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:

Microbial metabolite delta-valerobetaine is a diet-dependent obesogen

Abstract

Obesity and obesity-related metabolic disorders are linked to the intestinal microbiome. However, the causality of changes in the microbiome–host interaction affecting energy metabolism remains controversial. Here, we show the microbiome-derived metabolite δ-valerobetaine (VB) is a diet-dependent obesogen that is increased with phenotypic obesity and is correlated with visceral adipose tissue mass in humans. VB is absent in germ-free mice and their mitochondria but present in ex-germ-free conventionalized mice and their mitochondria. Mechanistic studies in vivo and in vitro show VB is produced by diverse bacterial species and inhibits mitochondrial fatty acid oxidation through decreasing cellular carnitine and mitochondrial long-chain acyl-coenzyme As. VB administration to germ-free and conventional mice increases visceral fat mass and exacerbates hepatic steatosis with a western diet but not control diet. Thus, VB provides a molecular target to understand and potentially manage microbiome–host symbiosis or dysbiosis in diet-dependent obesity.

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: δ-VB is a microbiome-derived mitochondrial metabolite.
Fig. 2: δ-VB is a microbiome-derived metabolite.
Fig. 3: VB decreases fatty acid oxidation.
Fig. 4: VB increases lipid accumulation in host tissues.
Fig. 5: Long-term VB treatment in GF and conventional mice.
Fig. 6: VB participates in microbiome–mitochondria communication to reprogramme host lipid metabolism.
Fig. 7: Clinical associations of plasma VB with microbiome manipulation and obesity-related phenotypes in people.

Similar content being viewed by others

Data availability

RNA-seq data are available at the NCBI Gene Expression Omnibus (GEO) GSE185525 and GSE145012. Source data are provided with this paper. Other data that support the findings of this study are available on request from the corresponding authors.

References

  1. Miranda, J. J. et al. Understanding the rise of cardiometabolic diseases in low- and middle-income countries. Nat. Med. 25, 1667–1679 (2019).

    CAS  PubMed  Google Scholar 

  2. Wang, Y. et al. Has the prevalence of overweight, obesity and central obesity levelled off in the United States? Trends, patterns, disparities, and future projections for the obesity epidemic. Int. J. Epidemiol. 49, 810–823 (2020).

    PubMed  PubMed Central  Google Scholar 

  3. Carding, S., Verbeke, K., Vipond, D. T., Corfe, B. M. & Owen, L. J. Dysbiosis of the gut microbiota in disease. Microb. Ecol. Health Dis. 26, 26191 (2015).

    PubMed  Google Scholar 

  4. Rosenbaum, M., Knight, R. & Leibel, R. L. The gut microbiota in human energy homeostasis and obesity. Trends Endocrinol. Metab. 26, 493–501 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  5. Ignacio, A. et al. Correlation between body mass index and faecal microbiota from children. Clin. Microbiol Infect. 22, 258 e251–e258 (2016).

    Google Scholar 

  6. Gonzalez, F. J., Jiang, C. & Patterson, A. D. An intestinal microbiota-farnesoid x receptor axis modulates metabolic disease. Gastroenterology 151, 845–859 (2016).

    CAS  PubMed  Google Scholar 

  7. San-Cristobal, R., Navas-Carretero, S., Martinez-Gonzalez, M. A., Ordovas, J. M. & Martinez, J. A. Contribution of macronutrients to obesity: implications for precision nutrition. Nat. Rev. Endocrinol. 16, 305–320 (2020).

    PubMed  Google Scholar 

  8. Holmes, E., Li, J. V., Athanasiou, T., Ashrafian, H. & Nicholson, J. K. Understanding the role of gut microbiome-host metabolic signal disruption in health and disease. Trends Microbiol. 19, 349–359 (2011).

    CAS  PubMed  Google Scholar 

  9. Spanogiannopoulos, P., Bess, E. N., Carmody, R. N. & Turnbaugh, P. J. The microbial pharmacists within us: a metagenomic view of xenobiotic metabolism. Nat. Rev. Microbiol. 14, 273–287 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Cani, P. D. et al. Metabolic endotoxemia initiates obesity and insulin resistance. Diabetes 56, 1761–1772 (2007).

