proximal tubule epithelial cells. The insets show a magnified view of the area outlined by the dotted box. Anti-TNAP and anti-HSP60 were used to visualize TNAP and mitochondria, respectively. Scale bars, 10 μm.

Extended Data Fig. 3 Mitochondrial localization of endogenous TNAP in BAT and non-thermogenic fat cells.

a, Confocal fluorescence images showing subcellular localization of endogenous TNAP in brown adipocytes (top and middle panels) and hepatocytes (bottom panels). Primary brown preadipocytes were prepared from Alplfl/fl mice, transduced with either AdGFP (WT) or AdCre (Alpl KO) on day 4 of differentiation, and fixed for imaging on day 8. Arrows denote selected peri-nuclear areas of TNAP signal that co-localize with mitochondria signal. Antibodies for TNAP (red) and HSP60 (green) were used to visualize TNAP and mitochondria. Scale bar, 5 μm. b, Pearson’s correlation coefficient (PCC) analysis showing the extent of co-localization of TNAP with mitochondria in indicated cell types; n = 10 cells per group; data are presented as mean ± s.e.m.; statistical significance was calculated by one-way ANOVA with Bonferroni’s multiple comparisons test. c, Western-blot analysis on TNAP in wild-type versus knockout cells. Vinculin (VCL) blot was used as a sample preparation control. d, Confocal fluorescence microscopic images showing subcellular localization of endogenous TNAP in different cell types. Anti-TNAP and anti-HSP60 were used to visualize TNAP and mitochondria, respectively. Scale bars, 5 μm. e, Western-blot analysis on TNAP and mitochondrial markers in mitochondrial preparations from BAT of cold-acclimated, wild-type vs adipo-Alpl knockout mice. Blots were processed in parallel with samples derived from the same experiment. f, Western-blot analysis of the insoluble fraction of mitochondria extract treated with phosphatidylinositol-specific phospholipase-C (PI-PLC), followed by ultracentrifugation, showing that PLC treatment releases TNAP from membranes. P, pellet; S, supernatant. Mitochondrial preparation was fragmented by sonication before treatment. Blots were processed in parallel with samples derived from the same experiment.

Source data

Extended Data Fig. 4 Proximity-based fluorescent labelling by TNAP–APEX2 and trypsin protection assay on mitochondria from BAT.

a, Confocal fluorescence microscopic images of immortalized brown adipocytes showing co-localization of the GFP signal from 3XHA-EGFP-OMP25 construct (mGFP, channel: 488 nm) with different mitochondria markers. Endogenous antibodies, OxPhos (upper red, channel: 561 nm) and HSP60 (lower red, channel: 640 nm), were used to visualize mitochondria. The insets show a magnified region of the image outlined by the dotted box. Scale bars, 5 μm. b, Illustration of how APEX2 reports subcellular localization of TNAP by its peroxidase activity. X indicates either Alexa Fluor 647-conjugated tyramide (for confocal microscopy) or 3,3′-diaminobenzidine (for TEM studies). c, Confocal fluorescence analysis of immortalized brown adipocytes (top) and hepatocytes (bottom) ectopically expressing a TNAP–APEX2 construct. Cells were fixed and treated with Alexa Fluor 647–tyramide/H2O2 for proximity-based fluorescent labelling facilitated by the peroxidase activity of APEX2. Stably expressed 3XHA-EGFP-OMP25 was used as mitochondria reporter. Scale bars, 10 μm. d, Western blot analysis of the protease protection assay on mitochondria derived from BAT of cold-acclimated mice. TOMM20, GPD2 (glycerol-3-phosphate dehydrogenase), Cyt C (cytochrome c) and CS (citrate synthase) are shown as markers of outer mitochondrial membrane (OMM), intermembrane space (IMS) and mitochondrial matrix. Blots were processed in parallel with samples derived from the same experiment. e, Relative protein abundances in trypsin-digested mitochondria derived from band intensities of intact protein quantified from d; n = 2 technical replicates. Data are presented as mean ± s.e.m.

Source data

Extended Data Fig. 5 Effect of Alpl silencing on cellular respiration.

a, Quantitative PCR with reverse transcription (qRT–PCR) of differentiated primary brown preadipocytes treated with shLacZ or shAlpl; n = 3 biologically independent samples per group. b, Western-blot analysis on TNAP in cells treated with shLacZ or shAlpl. Vinculin (VCL) blot was used as a sample processing control. c, Effect of Alpl knockdown (adenoviral shAlpl) on the oxygen consumption rate (OCR) of primary brown adipocytes. Treatments to initiate different respiration states are as follows: stimulated, noradrenaline; uncoupled, oligomycin; maximum, carbonyl cyanide m-chlorophenyl hydrazone; n = 11 biologically independent samples for the shLacZ group and 18 biologically independent samples for the shAlpl group. Data are presented as mean ± s.e.m. Statistical significance was calculated by unpaired Student’s two-sided t-test.

