Diet rapidly and reproducibly alters the human gut microbiome

Lawrence A David, Corinne F Maurice, Rachel N Carmody, David B Gootenberg, Julie E Button, Benjamin E Wolfe, Alisha V Ling, A Sloan Devlin, Yug Varma, Michael A Fischbach, Sudha B Biddinger, Rachel J Dutton, Peter J Turnbaugh, Lawrence A David, Corinne F Maurice, Rachel N Carmody, David B Gootenberg, Julie E Button, Benjamin E Wolfe, Alisha V Ling, A Sloan Devlin, Yug Varma, Michael A Fischbach, Sudha B Biddinger, Rachel J Dutton, Peter J Turnbaugh

Abstract

Long-term dietary intake influences the structure and activity of the trillions of microorganisms residing in the human gut, but it remains unclear how rapidly and reproducibly the human gut microbiome responds to short-term macronutrient change. Here we show that the short-term consumption of diets composed entirely of animal or plant products alters microbial community structure and overwhelms inter-individual differences in microbial gene expression. The animal-based diet increased the abundance of bile-tolerant microorganisms (Alistipes, Bilophila and Bacteroides) and decreased the levels of Firmicutes that metabolize dietary plant polysaccharides (Roseburia, Eubacterium rectale and Ruminococcus bromii). Microbial activity mirrored differences between herbivorous and carnivorous mammals, reflecting trade-offs between carbohydrate and protein fermentation. Foodborne microbes from both diets transiently colonized the gut, including bacteria, fungi and even viruses. Finally, increases in the abundance and activity of Bilophila wadsworthia on the animal-based diet support a link between dietary fat, bile acids and the outgrowth of microorganisms capable of triggering inflammatory bowel disease. In concert, these results demonstrate that the gut microbiome can rapidly respond to altered diet, potentially facilitating the diversity of human dietary lifestyles.

Figures

Fig. 1. Short-term diet alters the gut…
Fig. 1. Short-term diet alters the gut microbiota
Ten subjects were tracked across each diet arm. (A) Fiber intake on the plant-based diet rose from a median baseline value of 9.3±2.1 to 25.6±1.1 g/1,000kcal (p=0.007; two-sided Wilcoxon signed-rank test), but was negligible on the animal-based diet (p=0.005). (B) Daily fat intake doubled on the animal-based diet from a baseline of 32.5±2.2% to 69.5±0.4% kcal (p=0.005), but dropped on the plant-based diet to 22.1±1.7% (p=0.02). (C) Protein intake rose on the animal-based diet to 30.1±0.5% kcal from a baseline level of 16.2±1.3% (p=0.005) and decreased on the plant-based diet to 10.0±0.3% (p=0.005). (D) Within-sample species diversity (α-diversity, Shannon’s Diversity Index), did not significantly change during either diet. (E) The similarity of each individual’s gut microbiota to their baseline communities (β-diversity, Jensen-Shannon distance) decreased on the animal-based diet (dates with q<0.05 identified with asterisks; Bonferroni-corrected, two-sided Mann-Whitney U test). Community differences were apparent one day after a tracing dye showed the animal-based diet reached the gut (blue arrows depict appearance of food dyes added to first and last diet day meals; Extended Data Fig. 3a).
Fig. 2. Bacterial cluster responses to diet…
Fig. 2. Bacterial cluster responses to diet arms
Cluster log2 fold-changes on each diet arm were computed relative to baseline samples across all subjects and are drawn as circles. Clusters with significant fold-changes on the animal-based diet are colored in red, and clusters with significant fold-changes on both the plant- and animal-based diets are colored in both red and green. Uncolored clusters exhibited no significant fold-change on either the animal or plant-based diet (q

