Chapter 41
Bile Formation and the Enterohepatic Circulation Paul A. Dawson
41.1. INTRODUCTION Bile is a complex aqueous secretion that originates from hepatocytes and is modified distally by the biliary epithelium. As a basic “humor” in the body, the significance of bile had been recognized since antiquity. However, our understanding of bile was originally restricted to knowledge of its composition, and the mechanism of bile formation remained elusive until the mid-20th century with the advent of techniques to perfuse isolated livers and to study isolated hepatocytes.1 The concept that bile cycles between the liver and the gut, a “motus circularis bilis,” dates back more than 300 years to the work of Mauritius van Reverhorstand the elegant kinetic modeling by the Neapolitan mathematician, Giovanni Borelli.2 As the major biliary solute and driving force for bile flow, much attention has been focused on the mechanisms responsible for bile acid biosynthesis and enterohepatic cycling, and the relationship of those mechanisms to hepatic and gastrointestinal physiology.1,3
diversity in their chemical structures across the species, with modification to both the C19 steroid nucleus and the side chain. This large diversity is thought to be unique among classes of small molecule endobiotics, however, the evolutionary forces driving the variation remain poorly understood.8,9 In vivo, bile acids exist primarily as sulfate conjugates of bile alcohols and as taurine (or glycine) aminoacyl-amidated conjugates of bile acids.10 The general structure of the steroid nucleus and side chain, position of hydroxyl groups, and hydrophobicity for the major mammalian bile acid species are shown in Fig. 41.1.
41.3. MAJOR FUNCTIONS OF BILE AND BILE ACIDS
Bile acids are planar amphipathic molecules possessing a characteristic four-ring, 19-carbon perhydrocyclopentanophenanthrene nucleus and a multicarbon side chain. In all the vertebrates examined, cholesterol serves as the precursor for bile acid biosynthesis, whereby a waterinsoluble, hydrophobic membrane lipid is converted into water-soluble derivatives that can be excreted in bile.4 As a group, these molecules are termed “bile salts” or “bile acids,” and are included in the ST04 category (sterol lipids: Bile acids and derivatives) (http://www.lipidmaps.org) of the LIPID MAPS Lipid Classification System.5 Most bile acids can be assigned to three general structural classes according to the length of the side chain and functionality of the terminal polar group. The three classes are 27-carbon (C27) bile alcohols, C27 bile acids, and 24-carbon (C24) bile acids, with C24 bile acids being the predominant form in mammals.4,6 Bile acids are not known to be made by invertebrates.7 In vertebrates, bile acids show a remarkable
Bile formation and secretion is essential for life and fulfills a number of important functions in vertebrates.1 (1) Bile secretion is a major route for excretion of heavy metals, lipophilic endogenous compounds (endobiotics) such as bilirubin, cholesterol and steroids, and lipophilic exogenous compounds (xenobiotics) such as drugs, drug metabolites, and environmental toxins. (2) Bile is a critical digestive secretion and works in concert with saliva, gastric, and pancreatic secretions to facilitate the breakdown and assimilation of food.11 (3) Bile secretion plays a role in innate immunity and controlling intestinal microbes by serving as a conduit for the release of IgA antibodies.12 (4) Bile secretion is the vehicle for the excretion of bile acids, the major organic solutes in bile. As described in the next section, bile acids perform a variety of indispensable functions in the liver and gastrointestinal tract. Although best known for their ability to form micelles and facilitate absorption of lipids in the gut, the physiological functions of bile acids extend well beyond their role as simple detergents. The recognized functions ascribed to bile acids in the liver and gastrointestinal tract are summarized in Table 41.1.11,13 The major functions of bile acids include: (1) Inducing bile flow and hepatic secretion of biliary lipids (phospholipid
Physiology of the Gastrointestinal Tract. https://doi.org/10.1016/B978-0-12-809954-4.00041-4 © 2018 Elsevier Inc. All rights reserved.
931
41.2. STRUCTURE OF BILE ACIDS
932 SECTION | V Physiology of Secretion
FIG. 41.1 Structure and hydrophobicity/hydrophilicity profile of bile acids. (A) Structure of the common bile acids. In humans, cholic acid (CA) and chenodeoxycholic acid (CDCA) are primary bile acids synthesized by the hepatocyte. Primary bile acids in other species include muricholic acid (MCA) and ursodeoxycholic acid (UDCA). (B) Structure and hydrophobicity of bile acids. Hydroxyl groups that are oriented in the α-orientation are located below the steroid nucleus and are axial to the plane of the steroid nucleus. Hydroxyl groups that are in the β-orientation are located above the steroid nucleus and are equatorial to the plane of the steroid nucleus. The MCA escapes from this common rule as it contains a 6α-hydroxylated group that is equatorially positioned to the steroid nucleus plane. The equatorial location of hydroxyl groups confers polarity to the hydrophobic concave side of the steroid nucleus. Therefore, MCAs containing both 6α and 7β-oriented hydroxyl groups, and UDCA with its 7β-oriented hydroxyl group are more hydrophilic than other bile acids with the same number of hydroxyl groups axially positioned to the steroid nucleus. (Adapted with permission from Dawson PA. Bile acid metabolism. Biochemistry of lipids, lipoproteins and membranes. Amsterdam: Elsevier B.V.; 2015. p. 359.)
TABLE 41.1 Functions of Bile Acids in the Gastrointestinal Tract Tissue
Function
Whole body
• Elimination of cholesterol • Regulation of fat, glucose, and energy homeostasis by signaling through nuclear and G-protein-coupled receptors
Liver—hepatocyte
• Insertion of canalicular bile acid and phospholipid transporters • Induction of bile flow and biliary lipid secretion • Promotion of mitosis during hepatic regeneration • Regulation of gene expression via nuclear receptors (FXR, PXR, VDR)
Liver—endothelial cell
• Regulation of hepatic blood flow via activation of TGR5
Biliary tract— lumen
• Micellar solubilization of cholesterol • Micellar trapping of cholephilic xenobiotics • Antimicrobial actions • Calcium binding to prevent formation of calcium bilirubinate or salts of calcium phosphate, carbonate, or palmitate
Bile Formation and the Enterohepatic Circulation Chapter | 41 933
TABLE 41.1 Functions of Bile Acids in the Gastrointestinal Tract—cont’d Tissue
Function
Biliary tract— cholangiocytes
• Stimulation of bicarbonate secretion via CFTR and AE2
Gallbladder epithelium
• Modulation of cAMP-mediated secretion
Small intestine— lumen
• Micellar solubilization of dietary lipids, especially cholesterol, and fat-soluble vitamins
• Promotion of proliferation when bile duct is obstructed
• Promotion of mucin secretion
• Solubilization of lipophilic drugs and xenobiotics • Antimicrobial actions • Acceleration of protein hydrolysis by pancreatic proteases • Prevention of enteric hyperoxaluria
Small intestine— ileal enterocyte
• Regulation of gene expression via nuclear receptors (FXR, PXR, VDR) • Secretion of FGF19 to regulate hepatic bile acid synthesis • ASBT and CFTR-dependent induction of water secretion
Small intestine— other effects
• Secretion of antimicrobial factors by intestinal epithelium (FXR-mediated) • Activation of TGR5 on cholinergic neurons to inhibit intestinal contractility and delay small intestinal transit • Activation of TGR5 on enteroendocrine L-cells
Large intestine— colonic enterocyte
• Modulation of electrolyte absorption and secretion
Large intestine— other effects
• Alter intestinal motility
Gut microbiota
• Regulate microbial diversity and metabolism
• Activation of TGR5 on enteroendocrine L-cells
Adapted from Hofmann AF, Hagey LR. Cell Mol Life Sci 2008;65:2474.
and cholesterol). The active vectorial movement of bile acids from blood to the bile canaliculus creates an osmotic gradient, allowing water and small solutes to enter the biliary space by solvent drag. This is a major driving force for bile formation. (2) Digestion and absorption of dietary fats such as long-chain fatty acids, cholesterol, and fat-soluble vitamins. Bile acids form mix micelles with lipids and lipid digestion products to increase their aqueous solubility in the gut lumen, thereby enhancing their diffusion across the unstirred aqueous layer at the surface of the intestinal epithelium.14–16 Fat-soluble vitamins (A, D, E, K) are poorly absorbed from the intestinal lumen in the absence of bile acid micelles, and disturbances in the secretion or enterohepatic cycling of bile acids cause fat-soluble vitamin deficiency.17,18 Along with their major role in dietary lipid absorption, bile acids may facilitate intestinal absorption of dietary protein by promoting protein denaturation and accelerating hydrolysis by pancreatic proteases.19 (3) Bile acids play a complex role in maintaining cholesterol homeostasis. On one hand, bile acids increase cholesterol intake by promoting intestinal absorption of biliary and
dietary cholesterol. However, on the other hand, bile acids also promote cholesterol loss from the body. Bile acids are water-soluble end products of cholesterol catabolism and bile acid loss in the feces is quantitatively the second most important route for cholesterol elimination.20–22 Bile acids also promote hepatic secretion of cholesterol into bile by inducing bile flow and solubilizing biliary cholesterol, thereby enabling cholesterol to move from the liver to the intestinal lumen for elimination. (4) Bile acids contribute to the gut's antimicrobial defenses through direct bacteriostatic actions of bile acid-fatty acid mixed micelles and by signaling to induce expression of antimicrobial genes, thereby reducing small bowel bacterial translocation and intestinal inflammation.23–26 Bacterial overgrowth occurs in biliary fistula or bile duct-ligated animals, as well as in animals with experimental cirrhosis. In cirrhotic or cholestatic rats with bacterial overgrowth, feeding of conjugated bile acids or bile acid analogs ameliorates bacterial overgrowth, decreases bacterial translocation to intestinal lymph nodes, and decreases endotoxemia.23–26 (5) Bile acids regulate gut microbial diversity and vice versa under physiological and
934 SECTION | V Physiology of Secretion
pathophysiological conditions.27–30 (6) Bile acids act to prevent the formation of calcium gallstones and calcium oxalate kidney stones.31–33 Conjugated bile acids, which are fully water soluble as calcium salts, prevent the formation of gallstone-enucleating precipitates of calcium bilirubinate or salts of calcium phosphate, calcium carbonate, or calcium palmitate by binding calcium in the biliary tract and gallbladder.33,34 In the small intestinal lumen, dietary oxalate is usually precipitated by calcium. However, in patients with severe ileal resection and in obese patients after bariatric surgery, excess amounts of dietary fatty acids and bile acids passing into the colon act as a sink for calcium, greatly lowering its intraluminal activity.35–37 As a result, oxalate remains in solution and is absorbed from the colon, leading to hyperoxaluria and an increased risk of renal stone formation. In the presence of an intact enterohepatic circulation, bile acids are present at sufficient luminal concentrations in the small intestine to facilitate fat absorption, thereby reducing steatorrhea, colonic fatty acid concentrations, and oxalate absorption.38 (7) Bile acids act as hormones to signal through nuclear and G-protein-coupled receptors in order to regulate the bile acid enterohepatic circulation, hepatic function, gut motility, and fat, glucose, and energy homeostasis.39–42
41.3.1 Bile Acids as Signaling Molecules Beyond their roles as simple detergents to facilitate dietary lipid absorption and cholesterol homeostasis, bile acids also function as signaling molecules. Bile acids activate specific nuclear receptors (farnesoid X receptor alpha, FXR; pregnane X receptor, PXR; vitamin D receptor, VDR), G-proteincoupled receptors (Takeda G-protein-coupled receptor, TGR5; muscarinic receptors; sphingosine-1-phosphate receptor 2, S1PR2), integrins (α5,β1-integrin), and cellsignaling pathways (protein kinase C, PKC; c-jun N-terminal kinase 1/2, JNK 1/2; serine/threonine protein kinase, AKT/ PKB; extracellular signal-regulated kinase, ERK 1/2; p38 mitogen-activated protein kinase, p38 MAPK; epidermal growth factor receptor, EGFR),41–43 with FXR and TGR5 being the best understood examples. Evidence of their signaling properties began to emerge in the 1980s and 1990s, when bile acids were shown to activate protein kinase C isoforms and exhibit cell growthmodulatory effects.44–46 However, the role of bile acids as hormones/signaling molecules was not firmly established until 1999, when bile acids were identified as ligands for the orphan nuclear receptor FXR (gene symbol: NR1H4).47–49 The FXR (FXRα) should not be confused with the related nuclear receptor, FXRβ. FXRβ (gene symbol: NR1H5) is a widely expressed nuclear receptor that is activated by lanosterol, but not by bile acids. Although expressed by many vertebrates and mammalian species including mice, rats, rabbits, and dogs, FXRβ (NR1H5) is a nonexpressed
p seudogene in humans and primates.50,51 For FXR, many of the major mammalian bile acids (both unconjugated as well as glycine or taurine conjugated) function as ligands, with the following rank order of potency: chenodeoxycholic acid > lithocholic acid ≈ deoxycholic acid > cholic acid. Notable exceptions to the list of bile acid FXR agonists are ursodeoxycholic acid and 6-hydroxylated bile acids species such as muricholic acid, which do not activate FXR or function as FXR antagonists.52,53 There are four major isoforms of FXR in mice and humans, which are generated through the use of alternative promoters and alternative splicing.54 Although all four FXR isoforms encode identical ligand binding domains, the abundance of the isoforms vary between tissues and there are differences in their relative transcriptional activity.55,56 As noted above, vertebrates exhibit a remarkable diversity in their bile acid chemical structures, and the FXR ligand binding domain appears to have coevolved with its bile acid ligand.57,58 The FXR is mainly expressed in the liver intestine, kidney, and adrenal. Consistent with its gastrointestinal expression, FXR plays important roles in the regulation of enterohepatic cycling of bile acids, feedback regulation of bile acid biosynthesis, and protection against bile acid-associated toxicity.59–62 In the liver, these functions include stimulating bile acid conjugation and export across the canalicular membrane into bile. In the small intestine, activation of FXR protects the enterocyte from bile acid overload by inducing expression of the ileal cytosolic ileal bile acid binding protein (IBABP; gene symbol: FABP6), the basolateral bile acid transporter subunits, organic solute transporter (OST) alpha (OSTα), and OST beta (OSTβ),63,64 and the endocrine polypeptide hormone fibroblast growth factor (FGF) 19 (mouse ortholog: FGF15), a central regulator of hepatic bile acid synthesis.65,66 With regard to general functions in the gastrointestinal tract, FXR induces expression of genes important for intestinal barrier function and antimicrobial defense,24 and has important antiproliferative and antiinflammatory properties.67–69 Bile acids also signal via the nuclear receptors PXR (gene symbol: NR1I2)70,71 and VDR (gene symbol: NR1I1).72,73 These receptors are activated by lithocholic acid, a hydrophobic and potentially cytotoxic secondary bile acid produced from chenodeoxycholic acid by intestinal anaerobic bacteria. With regard to ligand specificity, bile acids activate PXR with a rank order of potency: lithocholic acid > deoxycholic acid > cholic acid, and activate VDR with a rank order of potency: 3-oxo-lithocholic acid > deoxycholic acid > cholic acid. With regard to bile acid homeostasis, PXR or VDR primarily function to induce expression of enzymes involved in bile acid metabolism and detoxification,71,72,74 and likely play only a minor role in regulating bile acid biosynthesis.75,76 In 2002, two groups independently identified TGR5 (also called membrane-type bile acid receptor, M-BAR; G-protein-coupled bile acid receptor 1, GPBAR1; gene
Bile Formation and the Enterohepatic Circulation Chapter | 41 935
symbol: GPBAR1) as a Gαs-coupled bile acid receptor, which stimulates adenylate cyclase and increases intracellular cAMP levels.77,78 TGR5 is activated by conjugated and unconjugated bile acids, with the following rank order of potency: deoxycholic acid > lithocholic acid > chenodeoxycholic acid> cholic acid.78 Notably, there are bile acid ligand specificity differences between FXR and TGR5, which have been exploited to generate bile acid derivatives that selectively activate the individual receptors or serve as agonists for both.79–81 TGR5 is ubiquitously expressed, with highest levels of expression in gallbladder and moderate levels of expression in liver and intestine. In the liver, TGR5 is not expressed by hepatocytes. However TGR5 is expressed by Kupffer cells, where it is thought to play an immunomodulatory role, and by sinusoidal endothelial cells, where TGR5 functions to induce nitric oxide synthesis and regulate the hepatic microcirculation.82 With the growing appreciation of bile acids as signaling molecules, considerable study is being directed toward understanding the physiological functions of TGR5.83 For example, bile acid activation of TGR5 can regulate gallbladder filling, intestinal motility, and may have a role in bile acid-induced itch and the analgesia associated with cholestatic liver disease.84–87 There are also metabolic effects associated with TGR5 signaling in brown adipose, muscle, and macrophages.41,88–90
41.4. BIOSYNTHESIS AND BIOTRANSFORMATION OF BILE ACIDS 41.4.1 Biosynthesis of Bile Acids Bile acids are synthesized from cholesterol in the pericentral hepatocytes of the hepatic acini.91 In this process, the hydrophobic substrate, cholesterol, is converted to a watersoluble, amphipathic product through a series of sterol ring hydroxylations and side chain oxidation steps. Bile acids synthesized by the hepatocyte are designated primary bile acids to distinguish them from the secondary bile acids that are formed by the reactions carried out by the host or gut microbiota, which include dehydroxylation, dehydrogenation (oxidation of a hydroxy group to an oxo group), oxidation, epimerization (changing an α-hydroxy group to a β-hydroxy group or vice versa), and esterification.3,92 Bile acid synthesis was originally thought to involve one major pathway, the “classical” or neutral pathway (cholesterol 7αhydroxylase pathway) that favors cholic acid biosynthesis.93 This paradigm was later modified by the discovery of a second pathway, the “alternative” or acidic pathway (oxysterol 7α-hydroxylase pathway) that favors the biosynthesis of chenodeoxycholic acid in humans and 6-hydroxylated bile acids such as muricholic acid and hyocholic acid in mice and rats.94,95 Details of the hepatocellular and biochemical mechanisms responsible for the metabolic channeling of cholesterol toward cholic acid versus chenodeoxycholic
acid/6-hydroxylated bile acids are still not clear, and the specificity is not absolute. The cholesterol 7α-hydroxylase pathway produces some chenodeoxycholic acid/6-hydroxylated bile acid whereas the alternative pathway can yield cholic acid.96–98 In the alternative pathway, the first step involves modification of the cholesterol side chain by C-24 (sterol 24-hydroxylase; gene symbol: CYP46A1), C-25 (sterol 25hydroxylase; gene symbol: CH25H), or C-27 (sterol 27-hydroxylase; gene symbol: CYP27A1) sterol hydroxylases present in liver and extra-hepatic tissues such as brain. This reaction is then followed by an oxysterol 7αhydroxylation, which is mediated primarily by CYP7B1 in liver.39,94,99–102 Of these alternative hydroxylation pathways, the contribution of 27-hydroxycholesterol to bile acid synthesis is quantitatively most important.93,94 Although quantitatively a minor contributor to bile acid synthesis, the conversion of cholesterol to 24S-hydroxycholesterol functions as a major mechanism for cholesterol elimination from brain by facilitating sterol transfer across the bloodbrain barrier into the systemic circulation for excretion by the liver.103–105 The overall process of bile acid biosynthesis is complex, requiring the action of 16 enzymes that catalyze as many as 17 different reactions.102 In the classical pathway, the steroid nucleus is modified before the side chain, whereas in the alternative pathways, side chain modifications occur before or coincident with changes to the steroid nucleus. Cholesterol 7α-hydroxylase (gene symbol: CYP7A1) is the rate-limiting enzyme for bile acid synthesis via the classical pathway. However, the step catalyzed by the sterol 12αhydroxylase (gene symbol, CYP8B1) controls the amount of cholic acid synthesized and is an important determinant of the ratio of cholic acid to chenodeoxycholic acid and cholic acid to muricholic acid in human and mouse bile, respectively.106 In this capacity, CYP8B1 plays a critical role in modulating the composition and hydrophobicity of the bile acid pool. It should be noted that humans and mice have substantially different bile acid pool compositions. This reflects differences in bile acid conjugation (discussed below), synthesis of ursodeoxycholic acid as a primary bile acid in mice, and hydroxylation at the 6-position of the bile acid steroid nucleus in mice.107 In contrast, hydroxylation of bile acids at the 6-position is rare in humans, and detectable under only specialized circumstances, such as in fetal/early neonatal development or in certain cholestatic conditions.108–110 It has long been known that 6hydroxylation of bile acids alters their physicochemical and detergent properties, with mice having a more hydrophilic bile acid pool.111,112 With the recognition of bile acids as signaling molecules that act through nuclear and G-protein-coupled receptors, the human-rodent bile acid structural differences have gained additional importance. The 6-hydroxylation of bile acids dramatically alters their
936 SECTION | V Physiology of Secretion
signaling properties, potentially limiting the human relevance of mouse models for the study of bile acid-related disease.113 The cytochrome P450 enzyme(s) (CYPs) responsible for 6-hydroxylation of bile acids were very recently identified.114 Correlative data had implicated members of the murine CYP3A family, particularly CYP3A11. However, analysis of knockout mice lacking CYP3A enzymes or all members of the CYP1A, CYP2C, CYP2D, or CYP3A families found that members of the murine CYP2C family, including at least CYP2C70, are required and directly mediate the first step in
6-hydroxylation of chenodeoxycholic acid to alpha- muricholic acid and ursodeoxycholic acid to beta-muricholic acid.114,115 In contrast, the major human CYP2C enzyme, CYP2C9, was unable to oxidize bile acids, in agreement with the finding that 6-hydroxylated bile acid species are absent in humans under physiological conditions.114 The major bile acid biosynthetic pathways are summarized in Fig. 41.2. After their biosynthesis, bile acids are conjugated via a two-step process involving the generation of a bile acid-CoA by bile acid-CoA synthase and then amidation with taurine
FIG. 41.2 Bile acid synthesis pathways. Primary bile acids are synthesized by the hepatocyte. The major classical (neutral) pathway for bile acid synthesis begins with cholesterol 7α-hydroxylase (CYP7A1). Bile acid intermediates synthesized via this pathway are substrates for the sterol 12α-hydroxylase (CYP8B1), the rate-determining step in the production of cholic acid. In the minor alternative (acidic) pathway for bile acid synthesis, cholesterol is first hydroxylated on its side chain by sterol 27-hydroxylase (CYP27A1), sterol 25-hydroxylase, or sterol 24-hydroxylase (CYP46A1). Subsequent hydroxylation of the steroid nucleus is catalyzed by oxysterol 7α-hydroxylase (CYP7B1) or to a lesser extent by the distinct oxysterol 7α-hydroxylase, CYP39A1. The classical and alternative pathways converge at the enzymatic steps for the reduction and dehydrogenation of the steroid ring. The alternative pathway preferentially produces chenodeoxycholic acid. In mice, ursodeoxycholic acid is also synthesized as a primary bile acid in the liver. The cytochrome P450 (CYP2C70) then converts chenodeoxycholic acid and ursodeoxycholic acid to alpha-muricholic acid and beta-muricholic acid, respectively. After side chain oxidation and cleavage, bile acids are aminoacyl amidated to taurine or glycine. (Adapted with permission from Chiang JY. Bile acid metabolism and signaling. Compr Physiol 2013;3(3):1191–1212.)
