• Open Access
    Review

    Cholestasis associated to inborn errors in bile acid synthesis

    Ricardo Espinosa-Escudero 1
    Elisa Herraez 1,2
    Anabel Sanchez-Martin 1
    Paula Sanchon-Sanchez 1,2
    Jose J. G. Marin 1,2
    Maria J. Monte 1,2*

    Explor Dig Dis. 2022;1:137–153 DOI: https://doi.org/10.37349/edd.2022.00010

    Received: June 22, 2022 Accepted: September 12, 2022 Published: December 7, 2022

    Academic Editor: Han Moshage, University of Groningen, The Netherlands

    This article belongs to the special issue CHOLESTASIS

    Abstract

    Several metabolic pathways are involved in the biotransformation of C27 neutral cholesterol to C24 primary bile acids (BAs), mainly cholic acid (CA) and chenodeoxycholic acid (CDCA), which are then conjugated with glycine or taurine. This process can start with the modification of the steroid ring or the shortening of the side chain and involves enzymes present in different subcellular compartments. Inborn errors affecting the biogenesis of organelles, such as peroxisomes, or the expression or function of specific enzymes of these convergent routes result in: i) the lack of mature C24-BAs, with the subsequent impairment in digestion and absorption of dietary fat and liposoluble vitamins, such as vitamin K, which may account for a deficient hepatic synthesis of several coagulation factors; ii) the accumulation of intermediate metabolites, which may affect hepatocyte physiology, causing cholestasis as a commonly shared alteration besides other deleterious hepatic events; and iii) extrahepatic clinical manifestations due to accumulation of toxic metabolites in other territories, such as the nervous system, causing neurological disorders. In general, diseases whose primary alteration is a genetic defect in BA synthesis are diagnosed in children or young individuals with a very low incidence. The symptomatology can markedly vary among individuals, ranging from mild to severe conditions. Oral therapy, based on the enrichment of the BA pool with natural C24-BAs, such as CA, CDCA, glyco-CA, or ursodeoxycholic acid (UDCA), depending on the exact deficiency causing the disease, may be beneficial in preventing life-threatening situations. In contrast, in other cases, a liver transplant is the only option for these patients. This review describes the updated information on the genetic and molecular bases of these diseases and the current approaches to achieve a selective diagnosis and specific treatment.

    Keywords

    Bile acid, cholestasis, genetic disorders, hepatitis, inborn errors, liver failure, metabolism

    Introduction

    Inborn errors in bile acid (BA) synthesis constitute a group of infrequent genetic disorders with an autosomal recessive transmission. Altogether, they are responsible for up to 2% of persistent cholestasis in children [1]. Although there is no consensus to define rare diseases, for instance, in the European Union these are defined as those that affect no more than 1 in 2,000 individuals while in the United States rare diseases are included in this group if they affect fewer than 200,000 people in the whole country, inborn defects in BA synthesis are globally considered rare diseases.

    The most common clinical presentation is cholestasis which develops progressively in childhood, although other clinical settings may also occur, such as neonatal hepatitis and advanced cirrhosis at birth. Cases of cholestasis development in late childhood or even adult-onset of liver disease have been described [2]. The severity of the conditions varies from mild to severe or even life-threatening, depending on the specific enzyme affected and the magnitude of the defect. Failure to synthesize BAs leads to reduced bile flow and decreased intestinal solubilization and absorption of fat and fat-soluble vitamins, resulting in steatorrhea and fat-soluble vitamin deficiency. When the alteration leads to an accumulation of monohydroxylated BAs or unsaturated oxo-BAs, many of which are toxic and potent cholestatic agents, the progression of liver damage is usually rapid [3].

    The diagnosis of these disorders is based on the analysis of the profile of BA molecular species and their precursors in body fluids by using mass spectrometric techniques, such as liquid chromatography coupled with tandem mass spectrometry (LC-MS/MS) [4, 5] and liquid secondary ionization mass spectrometry [6]. The most common biochemical findings in patients with BA synthetic defects are: i) the marked reduction or complete lack of primary BAs, cholic acid (CA), and chenodeoxycholic acid (CDCA), in serum, bile, and urine, and ii) the presence of elevated concentrations of atypical BAs and intermediate metabolites that retain the structure of the substrates that the deficient enzyme should have biotransformed. Confirmation of the genetic defect is usually achieved by DNA sequencing.

    Early diagnosis is critical since many of these patients respond well to oral treatment with BA sequestering resins, such as cholestyramine, but especially with natural BAs, such as CA, glycocholic acid (GCA), CDCA, or ursodeoxycholic acid (UDCA, Figure 1). BA replacement therapy has a dual objective. On the one hand, it is to provide the essential absent primary BAs necessary for the digestion of fats and fat-soluble vitamins. On the other hand, the aim is to suppress the endogenous synthesis of these molecules, thus reducing the formation of abnormal potentially toxic intermediate metabolites. BA therapy is lifelong for responsive conditions. In these cases, the patient’s adherence should be periodically evaluated.

    Molecular structure of some of the bile acid (BA) species mentioned in the text. DHCA: 3α,7α-dihydroxy-5β-cholestanoic acid; THCA: 3α,7α,12α-trihydroxy-5β-cholestanoic acid

    Regarding the existence of animal models for the study of these diseases, several murine strains have been generated and characterized (Tables 1 & 2). In general, with a few exceptions, these models only partially reproduce the features of the human condition. The differences are mainly accounted for by the dissimilarity in BA pool composition and metabolism between humans and rodents [7, 8].

    Inborn defects in BA synthesis: sterol nucleus modifications

    GeneProteinClinical featuresTreatmentRef.Animal models
    CYP7A1CYP7A1No signs of liver damage, hypercholesterolemia, hypertriglyceridemia, gallstone diseaseN.D.[24][2528]
    HSD3B7HSD3B7Neonatal cholestasis, cirrhosis, malabsorption of fats and fat-soluble vitaminsCA, CDCA[32, 33, 36][35]
    AKR1D1 (SR5B1)AKR1D1Neonatal hepatitis, cirrhosis, liver failureCA, CDCA[37, 40][38]
    CYP7B1CYP7B1Severe neonatal liver disease, cholestasis, fibrosis, cirrhosisCDCA, Liver transplantation[43, 46, 47][44]
    Display full size

    AKR1D1: Δ4-3-oxosteroid-5β-reductase; BA: bile acid; CYP7A1: cholesterol 7α-hydroxylase; CYP7B1: oxysterol 7α-hydroxylase; HSD3B7: 3β-hydroxy-Δ5-C27-steroid-oxidoreductase; N.D.: not defined

    Inborn defects in BA synthesis: side chain shortening and conjugation

    GeneProteinClinical featuresTreatmentRef.Animal models
    CYP27A1CYP27A1Children: neonatal cholestasis, diarrhea, cataracts, growth retardation
    Adults: xanthomas, neuropsychiatric symptoms (ataxia, dementia)
    CDCA[48, 54][5053]
    PEXSeveral peroxisomal proteinsLiver dysfunction, neurological abnormalities, hearing and vision impairment, adrenocortical dysfunctionSymptomatic or supportive therapies[56, 57][5860]
    ABCD3PMP70Hepatic dysfunction, hepatosplenomegalyN.D.[61][61, 87]
    AMACRAMACRChildren: malabsorption of fat-soluble vitamins, coagulation disorders, cholestasis, hepatitis
    Adults: sensory and motor degenerative neuropathy
    CA[6366][67]
    ACOX2ACOX2Hypertransaminasemia and severe hepatic and neurological alterationsCholestyramine
    UDCA
    [6972][73, 74]
    HSD17B4DBPHypotonia, delayed growth and psychomotor development, visual and auditory defects. Hepatomegaly, fibrosisN.D.[75][76, 77]
    SCP2SCPxFertility abnormalities. Brain lesions with motor disordersPhytanic acid-restricted diet[78, 79][80]
    SLC27A5BACSNo signs of liver diseaseN.D.[81][82, 83]
    BAATBAATNeonatal cholestasis. Malabsorption of fat-soluble vitamins. RicketsGCA[84, 86][85]
    Display full size

    ABCD3: ATP-binding cassette subfamily D member 3; ACOX2: acyl-coenzyme A (CoA) oxidase 2; AMACR: α-methyl acyl-CoA racemase; BA: bile acid; BAAT: BA CoA: amino acid N-acyl transferase; BACS: BA-CoA synthetase; CYP27A1: sterol 27-hydroxylase; DBP: D-bifunctional protein; PEX: peroxisomal biogenesis factor or peroxin; PMP70: peroxisomal membrane protein 70; SCP2: sterol carrier protein 2

    BA biosynthesis pathways

    Biosynthesis of primary BAs from cholesterol involves at least 17 enzymes, one or more intracellular transporters, and different subcellular compartments, including endoplasmic reticulum, mitochondria, peroxisomes, and cytosol. Four main pathways of BA biosynthesis have been described: the classic or neutral pathway, the alternative or acidic pathway, the Yamasaki pathway, and the 25-hydroxylation pathway.

