Review
Nuclear receptors CAR and PXR: Molecular, functional, and biomedical aspects

https://doi.org/10.1016/j.mam.2009.04.002Get rights and content

Abstract

Nuclear receptors (NRs) are ligand-activated transcription factors sharing a common evolutionary history and having similar sequence features at the protein level. Selective ligand(s) for some NRs is not known, therefore these NRs have been named “orphan receptors”. Whenever ligands have been recognized for any of the orphan receptor, it has been categorized and grouped as “adopted” orphan receptor. This group includes the constitutive androstane receptor (CAR) and the pregnane X receptor (PXR). They function as sensors of toxic byproducts derived from endogenous metabolites and of exogenous chemicals, in order to enhance their elimination. This unique function of CAR and PXR sets them apart from the steroid hormone receptors. The broad response profile has established that CAR and PXR are xenobiotic sensors that coordinately regulate xenobiotic clearance in the liver and intestine via induction of genes involved in drug and xenobiotic metabolism. In the past few years, research has revealed new and mostly unsuspected roles for CAR and PXR in modulating hormone, lipid, and energy homeostasis as well as cancer and liver steatosis. The purpose of this review is to highlight the structural and molecular bases of CAR and PXR impact on human health, providing information on mechanisms through which diet, chemical exposure, and environment ultimately impact health and disease.

Introduction

Living organisms are severely xenophobes in that they have an aversion to xenobiotics (from Greek, ξενoς “foreign” and βιoς “life”). This term was coined to cover all organic compounds that were foreign in the organism under study. The history of xenobiotics is interwoven with the birth of organic chemistry. In the early 1800s, the prevailing belief was that the composition of the human body, and indeed of all living things, was the result of a “vital force” or “internal flame” and that mere mortals were incapable of understanding these workings and would be particularly unable to synthesize constituent compounds of the human body (Murphy, 2001). Faced with this daunting concept, scientists were discouraged from thinking about the processes that dominated everyday digestion and nutrition. In 1828, Wöhler was able to shatter this myth of synthesizing urea, one of the compounds that he had examined in his studies on the urinary waste products, and sent the following note to Berzelius “I must tell you that I can prepare urea without requiring a kidney of an animal, either man or dog” (Whöler and Frerichs, 1848). In modern-day terminology, this was a true “paradigm shift” and led to revised thinking on what chemists could accomplish with regard to “organic” compounds. Wöhler test led to the first metabolism experiments performed in 1841 by Ure who demonstrated the conjugation of benzoic acid with glycine in man (Ure, 1841, Keller, 1842). The transformation of compounds such as cinnamic acid (Erdmann and Marchand, 1842) and benzaldehyde (Whöler and Frerichs, 1848) to benzoic acid, further emphasized by the discovery of the conversion of benzene to phenol (Schultzen and Naunyn, 1867), indicated the ability of the body to oxidize ingested compounds. A rapid development of chemical industries occurred soon after World War II, resulting in the manufacturing of a large number of chemical products. The increased use of fertilizers, insecticides, and herbicides led to a dramatic increase in world food production. This coupled with improved medicine and pharmacological science helped to improve general public health rising life expectancy to an average of 75 years (Yu, 2001). Whereas many people of the world were enjoying the benefits of technological and economic expansion and higher living standards, others perceived that these extraordinary developments were not without cost. Indeed, the impact of global environmental changes on human health has become a great concern. As early as the 1950s and 1960s, many urban dwellers and residents in the vicinity of industrial factories began to recognize undesirable changes in the environment (Yu, 2001).

Now it is clear that after a chemical pollutant gets into a mammalian organism, chemical reactions occur within the body, altering its reactivity and structural properties. The tome “Detoxication Mechanisms” by Williams (1947) heralded the birth of xenobiotic/drug metabolism as a distinct branch of science, providing a systematic approach based on the organic chemistry classification. Williams also expanded on his concepts of the principal biochemical reactions whereby drugs and other foreign compounds are metabolized in the body. Most importantly, he proposed that foreign compounds were metabolized in two distinct phases: one including oxidation, reduction, and hydrolysis processes, and the other involving conjugation reactions.

