Short Chain Fatty Acids (SCFAs)-Mediated Gut Epithelial and Immune Regulation and Its Relevance for Inflammatory Bowel Diseases

Daniela Parada Venegas, Marjorie K De la Fuente, Glauben Landskron, María Julieta González, Rodrigo Quera, Gerard Dijkstra, Hermie J M Harmsen, Klaas Nico Faber, Marcela A Hermoso, Daniela Parada Venegas, Marjorie K De la Fuente, Glauben Landskron, María Julieta González, Rodrigo Quera, Gerard Dijkstra, Hermie J M Harmsen, Klaas Nico Faber, Marcela A Hermoso

Abstract

Ulcerative colitis (UC) and Crohn's disease (CD), collectively known as Inflammatory Bowel Diseases (IBD), are caused by a complex interplay between genetic, immunologic, microbial and environmental factors. Dysbiosis of the gut microbiome is increasingly considered to be causatively related to IBD and is strongly affected by components of a Western life style. Bacteria that ferment fibers and produce short chain fatty acids (SCFAs) are typically reduced in mucosa and feces of patients with IBD, as compared to healthy individuals. SCFAs, such as acetate, propionate and butyrate, are important metabolites in maintaining intestinal homeostasis. Several studies have indeed shown that fecal SCFAs levels are reduced in active IBD. SCFAs are an important fuel for intestinal epithelial cells and are known to strengthen the gut barrier function. Recent findings, however, show that SCFAs, and in particular butyrate, also have important immunomodulatory functions. Absorption of SCFAs is facilitated by substrate transporters like MCT1 and SMCT1 to promote cellular metabolism. Moreover, SCFAs may signal through cell surface G-protein coupled receptors (GPCRs), like GPR41, GPR43, and GPR109A, to activate signaling cascades that control immune functions. Transgenic mouse models support the key role of these GPCRs in controlling intestinal inflammation. Here, we present an overview of microbial SCFAs production and their effects on the intestinal mucosa with specific emphasis on their relevance for IBD. Moreover, we discuss the therapeutic potential of SCFAs for IBD, either applied directly or by stimulating SCFAs-producing bacteria through pre- or probiotic approaches.

Keywords: IBD; IECs; SCFAs; dysbiosis; immune cells; intestinal mucosa.

Figures

Figure 1
Figure 1
Schematic representation of carboydrates fermentation pathways that lead acetate, propionate and butyrate production. The main enzymes involved in the butyrate production are indicated as (a) butyrate kinase and (b) butyryl CoA:acetate CoA transferase. Figure adapted from den Besten et al. (35).
Figure 2
Figure 2
SCFAs in healthy (A) and inflamed (B) colonic mucosa. In healthy mucosa, (1) bacterial fermentation of dietary fiber (DF) by SCFAs-producing bacteria (e.g., F. prausnitzii), increases luminal content of butyrate (green), propionate (blue) and acetate (purple) (2), forming a gradient along the crypt. In lamina propria (LP) macrophages under acute inflammatory stimulus (4), butyrate inhibits histone deacetylases (HDACs) thus; NF-κB-induced pro-inflammatory mediators (e.g. TNF-α, IL-6, IL-12 and iNOS) expression whereas increases anti-inflammatory mediators (e.g., IL-10). In colonocytes (5), butyrate is β-oxidized to Acetyl-CoA and constitutes the main source of energy by entering the TCA cycle. Alternatively, butyrate initiates signaling pathway activation (or repression) by GPCRs and/or directly inhibits HDACs, thus activating (e.g., HIF-1, STAT3 and SP1) or repressing (e.g., NF-κB) transcription factors (TFs), increasing epithelial barrier function, antimicrobial peptides (AMPs) production, cell proliferation and decreasing inflammation. In inflamed mucosa as IBD, (1) a decreased fermentation of DF by low levels of SCFAs-producing bacteria (e.g., F. prausnitzii) (2), reduces SCFAs luminal content (3). In LP inflammatory macrophages (4), butyrate-GPCRs activation and -HDACs inhibition are downregulated, thus, there is uncontrolled NF-κB-induced pro-inflammatory mediators' expression (e.g. TNF-α, IL-6, IL-12 and iNOS) and decrease of anti-inflammatory mediators (e.g., IL-10), although it appears that the inflammation increasesthe GPCRs and transporters expression. In inflamed colonocytes (5), butyrate uptake and oxidation are decreased and GPCRs and transporters are also downregulated. This contributes to decreased epithelial barrier integrity, AMPs production, cell proliferation and increased inflammation.

