©1996 by The American Society for Biochemistry and Molecular Biology, Inc.
Genomic Organization and Glucocorticoid Transcriptional Activation of the Rat Na/H Exchanger Nhe3 Gene (*)

(Received for publication, February 8, 1996)

Ramani A. Kandasamy John Orlowski (§)

From the Department of Physiology, McGill University, Montréal, Québec H3G 1Y6, Canada

ABSTRACT
INTRODUCTION
EXPERIMENTAL PROCEDURES
RESULTS
DISCUSSION
FOOTNOTES
ACKNOWLEDGEMENTS
REFERENCES

ABSTRACT

The activity of the apical membrane Na/H exchanger NHE3 isoform of renal or intestinal epithelial cells is chronically regulated by a wide variety of stimuli, including acidosis, cAMP, glucocorticoids, and thyroid hormone. To understand the molecular mechanisms responsible for long term regulation of this cation transporter, we have isolated and determined the structure of this gene from a rat genomic library. The Nhe3 gene spans >40 kilobases and contains 17 exons that are flanked by typical splice donor and acceptor sequences at the exon-intron boundaries. The transcription initiation site was mapped by S1 nuclease protection analyses of mRNA from rat kidney and intestine. Multiple start sites were clustered between nucleotides -100 and -96 relative to the translation initiation codon. An atypical TATA-box and CCAAT-box are centered 30 and 147 nucleotides, respectively, upstream of the predominant transcription initiation site. Sequence analysis of approximately 1.4 kilobases of the 5`-flanking promoter region also revealed the presence of other putative cis-acting elements recognized by various transcription factors (e.g. AP-1, AP-2, C/EBP, NF-I, OCT-1/OTF-1, PEA3, Sp1, glucocorticoid, and thyroid hormone receptors), some of which may participate in the chronic regulation of this gene. The glucocorticoid responsiveness of the Nhe3 gene was assessed by fusing its 5` regulatory region to the firefly luciferase reporter gene and then by measuring the expression of the chimeric gene in transiently transfected renal epithelial OK and LLC-PK(1) cells. Glucocorticoid treatment significantly increased the luciferase activity of the chimeric gene in both cell lines, thereby indicating that glucocorticoid regulation of Nhe3 is mediated primarily by a transcriptional mechanism.


INTRODUCTION

The Na/H exchanger (NHE) (^1)is an integral membrane protein present in all mammalian cells. Multiple isoforms (NHE1 to NHE5) have been identified by molecular cloning techniques(1, 2, 3, 4, 5, 6, 7) . They range in size between 81 and 93 kDa and appear to exist in the membrane as homodimers, at least in the cases of NHE1 and NHE3(8) . These isoforms exhibit differences in their patterns of tissue expression, biochemical and pharmacological characteristics, and physiological functions (reviewed in (9) and (10) ).

In addition, Na/H exchanger activity is influenced by a wide variety of molecular signals (e.g. neurotransmitters, growth factors, peptide hormones, phorbol esters, cAMP, chemotactic factors, lectins, osmotic shrinkage, acidosis, glucocorticoids, and thyroid hormone) that act either rapidly (seconds to minutes) or following a considerable latent period (hours to days) before the effects on the rate of transport are manifested (reviewed in (11) and (12) ). Recent studies have begun to identify the stimuli and signaling pathways that acutely modulate the individual NHE isoforms (13, 14, 15, 16, 17, 18, 19, 20) . The mechanisms responsible for these alterations, although not fully resolved, appear to involve both phosphorylation-dependent and -independent processes.

Considerably less is known about the molecular mechanisms involved in the chronic regulation of the NHE isoforms, although recent investigations are beginning to provide some insight. Prolonged exposure of cultured renal epithelial cells to acidic medium, which serves as a paradigm for studying physiological adaptations to chronic metabolic and respiratory acidosis, elevates the activities and/or mRNA abundances of NHE1 (21, 22) and NHE3(22, 23) , but inhibits those of NHE2(24) . Interestingly, acid-mediated stimulation of NHE1 in renal MCT cells is associated with activation of protein kinase C and the transcription factor AP-1(25) , whereas acid induction of NHE3 in renal OKP cells is linked to the c-src family of non-receptor protein-tyrosine kinases(26) . Aside from acidosis, chronic incubation of renal IMCD cells in hyperosmotic media also stimulates Na/H exchanger activity, which is associated with increased NHE2 and reduced NHE1 mRNA abundances (24) . Last, long-term administration of glucocorticoids to rabbits or sheep selectively elevates NHE3 activity and mRNA in renal proximal convoluted tubules (27, 28) and in ileum(29) , but has no effect on NHE1 or NHE2. At present, the underlying mechanisms for acid, hyperosmolarity, and glucocorticoid regulation of specific isoform mRNA abundances have yet to be resolved in detail, but are likely to involve altered rates of gene transcription and/or mRNA stability. Further progress in this area will require the isolation and characterization of the NHE genes and their associated cis-acting DNA regulatory elements.

Chromosomal mapping analyses have revealed that the members of the Na/H exchanger gene family are dispersed in the mammalian genome(4, 30, 31, 32) . To date, the complete gene encoding human NHE1 has been characterized (33) and shown to bind several transcription factors in its 5`-flanking region(34) , although their functional importance was not established. However, more recent studies have shown that binding of the transcription factor AP-2 to the mouse Nhe1 gene promoter increases gene transcription(35) . No information is available regarding the other isoforms.

In order to understand the molecular mechanisms involved in the tissue-specific, developmental, and hormonal regulation of the NHE isoforms, we have characterized the genomic organization of the rat Nhe3 gene. Furthermore, we demonstrate that the transcriptional activity of the 5`-flanking promoter region is significantly elevated in response to glucocorticoid stimulation of transiently transfected renal cell lines.


EXPERIMENTAL PROCEDURES

Materials

Restriction endonucleases, DNA-modifying enzymes, and reagents for dideoxy sequencing using T7 DNA polymerase were purchased from New England Biolabs or Pharmacia Biotech Inc. [alpha-S]dATP, [alpha-P]dCTP, and [-P]ATP were purchased from NEN Research Products (Du Pont Canada Inc., Mississauga, Ontario). All other chemicals and reagents used in these experiments were purchased from British Drug House Inc. (St. Laurent, Québec) or Fisher Scientific and were of the highest grade available.

Isolation and Characterization of Rat Genomic Clones

Approximately 1 times 10^6 phage recombinants containing partially digested (MboI) rat genomic DNA (15-20 kb) inserted into the BamHI site of the phage vector EMBL3 (a gift of Dr. Donald Back, Queen's University, Kingston, Ontario) were screened with a 4.5-kb NHE3 cDNA probe (KpnI-SmaI fragment). The rat genomic phage clones, 6-2, 11-2, and 12-1, were isolated and together contained the entire coding region of the gene but lacked the 5` promoter region. To isolate this region, a small 90-bp polymerase chain reaction fragment from the 5` end of the cDNA was hybridized to a second rat genomic Lambda Dash II phage library (1 times 10^6 recombinants) obtained from Stratagene. The phage recombinants contained partially digested (Sau3AI) rat genomic DNA (9-22 kb) inserted into the BamHI site of the phage vector. This screening resulted in the isolation of phage clone 3B which contained 12 kb of the 5` upstream regulatory region.

