cDNA Cloning, Expression, and Mutagenesis Study of Liver-type Prostaglandin F Synthase*

Toshiko SuzukiDagger §, Yutaka Fujii, Masashi Miyanoparallel , Lan-Ying Chen**, Tomohiro TakahashiDagger Dagger , and Kikuko WatanabeDagger §§

From the Dagger  Second Department, Osaka Bioscience Institute, 6-2-4 Furuedai, Suita, Osaka 565-0874,  Fukui Medical School, 23-3 Shimoaizuki, Matsuoka-cho, Yoshida, Fukui 910-1103, parallel  Central Pharmaceutical Research Institute, Japan Tobacco, Inc., 1-1 Murasaki-cho, Takatsuki, Osaka 569-1125, Japan, the ** Department of Biochemistry, Cardiovascular Institute and Fu-Wai Hospital, Beijing 100037, China, andDagger Dagger  Bioscience Research Laboratory, Mochida Pharmaceutical Co., 1-1-1Kamiya, Kita, Tokyo 115-0043, Japan

    ABSTRACT
TOP
ABSTRACT
INTRODUCTION
REFERENCES

Prostaglandin (PG) F synthase catalyzes the reduction of PGD2 to 9alpha ,11beta -PGF2 and that of PGH2 to PGF2alpha on the same molecule. PGF synthase has at least two isoforms, the lung-type enzyme (Km value of 120 µM for PGD2 (Watanabe, K., Yoshida, R., Shimizu, T., and Hayaishi, O. (1985) J. Biol. Chem. 260, 7035-7041) and the liver-type one (Km value of 10 µM for PGD2 (Chen, L. -Y., Watanabe, K., and Hayaishi, O. (1992) Arch. Biochem. Biophys. 296, 17-26)). The liver-type enzyme was presently found to consist of a 969-base pair open reading frame coding for a 323-amino acid polypeptide with a Mr of 36,742. Sequence analysis indicated that the bovine liver PGF synthase had 87, 79, 77, and 76% identity with the bovine lung PGF synthase and human liver dihydrodiol dehydrogenase (DD) isozymes DD1, DD2, and DD4, respectively. Moreover, the amino acid sequence of the liver-type PGF synthase was identical with that of bovine liver DD3. The liver-type PGF synthase was expressed in COS-7 cells, and its recombinant enzyme had almost the same properties as the native enzyme. Furthermore, to investigate the nature of catalysis and/or substrate binding of PGF synthase, we constructed and characterized various mutant enzymes as follows: R27E, R91Q, H170C, R223L, K225S, S301R, and N306Y. Although the reductase activities toward PGH2 and phenanthrenequinone (PQ) of almost all mutants were not inactivated, the Km values of R27E, R91Q, H170C, R223L, and N306Y for PGD2 were increased from 15 to 110, 145, 75, 180, and 100 µM, respectively, indicating that Arg27, Arg91, His170, Arg223, and Asn306 are essential to give a low Km value for PGD2 of the liver-type PGF synthase and that these amino acid residues serve in the binding of PGD2. Moreover, the R223L mutant among these seven mutants especially has a profound effect on kcat for PGD2 reduction. The Km values of R223L, K225S, and S301R for PQ were about 2-10-fold lower than the wild-type value, indicating that the amino acid residues at 223, 225 and 301 serve in the binding of PQ to the enzyme. On the other hand, the Km value of H170C for PGH2 was 8-fold lower than that of the wild type, indicating that the amino acid residue at 170 is related to the binding of PGH2 to the enzyme and that Cys170 confer high affinity for PGH2. Additionally, the 5-fold increase in kcat/Km value of the N306Y mutant for PGH2 compared with the wild-type value suggests that the amino acid at 306 plays an important role in catalytic efficiency for PGH2.

    INTRODUCTION
TOP
ABSTRACT
INTRODUCTION
REFERENCES

F series prostaglandins (PG)1 are widely distributed in various organs of mammals and exhibit a variety of biological activities including constriction of pulmonary arteries (1, 2). PGF2 is synthesized from PGE2, PGD2, or PGH2 by PGE 9-ketoreductase, PGD 11-ketoreductase, or PGH 9,11-endoperoxide reductase, respectively. PGF synthase (EC 1.1.1.188) was purified from bovine lung by Watanabe et al. (3). It forms 9alpha ,11beta -PGF2 (4) from PGD2 (PGD2 11-ketoreductase activity) and PGF2alpha from PGH2 (PGH2 9,11-endoperoxide reductase activity) on the same molecule in the presence of NADPH (3, 4). This enzyme catalyzes the reduction of other carbonyl compounds including 9,10-phenanthrenequinone (PQ) as well as that of PGD2 and PGH2 but does not catalyze the reduction of PGE2. Although PGD2 11-ketoreductase activity was competitively inhibited by PQ, PGH2 9,11-endoperoxide reductase activity was not inhibited by PGD2 or PQ (3). PGF synthase belongs to the aldo-keto reductase family. The bovine lung PGF synthase is a monomeric protein with a Mr of 36,666 consisting of 323 amino acids, and its amino acid sequence shows high homology compared with that of other aldo-keto reductase family members (5). PGF synthase has two isozymes, one in the lung (3) and the other one in the liver (6), with different Km values for PGD2 (120 and 10 µM, respectively). The regulation by metals, the sensitivity to chloride ions, the inhibition by CuSO4 and HgCl2, and the profile of immuno-precipitation with anti-bovine lung PGF synthase antibody are different between the two isozymes (6). Although Kuchinke et al. (7) isolated a clone (PGFS II) of PGF synthase from bovine liver and determined its amino acid sequence, the 99% similarity with the amino acid sequence of lung PGF synthase and the high Km value for PGD2 of this recombinant PGFS II indicated that its cDNA was that of the lung-type enzyme even though it had been isolated from liver. Until now, the primary structure of the liver-type PGF synthase and the amino acids related to the affinity for PGD2 have not yet been defined.