In the early 1950s, biochemical research unveiled the mechanisms of a wide variety of transformation reactions. The pioneering studies of James and Elizabeth Miller and Axelrod revealed the subcellular localization of metabolism reactions and brought to the identification of enzymes responsible for catalytic transformations (Mueller and Miller, 1949, Axelrod, 1955).

A major class of catalytic oxidative transformations was initially characterized in 1955 by Hayashi and Mason. Since these reactions need an oxidant (molecular oxygen) and a reductant (NAD(P)H), the enzymes catalyzing these processes were trivially named “mixed-function oxidases” (Hayashi et al., 1955, Mason et al., 1955).

The understanding of the biochemical nature of chemical transformations occurring in living organisms grew out from early studies on the liver pigments by Garfinkel and Klingenberg in 1958. These researchers observed in liver microsomes an unusual carbon monoxide-binding pigment with an absorbance maximum at 450 nm (Garfinkel, 1958, Klingenberg, 1958). This pigment was ultimately characterized as a cytochrome by Omura and Sato (1962). The function of this unique cytochrome (called P450; CYP) was initially revealed in 1963 by Estabrook, Cooper, and Rosenthal, while investigating the role of microsomes from the adrenal cortex in the catalytic conversion (i.e., hydroxylation) of 17-hydroxyprogesterone to deoxycorticosterone (Estabrook et al., 1963). In the clinical setting, the discovery of the CYP polymorphism role in debrisoquine and sparteine metabolism emphasized the importance of individual response to drugs (Thomas et al., 1976, Mahgoub et al., 1977, Eichelbaum et al., 1978). At present, 18 CYP families have been identified in mammals, among which only three are primarily responsible for most xenobiotic metabolism. Determining the molecular bases for the xenobiotic-mediated induction of CYP gene expression has become one of the most challenging dilemmas of modern toxicology (Dickins, 2004, Xu et al., 2005, Xu et al., 2005, Nakata et al., 2006, Plant, 2007, Graham and Lake, 2008).

Thus, the threat represented by xenobiotics for the organism is the target of a complex strategy. Metabolism of drugs and other xenobiotics in the liver is the body primary defence against accumulation of potentially toxic lipophilic compounds. This strategy comprises xenosensors, dedicated to the recognition of xenobiotics and dangerous endogenous molecules, as well as xenobiotic metabolizing and transporters systems. The first step of the detoxification mechanism is the xenobiotic detection by the constitutive androstane receptor (CAR) and/or the pregnane X receptor (PXR), both belonging to the nuclear receptor (NR) super-family (Fuhr, 2000, Dixit et al., 2005, Chang and Waxman, 2006, Moreau et al., 2008), and/or by the aryl hydrocarbon receptor (AhR), a member of the Per/ARNT/Sim (PAS) family of transcription factors (Francis et al., 2003).

CAR and PXR are activated by a variety of endogenous (i.e., steroids and bile acid salts), and exogenous ligands including drugs, insecticides, pesticides, and nutritional compounds (Fig. 1). Currently, it remains uncertain whether these receptors display a high ligand specificity or instead function as more generalized steroid/xenobiotic sensors. These xenosensors coordinate the expression of several genes encoding CYPs. Both CAR and PXR are able to elicit alterations in xeno- and endo-biotic metabolism, regulating the expression of phases I and II metabolizing enzymes and phase III transporters (Fig. 2) (Matic et al., 2007, Plant, 2007, Timsit and Negishi, 2007, Tompkins and Wallace, 2007, Graham and Lake, 2008, Lim and Huang, 2008, Ma and Lu, 2008, Moreau et al., 2008, Pascussi et al., 2008, Teng and Piquette-Miller, 2008, Köhle and Bock, 2009, Zhou et al., 2009). Although CAR and PXR play physiological roles (e.g., regulate cholesterol elimination pathways), no physiological ligands have been identified, therefore they still remain orphan NRs (Timsit and Negishi, 2007).