Bile Formation and the Enterohepatic Circulation Chapter | 41 937
or glycine by bile acid-CoA: amino acid N-acyltransferase (BAAT).116 In most nonmammalian vertebrate species, bile acids are typically modified on their side by sulfation (for C27 bile alcohols) or conjugation with taurine or a taurine derivative (for C27 and C24 bile acids).4 In mammals, bile acids are primarily conjugated on their side chain to either taurine or glycine.4,6,117 Notably, the conjugation pattern varies considerably between different mammalian species, ranging from almost exclusively taurine in the rat, cat, mouse, sheep, and dog, to mostly glycine in the pig, hamster, guinea pig, and human, to exclusively glycine in the rabbit.4,6 The amino acid specificity for the conjugation of bile acids is controlled by the BAAT enzyme,118,119 and to a lesser degree, by the availability of the taurine precursor. However, the evolutionary forces driving the selection of a particular amino acid in different animal species are unclear. Taurine conjugated bile acids have a lower pKa than their respective glycine conjugates and are more likely to remain ionized and membrane impermeant. However, both the glycine and taurine amide linkages are more resistant to hydrolysis by the pancreatic carboxypeptidases, as compared to other amino acids. As such, both taurine and glycine-conjugated bile acids largely escape cleavage by host proteases in the intestinal lumen during the digestive process.120 Of the two major biosynthetic pathways, the classical (CYP7A1) pathway is quantitatively more important in rodents and humans.93 In mice, the classical pathway accounts for ~70% of the total bile acid synthesis in adults, and is the predominant pathway in neonates.121–123 In humans, the classical pathway accounts for more than 90% of bile acid synthesis, as evidenced by approximately 96% reduction in fecal bile acid excretion in an adult patient with an inherited CYP7A1 defect.96 In contrast to mice and adult humans,124 the alternative pathway is the predominant biosynthetic pathway in human neonates, as evidenced by low to undetectable CYP7A1 expression in newborns and the finding of severe cholestatic liver disease in infants with inherited oxysterol 7α-hydroxylase (CYP7B1) gene defects.125–127
41.4.2 Regulation of Bile Acid Biosynthesis Bile acid biosynthesis is regulated by bile acids, hormones, cytokines, growth factors, oxysterols, xenobiotics, and diurnal rhythm,39,102,128 reflecting the need to tightly control the body's bile acid load. It was recognized for many years that feedback inhibition of the rate-limiting enzyme CYP7A1 plays a major role in controlling bile acid biosynthesis.129,130 The major mechanism responsible for the regulation of CYP7A1 expression and bile acid synthesis have been elucidated over the past decade and are summarized in Fig. 41.3.
The major pathway for feedback regulation of bile acid synthesis involves FXR131–133 and gut-liver signaling via the endocrine polypeptide hormone FGF 19 (mouse ortholog: FGF15).65,66 In this pathway, bile acids activate FXR in ileal enterocytes to induce synthesis of FGF15/19. After its release by the enterocyte, FGF15/19 travels in the portal circulation to the hepatocyte where it signals via its cell surface receptor, a complex of the βKlotho protein and fibroblast growth factor receptor-4 (FGFR4), to repress CYP7A1 expression and bile acid synthesis.65,66,134 The dominant role of this pathway as the major physiological mechanism responsible for feedback repression of CYP7A1 expression is strongly supported by results obtained using knockout mouse models, including FGFR4, β-Klotho, FGF15, and tissue-specific FXR-null mice.62,65,135–137 Moreover, identification of this pathway helped to explain a series of puzzling experimental findings, which included the observation that intravenous infusion of bile acids into the bile-fistula rat was ineffective at downregulating hepatic bile acid synthesis as compared to intraduodenal infusion of bile acids138 and that bile acids were relatively weak inhibitors of bile acid synthesis when added directly to isolated hepatocytes in culture.139,140 This regulatory pathway appears to be conserved in humans and nonhuman primates, since circulating FGF19 levels inversely correlate with markers of hepatic bile acid biosynthesis, administration of inhibitory anti-FGFR4 antibodies stimulated bile acid synthesis in nonhuman primates, and administration of recombinant FGF19 to human subjects strongly repressed bile acid synthesis.66,141,142 After binding FGF15/19, FGFR4/β-Klotho signals through the docking protein fibroblast growth factor substrate 2 (FRS2α) and tyrosine-protein phosphatase nonreceptor type 11 (Shp2; gene symbol: PTPN11). Activation of Shp2 stimulates extracellular-signal-regulated kinase (ERK1/2) activity and blocks the activation of CYP7A1 gene expression by hepatic nuclear factor 4-alpha (HNF4α) and liver receptor homolog-1 (LRH1).143–147 Bile acids also regulate the expression of CYP8B1 by similar, but not identical pathways.62 The regulation of both CYP7A1 and CYP8B1 by bile acids and FGF15/19 appears to involve the orphan nuclear receptor, small heterodimer partner (SHP; gene symbol: NR0B2). For example: (1) SHP can antagonize LRH-1 or HNF4α-mediated activation of CYP7A1 and CYP8B1 expression, and (2) FGF15/19-mediated regulation of CYP7A1 and CYP8B1 is blunted in SHP null mice.62,65,137,148 Finally, it was very recently shown that FXR indirectly represses expression of other genes involved in bile acid biosynthesis (but not CYP7A1) by inducing expression of the transcriptional repressor v-Maf avian musculoaponeurotic fibrosarcoma oncogene homolog G (MAFG), which interacts directly with the promoters of those genes.61 This complex network of receptors and
938 SECTION | V Physiology of Secretion
FIG. 41.3 Mechanisms responsible for feedback negative regulation of hepatic bile acid synthesis. In the major physiological pathway, intestinal bile acids are taken up by the ASBT, and activate FXR to induce FGF15/19 expression in ileal enterocytes. The basolateral secretion of FGF15/19 protein may be facilitated by the endosomal membrane glycoprotein, Diet1. FGF15/19 is then carried in the portal circulation to the liver where it binds to its cell surface receptor, a complex of the receptor tyrosine kinase, FGFR4 and the associated protein β-Klotho. FGFR4/β-Klotho then signals through the docking protein FRS2α and the tyrosine phosphatase Shp2 to stimulate ERK1/2 phosphorylation and block activation of CYP7A1 gene expression by the nuclear factors HNF4α and LRH1. In a direct pathway that may be more significant under pathophysiological conditions, bile acids can activate FXR in hepatocytes to induce expression of SHP, an atypical orphan nuclear receptor. The SHP interacts with LRH-1 and HNF4α to block activation of CYP7A1. (Adapted with permission from Dawson PA. Bile acid metabolism. Biochemistry of lipids, lipoproteins and membranes. Amsterdam: Elsevier B.V.; 2015. p. 359.)
regulatory factors links the control of bile acid synthesis and bile acid composition to changes in ileal as well as hepatic bile acid levels. The receptors and protein factors that participate in the negative feedback regulation of bile acid synthesis are summarized in Table 41.2. With regard to the alternative bile acid biosynthetic pathway, bile acids can signal via FXR-MAFG to repress expression of important biosynthetic genes such as CYP27A1 and CYP7B1.61 However, the major mechanism for control of the alternative pathway appears to be posttranscriptional and involve the regulation of cholesterol delivery to the mitochondrial inner membrane, the site of cholesterol 27-hydroxylation.149,150 The process is thought to be similar to that for steroid hormone biosynthesis in adrenal gland, where steroidogenic acute regulatory protein D1 (StARD1)mediated transfer of cholesterol to the mitochondria inner
membrane is rate-limiting step.151 The liver expresses several different StAR-related lipid transfer (START) domaincontaining proteins in addition to StARD1,152 and StARD5 has been shown to directly bind primary bile acids.153 However, details of the molecular mechanisms underlying cholesterol trafficking, intramitochondrial cholesterol delivery, and their regulatory consequences for bile acid biosynthesis in hepatocytes still remains poorly understood.154
41.4.3 Biotransformation of Bile Acids During Enterohepatic Cycling Under physiological conditions, essentially all bile acids secreted into bile are conjugated. Conjugation with taurine or glycine increases the hydrophilicity of bile acids and the acidic strength of the side chain, in essence converting a
Bile Formation and the Enterohepatic Circulation Chapter | 41 939
TABLE 41.2 Regulation of Bile Acid Synthesis and Enterohepatic Cycling Protein (Gene)
Tissue
Description and Function in Bile Acid Metabolism
FXR (NR1H4)
Intestine, liver, kidney
Bile acid-activated nuclear receptor; regulation of bile acid synthesis, transport, and metabolism
SHP (NR0B2)
Liver, intestine
Nuclear receptor; negative feedback regulation of hepatic bile acid synthesis by antagonizing HNF4a, LRH1; regulation of bile acid transport and metabolism
HNF4a (NR2a1)
Liver, intestine
Nuclear receptor; positive regulator of CYP7A1 expression and hepatic bile acid synthesis
LRH1 (NR5A2)
Liver, intestine
Nuclear receptor; positive regulator of CYP7A1 expression and hepatic bile acid synthesis
PXR (NR1I2)
Liver, intestine
Bile acid and xenobiotic-activated nuclear receptor involved in detoxification of secondary bile acids
VDR (NR1I1)
Intestine
Vitamin D and bile acid-activated nuclear receptor; involved in detoxification of LCA
FGFR4 (FGFR4)
Ubiquitous
Membrane receptor; negative feedback regulation of CYP7A1 and hepatic bile acid synthesis
β-klotho (KLB)
Liver
Membrane co-receptor associated with FGFR4; confers liver specificity to FGFR4-FGF19 pathway; negative feedback regulation of Cyp7a1 and hepatic bile acid synthesis
FGF19 (FGF19)
Intestine
Protein growth factor; secreted by intestine in response to bile acids; regulates hepatic bile acid synthesis via FGFR4:b-klotho
TGR5
Ubiquitous
Bile acid-activated G-protein coupled receptor; regulates intestinal motility, metabolism
MafG
Ubiquitous
Transcription factor; negative regulation of hepatic bile acid synthesis and bile acid transport
FXR, farnesoid X-receptor; FGF19, fibroblast growth factor 19; FGFR4, fibroblast growth factor receptor 4; HNF4a, hepatocyte nuclear factor 4alpha; PXR, pregnane X-receptor; SHP, small heterodimer partner; VDR, vitamin D receptor.
weak acid (pKa > 5.0) to a strong acid (pKa ~3.9 for the glycine conjugate; pKa <2.0 for the taurine conjugate).155–157 As a result, conjugated bile acids are almost completely ionized under the pH conditions present in the lumen of the biliary tract and small intestine. The physiological consequence of conjugation is to decrease the passive diffusion of bile acids across cell membranes during their transit through the biliary tree and small intestine.158 Conjugated bile acids are also more soluble at acidic pH and more resistant to precipitation in the presence of high concentrations of calcium than unconjugated species.32 The net effect of conjugation is to maintain high intralumenal concentrations of bile acids to solubilize cholesterol and fat-soluble vitamins, and to facilitate lipid digestion and absorption down the length of the small intestine. The importance of bile acid conjugation is underscored by the finding that patients with inherited bile acid conjugation defects present with malabsorption of dietary triglyceride and fat-soluble vitamins.17,18,159,160 Most of the conjugated bile acids secreted into the small intestine are efficiently absorbed intact. However, a fraction of the bile acids undergoes deconjugation (cleavage of the amide bond linking the glycine or taurine to the bile acid side chain) and biotransformation by the gut microbiota.3,92 The bacterial modifications of primary bile acids are important for several reasons. First, these modifications increase the hydrophobicity and decrease the aqueous solubility of bile acids, resulting in a marked lowering of the monomeric concentration of bile acids in aqueous solution.156,161 This
in turn reduces the flux of bile acids across the ileal or colonic epithelium and increases bile acid loss in the feces. Second, the composition of the circulating pool of bile acids is influenced by the input of secondary bile acids from the ileum and colon. Notably, these secondary bile acids have detergent properties, signaling activities, and toxicities that are distinct from their primary bile acid precursors.13,162,163 Third, the bile acid composition affects the diversity and composition of the gut microbiome, which can have pleiotropic metabolic and physiologic effects.27 In the small intestine, unconjugated bile acids are passively or actively absorbed and returned to the liver. After hepatocellular uptake, the unconjugated bile acids are efficiently reconjugated to taurine or glycine and resecreted into bile, a process termed “bile acid repair” by Hofmann.10 In the colon, gut microbial deconjugation of bile acids may proceed to near completion prior to being excreted in the feces.164 Bile acid deconjugation (removal of the glycine of taurine N-acyl-amidation) by bacterial bile salt hydrolases is a “gateway reaction” and precedes subsequent biotransformation by the gut microbiota.165,166 In particular, bile acids must be deconjugated prior to the removal of the C-7 hydroxy group because the first step in that process is the formation of a coenzyme A derivative, which requires a weak acid such as the unconjugated bile acid side chain.167 As such, deconjugationresistant bile acid analogs such as cholylsarcosine are also resistant to dehydroxylation.168 Bacterial 7α-dehydroxylation converts cholic acid to deoxycholic acid, a dihydroxy bile
940 SECTION | V Physiology of Secretion
acid with hydroxyl groups at the C-3 and C-12 positions, and chenodeoxycholic acid to lithocholic acid, a monohydroxy bile acid with a hydroxyl group at the C-3 position167 (see Fig. 41.4). In rats and mice, the 3,6,7-trihydroxy bile acids (α, β, ω-muricholic acid; hyocholic acid) are converted to the 3,6-dihydroxy bile acids, hyodeoxycholic acid (3α,6αdihydroxy-5β-cholanoic acid), and murideoxycholic acid (3α,6β-dihydroxy-5β-cholanoic acid).169 Besides undergoing 7-dehydroxylation during colonic transit, the hydroxy groups on the steroid nucleus may be modified by dehydrogenation, epimerization, or even elimination to form an unsaturated bile acid (with a double bond in the steroid nucleus). One of the more common bacterial modifications is epimerization of the 3α-hydroxy or 7α-hydroxy groups to their corresponding 3βor 7β-hydroxy forms.164,166,170 For example, the 7α-hydroxy group of chenodeoxycholic acid is epimerized to form the 3α,7β-dihydroxy bile acid, ursodeoxycholic acid, and lithocholic acid and deoxycholic acid are epimerized to their 3β-hydroxy-epimers, isolithocholic acid and isodeoxycholic acid.164,170 Note that in addition to these reactions, other gut microbiota-mediated modifications of bile acids have been detected such as fatty acyl esterification and polymerization. However, the quantitative significance of these reactions and the biochemical pathways responsible for their synthesis remain largely unexplored.3 A fraction of the 7-deoxy bile acids are absorbed from the colon and returned to the liver, where they are efficiently reconjugated with glycine or taurine and potentially 7αrehydroxylated. Hepatic bile acid 7α-rehydroxylation activity varies considerably between species and the enzyme(s) responsible has not been identified.171,172 The interspecies variation in hepatic bile acid 7α-rehydroxylation is reflected in the biliary deoxycholate concentration, which is low (from 0% to 10% of bile acids) in species that actively 7α-rehydroxylate deoxycholate such as rats, mice, guinea pigs, prairie dogs, and hamsters,173–176 and higher in species that cannot rehydroxylate deoxycholate, ranging from 15% to 30% in dogs and humans177,178 to greater than 90% in rabbits.179 Finally, in addition to reconjugation with taurine/ glycine, hepatocytes can epimerize iso(3β-hydroxy) bile acids to their 3α-hydroxy form, reduce the oxo groups on bile acids to hydroxyl groups, and modify bile acids by sulfation (sulfonation) or to a lesser extent, by glucuronidation.180–182 The secondary metabolism of bile acids in humans and mice is summarized in Fig. 41.4.
