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 network disturbances in relapsing refractory Crohn’s disease

A Publisher Correction to this article was published on 07 March 2019

This article has been updated

Abstract

Inflammatory bowel diseases (IBD) can be broadly divided into Crohn’s disease (CD) and ulcerative colitis (UC) from their clinical phenotypes. Over 150 host susceptibility genes have been described, although most overlap between CD, UC and their subtypes, and they do not adequately account for the overall incidence or the highly variable severity of disease. Replicating key findings between two long-term IBD cohorts, we have defined distinct networks of taxa associations within intestinal biopsies of CD and UC patients. Disturbances in an association network containing taxa of the Lachnospiraceae and Ruminococcaceae families, typically producing short chain fatty acids, characterize frequently relapsing disease and poor responses to treatment with anti-TNF-α therapeutic antibodies. Alterations of taxa within this network also characterize risk of later disease recurrence of patients in remission after the active inflamed segment of CD has been surgically removed.

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: Distinct microbial features associated with IBD patients.
Fig. 2: Microbial co-occurrence analysis of IBD and non-IBD subjects with gastrointestinal ecosystems.
Fig. 3: Clinical determinants correlate with intestinal microbiota composition of IBD patients.
Fig. 4: Intestinal microbiota composition of IBD patients with responsiveness to anti-TNF-α or corticosteroid therapies.
Fig. 5: Disease locations are critically important determinants that correlate with intestinal microbiota of IBD patients.
Fig. 6: Altered microbiota profiles in inactive CD patients following surgical resection of active disease segments.

Similar content being viewed by others

Data availability

All sequencing datasets from the current study have been deposited in a figshare repository and are publicly available. The Cohort 1 fasta file withthe mapping file are available at https://figshare.com/s/e9f2cffd0f0328ca5811 (https://doi.org/10.6084/m9.figshare.7335068) and the Cohort 2 fasta file with the mapping file are available at https://figshare.com/s/bbdd5dfb01e29484efa1 (https://doi.org/10.6084/m9.figshare.7335071). Associated codes for the analysis using R packages and QIIME can be found in these depositories. The genome-scale metabolic model script and dataset are available at https://figshare.com/s/a34f96698ca6fcd36ac2.

Change history

  • 07 March 2019

    Owing to an error during typesetting, a number of references were deleted from the Methods reference list. This altered all of the references in the Methods section and some of the references in Extended Data Fig. 5, making them inaccurate. References 121–134 were added back into the Methods reference list, and the references in the Methods section and in Extended Data Fig. 5 were renumbered accordingly. The error has been corrected in the PDF and HTML versions of this article.

References

  1. Strober, W., Fuss, I. J. & Blumberg, R. S. The immunology of mucosal models of inflammation. Annu. Rev. Immunol. 20, 495–549 (2002).

    Article  CAS  PubMed  Google Scholar 

  2. Blumberg, R. & Powrie, F. Microbiota, disease, and back to health: a metastable journey. Sci Transl Med 4, 137rv137 (2012).

    Article  CAS  Google Scholar 

  3. Hegazy, A. N. et al. Circulating and tissue-resident CD4(+) T cells with reactivity to intestinal microbiota are abundant in healthy individuals and function is altered during inflammation. Gastroenterology 153, 1320–1337 e1316 (2017).

    Article  CAS  PubMed  Google Scholar 

  4. Maloy, K. J. & Powrie, F. Intestinal homeostasis and its breakdown in inflammatory bowel disease. Nature 474, 298–306 (2011).

    Article  CAS  PubMed  Google Scholar 

  5. Rutgeerts, P. et al. Effect of faecal stream diversion on recurrence of Crohn’s disease in the neoterminal ileum. Lancet 338, 771–774 (1991).

    Article  CAS  PubMed  Google Scholar 

  6. Zachos, M., Tondeur, M. & Griffiths, A. M. Enteral nutritional therapy for induction of remission in Crohn’s disease.Cochrane Database Syst. Rev. 4, CD000542 (2007).

    Google Scholar 

  7. Cummings, J. H. & Kong, S. C. Probiotics, prebiotics and antibiotics in inflammatory bowel disease. Novartis. Found. Symp. 263, 99–111 (2004). discussion 111–114, 211–118.

    CAS  PubMed  Google Scholar 

  8. Gecse, K. B. et al. A global consensus on the classification, diagnosis and multidisciplinary treatment of perianal fistulising Crohn’s disease. Gut 63, 1381–1392 (2014).

    Article  PubMed  Google Scholar 

  9. Faith, J. J. et al. The long-term stability of the human gut microbiota. Science 341, 1237439 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Arumugam, M. et al. Enterotypes of the human gut microbiome. Nature 473, 174–180 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Schloissnig, S. et al. Genomic variation landscape of the human gut microbiome. Nature 493, 45–50 (2013).

    Article  CAS  PubMed  Google Scholar 

  12. Flint, H. J., Duncan, S. H. & Louis, P. The impact of nutrition on intestinal bacterial communities. Curr. Opin. Microbiol. 38, 59–65 (2017).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  14. 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 

  15. Jostins, L. et al. Host-microbe interactions have shaped the genetic architecture of inflammatory bowel disease. Nature 491, 119–124 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Zuk, O., Hechter, E., Sunyaev, S. R. & Lander, E. S. The mystery of missing heritability: genetic interactions create phantom heritability. Proc. Natl Acad. Sci. USA 109, 1193–1198 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Uhlig, H. H. Monogenic diseases associated with intestinal inflammation: implications for the understanding of inflammatory bowel disease. Gut 62, 1795–1805 (2013).

    Article  CAS  PubMed  Google Scholar 

  18. Lee, J. C. & Lennard-Jones, J. E. Inflammatory bowel disease in 67 families each with three or more affected first-degree relatives. Gastroenterology 111, 587–596 (1996).

    Article  CAS  PubMed  Google Scholar 

  19. Lee, J. C. et al. Gene expression profiling of CD8+ T cells predicts prognosis in patients with Crohn disease and ulcerative colitis. J. Clin. Invest. 121, 4170–4179 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Lee, J. C. et al. Genome-wide association study identifies distinct genetic contributions to prognosis and susceptibility in Crohn’s disease. Nat. Genet. 49, 262–268 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Park, S. H., Aniwan, S. & Loftus, E. V. Jr. Advances in the use of biologics and other novel drugs for managing inflammatory bowel disease. Curr. Opin. Pharmacol. 37, 65–71 (2017).

    Article  CAS  PubMed  Google Scholar 

  22. Pillai, N., Dusheiko, M., Burnand, B. & Pittet, V. A systematic review of cost-effectiveness studies comparing conventional, biological and surgical interventions for inflammatory bowel disease. PLoS ONE 12, e0185500 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Engels, M., Cross, R. K. & Long, M. D. Exercise in patients with inflammatory bowel diseases: current perspectives. Clin. Exp. Gastroenterol. 11, 1–11 (2018).

    Article  PubMed  Google Scholar 

  24. Eckburg, P. B. et al. Diversity of the human intestinal microbial flora. Science 308, 1635–1638 (2005).

    Article  PubMed  PubMed Central  Google Scholar 

  25. Sartor, R. B. & Wu, G. D. Roles for intestinal bacteria, viruses, and fungi in pathogenesis of inflammatory bowel diseases and therapeutic approaches.Gastroenterology 152, 327–339.e4 (2017).

    Article  CAS  PubMed  Google Scholar 

  26. Png, C. W. et al. Mucolytic bacteria with increased prevalence in IBD mucosa augment in vitro utilization of mucin by other bacteria. Am. J. Gastroenterol. 105, 2420–2428 (2010).