Screening of both genomic libraries involved filter hybridization with rat NHE3 cDNA probes radiolabeled with [alpha-P]dCTP to a specific activity of 0.5-1.0 times 10^9 cpm/µg of DNA using the random primer oligonucleotide labeling kit (Pharmacia). The filters were prehybridized at 65 °C for 3-5 h in 6 times SSC, 5 times Denhardt's solution, 0.1% SDS, and 100 mg of denatured salmon sperm DNA/ml (see (36) for composition of SSC and Denhardt's solution). The filters were then hybridized in the same solution with denatured cDNA probe for 48 h at 65 °C. The replica filters were washed twice for 30 min each at 22 °C in 2 times SSC, 0.1% SDS, and then once for 60 min at 65 °C in 2 times SSC, 0.1% SDS. The filters were examined by autoradiography for positive hybridization signals, and the corresponding phage colonies were purified by 3 successive screenings using standard procedures.

Rat genomic DNA from the purified positive phage clones was mapped by cleavage with different restriction endonucleases and Southern blot analysis(37) . Briefly, the digested DNA was fractionated by electrophoresis on 1% agarose gels, denatured, and transferred to nylon membranes by capillary transfer. These membranes were successively hybridized to [-P]ATP end-labeled, synthetic oligonucleotides corresponding to various regions of the NHE3 cDNA.

DNA Sequence Analysis

Restriction endonuclease-generated DNA fragments from the phage clones that hybridized to synthetic NHE3 oligonucleotides were subcloned into pBluescript. These oligonucleotides, as well as others, were used as sequencing primers to define the exon-intron boundaries. Sequencing of the genomic DNA inserts was performed by the double-stranded dideoxy chain termination technique (38) using [alpha-S]dATP and T7 DNA polymerase sequencing kits from Pharmacia or U. S. Biochemical Corp. The 5`-flanking region was sequenced in both directions.

Primer Extension Analysis

Primer extension analysis was performed according to a previously described method(36) . A synthetic oligonucleotide complementary to the 5`-untranslated sequence in the rat NHE3 cDNA, nucleotides -13 to +4 relative to the translation initiation site, was 5`-end-labeled using [-P]ATP and T4 polynucleotide kinase. The labeled primer (10^5 cpm) was annealed to 50 µg of total cellular RNA by incubating overnight at 30 °C in hybridization solution (80% formamide, 40 mM PIPES, pH 6.4, 400 mM NaCl, 1 mM EDTA). After precipitation of the annealed primer and template, the hybridized primers were extended by incubating with 50 units of avian myeloblastosis virus reverse transcriptase in 20 µl of reverse transcriptase buffer (50 mM Tris-Cl, pH 7.6, 60 mM KCl, 10 mM MgCl(2), 1 mM dithiothreitol, 1 unit/µl placental RNase inhibitor, 50 µg/ml actinomycin D, and 1 mM each dNTP). The reaction was terminated by the addition of 1 µl of 0.5 M EDTA and 1 µl of DNase-free pancreatic RNase (5 µg/ml) for 30 min at 37 °C. The samples were precipitated by the addition of ethanol, denatured, and analyzed on a 6% polyacrylamide sequencing gel.

S1 Nuclease Mapping

S1 nuclease protection of the 5` end of the NHE3 transcript was performed as described previously(36) . Briefly, a P 5`-end-labeled oligonucleotide complementary to nucleotides -13 to +4 (relative to the translation initiation site) was annealed to a denatured plasmid (SH23/6-2) containing a 1.0-kb SalI-HindIII insert isolated from 3B which contained the putative transcription initiation site. The primer was extended using the Klenow fragment of DNA polymerase I and the resulting double-stranded product cleaved with SalI. After fractionating the DNA on a denaturing alkaline agarose gel, a radiolabeled, single-stranded fragment of 416 nucleotides in length was eluted from the gel. Approximately 5 times 10^4 cpm of probe was mixed with 50 µg of total cellular RNA and incubated in 40 µl of S1 hybridization buffer (80% formamide, 40 mM PIPES, pH 6.4, 400 mM NaCl, 1 mM EDTA) at 65 °C for 10 min, then at 30 °C for 36 h. The samples were digested with 300 units of S1 nuclease for 1 h at 30 °C in 300 µl of S1 buffer (280 mM NaCl, 50 mM sodium acetate, pH 4.5, and 4.5 mM ZnSO(4)) containing 6 µg of denatured salmon sperm DNA. The protected fragments were precipitated by the addition of ethanol, denatured, and analyzed on a 6% polyacrylamide sequencing gel.

Cell Culture

Renal OK and LLC-PK(1) cell lines were propagated in alpha-minimum essential medium supplemented with 10% fetal bovine serum, 100 µg/ml kanamycin sulfate, and 25 mM NaHCO(3), pH 7.4, and incubated in a humidified atmosphere of 95% air, 5% CO(2) at 37 °C.

Plasmid Construction and Transient Gene Expression Assays

A chimeric Nhe3-luciferase gene was constructed by isolating a 1.44-kb BamHI-KpnI genomic DNA fragment from the phage clone 3B that contains the Nhe3 5`-flanking region (-1380 to +59 relative to the transcription initiation site) and inserting it into the polylinker region of the promoterless luciferase vector pXP1(39) . This new plasmid was named pNhe3-1380.

Following plating overnight, subconfluent monolayers of OK and LLC-PK(1) cells were washed twice with phosphate-buffered saline and then changed to standard alpha-minimum essential medium supplemented with 10% ``charcoal-stripped'' fetal bovine serum at least 2 h before transfection. Cells were transfected using the calcium phosphate-DNA coprecipitation technique of Chen and Okayama(40) . The functional activity of the chimeric gene was assessed by cotransfection of 20 µg of plasmid DNA into cells. The composition of the DNA included either the promoterless pXP1 vector (negative control) or the pRSV-Luc vector containing the constitutively active Rous sarcoma virus (RSV) promoter linked to luciferase (positive control) or the pNhe3-1380 vector, plus pRSV110 (a vector containing the RSV promoter linked to the beta-galactosidase gene) and pHG1 (expression plasmid containing the SV40 promoter and human glucocorticoid receptor; generously provided by Dr. John White, McGill University) in a weight ratio of 3:1:1, respectively. The activity of the beta-galactosidase gene served as an internal control to monitor for transfection efficiency. The plasmid containing the human glucocorticoid receptor was transfected to ensure that a sufficient concentration of receptor was available for binding to the overexpressed Nhe3-1380 gene. After a 20-h DNA precipitation period, the cells were washed and then cultured in fresh medium in the absence or presence of 100 nM dexamethasone. The cells were harvested 72 h later by washing twice with phosphate-buffered saline, followed by the addition of Triton glycylglycine lysis buffer and then scraping the cells off the dish with a rubber policeman. Cell extracts were diluted in 25 mM glycylglycine (pH 7.8) and assayed for luciferase and beta-galactosidase activities using protocols and reagents provided by Promega. Luciferase activity was measured using a BioOrbit 1250 luminometer.


