CC BY-NC-ND 4.0 · Semin Liver Dis 2023; 43(03): 267-278
DOI: 10.1055/a-2128-5538
Review Article

FXR Friend-ChIPs in the Enterohepatic System

Vik Meadows
1   Department of Pharmacology and Toxicology, Rutgers University, Piscataway, New Jersey
2   Environmental and Occupational Health Science Institute, Rutgers University, Piscataway, New Jersey
,
Zhenning Yang
1   Department of Pharmacology and Toxicology, Rutgers University, Piscataway, New Jersey
2   Environmental and Occupational Health Science Institute, Rutgers University, Piscataway, New Jersey
,
Veronia Basaly
1   Department of Pharmacology and Toxicology, Rutgers University, Piscataway, New Jersey
2   Environmental and Occupational Health Science Institute, Rutgers University, Piscataway, New Jersey
,
Grace L. Guo
1   Department of Pharmacology and Toxicology, Rutgers University, Piscataway, New Jersey
2   Environmental and Occupational Health Science Institute, Rutgers University, Piscataway, New Jersey
3   Department of Veterans Affairs, New Jersey Health Care System, East Orange, New Jersey
› Author Affiliations
Funding This work was supported by the National Institutes of Health (grants: ES007148, ES029258, DK122725, GM135258, AND GM093854), the Department of Veteran Affairs (BX002741), and the Momental Foundation Mistletoe Research Fellowship (FP00032129). The authors would like to thank Rulaiha Elizabeth Taylor, Zakiyah Henry, and Dr. Bo Kong for their support to this work and Dr. Saskia van Mil for her permission to reuse two figure panels of FXR binding motif sequences found in [Fig. 1].
 


Abstract

Chronic liver diseases encompass a wide spectrum of hepatic maladies that often result in cholestasis or altered bile acid secretion and regulation. Incidence and cost of care for many chronic liver diseases are rising in the United States with few Food and Drug Administration-approved drugs available for patient treatment. Farnesoid X receptor (FXR) is the master regulator of bile acid homeostasis with an important role in lipid and glucose metabolism and inflammation. FXR has served as an attractive target for management of cholestasis and fibrosis; however, global FXR agonism results in adverse effects in liver disease patients, severely affecting quality of life. In this review, we highlight seminal studies and recent updates on the FXR proteome and identify gaps in knowledge that are essential for tissue-specific FXR modulation. In conclusion, one of the greatest unmet needs in the field is understanding the underlying mechanism of intestinal versus hepatic FXR function.


#
Lay Summary

Occurrence and treatment cost of chronic liver disease are increasing in the United States with few Food and Drug Administration-approved drugs available for patients. A common symptom of liver disease is reduced or blocked bile flow from the liver, which is regulated by farnesoid X receptor (FXR), a nuclear receptor protein that is important for regulating liver function. FXR must bind other proteins to control bile acid synthesis and bile flow and has unique organ-dependent roles. Understanding how FXR activity is controlled in different organs is an urgent unmet need in liver and intestinal disease research. In this review, we summarize the first findings of FXR-associated proteins and highlight recent studies addressing the knowledge gap for organ-specific FXR research.

Chronic liver disease encompasses a spectrum of liver diseases with cost burden of $81.1 billion for its related care and hospitalizations in the United States.[1] Patients often experience long asymptomatic lapses and are diagnosed at a late stage leading to poor prognosis and high mortality. While great effort has been made in identifying biomarkers and tests to aid in earlier diagnosis,[2] the heterogeneity of chronic liver diseases and subsequent comorbidity complicate this endeavor.

One of the major symptoms of chronic liver disease is cholestasis or impaired bile flow and secretion. Bile is an aqueous heterogenous mixture that contains bile salts, bilirubin phospholipid, cholesterol, amino acids, bicarbonate, vitamins, exogenous drugs, and xenobiotics.[3] Bile salts possess strong detergent properties allowing for fatty acid micelle formation and intestinal absorption.[4] [5] Bile allows for the removal of harmful toxicants and serves as the major route of cholesterol elimination through bile acid formation and secretion.[3] Bile acids are amphipathic sterols and serve as the end-product of cholesterol catabolism, mainly synthesized by hepatocytes, to aid in fat and fat-soluble vitamin absorption. Bile is readily altered by cholangiocytes, bile duct epithelial cells, through secretion of water, bicarbonate, secretin, and other signaling hormones.[6] Cholangiocytes can circumvent normal bile acid circulation through the cholehepatic shunt prior to secretion to the gall bladder for storage or small intestine postprandially.[6] In this process, cholangiocytes transport bile acids from the bile duct lumen to hepatocytes for further modification. However, up to 95% of bile acids are recirculated through enterocyte absorption and secretion into portal circulation, a process called enterohepatic circulation. This circulation and synthesis of bile acids are tightly regulated by a ligand-activated nuclear receptor, FXR, and disruption leads to severe consequences. Much like other nuclear receptors, FXR function relies on interactions with various cofactors and transcriptional regulators. Cofactors are considered promising targets for liver disease therapeutics; however, ubiquitous expression, transient complex formation, and poor antibody performance all pose significant challenges that impede progress in this field of research.

FXR

FXR is a member of the nuclear receptor superfamily and is widely recognized as the master regulator of bile acid synthesis and transport.[7] [8] [9] [10] [11] First discovered as a binding partner for retinoid X receptor (RXR),[12] FXR is highly expressed in the liver and intestine, where it carries out a major role in suppressing bile acid synthesis via downstream effectors, fibroblast growth factor (FGF) 15 (murine ortholog of human FGF19), and to a less extent small heterodimer partner (SHP).[13] [14] Four isoforms of FXR (FXRα1–4) arise in humans and mice with alternative splicing of a 4-amino acid extension of the DNA binding domain, which separates FXRα1 and FXRα3 isoforms from FXRα2 and FXRα4.[15] [16] [17] Human liver preferentially expresses FXRα1 and FXRα2, while mouse liver tissue preferentially expresses FXRα2 or FXRα4, and human and mouse intestines preferentially express FXRα3 and FXRα4.[15] [17] In both human and mouse livers, hepatic FXRα2 is the dominant driver of FXR agonism functions.[15] [18] It is still unclear if FXR is a type I (cytoplasmic) or type II (nuclear) nuclear receptor, but its transcriptional activation has been extensively studied in the liver and intestine.


#

Canonical Function

In the gastrointestinal tract, FXR is highly expressed in the distal ileum and is critical in regulating enterohepatic bile acid homeostasis, including suppressing bile acid synthesis, and promoting bile acid transport. In the intestine, FXR is activated by bile acids to initiate the expression and secretion of FGF15/19 into portal circulation. Intestinal FXR activation regulates enterohepatic bile acid circulation through complex regulation of intestinal bile acid transporters, specifically promoting efflux and inhibiting influx of bile acids. In the ileum, apical sodium bile acid transporter expression is decreased while the expression of fatty acid binding protein 6 and organic solute transporter α and β expression are increased.[19] [20] [21] Further, FXR promotes epithelial layer integrity following activation through increased intestinal tight junction protein expression[22] and mucus production.[23] Moreover, FXR may modulate the ceramide production in the ileum to regulate metabolic diseases.[24] [25] It is important to note that bile acids are metabolized and modified by the intestinal microbiome and there is a mutual relationship between bile acids and microbiome composition.[26] Bacteria create secondary bile acids via deconjugation, dihydroxylation at carbon 7, oxidation, and epimerization of primary bile acids to dampen antimicrobial function, alter intestinal immune microenvironment, and improve bacterial fitness.[26] [27]

In the liver, circulating FGF15/FGF19 binds to hepatic β-klotho and FGF receptor 4 dimer to inhibit gene expression of cytochrome P450 7a1 (Cyp7a1/CYP7A1) and 8b1 (Cyp8b1/CYP8B1), ultimately suppressing bile acid synthesis.[13] [14] [28] [29] Circulating bile acids activate hepatic FXR leading to induction of SHP that mainly functions to inhibit Cyp8b1 expression.[14] [29] [30] [31] [32] Hepatocyte canalicular bile acid efflux transporter, bile salt export pump (BSEP), and sinusoidal uptake transporter, sodium taurocholate co-transporting polypeptide, are both regulated by hepatic FXR activation, serving as the main mechanism for hepatic bile acids to be transported from portal circulation into the bile canaliculi.[14] [33] Hepatic FXR activation also results in reduced fatty acid synthesis[30] [34] [35] and hepatic inflammation.[36] [37] [38] Since CYP7A1 mediates the rate-limiting step of bile acid synthesis, and CYP8B1 determines bile acid hydrophobicity, it is generally considered that intestinal FXR is critical for regulating the bile acid pool and hydrophobicity, while hepatic FXR is critical in determining hydrophobicity of bile acids.[14] [39] [40] There are major differences between the murine and human bile acid speciation, which lends complexity to current studies of bile acid effects in disease states.[41] Overall, humans display a hydrophobic bile acid pool and mice exhibit hydrophilic bile acid pool with unique bile acid species, muricholic acids.[42] CYP2C70 has been identified as the enzyme responsible for α- and β-muricholic acid formation from chenodeoxycholic acid.[41] Murine models of CYP2C70 deficiency demonstrate a humanized bile acid pool with increased hepatic damage that is ameliorated following FXR activation.[41] [43]


#

Noncanonical Function

In recent years, our understanding of the impact of FXR activation has expanded from the enterohepatic system. FXR activation has been found to reduce lung macrophage activation following nitrogen mustard exposure[44] and increase β-oxidative gene expression in cardiomyocytes.[45] In the brain, FXR expression is correlated with Alzheimer's disease and loss of FXR reduces β-amyloid-induced brain injury.[46] FXR increases water reabsorption and promotes renal medullary collecting duct cell survival, ultimately affecting urine concentration during dehydration.[47] Besides, adipose-specific overexpression of FXR promotes brown adipose tissue whitening and fibrosis.[48] There is little information on FXR function in important sensory cells like cholangiocytes, tuft cells in the intestine, or chromaffin cells of the adrenals (PMID: 24068255, PMID: 35245089, PMID: 17963822). The ubiquitous expression of FXR in various organs, while less than in hepatocytes and ileal enterocytes, makes it crucial to understand FXR activation in a whole-body setting (PMID: 36988391).[44] [45] [46] [47] [49]


#

FXR Function in Disease

The role of FXR in intestinal inflammation and fibrosis has been of increasing interest. Whole-body activation of FXR with obeticholic acid (OCA) in mice reduces dextran sodium sulfate (DSS) and trinitrobenzenesulfonic acid-induced colitis including immune cell infiltration and inflammatory cytokine expression.[22] Further, OCA, also known as INT-747, reduces proinflammatory cytokine secretion in activated mononuclear cells and monocytes derived from inflammatory bowel disease patients.[22] Notably, murine models of whole-body FXR loss demonstrate an enhanced inflammatory phenotype following DSS treatment with increased innate lymphoid cell presence within the damaged intestine and increased inflammatory cytokine expression.[50] Similarly, inhibition of ileal FXR by Parabacteroides distasonis improves hepatic fibrosis in mice fed methionine and choline-deficient diet.[51] Prophylactic FXR activation in the intestine, with tissue-specific FXR agonist fexaramine, prevents DSS-induced intestinal villus damage, serum interleukin 17 (IL-17) secretion, and immune cell infiltration of the intestine.[50] Function of fexaramine, and other fex-derivatives, is thought to be gastrointestinal-specific with heterogeneity of FXR activation depending on route of administration.[52] [53] Oral administration of fexaramine is able to activate ileal FXR, with little to no activation in other colon, liver, and kidney, which is likely due to its increased interactions with helix 3 of the FXR protein and deeper penetration and filling of the ligand binding pocket due to fexaramine's hydrophobic rings and larger volume.[52] [53] Fexaramine's intestine-specific activation is likely due to poor absorption into circulation.[53]

While ileal FXR activation is widely regarded to contribute to hepatic function, liver FXR activation may also influence gut permeability. Hepatic FXR loss results in increased colonic mucus secretion and enhanced bacterial response gene expression profile.[23] Further, loss of hepatic FXR shifts the microbiome toward mucosal protection by reducing abundance of mucin-degrading genera (Turicibacter) and increasing abundance of mucus barrier-enforcing bacteria (Roseburia, Bifidobacterium, and Clostridium sensu stricto 1).[23] FXR activation antagonizes nuclear factor kappa B (NF-κB) signaling which results in reduced hepatic inflammation.[37] Mice lacking FXR display increased hepatic inflammation following treatment with lipopolysaccharide, a bacterial cell wall component, which is ameliorated following transfection with FXRα2-adenovirus.[37] FXR activation prevents NF-κB activity through interference of NF-κB and DNA binding.[37]

The effect of FXR activation on hepatic fibrosis is considered disease-dependent.[54] [55] [56] [57] Loss of FXR has been shown to have no effect on hepatic fibrosis in mice following carbon tetrachloride treatment, a classical model of liver injury; however, in common bile duct ligated and 3,5-diethoxycarbonyl-1,4-dihydrocollidine-fed mice, loss of FXR directs protection against portal fibrosis in the liver.[54] FXR expression, as shown by immunohistochemistry, is found mainly in hepatocytes and cholangiocytes and minimally in murine myofibroblasts.[54] Conversely, it has been shown that FXR activation by OCA attenuates collagen deposition, α smooth muscle actin positive staining, and hepatic hydroxyproline content in mice treated with carbon tetrachloride and rats treated with thioacetamide.[55] [56] [57] The protective effects of FXR activation are thought to result from SMAD3 and FXR interaction following FXR activation.[55] Together these data suggest that FXR activation may have indirect effects on fibrosis, and loss of FXR improves portal fibrosis while global FXR activation improves noncholestatic hepatic fibrosis.