    The main, neutral or classical pathway

    In this metabolic route, the modification of the cholesterol ring precedes the cleavage of its side chain (Figure 2). The pathway starts with the hydroxylation at C7, which is catalyzed by the microsomal enzyme CYP7A1, a member of the cytochrome P450 family of proteins. CYP7A1 is exclusively expressed in hepatocytes and, owing to its critical role in this biosynthetic pathway, is the element on which most regulatory factors converge [9]. The product of the CYP7A1-mediated reaction is 7α-hydroxycholesterol, which is then converted into 7α-hydroxy-4-cholesten-3-one (7αC4) by the action of microsomal HSD3B7. Then the microsomal enzyme sterol 12α-hydroxylase (CYP8B1) can add a hydroxyl group at C12, generating a series of intermediates whose final product is CA. If 12α-hydroxylation does not occur, the final compound generated by this route is CDCA, which highlights the role of CYP8B1 in controlling the CA/CDCA ratio and justifies its tight regulation [10]. In both cases, the double bond at the C4 position is then reduced by the cytosolic enzyme AKR1D1, resulting in the presence of a hydrogen at the 5β position. The last steroid ring modification is catalyzed in the cytosol by 3α-hydroxysteroid-dehydrogenase (AKR1C4), which reduces the 3-oxo group to a 3-hydroxyl with α-stereochemistry.

    Enzymes involved in the main, neutral, or classical pathway of bile acid (BA) synthesis. Several inborn errors leading to autosomal recessive disorders have been described to affect some of these enzymes (red bars). Although all resulting diseases are infrequent, the level of incidence is indicated by the number of asterisks as low (***), very low (**), and extremely low (*). 7αC4:7α-hydroxy-4-cholesten-3-one; CA: cholic acid; CDCA: chenodeoxycholic acid; DHCA: 3α,7α-dihydroxy-5β-cholestanoic acid; THCA: 3α,7α,12α-trihydroxy-5β-cholestanoic acid

    The shortening of the cholesterol side chain begins with the 26-hydroxylation due to the mitochondrial enzyme CYP27A1 [11]. Subsequent oxidation to an aldehyde and then to a carboxylate group allows the formation of the C27-BA intermediates DHCA and THCA. The hydroxylation of the pro-chiral side chain by CYP27A1 results in the exclusive formation of the 25R-stereoisomer of both compounds, which must then be activated by a thioester bond with an acetyl-CoA molecule. Both BACS and very long-chain acyl-CoA synthetase (VLCS) are involved in the generation of C27-BA-CoAs, which are taken up by the peroxisome through the ABCD3 transporter, also called PMP70. The reactions in this organelle start with the racemization of the C25 chiral center from the R- to the S-configuration by AMACR. Next, the stereospecific enzyme ACOX2 oxidizes the 25S-isomers, generating a double bond at position C24, which is then hydroxylated and subsequently oxidized to generate a 24-keto group by the action of HSD17B4 or DBP. SCPx or SCP2 accounts for the thiolytic cleavage of this ketone into propionyl-CoA and CA-CoA or CDCA-CoA. Finally, the conjugation of these BA-CoAs with glycine or taurine is carried out by BAAT, generating the four primary human BAs, i.e., taurocholic acid (TCA), GCA, taurochenodeoxycholic acid (TCDCA) and glycochenodeoxycholic acid (GCDCA).

    The acidic pathway

    In this metabolic pathway, which is quantitatively more important during early life and infancy, the first step is the formation of a C27-carboxylic acid, which is then followed by modifications of the steroid nucleus and further peroxisomal chain shortening to finally generate C24-BAs, preferentially CDCA [12]. The acidic pathway is initiated by CYP27A1, which carries out the hydroxylation of the C26 position generating the 25R-stereoisomer. Despite catalyzing the first reaction of the process, CYP27A1 is not its limiting enzyme. Instead, the crucial step appears to be the steroidogenic acute regulatory protein-related lipid transfer domain 1 (STARD1)-mediated transport of cholesterol to the inner mitochondrial membrane [13]. Once (25R)-3β-hydroxy-5-cholestenoic acid is obtained, its 7α-hydroxylation is carried out by CYP7B1. This constitutes a difference from the neutral pathway in which this hydroxylation is mediated by CYP7A1. Next, the same enzymes involved in the classic pathway (HSD3B7, AKR1D1, and AKR1C4) complete the ring modifications. Other oxysterols such as 24- and 25-hydroxycholesterol, highly relevant in regulating lipid metabolism, can also be incorporated into the biosynthesis of BAs at the beginning of this pathway. Both metabolites can be obtained by the action of CYP27A1. Moreover, brain CYP46A1 can also minimally contribute to this metabolic route through the synthesis and release to the blood of 24-hydroxycholesterol [14].

    The Yamasaki pathway

    This route shares with the acidic pathway the initial formation of (25R)-3β-hydroxy-5-cholestenoic acid through CYP27A1. Subsequently, the shortening of the side chain is carried out at the peroxisome to obtain 3β-hydroxy-5-cholenoic acid, whose steroidal nucleus is then modified to finally generate C24-BAs, mainly CDCA. Although the contribution of this route to the biosynthesis of BAs is unknown, the presence of 3β-hydroxylated intermediates in amniotic fluid suggests its importance during intrauterine life [15].

    The 25-hydroxylation pathway

    This alternative route of BA synthesis, in which the side chain shortening is independent of peroxisomal β-oxidation, begins with the 25-hydroxylation of 5β-cholestane-3α,7α,12α-triol by the action of the microsomal enzyme CYP3A4 [16]. The product of this reaction, 5β-cholestane-3α,7α,12α,25-tetrol is then hydroxylated to 5β-cholestane-3α,7α,12α,24S,25-pentol [17]. This hydroxylation is likely catalyzed by CYP27A1, because patients with cerebrotendinous xanthomatosis (CTX), who lack active CYP27A1, accumulate cholestanetetrol when the direct precursor is administered [18, 19]. Finally, conversion to CA would be carried out rapidly through a 24-keto intermediate by one or more cytosolic enzymes with stereospecific selectivity for the 24S isomer [18].

    Additional conjugation reactions

    Besides the amidation with taurine or glycine, under certain circumstances mature C24-BAs can be additionally conjugated with other polar polyatomic moieties, which results in compounds with enhanced water-solubility, reduced cytotoxicity, and facilitated renal elimination. More than 70% of BAs in urine are conjugated at 3α-position with a sulfate group, being this reaction carried out by sulfotransferase family 2A member 1 (SULT2A1) [20]. This biotransformation is particularly relevant in the case of lithocholic acid (LCA), a very hydrophobic and toxic BA, which is sulfated in high proportion, thus increasing its water solubility and facilitating its renal and fecal excretion [21]. Glucuronidation is mediated by uridine-diphospho (UDP) glucuronosyltransferase family 1 member A3 (UGT1A3), as well as other enzymes of the UGT family. This conjugation reaction can be performed at different sites of the molecule, depending on the involved isoenzyme, e.g., UGT1A3 mainly generates glucuronide-BAs at C24 position [22]. These conjugated BA species are a minority in healthy individuals but can become quantitatively relevant in patients with cholestasis [20, 23].

