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:

Neuronal programming by microbiota regulates intestinal physiology

Abstract

Neural control of the function of visceral organs is essential for homeostasis and health. Intestinal peristalsis is critical for digestive physiology and host defence, and is often dysregulated in gastrointestinal disorders1. Luminal factors, such as diet and microbiota, regulate neurogenic programs of gut motility2,3,4,5, but the underlying molecular mechanisms remain unclear. Here we show that the transcription factor aryl hydrocarbon receptor (AHR) functions as a biosensor in intestinal neural circuits, linking their functional output to the microbial environment of the gut lumen. Using nuclear RNA sequencing of mouse enteric neurons that represent distinct intestinal segments and microbiota states, we demonstrate that the intrinsic neural networks of the colon exhibit unique transcriptional profiles that are controlled by the combined effects of host genetic programs and microbial colonization. Microbiota-induced expression of AHR in neurons of the distal gastrointestinal tract enables these neurons to respond to the luminal environment and to induce expression of neuron-specific effector mechanisms. Neuron-specific deletion of Ahr, or constitutive overexpression of its negative feedback regulator CYP1A1, results in reduced peristaltic activity of the colon, similar to that observed in microbiota-depleted mice. Finally, expression of Ahr in the enteric neurons of mice treated with antibiotics partially restores intestinal motility. Together, our experiments identify AHR signalling in enteric neurons as a regulatory node that integrates the luminal environment with the physiological output of intestinal neural circuits to maintain gut homeostasis and health.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Programming of the enteric neuron transcriptome by microbiota.
Fig. 2: Microbiota- and ligand-dependent activation of AHR signalling.
Fig. 3: AHR signalling in enteric neurons regulates intestinal peristalsis.

Similar content being viewed by others

Data availability

All RNA-seq data are available at Gene Expression Omnibus (GEO) under accession number GSE140293. Source Data for Figs. 2, 3 and Extended Data Fig. 15, 7 are provided with the paper. All datasets analysed during the current study are presented in this manuscript, or are available from the corresponding authors upon reasonable request.

Code availability

The source code and installation instructions for colonic migrating motor complex evaluation and Ca2+ imaging can be found at https://doi.org/10.7554/eLife.42914.039 (Ca2+ imaging analysis source code) and https://doi.org/10.7554/eLife.42914.040 (installation instructions and user guide). For more information, please contact pieter.vandenberghe@kuleuven.be. The code related to the RNAscope signal quantification is available at GitHub (https://github.com/FrancisCrickInstitute/Pachnis-lab/tree/master/Neuronal-programming-Nature).

References

  1. Rolig, A. S. et al. The enteric nervous system promotes intestinal health by constraining microbiota composition. PLoS Biol. 15, e2000689 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Obata, Y. & Pachnis, V. The effect of microbiota and the immune system on the development and organization of the enteric nervous system. Gastroenterology 151, 836–844 (2016).

    Article  CAS  PubMed  Google Scholar 

  3. Dey, N. et al. Regulators of gut motility revealed by a gnotobiotic model of diet-microbiome interactions related to travel. Cell 163, 95–107 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. De Palma, G. et al. Transplantation of fecal microbiota from patients with irritable bowel syndrome alters gut function and behavior in recipient mice. Sci. Transl. Med. 9, eaaf6397 (2017).

    Article  CAS  PubMed  Google Scholar 

  5. Hyland, N. P. & Cryan, J. F. Microbe–host interactions: influence of the gut microbiota on the enteric nervous system. Dev. Biol. 417, 182–187 (2016).

    Article  CAS  PubMed  Google Scholar 

  6. Yoo, B. B. & Mazmanian, S. K. The enteric network: interactions between the immune and nervous systems of the gut. Immunity 46, 910–926 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Furness, J. B. The enteric nervous system and neurogastroenterology. Nat. Rev. Gastroenterol. Hepatol. 9, 286–294 (2012).