Dihydrodiol dehydrogenase (DD, EC 1.3.1.20) catalyzes the NADP-linked oxidation of trans-dihydrodiols of aromatic hydrocarbons to the corresponding catechols and is distributed in various mammalian tissues. DD, also belonging to the aldo-keto reductase superfamily, has been purified from various animal tissues, i.e. human liver (8), rat liver (9, 10), rabbit liver (11), mouse liver (12), bovine liver (13), guinea pig testis (14), and so forth. Human liver DD exists in at least four multiple forms (DD1-DD4) with similar mass of about 36 kDa (8). Human DD3 was identified as an aldehyde reductase, and the other three forms exhibited 3alpha -hydroxysteroid dehydrogenase (HSD, EC 1.1.1.213) activity. Bovine liver DD also has three multiple forms (DD1-DD3), namely DD1 (3alpha -HSD), DD2 (high Km aldehyde reductase), and DD3 (dihydrodiol-specific enzyme) (13). Human liver DD1, DD2, and DD3 are not identical with bovine liver DD1, DD2, and DD3, respectively, on the basis of their enzymatic properties including the substrate specificity. The amino acid sequence of the bovine lung-type PGF synthase (3, 7) showed an identity of 81, 79, 78, and 87% with that of human liver DD1, DD2, and DD4 (15, 16) and bovine liver DD3 (17). Among human liver DDs, DD1 and DD2 exhibited PGF synthase activity with Km values of 12 and 79 µM, respectively, for PGD2, but this activity of DD4 was not detected (16). Moreover, PGF synthase activity of bovine liver DD3 has not yet been reported.

In the present study, we describe the primary structure of the bovine liver-type PGF synthase and the enzymatic properties of the recombinant enzyme in COS-7 cells. Based on the comparison of the amino acid sequences among the liver-type and the lung-type PGF synthases and human liver DDs, several mutants were constructed, and their enzymatic properties were examined. The results of mutagenesis indicated the amino acid residues related to the binding sites of PGD2, PGH2, and PQ.

    EXPERIMENTAL PROCEDURES

Materials-- [5,6,8,9,12,14,15-3H]PGD2 (3.7 TBq/mmol) was obtained from NEN Life Science Products. [1-14C]PGH2 was prepared as described previously (18), with acetone powder of sheep vesicular gland microsomes (Ran Biochemicals, Tel Aviv, Israel) used as a source of PG endoperoxide synthase. Authentic PGs were kindly donated by Ono Pharmaceutical Co. pEF-BOS mammalian expression vector was a generous gift from Dr. S. Nagata. Other materials and commercial sources were as follows: NADP, NADPH, glucose 6-phosphate, and glucose-6-phosphate dehydrogenase from bakers' yeast (type IX), from Sigma (Japan); Red Sepharose and NAP10, from Amersham Pharmacia Biotech (UK); precoated silica gel glass plates (F254), from Merck (Germany). Other chemicals were at least of reagent grade.

Internal Amino Acid Sequences of the Liver-type PGF Synthase-- The bovine liver PGF synthase was purified to apparent homogeneity as described previously (6). S-Carboxymethylation and lysyl-endopeptidase digestion of the purified enzyme and the purification of fragments digested by lysyl-endopeptidase were done as described previously (5). The purified peptide fragments were sequenced by automated Edman degradation with an Applied Biosystems Inc. sequencer (Perkin-Elmer).

cDNA Cloning of Bovine Liver PGF Synthase-- Total RNA was prepared from bovine liver by the acid guanidinium thiocyanate/phenol/chloroform extraction method (19). Poly(A)+ RNA was prepared by use of an mRNA Purification Kit (Amersham Pharmacia Biotech, UK) according to the manufacturer's manual. A (dT)12-18 (Life Technologies, Inc.) primed cDNA was synthesized from 2 µg of poly(A)+ RNA by SuperScript II reverse transcriptase (Life Technologies, Inc.).

Degenerative reverse transcriptase-polymerase chain reaction (PCR) using mixed oligonucleotide primers was performed to obtain a partial cDNA fragment for screening of a bovine liver cDNA library. Mixed oligonucleotide primers were designed according to the amino acid sequences of peptides I, II, III, and IV shown in Fig. 1 and to the nucleotide sequence of the lung-type PGF synthase (5). Each primer was synthesized with an Applied Biosystems Inc. DNA synthesizer (Perkin-Elmer), and two sets of the sequences of sense and antisense primers were used as follows: (i) 5'-TTCCGCCATAT(CTA)GACAGTGCT-3' (corresponding to peptide I, 21-mers) named P1 and 5'-GTCAAACACCTG(TGA)ATATTCTC-3' (corresponding to peptide III, 21-mers) named P2; (ii) 5'-GTGTCCAACTTCAACCACAAG-3' (corresponding to peptide II, 21-mers) named P3 and 5'-ATATTCTTC(AGCT)ACAAATGGGTA-3' (corresponding to peptide IV, 21-mers) named P4. The reverse transcriptase-PCR was conducted with mRNA used as a template and primers P1/P2 for one PCR and P3/P4 for the other one under the conditions of denaturation at 98 °C for 15 s, annealing at 65 °C for 30 s, and elongation at 74 °C for 30 s by KOD polymerase (Toyobo, Japan). After 20 cycles of PCR, the products were ethanol-precipitated and separated on 1% agarose gel. Two bands were recovered from the gel by use of a QIAEX II Gel Extraction Kit (Qiagen, Netherlands). Each band was ligated to a BlueScript SK II (+) vector (Toyobo, Japan) by use of a ligation kit (Takara, Japan), and the resulting constructs were transfected into Escherichia coli DH10B competent cells (Life Technologies, Inc.). Plasmids were purified with a Qiagen plasmid purification kit and sequenced with an Applied Biosystems Inc. automated DNA sequencer 373A (Perkin-Elmer). Two bands, one of 720 base pairs (P1/P2) and one of 477 base pairs (P3/P4), encoded the internal cDNA and were used as two different probes for screening of the library.

The bovine liver cDNA library was constructed from 5 µg of poly(A)+ RNA with SuperScriptTM Plasmid System (Life Technologies, Inc.) for cDNA synthesis and a Plasmid Cloning Kit (Life Technologies, Inc.) according to the manufacturer's manual. The resulting constructs were transformed into ElectroMax, DH12S competent cells (Life Technologies, Inc.) by the electroporation method using a Gene-Pulser (Bio-Rad). The library yielded 2.0 × 105 independent clones. Full-length cDNA clones were obtained by the colony hybridization method. All clones were spread on 20 sheets of nylon filters for the master filter, and then two filters were replicated from each master filter. The replicate filters were alkaline-denatured and fixed by baking at 80 °C for 2 h. Two reverse transcriptase-PCR products, 720 and 477 base pairs described above, were randomly labeled by [alpha -32P]dCTP (111 TBq/mmol, Amersham Pharmacia Biotech, UK) with a Megaprime random primer labeling kit (Amersham Pharmacia Biotech, UK) and used as two probes for hybridization. After hybridization in 5× SSCP, 0.1% SDS, 100 µg salmon sperm DNA, and 10× Denhardt's solution at 65 °C for overnight, each filter was washed extensively twice in 2× SSC, 0.1% SDS at room temperature for 5 min and twice in 0.5× SSC, 0.1% SDS at 60 °C for 30 min. Thirty two double-positive clones against the two different probes were obtained from 2.0 × 105 independent clones. Six clones were picked up and sequenced with an Applied Biosystems Inc. automated DNA sequencer 373A. All clones coded for full-length cDNAs of bovine liver PGF synthase. One of these six clones was named pSPORT-BLiFS27.

