A microbial metabolite remodels the gut-liver axis following bariatric surgery

Snehal N Chaudhari, James N Luo, David A Harris, Hassan Aliakbarian, Lina Yao, Donggi Paik, Renuka Subramaniam, Arijit A Adhikari, Ashley H Vernon, Ayse Kiliç, Scott T Weiss, Jun R Huh, Eric G Sheu, A Sloan Devlin, Snehal N Chaudhari, James N Luo, David A Harris, Hassan Aliakbarian, Lina Yao, Donggi Paik, Renuka Subramaniam, Arijit A Adhikari, Ashley H Vernon, Ayse Kiliç, Scott T Weiss, Jun R Huh, Eric G Sheu, A Sloan Devlin

Abstract

Bariatric surgery is the most effective treatment for type 2 diabetes and is associated with changes in gut metabolites. Previous work uncovered a gut-restricted TGR5 agonist with anti-diabetic properties-cholic acid-7-sulfate (CA7S)-that is elevated following sleeve gastrectomy (SG). Here, we elucidate a microbiome-dependent pathway by which SG increases CA7S production. We show that a microbial metabolite, lithocholic acid (LCA), is increased in murine portal veins post-SG and by activating the vitamin D receptor, induces hepatic mSult2A1/hSULT2A expression to drive CA7S production. An SG-induced shift in the microbiome increases gut expression of the bile acid transporters Asbt and Ostα, which in turn facilitate selective transport of LCA across the gut epithelium. Cecal microbiota transplant from SG animals is sufficient to recreate the pathway in germ-free (GF) animals. Activation of this gut-liver pathway leads to CA7S synthesis and GLP-1 secretion, causally connecting a microbial metabolite with the improvement of diabetic phenotypes.

Keywords: bariatric surgery; bile acids; gut-liver axis; microbiome.

Conflict of interest statement

Declaration of interests CA7S is a subject of patents held by HMS and BWH on which S.N.C., D.A.H., E.G.S., and A.S.D. are inventors. A.S.D. is a consultant for Takeda Pharmaceuticals and HP Hood. E.G.S. is a consultant for Vicarious Surgical, Inc. and was previously on the scientific advisory board of Kitotech, Inc.

Copyright © 2020 Elsevier Inc. All rights reserved.

Figures

Figure 1.. Sleeve Gastrectomy (SG)-mediated increase in…
Figure 1.. Sleeve Gastrectomy (SG)-mediated increase in levels of CA7S and GLP-1 requires a microbiome
(A) Schematic of SG and sham surgery in diet-induced obese (DIO) mice treated with or without antibiotics (Abx.). (B,C) CA7S levels in cecal contents of mice subjected to SG or Sham surgeries treated with or without antibiotics (Abx.) (B, Sham n=7, SG n=6, *p=0.01; C, n=5 in each group, ns=not significant p=0.26, Welch’s t test). (D,E) CA7S levels in cecal contents (D) and liver (E) of conventional DIO mice were significantly higher than antibiotic-treated (+ Abx.) or GF DIO mice (Cecum: DIO, n=10, DIO + Abx., n=9, DIO; GF, n=8, DIO vs. DIO + Abx. ****p<1.00x10−4, DIO vs. DIO;GF ****p<1.00x10−4; Liver: DIO, n=9, DIO + Abx., n=9, DIO; GF, n=8; DIO vs. DIO + Abx. ***p=2.00x10−4, DIO vs. DIO; GF ***p=5.00x10−4, one-way ANOVA followed by Dunnett’s multiple comparisons test). (F,G) GLP-1 levels in systemic circulation of mice subjected to SG or Sham surgeries treated with or without antibiotics. (Data not marked with asterisk(s) are not significant. For F, Sham n=3, SG n=6, *p=0.01; for G, n=5 in each group, p=0.26, Welch’s t test). (H) The mammalian sulfotransferase enzyme SULT2A (mSULT2A in mice and hSULT2A in humans) catalyzes the conversion of the primary bile acid cholic acid (CA) to cholic acid-7-sulfate (CA7S). (I) qRT-PCR quantification of mSult2A isoforms reported to sulfate bile acids in the murine liver. Expression levels were normalized to 18S (data not marked with asterisk(s) are not significant. n=11 in each group, mSult2A1 *p=0.04, mSult2A2 p=0.92, mSult2A8 p=0.78, Welch’s t test). (J) As measured by qRT-PCR, the hepatic expression of mSult2A1 was significantly higher in DIO mice than DIO + Abx. mice and DIO;GF mice. Expression levels were normalized to mouse ribosomal 18S (DIO, n=9, DIO + Abx., n=9, DIO;GF, n=8; DIO vs. DIO + Abx. ****p<1x10−4, DIO vs. DIO;GF ****p<1x10−4, one-way ANOVA followed by Dunnett’s multiple comparisons test). All data are presented as mean ± SEM.