    CAS  PubMed  Google Scholar 

  11. Cani, P. D. et al. Changes in gut microbiota control metabolic endotoxemia-induced inflammation in high-fat diet-induced obesity and diabetes in mice. Diabetes 57, 1470–1481 (2008).

    CAS  PubMed  Google Scholar 

  12. Hoyles, L. et al. Molecular phenomics and metagenomics of hepatic steatosis in non-diabetic obese women. Nat. Med. 24, 1070–1080 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Wang, Z. et al. Gut flora metabolism of phosphatidylcholine promotes cardiovascular disease. Nature 472, 57–63 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Koeth, R. A. et al. Intestinal microbiota metabolism of L-carnitine, a nutrient in red meat, promotes atherosclerosis. Nat. Med. 19, 576–585 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Wikoff, W. R. A. et al. Metabolomics analysis reveals large effects of gut microflora on mammalian blood metabolites. Proc. Natl Acad. Sci. USA 106, 3698–3703 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Peisl, B. Y. L., Schymanski, E. L. & Wilmes, P. Dark matter in host-microbiome metabolomics: tackling the unknowns-a review. Anal. Chim. Acta 1037, 13–27 (2018).

    CAS  PubMed  Google Scholar 

  17. Goo, E., An, J. H., Kang, Y. & Hwang, I. Control of bacterial metabolism by quorum sensing. Trends Microbiol. 23, 567–576 (2015).

    CAS  PubMed  Google Scholar 

  18. Bajpai, P., Darra, A. & Agrawal, A. Microbe-mitochondrion crosstalk and health: an emerging paradigm. Mitochondrion 39, 20–25 (2018).

    CAS  PubMed  Google Scholar 

  19. Donohoe, D. R. et al. The microbiome and butyrate regulate energy metabolism and autophagy in the mammalian colon. Cell Metab. 13, 517–526 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Mollica, M. P. et al. Butyrate regulates liver mitochondrial function, efficiency, and dynamics in insulin-resistant obese mice. Diabetes 66, 1405–1418 (2017).

    CAS  PubMed  Google Scholar 

  21. Yardeni, T. et al. Host mitochondria influence gut microbiome diversity: a role for ROS. Sci. Signal. 12, eaaw3159 (2019).

    PubMed  Google Scholar 

  22. Bäckhed, F., Manchester, J. K., Semenkovich, C. F. & Gordon, J. I. Mechanisms underlying the resistance to diet-induced obesity in germ-free mice. Proc. Natl Acad. Sci. USA 104, 979–984 (2007).

    PubMed  PubMed Central  Google Scholar 

  23. Bellissimo, M. P. et al. Plasma high-resolution metabolomics differentiates adults with normal weight. Obes. Lean Individ. Obes. 27, 1729–1737 (2019).

    CAS  Google Scholar 

  24. Cioffi, C. & Vos, M. B. Su1506—comparison of plasma metabolomics profiles of pediatric NASH vs. NAFLD. Gastroenterology 154, S-1161 (2018).

    Google Scholar 

  25. Mølstad, P. The efflux of l-carnitine from cells in culture (CCL27). Biochim. Biophys. Acta 597, 166–173 (1980).

    PubMed  Google Scholar 

  26. Nagata, N. et al. Effects of bowel preparation on the human gut microbiome and metabolome. Sci. Rep. 9, 4042 (2019).

    PubMed  PubMed Central  Google Scholar 

  27. Ringseis, R., Keller, J. & Eder, K. Role of carnitine in the regulation of glucose homeostasis and insulin sensitivity: evidence from in vivo and in vitro studies with carnitine supplementation and carnitine deficiency. Eur. J. Nutr. 51, 1–18 (2012).