RESULTS

Structure of the Rat Na/H Exchanger Nhe3 Gene

The genomic organization of the rat Na/H exchanger Nhe3 gene was determined by sequencing genomic DNA inserts from phage clones 3B, 6-2, 11-2, and 12-1 using synthetic oligonucleotide probes that spanned the entire published NHE3 cDNA sequence(2) . As illustrated in Fig. 1, the Nhe3 gene spans >40 kilobases and contains at least 17 exons. The exons are evenly distributed along the genomic DNA with the exception of exon 1 which is separated from the other exons by a large intron that is estimated to be >25 kilobases. The protein-coding exons generally range in size from 71 to 302 nucleotides, with the exception of exon 17 which is >1700 nucleotides and contains the TGA stop codon (starting at position 2 of the exon) and a large segment of the 3`-untranslated region. However, the most 3` end of this exon containing the remainder of the 3`-untranslated sequence, including the polyadenylation signal, were not identified in this genomic DNA phage clone and leave open the possibility of additional exons in this region.


Figure 1: Genomic organization of the rat Na/H exchanger Nhe3 gene. A, overlapping genomic inserts from recombinant phage clones that encode the rat Nhe3 gene are aligned relative to the exon-intron organization of the gene (shown in B). The overlap regions between 3B and 6-2 and between 11-2 and 12-1 were determined by DNA sequencing. An overlap between 6-2 and 11-2 was not identified, presumably due to the presence of a large intron. B, the relative locations of the Nhe3 exons within the gene, depicted by gray bars, and their sizes (in base pairs) are illustrated. C, the positions of the boundaries between exons in the NHE3 protein are indicated by the arrows. The hatched boxes represent the proposed membrane-spanning segments.



The sizes of the introns were determined either by DNA sequencing or were estimated by agarose gel electrophoresis of PCR-generated DNA fragments using oligonucleotide primers complementary to the flanking exons. They generally varied in length from 0.1 to 1.2 kilobases, excluding the first intron which is estimated to be >25 kilobases. This estimate is based on the sizes of the 6-2 and 11-2 phage clones which did not overlap. As shown in Table 1, the exon-intron boundaries conform to typical splice donor AG/GT(A/G)AGT and acceptor (T/C)(n)N(C/T)AG/G (n > 11) consensus sequences(41) . Each splice donor site begins with an invariant GT dinucleotide, whereas each splice acceptor site ends with an invariant AG dinucleotide and is preceded by a polypyrimidine tract.



It has generally been observed that members of gene families share similar genomic organizations, such as the Na,K-ATPase isoforms(42, 43) . This also applies to the NHE gene family members that have been characterized to date. A comparison of the exons within the protein coding sequences of rat NHE3 and human NHE1 (33) is illustrated in Fig. 2. With the exception of exons 2 and 3 in human NHE1, which are both split by an apparent intron in rat Nhe3, the exon-intron boundaries occur in exactly the same positions within the proposed N-terminal transmembranous region and the first 50 amino acids of the C-terminal cytoplasmic region. These regions share the highest degree of amino acid identity among the isoforms. However, little similarity in exon organization exists in the remainder of the C-terminal cytoplasmic regions of the isoforms which also share minimal sequence identity and are believed to encode the diverse regulatory elements of the transporters.


Figure 2: Comparison of the exon arrangement of the Na/H exchanger NHE1 and NHE3 isoforms from human and rat, respectively. The deduced amino acid sequences of human NHE1 and rat NHE3 were aligned using the CLUSTAL W alignment program(83) . The locations of the exon boundaries for rat NHE3 (see Table 1) and human NHE1 (33) are illustrated by arrows.



Location of the Transcription Initiation Site and Nucleotide Sequence of the 5`-Flanking Region

To identify the Nhe3 transcription initiation sites, both primer extension and S1 nuclease protection analyses were performed. For primer extension analysis, an oligonucleotide primer complementary to the 5`-untranslated sequence (nucleotides -13 to +4 relative to the translation initiation site) was used to analyze total RNA isolated from adult rat kidney. A single intense band was observed that extended to within 19 nucleotides from the 5`-end of the known cDNA sequence (data not shown). Other experimental conditions and primers were unsuccessful at increasing the extension product beyond this point. This premature termination likely indicates the presence of a strong RNA secondary structure at this position which impeded the progression of the avian myeloblastosis virus reverse transcriptase.

To resolve this difficulty, S1 nuclease protection studies were performed using total RNA isolated from adult rat intestine and kidney (Fig. 3). The S1 probe consisted of a 416-nucleotide single-stranded DNA fragment that extended from position -412 to +4 relative to the translation start site. A cluster of apparent transcription initiation sites located within a 5-bp region were observed in total RNA isolated from adult rat intestine and kidney, with a major site at -97 (T nucleotide) and two minor sites at -100 (A nucleotide) and -96 (G nucleotide). Since T is the most predominant start site in rat Nhe3, it is designated as +1 and all 5` elements have been numbered relative to this site.


Figure 3: Determination of the transcription initiation site of the rat Na/H exchanger Nhe3 gene by S1 nuclease protection analysis. A single-stranded P-labeled DNA fragment beginning 4 nucleotides 3` to the translation start site and extending 416 nucleotides in the 5` direction was hybridized to 50 µg of total RNA from rat kidney or intestine. S1 nuclease protection analysis was performed as described under ``Experimental Procedures.'' The protected fragments were analyzed on a 6% denaturing polyacrylamide sequencing gel. Lane 1, 50 µg of total RNA from rat kidney; lanes 2-5, sequencing reaction using the same primer that was used in generating the S1 probe; lane 6, 50 µg of total RNA from rat intestine. The nucleotide sequences flanking the transcription initiation sites (indicated by asterisks) is illustrated on the right of the figure.



The nucleotide sequence of the 5` end of the Nhe3 gene is shown in Fig. 4. The center of an atypical TATA sequence, ATTAAA, is located 30 bp upstream of the major initiation site and potentially binds the multiprotein complex, TFIID (reviewed in (44) ). As well, although a classical CCAAT consensus sequence capable of recognizing NF-Y/CP1 (45, 46) is not observed between -200 and -50 bp of the transcription start site, an atypical sequence, CCAAG, that resembles the core binding motif is located 147 bp 5` of the major transcription initiation site. However, a classical TATA-box and CCAAT-box are positioned further upstream at -428 and -501, respectively, but are considered too far from the promoter region to be functional. This premise is based on our inability to obtain a primer extension product following hybridization of specific oligonucleotides in this region to total RNA from adult rat intestine and kidney (data not shown).


Figure 4: Nucleotide sequence of the 5`-flanking region of the rat Na/H exchanger Nhe3 gene. The sequence of the 5`-flanking region (1.38 kilobases) of the rat Nhe3 gene is shown (GenBank accession number U49386). A SalI site used in generating the 416-nucleotide S1 nuclease protection probe is indicated. Transcription initiation sites are indicated by large (major site) and small (minor sites) arrows. The DNA sequence was scanned for elements that share homology to known transcription factor binding sites using the computer program SIGNAL SCAN(84) . An apparent TATA-box and sequences exhibiting similarity to transcription factor binding sites (AP-1, AP-2, C/EBP, NF-I, NF-Y, OCT-1/OTF-1, PEA3, Sp1) or hormone receptor response elements (glucocorticoid, GR; and triiodothyronine, T3R) are underlined. Numbers to the left of the figure refer to nucleotide positions relative to the major transcription initiation site.