Figure 2.. Portal vein BAs induce expression…
Figure 2.. Portal vein BAs induce expression of hSULT2A1 in hepatocytes in vitro
(A) Schematic of portal vein BA analysis in indicated groups of mice. (B) Portal vein BAs in DIO mice 6 weeks post-Sham and SG (n=21 in each group; data not marked with asterisk(s) are not significant). All bile acids with measurable concentrations above the limit of detection are shown. Lithocholic acid (LCA) was the only bile acid whose levels were significantly increased post-SG (*p=0.01). (Total bile acids, p=0.51, α/βMCA, alpha-muricholic acid and beta-muricholic acid, p=0.35, Tα/βMCA, tauro-alpha- and tauro-beta-muricholic acid, p=0.08; TCA, tauro-cholic acid, p=0.20; TCDCA, tauro-chenodeoxycholic acid, p=0.07; TωMCA, tauro-omega-muricholic acid, p=0.11; CA, cholic acid, p=0.78; CDCA, chenodeoxycholic acid, p=0.45; TDCA, tauro-deoxycholic acid, p=0.45; Welch’s t test). (C) Portal vein BAs in DIO mice treated with antibiotics post-sham and SG (n=5 in each group; data not marked with asterisk(s) are not significant. LCA, CA, TDCA, and CDCA were undetectable in portal veins of both groups. (Total bile acids, p=0.81, α/βMCA, alpha-muricholic acid and beta-muricholic acid, p=0.43; Tα/βMCA, tauro-alpha- and tauro-beta-muricholic acid, p=0.93; TCA, tauro-cholic acid, p=0.36; TCDCA, tauro-chenodeoxycholic acid, p=0.23; TωMCA, tauro-omega-muricholic acid, p=0.98; CA, cholic acid; CDCA, chenodeoxycholic acid; TDCA, tauro-deoxycholic acid; LCA, lithocholic acid, not detected (n.d.), Welch’s t test). (D-F) Concentrated pools of bile acids mimicking the mean physiological ratios of individual portal vein bile acids measured in conventional sham and SG and antibiotic-treated sham and SG mice were generated in vitro in DMSO. HepG2 cells were treated with dilutions of these pools (total bile acid concentrations of 100 μM, 500 μM, and 1000 μM). (D) As measured by qRT-PCR, the expression of hSULT2A was increased in conventional SG PV BA-treated cells compared to conventional sham PV BA-treated cells. For each concentration, expression of sham hSULT2A was normalized to 1. hSULT2A expression was normalized to human GAPDH (≥3 biological replicates per condition, SG PV bile acids, 100 μM *p=0.01, 500 μM *p=0.02, 1000 μM **p=1.00x10−3, Welch’s t test). (E) There were no significant differences in hSULT2A expression in cells incubated with Abx-treated SG and sham PV BA pools. For each concentration, expression of sham hSULT2A was normalized to 1. hSULT2A expression was normalized to human GAPDH (≥3 biological replicates per condition, SG-Abx. PV bile acids, ns=not significant, 100 μM p=0.82, 500 μM p=0.37, 1000 μM p=0.60, Welch’s t test). (F) As measured by qRT-PCR, the expression of hSULT2A was increased in conventional sham PV BA-treated cells and decreased in antibiotic sham PV BA-treated cells relative to DMSO control. hSULT2A expression was normalized to human GAPDH (≥3 biological replicates per condition, data not marked with asterisk(s) are not significant, 100 μM DMSO vs. Sham **p=5.60x10−3; 100 μM DMSO vs. Sham-Abx. p=0.82; 100 μM Sham vs. Sham-Abx. ***p=1.00x10−4; 500 μM DMSO vs. Sham p=0.29; 500 μM DMSO vs. Sham-Abx. p=0.98; 500 μM Sham vs. Sham-Abx. *p=0.04; 1000 μM DMSO vs. Sham *p=0.03; 1000 μM DMSO vs. Sham-Abx. p=0.97; 1000 μM Sham vs. Sham-Abx. **p=2.30x10−3, two-way ANOVA followed by Dunnett’s multiple comparisons test). All data are presented as mean ± SEM.
Figure 3.. LCA induces expression of SULT…
Figure 3.. LCA induces expression of SULT via the Vitamin D receptor (VDR), resulting in production of CA7S and GLP-1 secretion
(A) qRT-PCR quantification of hSULT2A expression level in HepG2 cells treated with indicated concentrations of CA, TDCA, CDCA, or LCA normalized to human GAPDH. LCA increased hSULT2A expression in a dose-dependent manner relative to DMSO control. (≥3 biological replicates per condition, data marked with asterisk(s) are only for induction of hSULT2A. CA, 0.1 μM **p=8.90x10−3, 1 μM **p=5.40x10−3, 10 μM **p=5.30x10−3, 100 μM **p=8.40x10−3; TDCA, 0.1 μM ****p=1.00x10−4, 1 μM ****p=1.00x10−4, 10 μM ****p=1.00x10−4, 100 μM ****p=1.00x10−4; CDCA, 0.1 μM **p=1.80x10−3, 1 μM **p=2.50x10−3, 10 μM **p=9.30x10−3, 100 μM p=0.21; LCA, 0.1 μM p=0.92, 1 μM p=0.95, 10 μM **p=5.86x10−3, 100 μM *p=0.04, one-way ANOVA followed by Dunnett’s multiple comparisons