    CAS  PubMed  Google Scholar 

  28. Fritz, I. B. Action of carnitine on long chain fatty acid oxidation by liver. Am. J. Physiol. 197, 297–304 (1959).

    CAS  PubMed  Google Scholar 

  29. Noland, R. C. et al. Carnitine insufficiency caused by aging and overnutrition compromises mitochondrial performance and metabolic control. J. Biol. Chem. 284, 22840–22852 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Bianchi, P. B. L., Davis, D. C. & Carnitine, A. T. Supplementation ameliorates the steatosis and ketosis induced by pivalate in rats. J. Nutr. 126, 2873–2879 (1996).

    CAS  PubMed  Google Scholar 

  31. Asai, T. et al. Combined therapy with PPARalpha agonist and l-carnitine rescues lipotoxic cardiomyopathy due to systemic carnitine deficiency. Cardiovasc. Res. 70, 566–577 (2006).

    CAS  PubMed  Google Scholar 

  32. Jiang, F. et al. L-carnitine ameliorates the liver inflammatory response by regulating carnitine palmitoyltransferase I-dependent PPARγ signaling. Mol. Med. Rep. 13, 1320–1328 (2016).

    CAS  PubMed  Google Scholar 

  33. Libert, D. M., Nowacki, A. S. & Natowicz, M. R. Metabolomic analysis of obesity, metabolic syndrome, and type 2 diabetes: amino acid and acylcarnitine levels change along a spectrum of metabolic wellness. PeerJ 6, https://doi.org/10.7717/peerj.5410 (2018).

  34. Koves, T. R. et al. Mitochondrial overload and incomplete fatty acid oxidation contribute to skeletal muscle insulin resistance. Cell Metab. 7, 45–56 (2008).

    CAS  PubMed  Google Scholar 

  35. Spaniol, M. et al. Mechanisms of liver steatosis in rats with systemic carnitine deficiency due to treatment with trimethylhydraziniumpropionate. J. Lipid Res. 44, 144–153 (2003).

    CAS  PubMed  Google Scholar 

  36. Karkkainen, O. et al. Diets rich in whole grains increase betainized compounds associated with glucose metabolism. Am. J. Clin. Nutr. 108, 971–979 (2018).

    PubMed  Google Scholar 

  37. Servillo, L. et al. Ruminant meat and milk contain delta-valerobetaine, another precursor of trimethylamine N-oxide (TMAO) like gamma-butyrobetaine. Food Chem. 260, 193–199 (2018).

    CAS  PubMed  Google Scholar 

  38. Servillo, L. et al. Carnitine precursors and short-chain acylcarnitines in water buffalo milk. J. Agric. Food Chem. 66, 8142–8149 (2018).

    CAS  PubMed  Google Scholar 

  39. Ye, E. Q., Chacko, S. A., Chou, E. L., Kugizaki, M. & Liu, S. Greater whole-grain intake is associated with lower risk of type 2 diabetes, cardiovascular disease, and weight gain. J. Nutr. 142, 1304–1313 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Wang, Y. & Beydoun, M. A. Meat consumption is associated with obesity and central obesity among US adults. Int J. Obes. 33, 621–628 (2009).

    CAS  Google Scholar 

  41. Li, X. S. et al. Untargeted metabolomics identifies trimethyllysine, a TMAO-producing nutrient precursor, as a predictor of incident cardiovascular disease risk. JCI Insight 3, https://doi.org/10.1172/jci.insight.99096 (2018).

  42. Koeth, R. A. et al. gamma-Butyrobetaine is a proatherogenic intermediate in gut microbial metabolism of l-carnitine to TMAO. Cell Metab. 20, 799–812 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Cho, I. et al. Antibiotics in early life alter the murine colonic microbiome and adiposity. Nature 488, 621–626 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Cox, L. M. Antibiotics shape microbiota and weight gain across the animal kingdom. Anim. Front. 6, 8–14 (2016).

    Google Scholar 

  45. Stark, C. M., Susi, A., Emerick, J. & Nylund, C. M. Antibiotic and acid-suppression medications during early childhood are associated with obesity. Gut 68, 62–69 (2019).

