首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 53 毫秒
1.
Renal 9-ketoprostaglandin reductase activity from rabbits fed 0.3 g or 2.5 g NaCl per 100 g chow was measured in both centrifuged homogenates and in purified enzyme fractions. There was no salt related increase in 9-ketoprostaglandin reductase activity. PGA1-glutathione, 9,10-phenanthrenequinone, and 4-nitrobenzaldehyde were better substrates for the purified 9-ketoprostaglandin reductases tha was PGE2. Several carbonyl reductases were isolated which used PGA1-glutathione, 9,10-phenanthrenequinone, and 4-nitrobenzaldehyde, but not PGE2, as substrates. Although PGA1-glutathione was a more faithful indicator of PGE2-related 9-ketoprostaglandin reductase activity than either 9,10-phenanthrenequinone or 4-nitrobenzaldehyde, it did not always provide an accurate estimate of that activity.  相似文献   

2.
Substrate specificity of three prostaglandin dehydrogenases   总被引:3,自引:0,他引:3  
Studies on the substrate specificity, kcat/Km, and effect of inhibitors on the human placental NADP-linked 15-hydroxyprostaglandin dehydrogenase (9-ketoprostaglandin reductase) indicate that it is very similar to a human brain carbonyl reductase which also possesses 9-ketoprostaglandin reductase activity. These observations led to a comparison of three apparently homogeneous 15-hydroxyprostaglandin dehydrogenases with varying amounts of 9-ketoprostaglandin reductase activity: an NAD- and an NADP-linked enzyme from human placenta and an NADP-linked enzyme from rabbit kidney. All three enzymes are carbonyl reductases for certain non-prostaglandin compounds. The placental NAD-linked enzyme, which has no 9-ketoprostaglandin reductase activity, is the most specific of the three. Although it has carbonyl reductase activity, a comparison of the Km and kcat/Km for prostaglandin and non-prostaglandin substrates of this enzyme suggests that its most likely function is as a 15-hydroxyprostaglandin dehydrogenase. The results of similar comparisons imply that the other two enzymes may function as less specific carbonyl reductases.  相似文献   

3.
The hydroxylation of prostaglandin (PG) E1, PGE2, and PGA1 was investigated in a reconstituted rabbit liver microsomal enzyme system containing phenobarbital-inducible isozyme 2 or 5,6-benzoflavone-inducible isoenzyme 4 of P-450, NADPH-cytochrome P-450 reductase, phosphatidylcholine, and NADPH. Significant metabolism of prostaglandins by isozyme 2 occurred only in the presence of cytochrome b5. Under these conditions, PGE1 hydroxylation was linear with time (up to 45 min) and protein concentration, and maximal rates were obtained with a 1:1:2 molar ratio of reductase: cytochrome b5:P-450LM2. Moreover, P-450LM2 catalyzed the conversion of PGE1, PGE2, and PGA1 to the respective 19- and 20-hydroxy metabolites in a ratio of about 5:1, and displayed comparable activities toward the three prostaglandins based on the total products formed in 60 min. Apocytochrome b5 or ferriheme could not substitute for intact cytochrome b5, while reconstitution of apocytochrome b5 with ferriheme led to activities similar to those obtained with the native cytochrome. Isozyme 4 of P-450 differed markedly from isozyme 2 in that it catalyzed prostaglandin hydroxylation at substantial rates in the absence of cytochrome b5, was regiospecific for position 19 of all three prostaglandins, and had an order of activity of PGA1 greater than PGE1 greater than PGE2. P-450LM4 preparations from untreated and induced animals had similar activities with PGE1 and PGE2, respectively. Addition of cytochrome b5 resulted in a 20 to 30% increase in the rate of PGE1 hydroxylation and an appreciably greater enhancement in the extent of all the P-450LM4-catalyzed reactions, the stimulation being greatest with PGE2 (3-fold) and least with PGA1 (1.6-fold). Cytochrome b5 was thus required for maximal metabolism of all three prostaglandins, but did not alter the regiospecificity or the order of activity of P-450 isozyme 4 with the individual substrates. In the presence of cytochrome b5, the prostaglandin hydroxylase activities of isozyme 4 were two to six times higher than those of isozyme 2.  相似文献   