test). (B) siRNA-mediated knockdown of VDR significantly reduced LCA-mediated induction of hSULT2A in HepG2 cells compared to negative control siRNA (≥3 biological replicates per condition, *p=0.01, Welch’s t test). (C) Vdr expression levels in mouse livers were increased in SG compared to sham mice as determined by qRT-PCR. Expression was normalized to mouse ribosomal 18S (n=11 in each group; *p=0.02, Welch’s t test). (D) Hepatic expression of Vdr was increased in DIO mice compared to DIO + Abx. mice and DIO; GF mice. Expression levels were normalized to mouse ribosomal 18S (DIO, n=9, DIO + Abx., n=10, DIO;GF, n=8; DIO vs. DIO + Abx. **p=3.50x10−3, DIO vs. DIO;GF **p=3.90x10−3, one-way ANOVA followed by Dunnett’s multiple comparisons test). (E) CA7S levels were reduced in feces of Vdr-knockout (KO) mice compared to wild-type (WT) mice (female WT, n=4, Vdr-KO, n=7, *p=0.01, male WT, n=7, Vdr-KO, n=9, *p=0.01, Welch’s t test). (F) Schematic of portal vein injection with LCA (GB=gallbladder). (G,H) Quantitative real time PCR quantification of mSult2A1 (G) and Vdr (H) expression levels in mouse livers injected with LCA or PBS normalized to mouse ribosomal 18S (n=3; for (G) *p=0.02, for (H) *p=0.03, Welch’s t test). (I) CA7S levels in the gallbladder of mice injected with LCA or PBS normalized to mouse ribosomal 18S (n=3; *p=0.02, Welch’s t test). (J) Synthesis of CA7S in HepG2 cells requires the cofactor PAPS, is induced upon incubation with LCA, and is dependent on VDR. Vdr siRNA was used to knockdown Vdr expression. 48 hours post-knockdown, substrate CA and ligand LCA was added as indicated. CA7S production was quantified by UPLC-MS after 16 hours. (≥3 biological replicates per condition, data marked with asterisk(s) are only for production of CA7S; CA+LCA+PAPS vs. CA+PAPS **p=2.70x10−3, CA+LCA+PAPS vs. CA+LCA+PAPS+Vdr siRNA **p=1.60x10−3, one-way ANOVA followed by Dunnett’s multiple comparisons test). (K) Schematic of co-culture study. Liver HepG2 cells were cultured in the basolateral chamber, while NCI-H716 enteroendocrine cells were cultured separately in transwell inserts, and combined to measure GLP-1 secretion for indicated treatments. (L) Secretion of GLP-1 by NCI-H716 cells co-cultured with HepG2 cells was induced by addition of LCA, CA, and PAPS to liver cells and was reduced by siRNA-mediated knockdown of VDR in liver cells. (≥3 biological replicates per condition, data not marked by asterisk(s) are not significant, DMSO vs. CA p=0.96; DMSO vs. CA+PAPS p=0.26; DMSO vs. CA+LCA p=0.12; DMSO vs. CA+LCA+PAPS ****p<1.00x10−4; DMSO vs. CA+LCA+PAPS+VDR siRNA p=0.49; CA vs. CA+LCA+PAPS ***p=3.00x10−3; CA+LCA vs. CA+LCA+PAPS *p=0.04; CA+PAPS vs. CA+LCA+PAPS *p=0.01; CA+LCA+PAPS vs. CA+LCA+PAPS+VDR siRNA *p=0.02, one-way ANOVA followed by Dunnett’s multiple comparisons test). All data are presented as mean ± SEM.
Figure 4.. The production of the secondary…
Figure 4.. The production of the secondary bile acid LCA is decreased in mice and humans post-SG
(A) Host-produced conjugated bile acids are released into the gut post-prandially and converted into secondary bile acids through the enzymatic activity of gut bacteria. Deconjugation by bacterial bile salt hydrolases followed by 7α-dehydroxylation by enzymes encoded by the bile acid inducible (bai) operon converts cholic acid (CA) and chenodeoxycholic acid (CDCA) into deoxycholic acid (DCA) and lithocholic acid (LCA), respectively. (B) Schematic and timeline of sham and SG mouse cecal samples subjected to 16S sequencing. (C) Stacked bar plot showing mean relative abundances of phylum-level taxa in mouse samples (UK = Unknown, Sham, n=15; SG, n=17). (D) Relative abundance of Clostridiales in sham and SG mice (Sham, n=15; SG, n=17, p=0.16, ns=not significant, Welch’s t test). (E) qRT-PCR quantification of baiCD gene cluster expression levels in sham and SG mouse cecal contents normalized to bacterial 16S ribosomal DNA (Sham, n=10, SG, n=14, *p=0.02, Welch’s t test). (F) Schematic and timeline of human pre- and post-SG fecal samples subjected to 16S sequencing. (G) Stacked bar plot showing mean relative abundances of phylum-level taxa in human samples (UK = Unknown, n=17 patients). (H) Relative abundance of Clostridiales in pre- and post-SG human feces (n=17 patients, **p=4.80x10−3, paired t test). All data are presented as mean ± SEM.
