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:

Ifnar gene variants influence gut microbial production of palmitoleic acid and host immune responses to tuberculosis

Abstract

Both host genetics and the gut microbiome have important effects on human health, yet how host genetics regulates gut bacteria and further determines disease susceptibility remains unclear. Here, we find that the gut microbiome pattern of participants with active tuberculosis is characterized by a reduction of core species found across healthy individuals, particularly Akkermansia muciniphila. Oral treatment of A. muciniphila or A. muciniphila-mediated palmitoleic acid strongly inhibits tuberculosis infection through epigenetic inhibition of tumour necrosis factor in mice infected with Mycobacterium tuberculosis. We use three independent cohorts comprising 6,512 individuals and identify that the single-nucleotide polymorphism rs2257167 ‘G’ allele of type I interferon receptor 1 (encoded by IFNAR1 in humans) contributes to stronger type I interferon signalling, impaired colonization and abundance of A. muciniphila, reduced palmitoleic acid production, higher levels of tumour necrosis factor, and more severe tuberculosis disease in humans and transgenic mice. Thus, host genetics are critical in modulating the structure and functions of gut microbiome and gut microbial metabolites, which further determine disease susceptibility.

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: Reduction of A. muciniphila abundance is associated with active tuberculosis infection and higher M. tuberculosis-specific tumour necrosis factor production in humans.
Fig. 2: A. muciniphila may confer anti-tuberculosis protection and reduce tumour necrosis factor expression in wild-type mice.
Fig. 3: A. muciniphila-mediated high levels of palmitoleic acid are associated with anti-TB protection.
Fig. 4: Dietary A. muciniphila-mediated palmitoleic acid reduces tuberculosis pathology, Bacillus burden and tumour necrosis factor production in mice.
Fig. 5: Ifnar1 rs2257167 G allele impairs colonization of A. muciniphila and promotes more severe TB infection.
Fig. 6: Oral gavage of A. muciniphila or palmitoleic acid reduces tuberculosis pathology and Bacillus burdens in M. tuberculosis-infected IFNAR1-p.Val168Val mice.
Fig. 7: Effects of IFNAR1 variants are observed in participants with tuberculosis.

Similar content being viewed by others

Data availability

Data of 16S rDNA sequencing are available in a public repository at https://dataview.ncbi.nlm.nih.gov/. The accession number of 16S rDNA sequencing data is PRJNA609532. Figure 1, Extended Data Figs. 1 and 2 have associated raw data. Source data are provided with this paper. All other data supporting the findings of this study are available from the corresponding authors on reasonable request.

Code availability

Code is available at https://github.com/ZhenhuangGe/Akkermansia-muciniphila/.

References

  1. Schirmer, M. et al. Dynamics of metatranscription in the inflammatory bowel disease gut microbiome. Nat. Microbiol. 3, 337–346 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Fujimura, K. E. & Lynch, S. V. Microbiota in allergy and asthma and the emerging relationship with the gut microbiome. Cell Host Microbe 17, 592–602 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Rothschild, D. et al. Environment dominates over host genetics in shaping human gut microbiota. Nature 555, 210–215 (2018).

    Article  CAS  PubMed  Google Scholar 

  4. David, L. A. et al. Diet rapidly and reproducibly alters the human gut microbiome. Nature 505, 559–563 (2014).

    Article  CAS  PubMed  Google Scholar 

  5. Wu, G. D. et al. Linking long-term dietary patterns with gut microbial enterotypes. Science 334, 105–108 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Muegge, B. D. et al. Diet drives convergence in gut microbiome functions across mammalian phylogeny and within humans. Science 332, 970–974 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Ley, R. E., Turnbaugh, P. J., Klein, S. & Gordon, J. I. Microbial ecology: human gut microbes associated with obesity. Nature 444, 1022–1023 (2006).

    Article  CAS  PubMed  Google Scholar 

  8. Kawamoto, S. et al. The inhibitory receptor PD-1 regulates IgA selection and bacterial composition in the gut. Science 336, 485–489 (2012).

    Article  CAS  PubMed  Google Scholar 

  9. Hapfelmeier, S. et al. Reversible microbial colonization of germ-free mice reveals the dynamics of IgA immune responses. Science 328, 1705–1709 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Donaldson, G. P. et al. Gut microbiota utilize immunoglobulin A for mucosal colonization. Science 360, 795–800 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Bonder, M. J. et al. The effect of host genetics on the gut microbiome. Nat. Genet. 48, 1407–1412 (2016).

    Article  CAS  PubMed  Google Scholar 

  12. Igartua, C. et al. Host genetic variation in mucosal immunity pathways influences the upper airway microbiome. Microbiome 5, 16 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  13. Goodrich, J. K. et al. Human genetics shape the gut microbiome. Cell 159, 789–799 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Lim, M. Y. et al. The effect of heritability and host genetics on the gut microbiota and metabolic syndrome. Gut 66, 1031–1038 (2017).