    Article  CAS  PubMed  Google Scholar 

  27. Sokol, H. et al. Faecalibacterium prausnitzii is an anti-inflammatory commensal bacterium identified by gut microbiota analysis of Crohn disease patients. Proc. Natl Acad. Sci. USA 105, 16731–16736 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Magnusdottir, S. et al. Generation of genome-scale metabolic reconstructions for 773 members of the human gut microbiota. Nat. Biotechnol. 35, 81–89 (2017).

    Article  CAS  PubMed  Google Scholar 

  29. Fujimoto, T. et al. Decreased abundance of Faecalibacterium prausnitzii in the gut microbiota of Crohn’s disease. J. Gastroenterol. Hepatol. 28, 613–619 (2013).

    Article  CAS  PubMed  Google Scholar 

  30. Lopez-Siles, M. et al. Mucosa-associated Faecalibacterium prausnitzii phylotype richness is reduced in patients with inflammatory bowel disease. Appl. Environ. Microbiol. 81, 7582–7592 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Imhann, F. et al. Interplay of host genetics and gut microbiota underlying the onset and clinical presentation of inflammatory bowel disease. Gut 67, 108–119 (2018).

    Article  CAS  PubMed  Google Scholar 

  32. Goodman, A. L. & Gordon, J. I. Our unindicted coconspirators: human metabolism from a microbial perspective. Cell. Metab. 12, 111–116 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. 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 

  34. Mainali, K. P. et al. Statistical analysis of co-occurrence patterns in microbial presence-absence datasets. PLoS ONE 12, e0187132 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Williams, R. J., Howe, A. & Hofmockel, K. S. Demonstrating microbial co-occurrence pattern analyses within and between ecosystems. Front. Microbiol. 5, 358 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  36. Rahnavard, G. F. et al. HAllA: Hierarchical All-against-All significance testing. http://huttenhower.sph.harvard.edu/halla (2017).

  37. Plevinsky, J. M., Wojtowicz, A. A., Poulopoulos, N., Schneider, K. L. & Greenley, R. N. Perceived impairment in sports participation in adolescents with inflammatory bowel diseases: a preliminary examination. J. Pediatr. Gastroenterol. Nutr. 66, 79–83 (2018).

    Article  PubMed  Google Scholar 

  38. Torres, J., Mehandru, S., Colombel, J. F. & Peyrin-Biroulet, L. Crohn’s disease. Lancet 389, 1741–1755 (2017).

    Article  PubMed  Google Scholar 

  39. Khanna, R. et al. Early combined immunosuppression for the management of Crohn’s disease (REACT): a cluster randomised controlled trial. Lancet 386, 1825–1834 (2015).

    Article  PubMed  Google Scholar 

  40. Ungaro, R., Mehandru, S., Allen, P. B., Peyrin-Biroulet, L. & Colombel, J. F. Ulcerative colitis. Lancet 389, 1756–1770 (2017).

    Article  PubMed  Google Scholar 

  41. Magnusson, M. K. et al. Anti-TNF therapy response in patients with ulcerative colitis is associated with colonic antimicrobial peptide expression and microbiota composition. J. Crohns Colitis 10, 943–952 (2016).

    Article  PubMed  Google Scholar 

  42. Ghouri, Y. A. et al. Systematic review of randomized controlled trials of probiotics, prebiotics, and synbiotics in inflammatory bowel disease. Clin. Exp. Gastroenterol. 7, 473–487 (2014).

    PubMed  PubMed Central  Google Scholar 

  43. Topping, D. L. & Clifton, P. M. Short-chain fatty acids and human colonic function: roles of resistant starch and nonstarch polysaccharides. Physiol. Rev. 81, 1031–1064 (2001).

    Article  CAS  PubMed  Google Scholar 

  44. 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 

  45. Wu, F. et al. Phascolarctobacterium faecium abundant colonization in human gastrointestinal tract. Exp. Ther. Med. 14, 3122–3126 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Chen, J. et al. Dysbiosis of intestinal microbiota and decrease in paneth cell antimicrobial peptide level during acute necrotizing pancreatitis in rats. PLoS ONE 12, e0176583 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Hancock, L. & Mortensen, N. J. How often do IBD patients require resection of their intestine? Inflamm. Bowel Dis. 14, S68–S69 (2008).

    Article  PubMed  Google Scholar 

  48. Olivera, P., Spinelli, A., Gower-Rousseau, C., Danese, S. & Peyrin-Biroulet, L. Surgical rates in the era of biological therapy: up, down or unchanged? Curr. Opin. Gastroenterol. 33, 246–253 (2017).

    Article  PubMed  Google Scholar 

  49. Rutgeerts, P. et al. Predictability of the postoperative course of Crohn’s disease. Gastroenterology 99, 956–963 (1990).

    Article  CAS  PubMed  Google Scholar 

  50. Patterson, A. M. et al. Human gut symbiont roseburia hominis promotes and regulates innate immunity. Front. Immunol. 8, 1166 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Machiels, K. et al. Specific members of the predominant gut microbiota predict pouchitis following colectomy and IPAA in UC. Gut 66, 79–88 (2017).

    Article  PubMed  Google Scholar 

  52. Machiels, K. et al. A decrease of the butyrate-producing species Roseburia hominis and Faecalibacterium prausnitzii defines dysbiosis in patients with ulcerative colitis. Gut 63, 1275–1283 (2014).

    Article  CAS  PubMed  Google Scholar 

  53. Flint, H. J., Duncan, S. H., Scott, K. P. & Louis, P. Links between diet, gut microbiota composition and gut metabolism. Proc. Nutr. Soc. 74, 13–22 (2015).

    Article  CAS  PubMed  Google Scholar 

  54. Byndloss, M. X. et al. Microbiota-activated PPAR-gamma signaling inhibits dysbiotic Enterobacteriaceae expansion. Science 357, 570–575 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Liguori, G. et al. Fungal dysbiosis in mucosa-associated microbiota of Crohn’s disease patients. J. Crohns Colitis 10, 296–305 (2016).

    Article  PubMed  Google Scholar 

  56. Papa, E. et al. Non-invasive mapping of the gastrointestinal microbiota identifies children with inflammatory bowel disease. PLoS ONE 7, e39242 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Santoru, M. L. et al. Cross-sectional evaluation of the gut-microbiome metabolome axis in an Italian cohort of IBD patients. Sci. Rep. 7, 9523 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Pedamallu, C. S. et al. Metagenomic characterization of microbial communities in situ within the deeper layers of the ileum in crohn’s disease. Cell. Mol. Gastroenterol. Hepatol. 2, 563–566 e565 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  59. Kellermayer, R. et al. Microbiota separation and C-reactive protein elevation in treatment-naive pediatric granulomatous Crohn disease. J. Pediatr. Gastroenterol. Nutr. 55, 243–250 (2012).

    Article  CAS  PubMed  Google Scholar 

  60. Frank, D. N. et al. Molecular-phylogenetic characterization of microbial community imbalances in human inflammatory bowel diseases. Proc. Natl Acad. Sci. USA 104, 13780–13785 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Willing, B. P. et al. A pyrosequencing study in twins shows that gastrointestinal microbial profiles vary with inflammatory bowel disease phenotypes. Gastroenterology 139, 1844–1854 e1841 (2010).

    Article  PubMed  Google Scholar 

  62. Wang, W. et al. Increased proportions of Bifidobacterium and the Lactobacillus group and loss of butyrate-producing bacteria in inflammatory bowel disease. J. Clin. Microbiol. 52, 398–406 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Rajilic-Stojanovic, M., Shanahan, F., Guarner, F. & de Vos, W. M. Phylogenetic analysis of dysbiosis in ulcerative colitis during remission. Inflamm. Bowel Dis. 19, 481–488 (2013).