Fig. 3. Diet alters microbial activity and…

Fig. 3. Diet alters microbial activity and gene expression

Fecal concentrations of SCFAs from (…

Fig. 3. Diet alters microbial activity and gene expression
Fecal concentrations of SCFAs from (A) carbohydrate and (B) amino acid fermentation (*p<0.05, two-sided Mann-Whitney U test; n=9–11 fecal samples/diet arm; Supplementary Table 11). The animal-based diet was associated with significant increases in gene expression (normalized to reads per kilobase per million mapped, or RPKM; n=13–21 datasets/diet arm) among (C) glutamine amidotransferases (K08681, vitamin B6 metabolism), (D) methyltransferases (K00599, polycyclic aromatic hydrocarbon degradation), and (E) beta-lactamases (K01467). (F) Hierarchical clustering of gut microbial gene expression profiles collected on the animal-based (red) and plant-based (green) diets. Expression profile similarity was significantly associated with diet (p<0.003; two-sided Fisher’s exact test excluding replicate samples), despite inter-individual variation that preceded the diet (Extended Data Figs. 6a,b). Enrichment on animal-based diet (red) and plant-based diet (green) for expression of genes involved in (G) amino acid metabolism and (H) central metabolism. Numbers indicate the mean fold-change between the two diets for each KEGG orthologous group assigned to a given enzymatic reaction (Supplementary Table 17). Enrichment patterns on the animal- and plant-based diets agree perfectly with patterns observed in carnivorous and herbivorous mammals, respectively2 (p<0.001, Binomial test). Note: Pyr Cx is represented by two groups, which showed divergent fold-changes. Asterisks in panels C-E andG,H indicate p<0.05, Student’st test. Values in panels A-E are mean±sem. Abbreviations: glutamate dehydrogenase (GDH), glutamate decarboxylase (Glu Dx), succinate-semialdehyde dehydrogenase (SSADH), phosphoenolpyruvate carboxylase (PEPCx), pyruvate carboxylase (Pyr Cx), phosphotransferase system (PTS), PEP carboxykinase (PEPCk), oxaloacetate decarboxylase (ODx), pyruvate, orthophosphate dikinase (PPDk).

Fig. 4. Foodborne microbes are detectable in…

Fig. 4. Foodborne microbes are detectable in the distal gut

(A) Common bacteria and fungi…
Fig. 4. Foodborne microbes are detectable in the distal gut
(A) Common bacteria and fungi associated with the animal-based diet menu items, as measured by 16S rRNA and ITS gene sequencing, respectively. Taxa are identified on the genus (g) and species (s) level. A full list of foodborne fungi and bacteria on the animal-based diet can be found in Supplementary Table 21. Foods on the plant-based diet were dominated by matches to the Streptophyta, which derive from chloroplasts within plant matter (Extended Data Fig. 7a). (B-E). Fecal RNA transcripts were significantly enriched (q<0.1, Kruskal-Wallis test; n=6–10 samples/diet arm) for several food-associated microbes on the animal-based diet relative to baseline (BL) periods, including (B) Lactococcus lactis, (C) Staphylococcus carnosus, (D)Pediococcus acidilactici, and (E) aPenicillium sp. A complete table of taxa with significant expression differences can be found in Supplementary Table 22. (F) Fungal concentrations in feces before and 1–2 days after the animal-based diet were also measured using culture media selective for fungal growth (plate count agar with milk, salt, and chloramphenicol). Post-diet fecal samples exhibit significantly higher fungal concentrations than baseline samples (p<0.02; two-sided Mann-Whitney U test; n=7–10 samples/diet arm). (G) Increased RNA transcripts from the plant-derived Rubus chlorotic mottle virus transcripts increase on the plant-based diet (q<0.1, Kruskal-Wallis test; n=6–10 samples/diet arm). Barplots (B-G) all display mean±sem.

Fig. 5. Changes in the fecal concentration…

Fig. 5. Changes in the fecal concentration of bile acids and biomarkers for Bilophila on…