Figure 5.. Intestinal BA transport proteins ASBT…
Figure 5.. Intestinal BA transport proteins ASBT and OSTα facilitate selective transport of LCA into the portal vein
(A) Schematic of proteins involved in BA transport from the intestinal lumen into the portal vein. (B) qRT-PCR quantification of BA transport protein expression levels in sham and SG mouse distal ileum normalized to mouse ribosomal 18S. The expression of Asbt (apical sodium-dependent bile acid transporter) and Ostα (organic solute transporter α) were significantly increased post-SG. (Sham, n=15; SG, n=17; Asbt **p=8.80x10−3, Ostα *p=0.04, Ostβ p=0.89, iBabp p=0.76, Bsep p=0.64, Mrp1 p=0.89, Mrp2 p=0.89, Mrp3 p=0.35, Oatp1 p=0.78, Oatp2 p=0.68, Oatp4 p=0.84, ns=not significant, Welch’s t test). (C) SEM images of undifferentiated and differentiated Caco-2 cells in transwells. Scale bars indicate (L to R): 400 μm, 20 μm, and 4 μm. (D) Schematic of BA transport study. Caco-2 cells differentiated in transwells were treated with a defined mixture of indicated BAs at 10 μM each, followed by measurement of BA transport to the basolateral chamber. (E,F) Transport of indicated BAs from the apical chamber across differentiated Caco-2 cells into the basolateral chamber over 12 hours as measured by UPLC-MS. Area-under-the-curve graphs in (F) correspond to timecourses directly above in (E). siRNA-mediated knockdown of ASBT, OSTα, or ASBT+OSTα reduced transport of LCA and TCA, increased transport of DCA, and did not affect transport of CA, CDCA, βMCA, or TβMCA. Treatment with U0126 (50 μM), a small molecule that increases expression of ASBT, increased transport of LCA but not the other BAs tested. (≥3 biological replicates per condition, *p<0.05, #p<0.01 †p<1.00x10−3, data not marked are not significant; CA (ASBT) p=0.99, CA (OSTα) p=0.91, CA (ASBT+OSTα) p=0.97, CA (U0126) p=0.65; CDCA (ASBT) p=0.78, CDCA (OSTα) p=0.73, CDCA (ASBT+OSTα) p=0.73, CDCA (U0126) p=0.38; βMCA (ASBT) p=0.81, βMCA (OSTα) p=0.74, βMCA (ASBT+OSTα) p=0.81, βMCA (U0126) p=0.99; TβMCA (ASBT) p=0.99, TβMCA (OSTα) p=0.66, TβMCA (ASBT+OSTα) p=0.83, TβMCA (U0126) p=0.31; TCA (ASBT) *p=0.01, TCA (OSTα) †p=7.00x10−4, TCA (ASBT+OSTα) †p=1.70x10−4, TCA (U0126) †p=1.00x10−4; DCA (ASBT) †p=6.00x10−4, DCA (OSTα) *p=0.02, DCA (ASBT+OSTα) *p=0.04, DCA (U0126) p=0.08; LCA (ASBT) *p=0.03, LCA (OSTα) *p=0.02, LCA (ASBT+OSTα) *p=0.01, LCA (U0126) #p=2.80x10−3, one-way ANOVA followed by Turkey’s (HSD) post-hoc test). All data are presented as mean ± SEM.
Figure 6.. LCA is sufficient to inhibit…
Figure 6.. LCA is sufficient to inhibit Asbt expression and induce production of CA7S
(A) Schematic of sham and SG intestine BA transport modulated by the production of LCA by Clostridia. Levels of Clostridia and LCA production are higher in sham mice. LCA inhibits Asbt expression, resulting in less transport of LCA into the portal vein. In contrast, levels of Clostridia and LCA are lower in SG mice, allowing for higher expression of Asbt and increased transport of LCA from the gut to the liver in SG animals. (B) The sham cecal pool of BAs inhibited expression of ASBT compared to the SG cecal pool of BAs (3 biological replicates per condition, Sham vs. SG cecal pool *p=0.02, one-way ANOVA followed by Dunnett’s multiple comparisons test (C) LCA (100 μM) inhibited expression of ASBT in Caco-2 cells. (3 biological replicates per condition, *p=0.02, Welch’s t test). (D) Schematic of GF mice administered 0.3% LCA (w/w) in chow for 1 week prior to harvesting tissues for analyses. (E) LCA feeding led to accumulation of LCA in the proximal small intestine (SI), distal ileum (DI) and cecum of GF mice. (n=5 in each group, SI *p=0.04; DI *p=0.04; cecum **p=1.70x10−3, Welch’s t test). (F) LCA inhibited expression of Asbt in SI and DI (n=5 in each group, SI p=0.11; DI *p=0.02, Welch’s t test). (G) Introduction of LCA in GF mice induced CA7S production and accumulation in the gallbladder (GB) and the distal ileum (DI) (GB, GF n=3, GF+LCA n=4, **p=6.50x10−3; DI, n=5 in each group, *p=0.03, Welch’s t test). (H) LCA feeding led to increased expression of mSult2A1 and Vdr in livers of mice fed 0.3% LCA in chow (n=5 in each group, mSult2A1 *p=0.04, Vdr p=0.09, Welch’s t test). All data are presented as mean ± SEM.