The promoter region also contains several potential GC-box motifs or Sp1-binding sites that resemble the core hexanucleotide motif, GGGCGG (47) , and are located at positions -588, -350, -139, -72, and -54 (reverse orientation) upstream of the apparent cap signal. Indeed, the first 200 bp of the promoter region is highly (G + C)-rich (67%), a characteristic that is commonly observed in the immediate 5`-flanking region of housekeeping genes.

The 5`-flanking region also contains a number of other potential cis-acting elements recognized by known transcription factors that may play a role in the basal or chronic regulation of the Nhe3 gene (see ``Discussion''), including four half-sites for the glucocorticoid responsive element (GRE) (TGTTCT; positions -1273, -1235, -479, and -459)(48) , six half-sites for the thyroid hormone responsive element (AGGTCA; positions -1020, -1014, -1007, -234, -225, -215)(48) , one PEA3 site (AGGAAGT; position -1309) (49) , four AP-1 sites (TGA(C/G)T(C/A)A; positions -1178, -1130, -1065, and -907)(50, 51) , two C/EBP sites (ATTGCGCAAT; positions -979 and -948)(52) , one OCT-1/OTF-1 site (ATGCAAAT; position -628) (53) , four AP-2 sites (CCCC(A/G)(G/C)(G/C)C; positions -412, -295, -48, and -4)(54) , and one NF-I site (TTGGC(N)(5)GCCAA; position -271)(46, 55) .

Glucocorticoid Transcription Control of the Rat Nhe3 Gene Promoter

Glucocorticoids have been reported to elevate NHE3 activity and mRNA abundance in epithelial cells from the ileum (29) , renal proximal tubules(27, 28) , and cultured renal OK cells (56) which retain certain phenotypic characteristics of proximal tubule cells. In view of the observed putative GRE half-sites in the Nhe3 promoter, it was of interest to assess whether glucocorticoid induction of NHE3 expression could be accounted for by alterations in gene transcription. To this end, a BamHI-KpnI DNA fragment extending from -1380 to +59 relative to the major transcription initiation site was inserted into the promoterless luciferase expression plasmid pXP1 (called pNhe3-1380) (Fig. 5A) and transfected into two renal proximal tubule cell lines, OK and LLC-PK(1). Of these cell lines, OK cells have been shown to express only native NHE3 (23, 57) whereas LLC-PK(1) cells express both NHE1 (58) and NHE3(59) . Two additional plasmids were also cotransfected with pNhe3-1380: (i) pRSV110 (a vector containing the RSV promoter linked to the beta-galactosidase gene) and (ii) pHG1 (pSG5 expression plasmid containing the human glucocorticoid receptor). The activity of the beta-galactosidase gene served as an internal control to monitor for transfection efficiency and the human glucocorticoid receptor was transfected to provide for maximal ligand-dependent transactivation of the overexpressed Nhe3-1380 gene.


Figure 5: Glucocorticoid induction of a chimeric rat Nhe3 promoter-firefly luciferase reporter gene in transiently transfected renal OK and LLC-PK(1) cell lines. A, schematic illustration of the promoterless firefly luciferase gene in the parent vector pXP1 and a chimeric Nhe3 gene (pNhe3-1380) containing 1380 bp (relative to the transcription start site) of the 5`-flanking region of Nhe3 (BamHI-KpnI fragment) fused to the luciferase reporter gene. Each recombinant plasmid was transiently cotransfected into renal OK (B) or LLC-PK(1) (C) in the presence of pRSV110 which expresses the beta-galactosidase gene (to normalize for transfection efficiency) and pHG1 which expresses the human glucocorticoid receptor (to provide a sufficient concentration of receptor for binding to the overexpressed Nhe3-1380 gene) in a weight ratio of 3:1:1, respectively. The cells were incubated in the absence (gray shaded box) or presence (black box) of 100 nM dexamethasone for 72 h. Cell extracts were then prepared for measurements of luciferase and beta-galactosidase activities. For comparative purposes, the corrected values obtained for luciferase activity in unstimulated cells transfected with pNhe3-1380 were normalized to a value of 1 and all other activities were adjusted accordingly. The data represent fold stimulation (mean ± S.D.) from two experiments, each performed in quadruplicate.



The promoterless expression plasmid, pXP1, showed negligible luciferase activity when transfected into both cell lines, whereas pNhe3-1380 exhibited a 69- and 7-fold increase in basal luciferase activity in OK and LLC-PK(1) cells, respectively (Fig. 5, B and C). This indicated that the inserted 5`-flanking region of the Nhe3 gene contains a functional promoter. However, the basal activity of pNhe3-1380 in OK and LLC-PK(1) cells is considerably less (691- and 16-fold, respectively) than that obtained with the expression plasmid pRSV-Luc that contains the more powerful Rous sarcoma virus promoter linked to the luciferase gene (data not shown). The weak promoter activity of the Nhe3 gene could account, at least in part, for the generally low abundance of NHE3 mRNA in rat tissues(2) .

The glucocorticoid responsiveness of the pNhe3-1380 was tested by incubating the transfected cells in the absence or presence of 100 nM dexamethasone for 72 h. This treatment substantially elevated luciferase activity in OK and LLC-PK(1) cells by 15- and 6.5-fold, respectively (Fig. 5, B and C). In contrast, transfected cells containing the promoterless pXP1 plasmid did not show a significant change in luciferase activity.


DISCUSSION

We have isolated and characterized the structural organization of the rat Na/H exchanger Nhe3 gene and its 5`-flanking region from a phage genomic library. This gene, which has previously been mapped to rat chromosome 1 (32) and human chromosome 5p15.1(31) , spans >40 kb in length and contains at least 17 exons. The organization of the first 9 exons encoding the entire proposed N-terminal transmembranous region (exons 1 to 7; 453 amino acids) and the first 50 amino acids of the adjacent C-terminal cytoplasmic region (exons 8 and 9) is conserved for the most part with that of the human NHE1 gene, the only other NHE gene to be characterized to date. The only exceptions are exons 2 and 3 of rat Nhe3 which are continuous in human NHE1 (forming exon 2) and exons 4 and 5 of rat Nhe3 which are also uninterrupted in human NHE1 (forming exon 3). The N-terminal transmembranous segments share high amino acid identity (55-95%) among the NHE isoforms and are believed to comprise the structural domains necessary for ion translocation. In contrast, the organization of exons 10 to 17 of the Nhe3 gene is dissimilar to the exon arrangement of the comparable region (exons 8 to 12) of human NHE1. This C-terminal region in both isoforms, which is proposed to reside on the cytoplasmic side of the plasma membrane, shares minimal amino acid identity (25-35%) and contains numerous putative regulatory motifs that respond in a distinctive manner to a variety of stimuli(14, 16, 17, 18, 60) .