    Inborn errors in the transformation of the sterol nucleus

    Up to now, four inborn errors affecting the enzymes involved in transforming the cholesterol ring during BA biosynthesis have been identified (Table 1, Figure 2).

    CYP7A1

    Deficient CYP7A1 activity was described in three siblings with statin-resistant hypercholesterolemia. Interestingly, two of these patients were male and had hypertriglyceridemia and a history of biliary lithiasis. One presented an accumulation of cholesterol in the liver biopsy, along with decreased fecal excretion of BAs and an increased proportion of CDCA in feces. All three individuals were homozygous for a mutation in the CYP7A1 gene, leading to a frameshift (L413fsX414) that markedly alters the active site of the enzyme and subsequent results is the lack of function [24]. CYP7A1 deficiency does not lead to acute liver disease, as the reduction in BA synthesis by the neutral pathway can be compensated by an increment of the alternative acidic route. However, pathological repercussions appear in adulthood. As its inheritance seems co-dominant, individuals with this defect in heterozygosis also have high cholesterol levels, early gallstones formation, and even premature coronary and peripheral vascular disease [24]. Mice lacking CYP7A1 display reduced BA synthesis, while cholesterol serum levels and mortality differ between models probably due to their different genetic backgrounds [2528].

    HSD3B7

    Deficiency in HSD3B7 activity [Online Mendelian Inheritance in Man (OMIM): 607765] is the most frequent genetic defect affecting BA synthesis. In an American study from 1987 to 2016, HSD3B7 deficiency was found in 34% out of 297 patients with inherited errors of BA metabolism included in that cohort [29]. The inheritance of this disorder, characterized by a diverse spectrum of mutations in the HSD3B7 gene, is autosomal recessive [30]. In these patients, there are low levels of mature BAs due to their inability to reduce the double bond in the Δ5 position of the cholesterol ring, preventing the epimerization of the hydroxyl in 3β and causing the accumulation of unsaturated 3β,7α-dihydroxy-5-cholenoic acid and 3β,7α,12α-trihydroxy-5-cholenoic acid, which are usually sulfated in the C3 position. It has been proven that 3β-hydroxy-5-cholenoic acids fail to act as agonists of the main nuclear receptors involved in BA homeostasis farnesoid X receptor (FXR), pregnane X receptor (PXR) and constitutive androstane receptor (CAR), resulting in their impairment to exert hepatoprotective actions and thus explaining the mechanism for progressive cholestatic liver disease [31]. Sulfation of these toxic Δ5-intermediates will favor their elimination, as is the case for LCA [21]. Serum aminotransferases [alanine aminotransferase (ALT), aspartate aminotransferase (AST)] are elevated, while gamma-glutamyl transpeptidase (GGT) levels are normal. Clinically, HSD3B7 deficiency is characterized by neonatal cholestasis with jaundice, impaired absorption of lipids and fat-soluble vitamins, and steatorrhea, which progresses to more severe liver alterations, such as cirrhosis and hepatomegaly, with or without splenomegaly [1, 32]. However, in some cases, the clinical presentation is atypical [33] or even asymptomatic [34]. HSD3B7 deficiency in mice leads to cholesterol malabsorption and a very high mortality rate [35]. Early treatment with CA or CDCA is effective in normalizing liver function in these patients [33, 36]. Accurate determination of 3β-hydroxy-Δ5-BA sulfates is critical for both diagnosis and dose titration and monitoring response to therapy with primary BAs [4, 5].

    AKR1D1

    Deficiency in AKR1D1 activity (OMIM: 235555) is the second most frequent disorder of BA metabolism, accounting for 20% of cases in the American cohort mentioned above [29]. Patients with this disorder show similar symptomatology to those with HSD3B7 deficiency. Still, in this case, there is an accumulation of 4-cholenoic BAs, which may have a 3-oxo group, along with the presence of 5α-BAs (flat “allo” BAs) [37]. Liver function tests show elevated serum aminotransferases and GGT. The typical clinical presentation resembles neonatal hepatitis, with rapid progression of liver disease, leading to cirrhosis and liver failure in infancy. In contrast to patients, mice with AKR1D1 deficiency do not show overt signs of cholestasis, hepatic inflammation, or liver damage [38]. Due to the reduced synthesis of primary BAs, treatment based on BA replacement therapy could be beneficial for these patients [37]. However, conversely to HSD3B7 deficiency, treatment of pronounced AKR1D1 deficiency with CDCA worsen cholestasis and liver function tests. This may be due to the enhanced generation in these patients of 3-oxo-4-cholenoic acids which cannot be sulfated at the C3 position, and which inhibit the secretion of exogenously administered CDCA [3]. Relief of cholestasis and improvement of liver function in these patients is achieved by administration of the less toxic CA in combination with CDCA [39]. Different autosomal recessive mutations in the AKR1D1 gene, both in homozygosis and heterozygosis, have been associated with this disorder. Patients with a longer prothrombin time have a worse prognosis and a poorer response to treatment with BAs [40]. Importantly, secondary AKR1D1 deficiency may occur due to several liver diseases, highlighting the need for genotyping AKR1D1 deficient patients to elucidate the actual etiology of the disease [41].

    CYP7B1

    This enzyme is specific to the alternative acidic pathway. The lack of CYP7B1 enzymatic activity (OMIM: 613812) is one of the rarest defects in BA synthesis, as it has been reported in few children [29]. However, a recent study on Taiwanese infants suggested that the incidence of this disorder may be higher than believed due to its misdiagnosis in children [42]. The accumulation of Δ5-monohydroxylated and saturated 3β-monohydroxylated BAs leads to severe neonatal liver disorders such as progressive cholestasis, fibrosis, and cirrhosis with fatal consequences [43]. The serious clinical consequences of CYP7B1 deficiency highlight the importance of the alternative acidic pathway in early life in humans. Conversely, in Cyp7b1−/− mice, BA metabolism as well as liver function are normal, and the loss of this enzyme is compensated by increases in the synthesis of BAs by other pathways [44]. Along with 3β-hydroxy-Δ5-cholenoic acid, which makes up 90% of total plasma BAs, various oxysterols such as 25-hydroxycholesterol and 27-hydroxycholesterol accumulate and help to differentiate this disorder from AKR1D1 deficiency [45]. Serum aminotransferases are markedly elevated, while GGT levels are normal. Unlike other defects in enzymes of BA synthesis involved in the biotransformation of the steroid ring, CYP7B1 deficiency is a particularly severe condition that does not respond to CA. As therapeutic options, liver transplantation [46] or CDCA treatment [47] has been effective in some patients.

    Defects in cholesterol side chain shortening

    Altered shortening of cholesterol side chain to biotransform C27 steroid molecules to mature C24-BAs can be due to dysfunctional organelles or specific enzymes involved in the cleavage process (Table 2).

    CYP27A1

    Deficient mitochondrial CYP27A1 activity (OMIM: 213700) causes CTX, a rare autosomal recessive lipid storage disease. Its clinical manifestations in affected children include neonatal cholestasis, diarrhea, cataracts, and growth retardation. The situation worsens in adults with the onset of xanthomas and neuropsychiatric symptoms such as ataxia and dementia [48]. This alteration is characterized by a deficit of BAs and a deposit of cholesterol and cholestanol in different tissues. The blockade of the neutral pathway at this level generates an accumulation of 5β-cholestane-3α,7α,12α-triol, which is diverted through the 25-hydroxylation route for its biotransformation into CA [19]. However, this alternative pathway cannot entirely compensate for the lack of CYP27A1-mediated side chain shortening for C24-BA production. Hence, bile alcohols are generated and accumulated, which is a helpful plasma biomarker for the diagnosis of this disorder [49]. On the contrary, mice lacking CYP27A1 show a milder phenotype with lower accumulation of oxysterols and BA intermediates, probably due to a more active 25-hydroxylation pathway in this species [50, 51]. Consistently, these animals do not develop metabolic and neurologic CTX-related pathological abnormalities [52, 53]. The treatment of CTX patients with CDCA restores normal C24-BA levels and decreases those of cholestanol and bile alcohols, thus preventing severe clinical symptoms if the treatment is started early enough [54].