    Article  CAS  PubMed  Google Scholar 

  8. Yano, J. M. et al. Indigenous bacteria from the gut microbiota regulate host serotonin biosynthesis. Cell 161, 264–276 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Vincent, A. D., Wang, X. Y., Parsons, S. P., Khan, W. I. & Huizinga, J. D. Abnormal absorptive colonic motor activity in germ-free mice is rectified by butyrate, an effect possibly mediated by mucosal serotonin. Am. J. Physiol. Gastrointest. Liver Physiol. 315, G896–G907 (2018).

    Article  CAS  PubMed  Google Scholar 

  10. Roberts, R. R., Murphy, J. F., Young, H. M. & Bornstein, J. C. Development of colonic motility in the neonatal mouse-studies using spatiotemporal maps. Am. J. Physiol. Gastrointest. Liver Physiol. 292, G930–G938 (2007).

    Article  CAS  PubMed  Google Scholar 

  11. De Vadder, F. et al. Gut microbiota regulates maturation of the adult enteric nervous system via enteric serotonin networks. Proc. Natl Acad. Sci. USA 115, 6458–6463 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Ge, X. et al. Antibiotics-induced depletion of mice microbiota induces changes in host serotonin biosynthesis and intestinal motility. J. Transl. Med. 15, 13 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. McVey Neufeld, K. A., Mao, Y. K., Bienenstock, J., Foster, J. A. & Kunze, W. A. The microbiome is essential for normal gut intrinsic primary afferent neuron excitability in the mouse. Neurogastroenterol. Motil. 25, 183-e88 (2013).

    Article  CAS  PubMed  Google Scholar 

  14. Donaldson, G. P., Lee, S. M. & Mazmanian, S. K. Gut biogeography of the bacterial microbiota. Nat. Rev. Microbiol. 14, 20–32 (2016).

    Article  CAS  PubMed  Google Scholar 

  15. Stockinger, B., Di Meglio, P., Gialitakis, M. & Duarte, J. H. The aryl hydrocarbon receptor: multitasking in the immune system. Annu. Rev. Immunol. 32, 403–432 (2014).

    Article  CAS  PubMed  Google Scholar 

  16. Schiering, C. et al. Feedback control of AHR signalling regulates intestinal immunity. Nature 542, 242–245 (2017).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  17. Gutiérrez-Vázquez, C. & Quintana, F. J. Regulation of the immune response by the aryl hydrocarbon receptor. Immunity 48, 19–33 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Stejskalova, L., Dvorak, Z. & Pavek, P. Endogenous and exogenous ligands of aryl hydrocarbon receptor: current state of art. Curr. Drug Metab. 12, 198–212 (2011).

    Article  CAS  PubMed  Google Scholar 

  19. Hibino, H. et al. Inwardly rectifying potassium channels: their structure, function, and physiological roles. Physiol. Rev. 90, 291–366 (2010).

    Article  CAS  PubMed  Google Scholar 

  20. Zholos, A. V., Baidan, L. V., Starodub, A. M. & Wood, J. D. Potassium channels of myenteric neurons in guinea-pig small intestine. Neuroscience 89, 603–618 (1999).