41.5. ENTEROHEPATIC CIRCULATION OF BILE ACIDS As mentioned in the introduction, the concept of an enterohepatic circulation dates back more than 300 years.10 Anatomically, the gut-liver circulation can be subdivided into a portal and an extra-portal pathway. The extra-portal pathway consists primarily of the lymphatic drainage from
the intestine into the superior vena cava. Although this process is important for chylomicron particle-mediated transport of cholesterol, triglycerides, fat-soluble vitamins, and phospholipids, it plays little role in bile acid absorption. Bile acids undergo a portal enterohepatic circulation, whereby they are: (1) secreted into bile by the liver, (2) pass into the duodenum, (3) absorbed from the intestinal lumen at the distal ileum, (4) pass into the portal circulation, and (5) efficiently extracted by the liver for resecretion into bile. The enterohepatic circulation of bile acids is an extremely efficient process; less than 5% of the intestinal bile acids escape reabsorption and are eliminated in the feces.183 Thus, most of the bile acids secreted by the hepatocyte were previously secreted into the small intestine and returned to the liver in the portal circulation. During fasting, about half the bile acid pool is sequestered and concentrated approximately 10-fold in the gallbladder, resulting in lower levels of bile acids in the small intestine, portal vein, liver, and serum. However, basal rates of hepatic bile acid secretion are still present and there is continuous enterohepatic cycling of that portion of the bile acid pool that is not sequestered in the gallbladder.184,185 The gallbladder empties its contents in response to a meal and causes the release of cholecystokinin, and the newly secreted plus stored bile acids pass directly into the duodenum. In the digestive phase, the bile acid concentration in the small intestine is approximately 5–10 mM. During the inter-digestive phase, the sphincter of Oddi contracts and the gallbladder relaxes, causing a larger fraction of the secreted bile acids to enter the gallbladder for storage. Interestingly, bile acids have a direct role in promoting gallbladder filling by signaling directly via TGR5 or by stimulating ileal synthesis and release of FGF15/19, a polypeptide hormone that induces gallbladder relaxation.85,186 Thus, the enterohepatic cycling of bile acids increases during digestion and slows between meals and during overnight fasting. This rhythm is maintained even after cholecystectomy, where the fraction of the bile acids stored in proximal intestine is increased but bile acid metabolism and enterohepatic cycling is largely intact.187–189 A fraction (10%–50%, depending on the bile acid species) of the bile acids returning in the portal circulation escapes first pass hepatic extraction and spills into the systemic circulation. Bile acid binding to plasma proteins such as albumin reduces their glomerular filtration and minimizes their urinary excretion. In healthy humans, the kidney filters approximately 100 μmol of bile acids each day, but only 1–2 μmol of bile acid are excreted in urine because of a highly efficient tubular reabsorption.190 Even in patients with cholestatic liver disease, in whom plasma bile acid concentrations are greatly elevated, the 24-h urinary excretion of nonsulfated bile acids is significantly less than the quantity that undergoes glomerular filtration.191 Subsequent studies showed that bile acids in the glomerular filtrate are actively reabsorbed from the renal tubules by a
Bile Formation and the Enterohepatic Circulation Chapter | 41 941
FIG. 41.4 Biotransformation of bile acids by the gut microbiome. Gut bacterial metabolism includes deconjugation, 7-dehydroxylation, oxidation, and epimerization of bile acids. Major reactions are indicated in bold. Reversible reactions are indicated by the double arrows. Oxidation yields the respective oxo-bile acid species, which can be epimerized, converting the 7α-hydroxy group to 7β-hydroxyepimers (ursocholic acid, ursodeoxycholic acid), the 6β-hydroxy group to 6α-hydroxyepimers (hyodeoxycholic acid), and the 3α-hydroxy group to 3β-hydroxyepimers (isocholic acid, isodeoxycholic acid, isochenodeoxycholic acid, and isolithocholic acid). In human cecum, a significant fraction of the bile acids present are converted to their respective 3βepimers (iso-bile acids). In rats and mice, muricholic acid species (α, β, ω-muricholic acid) are converted to the 3,6-dihydroxy bile acids, murideoxycholic acid (3α,6β-dihydroxy-5β-cholanoic acid) and hyodeoxycholic acid (3α,6α-dihydroxy-5β-cholanoic acid). (Adapted with permission from Dawson PA. Bile acid metabolism. Biochemistry of lipids, lipoproteins and membranes. Amsterdam: Elsevier B.V.; 2015. p. 359.)
sodium-dependent mechanism.192,193 As in the ileum, the renal proximal tubule epithelium expresses the apical sodiumdependent bile acid transporter (ASBT; also called ileal bile acid transporter, IBAT; gene symbol: SLC10A2) as a salvage mechanism to conserve bile acids.194,195 In addition to expressing the ASBT on the apical surface, renal epithelial cells express OSTα-OSTβ196 on the basolateral membrane, thereby completing the route for bile acids to be taken up from the tubule lumen and exported into the systemic circulation.
41.6. BILE SECRETION AND HEPATIC BILE ACID TRANSPORT 41.6.1 Overview of Bile Secretion The formation of canalicular bile is an osmotic process driven by active secretion of organic solutes into the canalicular lumen, followed by passive inflow of water, electrolytes, and other solutes.1,197 Canalicular bile flow is traditionally
942 SECTION | V Physiology of Secretion
divided into two components: bile acid-dependent and bile acid-independent flow.198 Solutes such as conjugated bile acids that are actively pumped across the canalicular membrane generate bile flow and are termed primary solutes. Other primary solutes include conjugated bilirubin, glutathione, bicarbonate, and the glucuronide or sulfate conjugates of endobiotics and xenobiotics. Water, plasma electrolytes, calcium, glucose, amino acids, and other low-molecularweight solutes that flow passively into the canaliculus in response to the osmotic gradient are termed secondary solutes. The choleretic activity of each primary solute is defined as the volume of bile flow induced per amount of solute secreted. The apparent choleretic activity for different conjugated bile acid species ranges from 8 to 25 μL of bile flow induced per μmol of bile acid secreted.199 This activity is also influenced by the osmotic properties of mixed micelles in bile, as well as the permeability of paracellular junctions to other solutes that enter canalicular bile by solvent drag.200 In addition, certain unconjugated bile acids such as ursodeoxycholic acid or C23 side chain-shortened bile acid analogs such as nor-ursodeoxycholic acid can induce a bicarbonaterich “hypercholeresis,” which is defined as bile flow greater than can be explained simply by the bile acid osmotic effects.201,202 In that proposed mechanism, unconjugated bile acids in bile becomes pronated and are passively absorbed by the biliary epithelial cells (cholangiocytes). The unconjugated bile acid is then exported from the biliary epithelium into the venous drainage of the biliary tract, which empties into the portal vein or directly enters the liver, thereby returning the bile acids to hepatocytes for uptake and resecretion into bile.201,203 This “cholehepatic shunt pathway” carries a proton each time a molecule of bile acid is absorbed, thereby generating an additional bicarbonate ion from biliary carbon dioxide per cycle. The majority of bile flow is bile acid dependent in humans, whereas most of the bile flow in rodents is induced by the secretion of other anions.200,204 For example, the bile acid-dependent and bile acid-independent flow in rats has been estimated to be approximately 50 μL/kg-min and 70 μL/kg-min, respectively.200,204 Mouse models with a genetic defect in hepatic bile acid secretion exhibit relatively normal levels of bile flow.205 In contrast, a similar defect in humans is associated with significantly impaired bile flow (cholestasis).206 However, even in humans, secretion of other primary solutes by the hepatocyte and biliary epithelium contributes significantly to bile formation. Hepatobiliary excretion of reduced glutathione (GSH) and bicarbonate (HCO3 - ) constitute major components of the bile acid-independent fraction of bile flow.207,208 The ATP-dependent canalicular secretion of GSH via the multidrug resistance-associated protein-2 (MRP2; gene symbol: ABCC2) plays a particularly important role.209 Besides the ATP-dependent secretion of organic anions into bile, hepatic and biliary ATP-independent secretion
of bicarbonate via the HCO3 - /Cl− anion exchanger AE2 contributes to the bile acid-independent bile flow.210 The majority of this HCO3 - secretion is mediated by cholangiocytes lining the biliary tract211 in response to stimulation by a variety of hormones and neuropeptides such as secretin and vasoactive intestinal peptide.212 Biliary HCO3− secretion in humans far exceeds that of rodents and may be responsible for as much as 30% of total bile flow versus only 5%–10% in rodent models.200 In addition to secretion of HCO3 - , other ductular modifications to hepatic bile include the absorption of solutes such as glucose, amino acids, and bile acids, chloride secretion by CFTR and non-CFTR pathways, hydrolysis of GSH by γ-glutamylpeptidase, and movement of water through specific channels (aquaporins) and paracellularly.212,213
41.6.2 Overview of Hepatic Bile Acid Transport In the fasting state, bile acids are taken up predominantly by the periportal hepatocytes (the first hepatocytes of the liver acinus), whereas during feeding, more hepatocytes in the liver acinus participate in bile acid uptake.214,215 Conversely, perivenous (pericentral) hepatocytes are primarily responsible for bile acid synthesis.91 As a generalization, periportal hepatocytes absorb and secrete recirculating bile acids, whereas perivenous (pericentral) cells secrete predominantly newly synthesized bile acids. Hepatocellular uptake of bile acids occurs against an unfavorable electrochemical ion gradient and results in a 5–10-fold concentration gradient between the plasma filtrate present in the space of Disse and the hepatocyte cytosol.216 In lower vertebrates such as the little skate (Leucoraja erinacea), the uptake of bile acids (primarily the C27 bile alcohol sulfate: scymnol sulfate) is Na+ independent and mediated by a member of the organic anion transporting polypeptide (OATP) family.217,218 However, in higher vertebrates and all mammals studied to date, hepatocellular uptake of most conjugated bile acids under physiological conditions is Na+ dependent and mediated by the Na+-taurocholate cotransporting polypeptide (NTCP; gene symbol: SLC10A1).219 In contrast to conjugated bile acids, hepatocellular uptake of unconjugated bile acids is mediated by members of the OATP family with NTCP playing only a minor role.220,221 The transporters responsible for the enterohepatic circulation of bile acids have been identified and are shown in Fig. 41.5.3,222,223 After uptake, conjugated bile acids are shuttled across the hepatocyte to the canalicular membrane for secretion into bile; unconjugated bile acids follow a similar path after first undergoing N-acyl amidation (conjugation) to taurine or glycine.13 Although bile acid concentrations within the hepatocyte are presumed to be in the micromolar range, canalicular bile acid concentrations can be 1000-fold higher, necessitating active transport across the canalicu-
Bile Formation and the Enterohepatic Circulation Chapter | 41 943
FIG. 41.5 Enterohepatic circulation of bile acids showing the individual transport proteins in hepatocytes, ileocytes (ileal enterocytes), and renal proximal tubule cells. After their synthesis or reconjugation, taurine and glycine-conjugated bile acids (T/G-BA) are secreted into bile by the canalicular bile salt export pump (BSEP; gene symbol ABCB11). A fraction of the bile acids secreted into bile acids undergo cholehepatic shunting. In this pathway, conjugated bile acids are taken up by cholangiocytes via the apical sodium-dependent bile acid transporter (ASBT; gene symbol SLC10A2), and exported across the basolateral membrane via the heteromeric organic solute transporter OSTα-OSTβ (gene symbols: SLC51A-SLC51B), and possibly the multidrug resistance-associated protein-3 (MRP3; gene symbol ABCC3), for return to the hepatocyte. Ultimately, bile acids empty into the intestinal lumen. Conjugated bile acids are poorly absorbed in the proximal small intestine, but efficiently taken up by the ileal enterocytes via the ASBT. The bile acids bind to the ileal bile acid binding protein (IBABP; also called the ileal lipid binding protein, ILBP; gene symbol FABP6) in the cytosol, and are efficiently exported across the basolateral membrane into the portal circulation by OSTα-OSTβ. The multidrug resistance-associated protein-3 (MRP3; gene symbol ABCC3) is a minor contributor to basolateral export of native bile acids from the enterocyte, but may have a more significant role in export of any glucuronidated (U-BA) or sulfated (S-BA) bile acids that may be formed. Although most bile acids are absorbed in the small intestine, colonocytes express appreciable levels of MRP3 and OSTα-OSTβ. These carriers may serve to export unconjugated bile acids (BA) that were taken up by passive diffusion from the lumen of the colon. After their absorption from the intestine, bile acids travel back to the liver where they are cleared by the Na+-taurocholate cotransporting polypeptide (NTCP; gene symbol SLC10A1). Members of the organic anion transport protein family, OATP1B1 (gene symbol SLCO1B1) and OATP1B3 (gene symbol SLCO1B3) (Oatp1b2; gene symbol Slco1b2, in mice) also participate, and are particularly important for unconjugated bile acids. Under cholestatic conditions, unconjugated, conjugated, or modified (glucuronidated or sulfated) bile acids are effluxed across the basolateral (sinusoidal) membrane of the hepatocyte by OSTα-OSTβ, MRP3, or multidrug resistance-associated protein-4 (MRP4; gene symbol ABCC4) into the systemic circulation. Under normal physiological conditions, a fraction of the bile acid escapes first pass hepatic clearance enters the systemic circulation. The free bile acids are filtered by the renal glomerulus, efficiently reclaimed by the ASBT in the proximal tubules, and exported back into the systemic circulation, thereby minimizing their excretion in the urine. A fraction of the glucuronidated or sulfated bile acids can also be exported across the apical membrane by multidrug resistance-associated protein-2 (MRP2; gene symbol ABCC2). The tetrahydroxylated bile acids (H-BA) formed under pathophysiological conditions can be secreted into bile by the multidrug resistance protein (MDR1; gene symbol ABCB1; Mdr1a/b in mice). (Adapted with permission from Dawson PA, Hubbert ML, Rao A. Getting the mOST from OST: role of organic solute transporter, OSTalpha-OSTbeta, in bile acid and steroid metabolism. Biochim Biophys Acta 2010;1801(9):994–1004.)
