Expression Cloning of a Novel Estrogenic Mouse 17ß-Hydroxysteroid Dehydrogenase/ 17-Ketosteroid Reductase (m17HSD7), Previously Described as a Prolactin Receptor-Associated Protein (PRAP) in Rat

Pasi Nokelainen, Hellevi Peltoketo, Reijo Vihko and Pirkko Vihko

Biocenter Oulu and World Health Organization Collaborating Centre for Research on Reproductive Health University of Oulu FIN-90220 Oulu, Finland


    ABSTRACT
 TOP
 ABSTRACT
 INTRODUCTION
 RESULTS
 DISCUSSION
 MATERIALS AND METHODS
 REFERENCES
 
17ß-Hydroxysteroid dehydrogenases/17-ketosteroid reductases (17HSDs) modulate the biological activity of certain estrogens and androgens by catalyzing reductase or dehydrogenase reactions between 17-keto- and 17ß-hydroxysteroids. In the present study, we demonstrate expression cloning of a novel type of 17HSD, chronologically named 17HSD type 7, from the HC11 cell line derived from mouse mammary gland. The cloned cDNA, 1.7 kb in size, encodes a protein of 334 amino acids with a calculated molecular mass of 37,317 Da. The primary structure contains segments characteristic of enzymes belonging to the short-chain dehydrogenase/reductase superfamily. Strikingly, mouse 17HSD type 7 (m17HSD7) shows 89% identity with a recently cloned rat protein called PRL receptor-associated protein (PRAP). The function of PRAP has not yet been demonstrated.

The enzymatic characteristics of m17HSD7 and RT-PCR-cloned rat PRAP (rPRAP) were analyzed in cultured HEK-293 cells, where both of the enzymes efficiently catalyzed conversion of estrone (E1) to estradiol (E2). With other substrates tested no detectable 17HSD or 20{alpha}-hydroxysteroid dehydrogenase activities were found. Kinetic parameters for m17HSD7 further indicate that E1 is a preferred substrate for this enzyme. Relative catalytic efficiencies (Vmax/Km values) for E1 and E2 are 244 and 48, respectively. As it is the case with rPRAP, m17HSD7 is most abundantly expressed in the ovaries of pregnant animals. Further studies show that the rat enzyme is primarily expressed in the middle and second half of pregnancy, in parallel with E2 secretion from the corpus luteum. The mRNA for m17HSD7 is also apparent in the placenta, and a slight signal for m17HSD7 is found in the ovaries of adult nonpregnant mice, in the mammary gland, liver, kidney, and testis.

Altogether, because of their similar primary structures, enzymatic characteristics, and the tissue distribution of m17HSD7 and rPRAP, we suggest that rPRAP is rat 17HSD type 7. Furthermore, the results indicate that 17HSD7 is an enzyme of E2 biosynthesis, which is predominantly expressed in the corpus luteum of the pregnant animal.


    INTRODUCTION
 TOP
 ABSTRACT
 INTRODUCTION
 RESULTS
 DISCUSSION
 MATERIALS AND METHODS
 REFERENCES
 
17ß-Hydroxysteroid dehydrogenases (17HSDs) catalyze the conversion of low-activity 17-ketosteroids [e.g. estrone (E1), androstenedione (A-dione), and 5{alpha}-androstanedione] to their more potent 17ß-hydroxysteroid forms [estradiol (E2), testosterone (T), and dihydrotestosterone] and vice versa. Reductive 17HSDs are essential for the biosynthesis of E2 and testosterone in the gonads and, in addition, they modulate the activity of these steroids in certain extragonadal tissues of several species, especially primates. Oxidative 17HSDs are widely expressed and are suggested to play a role in the inactivation of 17ß-hydroxysteroids shifted to the tissues from the gastrointestinal tract, blood circulation, and amniotic fluid (Refs. 1, 2, 3, 4 and references therein). Four different human 17HSDs (types 1–4) (5, 6, 7, 8, 9) and six rodent 17HSDs (types 1–6) have been cloned to date (10, 11, 12, 13, 14, 15, 16, 17, 18). In addition to their distinct primary structures, they are dissimilarly distributed, they differ in substrate specificity, and consequently they have apparently separate physiological functions. 17HSD types 1–4 and 6 belong to the short-chain dehydrogenase/reductase (SDR) protein family, whereas the type 5 enzyme belongs to the aldoketoreductase family.

Up to now, estrogens have mostly been found to be modulated by 17HSD type 1 (17HSD1) and type 2 (17HSD2). 17HSD1 is predominantly expressed in ovarian granulosa cells (10, 19, 20, 21, 22) and in the syncytiotrophoblasts of human placenta (23, 24), being essential for E2 biosynthesis. The human type 1 enzyme is also expressed in certain estrogen target tissues such as breast (25) and endometrium (26), where it converts circulating E1 to E2. While human 17HSD1 primarily catalyzes reactions between phenolic steroids, i.e. estrogens (27), rodent 17HSD1 enzymes are also able to catalyze androgens efficiently (10, 28). In contrast to type 1, the oxidative 17HSD2 decreases the biological activity of estrogens and androgens and may thus protect tissues from excessive hormone action (3, 4, 7, 29). The type 2 enzyme is particularly expressed in human and rodent placenta, liver, kidney, and small intestine (7, 11, 16, 29). It is the predominant 17HSD in human endometrium (30), and it is also expressed in human prostate, at least in neoplastic tissue (31).

The function of other 17HSDs in estrogen biosynthesis and metabolism is less well understood. 17HSD type 3 (17HSD3) is indispensable for testicular T biosynthesis and is thus crucial in male sexual differentiation and reproduction, although it also reduces E1 to E2 (8). 17HSD type 4 (17HSD4) is part of peroxisomal multifunctional enzyme II, whose role in steroid metabolism appears to be minor compared with the other activities of the enzyme (9, 17, 32, 33). 17HSD type 5 (17HSD5), in turn, which shows oxidative 17HSD activity toward androgens, E2, and xenobiotics, is mainly expressed in the liver and kidney, and the physiological role of the enzyme has remained open (14). Finally, the recently cloned 17HSD type 6 (17HSD6) takes part in an inactivation path of dihydrotestosterone and is most abundantly expressed in the prostate and liver (18).

In the present study we demonstrate cloning of a novel 17HSD which we suggest to be involved in E2 biosynthesis. The main source of E2 in cycling humans and rodents is the ovary. E2 is synthesized in the granulosa cells of developing follicles from theca cell-derived androgens, by P450 aromatase (P450arom), and 17HSD1 (15, 19, 20, 21, 34). After ovulation, the follicles luteinize and turn to corpora lutea, which secrete progesterone (P), E2, and peptide hormones, for example. When impregnation occurs, the rat corpus luteum (CL) further develops as a result of PRL stimulation being essential for maintaining the pregnancy (35). During human pregnancy the placenta develops as a major source of E2, whereas in rodents E2 is produced in the ovaries from ovarian and placental precursors. Throughout gestation, some rodent follicles mature and express 17HSD1 (36), but from midpregnancy until parturition E2 is secreted mainly from the CL (37). 17HSD1, however, has not been found in the CL of either cycling or pregnant rats (15, 36), in contrast to human granulosa-luteal cells (19, 20, 22), suggesting that, at least in rodents, another enzyme is responsible for E2 biosynthesis in the CL. We have now cloned a novel type of 17HSD from a cell line originating from mouse mammary gland epithelial cells, but which is most abundantly expressed in the ovary of the pregnant mouse, and which efficiently catalyzes the reaction from E1 to E2. The enzyme, which has chronologically been named 17HSD type 7 (17HSD7), shows great identity to PRAP, recently cloned from rat ovary (38).