    CAS  PubMed  Google Scholar 

  46. Peters, B. A. et al. A taxonomic signature of obesity in a large study of American adults. Sci. Rep. 8, 9749 (2018).

    PubMed  PubMed Central  Google Scholar 

  47. Zhao, M. et al. TMAVA, a metabolite of intestinal microbes, is increased in plasma from patients with liver steatosis, inhibits gamma-butyrobetaine hydroxylase, and exacerbates fatty liver in mice. Gastroenterology 158, https://doi.org/10.1053/j.gastro.2020.02.033 (2020).

  48. Sagan, L. On the origin of mitosing cells. J. Theor. Biol. 14, 225–274 (1967).

    CAS  Google Scholar 

  49. Ristow, M. Unraveling the truth about antioxidants: mitohormesis explains ROS-induced health benefits. Nat. Med. 20, 709–711 (2014).

    CAS  PubMed  Google Scholar 

  50. Khan, S. A. et al. ATGL-catalyzed lipolysis regulates SIRT1 to control PGC-1α/PPAR-α signaling. Diabetes 64, 418 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Montagner, A. et al. Liver PPARalpha is crucial for whole-body fatty acid homeostasis and is protective against NAFLD. Gut 65, 1202–1214 (2016).

    CAS  PubMed  Google Scholar 

  52. Azevedo, R. A. & Arruda, P. High-lysine maize: the key discoveries that have made it possible. Amino Acids 39, 979–989 (2010).

    CAS  PubMed  Google Scholar 

  53. Torre, D., Lachmann, A. & Ma’ayan, A. BioJupies: automated generation of interactive notebooks for RNA-seq data analysis in the cloud. Cell Syst. 7, 556–561.e553 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Go, Y. M. et al. Mitochondrial metabolomics using high-resolution Fourier-transform mass spectrometry. Methods Mol. Biol. 1198, 43–73 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Liu, K. H. et al. High-resolution metabolomics assessment of military personnel: evaluating analytical strategies for chemical detection. J. Occup. Environ. Med. 58, S53–S61 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Liu, K. H. et al. Reference standardization for quantification and harmonization of large-scale metabolomics. Anal. Chem. https://doi.org/10.1021/acs.analchem.0c00338 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  57. Yu, T. & Jones, D. P. Improving peak detection in high-resolution LC/MS metabolomics data using preexisting knowledge and machine learning approach. Bioinformatics 30, 2941–2948 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  58. Yu, T., Park, Y., Johnson, J. M. & Jones, D. P. apLCMS–adaptive processing of high-resolution LC/MS data. Bioinformatics 25, 1930–1936 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Uppal, K. et al. xMSanalyzer: automated pipeline for improved feature detection and downstream analysis of large-scale, non-targeted metabolomics data. BMC Bioinf. 14, 15 (2013).

    Google Scholar 

  60. Ritchie, M. E. et al. limma powers differential expression analyses for RNA-sequencing and microarray studies. Nucleic Acids Res. 43, e47–e47 (2015).

    PubMed  PubMed Central  Google Scholar 

  61. Li, S. et al. Predicting network activity from high throughput metabolomics. PLoS Comput. Biol. 9, e1003123 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  62. Uppal, K., Walker, D. I. & Jones, D. P. xMSannotator: an R package for network-based annotation of high-resolution metabolomics data. Anal. Chem. 89, 1063–1067 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Smith, C. A. et al. METLIN: a metabolite mass spectral database. Ther. Drug Monit. 27, 747–751 (2005).

    CAS  PubMed  Google Scholar 

  64. Tars, K. et al. Targeting carnitine biosynthesis: discovery of new inhibitors against γ-butyrobetaine hydroxylase. J. Med. Chem. 57, 2213–2236 (2014).

    CAS  PubMed  Google Scholar 

  65. Chen, C. M. F. & Leo Benoiton, N. A new method of quaternizing amines and its use in amino acid and peptide chemistry. Can. J. Chem. 54, 3310 (1976).