#

Known Mechanisms and Interactions of the FXR Proteome

Due to the synergistic roles of bile acids in lipid and glucose homeostasis, FXR regulation of bile acid synthesis and transport, and FXR antagonism effects on inflammation, FXR has been extensively researched as a therapeutic target for chronic liver diseases. This pursuit of global FXR agonists can be controversial in the context of disease treatment, largely due to our knowledge gaps in understanding mechanisms underlying tissue-specific FXR functions.[58] [59] [60]

Originally speculated to be an independent bile acid sensor, the complex role of FXR cofactors in directing tissue-specific FXR response has been of growing interest.[61] [62] [63] FXR can inhibit gene expression of apolipoprotein A-I (ApoA-I) as a monomer or homodimer[64]; however, FXR transcriptional activation is regarded to be a direct result of heterodimerization with other transcriptional regulators like RXR α (RXRα).[65] [66] Interestingly, several factors are now shown to interact with and regulate FXR function. Hereon, we summarize a few of the suspected, and confirmed, members of the FXR proteome in the liver and intestine.


#

FXR Binding Partners

RXRα

RXRα is a nuclear receptor and promiscuous binding partner discovered to be the “missing factor” in various nuclear receptor transcriptional activity.[67] RXR isoforms, α, β, and γ,[68] are activated by 9-cis-retinoic acid and can act as a homodimer to activate the transcription of target genes. Heterodimerization of RXRs with other nuclear receptors can result in nonpermissive or silent partner function, which cannot be activated by RXR agonists, or in a permissive function, responding to ligand activation of either RXR or its partner nuclear receptor.[67] [69] The FXR/RXR complex activated by RXR's endogenous ligands (e.g., 9-cis-retinoic acid) increased FXR-mediated transcriptional activation following activation by synthetic agonists (e.g., WAY-362450), suggesting that RXR activation promotes transcriptional activity of their permissive partners.[65] Like other RXR/nuclear receptor complexes, FXR/RXR heterodimer facilitates transactivation by binding to target sequences with RXR binding to 5′ half-site and its partner binding to the 3′ half-site of target sequences/response elements.[67] Interestingly, it has been found that FXR binding of the SHP promoter requires FXR interaction with the liver receptor homologue 1 (LRH-1) response element without LRH-1 binding; however, 9-cis-retinoic acid-dependent SHP expression requires RXRα occupation of the inverted repeat separated by 1 nucleotide (IR-1) site for subsequent SHP expression.[70] It was originally found that the FXR/RXR heterodimer can bind to the IR-1 sequence with high affinity; however, changes to the half-site sequences, spacing nucleotide, and flanking nucleotides are also bound by this heterodimer, shown pictorially in [Fig. 1].[71] While the FXR/RXR complex upholds many known behaviors of nuclear receptor interactions, their differential expressions may rely on unique site binding and cofactor recruitment at the time of ligand activation.[72]

Zoom Image
Fig. 1 Current understanding of the FXR proteome. The liver in human and mice preferentially expresses FXRα2 to perform ligand-activated transcriptional activity. It is unknown which FXR isoform, FXRα3 or FXRα4, is preferentially expressed in the intestine. All FXR isoforms bind to the IR-1 motif while only FXR2 and FXR4 have been shown to bind the ER-2 DNA binding motifs. Identification of confirmed binding partners of the hepatic and intestinal FXR proteome, and studies on confirmed DNA binding motifs the FXR isoforms, may provide ideal targets for tissue-specific FXR therapeutics. Figure created with biorender.com and confirmed DNA binding motif sequences are repurposed with permission.[15]

#

Hepatocyte Nuclear Factor 4 α

Hepatocyte nuclear factor 4 α (HNF4α) is an orphan nuclear receptor that is highly expressed in epithelial tissues of digestive organs such as the liver and intestine. HNF4α plays essential roles in enterohepatic development, hepatic metabolism, and regulation of hepatocyte cell fate of hepatic progenitor cells.[73] [74] [75] HNF4α is known to interact with other transcription factors to induce transcriptional regulation.[76] HNF4α and FXR share many target genes related to bile acid synthesis, albeit their actions are in an opposing manner, as HNF4α normally promotes, whereas FXR suppresses, the expression of genes in bile acid synthesis.[77] HNF4 regulates bile acid conjugation through expression of bile acid-CoA:amino acid N-acyltransferase and bile acid-CoA ligase.[78] The existence of the FXR/HNF4α complex has been established in mouse[77] and human hepatocytes.[79] Despite these findings, the mechanism or biological significance of the interaction between FXR and HNF4α remains unclear.

HNF4α not only interacts with FXR but also induces FXR gene expression. In the fasting state, peroxisome proliferator-activated receptor-gamma coactivator-1 α (PGC-1α) coactivates HNF4α to induce FXR transcription, favoring isoforms FXRα3 and FXRα4.[80] FXR competitively binds PGC-1α to inhibit transcriptional activation of sulfotransferase family 1E member 1 (Sult1e1) gene by HNF4α.[81] HNF4α and LRH-1 interaction keeps Cyp7a1 gene in a transcriptionally active state, which can be reversed by SHP-FGF15/19-mediated suppression. Specifically, SHP inhibits LRH1 activity to prevent FGF15/19 activation of ERK and JNK pathways that activate Cyp7a1 gene transcription.[14] [82] [83] [84] SHP has been shown to directly interact with HNF4α at the cysteine sulfinic acid decarboxylase promoter to inhibit its transcription and reduce downstream taurine production.[85] Liver zonation transcriptomics data found that Cyp7a1 expression is limited to pericentral hepatocytes, specifically referred to as layer 1, while HNF4α and FXR are equally expressed through all hepatocyte zones, layers 1 to 9.[86] Ubiquitous expression of HNF4α and FXR throughout the hepatocyte zones indicates that their downstream effects may rely on cofactor or ligand binding.


#
#

FXR Cofactors

It has been long suspected that FXR cofactors, unable to bind DNA but able to bind nuclear receptors, influence tissue-specific FXR activation. These regulatory cofactors often function in histone modification or chromatin remodeling capacities, inherently affecting the transcription of FXR target genes through their coactivator or corepressor function. Below we detail key studies that identify and investigate FXR/cofactor complexes in vitro and in vivo.

Cofactors via Posttranslational Modifications of FXR (PMRT, p300, SIRT1, SUMO1, SRC1, O-GlcNac Transferase)

Posttranslational modifications of proteins are important for cell homeostasis, proliferation, and stress response. Posttranslational modifications of FXR direct its function by altering DNA binding, ligand binding, heterodimer formation, and subcellular localization.[87] Protein arginine methyl-transferase type I (PMRT1), p300, and sirtuin 1 (SIRT1) can regulate transcription through methylation, acetylation, and deacetylation, respectively, of histone and nonhistone proteins.[88] [89] [90] Further, SIRT1 can interact with p300 to repress its transcriptional regulatory activity.[91] Small ubiquitin-like modifier (SUMO) proteins direct protein–protein interaction and cellular localization of nuclear receptors.[87] The steroid receptor coactivator 1 (SRC1) initiates p160 SRC family protein recruitment to regulate nuclear receptor function.[92] FXR is subject to methylation, phosphorylation, acetylation, SUMOylation, and O-GlcNAcylation at various sites including lysine 67, 122, and 127 and glutamate 277 and AF1 domain.[88] [93]

Methylation of FXR and FXR target gene histones by PRMT1 is essential for FXR activation.[88] Following treatment with a synthetic bile acid, OCA, FXR recruits PRMT1, which methylates histone H4 protein near promoter regions of BSEP and SHP.[88] This FXR activation is ablated in the presence of methylation inhibitors, indicating that methylation is important in regulating FXR transcriptional activity. Reduced methylation of FXR target promoter regions results in decreased FXR transcriptional activity and subsequently the level of conjugated bile acids in the liver.[94] Conversely, FXR transcriptional regulation of bile acid homeostasis requires phosphorylation by nonreceptor tyrosine kinase, Src, at tyrosine 67.[95] Phospho-defective FXR, or Src downregulation, disrupts the expression of FXR target genes and impairs bile acid homeostasis following cholic acid feeding in wild-type (WT) mice.[95]

Acetylation of FXR and histones at the Shp/SHP promoter initiates SHP gene expression following FXR activation in mouse livers and HepG2 cells.[89] The recruitment of p300 is FXR-dependent as shown in FXR null mice who lack p300 recruitment and its subsequent acetylation at the Shp promoter. Interestingly, acetylation of FXR at lysine 157 and lysine 217 by p300 prevents the FXR/RXRα dimer formation.[66] Mutations at these acetylation sites result in retained RXR binding and ablated p300 acetylation. Further, it has been found that inhibition of p300, in vitro, resulted in increased ApoA-I and reduced G-6-Pase and phosphoenolpyruvate carboxykinase expression, which was unaffected by FXR activation with GW4064.[89] SIRT1 deacetylation of FXR promotes FXR/RXRα dimer formation with increased FXR transactivation.[66] Deletion of intestinal SIRT1 decreases FXR/HNF1α complex formation resulting in reduced bile acid transport and increased hepatic bile acid synthesis.[96]

Hepatic fibrosis resolution remains a key goal in liver disease research. OCA has been shown to be an effective prophylactic treatment against fibrosis.[97] Activated hepatic stellate cells (HSCs) display increased FXR SUMOylation, which renders FXR unable to bind OCA. Prevention of FXR SUMOylation, in combination with OCA treatment, effectively reduces HSC activation and hepatic fibrosis formation in mice.[97] In addition, FXR/SUMO1 complex formation decreases FXR binding and recruitment to the BSEP and SHP promoters in HepG2 cells.[98]

SRC1, along with other co-activators such as PGC-1α, is responsible for hepatocyte differentiation, metabolism, and homeostasis via HNF4α regulation.[99] Due to HNF4α-directed expression of Cyp7a1, it is unsurprising that SRC1 impacts FXR activity. SRC1 interacts with the FXR ligand binding domain following the formation of the FXR/RXR complex.[100]

FXR transcriptional activity is regulated by glucose and O-linked-N-acetylglucosaminylation (O-GlcNAc) of the N-terminus of the AF1 domain.[93] O-GlcNAc transferase regulates FXR activity during fasting and feeding through O-GlcNAcylation at serine 72 in murine FXRα1 and human FXRα3 and serine 62 in human FXRα2.[93] Further, O-GlcNAc transferase can also modify carbohydrate-responsive element-binding protein (ChREBP) to interact with O-GlcNAc–FXR under high glucose concentrations to express glycolysis and lipogenesis genes.[101] However, in the presence of bile acids, regardless of high glucose levels, ChREBP-target gene expression is inhibited.[101] In human hepatocytes, ligand activation of FXR inhibits glucose transcription of ChREBP genes.[79]

Taken together, studies demonstrate that FXR transactivation is not only cofactor-dependent but driven by posttranslational modifications of FXR and target gene environments, including epigenetic modifications. Due to the complex nature of FXR regulation, targeting of individual cofactors will likely need to be disease- or cell-dependent. Understanding of FXR posttranslational modifications, as well as cofactors that induce them, will provide key insights into the regulation of FXR transcriptional activity in a tissue-specific manner.


#

Beta Catenin

Beta catenin (β-catenin) is a well-known and evolutionary conserved protein shown to be important in tight-junction formation, cell proliferation, and is integral to the Wnt signaling cascade.[102] In the liver, β-catenin regulates liver homeostasis, injury repair, and tumorigenesis, and protein expression is mainly found in pericentral hepatocytes.[103] [104] [105] [106] While a relationship has been identified, the molecular mechanism of β-catenin and FXR interactions is undefined. In mouse models of hepatocellular carcinoma (HCC), FXR and β-catenin expression patterns display an inverse relationship.[107] [108] HCC patients display decreased FXR expression,[108] while β-catenin expression increases in HCC patients and human-derived HCC cell lines[109] [110] compared with controls.

In mouse hepatocytes, it is thought that β-catenin sequesters FXR resulting in reduced FXR availability to promote bile acid efflux via regulating bile acid transporter expression and coactivating pregnane X receptor to regulate Cyp3a11 gene expression.[111] GW4064 treatment in β-catenin knockout (KO) mice, subjected to bile duct ligation, demonstrates increased RXRα and FXR binding in hepatocytes.[111] Similarly, GW4064 treatment, in an α-naphthyl isothiocyanate model of biliary injury, shows increased FXR binding to RXRα and reduced β-catenin binding to FXR.[112] Bile duct ligation of transgenic mice overexpressing hepatocyte S45D- β-catenin and low-density lipoprotein receptor 5/6 double KO mice with deficient hepatocyte Wnt signaling demonstrates similar FXR/β-catenin complex levels to WT mice following immunoprecipitation pulldown.[113] Contradictory to these findings, patients with primary sclerosing cholangitis display reduced β-catenin protein expression and mRNA expression of SHP, FXR target gene, and Cyp7a1, SHP target gene.[104] However, there could be other mechanisms to downregulate the expression of these genes, such as inflammation. Taken together, these studies indicate that the FXR/β-catenin complex inhibits hepatocyte FXR function, and due to peri-central protein expression pattern of β-catenin, exploration of interzonal and portal hepatocyte FXR should be further studied.[103] [106] However, the formation of this complex may be transient, depending on injury caused by experimental cholestasis model, hepatocyte zonation, or ligand activation.


#

G Protein Pathway Suppressor 2

G protein pathway suppressor 2 (GPS2) is an epigenetic modifier and is considered one of the core subunits, along with silencing mediator of retinoid and thyroid receptors and nuclear receptor corepressor (NCOR), of the chromatin corepressor complex.[114] [115] The role of GPS2 is critical in regulating transcription, e.g., the regulation of macrophage plasticity is conducted by tightly regulated chromatin remodeling and transcription regulation via the chromatin corepressor complex containing GPS2.[116] In murine models of nonalcoholic steatohepatitis (NASH), GPS2 has been shown to promote steatosis by antagonizing peroxisome proliferator-activated receptor α (PPARα) transcriptional activity with the corepressor, NCOR.[114] It has also been shown that hepatocyte GPS2 is required for hepatitis C virus replication in Huh-7 cell lines.[117]

For bile acid regulation, GPS2 manifested a gene-specific regulation of CYP7A1 and CYP8B1 expression. Where functions to enhance SHP-mediated suppression of CYP7A1 gene transcription, GPS2 can recruit P300/CREB binding protein complex to the HNF4α response element and interact with FXR to form an enhancer/promoter loop for increased expression of CYP8B1 in HepG2 cells.[118] While little is known about the role of hepatic GPS2 in cholestasis, further investigation of the FXR/GPS2 complex may provide insight into its regulation of FXR activity.


#

Glucocorticoid Receptor

Glucocorticoid receptor (GR), a member of the nuclear receptor superfamily, is activated by glucocorticoids in the cytoplasm and translocates to the nucleus to activate various transcriptional pathways.[119] GR activation promotes anti-inflammatory signaling but can lead to cholestasis and insulin resistance.[120] The formation of FXR/GR complex prevents FXR-directed SHP expression through recruitment of C-terminal binding protein to the SHP promoter in HepG2 cells.[120] Hepatic GR activation increases autocrine regulation of Cyp7a1 through FGF21 secretion[121] and activation of FXR increases glucocorticoid secretion in WT mice.[122] More investigation is required to understand the FXR and GR interaction during FXR activation.


#
#

Recent Advances of the FXR Proteome

Utilization of global FXR agonists in primary biliary cholangitis and NASH patients remains controversial due to severe adverse effects such as pruritus, fatigue, and increased serum low-density lipoprotein.[123] [124] [125] In preclinical settings, inhibition of mast cell FXR reduces serum histamine levels and prevents bile duct damage in a murine model of mast cell-induced cholestasis.[58] In a murine model of nonalcoholic fatty liver disease, caffeic acid phenethyl ester treatment reduces steatosis through decreased bacterial bile salt hydrolase activity and increased tauro-β-muricholic acid, an endogenous FXR antagonist.[126] To prevent off-target effects of FXR agonism, the field must turn to understanding the tissue- and cell-specific roles of FXR.

Various research groups have explored the FXR interactome through chromatin immunoprecipitation (ChIP) with a greater focus on hepatic[15] [127] [128] than intestinal FXR.[129] Below we briefly describe seminal studies on the FXR proteome.


#

ChIP Insights

In humans, the dominance of FXR isoforms in the liver affects FXR activation responses.[15] Diseased livers from patients with NASH, cirrhosis, and HCC have increased FXRα1 isoform expression with preferential binding to the IR-1 DNA motif.[15] IR-1 binding by FXRα1 regulates bile acid metabolism/transport and inflammatory signaling. Patients with healthy or steatotic livers express increased FXRα2 with increased binding to everted repeat spaced by 2 nucleotides (ER-2) binding motif, shown pictorially in [Fig. 1].[15] In vitro exploration of HepG2 cells overexpressing FXRα1 or FXRα2 confirms preferential binding to IR-1 or ER-2 regulatory regions, respectively.

In mice, FXR binds IR-1 motifs at intergenic and intron regions, with additional clusters of FXR binding within 1–2 kb of transcription start sites.[128] [129] [130] FXR Re-ChIP analysis demonstrates that FXR/RXR co-occupancy of the SHP promoter is unchanged following FXR activation, despite a marked increase in SHP mRNA expression.[128] In normal and obese mice treated with GW4064, activated FXR represses a large amount of binding motifs identified by ChIP sequencing (ChIP-seq), which challenges previous understanding that SHP represses genes following FXR activation.[128] In vitro, FXR/RXR transcriptional activity increases with LRH-1 transfection and FXR/LRH-1 complex has been detected following co-immunoprecipitation.[130] Based on these findings, FXR transcriptional activation may depend on isoform expression, cofactor interaction, disease setting, and ligand binding.

A recent study of the hepatic FXR proteome demonstrates that cistrome, epigenetic, and protein forces regulate the specific biological pathways studied in various disease models.[127] Based on analysis of publicly available databases, LRH-1, retinoic acid receptor α (RARα), and GA-binding protein (GABPA) interact with FXR to direct its intracellular protein trafficking, protein metabolism, and cell cycle functions.[127] Conversely, Foxa1/2, nuclear factor interleukin 3 (NFIL3), RAR-related orphan receptor α (RORα), GR, NCOR1, and HNF1α interact with FXR to regulate lipid and steroid, amino acid, and carbohydrate metabolism.[127] It is important to recognize that many transcriptional regulators are shared between these two sets of FXR functions. In the WT mouse liver, only complexes with CCAAT/enhancer-binding protein β (CEBP), GATA binding protein 4 (GATA4), HNF1α, GR, and RXRα are confirmed to interact with FXR following rapid immunoprecipitation mass spectrometry of endogenous proteins, also called RIME.[127]

One of the greatest unmet needs in the field is understanding the regulation of intestinal versus hepatic FXR function. Enterohepatic ChIP-seq reveals that only 11% of total FXR DNA binding sites are shared between the liver and intestine accounting for 1,713 genes.[129] Moreover, FXR binds 4,248 unique genes in the liver and 3,406 unique genes in the intestine.[129] The most enriched liver transcription pathways include metabolic and biosynthetic processes while the intestine is enriched for catalytic activity and oxidoreductase activity following FXR activation in WT mice. It has been found that mouse livers contain IR-1 DNA motifs while intestine presents with both IR-1 and ER-2.[129] These results suggest an organ-specific transcriptome is dependent on DNA regulatory element motifs. Further investigation of FXR proteome formation, duration, and ligand dependency, in liver and intestine, will allow researchers to develop targeted therapeutics to enhance specific FXR functions.


#

Discussion

We have outlined the concerted efforts of transcriptional regulators in the diverse functions of FXR activation in the liver and intestine, summarized in [Tables 1] and [2], and highlighted known binding partners of tissue-specific FXR isoforms, summarized in [Fig. 1]. However, to the best of our knowledge, few intestinal FXR proteome studies have been published to date. Increased focus on defining the intestinal FXR proteome may assist in identifying FXR protein complexes for therapeutic functions due to DNA binding heterogeneity and potential unique protein interactions. Similarly, there is little knowledge of the FXR proteome in key bile acid facing cells like cholangiocytes, endothelial cells, and renal cells. Understanding FXR function through its binding partners in these few but impactful cells will help researchers attenuate adverse effects of global FXR agonism. Continued practice of open-access ChIP-seq datasets, as done with the FXR super-signaling atlas that combines multiple single datasets into an interactive platform,[131] can inspire researchers to solve the FXR proteome puzzle.

Table 1

Posttranslational modifications of FXR

Enzyme, modification

Modification target

Function

Reference

PMRT1, methylation

Promoter region

Increases BSEP and SHP mRNA expression

Increases FXR transcriptional activity

Increases conjugated bile acids (liver)

[88] [94]

p300, acetylation

Promoter region,

FXR lysine 157 and lysine 217

Increases SHP expression

Prevents FXR/RXRα dimerization

[66] [89]

SIRT-1, deacetylation

FXR

Increases FXR transcriptional activity

Promotes FXR/RXRα dimerization

[66]

SUMO1, SUMOylation

FXR

Decreases FXR binding to BSEP and SHP promoters

[98]

Src kinase, phosphorylation

FXR tyrosine 67

Increases FXR transcriptional activity

[95]

O-linked-N-acetylglucosamine transferase, O-GlcNAc

FXR serine 62 or 72, isoform-dependent

Increases glycolytic and lipogenic gene expression (in absence of FXR ligands)

[79] [93] [101]

Abbreviation: FXR, farnesoid X receptor.


Table 2

FXR binding partners

Binding partners

Detection method

Function

Reference

RXRα

EMSA, ChIP, co-IP, ALPHA

Increases FXR transcriptional activity

[65] [67] [70]

HNF4α

ChIP-Seq

Unknown biological significance

[77]

β-catenin

ChIP

Inhibits FXR transcription through inhibitory complex formation

[104] [113]

GPS-2

Yeast two-hybrid interaction screening

Increases Cyp7a1 and Cyp8b1 expression

[118]

GR

ChIP, Co-IP

Represses FXR transcriptional activity and reduced hepatic gluconeogenesis

[120] [121]

SRC1

Protein crystallization

SRC1 binds FXR ligand binding domain in FXR/RXRα complex

[100]

Abbreviations: ALPHA, amplified luminescence proximity homogenous assay; ChIP, chromatin immunoprecipitation; ChIP-seq, ChIP sequencing; Co-IP, co-immunoprecipitation; EMSA, electrophoretic mobility shift assay; FXR, farnesoid X receptor.



#

Conclusion

While we believe that deciphering tissue-specific FXR proteomes is the key to understanding the tissue-specific FXR function, the role of chromatin structure, FXR isoform expression, hepatocyte liver zonation, and DNA binding affinity cannot be ignored. The recruitment of FXR activators results in histone modification and chromatin remodeling, beyond the initial euchromatin opening by tissue-specific pioneer factors, to allow the expression of target genes. Moreover, FXR isoform expression and their protein and DNA binding affinity also impact FXR transcriptional activity and ligand activation. Liver zonation may influence FXR function through cofactor expression, ligand secretion, and downstream FXR gene expression. Effort must be made to combine research in chromatin environment, DNA binding motifs, and proteome analysis to push the field of nuclear receptor biology forward.


#
#

Conflict of Interest

None declared.

  • References

  • 1 Hirode G, Saab S, Wong RJ. Trends in the burden of chronic liver disease among hospitalized US adults. JAMA Netw Open 2020; 3 (04) e201997
  • 2 Abeysekera KWM, Macpherson I, Glyn-Owen K. et al. Community pathways for the early detection and risk stratification of chronic liver disease: a narrative systematic review. Lancet Gastroenterol Hepatol 2022; 7 (08) 770-780
  • 3 Boyer JL. Bile formation and secretion. Compr Physiol 2013; 3 (03) 1035-1078
  • 4 Anson ML. The denaturation of proteins by synthetic detergents and bile salts. J Gen Physiol 1939; 23 (02) 239-246
  • 5 Hofmann AF. Micellar solubilization of fatty acids and monoglycerides by bile salt solutions. Nature 1961; 190: 1106-1107
  • 6 Jones H, Alpini G, Francis H. Bile acid signaling and biliary functions. Acta Pharm Sin B 2015; 5 (02) 123-128
  • 7 Chiang JY, Kimmel R, Weinberger C, Stroup D. Farnesoid X receptor responds to bile acids and represses cholesterol 7alpha-hydroxylase gene (CYP7A1) transcription. J Biol Chem 2000; 275 (15) 10918-10924
  • 8 Makishima M, Okamoto AY, Repa JJ. et al. Identification of a nuclear receptor for bile acids. Science 1999; 284 (5418): 1362-1365
  • 9 Parks DJ, Blanchard SG, Bledsoe RK. et al. Bile acids: natural ligands for an orphan nuclear receptor. Science 1999; 284 (5418): 1365-1368
  • 10 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 (05) 543-553
  • 11 Forman BM, Goode E, Chen J. et al. Identification of a nuclear receptor that is activated by farnesol metabolites. Cell 1995; 81 (05) 687-693
  • 12 Seol W, Choi HS, Moore DD. Isolation of proteins that interact specifically with the retinoid X receptor: two novel orphan receptors. Mol Endocrinol 1995; 9 (01) 72-85
  • 13 Inagaki T, Choi M, Moschetta A. et al. Fibroblast growth factor 15 functions as an enterohepatic signal to regulate bile acid homeostasis. Cell Metab 2005; 2 (04) 217-225
  • 14 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 (03) 1034-1043
  • 15 Ramos Pittol JM, Milona A, Morris I. et al. FXR isoforms control different metabolic functions in liver cells via binding to specific DNA motifs. Gastroenterology 2020; 159 (05) 1853.e10-1865.e10
  • 16 Huber RM, Murphy K, Miao B. et al. Generation of multiple farnesoid-X-receptor isoforms through the use of alternative promoters. Gene 2002; 290 (1–2): 35-43
  • 17 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 (01) 104-110
  • 18 Boesjes M, Bloks VW, Hageman J. et al. Hepatic farnesoid X-receptor isoforms α2 and α4 differentially modulate bile salt and lipoprotein metabolism in mice. PLoS One 2014; 9 (12) e115028
  • 19 Chen F, Ma L, Dawson PA. et al. Liver receptor homologue-1 mediates species- and cell line-specific bile acid-dependent negative feedback regulation of the apical sodium-dependent bile acid transporter. J Biol Chem 2003; 278 (22) 19909-19916
  • 20 Dawson PA, Haywood J, Craddock AL. et al. Targeted deletion of the ileal bile acid transporter eliminates enterohepatic cycling of bile acids in mice. J Biol Chem 2003; 278 (36) 33920-33927
  • 21 Grober J, Zaghini I, Fujii H. et al. Identification of a bile acid-responsive element in the human ileal bile acid-binding protein gene. Involvement of the farnesoid X receptor/9-cis-retinoic acid receptor heterodimer. J Biol Chem 1999; 274 (42) 29749-29754
  • 22 Gadaleta RM, van Erpecum KJ, Oldenburg B. et al. Farnesoid X receptor activation inhibits inflammation and preserves the intestinal barrier in inflammatory bowel disease. Gut 2011; 60 (04) 463-472
  • 23 Ijssennagger N, van Rooijen KS, Magnúsdóttir S. et al. Ablation of liver Fxr results in an increased colonic mucus barrier in mice. JHEP Rep 2021; 3 (05) 100344
  • 24 Jiang C, Xie C, Li F. et al. Intestinal farnesoid X receptor signaling promotes nonalcoholic fatty liver disease. J Clin Invest 2015; 125 (01) 386-402
  • 25 Jiang C, Xie C, Lv Y. et al. Intestine-selective farnesoid X receptor inhibition improves obesity-related metabolic dysfunction. Nat Commun 2015; 6: 10166
  • 26 Guzior DV, Quinn RA. Review: microbial transformations of human bile acids. Microbiome 2021; 9 (01) 140
  • 27 Zhang Y, Gao X, Gao S. et al. Effect of gut flora mediated-bile acid metabolism on intestinal immune microenvironment. Immunology 2023; DOI: 10.1111/imm.13672.
  • 28 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 (01) 297-305
  • 29 Kim I, Ahn SH, Inagaki T. et al. Differential regulation of bile acid homeostasis by the farnesoid X receptor in liver and intestine. J Lipid Res 2007; 48 (12) 2664-2672
  • 30 Goodwin B, Jones SA, Price RR. et al. A regulatory cascade of the nuclear receptors FXR, SHP-1, and LRH-1 represses bile acid biosynthesis. Mol Cell 2000; 6 (03) 517-526
  • 31 Lu TT, Makishima M, Repa JJ. et al. Molecular basis for feedback regulation of bile acid synthesis by nuclear receptors. Mol Cell 2000; 6 (03) 507-515
  • 32 Miao J, Choi SE, Seok SM. et al. Ligand-dependent regulation of the activity of the orphan nuclear receptor, small heterodimer partner (SHP), in the repression of bile acid biosynthetic CYP7A1 and CYP8B1 genes. Mol Endocrinol 2011; 25 (07) 1159-1169
  • 33 Rizzo G, Renga B, Mencarelli A, Pellicciari R, Fiorucci S. Role of FXR in regulating bile acid homeostasis and relevance for human diseases. Curr Drug Targets Immune Endocr Metabol Disord 2005; 5 (03) 289-303
  • 34 Clifford BL, Sedgeman LR, Williams KJ. et al. FXR activation protects against NAFLD via bile-acid-dependent reductions in lipid absorption. Cell Metab 2021; 33 (08) 1671.e4-1684.e4
  • 35 Watanabe M, Houten SM, Wang L. et al. Bile acids lower triglyceride levels via a pathway involving FXR, SHP, and SREBP-1c. J Clin Invest 2004; 113 (10) 1408-1418
  • 36 Gai Z, Visentin M, Gui T. et al. Effects of farnesoid X receptor activation on arachidonic acid metabolism, NF-kB signaling, and hepatic inflammation. Mol Pharmacol 2018; 94 (02) 802-811
  • 37 Wang YD, Chen WD, Wang M, Yu D, Forman BM, Huang W. Farnesoid X receptor antagonizes nuclear factor kappaB in hepatic inflammatory response. Hepatology 2008; 48 (05) 1632-1643
  • 38 Xu Z, Huang G, Gong W. et al. FXR ligands protect against hepatocellular inflammation via SOCS3 induction. Cell Signal 2012; 24 (08) 1658-1664
  • 39 Li T, Matozel M, Boehme S. et al. Overexpression of cholesterol 7α-hydroxylase promotes hepatic bile acid synthesis and secretion and maintains cholesterol homeostasis. Hepatology 2011; 53 (03) 996-1006
  • 40 Pandak WM, Bohdan P, Franklund C. et al. Expression of sterol 12alpha-hydroxylase alters bile acid pool composition in primary rat hepatocytes and in vivo. Gastroenterology 2001; 120 (07) 1801-1809
  • 41 de Boer JF, Verkade E, Mulder NL. et al. A human-like bile acid pool induced by deletion of hepatic Cyp2c70 modulates effects of FXR activation in mice. J Lipid Res 2020; 61 (03) 291-305
  • 42 Li J, Dawson PA. Animal models to study bile acid metabolism. Biochim Biophys Acta Mol Basis Dis 2019; 1865 (05) 895-911
  • 43 de Boer JF, de Vries HD, Palmiotti A. et al. Cholangiopathy and biliary fibrosis in Cyp2c70-deficient mice are fully reversed by ursodeoxycholic acid. Cell Mol Gastroenterol Hepatol 2021; 11 (04) 1045-1069
  • 44 Murray A, Banota T, Guo GL. et al. Farnesoid X receptor regulates lung macrophage activation and injury following nitrogen mustard exposure. Toxicol Appl Pharmacol 2022; 454: 116208
  • 45 Guo GL, Santamarina-Fojo S, Akiyama TE. et al. Effects of FXR in foam-cell formation and atherosclerosis development. Biochim Biophys Acta 2006; 1761 (12) 1401-1409
  • 46 Yan N, Yan T, Xia Y, Hao H, Wang G, Gonzalez FJ. The pathophysiological function of non-gastrointestinal farnesoid X receptor. Pharmacol Ther 2021; 226: 107867
  • 47 Guo Y, Xie G, Zhang X. Role of FXR in renal physiology and kidney diseases. Int J Mol Sci 2023; 24 (03) 24
  • 48 Yang J, de Vries HD, Mayeuf-Louchart A. et al. Role of bile acid receptor FXR in development and function of brown adipose tissue. Biochim Biophys Acta Mol Cell Biol Lipids 2023; 1868 (02) 159257
  • 49 Ding L, Yang L, Wang Z, Huang W. Bile acid nuclear receptor FXR and digestive system diseases. Acta Pharm Sin B 2015; 5 (02) 135-144
  • 50 Fu T, Li Y, Oh TG. et al. FXR mediates ILC-intrinsic responses to intestinal inflammation. Proc Natl Acad Sci U S A 2022; 119 (51) e2213041119
  • 51 Zhao Q, Dai MY, Huang RY. et al. Parabacteroides distasonis ameliorates hepatic fibrosis potentially via modulating intestinal bile acid metabolism and hepatocyte pyroptosis in male mice. Nat Commun 2023; 14 (01) 1829
  • 52 Downes M, Verdecia MA, Roecker AJ. et al. A chemical, genetic, and structural analysis of the nuclear bile acid receptor FXR. Mol Cell 2003; 11 (04) 1079-1092
  • 53 Fang S, Suh JM, Reilly SM. et al. Intestinal FXR agonism promotes adipose tissue browning and reduces obesity and insulin resistance. Nat Med 2015; 21 (02) 159-165
  • 54 Fickert P, Fuchsbichler A, Moustafa T. et al. Farnesoid X receptor critically determines the fibrotic response in mice but is expressed to a low extent in human hepatic stellate cells and periductal myofibroblasts. Am J Pathol 2009; 175 (06) 2392-2405
  • 55 Fan YY, Ding W, Zhang C, Fu L, Xu DX, Chen X. Obeticholic acid prevents carbon tetrachloride-induced liver fibrosis through interaction between farnesoid X receptor and Smad3. Int Immunopharmacol 2019; 77: 105911
  • 56 Zhou J, Huang N, Guo Y. et al. Combined obeticholic acid and apoptosis inhibitor treatment alleviates liver fibrosis. Acta Pharm Sin B 2019; 9 (03) 526-536
  • 57 Verbeke L, Mannaerts I, Schierwagen R. et al. FXR agonist obeticholic acid reduces hepatic inflammation and fibrosis in a rat model of toxic cirrhosis. Sci Rep 2016; 6: 33453
  • 58 Meadows V, Kennedy L, Ekser B. et al. Mast cells regulate ductular reaction and intestinal inflammation in cholestasis through farnesoid X receptor signaling. Hepatology 2021; 74 (05) 2684-2698
  • 59 Kjærgaard K, Frisch K, Sørensen M. et al. Obeticholic acid improves hepatic bile acid excretion in patients with primary biliary cholangitis. J Hepatol 2021; 74 (01) 58-65
  • 60 Mudaliar S, Henry RR, Sanyal AJ. et al. Efficacy and safety of the farnesoid X receptor agonist obeticholic acid in patients with type 2 diabetes and nonalcoholic fatty liver disease. Gastroenterology 2013; 145 (03) 574.e1-82.e1
  • 61 Eloranta JJ, Kullak-Ublick GA. Coordinate transcriptional regulation of bile acid homeostasis and drug metabolism. Arch Biochem Biophys 2005; 433 (02) 397-412
  • 62 Nettles KW, Greene GL. Nuclear receptor ligands and cofactor recruitment: is there a coactivator “on deck”?. Mol Cell 2003; 11 (04) 850-851
  • 63 Henry Z, Meadows V, Guo GL. FXR and NASH: an avenue for tissue-specific regulation. Hepatol Commun 2023; 7 (05) 7
  • 64 Claudel T, Sturm E, Duez H. et al. Bile acid-activated nuclear receptor FXR suppresses apolipoprotein A-I transcription via a negative FXR response element. J Clin Invest 2002; 109 (07) 961-971
  • 65 Zheng W, Lu Y, Tian S. et al. Structural insights into the heterodimeric complex of the nuclear receptors FXR and RXR. J Biol Chem 2018; 293 (32) 12535-12541
  • 66 Kemper JK, Xiao Z, Ponugoti B. et al. FXR acetylation is normally dynamically regulated by p300 and SIRT1 but constitutively elevated in metabolic disease states. Cell Metab 2009; 10 (05) 392-404
  • 67 Mangelsdorf DJ, Evans RM. The RXR heterodimers and orphan receptors. Cell 1995; 83 (06) 841-850
  • 68 Wagner CE, Jurutka PW, Marshall PA, Heck MC. Retinoid X receptor selective agonists and their synthetic methods. Curr Top Med Chem 2017; 17 (06) 742-767
  • 69 Shulman AI, Larson C, Mangelsdorf DJ, Ranganathan R. Structural determinants of allosteric ligand activation in RXR heterodimers. Cell 2004; 116 (03) 417-429
  • 70 Hoeke MO, Heegsma J, Hoekstra M, Moshage H, Faber KN. Human FXR regulates SHP expression through direct binding to an LRH-1 binding site, independent of an IR-1 and LRH-1. PLoS One 2014; 9 (02) e88011
  • 71 Laffitte BA, Kast HR, Nguyen CM, Zavacki AM, Moore DD, Edwards PA. Identification of the DNA binding specificity and potential target genes for the farnesoid X-activated receptor. J Biol Chem 2000; 275 (14) 10638-10647
  • 72 Jiang L, Liu X, Liang X. et al. Structural basis of the farnesoid X receptor/retinoid X receptor heterodimer on inverted repeat DNA. Comput Struct Biotechnol J 2023; 21: 3149-3157
  • 73 Thakur A, Wong JCH, Wang EY. et al. Hepatocyte nuclear factor 4-alpha is essential for the active epigenetic state at enhancers in mouse liver. Hepatology 2019; 70 (04) 1360-1376
  • 74 Bookout AL, Jeong Y, Downes M, Yu RT, Evans RM, Mangelsdorf DJ. Anatomical profiling of nuclear receptor expression reveals a hierarchical transcriptional network. Cell 2006; 126 (04) 789-799
  • 75 Yamagata K, Furuta H, Oda N. et al. Mutations in the hepatocyte nuclear factor-4alpha gene in maturity-onset diabetes of the young (MODY1). Nature 1996; 384 (6608): 458-460
  • 76 Yeh MM, Bosch DE, Daoud SS. Role of hepatocyte nuclear factor 4-alpha in gastrointestinal and liver diseases. World J Gastroenterol 2019; 25 (30) 4074-4091
  • 77 Thomas AM, Hart SN, Li G. et al. Hepatocyte nuclear factor 4 alpha and farnesoid X receptor co-regulates gene transcription in mouse livers on a genome-wide scale. Pharm Res 2013; 30 (09) 2188-2198
  • 78 Inoue Y, Yu AM, Inoue J, Gonzalez FJ. Hepatocyte nuclear factor 4alpha is a central regulator of bile acid conjugation. J Biol Chem 2004; 279 (04) 2480-2489
  • 79 Caron S, Huaman Samanez C, Dehondt H. et al. Farnesoid X receptor inhibits the transcriptional activity of carbohydrate response element binding protein in human hepatocytes. Mol Cell Biol 2013; 33 (11) 2202-2211
  • 80 Zhang Y, Castellani LW, Sinal CJ, Gonzalez FJ, Edwards PA. Peroxisome proliferator-activated receptor-gamma coactivator 1alpha (PGC-1alpha) regulates triglyceride metabolism by activation of the nuclear receptor FXR. Genes Dev 2004; 18 (02) 157-169
  • 81 Wang S, Yuan X, Lu D, Guo L, Wu B. Farnesoid X receptor regulates SULT1E1 expression through inhibition of PGC1α binding to HNF4α. Biochem Pharmacol 2017; 145: 202-209
  • 82 Kir S, Zhang Y, Gerard RD, Kliewer SA, Mangelsdorf DJ. Nuclear receptors HNF4α and LRH-1 cooperate in regulating Cyp7a1 in vivo. J Biol Chem 2012; 287 (49) 41334-41341
  • 83 Li T, Jahan A, Chiang JY. Bile acids and cytokines inhibit the human cholesterol 7 alpha-hydroxylase gene via the JNK/c-jun pathway in human liver cells. Hepatology 2006; 43 (06) 1202-1210
  • 84 Gupta S, Stravitz RT, Dent P, Hylemon PB. Down-regulation of cholesterol 7alpha-hydroxylase (CYP7A1) gene expression by bile acids in primary rat hepatocytes is mediated by the c-Jun N-terminal kinase pathway. J Biol Chem 2001; 276 (19) 15816-15822
  • 85 Wang Y, Matye D, Nguyen N, Zhang Y, Li T. HNF4α regulates CSAD to couple hepatic taurine production to bile acid synthesis in mice. Gene Expr 2018; 18 (03) 187-196
  • 86 Halpern KB, Shenhav R, Matcovitch-Natan O. et al. Single-cell spatial reconstruction reveals global division of labour in the mammalian liver. Nature 2017; 542 (7641): 352-356
  • 87 Appelman MD, van der Veen SW, van Mil SWC. Post-translational modifications of FXR; implications for cholestasis and obesity-related disorders. Front Endocrinol (Lausanne) 2021; 12: 729828
  • 88 Rizzo G, Renga B, Antonelli E, Passeri D, Pellicciari R, Fiorucci S. The methyl transferase PRMT1 functions as co-activator of farnesoid X receptor (FXR)/9-cis retinoid X receptor and regulates transcription of FXR responsive genes. Mol Pharmacol 2005; 68 (02) 551-558
  • 89 Fang S, Tsang S, Jones R. et al. The p300 acetylase is critical for ligand-activated farnesoid X receptor (FXR) induction of SHP. J Biol Chem 2008; 283 (50) 35086-35095
  • 90 Rahman S, Islam R. Mammalian Sirt1: insights on its biological functions. Cell Commun Signal 2011; 9: 11
  • 91 Bouras T, Fu M, Sauve AA. et al. SIRT1 deacetylation and repression of p300 involves lysine residues 1020/1024 within the cell cycle regulatory domain 1. J Biol Chem 2005; 280 (11) 10264-10276
  • 92 Walsh CA, Qin L, Tien JC, Young LS, Xu J. The function of steroid receptor coactivator-1 in normal tissues and cancer. Int J Biol Sci 2012; 8 (04) 470-485
  • 93 Berrabah W, Aumercier P, Gheeraert C. et al. Glucose sensing O-GlcNAcylation pathway regulates the nuclear bile acid receptor farnesoid X receptor (FXR). Hepatology 2014; 59 (05) 2022-2033
  • 94 Cabrerizo R, Castaño GO, Burgueño AL. et al. Promoter DNA methylation of farnesoid X receptor and pregnane X receptor modulates the intrahepatic cholestasis of pregnancy phenotype. PLoS One 2014; 9 (01) e87697
  • 95 Byun S, Kim DH, Ryerson D. et al. Postprandial FGF19-induced phosphorylation by Src is critical for FXR function in bile acid homeostasis. Nat Commun 2018; 9 (01) 2590
  • 96 Kazgan N, Metukuri MR, Purushotham A. et al. Intestine-specific deletion of SIRT1 in mice impairs DCoH2-HNF-1α-FXR signaling and alters systemic bile acid homeostasis. Gastroenterology 2014; 146 (04) 1006-1016
  • 97 Zhou J, Cui S, He Q. et al. SUMOylation inhibitors synergize with FXR agonists in combating liver fibrosis. Nat Commun 2020; 11 (01) 240
  • 98 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) 13850-13862
  • 99 Martínez-Jiménez CP, Gómez-Lechón MJ, Castell JV, Jover R. Underexpressed coactivators PGC1alpha and SRC1 impair hepatocyte nuclear factor 4 alpha function and promote dedifferentiation in human hepatoma cells. J Biol Chem 2006; 281 (40) 29840-29849
  • 100 Wang N, Zou Q, Xu J, Zhang J, Liu J. Ligand binding and heterodimerization with retinoid X receptor α (RXRα) induce farnesoid X receptor (FXR) conformational changes affecting coactivator binding. J Biol Chem 2018; 293 (47) 18180-18191
  • 101 Benhamed F, Filhoulaud G, Caron S, Lefebvre P, Staels B, Postic C. O-GlcNAcylation links ChREBP and FXR to glucose-sensing. Front Endocrinol (Lausanne) 2015; 5: 230
  • 102 Valenta T, Hausmann G, Basler K. The many faces and functions of β-catenin. EMBO J 2012; 31 (12) 2714-2736
  • 103 Goel C, Monga SP, Nejak-Bowen K. Role and regulation of Wnt/β-catenin in hepatic perivenous zonation and physiological homeostasis. Am J Pathol 2022; 192 (01) 4-17
  • 104 Ayers M, Liu S, Singhi AD, Kosar K, Cornuet P, Nejak-Bowen K. Changes in beta-catenin expression and activation during progression of primary sclerosing cholangitis predict disease recurrence. Sci Rep 2022; 12 (01) 206
  • 105 Zhang S, Zhang J, Evert K. et al. The hippo effector transcriptional coactivator with PDZ-binding motif cooperates with oncogenic β-catenin to induce hepatoblastoma development in mice and humans. Am J Pathol 2020; 190 (07) 1397-1413
  • 106 Zummo FP, Berthier A, Gheeraert C. et al. A time- and space-resolved nuclear receptor atlas in mouse liver. J Mol Endocrinol 2023; 71 (01) 71
  • 107 Kim I, Morimura K, Shah Y, Yang Q, Ward JM, Gonzalez FJ. Spontaneous hepatocarcinogenesis in farnesoid X receptor-null mice. Carcinogenesis 2007; 28 (05) 940-946
  • 108 Wolfe A, Thomas A, Edwards G, Jaseja R, Guo GL, Apte U. Increased activation of the Wnt/β-catenin pathway in spontaneous hepatocellular carcinoma observed in farnesoid X receptor knockout mice. J Pharmacol Exp Ther 2011; 338 (01) 12-21
  • 109 Xu C, Xu Z, Zhang Y, Evert M, Calvisi DF, Chen X. β-Catenin signaling in hepatocellular carcinoma. J Clin Invest 2022; 132 (04) 132
  • 110 Liu X, Zhang X, Ji L, Gu J, Zhou M, Chen S. Farnesoid X receptor associates with β-catenin and inhibits its activity in hepatocellular carcinoma. Oncotarget 2015; 6 (06) 4226-4238
  • 111 Thompson MD, Moghe A, Cornuet P. et al. β-Catenin regulation of farnesoid X receptor signaling and bile acid metabolism during murine cholestasis. Hepatology 2018; 67 (03) 955-971
  • 112 Liu J, Liu J, Meng C. et al. NRF2 and FXR dual signaling pathways cooperatively regulate the effects of oleanolic acid on cholestatic liver injury. Phytomedicine 2023; 108: 154529
  • 113 Zhang R, Nakao T, Luo J. et al. Activation of WNT/beta-catenin signaling and regulation of the farnesoid X receptor/beta-catenin complex after murine bile duct ligation. Hepatol Commun 2019; 3 (12) 1642-1655
  • 114 Liang N, Damdimopoulos A, Goñi S. et al. Hepatocyte-specific loss of GPS2 in mice reduces non-alcoholic steatohepatitis via activation of PPARα. Nat Commun 2019; 10 (01) 1684
  • 115 Huang Z, Liang N, Goñi S. et al. The corepressors GPS2 and SMRT control enhancer and silencer remodeling via eRNA transcription during inflammatory activation of macrophages. Mol Cell 2021; 81 (05) 953.e9-968.e9
  • 116 Fan R, Toubal A, Goñi S. et al. Loss of the co-repressor GPS2 sensitizes macrophage activation upon metabolic stress induced by obesity and type 2 diabetes. Nat Med 2016; 22 (07) 780-791
  • 117 Xu G, Xin X, Zheng C. GPS2 is required for the association of NS5A with VAP-A and hepatitis C virus replication. PLoS One 2013; 8 (11) e78195
  • 118 Sanyal S, Båvner A, Haroniti A. et al. Involvement of corepressor complex subunit GPS2 in transcriptional pathways governing human bile acid biosynthesis. Proc Natl Acad Sci U S A 2007; 104 (40) 15665-15670
  • 119 Petta I, Dejager L, Ballegeer M. et al. The interactome of the glucocorticoid receptor and its influence on the actions of glucocorticoids in combatting inflammatory and infectious diseases. Microbiol Mol Biol Rev 2016; 80 (02) 495-522
  • 120 Lu Y, Zhang Z, Xiong X. et al. Glucocorticoids promote hepatic cholestasis in mice by inhibiting the transcriptional activity of the farnesoid X receptor. Gastroenterology 2012; 143 (06) 1630-1640.e8
  • 121 Al-Aqil FA, Monte MJ, Peleteiro-Vigil A. et al. Interaction of glucocorticoids with FXR/FGF19/FGF21-mediated ileum-liver crosstalk. Biochim Biophys Acta Mol Basis Dis 2018; 1864 (9, Pt B): 2927-2937
  • 122 Hoekstra M, van der Sluis RJ, Li Z, Oosterveer MH, Groen AK, Van Berkel TJ. FXR agonist GW4064 increases plasma glucocorticoid levels in C57BL/6 mice. Mol Cell Endocrinol 2012; 362 (1–2): 69-75
  • 123 Younossi ZM, Ratziu V, Loomba R. et al; REGENERATE Study Investigators. Obeticholic acid for the treatment of non-alcoholic steatohepatitis: interim analysis from a multicentre, randomised, placebo-controlled phase 3 trial. Lancet 2019; 394 (10215): 2184-2196
  • 124 Nevens F, Andreone P, Mazzella G. et al; POISE Study Group. A placebo-controlled trial of obeticholic acid in primary biliary cholangitis. N Engl J Med 2016; 375 (07) 631-643
  • 125 Xu J, Wang Y, Khoshdeli M. et al. IL-31 levels correlate with pruritus in patients with cholestatic and metabolic liver diseases and is farnesoid X receptor responsive in NASH. Hepatology 2023; 77 (01) 20-32
  • 126 Zhong XC, Liu YM, Gao XX. et al. Caffeic acid phenethyl ester suppresses intestinal FXR signaling and ameliorates nonalcoholic fatty liver disease by inhibiting bacterial bile salt hydrolase activity. Acta Pharmacol Sin 2023; 44 (01) 145-156
  • 127 Dubois-Chevalier J, Dubois V, Dehondt H. et al. The logic of transcriptional regulator recruitment architecture at cis-regulatory modules controlling liver functions. Genome Res 2017; 27 (06) 985-996
  • 128 Lee J, Seok S, Yu P. et al. Genomic analysis of hepatic farnesoid X receptor binding sites reveals altered binding in obesity and direct gene repression by farnesoid X receptor in mice. Hepatology 2012; 56 (01) 108-117
  • 129 Thomas AM, Hart SN, Kong B, Fang J, Zhong XB, Guo GL. Genome-wide tissue-specific farnesoid X receptor binding in mouse liver and intestine. Hepatology 2010; 51 (04) 1410-1419
  • 130 Chong HK, Infante AM, Seo YK. et al. Genome-wide interrogation of hepatic FXR reveals an asymmetric IR-1 motif and synergy with LRH-1. Nucleic Acids Res 2010; 38 (18) 6007-6017
  • 131 Jungwirth E, Panzitt K, Marschall HU, Wagner M, Thallinger GG. A comprehensive FXR signaling atlas derived from pooled ChIP-seq data. Stud Health Technol Inform 2019; 260: 105-112