To begin addressing questions concerning the regulation of Nhe3 gene expression at the transcriptional level, the transcription initiation sites were determined. In addition, the 5`-flanking sequence was examined for potential promoter elements that are fixed close to the start of transcription (i.e. TATA-box and cap signal) and enhancer motifs (e.g. CCAAT-box, Sp1, AP-1, AP-2, C/EBP, OCT-1/OTF-1, and NF-I binding sites) and hormone receptor responsive elements (e.g. triiodothyronine and glucocorticoid receptor binding sites) that are less restricted in both position and orientation. Only those potential sites that may be of physiological relevance to Nhe3 are discussed below.

Multiple transcription initiation sites were identified for the Nhe3 gene using S1 nuclease protection analyses from rat kidney and intestine, the two major tissues where the gene is expressed. Although three initiation sites were found to cluster between nucleotides -100 and -96 relative to the translation ATG start codon, the predominant site used in both tissues occurred at position -97 (thymidine) with minor sites located at -100 (adenine) and -96 (guanine). The cap signal CCT(0)G closely matches the consensus cap signal NCA(0)(G/C/T)(61) . Although the site of initiation is frequently an A preceded by an invariant C nucleotide, the use of T (but never C or G) is observed in a minority of cases. However, this motif is only present in approximately 60% of all promoters, suggesting that this sequence is not essential for promoter function(61) . Multiple start sites scattered around the promoter region have also been reported for several genes (62, 63) that lack a canonical TATA sequence. The presence of an atypical TATA-box (discussed below) in rat Nhe3 may also account for the multiple start sites. Consequently, multiple transcription initiation sites may be a characteristic of genes that lack a TATA-box, or a well defined TATA motif, in the promoter region.

The promoter/enhancer region of the rat Nhe3 gene contains an atypical TATA-box and CCAAT-box that are centered 30 and 147 nucleotides, respectively, upstream of the major transcription initiation start site. The core and flanking residues of the atypical TATA-box, GATT(0)AAAGG, differ from the extended canonical sequence (G/C)TAT(0)AAA(A/T)(G/A) by having an A in the -2-position, a T in the -1-position, and a G in the +4 position. Nonetheless, a recent analysis of the promoter region of 400 genes has revealed that A, T, and G can be present in these positions, respectively, but at low frequencies (4%, 9%, and 11%, respectively)(61) . The atypical CCAAT-box, AACCA(0)AGTAG, is centered at position -147 and exhibits similarity (8/10 match) to the extended CCAAT consensus sequence (A/G)(G/A)CCA(0)AT(C/G)(A/G)G(45, 46, 61) . This pentanucleotide core sequence is recognized by the transcription factor NF-Y/CP1 and is typically located between 50 and 200 nucleotides upstream of the transcription start site. The atypical CCAAT-box in Nhe3 differs from the canonical sequence by substitution of G and T in the +2 and +3 positions, respectively. The presence of G and T in these positions is rare, being observed in only 9% and 2%, respectively, of genes examined(61) . However, since 6 out of 7 nucleotides that were identified to contact the NF-Y protein (based on methylation interference analysis) (45) are conserved in the proposed CCAAT-box, it is reasonable to suspect that the observed motif may be transcriptionally relevant. However, functional studies are required to verify this possibility.

In addition, other CCAAT-like elements are known to exist, such as the palindromic consensus sequence TTGGC(N)(5)GCCAA that is recognized by nuclear factor I (NF-I/CTF)(46, 55) . One potential NF-I binding site in the Nhe3 promoter region is located between nucleotides -271 and -257 and maintains all 6 essential contact points in the NF-I binding site that, based on methylation interference analysis, seem to contact the NF-I protein(46) . A third class of nuclear transcription factors that is able to bind to CCAAT-box and related enhancer motifs belongs to the CCAAT-box/enhancer binding protein (C/EBP) gene family(52, 64) . C/EBP DNA-binding proteins recognize the optimal dyad-symmetrical sequence ATTGCGCAAT and regulate the expression of genes during cellular differentiation (52, 64) and in response to cAMP(65, 66) . Two potential C/EBP elements that match at least 7 out of 10 nucleotides of the optimal motif are located in the Nhe3 gene at positions -979 and -948.

The promoter region also contains five potential binding sites for Sp1 that are positioned at nucleotides -588 (TTGGCGGGAG), -350 (AAGGCGGGAG), -139 (GGGGCGTGAG), -72 (GGGGCGGGAA), and -54 (GCCCCGCCCT; reverse orientation) upstream of the apparent cap signal. These sites closely match the consensus decanucleotide sequence for Sp1, (G/T)GGGCGG(G/A)(G/A)(C/T), which contains an optimized core hexanucleotide sequence GGGCGG and 5`- and 3`-flanking nucleotides that, although degenerate, can significantly influence Sp1 binding(47, 61) . Sp1 sites, which can function in either orientation, are often found within 200 nucleotides upstream of the start of transcription and can act in conjunction with NF-I to increase the rate of transcription (47) . It is also noted that the sequence of the promoter region is (G + C)-rich (67% from nucleotides -1 to -201) which is frequently found in the 5` promoter region of housekeeping genes. Thus, it appears that the Nhe3 promoter region is a mosaic of potential elements that are characteristic of both cell-specific and housekeeping genes.

Since the mRNA and/or activity of Na/H exchangers in renal or intestinal epithelial cells is elevated following prolonged exposure to various stimuli, including phorbol esters(67) , cAMP(68) , acidic medium(22, 23, 25, 26, 69, 70) , triiodothyronine(71) , and glucocorticoids(27, 28, 29, 56) , we examined the 5`-flanking region of the Nhe3 gene for potential cis-acting DNA response elements that may be involved in some of these responses.

The trans-acting factor AP-1 is a heterodimer composed of the c-jun and c-fos proto-oncogenes that influences basal transcription and is also required for induction of transcription by phorbol esters(50, 51) . The 5`-flanking region of the Nhe3 gene contains four potential AP-1 sequences; three match 6 out of 7 nucleotides of the AP-1 consensus sequence, TGA(G/C)T(C/A)A and the other exhibits a 5/7 match. In addition to the AP-1 site, other phorbol ester responsive elements have been identified (reviewed in (72) ). Insertion of an oligonucleotide containing the PEA3 consensus sequence, AGGAAGT, upstream of a heterologous promoter can confer responsiveness to phorbol esters(49) . Furthermore, PEA3 can act synergistically with AP-1 to achieve maximal induction of transcription by phorbol esters(73) . A PEA3 site (in the reverse orientation) that precisely matches the consensus sequence and is in close proximity to the AP-1 sites is located at nucleotide -1309 of the Nhe3 gene. Last, the regulatory region of this gene contains four potential sites at positions -412, -295, -48, and -4 that share homology with AP-2 elements, CCCC(A/G)(G/C)(G/C)C, that act as basal transcription enhancers but are also responsive to both phorbol esters and cAMP(54, 74) . The sequence of individual sites, however, can vary considerably from the canonical binding motif. The responsiveness of AP-2 sites, as well as C/EBP sites(66) , to cAMP is of particular relevance since cAMP chronically up-regulates NHE3 activity in renal OK cells(68) . Moreover, a consensus site for the classical cAMP-responsive binding protein CREB, TGACGT(A/C)A(74) , was not identified in the 1.4-kb 5`-flanking region, although potential CREB-binding sites may be present further upstream.