In contrast, AhR ligands are the hydrophobic environmental pollutants of polyhalogenated aromatic hydrocarbons, such as polychlorinated dibenzodioxins, dibenzofurans and coplanar biphenyls or polycyclic aromatic hydrocarbons (e.g., benzo[a]pyrene, 3-methylcholanthrene, benzoflavones, and omeprazole) (Francis et al., 2003, Nguyen and Bradfield, 2008, Gomez-Duran et al., 2009, Hahn et al., 2009).

Human CAR (hCAR), initially identified as the orphan receptor MB67 (Baes et al., 1994), was found as a heterodimer partner for the retinoic acid X receptor (RXR) (see Tompkins and Wallace, 2007, Graham and Lake, 2008, Lim and Huang, 2008, Ma and Lu, 2008, Moreau et al., 2008, Pascussi et al., 2008, Teng and Piquette-Miller, 2008, Köhle and Bock, 2009, Zhou et al., 2009). hCAR was able to bind to retinoic acid response elements (REs), but the extent of activation was much lower than that mediated by RXR. The isolation and characterization of mouse CAR (mCAR) was reported in 1997 (Choi et al., 1997); at that time, it was hypothesized to target a subset of retinoid acid REs, but its role in CYP regulation has not yet been realized (Stanley et al., 2006).

Meanwhile, separate lines of investigation were leading to the identification of mouse PXR (mPXR) that was identified in 1998 by using an expressed sequence tag to screen a mouse liver library (Kliewer et al., 1998). mPXR was named “pregnane X receptor” since it was shown to be activated by derivatives of dexamethasone and pregnenolone (Kliewer et al., 1998).

At approximately the same time, the human steroid X receptor was cloned in studies focused on the identification of the Xenopus laevis benzoate X receptor homologues. The human steroid X receptor was shown to be related to CAR and to the vitamin D receptor (VDR) (Blumberg et al., 1998), and then it was established to be the human homologue of mPXR (Lehmann et al., 1998). A parallel computer modeling approach confirmed that the human steroid X receptor is the human homologue of mPXR (Bertilsson et al., 1998). According to the literature (Bertilsson et al., 1998), the human steroid X receptor has been renamed human PXR (hPXR).

One of the most intriguing hypotheses to justify the existence of NR xenosensors was built on the basis of studies dealing with the broad implications of these receptors in steroid hormone homeostasis. It is intriguing to hypothesize that an ancestral protein could act as the endogenous metabolite (endobiotics) sensor. Starting from this hypothetical receptor progenitor, evolution proceeded by ligand exploitation and serial genome expansions. If this hypothesis is true, the conceptual distance between specific endogenous hormones and exogenous environmental molecules is probably closer than previously thought (Buters et al., 1994, Zhai et al., 2007).

NRs form a super-family of ligand-activated transcription factors implicated in various physiological functions from development to homoeostasis (Escriva et al., 2004, Ascenzi et al., 2006). NRs share a common evolutionary history as revealed by their conserved structure and by their high degree of sequence conservation (Escriva et al., 2004). The phylogenetic analysis of the NR super-family brought to its sub-division into six sub-families of unequal size (Fig. 3): (i) the large sub-family I contains the thyroid hormone receptors (THRs), the retinoic acid receptors (RARs), VDR, the ecdysone receptor, the peroxisome proliferator-activated receptors (PPARs), as well as numerous orphan receptors including CAR and PXR; (ii) the sub-family II includes RXR, the chicken ovalbumin upstream promoter-transcription factor (COUP-TF), and the hepatocyte nuclear factor 4 (HNF4); (iii) the steroid receptor sub-family III includes the estrogen receptors (ERs), the estrogen-related receptors (ERRs), the glucocorticoid receptors (GRs), the mineralocorticoid receptor (MR), the progesterone receptor (PR), as well as the androgen receptors (ARs); (iv) the sub-family IV contains the nerve growth factor inducible I-B group of orphan receptors (NGF1B); (v) the sub-family V includes the steroidogenic factor-1 (SF-1) and the Drosophila melanogaster “fushi tarazu” factor-1 receptor (FTZ-F1); and (vi) the small sub-family VI contains only the germ cell nuclear factor-1 receptor (GCNF) (Laudet et al., 1999, Escriva et al., 2004, Ascenzi et al., 2006).