References

    1. Ananthakrishnan AN. Epidemiology and risk factors for IBD. Nat Rev Gastroenterol Hepatol. (2015) 12:205–17. 10.1038/nrgastro.2015.34
    1. Cosnes J, Gowerrousseau C, Seksik P, Cortot A. Epidemiology and natural history of inflammatory bowel diseases. Gastroenterology. (2011) 140:1785–94. 10.1053/j.gastro.2011.01.055
    1. De Souza HSP, Fiocchi C. Immunopathogenesis of IBD: current state of the art. Nat Rev Gastroenterol Hepatol. (2016) 13:13–27. 10.1038/nrgastro.2015.186
    1. Sommer F, Bäckhed F. The gut microbiota-masters of host development and physiology. Nat Rev Microbiol. (2013) 11:227–38. 10.1038/nrmicro2974
    1. Thorburn AN, Macia L, Mackay CR. Diet, metabolites, and “Western-Lifestyle” inflammatory diseases. Immunity. (2014) 40:833–842. 10.1016/j.immuni.2014.05.014
    1. Agus A, Denizot J, Thévenot J, Martinez-Medina M, Massier S, Sauvanet P, et al. . Western diet induces a shift in microbiota composition enhancing susceptibility to Adherent-Invasive E. coli infection and intestinal inflammation. Sci Rep. (2016) 6:1–14. 10.1038/srep19032
    1. Scott NA, Andrusaite A, Andersen P, Lawson M, Alcon-Giner C, Leclaire C, et al. . Antibiotics induce sustained dysregulation of intestinal T cell immunity by perturbing macrophage homeostasis. Sci Transl Med. (2018) 10:eaao4755. 10.1126/scitranslmed.aao4755
    1. Neurath MF. New targets for mucosal healing and therapy in inflammatory bowel diseases. Mucosal Immunol. (2014) 7:6–19. 10.1038/mi.2013.73
    1. Nie Y, Lin Q, Luo F. Effects of non-starch polysaccharides on inflammatory bowel disease. Int J Mol Sci. (2017) 18:E1372. 10.3390/ijms18071372
    1. Koboziev I, Reinoso Webb C, Furr KL, Grisham MB. Role of the enteric microbiota in intestinal homeostasis and inflammation. Free Radic Biol Med. (2014) 68:122–33. 10.1016/j.freeradbiomed.2013.11.008
    1. Rakoff-Nahoum S, Paglino J, Eslami-varzaneh F, Edberg S, Medzhitov R. Recognition of commensal microflora by toll- like receptors is required for intestinal homeostasis recognition of commensal microflora by toll-like receptors is required for intestinal homeostasis. Cell. (2015) 118:229–41. 10.1016/j.cell.2004.07.002
    1. Kamada N, Chen GY, Inohara N, Núñez G. Control of pathogens and pathobionts by the gut microbiota. Nat Immunol. (2013) 14:685–90. 10.1038/ni.2608
    1. Madsen KL, Doyle JS, Tavernini MM, Jewell LD, Rennie RP, Fedorak RN. Antibiotic therapy attenuates colitis in interleukin 10 gene-deficient mice. Gastroenterology. (2000) 118:1094–105. 10.1016/S0016-5085(00)70362-3
    1. Hörmannsperger G, Schaubeck M, Haller D. Intestinal microbiota in animal models of inflammatory diseases. ILAR J. (2015) 56:179–91. 10.1093/ilar/ilv019
    1. Maslowski KM, Vieira AT, Ng A, Kranich J, Sierro F, Di Yu, et al. . Regulation of inflammatory responses by gut microbiota and chemoattractant receptor GPR43. Nature. (2009) 461:1282–6. 10.1038/nature08530
    1. Tominaga K, Kamimura K, Takahashi K, Yokoyama J, Yamagiwa S, Terai S. Diversion colitis and pouchitis: a mini-review. World J Gastroenterol. (2018) 24:1734–47. 10.3748/wjg.v24.i16.1734
    1. Donohoe DR, Garge N, Zhang X, Sun W, O'Connell TM, Bunger MK, et al. . The microbiome and butyrate regulate energy metabolism and autophagy in the mammalian colon. Cell Metab. (2011) 13:517–26. 10.1016/j.cmet.2011.02.018
    1. Corrêa-Oliveira R, Fachi JL, Vieira A, Sato FT, Vinolo MAR. Regulation of immune cell function by short-chain fatty acids. Clin Transl Immunol. (2016) 5:e73. 10.1038/cti.2016.17
    1. Rechkemmer G, von Engelhardt W. Concentration- and pH-dependence of short-chain fatty acid absorption in the proximal and distal colon of guinea pig (Cavia porcellus). Comp Biochem Physiol A Comp Physiol. (1988) 91:659–63.
    1. Ritzhaupt A, Wood IS, Ellis A, Hosie KB, Shirazi-Beechey SP. Identification and characterization of a monocarboxylate transporter (MCT1) in pig and human colon: its potential to transport l -lactate as well as butyrate. J Physiol. (1998) 513:719–32. 10.1111/j.1469-7793.1998.719ba.x
    1. Miyauchi S, Gopal E, Fei YJ, Ganapathy V. Functional identification of SLC5A8, a tumor suppressor down-regulated in colon cancer, as a Na+-coupled transporter for short-chain fatty acids. J Biol Chem. (2004) 279:13293–6. 10.1074/jbc.C400059200
    1. Brown AJ, Goldsworthy SM, Barnes AA, Eilert MM, Tcheang L, Daniels D, et al. . The orphan G protein-coupled receptors GPR41 and GPR43 are activated by propionate and other short chain carboxylic acids. J Biol Chem. (2003) 278:11312–9. 10.1074/jbc.M211609200
    1. Le Poul E, Loison C, Struyf S, Springael JY, Lannoy V, Decobecq ME, et al. . Functional characterization of human receptors for short chain fatty acids and their role in polymorphonuclear cell activation. J Biol Chem. (2003) 278:25481–9. 10.1074/jbc.M301403200
    1. Taggart AKP, Kero J, Gan X, Cai TQ, Cheng K, Ippolito M, et al. . (D)-β-hydroxybutyrate inhibits adipocyte lipolysis via the nicotinic acid receptor PUMA-G. J Biol Chem. (2005) 280:26649–52. 10.1074/jbc.C500213200
    1. Kumari R, Ahuja V, Paul J. Fluctuations in butyrate-producing bacteria in ulcerative colitis patients of North India. World J Gastroenterol. (2013) 19:3404–14. 10.3748/wjg.v19.i22.3404
    1. Wang W, Chen L, Zhou R, Wang X, Song L, Huang S, et al. . Increased proportions of Bifidobacterium and the Lactobacillus group and loss of butyrate-producing bacteria in inflammatory bowel disease. J Clin Microbiol. (2014) 52:398–406. 10.1128/JCM.01500-13
    1. Joossens M, Huys G, Cnockaert M, De Preter V, Verbeke K, Rutgeerts P, et al. . Dysbiosis of the faecal microbiota in patients with Crohn's disease and their unaffected relatives. Gut. (2011) 60:631–637. 10.1136/gut.2010.223263
    1. Pascal V, Pozuelo M, Borruel N, Casellas F, Campos D, Santiago A, et al. . A microbial signature for Crohn's disease. Gut. (2017) 66:813–22. 10.1136/gutjnl-2016-313235
    1. Takahashi K, Nishida A, Fujimoto T, Fujii M, Shioya M, Imaeda H, et al. . Reduced abundance of butyrate-producing bacteria species in the fecal microbial community in crohn's disease. Digestion. (2016) 93:59–65. 10.1159/000441768
    1. Paramsothy S, Paramsothy R, Rubin DT, Kamm MA, Kaakoush NO, Mitchell HM, et al. . Faecal microbiota transplantation for inflammatory bowel disease: a systematic review and meta-analysis. J Crohns Colitis. (2017) 11:1180–99. 10.1093/ecco-jcc/jjx063
    1. Joint FAO/WHO CODEX Alimentarius Commission. Guidelines on nutrition labelling CAC/GL 2-1985. Codex Alimentarius International Food Stand.