    RESULTS
 TOP
 ABSTRACT
 INTRODUCTION
 RESULTS
 DISCUSSION
 MATERIALS AND METHODS
 REFERENCES
 
Expression Cloning of m17HSD7
When screening different cell lines for 17HSD activity, we found that the HC11 cell line originating from the epithelial cells of the mammary gland of a pregnant mouse (39) possessed strong estrogenic 17HSD activity. E1 was effectively converted to E2 in the cultured cells but, in contrast to the activity of m17HSD1, the observed activity poorly catalyzed A-dione to T (Fig. 1Go). HC11 cells were not able to catalyze oxidative 17HSD reactions either. Probes for mouse 17HSD1 (m17HSD1) as well as those for mouse 17HSD type 2 and 5, human 17HSD3, and rat 17HSD4 did not recognize any mRNAs in HC11 cell samples (data not shown). The cell line was therefore assumed to express an unidentified type of 17HSD enzyme. An expression cDNA library was prepared from poly(A)+-enriched RNA of the HC11 cell line, and a total of 600,000 independent clones from the library were screened by monitoring their capability to convert E1 to E2.



View larger version (21K):
[in this window]
[in a new window]
 
Figure 1. Characterization of 17HSD Activity in Cultured HC11 Cells

The cells were seeded into six-well plates at a density of 230,000 cells per well and allowed to grow for 2 days. 17HSD activity measurements were then carried out in the cultured cells using estrogens (E1, E2) and androgens (A-dione, T) as substrates (100 nM) in the culture media. After incubation, the media were collected and the amount of product was measured.

 
Of the 243 plasmid pools, each containing DNA from 2500 individual colonies, two gave a positive signal, i.e. possessed 17HSD activity. These two plasmid pools were subdivided into smaller pools until two independent cDNA samples (m17HSD7.1 and m17HSD7.2) were obtained and sequenced. The two clones were found to be identical except that the m17HSD7.1 clone had a 42 bp longer 5'-region than the m17HSD7.2 clone and, in addition, the m17HSD7.2 clone had a G-to-T substitution at position 1414, found to be in the noncoding region (Fig. 2Go). The m17HSD7.1 clone had an open reading frame of 1005 bp, a 5'-region of 63 bp, and a 3'-region of 670 bp followed by a polyadenylate tract. The open reading frame encoded a peptide of 334 amino acids with a predicted molecular mass of 37,317 Da, and the protein was named m17HSD7. A stop codon situated in the frame upstream of the ATG triplet further indicated that the m17HSD7.1 clone contained the whole coding region. A putative polyadenylation signal (AATTAAA) was located at position 1703 in the 3'-noncoding region of the clone (Fig. 2Go).



View larger version (87K):
[in this window]
[in a new window]
 
Figure 2. Nucleotide Sequence and Deduced Amino Acid Sequence of m17HSD7

Stop codons are marked with a box, and potential polyadenylation signal is underlined. The nucleotide sequence data reported in this paper will appear in the EMBL, GenBank, and DDBJ Nucleotide Sequence Databases under accession number Y15733.

 
Comparison of the predicted amino acid sequence of m17HSD7 with the sequences in the protein databases revealed a significant 89% identity between m17HSD7 and a protein called PRAP recently cloned from the rat (38) (Fig. 3Go). Analysis of the amino acid sequence of m17HSD7 against the PROSITE database, allowing one mismatch, further showed that the protein belonged to the SDR superfamily. It contained the typical glycine pattern (Gly9-Xaa-Xaa-Xaa-Gly13-Xaa-Gly15 in m17HSD7) and the three amino acid residues (Ser180, Tyr193 and Lys197) (Fig. 3Go) that are involved in the catalytic reaction and are characteristic of SDR enzymes (40, 41). Identity of the predicted amino acid sequence to other mouse 17HSD enzymes was 18–28%. In addition, several putative posttranslational modification sites were found in m17HSD7: four N-glycosylation sites at positions 37, 127, 178, and 229, three casein kinase II phosphorylation sites at positions 118, 125, and 180, and two protein kinase C phosphorylation sites at positions 170 and 195 (Fig. 3Go).



View larger version (79K):
[in this window]
[in a new window]
 
Figure 3. Comparison of the Amino Acid Sequences of m17HSD7 (m7) and r17HSD7/PRAP (r7)

The glycine pattern and the conserved amino acids recognized as the SDR consensus sequence in the PROSITE database, in both enzymes, are marked with white and gray boxes, respectively. Putative sites for N-linked glycosylation (#), casein kinase II phosphorylation (- - -), and protein kinase C phosphorylation (+++) are marked above the sequences. These consensus sequences are present in both proteins, excluding the N-linked glycosylation site at position 127, which is present only in m17HSD7. A hydrophobic region suggested to be a membrane-spanning domain in r17HSD7/PRAP (38 ) (underlined) is also included in m17HSD7, according to a hydropathy plot (data not shown).

 
Characterization of the Enzymatic Properties of m17HSD7
To characterize the enzymatic activities of the cloned m17HSD7, the cDNA was expressed under a cytomegalovirus (CMV)-promoter in human embryonic kidney 293 (HEK-293) cells. As shown in Fig. 4AGo, m17HSD7 catalyzed only a reductive reaction from E1 to E2. No significant 17HSD or 20{alpha}-hydroxysteroid dehydrogenase (20HSD) activity was observed when E2, A-dione, T, P, or 20{alpha}-hydroxyprogesterone (20-OHP) were added to the reaction at concentrations of 100 nM (Fig. 4AGo) or 1 nM (data not shown). To compare the efficacies of m17HSD1 and m17HSD7, the cDNAs encoding them were separately transfected into HEK-293 cells. The comparison indicated that the 17HSD activity of the type 7 enzyme was relevant, i.e. of same magnitude as that of 17HSD1, even though it was not as great as that of the type 1 enzyme in the conditions used (Fig. 4BGo). As also described previously, m17HSD1 catalyzes reactions from E1 to E2 and from A-dione to T (Fig. 4BGo and Ref. 10), while m17HSD7 was able to catalyze only the reduction of E1 (Fig. 4Go).



View larger version (21K):
[in this window]
[in a new window]
 
Figure 4. Characterization of 17HSD Activity of m17HSD7 in Cultured HEK-293 Cells after Transient Transfection with m17HSD7 cDNA (A) and Comparison with m17HSD1 (B)

Cells grown in six-well plates (300,000 cells per well) were transfected with 0.5 µg of m17HSD7.1-cDNA3.1 plasmids per well (A), 0.5 µg of m17HSD7.1-cDNA3.1 plasmids (B) or m17HSD1-CMV6 plasmids per well. 17HSD activity measurements were carried out in cultured cells using estrogens (E1, E2), androgens (A-dione, T), or progestins (20-OHP, P) as substrates (100 nM) in the culture media. After incubation periods, the media were collected and the amount of product was measured. The background 17HSD activity observed was subtracted from the results. The experiments were repeated twice and typical reaction curves and columns are shown.