    Article  CAS  PubMed  Google Scholar 

  21. Wang, H. R. et al. Selective inhibition of the Kir2 family of inward rectifier potassium channels by a small molecule probe: the discovery, SAR, and pharmacological characterization of ML133. ACS Chem. Biol. 6, 845–856 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Walisser, J. A., Glover, E., Pande, K., Liss, A. L. & Bradfield, C. A. Aryl hydrocarbon receptor-dependent liver development and hepatotoxicity are mediated by different cell types. Proc. Natl Acad. Sci. USA 102, 17858–17863 (2005).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  23. Schmidt, J. V., Su, G. H., Reddy, J. K., Simon, M. C. & Bradfield, C. A. Characterization of a murine Ahr null allele: involvement of the Ah receptor in hepatic growth and development. Proc. Natl Acad. Sci. USA 93, 6731–6736 (1996).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  24. Bjeldanes, L. F., Kim, J. Y., Grose, K. R., Bartholomew, J. C. & Bradfield, C. A. Aromatic hydrocarbon responsiveness-receptor agonists generated from indole-3-carbinol in vitro and in vivo: comparisons with 2,3,7,8-tetrachlorodibenzo-p-dioxin. Proc. Natl Acad. Sci. USA 88, 9543–9547 (1991).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  25. Metidji, A. et al. The environmental sensor AHR protects from inflammatory damage by maintaining intestinal stem cell homeostasis and barrier integrity. Immunity 49, 353–362 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Roager, H. M. & Licht, T. R. Microbial tryptophan catabolites in health and disease. Nat. Commun. 9, 3294 (2018).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  27. Agus, A., Planchais, J. & Sokol, H. Gut microbiota regulation of tryptophan metabolism in health and disease. Cell Host Microbe 23, 716–724 (2018).

    Article  CAS  PubMed  Google Scholar 

  28. Jankipersadsing, S. A. et al. A GWAS meta-analysis suggests roles for xenobiotic metabolism and ion channel activity in the biology of stool frequency. Gut 66, 756–758 (2017).

    Article  CAS  PubMed  Google Scholar 

  29. Rothhammer, V. & Quintana, F. J. The aryl hydrocarbon receptor: an environmental sensor integrating immune responses in health and disease. Nat. Rev. Immunol. 19, 184–197 (2019).

    Article  CAS  PubMed  Google Scholar 

  30. Wilhelmsen, K., Ketema, M., Truong, H. & Sonnenberg, A. KASH-domain proteins in nuclear migration, anchorage and other processes. J. Cell Sci. 119, 5021–5029 (2006).

    Article  CAS  PubMed  Google Scholar 

  31. Henderson, C. J. et al. Application of a novel regulatable Cre recombinase system to define the role of liver and gut metabolism in drug oral bioavailability. Biochem. J. 465, 479–488 (2015).

    Article  CAS  PubMed  Google Scholar 

  32. Srinivas, S. et al. Cre reporter strains produced by targeted insertion of EYFP and ECFP into the ROSA26 locus. BMC Dev. Biol. 1, 4 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Danielian, P. S., Muccino, D., Rowitch, D. H., Michael, S. K. & McMahon, A. P. Modification of gene activity in mouse embryos in utero by a tamoxifen-inducible form of Cre recombinase. Curr. Biol. 8, 1323–1326 (1998).

    Article  CAS  PubMed  Google Scholar 

  34. Madisen, L. et al. Transgenic mice for intersectional targeting of neural sensors and effectors with high specificity and performance. Neuron 85, 942–958 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Gomez de Agüero, M. et al. The maternal microbiota drives early postnatal innate immune development. Science 351, 1296–1302 (2016).

    Article  ADS  CAS  PubMed  Google Scholar 

  36. Gombash, S. E. et al. Intravenous AAV9 efficiently transduces myenteric neurons in neonate and juvenile mice. Front. Mol. Neurosci. 7, 81 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  37. Jiang, W. et al. Persistent induction of cytochrome P450 (CYP)1A enzymes by 3-methylcholanthrene in vivo in mice is mediated by sustained transcriptional activation of the corresponding promoters. Biochem. Biophys. Res. Commun. 390, 1419–1424 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Sasselli, V. et al. Planar cell polarity genes control the connectivity of enteric neurons. J. Clin. Invest. 123, 1763–1772 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Kamentsky, L. et al. Improved structure, function and compatibility for CellProfiler: modular high-throughput image analysis software. Bioinformatics 27, 1179–1180 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Li, Z. et al. Regional complexity in enteric neuron wiring reflects diversity of motility patterns in the mouse large intestine. eLife 8, e42914 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  41. Rao, M. et al. Enteric glia express proteolipid protein 1 and are a transcriptionally unique population of glia in the mammalian nervous system. Glia 63, 2040–2057 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  42. Gabanyi, I. et al. Neuro-immune interactions drive tissue programming in intestinal macrophages. Cell 164, 378–391 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank the Crick Science Technology Platforms, the University of Bern FACSLab and the Bern Clean Mouse Facility for expert support; R. Lasrado and S.-H. Chng for assistance with tissue dissection; M. Shapiro for bioinformatic input; C. Schiering for useful advice; all members of the Pachnis laboratory for insightful comments on the manuscript and discussions; and M. D’Amato for insightful comments on the manuscript. Y.O. was supported by an EMBO long-term fellowship (ALTF 1214-2015), travel grants from Boehringer Ingelheim Fonds and the Society for Mucosal Immunology (SMI); he is currently supported by an HFSP postdoctoral fellowship (LT000176/2016). This work was supported by the Medical Research Council (MRC) and The Francis Crick Institute (which receives funding from the MRC, Cancer Research UK and the Wellcome Trust). V.P. was also funded by BBSRC (BB/L022974) and the Wellcome Trust (212300/Z/18/Z).