    1. Champ MMJ. Physiological aspects of resistant starch and in vivo measurements. J AOAC Int. (2004) 87:749–55.
    1. Louis P, Flint HJ. Diversity, metabolism and microbial ecology of butyrate-producing bacteria from the human large intestine. FEMS Microbiol Lett. (2009) 294:1–8. 10.1111/j.1574-6968.2009.01514.x
    1. Louis P, Flint HJ. Formation of propionate and butyrate by the human colonic microbiota. Environ Microbiol. (2017) 19:29–41. 10.1111/1462-2920.13589
    1. den Besten G, van Eunen K, Groen AK, Venema K, Reijngoud D-J, Bakker BM. The role of short-chain fatty acids in the interplay between diet, gut microbiota, and host energy metabolism. J Lipid Res. (2013) 54:2325–40. 10.1194/jlr.R036012
    1. Miller TL, Wolin MJ. Pathways of acetate, propionate, and butyrate formation by the human fecal microbial flora. Appl Environ Microbiol. (1996) 62:1589–92.
    1. Duncan SH, Holtrop G, Lobley GE, Calder AG, Stewart CS, Flint HJ. Contribution of acetate to butyrate formation by human faecal bacteria. Br J Nutr. (2004) 91:915. 10.1079/BJN20041150
    1. Duncan SH, Barcenilla A, Stewart CS, Pryde SE, Flint HJ. Acetate utilization and butyryl coenzyme A (CoA):acetate-CoA transferase in butyrate-producing bacteria from the human large intestine. Appl Environ Microbiol. (2002) 68:5186–90. 10.1128/AEM.68.10.5186-5190.2002
    1. McNeil NI, Cummings JH, James WP. Short chain fatty acid absorption by the human large intestine. Gut. (1978) 19:819–22. 10.1136/gut.19.9.819
    1. McOrist AL, Miller RB, Bird AR, Keogh JB, Noakes M, Topping DL, et al. . Fecal butyrate levels vary widely among individuals but are usually increased by a diet high in resistant starch. J Nutr. (2011) 141:883–9. 10.3945/jn.110.128504
    1. Haenen D, Zhang J, Souza da Silva C, Bosch G, van der Meer IM, van Arkel J, et al. . A diet high in resistant starch modulates microbiota composition, SCFA concentrations, and gene expression in pig intestine. J Nutr. (2013) 143:274–83. 10.3945/jn.112.169672
    1. Cummings JH, Pomare EW, Branch WJ, Naylor CP, Macfarlane GT. Short chain fatty acids in human large intestine, portal, hepatic and venous blood. Gut. (1987) 28:1221–7. 10.1136/gut.28.10.1221
    1. Topping DL, Clifton PM. Short-chain fatty acids and human colonic function: roles of resistant starch and nonstarch polysaccharides. Physiol Rev. (2001) 81:1031–64. 10.1152/physrev.2001.81.3.1031
    1. Huda-Faujan N, Abdulamir AS, Fatimah AB, Anas OM, Shuhaimi M, Yazid AM, et al. . The impact of the level of the intestinal short chain fatty acids in inflammatory bowel disease patients versus healthy subjects. Open Biochem J. (2010) 4:53–8. 10.2174/1874091X01004010053
    1. Machiels K, Joossens M, Sabino J, De Preter V, Arijs I, Eeckhaut V, et al. . A decrease of the butyrate-producing species roseburia hominis and faecalibacterium prausnitzii defines dysbiosis in patients with ulcerative colitis. Gut. (2014) 63:1275–83. 10.1136/gutjnl-2013-304833
    1. Haghikia A, Jörg S, Duscha A, Berg J, Manzel A, Waschbisch A, et al. . Dietary fatty acids directly impact central nervous system autoimmunity via the small intestine. Immunity. (2015) 43:817–29. 10.1016/j.immuni.2015.09.007
    1. Vital M, Howe AC, Tiedje JM. Revealing the bacterial butyrate synthesis pathways by analyzing (Meta)genomic data. MBio. (2014) 5:1–11. 10.1128/mBio.00889-14
    1. Rivière A, Selak M, Lantin D, Leroy F, De Vuyst L. Bifidobacteria and butyrate-producing colon bacteria: importance and strategies for their stimulation in the human gut. Front Microbiol. (2016) 7:979. 10.3389/fmicb.2016.00979
    1. Derrien M, Vaughan EE, Plugge CM, de Vos WM. Akkermansia municiphila gen. nov., sp. nov., a human intestinal mucin-degrading bacterium. Int J Syst Evol Microbiol. (2004) 54:1469–76. 10.1099/ijs.0.02873-0
    1. Kelly CJ, Colgan SP. Breathless in the Gut: implications of luminal O2for microbial pathogenicity. Cell Host Microbe. (2016) 19:427–8. 10.1016/j.chom.2016.03.014
    1. Rivera-Chávez F, Zhang LF, Faber F, Lopez CA, Byndloss MX, Olsan EE, et al. . Depletion of butyrate-producing clostridia from the gut microbiota drives an aerobic luminal expansion of Salmonella. Cell Host Microbe. (2016) 19:443–54. 10.1016/j.chom.2016.03.004
    1. Byndloss MX, Olsan EE, Rivera-Chávez F, Tiffany CR, Cevallos SA, Lokken KL, et al. . Microbiota-activated PPAR-γ signaling inhibits dysbiotic Enterobacteriaceae expansion. Science. (2017) 357:570–5. 10.1126/science.aam9949
    1. Gill RK. Expression and membrane localization of MCT isoforms along the length of the human intestine. AJP Cell Physiol. (2005) 289:C846–52. 10.1152/ajpcell.00112.2005
    1. Al-mosauwi H, Ryan E, McGrane A, Riveros-Beltran S, Walpole C, Dempsey E, et al. . Differential protein abundance of a basolateral MCT1 transporter in the human gastrointestinal tract. Cell Biol Int. (2016) 40:1303–12. 10.1002/cbin.10684