 
To further characterize the enzymatic properties of the novel 17HSD, HEK-293 cells were transiently transfected with m17HSD7 cDNA and then homogenized in phosphate buffer containing 20% glycerol and 0.08% nonionic detergent (Big Chap), which were found to stabilize the labile enzyme (data not shown). Kinetic parameters (Km and Vmax) were determined for E1 and E2 (Table 1Go). The Vmax values for E1 and E2 did not differ considerably, but the Km values, and consequently the catalytic efficiencies (Vmax/Km), clearly showed that of these two E1 is the preferred substrate for the enzyme. While 49% of E1 (1 µM) was converted to E2 in 20 min, the 17HSD activity of the homogenate toward A-dione, T, dehydroepiandrosterone (DHEA), and androst-5-ene-3ß,17ß-diol (A-diol) (0.5 µM) was 0.5–1.5% in 2 h and, therefore, kinetic constants could not be determined for these steroids. In similar conditions, conversions between 20-OHP and P were less than 2.5%. Finally, in the conditions used, the conversion of E1 to E2 was approximately 40 times faster when the phosphorylated form of cofactor, NADPH, was used as when NADH was added to the reaction. Likewise, the rate of reaction from E2 to E1 was 7-fold greater with NADP+ than with NAD+.


View this table:
[in this window]
[in a new window]
 
Table 1. Kinetic Parameters of m17HSD7

 
Tissue Distribution of m17HSD7
In Northern blot analysis, the m17HSD7 cDNA probe recognized mRNAs of 4.6, 3.6, 1.7, and 1.2 kb in size, of which the 4.6-kb transcript was the dominant form (Fig. 5AGo). The 4.6-kb mRNA was very abundantly expressed in the ovaries of pregnant (14–20 day) mice (Fig. 5Go, A and B). This mRNA was also expressed in the placenta and was detectable in the mammary gland, liver, kidney, testis, the HC11 cell line, and in the ovaries of nonpregnant mice. Since the major mRNA recognized (4.6 kb) and the cDNA cloned (1.7 kb) differed so remarkably in size, and also because 28S ribosomal RNA is approximately 4.7 kb in size, the authenticity of the weak signals detected was confirmed by RT-PCR. A specific fragment of 0.5 kb was amplified from all the tissues tested, including a barely detectable amplification signal from the spleen and uterine samples, and excluding the pancreatic sample, where the RNA was found to be degraded (Fig. 5CGo). The quality of the total RNA used for RT-PCR was tested by checking the integrity of ribosomal bands in agarose-formaldehyde gel (data not shown) and also by performing RT-PCR with mouse glyceraldehyde-3-phosphate dehydrogenase (GAPDH) primers (Fig. 5CGo).



View larger version (27K):
[in this window]
[in a new window]
 
Figure 5. Tissue Distribution of m17HSD7 Analyzed by Northern Blot (A and B) and RT-PCR (C) Analyses

A, Six micrograms of poly(A)+-enriched RNA extracted from adult female mouse tissues, from testis, and from the HC11 cell line (10 µg) were subjected to Northern blotting. Ovaries were collected from pregnant mice at gestation days 14–20, mammary glands were from gestation days 15 and 20, and placentae were from day 15. Note the variable exposure times. B, Expression of the enzyme in the ovaries of normal cycling (control) and pregnant mice (14–20 days) was analyzed in a separate experiment using 20 µg total RNA. The exposure time was 16 h. The sizes of the mRNA transcripts were determined using a commercial RNA ladder, and the amount of RNA loaded was controlled by hybridization of the filter with a rat GAPDH cDNA probe (bottom; exposure time, 20 h). (C) RT-PCR was carried out using 125 ng of the same total RNA preparation as used for the isolation of poly(A)+-enriched RNAs for the Northern blot analysis (panel A). Primers for m17HSD7 and for mouse GAPDH were used. Eighty percent of each PCR product was run in a agarose gel, blotted, and analyzed by hybridization with m17HSD7 or rat GAPDH probes. Cont., Negative control samples containing sterile H2O instead of RNA template.

 
Characterization of the Enzymatic Nature of rPRAP Suggested to be Rat 17HSD Type 7, and the Expression Pattern of the Enzyme in the Ovary during Gestation
rPRAP was cloned by RT-PCR from RNA isolated from the ovaries of pregnant rats, using the sequence described (38). The rPRAP cDNA was then subcloned to CMV6-vectors and transfected to HEK-293 cells. In a identical manner to the mouse type 7 enzyme, the rat enzyme catalyzed a reductive reaction from E1 to E2 in the cultured cells (Fig. 6Go). Neither 17HSD nor 20HSD activity was observed when E2, A-dione, T, 20-OHP, or P were used as substrates.



View larger version (19K):
[in this window]
[in a new window]
 
Figure 6. Characterization of the Activity of Transiently Transfected r17HSD7/PRAP in Cultured HEK-293 Cells

Cells grown in six-well plates (300,000 cells per well) were transfected with 1 µg of r17HSD7/PRAP-CMV6 plasmid per well. Estrogens (E1, E2), androgens (A-dione, T), or progestins (20-OHP, P) were added to the culture media at a concentration of 100 nM to determine 17HSD and 20HSD activities. After incubation, the media were collected and the amount of product was measured. Conversion of E1 to E2 was measured after 2, 4, 8, and 16 h, while other conversions were measured only after 20 h. The background activities observed were subtracted from the results. The experiments were repeated and typical conversions are shown.

 
m17HSD7 mRNA was most abundant in ovarian samples from pregnant mice (Fig. 5Go), and rPRAP has also been shown to be mostly expressed during pregnancy (42). Expression of the enzyme was therefore followed in more detail in rat ovaries during pregnancy. Northern blot analysis revealed two distinct PRAP mRNAs, 1.8 and 4.3 kb in size, of which the longer species is dominant (Fig. 7Go), similar to m17HSD7 mRNAs. After 40 h exposure the 4.3-kb mRNA could be detected in samples taken throughout the first part of pregnancy, on days 1–5, but the expression of PRAP was remarkably up-regulated at day 8. Expression of the 4.3-kb mRNA continued at a high level until close to parturition. On day 21 its expression was again barely detectable, after a long (40 h) exposure.



View larger version (68K):
[in this window]
[in a new window]
 
Figure 7. Expression of 17HSD7/PRAP in the Ovaries of Pregnant Rats

Thirty micrograms of total RNA from the ovaries of pregnant rats on days 1, 3, 5, 8, 12, 15, 18, and 21 were analyzed in Northern blot analysis. The upper and middle panels represent the same film after two different exposure times. The sizes of the mRNA transcripts were determined using a commercial RNA ladder, and the amount of RNA loaded was controlled by hybridization with rat GAPDH cDNA (bottom).

 

    DISCUSSION
 TOP
 ABSTRACT
 INTRODUCTION
 RESULTS
 DISCUSSION
 MATERIALS AND METHODS
 REFERENCES
 
In addition to being crucial for gonadal sex steroid synthesis, 17HSD activity has been found ubiquitously in human and rodent tissues (43, 44). Within the last few years, the 17HSD family has rapidly expanded to six different types, although not all of them have been linked primarily to steroidogenesis. In the present work we describe the expression cloning of a novel type of 17HSD, chronologically named 17HSD7. m17HSD7 is an enzyme of 334 amino acids and it belongs to the SDR protein family, as do 17HSD types 1–4 and 6. The identity of m17HSD7 with other mouse 17HSD enzymes is between 18–28%, which is in the range characteristic of SDR members (40). Like 17HSD1, the type 7 enzyme contains the three critical amino acid residues, Ser180, Tyr193, and Lys197, of the catalytic triad (28, 41, 45) that lie on the segment recognized as an SDR signal in the PROSITE database. The primary structure also includes the glycine pattern that forms part of the ß{alpha}ß Rossmann fold involved in the binding of a nucleotide cofactor (40).