Address for correspondence

Grace L. Guo, MBBS, PhD
170 Frelinghuysen Road, Room 322, Piscataway, NJ 08854

Publication History

Accepted Manuscript online:
13 July 2023

Article published online:
11 August 2023

© 2023. The Author(s). This is an open access article published by Thieme under the terms of the Creative Commons Attribution-NonDerivative-NonCommercial License, permitting copying and reproduction so long as the original work is given appropriate credit. Contents may not be used for commercial purposes, or adapted, remixed, transformed or built upon. (https://creativecommons.org/licenses/by-nc-nd/4.0/)

Thieme Medical Publishers, Inc.
333 Seventh Avenue, 18th Floor, New York, NY 10001, USA

  • References

  • 1 Hirode G, Saab S, Wong RJ. Trends in the burden of chronic liver disease among hospitalized US adults. JAMA Netw Open 2020; 3 (04) e201997
  • 2 Abeysekera KWM, Macpherson I, Glyn-Owen K. et al. Community pathways for the early detection and risk stratification of chronic liver disease: a narrative systematic review. Lancet Gastroenterol Hepatol 2022; 7 (08) 770-780
  • 3 Boyer JL. Bile formation and secretion. Compr Physiol 2013; 3 (03) 1035-1078
  • 4 Anson ML. The denaturation of proteins by synthetic detergents and bile salts. J Gen Physiol 1939; 23 (02) 239-246
  • 5 Hofmann AF. Micellar solubilization of fatty acids and monoglycerides by bile salt solutions. Nature 1961; 190: 1106-1107
  • 6 Jones H, Alpini G, Francis H. Bile acid signaling and biliary functions. Acta Pharm Sin B 2015; 5 (02) 123-128
  • 7 Chiang JY, Kimmel R, Weinberger C, Stroup D. Farnesoid X receptor responds to bile acids and represses cholesterol 7alpha-hydroxylase gene (CYP7A1) transcription. J Biol Chem 2000; 275 (15) 10918-10924
  • 8 Makishima M, Okamoto AY, Repa JJ. et al. Identification of a nuclear receptor for bile acids. Science 1999; 284 (5418): 1362-1365
  • 9 Parks DJ, Blanchard SG, Bledsoe RK. et al. Bile acids: natural ligands for an orphan nuclear receptor. Science 1999; 284 (5418): 1365-1368
  • 10 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 (05) 543-553
  • 11 Forman BM, Goode E, Chen J. et al. Identification of a nuclear receptor that is activated by farnesol metabolites. Cell 1995; 81 (05) 687-693
  • 12 Seol W, Choi HS, Moore DD. Isolation of proteins that interact specifically with the retinoid X receptor: two novel orphan receptors. Mol Endocrinol 1995; 9 (01) 72-85
  • 13 Inagaki T, Choi M, Moschetta A. et al. Fibroblast growth factor 15 functions as an enterohepatic signal to regulate bile acid homeostasis. Cell Metab 2005; 2 (04) 217-225
  • 14 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 (03) 1034-1043
  • 15 Ramos Pittol JM, Milona A, Morris I. et al. FXR isoforms control different metabolic functions in liver cells via binding to specific DNA motifs. Gastroenterology 2020; 159 (05) 1853.e10-1865.e10
  • 16 Huber RM, Murphy K, Miao B. et al. Generation of multiple farnesoid-X-receptor isoforms through the use of alternative promoters. Gene 2002; 290 (1–2): 35-43
  • 17 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 (01) 104-110
  • 18 Boesjes M, Bloks VW, Hageman J. et al. Hepatic farnesoid X-receptor isoforms α2 and α4 differentially modulate bile salt and lipoprotein metabolism in mice. PLoS One 2014; 9 (12) e115028
  • 19 Chen F, Ma L, Dawson PA. et al. Liver receptor homologue-1 mediates species- and cell line-specific bile acid-dependent negative feedback regulation of the apical sodium-dependent bile acid transporter. J Biol Chem 2003; 278 (22) 19909-19916
  • 20 Dawson PA, Haywood J, Craddock AL. et al. Targeted deletion of the ileal bile acid transporter eliminates enterohepatic cycling of bile acids in mice. J Biol Chem 2003; 278 (36) 33920-33927
  • 21 Grober J, Zaghini I, Fujii H. et al. Identification of a bile acid-responsive element in the human ileal bile acid-binding protein gene. Involvement of the farnesoid X receptor/9-cis-retinoic acid receptor heterodimer. J Biol Chem 1999; 274 (42) 29749-29754
  • 22 Gadaleta RM, van Erpecum KJ, Oldenburg B. et al. Farnesoid X receptor activation inhibits inflammation and preserves the intestinal barrier in inflammatory bowel disease. Gut 2011; 60 (04) 463-472
  • 23 Ijssennagger N, van Rooijen KS, Magnúsdóttir S. et al. Ablation of liver Fxr results in an increased colonic mucus barrier in mice. JHEP Rep 2021; 3 (05) 100344
  • 24 Jiang C, Xie C, Li F. et al. Intestinal farnesoid X receptor signaling promotes nonalcoholic fatty liver disease. J Clin Invest 2015; 125 (01) 386-402
  • 25 Jiang C, Xie C, Lv Y. et al. Intestine-selective farnesoid X receptor inhibition improves obesity-related metabolic dysfunction. Nat Commun 2015; 6: 10166
  • 26 Guzior DV, Quinn RA. Review: microbial transformations of human bile acids. Microbiome 2021; 9 (01) 140
  • 27 Zhang Y, Gao X, Gao S. et al. Effect of gut flora mediated-bile acid metabolism on intestinal immune microenvironment. Immunology 2023; DOI: 10.1111/imm.13672.
  • 28 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 (01) 297-305
  • 29 Kim I, Ahn SH, Inagaki T. et al. Differential regulation of bile acid homeostasis by the farnesoid X receptor in liver and intestine. J Lipid Res 2007; 48 (12) 2664-2672
  • 30 Goodwin B, Jones SA, Price RR. et al. A regulatory cascade of the nuclear receptors FXR, SHP-1, and LRH-1 represses bile acid biosynthesis. Mol Cell 2000; 6 (03) 517-526
  • 31 Lu TT, Makishima M, Repa JJ. et al. Molecular basis for feedback regulation of bile acid synthesis by nuclear receptors. Mol Cell 2000; 6 (03) 507-515
  • 32 Miao J, Choi SE, Seok SM. et al. Ligand-dependent regulation of the activity of the orphan nuclear receptor, small heterodimer partner (SHP), in the repression of bile acid biosynthetic CYP7A1 and CYP8B1 genes. Mol Endocrinol 2011; 25 (07) 1159-1169
  • 33 Rizzo G, Renga B, Mencarelli A, Pellicciari R, Fiorucci S. Role of FXR in regulating bile acid homeostasis and relevance for human diseases. Curr Drug Targets Immune Endocr Metabol Disord 2005; 5 (03) 289-303
  • 34 Clifford BL, Sedgeman LR, Williams KJ. et al. FXR activation protects against NAFLD via bile-acid-dependent reductions in lipid absorption. Cell Metab 2021; 33 (08) 1671.e4-1684.e4
  • 35 Watanabe M, Houten SM, Wang L. et al. Bile acids lower triglyceride levels via a pathway involving FXR, SHP, and SREBP-1c. J Clin Invest 2004; 113 (10) 1408-1418
  • 36 Gai Z, Visentin M, Gui T. et al. Effects of farnesoid X receptor activation on arachidonic acid metabolism, NF-kB signaling, and hepatic inflammation. Mol Pharmacol 2018; 94 (02) 802-811
  • 37 Wang YD, Chen WD, Wang M, Yu D, Forman BM, Huang W. Farnesoid X receptor antagonizes nuclear factor kappaB in hepatic inflammatory response. Hepatology 2008; 48 (05) 1632-1643
  • 38 Xu Z, Huang G, Gong W. et al. FXR ligands protect against hepatocellular inflammation via SOCS3 induction. Cell Signal 2012; 24 (08) 1658-1664
  • 39 Li T, Matozel M, Boehme S. et al. Overexpression of cholesterol 7α-hydroxylase promotes hepatic bile acid synthesis and secretion and maintains cholesterol homeostasis. Hepatology 2011; 53 (03) 996-1006
  • 40 Pandak WM, Bohdan P, Franklund C. et al. Expression of sterol 12alpha-hydroxylase alters bile acid pool composition in primary rat hepatocytes and in vivo. Gastroenterology 2001; 120 (07) 1801-1809
  • 41 de Boer JF, Verkade E, Mulder NL. et al. A human-like bile acid pool induced by deletion of hepatic Cyp2c70 modulates effects of FXR activation in mice. J Lipid Res 2020; 61 (03) 291-305
  • 42 Li J, Dawson PA. Animal models to study bile acid metabolism. Biochim Biophys Acta Mol Basis Dis 2019; 1865 (05) 895-911
  • 43 de Boer JF, de Vries HD, Palmiotti A. et al. Cholangiopathy and biliary fibrosis in Cyp2c70-deficient mice are fully reversed by ursodeoxycholic acid. Cell Mol Gastroenterol Hepatol 2021; 11 (04) 1045-1069
  • 44 Murray A, Banota T, Guo GL. et al. Farnesoid X receptor regulates lung macrophage activation and injury following nitrogen mustard exposure. Toxicol Appl Pharmacol 2022; 454: 116208
  • 45 Guo GL, Santamarina-Fojo S, Akiyama TE. et al. Effects of FXR in foam-cell formation and atherosclerosis development. Biochim Biophys Acta 2006; 1761 (12) 1401-1409
  • 46 Yan N, Yan T, Xia Y, Hao H, Wang G, Gonzalez FJ. The pathophysiological function of non-gastrointestinal farnesoid X receptor. Pharmacol Ther 2021; 226: 107867