A hypothetical evolutionary path that might have been taken by the first NR in the early Metazoan is represented in Fig. 3. Since several sub-families were present in early Metazoan, it appears that the super-family underwent an “explosive expansion” during early Metazoan evolution. The diversification of the super-family followed two waves of gene duplication. The first wave before the Protostome/Deuterostome split, during the emergence of Metazoan, led to the acquisition of the present six sub-families and the various groups of receptors within each sub-family. The second wave occurred after the arthropod/vertebrate split, specifically in vertebrates, producing the paralogous groups within each sub-family (see Laudet et al., 1999, Escriva et al., 2004, Ascenzi et al., 2006). In support of NR paralogous through genome or block duplications, mapping studies have demonstrated the presence of extensive “paralogy groups” which include NRs on different chromosomes (Owen and Zelent, 2000).

The first NR arose as one unit consisting of the currently recognized DNA-binding domain (DBD) and ligand-binding domain (LBD), based on the finding that the two domains existed together in lower Metazoan. However, such observation does not entirely dismiss the possibility that the two regions existed independently earlier in evolution, despite the fact that strong similarity to either domain has not yet been observed outside the Metazoan kingdom (Owen and Zelent, 2000). The ancestor DBD may have been similar to that of modern receptors, the LBD being present but not possessing a transactivation domain or the ability to dimerize (Owen and Zelent, 2000). The possibility that the VDR could be a chimera is in accordance with its clustering into different families based on the phylogenetic trees compiled from either DBD or LBD sequence (Laudet et al., 1999). The DBD of the VDR closely resembles that of FTZ-F1, while the LBD is closer to that of the RAR grouping. However, the more recently characterized CAR, PXR, and liver X receptor (LXR), along with several arthropod receptors, cluster together with the VDR in phylogenetic trees based on both the DBD and LBD, thus throwing doubt on the VDR being a chimera (Owen and Zelent, 2000).

Convergent evolution may be in part responsible for this phenomenon which is observed in other receptors such as GCNF and NGF1B. The possibility that each of these receptors diverged from a common chimeric ancestor still remains and a mechanism for such an event, originally suggested by Laudet et al. (1999), is represented in Fig. 3. In constructing the evolutionary tree represented in Fig. 3, three assumptions have been made: (i), the earliest Metazoan possessed COUP-TF, RXR, and FTZ-F1, all of which are well conserved to the present day and thus are close to the NR ancestor of subfamilies II and V; (ii) the ability to heterodimerize with RXR arose once and has diversified through sub-families I, II, and IV; and (iii) there is no correlation between the ligand recognized by a given receptor and its position in the family tree (see Laudet, 1997). The presence of RXR in the earliest branches of the Metazoan enables the extrapolation that an archaic family II member gave rise to families I and IV, which can also dimerize with RXR. It could be assumed from data in Cnidaria that a direct RXR ancestor developed the ability to bind the ligand 9-cis-retinoic acid with a subset diverging in RAR, thus acquiring the ability to bind all trans-retinoic acids (Owen and Zelent, 2000).

The gene comparison among human, chimpanzee, mouse, and rat suggests that the high degree of conservation within members of the NR super-family may have arisen from negative evolution selection against changes in the protein amino acid sequence (Clark et al., 2003, Zhang et al., 2004a). Indeed, the amino acid sequence identity between human and mouse ortholog NR genes is typically greater than 95% and 85% in the DBD and in the LBD, respectively (Zhang et al., 2004a, Iyer et al., 2006). The only two exceptions of such conservation are CAR-LBD and PXR-LBD, both characterized by a high cross-species amino acid sequence difference, indicating divergence during the course of evolution (Fig. 3) (Jones et al., 2000, Moore et al., 2002, Zhang et al., 2004a, Krasowski et al., 2005). Indeed, CAR-LBD and PXR-LBD show considerable differences in ligand recognition (Watkins et al., 2001, Watkins et al., 2003a, Watkins et al., 2003b, Shan et al., 2004, Suino et al., 2004, Xu et al., 2004, Chrencik et al., 2005, Xue et al., 2007a, Xue et al., 2007b, Teotico et al., 2008, Wang et al., 2008).