    1. Iwanaga T, Takebe K, Kato I, Karaki S-I, Kuwahara A. Cellular expression of monocarboxylate transporters (MCT) in the digestive tract of the mouse, rat, and humans, with special reference to slc5a8. Biomed Res. (2006) 27:243–54. 10.2220/biomedres.27.243
    1. Kirat D, Kondo K, Shimada R, Kato S. Dietary pectin up-regulates monocaboxylate transporter 1 in the rat gastrointestinal tract. Exp Physiol. (2009) 94:422–33. 10.1113/expphysiol.2009.046797
    1. Merezhinskaya N, Ogunwuyi SA, Mullick FG, Fishbein WN. Presence and localization of three lactic acid transporters (MCT1,−2, and −4) in separated human granulocytes, lymphocytes, and monocytes. J Histochem Cytochem. (2004) 52:1483–93. 10.1369/jhc.4A6306.2004
    1. Murray CM, Hutchinson R, Bantick JR, Belfield GP, Benjamin AD, Brazma D, et al. . Monocarboxylate transporter Mctl is a target for immunosuppression. Nat Chem Biol. (2005) 1:371–6. 10.1038/nchembio744
    1. Hahn EL, Halestrap AP, Gamelli RL. Expression of the lactate transporter MCT1 in macrohages. Shock. (2000) 13:253–260. 10.1097/00024382-200004000-00001
    1. Borthakur A, Anbazhagan AN, Kumar A, Raheja G, Singh V, Ramaswamy K, et al. . The probiotic Lactobacillus plantarum counteracts TNF- -induced downregulation of SMCT1 expression and function. AJP Gastrointest Liver Physiol. (2010) 299:G928–34. 10.1152/ajpgi.00279.2010
    1. Cox MA, Jackson J, Stanton M, Rojas-Triana A, Bober L, Laverty M, et al. Short-chain fatty acids act as antiinflammatory mediators by regulating prostaglandin E2and cytokines. World J Gastroenterol. (2009) 15:5549–57. 10.3748/wjg.15.5549
    1. Nastasi C, Candela M, Bonefeld CM, Geisler C, Hansen M, Krejsgaard T, et al. . The effect of short-chain fatty acids on human monocyte-derived dendritic cells. Sci Rep. (2015) 5:1–10. 10.1038/srep16148
    1. Ang Z, Er JZ, Ding JL. The short-chain fatty acid receptor GPR43 is transcriptionally regulated by XBP1 in human monocytes. Sci Rep. (2015) 5:1–9. 10.1038/srep08134
    1. Smith PM, Howitt MR, Panikov N, Michaud M, Gallini CA, Bohlooly-YM, et al. . The microbial metabolites, short-chain fatty acids, regulate colonic Treg cell homeostasis. Science. (2013) 341:569–73. 10.1126/science.1241165
    1. Tunaru S, Kero J, Schaub A, Wufka C, Blaukat A, Pfeffer K, et al. . PUMA-G and HM74 are receptors for nicotinic acid and mediate its anti-lipolytic effect. Nat Med. (2003) 9:352–5. 10.1038/nm824
    1. Thangaraju M, Cresci GA, Liu K, Ananth S, Gnanaprakasam JP, Browning DD, et al. . GPR109A is a g-protein–coupled receptor for the bacterial fermentation product butyrate and functions as a tumor suppressor in colon. Cancer Res. (2009) 69:2826–32. 10.1158/0008-5472.CAN-08-4466
    1. Chang PV, Hao L, Offermanns S, Medzhitov R. The microbial metabolite butyrate regulates intestinal macrophage function via histone deacetylase inhibition. Proc Natl Acad Sci USA. (2014) 111:2247–52. 10.1073/pnas.1322269111
    1. Singh N, Gurav A, Sivaprakasam S, Brady E, Padia R, Shi H, et al. . Activation of Gpr109a, receptor for niacin and the commensal metabolite butyrate, suppresses colonic inflammation and carcinogenesis. Immunity. (2014) 40:128–39. 10.1016/j.immuni.2013.12.007
    1. Park J, Kotani T, Konno T, Setiawan J, Kitamura Y, Imada S, et al. . Promotion of intestinal epithelial cell turnover by commensal bacteria: role of short-chain fatty acids. PLoS ONE. (2016) 11:e0156334. 10.1371/journal.pone.0156334
    1. Lukovac S, Belzer C, Pellis L, Keijser BJ, de Vos WM, Montijn RC, et al. . Differential modulation by Akkermansia muciniphila and Faecalibacterium prausnitzii of host peripheral lipid metabolism and histone acetylation in mouse gut organoids. MBio. (2014) 5:1–10. 10.1128/mBio.01438-14
    1. Kaiko GE, Ryu SH, Koues OI, Collins PL, Solnica-Krezel L, Pearce EJ, et al. . The colonic crypt protects stem cells from microbiota-derived metabolites. Cell. (2016) 165:1708–20. 10.1016/j.cell.2016.05.018
    1. Matthews GM, Howarth GS, Butler RN. Short-chain fatty acids induce apoptosis in colon cancer cells associated with changes to intracellular redox state and glucose metabolism. Chemotherapy. (2012) 58:102–9. 10.1159/000335672
    1. Tang Y, Chen Y, Jiang H, Nie D. Short-chain fatty acids induced autophagy serves as an adaptive strategy for retarding mitochondria-mediated apoptotic cell death. Cell Death Differ. (2011) 18:602–18. 10.1038/cdd.2010.117
    1. Zhang J, Yi M, Zha L, Chen S, Li Z, Li C, et al. . Sodium butyrate induces endoplasmic reticulum stress and autophagy in colorectal cells: implications for apoptosis. PLoS ONE. (2016) 11:1–25. 10.1371/journal.pone.0147218
    1. Wu X, Wu Y, He L, Wu L, Wang X, Liu Z. Effects of the intestinal microbial metabolite butyrate on the development of colorectal cancer. J Cancer. (2018) 9:2510–7. 10.7150/jca.25324