    Peroxisomal defects

    Peroxisomes are involved in many cellular functions, e.g., long-chain fatty acid β-oxidation, pipecolic and phytanic acid oxidation, synthesis of plasmalogens, cholesterol, BAs, and regulation of the oxidation-reduction status. Therefore, disorders accounting for impaired peroxisome biogenesis or specific defects in any particular peroxisomal enzyme involved in these metabolic pathways lead to the accumulation of various metabolites, which may be deposited in several tissues, consequently causing pathophysiological alterations [55]. These defects affecting BA metabolism are described below.

    Impaired peroxisome biogenesis

    Zellweger spectrum disorders (ZSDs) are a group of diseases characterized by the absence of functional peroxisomes. Several mutations affecting members of the PEX gene family are involved in ZSDs. These mutations impair the transport of certain proteins to the peroxisomal membrane and matrix during the biogenesis of these organelles. This group of disorders includes Zellweger syndrome, neonatal adrenoleukodystrophy, and infantile Refsum disease. Although ZSDs are clinically heterogeneous, they share features such as liver dysfunction, developmental delay, adrenocortical dysfunction, and neurological abnormalities (e.g., hearing and vision impairment). ZSD patients lack mature C24-BAs, whereas their precursors DHCA and THCA accumulate. There are also increased levels of C29-dicarboxylic acid, which is likely generated by elongation of THCA-CoA. The detection of these BA intermediates in serum can be used in the diagnosis of these patients [56]. Their treatment is based on symptomatic and supportive therapies [57]. Several animal models lacking different PEX proteins have been developed. Their phenotypes resemble those typical of patients with ZSDs. Thus, Pex5 and Pex13 knockout mice show intrauterine growth retardation, hypotonia and high neonatal mortality, associated with alterations in neuronal maturation and migration, together with biochemical alterations similar to those of ZSDs [58, 59]. Mice lacking PEX2 develop early intrahepatic cholestasis leading to steatohepatitis, with reduction of C24-BAs and accumulation of C27 intermediates [60].

    Impaired substrate transport across the peroxisomal membrane

    Other peroxisomal disorders are characterized by the alteration in the transport of substrates into the peroxisomal matrix via the ATP-binding cassette (ABC) pump ABCD3, also known as PMP70, accounting for the peroxisomal uptake of THCA-CoA and DHCA-CoA. This disorder has been described in a patient with hepatosplenomegaly who presented plasma accumulation of THCA-CoA, DHCA-CoA, and C29-dicarboxylic acid [61]. Abcd3 knockout mice also show high levels of these BA intermediates in serum and an increase of liver weight [61]. On the other hand, although X-linked adrenoleukodystrophy (X-ALD) due to mutations in the ABCD1 gene has never been associated with alterations in BA metabolism, a decrease in mature C24-BA levels has been reported in a recent case of X-ALD [62]. However, the absence of inactivating mutations in other peroxisomal enzymes of the BA biosynthetic pathway remains to be demonstrated and the actual link between BA synthesis and X-ALD elucidated.

    AMACR

    Deficiency in AMACR activity (OMIM: 614307) is an autosomal recessive disorder characterized by plasma accumulation of pristanic acid and C27-BAs. A feature that facilitates the diagnosis is that, due to the lack of racemase enzymatic activity, only the 25R-isomers of THCA, DHCA, and pristanic acid are accumulated. The clinical presentation of this disease is highly variable. Affected children have malabsorption of fat-soluble vitamins, vitamin K deficiency-associated coagulation disorders, cholestasis, and hepatitis, although they respond favorably to treatment with primary C24-BAs [63, 64]. Adult presentation is characterized by sensory and motor degenerative neuropathy, with high interindividual variability. Thus, peripheral neuropathy, cognitive deficits, retinitis pigmentosa, and cerebellar dysarthria, among other alterations, may appear [65, 66]. Surprisingly, AMACR deficient mice do not develop clinical symptoms, probably due to a metabolic adaptation that reduces intestinal absorption of cholesterol and liver biosynthesis of different sterols, keeping the overall cholesterol levels in normal ranges [67].

    ACOX2

    Deficiency of ACOX2 activity (OMIM: 617308) results in impaired peroxisomal metabolism of branched-chain fatty acids and BA intermediates. A few cases of ACOX2 deficiency have been described so far. The first patient reported was a 16-year-old male diagnosed with persistent hypertransaminasemia upon exposure to non-steroidal anti-inflammatory drugs [68]. The finding of the homozygous mutation c.673C>T (p.Arg225Trp) in the ACOX2 gene was associated with decreased activity of the enzyme, leading to serum and urine accumulation of THCA and DHCA [69]. Different mutations in the same gene reported in two additional patients have been associated with more severe clinical manifestations. Both children had consanguineous parents and died suffering from severe hepatic and neurological alterations. One of these patients carried a homozygous premature termination mutation in ACOX2 (p.Y69*) [70], whereas the other one was homozygous for a four nucleotide deletion leading to a premature stop codon (p.Thr154Serfs*25) [71]. Both patients presented elevated serum levels of C27-BAs. A more recent study has detected a new mutation and identified several mutations with a potential dysfunctional impact on ACOX2 activity. Moreover, in two patients, the coexistence of two different mutations in heterozygosis but on different alleles also resulted in metabolic disorder leading to C27-BA accumulation [72]. Interestingly, this study revealed that patients with ACOX2 deficiency-associated hypertransaminasemia (ADAH) could benefit from treatment with UDCA, which normalizes serum levels of aminotransferases [72]. Two Acox2 knockout mice models have recently been reported. In one of these studies, the animals spontaneously developed steatosis, liver inflammation with fibrosis, and eventually liver cancer [73]. Unfortunately, the authors did not provide information regarding BA metabolism in this case. In contrast, in the other murine model of ACOX2 deficiency, developed in our laboratory, young animals do not show signs of severe liver disease. These mice partially mimic the phenotype observed in ADAH patients as far as BA metabolism is concerned [74].

    HSD17B4

    Because DBP is a complex enzyme with more than one activity, which justifies its name, inborn errors affecting the HSD17B4 gene (OMIM: 261515) can impair the two sequential reactions catalyzed by this enzyme during the peroxisomal step of BA metabolism. Consequently, DBP disorders can be classified into three groups: type I, characterized by the absence of both hydratase and dehydrogenase activity; type II, in which only the hydratase activity is impaired; and type III, in which only dehydrogenase activity is deficient. The three phenotypes have similar severe clinical symptomatology, with most patients not surviving beyond the age of two years. Although the clinical manifestations show marked interindividual variability, the most frequent alterations are neonatal hypotonia accompanied by seizures and cranial dysmorphia, delayed growth and psychomotor development, visual and auditory system failures, and hepatomegaly. Postmortem liver biopsy carried out in some cases has shown fibrosis, steatosis, hemosiderosis, and signs of cholestasis. The few patients with prolonged survival reported so far had a severe delay in acquiring motor skills, lack of the ability to use language, audiovisual defects, and signs of demyelination. Regarding biochemistry data, elevated plasma levels of very long-chain fatty acids, THCA and DHCA have been found. Although the accumulation of these C27-BA intermediates does not occur in all patients, there seems to be an inverse correlation between their plasma levels and patient survival [75]. Concerning mice models of HSD17B4 deficiency, the absence of this protein and the accumulation of the mentioned metabolites cause growth retardation during the first weeks of life and lead to premature death [76, 77].