The percentage of identity among CAR and PXR expressed in several vertebrate species has been calculated by aligning their complete amino acidic sequences, and are summarized in Table 1, Table 2, respectively. While hCAR and mCAR share approximately 93.9% amino acid identity in their whole sequence (Table 1), only approximately 70% amino acid identity has been observed in their LBDs (Escriva et al., 2002, Robinson-Rechavi et al., 2004). Moreover marked differences have been observed in their responses to xenobiotics. Indeed, clotrimazole is an efficacious deactivator of hCAR but has little or no effect on the mCAR activity. Conversely, 3,5-dichloro-2-{4-[(3,5-dichloropyridin-2yl)oxy]phenoxy}pyridine (TCPOBOP) is a potent activator of mCAR but lacks any activity on hCAR. The divergence in the amino acid sequence across CAR orthologs undoubtedly contributes to cross-species differences in the physiological effects of xenobiotics (Moore et al., 2000a).

The analysis of the complete amino acid sequence of PXR of different species indicates an identity of 76.5% and 76% between hPXR and mouse and rat PXRs, respectively (Table 2). Similarly to CAR, the human, rabbit, mouse, and rat PXRs are all roughly equally divergent, sharing only approximately 80% amino acid identity in their LBDs (Moore et al., 2002, Zhang et al., 2004a). Indeed, although both the rabbit and human PXR are activated by rifampicin, there are marked differences in their responsiveness to synthetic steroids such as dexamethasone and cyproterone, as indicated by the level of CYPs gene expression in vivo (Jones et al., 2000). The PXR divergence represents an important component of cross-species differences in the regulation of CYP3A expression by xenobiotics, and could correspond to either an adaptive response to different environmental xenobiotic challenges or the different specificities of natural PXR ligands within vertebrate species. Despite their divergence, several lines of evidence suggest that PXRs are orthologs and not closely related members of sub-family 1I of NRs (Jones et al., 2000).

The cross-species variation in CAR-LBD and PXR-LBD is even more striking when DNA sequences are compared; in fact, the ratio between the rate of non-synonymous (nucleotidic substitutions that result in amino acid replacement) and synonymous (nucleotidic substitutions that do not cause an amino acid replacement) nucleotide variation provides information concerning evolutionary selective forces acting on a given gene (Yang and Bielawski, 2000). The analysis of this parameter strongly suggests that the natural selection has favored sequence diversity in CAR-LBD and PXR-LBD, possibly to adapt to cross-species differences in ligand recognition. CAR and PXR may thus represent unusual examples of NR genes that have changed their ligand specificity across vertebrate species to adapt to cross-species differences in the recognition of exogenous and/or endogenous toxic compounds (Kliewer et al., 1998, Schuetz et al., 2001, Xie et al., 2001, Moore et al., 2002, Krasowski et al., 2005, Reschly and Krasowski, 2006).

The different CAR and PXR ligand recognition properties constitute an especially unusual finding in the NR super-family, representing the more extreme divergence of ligand-binding residues of any of the ligand-activated NRs in vertebrates (Krasowski et al., 2005, Reschly and Krasowski, 2006). Indeed, residues involved in ligand binding are strongly conserved in other NRs as demonstrated by VDR, an endocrine NR closely related to PXR; only four residues are not conserved across vertebrate species, ranging from sea lamprey to human (Reschly and Krasowski, 2006).