    Article  CAS  PubMed  Google Scholar 

  15. Wang, J. et al. Genome-wide association analysis identifies variation in vitamin D receptor and other host factors influencing the gut microbiota. Nat. Genet. 48, 1396–1406 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. WHO. Global tuberculosis report. Geneva: World Health Organization (2019).

  17. Bradley, C. P. et al. Segmented filamentous bacteria provoke lung autoimmunity by inducing gut–lung axis TH17 cells expressing dual TCRs. Cell Host Microbe 22, 697–704 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Gauguet, S. et al. Intestinal microbiota of mice influences resistance to Staphylococcus aureus pneumonia. Infect. Immun. 83, 4003–4014 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Schuijt, T. J. et al. The gut microbiota plays a protective role in the host defence against pneumococcal pneumonia. Gut 65, 575–583 (2016).

    Article  CAS  PubMed  Google Scholar 

  20. Budden, K. F. et al. Emerging pathogenic links between microbiota and the gut–lung axis. Nat. Rev. Microbiol. 15, 55–63 (2017).

    Article  CAS  PubMed  Google Scholar 

  21. Wypych, T. P., Wickramasinghe, L. C. & Marsland, B. J. The influence of the microbiome on respiratory health. Nat. Immunol. 20, 1279–1290 (2019).

    Article  CAS  PubMed  Google Scholar 

  22. Roca, F. J. & Ramakrishnan, L. TNF dually mediates resistance and susceptibility to mycobacteria via mitochondrial reactive oxygen species. Cell 153, 521–534 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Stallings, C. L. Host response: inflammation promotes TB growth. Nat. Microbiol. 2, 17102 (2017).

    Article  CAS  PubMed  Google Scholar 

  24. Mishra, B. B. et al. Nitric oxide prevents a pathogen-permissive granulocytic inflammation during tuberculosis. Nat. Microbiol. 2, 17072 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Zumla, A. et al. Host-directed therapies for infectious diseases: current status, recent progress, and future prospects. Lancet Infect. Dis. 16, e47–63 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  26. Orme, I. M., Robinson, R. T. & Cooper, A. M. The balance between protective and pathogenic immune responses in the TB-infected lung. Nat. Immunol. 16, 57–63 (2015).

    Article  CAS  PubMed  Google Scholar 

  27. Botero, L. E. et al. Respiratory tract clinical sample selection for microbiota analysis in patients with pulmonary tuberculosis. Microbiome 2, 29 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  28. Namasivayam, S. et al. Longitudinal profiling reveals a persistent intestinal dysbiosis triggered by conventional anti-tuberculosis therapy. Microbiome 5, 71 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  29. Khan, N. et al. Alteration in the gut microbiota provokes susceptibility to tuberculosis. Front. Immunol. 7, 529 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  30. Derrien, M., Vaughan, E. E., Plugge, C. M. & de Vos, W. M. Akkermansia muciniphila gen. nov., sp. nov., a human intestinal mucin-degrading bacterium. Int. J. Syst. Evol. Microbiol. 54, 1469–1476 (2004).

    Article  CAS  PubMed  Google Scholar 

  31. Kursa, M. B. & Rudnicki, W. R. Feature selection with the Boruta package. J. Stat. Softw. 36, 1–13 (2010).

    Article  Google Scholar 

  32. Hu, Y. et al. The gut microbiome signatures discriminate healthy from pulmonary tuberculosis patients. Front. Cell. Infect. Microbiol. 9, 90 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Behar, S. M. & Sassetti, C. M. Immunology: fixing the odds against tuberculosis. Nature 511, 39–40 (2014).

    Article  CAS  PubMed  Google Scholar 

  34. Ernst, J. D. The immunological life cycle of tuberculosis. Nat. Rev. Immunol. 12, 581–591 (2012).

    Article  CAS  PubMed  Google Scholar 

  35. Flynn, J. L. & Chan, J. Immunology of tuberculosis. Annu. Rev. Immunol. 19, 93–129 (2001).

    Article  CAS  PubMed  Google Scholar 

  36. Kaufmann, S. H. E., Dorhoi, A., Hotchkiss, R. S. & Bartenschlager, R. Host-directed therapies for bacterial and viral infections. Nat. Rev. Drug Discov. 17, 35–56 (2018).

    Article  CAS  PubMed  Google Scholar 

  37. O’Garra, A. et al. The immune response in tuberculosis. Annu. Rev. Immunol. 31, 475–527 (2013).

    Article  PubMed  Google Scholar 

  38. Zeng, G., Zhang, G. & Chen, X. TH1 cytokines, true functional signatures for protective immunity against TB? Cell. Mol. Immunol. 15, 206–215 (2018).