Northern Blot Analysis-- For Northern blot analysis, total RNA (20 µg), which was isolated from bovine liver with a total RNA purification kit (Amersham Pharmacia Biotech, UK) according to the manufacturer's manual, was separated on 1.0% agarose gel and transferred to nylon membranes. The 477-base pair fragment (P3/P4 described above) was labeled with a BcaBEST Labeling Kit (Takara, Japan) according to the manufacturer's manual using [alpha -32P]dCTP and was used as a probe. The membranes were prehybridized in Rapid Hybridization buffer (Amersham Pharmacia Biotech, UK) at 65 °C for 1 h and then hybridized for 2.5 h after addition of the radiolabeled probe. The membranes were washed for 5 min once at room temperature in 2× SSC, 0.1% SDS, twice for 30 min at 65 °C in 0.5× SSC, 0.1% SDS, and twice for 30 min at 65 °C in 0.2× SSC, 0.1% SDS. Autoradiograms were examined with a BAS-2000 system analyzer (Fuji Film, Japan).

Expression of Bovine Liver PGF Synthase in COS-7 Cells and Purification of Its Expressed Protein-- The bovine liver cDNA insert was removed from pSPORT-BLiFS27 by digestion with EcoRI and AflII, and its ends were blunted. The blunt-ended insert containing the complete coding region was subcloned into the blunt-ended BstXI sites of the pEF-BOS mammalian expression vector (20). Monkey COS-7 cells were cultured in Dulbecco's modified Eagle's medium (Nissui Co., Tokyo) containing 10% fetal calf serum. COS-7 cells (5 × 106 cells) were transfected with 20 µg of plasmid DNA by the electroporation method using a Gene Pulser (Bio-Rad). These cells were incubated in a 5% CO2 incubator at 37 °C for 72 h. The transfected COS-7 cells were sonicated in 3 volumes of 10 mM potassium phosphate buffer (KPB) (pH 7.0). The recombinant enzyme was purified by the method of Chen et al. (6) with a minor modification. The cytosol fraction of the homogenated cells, which was centrifuged at 100,000 × g, was subjected to ammonium sulfate fractionation between 40 and 75% saturation. The precipitate formed was suspended in 500 µl of 10 mM KPB (pH 7.0) and desalted by passage through a NAP-10 column. The desalted sample was applied to a Red Sepharose column, and the enzyme was eluted with 10 mM KPB (pH 7.0) containing M KCl and 1 mM NADP. About 2.8-fold purification of the recombinant protein was achieved, and the apparent homogeneity was concluded following SDS-polyacrylamide gel electrophoresis (PAGE) and staining with Two-dimensional Silver Stain·II "Dai-ichi" (Dai-ichi Pure Chemicals Co., Ltd., Japan). A polyclonal antibody against PGF synthase was raised in a rabbit by the same procedure as described previously (3), with the enzyme purified from bovine liver used as the immunogen (6). For Western blot analysis, the purified enzyme was subjected to SDS-PAGE and electrophoretically transferred to a polyvinylidene difluoride membrane (Amersham Pharmacia Biotech, UK). Protein bands were immunostained with the anti-bovine liver PGF synthase antibody and reagents from a Vectastain ABC kit (Vector Laboratories) and visualized with an Enhanced Chemiluminescence Kit (ECLTM, Amersham Pharmacia Biotech, UK).

Enzyme Assay-- The PGD2 11-ketoreductase, PGH2 9,11-endoperoxide reductase, and PQ reductase activities of the recombinant protein were measured as described previously (3). The standard assay mixture for PGD2 11-ketoreductase contained 0.1 M KPB (pH 6.5), 0.5 mM NADP, 5 mM glucose 6-phosphate, glucose-6-phosphate dehydrogenase (1 unit), 1.5 mM [3H]PGD2 (3.7 KBq), and enzyme in a total volume of 50 µl. Incubation was carried out at 37 °C for 30 min. The PGH2 9,11-endoperoxide reductase activity was assayed under the same conditions as those of the PGD2 11-ketoreductase activity except that 40 µM [1-14C]PGH2 (4 MBq) was used as a substrate in place of 1.5 mM [3H]PGD2 and that the incubation time was 2 min. The PQ reductase activity was measured spectrophotometrically at 37 °C by following a decrease in absorbance at 340 nm in the assay mixture consisting of 0.1 M KPB (pH6.5), 80 µM NADPH, 10 µM PQ, and enzyme in a total volume of 0.5 ml. One unit of enzyme activity was defined as the amount that produced 1 µmol of PGF2 per min at 37 °C. Specific activity was expressed as the number of units/mg of protein. Protein was determined according to the method of Lowry et al. (21).

Site-directed Mutagenesis-- A mutagenesis study was performed by the method of Jones et al. (22). The mutagenesis primers for the mutant R27E were designed as follows: LiFSPN, 5'-CAAACAATGGATCC-3'; LuFSP1, 5'-TTGCACCTGAGGAGGTTCC-3'; LuFSP2, 5'-CCTCCTCAGGTGCAAAGGTT-3'; LiFSPC, 5'-GAATGCACGTGTACAGCT-3'. The bold letters of LuFSP1 and LuFSP2 were the sites of mutagenesis. As shown in Fig. 1, one PCR between LiFSPN and LuFSP2 and the other one between LuFSP1 and LiFSPC were conducted with pSPORT-BLiFS27 harboring the full-length cDNA for bovine liver PGF synthase as a template. The conditions of PCR were described above. After 20 cycles of PCR, the products were purified by 1% agarose gel electrophoresis, and the purified products were treated with Klenow fragment of DNA polymerase I after recovery. These products were mixed at the ratio of 1 to 1, and the second PCR was conducted using LiFSPN and LiFSPC as described above. The product of the second PCR was blunt-ended and then was ligated to the blunted BstXI sites of the pEF-BOS expression vector. Consequently, the mutant of R27E was formed. The other mutant enzymes were formed by the same procedure as used for the R27E mutant.