Figure 7.. SG microbiota transfer recreates the…
Figure 7.. SG microbiota transfer recreates the CA7S pathway in GF mice
(A) Schematic of cecal microbial transplant (CMT). GF mice were fed a high-fat diet (HFD) for 5 days prior to CMT. Six weeks post-op sham and SG cecal stool were anaerobically homogenized and gavaged into GF animals. Two weeks post-gavage, mice were sacrificed, and their tissues were harvested for analyses. (B,C) qRT-PCR quantification of baiCD gene cluster expression levels in donor sham and SG mouse cecal contents (B) and recipient sham-CMT and SG-CMT mouse cecal contents (C) normalized to bacterial 16S ribosomal DNA (Sham, n=4, SG, n=4, **p=7.00x10−3; Sham-CMT, n=10, SG-CMT, n=10, *p=0.01, Welch’s t test). (D) LCA levels were reduced in cecal contents of SG-CMT compared to sham-CMT mice. DCA and total bile acid levels did not differ between the two groups (n=10 in each group; LCA, lithocholic acid *p=0.04; DCA, deoxycholic acid p=0.95; Total bile acids p=0.95, ns=not significant, Welch’s t test). (E) qRT-PCR quantification of BA transport protein expression levels in sham-CMT and SG-CMT mouse distal ileum (DI) normalized to mouse ribosomal 18S. The expression of Asbt (apical sodium-dependent bile acid transporter) and Ostα (organic solute transporter α) were significantly increased in SG-CMT (n=10 in each group, Asbt *p=0.02; Ostα *p=0.02, Welch’s t test). (F) Portal vein LCA in sham-CMT and SG-CMT mice. LCA was the only bile acid whose levels were significantly increased post-SG (n=10 in each group, LCA *p=0.03, Welch’s t test) (G) qRT-PCR quantification of mSult2A1 and Vdr expression levels in sham-CMT and SG-CMT mouse livers normalized to mouse ribosomal 18S. mSult2A1 and Vdr expression was significantly increased in SG-CMT (n=10 in each group, mSult2A1 *p=0.02; Vdr *p=0.02, Welch’s t test) (H) CA7S levels were increased in cecal contents of SG-CMT compared to sham-CMT mice. (n=10 in each group; CA7S, cholic acid-7-sulfate *p=0.04, Welch’s t test). (I) SG-CMT (cecal microbial transplant) mice have higher levels of GLP-1 compared to Sham-CMT (n=10 in each group, p=0.25, not significant, Welch’s t test). (J) qRT-PCR quantification of Tgr5 expression levels in sham-CMT and SG-CMT mouse distal ileum (DI) normalized to 18S. Tgr5 expression was significantly increased in SG-CMT (n=10 in each group, Tgr5 *p=0.03, Welch’s t test). All data are presented as mean ± SEM. (K) Model for LCA-mediated induction of CA7S synthesis and downstream GLP-1 secretion post-SG. Sleeve gastrectomy (1) results in a shift in the microbiome, specifically a decrease in Clostridia and LCA production, thus inducing an increase of the BA transporters ASBT and OSTα in the intestine (2), which in turn selectively transport LCA into the portal vein (3); LCA is delivered to the liver, where this molecule agonizes VDR, thereby increasing the expression of hSULT2A/mSult2A1 and CA7S synthesis (4). CA7S is then stored in the gallbladder and secreted into the gut. In the colon, CA7S agonizes and increases expression of TGR5 in intestinal L cells, resulting in secretion of GLP-1 (5), an incretin hormone that exerts global glucoregulatory effects.

References

    1. Abbasi J (2017). Unveiling the "Magic" of Diabetes Remission After Weight-Loss Surgery. JAMA 317, 571–574.
    1. Adhikari AA, Seegar TCM, Ficarro SB, McCurry MD, Ramachandran D, Yao L, Chaudhari SN, Ndousse-Fetter S, Banks AS, Marto JA, et al. (2020). Development of a covalent inhibitor of gut bacterial bile salt hydrolases. Nat Chem Biol 16, 318–326.
    1. Alnouti Y (2009). Bile Acid sulfation: a pathway of bile acid elimination and detoxification. Toxicol Sci 108, 225–246.
    1. Bachmanov AA, Reed DR, Beauchamp GK, and Tordoff MG (2002). Food intake, water intake, and drinking spout side preference of 28 mouse strains. Behav Genet 32, 435–443.
    1. Besnard P, Landrier JF, Grober J, and Niot I (2004). [Is the ileal bile acid-binding protein (I-BABP) gene involved in cholesterol homeostasis?]. Med Sci (Paris) 20, 73–77.
    1. Callahan BJ, McMurdie PJ, Rosen MJ, Han AW, Johnson AJ, and Holmes SP (2016). DADA2: High-resolution sample inference from Illumina amplicon data. Nat Methods 13, 581–583.
    1. Caporaso JG, Kuczynski J, Stombaugh J, Bittinger K, Bushman FD, Costello EK, Fierer N, Pena AG, Goodrich JK, Gordon JI, et al. (2010). QIIME allows analysis of high-throughput community sequencing data. Nat Methods 7, 335–336.
    1. Chaudhari SN, Harris DA, Aliakbarian H, Luo JN, Henke MT, Subramaniam R, Vernon AH, Tavakkoli A, Sheu EG, and Devlin AS (2020). Bariatric surgery reveals a gut-restricted TGR5 agonist with anti-diabetic effects. Nat Chem Biol.
    1. Craddock AL, Love MW, Daniel RW, Kirby LC, Walters HC, Wong MH, and Dawson PA (1998). Expression and transport properties of the human ileal and renal sodium-dependent bile acid transporter. Am J Physiol 274, G157–169.
    1. Cristina ML, Lehy T, Zeitoun P, and Dufougeray F (1978). Fine structural classification and comparative distribution of endocrine cells in normal human large intestine. Gastroenterology 75, 20–28.