    Article  PubMed  Google Scholar 

  64. Seksik, P. et al. Alterations of the dominant faecal bacterial groups in patients with Crohn’s disease of the colon. Gut 52, 237–242 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Halfvarson, J. et al. Dynamics of the human gut microbiome in inflammatory bowel disease. Nat. Microbiol. 2, 17004 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Erickson, A. R. et al. Integrated metagenomics/metaproteomics reveals human host-microbiota signatures of Crohn’s disease. PLoS ONE 7, e49138 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Eun, C. S. et al. Does the intestinal microbial community of Korean Crohn’s disease patients differ from that of western patients? BMC. Gastroenterol. 16, 28 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Forbes, J. D., Van Domselaar, G. & Bernstein, C. N. Microbiome survey of the inflamed and noninflamed gut at different compartments within the gastrointestinal tract of inflammatory bowel disease patients. Inflamm. Bowel Dis. 22, 817–825 (2016).

    Article  PubMed  Google Scholar 

  69. Andoh, A. et al. Multicenter analysis of fecal microbiota profiles in Japanese patients with Crohn’s disease. J. Gastroenterol. 47, 1298–1307 (2012).

    Article  PubMed  Google Scholar 

  70. Rehman, A. et al. Transcriptional activity of the dominant gut mucosal microbiota in chronic inflammatory bowel disease patients. J. Med. Microbiol. 59, 1114–1122 (2010).

    Article  CAS  PubMed  Google Scholar 

  71. Martinez-Medina, M., Aldeguer, X., Gonzalez-Huix, F., Acero, D. & Garcia-Gil, L. J. Abnormal microbiota composition in the ileocolonic mucosa of Crohn’s disease patients as revealed by polymerase chain reaction-denaturing gradient gel electrophoresis. Inflamm. Bowel Dis. 12, 1136–1145 (2006).

    Article  PubMed  Google Scholar 

  72. Norman, J. M. et al. Disease-specific alterations in the enteric virome in inflammatory bowel disease. Cell 160, 447–460 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Ashton, J. J. et al. 16S sequencing and functional analysis of the fecal microbiome during treatment of newly diagnosed pediatric inflammatory bowel disease. Medicine. 96, e7347 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Kabeerdoss, J., Jayakanthan, P., Pugazhendhi, S. & Ramakrishna, B. S. Alterations of mucosal microbiota in the colon of patients with inflammatory bowel disease revealed by real time polymerase chain reaction amplification of 16 S ribosomal ribonucleic acid. Indian J. Med. Res. 142, 23–32 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  75. Kolho, K. L. et al. Fecal microbiota in pediatric inflammatory bowel disease and its relation to inflammation. Am. J. Gastroenterol. 110, 921–930 (2015).

    Article  PubMed  Google Scholar 

  76. Martinez-Medina, M. et al. Molecular diversity of Escherichia coli in the human gut: new ecological evidence supporting the role of adherent-invasive E. coli (AIEC) in Crohn’s disease. Inflamm. Bowel Dis. 15, 872–882 (2009).

    Article  PubMed  Google Scholar 

  77. Duboc, H. et al. Connecting dysbiosis, bile-acid dysmetabolism and gut inflammation in inflammatory bowel diseases. Gut 62, 531–539 (2013).

    Article  CAS  PubMed  Google Scholar 

  78. Schwiertz, A. et al. Microbiota in pediatric inflammatory bowel disease. J. Pediatr. 157, 240–244 e241 (2010).

    Article  PubMed  Google Scholar 

  79. Kaakoush, N. O. et al. Microbial dysbiosis in pediatric patients with Crohn’s disease. J. Clin. Microbiol. 50, 3258–3266 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  80. Walker, A. W. et al. High-throughput clone library analysis of the mucosa-associated microbiota reveals dysbiosis and differences between inflamed and non-inflamed regions of the intestine in inflammatory bowel disease. BMC. Microbiol. 11, 7 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  81. Gophna, U., Sommerfeld, K., Gophna, S., Doolittle, W. F. & van Zanten, S. J. O. V. Differences between tissue-associated intestinal microfloras of patients with Crohn’s disease and ulcerative colitis. J. Clin. Microbiol. 44, 4136–4141 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Knoll, R. L. et al. Gut microbiota differs between children with Inflammatory Bowel Disease and healthy siblings in taxonomic and functional composition: a metagenomic analysis. Am. J. Physiol-Gastr. L 312, G327–G339 (2017).

    Google Scholar 

  83. Hansen, R. et al. Microbiota of de-novo pediatric IBD: increased Faecalibacterium prausnitzii and reduced bacterial diversity in Crohn’s but not in ulcerative colitis. Am. J. Gastr. 107, 1913–1922 (2012).

    Article  CAS  Google Scholar 

  84. Jacobs, J. P. et al. A disease-associated microbial and metabolomics state in relatives of pediatric inflammatory bowel disease patients. Cell. Mol. Gastroenterol. Hepatol. 2, 750–766 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  85. Bajer, L. et al. Distinct gut microbiota profiles in patients with primary sclerosing cholangitis and ulcerative colitis. World J. Gastr. 23, 4548–4558 (2017).

    Article  CAS  Google Scholar 

  86. Baumgart, M. et al. Culture independent analysis of ileal mucosa reveals a selective increase in invasive Escherichia coli of novel phylogeny relative to depletion of Clostridiales in Crohn’s disease involving the ileum. ISME. J. 1, 403–418 (2007).

    Article  CAS  PubMed  Google Scholar 

  87. Gevers, D. et al. The treatment-naive microbiome in new-onset Crohn’s disease. Cell. Host. Microbe. 15, 382–392 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Tong, M. et al. A modular organization of the human intestinal mucosal microbiota and its association with inflammatory bowel disease. PLoS ONE 8, e80702 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Pascal, V. et al. A microbial signature for Crohn’s disease. Gut 66, 813–822 (2017).

    Article  CAS  PubMed  Google Scholar 

  90. Mottawea, W. et al. Altered intestinal microbiota-host mitochondria crosstalk in new onset Crohn’s disease. Nat. Commun. 7, 13419 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Morgan, X. C. et al. Dysfunction of the intestinal microbiome in inflammatory bowel disease and treatment. Genome. Biol. 13, R79 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Verma, R., Verma, A. K., Ahuja, V. & Paul, J. Real-time analysis of mucosal flora in patients with inflammatory bowel disease in India. J. Clin. Microbiol. 48, 4279–4282 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  93. Arun Gupta, S. K., Wagner, Josef, Kirkwood, Carl, Morrison, Mark & McSweeney, Chris and finlay macrae. analysis of mucosal microbiota in in ammatory bowel disease using a custom phylogenetic microarray. Austin J. Gastr. 1, 1–6 (2014).

    Google Scholar 

  94. Bibiloni, R., Mangold, M., Madsen, K. L., Fedorak, R. N. & Tannock, G. W. The bacteriology of biopsies differs between newly diagnosed, untreated, Crohn’s disease and ulcerative colitis patients. J. Med. Microbiol. 55, 1141–1149 (2006).

    Article  PubMed  Google Scholar 

  95. Sokol, H. et al. Low counts of Faecalibacterium prausnitzii in colitis microbiota. Inflamm. Bowel Dis. 15, 1183–1189 (2009).

    Article  CAS  PubMed  Google Scholar 

  96. Andoh, A. et al. Comparison of the fecal microbiota profiles between ulcerative colitis and Crohn’s disease using terminal restriction fragment length polymorphism analysis. J. Gastroenterol. 46, 479–486 (2011).