Source data

Extended Data Fig. 6 Effect of TNAP inhibition on the futile creatine cycle.

a, Effect of SBI-425 treatment (10 μM) on the OCR of beige fat-derived mitochondria from wild-type vs adipo-Alpl knockout mice in the presence of 0.01 mM creatine and 0.1 mM ADP (limiting ADP) or 1 mM ADP (saturating ADP), as measured by a Seahorse XF24 Extracellular Flux Analyzer; n = 7 independent measurements per group. b, Effect of SBI-425 treatment (10 μM) on OCR of beige fat-derived mitochondria in the presence of 0.1 mM ADP (limiting ADP) or 1 mM ADP (saturating ADP), but in the absence of creatine, as measured by a Seahorse XF24 Extracellular Flux Analyzer; n = 6 independent measurements per group. Data are presented as mean ± s.e.m. Statistical significance was calculated by either two-way ANOVA with Bonferroni’s multiple comparisons test (a) or unpaired Student’s two-sided t-test (b).

Source data

Extended Data Fig. 7 Rapid mitochondria purification enriched mitochondrial metabolites and proteins.

a, Relative abundances of citric acid cycle intermediates in mitochondria vs whole-cell metabolomics; n = 6 biologically independent samples. Data are presented as mean ± s.e.m. b, Western-blot analysis on mitochondrial markers in immunoprecipitated mitochondria for metabolomics study. Blots were processed in parallel with samples derived from the same experiment.

Source data

Extended Data Fig. 8 Movement and food intake of mice upon SBI-425 treatment.

a, b, Cumulative movement and food intake of wild-type mice (a) and adipo-Alpl knockout mice (b) for 24 h after treatment of SBI-425 versus vehicle; n = 10 mice for wild-type and 4 mice for adipo-Alpl knockout. Data are presented as mean ± s.e.m. Statistical significance was calculated using an unpaired Student’s two-sided t-test.

Source data

Extended Data Fig. 9 Energy expenditure and movement of wild-type and adipo-Alpl knockout mice on a HFD.

a, Indirect calorimetric measurements taken for wild-type and adipo-Alpl knockout mice that had been on a HFD for 4 weeks at 22 °C; n = 7 mice for wild-type and 6 mice for adipo-Alpl knockout; the grey area indicates the dark period. b. Averaged respiration rates over 24 h as measured in a; n = 7 mice for wild-type and 6 mice for adipo-Alpl knockout. c, Cumulative movement of mice over 24 h; n = 7 mice for wild-type and 6 mice for adipo-Alpl knockout. Data are presented as mean ± s.e.m. Statistical significance was calculated by either two-way ANOVA (a) or unpaired Student’s two-sided t-test (b, c).

Source data

Extended Data Fig. 10 Compensatory thermogenesis in adipo-Alpl knockout mice.

a, Indirect calorimetric measurement taken for wild-type and adipo-Alpl knockout mice kept in metabolic cages at 22 °C, showing stimulation of respiration by CL 316,243 administration (1.0 mg kg−1). Arrow denotes the time point of drug administration; n = 12 mice for wild-type and 9 mice for adipo-Alpl knockout; all mice were pre-treated with CL 316,243 (1.0 mg kg−1 day−1) for 5 days. b, Western-blot analysis of BAT and iWAT from wild-type or adipo-Alpl knockout mice pre-treated with CL 316,243 (1.0 mg kg−1 day−1) for 5 days. c, d, Gene set enrichment plot of quantitative mass spectrometric analyses of BAT (c) and iWAT (d) from wild-type or adipo-Alpl knockout mice treated with CL 316,243 (1.0 mg kg−1 day−1) for 5 days; n = 4 mice for wild-type and 6 mice for adipo-Alpl knockout. Enrichment analysis was performed with GSEA 4.1.051,52. The hallmark gene sets were surveyed, and oxidative phosphorylation is the top hit for both BAT and iWAT. Family-wise error rate P value was presented for statistical significance (number of permutations, 1,000); NES, normalized enrichment score (enrichment statistics, ‘classic’). e, f, qRT–PCR of BAT (e) and iWAT (f) from wild-type and adipo-Alpl knockout mice treated with CL 316,243 (1.0 mg kg−1 day−1) for 5 days; n = 4 mice for wild-type and 6 mice for adipo-Alpl knockout. Data are presented as mean ± s.e.m. Statistical significance was calculated by either two-way ANOVA (a) or unpaired Student’s two-sided t-test (e, f).

Source data

Supplementary information

Supplementary Figure 1

This file contains the uncropped images of western blots represented in Extended Data Figure 1c, 3c, 3e-f, 4d, 5b, 7b, and 10b. Red boxes indicate regions displayed in the final figures.

Reporting Summary

Supplementary Table 1

This file contains the mitochondrial metabolomics data of brown adipocytes treated with vehicle vs SBI-425 (10 μM) for 1 hour.

Supplementary Table 2

This file contains the proteomics data of BAT and iWAT from WT vs Adipo-Alpl KO mice housed at room-temperature and pre-treated with CL 316,243 (1.0 mg/kg/day) for 5 days.

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sun, Y., Rahbani, J.F., Jedrychowski, M.P. et al. Mitochondrial TNAP controls thermogenesis by hydrolysis of phosphocreatine. Nature 593, 580–585 (2021). https://doi.org/10.1038/s41586-021-03533-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-021-03533-z

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

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing