Skip to main content
Nuclear Receptor Signaling logoLink to Nuclear Receptor Signaling
. 2007 Nov 30;5:e012. doi: 10.1621/nrs.05012

The mineralocorticoid receptor: insights into its molecular and (patho)physiological biology

Say Viengchareun 1, Damien Le Menuet 1, Laetitia Martinerie 1, Mathilde Munier 1, Laurent Pascual-Le Tallec 1, Marc Lombès 1,
PMCID: PMC2121322  PMID: 18174920

Abstract

The last decade has witnessed tremendous progress in the understanding of the mineralocorticoid receptor (MR), its molecular mechanism of action, and its implications for physiology and pathophysiology. After the initial cloning of MR, and identification of its gene structure and promoters, it now appears as a major actor in protein-protein interaction networks. The role of transcriptional coregulators and the determinants of mineralocorticoid selectivity have been elucidated. Targeted oncogenesis and transgenic mouse models have identified unexpected sites of MR expression and novel roles for MR in non-epithelial tissues. These experimental approaches have contributed to the generation of new cell lines for the characterization of aldosterone signaling pathways, and have also facilitated a better understanding of MR physiology in the heart, vasculature, brain and adipose tissues. This review describes the structure, molecular mechanism of action and transcriptional regulation mediated by MR, emphasizing the most recent developments at the cellular and molecular level. Finally, through insights obtained from mouse models and human disease, its role in physiology and pathophysiology will be reviewed. Future investigations of MR biology should lead to new therapeutic strategies, modulating cell-specific actions in the management of cardiovascular disease, neuroprotection, mineralocorticoid resistance, and metabolic disorders.

A brief history

In the late 1960s, evidence for the presence of specific receptors mediating corticosteroid action in the toad bladder was initially proposed by the group of Edelman [Porter and Edelman, 1964]. Subsequently, Type I and Type II corticosteroid receptors were described and identified as mineralocorticoid (MR) and glucocorticoid receptors (GR) [Marver et al., 1974]. MR was characterized as a high affinity (Kd~1 nM), low capacity (20-50 fmol/mg protein) receptor and demonstrated to be a major regulator of sodium reabsorption in the kidney [Funder et al., 1972]. Fifteen years later, the human MR (hMR) cDNA was cloned by the Evans laboratory by screening a human kidney cDNA library at low stringency with a probe encompassing the DNA binding domain of the GR [Arriza et al., 1987]. MR was subsequently cloned and characterized in many species including Xenopus, fish (zebra fish, teleost fish [Greenwood et al., 2003], rainbow trout [Sturm et al., 2005]), bird [Hodgson et al., 2007; Porter et al., 2007] and mammals (mouse, rat [Patel et al., 1989], mole, pig, cow, monkey [Patel et al., 2000; Pryce et al., 2005]). In the late 1990s came the identification of multiple transcription coregulators that mediate MR transcriptional potency at aldosterone-target genes, as reviewed by [O'Malley, 2007]. MR is now recognized as a crucial transcription factor involved in many physiological processes and pathological disorders.

Structure

Structure of the gene

The gene NR3C2 encoding the hMR is located on chromosome 4 in the q31.1 region and spans approximately 450 kb [Morrison et al., 1990; Zennaro et al., 1995]. As illustrated in Figure 1, the gene is composed of ten exons; the first two exons, 1α and 1β, are untranslated, and the following eight exons encode the entire MR protein of 984 amino acids (aa). The rat MR gene is located on chromosome 19q11 and differs slightly in having three untranslated exons (1α, 1β and 1γ) and encoding a 981 aa protein [Kwak et al., 1993]; a similar genomic structure is found for mouse MR gene, which encodes a 978 aa protein. In addition, it now appears that the MR gene does not encode only one protein, but gives rise to multiple mRNA isoforms and protein variants [Pascual-Le Tallec and Lombes, 2005], thus allowing combinatorial patterns of receptor expression potentially responsible for distinct cellular and physiological responses in a tissue-specific manner.

Figure 1. Schematic representation of human MR structure.

Figure 1

MR gene, mRNA, protein, functional domains and associated posttranslational modifications are depicted. The hMR gene is composed of ten exons, including two untranslated first exons (1α and 1β). The AUG translational initiation start codon is located 2 bp after the beginning of exon 2, while the stop codon is located in exon 9. Multiple mRNA isoforms generated by alternative transcription or splicing events are translated into various protein variants, including those generated by utilization of alternative translation initiation sites (not shown). The receptor is comprised of distinct functional domains (activation function AF-1a, AF-1b and AF-2) and nuclear localization signals (NLS0, NLS1 and NSL2), as well as one nuclear export signal (NES). The positioning of amino acids targeted for phosphorylation, sumoylation, acetylation and ubiquitylation is indicated for the human MR sequence.

Structure of the protein

Like all members of the nuclear receptor superfamily, MR has three major functional domains; a N-terminal domain (NTD), followed by a central DNA-binding domain (DBD), and a hinge region linking them to a C-terminal ligand-binding domain (LBD). Exon 2 encodes most of the NTD, small exons 3 and 4 for each of the two zinc fingers of the DBD, and the last five exons for the LBD (Figure 1).

The MR NTD is the longest among all the steroid receptors (SR), (602 aa). The NTD is highly variable among SR, showing less than 15% identity, but for a given receptor, highly conserved between species (more than 50% homology), strongly suggesting a crucial functional importance. The NTD possesses several functional domains responsible for ligand-independent transactivation or transrepression, as shown schematically in Figure 1. Two distinct activation function 1 domains (AF1), referred to as AF1a (residues 1-167) and AF1b (residues 445-602), have been demonstrated in both rat [Fuse et al., 2000] and human MR [Pascual-Le Tallec et al., 2003]. A central inhibitory domain (residues 163-437) has also been characterized and seems to be sufficient to attenuate the overall transactivation strength of the NTD fused either to AF-1a or AF-1b [Pascual-Le Tallec et al., 2003]. These different domains of the NTD recruit various coregulators responsible for modulating the transcriptional activity of MR in a highly selective manner compared with other SR, and are now considered to be important determinants of mineralocorticoid selectivity [Pascual-Le Tallec and Lombes, 2005].

The DBD has the ability to recognize specific target DNA sequences or hormone response elements (HRE). The MR DBD is 94% identical with that of GR, and more than 90% compared with the progesterone receptor (PR) and the androgen receptor (AR). The MR DBD is a 66 aa domain encoded by exons 3 and 4: by analogy with the crystal structure obtained of the GR DBD, it contains two perpendicular α helices, structurally coordinated by a zinc ion that interacts with four cysteine residues and is thus responsible for the zinc finger’s structure. The first zinc finger contains the “P box” (defined by the three residues Gly621Ser-Val625) responsible for tight binding to the minor groove of the DNA double helix. The second zinc finger facilitates receptor dimerization through the so-called “D box” (Ala640GlyArgAsnAsp645), located in the N-terminal part of the DBD. In this context, it is interesting to note that MR is able to heterodimerize with other members of the SR subgroup, most notably GR and AR [Liu et al., 1995], consistent with the possibility that heterodimerization might play a role in some physiological responses at the level of transcriptional regulation.

The MR LBD is a complex and multifunctional domain that spans 251 aa. It is relatively conserved among SR (~55% homology) and highly conserved across species (80-97% homology), and allows selective hormone binding, thus transducing endocrine messages into specific transcriptional responses. The MR LBD crystal structure has recently been solved, thus confirming the remarkable similarity in structure among all SR [Bledsoe et al., 2005; Fagart et al., 2005; Li et al., 2005]. Basically, the MR LBD consists of 11 α helices and four small antiparallel β strands that fold into a three layer helical sandwich. Solving the crystal structure allowed identification of the crucial amino acid residues interacting with the functional group of steroid ligands. For instance, Gln776 of helix H3 and Arg817 at the end of helix 5 directly contact the 3-ketone group of aldosterone, and Asn770 of helix H3 stabilizes aldosterone 18-hydroxyl group. Other residues in helices 6 and 7, as well as Thr945 on helix 10, also directly contact the steroid ligand. A single residue at position 848 in helix H7 switches hormone specificity between MR and GR [Li et al., 2005], and amino acids 820-844 (which are not part of the ligand binding pocket) are also critical for aldosterone binding and ligand binding selectivity [Rogerson et al., 2007]. The role of Met852 in accommodating the C7 substituents of antimineralocorticoid spirolactones has very recently been described [Huyet et al., 2007]. Finally, on the basis of the high similarity between the LBD of MR and GR, and considering the evolutionary tree of this receptor subgroup, it has been proposed that MR was closer to the primordial ancestral corticosteroid receptor [Hu and Funder, 2006], which has been proposed as having high affinity for aldosterone, well before the hormone appeared [Bridgham et al., 2006]. In this context, it is also interesting to note that the Ser949 in human MR is deleted in almost all GR and that the His950 in human MR, conserved in MR in Old World monkeys, is a glutamine in all teleost and land vertebrate MR and may thus represent a breaking point during evolution [Baker et al., 2007].

The MR LBD possesses a ligand-dependent AF-2 constituted by the helices H3, H4, H5 and H12. Upon ligand binding, a rearrangement of the LBD occurs: H12 closes over the ligand pocket which, in combination with the bending of helices H3, H5 and H11, forms a hydrophobic cleft on the surface of the LBD. This groove serves as a docking surface for transcriptional coactivators possessing a NR box defined by the LXXLL motif, an interaction essential for activation of MR transcriptional activity. Given the high level of homology between their LBDs, it is not surprising that GR and MR recruit almost identical coactivators through their AF2 domains. A recent study, using an isolated MR LBD as bait to screen interaction with a LXXLL peptide library showed that MR interacts with a restricted number of coactivator peptides including SRC-1, ASC2, PGC-1α and PGC-1β [Hultman et al., 2005].

Although the relative contributions of AF1a, AF1b and the AF2 to MR transcriptional activity appears to be highly dependent on cellular and promoter contexts, the NTD appears to account for ~40-50% of total transactivation and represents a key determinant of MR specificity [Pascual-Le Tallec and Lombes, 2005].

As noted above, MR is expressed as at least two different proteins, MRA and MRB [Pascual-Le Tallec et al., 2004], resulting from strong Kozak sequences initiating alternative translation. These variants display distinct transactivation capacities in vitro; it remains to be established whether they are differentially expressed in vivo, and the extent to which they contribute to fine-tuning of MR transcriptional activity and the differential patterns of gene expression in different cellular contexts, as previously described for GR [Lu and Cidlowski, 2005].

Interactions

In its non-liganded state, MR interacts with a large variety of proteins, most notably in the cytoplasmic compartment, thus forming part of a hetero-oligomer [Rafestin-Oblin et al., 1989]. MR contacts chaperone proteins such as the heat shock protein hsp90 [Binart et al., 1995], and indirectly interacts with hsp70, the p23 and p48 proteins and the FKBP-59 immunophilins or CYP40 cyclophillin [Bruner et al., 1997; Pratt and Toft, 1997]. These chaperones play a pivotal role in maintaining MR in an appropriate conformation for ligand binding. Besides the chaperone proteins, the MR also interacts with actin [Jalaguier et al., 1996], which may thus play a role in the ligand-dependent nuclear translocation. Upon hormone binding, the MR dissociates from chaperone proteins, undergoes nuclear translocation and interacts with numerous molecular partners in a coordinate and sequential manner to ensure appropriate transcriptional regulation. For over a decade, yeast two-hybrid screening, GST pulldown and coimmunoprecipitation assays have been used to identify various MR-interacting nuclear proteins.

From a functional point of view, the most important are the transcriptional coregulators acting either as coactivators or corepressors of MR transactivation. Since the initial description of coactivators in the mid 1990s, our knowledge of the complexity of transcriptional regulation has considerably increased. SR are now considered as platforms recruiting in an ordered and cyclical manner different coregulators [Metivier et al., 2003], which exhibit various enzymatic activities to play the role of transcriptional master switches [O'Malley, 2007]. The first member of the large coactivator family identified was steroid receptor coactivator-1 (SRC-1) [Onate et al., 1995], postulated to initiate transcription by recruiting a series of proteins involved in chromatin remodeling, histone acetylation and methylation [Freiman and Tjian, 2003; Rosenfeld et al., 2006]. Since then, a dozen coregulators have been demonstrated to interact with MR and modulate its activity (for an exhaustive review on this topic, see [Pascual-Le Tallec and Lombes, 2005]). As summarized and referenced in Table 1, the most important coactivators for MR transactivation appeared to be the histone acetylase CBP/p300; the helicase RHA; the transcriptional coactivators SRC-1, SRC-1e, and PGC-1; and finally, the Pol II elongation factor ELL, that constitutes the first example of a selective transcriptional coregulator of MR [Pascual-Le Tallec et al., 2005]. Corepressors able to bind MR and repress its transcriptional function include the widely repressive SMRT and NCoR; the apoptosis regulator DAXX; and the specific SUMO-ligase PIAS proteins. Of interest, sumoylation now emerges as an important posttranslational modification for many nuclear receptors and coregulators. SUMO-E3 ligase PIAS proteins repress MR, potentially in collaboration with DAXX protein [Lin et al., 2006]; it has recently been shown, however, that the SUMO-E2 activating enzyme Ubc9 interacts with the MR NTD/DBD (1-670 aa) to potentiate aldosterone-dependent MR transactivation [Yokota et al., 2007], further increasing the complexity of sumoylation-mediated regulation. The MR interaction network per se appears not to determine functional MR activity; rather, the coordination and sequential interaction of molecular partners, directly or indirectly with MR, controls activation and function of cooperative transcriptional complexes, probably in a cell- and promoter-dependent manner. It should be mentioned that MR also heterodimerizes with other SR, notably GR and AR [Liu et al., 1995; Savory et al., 2001], thus providing additional support for MR functional diversity of action.