    1. Damms-Machado A, Mitra S, Schollenberger AE, Kramer KM, Meile T, Konigsrainer A, Huson DH, and Bischoff SC (2015). Effects of surgical and dietary weight loss therapy for obesity on gut microbiota composition and nutrient absorption. Biomed Res Int 2015, 806248.
    1. Dawson PA, Lan T, and Rao A (2009). Bile acid transporters. J Lipid Res 50, 2340–2357.
    1. Dawson PA, and Setchell KDR (2017). Will the real bile acid sulfotransferase please stand up? Identification of Sult2a8 as a major hepatic bile acid sulfonating enzyme in mice. J Lipid Res 58, 1033–1035.
    1. Ding L, Sousa KM, Jin L, Dong B, Kim BW, Ramirez R, Xiao Z, Gu Y, Yang Q, Wang J, et al. (2016). Vertical sleeve gastrectomy activates GPBAR-1/TGR5 to sustain weight loss, improve fatty liver, and remit insulin resistance in mice. Hepatology 64, 760–773.
    1. Duboc H, Tache Y, and Hofmann AF (2014). The bile acid TGR5 membrane receptor: from basic research to clinical application. Dig Liver Dis 46, 302–312.
    1. Elam MB, Cowan GS Jr., Rooney RJ, Hiler ML, Yellaturu CR, Deng X, Howell GE, Park EA, Gerling IC, Patel D, et al. (2009). Hepatic gene expression in morbidly obese women: implications for disease susceptibility. Obesity (Silver Spring) 17, 1563–1573.
    1. Feng L, Yuen YL, Xu J, Liu X, Chan MY, Wang K, Fong WP, Cheung WT, and Lee SS (2017). Identification and characterization of a novel PPARalpha-regulated and 7alpha-hydroxyl bile acid-preferring cytosolic sulfotransferase mL-STL (Sult2a8). J Lipid Res 58, 1114–1131.
    1. Ferruzza S, Rossi C, Scarino ML, and Sambuy Y (2012). A protocol for differentiation of human intestinal Caco-2 cells in asymmetric serum-containing medium. Toxicol In Vitro 26, 1252–1255.
    1. Fiorucci S, and Distrutti E (2015). Bile Acid-Activated Receptors, Intestinal Microbiota, and the Treatment of Metabolic Disorders. Trends Mol Med 21, 702–714.
    1. Fukushima K, Okada A, Hayashi Y, Ichikawa H, Nishimura A, Shibata N, and Sugioka N (2015). Enhanced oral bioavailability of vancomycin in rats treated with long-term parenteral nutrition. Springerplus 4, 442.
    1. Funabashi M, Grove TL, Wang M, Varma Y, McFadden ME, Brown LC, Guo C, Higginbottom S, Almo SC, and Fischbach MA (2020). A metabolic pathway for bile acid dehydroxylation by the gut microbiome. Nature 582, 566–570.
    1. Ghosh A, Chen F, Banerjee S, Xu M, and Shneider BL (2014). c-Fos mediates repression of the apical sodium-dependent bile acid transporter by fibroblast growth factor-19 in mice. Am J Physiol Gastrointest Liver Physiol 306, G163–171.
    1. Gralka E, Luchinat C, Tenori L, Ernst B, Thurnheer M, and Schultes B (2015). Metabolomic fingerprint of severe obesity is dynamically affected by bariatric surgery in a procedure-dependent manner. Am J Clin Nutr 102, 1313–1322.
    1. Hall AB, Yassour M, Sauk J, Garner A, Jiang X, Arthur T, Lagoudas GK, Vatanen T, Fornelos N, Wilson R, et al. (2017). A novel Ruminococcus gnavus clade enriched in inflammatory bowel disease patients. Genome Med 9, 103.
    1. Han S, and Chiang JY (2009). Mechanism of vitamin D receptor inhibition of cholesterol 7alpha-hydroxylase gene transcription in human hepatocytes. Drug Metab Dispos 37, 469–478.
    1. Heshmati K, Harris DA, Aliakbarian H, Tavakkoli A, and Sheu EG (2019). Comparison of early type 2 diabetes improvement after gastric bypass and sleeve gastrectomy: medication cessation at discharge predicts 1-year outcomes. Surg Obes Relat Dis 15, 2025–2032.
    1. Jahansouz C, Staley C, Bernlohr DA, Sadowsky MJ, Khoruts A, and Ikramuddin S (2017). Sleeve gastrectomy drives persistent shifts in the gut microbiome. Surg Obes Relat Dis 13, 916–924.
    1. Jahansouz C, Staley C, Kizy S, Xu H, Hertzel AV, Coryell J, Singroy S, Hamilton M, DuRand M, Bernlohr DA, et al. (2018). Antibiotic-induced Disruption of Intestinal Microbiota Contributes to Failure of Vertical Sleeve Gastrectomy. Ann Surg.
    1. Kakizaki S, Takizawa D, Tojima H, Yamazaki Y, and Mori M (2009). Xenobiotic-sensing nuclear receptors CAR and PXR as drug targets in cholestatic liver disease. Curr Drug Targets 10, 1156–1163.
    1. Kaska L, Sledzinski T, Chomiczewska A, Dettlaff-Pokora A, and Swierczynski J (2016). Improved glucose metabolism following bariatric surgery is associated with increased circulating bile acid concentrations and remodeling of the gut microbiome. World J Gastroenterol 22, 8698–8719.