    Article  CAS  PubMed  Google Scholar 

  39. Wang, L. et al. Oxidization of TGFβ-activated kinase by MPT53 is required for immunity to Mycobacterium tuberculosis. Nat. Microbiol. 4, 1378–1388 (2019).

    Article  CAS  PubMed  Google Scholar 

  40. Wang, Y. et al. Long noncoding RNA derived from CD244 signaling epigenetically controls CD8+ T cell immune responses in tuberculosis infection. Proc. Natl Acad. Sci. USA 112, E3883–3892 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Wang, J. et al. Sequence features and chromatin structure around the genomic regions bound by 119 human transcription factors. Genome Res. 22, 1798–1812 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Qiu, L. et al. Severe tuberculosis induces unbalanced up-regulation of gene networks and overexpression of IL-22, MIP-1α, CCL27, IP-10, CCR4, CCR5, CXCR3, PD1, PDL2, IL-3, IFN-β, TIM1 and TLR2 but low antigen-specific cellular responses. J. Infect. Dis. 198, 1514–1519 (2008).

    Article  PubMed  Google Scholar 

  43. Teles, R. M. et al. Type I interferon suppresses type II interferon-triggered human anti-mycobacterial responses. Science 339, 1448–1453 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Berry, M. P. et al. An interferon-inducible neutrophil-driven blood transcriptional signature in human tuberculosis. Nature 466, 973–977 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Mayer-Barber, K. D. et al. Host-directed therapy of tuberculosis based on interleukin-1 and type I interferon crosstalk. Nature 511, 99–103 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Tschurtschenthaler, M. et al. Type I interferon signalling in the intestinal epithelium affects Paneth cells, microbial ecology and epithelial regeneration. Gut 63, 1921–1931 (2014).

    Article  CAS  PubMed  Google Scholar 

  47. Sun, L. et al. Type I interferons link viral infection to enhanced epithelial turnover and repair. Cell Host Microbe 17, 85–97 (2015).

    Article  CAS  PubMed  Google Scholar 

  48. Zhang, G. et al. A proline deletion in IFNAR1 impairs IFN-signaling and underlies increased resistance to tuberculosis in humans. Nat. Commun. 9, 85 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  49. Zhang, G. et al. A functional single-nucleotide polymorphism in the promoter of the gene encoding interleukin 6 is associated with susceptibility to tuberculosis. J. Infect. Dis. 205, 1697–1704 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Zhang, G. et al. Allele-specific induction of IL-1β expression by C/EBPβ and PU.1 contributes to increased tuberculosis susceptibility. PLoS Pathog. 10, e1004426 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  51. Ors, F. et al. High-resolution CT findings in patients with pulmonary tuberculosis: correlation with the degree of smear positivity. J. Thorac. Imaging 22, 154–159 (2007).

    Article  PubMed  Google Scholar 

  52. Khan, N. et al. Intestinal dysbiosis compromises alveolar macrophage immunity to Mycobacterium tuberculosis. Mucosal Immunol. 12, 772–783 (2019).

    Article  CAS  PubMed  Google Scholar 

  53. Negi, S., Pahari, S., Bashir, H. & Agrewala, J. N. Gut microbiota regulates mincle-mediated activation of lung dendritic cells to protect against Mycobacterium tuberculosis. Front. Immunol. 10, 1142 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Falony, G. et al. Population-level analysis of gut microbiome variation. Science 352, 560–564 (2016).

    Article  CAS  PubMed  Google Scholar 

  55. Vandeputte, D. et al. Stool consistency is strongly associated with gut microbiota richness and composition, enterotypes and bacterial growth rates. Gut 65, 57–62 (2016).

    Article  CAS  PubMed  Google Scholar 

  56. Giosue, S. et al. Effects of aerosolized interferon-alpha in patients with pulmonary tuberculosis. Am. J. Respir. Crit. Care Med. 158, 1156–1162 (1998).

    Article  CAS  PubMed  Google Scholar 

  57. Boxx, G. M. & Cheng, G. The roles of type I interferon in bacterial infection. Cell Host Microbe 19, 760–769 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. McNab, F., Mayer-Barber, K., Sher, A., Wack, A. & O’Garra, A. Type I interferons in infectious disease. Nat. Rev. Immunol. 15, 87–103 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Moreira-Teixeira, L., Mayer-Barber, K., Sher, A. & O’Garra, A. Type I interferons in tuberculosis: foe and occasionally friend. J. Exp. Med. 215, 1273–1285 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Ji, D. X. et al. Type I interferon-driven susceptibility to Mycobacterium tuberculosis is mediated by IL-1Ra. Nat. Microbiol. 4, 2128–2135 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Flynn, J. L. et al. Tumor necrosis factor-alpha is required in the protective immune response against Mycobacterium tuberculosis in mice. Immunity 2, 561–572 (1995).