Table 1. MR-interacting proteins.

This table presents several categories of proteins interacting with the MR (chaperones, coactivators, corepressors). Their names, functions and putative interacting domains are shown, together with the corresponding references.

graphic file with name nrs05012.t1.jpg

MR expression

Until the late 1970s, MR expression measured by binding assays was considered restricted to polarized tight epithelia, which show aldosterone-dependent transepithelial sodium transport [Marver et al., 1974]. MR expression was located by immunohistochemistry in the kidney, most notably in the distal convoluted tubules and cortical collecting ducts [Krozowski et al., 1989; Lombes et al., 1990]. MR seems also to be expressed at the messenger and the protein levels in glomeruli, especially in mesangial cells [Miyata et al., 2005; Nishiyama et al., 2005] and podocytes [Shibata et al., 2007], where aldosterone has been reported to modulate podocyte function, possibly through the induction of oxidative stress and of the serum and glucocorticoid-regulated kinase 1 (SGK1). MR expression was also detected, by specific binding of [3H]-aldosterone, in the distal colon of rat [Pressley and Funder, 1975], human [Lombes et al., 1984] and chick [Rafestin-Oblin et al., 1989]. The lung may represent another aldosterone target tissue, in that MR binding sites were demonstrated in airway epithelia from bronchiole to trachea [Krozowski and Funder, 1981]. MR expression (transcript and/or protein) was clearly revealed in the salivary [Funder et al., 1972] and sweat glands [Kenouch et al., 1994], in the liver [Duval and Funder, 1974] and in the inner ear [Furuta et al., 1994; Pitovski et al., 1993; Teixeira et al., 2006]. Importantly, epithelial expression of MR was always associated with expression of the 11β-hydroxysteroid dehydrogenase 2 (11β-HSD2), the enzyme that allows aldosterone to selectively activate MR, by converting glucocorticoid hormones to their 11-keto analogs, unable to bind MR [Edwards et al., 1988; Funder et al., 1988].

Subsequently, MR expression was detected in non-epithelial tissues in which the expression of 11β-HSD2 was absent or extremely low. For instance, specific binding sites for aldosterone were identified in mononuclear leucocytes [Armanini et al., 1985] and in the heart [Barnett and Pritchett, 1988; Pearce and Funder, 1987] and MR transcripts were detected in specific structures of the hippocampus (dentate gyrus and CA1, 2, and 3 nuclei) and in the hypothalamus [Han et al., 2005; Herman et al., 1989; Van Eekelen et al., 1988]. In 1992, MR was localized at the cellular level by immunohistochemistry in cardiomyocytes, endothelial cells and large vessels [Lombes et al., 1992]. Some years later, MR expression was confirmed, at the messenger and protein level in the skin, not restricted to sweat and sebaceous glands, but also in keratinocytes constituting the stratified epithelium [Kenouch et al., 1994]. Recent studies have shown MR to be expressed at the transcript and protein level in adipose tissues, both in white [Caprio et al., 2007; Fu et al., 2005; Rondinone et al., 1993] and brown adipocytes [Penfornis et al., 2000; Viengchareun et al., 2001; Zennaro et al., 1998]. This is of particular interest considering the functional interaction between MR and PGC-1 [Hultman et al., 2005] and the central role of this coactivator for brown adipocyte differentiation [Lin et al., 2005]. Of note, MR is also expressed at the protein level in ocular tissues, such as retina [Mirshahi et al., 1997] and iris-ciliary body [Schwartz and Wysocki, 1997], in placenta [Hirasawa et al., 2000], and at the messenger level in uterus, ovaries and testis [Le Menuet et al., 2000], with no clear roles reported to date.

This widespread expression of MR suggests novel functions for this receptor in these target tissues. It also raises questions regarding the role played by glucocorticoid hormones (cortisol or corticosterone) in MR activation, considering the absence of 11β-HSD2 in non-epithelial tissues, except in certain brain areas such as the nucleus of the solitary tract, as recently reported [Geerling et al., 2006; Naray-Fejes-Toth and Fejes-Toth, 2007].

MR is now considered an ubiquitous transcription factor, and real-time PCR quantification of MR and GR transcripts reveals interesting anatomical expression patterns [Bookout et al., 2006]. MR and GR expression are equivalent and high in the gastrointestinal system, and moderate in the endocrine, reproductive, metabolic and cardiovascular systems. MR expression is higher than that of GR in the central nervous system (CNS) and the structural system (skeleton), whereas GR expression is, as expected, more pronounced than that of MR in the immune system. According to the classification based on hierarchical clustering of gene expression, MR seems to belong to the same cluster as LXRβ and RXRβ, the expression of which is most abundant in the CNS and is crucial to the global basal metabolism, linked to circadian clocks, metabolism and cardiovascular control [Bookout et al., 2006]. Thus, distinct MR and GR expression patterns strongly support that MR and GR differentially affect transcriptional programs governing distinct physiological processes and pathophysiological disorders.

Regulation of MR expression

A key step in the mineralocorticoid response is the regulation of the MR expression level. Indeed, in non-epithelial tissues, such as the hippocampus, MR exerts its action in balance with the GR [de Kloet, 2003]. In the brain, some compounds, such as serotonin [Lai et al., 2003] and progesterone [Castren et al., 1995], have been reported to modulate MR mRNA expression. In other tissues, MR expression levels positively correlate with the severity of heart [Yoshida et al., 2005] and kidney failure [Quinkler et al., 2005]. Of major interest, a recent report showed that lowering renal expression of MR by a RNA interference strategy in cold-induced hypertensive rats both prevented the progression of hypertension and attenuated renal damage [Wang et al., 2006], emphasizing the potential importance of regulation of MR expression. The hMR gene has two 5’ untranslated exons (1α and 1β) alternatively spliced onto exon 2. Their 5’-flanking regions were identified as functional promoters, referred to as P1 and P2, respectively, and these were studied in transient transfection assays [Zennaro et al., 1996]. The basal transcriptional activity of the P1 (proximal) promoter is stronger than that of the P2 (distal) promoter. In vitro, while both promoters are stimulated by glucocorticoids, only P2 is activated by aldosterone. Transgenic mouse models were established by targeted oncogenesis with each promoter fused to the simian virus 40 large T antigen as a reporter gene, providing a unique opportunity to examine tissue-specific utilization of these promoters in vivo [Le Menuet et al., 2000] and to establish novel mineralocorticoid-sensitive cell lines [Le Menuet et al., 2004]. The P1 promoter was shown to be active in vivo in all MR-expressing tissues, whereas the P2 promoter activity was much lower and appeared to be restricted to development, questioning the role of the P2 promoter and the molecular events and transcription factors involved in regulation of the MR promoter in vivo.

Cellular mechanisms and cellular biology

Subcellular distribution

In the absence of ligand, MR is located mainly in the cytoplasm [Binart et al., 1991; Lombes et al., 1994a], associated with chaperone proteins. Upon ligand binding and dissociation of receptor-associated proteins, activated MR translocates into the nuclear compartment, in response to nuclear localization signals (NLS) present in the receptor protein sequence. Three functional NLS have been described so far [Walther et al., 2005] and are depicted in Figure 1. The first one, (NLS0), is located at the end of the NTD, between amino acids 590 and 602. It possesses five serine and one threonine residues that play a crucial role in the MR nuclear import dependent upon Ser601 phosphorylation. The NLS2 is located in the receptor LBD, as a sequence without any basic amino acids, a feature of NLS sequences in other SR. Nuclear translocation mediated by NLS2 seems to depend on the nature of the ligand; only MR agonists induce rapid translocation of the receptor, with antagonists less effective [Lombes et al., 1994a]. The third sequence, (NLS1), is located in the C-terminal part of the DBD; NLS1 acts cooperatively with NLS0 and NLS2 to facilitate nuclear translocation of the unbound MR [Walther et al., 2005]. In addition, a nuclear export signal (NES) is located between the two zinc fingers of the DBD near the NLS1 [Black et al., 2001]. Altogether, these data indicate that MR is a receptor, the distribution of which is an active process under the control of functional NLS.

Although it has been suggested that a membrane receptor for corticosteroids might mediate their rapid effects on cortical collecting duct cells [Le Moellic et al., 2004] and on CA1 pyramidal neurons [Karst et al., 2005] through a nongenomic signaling pathway, evidence for a membrane-bound MR is lacking.

Promoter binding and transcriptional activation

Upon ligand binding, MR moves into the nucleus acting as a transcription factor by binding to specific HRE present in target genes potentially located up to 10 kb upstream or downstream from transcriptional start sites, as recently described for GR and GREs [So et al., 2007]. This study reported that GRE sequences vary extensively around a consensus, but are strikingly conserved for a given site across species, raising the possibility of distinct sets of MREs versus GREs for MR target genes. MR, when bound to response elements, recruits chromatin remodeling complexes to release the nucleosome structure, and components of the transcriptional machinery to activate Pol II transcription. A complete and sequential picture of the transcriptional events remains to be clearly established for MR.

Posttranslational modifications

Even though phosphorylation of SR has been shown to play a major role in modulating their intrinsic function, MR phosphorylation has received little attention so far, despite an early report describing MR as a phosphoprotein [Alnemri et al., 1991]. Several tyrosine, serine and threonine residues are present throughout the MR protein (Figure 1). A tyrosine to cysteine substitution (Y73C), at a potential phosphorylation site, was identified in the NTD of the Brown Norway rat as opposed to Fischer 344 rat, and was associated with increased MR transactivation and unexpectedly, partial activation by progesterone [Marissal-Arvy et al., 2004]. This tyrosine phosphorylation site is also present in hMR and could thus be implicated in potential pathophysiological dysregulation. Phosphorylation sites at Thr735 and Ser737 were recently identified in the hMR LBD by proteomic analysis [Hirschberg et al., 2004], and may thus participate in LBD conformational change. In addition, as mentioned above, mutations in the serine/threonine-rich sequence in the MR NLS0, most notably Ser601, have highlighted their role in receptor subcellular shuttling [Walther et al., 2005]. MR is phosphorylated within minutes after aldosterone exposure on serine and threonine residues by the protein kinase C α, leading to subsequent ionic transport [Le Moellic et al., 2004], and providing strong evidence for cross-talk between early (<30 min) non-genomic and late (>2h) genomic effects of MR and aldosterone (for detailed review see [Funder, 2005]).

Recently, sumoylation has emerged as an extremely important posttranslational modification for regulation of transcription factors, most notably members of the nuclear receptor superfamily [Seeler and Dejean, 2003]. Sumoylation entails a covalent link with small ubiquitin-related modifier (SUMO). MR possesses several sumoylation consensus motifs defined by the peptide sequence site ΨKXE, where Ψ is an aliphatic residue, K the target lysine for sumoylation and X any residue. Four SUMO sites have been identified in the NTD at positions K89, K399, K428, K494 and one in the LBD at position K953 [Pascual-Le Tallec et al., 2003] (see Figure 1). The consensus sites, also named synergy control motif [Iniguez-Lluhi and Pearce, 2000], are highly conserved throughout evolution, supporting an important functional role for sumoylation. Indeed, sumoylation is now considered a general repressive mechanism for many transcription factor functions. MR-mediated transcription is repressed by sumoylation to an extent dependent on the nature of the response element bound, in that it may impair protein-protein interactions with the transcriptional initiation complex. Acetylation mediated by the p300 and p/CAF proteins has been reported for other SR. The KXKK/RXKK acetylation motif is a perfect match with the MR NLS1, most notably the K677, suggesting that MR acetylation might affect not only ligand-dependent nucleo-cytoplasmic shuttling, but also protein-protein interaction and subsequent transcriptional regulation. The exact impact of acetylation on MR function, and the relative contribution of CBP/p300 and HDAC warrant further investigation.

Very recently, ubiquitylation (or ubiquitination), another posttranslational modification of MR, has been reported. MR can be poly-ubiquitylated and targeted to the proteasome, a mechanism which seems to be essential for the regulation of MR-mediated transcriptional activation [Tirard et al., 2007].

Target genes and their biological functions

Epithelial tissues

In classical polarized epithelial tissues (i.e., kidney, colon), MR regulates salt balance and water homeostasis by directly stimulating expression of specific ionic transporters active at the cellular membrane: the amiloride-sensitive Epithelial Na Channel (ENaC), located at the apical membrane [Canessa et al., 1994; Rossier et al., 2002], and the basolateral Na+,K+-ATPase pump [Horisberger et al., 1991; Jorgensen, 1986]. These transporters are responsible for unidirectional transepithelial sodium transport from the lumen to the interstitium. Figure 2 illustrates a schematized model of aldosterone action in an epithelial renal cell and Table 2 summarizes early aldosterone-induced genes, identified in epithelial tissues. They include the ENaC subunits [Bens et al., 1999; Epple et al., 2000; Masilamani et al., 1999; Teixeira et al., 2006], the Na+,K+-ATPase pump subunits [Kolla and Litwack, 2000] the channel-inducing factor (CHIF), a member of the FXDY family, which regulates pump activity in the colon [Brennan and Fuller, 1999; Wald et al., 1996], and the K-ras2 gene, expression of which is also stimulated by aldosterone in the colon [Brennan and Fuller, 2006]. The best studied aldosterone-induced gene in epithelial tissues is the serine/threonine kinase SGK1, which plays a pivotal role in sodium homeostasis, in that aldosterone-induced expression of SGK1 strongly activates ENaC by phosphorylating the ubiquitin-ligase Nedd4-2 [Bhargava et al., 2001; Chen et al., 1999; Naray-Fejes-Toth et al., 1999]. This phosphorylation impairs the interaction of Nedd4-2 and ENaC, leading to degradation of ENaC by the proteasome, and thus allowing a pool of channels to remain at the apical membrane [Debonneville et al., 2001].