    1. Kelly CJ, Zheng L, Campbell EL, Saeedi B, Scholz CC, Bayless AJ, et al. . Crosstalk between microbiota-derived short-chain fatty acids and intestinal epithelial HIF augments tissue barrier function. Cell Host Microbe. (2015) 17:662–71. 10.1016/j.chom.2015.03.005
    1. Peng L, Li Z-R, Green RS, Holzman IR, Lin J. Butyrate enhances the intestinal barrier by facilitating tight junction assembly via activation of AMP-activated protein kinase in Caco-2 cell monolayers. J Nutr. (2009) 139:1619–25. 10.3945/jn.109.104638
    1. Miao W, Wu X, Wang K, Wang W, Wang Y, Li Z, et al. . Sodium butyrate promotes reassembly of tight junctions in Caco-2 monolayers involving inhibition of MLCK/MLC2 pathway and phosphorylation of PKCβ2. Int J Mol Sci. (2016) 17:1–12. 10.3390/ijms17101696
    1. Valenzano MC, DiGuilio K, Mercado J, Teter M, To J, Ferraro B, et al. . Remodeling of tight junctions and enhancement of barrier integrity of the CACO-2 intestinal epithelial cell layer by micronutrients. PLoS ONE. (2015) 10:e0133926. 10.1371/journal.pone.0133926
    1. Zheng L, Kelly CJ, Battista KD, Schaefer R, Lanis JM, Alexeev EE, et al. . Microbial-derived butyrate promotes epithelial barrier function through IL-10 receptor–dependent repression of claudin-2. J Immunol. (2017) 199:2976–84. 10.4049/jimmunol.1700105
    1. Wang HB, Wang PY, Wang X, Wan YL, Liu YC. Butyrate enhances intestinal epithelial barrier function via up-regulation of tight junction protein claudin-1 transcription. Dig Dis Sci. (2012) 57:3126–35. 10.1007/s10620-012-2259-4
    1. Yan H, Ajuwon KM. Butyrate modifies intestinal barrier function in IPEC-J2 cells through a selective upregulation of tight junction proteins and activation of the Akt signaling pathway. PLoS ONE. (2017) 12:1–20. 10.1371/journal.pone.0179586
    1. Park JS, Lee EJ, Lee JC, Kim WK, Kim HS. Anti-inflammatory effects of short chain fatty acids in IFN-γ-stimulated RAW 264.7 murine macrophage cells: involvement of NF-κB and ERK signaling pathways. Int Immunopharmacol. (2007) 7:70–77. 10.1016/j.intimp.2006.08.015
    1. Lührs H, Gerke T, Müller JG, Melcher R, Schauber J, Boxberger F, et al. . Butyrate inhibits NF-κB activation in lamina propria macrophages of patients with ulcerative colitis. Scand J Gastroenterol. (2002) 37:458–66. 10.1080/003655202317316105
    1. Arpaia N, Campbell C, Fan X, Dikiy S, Van Der Veeken J, Deroos P, et al. . Metabolites produced by commensal bacteria promote peripheral regulatory T-cell generation. Nature. (2013) 504:451–5. 10.1038/nature12726
    1. Furusawa Y, Obata Y, Fukuda S, Endo TA, Nakato G, Takahashi D, et al. . Commensal microbe-derived butyrate induces the differentiation of colonic regulatory T cells. Nature. (2013) 504:446–50. 10.1038/nature12721
    1. Geirnaert A, Calatayud M, Grootaert C, Laukens D, Devriese S, Smagghe G, et al. . Butyrate-producing bacteria supplemented in vitro to Crohn's disease patient microbiota increased butyrate production and enhanced intestinal epithelial barrier integrity. Sci Rep. (2017) 7:11450. 10.1038/s41598-017-11734-8
    1. Zhao Y, Chen F, Wu W, Sun M, Bilotta AJ, Yao S, et al. GPR43 mediates microbiota metabolite SCFA regulation of antimicrobial peptide expression in intestinal epithelial cells via activation of mTOR and STAT3. Mucosal Immunol. (2018) 1–11: 752–62. 10.1038/mi.2017.118
    1. Roediger WEW. Role of anaerobic bacteria in the metabolic welfare of the colonic mucosa in man. Gut. (1980) 21:793–8. 10.1136/gut.21.9.793
    1. Macia L, Tan J, Vieira AT, Leach K, Stanley D, Luong S, et al. . Metabolite-sensing receptors GPR43 and GPR109A facilitate dietary fibre-induced gut homeostasis through regulation of the inflammasome. Nat Commun. (2015) 6:1–15. 10.1038/ncomms7734
    1. Borthakur A, Priyamvada S, Kumar A, Natarajan AA, Gill RK, Alrefai WA, et al. . A novel nutrient sensing mechanism underlies substrate-induced regulation of monocarboxylate transporter-1. AJP Gastrointest Liver Physiol. (2012) 303:G1126–33. 10.1152/ajpgi.00308.2012
    1. Sahasrabudhe NM, Beukema M, Tian L, Troost B, Scholte J, Bruininx E, et al. . Dietary fiber pectin directly blocks toll-like receptor 2-1 and prevents doxorubicin-induced ileitis. Front Immunol. (2018) 9:1–19. 10.3389/fimmu.2018.00383
    1. Ji J, Shu D, Zheng M, Wang J, Luo C, Wang Y, et al. . Microbial metabolite butyrate facilitates M2 macrophage polarization and function. Sci Rep. (2016) 6:1–10. 10.1038/srep24838
    1. Fernando MR, Saxena A, Reyes J-L, McKay DM. Butyrate enhances antibacterial effects while suppressing other features of alternative activation in IL-4-induced macrophages. Am J Physiol Gastrointest Liver Physiol. (2016) 310:G822–31. 10.1152/ajpgi.00440.2015
    1. Riggs MG, Whittaker RG, Neumann JR, Ingram VM. n-Butyrate causes histone modification in HeLa and Friend erythroleukaemia cells. Nature. (1977) 268:462–4. 10.1038/268462a0
    1. Boffa LC, Vidali G, Mann RS, Allfrey VG. Suppression of histone deacetylation in vivo and in vitro by sodium butyrate. J Biol Chem. (1978) 253:3364–6.