Fig. 5. Changes in the fecal concentration of bile acids and biomarkers for Bilophila on the animal-based diet
(A) Deoxycholic acid, a secondary bile acid known to promote DNA damage and hepatic carcinomas, accumulates significantly on the animal-based diet (p<0.01, two-sided Wilcoxon signed-rank test; see Supplementary Table 23 for the diet response of other secondary bile acids). (B) RNA-Seq data also supports increased microbial metabolism of bile acids on the animal-based diet, as we observe significantly increased expression of microbial bile salt hydrolases (K01442) during that diet arm (q<0.05, Kruskal-Wallis test; normalized to reads per kilobase per million mapped, or RPKM; n=8–21 samples/diet arm). (C) Total fecal bile acid concentrations also increase significantly on the animal-based diet, relative to the preceding baseline period (p<0.05, two-sided Wilcoxon signed-rank test), but do not change on the plant-based diet (Extended Data Fig. 9). Bile acids have been shown to cause IBD in mice by stimulating the growth of the bacterium Bilophila, which is known to reduce sulfite to hydrogen sulfide via the sulfite reductase enzyme (dsrA; Extended Data Fig. 10). (D) Quantitative PCR showed a significant increase in microbial DNA coding for dsrA on the animal-based diet (p<0.05; two-sided Wilcoxon signed-rank test), and (E) RNA-Seq identified a significant increase in sulfite reductase expression (q<0.05, Kruskal-Wallis test; n=8–21 samples/diet arm). Barplots (B,E) display mean±sem.
Fig. 3. Diet alters microbial activity and…
Fig. 3. Diet alters microbial activity and gene expression
Fecal concentrations of SCFAs from (A) carbohydrate and (B) amino acid fermentation (*p<0.05, two-sided Mann-Whitney U test; n=9–11 fecal samples/diet arm; Supplementary Table 11). The animal-based diet was associated with significant increases in gene expression (normalized to reads per kilobase per million mapped, or RPKM; n=13–21 datasets/diet arm) among (C) glutamine amidotransferases (K08681, vitamin B6 metabolism), (D) methyltransferases (K00599, polycyclic aromatic hydrocarbon degradation), and (E) beta-lactamases (K01467). (F) Hierarchical clustering of gut microbial gene expression profiles collected on the animal-based (red) and plant-based (green) diets. Expression profile similarity was significantly associated with diet (p<0.003; two-sided Fisher’s exact test excluding replicate samples), despite inter-individual variation that preceded the diet (Extended Data Figs. 6a,b). Enrichment on animal-based diet (red) and plant-based diet (green) for expression of genes involved in (G) amino acid metabolism and (H) central metabolism. Numbers indicate the mean fold-change between the two diets for each KEGG orthologous group assigned to a given enzymatic reaction (Supplementary Table 17). Enrichment patterns on the animal- and plant-based diets agree perfectly with patterns observed in carnivorous and herbivorous mammals, respectively2 (p<0.001, Binomial test). Note: Pyr Cx is represented by two groups, which showed divergent fold-changes. Asterisks in panels C-E andG,H indicate p<0.05, Student’st test. Values in panels A-E are mean±sem. Abbreviations: glutamate dehydrogenase (GDH), glutamate decarboxylase (Glu Dx), succinate-semialdehyde dehydrogenase (SSADH), phosphoenolpyruvate carboxylase (PEPCx), pyruvate carboxylase (Pyr Cx), phosphotransferase system (PTS), PEP carboxykinase (PEPCk), oxaloacetate decarboxylase (ODx), pyruvate, orthophosphate dikinase (PPDk).