Triiodothyronine, estrogen, retinoids, and vitamin D bind to specific intracellular receptors that are part of a large family of ligand-dependent transcription factors. Interestingly, the DNA recognition site for each of these receptors, which may bind as monomers, homodimers, or heterodimers, is a variant of the optimal hexanucleotide half-site AGGTCA (reviewed in (48) ). However, the specificity of hormone receptor binding to the cognate response element and the magnitude of transcriptional activity is highly dependent upon the number, orientation, and spacing between half-sites(75, 76, 77) . For example, direct repeats separated by 3, 4, and 5 bp are selectively responsive to vitamin D, triiodothyronine, and retinoic acid, respectively(76) . Two clusters of AGGTCA-like elements (each containing 3 core binding motifs) are present in the rat Nhe3 gene at positions -1025 to -1002 and -239 to -210. Each cluster contains two direct repeats in one orientation and a single half-site in the reverse orientation. The selective hormone responsiveness, if any, awaits further characterization.

The glucocorticoid hormone receptor complex binds to an inverted palindromic sequence AGAACA(N)(3)TGTTCT that is distinct from the triiodothyronine/estrogen/retinoid/vitamin D core binding motif (48) . Within this consensus glucocorticoid responsive element (GRE), the hexanucleotide core TGTTCT is generally highly conserved, whereas the leftmost hexanucleotide sequence (AGAACA) can be quite variable. The TGTTCT half-site is also recognized by mineralocorticoid, androgen, and progesterone receptors(48) . Similar to the core binding motif for the thyroid hormone class of receptors, hormone selectivity for the GRE is influenced by the number, spacing, orientation, and DNA sequence of the cognate responsive element. There are 4 half-site sequences segregated into two clusters (each containing two half-sites) within the 1.4-kb 5`-flanking region of Nhe3. The distal cluster is located between nucleotides -1273 and -1230, and the more proximal cluster is found between nucleotides -479 and -454. Each cluster contains a half-site that exhibits either a 5/6 match (positions -1273 and -459) or 6/6 match (positions -1235 and -479) to the conserved hexanucleotide core element. However, homology to the entire 15-bp palindromic sequence is low. Although GREs can function as independent elements, it is now becoming apparent that the maximal activity of many GREs is greatly influenced by the presence of additional cis-elements/trans-acting factors. For example, both NF-I (78) and the ubiquitous transcription factor OCT-1/OTF-1 (53) participate in the glucocorticoid regulation of the mouse mammary tumor virus promoter. In addition, the AP-1 complex has been found to influence the ability of the glucocorticoid receptor to stimulate or inhibit gene transcription(79) . In many respects, these factors can be viewed as components of a larger composite hormone responsive unit that creates the potential for greater flexibility in the hormonal regulation of gene transcription. Interestingly, potential NF-I, OCT-1/OTF-1, and AP-1 binding sites are also located in the general vicinity of the GRE clusters in the rat Nhe3 gene, although their functional importance, if any, remains to be determined.

As mentioned previously, glucocorticoids elevate Na/H exchanger NHE3 activity and/or mRNA abundance in ileum(29) , renal proximal tubules(27, 28) , and OK cells (56) . The studies described in this report extend these observations by demonstrating that the transcriptional activity of the 5`-flanking region of the rat Nhe3 gene fused to the luciferase reporter gene is activated by glucocorticoids in transiently transfected OK and LLC-PK(1) cells. This suggests that the observed in vivo induction of native NHE3 mRNA by glucocorticoids is most probably mediated primarily at the transcriptional level. More detailed studies will be required to precisely delineate the specific cis-acting elements involved.

In summary, we have isolated and characterized the 5`-flanking promoter region and exon-intron organization of the rat Nhe3 gene. In addition, we have demonstrated that this gene can be transcriptionally activated by glucocorticoids. This information provides the framework for further investigations on the mechanisms involved in the chronic regulation of this gene by glucocorticoids as well as other stimuli and its possible involvement in pathophysiological conditions such as systemic acidosis(69) , hypertension(80, 81) , and congenital secretory diarrhea(82) .


FOOTNOTES

*
This work was supported in part by Grant MT-11221 from the Medical Research Council (MRC) of Canada and grants from the Kidney Foundation of Canada (to J. O.). The costs of publication of this article were defrayed in part by the payment of page charges. This article must therefore by hereby marked ``advertisement'' in accordance with 18 U.S.C. Section 1734 solely to indicate this fact.

The nucleotide sequence(s) reported in this paper has been submitted to the GenBank(TM)/EMBL Data Bank with accession number(s) U49386[GenBank].

§
Recipient of a scholarship from the Fond de la Recherche en Santé du Québec (FRSQ). To whom correspondence should be addressed: Dept. of Physiology, McGill University, McIntyre Medical Science Bldg., 3655 Drummond St., Montreal, Quebec H3G 1Y6, Canada. Tel.: 514-398-8335; Fax: 514-398-7452; orlowski{at}physio.mcgill.ca.

(^1)
The abbreviations used are: NHE, Na/H exchanger; kb, kilobase(s); bp, base pair(s); PIPES, 1,4-piperazinediethanesulfonic acid; RSV, Rous sarcoma virus; GRE, glucocorticoid response element.


ACKNOWLEDGEMENTS

We thank Dr. John White (McGill University) for critical reading of the manuscript.