The origin and evolution of the NR1I sub-family (which currently includes VDR, CAR, and PXR) is still a matter of debate, although it appears probable that a single NR gene duplicated early in vertebrate evolution gave origin to VDR and PXR genes (Reschly and Krasowski, 2006). These two genes then diverged from each other, and additional duplications have resulted in multiple VDR and PXR genes in some non-mammals species. Interestingly, distinct CAR and PXR genes appear to be solely found in mammals (Reschly and Krasowski, 2006). Indeed, the chicken has only a single “xenobiotic-responsive” NR1I gene, currently classified as CAR, the product of which has properties similar to both CAR and PXR (Fig. 3) (Handshin et al., 2004). A single “xenobiotic-responsive” gene is also present in Danio renio (Fig. 3) and Xenopus tropicalis (Reschly and Krasowski, 2006). A likely explanation is that an ancestral gene, similar to that of chicken CAR, duplicated just prior to or early in mammalian evolution. The two genes then diverged from each other to become the modern-day CAR and PXR genes found in all mammalian genomes sequenced so far (including opossum, seal, dog, pig, mouse, rat, and rhesus monkey) (Reschly and Krasowski, 2006).

The different amino acid sequences of vertebrate CAR-LBD and PXR-LBD contrast with the fact that non-synonymous substitutions are rare in hCAR and hPXR (Lamba et al., 2005). Interestingly, the hCAR and hPXR sequencing of 70 and 100 individuals from three different ethnic groups, respectively, shows very low nucleotide diversity (lower than the genome-wide averages for human genes) and no non-synonymous substitutions in the LBD coded by either gene (Zhang et al., 2001, Thompson et al., 2005). Overall, NR1I sub-family members show little variation in the LBD between human individuals. In addition, the nucleotide divergence between human and chimpanzee CAR and PXR is lower than the average of the nucleotide divergence for other genes in the human genome (Ebersberger et al., 2002, Clark et al., 2003, Thompson et al., 2005). This suggests that CAR and PXR ligands, relevant at least in terms of influencing reproductive fitness, do not vary between humans, and perhaps not even between humans and primates, but do vary between different animals (Ebersberger et al., 2002, Clark et al., 2003, Thompson et al., 2005).

Section snippets

Gene organization

A common paradigm for many genes, including those of NRs, is the generation of several splice variants determined by differential promoter usage, insertion or deletion of amino acid residues, shift of the reading frame, and/or introduction of premature termination codons (Lander et al., 2001, Venter, 2001). Approximately 80% of alternative splicing results in protein isoforms that could vary in structure and functional properties (Modrek and Lee, 2002, Lamba et al., 2004a). In addition, single

CAR and PXR expression patterns

Due to their role in protecting living organisms from exogenous insults, the expression of xenosensors CAR and PXR at the points of xenobiotic entry and metabolism into the body is not unexpected. In fact, CAR is highly expressed in the liver and in the epithelial cells of the small intestine villi; moreover, CAR is expressed at low levels in both mouse and human heart, skeletal muscle, brain, kidney, and in human lung (Baes et al., 1994, Choi et al., 1997, Doherty and Charman, 2002, Swales and

CAR and PXR functions

Human and animals are exposed to potentially toxic chemicals from both endogenous and foreign sources including bile salts, cholesterol and oxysterols, steroid hormones, bilirubin and fatty acids (endobiotics), as well as toxins, carcinogens, pollutants, drugs, dietary components, and herbal remedies (xenobiotics). To counter toxic insults and to maintain the homeostatic balance in important metabolic pathways, defense systems have been developed comprising enzymes and transport proteins

PXR-mediated drug–drug interactions

The most common clinical implication for the activation of PXR is the occurrence of drug–drug interactions especially in patients affected by tuberculosis, cancer, AIDS, cardiovascular diseases, and diabetes. The clinical consequences of drug–drug interactions are generally a decreased therapeutic efficacy and, occasionally, an increased drug toxicity. Therefore, understanding the mechanism(s) involved in drug–drug interactions represents an important goal for the improvement of drug safety.