While the identity between m17HSD7 and other 17HSDs is less than 30%, the cloned enzyme shares 89% identity with a rat protein called PRAP (38). rPRAP also contains the SDR consensus sequence, including the three highly conserved amino acid residues, and the glycine pattern (see Fig. 3Go), which are needed for 17HSD activity in other 17HSD enzymes (45). In the present work we further demonstrate the 17HSD activity of rPRAP, hereafter called r17HSD7/PRAP, thus confirming the identity between 17HSD7 and PRAP. r17HSD7/PRAP is a microsomal phosphoprotein that has been shown to be associated with a short form of PRL receptor (PRL-R) (38) and, like r17HSD7/PRAP, m17HSD7 contains several putative N-linked glycosylation and phosphorylation sites. Lability of m17HSD7, which can be decreased with a nonionic detergent, and a putative transmembrane domain of the enzyme similar to one in r17HSD7/PRAP (38) also point to the membrane-associated character of the mouse type 7 enzyme.

Characterization of 17HSD activities of m17HSD7 and r17HSD7/PRAP in cultured HEK-293 cells showed that in in vivo conditions both enzymes exclusively catalyze the reduction of E1 to E2. Thus the rodent type 7 enzymes resemble human 17HSD1, but differ from rodent type 1 enzymes, which also catalyze androgens efficiently (10, 28). In addition, the activity of mouse and rat 17HSD7 enzymes corresponds to the activity originally found in the HC11 cell line. Comparison of the activities of m17HSD1 and m17HSD7 further indicates that the type 1 and 7 enzymes catalyze E1 to E2 to the same degree and thus a noteworthy role for the type 7 enzyme in E2 biosynthesis can be presented. Cell homogenate enriched with 17HSD7 is also able to catalyze, to some extent, the reverse reaction from E2 to E1 in the presence of excess NADP+. As in the case of human 17HSD1 (28), the direction of the reaction may be reversed by an excess of cofactor in vitro. Km and Vmax/Km values and particularly the results obtained in vivo, however, show that the type 7 enzyme evidently prefers E1 over E2 as a substrate. Hence, 17HSD7 belongs to the group of reductive enzymes, as do types 1 and 3, while the other isoenzymes, 2, 4, 5, and 6, favor 17ß-hydroxy substrates. As suggested to be typical of reductive enzymes (46), m17HSD7 also utilizes the phosphorylated forms of cofactor, NADPH and NADP+, more efficiently than the nonphosphorylated forms.

Northern blot analysis of the mouse type 7 enzyme revealed a remarkable difference between the size of the cDNA (1.7 kb) and the major mRNA (4.6 kb). The 1.7-kb mRNA is barely detectable as compared with the 4.6-kb transcript. This is also the case with r17HSD7/PRAP cDNA (1.8 kb in size), which recognizes two to three mRNAs of 1.8 kb, 4.3 kb, and 5.5 kb (present study and Ref. 47). In rat tissues, the mRNAs longer than 4 kb are also the dominant forms. A possible explanation for the difference in lengths of the cDNAs and major mRNAs is an exceptionally long 5'-noncoding sequence, or a 3'-noncoding region including more than one polyadenylation signal. Thorough studies, most likely structural analyses of the HSD17B7 genes, are needed to solve the size differences. Nevertheless, the results of functional studies and an in-frame stop codon in the 5'-noncoding area confirm that the cDNA contains the whole coding area, and correspondence of results of Northern blotting and RT-PCRs indicates that the mRNA, 4.6 kb in size, is the major m17HSD7 transcript.

In agreement with the results published by Duan and co-workers (47), our results show that 17HSD7/PRAP is most abundantly expressed in the ovaries of pregnant rodents, even though the 17HSD7/PRAP mRNA and protein is also detectable in ovarian samples from nonpregnant animals (present study and Ref. 42). 17HSD7/PRAP is expressed in the ovaries throughout pregnancy but it is particularly strongly up-regulated around day 8 of pregnancy, and this is sustained at least until day 18. Data of Parmer and co-workers (42) and Duan and co-workers (47) confirmed that the expression of 17HSD7/PRAP occurs in the CL. During rodent pregnancy, biosynthesis of E2 especially takes place in the CL, even though some follicles continue to mature and the granulosa cells are therefore also secreting E2. In the first half of rodent pregnancy, steroidogenesis in the CL is stimulated by PRL and LH, and E2 is synthesized from ovarian androgen precursors (Ref. 35 and references therein). LH action on the CL of the pregnant rat is known to be mediated by LH stimulation of E2 biosynthesis. E2, together with decidual luteotropin (DLt), further maintains P production. This is in line with the data showing that 17HSD7/PRAP is tightly regulated by E2 and that removal of the effect of tropic hormones on the CL leads to disappearance of r17HSD7/PRAP (47).

At midpregnancy, remarkable endocrinological changes take place, such as down-regulation of LH and pituitary PRL, and a luteal-placental shift as a result of which the placenta starts production of the androgen precursors needed for increased E2 synthesis in the ovaries. E2 then acts in an autocrine and intracrine manner together with placental lactogen (PL), causing differentiation of the CL, characterized by hypertrophy of luteal cells, vascularization of the CL, and stimulation of steroidogenesis [synthesis of progesterone in particular (35)]. High expression of 17HSD7/PRAP from day 8 onward and during the second half of pregnancy coincides first with LH-induced E2 production from the CL and then enhanced A-dione secretion from the placenta, the expression of P450arom in the CL (48), and elevated serum E2 concentrations (37). Altogether, our results, in combination with those reported by Duan and co-workers (38, 47), strongly suggest that 17HSD7/PRAP, not 17HSD1, is the enzyme required for E2 biosynthesis in the CL.

We demonstrate here a steroidogenic function for 17HSD7/PRAP. Moreover, the enzyme has been shown to be associated with the short form of PRL-R, and a potential role for 17HSD7/PRAP in PRL signaling has been much discussed (38, 47). Both long and short forms of PRL-R are expressed in the CL throughout rodent pregnancy (49). While the longer form of the receptor mediates the PRL signal via the JAK2/Stat5 system, the function of the short form is known only superficially (50, 51). PRL or PRL-like hormones, DLt and PL, and E2 act together on the CL throughout rodent pregnancy. Furthermore, PRL and PRL-like hormones maintain high concentrations of estrogen receptor in the CL (Ref. 35 and references therein), and 17HSD7/PRAP is up-regulated not only by its product E2, but also by PRL (47). It will thus be of great interest to further investigate the cross-talk between E2- and PRL-signaling systems in the CL and the possible role of 17HSD7/PRAP. One question of interest is the association between 17HSD7/PRAP and the cell membrane (38) in terms of its steroidogenic activity and possible role in PRL signaling.