REFERENCES

  1. Sardet, C., Franchi, A., and Pouysségur, J. (1989) Cell 56, 271-280 [Medline] [Order article via Infotrieve]
  2. Orlowski, J., Kandasamy, R. A., and Shull, G. E. (1992) J. Biol. Chem. 267, 9331-9339 [Abstract/Free Full Text]
  3. Wang, Z., Orlowski, J., and Shull, G. E. (1993) J. Biol. Chem. 268, 11925-11928 [Abstract/Free Full Text]
  4. Klanke, C. A., Su, Y. R., Callen, D. F., Wang, Z., Meneton, P., Baird, N., Kandasamy, R. A., Orlowski, J., Otterud, B. E., Leppert, M., Shull, G. E., and Menon, A. G. (1995) Genomics 25, 615-622 [CrossRef][Medline] [Order article via Infotrieve]
  5. Tse, C.-M., Ma, A. I., Yang, V. W., Watson, A. J. M., Levine, S., Montrose, M. H., Potter, J., Sardet, C., Pouysségur, J., and Donowitz, M. (1991) EMBO J. 10, 1957-1967 [Abstract]
  6. Tse, C.-M., Levine, S. A., Yun, C. H. C., Montrose, M. H., Little, P. J., Pouysségur, J., and Donowitz, M. (1993) J. Biol. Chem. 268, 11917-11924 [Abstract/Free Full Text]
  7. Tse, C.-M., Brant, S. R., Walker, M. S., Pouysségur, J., and Donowitz, M. (1992) J. Biol. Chem. 267, 9340-9346 [Abstract/Free Full Text]
  8. Fafournoux, P., Noël, J., and Pouysségur, J. (1994) J. Biol. Chem. 269, 2589-2596 [Abstract/Free Full Text]
  9. Noël, J., and Pouysségur, J. (1995) Am. J. Physiol. 268, C283-C296
  10. Yun, C. H. C., Tse, C.-M., Nath, S. K., Levine, S. A., Brant, S. R., and Donowitz, M. (1995) Am. J. Physiol. 269, G1-G11
  11. Grinstein, S., and Rothstein, A. (1986) J. Membr. Biol. 90, 1-12 [Medline] [Order article via Infotrieve]
  12. Grinstein, S., Rotin, D., and Mason, M. J. (1989) Biochim. Biophys. Acta 988, 73-97 [Medline] [Order article via Infotrieve]
  13. Grinstein, S., Woodside, M., Sardet, C., Pouysségur, J., and Rotin, D. (1992) J. Biol. Chem. 267, 23823-23828 [Abstract/Free Full Text]
  14. Wakabayashi, S., Bertrand, B., Shigekawa, M., Fafournoux, P., and Pouysségur, J. (1994) J. Biol. Chem. 269, 5583-5588 [Abstract/Free Full Text]
  15. Wakabayashi, S., Bertrand, B., Ikeda, T., Pouysségur, J., and Shigekawa, M. (1994) J. Biol. Chem. 269, 13710-13715 [Abstract/Free Full Text]
  16. Levine, S. A., Montrose, M. H., Tse, C. M., and Donowitz, M. (1993) J. Biol. Chem. 268, 25527-25535 [Abstract/Free Full Text]
  17. Kapus, A., Grinstein, S., Wasan, S., Kandasamy, R. A., and Orlowski, J. (1994) J. Biol. Chem. 269, 23544-23552 [Abstract/Free Full Text]
  18. Kandasamy, R. A., Yu, F. H., Harris, R., Boucher, A., Hanrahan, J. W., and Orlowski, J. (1995) J. Biol. Chem. 270, 29209-29216 [Abstract/Free Full Text]
  19. Moe, O. W., Amemiya, M., and Yamaji, Y. (1995) J. Clin. Invest. 96, 2187-2194 [Medline] [Order article via Infotrieve]
  20. Weinman, E. J., Steplock, D., Wang, Y., and Shenolikar, S. (1995) J. Clin. Invest. 95, 2143-2149 [Medline] [Order article via Infotrieve]
  21. Moe, O. W., Miller, R. T., Horie, S., Cano, A., Preisig, P. A., and Alpern, R. J. (1991) J. Clin. Invest. 88, 1703-1708 [Medline] [Order article via Infotrieve]
  22. Igarashi, P., Freed, M. I., Ganz, M. B., and Reilly, R. F. (1992) Am. J. Physiol. 263, F83-F88
  23. Amemiya, M., Yamaji, Y., Cano, A., Moe, O. W., and Alpern, R. J. (1995) Am. J. Physiol. 269, C126-C133
  24. Soleimani, M., Singh, G., Bizal, G. L., Gullans, S. R., and McAteer, J. A. (1994) J. Biol. Chem. 269, 27973-27978 [Abstract/Free Full Text]
  25. Horie, S., Moe, O., Yamaji, Y., Cano, A., Miller, R. T., and Alpern, R. J. (1992) Proc. Natl. Acad. Sci. U. S. A. 89, 5236-5240 [Abstract]
  26. Yamaji, Y., Amemiya, M., Cano, A., Preisig, P. A., Miller, R. T., Moe, O. W., and Apern, R. J. (1995) Proc. Natl. Acad. Sci. U. S. A. 92, 6274-6278 [Abstract]
  27. Baum, M., Moe, O. W., Gentry, D. L., and Alpern, R. J. (1994) Am. J. Physiol. 267, F437-F442
  28. Guillery, E. N., Karniski, L. P., Mathews, M. S., Page, W. V., Orlowski, J., Jose, P. A., and Robillard, J. E. (1995) Am. J. Physiol. 268, F710-F717
  29. Yun, C. H. C., Gurubhagavatula, S., Levine, S. A., Montgomery, J. L. M., Brant, S. R., Cohen, M. E., Cragoe, E. J., Jr., Pouysségur, J., Tse, C.-M., and Donowitz, M. (1993) J. Biol. Chem. 268, 206-211 [Abstract/Free Full Text]
  30. Lifton, R. P., Sardet, C., Pouysségur, J., and Lalouel, J. M. (1990) Genomics 7, 131-135 [Medline] [Order article via Infotrieve]
  31. Brant, S. R., Bernstein, M., Wasmuth, J. J., Taylor, E. W., McPherson, J. D., Li, X., Walker, S., Pouysségur, J., Donowitz, M., Tse, C.-M., and Wang Jabs, E. (1993) Genomics 15, 668-672 [CrossRef][Medline] [Order article via Infotrieve]
  32. Szpirer, C., Szpirer, J., Rivière, M., Levan, G., and Orlowski, J. (1994) Mamm. Genome 5, 153-159 [Medline] [Order article via Infotrieve]
  33. Miller, R. T., Counillon, L., Pages, G., Lifton, R. P., Sardet, C., and Pouysségur, J. (1991) J. Biol. Chem. 266, 10813-10819 [Abstract/Free Full Text]
  34. Kolyada, A. Y., Lebedeva, T. V., Johns, C. A., and Madias, N. E. (1994) Biochim. Biophys. Acta 1217, 54-64 [Medline] [Order article via Infotrieve]
  35. Dyck, J. R. B., Silva, N. L. C. L., and Fliegel, L. (1995) J. Biol. Chem. 270, 1375-1381 [Abstract/Free Full Text]
  36. Sambrook, J., Fritsch, E. F., and Maniatis, T. (1989) Molecular Cloning: A Laboratory Manual , Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY
  37. Southern, E. M. (1975) J. Mol. Biol. 98, 503-517 [Medline] [Order article via Infotrieve]
  38. Sanger, F., Nicklen, S., and Coulson, A. R. (1977) Proc. Natl. Acad. Sci. U. S. A. 74, 5463-5467 [Abstract]
  39. Nordeen, S. K. (1988) BioTechniques 6, 454-456 [Medline] [Order article via Infotrieve]
  40. Chen, C., and Okayama, H. (1987) Mol. Cell. Biol. 7, 2745-2752 [Medline] [Order article via Infotrieve]