Author information

Authors and Affiliations

Authors

Contributions

Y.O. and V.P. conceived the study and together with B.S. and A. J. Macpherson designed the experiments. Á.C., A.C.B.-F., C.F., M.G.d.A., B.Y., M.R.M., W.B. and B.Y. helped with the experiments. Á.C. performed the RNAscope in situ hybridization experiments; T.F. helped with the quantification of RNAscope data; M.G.d.A. and B.Y. helped to organize experiments with germ-free and exGF mice; A.C.B.-F., W.B. and P.V.B. provided help with the spatiotemporal mapping experiments and the analysis. C.F. carried out and analysed the Ca2+ imaging experiments with help from W.B. and P.V.B. A.H. prepared the cDNA library for the bulk nRNA-seq. S.B. performed bioinformatics analysis. S.H. performed statistical analysis. R.L. generated ChAT-TVA-mCherry mice. Y.O. generated the AAV-CaMKII-eGFP-KASH construct with help from A. J. Murray. V.P. and Y.O. wrote the manuscript with help from B.S. and A. J. Macpherson, and contributions from all authors.

Corresponding authors

Correspondence to Yuuki Obata or Vassilis Pachnis.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature thanks John Cryan, Michael D. Gershon, Sven Petterson and Harry Sokol 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 AAV-based transcriptional profiling of enteric neurons.

ag, Representative images of myenteric ganglia from mice injected with the AAV9-CaMKII- eGFP-KASH vector. Colon (a) and small intestine (bg) myenteric-plexus preparations were immunostained with antibodies against eGFP (ag), PGP9.5 (a, b), S100β (c), SOX10 (d), nNOS and calretinin (CALR) (e), nNOS and calbindin (CALB) (f), and VIP and ChAT (g). Data represent two independent experiments. Scale bars, 100 μm (a, b) and 30 μm (cg). h, Percentage of PGP9.5+ enteric neurons (mean ± s.d.) in the proximal small intestine and colon expressing eGFP, following intravenous administration of the AAV9-CaMKII-eGFP-KASH vector. n = 1,308 colon neurons and 784 small-intestine neurons from 3 mice. i, FACS plots indicating the gating parameters for the isolation of muscularis externa nuclei (gated on DAPI) from the colon (left) and small intestine (right) of mice injected intravenously with AAV9-CaMKII-eGFP-KASH. j, k, Peripherin (red) and eGFP (green) whole-mount immunostaining of the colon (j) and small intestine (k) of mice injected with AAV9-CaMKII-eGFP-KASH mice following dissection of the muscularis externa. The identification of an intact submucosal plexus demonstrates that our transcriptomic analysis is specific for myenteric neurons. Images represent two independent experiments. Scale bars, 100 μm. l, Volcano plots showing mean log2-transformed fold change (x axis) and significance (−log10(adjusted P value)) of differentially expressed genes between eGFP+ and eGFP nuclei isolated from the colon (left) and small intestine (right) of mice injected with AAV9-CaMKII-eGFP-KASH vector. Coloured dots indicate genes specific to enteric neurons (Ret, Chat, Camk2a, Elavl3, Elavl4, Nos1 and Tubb3) in red, glial cells (Sox10, Gfap, Cdh19, Entpd2, S100b and Plp1)41 in blue and muscular macrophages (Itgam, Cd163, H2-Ab1, Mrc1 and Retnla)42 in green. n = 4 mice (Crick). m, Principal component analysis of the transcriptomes of eGFP+ (neuronal) and eGFP (non-neuronal) nuclei isolated from the muscularis externa of the colon and small intestine of mice injected with AAV9-CaMKII-eGFP-KASH vectors. Segregation of nuclear transcriptomes according to their neuronal versus non-neuronal origin and anatomical location along the gut. n = 4 mice (Crick).