944 SECTION | V Physiology of Secretion
lar membrane via the ATP-dependent bile salt export pump (BSEP; gene symbol: ABCB11).224–226 A small amount of bile acid can be modified by the addition of sulfate or glucuronide, and is secreted into bile by other ABC transporters, primarily MRP2.227 Sulfation is particularly important for the monohydroxy bile acid, lithocholic acid. Under pathophysiological conditions, bile acids also undergo additional hydroxylation reactions in hepatocytes and these unusual tri- or tetrahydroxylated bile acid species are secreted into bile by MRP2228 and possibly by the breast cancer-related protein (BCRP; gene symbol: ABCG2)228 or P-glycoprotein (MDR1; gene symbol: ABCB1).229 Although sinusoidal membrane bile acid transport is overwhelming in the direction of uptake under normal physiological conditions, bile acids can also be effluxed from the interior of the hepatocyte into the space of Disse as a protective mechanism to reduce bile acid overload.230 This process may be analogous to the “hepatocyte-hopping” that was recently demonstrated for conjugated bilirubin and drugs.220 Under conditions when the bile acid load returning to periportal hepatocytes exceeds their canalicular bile acid export capacity, a fraction of the bile acids can be rerouted to sinusoidal blood by basolateral membrane efflux, and carried to more pericentral hepatocytes for reuptake and secretion into bile. This mechanism would dynamically recruit additional hepatocytes within the liver lobule for bile acid clearance and serve as an additional safeguard to protect the periportal hepatocytes from bile acid overload. The major transporters involved in hepatocyte sinusoidal bile acid efflux includes members of the MRP (ABCC) family of ATP-dependent transporters, MRP3231 and MRP4,232 and OSTα-OSTβ.233,234 MRP3 and MRP4 are expressed on the sinusoidal membrane of hepatocytes235,236 and have bile acid transport237 or glutathione-bile acid cotransport activity.238 The expression of these transporters is induced under cholestatic conditions to promote bile acid efflux,230,233,239 thus lowering the concentration in the hepatocyte and decreasing the likelihood of apoptosis or necrosis. The relative contribution of the different efflux transporters remains to be determined under different pathophysiological conditions. However, current results indicate especially significant roles for MRP4 and OSTα-OSTβ,232,234 whereas MRP3 may be more important for glucuronidated or sulfated bile acids.240 Overall, the increased expression of these transporters appears to be an important part of the adaptive response to conditions of bile acid overload, along with downregulation of the major hepatic bile acid uptake transporters, NTCP, and members of the OATP family.230
41.6.3 Bile Acid Uptake Across the Sinusoidal Membrane of the Hepatocyte The NTCP is the founding member of the SLC10 family of solute carrier proteins, which includes two bile acid transporters (SLC10A1/NTCP and SLC10A2/ASBT),
one steroid sulfate transporter (SLC10A6/SOAT), and four orphan carriers (SLC10A3, SLC10A4, SLC10A5, SLC10A7).241–243 The NTCP encodes a 349-amino acid (approximately 45 kDa) membrane glycoprotein that functions as an electrogenic sodium-solute cotransporter to mediate the uptake of all the major glycine/taurine-conjugated bile acids.244,245 Depending on the bile acid species and NTCP ortholog, unconjugated bile acids may be moderate or weak substrates,246 and sulfated bile acids appear to be only weakly transported.195 The NTCP accounts for most, if not all, hepatocellular sinusoidal membrane Na+dependent conjugated bile acid transport under physiological conditions in humans.247 The primary role of NTCP in the clearance of conjugated bile acids is strongly supported by recent genetic evidence in humans and mice.248–251 In 2015, a single pediatric patient with an isolated inherited NTCP deficiency and greatly elevated (25–100-fold) plasma levels of conjugated bile acids (hypercholanemia) was reported. In contrast to the inherited defects in bile acid synthesis, the patient exhibited no jaundice, itching (pruritus), fat malabsorption (steatorrhea), or obvious liver disease.248 Alternative transport mechanisms were unable to compensate for the loss of NTCP in this pediatric patient, but subsequent reports of additional NTCP-deficient adults with only mild hypercholanemia suggests that other transporters may contribute in a limited fashion to hepatocellular conjugated bile acid uptake.251 In contrast to humans, mice lacking NTCP maintain normal plasma levels of bile acids.250 This species difference is explained by efficient hepatic uptake of conjugated bile acids by mouse Oatp1a/ Oatp1b family members, including Oatp1a1.249 In addition to being the major hepatic bile acid uptake transporter, NTCP was recently identified as a cell surface receptor for binding of hepatitis B virus (HBV) and hepatitis delta virus (HDV).252 The HBV infection is initiated via lowaffinity binding of the HBV small surface protein to heparinsulfate proteoglycans, followed by high-affinity binding to a host receptor that allows for viral entry. The identity of the HBV entry receptor(s) had long remained a mystery, and a major breakthrough came in 2012 when Yan et al. demonstrated that binding of HBV large surface protein pre-S1 domain myristoylated peptide (amino acids 2–48) to NTCP is essential for infection of human hepatocytes.252 This finding and subsequent analyses of NTCP-HBV interactions has led to rapid basic science and clinical advances in the field.253 In contrast to conjugated bile acids, the hepatic uptake of unconjugated bile acids is mediated primarily by an Na+independent mechanism.219,221,254 This Na+-independent bile acid transport involves several different OATPs, members of a gene family encoding 12 potential transmembrane domain proteins that share no sequence identity with the Na+-dependent bile acid transporters. The driving force for OATP-mediated organic anion uptake remains to be conclusively established, but may involve anion exchange
Bile Formation and the Enterohepatic Circulation Chapter | 41 945
or facilitative diffusion.255 The original Human Genome Organization (HUGO) Gene Nomenclature Committee designation for the supergene family of OATP solute carriers was SLC21A. However, confusion related to species differences led to the adoption of a species-independent classification and nomenclature system in 2004, designated OATP/ SLCO where OATP refers to the protein isoforms and SLCO (“SLC” = solute carrier, “O” = member of the OATP family) refers to the genes.256 The OATP/SLCO-type transporters include 11 human and 16 rat/mouse genes that fall within six subgroups.256,257 In humans, the liver expresses several different OATPs, including OATP1B1, OATP1B3, and OATP2B1. OATP1B1 (gene symbol: SLCO1B1; original protein name: OATP-C) and OATP1B3 (gene symbol: SLCO1B3; original protein name: OATP8) exhibit partially overlapping substrate specificities and account for the majority of hepatic Na+-independent bile acid clearance, as well as uptake of bilirubin glucuronides. OATP2B1 (gene symbol: SLCO2B1; original protein name OATP-B) does not transport bile acids, but functions in human liver along with OATP1B1 and OATP1B3 to transport organic anions and a variety of drugs or drug metabolites. Mice and rats encode only a single OATP1B ortholog (Oatp1b2; gene symbol: Slco1b2), and OATP1B1 and OATP1B3 arose in primates by duplication after divergence from rodents.256 Conversely, the human genome encodes only one member of the OATP1A family (OATP1A2), whereas the rodent genome includes multiple members (Oatp1a1, Oatp1a3, Oatp1a4, Oatp1a5, Oatp1a6) that arose from gene duplication and are clustered in the same chromosomal region (chromosome 6G2 in mouse). In mouse, several different bile acid-transporting Oatps are expressed on the hepatic sinusoidal membrane, including Oatp1a1, Oatp1a4, and Oatp1b2. Compelling genetic evidence comes from recent knockout mouse studies for a significant role of the Oatp1a/1b transporters, particularly Oatp1b2, in hepatic clearance of unconjugated bile acids,220,258 and these transporters participate in the clearance of conjugated bile acids when NTCP is inhibited.249 In humans, OATP1B1/ OATP1B3 plays a primary role in the hepatic clearance of unconjugated bile acids,259 and combined loss of the adjacent SLCO1B1 and SLCO1B3 genes on chromosome 12 causes Rotor Syndrome, a rare benign disorder characterized by elevated plasma levels of unconjugated bile acids and conjugated bilirubin.259
41.6.4 Canalicular Bile Acid Transport Since bile acids are potent detergents, their hepatocellular uptake and export must be carefully balanced to avoid intracellular accumulation and cytotoxicity. Regardless of their uptake mechanism, the returning as well as newly synthesized bile acids are shuttled to the canalicular membrane for secretion into bile by the ATP-dependent bile salt export pump (BSEP; gene symbol: ABCB11).225 BSEP's role as the
major canalicular bile acid efflux pump was confirmed with the identification of ABCB11 mutations in patients with progressive familial intrahepatic cholestasis (PFIC) type 2, a hepatic disorder characterized by biliary bile acid concentrations less than 1% of normal.206,260 When analyzed in transfected mammalian cells, BSEP transports conjugated as well as unconjugated bile acids,246 and in some species, such as humans but not rodents, BSEP also transports sulfated bile acids such as taurolithosulfocholate.261 However under physiological conditions, unconjugated bile acids are first conjugated to glycine or taurine prior to secretion, as evidenced by the low proportions of unconjugated bile acids typically found in bile (less than 5%). Although bile acids are the major physiological substrate, a variety of drugs such as cyclosporin, rifamycin, troglitazone, and glibenclamide also interact with BSEP as nonsubstrate inhibitors to impede bile acid export. This direct inhibition of BSEP is thought to be an important mechanism underlying druginduced hepatoxicity or cholestasis.225,262–264 The BSEP expression is induced when hepatocyte bile acid levels are elevated, such as following dietary challenge with bile acids,265 or under certain cholestatic conditions.266–268 This is due to a direct activation of the human and rodent BSEP gene expression by bile acids signaling via FXR.269–272 The induction of BSEP expression is not a universal property of all bile acid species and appears to correlate with FXR ligand specificity.48,273,274 In addition to transcriptional regulation, there is also posttranscriptional regulation of BSEP protein trafficking and localization to the canalicular membrane.225,275 This short-term regulation enables the hepatocyte to rapidly modulate bile acid secretion in response to bile acid load and to pathophysiological conditions such as cholestasis.276
41.6.5 Cholehepatic Shunt Pathway The cholehepatic shunt pathway was proposed to describe the cycle whereby unconjugated dihydroxy bile acids secreted into bile are passively absorbed by cholangiocytes lining the bile ducts, returned to the hepatocyte via the venous drainage, transported across the sinusoidal membrane, and resecreted into bile.201 Absorption of the protonated unconjugated bile acid molecule generates a bicarbonate anion, resulting in a bicarbonate-rich hypercholeresis (increased bile secretion).277 Cholehepatic shunting is important for the hypercholeresis induced by therapeutic doses of unconjugated C-24 dihydroxy bile acid ursodeoxycholic acid and C-23 bile acid analogs such as nor-ursodeoxycholic acid.201,278,279 At the doses administered, ursodeoxycholate exceeds the capacity for bile acid conjugation, whereas the side chain-shorted bile acid analog nor-ursodeoxycholic acid is resistant to conjugation with glycine or taurine, resulting in hepatic biliary secretion of the unconjugated bile acids at high concentrations.280,281 Because hepatic bile contains primarily conjugated bile acids, the contribution
946 SECTION | V Physiology of Secretion
of cholehepatic shunting of unconjugated bile acids to bile flow was thought to be negligible under normal physiological conditions.201 However, the subsequent discovery that the biliary epithelium and gallbladder expresses the ASBT and OSTα-OSTβ provided a physiological mechanism for cholehepatic or cholecysthepatic shunting of ionized conjugated bile acids.282–286 In vivo data documenting cholehepatic or cholecysthepatic shunting for conjugated bile acids has been obtained using rat and mouse models,285,286 but the quantitative significance in humans or under different physiological or pathophysiological conditions still remains to be determined.287,288
41.7. INTESTINAL ABSORPTION OF BILE ACIDS 41.7.1 Overview Bile acids are reclaimed through a combination of passive absorption in the proximal small intestine, active transport in the distal ileum, and passive absorption in the colon.3 In many species including humans, the bile acid pool includes glycine conjugates, unconjugated bile acids, and hydrophobic bile acid species.13 A fraction of the glycine conjugates and unconjugated bile acids are protonated at the acidic cell surface pH conditions of the intestinal mucosa and are absorbed by passive diffusion across the apical brush border membrane.289 Besides passive membrane diffusion, there is limited evidence of carrier-mediated transport of bile acids in the proximal small intestine of some species.290 In terminal ileum, the ASBT actively transports bile acids from the intestinal lumen across the apical brush border membrane.247,291 Bile acids are then shuttled to the basolateral membrane and effluxed into the portal circulation by OSTα-OSTβ.63 Several observations support the concept that terminal ileum is the major site of bile acid reabsorption in man and experimental animal models. These observations include the finding that there is little decrease in intraluminal bile acid concentration prior to the ileum292 and the appearance of bile acid malabsorption after ileal resection.38 Subsequent studies using in situ perfused intestinal segments to measure bile acid absorption289 demonstrated that ileal bile acid transport is a high capacity absorptive system sufficient to account for the biliary bile acid output. In mouse, bile acids are taurine-conjugated and hydrophilic,293 thereby limiting their nonionic diffusion. Inactivation of the ASBT in mice largely abolished intestinal bile acid absorption, suggesting that alternative uptake mechanisms are of only minor importance.294 Similarly in humans, inherited mutations in ASBT causes significant bile acid malabsorption.295 The general consensus from these studies was that ileal active transport is the major route for conjugated bile acid uptake, whereas passive or facilitative absorption down the length
of the small intestine is significant for unconjugated and some glycine-conjugated bile acids. However, the quantitative contribution of jejunal bile acid absorption in different species is still being debated.296 Little is known about intracellular transport of bile acids in the enterocyte. The ileal bile acid binding protein (IBABP, also called the ileal lipid binding protein, ILBP; gene symbol: FABP6) is a member of the fatty acid binding protein family and very abundant cytosolic protein in ileal enterocytes.297,298 The IBABP expression parallels that of the ASBT and is induced by bile acids acting through FXR.299 The IBABP preferentially binds bile acids, with a stoichiometry of two or three bile acids per molecular of IBABP and with highest affinity for taurine-conjugated bile acids versus glycine-conjugated and unconjugated species.300,301 It is likely that IBABP functions as an intracellular buffer to bind bile acids and facilitate their transcellular transport. In doing so, IBABP reduces bile acid-membrane interactions and protects the ileal enterocyte from bile acid cytotoxicity.302 Genetic evidence that IBABP plays a role in intestinal bile acid transport comes from studies of an IBABP-null mouse.303 In that model, apical to basolateral transport of taurocholate was reduced when analyzed using the inverted gut sac method of Wilson and Wiseman. In addition, bile acid metabolism was altered suggesting that IBABP may also be involved in enterocyte bile acid sensing.
41.7.2 Ileal Na+-Dependent Bile Acid Uptake Bile acids are actively transported across the ileal brush border membrane by the apical sodium-dependent bile acid transporter (ASBT).195,295 The ASBT transports all major species of bile acids, but favors trihydroxy over dihydroxy bile acids, and conjugated over unconjugated species.195 The intestinal absorption efficiency for an individual bile acid species depends on the affinity of ASBT and the length of time the bile acid is exposed to ASBT-expressing epithelium. As such, the substrate specificity and gradient of ASBT intestinal expression are important determinants of the bile acid pool composition,304 along with the hepatic expression of 12α-hydroxylase (CYP8B1) and the gut microbiota flora (levels of 7α-dehydroxlating bacteria). The properties of the ASBT satisfied all the functional criteria for ileal active bile acid uptake, including: (1) a strict sodium dependence for bile acid transport195; (2) specific transport of all the major species of bile acids with negligible uptake of nonbile acid solutes195,245; (3) specific intestinal expression in the terminal ileum; (4) similar ontogeny for ileal sodium-dependent taurocholate uptake and ASBT expression at fetal day 22 and postnatal day 17305,306; (5) targeted inactivation of the ASBT gene eliminates enterohepatic cycling of bile acids in mice294; (6) loss-offunction mutations in the human ASBT gene are associated
Bile Formation and the Enterohepatic Circulation Chapter | 41 947
with intestinal bile acid malabsorption.295 In addition to the ileal enterocyte, the ASBT is expressed in other tissues that serve to facilitate the enterohepatic circulation of bile acids, including the apical membrane of proximal renal convoluted tubule cells, large cholangiocytes, and gallbladder epithelial cells.247 Considering its central role in the enterohepatic circulation, inherited defects or dysfunctional regulation of the ASBT may play a role in the pathogenesis of a number of gastrointestinal disorders. For example, ASBT mutations were identified as a cause of Primary Bile Acid Malabsorption, a rare idiopathic disorder associated with chronic diarrhea beginning in early infancy, steatorrhea, interruption of the enterohepatic circulation of bile acids, and reduced plasma cholesterol levels.295 Other disorders associated with intestinal bile acid malabsorption that could potentially involve the ASBT in their pathogenesis or clinical phenotype include hypertriglyceridemia,307,308 idiopathic chronic diarrhea,309 chronic ileitis,310 gallstone disease,311,312 postcholecystectomy diarrhea,313 and irritable bowel syndrome.314
41.7.3 Ileal Basolateral Bile Acid Transport The unusual heteromeric transporter, OSTα-OSTβ was identified as the carrier responsible for bile acid export across the basolateral membrane of the ileal enterocyte, cholangiocytes, and renal proximal tubule cell.63 The breakthrough in this area came with the elegant expression cloning of an unusual transporter, OSTα-OSTβ, from the little skate (Raja erinacea)315 followed by cloning of the mammalian orthologs.316 In contrast to the previously identified bile acid carriers, co-expression of two distinct subunits: OSTα, a 340 amino acid polytopic membrane protein, and OSTβ, a 128 amino acid predicted type Ib membrane protein, was required for bile acid transport. Functional studies of the individual subunits to date indicate that co-expression and assembly of both subunits into a complex is required for their trafficking to the plasma membrane and solute transport activity.317–320 Soon after their cloning, OSTα-OSTβ was identified as a candidate ileal basolateral bile acid transporter using a transcriptional profiling approach.317 Support for a role of OSTα-OSTβ in basolateral bile acid transport includes: (1) intestinal expression of OSTα and OSTβ mRNA that generally follows that of the ASBT, with highest levels in ileum,196,317,321 (2) appropriate cellular localization on the lateral and basal membranes of ileal enterocyte,317 (3) expression of OSTα-OSTβ on the basolateral membrane of hepatocytes, cholangiocytes, and renal proximal tubule cells, other membranes exhibiting bile acid efflux,196 (4) efficient transport of the major bile acid species,196,317 (5) positive regulation of OSTα-OSTβ expression by bile acids acting via FXR,64,322 (6) targeted inactivation of the OSTα gene in mice resulted in impaired intestinal bile acid absorption and
altered bile acid metabolism,323–325 and (7) inherited mutations in the human OSTβ gene (SLC51B) cause congenital chronic diarrhea, a phenotype similar to that observed in patients with ASBT mutations.326
REFERENCES 1. Boyer JL. Bile formation and secretion. Compr Physiol 2013;3(3):1035–78. 2. Reuben A. The biliary cycle of Moritz Schiff. Hepatology 2005;42(2):500–5. 3. Dawson PA, Karpen SJ. Intestinal transport and metabolism of bile acids. J Lipid Res 2015;56(6):1085–99. 4. Hofmann AF, Hagey LR, Krasowski MD. Bile salts of vertebrates: structural variation and possible evolutionary significance. J Lipid Res 2010;51(2):226–46. 5. Fahy E, Subramaniam S, Brown HA, Glass CK, Merrill Jr. AH, Murphy RC, et al. A comprehensive classification system for lipids. J Lipid Res 2005;46(5):839–61. 6. Moschetta A, Xu F, Hagey LR, van Berge Henegouwen GP, van Erpecum KJ, Brouwers JF, et al. A phylogenetic survey of biliary lipids in vertebrates. J Lipid Res 2005;46(10):2221–32. 7. Haslewood GA. Bile salt evolution. J Lipid Res 1967;8(6):535–50. 8. Hagey LR, Moller PR, Hofmann AF, Krasowski MD. Diversity of bile salts in fish and amphibians: evolution of a complex biochemical pathway. Physiol Biochem Zool 2010;83(2):308–21. 9. Hagey LR, Vidal N, Hofmann AF, Krasowski MD. Evolutionary diversity of bile salts in reptiles and mammals, including analysis of ancient human and extinct giant ground sloth coprolites. BMC Evol Biol 2010;10:133. 10. Hofmann AF, Hagey LR. Key discoveries in bile acid chemistry and biology and their clinical applications: history of the last eight decades. J Lipid Res 2014;55(8):1553–95. 11. Hofmann AF. Biliary secretion and excretion in health and disease: current concepts. Ann Hepatol 2007;6(1):15–27. 12. Moro-Sibilot L, Blanc P, Taillardet M, Bardel E, Couillault C, Boschetti G, et al. Mouse and human liver contain immunoglobulin A-secreting cells originating from Peyer's patches and directed against intestinal antigens. Gastroenterology 2016;151(2):311–23. 13. Hofmann AF, Hagey LR. Bile acids: chemistry, pathochemistry, biology, pathobiology, and therapeutics. Cell Mol Life Sci 2008; 65(16):2461–83. 14. Wilson FA, Sallee VL, Dietschy JM. Unstirred water layers in intestine: rate determinant of fatty acid absorption from micellar solutions. Science 1971;174(13):1031–3. 15. Hernell O, Staggers JE, Carey MC. Physical-chemical behavior of dietary and biliary lipids during intestinal digestion and absorption. 2. Phase analysis and aggregation states of luminal lipids during duodenal fat digestion in healthy adult human beings. Biochemistry 1990;29(8):2041–56. 16. Hofmann AF, Borgstroem B. The intraluminal phase of fat digestion in man: the lipid content of the micellar and oil phases of intestinal content obtained during fat digestion and absorption. J Clin Invest 1964;43:247–57. 17. Setchell KD, Heubi JE, Shah S, Lavine JE, Suskind D, AlEdreesi M, et al. Genetic defects in bile acid conjugation cause fatsoluble vitamin deficiency. Gastroenterology 2013;144(5):945–55. e6; quiz e14-5.