  41. Mount, S. M. (1982) Nucleic Acids Res. 10, 459-472 [Abstract]
  42. Shull, M. M., Pugh, D., and Lingrel, J. B. (1989) J. Biol. Chem. 264, 17532-17543 [Abstract/Free Full Text]
  43. Ovchinnikov, Y. A., Monastyrskaya, G. S., Broude, N. E., Ushkaryov, Y. A., Melkov, A. M., Smirnov, Y. V., Malyshev, I. V., Allikmets, R. L., Kostina, M. B., Dulubova, I. E., Kiyatkin, N. I., Grishin, A. V., Modyanov, N. N., and Sverdlov, E. D. (1988) FEBS Lett. 233, 87-94 [CrossRef][Medline] [Order article via Infotrieve]
  44. Zawel, L., and Reinberg, D. (1995) Annu. Rev. Biochem. 64, 533-561 [CrossRef][Medline] [Order article via Infotrieve]
  45. Dorn, A., Bollenkens, J., Staub, A., Benoist, C., and Mathis, D. (1987) Cell 50, 863-872 [Medline] [Order article via Infotrieve]
  46. Chodosh, L. A., Baldwin, A. S., Carthew, R. W., and Sharp, P. A. (1988) Cell 53, 11-24 [Medline] [Order article via Infotrieve]
  47. Kadonaga, J. T., Jones, K. A., and Tjian, R. (1986) Trends Biochem. Sci. 11, 20-23 [CrossRef]
  48. Lucas, P. C., and Granner, D. K. (1992) Annu. Rev. Biochem. 61, 1131-1173 [CrossRef][Medline] [Order article via Infotrieve]
  49. Wasylyk, C., Flores, P., Gutman, A., and Wasylyk, B. (1989) EMBO J. 8, 3371-3378 [Abstract]
  50. Lee, W., Mitchell, P., and Tjian, R. (1987) Cell 49, 741-752 [Medline] [Order article via Infotrieve]
  51. Rauscher, F. J., III, Sambucetti, L. C., Curran, T., Distel, R. J., and Spiegelman, B. M. (1988) Cell 52, 471-480 [Medline] [Order article via Infotrieve]
  52. Johnson, P. F. (1993) Mol. Cell. Biol. 13, 6919-6930 [Abstract]
  53. Brüggemeier, U., Kalff, M., Franke, S., Scheidereit, C., and Beato, M. (1991) Cell 64, 565-572 [Medline] [Order article via Infotrieve]
  54. Imagawa, M., Chiu, R., and Karin, M. (1987) Cell 51, 251-260 [Medline] [Order article via Infotrieve]
  55. Jones, K. A., Kadonaga, J. T., Rosenfeld, P. J., Kelly, T. J., and Tjian, R. (1987) Cell 48, 79-89 [Medline] [Order article via Infotrieve]
  56. Baum, M., Cano, A., and Alpern, R. J. (1993) Am. J. Physiol. 264, F1027-F1031
  57. Azarani, A., Goltzman, D., and Orlowski, J. (1995) J. Biol. Chem. 270, 20004-20010 [Abstract/Free Full Text]
  58. Reilly, R. F., Hildebrandt, F., Biemesderfer, D., Sardet, C., Pouysségur, J., Aronson, P. S., Slayman, C. W., and Igarashi, P. (1991) Am. J. Physiol. 261, F1088-F1094
  59. Soleimani, M., Bookstein, C., McAteer, J. A., Hattabaugh, Y. J., Bizal, G. L., Musch, M. W., Villereal, M., Rao, M. C., Howard, R. L., and Chang, E. B. (1994) J. Biol. Chem. 269, 15613-15618 [Abstract/Free Full Text]
  60. Levine, S. A., Nath, S. K., Yun, C. H. C., Yip, J. W., Montrose, M., Donowitz, M., and Tse, C. M. (1995) J. Biol. Chem. 270, 13716-13725 [Abstract/Free Full Text]
  61. Bucher, P. (1990) J. Mol. Biol. 212, 563-578 [Medline] [Order article via Infotrieve]
  62. Pathak, B. G., Pugh, D. G., and Lingrel, J. B. (1990) Genomics 8, 641-647 [Medline] [Order article via Infotrieve]
  63. Juang, H. H., Costello, L. C., and Franklin, R. B. (1995) J. Biol. Chem. 270, 12629-12634 [Abstract/Free Full Text]
  64. Johnson, P. F. and McKnight, S. L. (1989) Annu. Rev. Biochem. 58, 799-839 [CrossRef][Medline] [Order article via Infotrieve]
  65. Vallejo, M., Ron, D., Miller, C. P., and Habener, J. F. (1993) Proc. Natl. Acad. Sci. U. S. A. 90, 4679-4683 [Abstract]
  66. Pittman, N., Shue, G., LeLeiko, N. S., and Walsh, M. J. (1995) J. Biol. Chem. 270, 28848-28857 [Abstract/Free Full Text]
  67. Horie, S., Moe, O., Miller, R. T., and Alpern, R. J. (1992) J. Clin. Invest. 89, 365-372 [Medline] [Order article via Infotrieve]
  68. Cano, A., Preisig, P., and Alpern, R. J. (1993) J. Clin. Invest. 92, 1632-1638 [Medline] [Order article via Infotrieve]
  69. Kinsella, J., Cujdik, T., and Sacktor, B. (1984) J. Biol. Chem. 259, 13224-13227 [Abstract/Free Full Text]
  70. Horie, S., Moe, O., Tejedor, A., and Alpern, R. J. (1990) Proc. Natl. Acad. Sci. U. S. A. 87, 4742-4745 [Abstract]
  71. Kinsella, J., and Sacktor, B. (1985) Proc. Natl. Acad. Sci. U. S. A. 82, 3606-3610 [Abstract]
  72. Rahmsdorf, H. J., and Herrlich, P. (1990) Pharmacol. Ther. 48, 157-188 [CrossRef][Medline] [Order article via Infotrieve]
  73. Gutman, A., and Wasylyk, B. (1990) EMBO J. 9, 2241-2246 [Abstract]
  74. Roesler, W. J., Vandenbark, G. R., and Hanson, R. W. (1988) J. Biol. Chem. 263, 9063-9066 [Free Full Text]
  75. Näär, A. N., Boutin, J.-M., Lipkin, S. M., Yu, V. C., Holloway, J. M., and Rosenfeld, M. G. (1991) Cell 65, 1267-1279 [Medline] [Order article via Infotrieve]
  76. Umesono, K., Murakami, K. K., Thompson, C. C., and Evans, R. M. (1991) Cell 65, 1255-1266 [Medline] [Order article via Infotrieve]
  77. Forman, B. M., Casanova, J., Raaka, B. M., Ghysdael, J., and Samuels, H. H. (1992) Mol. Endocrinol. 6, 429-442 [Abstract]
  78. Brüggemeier, U., Rogge, L., Winnacker, E.-L., and Beato, M. (1990) EMBO J. 9, 2233-2239 [Abstract]
  79. Miner, J. N., Diamond, M. I., and Yamamoto, K. R. (1991) Cell Growth Differ. 2, 525-530 [Medline] [Order article via Infotrieve]
  80. Morduchowicz, G. A., Sheikh-Hamad, D., Jo, O. D., Nord, E. P., Lee, D. B. N., and Yanagawa, N. (1989) Kidney Int. 36, 576-581 [Medline] [Order article via Infotrieve]
  81. Mahnensmith, R. L., and Aronson, P. S. (1985) Circ. Res. 56, 773-788 [Abstract]
  82. Booth, I. W., Stange, G., Murer, H., Fenton, T. R., and Milla, P. J. (1985) Lancet i, 1066-1069
  83. Thompson, J. D., Higgins, D. G., and Gibson, T. J. (1994) Nucleic Acids Res. 22, 4673-4680 [Abstract]
  84. Prestridge, D. S. (1991) Comput. Appl. Biosci. 7, 203-206 [Abstract]

©1996 by The American Society for Biochemistry and Molecular Biology, Inc.