Source data

Extended Data Fig. 2 Differential expression of enteric-neuron-specific genes in myenteric neurons from the small intestine and colon of SPF and germ-free mice.

a, b, Representative images of myenteric ganglia (outlined by dotted line) from small intestine (left) and colon (right) of SPF (a) and germ-free (b) mice hybridized with the indicated fluorescence RNAscope probes and counterstained for the pan-neuronal marker HuC/D. Ret (positive control for RNAscope detection) is expressed in neurons of myenteric ganglia of both the small intestine and the colon. Pou3f3, Pde1c, Pantr2, Ano5, Unc5d and Col25a1 are expressed at higher levels in colonic versus small intestine neurons in both SPF (a) and germ-free (b) mice. Data represent three independent experiments. Transcripts per kilobase million (TPM) values (mean ± s.d.) for each transcript in small intestine and colon neurons from SPF (a) and germ-free (b) mice. n = 8 SPF and 3 germ-free mice. Scale bars, 30 μm.

Source data

Extended Data Fig. 3 Molecular and neurochemical characterization of colonic neurons in germ-free mice.

a, TPM values (mean ± s.d.) for neuronal gene markers Elavl4, Uchl1, Prph, Chat, Vip, Nos1, Calb2 and Nefm in the muscularis externa of the colon of SPF and germ-free mice. n = 4 SPF and 3 germ-free mice. bf, Immunostaining of colonic myenteric ganglia from germ-free (top) and SPF (bottom) mice with VIP, CALR and HuC/D (b), CALR, nNOS and TuJ1 (c), VIP, nNOS and CALB (d), PGP9.5, ChAT and TuJ1 (e) and peripherin and NF-M (f). Scale bars, 30 μm. Data represent three independent experiments.

Source data

Extended Data Fig. 4 Microbiota-dependent expression of AHR in colonic neurons.