    Article  PubMed  Google Scholar 

  97. Swidsinski, A., Loening-Baucke, V., Vaneechoutte, M. & Doerffel, Y. Active Crohn’s disease and ulcerative colitis can be specifically diagnosed and monitored based on the biostructure of the fecal flora. Inflamm. Bowel Dis. 14, 147–161 (2008).

    Article  PubMed  Google Scholar 

  98. Nishino, K. et al. Analysis of endoscopic brush samples identified mucosa-associated dysbiosis in inflammatory bowel disease. J. Gastroenterol. 53, 95–106 (2017).

    Article  PubMed  Google Scholar 

  99. Chen, L. et al. Characteristics of fecal and mucosa-associated microbiota in Chinese patients with inflammatory bowel disease. Medicine (Baltimore) 93, e51 (2014).

    Article  CAS  Google Scholar 

  100. Suchodolski, J. S., Dowd, S. E., Wilke, V., Steiner, J. M. & Jergens, A. E. 16S rRNA gene pyrosequencing reveals bacterial dysbiosis in the duodenum of dogs with idiopathic inflammatory bowel disease. PLoS ONE 7, e39333 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Xenoulis, P. G. et al. Molecular-phylogenetic characterization of microbial communities imbalances in the small intestine of dogs with inflammatory bowel disease. FEMS Microbiol. Ecol. 66, 579–589 (2008).

    Article  CAS  PubMed  Google Scholar 

  102. Omori, M. et al. Fecal microbiome in dogs with inflammatory bowel disease and intestinal lymphoma. J. Vet. Med. Sci. 79, 1840–1847 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Allenspach, K. et al. Evaluation of mucosal bacteria and histopathology, clinical disease activity and expression of Toll-like receptors in German shepherd dogs with chronic enteropathies. Vet. Microbiol. 146, 326–335 (2010).

    Article  CAS  PubMed  Google Scholar 

  104. Suchodolski, J. S., Xenoulis, P. G., Paddock, C. G., Steiner, J. M. & Jergens, A. E. Molecular analysis of the bacterial microbiota in duodenal biopsies from dogs with idiopathic inflammatory bowel disease. Vet. Microbiol. 142, 394–400 (2010).

    Article  CAS  PubMed  Google Scholar 

  105. Suchodolski, J. S. et al. The fecal microbiome in dogs with acute diarrhea and idiopathic inflammatory bowel disease. PLoS ONE 7, e51907 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Rossi, G. et al. Comparison of microbiological, histological, and immunomodulatory parameters in response to treatment with either combination therapy with prednisone and metronidazole or probiotic VSL#3 strains in dogs with idiopathic inflammatory bowel disease. PLoS ONE 9, e94699 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Minamoto, Y. et al. Alteration of the fecal microbiota and serum metabolite profiles in dogs with idiopathic inflammatory bowel disease. Gut Microbes 6, 33–47 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Jergens, A. E. Feline idiopathic inflammatory bowel disease: what we know and what remains to be unraveled. J. Feline. Med. Surg. 14, 445–458 (2012).

    Article  PubMed  Google Scholar 

  109. Vazquez-Baeza, Y., Hyde, E. R., Suchodolski, J. S. & Knight, R. Dog and human inflammatory bowel disease rely on overlapping yet distinct dysbiosis networks. Nat. Microbiol. 1, 16177 (2016).

    Article  CAS  PubMed  Google Scholar 

  110. Inness, V. L., McCartney, A. L., Khoo, C., Gross, K. L. & Gibson, G. R. Molecular characterisation of the gut microflora of healthy and inflammatory bowel disease cats using fluorescence in situ hybridisation with special reference to Desulfovibrio spp. J. Anim. Physiol. Anim. Nutr. 91, 48–53 (2007).

    Article  CAS  Google Scholar 

  111. Elinav, E. et al. NLRP6 inflammasome regulates colonic microbial ecology and risk for colitis. Cell 145, 745–757 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Li, M. et al. Upregulation of intestinal barrier function in mice with DSS-induced colitis by a defined bacterial consortium is associated with expansion of IL-17A producing gamma delta T cells. Front. Immunol. 8, 824 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Rooks, M. G. et al. Gut microbiome composition and function in experimental colitis during active disease and treatment-induced remission. ISME. J. 8, 1403–1417 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Simpson, K. W. et al. Adherent and invasive Escherichia coli is associated with granulomatous colitis in boxer dogs. Infect. Immun. 74, 4778–4792 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Janeczko, S. et al. The relationship of mucosal bacteria to duodenal histopathology, cytokine mRNA, and clinical disease activity in cats with inflammatory bowel disease. Vet. Microbiol. 128, 178–193 (2008).

    Article  CAS  PubMed  Google Scholar 

  116. Larmonier, C. B. et al. Reduced colonic microbial diversity is associated with colitis in NHE3-deficient mice. Am. J. Physiol. Gastrointest. Liver Physiol. 305, G667–G677 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Spalinger, M. R. et al. PTPN2 controls differentiation of CD4(+) T cells and limits intestinal inflammation and intestinal dysbiosis. Mucosal Immunol 8, 918–929 (2015).

    Article  CAS  PubMed  Google Scholar 

  118. Munyaka, P. M., Rabbi, M. F., Khafipour, E. & Ghia, J. E. Acute dextran sulfate sodium (DSS)-induced colitis promotes gut microbial dysbiosis in mice. J. Basic Microbiol. 56, 986–998 (2016).

    Article  CAS  PubMed  Google Scholar 

  119. Okayasu, I. et al. A novel method in the induction of reliable experimental acute and chronic ulcerative colitis in mice. Gastroenterology 98, 694–702 (1990).

    Article  CAS  PubMed  Google Scholar 

  120. Robinson, A. M. et al. Alterations of colonic function in the Winnie mouse model of spontaneous chronic colitis. Am. J. Physiol. Gastrointest. Liver Physiol. 312, G85–G102 (2017).

    Article  PubMed  Google Scholar 

  121. Pittet, V. et al. Cohort profile: the Swiss Inflammatory Bowel Disease Cohort Study (SIBDCS). Int. J. Epidemiol. 38, 922–931 (2009).

    Article  PubMed  Google Scholar 

  122. Lennard-Jones, J. E. Classification of inflammatory bowel disease. Scand. J. Gastroenterol. Suppl. 170, 2–6, discussion 16–19 (1989).

  123. Harris, P. A. et al. Research Electronic Data Capture (REDCap)—a metadata-driven methodology and workflow process for providing translational research informatics support. J. Biomed. Inform. 42, 377–381 (2009).

    Article  PubMed  Google Scholar 

  124. Sundquist, A. et al. Bacterial flora-typing with targeted, chip-based Pyrosequencing. BMC Microbiol. 7, 108 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Yilmaz, B. et al. The presence of genetic risk variants within PTPN2 and PTPN22 is associated with intestinal microbiota alterations in Swiss IBD cohort patients. PLoS One 13, e0199664 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Whiteley, A. S. et al. Microbial 16S rRNA Ion Tag and community metagenome sequencing using the Ion Torrent (PGM) Platform. J. Microbiol. Methods 91, 80–88 (2012).

    Article  CAS  PubMed  Google Scholar 

  127. Caporaso, J. G. et al. QIIME allows analysis of high-throughput community sequencing data. Nat. Methods 7, 335–336 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. McMurdie, P. J. & Holmes, S. phyloseq: an R package for reproducible interactive analysis and graphics of microbiome census data. PLoS One 8, e61217 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Callahan, B. J., Sankaran, K., Fukuyama, J. A., McMurdie, P. J. & Holmes, S. P. Bioconductor Workflow for Microbiome Data Analysis: from raw reads to community analyses. F1000Res 5, 1492 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  130. Edgar, R. C. Search and clustering orders of magnitude faster than BLAST. Bioinformatics 26, 2460–2461 (2010).