    CAS  Google Scholar 

  66. R Core Team. R: A Language Environment for Statistical Computing (R Foundation for Statistical Computing, 2014); https://www.R-project.org/

  67. Rask, K. J., Brigham, K. l., Fau-Johns, M. M. E. & Johns, M. M. Integrating comparative effectiveness research programs into predictive health: a unique role for academic health centers. Acad. Med. 86, 718–723 (2011).

    PubMed  Google Scholar 

  68. Brigham, K. L. Predictive health: the imminent revolution in health care. J. Am. Geriatr Soc. 58, S298–S302 (2010).

    PubMed  PubMed Central  Google Scholar 

  69. Soltow, Q. A. et al. High-performance metabolic profiling with dual chromatography-Fourier-transform mass spectrometry (DC-FTMS) for study of the exposome. Metabolomics 9, 132–143 (2011).

    Google Scholar 

  70. Kaul, S. et al. Dual-energy X-ray absorptiometry for quantification of visceral fat. Obesity (Silver Spring) 20, 1313–1318 (2012).

    Google Scholar 

Download references

Acknowledgements

This work was supported by grant nos. T32 GM008602 (K.H.L.), P30 ES019776 (D.P.J.), R01 ES023485 (D.P.J., Y.M.G.), U2C ES030163 (D.P.J.), R2C DK118619 (D.P.J.), S10 OD018006 (D.P.J.), K01 DK102851 (J.A.A.), R03 DK117246 (J.A.A.), K24 DK096574 (T.R.Z., J.A.A.), R21HD089056 (M.B.V.), R00AA021803 (S.M.Y.), R01AA026086 (S.M.Y.) and R01AI064462 (A.S.N.). Additionally, we acknowledge support from the Georgia Clinical and Translational Science Alliance grant no. UL1 TR002378 and the Emory-Georgia Tech Predictive Health Institute and Center for Health Discovery and Well-Being. The funders had no role in study design, data collection and analysis, decision to publish or preparation of the paper.

Author information

Authors and Affiliations

Authors

Contributions

K.H.L., J.A.O., B.S., D.P.J. and A.S.N. conceived the work. K.H.L., B.S., D.P.J., A.S.N., J.A.O., C.N., Y.-M.G., M.H.W., C.S.K., S.M.Y. and E.O. designed experiments. K.H.L., B.S., J.A.O., C.E.C., M.P.B., C.N., T.D., S.D., K.M.-S., M.O., X.H., J.F., M.C.C., S.H.-C., T.G., T.R.Z., M.B.V., J.A.A., D.v.I., M.H.W., C.S.K., C.M., R.M.J. and K.U. performed experiments, acquired or analysed data for this study. K.H.L., B.S., J.A.O., Y.-M.G., D.P.J. and A.S.N. prepared the paper.

Corresponding authors

Correspondence to Andrew S. Neish or Dean P. Jones.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature Metabolism thanks the anonymous reviewer(s) for their contribution to the peer review of this work. Primary Handling Editors: Ashley Castellanos-Jankiewicz and Pooja Jha.

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

Extended data

Extended Data Fig. 1 Additional mass spectrometry analysis of VB.