Figure 2. Model of MR action in a renal polarized epithelial cell.

Figure 2

Aldosterone enters a target cell and binds MR, which translocates into the nucleus. MR interacts with a HRE, recruits various transcriptional coregulators (Coregulators) to activate the transcriptional machinery (TM), and thus alters expression of aldosterone target genes (in blue). At the apical membrane, ENaC (epithelial sodium channel), composed of three subunits (α, β and γ), constitutes the rate-limiting step of apical Na+ entry. Na+ is then extruded into the basolateral space by the Na+/K+-ATPase pump, the activity of which is modulated in the colon by the regulatory protein CHIF (corticosteroid hormone-induced factor). In the absence of aldosterone, ENaC proteins interact with Nedd4-2, an ubiquitin-ligase which targets ENAC to proteosomal degradation. SGK1 (serum and glucocorticoid-regulated kinase) is a key aldosterone-regulated target gene that plays a central role in sodium reabsorption. Upon aldosterone exposure, PDK1-activated kinase SGK1 phosphorylates Nedd4-2, which in turn dissociates from ENaC, increasing its apical membrane abundance. Usp2-45, a novel early aldosterone-induced mRNA, is an ubiquitin-specific protease which deubiquitylates ENaC and thereby increases ENaC-mediated sodium transport. Other aldosterone-induced genes included kidney specific KS-WNK1 (serine/threonine kinase With No K), K-Ras2, NDRG2 (N-myc down-stream regulated gene 2), GILZ (glucocorticoid-induced leucine zipper), endothelin ET-1 and plasminogen activator inhibitor-1 (PAI-1).

Table 2. MR target genes and their biological functions in epithelial tissues.

This table presents the names of target genes, the expression of which is regulated by MR in epithelial tissues. The function of the target gene products is listed, together with the corresponding references.

graphic file with name nrs05012.t2.jpg

Several lines of evidence suggest that SGK1 is not the only mediator of aldosterone action. The glucocorticoid-induced leucine zipper protein (GILZ), initially identified as a transcription factor [Robert-Nicoud et al., 2001], was subsequently shown to stimulate ENaC-mediated Na+ transport in the kidney by inhibiting extracellular signal-regulated kinase (ERK) signaling [Soundararajan et al., 2005]. A recent study has identified novel early aldosterone-regulated mRNA in microdissected mouse distal nephron by microarray. Besides SGK1, the induced mRNA species include Grem2, activating transcription factor 3, and the ubiquitin-specific protease Usp2-45, which deubiquitylates ENaC, and thus stimulates ENaC-mediated sodium transport [Fakitsas et al., 2007]. Of importance, several downregulated genes have also been identified, but their roles in the regulation of ionic transport remain to be established. KS-WNK1 (With No lysine K) is another serine/threonine kinase which plays a crucial role in ENaC activation [Naray-Fejes-Toth et al., 2004] upon aldosterone exposure, stimulating the PI3 kinase pathway, which directly activates SGK1 [Xu et al., 2005a; Xu et al., 2005b]. Interestingly, other aldosterone-induced genes have been reported so far, but their biological function remains to be elucidated. For instance, NDRG2 (N-Myc downstream regulated gene 2) [Boulkroun et al., 2002; Wielputz et al., 2007], Endothelin-1 (ET-1) [Wong et al., 2007] and the Plasminogen Activator Inhibitor-1 (PAI-1) [Yuan et al., 2007] transcripts have been shown to be specifically increased upon aldosterone exposure in the kidney or distal colon.

Non-epithelial tissues

The discovery of new sites of MR expression in non-epithelial tissues such as the heart, vasculature, brain and adipocytes has led to identification of potential new aldosterone target genes in these tissues with unexpected biological functions (see Table 3). For instance, in aortic endothelial cells, aldosterone was shown to increase osteopontin and angiotensin-converting enzyme (ACE) gene expression, which may be involved in the development of endothelial dysfunction and vascular injury induced by this steroid [Sugiyama et al., 2005a; Sugiyama et al., 2005b]. Leopold et al. recently reported that aldosterone impairs vascular reactivity by decreasing endothelial glucose-6-phosphate dehydrogenase (G6PD) expression and activity, which results in increased oxidative stress and decreased NO levels [Leopold et al., 2007]. In vascular smooth muscle cells, MDM2 was identified as a novel mineralocorticoid-responsive gene involved in aldosterone-induced vascular remodeling [Nakamura et al., 2006]. In addition, MR activation in vitro also directly enhances EGF-R gene expression, ultimately contributing to an aldosterone-induced increase in fibronectin abundance in aorta [Grossmann et al., 2007]. Several studies have demonstrated that aldosterone can stimulate collagen (I, III and IV) gene expression in cardiac [Brilla et al., 1994] or renal fibroblasts via MR-mediated ERK1/2 activation pathway [Nagai et al., 2005], which may contribute to the progression of aldosterone-induced myocardial or tubulointerstitial fibrosis. These adverse effects of aldosterone have prompted a renewed interest in the use of MR antagonists and for the identification of MR target genes, particularly in the heart.

Table 3. MR target genes and their biological functions in non-epithelial tissues.

This table shows the target genes, the expression of which is upregulated or downregulated by MR in non-epithelial tissues. The functions of the target gene products are listed, together with the corresponding references.

graphic file with name nrs05012.t3.jpg

Very recently, a large number of target genes have been identified in a cardiomyocyte cell line stably expressing MR, including genes related to extracellular matrix regulation (tenascin-X, ADAMTS1, PAI-1, UPAR, and hyaluronic acid synthase-2), signaling, regulation of vascular tone (RGS2, adrenomedullin) and inflammation (orosomucoid) [Fejes-Toth and Naray-Fejes-Toth, 2007].

The role of aldosterone and MR in cell proliferation [Stockand and Meszaros, 2003] and differentiation [Caprio et al., 2007; Penfornis et al., 2000] is now well documented. Physiological studies have demonstrated the anti-apoptotic role of aldosterone and MR in hippocampus degeneration. Indeed, the balance between GR and MR seems to control limbic neuron fates [de Kloet et al., 2005] and efforts have been made to identify corticosteroid-responsive genes in the hippocampus [Datson et al., 2001]. Interestingly, studies performed on hippocampal neurons confirmed the protective effects of MR on GR-induced apoptosis [Almeida et al., 2000; Crochemore et al., 2005], and are in accordance with the knockout mouse model for MR, which shows degeneration of the hippocampus granule cells in adulthood [Gass et al., 2000]. Recently, light has been shed on the molecular basis of MR-prevented apoptosis with molecular data showing that the MR NTD was able to prevent all the features (modulation of pro versus anti-apoptotic genes, and caspase inhibition) of glucocorticoid-induced apoptosis in lymphocytes [Planey et al., 2002], potentially through competition for common coregulators (like ELL, FAF, or FLASH). In brown adipocytes, it has been lately shown that aldosterone treatment induced a marked decrease in expression and function of mitochondrial uncoupling proteins UCP1 and UCP3 [Viengchareun et al., 2001], providing additional support for a role of MR in the control of energy expenditure. Finally, aldosterone-activated MR also promotes osteoblastic differentiation and mineralization of vascular smooth muscle cells independently of BMP2 signaling [Jaffe et al., 2007], with the genes involved in such effects yet to be identified.

Additional studies are needed to assess the contribution of these potential target genes to the physiological and pathophysiological effects of aldosterone in non-epithelial tissues.

MR transgenic models

A better understanding of physiological MR actions has been achieved by its transgenic expression in various animal models. Such genetically engineered mice, including MR gene inactivation, RNA interference and MR overexpression, have been generated by several groups. MR knockout mice, in which the MR DBD has been disrupted, die around day 10 after birth from renal sodium loss [Berger et al., 1998], prevented by exogenous NaCl administration [Berger et al., 2000]. The animals so rescued show impaired neurogenesis and degeneration of the hippocampus granule cells in adulthood, as previously noted [Gass et al., 2000]. This finding emphasizes the role of MR in cell proliferation/apoptosis balance in limbic structures. At variance with the MR KO mouse model, the mouse aldosterone synthase gene knockout is not lethal. As expected, ionic homeostasis is altered in the absence of aldosterone, but as a consequence, a high level of corticosterone and angiotensin II seems to partially rescue sodium balance [Makhanova et al., 2006], underscoring in this context the importance of MR over aldosterone.

More recently, AQP2-Cre-MR mice with MR inactivation in renal principal cells exhibited normal renal sodium excretion associated with elevated aldosterone levels on a standard diet. On a low-sodium diet, however, the mice rapidly lost sodium consistent with the impaired induction of ENaC channels in principal cells of the collecting duct and late connecting tubule, in part functionally offset by the late distal convoluted tubule and early connecting tubule [Ronzaud et al., 2007]. Another MR conditional knockout mouse with forebrain specific inactivation of MR showed impaired spatial learning ability linked to behavioral stereotype [Berger et al., 2006]. Brain-specific MR inactivation compromised the rapid effect by corticosterone, providing evidence that this nongenomic effect is MR-mediated [Karst et al., 2005].

Mouse models of MR overexpression have also been generated. The first used the human P1 proximal MR promoter to drive human MR transgene expression in most aldosterone target tissues (distal nephron, brain, heart). These animals presented with enlarged kidneys associated with renal tubular dilatation and normal blood pressure, and developed mild dilated cardiomyopathy associated with arrhythmia [Le Menuet et al., 2001]. In addition, gene expression studies of these mice revealed several alterations of cardiac- and renal-specific gene expression such as ANF, SGK1, egr-1 [Le Menuet et al., 2001]. Of interest, conditional cardiac MR overexpression controlled by the α-myosin heavy chain promoter led to life-threatening arrhythmias [Ouvrard-Pascaud et al., 2005], consistent with the previous model. None of these MR-overexpressing mice developed cardiac fibrosis, clearly different from the cardiac phenotype previously described for salt and aldosterone excess animal models [Young et al., 1994], questioning the direct role of cardiomyocyte MR in cardiac fibrosis. MR knockdown experiments using the conditional expression of a MR antisense mRNA was reported to induce cardiac fibrosis and heart failure [Beggah et al., 2002]. This phenotype was somewhat unexpected, but could be related to an imbalance between cardiac MR and GR levels, which could ultimately lead to inappropriate corticosteroid responses, reminiscent of the cardiac phenotype observed with heart-specific overexpression of 11β-HSD2 [Qin et al., 2003]. Another transgenic mouse model with a targeted overexpression of MR in the forebrain, using the forebrain-specific calcium/calmodulin-dependent protein kinase II α promoter, exhibited decreased anxiety-like behavior associated with a diminution of hippocampal GR protein and increased in serotonin 5HT-1a receptor levels [Rozeboom et al., 2007]. A conditional targeted skin overexpression of MR resulted in early postnatal death when the transgene was expressed during gestation, associated with excessive keratinocyte apoptosis. Additionally, postnatal transgene expression led to alopecia and hair follicle dystrophy [Sainte Marie et al., 2007].

MR and disease

Since the aldosterone-MR signaling pathway plays a major role not only in the control of sodium and potassium homeostasis, but also in other important physiological processes, its involvement in numerous human diseases is not surprising. Here, we will focus on diseases for which a direct implication of MR has been documented. Among them, the most important genetic disorder is the autosomal dominant pseudohypoaldosteronism type 1 (adPHA1), which is caused, in most cases, by heterozygous loss-of-function mutations of MR. First described in 1958 [Cheek and Perry, 1958], this rare life-threatening condition is characterized by renal resistance to aldosterone action. Early in infancy, usually during the first months of life, patients present with dehydration and failure to thrive associated with massive salt wasting, hyperkalemia and metabolic acidosis, despite elevated plasma aldosterone and renin levels. They require sodium supplementation during their early years, which can be subsequently reduced, as adults are usually asymptomatic in spite of high plasma aldosterone and renin levels [Geller, 2005]. To date, approximately 50 distinct mutations in the human MR gene responsible for this syndrome have been described [Geller, 2005; Geller et al., 1998; Pujo et al., 2007] and are summarized in Figure 3. They include missense, nonsense, frameshift and splice site mutations, as well as deletions spread throughout the gene. These mutations are responsible for either an early termination of translation with MR truncation or a defect in MR activity (loss of LBD or DBD), disruption of nucleocytoplasmic shuttling or alteration in some transcriptional coregulator recruitment. Only heterozygous mutations have been reported in humans, suggesting that the homozygous state may be lethal in utero. The loss of one allele results in haploinsufficiency, sufficient to generate adPHA1 symptoms, thus underlining the importance of a substantial MR protein level, most notably during the neonatal period [Geller et al., 2006].

Figure 3. Schematic localization of MR mutations causing pseudohypoaldosteronism type 1 and polymorphisms in the human NR3C2 gene.

Figure 3

Nucleotide and amino acid numbering is indicated in reference to the published cDNA, where +1 is the A of the translational initiation codon (Arriza et al., 1987). Fs, frameshift (red); X, stop codon (blue); missense (green); sp alt, splice alteration (purple); polymorphism (black).

To date, there is only one mutation known to result in a gain-of-function of MR, and it leads to a severe inherited form of early-onset hypertension, which is exacerbated during pregnancy [Geller et al., 2000]. This single mutation S810L in the LBD causes a constitutive MR activation, as well as illicit activation of MR by progesterone and other steroids, including cortisone, 11-dehydrocorticosterone and the mineralocorticoid antagonist spironolactone [Geller et al., 2000; Pinon et al., 2004; Rafestin-Oblin et al., 2003].

Besides pathogenic mutations of the hMR gene, common MR polymorphisms have been described and seem to be mostly related to modulation of stress responsiveness. In the brain, where 11β-HSD2 is not expressed except in discrete brain areas, MR may be sensitive to cortisol level variations induced by stress. Thus, the MR allele I180V, which has a frequency of approximately 15% varying with ethnic background [Arai et al., 1999; Balsamo et al., 2007], has been found associated with enhanced endocrine and autonomic responses to a psychological stressor, without any modification of sodium homeostasis or blood pressure [DeRijk et al., 2006]. Given this finding, other polymorphisms deserve to be characterized in terms of functional consequences, both for sodium handling or clinical implications.

Besides genetic diseases, the prominent role played by MR in the pathogenesis of cardiac dysfunction in humans is now well documented. Indeed, large clinical trials (RALES, EPHESUS, 4E) have demonstrated the major benefits of anti-mineralocorticoid therapy, such as spironolactone and eplerenone, in reducing the mortality and improving the prognosis of heart failure patients [Pitt et al., 1999] or after post-acute myocardial infarction [Pitt et al., 2003]. However, the molecular mechanisms involved in the protective effects of these MR antagonists remain to be fully characterized. It is noteworthy that a recent report shows that MR activation is a critical factor in the early pathogenesis of renal disease in both type 1 and type 2 diabetes mellitus [Guo et al., 2006].

Ligands

In human, aldosterone is the physiological ligand of MR, most notably for epithelial MR. This hormone is synthesized in the zona glomerulosa of adrenal cortex under the regulation of angiotensin II, ACTH and serum potassium concentration. Another important physiological ligand of MR is cortisol (corticosterone in rodents), which has the same affinity for MR even though the dynamic binding parameters and the functional properties of glucocorticoid-MR complexes are somehow different from that of aldosterone-MR [Kitagawa et al., 2002; Lombes et al., 1994b]. The most potent MR-protective mechanism against MR activation by glucocorticoid hormones is 11β-HSD2, the NAD-dependent enzyme which converts cortisol into its inactive metabolite cortisone, in epithelial and vascular target tissues. Deoxycorticosterone (most notably in fish) [Sturm et al., 2005], as well as other glucocorticosteroid compounds (dexamethasone, fludrocortisone), act as mineralocorticoid agonists, at least in vitro. It is now clear that progesterone and androgens and their derivatives could also bind MR and exert partial agonist and antagonist effects depending on the cellular context [Marissal-Arvy et al., 2004; Quinkler et al., 2002; Quinkler et al., 2003; Takeda et al., 2007].

Several mineralocorticoid antagonists, such as spironolactone and the more selective eplerenone [Pitt et al., 2003], have been developed and are currently used as antihypertensive and cardiovascular protective agents. Drospirenone is a novel progestogen with considerable MR antagonist activity [Oelkers, 2004].

Selective mineralocorticoid receptor modulators, either steroid-derived compounds or alternatively nonsteroidal scaffold ligands, should ultimately produce beneficial effects mediated by transactivation or transrepression of MR in a tissue-dependent manner. Ideally, a MR modulator would have neuronal anti-apoptotic effects and efficient protective effects on renal and cardiovascular systems.

Conclusions

It now appears that the roles of MR extend far beyond control of fluid and electrolyte homeostasis, being involved in (patho)physiological processes as diverse as cardiovascular remodeling, fat storage, energy balance, and mammalian behavior. MR transgenic mouse models have been and will continue to be extremely useful in unraveling these new aspects of MR biology. MR action must be seen in the context of its functional interaction with transcriptional coregulators, for which physiological roles are not redundant as initially thought, but are quite specific, as exemplified by members of the SRC family. The combinatorial tissue distribution of MR and its coregulators thus provides subtle cell-dependent tuning of MR target gene regulation and should also offer new insights into the molecular mechanisms of MR selectivity. Furthermore, recent evidence supports cross-talk between SR and other signaling pathways (like MAPK, src, STAT); the implications for MR remain to be established. The regulation of MR expression in non-epithelial tissues may differ from that in epithelia. Determining the transcriptional factors orchestrating MR expression constitutes an interesting challenge in terms of better understanding MR-mediated responses, in addition to its potential relevance to pathophysiological disorders. Altogether, the aim of continuing research is to improve comprehension of MR-mediated signaling, both at the genomic and non-genomic levels, in order to propose new therapeutic strategies for human diseases.

Acknowledgments

We acknowledge John Funder for helpful discussions, support over the years and his critical reading and editing of this manuscript. We also want to thank our colleagues and collaborators for their help and encouragement in this exciting research area of nuclear receptor biology. We would like to apologize to any investigators who have made contributions in the mineralocorticoid receptor field and whose works may have been omitted. Finally, we are indebted to INSERM and Université Paris-Sud 11 for funding.

Abbreviations

11β-HSD2

11β-hydroxysteroid dehydrogenase 2

ACE

angiotensin converting enzyme

ACTH

adrenocorticotrophic hormone

ADAMTS1

a disintegrin and metalloproteinase with thrombospondin-like motifs 1

adPHA1

autosomal dominant pseudohypoaldosteronism type 1

AF1

activation function 1

AF2

activation function 2

ANF

atrial natriuretic factor

AR

androgen receptor

ASC2

activating signal cointegrator 2

BMP2

bone morphogenetic protein 2

CBP

CREB binding protein

CHIF

channel-inducing factor

CNS

central nervous system

DAXX

death-associated protein 6

DBD

DNA binding domain

EGF-R

epidermal growth factor receptor

Egr-1

early growth response gene-1

ELL

eleven-nineteen lysine-rich leukemia

ENaC

epithelial sodium channel

ERK

extracellular signal-regulated kinase

ET-1

endothelin-1

FAF1

Fas associated factor 1

FLASH

FLICE associated huge

G6PD

glucose-6-phosphate dehydrogenase

GILZ

glucocorticoid-induced leucine zipper protein

GR

glucocorticoid receptor

GRE

glucocorticoid responsive element

HAS2

hyaluronic acid synthase 2

HDAC

histone deacetylase

hMR

human mineralocorticoid receptor

HRE

hormone responsive element

hsp

heat shock protein

KS-WNK1

kidney specific with no lysine [K] kinase 1

LBD

ligand binding domain

LXRβ

liver X receptor β

MAPK

mitogen-activated protein kinase

MDM2

murine double minute gene 2

MR

mineralocorticoid receptor

MRE

mineralocorticoid responsive element

NAD

nicotinamide adenine dinucleotide

NCoR

nuclear receptor corepressor

NDRG2

N-Myc downstream regulated gene 2

Nedd

neuronal precursor cell-expressed, developmentally down-regulated gene

NES

nuclear export signal

NLS

nuclear localization signal

NR

nuclear receptor

NTD

N-terminal domain

Orm

orosomucoid

p/CAF

p300/CBP-associated protein

PAI-1

plasminogen activator inhibitor-1

PDK1

3-phosphoinositide-dependent kinase 1

PGC-1α

peroxisome proliferator-activated receptors γ (PPARgamma) coactivator-1 α

PGC-1β

PPAR γ coactivator-1 β

PIAS

protein inhibitor of activated signal transducer and activator of transcription

PR

progesterone receptor

RGS2

regulator of G-protein signaling 2

RHA

RNA Helicase A

RIP140

receptor-interacting protein 140

RXRβ

retinoid X receptor β

SGK1

serum and glucocorticoid-regulated kinase 1

SMRT

silencing mediator of retinoid and thyroid hormone receptors

SR

steroid receptor

SRC-1

steroid receptor coactivator-1

STAT

signal transducer and activator of transcription

SUMO

small ubiquitin-like modifier

TIF

transcription-intermediary-factor

TM

transcriptional machinery

TNX

Tenascin-X

Ubc9

ubiquitin-like protein SUMO-1 E2-conjugating enzyme 9

UCP-1

uncoupling protein-1

UPAR

urokinase-type plasminogen activator receptor

Usp2-45

ubiquitin-specific protease 2-45

References

  1. Almeida O. F., Conde G. L., Crochemore C., Demeneix B. A., Fischer D., Hassan A. H., Meyer M., Holsboer F., Michaelidis T. M. Subtle shifts in the ratio between pro- and antiapoptotic molecules after activation of corticosteroid receptors decide neuronal fate. Faseb J. 2000;14:779–90. doi: 10.1096/fasebj.14.5.779. [DOI] [PubMed] [Google Scholar]
  2. Alnemri E. S., Maksymowych A. B., Robertson N. M., Litwack G. Overexpression and characterization of the human mineralocorticoid receptor. J Biol Chem. 1991;266:18072–81. [PubMed] [Google Scholar]
  3. Arai K., Zachman K., Shibasaki T., Chrousos G. P. Polymorphisms of amiloride-sensitive sodium channel subunits in five sporadic cases of pseudohypoaldosteronism: do they have pathologic potential? J Clin Endocrinol Metab. 1999;84:2434–7. doi: 10.1210/jcem.84.7.5857. [DOI] [PubMed] [Google Scholar]
  4. Armanini D., Strasser T., Weber P. C. Characterization of aldosterone binding sites in circulating human mononuclear leukocytes. Am J Physiol. 1985;248:E388–90. doi: 10.1152/ajpendo.1985.248.3.E388. [DOI] [PubMed] [Google Scholar]
  5. Arriza J. L., Weinberger C., Cerelli G., Glaser T. M., Handelin B. L., Housman D. E., Evans R. M. Cloning of human mineralocorticoid receptor complementary DNA: structural and functional kinship with the glucocorticoid receptor. Science. 1987;237:268–75. doi: 10.1126/science.3037703. [DOI] [PubMed] [Google Scholar]
  6. Baker M. E., Chandsawangbhuwana C., Ollikainen N. Structural analysis of the evolution of steroid specificity in the mineralocorticoid and glucocorticoid receptors. BMC Evol Biol. 2007;7:24. doi: 10.1186/1471-2148-7-24. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Balsamo A., Cicognani A., Gennari M., Sippell W. G., Menabo S., Baronio F., Riepe F. G. Functional characterization of naturally occurring NR3C2 gene mutations in Italian patients suffering from pseudohypoaldosteronism type 1. Eur J Endocrinol. 2007;156:249–56. doi: 10.1530/eje.1.02330. [DOI] [PubMed] [Google Scholar]
  8. Barnett C. A., Pritchett E. L. Detection of corticosteroid type I binding sites in heart. Mol Cell Endocrinol. 1988;56:191–8. doi: 10.1016/0303-7207(88)90060-3. [DOI] [PubMed] [Google Scholar]
  9. Beggah A. T., Escoubet B., Puttini S., Cailmail S., Delage V., Ouvrard-Pascaud A., Bocchi B., Peuchmaur M., Delcayre C., Farman N., Jaisser F. Reversible cardiac fibrosis and heart failure induced by conditional expression of an antisense mRNA of the mineralocorticoid receptor in cardiomyocytes. Proc Natl Acad Sci U S A. 2002;99:7160–5. doi: 10.1073/pnas.102673599. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Bens M., Vallet V., Cluzeaud F., Pascual-Letallec L., Kahn A., Rafestin-Oblin M. E., Rossier B. C., Vandewalle A. Corticosteroid-dependent sodium transport in a novel immortalized mouse collecting duct principal cell line. J Am Soc Nephrol. 1999;10:923–34. doi: 10.1681/ASN.V105923. [DOI] [PubMed] [Google Scholar]
  11. Berger S., Wolfer D. P., Selbach O., Alter H., Erdmann G., Reichardt H. M., Chepkova A. N., Welzl H., Haas H. L., Lipp H. P., Schutz G. Loss of the limbic mineralocorticoid receptor impairs behavioral plasticity. Proc Natl Acad Sci U S A. 2006;103:195–200. doi: 10.1073/pnas.0503878102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Berger S., Bleich M., Schmid W., Greger R., Schutz G. Mineralocorticoid receptor knockout mice: lessons on Na+ metabolism. Kidney Int. 2000;57:1295–8. doi: 10.1046/j.1523-1755.2000.00965.x. [DOI] [PubMed] [Google Scholar]
  13. Berger S., Bleich M., Schmid W., Cole T. J., Peters J., Watanabe H., Kriz W., Warth R., Greger R., Schutz G. Mineralocorticoid receptor knockout mice: pathophysiology of Na+ metabolism. Proc Natl Acad Sci U S A. 1998;95:9424–9. doi: 10.1073/pnas.95.16.9424. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Bhargava A., Fullerton M. J., Myles K., Purdy T. M., Funder J. W., Pearce D., Cole T. J. The serum- and glucocorticoid-induced kinase is a physiological mediator of aldosterone action. Endocrinology. 2001;142:1587–94. doi: 10.1210/endo.142.4.8095. [DOI] [PubMed] [Google Scholar]
  15. Binart N., Lombes M., Rafestin-Oblin M. E., Baulieu E. E. Characterization of human mineralocorticosteroid receptor expressed in the baculovirus system. Proc Natl Acad Sci U S A. 1991;88:10681–5. doi: 10.1073/pnas.88.23.10681. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Binart N., Lombes M., Baulieu E. E. Distinct functions of the 90 kDa heat-shock protein (hsp90) in oestrogen and mineralocorticosteroid receptor activity: effects of hsp90 deletion mutants. Biochem J. 1995;311 ( Pt 3):797–804. doi: 10.1042/bj3110797. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Black B. E., Holaska J. M., Rastinejad F., Paschal B. M. DNA binding domains in diverse nuclear receptors function as nuclear export signals. Curr Biol. 2001;11:1749–58. doi: 10.1016/s0960-9822(01)00537-1. [DOI] [PubMed] [Google Scholar]
  18. Bledsoe R. K., Madauss K. P., Holt J. A., Apolito C. J., Lambert M. H., Pearce K. H., Stanley T. B., Stewart E. L., Trump R. P., Willson T. M., Williams S. P. A ligand-mediated hydrogen bond network required for the activation of the mineralocorticoid receptor. J Biol Chem. 2005;280:31283–93. doi: 10.1074/jbc.M504098200. [DOI] [PubMed] [Google Scholar]
  19. Bookout A. L., Jeong Y., Downes M., Yu R. T., Evans R. M., Mangelsdorf D. J. Anatomical profiling of nuclear receptor expression reveals a hierarchical transcriptional network. Cell. 2006;126:789–99. doi: 10.1016/j.cell.2006.06.049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Boulkroun S., Fay M., Zennaro M. C., Escoubet B., Jaisser F., Blot-Chabaud M., Farman N., Courtois-Coutry N. Characterization of rat NDRG2 (N-Myc downstream regulated gene 2), a novel early mineralocorticoid-specific induced gene. J Biol Chem. 2002;277:31506–15. doi: 10.1074/jbc.M200272200. [DOI] [PubMed] [Google Scholar]
  21. Brennan F. E., Fuller P. J. Acute regulation by corticosteroids of channel-inducing factor gene messenger ribonucleic acid in the distal colon. Endocrinology. 1999;140:1213–8. doi: 10.1210/endo.140.3.6582. [DOI] [PubMed] [Google Scholar]
  22. Brennan F. E., Fuller P. J. Mammalian K-ras2 is a corticosteroid-induced gene in vivo . Endocrinology. 2006;147:2809–16. doi: 10.1210/en.2005-1481. [DOI] [PubMed] [Google Scholar]
  23. Bridgham J. T., Carroll S. M., Thornton J. W. Evolution of hormone-receptor complexity by molecular exploitation. Science. 2006;312:97–101. doi: 10.1126/science.1123348. [DOI] [PubMed] [Google Scholar]
  24. Brilla C. G., Zhou G., Matsubara L., Weber K. T. Collagen metabolism in cultured adult rat cardiac fibroblasts: response to angiotensin II and aldosterone. J Mol Cell Cardiol. 1994;26:809–20. doi: 10.1006/jmcc.1994.1098. [DOI] [PubMed] [Google Scholar]
  25. Bruner K. L., Derfoul A., Robertson N. M., Guerriero G., Fernandes-Alnemri T., Alnemri E. S., Litwack G. The unliganded mineralocorticoid receptor is associated with heat shock proteins 70 and 90 and the immunophilin FKBP-52. Recept Signal Transduct. 1997;7:85–98. [PubMed] [Google Scholar]
  26. Canessa C. M., Schild L., Buell G., Thorens B., Gautschi I., Horisberger J. D., Rossier B. C. Amiloride-sensitive epithelial Na+ channel is made of three homologous subunits. Nature. 1994;367:463–7. doi: 10.1038/367463a0. [DOI] [PubMed] [Google Scholar]
  27. Caprio M., Feve B., Claes A., Viengchareun S., Lombes M., Zennaro M. C. Pivotal role of the mineralocorticoid receptor in corticosteroid-induced adipogenesis. Faseb J. 2007;21:2185–94. doi: 10.1096/fj.06-7970com. [DOI] [PubMed] [Google Scholar]
  28. Castren M., Patchev V. K., Almeida O. F., Holsboer F., Trapp T., Castren E. Regulation of rat mineralocorticoid receptor expression in neurons by progesterone. Endocrinology. 1995;136:3800–6. doi: 10.1210/endo.136.9.7649087. [DOI] [PubMed] [Google Scholar]
  29. Cheek D. B., Perry J. W. A salt wasting syndrome in infancy. Arch Dis Child. 1958;33:252–6. doi: 10.1136/adc.33.169.252. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Chen S. Y., Bhargava A., Mastroberardino L., Meijer O. C., Wang J., Buse P., Firestone G. L., Verrey F., Pearce D. Epithelial sodium channel regulated by aldosterone-induced protein sgk. Proc Natl Acad Sci U S A. 1999;96:2514–9. doi: 10.1073/pnas.96.5.2514. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Crochemore C., Lu J., Wu Y., Liposits Z., Sousa N., Holsboer F., Almeida O. F. Direct targeting of hippocampal neurons for apoptosis by glucocorticoids is reversible by mineralocorticoid receptor activation. Mol Psychiatry. 2005;10:790–8. doi: 10.1038/sj.mp.4001679. [DOI] [PubMed] [Google Scholar]
  32. Datson N. A., van der Perk J., de Kloet E. R., Vreugdenhil E. Identification of corticosteroid-responsive genes in rat hippocampus using serial analysis of gene expression. Eur J Neurosci. 2001;14:675–89. doi: 10.1046/j.0953-816x.2001.01685.x. [DOI] [PubMed] [Google Scholar]
  33. de Kloet E. R. Hormones, brain and stress. Endocr Regul. 2003;37:51–68. [PubMed] [Google Scholar]
  34. de Kloet E. R., Joels M., Holsboer F. Stress and the brain: from adaptation to disease. Nat Rev Neurosci. 2005;6:463–75. doi: 10.1038/nrn1683. [DOI] [PubMed] [Google Scholar]
  35. Debonneville C., Flores S. Y., Kamynina E., Plant P. J., Tauxe C., Thomas M. A., Munster C., Chraibi A., Pratt J. H., Horisberger J. D., Pearce D., Loffing J., Staub O. Phosphorylation of Nedd4-2 by Sgk1 regulates epithelial Na(+) channel cell surface expression. Embo J. 2001;20:7052–9. doi: 10.1093/emboj/20.24.7052. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. DeRijk R. H., Wust S., Meijer O. C., Zennaro M. C., Federenko I. S., Hellhammer D. H., Giacchetti G., Vreugdenhil E., Zitman F. G., de Kloet E. R. A common polymorphism in the mineralocorticoid receptor modulates stress responsiveness. J Clin Endocrinol Metab. 2006;91:5083–9. doi: 10.1210/jc.2006-0915. [DOI] [PubMed] [Google Scholar]
  37. Duval D., Funder J. W. The binding of tritiated aldosterone in the rat liver cytosol. Endocrinology. 1974;94:575–9. doi: 10.1210/endo-94-2-575. [DOI] [PubMed] [Google Scholar]
  38. Edwards C. R., Stewart P. M., Burt D., Brett L., McIntyre M. A., Sutanto W. S., de Kloet E. R., Monder C. Localisation of 11 β-hydroxysteroid dehydrogenase--tissue specific protector of the mineralocorticoid receptor. Lancet. 1988;2:986–9. doi: 10.1016/s0140-6736(88)90742-8. [DOI] [PubMed] [Google Scholar]
  39. Epple H. J., Amasheh S., Mankertz J., Goltz M., Schulzke J. D., Fromm M. Early aldosterone effect in distal colon by transcriptional regulation of ENaC subunits. Am J Physiol Gastrointest Liver Physiol. 2000;278:G718–24. doi: 10.1152/ajpgi.2000.278.5.G718. [DOI] [PubMed] [Google Scholar]
  40. Fagart J., Huyet J., Pinon G. M., Rochel M., Mayer C., Rafestin-Oblin M. E. Crystal structure of a mutant mineralocorticoid receptor responsible for hypertension. Nat Struct Mol Biol. 2005;12:554–5. doi: 10.1038/nsmb939. [DOI] [PubMed] [Google Scholar]
  41. Fakitsas P., Adam G., Daidie D., van Bemmelen M. X., Fouladkou F., Patrignani A., Wagner U., Warth R., Camargo S. M., Staub O., Verrey F. Early aldosterone-induced gene product regulates the epithelial sodium channel by deubiquitylation. J Am Soc Nephrol. 2007;18:1084–92. doi: 10.1681/ASN.2006080902. [DOI] [PubMed] [Google Scholar]
  42. Fejes-Toth G., Naray-Fejes-Toth A. Early aldosterone-regulated genes in cardiomyocytes: clues to cardiac remodeling? Endocrinology. 2007;148:1502–10. doi: 10.1210/en.2006-1438. [DOI] [PubMed] [Google Scholar]
  43. Freiman R. N., Tjian R. Regulating the regulators: lysine modifications make their mark. Cell. 2003;112:11–7. doi: 10.1016/s0092-8674(02)01278-3. [DOI] [PubMed] [Google Scholar]
  44. Fu M., Sun T., Bookout A. L., Downes M., Yu R. T., Evans R. M., Mangelsdorf D. J. A Nuclear Receptor Atlas: 3T3-L1 adipogenesis. Mol Endocrinol. 2005;19:2437–50. doi: 10.1210/me.2004-0539. [DOI] [PubMed] [Google Scholar]
  45. Funder J. W., Pearce P. T., Smith R., Smith A. I. Mineralocorticoid action: target tissue specificity is enzyme, not receptor, mediated. Science. 1988;242:583–5. doi: 10.1126/science.2845584. [DOI] [PubMed] [Google Scholar]
  46. Funder J. W., Feldman D., Edelman I. S. Specific aldosterone binding in rat kidney and parotid. J Steroid Biochem. 1972;3:209–18. doi: 10.1016/0022-4731(72)90052-0. [DOI] [PubMed] [Google Scholar]
  47. Funder J. W. The nongenomic actions of aldosterone. Endocr Rev. 2005;26:313–21. doi: 10.1210/er.2005-0004. [DOI] [PubMed] [Google Scholar]
  48. Furuta H., Mori N., Sato C., Hoshikawa H., Sakai S., Iwakura S., Doi K. Mineralocorticoid type I receptor in the rat cochlea: mRNA identification by polymerase chain reaction (PCR) and in situ hybridization. Hear Res. 1994;78:175–80. doi: 10.1016/0378-5955(94)90023-x. [DOI] [PubMed] [Google Scholar]
  49. Fuse H., Kitagawa H., Kato S. Characterization of transactivational property and coactivator mediation of rat mineralocorticoid receptor activation function-1 (AF-1) Mol Endocrinol. 2000;14:889–99. doi: 10.1210/mend.14.6.0467. [DOI] [PubMed] [Google Scholar]
  50. Gass P., Kretz O., Wolfer D. P., Berger S., Tronche F., Reichardt H. M., Kellendonk C., Lipp H. P., Schmid W., Schutz G. Genetic disruption of mineralocorticoid receptor leads to impaired neurogenesis and granule cell degeneration in the hippocampus of adult mice. EMBO Rep. 2000;1:447–51. doi: 10.1093/embo-reports/kvd088. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Geerling J. C., Kawata M., Loewy A. D. Aldosterone-sensitive neurons in the rat central nervous system. J Comp Neurol. 2006;494:515–27. doi: 10.1002/cne.20808. [DOI] [PubMed] [Google Scholar]
  52. Geller D. S., Farhi A., Pinkerton N., Fradley M., Moritz M., Spitzer A., Meinke G., Tsai F. T., Sigler P. B., Lifton R. P. Activating mineralocorticoid receptor mutation in hypertension exacerbated by pregnancy. Science. 2000;289:119–23. doi: 10.1126/science.289.5476.119. [DOI] [PubMed] [Google Scholar]
  53. Geller D. S., Zhang J., Zennaro M. C., Vallo-Boado A., Rodriguez-Soriano J., Furu L., Haws R., Metzger D., Botelho B., Karaviti L., Haqq A. M., Corey H., Janssens S., Corvol P., Lifton R. P. Autosomal dominant pseudohypoaldosteronism type 1: mechanisms, evidence for neonatal lethality, and phenotypic expression in adults. J Am Soc Nephrol. 2006;17:1429–36. doi: 10.1681/ASN.2005111188. [DOI] [PubMed] [Google Scholar]
  54. Geller D. S. Mineralocorticoid resistance. Clin Endocrinol (Oxf) 2005;62:513–20. doi: 10.1111/j.1365-2265.2005.02229.x. [DOI] [PubMed] [Google Scholar]
  55. Geller D. S., Rodriguez-Soriano J., Vallo Boado A., Schifter S., Bayer M., Chang S. S., Lifton R. P. Mutations in the mineralocorticoid receptor gene cause autosomal dominant pseudohypoaldosteronism type I. Nat Genet. 1998;19:279–81. doi: 10.1038/966. [DOI] [PubMed] [Google Scholar]
  56. Greenwood A. K., Butler P. C., White R. B., DeMarco U., Pearce D., Fernald R. D. Multiple corticosteroid receptors in a teleost fish: distinct sequences, expression patterns, and transcriptional activities. Endocrinology. 2003;144:4226–36. doi: 10.1210/en.2003-0566. [DOI] [PubMed] [Google Scholar]
  57. Grossmann C., Krug A. W., Freudinger R., Mildenberger S., Voelker K., Gekle M. Aldosterone-induced EGFR expression: interaction between the human mineralocorticoid receptor and the human EGFR promoter. Am J Physiol Endocrinol Metab. 2007;292:E1790–800. doi: 10.1152/ajpendo.00708.2006. [DOI] [PubMed] [Google Scholar]
  58. Guo C., Martinez-Vasquez D., Mendez G. P., Toniolo M. F., Yao T. M., Oestreicher E. M., Kikuchi T., Lapointe N., Pojoga L., Williams G. H., Ricchiuti V., Adler G. K. Mineralocorticoid receptor antagonist reduces renal injury in rodent models of types 1 and 2 diabetes mellitus. Endocrinology. 2006;147:5363–73. doi: 10.1210/en.2006-0944. [DOI] [PubMed] [Google Scholar]
  59. Han F., Ozawa H., Matsuda K., Nishi M., Kawata M. Colocalization of mineralocorticoid receptor and glucocorticoid receptor in the hippocampus and hypothalamus. Neurosci Res. 2005;51:371–81. doi: 10.1016/j.neures.2004.12.013. [DOI] [PubMed] [Google Scholar]
  60. Herman J. P., Patel P. D., Akil H., Watson S. J. Localization and regulation of glucocorticoid and mineralocorticoid receptor messenger RNAs in the hippocampal formation of the rat. Mol Endocrinol. 1989;3:1886–94. doi: 10.1210/mend-3-11-1886. [DOI] [PubMed] [Google Scholar]
  61. Hirasawa G., Takeyama J., Sasano H., Fukushima K., Suzuki T., Muramatu Y., Darnel A. D., Kaneko C., Hiwatashi N., Toyota T., Nagura H., Krozowski Z. S. 11Beta-hydroxysteroid dehydrogenase type II and mineralocorticoid receptor in human placenta. J Clin Endocrinol Metab. 2000;85:1306–9. doi: 10.1210/jcem.85.3.6429. [DOI] [PubMed] [Google Scholar]
  62. Hirschberg D., Jagerbrink T., Samskog J., Gustafsson M., Stahlberg M., Alvelius G., Husman B., Carlquist M., Jornvall H., Bergman T. Detection of phosphorylated peptides in proteomic analyses using microfluidic compact disk technology. Anal Chem. 2004;76:5864–71. doi: 10.1021/ac040044g. [DOI] [PubMed] [Google Scholar]
  63. Hodgson Z. G., Meddle S. L., Roberts M. L., Buchanan K. L., Evans M. R., Metzdorf R., Gahr M., Healy S. D. Spatial ability is impaired and hippocampal mineralocorticoid receptor mRNA expression reduced in zebra finches (Taeniopygia guttata) selected for acute high corticosterone response to stress. Proc Biol Sci. 2007;274:239–45. doi: 10.1098/rspb.2006.3704. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Horisberger J. D., Lemas V., Kraehenbuhl J. P., Rossier B. C. Structure-function relationship of Na,K-ATPase. Annu Rev Physiol. 1991;53:565–84. doi: 10.1146/annurev.ph.53.030191.003025. [DOI] [PubMed] [Google Scholar]
  65. Hultman M. L., Krasnoperova N. V., Li S., Du S., Xia C., Dietz J. D., Lala D. S., Welsch D. J., Hu X. The ligand-dependent interaction of mineralocorticoid receptor with coactivator and corepressor peptides suggests multiple activation mechanisms. Mol Endocrinol. 2005;19:1460–73. doi: 10.1210/me.2004-0537. [DOI] [PubMed] [Google Scholar]
  66. Hu X., Funder J. W. The evolution of mineralocorticoid receptors. Mol Endocrinol. 2006;20:1471–8. doi: 10.1210/me.2005-0247. [DOI] [PubMed] [Google Scholar]
  67. Huyet J., Pinon G. M., Fay M. R., Fagart J., Rafestin-Oblin M. E. Structural basis of spirolactone recognition by the mineralocorticoid receptor. Mol Pharmacol. 2007;72:563–71. doi: 10.1124/mol.107.036459. [DOI] [PubMed] [Google Scholar]
  68. Iniguez-Lluhi J. A., Pearce D. A common motif within the negative regulatory regions of multiple factors inhibits their transcriptional synergy. Mol Cell Biol. 2000;20:6040–50. doi: 10.1128/mcb.20.16.6040-6050.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Jaffe I. Z., Tintut Y., Newfell B. G., Demer L. L., Mendelsohn M. E. Mineralocorticoid receptor activation promotes vascular cell calcification. Arterioscler Thromb Vasc Biol. 2007;27:799–805. doi: 10.1161/01.ATV.0000258414.59393.89. [DOI] [PubMed] [Google Scholar]
  70. Jalaguier S., Mornet D., Mesnier D., Leger J. J., Auzou G. Human mineralocorticoid receptor interacts with actin under mineralocorticoid ligand modulation. FEBS Lett. 1996;384:112–6. doi: 10.1016/0014-5793(96)00295-5. [DOI] [PubMed] [Google Scholar]
  71. Jorgensen P. L. Structure, function and regulation of Na,K-ATPase in the kidney. Kidney Int. 1986;29:10–20. doi: 10.1038/ki.1986.3. [DOI] [PubMed] [Google Scholar]
  72. Karst H., Berger S., Turiault M., Tronche F., Schutz G., Joels M. Mineralocorticoid receptors are indispensable for nongenomic modulation of hippocampal glutamate transmission by corticosterone. Proc Natl Acad Sci U S A. 2005;102:19204–7. doi: 10.1073/pnas.0507572102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Kenouch S., Lombes M., Delahaye F., Eugene E., Bonvalet J. P., Farman N. Human skin as target for aldosterone: coexpression of mineralocorticoid receptors and 11 β-hydroxysteroid dehydrogenase. J Clin Endocrinol Metab. 1994;79:1334–41. doi: 10.1210/jcem.79.5.7962326. [DOI] [PubMed] [Google Scholar]
  74. Kitagawa H., Yanagisawa J., Fuse H., Ogawa S., Yogiashi Y., Okuno A., Nagasawa H., Nakajima T., Matsumoto T., Kato S. Ligand-selective potentiation of rat mineralocorticoid receptor activation function 1 by a CBP-containing histone acetyltransferase complex. Mol Cell Biol. 2002;22:3698–706. doi: 10.1128/MCB.22.11.3698-3706.2002. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  75. Kolla V., Litwack G. Transcriptional regulation of the human Na/K ATPase via the human mineralocorticoid receptor. Mol Cell Biochem. 2000;204:35–40. doi: 10.1023/a:1007009700377. [DOI] [PubMed] [Google Scholar]
  76. Krozowski Z. S., Rundle S. E., Wallace C., Castell M. J., Shen J. H., Dowling J., Funder J. W., Smith A. I. Immunolocalization of renal mineralocorticoid receptors with an antiserum against a peptide deduced from the complementary deoxyribonucleic acid sequence. Endocrinology. 1989;125:192–8. doi: 10.1210/endo-125-1-192. [DOI] [PubMed] [Google Scholar]
  77. Krozowski Z., Funder J. W. Mineralocorticoid receptors in the rat lung. Endocrinology. 1981;109:1811–3. doi: 10.1210/endo-109-6-1811. [DOI] [PubMed] [Google Scholar]
  78. Kwak S. P., Patel P. D., Thompson R. C., Akil H., Watson S. J. 5'-Heterogeneity of the mineralocorticoid receptor messenger ribonucleic acid: differential expression and regulation of splice variants within the rat hippocampus. Endocrinology. 1993;133:2344–50. doi: 10.1210/endo.133.5.8404687. [DOI] [PubMed] [Google Scholar]
  79. Lai M., McCormick J. A., Chapman K. E., Kelly P. A., Seckl J. R., Yau J. L. Differential regulation of corticosteroid receptors by monoamine neurotransmitters and antidepressant drugs in primary hippocampal culture. Neuroscience. 2003;118:975–84. doi: 10.1016/s0306-4522(03)00038-1. [DOI] [PubMed] [Google Scholar]
  80. Le Menuet D., Isnard R., Bichara M., Viengchareun S., Muffat-Joly M., Walker F., Zennaro M. C., Lombes M. Alteration of cardiac and renal functions in transgenic mice overexpressing human mineralocorticoid receptor. J Biol Chem. 2001;276:38911–20. doi: 10.1074/jbc.M103984200. [DOI] [PubMed] [Google Scholar]
  81. Le Menuet D., Viengchareun S., Muffat-Joly M., Zennaro M. C., Lombes M. Expression and function of the human mineralocorticoid receptor: lessons from transgenic mouse models. Mol Cell Endocrinol. 2004;217:127–36. doi: 10.1016/j.mce.2003.10.045. [DOI] [PubMed] [Google Scholar]
  82. Le Menuet D., Viengchareun S., Penfornis P., Walker F., Zennaro M. C., Lombes M. Targeted oncogenesis reveals a distinct tissue-specific utilization of alternative promoters of the human mineralocorticoid receptor gene in transgenic mice. J Biol Chem. 2000;275:7878–86. doi: 10.1074/jbc.275.11.7878. [DOI] [PubMed] [Google Scholar]
  83. Le Moellic C., Ouvrard-Pascaud A., Capurro C., Cluzeaud F., Fay M., Jaisser F., Farman N., Blot-Chabaud M. Early nongenomic events in aldosterone action in renal collecting duct cells: PKCalpha activation, mineralocorticoid receptor phosphorylation, and cross-talk with the genomic response. J Am Soc Nephrol. 2004;15:1145–60. [PubMed] [Google Scholar]
  84. Leopold J. A., Dam A., Maron B. A., Scribner A. W., Liao R., Handy D. E., Stanton R. C., Pitt B., Loscalzo J. Aldosterone impairs vascular reactivity by decreasing glucose-6-phosphate dehydrogenase activity. Nat Med. 2007;13:189–97. doi: 10.1038/nm1545. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Lin J., Handschin C., Spiegelman B. M. Metabolic control through the PGC-1 family of transcription coactivators. Cell Metab. 2005;1:361–70. doi: 10.1016/j.cmet.2005.05.004. [DOI] [PubMed] [Google Scholar]