    1. Kastl AJ Jr., Terry NA, Wu GD, and Albenberg LG (2020). The Structure and Function of the Human Small Intestinal Microbiota: Current Understanding and Future Directions. Cell Mol Gastroenterol Hepatol 9, 33–45.
    1. Kohli R, Bradley D, Setchell KD, Eagon JC, Abumrad N, and Klein S (2013). Weight loss induced by Roux-en-Y gastric bypass but not laparoscopic adjustable gastric banding increases circulating bile acids. J Clin Endocrinol Metab 98, E708–712.
    1. Lamp KC, Freeman CD, Klutman NE, and Lacy MK (1999). Pharmacokinetics and pharmacodynamics of the nitroimidazole antimicrobials. Clin Pharmacokinet 36, 353–373.
    1. Larraufie P, Roberts GP, McGavigan AK, Kay RG, Li J, Leiter A, Melvin A, Biggs EK, Ravn P, Davy K, et al. (2019). Important Role of the GLP-1 Axis for Glucose Homeostasis after Bariatric Surgery. Cell Rep 26, 1399–1408 e1396.
    1. Lea T (2015). Caco-2 Cell Line. In The Impact of Food Bioactives on Health: in vitro and ex vivo models, Verhoeckx K, Cotter P, Lopez-Exposito I, Kleiveland C, Lea T, Mackie A, Requena T, Swiatecka D, and Wichers H, eds. (Cham (CH)), pp. 103–111.
    1. Lespessailles E, and Toumi H (2017). Vitamin D alteration associated with obesity and bariatric surgery. Exp Biol Med (Maywood) 242, 1086–1094.
    1. Li YC, Amling M, Pirro AE, Priemel M, Meuse J, Baron R, Delling G, and Demay MB (1998). Normalization of mineral ion homeostasis by dietary means prevents hyperparathyroidism, rickets, and osteomalacia, but not alopecia in vitamin D receptor-ablated mice. Endocrinology 139, 4391–4396.
    1. Liu H, Hu C, Zhang X, and Jia W (2018). Role of gut microbiota, bile acids and their cross-talk in the effects of bariatric surgery on obesity and type 2 diabetes. J Diabetes Investig 9, 13–20.
    1. Magouliotis DE, Tasiopoulou VS, Sioka E, Chatedaki C, and Zacharoulis D (2017). Impact of Bariatric Surgery on Metabolic and Gut Microbiota Profile: a Systematic Review and Meta-analysis. Obes Surg 27, 1345–1357.
    1. Manchanda PK, and Bid HK (2012). Vitamin D receptor and type 2 diabetes mellitus: Growing therapeutic opportunities. Indian J Hum Genet 18, 274–275.
    1. Marion S, Studer N, Desharnais L, Menin L, Escrig S, Meibom A, Hapfelmeier S, and Bernier-Latmani R (2019). In vitro and in vivo characterization of Clostridium scindens bile acid transformations. Gut Microbes 10, 481–503.
    1. Martinez-Augustin O, and Sanchez de Medina F (2008). Intestinal bile acid physiology and pathophysiology. World J Gastroenterol 14, 5630–5640.
    1. Martinez-Guryn K, Hubert N, Frazier K, Urlass S, Musch MW, Ojeda P, Pierre JF, Miyoshi J, Sontag TJ, Cham CM, et al. (2018). Small Intestine Microbiota Regulate Host Digestive and Absorptive Adaptive Responses to Dietary Lipids. Cell Host Microbe 23, 458–469 e455.
    1. McGavigan AK, Garibay D, Henseler ZM, Chen J, Bettaieb A, Haj FG, Ley RE, Chouinard ML, and Cummings BP (2017). TGR5 contributes to glucoregulatory improvements after vertical sleeve gastrectomy in mice. Gut 66, 226–234.
    1. Medina DA, Pedreros JP, Turiel D, Quezada N, Pimentel F, Escalona A, and Garrido D (2017). Distinct patterns in the gut microbiota after surgical or medical therapy in obese patients. PeerJ 5, e3443.
    1. Out C, Patankar JV, Doktorova M, Boesjes M, Bos T, de Boer S, Havinga R, Wolters H, Boverhof R, van Dijk TH, et al. (2015). Gut microbiota inhibit Asbt-dependent intestinal bile acid reabsorption via Gata4. J Hepatol 63, 697–704.
    1. Patti ME, Houten SM, Bianco AC, Bernier R, Larsen PR, Holst JJ, Badman MK, Maratos-Flier E, Mun EC, Pihlajamaki J, et al. (2009). Serum bile acids are higher in humans with prior gastric bypass: potential contribution to improved glucose and lipid metabolism. Obesity (Silver Spring) 17, 1671–1677.
    1. Pournaras DJ, Glicksman C, Vincent RP, Kuganolipava S, Alaghband-Zadeh J, Mahon D, Bekker JH, Ghatei MA, Bloom SR, Walters JR, et al. (2012). The role of bile after Roux-en-Y gastric bypass in promoting weight loss and improving glycaemic control. Endocrinology 153, 3613–3619.
    1. Ridlon JM, Kang DJ, and Hylemon PB (2006). Bile salt biotransformations by human intestinal bacteria. J Lipid Res 47, 241–259.