a, TPM values (mean ± s.d.) for AHR transcripts in neuronal and non-neuronal nuclear preparations from muscularis externa from colon and small intestine of SPF and germ-free mice. n = 4 (SPF) and 3 (germ-free) mice. b, c, Myenteric ganglia immunostained for KIT (which identifies interstitial cells of Cajal) and AHR (b) or SOX10 (enteric glial cells) and AHR (c). AHR+ cells are distinct from intestitial cells of Cajal and enteric glia. Scale bars, 30 μm. df, Immunostaining of neurons from the colon of wild-type mice for AHR (df) and the neuronal markers peripherin and NF-M (d), calbindin and nNOS (e), and calretinin and HuC/D (f). AHR signal was detected in all subtypes of myenteric neurons (arrowheads). Scale bars, 30 μm. g, Immunostaining of neurons from the colon of ChAT-mCherry-TVA reporter mice for mCherry (red), AHR (green) and HuC/D (blue). Arrowhead indicates an enteric neuron positive for ChAT and AHR. Scale bar, 30 μm. h, i, Immunostaining of myenteric ganglia from the jejunum (h) and ileum (i) with the pan-neuronal marker peripherin (blue) and AHR (red). Scale bar, 30 μm. j, k, Representative images of enteric ganglia from duodenum (j) and colon (k) hybridized with RNAscope probe for Ahr (green). Dotted line defines the borders of myenteric ganglia. Scale bar, 30 μm. l, Quantification (mean ± s.d.) of RNAscope signal per neuron is shown (two-sided non-parametric Mann–Whitney U-test). n = 91 small-intestine and 254 colon neurons from 6 mice. mo, Immunostaining of ganglia from the colon of control (m), antibiotic-treated (n) and microbiota-colonized, antibiotic-treated (o) mice with peripherin (blue) and AHR (red). Small panels show signal for AHR (top) and peripherin (bottom). n = 3 mice for each condition. Scale bars, 30 μm. p, q, Representative images of enteric ganglia from the colon of control (p) and antibiotic-treated (q) mice hybridized with RNAscope probe for Ahr (green). Dotted line defines the borders of myenteric ganglia and arrows indicate positive cells. Scale bars, 30 μm. r, Quantification (mean ± s.d.) of RNAscope signal per neuron is also shown (two-sided non-parametric Mann–Whitney U-test). n = 518 neurons from 4 control and 468 neurons from 4 antibiotic-treated mice. Data represent two (j, k, mq) or three (bi) independent experiments.

Source data

Extended Data Fig. 5 AHR-dependent gene expression and effects on colon myenteric neurons.

a, The top 30 genes upregulated in colonic neurons by AHR-ligand treatment (AHR-induced CUEGs) were identified on the basis of fold-change criteria (log2-transformed fold change = 2 < maximum). b, Cyp1a1::cre;Rosa26eYFP reporter mice were intraperitonially injected with 3MC five days before GFP immunostaining. CYP1A1 induction in response to ligand-activated AHR signalling is expected to induce expression of eYFP. c, d, Immunostaining of myenteric ganglia from the colon (c) and small intestine (d) of 3MC-treated Cyp1a1::cre;Rosa26 mice for peripherin (red), HuC/D (blue) and eGFP (green). Scale bars, 100 μm. Data represent three independent experiments. eg, Live calcium imaging of colonic myenteric plexus preparations from Wnt1::cre;Rosa26-GCaMP6f mice. Electrically stimulated Ca2+ transients in enteric neurons under control conditions (e) or in the presence of the ML-133 blocker21 (10 μM) (f). Data represent four independent experiments. The greyscale images depict a proximal colon myenteric plexus preparation in which enteric neurons were stimulated by a single electrical pulse (top panels) or an electrical pulse train (1 s, 20 Hz; bottom panels) via a focal electrode positioned on an internodal strand leading into the myenteric ganglion in the field of view. Left, baseline before stimulation. Middle, peak GCaMP6f fluorescence of the same ganglion upon electrical stimulation. Scale bars, 20 μm. Right, Ca2+ transients of individual enteric neurons (indicated by colour-coded arrows shown in the middle panels) induced by electrical stimulation. The electrical stimulus was applied at 10 s as marked by the black arrows. Comparison of the average maximal GCaMP6f fluorescence amplitudes of neuronal Ca2+ responses (mean ± s.e.m.) under control conditions (e) or the presence of ML-133 (f) upon single pulse (top) (n = 457 neurons) and pulse train (bottom) (n = 526 neurons) electrical stimulation is shown in g (two-sided paired t-test). h, i, Myenteric ganglia from colon of control (h) and antibiotic-treated (i) mice hybridized with the Kcnj12 RNAscope probe. Dotted line defines the borders of myenteric ganglia and arrows indicate Kcnj12-expressing cells. Scale bars, 30 μm. Data represent two independent experiments. j, Quantification of RNAscope signal (mean ± s.e.m.) shown in h and i (two-sided non-parametric Mann–Whitney U-test). n = 421 neurons from 4 control and 468 neurons from 4 antibiotic-treated mice. Abx, antibiotics. k, l, Myenteric ganglia from colon of SPF mice hybridized with the Ahr (green) (k)and Kcnj12 (blue) (l) RNAscope probes and immunostained with HuC/D (data not shown). Dotted line defines the borders of myenteric ganglia and arrows indicate AHR- and KCNJ12-expressing neurons. Scale bars, 30 μm. Data represent two independent experiments. m, Scatter plot shows positive correlation in RNAscope signal for Ahr (k) and Kcnj12 (l) in myenteric neurons (F-test). n = 1,037 neurons from 3 mice. n, o, Immunostaining of myenteric ganglia from control (Ahr+/+;Rosa26eYFP injected with the AAV9-CaMKII-Cre vector) (n) and AhrEN-KO (o) mice for AHR (red) and eYFP (green). Note the lack of overlap between green and red signal in the case of AhrEN-KO (o). Data are representative of two independent experiments. Scale bars, 30 μm. p, Percentage of AHR+ neurons in myenteric ganglia of control (Ahr+/+;Rosa26eYFP mice injected with the AAV9-CaMKII-Cre vector) and AhrEN-KO mice. Random images were acquired from the colon of each biological replicate (n = 9 for control, n = 13 for AhrEN-KO), and the average percentage (mean ± s.d.) of AHR+ HuC/D + cells among the total population of HuC/D+ neurons was calculated (two-sided Student’s t-test).