  86. Lin D. Y., Huang Y. S., Jeng J. C., Kuo H. Y., Chang C. C., Chao T. T., Ho C. C., Chen Y. C., Lin T. P., Fang H. I., Hung C. C., Suen C. S., Hwang M. J., Chang K. S., Maul G. G., Shih H. M. Role of SUMO-interacting motif in Daxx SUMO modification, subnuclear localization, and repression of sumoylated transcription factors. Mol Cell. 2006;24:341–54. doi: 10.1016/j.molcel.2006.10.019. [DOI] [PubMed] [Google Scholar]
  87. Li Y., Suino K., Daugherty J., Xu H. E. Structural and biochemical mechanisms for the specificity of hormone binding and coactivator assembly by mineralocorticoid receptor. Mol Cell. 2005;19:367–80. doi: 10.1016/j.molcel.2005.06.026. [DOI] [PubMed] [Google Scholar]
  88. Liu W., Wang J., Sauter N. K., Pearce D. Steroid receptor heterodimerization demonstrated in vitro and in vivo . Proc Natl Acad Sci U S A. 1995;92:12480–4. doi: 10.1073/pnas.92.26.12480. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Lombes M., Claire M., Pinto M., Michaud A., Rafestin-Oblin M. E. Aldosterone binding in the human colon carcinoma cell line HT29: correlation with cell differentiation. J Steroid Biochem. 1984;20:329–33. doi: 10.1016/0022-4731(84)90227-9. [DOI] [PubMed] [Google Scholar]
  90. Lombes M., Binart N., Delahaye F., Baulieu E. E., Rafestin-Oblin M. E. Differential intracellular localization of human mineralocorticosteroid receptor on binding of agonists and antagonists. Biochem J. 1994a;302 ( Pt 1):191–7. doi: 10.1042/bj3020191. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Lombes M., Oblin M. E., Gasc J. M., Baulieu E. E., Farman N., Bonvalet J. P. Immunohistochemical and biochemical evidence for a cardiovascular mineralocorticoid receptor. Circ Res. 1992;71:503–10. doi: 10.1161/01.res.71.3.503. [DOI] [PubMed] [Google Scholar]
  92. Lombes M., Farman N., Oblin M. E., Baulieu E. E., Bonvalet J. P., Erlanger B. F., Gasc J. M. Immunohistochemical localization of renal mineralocorticoid receptor by using an anti-idiotypic antibody that is an internal image of aldosterone. Proc Natl Acad Sci U S A. 1990;87:1086–8. doi: 10.1073/pnas.87.3.1086. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Lombes M., Kenouch S., Souque A., Farman N., Rafestin-Oblin M. E. The mineralocorticoid receptor discriminates aldosterone from glucocorticoids independently of the 11 β-hydroxysteroid dehydrogenase. Endocrinology. 1994b;135:834–40. doi: 10.1210/endo.135.3.8070376. [DOI] [PubMed] [Google Scholar]
  94. Lu N. Z., Cidlowski J. A. Translational regulatory mechanisms generate N-terminal glucocorticoid receptor isoforms with unique transcriptional target genes. Mol Cell. 2005;18:331–42. doi: 10.1016/j.molcel.2005.03.025. [DOI] [PubMed] [Google Scholar]
  95. Makhanova N., Sequeira-Lopez M. L., Gomez R. A., Kim H. S., Smithies O. Disturbed homeostasis in sodium-restricted mice heterozygous and homozygous for aldosterone synthase gene disruption. Hypertension. 2006;48:1151–9. doi: 10.1161/01.HYP.0000249902.09036.e7. [DOI] [PubMed] [Google Scholar]
  96. Marissal-Arvy N., Lombes M., Petterson J., Moisan M. P., Mormede P. Gain of function mutation in the mineralocorticoid receptor of the Brown Norway rat. J Biol Chem. 2004;279:39232–9. doi: 10.1074/jbc.M407436200. [DOI] [PubMed] [Google Scholar]
  97. Marver D., Stewart J., Funder J. W., Feldman D., Edelman I. S. Renal aldosterone receptors: studies with (3H)aldosterone and the anti-mineralocorticoid (3H)spirolactone (SC-26304) Proc Natl Acad Sci U S A. 1974;71:1431–5. doi: 10.1073/pnas.71.4.1431. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Masilamani S., Kim G. H., Mitchell C., Wade J. B., Knepper M. A. Aldosterone-mediated regulation of ENaC α, β, and γ subunit proteins in rat kidney. J Clin Invest. 1999;104:R19–23. doi: 10.1172/JCI7840. [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Metivier R., Penot G., Hubner M. R., Reid G., Brand H., Kos M., Gannon F. Estrogen receptor-α directs ordered, cyclical, and combinatorial recruitment of cofactors on a natural target promoter. Cell. 2003;115:751–63. doi: 10.1016/s0092-8674(03)00934-6. [DOI] [PubMed] [Google Scholar]
  100. Mirshahi M., Mirshahi A., Sedighian R., Hecquet C., Faure J. P., Agarwal M. K. Immunochemical demonstration of the mineralocorticoid receptor in ocular tissues. Neuroendocrinology. 1997;65:70–8. doi: 10.1159/000127166. [DOI] [PubMed] [Google Scholar]
  101. Miyata K., Rahman M., Shokoji T., Nagai Y., Zhang G. X., Sun G. P., Kimura S., Yukimura T., Kiyomoto H., Kohno M., Abe Y., Nishiyama A. Aldosterone stimulates reactive oxygen species production through activation of NADPH oxidase in rat mesangial cells. J Am Soc Nephrol. 2005;16:2906–12. doi: 10.1681/ASN.2005040390. [DOI] [PubMed] [Google Scholar]
  102. Morrison N., Harrap S. B., Arriza J. L., Boyd E., Connor J. M. Regional chromosomal assignment of the human mineralocorticoid receptor gene to 4q31.1. Hum Genet. 1990;85:130–2. doi: 10.1007/BF00276340. [DOI] [PubMed] [Google Scholar]
  103. Nagai Y., Miyata K., Sun G. P., Rahman M., Kimura S., Miyatake A., Kiyomoto H., Kohno M., Abe Y., Yoshizumi M., Nishiyama A. Aldosterone stimulates collagen gene expression and synthesis via activation of ERK1/2 in rat renal fibroblasts. Hypertension. 2005;46:1039–45. doi: 10.1161/01.HYP.0000174593.88899.68. [DOI] [PubMed] [Google Scholar]
  104. Nakamura Y., Suzuki S., Suzuki T., Ono K., Miura I., Satoh F., Moriya T., Saito H., Yamada S., Ito S., Sasano H. MDM2: a novel mineralocorticoid-responsive gene involved in aldosterone-induced human vascular structural remodeling. Am J Pathol. 2006;169:362–71. doi: 10.2353/ajpath.2006.051351. [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Naray-Fejes-Toth A., Fejes-Toth G. Novel mouse strain with Cre recombinase in 11beta-hydroxysteroid dehydrogenase-2-expressing cells. Am J Physiol Renal Physiol. 2007;292:F486–94. doi: 10.1152/ajprenal.00188.2006. [DOI] [PubMed] [Google Scholar]
  106. Naray-Fejes-Toth A., Canessa C., Cleaveland E. S., Aldrich G., Fejes-Toth G. sgk is an aldosterone-induced kinase in the renal collecting duct. Effects on epithelial na+ channels. J Biol Chem. 1999;274:16973–8. doi: 10.1074/jbc.274.24.16973. [DOI] [PubMed] [Google Scholar]
  107. Naray-Fejes-Toth A., Snyder P. M., Fejes-Toth G. The kidney-specific WNK1 isoform is induced by aldosterone and stimulates epithelial sodium channel-mediated Na+ transport. Proc Natl Acad Sci U S A. 2004;101:17434–9. doi: 10.1073/pnas.0408146101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  108. Nishiyama A., Yao L., Fan Y., Kyaw M., Kataoka N., Hashimoto K., Nagai Y., Nakamura E., Yoshizumi M., Shokoji T., Kimura S., Kiyomoto H., Tsujioka K., Kohno M., Tamaki T., Kajiya F., Abe Y. Involvement of aldosterone and mineralocorticoid receptors in rat mesangial cell proliferation and deformability. Hypertension. 2005;45:710–6. doi: 10.1161/01.HYP.0000154681.38944.9a. [DOI] [PubMed] [Google Scholar]
  109. Oelkers W. Drospirenone, a progestogen with antimineralocorticoid properties: a short review. Mol Cell Endocrinol. 2004;217:255–61. doi: 10.1016/j.mce.2003.10.030. [DOI] [PubMed] [Google Scholar]
  110. O'Malley B. W. Coregulators: from whence came these "master genes". Mol Endocrinol. 2007;21:1009–13. doi: 10.1210/me.2007-0012. [DOI] [PubMed] [Google Scholar]
  111. Onate S. A., Tsai S. Y., Tsai M. J., O'Malley B. W. Sequence and characterization of a coactivator for the steroid hormone receptor superfamily. Science. 1995;270:1354–7. doi: 10.1126/science.270.5240.1354. [DOI] [PubMed] [Google Scholar]
  112. Ouvrard-Pascaud A., Sainte-Marie Y., Benitah J. P., Perrier R., Soukaseum C., Cat A. N., Royer A., Le Quang K., Charpentier F., Demolombe S., Mechta-Grigoriou F., Beggah A. T., Maison-Blanche P., Oblin M. E., Delcayre C., Fishman G. I., Farman N., Escoubet B., Jaisser F. Conditional mineralocorticoid receptor expression in the heart leads to life-threatening arrhythmias. Circulation. 2005;111:3025–33. doi: 10.1161/CIRCULATIONAHA.104.503706. [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Pascual-Le Tallec L., Demange C., Lombes M. Human mineralocorticoid receptor A and B protein forms produced by alternative translation sites display different transcriptional activities. Eur J Endocrinol. 2004;150:585–90. doi: 10.1530/eje.0.1500585. [DOI] [PubMed] [Google Scholar]
  114. Pascual-Le Tallec L. , Kirsh O., Lecomte M. C., Viengchareun S., Zennaro M. C., Dejean A., Lombes M. Protein inhibitor of activated signal transducer and activator of transcription 1 interacts with the N-terminal domain of mineralocorticoid receptor and represses its transcriptional activity: implication of small ubiquitin-related modifier 1 modification. Mol Endocrinol. 2003;17:2529–42. doi: 10.1210/me.2003-0299. [DOI] [PubMed] [Google Scholar]
  115. Pascual-Le Tallec L., Simone F., Viengchareun S., Meduri G., Thirman M. J., Lombes M. The elongation factor ELL (eleven-nineteen lysine-rich leukemia) is a selective coregulator for steroid receptor functions. Mol Endocrinol. 2005b;19:1158–69. doi: 10.1210/me.2004-0331. [DOI] [PubMed] [Google Scholar]
  116. Pascual-Le Tallec L., Lombes M. The mineralocorticoid receptor: a journey exploring its diversity and specificity of action. Mol Endocrinol. 2005a;19:2211–21. doi: 10.1210/me.2005-0089. [DOI] [PubMed] [Google Scholar]
  117. Patel P. D., Lopez J. F., Lyons D. M., Burke S., Wallace M., Schatzberg A. F. Glucocorticoid and mineralocorticoid receptor mRNA expression in squirrel monkey brain. J Psychiatr Res. 2000;34:383–92. doi: 10.1016/s0022-3956(00)00035-2. [DOI] [PubMed] [Google Scholar]
  118. Patel P. D., Sherman T. G., Goldman D. J., Watson S. J. Molecular cloning of a mineralocorticoid (type I) receptor complementary DNA from rat hippocampus. Mol Endocrinol. 1989;3:1877–85. doi: 10.1210/mend-3-11-1877. [DOI] [PubMed] [Google Scholar]
  119. Pearce P., Funder J. W. High affinity aldosterone binding sites (type I receptors) in rat heart. Clin Exp Pharmacol Physiol. 1987;14:859–66. doi: 10.1111/j.1440-1681.1987.tb02422.x. [DOI] [PubMed] [Google Scholar]
  120. Penfornis P., Viengchareun S., Le Menuet D., Cluzeaud F., Zennaro M. C., Lombes M. The mineralocorticoid receptor mediates aldosterone-induced differentiation of T37i cells into brown adipocytes. Am J Physiol Endocrinol Metab. 2000;279:E386–94. doi: 10.1152/ajpendo.2000.279.2.E386. [DOI] [PubMed] [Google Scholar]
  121. Pinon G. M., Fagart J., Souque A., Auzou G., Vandewalle A., Rafestin-Oblin M. E. Identification of steroid ligands able to inactivate the mineralocorticoid receptor harboring the S810L mutation responsible for a severe form of hypertension. Mol Cell Endocrinol. 2004;217:181–8. doi: 10.1016/j.mce.2003.10.053. [DOI] [PubMed] [Google Scholar]
  122. Pitovski D. Z., Drescher M. J., Drescher D. G. High affinity aldosterone binding sites (type I receptors) in the mammalian inner ear. Hear Res. 1993;69:10–4. doi: 10.1016/0378-5955(93)90088-i. [DOI] [PubMed] [Google Scholar]
  123. Pitt B., Remme W., Zannad F., Neaton J., Martinez F., Roniker B., Bittman R., Hurley S., Kleiman J., Gatlin M. Eplerenone, a selective aldosterone blocker, in patients with left ventricular dysfunction after myocardial infarction. N Engl J Med. 2003;348:1309–21. doi: 10.1056/NEJMoa030207. [DOI] [PubMed] [Google Scholar]
  124. Pitt B., Zannad F., Remme W. J., Cody R., Castaigne A., Perez A., Palensky J., Wittes J. The effect of spironolactone on morbidity and mortality in patients with severe heart failure. Randomized Aldactone Evaluation Study Investigators. N Engl J Med. 1999;341:709–17. doi: 10.1056/NEJM199909023411001. [DOI] [PubMed] [Google Scholar]
  125. Planey S. L., Derfoul A., Steplewski A., Robertson N. M., Litwack G. Inhibition of glucocorticoid-induced apoptosis in 697 pre-B lymphocytes by the mineralocorticoid receptor N-terminal domain. J Biol Chem. 2002;277:42188–96. doi: 10.1074/jbc.M205085200. [DOI] [PubMed] [Google Scholar]
  126. Porter T. E., Ghavam S., Muchow M., Bossis I., Ellestad L. Cloning of partial cDNAs for the chicken glucocorticoid and mineralocorticoid receptors and characterization of mRNA levels in the anterior pituitary gland during chick embryonic development. Domest Anim Endocrinol. 2007;33:226–39. doi: 10.1016/j.domaniend.2006.05.006. [DOI] [PubMed] [Google Scholar]
  127. Porter G. A., Edelman I. S. The Action of Aldosterone and Related Corticosteroids on Sodium Transport across the Toad Bladder. J Clin Invest. 1964;43:611–20. doi: 10.1172/JCI104946. [DOI] [PMC free article] [PubMed] [Google Scholar]
  128. Pratt W. B., Toft D. O. Steroid receptor interactions with heat shock protein and immunophilin chaperones. Endocr Rev. 1997;18:306–60. doi: 10.1210/edrv.18.3.0303. [DOI] [PubMed] [Google Scholar]
  129. Pressley L., Funder J. W. Glucocorticoid and mineralocorticoid receptors in gut mucosa. Endocrinology. 1975;97:588–96. doi: 10.1210/endo-97-3-588. [DOI] [PubMed] [Google Scholar]