    RESULTS

cDNA Cloning of Bovine Liver PGF Synthase-- Screening of 2.0 × 105 clones with the two probes (P1/P2 and P3/P4 products described under "Experimental Procedures") gave 32 double-positive clones, and six independent clones of these 32 clones were picked up. DNA sequencing confirmed that these clones coded for full-length cDNAs of bovine liver PGF synthase. The deduced amino acid sequences of bovine liver PGF synthase contained all the amino acid sequences of the nine peptide fragments obtained from the native bovine liver enzyme (Fig. 1). The bovine liver PGF synthase cDNA clone BLiFS27 contained a polyadenylation signal after the stop codon (Fig. 1), showing that it coded for a full-length PGF synthase. BLiFS27 contained an open reading frame of 969 base pairs coding for 323 amino acids. The calculated Mr of the bovine liver enzyme was 36,742, a value similar to that of the native enzyme, which is about 36 kDa (6). The identity between the liver and lung enzymes was 87% at the amino acid level and 90% at the nucleotide level. As shown in Fig. 2, this enzyme showed a high identity in amino acid sequence with not only bovine lung PGF synthase (87%) but also human liver DD1 (79%), DD2 (77%), and DD4 (76%). Moreover, its sequence was identical with that of bovine liver DD3.


View larger version (60K):
[in this window]
[in a new window]
 
Fig. 1.   Nucleotide sequence and deduced amino acid sequence of bovine liver PGF synthase. The underlined letters indicate the peptide sequences from the purified native enzyme. The arrows indicate the primers used for PCR or mutagenesis as described under "Experimental Procedures." The bold letters in the nucleotide sequence show the polyadenylation signal. The dashed lines show the amino acid sequences of peptides I, II, III, and IV for design of the mixed oligonucleotide primers.


View larger version (51K):
[in this window]
[in a new window]
 
Fig. 2.   Amino acid alignment of bovine PGF synthase isozymes and human liver DD isozymes. The amino acid sequences of the bovine liver-type PGF synthase (LiPGFS), the bovine lung-type ones (LuPGFSI (5) and LuPGFSII (7)), and human liver DD isozymes (hDD1 (16), hDD2 (16), and hDD4 (15)) are aligned. The arrowheads show the positions of site-directed mutagenesis. Arg27, Arg91, His170, Arg223, Lys225, Ser301, and Asn306 were converted to Glu, Gln, Cys, Leu, Ser, Arg, and Tyr (R27E, R91Q, H170C, R223L, K225S, S301R, and N306Y), respectively. Dots indicate the same amino acid residues as LiPGFS, and asterisks indicate the same amino acid residues among the six proteins.

Identification and Size Determination of Bovine Liver PGF Synthase mRNA by Blot Hybridization Analysis-- Fig. 3A shows the result of the Northern blot analysis of bovine liver mRNA with the P3/P4 PCR product (477 base pairs) used as a probe. From its migration in a denaturing gel system, the sequence of PGF synthase mRNA from bovine liver was estimated to be 1400 nucleotides. Therefore, assuming a length for the poly(A) tail of 100-150 nucleotides, the insert cDNA sequence of 1139 nucleotides extended nearly the full length of the mRNA.


View larger version (27K):
[in this window]
[in a new window]
 
Fig. 3.   Northern blot analysis of bovine liver (A), SDS-PAGE (B), and Western blot analysis (C) of the cytosol and the purified recombinant bovine liver PGF synthase. A, total RNA (10 µg) obtained from bovine liver was hybridized with a [alpha -32P]dCTP-labeled P3/P4 probe. The positions at 28 S and 18 S of ribosomal RNA are shown. B, SDS-PAGE silver stain and (C) Western blot analysis using the antiserum against the purified native bovine liver PGF synthase: the native enzyme purified from bovine liver (lane 1, 0.1 µg for B and 0.01 µg for C), and the cytosol (lane 2, 0.5 µg for B and 0.05 µg for C), the ammonium sulfate fraction (lane 3, 0.5 µg for B and 0.05 µg for C), and the Red Sepharose fraction (lane 4, 0.1 µg for B and 0.01 µg for C) of expressed protein in COS-7 cells were loaded into the indicated lanes. The positions of the molecular mass standards are shown: phosphorylase b (97,400), bovine serum albumin (66,300), ovalbumin (42,400), carbonic anhydrase (30,000), soybean trypsin inhibitor (20,100).

Expression of Bovine Liver PGF Synthase in COS-7 Cells and Purification of Its Expressed Protein-- pBOS-BLiFS27 carrying the full-length bovine liver PGF synthase in the mammalian expression vector pEF-BOS was prepared by use of the strategy described under "Experimental Procedures." COS-7 cells were transfected with pBOS-BLiFS27. The recombinant protein was expressed transiently in COS-7 cells and was located in the cytosol fraction. The recombinant protein was purified to an apparent homogeneity by the following purification steps: ammonium sulfate fractionation between 40 and 75% saturation and Red Sepharose column chromatography, as described under "Experimental Procedures." About 2.8-fold purification of the PGD2 11-ketoreductase activity was achieved from the cytosol of COS-7 cells with a yield of 44% (Table I). A sample of each of the purification steps was subjected to SDS-PAGE. Silver staining of the gel indicated that an approximately 36.7-kDa protein was produced in the cells harboring pBOS-BLiFS27, and this protein was purified to an apparent homogeneity (Fig. 3B). Western blot analysis of each sample revealed that the 36.7-kDa protein was recognized by anti-bovine liver PGF synthase antibody (Fig. 3C). The molecular weight of the expressed enzyme was the same as that of the native enzyme, as shown in Fig. 3, B and C. No protein from the control COS-7 cells bearing pEF-BOS without the insert DNA interacted with this antibody (data not shown).

                              
View this table:
[in this window]
[in a new window]
 
Table I
Purification of recombinant wild-type PGF synthase and R27E, R91Q, H170C, R223L, K225S, S301R, and N306Y mutants
All activity measurements were performed under standard assay condition for the PGD2 11-ketoreductase activity as described under "Experimental Procedures." One milliunit of enzyme activity was determined as the amount that produced 1 nmol of PGF2/min at 37 °C.