Fig. 4. Foodborne microbes are detectable in…
Fig. 4. Foodborne microbes are detectable in the distal gut
(A) Common bacteria and fungi associated with the animal-based diet menu items, as measured by 16S rRNA and ITS gene sequencing, respectively. Taxa are identified on the genus (g) and species (s) level. A full list of foodborne fungi and bacteria on the animal-based diet can be found in Supplementary Table 21. Foods on the plant-based diet were dominated by matches to the Streptophyta, which derive from chloroplasts within plant matter (Extended Data Fig. 7a). (B-E). Fecal RNA transcripts were significantly enriched (q<0.1, Kruskal-Wallis test; n=6–10 samples/diet arm) for several food-associated microbes on the animal-based diet relative to baseline (BL) periods, including (B) Lactococcus lactis, (C) Staphylococcus carnosus, (D)Pediococcus acidilactici, and (E) aPenicillium sp. A complete table of taxa with significant expression differences can be found in Supplementary Table 22. (F) Fungal concentrations in feces before and 1–2 days after the animal-based diet were also measured using culture media selective for fungal growth (plate count agar with milk, salt, and chloramphenicol). Post-diet fecal samples exhibit significantly higher fungal concentrations than baseline samples (p<0.02; two-sided Mann-Whitney U test; n=7–10 samples/diet arm). (G) Increased RNA transcripts from the plant-derived Rubus chlorotic mottle virus transcripts increase on the plant-based diet (q<0.1, Kruskal-Wallis test; n=6–10 samples/diet arm). Barplots (B-G) all display mean±sem.
Fig. 5. Changes in the fecal concentration…
Fig. 5. Changes in the fecal concentration of bile acids and biomarkers for Bilophila on the animal-based diet
(A) Deoxycholic acid, a secondary bile acid known to promote DNA damage and hepatic carcinomas, accumulates significantly on the animal-based diet (p<0.01, two-sided Wilcoxon signed-rank test; see Supplementary Table 23 for the diet response of other secondary bile acids). (B) RNA-Seq data also supports increased microbial metabolism of bile acids on the animal-based diet, as we observe significantly increased expression of microbial bile salt hydrolases (K01442) during that diet arm (q<0.05, Kruskal-Wallis test; normalized to reads per kilobase per million mapped, or RPKM; n=8–21 samples/diet arm). (C) Total fecal bile acid concentrations also increase significantly on the animal-based diet, relative to the preceding baseline period (p<0.05, two-sided Wilcoxon signed-rank test), but do not change on the plant-based diet (Extended Data Fig. 9). Bile acids have been shown to cause IBD in mice by stimulating the growth of the bacterium Bilophila, which is known to reduce sulfite to hydrogen sulfide via the sulfite reductase enzyme (dsrA; Extended Data Fig. 10). (D) Quantitative PCR showed a significant increase in microbial DNA coding for dsrA on the animal-based diet (p<0.05; two-sided Wilcoxon signed-rank test), and (E) RNA-Seq identified a significant increase in sulfite reductase expression (q<0.05, Kruskal-Wallis test; n=8–21 samples/diet arm). Barplots (B,E) display mean±sem.