948 SECTION | V Physiology of Secretion
18. Heubi JE, Setchell KD, Jha P, Buckley D, Zhang W, Rosenthal P, et al. Treatment of bile acid amidation defects with glycocholic acid. Hepatology 2015;61(1):268–74. 19. Gass J, Vora H, Hofmann AF, Gray GM, Khosla C. Enhancement of dietary protein digestion by conjugated bile acids. Gastroenterology 2007;133(1):16–23. 20. Dietschy JM, Turley SD, Spady DK. Role of liver in the maintenance of cholesterol and low density lipoprotein homeostasis in different animal species, including humans. J Lipid Res 1993;34(10):1637–59. 21. Dietschy JM, Turley SD. Control of cholesterol turnover in the mouse. J Biol Chem 2002;277(6):3801–4. 22. Dawson PA. Impact of inhibiting ileal apical versus basolateral bile acid transport on cholesterol metabolism and atherosclerosis in mice. Dig Dis 2015;33(3):382–7. 23. Lorenzo-Zuniga V, Bartoli R, Planas R, Hofmann AF, Vinado B, Hagey LR, et al. Oral bile acids reduce bacterial overgrowth, bacterial translocation, and endotoxemia in cirrhotic rats. Hepatology 2003;37(3):551–7. 24. Inagaki T, Moschetta A, Lee YK, Peng L, Zhao G, Downes M, et al. Regulation of antibacterial defense in the small intestine by the nuclear bile acid receptor. Proc Natl Acad Sci U S A 2006;103(10):3920–5. 25. Verbeke L, Farre R, Verbinnen B, Covens K, Vanuytsel T, Verhaegen J, et al. The FXR agonist obeticholic acid prevents gut barrier dysfunction and bacterial translocation in cholestatic rats. Am J Pathol 2015;185(2):409–19. 26. Ubeda M, Lario M, Munoz L, Borrero MJ, Rodriguez-Serrano M, Sanchez-Diaz AM, et al. Obeticholic acid reduces bacterial translocation and inhibits intestinal inflammation in cirrhotic rats. J Hepatol 2016;64(5):1049–57. 27. Ridlon JM, Kang DJ, Hylemon PB, Bajaj JS. Bile acids and the gut microbiome. Curr Opin Gastroenterol 2014;30(3):332–8. 28. Long SL, Gahan CGM, Joyce SA. Interactions between gut bacteria and bile in health and disease. Mol Aspects Med 2017;56:54–65. 29. Kakiyama G, Pandak WM, Gillevet PM, Hylemon PB, Heuman DM, Daita K, et al. Modulation of the fecal bile acid profile by gut microbiota in cirrhosis. J Hepatol 2013;58(5):949–55. 30. David LA, Maurice CF, Carmody RN, Gootenberg DB, Button JE, Wolfe BE, et al. Diet rapidly and reproducibly alters the human gut microbiome. Nature 2014;505(7484):559–63. 31. Emmett M, Guirl MJ, Santa Ana CA, Porter JL, Neimark S, Hofmann AF, et al. Conjugated bile acid replacement therapy reduces urinary oxalate excretion in short bowel syndrome. Am J Kidney Dis 2003;41(1):230–7. 32. JJ G, Hofmann AF, Ton-Nu HT, Schteingart CD, Mysels KJ. Solubility of calcium salts of unconjugated and conjugated natural bile acids. J Lipid Res 1992;33(5):635–46. 33. Moore EW, Celic L, Ostrow JD. Interactions between ionized calcium and sodium taurocholate: bile salts are important buffers for prevention of calcium-containing gallstones. Gastroenterology 1982;83(5):1079–89. 34. Moore EW. The role of calcium in the pathogenesis of gallstones: Ca++ electrode studies of model bile salt solutions and other biologic systems. With an hypothesis on structural requirements for Ca++ binding to proteins and bile acids. Hepatology 1984;4(5 Suppl):228S–43S. 35. Dobbins JW, Binder HJ. Effect of bile salts and fatty acids on the colonic absorption of oxalate. Gastroenterology 1976;70(6):1096–100.
36. Moreland AM, Santa Ana CA, Asplin JR, Kuhn JA, Holmes RP, Cole JA, et al. Steatorrhea and hyperoxaluria in severely obese patients before and after Roux-en-Y gastric bypass. Gastroenterology 2017;152(5):1055–67. e3. 37. Chadwick VS, Modha K, Dowling RH. Mechanism for hyperoxaluria in patients with ileal dysfunction. N Engl J Med 1973;289(4): 172–6. 38. Hofmann AF, Poley JR. Role of bile acid malabsorption in pathogenesis of diarrhea and steatorrhea in patients with ileal resection. I. Response to cholestyramine or replacement of dietary long chain triglyceride by medium chain triglyceride. Gastroenterology 1972;62(5):918–34. 39. Chiang JY. Bile acid metabolism and signaling. Compr Physiol 2013;3(3):1191–212. 40. Kuipers F, Bloks VW, Groen AK. Beyond intestinal soap–bile acids in metabolic control. Nat Rev Endocrinol 2014;10(8):488–98. 41. Copple BL, Li T. Pharmacology of bile acid receptors: evolution of bile acids from simple detergents to complex signaling molecules. Pharmacol Res 2016;104:9–21. 42. de Aguiar Vallim TQ, Tarling EJ, Edwards PA. Pleiotropic roles of bile acids in metabolism. Cell Metab 2013;17(5):657–69. 43. Zhou H, Hylemon PB. Bile acids are nutrient signaling hormones. Steroids 2014;86:62–8. 44. Craven PA, Pfanstiel J, DeRubertis FR. Role of activation of protein kinase C in the stimulation of colonic epithelial proliferation and reactive oxygen formation by bile acids. J Clin Invest 1987;79(2):532–41. 45. Beuers U, Throckmorton DC, Anderson MS, Isales CM, Thasler W, Kullak-Ublick GA, et al. Tauroursodeoxycholic acid activates protein kinase C in isolated rat hepatocytes. Gastroenterology 1996;110(5):1553–63. 46. Jones BA, Rao YP, Stravitz RT, Gores GJ. Bile salt-induced apoptosis of hepatocytes involves activation of protein kinase C. Am J Physiol 1997;272(5 Pt 1):G1109–15. 47. Wang H, Chen J, Hollister K, Sowers LC, Forman BM. Endogenous bile acids are ligands for the nuclear receptor FXR/BAR. Mol Cell 1999;3(5):543–53. 48. Parks DJ, Blanchard SG, Bledsoe RK, Chandra G, Consler TG, Kliewer SA, et al. Bile acids: natural ligands for an orphan nuclear receptor. Science 1999;284(5418):1365–8. 49. Makishima M, Okamoto AY, Repa JJ, Tu H, Learned RM, Luk A, et al. Identification of a nuclear receptor for bile acids. Science 1999;284(5418):1362–5. 50. Gomez-Ospina N, Potter CJ, Xiao R, Manickam K, Kim MS, Kim KH, et al. Mutations in the nuclear bile acid receptor FXR cause progressive familial intrahepatic cholestasis. Nat Commun 2016;7:10713. https://doi.org/10.1038/ncomms10713. 51. Zhang ZD, Cayting P, Weinstock G, Gerstein M. Analysis of nuclear receptor pseudogenes in vertebrates: how the silent tell their stories. Mol Biol Evol 2008;25(1):131–43. 52. Sayin SI, Wahlstrom A, Felin J, Jantti S, Marschall HU, Bamberg K, et al. Gut microbiota regulates bile acid metabolism by reducing the levels of tauro-beta-muricholic acid, a naturally occurring FXR antagonist. Cell Metab 2013;17(2):225–35. 53. Mueller M, Thorell A, Claudel T, Jha P, Koefeler H, Lackner C, et al. Ursodeoxycholic acid exerts farnesoid X receptor-antagonistic effects on bile acid and lipid metabolism in morbid obesity. J Hepatol 2015;62(6):1398–404.
Bile Formation and the Enterohepatic Circulation Chapter | 41 949
54. Zhang Y, Kast-Woelbern HR, Edwards PA. Natural structural variants of the nuclear receptor farnesoid X receptor affect transcriptional activation. J Biol Chem 2003;278(1):104–10. 55. Vaquero J, Monte MJ, Dominguez M, Muntane J, Marin JJ. Differential activation of the human farnesoid X receptor depends on the pattern of expressed isoforms and the bile acid pool composition. Biochem Pharmacol 2013;86(7):926–39. 56. Boesjes M, Bloks VW, Hageman J, Bos T, van Dijk TH, Havinga R, et al. Hepatic farnesoid X-receptor isoforms alpha2 and alpha4 differentially modulate bile salt and lipoprotein metabolism in mice. PLoS ONE 2014;9(12):e115028. 57. Reschly EJ, Ai N, Ekins S, Welsh WJ, Hagey LR, Hofmann AF, et al. Evolution of the bile salt nuclear receptor FXR in vertebrates. J Lipid Res 2008;49(7):1577–87. 58. Krasowski MD, Ai N, Hagey LR, Kollitz EM, Kullman SW, Reschly EJ, et al. The evolution of farnesoid X, vitamin D, and pregnane X receptors: insights from the green-spotted pufferfish (Tetraodon nigriviridis) and other non-mammalian species. BMC Biochem 2011;12:5. 59. Gadaleta RM, van Mil SW, Oldenburg B, Siersema PD, Klomp LW, van Erpecum KJ. Bile acids and their nuclear receptor FXR: relevance for hepatobiliary and gastrointestinal disease. Biochim Biophys Acta 2010;1801(7):683–92. 60. Modica S, Bellafante E, Moschetta A. Master regulation of bile acid and xenobiotic metabolism via the FXR, PXR and CAR trio. Front Biosci 2009;(14)4719–45. 61. de Aguiar Vallim TQ, Tarling EJ, Ahn H, Hagey LR, Romanoski CE, Lee RG, et al. MAFG is a transcriptional repressor of bile acid synthesis and metabolism. Cell Metab 2015;21(2):298–310. 62. Kong B, Wang L, Chiang JY, Zhang Y, Klaassen CD, Guo GL. Mechanism of tissue-specific farnesoid X receptor in suppressing the expression of genes in bile-acid synthesis in mice. Hepatology 2012;56(3):1034–43. 63. Dawson PA, Hubbert ML, Rao A. Getting the mOST from OST: role of organic solute transporter, OSTalpha-OSTbeta, in bile acid and steroid metabolism. Biochim Biophys Acta 2010;1801(9):994–1004. 64. Frankenberg T, Rao A, Chen F, Haywood J, Shneider BL, Dawson PA. Regulation of the mouse organic solute transporter alphabeta, Ostalpha-Ostbeta, by bile acids. Am J Physiol Gastrointest Liver Physiol 2006;290(5):G912–22. 65. Inagaki T, Choi M, Moschetta A, Peng L, Cummins CL, McDonald JG, et al. Fibroblast growth factor 15 functions as an enterohepatic signal to regulate bile acid homeostasis. Cell Metab 2005;2(4):217–25. 66. Lundasen T, Galman C, Angelin B, Rudling M. Circulating intestinal fibroblast growth factor 19 has a pronounced diurnal variation and modulates hepatic bile acid synthesis in man. J Intern Med 2006;260(6):530–6. 67. Gadaleta RM, van Erpecum KJ, Oldenburg B, Willemsen EC, Renooij W, Murzilli S, et al. Farnesoid X receptor activation inhibits inflammation and preserves the intestinal barrier in inflammatory bowel disease. Gut 2011;60(4):463–72. 68. Chavez-Talavera O, Tailleux A, Lefebvre P, Staels B. Bile acid control of metabolism and inflammation in obesity, type 2 diabetes, dyslipidemia, and nonalcoholic fatty liver disease. Gastroenterology 2017;152(7):1679–94. e3. 69. Modica S, Murzilli S, Salvatore L, Schmidt DR, Moschetta A. Nuclear bile acid receptor FXR protects against intestinal tumorigenesis. Cancer Res 2008;68(23):9589–94.
70. Goodwin B, Gauthier KC, Umetani M, Watson MA, Lochansky MI, Collins JL, et al. Identification of bile acid precursors as endogenous ligands for the nuclear xenobiotic pregnane X receptor. Proc Natl Acad Sci U S A 2003;100(1):223–8. 71. Xie W, Radominska-Pandya A, Shi Y, Simon CM, Nelson MC, Ong ES, et al. An essential role for nuclear receptors SXR/PXR in detoxification of cholestatic bile acids. Proc Natl Acad Sci U S A 2001;98(6):3375–80. 72. Makishima M, TT L, Xie W, Whitfield GK, Domoto H, Evans RM, et al. Vitamin D receptor as an intestinal bile acid sensor. Science 2002;296(5571):1313–6. 73. Kollitz EM, Zhang G, Hawkins MB, Whitfield GK, Reif DM, Kullman SW. Evolutionary and functional diversification of the vitamin D receptor-lithocholic acid partnership. PLoS One 2016;11(12):e0168278. 74. Hofmann AF. Detoxification of lithocholic acid, a toxic bile acid: relevance to drug hepatotoxicity. Drug Metab Rev 2004;36(3–4):703–22. 75. Han S, Li T, Ellis E, Strom S, Chiang JY. A novel bile acid-activated vitamin D receptor signaling in human hepatocytes. Mol Endocrinol 2010;24(6):1151–64. 76. Li T, Chiang JY. Mechanism of rifampicin and pregnane X receptor inhibition of human cholesterol 7 alpha-hydroxylase gene transcription. Am J Physiol Gastrointest Liver Physiol 2005;288(1):G74–84. 77. Maruyama T, Miyamoto Y, Nakamura T, Tamai Y, Okada H, Sugiyama E, et al. Identification of membrane-type receptor for bile acids (M-BAR). Biochem Biophys Res Commun 2002;298(5): 714–9. 78. Kawamata Y, Fujii R, Hosoya M, Harada M, Yoshida H, Miwa M, et al. A G protein-coupled receptor responsive to bile acids. J Biol Chem 2003;278(11):9435–40. 79. Sato H, Macchiarulo A, Thomas C, Gioiello A, Une M, Hofmann AF, et al. Novel potent and selective bile acid derivatives as TGR5 agonists: biological screening, structure-activity relationships, and molecular modeling studies. J Med Chem 2008;51(6):1831–41. 80. Pellicciari R, Costantino G, Camaioni E, Sadeghpour BM, Entrena A, Willson TM, et al. Bile acid derivatives as ligands of the farnesoid X receptor. Synthesis, evaluation, and structure-activity relationship of a series of body and side chain modified analogues of chenodeoxycholic acid. J Med Chem 2004;47(18):4559–69. 81. Rizzo G, Passeri D, De Franco F, Ciaccioli G, Donadio L, Rizzo G, et al. Functional characterization of the semisynthetic bile acid derivative INT-767, a dual farnesoid X receptor and TGR5 agonist. Mol Pharmacol 2010;78(4):617–30. 82. Keitel V, Reinehr R, Gatsios P, Rupprecht C, Gorg B, Selbach O, et al. The G-protein coupled bile salt receptor TGR5 is expressed in liver sinusoidal endothelial cells. Hepatology 2007;45(3):695–704. 83. Li T, Chiang JY. Bile acid signaling in metabolic disease and drug therapy. Pharmacol Rev 2014;66(4):948–83. 84. Alemi F, Poole DP, Chiu J, Schoonjans K, Cattaruzza F, Grider JR, et al. The receptor TGR5 mediates the prokinetic actions of intestinal bile acids and is required for normal defecation in mice. Gastroenterology 2013;144(1):145–54. 85. Li T, Holmstrom SR, Kir S, Umetani M, Schmidt DR, Kliewer SA, et al. The G protein-coupled bile acid receptor, TGR5, stimulates gallbladder filling. Mol Endocrinol 2011;25(6):1066–71. 86. Alemi F, Kwon E, Poole DP, Lieu T, Lyo V, Cattaruzza F, et al. The TGR5 receptor mediates bile acid-induced itch and analgesia. J Clin Invest 2013;123(4):1513–30.