The purified recombinant protein exhibited enzymatic properties similar to those of the native enzyme. The Km values for PGD2, PGH2, and PQ were 15, 25, and 1.1 µM, respectively (Table II). The low Km value for PGD2 was essentially identical to that of the native liver-type enzyme and was different from the high Km value (120 µM) of the lung-type one. The specific activities of the recombinant protein for PGD2, PGH2, and PQ were 23, 9, and 186 milliunits/mg of protein, respectively. However, PGE2 was not reduced. Moreover, the IC50 value for inhibition of PGD2 11-ketoreductase activity of the expressed enzyme by CuSO4 was 0.5 mM (data not shown), which was almost the same as that for inhibition of the native liver-type enzyme (0.4 mM) but unlike that for the lung-type enzyme (0.003 mM) (6). These results indicate that the cloned cDNA coded for the liver-type PGF synthase and not for the lung-type synthase.

                              
View this table:
[in this window]
[in a new window]
 
Table II
Comparison of kinetic constants for PGD2, PGH2, and phenanthrenequinone among expressed liver PGF synthase, native liver PGF synthase, and native lung PGF synthase
Enzyme assays of expressed liver PGF synthase with PGD2, PGH2, and phenanthrenequinone were conducted under standard assay conditions including the enzyme (3.4 µg for PGD2, 17 µg for PGH2, and 3.4 µg for PQ), and measured by the radioisotope method for PGD2 and PGH2 and by the spectrophotometric method for phenantherenequinone at 37 °C as described under "Experimental Procedures." One milliunit of enzyme activity was determined as the amount that produced 1 nmol of PGF2/min at 37 °C. Enzymes purified to apparent homogeneity were used as enzyme sources.

Site-directed Mutagenesis-- The comparison of amino acid sequences among PGF synthases of bovine liver and lung, and human DD1, DD2, and DD4 is shown in Fig. 2. The Km values of the lung-type (3) and the liver-type (6) PGF synthases and the human liver DD1 and DD2 for PGD2 (16) were 120, 10, 12, and 79 µM, respectively. DD4 does not catalyze the reduction of PGD2 (16). To determine which amino acid residues were related to PGF synthase activity, especially to PGD2 11-ketoreductase activity, we conducted a site-directed mutagenesis study. Arg27, Arg91, Arg223, Lys225, Ser301, or Asn306 of the liver-type PGF synthase with a low Km value for PGD2 was changed to Glu, Gln, Leu, Ser, Arg, or Tyr, respectively, the latter of which are the residues of the lung-type PGF synthase with a high Km value for PGD2. In addition to these mutations, His170 was changed to the Cys of DD4 (Fig. 2), which has no PGD2 11-ketoreductase activity. The mutant enzyme, R27E, R91Q, H170C, R223L, K225S, S301R, or N306Y, was expressed in COS-7 cells and was purified to an apparent homogeneity (Fig. 4) as described under "Experimental Procedures." The results of their final purification step are shown in Table I. The kcat, Km, and kcat/Km values for three representative substrates, i.e. PGD2, PGH2, and PQ, of the purified mutant enzymes are shown in Table III.


View larger version (19K):
[in this window]
[in a new window]
 
Fig. 4.   SDS-PAGE (A) and Western blot analysis (B) of PGF synthase mutants. Wild-type and the mutant PGF synthases were purified in an identical manner and analyzed by SDS-PAGE silver stain (0.1 µg for each lane) (A) and Western blot analysis (0.01 µg for each lane) (B) using the antiserum against the purified native bovine liver PGF synthase: the native enzyme purified from bovine liver (lane 1), the purified wild-type recombinant enzyme (lane 2), and the purified R27E (lane 3), R91Q (lane 4), H170C (lane 5), R223L (lane 6), K225S (lane 7), S301R (lane 8), and N306 (lane 9) mutants cells were loaded into the indicated lanes. The positions of the molecular mass standards are shown in Fig. 3.

                              
View this table:
[in this window]
[in a new window]
 
Table III
Comparison of kinetics for PGD2, PGH2, and phenanthrenequinone among expressed liver PGF synthase and mutants
The enzyme activities for PGD2, PGH2, and phenanthrenequinone were measured in the presence of wild-type PGF synthase and seven mutants (0.06-15 µg for PGD2, 0.3-19 µg for PGH2, and 0.3-10 µg for PQ) by the methods shown in Table II.

Although the kcat/Km values of almost all mutants for PGH2 and PQ were retained above 50% of the wild-type value, these values of all mutants for PGD2 decreased. The kcat/Km values of R27E, R91Q, H170C, and N306Y for PGD2 were only about 10% that of the wild type, and the Km values of R27E, R91Q, H170C, R223L, and N306Y for PGD2 were 110, 145, 75, 180, and 100 µM, respectively. These Km values were 5-10-fold higher than the wild-type value and were almost the same as that of the lung-type PGF synthase. Considering the amino acid residues of these mutants were changed from the liver-type PGF synthase to the lung-type synthase for R27E, R91Q, R223L, and N306Y or to DD4 for H170C, these results suggest that Arg27, Arg91, His170, Arg223, and Asn306 are essential to give a low Km value for PGD2 and that these amino acid residues play an important role on the binding for PGD2 to PGF synthase. In addition, the R223L mutant increased kcat for PGD2 5-fold, indicating that the amino acid residue at 223 has a profound effect on kcat for PGD2 reduction. The kcat/Km values of R27E, R91Q, and N306Y mutants for PQ were 50-80% of that value of the wild type, indicating that Arg27, Arg91, and Asn306 have little effect on the catalytic efficiency for PQ, less than that of these residues on the catalytic efficiency for PGD2. Moreover, the Km values of R223L, K225S, and S301R for PQ were about 2-10-fold lower than the wild-type value, and the kcat/Km values of these mutants for PQ were 3-15-fold higher than the value of the wild type. These results suggest that the amino acid residues at 223, 225, and 301 are related to the binding for PQ to the enzyme and that the binding to the carbonyl group is different between PG and quinone compounds. On the other hand, the Km value of H170C for PGH2 was 8-fold lower than that of the wild type, and the kcat/Km values of H170C and N306Y for PGH2 were 5-6-fold higher than that of the wild type. These results indicate that the amino acid residue at 170 is related to the binding for PGH2 and that Cys170 seems to confer greater affinity for PGH2 than His. Moreover, the amino acid residue at 306 plays an important role in catalytic efficiency for PGH2.