    1. Roda A, Cappelleri G, Aldini R, Roda E, and Barbara L (1982). Quantitative aspects of the interaction of bile acids with human serum albumin. J Lipid Res 23, 490–495.
    1. Runge-Morris M, Kocarek TA, and Falany CN (2013). Regulation of the cytosolic sulfotransferases by nuclear receptors. Drug Metab Rev 45, 15–33.
    1. Ryan KK, Tremaroli V, Clemmensen C, Kovatcheva-Datchary P, Myronovych A, Karns R, Wilson-Perez HE, Sandoval DA, Kohli R, Backhed F, et al. (2014). FXR is a molecular target for the effects of vertical sleeve gastrectomy. Nature 509, 183–188.
    1. Samczuk P, Ciborowski M, and Kretowski A (2018). Application of Metabolomics to Study Effects of Bariatric Surgery. J Diabetes Res 2018, 6270875.
    1. Schloss PD, Westcott SL, Ryabin T, Hall JR, Hartmann M, Hollister EB, Lesniewski RA, Oakley BB, Parks DH, Robinson CJ, et al. (2009). Introducing mothur: open-source, platform-independent, community-supported software for describing and comparing microbial communities. Appl Environ Microbiol 75, 7537–7541.
    1. Shang Q, Saumoy M, Holst JJ, Salen G, and Xu G (2010). Colesevelam improves insulin resistance in a diet-induced obesity (F-DIO) rat model by increasing the release of GLP-1. Am J Physiol Gastrointest Liver Physiol 298, G419–424.
    1. Sisley SR, Arble DM, Chambers AP, Gutierrez-Aguilar R, He Y, Xu Y, Gardner D, Moore DD, Seeley RJ, and Sandoval DA (2016). Hypothalamic Vitamin D Improves Glucose Homeostasis and Reduces Weight. Diabetes 65, 2732–2741.
    1. Sun AQ, Balasubramaniyan N, Chen H, Shahid M, and Suchy FJ (2006). Identification of functionally relevant residues of the rat ileal apical sodium-dependent bile acid cotransporter. J Biol Chem 281, 16410–16418.
    1. Tan HY, Trier S, Rahbek UL, Dufva M, Kutter JP, and Andresen TL (2018). A multi-chamber microfluidic intestinal barrier model using Caco-2 cells for drug transport studies. PLoS One 13, e0197101.
    1. Tremaroli V, Karlsson F, Werling M, Stahlman M, Kovatcheva-Datchary P, Olbers T, Fandriks L, le Roux CW, Nielsen J, and Backhed F (2015). Roux-en-Y Gastric Bypass and Vertical Banded Gastroplasty Induce Long-Term Changes on the Human Gut Microbiome Contributing to Fat Mass Regulation. Cell Metab 22, 228–238.
    1. Wahlstrom A, Kovatcheva-Datchary P, Stahlman M, Khan MT, Backhed F, and Marschall HU (2017). Induction of farnesoid X receptor signaling in germ-free mice colonized with a human microbiota. J Lipid Res 58, 412–419.
    1. Wahlstrom A, Sayin SI, Marschall HU, and Backhed F (2016). Intestinal Crosstalk between Bile Acids and Microbiota and Its Impact on Host Metabolism. Cell Metab 24, 41–50.
    1. Wang W, Cheng Z, Wang Y, Dai Y, Zhang X, and Hu S (2019). Role of Bile Acids in Bariatric Surgery. Front Physiol 10, 374.
    1. Weber N, Liou D, Dommer J, MacMenamin P, Quinones M, Misner I, Oler AJ, Wan J, Kim L, Coakley McCarthy M, et al. (2018). Nephele: a cloud platform for simplified, standardized and reproducible microbiome data analysis. Bioinformatics 34, 1411–1413.
    1. Wells JE, Williams KB, Whitehead TR, Heuman DM, and Hylemon PB (2003). Development and application of a polymerase chain reaction assay for the detection and enumeration of bile acid 7alpha-dehydroxylating bacteria in human feces. Clin Chim Acta 331, 127–134.
    1. Wu WK, Hsu CC, Sheen LY, and Wu MS (2019). Measurement of gut microbial metabolites in cardiometabolic health and translational research. Rapid Commun Mass Spectrom.
    1. Xie G, Wang X, Jiang R, Zhao A, Yan J, Zheng X, Huang F, Liu X, Panee J, Rajani C, et al. (2018). Dysregulated bile acid signaling contributes to the neurological impairment in murine models of acute and chronic liver failure. EBioMedicine 37, 294–306.
    1. Yalcin EB, More V, Neira KL, Lu ZJ, Cherrington NJ, Slitt AL, and King RS (2013). Downregulation of sulfotransferase expression and activity in diseased human livers. Drug Metab Dispos 41, 1642–1650.
    1. Yao L, Seaton SC, Ndousse-Fetter S, Adhikari AA, DiBenedetto N, Mina AI, Banks AS, Bry L, and Devlin AS (2018). A selective gut bacterial bile salt hydrolase alters host metabolism. Elife 7.
    1. Ye ZQ, Niu S, Yu Y, Yu H, Liu BH, Li RX, Xiao HS, Zeng R, Li YX, Wu JR, et al. (2010). Analyses of copy number variation of GK rat reveal new putative type 2 diabetes susceptibility loci. PLoS One 5, e14077.

Source: PubMed

Подписаться