    Article  CAS  PubMed  Google Scholar 

  62. Miller, E. A. & Ernst, J. D. Anti-TNF immunotherapy and tuberculosis reactivation: another mechanism revealed. J. Clin. Invest. 119, 1079–1082 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Keane, J. et al. Tuberculosis associated with infliximab, a tumor necrosis factor alpha-neutralizing agent. N. Engl. J. Med. 345, 1098–1104 (2001).

    Article  CAS  PubMed  Google Scholar 

  64. Souza, C. O. et al. Palmitoleic acid reduces high-fat diet-induced liver inflammation by promoting PPAR-gamma-independent M2a polarization of myeloid cells. Biochim Biophys. Acta Mol. Cell Biol. Lipids 1865, 158776 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Guo, X. et al. Palmitoleate induces hepatic steatosis but suppresses liver inflammatory response in mice. PLoS ONE 7, e39286 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. de Souza, C. O., Vannice, G. K., Rosa Neto, J. C. & Calder, P. C. Is palmitoleic acid a plausible nonpharmacological strategy to prevent or control chronic metabolic and inflammatory disorders? Mol. Nutr. Food Res. 62 (2018).

  67. Aden, K. et al. Metabolic functions of gut microbes associate with efficacy of tumor necrosis factor antagonists in patients with inflammatory bowel diseases. Gastroenterology 157, 1279–1292 (2019).

    Article  CAS  PubMed  Google Scholar 

  68. Scott, N. A. et al. Antibiotics induce sustained dysregulation of intestinal T cell immunity by perturbing macrophage homeostasis. Sci. Transl. Med. 10, eaao4755 (2018).

  69. Lachmandas, E. et al. Diabetes mellitus and increased tuberculosis susceptibility: the role of short-chain fatty acids. J. Diabetes Res. 2016, 6014631 (2016).

    Article  PubMed  Google Scholar 

  70. Arpaia, N. et al. Metabolites produced by commensal bacteria promote peripheral regulatory T cell generation. Nature 504, 451–455 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Collado, M. C., Derrien, M., Isolauri, E., de Vos, W. M. & Salminen, S. Intestinal integrity and Akkermansia muciniphila, a mucin-degrading member of the intestinal microbiota present in infants, adults and the elderly. Appl. Environ. Microbiol. 73, 7767–7770 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Kind, T. et al. FiehnLib: mass spectral and retention index libraries for metabolomics based on quadrupole and time-of-flight gas chromatography/mass spectrometry. Anal. Chem. 81, 10038–10048 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Amann, R. I. et al. Combination of 16S rRNA-targeted oligonucleotide probes with flow cytometry for analyzing mixed microbial populations. Appl Environ. Microbiol. 56, 1919–1925 (1990).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

This work was supported by grants from the Natural Science Foundation of China (82072250 and 81361120379 to G. Zeng and 81873958 to G. Zhang), the National Science and Technology Major Project (2017ZX10201301, 2017ZX106019 and 2017ZX10103004 to G. Zhang), the Shenzhen Scientific and Technological Foundation (JCYJ20180228162511084 to G. Zhang) and the Sanming Project of Medicine in Shenzhen (SZSM201911009 to G. Zhang).

Author information

Authors and Affiliations

Authors

Contributions

L.C., Z.G., Y.Y., Z.Z., Y.C., Q.M., Z.C., S.Y., L.C.Z. and G.B.L. performed the experiments and analysed the data. G.L.Z., G.L., Z.L., W.W., X.H. and L. Zhou collected clinical samples and conducted analyses. S.Y. and J.C. helped with signalling analysis experiments. X.H. and G.L.Z. performed SNP and clinical analysis works and X.H., Z.L., J.P., J.C., L.S., L.C.Z., H.L., X.C. and B.G. assisted with the data analysis and manuscript preparation. L.C., G.C.Z., Z.W.C., X.C. and B.G. drafted, discussed and revised the manuscript. G.C.Z. conceived the study. L.S. and Z.W.C. used their overtimes at night to help project design, data evaluation/discussion and writing before 2019. This study addressed scientific questions of the US-China Collaborative Biomedical Research program approved by NIH/UIC/NSFC.

Corresponding author

Correspondence to Gucheng Zeng.

Ethics declarations

Competing interests

G.C.Z. is a founder of Revaissant Bioscience. All other authors declare no competing interests.

Peer review

Peer review information

Nature Metabolism thanks the anonymous reviewers for their contribution to the peer review of this work. Primary Handling Editor: Alfredo Giménez-Cassina.