    DISCUSSION

We isolated a clone of the liver-type PGF synthase with a low Km value for PGD2 from the cDNA library of bovine liver, expressed the enzyme in COS-7 cells, and constructed seven mutants. Moreover, we examined the enzymatic properties of the wild-type enzyme and of the mutant enzymes, and we investigated the amino acid residue(s) related to the affinity of PGF synthase for the substrates.

The amino acid sequence of the liver-type PGF synthase consisted of 323 amino acid residues with a Mr of 36,742 (Fig. 1) and showed 87% identity with that of the lung-type synthase (Fig. 2). When the liver-type enzyme was expressed in COS-7 cells (Fig. 3), the recombinant purified protein (Fig. 3, B and C, and Table I) was essentially identical to the native liver-type enzyme and not to the lung-type enzyme, based on the enzymatic properties (Table II) including the low Km value (15 µM) for PGD2 and the results of Western blot analysis (Fig. 3). These results suggest that this liver-type PGF synthase is distinct from the lung-type PGF synthase isolated from the cDNA library of bovine lung (5) or liver (7).

Recently, Jez et al. (23) reported on a structure/function analysis of the aldo-keto reductase superfamily. They reported that five amino acid residues, i.e. Asp50, Asn167, Gln190, Ser271, and Arg276, and three residues, i.e. Asp50, Tyr55, and Lys84, function in cofactor binding and in the active site, respectively. Based on the locations of the cofactor-binding pocket and the active site, a putative substrate-binding site was also proposed. However, the binding site for PGs has not yet been reported. The amino acid sequence of the liver-type PGF synthase was highly homologous with those sequences of DD1, DD2, and DD4 of human liver (Fig. 2), and DD1 and DD2 exhibited PGD2 11-ketoreductase activity (16). Based on the comparison among the amino acid sequences of human liver DDs (15, 16) and the lung-type and the liver-type PGF synthases, we studied the site-directed mutagenesis to change seven amino acid residues as follows: R27E, R91Q, H170C, R223L, K225S, S301R, and N306Y. All mutants expressed the proteins in COS-7 cells to almost the same extent as the wild-type protein, suggesting that the tertiary structures of these mutants were not drastically changed. Moreover, all mutants retained the activity with above 50% of kcat/Km values of the wild type for PGH2 and PQ, indicating that the structures of all mutants were conserved. Therefore, the change in the enzymatic properties of the mutants reflected the mutation of the amino acid residue and not a change in the structure. Judging from the results of this study, Arg27, Arg91, His170, Arg223, and Asn306 are essential to give a low Km value for PGD2 and play an important role in the binding of PGD2; and the amino acid residue at 223 has a significant effect on kcat for PGD2 reduction. Moreover, the results of Km of the R223L, K225S, and S301R mutants for PQ and PGD2 show that R223L, K225S, and S301R mutations acquired high binding for PQ and low for PGD2. Leu223, Ser225, and Arg301 of the lung-type PGF synthase have a positive effect on the binding of PQ. In terms of the binding site for PGH2, the amino acid residue at 170 seems to be related to the binding of this PG and that at 306 plays an important role in kcat/Km of this PG.

Twelve amino acid residues among the 42 amino acid residues of the liver-type enzyme that differed from those of the lung-type one were located in the alpha -helix or beta -sheet, and the other 30 amino acid residues were located in loop structures of the tertiary structure of the liver-type PGF synthase, as inferred from data on human aldose reductase (24, 25) and rat 3alpha -HSD (26), which showed 46 and 70% identity, respectively, in terms of amino acid sequence with bovine liver-type PGF synthase. As a general rule, an amino acid(s) located in an alpha -helix or a beta -sheet is involved in supporting the tertiary structure and that in loop structures of aldo-keto reductases is related to define substrate specificity (23, 27). Although His170 is located in an alpha -helix of the inferred tertiary structure of the liver-type PGF synthase, the amino acid residue at this position in the aldo-keto reductase family shows variation and is located near Asn167, which makes hydrogen bonds with the carboxyamide moiety of the cofactor (23). His170 was changed to Cys of DD4, which has no PGD2 11-ketoreductase activity. The kcat/Km values of this mutant for PGD2 was 16% that of the wild type, and its 1/Km value decreased to about 20% that of the wild type. On the other hand, the kcat/Km values of this mutant for PGH2 were about 6-fold that of the wild type, and its 1/Km value of this mutant increased to 8-fold that of the wild type. These results indicate that the amino acid residue at 170 seems to be related to the binding for PGD2 and to that for PGH2. The reverse effects of the H170C mutant on PGD2 and PGH2 reductase activities indicate that His and Cys at 170 play an important role in the binding of PGD2 and PGH2.

Arg27, Arg91, Arg223, Lys225, Ser301, and Asn306 located in a loop structure are related to the binding site of PGD2, PQ, or PGH2. Jez et al. (27) reported that Trp86 and Trp227 of the 3alpha -HSD near the active site may have roles in substrate binding. They proposed that Trp86 is important in binding to a steroid ligand, whose A-ring lies between this Trp and the cofactor, and that Trp227 interacts with the C- and/or D-rings of steroid ligands. The effect of R91Q on PGD2 taken together with the report on Trp86 of the 3alpha -HSD suggests that Arg91 near Trp86 is a part of the substrate-binding pocket for PGD2. Moreover, although the effects of W227Y on the kcat/Km values for steroids are dramatic, smaller effects of this mutant on those for one-, two-, and three-ring substrates are also shown in this order (27). PGD2 has one cyclopentane ring and two long tails (omega - and alpha -chains), and Arg223 and Lys225 are also located near Trp227. The results of the R223L mutant for PGD2 suggest that Arg223 is an essential amino acid residue to give a low Km value for PGD2 of the liver-type PGF synthase and that the amino acid residue at 223 plays an important role in kcat of PGD2 reduction. The results of the R223L and K225S mutants for PQ suggest that Leu223 and Ser225 have a significant effect on the binding of PQ. PGD2 and PQ seem to bind to the same apolar pocket of PGF synthase differently, like steroids, non-steroidal anti-inflammatory drugs, and aldose reductase inhibitors of 3alpha -HSD (27). Moreover, the effect of S301R on PQ suggests that Ser301 also has a role in the binding of PQ to the enzyme. Asn306, located in the C-terminal region, may have structural importance for the binding of the substrate. Members of the NADPH-dependent aldo-keto reductase superfamily are in part distinguished by unique C-terminal loops (28-30). C-terminal loops of aldo-keto reductase are unique for each member and differ drastically in length and amino acid composition; and the C-terminal loop is critical for catalytic efficiency and for substrate and inhibitor specificity (28-30). In the case of PGF synthases, the C-terminal region may be also critical for the affinity and specificity for the substrate.2 The effect on the binding of the N306Y mutant for PGD2 and that on kcat/Km of this mutant for PGH2 also suggest that Asn306 is a critical determinant of PGs.