  • 47 Guo Y, Xie G, Zhang X. Role of FXR in renal physiology and kidney diseases. Int J Mol Sci 2023; 24 (03) 24
  • 48 Yang J, de Vries HD, Mayeuf-Louchart A. et al. Role of bile acid receptor FXR in development and function of brown adipose tissue. Biochim Biophys Acta Mol Cell Biol Lipids 2023; 1868 (02) 159257
  • 49 Ding L, Yang L, Wang Z, Huang W. Bile acid nuclear receptor FXR and digestive system diseases. Acta Pharm Sin B 2015; 5 (02) 135-144
  • 50 Fu T, Li Y, Oh TG. et al. FXR mediates ILC-intrinsic responses to intestinal inflammation. Proc Natl Acad Sci U S A 2022; 119 (51) e2213041119
  • 51 Zhao Q, Dai MY, Huang RY. et al. Parabacteroides distasonis ameliorates hepatic fibrosis potentially via modulating intestinal bile acid metabolism and hepatocyte pyroptosis in male mice. Nat Commun 2023; 14 (01) 1829
  • 52 Downes M, Verdecia MA, Roecker AJ. et al. A chemical, genetic, and structural analysis of the nuclear bile acid receptor FXR. Mol Cell 2003; 11 (04) 1079-1092
  • 53 Fang S, Suh JM, Reilly SM. et al. Intestinal FXR agonism promotes adipose tissue browning and reduces obesity and insulin resistance. Nat Med 2015; 21 (02) 159-165
  • 54 Fickert P, Fuchsbichler A, Moustafa T. et al. Farnesoid X receptor critically determines the fibrotic response in mice but is expressed to a low extent in human hepatic stellate cells and periductal myofibroblasts. Am J Pathol 2009; 175 (06) 2392-2405
  • 55 Fan YY, Ding W, Zhang C, Fu L, Xu DX, Chen X. Obeticholic acid prevents carbon tetrachloride-induced liver fibrosis through interaction between farnesoid X receptor and Smad3. Int Immunopharmacol 2019; 77: 105911
  • 56 Zhou J, Huang N, Guo Y. et al. Combined obeticholic acid and apoptosis inhibitor treatment alleviates liver fibrosis. Acta Pharm Sin B 2019; 9 (03) 526-536
  • 57 Verbeke L, Mannaerts I, Schierwagen R. et al. FXR agonist obeticholic acid reduces hepatic inflammation and fibrosis in a rat model of toxic cirrhosis. Sci Rep 2016; 6: 33453
  • 58 Meadows V, Kennedy L, Ekser B. et al. Mast cells regulate ductular reaction and intestinal inflammation in cholestasis through farnesoid X receptor signaling. Hepatology 2021; 74 (05) 2684-2698
  • 59 Kjærgaard K, Frisch K, Sørensen M. et al. Obeticholic acid improves hepatic bile acid excretion in patients with primary biliary cholangitis. J Hepatol 2021; 74 (01) 58-65
  • 60 Mudaliar S, Henry RR, Sanyal AJ. et al. Efficacy and safety of the farnesoid X receptor agonist obeticholic acid in patients with type 2 diabetes and nonalcoholic fatty liver disease. Gastroenterology 2013; 145 (03) 574.e1-82.e1
  • 61 Eloranta JJ, Kullak-Ublick GA. Coordinate transcriptional regulation of bile acid homeostasis and drug metabolism. Arch Biochem Biophys 2005; 433 (02) 397-412
  • 62 Nettles KW, Greene GL. Nuclear receptor ligands and cofactor recruitment: is there a coactivator “on deck”?. Mol Cell 2003; 11 (04) 850-851
  • 63 Henry Z, Meadows V, Guo GL. FXR and NASH: an avenue for tissue-specific regulation. Hepatol Commun 2023; 7 (05) 7
  • 64 Claudel T, Sturm E, Duez H. et al. Bile acid-activated nuclear receptor FXR suppresses apolipoprotein A-I transcription via a negative FXR response element. J Clin Invest 2002; 109 (07) 961-971
  • 65 Zheng W, Lu Y, Tian S. et al. Structural insights into the heterodimeric complex of the nuclear receptors FXR and RXR. J Biol Chem 2018; 293 (32) 12535-12541
  • 66 Kemper JK, Xiao Z, Ponugoti B. et al. FXR acetylation is normally dynamically regulated by p300 and SIRT1 but constitutively elevated in metabolic disease states. Cell Metab 2009; 10 (05) 392-404
  • 67 Mangelsdorf DJ, Evans RM. The RXR heterodimers and orphan receptors. Cell 1995; 83 (06) 841-850
  • 68 Wagner CE, Jurutka PW, Marshall PA, Heck MC. Retinoid X receptor selective agonists and their synthetic methods. Curr Top Med Chem 2017; 17 (06) 742-767
  • 69 Shulman AI, Larson C, Mangelsdorf DJ, Ranganathan R. Structural determinants of allosteric ligand activation in RXR heterodimers. Cell 2004; 116 (03) 417-429
  • 70 Hoeke MO, Heegsma J, Hoekstra M, Moshage H, Faber KN. Human FXR regulates SHP expression through direct binding to an LRH-1 binding site, independent of an IR-1 and LRH-1. PLoS One 2014; 9 (02) e88011
  • 71 Laffitte BA, Kast HR, Nguyen CM, Zavacki AM, Moore DD, Edwards PA. Identification of the DNA binding specificity and potential target genes for the farnesoid X-activated receptor. J Biol Chem 2000; 275 (14) 10638-10647
  • 72 Jiang L, Liu X, Liang X. et al. Structural basis of the farnesoid X receptor/retinoid X receptor heterodimer on inverted repeat DNA. Comput Struct Biotechnol J 2023; 21: 3149-3157
  • 73 Thakur A, Wong JCH, Wang EY. et al. Hepatocyte nuclear factor 4-alpha is essential for the active epigenetic state at enhancers in mouse liver. Hepatology 2019; 70 (04) 1360-1376
  • 74 Bookout AL, Jeong Y, Downes M, Yu RT, Evans RM, Mangelsdorf DJ. Anatomical profiling of nuclear receptor expression reveals a hierarchical transcriptional network. Cell 2006; 126 (04) 789-799
  • 75 Yamagata K, Furuta H, Oda N. et al. Mutations in the hepatocyte nuclear factor-4alpha gene in maturity-onset diabetes of the young (MODY1). Nature 1996; 384 (6608): 458-460
  • 76 Yeh MM, Bosch DE, Daoud SS. Role of hepatocyte nuclear factor 4-alpha in gastrointestinal and liver diseases. World J Gastroenterol 2019; 25 (30) 4074-4091
  • 77 Thomas AM, Hart SN, Li G. et al. Hepatocyte nuclear factor 4 alpha and farnesoid X receptor co-regulates gene transcription in mouse livers on a genome-wide scale. Pharm Res 2013; 30 (09) 2188-2198
  • 78 Inoue Y, Yu AM, Inoue J, Gonzalez FJ. Hepatocyte nuclear factor 4alpha is a central regulator of bile acid conjugation. J Biol Chem 2004; 279 (04) 2480-2489
  • 79 Caron S, Huaman Samanez C, Dehondt H. et al. Farnesoid X receptor inhibits the transcriptional activity of carbohydrate response element binding protein in human hepatocytes. Mol Cell Biol 2013; 33 (11) 2202-2211
  • 80 Zhang Y, Castellani LW, Sinal CJ, Gonzalez FJ, Edwards PA. Peroxisome proliferator-activated receptor-gamma coactivator 1alpha (PGC-1alpha) regulates triglyceride metabolism by activation of the nuclear receptor FXR. Genes Dev 2004; 18 (02) 157-169
  • 81 Wang S, Yuan X, Lu D, Guo L, Wu B. Farnesoid X receptor regulates SULT1E1 expression through inhibition of PGC1α binding to HNF4α. Biochem Pharmacol 2017; 145: 202-209
  • 82 Kir S, Zhang Y, Gerard RD, Kliewer SA, Mangelsdorf DJ. Nuclear receptors HNF4α and LRH-1 cooperate in regulating Cyp7a1 in vivo. J Biol Chem 2012; 287 (49) 41334-41341
  • 83 Li T, Jahan A, Chiang JY. Bile acids and cytokines inhibit the human cholesterol 7 alpha-hydroxylase gene via the JNK/c-jun pathway in human liver cells. Hepatology 2006; 43 (06) 1202-1210
  • 84 Gupta S, Stravitz RT, Dent P, Hylemon PB. Down-regulation of cholesterol 7alpha-hydroxylase (CYP7A1) gene expression by bile acids in primary rat hepatocytes is mediated by the c-Jun N-terminal kinase pathway. J Biol Chem 2001; 276 (19) 15816-15822
  • 85 Wang Y, Matye D, Nguyen N, Zhang Y, Li T. HNF4α regulates CSAD to couple hepatic taurine production to bile acid synthesis in mice. Gene Expr 2018; 18 (03) 187-196
  • 86 Halpern KB, Shenhav R, Matcovitch-Natan O. et al. Single-cell spatial reconstruction reveals global division of labour in the mammalian liver. Nature 2017; 542 (7641): 352-356
  • 87 Appelman MD, van der Veen SW, van Mil SWC. Post-translational modifications of FXR; implications for cholestasis and obesity-related disorders. Front Endocrinol (Lausanne) 2021; 12: 729828
  • 88 Rizzo G, Renga B, Antonelli E, Passeri D, Pellicciari R, Fiorucci S. The methyl transferase PRMT1 functions as co-activator of farnesoid X receptor (FXR)/9-cis retinoid X receptor and regulates transcription of FXR responsive genes. Mol Pharmacol 2005; 68 (02) 551-558
  • 89 Fang S, Tsang S, Jones R. et al. The p300 acetylase is critical for ligand-activated farnesoid X receptor (FXR) induction of SHP. J Biol Chem 2008; 283 (50) 35086-35095
  • 90 Rahman S, Islam R. Mammalian Sirt1: insights on its biological functions. Cell Commun Signal 2011; 9: 11
  • 91 Bouras T, Fu M, Sauve AA. et al. SIRT1 deacetylation and repression of p300 involves lysine residues 1020/1024 within the cell cycle regulatory domain 1. J Biol Chem 2005; 280 (11) 10264-10276
  • 92 Walsh CA, Qin L, Tien JC, Young LS, Xu J. The function of steroid receptor coactivator-1 in normal tissues and cancer. Int J Biol Sci 2012; 8 (04) 470-485