Additional information

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

Extended data

Extended Data Fig. 1 Healthy controls (HC)and active TB patients (TB) exhibit significant differences in diversity, composition, and abundance of gut bacteria at Shenzhen cohort.

(a) PCoA plot (based on Weighted UniFrac distances). Red points indicated the microbiota enriched in HC (n = 28), blue points indicated the microbiota enriched in TB (n = 26). (b) The relative abundance of gut bacteria at phylum level in the fecal samples from HC (n = 28) and TB (n = 26). (c) The top 10 bacterial species with enriched relative abundance in HC were adjusted P < 0.05 and log2 (HC/TB) > 0, species with enriched relative abundance in TB were adjusted P < 0.05 and log2 (HC/TB) < 0 (HC, n = 28; TB, n = 26). The red box indicates A. muciniphila. (d) The absolute abundance of detected A. muciniphila in the fecal samples from HC (n = 28) and TB (n = 26). (e) Predictive power of top 10 species enriched in HC (that is the top 10 most reduced species in TB) assessed by random forest analysis. Blue boxplots acted as benchmarks. Green boxplots represent confirmed species and red boxplots represent rejected species. Red arrowhead marks A. muciniphila. HC = 28, TB = 26. (f) Histogram of the linear discriminant analysis (LDA) coupled with effect size measurements (LEfSe) identified the species with different abundance between HC and TB. Higher abundant species in TB are shaded with green, higher abundant species in HC are shaded with red. Red arrowhead marks the A. muciniphila, black arrowhead marks B. vulgatus. (g) Circle charts marked the relative abundance of six bacteria with differentiated relative abundance between HC and TB in Shenzhen cohort, and these differentiated six bacteria with differentiated relative abundance were also observed at both Shenzhen and Foshan cohorts. Histogram showed the fold changes of relative abundance of six bacteria, including B. vulgatus, B. uniformis, A. muciniphila, B. caccae, P. merdae, and E. ramosun between HC and TB in Shenzhen cohort (fold changes were calculated as HC/TB). Bacteria are identified by color bars above the chart. Data are presented as mean + /- SD. Pvalues were calculated by PERMANOVA test (a), Mann–Whitney U test [(c) and (d)].

Source data

Extended Data Fig. 2 HC and TB show significantly different diversity, composition and abundance of gut bacteria, and abundance of A. muciniphila and B. vulgatus are significantly higher in HC than TB at Foshan cohort.

(a) PCoAplot (based on Weighted UniFrac distances (HC = 17, TB = 19, Foshan cohort). (b) The relative abundance of gut bacteria in phylum level in the fecal samples from HC (n = 17) and TB (n = 19). (c) A. muciniphila was belonged to top 10 bacterial species in fecal microbiota of HC. Species with enriched relative abundance in HC are adjusted P < 0.05 and log2 (HC/TB) > 0, species with enriched relative abundance in TB are adjusted P < 0.05 and log2 (HC/TB) < 0. The red-boxed area marks the A. muciniphila. (d) The absolute abundance of A. muciniphila in the fecal microbiota from HC (n = 17) and TB (n = 19). (e)Predictive power of top 10 species enriched in HC (that is the top 10 most reduced species in TB) assessed by random forest analysis. Blue boxplots correspond to minimal, average, and maximum Z-score of shadow species, which were shuffled version of real species introduced to random forest classifier and act as benchmarks to detect truly predictive species. Red boxplots represent rejected species, yellow boxplots represent suggestive species, and green boxplots represent confirmed species. The red arrowhead marks the A. muciniphila. HC = 17, TB = 19. (f) Histogram of the Linear discriminant analysis (LDA) coupled with effect size measurements (LEfSe) identified the species with different abundance in HC and TB. Higher abundant species in TB are shaded in green, higher abundant species in HC are shaded in red. Red arrowhead pointed A. muciniphila, black arrowhead pointed B. vulgatus. (g) Circle charts showed the relative abundance of six bacteria with differentiated relative abundance between HC and TB in Foshan cohort, and these six bacteria were also observed at both Shenzhen and Foshan cohorts. Histogram showed the fold change of relative abundance of six bacteria, including B. vulgatus, B. uniformis, A. muciniphila, B. caccae, P. merdae and E. ramosun between HC and TB (calculated as HC/TB) in Foshan cohort. Bacteria are identified by color bars above the chart. Data are presented as mean + /- SD. Pvalue was calculated by PERMANOVA test (a), Mann-Whitney test [(c)and (d)].

Source data

Extended Data Fig. 3 The association of M. tuberculosis-specific cytokinesex vivo produced by PBMCs and B. vulgatus in fecal samples of TB patients.