950 SECTION | V Physiology of Secretion
87. Reich M, Deutschmann K, Sommerfeld A, Klindt C, Kluge S, Kubitz R, et al. TGR5 is essential for bile acid-dependent cholangiocyte proliferation in vivo and in vitro. Gut 2016;65(3):487–501. 88. Thomas C, Auwerx J, Schoonjans K. Bile acids and the membrane bile acid receptor TGR5--connecting nutrition and metabolism. Thyroid 2008;18(2):167–74. 89. van Nierop FS, Scheltema MJ, Eggink HM, Pols TW, Sonne DP, Knop FK, et al. Clinical relevance of the bile acid receptor TGR5 in metabolism. Lancet Diabetes Endocrinol 2017;5(3):224–33. 90. Ding L, Sousa KM, Jin L, Dong B, Kim BW, Ramirez R, et al. Vertical sleeve gastrectomy activates GPBAR-1/TGR5 to sustain weight loss, improve fatty liver, and remit insulin resistance in mice. Hepatology 2016;64(3):760–73. 91. Twisk J, Hoekman MF, Mager WH, Moorman AF, de Boer PA, Scheja L, et al. Heterogeneous expression of cholesterol 7 alphahydroxylase and sterol 27-hydroxylase genes in the rat liver lobulus. J Clin Invest 1995;95(3):1235–43. 92. Ridlon JM, Harris SC, Bhowmik S, Kang DJ, Hylemon PB. Consequences of bile salt biotransformations by intestinal bacteria. Gut Microbes 2016;7(1):22–39. 93. Russell DW. Fifty years of advances in bile acid synthesis and metabolism. J Lipid Res 2009;50(Suppl):S120–5. 94. Russell DW. The enzymes, regulation, and genetics of bile acid synthesis. Annu Rev Biochem 2003;72:137–74. 95. Botham KM, Boyd GS. The metabolism of chenodeoxycholic acid to beta-muricholic acid in rat liver. Eur J Biochem 1983;134(1):191–6. 96. Pullinger CR, Eng C, Salen G, Shefer S, Batta AK, Erickson SK, et al. Human cholesterol 7alpha-hydroxylase (CYP7A1) deficiency has a hypercholesterolemic phenotype. J Clin Invest 2002;110(1):109–17. 97. Ferrell JM, Boehme S, Li F, Chiang JY. Cholesterol 7alphahydroxylase-deficient mice are protected from high-fat/ high-cholesterol diet-induced metabolic disorders. J Lipid Res 2016;57(7):1144–54. 98. Rosen H, Reshef A, Maeda N, Lippoldt A, Shpizen S, Triger L, et al. Markedly reduced bile acid synthesis but maintained levels of cholesterol and vitamin D metabolites in mice with disrupted sterol 27-hydroxylase gene. J Biol Chem 1998;273(24):14805–12. 99. Li-Hawkins J, Lund EG, Turley SD, Russell DW. Disruption of the oxysterol 7alpha-hydroxylase gene in mice. J Biol Chem 2000;275(22):16536–42. 100. Repa JJ, Lund EG, Horton JD, Leitersdorf E, Russell DW, Dietschy JM, et al. Disruption of the sterol 27-hydroxylase gene in mice results in hepatomegaly and hypertriglyceridemia. Reversal by cholic acid feeding. J Biol Chem 2000;275(50):39685–92. 101. Lund EG, Xie C, Kotti T, Turley SD, Dietschy JM, Russell DW. Knockout of the cholesterol 24-hydroxylase gene in mice reveals a brain-specific mechanism of cholesterol turnover. J Biol Chem 2003;278(25):22980–8. 102. Dawson P. Bile acid metabolism. In: Ridgway N, McLeod R, editors. Biochemistry of lipids, lipoproteins, and membranes. 6th ed. Amsterdam: Elsevier; 2016. p. 359–89. 103. Bjorkhem I, Meaney S. Brain cholesterol: long secret life behind a barrier. Arterioscler Thromb Vasc Biol 2004;24(5):806–15. 104. Meaney S, Heverin M, Panzenboeck U, Ekstrom L, Axelsson M, Andersson U, et al. Novel route for elimination of brain oxysterols across the blood-brain barrier: conversion into 7alpha-hydroxy3-oxo-4-cholestenoic acid. J Lipid Res 2007;48(4):944–51. 105. Bjorkhem I. Five decades with oxysterols. Biochimie 2013;95(3):448–54.
106.
Li-Hawkins J, Gafvels M, Olin M, Lund EG, Andersson U, Schuster G, et al. Cholic acid mediates negative feedback regulation of bile acid synthesis in mice. J Clin Invest 2002;110(8):1191–200. 107. Zhang Y, Klaassen CD. Effects of feeding bile acids and a bile acid sequestrant on hepatic bile acid composition in mice. J Lipid Res 2010;51(11):3230–42. 108. Bathena SP, Mukherjee S, Olivera M, Alnouti Y. The profile of bile acids and their sulfate metabolites in human urine and serum. J Chromatogr B Analyt Technol Biomed Life Sci 2013;942–943:53–62. 1 09. Setchell KD, Dumaswala R, Colombo C, Ronchi M. Hepatic bile acid metabolism during early development revealed from the analysis of human fetal gallbladder bile. J Biol Chem 1988;263(32):16637–44. 1 10. Lee CS, Kimura A, JF W, Ni YH, Hsu HY, Chang MH, et al. Prognostic roles of tetrahydroxy bile acids in infantile intrahepatic cholestasis. J Lipid Res 2017;58(3):607–14. 1 11. Wang DQ, Tazuma S, Cohen DE, Carey MC. Feeding natural hydrophilic bile acids inhibits intestinal cholesterol absorption: studies in the gallstone-susceptible mouse. Am J Physiol Gastrointest Liver Physiol 2003;285(3):G494–502. 1 12. Kunne C, Acco A, Hohenester S, Duijst S, de Waart DR, Zamanbin A, et al. Defective bile salt biosynthesis and hydroxylation in mice with reduced cytochrome P450 activity. Hepatology 2013;57(4):1509–17. 1 13. Rudling M. Understanding mouse bile acid formation: is it time to unwind why mice and rats make unique bile acids? J Lipid Res 2016;57(12):2097–8. 1 14. Takahashi S, Fukami T, Masuo Y, Brocker CN, Xie C, Krausz KW, et al. Cyp2c70 is responsible for the species difference in bile acid metabolism between mice and humans. J Lipid Res 2016;57(12): 2130–7. 1 15. Wahlström A, Al-Dury S, Ståhlman M, Bäckhed F, Marschall HU. Cyp3a11 is not essential for the formation of murine bile acids. Biochem Biophys Rep 2017;10:70–5. 1 16. Hubbard B, Doege H, Punreddy S, Wu H, Huang X, Kaushik VK, et al. Mice deleted for fatty acid transport protein 5 have defective bile acid conjugation and are protected from obesity. Gastroenterology 2006;130(4):1259–69. 1 17. Alnouti Y, Csanaky IL, Klaassen CD. Quantitative-profiling of bile acids and their conjugates in mouse liver, bile, plasma, and urine using LC-MS/MS. J Chromatogr B Analyt Technol Biomed Life Sci 2008;873(2):209–17. 1 18. Falany CN, Johnson MR, Barnes S, Diasio RB. Glycine and taurine conjugation of bile acids by a single enzyme. Molecular cloning and expression of human liver bile acid CoA:amino acid N-acyltransferase. J Biol Chem 1994;269(30):19375–9. 1 19. Falany CN, Fortinberry H, Leiter EH, Barnes S. Cloning, expression, and chromosomal localization of mouse liver bile acid CoA:amino acid N-acyltransferase. J Lipid Res 1997;38(6):1139–48. 1 20. Huijghebaert SM, Hofmann AF. Pancreatic carboxypeptidase hydrolysis of bile acid-amino conjugates: selective resistance of glycine and taurine amidates. Gastroenterology 1986;90(2):306–15. 1 21. Ishibashi S, Schwarz M, Frykman PK, Herz J, Russell DW. Disruption of cholesterol 7alpha-hydroxylase gene in mice. I. Postnatal lethality reversed by bile acid and vitamin supplementation. J Biol Chem 1996;271(30):18017–23. 1 22. Schwarz M, Lund EG, Setchell KD, Kayden HJ, Zerwekh JE, Bjorkhem I, et al. Disruption of cholesterol 7alpha-hydroxylase gene in mice. II. Bile acid deficiency is overcome by induction of oxysterol 7alpha-hydroxylase. J Biol Chem 1996;271(30):18024–31.
Bile Formation and the Enterohepatic Circulation Chapter | 41 951
123.
124.
125.
126.
127. 128.
129.
130.
131.
132.
133.
134.
135.
136.
137.
138.
139.
140.
Post SM, Groenendijk M, Solaas K, Rensen PC, Princen HM. Cholesterol 7alpha-hydroxylase deficiency in mice on an APOE*3Leiden background impairs very-low-density lipoprotein production. Arterioscler Thromb Vasc Biol 2004;24(4):768–74. Schwarz M, Lund EG, Lathe R, Bjorkhem I, Russell DW. Identification and characterization of a mouse oxysterol 7alphahydroxylase cDNA. J Biol Chem 1997;272(38):23995–4001. Setchell KD, Schwarz M, O'Connell NC, Lund EG, Davis DL, Lathe R, et al. Identification of a new inborn error in bile acid synthesis: mutation of the oxysterol 7alpha-hydroxylase gene causes severe neonatal liver disease. J Clin Invest 1998;102(9):1690–703. Ueki I, Kimura A, Nishiyori A, Chen HL, Takei H, Nittono H, et al. Neonatal cholestatic liver disease in an Asian patient with a homozygous mutation in the oxysterol 7alpha-hydroxylase gene. J Pediatr Gastroenterol Nutr 2008;46(4):465–9. Vaz FM, Ferdinandusse S. Bile acid analysis in human disorders of bile acid biosynthesis. Mol Aspects Med 2017;56:10–24. Marin JJ, Macias RI, Briz O, Banales JM, Monte MJ. Bile acids in physiology, pathology and pharmacology. Curr Drug Metab 2015;17(1):4–29. Shefer S, Hauser S, Bekersky I, Mosbach EH. Biochemical site of regulation of bile acid biosynthesis in the rat. J Lipid Res 1970;11(5):404–11. Eriksson S. Biliary excretion of bile acids and cholesterol in bile fistula rats; bile acids and steroids. Proc Soc Exp Biol Med 1957;94(3):578–82. TT L, Makishima M, Repa JJ, Schoonjans K, Kerr TA, Auwerx J, et al. Molecular basis for feedback regulation of bile acid synthesis by nuclear receptors. Mol Cell 2000;6(3):507–15. Goodwin B, Jones SA, Price RR, Watson MA, McKee DD, Moore LB, et al. A regulatory cascade of the nuclear receptors FXR, SHP-1, and LRH-1 represses bile acid biosynthesis. Mol Cell 2000;6(3):517–26. Sinal CJ, Tohkin M, Miyata M, Ward JM, Lambert G, Gonzalez FJ. Targeted disruption of the nuclear receptor FXR/BAR impairs bile acid and lipid homeostasis. Cell 2000;102(6):731–44. Kurosu H, Choi M, Ogawa Y, Dickson AS, Goetz R, Eliseenkova AV, et al. Tissue-specific expression of betaKlotho and fibroblast growth factor (FGF) receptor isoforms determines metabolic activity of FGF19 and FGF21. J Biol Chem 2007;282(37):26687–95. Yu C, Wang F, Kan M, Jin C, Jones RB, Weinstein M, et al. Elevated cholesterol metabolism and bile acid synthesis in mice lacking membrane tyrosine kinase receptor FGFR4. J Biol Chem 2000;275(20):15482–9. Ito S, Fujimori T, Furuya A, Satoh J, Nabeshima Y, Nabeshima Y. Impaired negative feedback suppression of bile acid synthesis in mice lacking betaKlotho. J Clin Invest 2005;115(8):2202–8. Kim I, Ahn SH, Inagaki T, Choi M, Ito S, Guo GL, et al. Differential regulation of bile acid homeostasis by the farnesoid X receptor in liver and intestine. J Lipid Res 2007;48(12):2664–72. Pandak WM, Heuman DM, Hylemon PB, Chiang JY, Vlahcevic ZR. Failure of intravenous infusion of taurocholate to downregulate cholesterol 7 alpha-hydroxylase in rats with biliary fistulas. Gastroenterology 1995;108(2):533–44. Kubaska WM, Gurley EC, Hylemon PB, Guzelian PS, Vlahcevic ZR. Absence of negative feedback control of bile acid biosynthesis in cultured rat hepatocytes. J Biol Chem 1985;260(25):13459–63. Trawick JD, Lewis KD, Dueland S, Moore GL, Simon FR, Davis RA. Rat hepatoma L35 cells, a liver-differentiated cell line,
display resistance to bile acid repression of cholesterol 7 alphahydroxylase. J Lipid Res 1996;37(3):588–98. 141. Pai R, French D, Ma N, Hotzel K, Plise E, Salphati L, et al. Antibody-mediated inhibition of fibroblast growth factor 19 results in increased bile acids synthesis and ileal malabsorption of bile acids in cynomolgus monkeys. Toxicol Sci 2012;126(2):446–56. 142. Luo J, Ko B, Elliott M, Zhou M, Lindhout DA, Phung V, et al. A nontumorigenic variant of FGF19 treats cholestatic liver diseases. Sci Transl Med 2014;6(247):247ra100. 143. Song KH, Li T, Owsley E, Strom S, Chiang JY. Bile acids activate fibroblast growth factor 19 signaling in human hepatocytes to inhibit cholesterol 7alpha-hydroxylase gene expression. Hepatology 2009;49(1):297–305. 144. Chiang JY. Bile acids: regulation of synthesis. J Lipid Res 2009;50(10):1955–66. 145. Li S, Hsu DD, Li B, Luo X, Alderson N, Qiao L, et al. Cytoplasmic tyrosine phosphatase Shp2 coordinates hepatic regulation of bile acid and FGF15/19 signaling to repress bile acid synthesis. Cell Metab 2014;20(2):320–32. 146. Xu Z, Tavares-Sanchez OL, Li Q, Fernando J, Rodriguez CM, Studer EJ, et al. Activation of bile acid biosynthesis by the p38 mitogen-activated protein kinase (MAPK): hepatocyte nuclear factor-4alpha phosphorylation by the p38 MAPK is required for cholesterol 7alpha-hydroxylase expression. J Biol Chem 2007;282(34): 24607–14. 1 47. Yu C, Wang F, Jin C, Huang X, McKeehan WL. Independent repression of bile acid synthesis and activation of c-Jun N-terminal kinase (JNK) by activated hepatocyte fibroblast growth factor receptor 4 (FGFR4) and bile acids. J Biol Chem 2005;280(18):17707–14. 1 48. Lee YK, Schmidt DR, Cummins CL, Choi M, Peng L, Zhang Y, et al. Liver receptor homolog-1 regulates bile acid homeostasis but is not essential for feedback regulation of bile acid synthesis. Mol Endocrinol 2008;22(6):1345–56. 1 49. Pandak WM, Ren S, Marques D, Hall E, Redford K, Mallonee D, et al. Transport of cholesterol into mitochondria is rate-limiting for bile acid synthesis via the alternative pathway in primary rat hepatocytes. J Biol Chem 2002;277(50):48158–64. 1 50. Hall EA, Ren S, Hylemon PB, Rodriguez-Agudo D, Redford K, Marques D, et al. Detection of the steroidogenic acute regulatory protein, StAR, in human liver cells. Biochim Biophys Acta 2005;1733(2-3):111–9. 1 51. Strauss 3rd JF, Kishida T, Christenson LK, Fujimoto T, Hiroi H. START domain proteins and the intracellular trafficking of cholesterol in steroidogenic cells. Mol Cell Endocrinol 2003;202(1-2):59–65. 1 52. Ren S, Hylemon P, Marques D, Hall E, Redford K, Gil G, et al. Effect of increasing the expression of cholesterol transporters (StAR, MLN64, and SCP-2) on bile acid synthesis. J Lipid Res 2004;45(11):2123–31. 1 53. Letourneau D, Lorin A, Lefebvre A, Frappier V, Gaudreault F, Najmanovich R, et al. StAR-related lipid transfer domain protein 5 binds primary bile acids. J Lipid Res 2012;53(12):2677–89. 1 54. Midzak A, Papadopoulos V. Binding domain-driven intracellular trafficking of sterols for synthesis of steroid hormones, bile acids and oxysterols. Traffic 2014;15(9):895–914. 155. Fini A, Roda A. Chemical properties of bile acids. IV. Acidity constants of glycine-conjugated bile acids. J Lipid Res 1987;28(7):755–9. 1 56. Carey MC. Bile acids and bile salts: ionization and solubility properties. Hepatology 1984;4(5 Suppl):66S–71S.
952 SECTION | V Physiology of Secretion
157.
Roda A, Grigolo B, Pellicciari R, Natalini B. Structure-activity relationship studies on natural and synthetic bile acid analogs. Dig Dis Sci 1989;34(12 Suppl):24S–35S. 158. Kamp F, Hamilton JA, Kamp F, Westerhoff HV, Hamilton JA. Movement of fatty acids, fatty acid analogues, and bile acids across phospholipid bilayers. Biochemistry 1993;32(41):11074–86. 159. Carlton VE, Harris BZ, Puffenberger EG, Batta AK, Knisely AS, Robinson DL, et al. Complex inheritance of familial hypercholanemia with associated mutations in TJP2 and BAAT. Nat Genet 2003;34(1):91–6. 160. Morton DH, Salen G, Batta AK, Shefer S, Tint GS, Belchis D, et al. Abnormal hepatic sinusoidal bile acid transport in an Amish kindred is not linked to FIC1 and is improved by ursodiol. Gastroenterology 2000;119(1):188–95. 161. Roda A, Fini A. Effect of nuclear hydroxy substituents on aqueous solubility and acidic strength of bile acids. Hepatology 1984;4(5 Suppl):72S–6S. 162. Hofmann AF, Molino G, Milanese M, Belforte G. Description and simulation of a physiological pharmacokinetic model for the metabolism and enterohepatic circulation of bile acids in man. Cholic acid in healthy man. J Clin Invest 1983;71(4):1003–22. 163. Fujino T, Une M, Imanaka T, Inoue K, Nishimaki-Mogami T. Structure-activity relationship of bile acids and bile acid analogs in regard to FXR activation. J Lipid Res 2004;45(1):132–8. 164. Hamilton JP, Xie G, Raufman JP, Hogan S, Griffin TL, Packard CA, et al. Human cecal bile acids: concentration and spectrum. Am J Physiol Gastrointest Liver Physiol 2007;293(1):G256–63. 165. Jones BV, Begley M, Hill C, Gahan CG, Marchesi JR. Functional and comparative metagenomic analysis of bile salt hydrolase activity in the human gut microbiome. Proc Natl Acad Sci U S A 2008;105(36):13580–5. 166. Ridlon JM, Kang DJ, Hylemon PB. Bile salt biotransformations by human intestinal bacteria. J Lipid Res 2006;47(2):241–59. 167. Hylemon PB, Harder J. Biotransformation of monoterpenes, bile acids, and other isoprenoids in anaerobic ecosystems. FEMS Microbiol Rev 1998;22(5):475–88. 168. Schmassmann A, Angellotti MA, Ton-Nu HT, Schteingart CD, Marcus SN, Rossi SS, et al. Transport, metabolism, and effect of chronic feeding of cholylsarcosine, a conjugated bile acid resistant to deconjugation and dehydroxylation. Gastroenterology 1990;98(1):163–74. 169. Eyssen HJ, De Pauw G, Van Eldere J. Formation of hyodeoxycholic acid from muricholic acid and hyocholic acid by an unidentified gram-positive rod termed HDCA-1 isolated from rat intestinal microflora. Appl Environ Microbiol 1999;65(7):3158–63. 170. Devlin AS, Fischbach MA. A biosynthetic pathway for a prominent class of microbiota-derived bile acids. Nat Chem Biol 2015;11(9):685–90. 171. Hanson RF, Williams G. Metabolism of deoxycholic acid in bile fistula patients. J Lipid Res 1971;12(6):688–91. 172. Hepner GW, Hofmann AF, Thomas PJ. Metabolism of steroid and amino acid moieties of conjugated bile acids in man. II. Glycine-conjugated dihydroxy bile acids. J Clin Invest 1972;51(7):1898–905. 1 73. Subbiah MT, Kuksis A, Mookerjea S. Secretion of bile salts by intact and isolated rat livers. Can J Biochem 1969;47(9):847–54. 174. Wang DQ, Paigen B, Carey MC. Genetic factors at the e nterocyte level account for variations in intestinal cholesterol absorption efficiency among inbred strains of mice. J Lipid Res 2001;42(11):1820–30. 1 75. Schoenfield LJ, Sjovall J. Identification of bile acids and neutral sterols in guinea pig bile. Bile acids and steroids 163. Acta Chem Scand 1966;20(5):1297–303.