A Mass spectrometric analysis of VB in a) lung and brain of conventional (C) and germ-free (GF) mice and b) Ion dissociation spectra (MS/MS) analysis of 160.1332 m/z (hcd 30) in caecal contents and mouse diet. VB was not present in GF cecal contents, or the control chow (Teklad) diet used for GF mouse experiments. The fragmentation of 160.1332 m/z in GF cecal contents and GF control diet is consistent with valine betaine (Supplementary Information). c) Ion fragmentation spectra of 160.1332 m/z in GF control diet and extracted ion chromatograms of 160.1332 m/z in all diets used in this study. There was no peak for 160.1332 m/z observed in the Western diet. c) Extracted ion chromatograms of 160.1332 m/z from mouse diets used for this study. VB is not present in the sterilized chow (Teklad) used as the control diet for GF and conventionalization experiments. The conventional chow (Labdiet), which was not autoclavable, was used as the control for conventional mouse experiments. This diet contained a minor peak for VB (red brackets), but the later eluting major peak (0.8m) had a fragmentation pattern observed for 160.1332 m/z (valine betaine) in the GF diet. The small amount of VB present in the Labdiet chow is not the major source of VB in conventional mice since serum VB was equivalent between WD-fed (which does not contain VB) and Labdiet chow-fed conventional mice (Extended data 6c). d) (Nε, Nε, Nε)-trimethyllysine (TML) is a precursor to δ-valerobetaine (VB). (Nε, Nε, Nε)-trimethyllysine (100 mg/kg) or (13C Nε, 13C Nε, 13C Nε)-trimethyllysine (25 mg/kg) was gavaged into conventional mice for 3 days. The formation of unlabelled VB (160.1332 m/z) from TML and labeled VB (163.1431 m/z) from isotopically labeled TML is shown here along with accurate mass MS1 and ion fragmentation spectra consistent with VB and labeled VB.

Extended Data Fig. 2 VB inhibits mitochondrial fatty acid oxidation.

a) Tracing the oxidation of 13C16 palmitic acid in HepG2 cells to examine the effect of VB treatment on cellular mitochondrial fatty acid oxidation. 12 hour pretreatment with VB (green) decreased (p = 0.0015 (AUC - Area under the curve), two-tailed student’s t-test) the formation of labeled acetyl-CoA (bottom middle) by approximately 75% compared to vehicle (blue). Addition of carnitine back to cells pretreated with VB for 12 hours (purple) restored the carnitine-dependent formation of mitochondrial acetyl-CoA (p = 0.49 (AUC), two-tailed student’s t-test). Co-treatment of VB with the addition of stable isotope-labeled palmitate (red) decreased the formation of labeled acetyl-CoA by approximately 25% compared to vehicle (p = 0.04 (AUC), two-tailed student’s t-test). Data for VB, carnitine, and other metabolites are shown to illustrate VB treatment does not affect uptake of labeled 13C16 palmitate (middle left), or the conjugation of labeled palmitate to CoA (middle middle). VB treatment decreased carnitine approximately 20% after one hour (top middle) and these changes drive decreased formation of labeled palmitoylcarnitine (middle right), labeled acetyl-CoA (bottom middle), labeled acetylcarnitine (bottom left), and labeled citrate (bottom right). Each data point represents the average of 3 biological replicates ± standard deviation. b) The effects of VB on mitochondrial respiration is dependent on the availability of fuel substrates. In the presence of glucose and glutamine, VB does not decrease basal respiration (ANOVA (F = 6.239, p = 0.0014) with post-hoc test for linear trend (R2 = 0.0078, p = 0.50)) or maximum respiratory capacity after addition of FCCP (ANOVA (F = 2.223, p = 0.10). VB does not influence basal respiration with culture conditions with palmitate but without glucose and glutamine (ANOVA (F = 9.729, p = 0.001) with post-hoc test for linear trend (R2 = 0.075, p < 0.096). Each data point represents at least 8 biological replicates.

Source data

Extended Data Fig. 3 VB alters carnitine metabolism in mice.

VB treatment increased a) circulating and b) hepatic VB in conventional mice [Kruskal-Wallis analysis to determine group-wise differences with Dunn’s multiple comparisons test (serum VB males (n = 3 per treatment) – KW statistic 7.2, p = 0.0036; serum VB females (Vehicle n = 5, 10 mg/kg n = 6, 100 mg/kg n = 6) – KW statistic 10.84, p = 0.0008; liver VB males – KW statistic 6.489, p = 0.0107; liver VB females – KW statistic 12.94, p < 0.0001)]. c, d) Pathway enrichment analysis of serum metabolites correlated with VB treatment shows VB alters carnitine shuttle metabolism in male and female mice (p < 0.05, mummichog p-value based on permutation analysis). e) VB decreases circulating carnitines in mice. Correlation heatmaps are based on Spearman’s correlation values with red corresponding to Spearman’s rho = 1 and blue Spearmen’s rho = -1. f) VB treatment increases urinary carnitine in mice (n = 3, time 0 p = 0.126, time 2h p = 0.001, time 6h p = 5.86e-005, two-tailed student’s t-test). g) Fasting does not change circulating (n = 6, p = 0.95, two-tailed student’s t-test) or hepatic (n = 5, p = 0.24, two-tailed student’s t-test) VB compared to fed mice.