    1. Vidali G, Boffa LC, Bradbury EM, Allfrey VG. Butyrate suppression of histone deacetylation leads to accumulation of multiacetylated forms of histones H3 and H4 and increased DNase I sensitivity of the associated DNA sequences. Proc Natl Acad Sci USA. (1978) 75:2239–43. 10.1073/pnas.75.5.2239
    1. Candido EPM, Reeves R, Davie JR. Sodium butyrate inhibits histone deacetylation in cultured cells. Cell. (1978) 14:105–13. 10.1016/0092-8674(78)90305-7
    1. Davie JR. Inhibition of histone deacetylase activity by butyrate. J Nutr. (2003) 133:2485S−93S. 10.1093/jn/133.7.2485S
    1. Shakespear MR, Halili MA, Irvine KM, Fairlie DP, Sweet MJ. Histone deacetylases as regulators of inflammation and immunity. Trends Immunol. (2011) 32:335–43. 10.1016/j.it.2011.04.001
    1. Glauben R, Batra A, Fedke I, Zeitz M, Lehr HA, Leoni F, et al. . Histone hyperacetylation is associated with amelioration of experimental colitis in mice. J Immunol. (2006) 176:5015–22. 10.4049/jimmunol.176.8.5015
    1. Cresci GA, Thangaraju M, Mellinger JD, Liu K, Ganapathy V. Colonic gene expression in conventional and germ-free mice with a focus on the butyrate receptor GPR109A and the butyrate transporter SLC5A8. J Gastrointest Surg. (2010) 14:449–61. 10.1007/s11605-009-1045-x
    1. Villodre Tudela C, Boudry C, Stumpff F, Aschenbach JR, Vahjen W, Zentek J, et al. . Down-regulation of monocarboxylate transporter 1 (MCT1) gene expression in the colon of piglets is linked to bacterial protein fermentation and pro-inflammatory cytokine-mediated signalling. Br J Nutr. (2015) 113:610–7. 10.1017/S0007114514004231
    1. Cuff MA, Lambert DW, Shirazi-Beechey SP. Substrate-induced regulation of the human colonic monocarboxylate transporter, MCT1. J Physiol. (2002) 539:361–71. 10.1113/jphysiol.2001.014241
    1. Thibault R, De Coppet P, Daly K, Bourreille A, Cuff M, Bonnet C, et al. . Down-regulation of the monocarboxylate transporter 1 is involved in butyrate deficiency during intestinal inflammation. Gastroenterology. (2007) 133:1916–27. 10.1053/j.gastro.2007.08.041
    1. De Preter V, Arijs I, Windey K, Vanhove W, Vermeire S, Schuit F, et al. . Impaired butyrate oxidation in ulcerative colitis is due to decreased butyrate uptake and a defect in the oxidation pathway. Inflamm Bowel Dis. (2012) 18:1127–36. 10.1002/ibd.21894
    1. Planell N, Lozano JJ, Mora-Buch R, Masamunt MC, Jimeno M, Ordás I, et al. . Transcriptional analysis of the intestinal mucosa of patients with ulcerative colitis in remission reveals lasting epithelial cell alterations. Gut. (2013) 62:967–76. 10.1136/gutjnl-2012-303333
    1. Palmieri O, Creanza TM, Bossa F, Palumbo O, Maglietta R, Ancona N, et al. . Genome-wide pathway analysis using gene expression data of colonic mucosa in patients with inflammatory bowel disease. Inflamm Bowel Dis. (2015) 21:1260–8. 10.1097/MIB.0000000000000370
    1. Nancey S, Moussata D, Graber I, Claudel S, Saurin J-C, Flourié B. Tumor necrosis factor alpha reduces butyrate oxidation in vitro in human colonic mucosa: a link from inflammatory process to mucosal damage? Inflamm Bowel Dis. (2005) 11:559–66. 10.1097/01.MIB.0000161918.04760.f3
    1. Masui R, Sasaki M, Funaki Y, Ogasawara N, Mizuno M, Iida A, et al. . G protein-coupled receptor 43 moderates gut inflammation through cytokine regulation from mononuclear cells. Inflamm Bowel Dis. (2013) 19:2848–56. 10.1097/01.MIB.0000435444.14860.ea
    1. Sina C, Gavrilova O, Forster M, Till A, Derer S, Hildebrand F, et al. . G protein-coupled receptor 43 is essential for neutrophil recruitment during intestinal inflammation. J Immunol. (2009) 183:7514–22. 10.4049/jimmunol.0900063
    1. Kim MH, Kang SG, Park JH, Yanagisawa M, Kim CH. Short-chain fatty acids activate GPR41 and GPR43 on intestinal epithelial cells to promote inflammatory responses in mice. Gastroenterology. (2013) 145:396–406.e10. 10.1053/j.gastro.2013.04.056
    1. Alex P, Zachos NC, Nguyen T, Gonzales L, Chen TE, Conklin LS, et al. . Distinct cytokine patterns identified from multiplex profiles of murine DSS and TNBS-induced colitis. Inflamm Bowel Dis. (2009) 15:341–52. 10.1002/ibd.20753
    1. Wirtz S, Popp V, Kindermann M, Gerlach K, Weigmann B, Fichtner-Feigl S, et al. . Chemically induced mouse models of acute and chronic intestinal inflammation. Nat Protoc. (2017) 12:1295–309. 10.1038/nprot.2017.044
    1. Higgins LM, Frankel G, Douce G, Dougan G, MacDonald TT. Citrobacter rodentium infection in mice elicits a mucosal Th1 cytokine response and lesions similar to those in murine inflammatory bowel disease. Infect Immun. (1999) 67:3031–9.