  130. Pryce C. R., Feldon J., Fuchs E., Knuesel I., Oertle T., Sengstag C., Spengler M., Weber E., Weston A., Jongen-Relo A. Postnatal ontogeny of hippocampal expression of the mineralocorticoid and glucocorticoid receptors in the common marmoset monkey. Eur J Neurosci. 2005;21:1521–35. doi: 10.1111/j.1460-9568.2005.04003.x. [DOI] [PubMed] [Google Scholar]
  131. Pujo L., Fagart J., Gary F., Papadimitriou D. T., Claes A., Jeunemaitre X., Zennaro M. C. Mineralocorticoid receptor mutations are the principal cause of renal type 1 pseudohypoaldosteronism. Hum Mutat. 2007;28:33–40. doi: 10.1002/humu.20371. [DOI] [PubMed] [Google Scholar]
  132. Qin W., Rudolph A. E., Bond B. R., Rocha R., Blomme E. A., Goellner J. J., Funder J. W., McMahon E. G. Transgenic model of aldosterone-driven cardiac hypertrophy and heart failure. Circ Res. 2003;93:69–76. doi: 10.1161/01.RES.0000080521.15238.E5. [DOI] [PubMed] [Google Scholar]
  133. Quinkler M., Meyer B., Bumke-Vogt C., Grossmann C., Gruber U., Oelkers W., Diederich S., Bahr V. Agonistic and antagonistic properties of progesterone metabolites at the human mineralocorticoid receptor. Eur J Endocrinol. 2002;146:789–99. doi: 10.1530/eje.0.1460789. [DOI] [PubMed] [Google Scholar]
  134. Quinkler M., Zehnder D., Eardley K. S., Lepenies J., Howie A. J., Hughes S. V., Cockwell P., Hewison M., Stewart P. M. Increased expression of mineralocorticoid effector mechanisms in kidney biopsies of patients with heavy proteinuria. Circulation. 2005;112:1435–43. doi: 10.1161/CIRCULATIONAHA.105.539122. [DOI] [PubMed] [Google Scholar]
  135. Quinkler M., Meyer B., Oelkers W., Diederich S. Renal inactivation, mineralocorticoid generation, and 11beta-hydroxysteroid dehydrogenase inhibition ameliorate the antimineralocorticoid effect of progesterone in vivo . J Clin Endocrinol Metab. 2003;88:3767–72. doi: 10.1210/jc.2003-030092. [DOI] [PubMed] [Google Scholar]
  136. Rafestin-Oblin M. E., Couette B., Radanyi C., Lombes M., Baulieu E. E. Mineralocorticosteroid receptor of the chick intestine. Oligomeric structure and transformation. J Biol Chem. 1989;264:9304–9. [PubMed] [Google Scholar]
  137. Rafestin-Oblin M. E., Souque A., Bocchi B., Pinon G., Fagart J., Vandewalle A. The severe form of hypertension caused by the activating S810L mutation in the mineralocorticoid receptor is cortisone related. Endocrinology. 2003;144:528–33. doi: 10.1210/en.2002-220708. [DOI] [PubMed] [Google Scholar]
  138. Robert-Nicoud M., Flahaut M., Elalouf J. M., Nicod M., Salinas M., Bens M., Doucet A., Wincker P., Artiguenave F., Horisberger J. D., Vandewalle A., Rossier B. C., Firsov D. Transcriptome of a mouse kidney cortical collecting duct cell line: effects of aldosterone and vasopressin. Proc Natl Acad Sci U S A. 2001;98:2712–6. doi: 10.1073/pnas.051603198. [DOI] [PMC free article] [PubMed] [Google Scholar]
  139. Rogerson F. M., Yao Y. Z., Elsass R. E., Dimopoulos N., Smith B. J., Fuller P. J. A critical region in the mineralocorticoid receptor for aldosterone binding and activation by cortisol: evidence for a common mechanism governing ligand binding specificity in steroid hormone receptors. Mol Endocrinol. 2007;21:817–28. doi: 10.1210/me.2006-0246. [DOI] [PubMed] [Google Scholar]
  140. Rondinone C. M., Rodbard D., Baker M. E. Aldosterone stimulated differentiation of mouse 3T3-L1 cells into adipocytes. Endocrinology. 1993;132:2421–6. doi: 10.1210/endo.132.6.8504747. [DOI] [PubMed] [Google Scholar]
  141. Ronzaud C., Loffing J., Bleich M., Gretz N., Grone H. J., Schutz G., Berger S. Impairment of sodium balance in mice deficient in renal principal cell mineralocorticoid receptor. J Am Soc Nephrol. 2007;18:1679–87. doi: 10.1681/ASN.2006090975. [DOI] [PubMed] [Google Scholar]
  142. Rosenfeld M. G., Lunyak V. V., Glass C. K. Sensors and signals: a coactivator/corepressor/epigenetic code for integrating signal-dependent programs of transcriptional response. Genes Dev. 2006;20:1405–28. doi: 10.1101/gad.1424806. [DOI] [PubMed] [Google Scholar]
  143. Rossier B. C., Pradervand S., Schild L., Hummler E. Epithelial sodium channel and the control of sodium balance: interaction between genetic and environmental factors. Annu Rev Physiol. 2002;64:877–97. doi: 10.1146/annurev.physiol.64.082101.143243. [DOI] [PubMed] [Google Scholar]
  144. Rozeboom A. M., Akil H., Seasholtz A. F. Mineralocorticoid receptor overexpression in forebrain decreases anxiety-like behavior and alters the stress response in mice. Proc Natl Acad Sci U S A. 2007;104:4688–93. doi: 10.1073/pnas.0606067104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  145. Sainte Marie Y., Toulon A., Paus R., Maubec E., Cherfa A., Grossin M., Descamps V., Clemessy M., Gasc J. M., Peuchmaur M., Glick A., Farman N., Jaisser F. Targeted skin overexpression of the mineralocorticoid receptor in mice causes epidermal atrophy, premature skin barrier formation, eye abnormalities, and alopecia. Am J Pathol. 2007;171:846–60. doi: 10.2353/ajpath.2007.060991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  146. Savory J. G., Prefontaine G. G., Lamprecht C., Liao M., Walther R. F., Lefebvre Y. A., Hache R. J. Glucocorticoid receptor homodimers and glucocorticoid-mineralocorticoid receptor heterodimers form in the cytoplasm through alternative dimerization interfaces. Mol Cell Biol. 2001;21:781–93. doi: 10.1128/MCB.21.3.781-793.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  147. Schwartz B., Wysocki A. Mineralocorticoid receptors in the rabbit iris-ciliary body. Ophthalmic Res. 1997;29:42–7. doi: 10.1159/000267990. [DOI] [PubMed] [Google Scholar]
  148. Seeler J. S., Dejean A. Nuclear and unclear functions of SUMO. Nat Rev Mol Cell Biol. 2003;4:690–9. doi: 10.1038/nrm1200. [DOI] [PubMed] [Google Scholar]
  149. Shibata S., Nagase M., Yoshida S., Kawachi H., Fujita T. Podocyte as the target for aldosterone: roles of oxidative stress and Sgk1. Hypertension. 2007;49:355–64. doi: 10.1161/01.HYP.0000255636.11931.a2. [DOI] [PubMed] [Google Scholar]
  150. So A. Y., Chaivorapol C., Bolton E. C., Li H., Yamamoto K. R. Determinants of cell- and gene-specific transcriptional regulation by the glucocorticoid receptor. PLoS Genet. 2007;3:e94. doi: 10.1371/journal.pgen.0030094. [DOI] [PMC free article] [PubMed] [Google Scholar]
  151. Soundararajan R., Zhang T. T., Wang J., Vandewalle A., Pearce D. A novel role for glucocorticoid-induced leucine zipper protein in epithelial sodium channel-mediated sodium transport. J Biol Chem. 2005;280:39970–81. doi: 10.1074/jbc.M508658200. [DOI] [PubMed] [Google Scholar]
  152. Stockand J. D., Meszaros J. G. Aldosterone stimulates proliferation of cardiac fibroblasts by activating Ki-RasA and MAPK1/2 signaling. Am J Physiol Heart Circ Physiol. 2003;284:H176–84. doi: 10.1152/ajpheart.00421.2002. [DOI] [PubMed] [Google Scholar]
  153. Sturm A., Bury N., Dengreville L., Fagart J., Flouriot G., Rafestin-Oblin M. E., Prunet P. 11-deoxycorticosterone is a potent agonist of the rainbow trout (Oncorhynchus mykiss) mineralocorticoid receptor. Endocrinology. 2005;146:47–55. doi: 10.1210/en.2004-0128. [DOI] [PubMed] [Google Scholar]
  154. Sugiyama T., Yoshimoto T., Hirono Y., Suzuki N., Sakurada M., Tsuchiya K., Minami I., Iwashima F., Sakai H., Tateno T., Sato R., Hirata Y. Aldosterone increases osteopontin gene expression in rat endothelial cells. Biochem Biophys Res Commun. 2005a;336:163–7. doi: 10.1016/j.bbrc.2005.08.056. [DOI] [PubMed] [Google Scholar]
  155. Sugiyama T., Yoshimoto T., Tsuchiya K., Gochou N., Hirono Y., Tateno T., Fukai N., Shichiri M., Hirata Y. Aldosterone induces angiotensin converting enzyme gene expression via a JAK2-dependent pathway in rat endothelial cells. Endocrinology. 2005b;146:3900–6. doi: 10.1210/en.2004-1674. [DOI] [PubMed] [Google Scholar]
  156. Takeda A. N., Pinon G. M., Bens M., Fagart J., Rafestin-Oblin M. E., Vandewalle A. The synthetic androgen methyltrienolone (r1881) acts as a potent antagonist of the mineralocorticoid receptor. Mol Pharmacol. 2007;71:473–82. doi: 10.1124/mol.106.031112. [DOI] [PubMed] [Google Scholar]
  157. Teixeira M., Viengchareun S., Butlen D., Ferreira C., Cluzeaud F., Blot-Chabaud M., Lombes M., Ferrary E. Functional IsK/KvLQT1 potassium channel in a new corticosteroid-sensitive cell line derived from the inner ear. J Biol Chem. 2006;281:10496–507. doi: 10.1074/jbc.M512254200. [DOI] [PubMed] [Google Scholar]
  158. Tirard M., Almeida O. F., Hutzler P., Melchior F., Michaelidis T. M. Sumoylation and proteasomal activity determine the transactivation properties of the mineralocorticoid receptor. Mol Cell Endocrinol. 2007;268:20–9. doi: 10.1016/j.mce.2007.01.010. [DOI] [PubMed] [Google Scholar]
  159. Van Eekelen J. A., Jiang W., De Kloet E. R., Bohn M. C. Distribution of the mineralocorticoid and the glucocorticoid receptor mRNAs in the rat hippocampus. J Neurosci Res. 1988;21:88–94. doi: 10.1002/jnr.490210113. [DOI] [PubMed] [Google Scholar]
  160. Viengchareun S., Penfornis P., Zennaro M. C., Lombes M. Mineralocorticoid and glucocorticoid receptors inhibit UCP expression and function in brown adipocytes. Am J Physiol Endocrinol Metab. 2001;280:E640–9. doi: 10.1152/ajpendo.2001.280.4.E640. [DOI] [PubMed] [Google Scholar]
  161. Wald H., Goldstein O., Asher C., Yagil Y., Garty H. Aldosterone induction and epithelial distribution of CHIF. Am J Physiol. 1996;271:F322–9. doi: 10.1152/ajprenal.1996.271.2.F322. [DOI] [PubMed] [Google Scholar]
  162. Walther R. F., Atlas E., Carrigan A., Rouleau Y., Edgecombe A., Visentin L., Lamprecht C., Addicks G. C., Hache R. J., Lefebvre Y. A. A serine/threonine-rich motif is one of three nuclear localization signals that determine unidirectional transport of the mineralocorticoid receptor to the nucleus. J Biol Chem. 2005;280:17549–61. doi: 10.1074/jbc.M501548200. [DOI] [PubMed] [Google Scholar]
  163. Wang X., Skelley L., Cade R., Sun Z. AAV delivery of mineralocorticoid receptor shRNA prevents progression of cold-induced hypertension and attenuates renal damage. Gene Ther. 2006;13:1097–103. doi: 10.1038/sj.gt.3302768. [DOI] [PubMed] [Google Scholar]
  164. Wielputz M. O., Lee I. H., Dinudom A., Boulkroun S., Farman N., Cook D. I., Korbmacher C., Rauh R. (NDRG2) stimulates amiloride-sensitive Na+ currents in Xenopus laevis oocytes and fisher rat thyroid cells. J Biol Chem. 2007;282:28264–73. doi: 10.1074/jbc.M702168200. [DOI] [PubMed] [Google Scholar]
  165. Wong S., Brennan F. E., Young M. J., Fuller P. J., Cole T. J. A direct effect of aldosterone on endothelin-1 gene expression in vivo . Endocrinology. 2007;148:1511–7. doi: 10.1210/en.2006-0965. [DOI] [PubMed] [Google Scholar]
  166. Xu B. E., Stippec S., Lazrak A., Huang C. L., Cobb M. H. WNK1 activates SGK1 by a phosphatidylinositol 3-kinase-dependent and non-catalytic mechanism. J Biol Chem. 2005b;280:34218–23. doi: 10.1074/jbc.M505735200. [DOI] [PubMed] [Google Scholar]
  167. Xu B. E., Stippec S., Chu P. Y., Lazrak A., Li X. J., Lee B. H., English J. M., Ortega B., Huang C. L., Cobb M. H. WNK1 activates SGK1 to regulate the epithelial sodium channel. Proc Natl Acad Sci U S A. 2005a;102:10315–20. doi: 10.1073/pnas.0504422102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  168. Yokota K., Shibata H., Kurihara I., Kobayashi S., Suda N., Murai-Takeda A., Saito I., Kitagawa H., Kato S., Saruta T., Itoh H. Coactivation of the N-terminal transactivation of mineralocorticoid receptor by Ubc9. J Biol Chem. 2007;282:1998–2010. doi: 10.1074/jbc.M607741200. [DOI] [PubMed] [Google Scholar]
  169. Yoshida M., Ma J., Tomita T., Morikawa N., Tanaka N., Masamura K., Kawai Y., Miyamori I. Mineralocorticoid receptor is overexpressed in cardiomyocytes of patients with congestive heart failure. Congest Heart Fail. 2005;11:12–6. doi: 10.1111/j.1527-5299.2005.03722.x. [DOI] [PubMed] [Google Scholar]
  170. Young M., Fullerton M., Dilley R., Funder J. Mineralocorticoids, hypertension, and cardiac fibrosis. J Clin Invest. 1994;93:2578–83. doi: 10.1172/JCI117269. [DOI] [PMC free article] [PubMed] [Google Scholar]
  171. Yuan J., Jia R., Bao Y. Aldosterone up-regulates production of plasminogen activator inhibitor-1 by renal mesangial cells. J Biochem Mol Biol. 2007;40:180–8. doi: 10.5483/bmbrep.2007.40.2.180. [DOI] [PubMed] [Google Scholar]
  172. Zennaro M. C., Le Menuet D., Lombes M. Characterization of the human mineralocorticoid receptor gene 5'-regulatory region: evidence for differential hormonal regulation of two alternative promoters via nonclassical mechanisms. Mol Endocrinol. 1996;10:1549–60. doi: 10.1210/mend.10.12.8961265. [DOI] [PubMed] [Google Scholar]
  173. Zennaro M. C., Le Menuet D., Viengchareun S., Walker F., Ricquier D., Lombes M. Hibernoma development in transgenic mice identifies brown adipose tissue as a novel target of aldosterone action. J Clin Invest. 1998;101:1254–60. doi: 10.1172/JCI1915. [DOI] [PMC free article] [PubMed] [Google Scholar]
  174. Zennaro M. C., Keightley M. C., Kotelevtsev Y., Conway G. S., Soubrier F., Fuller P. J. Human mineralocorticoid receptor genomic structure and identification of expressed isoforms. J Biol Chem. 1995;270:21016–20. doi: 10.1074/jbc.270.36.21016. [DOI] [PubMed] [Google Scholar]

Articles from Nuclear Receptor Signaling are provided here courtesy of SAGE Publications

RESOURCES