References

    1. Wu GD, et al. Linking long-term dietary patterns with gut microbial enterotypes. Science. 2011;334:105–108.
    1. Muegge BD, et al. Diet drives convergence in gut microbiome functions across mammalian phylogeny and within humans. Science. 2011;332:970–974.
    1. Duncan SH, et al. Reduced dietary intake of carbohydrates by obese subjects results in decreased concentrations of butyrate and butyrate-producing bacteria in feces. Appl Environ Microbiol. 2007;73:1073–1078.
    1. Ley RE, Turnbaugh PJ, Klein S, Gordon JI. Microbial ecology: human gut microbes associated with obesity. Nature. 2006;444:1022–1023.
    1. Walker AW, et al. Dominant and diet-responsive groups of bacteria within the human colonic microbiota. ISME J. 2011;5:220–230.
    1. Devkota S, et al. Dietary-fat-induced taurocholic acid promotes pathobiont expansion and colitis in Il10−/− mice. Nature. 2012;487:104–108.
    1. Turnbaugh PJ, et al. The effect of diet on the human gut microbiome: a metagenomic analysis in humanized gnotobiotic mice. Sci Transl Med. 2009;1:6ra14.
    1. Turnbaugh PJ, et al. An obesity-associated gut microbiome with increased capacity for energy harvest. Nature. 2006;444:1027–1031.
    1. Faith JJ, McNulty NP, Rey FE, Gordon JI. Predicting a human gut microbiota's response to diet in gnotobiotic mice. Science. 2011;333:101–104.
    1. Russell WR, et al. High-protein, reduced-carbohydrate weight-loss diets promote metabolite profiles likely to be detrimental to colonic health. Am J Clin Nutr. 2011;93:1062–1072.
    1. Cordain L, et al. Plant-animal subsistence ratios and macronutrient energy estimations in worldwide hunter-gatherer diets. Am J Clin Nutr. 2000;71:682–692.
    1. Arumugam M, et al. Enterotypes of the human gut microbiome. Nature. 2011;473:174–180.
    1. De Filippo C, et al. Impact of diet in shaping gut microbiota revealed by a comparative study in children from Europe and rural Africa. Proc Natl Acad Sci U S A. 2010;107:14691–14696.
    1. Reddy BS. Diet and excretion of bile acids. Cancer Res. 1981;41:3766–3768.
    1. Smith EA, Macfarlane GT. Enumeration of amino acid fermenting bacteria in the human large intestine: effects of pH and starch on peptide metabolism and dissimilation of amino acids. FEMS Microbiol Ecol. 1998;25:355–368.
    1. Smith EA, Macfarlane GT. Enumeration of human colonic bacteria producing phenolic and indolic compounds: effects of pH, carbohydrate availability and retention time on dissimilatory aromatic amino acid metabolism. J Appl Bacteriol. 1996;81:288–302.
    1. Sinha R, et al. High concentrations of the carcinogen 2-amino-1-methyl-6-phenylimidazo[4,5-b]pyridine (PhIP) occur in chicken but are dependent on the cooking method. Cancer Res. 1995;55:4516–4519.
    1. PICRUSt ( )
    1. Kanehisa M, Goto S. KEGG: kyoto encyclopedia of genes and genomes. Nucleic Acids Res. 2000;28:27–30.
    1. Pittard J, Wallace BJ. Distribution and function of genes concerned with aromatic biosynthesis in Escherichia coli. J Bacteriol. 1966;91:1494–1508.
    1. Hawkes K, O'Connell JF, Jones NG. Hunting income patterns among the Hadza: big game, common goods, foraging goals and the evolution of the human diet. Philos Trans R Soc Lond B Biol Sci. 1991;334:243–250. discussion 250–241.
    1. Bourdichon F, Berger B, Casaregola S. Safety Demonstration of Microbial Food Cultures (MFC) in Fermented Food Products. Bulletin of the International Dairy Federation. 2012;455
    1. Nychas GJ, Arkoudelos JS. Staphylococci: their role in fermented sausages. Soc Appl Bacteriol Symp Ser. 1990;19:167S–188S.
    1. McGavin WJ, Macfarlane SA. Rubus chlorotic mottle virus, a new sobemovirus infecting raspberry and bramble. Virus Res. 2009;139:10–13.
    1. Zhang T, et al. RNA viral community in human feces: prevalence of plant pathogenic viruses. PLoS Biol. 2006;4:e3.
    1. Yoshimoto S, et al. Obesity-induced gut microbial metabolite promotes liver cancer through senescence secretome. Nature. 2013;499:97–101.
    1. Ridlon JM, Kang DJ, Hylemon PB. Bile salt biotransformations by human intestinal bacteria. J Lipid Res. 2006;47:241–259.
    1. Islam KB, et al. Bile acid is a host factor that regulates the composition of the cecal microbiota in rats. Gastroenterology. 2011;141:1773–1781.
    1. Maurice CF, Haiser HJ, Turnbaugh PJ. Xenobiotics shape the physiology and gene expression of the active human gut microbiome. Cell. 2013;152:39–50.