    1. Mamontov P, Neiman E, Cao T, Perrigoue J, Friedman J, Das A MJ. Effects of short chain fatty acids and GPR43 stimulation on human Treg function (IRC5P.631). J Immunol. (2015) 194:58.14.
    1. Park J, Kim M, Kang SG, Jannasch AH, Cooper B, Patterson J, et al. . Short-chain fatty acids induce both effector and regulatory T cells by suppression of histone deacetylases and regulation of the mTOR-S6K pathway. Mucosal Immunol. (2014) 8:80–93. 10.1038/mi.2014.44
    1. Kespohl M, Vachharajani N, Luu M, Harb H, Pautz S, Wolff S, et al. . The microbial metabolite butyrate induces expression of Th1- associated factors in cD4+T cells. Front Immunol. (2017) 8:1–12. 10.3389/fimmu.2017.01036
    1. Feingold KR, Moser A, Shigenaga JK, Grunfeld C. Inflammation stimulates niacin receptor (GPR109A/HCA2) expression in adipose tissue and macrophages. J Lipid Res. (2014) 55:2501–8. 10.1194/jlr.M050955
    1. Frank DN, St Amand AL, Feldman RA, Boedeker CE, Harpaz N, Pace NR. Molecular-phylogenetic characterization of microbial community imbalances in human inflammatory bowel diseases. Proc Natl Am Sci USA. (2013) 104:13780–5. 10.1073/pnas.0706625104
    1. Imhann F, Vila AV, Bonder MJ, Fu J, Gevers D, Visschedijk MC, et al. . Interplay of host genetics and gut microbiota underlying the onset and clinical presentation of inflammatory bowel disease. Gut. (2016) 67:108–19. 10.1136/gutjnl-2016-312135
    1. Prosberg M, Bendtsen F, Vind I, Petersen AM, Gluud LL. The association between the gut microbiota and the inflammatory bowel disease activity: a systematic review and meta-analysis. Scand J Gastroenterol. (2016) 51:1407–15. 10.1080/00365521.2016.1216587
    1. Gupta VK, Paul S, Dutta C. Geography, ethnicity or subsistence-specific variations in human microbiome composition and diversity. Front Microbiol. (2017) 8:1162. 10.3389/fmicb.2017.01162
    1. Gill PA, van Zelm MC, Muir JG, Gibson PR. Review article: short chain fatty acids as potential therapeutic agents in human gastrointestinal and inflammatory disorders. Aliment Pharmacol Ther. (2018) 48:15–34. 10.1111/apt.14689
    1. Hamer HM, Jonkers D, Venema K, Vanhoutvin S, Troost FJ, Brummer RJ. Review article: the role of butyrate on colonic function. Aliment Pharmacol Ther. (2008) 27:104–19. 10.1111/j.1365-2036.2007.03562.x
    1. Scheppach W, Sommer H, Kirchner T, Paganelli GM, Bartram P, Christl S, et al. . Effect of butyrate enemas on the colonic mucosa in distal ulcerative colitis. Gastroenterology. (1992) 103:51–56. 10.1016/0016-5085(92)91094-K
    1. Harig JM, Soergel KH, Komorowski RA, Wood CM. Treatment of diversion colitis with short-chain-fatty acid irrigation. N Engl J Med. (1989) 320:23–8. 10.1056/NEJM198901053200105
    1. Segain JP, Raingeard de la Blétière D, Bourreille A, Leray V, Gervois N, Rosales C, et al. . Butyrate inhibits inflammatory responses through NFkappaB inhibition: implications for Crohn's disease. Gut. (2000) 47:397–403. 10.1136/gut.47.3.397
    1. Breuer RI, Soergel KH, Lashner BA, Christ ML, Hanauer SB, Vanagunas A, et al. . Short chain fatty acid rectal irrigation for left-sided ulcerative colitis: a randomised, placebo controlled trial. Gut. (1997) 40:485–91. 10.1136/gut.40.4.485
    1. Hamer HM, Jonkers DMAE, Vanhoutvin SALW, Troost FJ, Rijkers G, de Bruïne A, et al. . Effect of Butyrate enemas on inflammation and antioxidant status in the colonic mucosa of patients with ulcerative colitis in remission. Clin Nutr. (2010) 29:738–44. 10.1016/j.clnu.2010.04.002
    1. Guillemot FÇ, Colombel JF, Neut C, Verplanck N, Lecomte M, Romond C, et al. . Treatment of diversion colitis by short-chain fatty acids - Prospective and double-blind study. Dis Colon Rectum. (1991) 34:861–4. 10.1007/BF02049697
    1. Tarrerias AL, Millecamps M, Alloui A, Beaughard C, Kemeny JL, Bourdu S, et al. . Short-chain fatty acid enemas fail to decrease colonic hypersensitivity and inflammation in TNBS-induced colonic inflammation in rats. Pain. (2002) 100:91–7. 10.1016/S0304-3959(02)00234-8
    1. Sokol H, Pigneur B, Watterlot L, Lakhdari O, Bermudez-Humaran LG, Gratadoux J-J, et al. . Faecalibacterium prausnitzii is an anti-inflammatory commensal bacterium identified by gut microbiota analysis of Crohn disease patients. Proc Natl Acad Sci USA. (2008) 105:16731–6. 10.1073/pnas.0804812105