    SCPx

    Only two cases of patients with a deficiency of SCPx activity (OMIM: 613724), a peroxisomal enzyme required for the last step of β-oxidation of branched-chain fatty acids, have been reported [78]. The first patient was a male with a medical history of spasmodic torticollis, dystonic head tremor, dysphemia, and fertility disturbances such as hypergonadotrophic hypogonadism and azoospermia. Worsening neurological symptoms revealed hypothalamic, pontine, and occipital brain lesions, as well as hyposmia, slight hypoacusis, saccadic eye movements, diminished reflexes of lower extremities, slight cerebellar ataxia, and gait and balance impairment. Nevertheless, liver biochemical markers were normal, and traces of C27-BAs (DHCA and THCA) were barely detected in plasma. In contrast, serum levels of pristanic acid were elevated. In addition, large amounts of bile alcohols were detected in urine [78]. Although the second patient carried in heterozygosis a different mutation affecting SCPx, the clinical features of the disease were similar, with motor disturbances in the hands, balance problems, and deafness. Subsequent medical examinations revealed alterations in proprioception, mild dysmetria, and dysdiadochokinesis. In addition, serological tests showed an increase in liver markers such as aminotransferases, GGT, and creatine kinase, as well as the accumulation of pristanic acid [79]. In both cases, symptoms improved after starting a phytanic acid (a precursor of pristanic acid) restricted diet. Cholesterol and BA metabolism alterations, without signs of neuropathy, have been reported in SCP2 knockout mice [80].

    Defects in amidation

    Conjugation of the C24-BAs with the amino acids glycine and taurine is the final step of BA synthesis by hepatocytes. This process requires the sequential action of two enzymes, BACS (encoded by SLC27A5), essential for re-conjugation of BAs deconjugated by gut bacteria, and BAAT (encoded by BAAT). Firstly, BACS catalyzes the activation of BAs via the formation of BA-CoA thioesters, which are needed for the glycine or taurine addition to BAs, a reaction catalyzed by BAAT [9]. Inborn errors affecting these two enzymes have been reported (Table 2).

    BACS

    So far, the only patients described with deficiency concerning BACS activity were two sisters carrying a homozygous mutation in the SCL27A5 gene. Although both girls had a similar abundance (85%) of unconjugated BAs in plasma and urine, only one of them developed liver disease. The reason for this discrepancy likely was that in the patient with more severe manifestations, BACS deficiency coexisted with another homozygous deleterious mutation in the ABCB11 gene, encoding the bile salt export pump (BSEP), which plays an essential role in the canalicular secretion of BAs. In contrast, her sister was heterozygous for this mutation [81]. Mice lacking BACS show a high proportion of unconjugated BAs in bile, without signs of disease. However, these animals do no gain extra weight when fed with a high-fat diet [82, 83].

    BAAT

    In ten pediatric patients harboring a lack of enzymatic activity in BAAT (OMIM: 619232), this deficiency caused malabsorption of fat-soluble vitamins and increased prothrombin time. The rest of the clinical manifestations showed marked interindividual variability. Thus, neonatal cholestasis ranged from mild to very severe, with liver transplantation being necessary for one of the patients who developed biliary cholangiopathy, eventually leading to cirrhosis and liver failure. Data obtained from the liver biopsy of several of these patients showed idiopathic neonatal hepatitis with ductular proliferation and fibrosis. In addition, four patients suffered from rickets because of vitamin D deficiency. As could be expected, biochemical signs included the absence of glyco- and tauroconjugated BAs in plasma, urine, and bile, where unconjugated CA was the most abundant BA [84]. BAs in the liver of BAAT-deficient mice are mostly unconjugated. As a response to facilitate their detoxification, BA polyhydroxylation is markedly enhanced. These mice also display defective body weight gain [85]. The similarity in the BA pattern of patients with BAAT or BACS deficiency determines that the only way to discriminate between both disorders is through genotyping or immunolabeling analysis of liver tissue obtained by biopsy. Therapy based on the oral administration of conjugated CA (GCA) has improved growth rate and fat-soluble vitamin absorption in patients with BA amidation defects [86].

    Conclusions and perspectives

    The general conclusion of this review is that, in patients with the clinical characteristics described above, after ruling out the presence of more frequent pediatric liver diseases, the existence of an inborn error in BA metabolism must be elucidated. Despite the difficulty in discriminating between different types of cholestatic conditions associated with these rare genetic defects, it is evident that a crucial goal in the diagnosis of these patients is to identify the exact cause of the failure in this aspect of liver function as soon as possible. An accurate and early diagnosis will direct the specific pharmacological treatment, which can prevent the progression to more severe stages of liver damage. Fortunately, the advance in analytical techniques such as LC-MS/MS to determine BA profiles, particularly measuring in a sensitive manner these rare and less abundant species in the serum of normal individuals, and available sequencing methods to identify the underlying genetic defect permit us to carry out a rapid and helpful screening among patients with typical clinical manifestations and biochemical signs of cholestasis and liver damage.

    Abbreviations

    ABCD3:

    ATP-binding cassette subfamily D member 3

    ACOX2:

    acyl-coenzyme A oxidase 2

    AKR1C4:

    3α-hydroxysteroid-dehydrogenase

    AKR1D1:

    Δ4-3-oxosteroid-5β-reductase

    AMACR:

    α-methyl acyl-coenzyme A racemase

    BAAT:

    bile acid coenzyme A: amino acid N-acyl transferase

    BACS:

    bile acid-coenzyme A synthetase

    BAs:

    bile acids

    CA:

    cholic acid

    CDCA:

    chenodeoxycholic acid

    CoA:

    coenzyme A

    CTX:

    cerebrotendinous xanthomatosis

    CYP27A1:

    sterol 27-hydroxylase

    CYP7A1:

    cholesterol 7α-hydroxylase

    CYP7B1:

    oxysterol 7α-hydroxylase

    CYP8B1:

    sterol 12α-hydroxylase

    DBP:

    D-bifunctional protein

    DHCA:

    3α,7α-dihydroxy-5β-cholestanoic acid

    GCA:

    glycocholic acid

    GGT:

    gamma-glutamyl transpeptidase

    HSD3B7:

    3β-hydroxy-Δ5-C27-steroid-oxidoreductase

    OMIM:

    Online Mendelian Inheritance in Man

    PEX:

    peroxisomal biogenesis factor or peroxin

    PMP70:

    peroxisomal membrane protein 70

    SCP2:

    sterol carrier protein 2

    THCA:

    3α,7α,12α-trihydroxy-5β-cholestanoic acid

    UDCA:

    ursodeoxycholic acid

    UGT1A3:

    uridine-diphospho glucuronosyltransferase family 1 member A3

    X-ALD:

    X-linked adrenoleukodystrophy

    ZSDs:

    Zellweger spectrum disorders

    Declarations

    Author contributions

    REE, EH, ASM and PSS played a main role in collecting information; JJGM coordinated the study; MJM wrote the first draft.

    Conflicts of interest

    The authors declare that they have no conflicts of interest.

    Ethical approval

    Not applicable.

    Consent to participate

    Not applicable.

    Consent to publication

    Not applicable.

    Availability of data and materials

    Not applicable.

    Funding

    This study was funded by the CIBERehd (EHD15PI05/2016) and “Fondo de Investigaciones Sanitarias, Instituto de Salud Carlos III”, Spain (PI19/00819 and PI20/00189, co-funded by European Regional Development Fund/European Social Fund, “Investing in your future”); “Junta de Castilla y Leon” (SA074P20); Fundació Marato TV3 (Ref. 201916-31), Spain; AECC Scientific Foundation (2017/2020), Spain; “Centro Internacional sobre el Envejecimiento” (OLD-HEPAMARKER, 0348_CIE_6_E), Spain. REE was supported by a predoctoral contract funded by “Junta de Castilla y Leon” and “Fondo Social Europeo” (EDU/574/2018). ASM and PSS are supported by a predoctoral scholarship (FPU) funded by the Ministry of Science, Innovation and Universities, Spain. The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

    Copyright

    © The Author(s) 2022.