In addition to the ovaries, 17HSD7 was found to be expressed in mouse placenta, and weak signals for m17HSD7 were also detected in the mammary gland, liver, kidney, and testis of the mouse. Moreover, m17HSD7-PCR-amplicons were detected in samples from the uterus, brain, and small intestine. r17HSD7/PRAP mRNAs have been detected only in the ovaries (47), which may be due to species-specific differences or, more likely, differences in sensitivity of the methods used in recognition of the mRNAs. The physiological significance, if any, of low expression of 17HSD7 in the mouse tissues remains open. Transcripts of m17HSD3, for example, have also been detected in several male and female tissues, in addition to its major expression in the testis, when a very sensitive method, RT-PCR, has been used (12). The presence of 17HSD7 in HC11 cells and low but detectable concentrations of 17HSD7 mRNA in samples from the mammary glands of pregnant mice suggest, however, that 17HSD7 is expressed in the mouse mammary gland, at least during pregnancy. 17HSD7 might therefore catalyze circulating E1 to E2 locally in the mammary gland, in a similar manner as 17HSD1 has been suggested to do in human breast epithelial cells (1, 52). Interestingly, the mammary gland is also a well known target of both E2 and PRL action.

The expression of 17HSD7 in mouse placenta introduces new issues in placental steroidogenesis. In contrast to the human placenta, rodent placentae have been assumed to be unable to synthesize E2 as a result of their lack of P450arom (53) and 17HSD1 (10, 11). The mouse placenta is rich in 17HSD2 (4, 11), but Blomquist and co-workers (54) have shown that it may also express multiple forms of 17HSDs. 17HSD7 might thus catalyze circulating E1 to E2 in the placenta for local needs, while the oxidative type 2 enzyme is suggested to prevent the transfer of active 17ß-hydroxy forms of sex steroids between the fetus and the maternal circulation (4).

In summary, a novel type of 17HSD has been cloned from the HC11 cell line. 17HSD7, also known as PRAP, is primarily expressed in the ovaries of pregnant rodents. Based on the enzymatic character of 17HSD7/PRAP and its expression pattern, which shows developmental regulation in the ovary during gestation, we suggest that 17HSD7/PRAP is responsible for the final step in the biosynthesis of E2 in the CL.


    MATERIALS AND METHODS
 TOP
 ABSTRACT
 INTRODUCTION
 RESULTS
 DISCUSSION
 MATERIALS AND METHODS
 REFERENCES
 
Chemicals and Reagents
[A-dione, A-diol, DHEA, E2, E1, 20-OHP, P, and T were purchased from Steraloids, Inc. (Wilton, NH). [1,2,6,7-3H]A-dione (110 Ci/mmol), [2,4,6,7-3H]E2 (81 Ci/mmol), [2,4,6,7-3H]E1 (99 Ci/mmol), [1,2,6,7,16,17-3H]P (97 Ci/mmol), [1,2,6,7-3H]T (105 Ci/mmol), and [{alpha}32P]deoxy-CTP (3000 Ci/mmol) were obtained from Amersham Life Science (Little Chalfont, Buckinghamshire, UK). [1,2-3H(N)]A-diol (56 Ci/mmol), [1,2,6,7-3H(N)]DHEA (92 Ci/mmol), and [1,2-3H(N)]20-OHP (52 Ci/mmol) were from DuPont-New England Nuclear Life Science (Boston, MA). The transfection reagent, N-[1-(2, 3-dioleoyloxy)propyl]-N,N,N-trimethylammoniummethylsulfate (DOTAP) was a product of Boehringer Mannheim GmbH (Mannheim, Germany). Other reagents not mentioned in the text were purchased from Boehringer Mannheim GmbH, Sigma Chemical Co. (St. Louis, MO), or Merck AG (Darmstadt, Germany) and were of the highest purity grade available.

Expression Cloning
For the construction of a cDNA library, poly(A)+-enriched RNA was isolated using FastTrack 2.0 kit (Invitrogen, San Diego, CA) from the HC11 cell line, derived from the epithelial cells of the mammary gland of a pregnant mouse (39). The cDNA library was generated in cDNA3.1 plasmids containing a cytomegalovirus promoter and transformed to Escherichia coli TOP10F' cells by Invitrogen. The cDNA library obtained was then divided into pools of 2500 colonies on LB-ampicillin (50 µg/ml) plates. Colonies from each pool were collected and combined, and the plasmid fractions were isolated using a Qiagen Midi plasmid purification kit (Hilden, Germany).

The cDNA library was screened based on the method described by Wu et al. (7). Briefly, 5 µg of plasmid DNA were transfected from each pool together with 0.5 µg of pAdVAntage plasmid (Promega, Madison, WI) into HEK-293 cells (300,000 cells per well of a six-well plate) using DOTAP lipofection reagent. After transfection (18 h) the cells were allowed to grow for 30 h, and the media were then replaced with medium containing 100 nM unlabeled E1 + 200,000 cpm of 3H-labeled E1/ml. After additional incubation for 18 h, the media were collected, frozen in dry ice, and kept at -20 C until the steroids were extracted and analyzed as previously described (45).

Altogether, 243 pools, each containing cDNAs from 2500 colonies, were screened. Transfection of two plasmid pools resulted in the conversion of E1 to E2, which was 2–2.5 times higher than the conversion obtained in mock-transfected cells. The DNA from the two positive pools was transformed into Epicurian coli XL1-Blue MRF' supercompetent cells (Stratagene, La Jolla, CA) to generate 20 pools of approximately 1000 transformants from both original positive pools. Plasmid DNA samples isolated from these 40 pools were then transfected into HEK-293 cells, and conversion of E1 to E2 was again analyzed. One positive pool from both original lines was further subfractionated into 20 pools of 100 transformants and analyzed. At the final stage, a positive DNA pool (containing 100 cDNA clones) was transformed into the supercompetent cells, and 100 single colonies were picked up for plasmid isolation on a miniprep scale to identify a pure positive clone. The two enriched clones originating from the two independent positive pools were named m17HSD7.1 and m17HSD7.2. The nucleotide sequence of the cDNAs was determined from both their strands using an automatic DNA sequencer (ABI Prism 377 DNA Sequencer, Perkin-Elmer, Foster City, CA). Fasta and PROSITE searches for the sequences, homology comparisons, and determination of the hydropathicity profile were carried out with a GCG package, version 8.1 (Genetics Computer Group, Madison, WI).

Isolation of RNA and Cloning of r17HSD7/PRAP cDNA
Total and poly(A)+-enriched RNAs were extracted using standard methods (55, 56). r17HSD7/PRAP cDNA was derived from total RNA isolated from the ovaries of pregnant rats (12 day) using SuperScriptII RT (GIBCO BRL, Gaithersburg, MD). The primer needed for generation of the antisense strand corresponded to nucleotides 1021–1003 in the PRAP cDNA (38) and the RT reaction was carried out at 42 C for 50 min. The RNA strand was next degraded by ribonuclease H (RNase H) (Pharmacia Biotech, Uppsala, Sweden) at 37 C for 15 min. r17HSD7/PRAP cDNA was then amplified with Pyrococcus furiosus polymerase (Stratagene) using primers containing the EcoRI site and corresponding to nucleotides from 9–27 and from 1021–1003 in the PRAP cDNA. The PCR consisted of denaturation at 94 C for 1 min, annealing for 1 min, and extension at 72 C for 2 min, the total number of cycles being 35. In the first five cycles, the annealing temperature was 57 C, and in the rest of the cycles 62 C. The PCR product was digested with EcoRI, cloned into CMV6 and Bluescript KS+ plasmids, and sequenced as mentioned above.

Measurement of 17HSD and 20HSD Activities in Cultured Cells and in Vitro
HC11 cells were seeded into six-well plates at a density of 230,000 cells per well. After 48 h, the media were replaced with medium containing 100 nM unlabeled substrate (E1, E2, A-dione, or T) + 200,000 cpm of 3H-labeled substrate per ml. After additional incubations from 2–24 h, the media were collected, frozen in dry ice, and maintained at -20 C. Steroids were extracted and separated from each other, and the conversion of substrate to product was determined as previously described (45).