176.
Kuroki S, Muramoto S, Kuramoto T, Hoshita T. Sex differences in gallbladder bile acid composition and hepatic steroid 12 alpha-hydroxylase activity in hamsters. J Lipid Res 1983;24(12): 1543–9. 177. Washizu T, Ikenaga H, Washizu M, Ishida T, Tomoda I, Kaneko JJ. Bile acid composition of dog and cat gall-bladder bile. Nippon Juigaku Zasshi 1990;52(2):423–5. 178. Carulli N, Bertolotti M, Carubbi F, Concari M, Martella P, Carulli L, et al. Review article: effect of bile salt pool composition on hepatic and biliary functions. Aliment Pharmacol Ther 2000;14(Suppl 2): 14–8. 1 79. Hagey LR, Schteingart CD, Rossi SS, Ton-Nu HT, Hofmann AF. An N-acyl glycyltaurine conjugate of deoxycholic acid in the biliary bile acids of the rabbit. J Lipid Res 1998;39(11):2119–24. 1 80. Alnouti Y. Bile acid sulfation: a pathway of bile acid elimination and detoxification. Toxicol Sci 2009;108(2):225–46. 1 81. Hofmann AF. Why bile acid glucuronidation is a minor pathway for conjugation of endogenous bile acids in man. Hepatology 2007;45(4):1083–4. author reply 4–5. 1 82. Dawson PA, Setchell KDR. 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 2017;58(6):1033–5. 1 83. Hofmann AF. The enterohepatic circulation of bile acids in mammals: form and functions. Front Biosci 2009;(14)2584–98. 1 84. van Berge Henegouwen GP, Hofmann AF. Nocturnal gallbladder storage and emptying in gallstone patients and healthy subjects. Gastroenterology 1978;75(5):879–85. 1 85. O'Brien JJ, Shaffer EA, Williams Jr. LF, Small DM, Lynn J, Wittenberg JA. Physiological model to study gallbladder function in primates. Gastroenterology 1974;67(1):119–25. 1 86. Choi M, Moschetta A, Bookout AL, Peng L, Umetani M, Holmstrom SR, et al. Identification of a hormonal basis for gallbladder filling. Nat Med 2006;12(11):1253–5. 1 87. Roda E, Aldini R, Mazzella G, Roda A, Sama C, Festi D, et al. Enterohepatic circulation of bile acids after cholecystectomy. Gut 1978;19(7):640–9. 1 88. Kullak-Ublick GA, Paumgartner G, Berr F. Long-term effects of cholecystectomy on bile acid metabolism. Hepatology 1995;21(1):41–5. 1 89. Berr F, Stellaard F, Pratschke E, Paumgartner G. Effects of cholecystectomy on the kinetics of primary and secondary bile acids. J Clin Invest 1989;83(5):1541–50. 1 90. Stiehl A. Bile salt sulphates in cholestasis. Eur J Clin Invest 1974;4(1):59–63. 1 91. Stiehl A, Earnest DL, Admirant WH. Sulfation and renal excretion of bile salts in patients with cirrhosis of the liver. Gastroenterology 1975;68(3):534–44. 1 92. Wilson FA, Burckhardt G, Murer H, Rumrich G, Ullrich KJ. Sodiumcoupled taurocholate transport in the proximal convolution of the rat kidney in vivo and in vitro. J Clin Invest 1981;67(4):1141–50. 1 93. Weiner IM, Glasser JE, Lack L. Renal excretion of bile acids: taurocholic, glycocholic, and colic acids. Am J Physiol 1964;207:964–70. 1 94. Christie DM, Dawson PA, Thevananther S, Shneider BL. Comparative analysis of the ontogeny of a sodium-dependent bile acid transporter in rat kidney and ileum. Am J Physiol 1996;271 (2 Pt 1):G377–85. 1 95. Craddock AL, Love MW, Daniel RW, Kirby LC, Walters HC, Wong MH, et al. Expression and transport properties of the human ileal and renal sodium-dependent bile acid transporter. Am J Physiol 1998;274(1 Pt 1):G157–69.
Bile Formation and the Enterohepatic Circulation Chapter | 41 953
196.
197. 198. 199.
200. 201.
202.
203. 204. 205.
206.
207. 208.
209.
210.
211.
212.
213.
214.
Ballatori N, Christian WV, Lee JY, Dawson PA, Soroka CJ, Boyer JL, et al. OSTalpha-OSTbeta: a major basolateral bile acid and steroid transporter in human intestinal, renal, and biliary epithelia. Hepatology 2005;42(6):1270–9. Trauner M, Boyer JL. Bile salt transporters: molecular characterization, function, and regulation. Physiol Rev 2003;83(2):633–71. Sellinger M, Boyer JL. Physiology of bile secretion and cholestasis. Prog Liver Dis 1990;9:237–59. Gurantz D, Hofmann AF. Influence of bile acid structure on bile flow and biliary lipid secretion in the hamster. Am J Physiol 1984;247(6 Pt 1): G736–48. Nathanson MH, Boyer JL. Mechanisms and regulation of bile secretion. Hepatology 1991;14(3):551–66. Gurantz D, Schteingart CD, Hagey LR, Steinbach JH, Grotmol T, Hofmann AF. Hypercholeresis induced by unconjugated bile acid infusion correlates with recovery in bile of unconjugated bile acids. Hepatology 1991;13(3):540–50. Yoon YB, Hagey LR, Hofmann AF, Gurantz D, Michelotti EL, Steinbach JH. Effect of side-chain shortening on the physiologic properties of bile acids: hepatic transport and effect on biliary secretion of 23-nor-ursodeoxycholate in rodents. Gastroenterology 1986;90(4):837–52. Babu CSR, Sharma M. Biliary tract anatomy and its relationship with venous drainage. J Clin Exp Hepatol 2014;4:S18–26. Hofmann AF. Bile secretion in mice and men. Hepatology 2001;34(4 Pt 1):848–50. Wang R, Salem M, Yousef IM, Tuchweber B, Lam P, Childs SJ, et al. Targeted inactivation of sister of P-glycoprotein gene (spgp) in mice results in nonprogressive but persistent intrahepatic cholestasis. Proc Natl Acad Sci U S A 2001;98(4):2011–6. Strautnieks SS, Bull LN, Knisely AS, Kocoshis SA, Dahl N, Arnell H, et al. A gene encoding a liver-specific ABC transporter is mutated in progressive familial intrahepatic cholestasis. Nat Genet 1998;20(3):233–8. Ballatori N, Truong AT. Glutathione as a primary osmotic driving force in hepatic bile formation. Am J Physiol 1992;263(5 Pt 1):G617–24. Hardison WG, Wood CA. Importance of bicarbonate in bile salt independent fraction of bile flow. Am J Physiol 1978;235(2):E158–64. Ballatori N, Hammond CL, Cunningham JB, Krance SM, Marchan R. Molecular mechanisms of reduced glutathione transport: role of the MRP/CFTR/ABCC and OATP/SLC21A families of membrane proteins. Toxicol Appl Pharmacol 2005;204(3):238–55. Meier PJ, Knickelbein R, Moseley RH, Dobbins JW, Boyer JL. Evidence for carrier-mediated chloride/bicarbonate exchange in canalicular rat liver plasma membrane vesicles. J Clin Invest 1985;75(4):1256–63. Hirata K, Nathanson MH. Bile duct epithelia regulate biliary bicarbonate excretion in normal rat liver. Gastroenterology 2001;121(2):396–406. Kanno N, LeSage G, Glaser S, Alpini G. Regulation of cholangiocyte bicarbonate secretion. Am J Physiol Gastrointest Liver Physiol 2001;281(3):G612–25. Feranchak AP, Fitz JG. Thinking outside the cell: the role of extracellular adenosine triphosphate in bile formation. Gastroenterology 2007;133(5):1726–8. Jones AL, Hradek GT, Renston RH, Wong KY, Karlaganis G, Paumgartner G. Autoradiographic evidence for hepatic lobular concentration gradient of bile acid derivative. Am J Physiol 1980;238(3):G233–7.
215.
216.
217.
218.
219.
220.
221. 222. 223.
224.
225. 226.
227. 228.
229.
230.
231.
232.
233.
Baumgartner U, Baier P, Mappes HJ, Farthmann EH. Pericentral hepatocytes translocate hydrophilic bile acids more rapidly than hydrophobic ones. Dig Dis Sci 2001;46(10):2098–103. Weinman SA, Maglova LM. Free concentrations of intracellular fluorescent anions determined by cytoplasmic dialysis of isolated hepatocytes. Am J Physiol 1994;267(5 Pt 1):G922–31. Yu D, Zhang H, Lionarons DA, Boyer JL, Cai SY. Na+taurocholate cotransporting polypeptide (NTCP/SLC10A1) ortholog in the marine skate Leucoraja erinacea is not a physiological bile salt transporter. Am J Physiol Regul Integr Comp Physiol 2017;312(4):R477–84. Fricker G, Hugentobler G, Meier PJ, Kurz G, Boyer JL. Identification of a single sinusoidal bile salt uptake system in skate liver. Am J Physiol 1987;253(6 Pt 1):G816–22. Kouzuki H, Suzuki H, Ito K, Ohashi R, Sugiyama Y. Contribution of sodium taurocholate co-transporting polypeptide to the uptake of its possible substrates into rat hepatocytes. J Pharmacol Exp Ther 1998;286(2):1043–50. van de Steeg E, Wagenaar E, van der Kruijssen CM, Burggraaff JE, de Waart DR, Elferink RP, et al. Organic anion transporting polypeptide 1a/1b-knockout mice provide insights into hepatic handling of bilirubin, bile acids, and drugs. J Clin Invest 2010;120(8):2942–52. Reichen J, Paumgartner G. Uptake of bile acids by perfused rat liver. Am J Physiol 1976;231(3):734–42. Wolkoff AW. Organic anion uptake by hepatocytes. Compr Physiol 2014;4(4):1715–35. Halilbasic E, Claudel T, Trauner M. Bile acid transporters and regulatory nuclear receptors in the liver and beyond. J Hepatol 2013;58(1):155–68. Stieger B, O'Neill B, Meier PJ. ATP-dependent bile-salt transport in canalicular rat liver plasma-membrane vesicles. Biochem J 1992;284(Pt 1):67–74. Stieger B, Meier Y, Meier PJ. The bile salt export pump. Pflugers Arch 2007;453(5):611–20. Setchell KD, Rodrigues CM, Clerici C, Solinas A, Morelli A, Gartung C, et al. Bile acid concentrations in human and rat liver tissue and in hepatocyte nuclei. Gastroenterology 1997;112(1):226–35. Nies AT, Keppler D. The apical conjugate efflux pump ABCC2 (MRP2). Pflugers Arch 2007;453(5):643–59. Megaraj V, Iida T, Jungsuwadee P, Hofmann AF, Vore M. Hepatobiliary disposition of 3alpha,6alpha,7alpha,12alphatetrahydroxy-cholanoyl taurine: a substrate for multiple canalicular transporters. Drug Metab Dispos 2010;38(10):1723–30. Wang R, Chen HL, Liu L, Sheps JA, Phillips MJ, Ling V. Compensatory role of P-glycoproteins in knockout mice lacking the bile salt export pump. Hepatology 2009;50(3):948–56. Zollner G, Marschall HU, Wagner M, Trauner M. Role of nuclear receptors in the adaptive response to bile acids and cholestasis: pathogenetic and therapeutic considerations. Mol Pharm 2006;3(3):231–51. Belinsky MG, Dawson PA, Shchaveleva I, Bain LJ, Wang R, Ling V, et al. Analysis of the in vivo functions of Mrp3. Mol Pharmacol 2005;68(1):160–8. Mennone A, Soroka CJ, Cai SY, Harry K, Adachi M, Hagey L, et al. Mrp4-/- mice have an impaired cytoprotective response in obstructive cholestasis. Hepatology 2006;43(5):1013–21. Boyer JL, Trauner M, Mennone A, Soroka CJ, Cai SY, Moustafa T, et al. Upregulation of a basolateral FXR-dependent bile acid efflux transporter OSTalpha-OSTbeta in cholestasis in humans and rodents. Am J Physiol Gastrointest Liver Physiol 2006;290(6):G1124–30.
954 SECTION | V Physiology of Secretion
234.
Soroka CJ, Mennone A, Hagey LR, Ballatori N, Boyer JL. Mouse organic solute transporter alpha deficiency enhances renal excretion of bile acids and attenuates cholestasis. Hepatology 2010;51(1):181–90. 235. Soroka CJ, Lee JM, Azzaroli F, Boyer JL. Cellular localization and up-regulation of multidrug resistance-associated protein 3 in hepatocytes and cholangiocytes during obstructive cholestasis in rat liver. Hepatology 2001;33(4):783–91. 236. Rius M, Nies AT, Hummel-Eisenbeiss J, Jedlitschky G, Keppler D. Cotransport of reduced glutathione with bile salts by MRP4 (ABCC4) localized to the basolateral hepatocyte membrane. Hepatology 2003;38(2):374–84. 237. Hirohashi T, Suzuki H, Takikawa H, Sugiyama Y. ATP-dependent transport of bile salts by rat multidrug resistance-associated protein 3 (Mrp3). J Biol Chem 2000;275(4):2905–10. 238. Rius M, Hummel-Eisenbeiss J, Hofmann AF, Keppler D. Substrate specificity of human ABCC4 (MRP4)-mediated cotransport of bile acids and reduced glutathione. Am J Physiol Gastrointest Liver Physiol 2006;290(4):G640–9. 239. Marschall HU, Wagner M, Zollner G, Fickert P, Diczfalusy U, Gumhold J, et al. Complementary stimulation of hepatobiliary transport and detoxification systems by rifampicin and ursodeoxycholic acid in humans. Gastroenterology 2005;129(2):476–85. 240. Zelcer N, van de Wetering K, de Waart R, Scheffer GL, Marschall HU, Wielinga PR, et al. Mice lacking Mrp3 (Abcc3) have normal bile salt transport, but altered hepatic transport of endogenous glucuronides. J Hepatol 2006;44(4):768–75. 241. Geyer J, Wilke T, Petzinger E. The solute carrier family SLC10: more than a family of bile acid transporters regarding function and phylogenetic relationships. Naunyn Schmiedebergs Arch Pharmacol 2006;372(6):413–31. 242. Godoy JR, Fernandes C, Doring B, Beuerlein K, Petzinger E, Geyer J. Molecular and phylogenetic characterization of a novel putative membrane transporter (SLC10A7), conserved in vertebrates and bacteria. Eur J Cell Biol 2007;86(8):445–60. 243. Doring B, Lutteke T, Geyer J, Petzinger E. The SLC10 carrier family: transport functions and molecular structure. Curr Top Membr 2012;70:105–68. 244. Weinman SA. Electrogenicity of Na(+)-coupled bile acid transporters. Yale J Biol Med 1997;70(4):331–40. 245. Kramer W, Stengelin S, Baringhaus KH, Enhsen A, Heuer H, Becker W, et al. Substrate specificity of the ileal and the hepatic Na(+)/bile acid cotransporters of the rabbit. I. Transport studies with membrane vesicles and cell lines expressing the cloned transporters. J Lipid Res 1999;40(9):1604–17. 246. Mita S, Suzuki H, Akita H, Hayashi H, Onuki R, Hofmann AF, et al. Vectorial transport of unconjugated and conjugated bile salts by monolayers of LLC-PK1 cells doubly transfected with human NTCP and BSEP or with rat Ntcp and Bsep. Am J Physiol Gastrointest Liver Physiol 2006;290(3):G550–6. 247. Dawson PA, Lan T, Rao A. Bile acid transporters. J Lipid Res 2009;50(12):2340–57. 248. Vaz FM, Paulusma CC, Huidekoper H, de Ru M, Lim C, Koster J, et al. Sodium taurocholate cotransporting polypeptide (SLC10A1) deficiency: conjugated hypercholanemia without a clear clinical phenotype. Hepatology 2015;61(1):260–7. 249. Slijepcevic D, Roscam Abbing RLP, Katafuchi T, Blank A, Donkers JM, van Hoppe S, et al. Hepatic uptake of conjugated bile acids is mediated by both NTCP and OATPs and modulated by intestinal sensing of plasma bile acid levels in mice. Hepatology 2017;.
250.