    References

    Bove KE, Heubi JE, Balistreri WF, Setchell KDR. Bile acid synthetic defects and liver disease: a comprehensive review. Pediatr Dev Pathol. 2004;7:31534. [DOI] [PubMed]
    Fischler B, Bodin K, Stjernman H, Olin M, Hansson M, Sjovall J, et al. Cholestatic liver disease in adults may be due to an inherited defect in bile acid biosynthesis. J Intern Med. 2007;262:25462. [DOI] [PubMed]
    Stieger B, Zhang J, O’Neill B, Sjovall J, Meier PJ. Differential interaction of bile acids from patients with inborn errors of bile acid synthesis with hepatocellular bile acid transporters. Eur J Biochem. 1997;244:3944. [DOI] [PubMed]
    Zhang W, Jha P, Wolfe B, Gioiello A, Pellicciari R, Wang J, et al. Tandem mass spectrometric determination of atypical 3β-hydroxy-Δ5-bile acids in patients with 3β-hydroxy-Δ5-C27-steroid oxidoreductase deficiency: application to diagnosis and monitoring of bile acid therapeutic response. Clin Chem. 2015;61:95563. [DOI] [PubMed]
    Maekawa M, Shimada M, Ohno K, Togawa M, Nittono H, Iida T, et al. Focused metabolomics using liquid chromatography/electrospray ionization tandem mass spectrometry for analysis of urinary conjugated cholesterol metabolites from patients with Niemann-Pick disease type C and 3β-hydroxysteroid dehydrogenase deficiency. Ann Clin Biochem. 2015;52:57687. [DOI] [PubMed]
    Setchell KDR, Heubi JE. Defects in bile acid biosynthesis--diagnosis and treatment. J Pediatr Gastroenterol Nutr. 2006;43:S1722. [DOI] [PubMed]
    Lorbek G, Lewinska M, Rozman D. Cytochrome P450s in the synthesis of cholesterol and bile acids--from mouse models to human diseases. FEBS J. 2012;279:151633. [DOI] [PubMed]
    Li J, Dawson PA. Animal models to study bile acid metabolism. Biochim Biophys Acta Mol Basis Dis. 2019;1865:895911. [DOI] [PubMed] [PMC]
    Chiang JYL. Bile acids: regulation of synthesis. J Lipid Res. 2009;50:195566. [DOI] [PubMed] [PMC]
    Zhang M, Chiang JY. Transcriptional regulation of the human sterol 12α-hydroxylase gene (CYP8B1): roles of hepatocyte nuclear factor 4α in mediating bile acid repression. J Biol Chem. 2001;276:416909. [DOI] [PubMed]
    Fakheri RJ, Javitt NB. 27-Hydroxycholesterol, does it exist? On the nomenclature and stereochemistry of 26-hydroxylated sterols. Steroids. 2012;77:5757. [DOI] [PubMed]
    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;263:1663744. [DOI] [PubMed]
    Ren S, Hylemon PB, Marques D, Gurley E, Bodhan P, Hall E, et al. Overexpression of cholesterol transporter StAR increases in vivo rates of bile acid synthesis in the rat and mouse. Hepatology. 2004;40:9107. [DOI] [PubMed]
    Pandak WM, Kakiyama G. The acidic pathway of bile acid synthesis: not just an alternative pathway. Liver Res. 2019;3:8898. [DOI] [PubMed] [PMC]
    Nakagawa M, Setchell KD. Bile acid metabolism in early life: studies of amniotic fluid. J Lipid Res. 1990;31:108998. [DOI] [PubMed]
    Furster C, Wikvall K. Identification of CYP3A4 as the major enzyme responsible for 25-hydroxylation of 5β-cholestane-3α,7α,12α-triol in human liver microsomes. Biochim Biophys Acta. 1999;1437:4652. [DOI] [PubMed]
    Cheng FW, Shefer S, Dayal B, Tint GS, Setoguchi T, Salen G, et al. Cholic acid biosynthesis: conversion of 5beta-cholestane-3alpha,7alpha,12alpha,25-tetrol into 5beta-cholestane-3alpha,7alpha, 12alpha,24beta,25-pentol by human and rat liver microsomes. J Lipid Res. 1977;18:613. [DOI] [PubMed]
    Shefer S, Cheng FW, Dayal B, Hauser S, Tint GS, Salen G, et al. A 25-hydroxylation pathway of cholic acid biosynthesis in man and rat. J Clin Invest. 1976;57:897903. [DOI] [PubMed] [PMC]
    Salen G, Shefer S, Cheng FW, Dayal B, Batta AK, Tint GS. Cholic acid biosynthesis: the enzymatic defect in cerebrotendinous xanthomatosis. J Clin Invest. 1979;63:3844. [DOI] [PubMed] [PMC]
    Alnouti Y. Bile acid sulfation: a pathway of bile acid elimination and detoxification. Toxicol Sci. 2009;108:22546. [DOI] [PubMed]
    Hofmann AF. Detoxification of lithocholic acid, a toxic bile acid: relevance to drug hepatotoxicity. Drug Metab Rev. 2004;36:70322. [DOI] [PubMed]
    Gall WE, Zawada G, Mojarrabi B, Tephly TR, Green MD, Coffman BL, et al. Differential glucuronidation of bile acids, androgens and estrogens by human UGT1A3 and 2B7. J Steroid Biochem Mol Biol. 1999;70:1018. [DOI] [PubMed]
    Perreault M, Bialek A, Trottier J, Verreault M, Caron P, Milkiewicz P, et al. Role of glucuronidation for hepatic detoxification and urinary elimination of toxic bile acids during biliary obstruction. PLoS One. 2013;8:e80994. [DOI] [PubMed] [PMC]
    Pullinger CR, Eng C, Salen G, Shefer S, Batta AK, Erickson SK, et al. Human cholesterol 7α-hydroxylase (CYP7A1) deficiency has a hypercholesterolemic phenotype. J Clin Invest. 2002;110:10917. [DOI] [PubMed] [PMC]
    Ishibashi S, Schwarz M, Frykman PK, Herz J, Russell DW. Disruption of cholesterol 7α-hydroxylase gene in mice. I. Postnatal lethality reversed by bile acid and vitamin supplementation. J Biol Chem. 1996;271:1801723. [DOI] [PubMed]
    Schwarz M, Lund EG, Setchell KDR, Kayden HJ, Zerwekh JE, Bjorkhem I, et al. Disruption of cholesterol 7α-hydroxylase gene in mice. II. Bile acid deficiency is overcome by induction of oxysterol 7α-hydroxylase. J Biol Chem. 1996;271:1802431. [DOI] [PubMed] [PMC]
    Erickson SK, Lear SR, Deane S, Dubrac S, Huling SL, Nguyen L, et al. Hypercholesterolemia and changes in lipid and bile acid metabolism in male and female cyp7A1-deficient mice. J Lipid Res. 2003;44:10019. [DOI] [PubMed]
    Ferrell JM, Boehme S, Li F, Chiang JYL. Cholesterol 7α-hydroxylase-deficient mice are protected from high-fat/high-cholesterol diet-induced metabolic disorders. J Lipid Res. 2016;57:114454. [DOI] [PubMed] [PMC]
    Heubi JE, Setchell KDR, Bove KE. Inborn errors of bile acid metabolism. Clin Liver Dis. 2018;22:67187. [DOI] [PubMed]
    Cheng JB, Jacquemin E, Gerhardt M, Nazer H, Cresteil D, Heubi JE, et al. Molecular genetics of 3β-hydroxy-Δ5-C27-steroid oxidoreductase deficiency in 16 patients with loss of bile acid synthesis and liver disease. J Clin Endocrinol Metab. 2003;88:183341. [DOI] [PubMed]
    Gioiello A, Cerra B, Zhang W, Vallerini GP, Costantino G, De Franco F, et al. Synthesis of atypical bile acids for use as investigative tools for the genetic defect of 3β-hydroxy-Δ5-C27-steroid oxidoreductase deficiency. J Steroid Biochem Mol Biol. 2014;144:34860. [DOI] [PubMed]