    1. Araki Y, Andoh A, Takizawa J, Takizawa W, Fujiyama Y. Clostridium butyricum, a probiotic derivative, suppresses dextran sulfate sodium-induced experimental colitis in rats. Int J Mol Med. (2004) 13:577–80. 10.3892/ijmm.13.4.577
    1. Gibson GR, Hutkins R, Sanders ME, Prescott SL, Reimer RA, Salminen SJ, et al. . Expert consensus document: the international scientific association for probiotics and prebiotics (ISAPP) consensus statement on the definition and scope of prebiotics. Nat Rev Gastroenterol Hepatol. (2017) 14:491–502. 10.1038/nrgastro.2017.75
    1. Gibson GR, Probert HM, Van Loo J, Rasrall RA, Roberfroid MB. Dietary modulation of the human colonic microbiota: introducing the concept of prebiotics. Nutr Res Rev. (2004) 17:259–75. 10.1079/NRR200479
    1. Hallert C, Björck I, Nyman M, Pousette A, Grännö C, Svensson H. Increasing fecal butyrate in ulcerative colitis patients by diet: controlled pilot study. Inflamm Bowel Dis. (2003) 9:116–21. 10.1097/00054725-200303000-00005
    1. Casellas F, Borruel N, Torrejon A, Varela E, Antolin M, Guarner F, et al. . Oral oligofructose-enriched inulin supplementation in acute ulcerative colitis is well tolerated and associated with lowered faecal calprotectin. Aliment Pharmacol Ther. (2007) 25:1061–7. 10.1111/j.1365-2036.2007.03288.x
    1. Swennen K, Courtin CM, Delcour JA. Non-digestible oligosaccharides with prebiotic properties. Crit Rev Food Sci Nutr. (2006) 46:459–71. 10.1080/10408390500215746
    1. Looijer–Van Langen MAC, Dieleman LA. Prebiotics in Chronic Intestinal Inflammation. Inflamm Bowel Dis. (2009) 15:454–62. 10.1002/ibd.20737
    1. Rasmussen HE, Hamaker BR. Prebiotics and Inflammatory Bowel Disease. Gastroenterol Clin North Am. (2017) 46:783–95. 10.1016/j.gtc.2017.08.004
    1. Wéra O, Lancellotti P, Oury C. The dual role of neutrophils in inflammatory bowel diseases. J Clin Med. (2016) 5:118. 10.3390/jcm5120118
    1. Zhao M, Zhu W, Gong J, Zuo L, Zhao J, Sun J, et al. . Dietary fiber intake is associated with increased colonic mucosal GPR43+ polymorphonuclear infiltration in active Crohn's disease. Nutrients. (2015) 7:5327–46. 10.3390/nu7075223
    1. Sasaki D, Sasaki K, Ikuta N, Yasuda T, Fukuda I, Kondo A, et al. . Low amounts of dietary fibre increase in vitro production of short-chain fatty acids without changing human colonic microbiota structure. Sci Rep. (2018) 8:1–7. 10.1038/s41598-017-18877-8
    1. Hill C, Guarner F, Reid G, Gibson GR, Merenstein DJ, Pot B, et al. . Expert consensus document: the international scientific association for probiotics and prebiotics consensus statement on the scope and appropriate use of the term probiotic. Nat Rev Gastroenterol Hepatol. (2014) 11:506–14. 10.1038/nrgastro.2014.66
    1. Ganji-Arjenaki M, Rafieian-Kopaei M. Probiotics are a good choice in remission of inflammatory bowel diseases: a meta analysis and systematic review. J Cell Physiol. (2018) 233:2091–103. 10.1002/jcp.25911
    1. Derwa Y, Gracie DJ, Hamlin PJ, Ford AC. Systematic review with meta-analysis: the efficacy of probiotics in inflammatory bowel disease. Aliment Pharmacol Ther. (2017) 46:389–400. 10.1111/apt.14203
    1. Bibiloni R, Fedorak RN, Tannock GW, Madsen KL, Gionchetti P, Campieri M, et al. VSL#3 probiotic-mixture induces remission in patients with active ulcerative colitis. Am J Gastroenterol. (2005) 100:1539–46. 10.1111/j.1572-0241.2005.41794.x
    1. Sood A, Midha V, Makharia GK, Ahuja V, Singal D, Goswami P, et al. The probiotic preparation, VSL#3 induces remission in patients with mild-to-moderately active ulcerative colitis. Clin Gastroenterol Hepatol. (2009) 7:1202–9.e1. 10.1016/j.cgh.2009.07.016
    1. Kruis W, Frič P, Pokrotnieks J, Lukáš M, Fixa B, Kaščák M, et al. . Maintaining remission of ulcerative colitis with the probiotic Escherichia coli Nissle 1917 is as effective as with standard mesalazine. Gut. (2004) 53:1617–23. 10.1136/gut.2003.037747
    1. Groeger D, O'Mahony L, Murphy EF, Bourke JF, Dinan TG, Kiely B, et al. . Bifidobacterium infantis 35624 modulates host inflammatory processes beyond the gut. Gut Microbes. (2013) 4:325–339. 10.4161/gmic.25487

Source: PubMed

3
订阅