Source data

Extended Data Fig. 4 VB decreases circulating and hepatic beta-hydroxybutyrate.

VB decreases circulating and hepatic beta-hydroxybutyrate, produced from mitochondrial fatty acid oxidation during fasting. Kruskal-Wallis with Dunn’s multiple comparisons test was used for fasted serum and liver analyses (n = 8 vehicle, n = 3 10 mg/kg, n = 5 100 mg/kg for male and female). Male serum KW statistic 7.864, p = 0.0085; female serum KW statistic 10.46, p = 0.0006. Male liver KW statistic 7.864, p = 0.0085; female liver KW statistic 11.73, p = 0.0001. Control vs. 100 mg/kg was significantly different for all comparisons in fasted mice. For fed mice, (n = 3 vehicle, n = 3 10 mg/kg, n = 3 100 mg/kg for male and female), one-way ANOVA with Dunnett’s multiple comparison tests was used. Male serum (ANOVA F = 0.9472, p = 0.4390); Female serum (ANOVA F = 6.8, p = 0.028, vehicle vs. 100 mg/kg significant); Male liver (ANOVA F = 0.9032, p = 0.454); Female liver (ANOVA F = 4.295, p = 0.0695).

Extended Data Fig. 5 Untargeted lipidomics of VB treated mice.

VB alters neutral lipid profiles liver, heart, and brain of male and female mice. Neutral lipids from untargeted lipidomic profiling with average fold-change greater than 2 in 100 mg/kg VB-treated mice (n = 5 male, n = 5 female) versus control (n = 5 male, n = 5 female).

Source data

Extended Data Fig. 6 Baseline data for long-term mouse experiments with VB.

a) Comparison of VB in GF mouse serum and liver after treatment with 25 mg/kg VB with conventional mice on respective control diets. Samples were analyzed after 6 week treatment with VB and normal conventional mice. GF mice control diet was Teklad 2019S chow and conventional mice control diet was LabDiets 5001 chow. Serum GF+VB (n = 5) and conventional (n = 5) comparison (p = 0.01, two-tailed t-test with conventional mice having approximately 1.7x circulating VB compared to GF mice treated with VB. Liver GF+VB (n = 5) and conventional (n = 5) comparison (p = 0.0004, two-tailed t-test with conventional mice having approximately 0.74x liver VB compared to GF mice treated with VB. b) Comparison of carnitine in mouse serum and liver after treatment with 25 mg/kg in GF mice with conventional mouse (control diet) data for comparison. Carnitine is decreased in conventional mice compared to GF mice. VB treatment to GF mice led to serum and liver carnitine concentrations equivalent to conventional mice. (Two-tailed unpaired student t-tests: Serum GFCD vs. GFCD+VB p = 0.0001, GFCD vs. Conventional CD p < 0.0001, GFCD+VB vs. Conventional CD p = 0.395; Liver GFCD vs. GFCD+VB p < 0.0001, GFCD vs. Conventional CD p < 0.0001, GFCD+VB vs. Conventional CD p = 0.6559) c) Comparison of circulating VB in conventional mice (male M, female F) between Western diet (WD) and control chow (Two-tailed unpaired student t-tests: CD male vs. WD male p = 0.4982; CD female vs. WD female p = 0.1653).

Source data

Extended Data Fig. 7 Effects of 8-week VB treatment on weight gain and adipose tissue mass in control or Western diet in conventional female mice.