Spearman correlation analysis showed the association between TNF-α, IFN-γ, or IL-10 and the relative abundance of B. vulgatus. There were no significant correlation between B. vulgatus and M. tuberculosis-specific TNF-α, IFN-γ or IL-10. P-value was calculated by Spearman correlation (two-tailed).

Source data

Extended Data Fig. 4 Antibiotics-treated mice show more severe pulmonary pathology and higher bacillus burdens during M. tuberculosis infection.

(a) Two representative lungs derived from M. tuberculosis-infected mice with antibiotics or water control treatment. Red circles mark the cystic changes, hemorrhage or necrosis on the lungs of infected mice. (b-c) Hematoxylin and eosin (H&E) staining of two representative lungs (b) and the histological scores (c) of M. tuberculosis-infected mice with antibiotics or water control treatment at 5 weeks post infection. Scale bars, 500 μm (top: original magnification), 200 μm (middle: 2.5×) and 50 μm (bottom: 10×). The red-boxed areas at the top are enlarged below. As marked by yellow arrowheads, antibiotics-treated mice showed more severe lesions and more infiltration of inflammatory cells than untreated mice. (d) Quantification analysis of M. tuberculosis CFU in the lung homogenates of M. tuberculosis-infected mice at 5 weeks post infection. (e) The acid-fast staining of M. tuberculosis in lung section of mice at 5 weeks post infection. Note that more acid-fast staining-positive bacilli in lung sections derived from antibiotics-treated mice. Scale bars, 5 μm (top: 100× of original magnification). The red-boxed areas at the top are enlarged below. N = 6 mice per group. Data are representative at least two biological replicates.Data are presented as mean + /- SD. P values were calculated by Student’s two-tailed unpaired t-test [(c) and (d)].

Source data

Extended Data Fig. 5 A. muciniphila-mediated effects of reduced TB pathology and bacillus burdens were impaired in the absence of TNF-α signaling.

(a) Experimental diagram for determining whether the effect of A. muciniphila depends on TNF-α signaling during M. tuberculosis infection. Antibiotics pre-treated mice were gavaged with A. muciniphila (2 × 108 CFU) or saline for 3 times per week followed by aerosol M. tuberculosis infection. Mice were injected with 500 μg anti-TNF-α MAb or IgG antibody control at the third day after M. tuberculosis infection and injections were performed every third day for a total of six injections. (b) Gross pathology shows the hemorrhage and necrosis in lungs of M. tuberculosis-infected mice at 5 weeks post infection. Red circles mark the severe, unresolved hemorrhage, massive disruption or caseous necrosis on the lungs of M. tuberculosis-infected mice. (c-d) Hematoxylin and eosin (H&E) staining of four representative lungs (c) and the histological scores (d) of M. tuberculosis-infected mice at 5 weeks post infection. Scale bars, 500 μm (top: original magnification), 200 μm (middle: 2.5×) and 50 μm (bottom: 10×). The red-boxed areas at the top are enlarged below. Yellow arrowheads mark lesions and infiltration of inflammatory cells. (e) Quantification analysis of M. tuberculosis CFU in the lung homogenates of M. tuberculosis-infected mice at 5 weeks post infection. N = 6 mice per group. Data are representative at least two biological replicates.Data are presented as mean + /- SD. Pvalues were calculated by Student’s two-tailed unpaired t-test [(d) and (e)].

Source data

Extended Data Fig. 6 Intestine-colonized A. muciniphila and anA. muciniphila-mediated metabolite palmitoleic acid,inhibit M. tuberculosis-specific TNF-α expression via an epigenetic mechanism.

(a) The experimental diagram of transwell assay to mimic the effects of epithelium-colonized gut bacteria on cytokine production. (b) Representative CBA analysis of culture supernatants of CD3 + T cells in the lower chamber for transwell assay. Red and black-boxed areas mark the fluorescent clusters of TNF-α and IFN-γ, respectively, and dashed line mark the shift of fluorescent clusters of TNF-α and IFN-γ, respectively. (c) Pooled bar graphic data show the expression of TNF-α/IFN-γ/IL-2/IL-4/IL-6/IL-10 in culture supernatants of CD3 + T cells in the lower chamber. N = 7 mice per group. One-way ANOVA, P = 0.0064. (d) Representative CBA analysis of culture supernatants of CD3 + T cells of mice. CD3 + T cells were co-cultured with A. muciniphila lysates (10 μg/ml) or BHI in presence of M. tuberculosis lysates (10 μg/ml). Red and black-boxed areas mark the fluorescent clusters of TNF-α and IFN-γ, respectively, and dashed line mark the shift of fluorescent clusters of TNF-α and IFN-γ, respectively. (e) Pooled bar graphic data show the in vitro cytokines expression in culture supernatants of CD3 + T cells. N = 7 mice per group. (f) CHIP-qPCR analysis of IgG, H3K4Me3, H3K9Me1, H3K9Me3, H3K27Me2 and H3K27Me3, and control antibodies at the promoter of tnfα. N = 7 biologically independent samples. (g) Immunoblot analysis of H3K4Me3, H3K9Me1, H3K9Me3, H3K27Me2 and H3K27Me3. (h) Quantitative immunoblot analysis of expression of H3K4Me3, H3K9Me1, H3K9Me3, H3K27Me2 and H3K27Me3. ImageJ was used for quantitative analysis of immunoblot. N = 3 biologically independent samples. Data in (d) to (h) are from one experiment. Box and whisker plots depict the following: line, median; box limits, first and third quartiles; whiskers, Min to Max and Show all points. Data are representative at least two biological replicates.Data are presented as mean + /- SD. P values were calculated by one-way ANOVA with Tukey’s multiple comparison test (c) and Student’s two-tailed unpaired t-test [(e), (f) and (h)].