    Article  CAS  PubMed  Google Scholar 

  131. Good, I. J. The population frequencies of species and the estimation of population parameters. Biometrika 40, 237–264 (1953).

    Article  Google Scholar 

  132. Su, G., Kuchinsky, A., Morris, J. H., States, D. J. & Meng, F. GLay: community structure analysis of biological networks. Bioinformatics 26, 3135–3137 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Benjamini, Y. & Hochberg, Y. Controlling the false discovery rate - a practical and powerful approach to multiple testing. J. R. Stat. Soc. Series B Stat. Methodol. 57, 289–300 (1995).

    Google Scholar 

  134. Yang, I. et al. Intestinal microbiota composition of interleukin-10 deficient C57BL/6J mice and susceptibility to Helicobacter hepaticus-induced colitis. PLoS ONE 8, e70783 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Ren, Y. et al. Polysaccharide of Hericium erinaceus attenuates colitis in C57BL/6 mice via regulation of oxidative stress, inflammation-related signaling pathways and modulating the composition of the gut microbiota. J. Nutr. Biochem. 57, 67–76 (2018).

    Article  CAS  PubMed  Google Scholar 

  136. Nunberg, M. et al. Interleukin 1alpha-deficient mice have an altered gut microbiota leading to protection from dextran sodium sulfate-induced colitis. mSystems 3, e00213–17 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  137. Schaubeck, M. et al. Dysbiotic gut microbiota causes transmissible Crohn’s disease-like ileitis independent of failure in antimicrobial defence. Gut 65, 225–237 (2016).

    Article  CAS  PubMed  Google Scholar 

  138. Lupp, C. et al. Host-mediated inflammation disrupts the intestinal microbiota and promotes the overgrowth of Enterobacteriaceae. Cell. Host. Microbe. 2, 204 (2007).

    Article  CAS  PubMed  Google Scholar 

  139. Nones, K. et al. Multidrug resistance gene deficient (mdr1a−/−) mice have an altered caecal microbiota that precedes the onset of intestinal inflammation. J. Appl. Microbiol. 107, 557–566 (2009).

    Article  CAS  PubMed  Google Scholar 

  140. Manchester, A. C. et al. Association between granulomatous colitis in french bulldogs and invasive Escherichia coli and response to fluoroquinolone antimicrobials. J. Vet. Intern. Med. 27, 56–61 (2013).

    Article  CAS  PubMed  Google Scholar 

  141. Roy, U. et al. Distinct microbial communities trigger colitis development upon intestinal barrier damage via innate or adaptive immune cells. Cell Rep 21, 994–1008 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  142. Johnston, D. G. W. et al. Loss of microRNA-21 influences the gut microbiota causing reduced susceptibility in a murine model of colitis. J. Crohns Colitis 12, 835–848 (2018).

    Article  PubMed  Google Scholar 

  143. Berry, D. et al. Phylotype-level 16S rRNA analysis reveals new bacterial indicators of health state in acute murine colitis. ISME. J. 6, 2091–2106 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. Zhang, Q. et al. Accelerated dysbiosis of gut microbiota during aggravation of DSS-induced colitis by a butyrate-producing bacterium. Sci. Rep. 6, 27572 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Schwab, C. et al. Longitudinal study of murine microbiota activity and interactions with the host during acute inflammation and recovery. ISME. J. 8, 1101–1114 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Zimmermann, J. et al. The intestinal microbiota determines the colitis-inducing potential of T-bet-deficient Th cells in mice. Eur. J. Immunol. 48, 161–167 (2018).

    Article  CAS  PubMed  Google Scholar 

  147. Garrett, W. S. et al. Enterobacteriaceae act in concert with the gut microbiota to induce spontaneous and maternally transmitted colitis. Cell. Host. Microbe. 8, 292–300 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  148. Palm, N. W. et al. Immunoglobulin A coating identifies colitogenic bacteria in inflammatory bowel disease. Cell 158, 1000–1010 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  149. Maharshak, N. et al. Altered enteric microbiota ecology in interleukin 10-deficient mice during development and progression of intestinal inflammation. Gut Microbes 4, 316–324 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  150. Chassaing, B. et al. Dietary emulsifiers impact the mouse gut microbiota promoting colitis and metabolic syndrome. Nature 519, 92–96 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  151. Osaka, T. et al. Meta-analysis of fecal microbiota and metabolites in experimental colitic mice during the inflammatory and healing phases. Nutrients 9, 1329 (2017).

    Article  CAS  PubMed Central  Google Scholar 

  152. Perez-Munoz, M. E. et al. Discordance between changes in the gut microbiota and pathogenicity in a mouse model of spontaneous colitis. Gut Microbes 5, 286–295 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  153. Nagalingam, N. A., Kao, J. Y. & Young, V. B. Microbial ecology of the murine gut associated with the development of dextran sodium sulfate-induced colitis. Inflamm. Bowel Dis. 17, 917–926 (2011).

    Article  PubMed  Google Scholar 

  154. Alkadhi, S., Kunde, D., Cheluvappa, R., Randall-Demllo, S. & Eri, R. The murine appendiceal microbiome is altered in spontaneous colitis and its pathological progression. Gut Pathog. 6, 25 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  155. Carvalho, F. A. et al. Interleukin-1beta (IL-1beta) promotes susceptibility of Toll-like receptor 5 (TLR5) deficient mice to colitis. Gut. 61, 373–384 (2012).

    Article  CAS  PubMed  Google Scholar 

  156. Seregin, S. S. et al. NLRP6 protects Il10(−/−) mice from colitis by limiting colonization of Akkermansia muciniphila. Cell Rep. 19, 2174 (2017).

    Article  CAS  PubMed  Google Scholar 

  157. Selvanantham, T. et al. NKT cell-deficient mice harbor an altered microbiota that fuels intestinal inflammation during chemically induced colitis. J. Immunol. 197, 4464–4472 (2016).

    Article  CAS  PubMed  Google Scholar 

  158. Lamas, B. et al. CARD9 impacts colitis by altering gut microbiota metabolism of tryptophan into aryl hydrocarbon receptor ligands. Nat. Med. 22, 598 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  159. Madsen, K. L. et al. Antibiotic therapy attenuates colitis in interleukin 10 gene-deficient mice. Gastroenterology 118, 1094–1105 (2000).

    Article  CAS  PubMed  Google Scholar 

  160. Mitchell, J. et al. Colonic inhibition of phosphatase and tensin homolog increases colitogenic bacteria, causing development of colitis in Il10/ mice. Inflamm. Bowel Dis. 24, 1718–1732 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  161. Hakansson, A. et al. Immunological alteration and changes of gut microbiota after dextran sulfate sodium (DSS) administration in mice. Clin. Exp. Med. 15, 107–120 (2015).

    Article  CAS  PubMed  Google Scholar 

  162. Heimesaat, M. M. et al. Shift towards pro-inflammatory intestinal bacteria aggravates acute murine colitis via Toll-like receptors 2 and 4. PLoS ONE 2, e662 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  163. Wohlgemuth, S., Haller, D., Blaut, M. & Loh, G. Reduced microbial diversity and high numbers of one single Escherichia coli strain in the intestine of colitic mice. Environ. Microbiol. 11, 1562–1571 (2009).

    Article  PubMed  Google Scholar 

  164. Arthur, J. C. et al. Intestinal inflammation targets cancer-inducing activity of the microbiota. Science 338, 120–123 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  165. Schuppler, M., Lotzsch, K., Waidmann, M. & Autenrieth, I. B. An abundance of Escherichia coli is harbored by the mucosa-associated bacterial flora of interleukin-2-deficient mice. Infect. Immun. 72, 1983–1990 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  166. Hoentjen, F. et al. Antibiotics with a selective aerobic or anaerobic spectrum have different therapeutic activities in various regions of the colon in interleukin 10 gene deficient mice. Gut. 52, 1721–1727 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  167. Zenewicz, L. A. et al. IL-22 deficiency alters colonic microbiota to be transmissible and colitogenic. J. Immunol. 190, 5306–5312 (2013).