Effects of 8-week VB treatment on weight gain and adipose tissue mass in control or Western diet in conventional female mice (n = 5 per treatment). The combination of Western Diet with VB led to approximately a 3-6% increase in body weight compared to Western Diet alone in female mice, however these results were not statistically significant at a p-value threshold of 0.05 (p = 0.14, one-tailed t-test). In control-diet fed female conventional mice (n = 5 per treatment), VB did not increase weight gain (p-value for increase in weight = 0.9803 (1 – 0.0197), one-tailed t-test). VB treatment increased perigonadal visceral adipose tissue (VAT) and posterior subcutaneous adipose (SubQ) tissue mass in conventional female mice fed a Western diet. VB treatment increased interscapular brown adipose tissue (BAT) mass on a control diet but did not increase BAT mass with the Western diet in conventional female mice. One-tailed t-tests with p < 0.05 used to test for an increase in adipose tissue mass following VB treatment.

Source data

Extended Data Fig. 8 Cytokine panel following long-term VB treatment to male and female conventional mice.

a) Plasma biomarkers of glucose tolerance/insulin resistance (insulin, glucagon, resistin) in conventional mice after 8 weeks of VB treatment (n = 5 per treatment, outliers removed by Robust regression and Outlier removal (ROUT) in Prism 6.0). (One-way ANOVA with Sidak’s multiple comparisons: Insulin male F = 1.648, p = 0.2205; Insulin female F = 1.034, p = 0.4059; Glucagon male F = 0.5165, p = 0.6772; Glucagon female F = 1.723, p = 0.2025; Resistin male F = 4.747, p = 0.0149 [CD vs. CDVB not significant, WD vs. WDVB not significant]; Resistin female F = 2.687, p = 0.0814) b) Plasma biomarkers of inflammation (IL-6, TNF-alpha, MCP-1) in conventional mice after 8 weeks of VB treatment. (One-way ANOVA with Sidak’s multiple comparisons: IL-6 male F = 0.9316, p = 0.4497; IL-6 female F = 1.646, p = 0.2185; TNF-alpha male F = 1.645, p = 0.2186; TNF-alpha female F = 3.241, p = 0.0499 [CD vs. CDVB not significant, WD vs. WDVB not significant]; MCP-1 male F = 2.467, p = 0.1021; MCP-1 female F = 2.048, p = 0.1477).

Source data

Extended Data Fig. 9 Transcription factor analysis for Western Diet-fed GF mice.

Transcription factor enrichment analysis shows that Ppara target genes are upregulated by a) microbiome acquisition and VB treatment in control diet fed GF mice. Ppara target genes are downregulated by VB treatment in Western diet (WD) fed GF mice. b) VB treatment in WD fed mice downregulates genes linked to Ppara which function in mitochondria and lipid processing pathways (p-value < 0.005, FC < 0.8).

Source data

Extended Data Fig. 10 Urinary carnitine correlations with urinary VB.

Urinary carnitine is correlated with urinary VB in humans (n = 143, Pearson’s R = 0.754, p < 0.00001).

Source data

Supplementary information

Supplementary Information

Supplementary Figs. 1–3.

Reporting Summary

Supplementary Table 1

Diet composition for diets used in this study.

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 Fig. 7

Source data for Fig. 7.

Source Data Extended Data Fig. 2

Source data for Extended Data 2.

Source Data Extended Data Fig. 3

Source data for Extended Data 3.

Source Data Extended Data Fig. 5

Source data for Extended Data 5.

Source Data Extended Data Fig. 6

Source data for Extended Data 6.

Source Data Extended Data Fig. 7

Source data for Extended Data 7.

Source Data Extended Data Fig. 8

Source data for Extended Data 8.

Source Data Extended Data Fig. 9

Source data for Extended Data 9.

Source Data Extended Data Fig. 10

Source data for Extended Data 10.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Liu, K.H., Owens, J.A., Saeedi, B. et al. Microbial metabolite delta-valerobetaine is a diet-dependent obesogen. Nat Metab 3, 1694–1705 (2021). https://doi.org/10.1038/s42255-021-00502-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s42255-021-00502-8

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