17HSD/20HSD activity measurements after transient transfections with m17HSD1, m17HSD7, and r17HSD7/PRAP cDNAs under CMV-promoter were carried out in cultured HEK-293 cells. The cells were plated into six-well plates at a density of 300,000 cells per well 1 day before transfection. One half or 1 µg of plasmids was transfected using DOTAP. After transfection (18 h), the cells were allowed to grow for 30 h, and the medium was then replaced with one containing 100 nM or 1 nM unlabeled substrate + 200,000 cpm of 3H-labeled substrate/ml. After incubations from 2–24 h, the media were collected and stored, and steroids were analyzed as above.

For activity measurements of m17HSD7 in vitro, HEK-293 cells were transfected with m17HSD7.1-cDNA3.1 plasmids, after which they were collected and lysed by sonication in 10 mM potassium phosphate buffer, pH 8, containing 1 mM EDTA, 20% glycerol, 0.08% N,N-bis-(3-D-gluconamidopropyl)cholamide (Big Chap, Calbiochem-Novabiochem International, San Diego, Ca), 0.05% BSA, 0.02% NaN3, and PMSF, 87 ng/ml. Cell debris was removed by centrifugation at 1500 x g for 10 min, and the supernatant was stored at -70 C. For determination of Km and Vmax values, cell extracts were diluted in 10 mM potassium phosphate buffer, pH 8, containing 1 mM EDTA and 0.05% BSA, and the samples were then mixed with a substrate (0.1, 1, and 5 µM E1, and 3.3, 5, and 10 µM E2). When the conversion rates for A-dione, T, A-diol, DHEA, 20-OHP, and P were measured, a 0.5 µM substrate concentration was used. The enzymatic reactions were started by adding a cofactor (NADP+/NADPH, Boehringer Mannheim) to a final concentration of 1 mM, and the samples were incubated for an appropriate time at 37 C. The reactions were stopped by freezing the reaction mixture quickly in an ethanol-dry ice bath. Steroids were extracted and separated from each other, and the amount of substrate formed per min and per total amount of protein was calculated. Kinetic parameters (Km and Vmax) were calculated by using a GraFit-program (Erithacus Software Ltd., Staines, UK). The program fits data to the Michaelis-Menten equation using nonlinear regression analysis. The values presented represent the average ± SD of three independent experiments.

Northern Blotting and RT-PCR Analyses
Total or poly(A)+-enriched RNAs were subjected to electrophoresis in 1% (wt/vol) agarose-formaldehyde gel (56), which was blotted with a positively charged nylon membrane (Boehringer Mannheim). The fixed membranes were hybridized with a 32P-labeled fragment of either m17HSD7 cDNA (corresponding to nucleotides 33–849 in the cDNA, Fig. 2Go) or r17HSD7/PRAP cDNA [nucleotides 9–897 in the cDNA (38)]. Hybridization was performed at 42 C in 5 x NaCl/Pi,/EDTA-buffer (0.75 M NaCl, 50 mM NaH2PO4, 5 mM EDTA, pH 7.0), containing 50% formamide, 1% BSA, 1% Ficoll 400, 1% polyvinyl pyrrolidone, 0.5% SDS, and 100 µg salmon sperm DNA/ml. To control the amount of RNA applied to the gels, the membranes were also hybridized with a rat GAPDH cDNA probe.

RT-PCR analysis was run as previously described (10) using a Thermostable recombinant Thermus thermophilus RNA RT-PCR kit (Perkin Elmer-Roche Molecular Systems, Inc., Branchburg, NJ). The primers corresponded to nucleotides from 351–370 and from 849–831 in m17HSD7 cDNA and nucleotides from 54–75 and from 912–849 in mouse GAPDH cDNA (57). After PCR, 80 µl aliquots of the PCR mixtures were subjected to electrophoresis in 1.5% agarose gel and blotted with a nylon membrane. Hybridization of the membrane was performed similarly to Northern blot analysis, with m17HSD7 or rat GAPDH cDNA probes.


    ACKNOWLEDGMENTS
 
We thank Ms. Helmi Konola and Mrs. Eeva Holopainen for their skillful technical assistance. We are also grateful to Dr. Bernd Groner for providing the HC11 cell line and Dr. Lateef Akinola for preparing the filters containing RNA samples from rat tissues.


    FOOTNOTES
 
Address requests for reprints to: Pirkko Vihko, Professor, Biocenter Oulu and WHO Collaborating Centre for Research on Reproductive Health, University of Oulu, FIN-90220, Oulu, Finland.

This work was supported by the Research Council for Health of the Academy of Finland (project no. 3314 and 40990). The World Health Organization Collaborating Centre for Research on Reproductive Health is supported by the Ministries of Education, Social Affairs and Health, and Foreign Affairs, Finland.

Received for publication January 30, 1998. Revision received March 19, 1998. Accepted for publication March 19, 1998.