Slijepcevic D, Kaufman C, Wichers CG, Gilglioni EH, Lempp FA, Duijst S, et al. Impaired uptake of conjugated bile acids and hepatitis b virus pres1-binding in na(+) -taurocholate cotransporting polypeptide knockout mice. Hepatology 2015;62(1):207–19. 251. Dawson PA. Hepatic bile acid uptake in humans and mice: multiple pathways and expanding potential role for gut-liver signaling. Hepatology 2017;66(5):1384–6. 252. Yan H, Zhong G, Xu G, He W, Jing Z, Gao Z, et al. Sodium taurocholate cotransporting polypeptide is a functional receptor for human hepatitis B and D virus. eLife 2012;1:e00049. 253. Li W, Urban S. Entry of hepatitis B and hepatitis D virus into hepatocytes: basic insights and clinical implications. J Hepatol 2016;64(1 Suppl):S32–40. 254. Van Dyke RW, Stephens JE, Scharschmidt BF. Bile acid transport in cultured rat hepatocytes. Am J Physiol 1982;243(6):G484–92. 255. Mahagita C, Grassl SM, Piyachaturawat P, Ballatori N. Human organic anion transporter 1B1 and 1B3 function as bidirectional carriers and do not mediate GSH-bile acid cotransport. Am J Physiol Gastrointest Liver Physiol 2007;293(1):G271–8. 256. Hagenbuch B, Meier PJ. Organic anion transporting polypeptides of the OATP/ SLC21 family: phylogenetic classification as OATP/ SLCO superfamily, new nomenclature and molecular/functional properties. Pflugers Arch 2004;447(5):653–65. 257. Gui C, Hagenbuch B. Cloning/characterization of the canine organic anion transporting polypeptide 1b4 (Oatp1b4) and classification of the canine OATP/SLCO members. Comp Biochem Physiol C Toxicol Pharmacol 2010;151(3):393–9. 258. Csanaky IL, Lu H, Zhang Y, Ogura K, Choudhuri S, Klaassen CD. Organic anion-transporting polypeptide 1b2 (Oatp1b2) is important for the hepatic uptake of unconjugated bile acids: studies in Oatp1b2-null mice. Hepatology 2011;53(1):272–81. 259. van de Steeg E, Stranecky V, Hartmannova H, Noskova L, Hrebicek M, Wagenaar E, et al. Complete OATP1B1 and OATP1B3 deficiency causes human Rotor syndrome by interrupting conjugated bilirubin reuptake into the liver. J Clin Invest 2012;122(2): 519–28. 260. Strautnieks SS, Byrne JA, Pawlikowska L, Cebecauerova D, Rayner A, Dutton L, et al. Severe bile salt export pump deficiency: 82 different ABCB11 mutations in 109 families. Gastroenterology 2008;134(4):1203–14. 261. Hayashi H, Takada T, Suzuki H, Onuki R, Hofmann AF, Sugiyama Y. Transport by vesicles of glycine- and taurine-conjugated bile salts and taurolithocholate 3-sulfate: a comparison of human BSEP with rat Bsep. Biochim Biophys Acta 2005;1738(1-3):54–62. 262. Byrne JA, Strautnieks SS, Mieli-Vergani G, Higgins CF, Linton KJ, Thompson RJ. The human bile salt export pump: characterization of substrate specificity and identification of inhibitors. Gastroenterology 2002;123(5):1649–58. 263. Aleo MD, Luo Y, Swiss R, Bonin PD, Potter DM, Will Y. Human drug-induced liver injury severity is highly associated with dual inhibition of liver mitochondrial function and bile salt export pump. Hepatology 2014;60(3):1015–22. 264. Yucha RW, He K, Shi Q, Cai L, Nakashita Y, Xia CQ, et al. In vitro drug-induced liver injury prediction: criteria optimization of efflux transporter IC50 and physicochemical properties. Toxicol Sci 2017;157(2):487–99. 265. Wolters H, Elzinga BM, Baller JF, Boverhof R, Schwarz M, Stieger B, et al. Effects of bile salt flux variations on the expression of hepatic bile salt transporters in vivo in mice. J Hepatol 2002;37(5): 556–63.
Bile Formation and the Enterohepatic Circulation Chapter | 41 955
266.
Schaap FG, van der Gaag NA, Gouma DJ, Jansen PL. High expression of the bile salt-homeostatic hormone fibroblast growth factor 19 in the liver of patients with extrahepatic cholestasis. Hepatology 2009;49(4):1228–35. 267. Zollner G, Fickert P, Silbert D, Fuchsbichler A, Marschall HU, Zatloukal K, et al. Adaptive changes in hepatobiliary transporter expression in primary biliary cirrhosis. J Hepatol 2003;38(6): 717–27. 268. Wagner M, Fickert P, Zollner G, Fuchsbichler A, Silbert D, Tsybrovskyy O, et al. Role of farnesoid X receptor in determining hepatic ABC transporter expression and liver injury in bile ductligated mice. Gastroenterology 2003;125(3):825–38. 269. Ananthanarayanan M, Balasubramanian N, Makishima M, Mangelsdorf DJ, Suchy FJ. Human bile salt export pump promoter is transactivated by the farnesoid X receptor/bile acid receptor. J Biol Chem 2001;276(31):28857–65. 270. Plass JR, Mol O, Heegsma J, Geuken M, Faber KN, Jansen PL, et al. Farnesoid X receptor and bile salts are involved in transcriptional regulation of the gene encoding the human bile salt export pump. Hepatology 2002;35(3):589–96. 271. Ananthanarayanan M, Li S, Balasubramaniyan N, Suchy FJ, Walsh MJ. Ligand-dependent activation of the farnesoid X-receptor directs arginine methylation of histone H3 by CARM1. J Biol Chem 2004;279(52):54348–57. 272. Balasubramaniyan N, Luo Y, Sun AQ, Suchy FJ. SUMOylation of the farnesoid X receptor (FXR) regulates the expression of FXR target genes. J Biol Chem 2013;288(19):13850–62. 273. Lew JL, Zhao A, Yu J, Huang L, De Pedro N, Pelaez F, et al. The farnesoid X receptor controls gene expression in a ligand- and promoter-selective fashion. J Biol Chem 2004;279(10):8856–61. 274. Song X, Chen Y, Valanejad L, Kaimal R, Yan B, Stoner M, et al. Mechanistic insights into isoform-dependent and species-specific regulation of bile salt export pump by farnesoid X receptor. J Lipid Res 2013;54(11):3030–44. 275. Schmitt M, Kubitz R, Lizun S, Wettstein M, Haussinger D. Regulation of the dynamic localization of the rat Bsep gene-encoded bile salt export pump by anisoosmolarity. Hepatology 2001;33(3): 509–18. 2 76. Anwer MS. Cellular regulation of hepatic bile acid transport in health and cholestasis. Hepatology 2004;39(3):581–90. 2 77. Halilbasic E, Fiorotto R, Fickert P, Marschall HU, Moustafa T, Spirli C, et al. Side chain structure determines unique physiologic and therapeutic properties of norursodeoxycholic acid in Mdr2-/mice. Hepatology 2009;49(6):1972–81. 2 78. Glaser SS, Alpini G. Activation of the cholehepatic shunt as a potential therapy for primary sclerosing cholangitis. Hepatology 2009;49(6):1795–7. 2 79. Beuers U, Trauner M, Jansen P, Poupon R. New paradigms in the treatment of hepatic cholestasis: from UDCA to FXR, PXR and beyond. J Hepatol 2015;62(1 Suppl):S25–37. 2 80. Hofmann AF. Pharmacology of ursodeoxycholic acid, an enterohepatic drug. Scand J Gastroenterol Suppl 1994;204:1–15. 2 81. Hofmann AF, Zakko SF, Lira M, Clerici C, Hagey LR, Lambert KK, et al. Novel biotransformation and physiological properties of norursodeoxycholic acid in humans. Hepatology 2005;42(6): 1391–8. 2 82. Lazaridis KN, Pham L, Tietz P, Marinelli RA, deGroen PC, Levine S, et al. Rat cholangiocytes absorb bile acids at their apical domain via the ileal sodium-dependent bile acid transporter. J Clin Invest 1997;100(11):2714–21.
283.
Alpini G, Glaser SS, Rodgers R, Phinizy JL, Robertson WE, Lasater J, et al. Functional expression of the apical Na+-dependent bile acid transporter in large but not small rat cholangiocytes. Gastroenterology 1997;113(5):1734–40. 284. Benedetti A, Di Sario A, Marucci L, Svegliati-Baroni G, Schteingart CD, Ton-Nu HT, et al. Carrier-mediated transport of conjugated bile acids across the basolateral membrane of biliary epithelial cells. Am J Physiol 1997;272(6 Pt 1):G1416–24. 285. Debray D, Rainteau D, Barbu V, Rouahi M, El Mourabit H, Lerondel S, et al. Defects in gallbladder emptying and bile acid homeostasis in mice with cystic fibrosis transmembrane conductance regulator deficiencies. Gastroenterology 2012;142(7):1581–91. e6. 286. Alpini G, Glaser S, Baiocchi L, Francis H, Xia X, Lesage G. Secretin activation of the apical Na+-dependent bile acid transporter is associated with cholehepatic shunting in rats. Hepatology 2005;41(5):1037–45. 287. Xia X, Francis H, Glaser S, Alpini G, LeSage G. Bile acid interactions with cholangiocytes. World J Gastroenterol 2006;12(22):3553–63. 288. Boyer JL, Soroka CJ. A cholecystohepatic shunt pathway: does the gallbladder protect the liver? Gastroenterology 2012;142(7):1416–9. 289. Aldini R, Montagnani M, Roda A, Hrelia S, Biagi PL, Roda E. Intestinal absorption of bile acids in the rabbit: different transport rates in jejunum and ileum. Gastroenterology 1996;110(2):459–68. 290. Amelsberg A, Schteingart CD, Ton-Nu HT, Hofmann AF. Carriermediated jejunal absorption of conjugated bile acids in the guinea pig. Gastroenterology 1996;110(4):1098–106. 291. Alrefai WA, Gill RK. Bile acid transporters: structure, function, regulation and pathophysiological implications. Pharm Res 2007;24(10):1803–23. 292. Dietschy JM. Mechanisms for the intestinal absorption of bile acids. J Lipid Res 1968;9(3):297–309. 293. Zhang Y, Klaassen CD. Effects of feeding bile acids and a bile acid sequestrant on hepatic bile acid composition in mice. J Lipid Res 2010;51(11):3230–42. 294. Dawson PA, Haywood J, Craddock AL, Wilson M, Tietjen M, Kluckman K, et al. Targeted deletion of the ileal bile acid transporter eliminates enterohepatic cycling of bile acids in mice. J Biol Chem 2003;278(36):33920–7. 295. Oelkers P, Kirby LC, Heubi JE, Dawson PA. Primary bile acid malabsorption caused by mutations in the ileal sodiumdependent bile acid transporter gene (SLC10A2). J Clin Invest 1997;99(8):1880–7. 2 96. Gui X, Carraway RE. Involvement of mast cells in basal and neurotensin-induced intestinal absorption of taurocholate in rats. Am J Physiol Gastrointest Liver Physiol 2004;287(2):G408–16. 2 97. Gong YZ, Everett ET, Schwartz DA, Norris JS, Wilson FA. Molecular cloning, tissue distribution, and expression of a 14-kDa bile acid-binding protein from rat ileal cytosol. Proc Natl Acad Sci U S A 1994;91(11):4741–5. 2 98. Kramer W, Girbig F, Gutjahr U, Kowalewski S, Jouvenal K, Muller G, et al. Intestinal bile acid absorption. Na(+)-dependent bile acid transport activity in rabbit small intestine correlates with the coexpression of an integral 93-kDa and a peripheral 14-kDa bile acid-binding membrane protein along the duodenum-ileum axis. J Biol Chem 1993;268(24):18035–46. 2 99. Kok T, Hulzebos CV, Wolters H, Havinga R, Agellon LB, Stellaard F, et al. Enterohepatic circulation of bile salts in farnesoid X receptor-deficient mice: efficient intestinal bile salt absorption in the absence of ileal bile acid-binding protein. J Biol Chem 2003;278(43):41930–7.
956 SECTION | V Physiology of Secretion
300.
301.
302. 303.
304.
305.
306.
307.
308.
309.
310.
311.
312.
313.
Turpin ER, Fang HJ, Thomas NR, Hirst JD. Cooperativity and site selectivity in the ileal lipid binding protein. Biochemistry 2013;52(27):4723–33. Horvath G, Bencsura A, Simon A, Tochtrop GP, DeKoster GT, Covey DF, et al. Structural determinants of ligand binding in the ternary complex of human ileal bile acid binding protein with glycocholate and glycochenodeoxycholate obtained from solution NMR. FEBS J 2016;283(3):541–55. Perez MJ, Briz O. Bile-acid-induced cell injury and protection. World J Gastroenterol 2009;15(14):1677–89. Praslickova D, Torchia EC, Sugiyama MG, Magrane EJ, Zwicker BL, Kolodzieyski L, et al. The ileal lipid binding protein is required for efficient absorption and transport of bile acids in the distal portion of the murine small intestine. PLoS ONE 2012;7(12):e50810. Beuling E, Kerkhof IM, Nicksa GA, Giuffrida MJ, Haywood J, aan de Kerk DJ, et al. Conditional Gata4 deletion in mice induces bile acid absorption in the proximal small intestine. Gut 2010;59(7): 888–95. Shneider BL, Dawson PA, Christie DM, Hardikar W, Wong MH, Suchy FJ. Cloning and molecular characterization of the ontogeny of a rat ileal sodium-dependent bile acid transporter. J Clin Invest 1995;95(2):745–54. Cui JY, Aleksunes LM, Tanaka Y, ZD F, Guo Y, Guo GL, et al. Bile acids via FXR initiate the expression of major transporters involved in the enterohepatic circulation of bile acids in newborn mice. Am J Physiol Gastrointest Liver Physiol 2012;302(9):G979–96. Love MW, Craddock AL, Angelin B, Brunzell JD, Duane WC, Dawson PA. Analysis of the ileal bile acid transporter gene, SLC10A2, in subjects with familial hypertriglyceridemia. Arterioscler Thromb Vasc Biol 2001;21(12):2039–45. Duane WC, Hartich LA, Bartman AE, Ho SB. Diminished gene expression of ileal apical sodium bile acid transporter explains impaired absorption of bile acid in patients with hypertriglyceridemia. J Lipid Res 2000;41(9):1384–9. Schiller LR, Hogan RB, Morawski SG, Santa Ana CA, Bern MJ, Norgaard RP, et al. Studies of the prevalence and significance of radiolabeled bile acid malabsorption in a group of patients with idiopathic chronic diarrhea. Gastroenterology 1987;92(1):151–60. Meihoff WE, Kern Jr. F. Bile salt malabsorption in regional ileitis, ileal resection and mannitol-induced diarrhea. J Clin Invest 1968;47(2):261–7. Vitek L, Carey MC. Enterohepatic cycling of bilirubin as a cause of 'black' pigment gallstones in adult life. Eur J Clin Invest 2003;33(9):799–810. Holzer A, Harsch S, Renner O, Strohmeyer A, Schimmel S, Wehkamp J, et al. Diminished expression of apical sodiumdependent bile acid transporter in gallstone disease is independent of ileal inflammation. Digestion 2008;78(1):52–9. Arlow FL, Dekovich AA, Priest RJ, Beher WT. Bile acid-mediated postcholecystectomy diarrhea. Arch Intern Med 1987;147(7): 1327–9.
314.
Camilleri M, Nadeau A, Tremaine WJ, Lamsam J, Burton D, Odunsi S, et al. Measurement of serum 7alpha-hydroxy-4-cholesten3-one (or 7alphaC4), a surrogate test for bile acid malabsorption in health, ileal disease and irritable bowel syndrome using liquid chromatography- tandem mass spectrometry. Neurogastroenterol Motil 2009;21(7). 734–e43. 315. Wang W, Seward DJ, Li L, Boyer JL, Ballatori N. Expression cloning of two genes that together mediate organic solute and steroid transport in the liver of a marine vertebrate. Proc Natl Acad Sci U S A 2001;98(16):9431–6. 316. Seward DJ, Koh AS, Boyer JL, Ballatori N. Functional complementation between a novel mammalian polygenic transport complex and an evolutionarily ancient organic solute transporter, OSTalphaOSTbeta. J Biol Chem 2003;278(30):27473–82. 317. Dawson PA, Hubbert M, Haywood J, Craddock AL, Zerangue N, Christian WV, et al. The heteromeric organic solute transporter alpha-beta, Ostalpha-Ostbeta, is an ileal basolateral bile acid transporter. J Biol Chem 2005;280(8):6960–8. 318. Li N, Cui Z, Fang F, Lee JY, Ballatori N. Heterodimerization, trafficking and membrane topology of the two proteins, Ost alpha and Ost beta, that constitute the organic solute and steroid transporter. Biochem J 2007;407(3):363–72. 319. Christian WV, Li N, Hinkle PM, Ballatori N. beta-Subunit of the Ostalpha-Ostbeta organic solute transporter is required not only for heterodimerization and trafficking but also for function. J Biol Chem 2012;287(25):21233–43. 320. Christian WV, Hinkle PM. Global functions of extracellular, transmembrane and cytoplasmic domains of organic solute transporter beta-subunit. Biochem J 2017;474(12):1981–92. 321. Balesaria S, Pell RJ, Abbott LJ, Tasleem A, Chavele KM, Barley NF, et al. Exploring possible mechanisms for primary bile acid malabsorption: evidence for different regulation of ileal bile acid transporter transcripts in chronic diarrhoea. Eur J Gastroenterol Hepatol 2008;20(5):413–22. 3 22. Landrier JF, Eloranta JJ, Vavricka SR, Kullak-Ublick GA. The nuclear receptor for bile acids, FXR, transactivates human organic solute transporter-alpha and -beta genes. Am J Physiol Gastrointest Liver Physiol 2006;290(3):G476–85. 3 23. Rao A, Haywood J, Craddock AL, Belinsky MG, Kruh GD, Dawson PA. The organic solute transporter alpha-beta, OstalphaOstbeta, is essential for intestinal bile acid transport and homeostasis. Proc Natl Acad Sci U S A 2008;105(10):3891–6. 3 24. Ballatori N, Fang F, Christian WV, Li N, Hammond CL. OstalphaOstbeta is required for bile acid and conjugated steroid disposition in the intestine, kidney, and liver. Am J Physiol Gastrointest Liver Physiol 2008;295(1):G179–86. 3 25. Lan T, Rao A, Haywood J, Kock ND, Dawson PA. Mouse organic solute transporter alpha deficiency alters FGF15 expression and bile acid metabolism. J Hepatol 2012;57(2):359–65. 326. Sultan M, Rao A, Elpeleg O, Vaz FM, Abu-Libdeh B, Karpen SJ, et al. Organic solute transporter-b (SLC51B) deficiency in two brothers with congenital diarrhea and features of cholestasis. Hepatology 2018;67.