    Article  CAS  PubMed  Google Scholar 

  168. Dennis, K. L. et al. Adenomatous polyps are driven by microbe-instigated focal inflammation and are controlled by IL-10-producing T cells. Cancer Res. 73, 5905–5913 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  169. Abecia, L. H. L., Khoo, C., Frantz, N. & McCartney, A. Effects of a novel galactooligosaccharide on the faecal microbiota of healthy and inflammatory bowel disease. Int. J. Probiotic. Prebiotics 5, 61–68 (2010).

    Google Scholar 

  170. Xu, J. et al. Does canine inflammatory bowel disease influence gut microbial profile and host metabolism? BMC Vet. Res. 12, 114 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  171. Jones-Hall, Y. L., Kozik, A. & Nakatsu, C. Ablation of tumor necrosis factor is associated with decreased inflammation and alterations of the microbiota in a mouse model of inflammatory bowel disease. PLoS ONE 10, e0119441 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  172. Bel, S. et al. Reprogrammed and transmissible intestinal microbiota confer diminished susceptibility to induced colitis in TMF/ mice. Proc. Natl Acad. Sci. USA 111, 4964–4969 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  173. Bloom, S. M. et al. Commensal Bacteroides species induce colitis in host-genotype-specific fashion in a mouse model of inflammatory bowel disease. Cell. Host. Microbe. 9, 390–403 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  174. He, Q. et al. Dysbiosis of the fecal microbiota in the TNBS-induced Crohn’s disease mouse model. Appl. Microbiol. Biotechnol. 100, 4485–4494 (2016).

    Article  CAS  PubMed  Google Scholar 

  175. Couturier-Maillard, A. et al. NOD2-mediated dysbiosis predisposes mice to transmissible colitis and colorectal cancer. J. Clin. Invest. 123, 700–711 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  176. Ettreiki, C. et al. Juvenile ferric iron prevents microbiota dysbiosis and colitis in adult rodents. W. J. Gastr. 18, 2619–2629 (2012).

    Article  CAS  Google Scholar 

  177. Devkota, S. et al. Dietary-fat-induced taurocholic acid promotes pathobiont expansion and colitis in Il10−/− mice. Nature 487, 104–108 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  178. Zhang, Z. et al. Chlorogenic acid ameliorates experimental colitis by promoting growth of akkermansia in mice. Nutrients 9, 677 (2017).

    Article  CAS  PubMed Central  Google Scholar 

  179. Ghosh, S., Molcan, E., DeCoffe, D., Dai, C. & Gibson, D. L. Diets rich in n-6 PUFA induce intestinal microbial dysbiosis in aged mice. Br. J. Nutr. 110, 515–523 (2013).

    Article  CAS  PubMed  Google Scholar 

  180. Ye, J. et al. Bacteria and bacterial rRNA genes associated with the development of colitis in IL-10(−/−) mice. Inflamm. Bowel Dis. 14, 1041–1050 (2008).

    Article  PubMed  Google Scholar 

  181. Vijay-Kumar, M. et al. Metabolic syndrome and altered gut microbiota in mice lacking Toll-like receptor 5. Science 328, 228–231 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  182. Zhu, W. et al. Precision editing of the gut microbiota ameliorates colitis. Nature 553, 208–211 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  183. Moon, C., Stupp, G. S., Su, A. I. & Wolan, D. W. Metaproteomics of colonic microbiota unveils discrete protein functions among colitic mice and control groups.Proteomics 18, 1700391 (2018).

    Article  CAS  Google Scholar 

  184. Du, Z. et al. Development of gut inflammation in mice colonized with mucosa-associated bacteria from patients with ulcerative colitis. Gut Pathog. 7, 32 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  185. Bibiloni, R., Simon, M. A., Albright, C., Sartor, B. & Tannock, G. W. Analysis of the large bowel microbiota of colitic mice using PCR/DGGE. Lett. Appl. Microbiol. 41, 45–51 (2005).

    Article  CAS  PubMed  Google Scholar 

  186. Kudelka, M. R. et al. Cosmc is an X-linked inflammatory bowel disease risk gene that spatially regulates gut microbiota and contributes to sex-specific risk. Proc. Natl Acad. Sci. USA 113, 14787–14792 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  187. Vereecke, L. et al. A20 controls intestinal homeostasis through cell-specific activities. Nat. Commun. 5, 5103 (2014).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank all patients and the members of the Swiss IBD cohort and Bern cohort for their commitment. We also thank the staff of the University Hospital of Bern, Clinic of Visceral Medicine and Surgery, and the Bern City Hospitals led by F. Seibold and R. Tutuian for obtaining samples in Cohort 2. This research was supported by Systems X (GutX) to A.J.M. and J.S., and the Swiss IBD cohort (grant no. 33CS30-148422) to G.R., A.J.M and C.M. The founding institutions had no role in the study design, analysis or interpretation of the results. We thank G. Rahnavard and C. Huttenhower (Department of Biostatistics, Harvard T.H. Chan School of Public Health, USA) for their help in using the HAllA pipeline. We also thank J. Harrell Rieder, A. Suter, S. Brand, C. Mooser, W. Kwong Chung and J. Hugenschmidt for helping B.Y. during the process of sample preparation. We also thank G. Weingart (Department of Biostatistics, Harvard T.H. Chan School of Public Health, USA) for his enormous help in the optimization of MaAsLin running on the MacOS platform using R.

Author information

Authors and Affiliations

Authors

Consortia

Contributions

A.J.M. conceived, designed and supervised the study. B.Y. performed all the experiments, analyzed the data and wrote the manuscript with A.J.M. P.J. organized and collected the samples of the second cohort. P.J. F.D.B., Y.F., N.F. and M.G. were involved in data curation. O.O. and C.R. carried out metabolic reaction analysis, and J.S. supervised these analyses. A.J.M., P.J. P.M., C.M., V.E.H.P, M.H.M., G.R., R.W. and Swiss IBD cohort investigators acquired patient samples and detailed structured clinical phenotypes.

Corresponding author

Correspondence to Andrew J. Macpherson.

Ethics declarations

Competing interests

The authors declare no competing interests.

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 Unique microbial taxa identified as IBD signatures using unsupervised meta-analysis of published human IBD studies.

a, The heatmap shows significant microbial changes in CD, UC or IBD (CD + UC were combined as IBD) disease groups compared to non-IBD subjects. Each disease group (CD, UC or CD + UC) was compared independently to non-IBD and each color code reports the direction of microbial changes in each respective disease status against non-IBD subjects. Euclidean clustering was performed for sample annotations (vertical) including race/ethnicity, gender, median age, patient number, sample type, sequencing method and microbial taxa (horizontal) at different taxonomic ranks. Taxa in black bold label with asterisk demonstrate those findings verified by a subset of the findings in Fig. 1. Taxa in gray bold label with gray asterisk demonstrate findings in some of the studies verified by of the findings in Fig. 1 specifically, Lachnospiraceae family, Lachnospira, Coprococcus, Clostridiales order, Faecalibacterium, Ruminococcus, Roseburia and Ruminococcaceae family for Group 1, and Actinobacteria, Proteobacteria phyla and Enterobacteriaceae family for Group 2. 18 significant additional replicated taxa in the heterogeneous Group 3 are Bacteroidetes phylum and genera from this phylum including Bacteroides, Odoribacter, Butyricimonas, Parabacteroides, Sutterella, Prevotella (of Prevotellaceae), Prevotella (of Paraprevotellaceae) and Rikenellaceae family; also Firmicutes and genera from this phylum including Phascolarctobacterium, Dialister, Eubacterium∙∙∙ and Ruminococcus∙∙∙; Blautia, Collinsella and Bifidobacterium; Sutturella from Proteobacteria and Tenericutes. Underlined taxa are matching with Cluster CDA in Fig. 2. b, The heatmap (studies are in order as in a shows clinically relevant information collected through the studies analyzed in a. Recorded clinical phenotyping information in a given study is shown in green color (Identified) and the lack of clinical data is represented in white color (non-identified).