    REFERENCES
 TOP
 ABSTRACT
 INTRODUCTION
 RESULTS
 DISCUSSION
 MATERIALS AND METHODS
 REFERENCES
 

  1. Poutanen M, Isomaa V, Peltoketo H, Vihko R 1995 Role of 17ß-hydroxysteroid dehydrogenase type 1 in endocrine and intracrine estradiol biosynthesis. J Steroid Biochem Mol Biol 55:525–532[CrossRef][Medline]
  2. Labrie F, Luu-The V, Lin S-X, Labrie C, Simard J, Breton R, Bélanger A 1997 The key role of 17ß-hydroxysteroid dehydrogenases in sex steroid biology. Steroids 62:148–158[CrossRef][Medline]
  3. Mustonen MVJ, Poutanen MH, Kellokumpu S, de Launoit Y, Isomaa VV, Vihko RK, Vihko PT 1998 Mouse 17ß-hydroxysteroid dehydrogenase type 2 mRNA is predominantly expressed in hepatocytes and in surface epithelial cells of the gastrointestinal and urinary tracts. J Mol Endocrinol 20:67–74[Abstract/Free Full Text]
  4. Mustonen M, Poutanen M, Chotteau-Lelievre A, de Launoit Y, Isomaa V, Vainio S, Vihko R, Vihko P 1997 Ontogeny of 17ß-hydroxysteroid dehydrogenase type 2 mRNA expression in the developing mouse placenta and fetus. Mol Cell Endocrinol 134:33–40[CrossRef][Medline]
  5. Peltoketo H, Isomaa V, Mäentausta O, Vihko R 1988 Complete amino acid sequence of human placental 17ß-hydroxysteroid dehydrogenase deduced from cDNA. FEBS Lett 239:73–77[CrossRef][Medline]
  6. Luu The V, Labrie C, Zhao HF, Couët J, Lachance Y, Simard J, Leblanc G, Côté J, Bérubé D, Gagné R, Labrie F 1989 Characterization of cDNAs for human estradiol 17ß-dehydrogenase and assignment of the gene to chromosome 17: evidence of two mRNA species with distinct 5'-termini in human placenta. Mol Endocrinol 3:1301–1309[Abstract]
  7. Wu L, Einstein M, Geissler WM, Chan HK, Elliston KO, Andersson S 1993 Expression cloning and characterization of human 17ß-hydroxysteroid dehydrogenase type 2, a microsomal enzyme possessing 20{alpha}-hydroxysteroid dehydrogenase activity. J Biol Chem 268:12964–12969[Abstract/Free Full Text]
  8. Geissler WM, Davis DL, Wu L, Bradshaw KD, Patel S, Mendonca BB, Elliston KO, Wilson JD, Russell DW, Andersson S 1994 Male pseudohermaphroditism caused by mutations of testicular 17ß-hydroxysteroid dehydrogenase 3. Nat Genet 7:34–39[Medline]
  9. Adamski J, Normand T, Leenders F, Monté D, Begue A, Stéhelin D, Jungblut PW, de Launoit Y 1995 Molecular cloning of a novel widely expressed human 80 kDa 17ß-hydroxysteroid dehydrogenase IV. Biochem J 311:437–443[Medline]
  10. Nokelainen P, Puranen T, Peltoketo H, Orava M, Vihko P, Vihko R 1996 Molecular cloning of mouse 17ß-hydroxy-steroid dehydrogenase type 1 and characterization of enzyme activity. Eur J Biochem 236:482–490[Abstract]
  11. Mustonen MVJ, Poutanen MH, Isomaa VV, Vihko PT, Vihko RK 1997 Cloning of mouse 17ß-hydroxysteroid dehydrogenase type 2, and analysing expression of the mRNAs for types 1, 2, 3, 4 and 5 in mouse embryos and adult tissues. Biochem J 325:199–205[Medline]
  12. Sha JA, Dudley K, Rajapaksha WRAKJS, O’ShaughnessyPJ 1997 Sequence of mouse 17ß-hydroxysteroid dehydrogenase type 3 cDNA and tissue distribution of the type 1 and type 3 isoform mRNAs. J Steroid Biochem Mol Biol 60:19–24[CrossRef][Medline]
  13. Normand T, Husen B, Leenders F, Pelczar H, Baert J-L, Begue A, Flourens A-C, Adamski J, de Launoit Y 1995 Molecular characterization of mouse 17ß-hydroxysteroid dehydrogenase IV. J Steroid Biochem Mol Biol 55:541–548[CrossRef][Medline]
  14. Deyashiki Y, Ohshima K, Nakanishi M, Sato K, Matsuura K, Hara A 1995 Molecular cloning and characterization of mouse estradiol 17ß-dehydrogenase (A-specific), a member of the aldoketoreductase family. J Biol Chem 270:10461–10467[Abstract/Free Full Text]
  15. Ghersevich S, Nokelainen P, Poutanen M, Orava M, Autio-Harmainen H, Rajaniemi H, Vihko R 1994 Rat 17ß-hydroxysteroid dehydrogenase type 1: primary structure and regulation of enzyme expression in rat ovary by diethylstilbestrol and gonadotropins in vivo. Endocrinology 135:1477–1487[Abstract]
  16. Akinola LA, Poutanen M, Vihko R 1996 Cloning of rat 17ß-hydroxysteroid dehydrogenase type 2 and characterization of tissue distribution and catalytic activity of rat type 1 and type 2 enzymes. Endocrinology 137:1572–1579[Abstract]
  17. Qin Y-M, Poutanen MH, Helander HM, Kvist A-P, Siivari KM, Schmitz W, Conzelmann E, Hellman U, Hiltunen JK 1997 Peroxisomal multifunctional enzyme of ß-oxidation metabolizing D-3-hydroxyacyl-CoA esters in rat liver: molecular cloning, expression and characterization. Biochem J 321:21–28[Medline]
  18. Biswas MG, Russell DW 1997 Expression cloning and characterization of oxidative 17ß- and 3{alpha}-hydroxysteroid dehydrogenases from rat and human prostate. J Biol Chem 272:15959–15966[Abstract/Free Full Text]
  19. Sawetawan C, Milewich L, Word RA, Carr BR, Rainey WE 1994 Compartmentalization of type I 17ß-hydroxysteroid oxidoreductase in the human ovary. Mol Cell Endocrinol 99:161–168[CrossRef][Medline]
  20. Ghersevich SA, Poutanen MH, Martikainen HK, Vihko RK 1994 Expression of 17ß-hydroxysteroid dehydrogenase in human granulosa cells: correlation with follicular size, cytochrome P450 aromatase activity and oestradiol production. J Endocrinol 143:139–150[Abstract]
  21. Ghersevich SA, Poutanen MH, Rajaniemi HJ, Vihko RK 1994 Expression of 17ß-hydroxysteroid dehydrogenase in the rat ovary during follicular development and luteinization induced with pregnant mare serum gonadotrophin and human chorionic gonadotrophin. J Endocrinol 140:409–417[Abstract]
  22. Tremblay Y, Ringler GE, Morel Y, Mohandas TK, Labrie F, Strauss III JF, Miller WL 1989 Regulation of the gene for estrogenic 17-ketosteroid reductase lying on chromosome 17cen->q25. J Biol Chem 264:20458–20462[Abstract/Free Full Text]
  23. Fournet-Dulguerov N, MacLusky NJ, Leranth CZ, Todd R, Mendelson CR, Simpson ER, Naftolin F 1987 Immunohistochemical localization of aromatase cytocrome P-450 and estradiol dehydrogenase in the syncytiotrophoblast of the human placenta. J Clin Endocrinol Metab 65:757–764[Abstract]
  24. Mäentausta O, Sormunen R, Isomaa V, Lehto V-P, Jouppila P, Vihko R 1991 Immunohistochemical localization of 17ß-hydroxysteroid dehydrogenase in the human endometrium during the menstrual cycle. Lab Invest 65:582–587[Medline]
  25. Poutanen M, Isomaa V, Lehto V-P, Vihko R 1992 Immunological analysis of 17ß-hydroxysteroid dehydrogenase in benign and malignant human breast tissue. Int J Cancer 50:386–390[Medline]
  26. Mäentausta O, Boman K, Isomaa V, Stendahl U, Bäckström T, Vihko R 1992 Immunohistochemical study of the human 17ß-hydroxysteroid dehydrogenase and steroid receptors in endometrial adenocarcinoma. Cancer 70:1551–1555[Medline]
  27. Poutanen M, Miettinen M, Vihko R 1993 Differential estrogen substrate specificities for transiently expressed human placental 17ß-hydroxysteroid dehydrogenase and an endogenous enzyme expressed in cultured COS-m6 cells. Endocrinology 133:2639–2644[Abstract]