Source data

Extended Data Fig. 6 Deletion of Ahr does not alter the organization and composition of myenteric ganglia.

a, Immunostaining of muscularis externa preparations from the colon of control (top) and AhrEN-KO (bottom) mice with nNOS, eYFP and HuC/D (left) or peripherin, eYFP and VIP (right). Scale bars, 30 μm. b, Immunostaining of muscularis externa preparations from the colon of wild-type (top) and Ahr−/− (bottom) mice with PGP9.5 and HuC/D (left), VIP, peripherin and HuC/D (middle) and PGP9.5 and nNOS (right). Scale bars, 100 μm. Data represent three independent experiments.

Extended Data Fig. 7 Intravenous administration of AAV vectors does not elicit an inflammatory response or intestinal dysmotility.

a, Cross-sections from the colon of wild-type (top), wild-type infected with AAV9-CaMKII-Cre (middle) and Ahr fl/fl (bottom) mice stained with Alcian blue–PAS (left) or haematoxylin and eosin (H&E) (right). Data represent two independent experiments. b, Graph (mean ± s.d.) shows that administration of AAV-CaMKII-Cre vector into wild-type mice is not sufficient to alter intestinal transit time. n = 3 (wild type), 4 (WT + AAV) or 3 (Ahr fl/fl) mice. Statistical test is a two-sided non-parametric Mann–Whitney U-test. Scale bars, 50 μm.

Source data

Supplementary information

Reporting Summary

41586_2020_1975_MOESM2_ESM.xlsx

Supplementary Table 1 SPF colon-upregulated ENS genes (SPF-CUEGs). 254 genes expressed higher in colon relative to SI neurons in SPF mice. The genes validated by RNAscope experiment are highlighted in yellow.

41586_2020_1975_MOESM3_ESM.xlsx

Supplementary Table 2 GF colon-upregulated ENS genes (GF-CUEGs). 122 genes expressed higher in colon neurons relative to SI neurons in GF mice. The genes validated by RNAscope experiment are highlighted in yellow.

41586_2020_1975_MOESM4_ESM.xlsx

Supplementary Table 3 Microbiota-dependent colon-upregulated ENS genes (M-CUEGs). 25 genes expressed higher in colonic neurons from SPF relative to GF mice. The genes validated by RNAscope experiment are highlighted in yellow. Kcnj12 identified by marginal relaxation of the p-value criteria (p<0.06) is highlighted in green.

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Obata, Y., Castaño, Á., Boeing, S. et al. Neuronal programming by microbiota regulates intestinal physiology. Nature 578, 284–289 (2020). https://doi.org/10.1038/s41586-020-1975-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-020-1975-8

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