Considering the one binding site for NADPH in the inferred tertiary structure of the liver-type PGF synthase, the catalytic site for the various substrates of aldo-keto reductase including PGF synthase seems to be the same. Tyr55 of aldo-keto reductase is favored as the catalytic acid in the reaction mechanism (23). Tyr55 of PGF synthase also plays the same role as that of the corresponding position in the other aldo-keto reductases.2 The Y55F mutation eliminated PGD2, PGH2, and PQ reductase activities of PGF synthase completely.2 However, although PGD2 and PQ reductase activities of Y55Q were eliminated completely, only PGH2 reductase activity of this mutant was stimulated, 27-fold.2 These results suggest that Tyr55 is favored as the catalytic residue, but the reduction mechanism, which involves a hydride transfer from NADPH to the substrate and protonation of the oxygen by a residue of the enzyme acting as a general acid, may be different between PGH2 and PGD2/PQ. The reduction site of PGD2 is the keto group like other carbonyl compounds but that of PGH2 is the endoperoxide group. Moreover, the results of the reverse effects of H170C and N306Y on PGD2 and PGH2 reductase activities shown in this paper, taken together with the results on the competitive inhibition between PGD2 and PQ (3) and on the lack of inhibition between PGH2 and PGD2/PQ (3, 6), indicate that the reduction mechanism for PGD2 may be the same as that for PQ but not the same as that for PGH2.

3alpha -HSD, which catalyzes the NADP-dependent reversible oxidation of the 3alpha -hydroxy group of various steroids, interacts extensively with bile acids (31); and it and human liver high affinity bile acid-binding protein with minimal 3alpha -HSD activity are multifunctional proteins in bile acid transport and xenobiotic metabolism (32). The amino acid sequences of HSD and high affinity bile acid-binding protein are similar to the sequence of bovine liver PGF synthase, and DD2 and DD4 exhibited binding activity for bile acids (33). Therefore, bovine liver PGF synthase may also be expected to exhibit the ability to bind bile acids. In the liver, PGF2alpha , PGE2, and PGD2 reduce bile flow and bile acid secretion, and especially the effect of PGF2alpha is more potent than that of PGE2 or PGD2 (34). PGF synthase may be multifunctionally involved in the biosynthesis of PGF and in the binding of bile acids, and PGF synthase in the liver may reduce bile flow. Moreover, PGF2alpha stimulates hepatocyte DNA synthesis and may have a role in promoting hepatocyte proliferation (35, 36). Furthermore, PGF2alpha and PGD2 are released from primary Ito cell cultures after stimulation by noradrenaline and ATP (37). PGF2alpha plays an important physiological role in the liver, and the liver-type PGF synthase mainly contributes to the biosynthesis of PGF2alpha there.

Recently, the sequence of the bovine liver DD3 (17) was reported. The sequence of the liver-type PGF synthase reported in this paper was found to be identical to that of the bovine liver DD3. The liver-type PGF synthase showed the enzyme activity for (S)-(+)-indanol (6.3 µmol/min/mg),3 which is a typical substrate of DD3 (13, 17). However, PG(s) is the naturally occurring substrate(s) for this enzyme, as indanole is a xenobiotic compound.

    ACKNOWLEDGEMENTS

We are grateful to Professor S. Nagata of the Department of Genetics, Medical School of Osaka University, for a generous gift of the pEF-BOS mammalian expression vector. We are indebted to Professor S. Kuramitsu of the Department of Biology, Graduate School of Science, Osaka University, for the guidance of kinetics of enzyme and for critical reading of this manuscript.

    FOOTNOTES

* This work was supported in part by a grant-in-aid for scientific research from the Ministry of Education, Science, and Culture of Japan and by a grant from Sankyo Foundation of Life Science.The costs of publication of this article were defrayed in part by the payment of page charges. The article must therefore be 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 GenBankTM/EMBL Data Bank with accession number(s) D88749.

§ Present address: Dept. of Anatomy and Cell Biology, School of Medicine, The University of Tokushima, 3-18-15 Kuramoto-cho, Tokushima 770-8503, Japan.

§§ To whom correspondence should be addressed: Osaka Bioscience Institute, 6-2-4 Furuedai, Suita, Osaka 565-0874, Japan. Tel.: 81-6-872-4812; Fax: 81-6-872-4818; E-mail: watanaki{at}obisun1.obi.or.jp.

2 M. Sato, M. Tanaka, K. Ikehara, T. Suzuki, and K. Watanabe, manuscript in preparation.

3 T. Suzuki, and K. Watanabe, unpublished results.

    ABBREVIATIONS

The abbreviations used are: PG, prostaglandin; PQ, 9, 10-phenanthrene quinone; DD, dihydrodiol dehydrogenase; HSD, 3alpha -hydroxysteroid dehydrogenase; KPB, potassium phosphate buffer; PAGE, polyacrylamide gel electrophoresis; PCR, polymerase chain reaction..