    1. Caporaso JG, et al. QIIME allows analysis of high-throughput community sequencing data. Nat Methods. 2010;7:335–336.
References for Methods
    1. Lewis SJ, Heaton KW. Stool form scale as a useful guide to intestinal transit time. Scand J Gastroenterol. 1997;32:920–924.
    1. Diet History Questionnaire, Version 2.0. National Institutes of Health, Applied Research Program, National Cancer Institute; 2010.
    1. Caporaso JG, et al. Global patterns of 16S rRNA diversity at a depth of millions of sequences per sample. Proc Natl Acad Sci U S A. 2011;108(Suppl 1):4516–4522.
    1. Caporaso JG, et al. Ultra-high-throughput microbial community analysis on the Illumina HiSeq and MiSeq platforms. ISME J. 2012;6:1621–1624.
    1. Caporaso JG, et al. QIIME allows analysis of high-throughput community sequencing data Nat. Methods. 2010;7:335–336.
    1. DeSantis TZ, et al. Greengenes, a chimera-checked 16S rRNA gene database and workbench compatible with ARB. Appl Environ Microbiol. 2006;72:5069–5072.
    1. SciPy: Open source scientific tools for Python.
    1. McKinney W. Proceedings of the 9th Python in Science Conference. 2010
    1. Strimmer K. fdrtool: a versatile R package for estimating local and tail areabased false discovery rates. Bioinformatics. 2008;24:1461–1462.
    1. Friedman J, Alm EJ. Inferring correlation networks from genomic survey data. PLoS Comput Biol. 2012;8:e1002687.
    1. Turnbaugh PJ, et al. Organismal, genetic, and transcriptional variation in the deeply sequenced gut microbiomes of identical twins. Proc Natl Acad Sci U S A. 2010;107:7503–7508.
    1. Maurice CF, Haiser HJ, Turnbaugh PJ. Xenobiotics shape the physiology and gene expression of the active human gut microbiome. Cell. 2013;152:39–50.
    1. Turnbaugh PJ, et al. The effect of diet on the human gut microbiome: a metagenomic analysis in humanized gnotobiotic mice. Sci Transl Med. 2009;1:6ra14.
    1. Rey FE, et al. Dissecting the in vivo metabolic potential of two human gut acetogens. J Biol Chem. 2010;285:22082–22090.
    1. Nelson KE, et al. A catalog of reference genomes from the human microbiome. Science. 2010;328:994–999.
    1. Kanehisa M, Goto S. KEGG: kyoto encyclopedia of genes and genomes. Nucleic Acids Res. 2000;28:27–30.
    1. Ning Z, Cox AJ, Mullikin JC. SSAHA: a fast search method for large DNA databases. Genome Res. 2001;11:1725–1729.
    1. Abubucker S, et al. Metabolic reconstruction for metagenomic data and its application to the human microbiome. PLoS Comput Biol. 2012;8:e1002358.
    1. Segata N, et al. Metagenomic biomarker discovery and explanation. Genome Biol. 2011;12:R60.
    1. Langmead B, Salzberg SL. Fast gapped-read alignment with Bowtie 2. Nat Methods. 2012;9:357–359.
    1. Markowitz VM, et al. IMG: the Integrated Microbial Genomes database and comparative analysis system. Nucleic Acids Res. 2012;40:D115–122.
    1. Martin J, et al. Optimizing read mapping to reference genomes to determine composition and species prevalence in microbial communities. PLoS ONE. 2012;7:e36427.
    1. Devkota S, et al. Dietary-fat-induced taurocholic acid promotes pathobiont expansion and colitis in Il10−/− mice. Nature. 2012;487:104–108.
    1. Deplancke B, et al. Molecular ecological analysis of the succession and diversity of sulfate-reducing bacteria in the mouse gastrointestinal tract. Appl Environ Microbiol. 2000;66:2166–2174.
    1. Stewart JA, Chadwick VS, Murray A. Carriage, quantification, and predominance of methanogens and sulfate-reducing bacteria in faecal samples. Lett Appl Microbiol. 2006;43:58–63.
    1. Porter JL, et al. Accurate enzymatic measurement of fecal bile acids in patients with malabsorption. J Lab Clin Med. 2003;141:411–418.
    1. Setchell KD, Lawson AM, Tanida N, Sjovall J. General methods for the analysis of metabolic profiles of bile acids and related compounds in feces. J Lipid Res. 1983;24:1085–1100.
    1. Schoch CL, et al. Nuclear ribosomal internal transcribed spacer (ITS) region as a universal DNA barcode marker for Fungi . Proc Natl Acad Sci U S A. 2012;109:6241–6246.
    1. Gardes M, Bruns TD. ITS primers with enhanced specificity for basidiomycetes--application to the identification of mycorrhizae and rusts. Mol Ecol. 1993;2:113–118.
    1. White TJ, Bruns T, Lee S, Taylor J. Gelfand Ma Innis Dh, Shinsky Jj, White Tj., editors. PCR Protocols: A Guide to. Methods and Applications. 1990:315–322.

Source: PubMed

3
S'abonner