Source data

Extended Data Fig. 7 A. muciniphila played a dominant role in contributing to increased levels of palmitoleic acid in mice.

(a) Representative total ion chromatograms (TICs) of metabolites profiles in culture supernatants of A. muciniphila by GC-TOF-MS-based metabolic analysis. Black arrowhead marks the identification of palmitoleic acid (CAS: 373-49-9). (b) Comparative KEGG pathway analyses of palmitoleic acid synthesis and degradation between A. muciniphila (marked with oval black dashed line) and B. vulgatus (marked with green solid line). The green box mark the presence of long-chain-fatty-acid-CoA ligase for degrading/processing palmitoleic acid in B. vulgatus, and the gray box point the lack of presence (or at least lower level) of long-chain-fatty-acid-CoA ligase in A. muciniphila. The solid green arrow mark the downstream metabolic pathways for degrading/processing palmitoleic acid in B. vulgatus, and the black dashed arrow point that no (or at least less) long-chain-fatty-acid-CoA ligase to degrade/process palmitoleic acid in A. muciniphila. (c) Schematic diagram for palmitoleic acid biosynthesis pathway in A. muciniphila. (d) Genes involved in palmitoleic acid biosynthesis pathway in A. muciniphila and B. vulgatus, respectively. (e-f) The existences of genes involved in palmitoleic acid biosynthesis pathway in A. muciniphila and B. vulgatus were confirmed from the bacterial genome DNA and bacterial mRNA. (g) The qPCR-based analysis of fatty acyl-ACP thioesterase (FAT) mRNA expression in fecal samples from M. tuberculosis-infected mice. (h) Genes involved in palmitoleic acid metabolism pathway in gut bacteria, which showed significantly different abundance between HC and TB. These bacteria includeAcinetobacter johnsonii (AJ), Lactobacillus mucosae (LM), Lactobacillus salivarius (LS), Clostridium innocuum (CI), Plesiomonas shigelloides (PS), Comamonas kerstersii (CK), Alistipes onderdonkii (AO), Lactobacillus ruminis (LR), Streptococcus anginosus (SA), Lactobacillus fermentum (LF), Faecalibacterium prausnitzii (FP), Bifidobacterium adolescentis (BA), Bacteroides uniformis (BU) and Bacteroides caccae (BC). (i-j)The relative abundance of A. muciniphila, A. johnsonii (AJ) and L. salivarius (LS) in stool samples, and palmitoleic acid concentrations in plasmaand fecal samples fromantibiotics-treated mice at the third day after bacteria gavage. N = 6 mice per group. Data are representative at least two biological replicates.Data are presented as mean + /- SD. P values were calculated by Student’s two-tailed unpaired t-test [(g) and (i)] and one-way ANOVA with Tukey’s multiple comparison test (j).

Source data

Extended Data Fig. 8 Live but not inactive A. muciniphila reduces TB pathology and bacillus burdens.

(a) Three representative lungs of M. tuberculosis-infected mice treated with live A. muciniphila, inactive A. muciniphila or saline. Red circles mark the severe, unresolved hemorrhage, massive disruption or caseous necrosis on the lungs of M. tuberculosis-infected mice. (b-c) Hematoxylin and eosin (H&E) staining of three representative lungs (b) and histological score (c) of M. tuberculosis-infected mice treated with live A. muciniphila, inactive A. muciniphila and saline at 5 weeks post infection. Scale bars, 500 μm (top: original magnification), 200 μm (middle: 2.5×) and 50 μm (bottom: 10×). The red-boxed areas at the top are enlarged below. Yellow arrowheads mark lesions and infiltration of inflammatory cells. (d) Quantitative analysis of bacillus CFU in lung homogenates of M. tuberculosis-infected mice treated with live A. muciniphila, inactive A. muciniphila and saline at 5 weeks post infection. N = 6 mice per group. Data are representative at least two biological replicates.Data are presented as mean + /- SD. P values were calculated by one-way ANOVA with Tukey’s multiple comparison test [(c) and (d)].