Concluding remarks and future directions

CAR (also called MB67) was originally described in 1994 as a new orphan member of the NR super-family predominantly expressed in the liver (Baes et al., 1994). Remarkably, an important role of CAR in the complex network of proteins involved in the response to retinoic acid and its metabolites was proposed. At the start of the new century the interest for this receptor was very low (just 27 articles cited in PubMed data base from 1994 to January 2000) to grow up exponentially in the recent years

Acknowledgements

The authors wish to thank Dr. Alessandro Bolli, Dr. Fabio Polticelli, and Dr. Daniele Salvi (University Roma Tre, Roma, Italy) for helpful discussions. This work was partially supported by grants from the Ministry for Health of Italy (National Institute for Infectious Diseases I.R.C.C.S. “Lazzaro Spallanzani”, Roma, Italy, Ricerca corrente 2008 to P.A.) and from the Ministry of Education, University, and Research of Italy (CLAR 2008 to P.A.).

References (341)

  • A. Dobrzyn et al.

    The role of stearoyl-CoA desaturase in the control of metabolism

    Prostaglandins Leukot Essent Fatty Acids

    (2005)
  • M.M. Dring et al.

    The pregnane X receptor locus is associated with susceptibility to inflammatory bowel disease

    Gastroenterology

    (2006)
  • Z. Duniec-Dmuchowski et al.

    Regulation of CYP3A4 and CYP2B6 expression by liver X receptor agonists

    Biochem. Pharmacol.

    (2007)
  • I. Ebersberger et al.

    Genomewide comparison of DNA sequences between humans and chimpanzees

    Am. J. Hum. Genet.

    (2002)
  • P.F. Egea et al.

    Ligand–protein interactions in nuclear receptors of hormones

    FEBS Lett.

    (2000)
  • J.L. Fayer et al.

    A novel testosterone 6 beta-hydroxylase activity assay for the study of CYP3A-mediated metabolism, inhibition, and induction in vitro

    J. Pharmacol. Toxicol. Methods

    (2001)
  • C. Frank et al.

    Characterization of DNA complexes formed by the nuclear receptor constitutive androstane receptor

    J. Biol. Chem.

    (2003)
  • C. Frank et al.

    Identification of pregnane X receptor binding sites in the regulatory regions of genes involved in bile acid homeostasis

    J. Mol. Biol.

    (2005)
  • R.T. Gampe et al.

    Asymmetry in the PPARγ/RXRα crystal structure reveals the molecular basis of heterodimerization among nuclear receptors

    Mol. Cell.

    (2000)
  • V. Ganapathy et al.

    Role of transporters in placental transfer of drugs

    Toxicol. Appl. Pharmacol.

    (2005)
  • L.G. Ganem et al.

    Phenobarbital induction of CYP2B1/2 in primary hepatocytes: endocrine regulation and evidence for a single pathway for multiple inducers

    Toxicol. Appl. Pharmacol.

    (1999)
  • D. Garfinkel

    Studies on pig liver microsomes. I. Enzymic and pigment composition of different microsomal fractions

    Arch. Biochem. Biophys.

    (1958)
  • S. Gerbal-Chaloin et al.

    Transcriptional regulation of CYP2C9 gene. Role of glucocorticoid receptor and constitutive androstane receptor

    J. Biol. Chem.

    (2002)
  • A. Gomez-Duran et al.

    Fitting a xenobiotic receptor into cell homeostasis: how the dioxin receptor interacts with TGFbeta signaling

    Biochem. Pharmacol.

    (2009)
  • M.J. Graham et al.

    Induction of drug metabolism: species differences and toxicological relevance

    Toxicology

    (2008)
  • G.L. Guo et al.

    Complementary roles of farnesoid X receptor, pregnane X receptor, and constitutive androstane receptor in protection against bile acid toxicity

    J. Biol. Chem.

    (2003)
  • M.E. Hahn et al.

    Regulation of constitutive and inducible AHR signaling: complex interactions involving the AHR repressor

    Biochem. Pharmacol.

    (2009)
  • S. Harmsen et al.

    The role of nuclear receptors in pharmacokinetic drug–drug interactions in oncology

    Cancer Treat. Rev.

    (2007)
  • E.L. Haughton et al.