  28. Puranen T, Poutanen M, Ghosh D, Vihko R, Vihko P 1997 Origin of substrate specificity of human and rat 17ß-hydroxysteroid dehydrogenase type 1, using chimeric enzymes and site-directed substitutions. Endocrinology 138:3532–3539[Abstract/Free Full Text]
  29. Miettinen MM, Mustonen MVJ, Poutanen MH, Isomaa VV, Vihko RK 1996 Human 17ß-hydroxysteroid dehydrogenase type 1 and type 2 isoenzymes have opposite activities in cultured cells and characteristic cell- and tissue-specific expression. Biochem J 314:839–845[Medline]
  30. Casey ML, MacDonald PC, Andersson S 1994 17ß-Hydroxysteroid dehydrogenase type 2: chromosomal assignment and progestin regulation of gene expression in human endometrium. J Clin Invest 94:2135–2141[Medline]
  31. Elo JP, Akinola LA, Poutanen M, Vihko P, Kyllönen AP, Lukkarinen O, Vihko R 1996 Characterization of 17ß-hydroxysteroid dehydrogenase isoenzyme expression in benign and malignant human prostate. Int J Cancer 66:37–41[CrossRef][Medline]
  32. Dieuaide-Noubhani M, Novikov D, Baumgart E, Vanhooren JCT, Fransen M, Goethals M, Vandekerckhove J, van Veldhoven PP, Mannaerts GP 1996 Further characterization of the peroxisomal 3-hydroxyacyl-CoA dehydrogenases from rat liver. Relationship between the different dehydrogenases and evidence that fatty acids and the C27 bile acids di- and tri-hydroxycoprostanic acids are metabolized by separate multifunctional proteins. Eur J Biochem 240:660–666[Abstract]
  33. Corton JC, Bocos C, Moreno ES, Merritt A, Marsman DS, Sausen PJ, Cattley RC, Gustafsson JÅ 1996 Rat 17ß-hydroxysteroid dehydrogenase type IV is a novel peroxisome proliferator-inducible gene. Mol Pharmacol 50:1157–1166[Abstract]
  34. Simpson ER, Mahendroo MS, Means GD, Kilgore MW, Hinshelwood MM, Graham-Lorence S, Amarneh B, Ito Y, Fisher CR, Michael MD, Mendelson CR, Bulun SE 1994 Aromatase cytochrome P450, the enzyme responsible for estrogen biosynthesis. Endocr Rev 15:342–355[Medline]
  35. Gibori G 1993 The corpus luteum of pregnancy. In: Adashi EY, Leung PCK (eds) The Ovary. Raven Press, New York, pp 261–317
  36. Akinola LA, Poutanen M, Vihko R, Vihko P 1997 Expression of 17ß-hydroxysteroid dehydrogenase type 1 and type 2, P450 aromatase, and 20{alpha}-hydroxysteroid dehydrogenase enzymes in immature, mature, and pregnant rats. Endocrinology 138:2886–2892[Abstract/Free Full Text]
  37. Taya K, Greenwald GS 1981 In vivo and in vitro ovarian steroidogenesis in the pregnant rat. Biol Reprod 25:683–691[Medline]
  38. Duan WR, Linzer DIH, Gibori G 1996 Cloning and characterization of an ovarian-specific protein that associates with the short form of the prolactin receptor. J Biol Chem 271:15602–15607[Abstract/Free Full Text]
  39. Ball RK, Friis RR, Schoenenberger CA, Doppler W, Groner B 1988 Prolactin regulation of ß-casein gene expression and of a cytosolic 120-kd protein in a cloned mouse mammary epithelial cell line. EMBO J 7:2089–2095[Abstract]
  40. Jörnvall H, Persson B, Krook M, Atrian S, Gonzàlez-Duarte R, Jeffery J, Ghosh D 1995 Short-chain dehydrogenases/reductases (SDR). Biochemistry 34:6003–6013[Medline]
  41. Ghosh D, Pletnev VZ, Zhu D-W, Wawrzak Z, Duax WL, Pangborn W, Labrie F, Lin S-X 1995 Structure of human estrogenic 17ß-hydroxysteroid dehydrogenase at 2.20 Å resolution. Structure 3:503–513[Medline]
  42. Parmer TG, McLean MP, Duan WR, Nelson SE, Albarracin CT, Khan I, Gibori G 1992 Hormonal and immunological characterization of the 32 kilodalton ovarian-specific protein. Endocrinology 131:2213–2221[Abstract]
  43. Martel C, Rhéaume E, Takahashi M, Trudel C, Couët J, Luu-The V, Simard J, Labrie F 1992 Distribution of 17ß-hydroxysteroid dehydrogenase gene expression and activity in rat and human tissues. J Steroid Biochem Mol Biol 41:597–603[CrossRef][Medline]
  44. Milewich L, Garcia RL, Gerrity LW 1985 17ß-Hydroxy-steroid oxidoreductase: a ubiquitous enzyme. Interconversion of estrone and estradiol-17ß in BALB/c mouse tissues. Metabolism 34:938–944[Medline]
  45. Puranen T, Poutanen M, Ghosh D, Vihko P, Vihko R 1997 Characterization of structural and functional properties of human 17ß-hydroxysteroid dehydrogenase type 1 using recombinant enzymes and site-directed mutagenesis. Mol Endocrinol 11:77–86[Abstract/Free Full Text]
  46. Luu-The V, Zhang Y, Poirier D, Labrie F 1995 Characteristics of human types 1, 2 and 3 17ß-hydroxysteroid dehydrogenase activities: oxidation/reduction and inhibition. J Steroid Biochem Mol Biol 55:581–587[CrossRef][Medline]
  47. Duan WR, Parmer TG, Albarracin CT, Zhong L, Gibori G 1997 PRAP, a prolactin receptor associated protein: its gene expression and regulation in the corpus luteum. Endocrinology 138:3216–3221[Abstract/Free Full Text]
  48. Hickey GJ, Chen S, Besman MJ, Shively JE, Hall PF, Gaddy-Kurten D, Richards JS 1988 Hormonal regulation, tissue distribution, and content of aromatase cytochrome P450 messenger ribonucleic acid and enzyme in rat ovarian follicles and corpora lutea: relationship to estradiol biosynthesis. Endocrinology 122:1426–1436[Abstract]
  49. Telleria CM, Parmer TG, Zhong L, Clarke DL, Albarracin CT, Duan WR, Linzer DIH, Gibori G 1997 The different forms of the prolactin receptor in the rat corpus luteum: developmental expression and hormonal regulation in pregnancy. Endocrinology 138:4812–4820[Abstract/Free Full Text]
  50. Gouilleux F, Wakao H, Mundt M, Groner B 1994 Prolactin induces phosphorylation of Tyr694 of Stat 5 (MGF), a prerequisite for DNA binding and induction of transcription. EMBO J 13:4361–4369[Abstract]
  51. O’Neal KD, Yu-Lee L 1994 Differential signal transduction of the short, Nb2, and long prolactin receptors. J Biol Chem 269:26076–26082[Abstract/Free Full Text]
  52. Miettinen MM, Poutanen MH, Vihko RK 1996 Characterization of estrogen-dependent growth of cultured MCF-7 human breast-cancer cells expressing 17ß-hydroxy-steroid dehydrogenase type 1. Int J Cancer 68:600–604[CrossRef][Medline]
  53. Rembiesa R, Marchut M, Warchol A 1971 Formation and metabolism of progesterone by the mouse placenta in vitro. J Steroid Biochem 2:111–119[CrossRef]
  54. Blomquist CH, Hensleigh HC, Block DL, Feeney LA 1993 Placental 17ß-hydroxysteroid oxidoreductase, lactate dehydrogenase and malate dehydrogenase during the latter half of pregnancy in the mouse. J Steroid Biochem Mol Biol 46:61–67[CrossRef][Medline]
  55. Davis LG, Dibner MD, Battey JF (eds) 1986 Preparation and analysis of RNA from eukaryotic cells. In: Basic Methods in Molecular Biology. Elsevier, New York, pp 130–135
  56. Sambrook J, Fritsch EF, Maniatis T (eds) 1989 Extraction, purification, and analysis of messenger RNA from eukaryotic cells. In: Molecular Cloning: A Laboratory Manual. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY, pp 7.1–7.84
  57. Sabath DE, Broome HE, Prystowsky MB 1990 Glyceraldehyde-3-phosphate dehydrogenase mRNA is a major interleukin 2-induced transcript in a cloned T-helper lymphocyte. Gene 91:185–191[CrossRef][Medline]