Extended Data Fig. 2 Microbial taxa comparison of CD with UC in published IBD studies.

The heatmap shows the comparison of microbial changes between CD and UC. Taxa higher in CD (lower in UC) are in red, while taxa higher in UC (lower CD) are in blue. Euclidean clustering was performed for sample annotations (vertical) including race/ethnicity, gender, median age, patient number, sample type and sequencing method and microbial taxa (horizontal) at different taxonomic ranks. Taxa names in bold demonstrate those findings verified by a subset of the findings in Fig. 1 and taxa names in gray bold shows the findings that partially match with our findings in Fig. 1.

Extended Data Fig. 3 The dominant bacterial phyla along the gastrointestinal tract of IBD patients and dysbiosis in IBD patients.

a,b, Distribution of predominant bacterial phylotypes along the cephalocaudal axis of the gut in CD, UC and non-IBD subjects of Cohort 1 (a) and Cohort 2 (b) are depicted after stratification according to the relative abundance of Firmicutes at each sampling site. The dominant bacterial phylotypes are Bacteroidetes (51.6% IBD Cohort 1; 56% IBD Cohort 2 and 56% non-IBD), Firmicutes (34.9, 29 and 25.7%, respectively) and Proteobacteria (9.1, 17 and 14%); with a smaller proportion of Fusobacteria (0.8, 0.4 and 1.1%), Actinobacteria (0.79, 0.46 and 0.33%) and Tenericutes (0.2, 0.04 and 0.09%). cf, Microbial composition differences between IBD patients and non-IBD subjects were identified by species richness (Observed OTUs, Shannon and Simpson indices) in Cohort 1 (c) and Cohort 2 (d) and microbiome clustering based on unweighted and weighted UniFrac PCoA metrics for Cohort 1 (e) and Cohort 2 (f). Box-and-whisker plots in c and d display first and third quartiles and whiskers are from each quartile to the minimum or maximum. g,h, Beta dispersion statistics were performed by analyzing the sampling distance to centroids for Cohort 1 (g) and Cohort 2 (h) and there is no significant differences between compared groups in g and h. i,j, Only significant taxa associated with CD or UC shown as relative abundance ratio in Cohort 1 (i) or Cohort 2 (j) were identified using MaAsLin pipeline with BH-FDR correction (q value). q < 0.05 was considered significant. Significant differences were determined by either non-parametric two-sided Mann–Whitney U-test (c,d,g,h) or Adonis test for multiple comparisons (e,f) and P < 0.05 was considered significant. Box-and-whisker plots in c,d,g,h display first and third quartiles and whiskers are from each quartile to the minimum or maximum. 494 CD and 447 UC samples in Cohort 1 and 230 CD, 195 UC and 770 non-IBD samples in Cohort 2 were used for analysis (aj).

Extended Data Fig. 4 Microbial taxa and functional metabolic subsystem differences in IBD patients.

a,b, Significant taxonomic differences are depicted as relative abundance ratios between CD and non-IBD samples (a) and between UC and non-IBD samples (b) identified in MaAsLin pipeline with BH-FDR correction. q < 0.05 was considered significant. c,d, Relative abundance of the most important matching OTUs were identified using machine learning algorithm and are depicted for Cohort 1 (c) and Cohort 2 (d) using notched box whisker showing first and third quartiles with median value. Each dot represents a single sample (c,d). e,f,h,i, After mapping OTUs to metabolic reactions, calculated the metabolic distances between all pairs of patients based on raw reaction counts are shown on PCA plots based on L2 distance of total reaction counts between UC and CD and boxplots show the respective coefficients PC1 and PC2 axis in Cohort 1 (e,f) and Cohort 2 (h,i). PC1 and PC2 are the first two principal components. g,j, The principal component analysis (PCA) analysis illustrates robust data at the metabolic reaction level: (60% variance explained by PC1/2 Blue indicates UC and red indicates CD patients and are shown using notched box whisker plots showing first and third quartiles with median value. Metabolic subsystems different between in CD and UC patients were identified in Cohort 1 (g) and in Cohort 2 (j). Similar significant metabolic pathway enrichment was detected in both cohorts. Red color represents the enrichment in CD and blue color represents the enrichment in UC. Box-and-whisker plots in g,j display first and third quartiles and whiskers are from each quartile to the minimum or maximum and possible outliers. Consistent metabolic subsystems increased in CD belonged to B-vitamin and LPS biosynthesis, heparan sulfate and chondroitin sulfate degradation and fatty acid oxidation. The BH-FDR was applied to correct for multiple testing and q < 0.05 between groups was considered significant (f,i). Fisher’s exact test was performed to determine if the subsystem was overrepresented among the significantly different reactions; subsystems with P < 0.05 were considered enriched (g,j). 494 CD and 447 UC samples in Cohort 1 and 230 CD, 195 UC and 770 non-IBD samples in Cohort 2 were used for analysis (ad).

Extended Data Fig. 5 Unique microbial taxa identified as IBD signatures across different species using unsupervised meta-analysis of the published IBD studies.

a, The Euclidean clustering of IBD patients and animal models of IBD including dogs/cats diagnosed with IBD and mice with genetically and/or chemically (DSS or TNBS) induced colitis was performed using information of taxa identified significantly changing between disease groups and is plotted using the categorical information of the study models such as according to race/ethnicity, gender, median age, species, subject number, sample type, sample size, sequencing method and experimental model of IBD induction. b, The Spearman correlation heatmap shows the correlation between 123 different human and animal IBD studies based on identified 96 differentially abundant microbial taxa that are characterized in a. The correlation values ranging from 0 to 1 show positive correlation (in red) and the values ranging from −1 to 0 show negative correlation (in blue) between compared IBD studies. c, Statistical information of data of 123 independent human and animal IBD studies in total (a,b) is shown on the same matrix. Non-parametric two-tailed Spearman correlation test was performed and P < 0.05 was considered significant. Green color shows significant correlation between taxa plotted in a.

Extended Data Fig. 6 Microbial stability over time in longitudinally studied IBD patients and correlation of intestinal inflammation with microbial abundance.

a, Biopsies were collected from 22 individuals in Cohort 1 and 12 individuals in Cohort 2 over several years (1–9 years and 0.25–2 years, respectively). Each row corresponds to the time course of an individual patient. The resulting data comprised 176 biopsy samples. b, PCoA on Bray–Curtis dissimilarity distance matrix for longitudinally collected (as shown in a) 77 CD and 44 UC samples in Cohort 1 and 49 CD and 14 UC samples in Cohort 2 are plotted. Each color in represents an individual IBD patient. Ellipsoids represent a 95% confidence interval surrounding each disease group. c, The relative abundance changes for Bacteroides, Firmicutes and Proteobacteria phyla in IBD patients are plotted based on their disease severity changes over time. The x axis shows the relative abundance difference of a given phylum compared to previous sampling time point. A value higher than zero indicates that the phylum increases and a value lower than zero indicates that the phylum decreases compared to the previous sampling time point. Disease severity worsening over time is labeled as ‘decreasing’, improving over time is labeled as ‘increasing’ and stable disease severity is identified as ‘steady’ on the y axis. d,e, Fecal calprotectin that is positively correlated with Enterobacteriaceae∙ and Klebsiella for 79 CD patients (d) and negatively correlated with Ruminococcus∙∙∙ and Prevotella in 42 UC patients are shown on continuous data plot generated in MaAsLin pipeline with q < 0.05 (e). Spearman’s rank correlation coefficient for taxa in CD: 0.284 and 0.147 and for taxa in UC: −0.322 and −0.2, respectively. Adonis test was used to determine significant differences between the distance matrix of each group (b). Data shown in b was not significant when longitudinal samples were compared for individuals (P > 0.05). However, significant microbial differences were only observed between patients (P < 0.05). Taxa significantly associated with disease severity and fecal calprotectin were identified in MaAsLin pipeline (c,d,e) with BH-FDR correction and significant taxa are plotted. The q < 0.05 was considered significant.