    Clayton PT, Leonard JV, Lawson AM, Setchell KD, Andersson S, Egestad B, et al. Familial giant cell hepatitis associated with synthesis of 3 beta, 7 alpha-dihydroxy-and 3 beta,7 alpha, 12 alpha-trihydroxy-5-cholenoic acids. J Clin Invest. 1987;79:10318. [DOI] [PubMed] [PMC]
    Bossi G, Giordano G, Rispoli GA, Maggiore G, Naturale M, Marchetti D, et al. Atypical clinical presentation and successful treatment with oral cholic acid of a child with defective bile acid synthesis due to a novel mutation in the HSD3B7 gene. Pediatr Rep. 2017;9:7266. [DOI] [PubMed] [PMC]
    Subramaniam P, Clayton PT, Portmann BC, Mieli-Vergani G, Hadzic N. Variable clinical spectrum of the most common inborn error of bile acid metabolism—3β-hydroxy-Δ5-C27-steroid dehydrogenase deficiency. J Pediatr Gastroenterol Nutr. 2010;50:616. [DOI] [PubMed]
    Shea HC, Head DD, Setchell KDR, Russell DW. Analysis of HSD3B7 knockout mice reveals that a 3α-hydroxyl stereochemistry is required for bile acid function. Proc Natl Acad Sci U S A. 2007;104:1152633. [DOI] [PubMed] [PMC]
    Mizuochi T, Kimura A, Ueki I, Takahashi T, Hashimoto T, Takao A, et al. Molecular genetic and bile acid profiles in two Japanese patients with 3β-hydroxy-Δ5-C27-steroid dehydrogenase/isomerase deficiency. Pediatr Res. 2010;68:25863. [DOI] [PubMed]
    Setchell KD, Suchy FJ, Welsh MB, Zimmer-Nechemias L, Heubi J, Balistreri WF. Delta 4-3-oxosteroid 5 beta-reductase deficiency described in identical twins with neonatal hepatitis. A new inborn error in bile acid synthesis. J Clin Invest. 1988;82:214857. [DOI] [PubMed] [PMC]
    Gathercole LL, Nikolaou N, Harris SE, Arvaniti A, Poolman TM, Hazlehurst JM, et al. AKR1D1 knockout mice develop a sex-dependent metabolic phenotype. J Endocrinol. 2022;253:97113. [DOI] [PubMed] [PMC]
    Clayton PT, Mills KA, Johnson AW, Barabino A, Marazzi MG. Delta 4-3-oxosteroid 5 beta-reductase deficiency: failure of ursodeoxycholic acid treatment and response to chenodeoxycholic acid plus cholic acid. Gut. 1996;38:6238. [DOI] [PubMed] [PMC]
    Wang HH, Wen FQ, Dai DL, Wang JS, Zhao J, Setchell KDR, et al. Infant cholestasis patient with a novel missense mutation in the AKR1D1 gene successfully treated by early adequate supplementation with chenodeoxycholic acid: a case report and review of the literature. World J Gastroenterol. 2018;24:408692. [DOI] [PubMed] [PMC]
    Ueki I, Kimura A, Chen HL, Yorifuji T, Mori J, Itoh S, et al. SRD5B1 gene analysis needed for the accurate diagnosis of primary 3-oxo-Δ4-steroid 5β-reductase deficiency. J Gastroenterol Hepatol. 2009;24:77685. [DOI] [PubMed]
    Chen JY, Wu JF, Kimura A, Nittono H, Liou BY, Lee CS, et al. AKR1D1 and CYP7B1 mutations in patients with inborn errors of bile acid metabolism: possibly underdiagnosed diseases. Pediatr Neonatol. 2020;61:7583. [DOI] [PubMed]
    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;102:1690703. [DOI] [PubMed] [PMC]
    Li-Hawkins J, Lund EG, Turley SD, Russell DW. Disruption of the oxysterol 7α-hydroxylase gene in mice. J Biol Chem. 2000;275:1653642. [DOI] [PubMed]
    Kimura A, Nittono H, Mizuochi T, Ueki I, Kurosawa T, Muto A, et al. Neonatal cholestasis with increased 3β-monohydroxy-Δ5 bile acids in serum and urine: not necessarily primary oxysterol 7α hydroxylase deficiency. Clin Chim Acta. 2012;413:1700-4. [DOI] [PubMed]
    Mizuochi T, Kimura A, Suzuki M, Ueki I, Takei H, Nittono H, et al. Successful heterozygous living donor liver transplantation for an oxysterol 7α-hydroxylase deficiency in a Japanese patient. Liver Transpl. 2011;17:105965. [DOI] [PubMed]
    Dai D, Mills PB, Footitt E, Gissen P, McClean P, Stahlschmidt J, et al. Liver disease in infancy caused by oxysterol 7α-hydroxylase deficiency: successful treatment with chenodeoxycholic acid. J Inherit Metab Dis. 2014;37:85161. [DOI] [PubMed]
    Verrips A, Hoefsloot LH, Steenbergen GCH, Theelen JP, Wevers RA, Gabreels FJM, et al. Clinical and molecular genetic characteristics of patients with cerebrotendinous xanthomatosis. Brain. 2000;123:90819. [DOI] [PubMed]
    Shimazu K, Kuwabara M, Yoshii M, Kihira K, Takeuchi H, Nakano I, et al. Bile alcohol proffles in bile, urine, and feces of a patient with cerebrotendinous xanthomatosis. J Biochem. 1986;99:47783. [DOI] [PubMed]
    Honda A, Salen G, Matsuzaki Y, Batta AK, Xu G, Leitersdorf E, et al. Differences in hepatic levels of intermediates in bile acid biosynthesis between Cyp27−/− mice and CTX. J Lipid Res. 2001;42:291300. [DOI] [PubMed]
    Honda A, Salen G, Matsuzaki Y, Batta AK, Xu G, Leitersdorf E, et al. Side chain hydroxylations in bile acid biosynthesis catalyzed by CYP3A are markedly up-regulated in Cyp27−/− mice but not in cerebrotendinous xanthomatosis. J Biol Chem. 2001;276:3457985. [DOI] [PubMed]
    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;273:1480512. [DOI] [PubMed]
    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;275:3968592. [DOI] [PubMed]
    Berginer VM, Salen G, Shefer S. Long-term treatment of cerebrotendinous xanthomatosis with chenodeoxycholic acid. N Engl J Med. 1984;311:164952. [DOI] [PubMed]
    Waterham HR, Ferdinandusse S, Wanders RJA. Human disorders of peroxisome metabolism and biogenesis. Biochim Biophys Acta. 2016;1863:92233. [DOI] [PubMed]
    Van Eldere JR, Parmentier GG, Eyssen HJ, Wanders RJ, Schutgens RB, Vamecq J, et al. Bile acids in peroxisomal disorders. Eur J Clin Invest. 1987;17:38690. [DOI] [PubMed]
    Braverman NE, Raymond GV, Rizzo WB, Moser AB, Wilkinson ME, Stone EM, et al. Peroxisome biogenesis disorders in the Zellweger spectrum: an overview of current diagnosis, clinical manifestations, and treatment guidelines. Mol Genet Metab. 2016;117:31321. [DOI] [PubMed] [PMC]
    Baumgart E, Vanhorebeek I, Grabenbauer M, Borgers M, Declercq PE, Fahimi HD, et al. Mitochondrial alterations caused by defective peroxisomal biogenesis in a mouse model for Zellweger syndrome (PEX5 knockout mouse). Am J Pathol. 2001;159:147794. [DOI] [PubMed] [PMC]
    Maxwell M, Bjorkman J, Nguyen T, Sharp P, Finnie J, Paterson C, et al. Pex13 inactivation in the mouse disrupts peroxisome biogenesis and leads to a Zellweger syndrome phenotype. Mol Cell Biol. 2003;23:594757. [DOI] [PubMed] [PMC]
    Keane MH, Overmars H, Wikander TM, Ferdinandusse S, Duran M, Wanders RJA, et al. Bile acid treatment alters hepatic disease and bile acid transport in peroxisome-deficient PEX2 Zellweger mice. Hepatology. 2007;45:98297. [DOI] [PubMed]