    REFERENCES
TOP
ABSTRACT
INTRODUCTION
REFERENCES
  1. Hyman, A. L. (1969) J. Pharmacol. Exp. Ther. 165, 267-273[Medline] [Order article via Infotrieve]
  2. Mathé, A. A. (1977) in The Prostaglandins (Ramwell, P. W., ed), Vol. 3, pp. 169-224, Plenum Publishing Corp., New York
  3. Watanabe, K., Yoshida, R., Shimizu, T., and Hayaishi, O. (1985) J. Biol. Chem. 260, 7035-7041[Abstract/Free Full Text]
  4. Watanabe, K., Iguchi, Y., Iguchi, S., Arai, Y., Hayaishi, O., and Roberts, L. J., II. (1986) Proc. Natl. Acad. Sci. U. S. A 83, 1583-1587[Abstract]
  5. Watanabe, K., Fujii, Y., Nakayama, K., Ohkubo, H., Kuramitsu, S., Kagamiyama, H., Nakanishi, S., and Hayaishi, O. (1988) Proc. Natl. Acad. Sci. U. S. A. 85, 11-15[Abstract]
  6. Chen, L.-Y., Watanabe, K., and Hayaishi, O. (1992) Arch. Biochem. Biophys. 296, 17-26[Medline] [Order article via Infotrieve]
  7. Kuchinke, W., Barski, O., Watanabe, K., and Hayaishi, O. (1992) Biochem. Biophys. Res. Commun. 183, 1238-1246[Medline] [Order article via Infotrieve]
  8. Hara, A., Taniguchi, T., Nakayama, T., and Sawada, H. (1990) J. Biochem. (Tokyo) 108, 250-254[Abstract]
  9. Vogel, K., Bentley, P., Platt, K.-L., and Oesch, F. (1980) J. Biol. Chem. 255, 9621-9625[Abstract/Free Full Text]
  10. Penning, T. M., Mukharji, I., Barrows, S., and Talalay, P. (1984) Biochem. J. 222, 601-611[Medline] [Order article via Infotrieve]
  11. Klein, J., Thomas, H., Post, K., Worner, W., and Oesch, F. (1992) Eur. J. Biochem. 205, 1155-1162[Abstract]
  12. Bolcsak, L. E., and Nerland, D. E. (1983) J. Biol. Chem. 258, 7252-7255[Abstract/Free Full Text]
  13. Nanjo, H., Adachi, H., Morihana, S., Mizoguchi, T., Nishihara, T., and Terada, T. (1995) Biochim. Biophys. Acta 1244, 53-61[Medline] [Order article via Infotrieve]
  14. Matsuura, K., Hara, A., Nakayama, T., Nakagawa, M., and Sawada, H. (1987) Biochim. Biophys. Acta 912, 270-277[Medline] [Order article via Infotrieve]
  15. Deyashiki, Y., Ogasawara, A., Nakayama, T., Nakanishi, M., Miyabe, Y., Sato, K., and Hara, A. (1994) Biochem. J. 299, 545-552[Medline] [Order article via Infotrieve]
  16. Hara, A., Matsuura, K., Tamada, Y., Sato, K., Miyabe, Y., Deyashiki, Y., and Ishida, N. (1996) Biochem. J. 313, 373-376[Medline] [Order article via Infotrieve]
  17. Terada, T., Adachi, H., Nanjo, H., Fujita, N., Takagi, T., Nishikawa, M., Imagawa, M., Nishihara, T., and Maeda, M. (1996) Adv. Exp. Med. Biol 414, 545-553
  18. Yoshimoto, T., Yamamoto, S., Okuma, M., and Hayaishi, O. (1977) J. Biol. Chem. 252, 5871-5874[Abstract]
  19. Chomczynski, P., and Sacchi, N. (1987) Anal. Biochem. 162, 156-159[CrossRef][Medline] [Order article via Infotrieve]
  20. Mizushima, S., and Nagata, S. (1990) Nucleic Acids Res. 18, 5322[Medline] [Order article via Infotrieve]
  21. Lowry, O. H., Rosebrough, N. J., Farr, A. L., and Randall, R. J. (1951) J. Biol. Chem. 193, 265-275[Free Full Text]
  22. Jones, D. H., Sakamoto, K., Vorce, R. L., and Howard, B. H. (1990) Nature 344, 793-794[Medline] [Order article via Infotrieve]
  23. Jez, J. M., Bennett, M. J., Schlegel, B. P., Lewis, M., and Penning, T. M. (1997) Biochem. J. 326, 625-636[Medline] [Order article via Infotrieve]
  24. Wilson, D. K., Bohren, K. M., Gabbay, K. H., and Quiocho, F. A. (1992) Science 257, 81-84[Medline] [Order article via Infotrieve]
  25. Rondeau, J.-M., Tete-Favier, F., Podjarny, A., Reymann, J.-M., Barth, P., Biellmann, J.-F., and Moras, D. (1992) Nature 355, 469-472[CrossRef][Medline] [Order article via Infotrieve]
  26. Hoog, S. S., Pawlowski, J. E., Alzari, P. M., Penning, T. M., and Lewis, M. (1994) Proc. Natl. Acad. Sci. U. S. A 91, 2517-2521[Abstract]
  27. Jez, J. M., Schlegel, B. P., and Penning, T. M. (1996) J. Biol. Chem. 271, 30190-30198[Abstract/Free Full Text]
  28. Oleg, A. B., Kenneth, H. G., and Bohren, K. M. (1996) Biochemistry 35, 14276-14280[CrossRef][Medline] [Order article via Infotrieve]
  29. Bohren, K. M., Grimshaw, C. E., and Gabbay, K. H. (1992) J. Biol. Chem. 267, 20965-20970[Abstract/Free Full Text]
  30. Bohren, K. M., Grimshaw, C. E., Lai, C.-J., Harrison, D. H., Ringe, D., Petsko, G. A., and Gabbay, K. H. (1994) Biochemistry 33, 2021-2032[Medline] [Order article via Infotrieve]
  31. Stolz, A., Takikawa, H., Ookhtens, M., and Kaplowitz, N. (1989) Annu. Rev. Physiol. 51, 161-176[CrossRef][Medline] [Order article via Infotrieve]
  32. Stolz, A., Hammond, L., Lou, H., Takikawa, H., Ronk, M., and Shively, J. E. (1993) J. Biol. Chem. 268, 10448-10457[Abstract/Free Full Text]
  33. Deyashiki, Y., Taniguchi, H., Amano, T., Nakayama, T., Hara, A., and Sawada, H. (1992) Biochem. J. 282, 741-746[Medline] [Order article via Infotrieve]
  34. Beckh, K., Kneip, S., and Arnold, R. (1994) Hepatology 19, 1208-1213[CrossRef][Medline] [Order article via Infotrieve]
  35. Refsnes, M., Thoresen, G. H., Dajani, O. F., and Christoffersen, T. (1994) J. Cell. Physiol. 159, 35-40[Medline] [Order article via Infotrieve]
  36. Dajani, O. F., Rottingen, J. A., Sandnes, D., Horn, R. S., Refsnes, M., Thoresen, G. H., Iversen, J.-G., and Christoffersen, T. (1996) J. Cell. Physiol. 168, 608-617[CrossRef][Medline] [Order article via Infotrieve]
  37. Athari, A., Hanecke, K., and Jungermann, K. (1994) Hepatology 20, 142-148[Medline] [Order article via Infotrieve]


Copyright © 1999 by The American Society for Biochemistry and Molecular Biology, Inc.