Extended Data Fig. 7 Microbial profile along the gut in IBD patients.

a,b, Species richness of samples collected along the gut including Ileum (I), right colon (RC), transverse colon (CT), left colon (CL) and rectum (R) were calculated with Shannon index for 494 CD and 447 UC samples shown in Cohort 1 (a) and 230 CD and 195 UC in Cohort 2 (b). c,d, Beta diversity of these samples are shown for Cohort 1 (c) and Cohort 2 (d). e,f, Samples collected from same patients cluster intra-individually, as depicted CD (e) and UC (f) patients. gj, Species richness calculated with Shannon and Simpson indices for 494 CD and 447 UC samples shown in Cohort 1 and 230 CD and 195 UC in Cohort 2 with different inflammation status are shown in g for Cohort 1 and in h for Cohort 2. Beta diversity of these samples individually analyzed for CD and UC are shown for Cohort 1 (i) and Cohort 2 (j). Significant differences between groups were determined by one-way ANOVA corrected for multiple comparisons using BH-FDR and there is no significance between compared groups (q > 0.05) (a,b,g,h). Lines indicate mean values and error bars are standard deviations (a,b,g,h). Adonis test was used to determine significant differences between the dissimilarity distance matrix of each group and groups are not significantly different than each other (c,d,i,j). The edges strongly similar to each other are connected with a solid line (pure edge) and the edges partially similar to each other are connected with dashed lines (mixed edge) (e,f).

Extended Data Fig. 8 Co-occurrence patterns and degree centrality scores identify the important components of IBD.

a,b, Ecosystem-specific co-occurrence patterns are visualized using network diagrams where microbial phyla represent nodes and the presence of a positive co-occurrence relationship based on correlation is represented by an edge in Cohort 1 (a) and Cohort 2 (b). Co-occurrence relationships with less strong Spearman’s correlation coefficients (ρ value > 0.25 and P < 0.05) are depicted with network diagram for each disease. c,d, The value of eigenvector and betweenness centralities for CD, UC and non-IBD samples were calculated in Cohort 1 (c) and in Cohort 2 (d). Significant differences between groups were determined by non-parametric two-sided Mann–Whitney U-test (c) and ordinary one-way ANOVA corrected for multiple comparisons using BH-FDR correction (d) and P < 0.05 was considered significant and significant results are shown on the plot. Lines indicate mean values and error bars are standard deviations. 65 CD and 61 UC taxa from Cohort 1 and 48 CD, 44 UC and 41 non-IBD taxa from Cohort 2 were used for analysis in c and d. e,f, Important genera based on their between centrality score for Cohort 1 (e) and Cohort 2 (f) are depicted for each cohort (CD in red, UC in blue and Non-IBD in green). gj, Prominent and influential taxa were identified using in- and out-degree scores are shown for CD (g,i) and UC (h,j) for corresponding cohorts: Cohort 1 (g,h) and Cohort 2 (i,j). Taxa with ρ value > 0.25 and P < 0.05 are plotted (ej). (Cohort 1, CD phyla 9 nodes (N) and 13 edges (E); UC phyla 8 N and 12E; CD genera 64 N, 473E; UC genera, 60 N, 440E and Cohort 2: CD phyla 6 N, 7E; UC 9 N, 11E; CD genera 40 N, 273E; UC genera 38 N, 276E).

Extended Data Fig. 9 Gut microbiota differences in IBD patients with different lifestyles and different responsiveness to disease.

ad, Major taxonomic changes were observed in 494 CD samples in Cohort 1 when samples were analyzed for sport activities (a), smoking status (b), alcohol abuse (c) and family history of disease (d). eg, Species richness biopsy samples obtained from patients responsive (success) or unresponsive (failure) to anti-TNF-α therapies (e,f) and corticosteroid therapies (g,h) are shown in e,g for Cohort 1 and in f,h for Cohort 2. i,j, Microbial clustering of intestinal biopsy samples from IBD patients responding or non-responding to corticosteroids therapies for Cohort 1 (i) and Cohort 2 (j) is shown with PCoA on Bray–Curtis distance dissimilarity metrics. CD (solid line) and UC (dashed line) are used to identify the disease groups on PCoA plot. km, Unique microbial taxa identified as a signature of responding and non-responding groups in CD (182 with success and 47 with failure) (k) and in UC (l) are shown for Cohort 1 (131 with success and 36 with failure) and in UC (146 with success and 16 with failure) for Cohort 2 (m). n,o, Species richness (of biopsy samples obtained from patients with different disease activities, characterized by the frequency of exacerbations (active) and remissions (quiescent) are shown for Cohort 1 (n) and for Cohort 2 (o). Data is not significant in n and o. p,q, Microbial clustering of intestinal biopsy samples from IBD patients with different disease activities for Cohort 1 (p) and Cohort 2 (q) is shown with PCoA on Bray–Curtis distance dissimilarity metrics. CD (solid line) and UC (dashed line) are used to identify the disease groups on PCoA plot. 494 samples in CD and 447 samples in UC for Cohort 1 (p) and 226 samples in CD and 195 samples in UC for Cohort 2 (q) were analyzed. Microbial profiles were analyzed using MaAsLin pipeline with BH-FDR correction (q value) and q < 0.05 was considered significant (ad,km). Significant taxa (ad,km) are plotted using notched box whisker showing first and third quartiles and median value. (a) Sport: actively (several times per week), sometimes (once or twice a week) and rarely (less than once a week); (c) alcohol abuse; (d) family history: N (None), Y (Yes). Mann–Whitney U-test was used for statistical analysis of alpha diversity (eh,n,o). No significant differences in species richness observed between groups (P > 0.05). Box-and-whisker plots display quartiles and range with standard deviations with possible outlier shown with dots in eh,no. Adonis test assessed the significant difference differences between the dissimilarity distance matrix of each group (i,j,p,q). P < 0.05 for each compared group in each cohort (i,j,p,q).

Extended Data Fig. 10 The relative abundance changes in Cluster CDA taxa with disease activity and intestinal inflammation.

a, The relative abundance changes of taxa in CDA cluster in longitudinally studied 34 IBD patients (77 CD and 44 UC samples in Cohort 1 and 49 CD and 14 UC samples in Cohort 2) based on the clinically defined changes in disease activity over time as described in Fig. 6c. b, The correlation between fecal calprotectin of 78 CD patients and Cluster CDA taxa. Data was analyzed using MaAsLin with BH-FDR correction and data was not significant (NS; q > 0.05).

Supplementary information

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yilmaz, B., Juillerat, P., Øyås, O. et al. Microbial network disturbances in relapsing refractory Crohn’s disease. Nat Med 25, 323–336 (2019). https://doi.org/10.1038/s41591-018-0308-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41591-018-0308-z

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