  • 93 Berrabah W, Aumercier P, Gheeraert C. et al. Glucose sensing O-GlcNAcylation pathway regulates the nuclear bile acid receptor farnesoid X receptor (FXR). Hepatology 2014; 59 (05) 2022-2033
  • 94 Cabrerizo R, Castaño GO, Burgueño AL. et al. Promoter DNA methylation of farnesoid X receptor and pregnane X receptor modulates the intrahepatic cholestasis of pregnancy phenotype. PLoS One 2014; 9 (01) e87697
  • 95 Byun S, Kim DH, Ryerson D. et al. Postprandial FGF19-induced phosphorylation by Src is critical for FXR function in bile acid homeostasis. Nat Commun 2018; 9 (01) 2590
  • 96 Kazgan N, Metukuri MR, Purushotham A. et al. Intestine-specific deletion of SIRT1 in mice impairs DCoH2-HNF-1α-FXR signaling and alters systemic bile acid homeostasis. Gastroenterology 2014; 146 (04) 1006-1016
  • 97 Zhou J, Cui S, He Q. et al. SUMOylation inhibitors synergize with FXR agonists in combating liver fibrosis. Nat Commun 2020; 11 (01) 240
  • 98 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) 13850-13862
  • 99 Martínez-Jiménez CP, Gómez-Lechón MJ, Castell JV, Jover R. Underexpressed coactivators PGC1alpha and SRC1 impair hepatocyte nuclear factor 4 alpha function and promote dedifferentiation in human hepatoma cells. J Biol Chem 2006; 281 (40) 29840-29849
  • 100 Wang N, Zou Q, Xu J, Zhang J, Liu J. Ligand binding and heterodimerization with retinoid X receptor α (RXRα) induce farnesoid X receptor (FXR) conformational changes affecting coactivator binding. J Biol Chem 2018; 293 (47) 18180-18191
  • 101 Benhamed F, Filhoulaud G, Caron S, Lefebvre P, Staels B, Postic C. O-GlcNAcylation links ChREBP and FXR to glucose-sensing. Front Endocrinol (Lausanne) 2015; 5: 230
  • 102 Valenta T, Hausmann G, Basler K. The many faces and functions of β-catenin. EMBO J 2012; 31 (12) 2714-2736
  • 103 Goel C, Monga SP, Nejak-Bowen K. Role and regulation of Wnt/β-catenin in hepatic perivenous zonation and physiological homeostasis. Am J Pathol 2022; 192 (01) 4-17
  • 104 Ayers M, Liu S, Singhi AD, Kosar K, Cornuet P, Nejak-Bowen K. Changes in beta-catenin expression and activation during progression of primary sclerosing cholangitis predict disease recurrence. Sci Rep 2022; 12 (01) 206
  • 105 Zhang S, Zhang J, Evert K. et al. The hippo effector transcriptional coactivator with PDZ-binding motif cooperates with oncogenic β-catenin to induce hepatoblastoma development in mice and humans. Am J Pathol 2020; 190 (07) 1397-1413
  • 106 Zummo FP, Berthier A, Gheeraert C. et al. A time- and space-resolved nuclear receptor atlas in mouse liver. J Mol Endocrinol 2023; 71 (01) 71
  • 107 Kim I, Morimura K, Shah Y, Yang Q, Ward JM, Gonzalez FJ. Spontaneous hepatocarcinogenesis in farnesoid X receptor-null mice. Carcinogenesis 2007; 28 (05) 940-946
  • 108 Wolfe A, Thomas A, Edwards G, Jaseja R, Guo GL, Apte U. Increased activation of the Wnt/β-catenin pathway in spontaneous hepatocellular carcinoma observed in farnesoid X receptor knockout mice. J Pharmacol Exp Ther 2011; 338 (01) 12-21
  • 109 Xu C, Xu Z, Zhang Y, Evert M, Calvisi DF, Chen X. β-Catenin signaling in hepatocellular carcinoma. J Clin Invest 2022; 132 (04) 132
  • 110 Liu X, Zhang X, Ji L, Gu J, Zhou M, Chen S. Farnesoid X receptor associates with β-catenin and inhibits its activity in hepatocellular carcinoma. Oncotarget 2015; 6 (06) 4226-4238
  • 111 Thompson MD, Moghe A, Cornuet P. et al. β-Catenin regulation of farnesoid X receptor signaling and bile acid metabolism during murine cholestasis. Hepatology 2018; 67 (03) 955-971
  • 112 Liu J, Liu J, Meng C. et al. NRF2 and FXR dual signaling pathways cooperatively regulate the effects of oleanolic acid on cholestatic liver injury. Phytomedicine 2023; 108: 154529
  • 113 Zhang R, Nakao T, Luo J. et al. Activation of WNT/beta-catenin signaling and regulation of the farnesoid X receptor/beta-catenin complex after murine bile duct ligation. Hepatol Commun 2019; 3 (12) 1642-1655
  • 114 Liang N, Damdimopoulos A, Goñi S. et al. Hepatocyte-specific loss of GPS2 in mice reduces non-alcoholic steatohepatitis via activation of PPARα. Nat Commun 2019; 10 (01) 1684
  • 115 Huang Z, Liang N, Goñi S. et al. The corepressors GPS2 and SMRT control enhancer and silencer remodeling via eRNA transcription during inflammatory activation of macrophages. Mol Cell 2021; 81 (05) 953.e9-968.e9
  • 116 Fan R, Toubal A, Goñi S. et al. Loss of the co-repressor GPS2 sensitizes macrophage activation upon metabolic stress induced by obesity and type 2 diabetes. Nat Med 2016; 22 (07) 780-791
  • 117 Xu G, Xin X, Zheng C. GPS2 is required for the association of NS5A with VAP-A and hepatitis C virus replication. PLoS One 2013; 8 (11) e78195
  • 118 Sanyal S, Båvner A, Haroniti A. et al. Involvement of corepressor complex subunit GPS2 in transcriptional pathways governing human bile acid biosynthesis. Proc Natl Acad Sci U S A 2007; 104 (40) 15665-15670
  • 119 Petta I, Dejager L, Ballegeer M. et al. The interactome of the glucocorticoid receptor and its influence on the actions of glucocorticoids in combatting inflammatory and infectious diseases. Microbiol Mol Biol Rev 2016; 80 (02) 495-522
  • 120 Lu Y, Zhang Z, Xiong X. et al. Glucocorticoids promote hepatic cholestasis in mice by inhibiting the transcriptional activity of the farnesoid X receptor. Gastroenterology 2012; 143 (06) 1630-1640.e8
  • 121 Al-Aqil FA, Monte MJ, Peleteiro-Vigil A. et al. Interaction of glucocorticoids with FXR/FGF19/FGF21-mediated ileum-liver crosstalk. Biochim Biophys Acta Mol Basis Dis 2018; 1864 (9, Pt B): 2927-2937
  • 122 Hoekstra M, van der Sluis RJ, Li Z, Oosterveer MH, Groen AK, Van Berkel TJ. FXR agonist GW4064 increases plasma glucocorticoid levels in C57BL/6 mice. Mol Cell Endocrinol 2012; 362 (1–2): 69-75
  • 123 Younossi ZM, Ratziu V, Loomba R. et al; REGENERATE Study Investigators. Obeticholic acid for the treatment of non-alcoholic steatohepatitis: interim analysis from a multicentre, randomised, placebo-controlled phase 3 trial. Lancet 2019; 394 (10215): 2184-2196
  • 124 Nevens F, Andreone P, Mazzella G. et al; POISE Study Group. A placebo-controlled trial of obeticholic acid in primary biliary cholangitis. N Engl J Med 2016; 375 (07) 631-643
  • 125 Xu J, Wang Y, Khoshdeli M. et al. IL-31 levels correlate with pruritus in patients with cholestatic and metabolic liver diseases and is farnesoid X receptor responsive in NASH. Hepatology 2023; 77 (01) 20-32
  • 126 Zhong XC, Liu YM, Gao XX. et al. Caffeic acid phenethyl ester suppresses intestinal FXR signaling and ameliorates nonalcoholic fatty liver disease by inhibiting bacterial bile salt hydrolase activity. Acta Pharmacol Sin 2023; 44 (01) 145-156
  • 127 Dubois-Chevalier J, Dubois V, Dehondt H. et al. The logic of transcriptional regulator recruitment architecture at cis-regulatory modules controlling liver functions. Genome Res 2017; 27 (06) 985-996
  • 128 Lee J, Seok S, Yu P. et al. Genomic analysis of hepatic farnesoid X receptor binding sites reveals altered binding in obesity and direct gene repression by farnesoid X receptor in mice. Hepatology 2012; 56 (01) 108-117
  • 129 Thomas AM, Hart SN, Kong B, Fang J, Zhong XB, Guo GL. Genome-wide tissue-specific farnesoid X receptor binding in mouse liver and intestine. Hepatology 2010; 51 (04) 1410-1419
  • 130 Chong HK, Infante AM, Seo YK. et al. Genome-wide interrogation of hepatic FXR reveals an asymmetric IR-1 motif and synergy with LRH-1. Nucleic Acids Res 2010; 38 (18) 6007-6017
  • 131 Jungwirth E, Panzitt K, Marschall HU, Wagner M, Thallinger GG. A comprehensive FXR signaling atlas derived from pooled ChIP-seq data. Stud Health Technol Inform 2019; 260: 105-112

Zoom Image
Fig. 1 Current understanding of the FXR proteome. The liver in human and mice preferentially expresses FXRα2 to perform ligand-activated transcriptional activity. It is unknown which FXR isoform, FXRα3 or FXRα4, is preferentially expressed in the intestine. All FXR isoforms bind to the IR-1 motif while only FXR2 and FXR4 have been shown to bind the ER-2 DNA binding motifs. Identification of confirmed binding partners of the hepatic and intestinal FXR proteome, and studies on confirmed DNA binding motifs the FXR isoforms, may provide ideal targets for tissue-specific FXR therapeutics. Figure created with biorender.com and confirmed DNA binding motif sequences are repurposed with permission.[15]