Source data

Extended Data Fig. 9 Knock-out of an IFN-I receptor (Ifnar1) increases the intestinal colonization and abundance of bulk anaerobic bacteria and A. muciniphila.

(a) Experimental design for determining the effect of IFN-I signalingon A. muciniphilaabundance. (b) The proportion analysis of anaerobic bacteria in jejunum, ileum, cecum or stool of Ifnar1-/- mice and wild-type miceat day 18. (c) Pooled data of relative amounts of anaerobic bacteria recovered from jejunum, ileum or cecum of Ifnar1-/- mice (red bar) and wild-type mice (black bar), respectively. Mice were administrated with fecal samples of HC (Left panel) or TB (Right panel) as described in (a), and then tissues of jejunum, ileum or cecum and fecal samples of mice were subjected to anaerobic culture, followed by quantitative analysis of anaerobic bacteria CFU. Fold changes were calculated as CFU of anaerobic bacteria recovered from jejunum, ileum or cecum and fecal samples of mice administrated with fecal samples of HC or TB normalized to those of saline-treated Ifnar1-/- mice and wild-type mice. (d) The abundance analysis of A. muciniphila in jejunum, ileum, or cecum of Ifnar1-/- mice and wild-type mice, respectively. Tissues of jejunum, ileum and cecum of mice were collected at day 18 to analyze the copies of A. muciniphila per gram feces. (e) Pooled data of relative gene expression levels of A. muciniphila recovered from jejunum, ileum, or cecum of Ifnar1-/- mice (red bar) and wild-type mice (black bar), respectively. Mice were administrated with fecal samples of HC (Left panel) or TB (Right panel) as described in (a), and then tissues of jejunum, ileum, or cecum were subjected to qPCR analysis of DNA extracted from jejunum, ileum, and cecum of mice. Fold changes were calculated as gene expression levels of A. muciniphila recovered from jejunum, ileum, or cecum of mice administrated with fecal samples of HC or TB normalized to those of same genotype mice with oral saline administration. N = 6 mice per group. Data are representative at least two biological replicates.Data are presented as mean + /- SD.P values were calculated byStudent’s two-tailed unpaired t-test (d).

Source data

Extended Data Fig. 10 Pooled bar graphic data shows the expression of IL-10, IL-6, IL-2, and IL-4 in culture supernatants of PBMCs derived from TB patients carrying genotype GG, GC, CC.

sPBMCs derived from TB patients carrying genotype GG (n = 185), GC (n = 206) or CC (n = 62) were ex vivo stimulated with M. tuberculosis lysates and analyzed as in Fig. 7. Data are presented as mean + /- SD. P values were calculated by one-way ANOVA with Tukey’s multiple comparison test.

Source data

Supplementary information

Supplementary Information

Supplementary Figs. 1–11, Additional Supplementary Fig. 1, Supplementary Tables 1–5 and Supplementary Methods

Reporting Summary

Supplementary Data 1

Statistical source data for Supplementary Figs 1–4, 6, 7 and 10.

Source data

Source Data Fig. 1

Statistical source data.

Source Data Fig. 2

Statistical source data.

Source Data Fig. 3

Statistical source data.

Source Data Fig. 4

Statistical source data.

Source Data Fig. 5

Statistical source data.

Source Data Fig. 5

Unprocessed western blots and/or gels.

Source Data Fig. 6

Statistical source data.

Source Data Fig. 7

Statistical source data.

Source Data Extended Data Fig. 1

Statistical source data.

Source Data Extended Data Fig. 2

Statistical source data.

Source Data Extended Data Fig. 3

Statistical source data.

Source Data Extended Data Fig. 4

Statistical source data.

Source Data Extended Data Fig. 5

Statistical source data.

Source Data Extended Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 6

Unprocessed western blots and/or gels.

Source Data Extended Data Fig. 7

Statistical source data.

Source Data Extended Data Fig. 8

Statistical source data.

Source Data Extended Data Fig. 9

Statistical source data.

Source Data Extended Data Fig. 10

Statistical source data.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chen, L., Zhang, G., Li, G. et al. Ifnar gene variants influence gut microbial production of palmitoleic acid and host immune responses to tuberculosis. Nat Metab 4, 359–373 (2022). https://doi.org/10.1038/s42255-022-00547-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s42255-022-00547-3

This article is cited by

Search

Quick links

Nature Briefing Microbiology

Sign up for the Nature Briefing: Microbiology newsletter — what matters in microbiology research, free to your inbox weekly.

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