    Pregnane X receptor activators inhibit human hepatic stellate cell transdifferentiation in vitro

    Gastroenterology

    (2006)
  • P. Honkakoski et al.

    Effects of triaryl phosphates on mouse and human nuclear receptors

    Biochem. Pharmacol.

    (2004)
  • C.H. Hurst et al.

    Environmental phthalate monoesters activate pregnane X receptor-mediated transcription

    Toxicol. Appl. Pharmacol.

    (2004)
  • A.K. Agrawal et al.

    Phenobarbital induction of hepatic CYP2B1 and CYP2B2: pretranscriptional and post-transcriptional effects of gender, adult age, and phenobarbital dose

    Mol. Pharmacol.

    (1996)
  • D. Argaud

    Inhibition of gluconeogenesis in isolated rat hepatocytes after chronic treatment with phenobarbital

    Biochem. J.

    (1991)
  • K.A. Arnold et al.

    Alternative splicing affects the function and tissue-specific expression of the human constitutive androstane receptor

    Nucl. Recept.

    (2004)
  • S.S. Auerbach et al.

    Alternatively spliced isoforms of the human constitutive androstane receptor

    Nucl. Acid. Res.

    (2003)
  • J.T. Backman et al.

    Rifampin drastically reduces plasma concentrations and effects of oral midazolam

    Clin. Pharmacol. Ther.

    (1996)
  • M. Baes et al.

    A new orphan member of the nuclear hormone receptor superfamily that interacts with a subset of retinoic acid response elements

    Mol. Cell. Biol.

    (1994)
  • B. Bauer et al.

    Pregnane X receptor up-regulation of P-glycoprotein expression and transport function at the blood–brain barrier

    Mol. Pharmacol.

    (2004)
  • J. Behravan et al.

    Drug transport across the placenta, role of the ABC drug efflux transporters

    Expert Opin. Drug Metab. Toxicol.

    (2007)
  • G. Bertilsson et al.

    Identification of a human nuclear receptor defines a new signaling pathway for CYP3A induction

    Proc. Natl. Acad. Sci. USA

    (1998)
  • D. Blizard et al.

    Mechanism of induction of cytochrome p450 enzymes by the proestrogenic endocrine disruptor pesticide-methoxychlor: interactions of methoxychlor metabolites with the constitutive androstane receptor system

    Drug Metab. Dispos.

    (2001)
  • B. Blumberg et al.

    SXR, a novel steroid and xenobiotic-sensing nuclear receptor

    Genes Dev.

    (1998)
  • W. Bourguet et al.

    Crystal structure of the ligand-binding domain of the human nuclear receptor RXR-alpha

    Nature

    (1995)
  • U. Brinkmann et al.

    Polymorphisms in the ABC drug transporter gene MDR1

    Pharmacogenom. J.

    (2001)
  • A.M. Brzozowski et al.

    Molecular basis of agonism and antagonism in the oestrogen receptor

    Nature

    (1997)
  • O. Burk et al.

    Antimalarial artemisinin drugs induce cytochrome P450 and MDR1 expression by activation of xenosensors pregnane X receptor and constitutive androstane receptor

    Mol. Pharmacol.

    (2005)
  • J.T.M. Buters et al.

    cDNA-directed expression of human cytochrome-p450 cyp3a4 using baculovirus

    Drug Metab. Dispos.

    (1994)
  • V.E. Carnahan et al.

    Structure and function of the human nuclear xenobiotic receptor PXR

    Curr. Drug Metab.

    (2005)
  • T.K. Chang et al.

    Synthetic drugs and natural products as modulators of constitutive androstane receptor (CAR) and pregnane X receptor (PXR)

    Drug Metab. Rev.

    (2006)
  • T.K.H. Chang et al.

    Effect of ovariectomy and androgen on phenobarbital induction of hepatic CYP2B1 and CYP2B2 in Sprague-Dawley rats

    Drug Metab. Dispos.

    (1997)
  • Cited by (251)

    • Nuclear Receptors

      2022, Comprehensive Pharmacology
    View all citing articles on Scopus
    1

    These authors contributed equally to this work.

    View full text