    Ferdinandusse S, Jimenez-Sanchez G, Koster J, Denis S, Van Roermund CW, Silva-Zolezzi I, et al. A novel bile acid biosynthesis defect due to a deficiency of peroxisomal ABCD3. Hum Mol Genet. 2015;24:36170. [DOI] [PubMed]
    Platek T, Orso E, Zapala B, Polus A, Kiec-Wilk B, Piwowar M, et al. Case report of dysregulation of primary bile acid synthesis in a family with X-linked adrenoleukodystrophy. Medicine (Baltimore). 2018;97:e13353. [DOI] [PubMed] [PMC]
    Setchell KDR, Heubi JE, Bove KE, O’Connell NC, Brewsaugh T, Steinberg SJ, et al. Liver disease caused by failure to racemize trihydroxycholestanoic acid: gene mutation and effect of bile acid therapy. Gastroenterology. 2003;124:21732. [DOI] [PubMed]
    Van Veldhoven PP, Meyhi E, Squires RH, Fransen M, Fournier B, Brys V, et al. Fibroblast studies documenting a case of peroxisomal 2-methylacyl-CoA racemase deficiency: possible link between racemase deficiency and malabsorption and vitamin K deficiency. Eur J Clin Invest. 2001;31:71422. [DOI] [PubMed]
    Ferdinandusse S, Denis S, Clayton PT, Graham A, Rees JE, Allen JT, et al. Mutations in the gene encoding peroxisomal α-methylacyl-CoA racemase cause adult-onset sensory motor neuropathy. Nat Genet. 2000;24:18891. [DOI] [PubMed]
    Smith EH, Gavrilov DK, Oglesbee D, Freeman WD, Vavra MW, Matern D, et al. An adult onset case of alpha-methyl-acyl-CoA racemase deficiency. J Inherit Metab Dis. 2010;33:34953. [DOI] [PubMed]
    Selkala EM, Kuusisto SM, Salonurmi T, Savolainen MJ, Jauhiainen M, Pirila PL, et al. Metabolic adaptation allows Amacr-deficient mice to remain symptom-free despite low levels of mature bile acids. Biochim Biophys Acta. 2013;1831:133543. [DOI] [PubMed]
    Monte MJ, Alonso M, Briz O, Herraez E, Macias RR, Perez MJ, et al. Identification of a new inborn error in the peroxisomal steps of bile acid synthesis resulting in accumulation of toxic bile acid intermediates and hepatocellular damage. J Hepatol. 2016;64:S301. [DOI]
    Monte MJ, Alonso-Pena M, Briz O, Herraez E, Berasain C, Argemi J, et al. ACOX2 deficiency: an inborn error of bile acid synthesis identified in an adolescent with persistent hypertransaminasemia. J Hepatol. 2017;66:5818. [DOI] [PubMed]
    Vilarinho S, Sari S, Mazzacuva F, Bilguvar K, Esendagli-Yilmaz G, Jain D, et al. ACOX2 deficiency: a disorder of bile acid synthesis with transaminase elevation, liver fibrosis, ataxia, and cognitive impairment. Proc Natl Acad Sci U S A. 2016;113:1128993. [DOI] [PubMed] [PMC]
    Ferdinandusse S, Denis S, van Roermund CWT, Preece MA, Koster J, Ebberink MS, et al. A novel case of ACOX2 deficiency leads to recognition of a third human peroxisomal acyl-CoA oxidase. Biochim Biophys Acta Mol Basis Dis. 2018;1864:9528. [DOI] [PubMed]
    Alonso-Pena M, Espinosa-Escudero R, Herraez E, Briz O, Cagigal ML, Gonzalez-Santiago JM, et al. Beneficial effect of ursodeoxycholic acid in patients with acyl-CoA oxidase 2 (ACOX2) deficiency–associated hypertransaminasemia. Hepatology. 2022;76:125974. [DOI] [PubMed]
    Zhang Y, Chen Y, Zhang Z, Tao X, Xu S, Zhang X, et al. Acox2 is a regulator of lysine crotonylation that mediates hepatic metabolic homeostasis in mice. Cell Death Dis. 2022;13:279. [DOI] [PubMed] [PMC]
    Monte MJ, Espinosa-Escudero R, Alonso-Peña M, Herraez E, Macias RIR, Briz O, et al. Development of in vitro and in vivo models to study ACOX2 deficiency-associated hypertransaminasemia (ADAH). XXVI International Bile Acid Meeting: Bile Acids in Health and Disease; 2022 Jul 8–9; Amsterdam, The Netherlands: 2022. p. 55.
    Ferdinandusse S, Denis S, Mooyer PAW, Dekker C, Duran M, Soorani-Lunsing RJ, et al. Clinical and biochemical spectrum of D-bifunctional protein deficiency. Ann Neurol. 2006;59:92104. [DOI] [PubMed]
    Baes M, Huyghe S, Carmeliet P, Declercq PE, Collen D, Mannaerts GP, et al. Inactivation of the peroxisomal multifunctional protein-2 in mice impedes the degradation of not only 2-methyl-branched fatty acids and bile acid intermediates but also of very long chain fatty acids. J Biol Chem. 2000;275:1632936. [DOI] [PubMed]
    Huyghe S, Schmalbruch H, Hulshagen L, Veldhoven PV, Baes M, Hartmann D. Peroxisomal multifunctional protein-2 deficiency causes motor deficits and glial lesions in the adult central nervous system. Am J Pathol. 2006;168:132134. [DOI] [PubMed] [PMC]
    Ferdinandusse S, Kostopoulos P, Denis S, Rusch H, Overmars H, Dillmann U, et al. Mutations in the gene encoding peroxisomal sterol carrier protein X (SCPx) cause leukencephalopathy with dystonia and motor neuropathy. Am J Hum Genet. 2006;78:104652. [DOI] [PubMed] [PMC]
    Horvath R, Lewis-Smith D, Douroudis K, Duff J, Keogh M, Pyle A, et al. SCP2 mutations and neurodegeneration with brain iron accumulation. Neurology. 2015;85:190911. [DOI] [PubMed] [PMC]
    Fuchs M, Hafer A, Munch C, Kannenberg F, Teichmann S, Scheibner J, et al. Disruption of the sterol carrier protein 2 gene in mice impairs biliary lipid and hepatic cholesterol metabolism. J Biol Chem. 2001;276:4805865. [DOI] [PubMed]
    Chong CPK, Mills PB, McClean P, Gissen P, Bruce C, Stahlschmidt J, et al. Bile acid-CoA ligase deficiency—a new inborn error of bile acid metabolism. J Inherit Metab Dis. 2012;35:52130. [DOI] [PubMed]
    Doege H, Baillie RA, Ortegon AM, Tsang B, Wu Q, Punreddy S, et al. Targeted deletion of FATP5 reveals multiple functions in liver metabolism: alterations in hepatic lipid homeostasis. Gastroenterology. 2006;130:124558. [DOI] [PubMed]
    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;130:125969. [DOI] [PubMed]
    Setchell KDR, Heubi JE, Shah S, Lavine JE, Suskind D, Al-Edreesi M, et al. Genetic defects in bile acid conjugation cause fat-soluble vitamin deficiency. Gastroenterology. 2013;144:94555.e6. [DOI] [PubMed] [PMC]
    Alrehaili BD, Lee M, Takahashi S, Novak R, Rimal B, Boehme S, et al. Bile acid conjugation deficiency causes hypercholanemia, hyperphagia, islet dysfunction, and gut dysbiosis in mice. Hepatol Commun. 2022;6:276580. [DOI] [PubMed] [PMC]
    Heubi JE, Setchell KDR, Jha P, Buckley D, Zhang W, Rosenthal P, et al. Treatment of bile acid amidation defects with glycocholic acid. Hepatology. 2015;61:26874. [DOI] [PubMed] [PMC]
    Ranea-Robles P, Chen H, Stauffer B, Yu C, Bhattacharya D, Friedman SL, et al. The peroxisomal transporter ABCD3 plays a major role in hepatic dicarboxylic fatty acid metabolism and lipid homeostasis. J Inherit Metab Dis. 2021;44:141933. [DOI] [PubMed] [PMC]