4.
An enzyme with 6-pyruvoyl tetrahydropterin (6PPH4) (2'-oxo)reductase activity was purified to near homogeneity from whole rat brains by a rapid method involving affinity chromatography on Cibacron blue F3Ga-agarose followed by high performance ion exchange chromatography and high performance gel filtration. The enzyme has a single subunit of Mr 37,000 and has a similar amino acid composition to previously described aldoketo reductases. The reductase activity is absolutely dependent on NADPH, will only catalyze the reduction of the C-2'-oxo group of 6PPH4, and is inactive towards the C-1'-oxo group. However, the enzyme also shows high activity towards nonspecific substrates, such as 4-nitrobenzaldehyde, phenanthrenequinone, and menadione. The role of this 6PPH4 reductase in the formation of tetrahydrobiopterin (BH4) was investigated. Measurements were made of the rate of conversion of 6PPH4, generated from dihydroneopterin triphosphate with purified 6PPH4 synthase, to BH4 in the presence of mixtures of pure sepiapterin reductase and the 6PPH4 (2'-oxo)reductase purified from rat brains. The results suggest that when sepiapterin reductase activity is limiting, a large proportion of BH4 synthesis proceeds through the 6-lactoyl intermediate. However, when sepiapterin reductase is not limiting, most of the BH4 is probably formed via reduction of the other mono-reduced intermediate which is produced from 6PPH4 by sepiapterin reductase alone.  相似文献   

5.
H Chung  J Fried  J Jarabak 《Prostaglandins》1987,33(3):391-402
Oxidation of glutathione disulfide by a mixture of performic and hydrochloric acids leads to the formation of several compounds that are stronger inhibitors than glutathione disulfide of the placental enzyme that possess both NADP-linked 15-hydroxyprostaglandin dehydrogenase and 9-ketoprostaglandin reductase activities. The only one of these inhibitors that has been identified is glutathione thiosulfonate. The others are unstable and may include glutathione sulfinyl sulfone and glutathione disulfone. Since the enzyme appears to have a glutathione binding site in close proximity to its active site and glutathione thiosulfonate reacts with free sulfhydryl groups, the effects of this thiosulfonate on the enzyme were examined in more detail. Glutathione thiosulfonate and methyl methanethiosulfonate cause a time-dependent irreversible inhibition of both the hydroxyprostaglandin dehydrogenase and the ketoprostaglandin reductase activities, presumably by reacting with a free sulfhydryl at the prostaglandin binding site. Experiments with PGA1-glutathione show that this sulfhydryl is not necessary for the catalytic activity of the enzyme as long as the substrate can bind at the glutathione site.  相似文献   

6.
The primary structure of prostaglandin (PG) F synthetase from bovine lung shows 62% similarity with that of human liver aldehyde reductase (EC 1.1.1.2) (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). We therefore purified human liver aldehyde reductase to homogeneity and compared the immunological and catalytic properties of aldehyde reductase and PGF synthetase. Although both enzymes belong to a group of aldoketoreductases and their molecular weights are essentially identical, aldehyde reductase had no cross-reactivity to anti-PGF synthetase antiserum. Furthermore, there was a difference in the substrate specificity for reduction of PGs between the two enzymes. Aldehyde reductase catalyzed the reduction of PGJ2, delta 12-PGJ2, PGH2, or PGA2, but not that of PGB2, PGD2, or PGE2, whereas PGF synthetase reduced PGD2. The optimum pH, Km value for PGH2, and the turnover number were 6.5, 100 microM, and 3.1 min-1, respectively. The PGH2 9,11-endoperoxide reductase activity of aldehyde reductase was not affected in the presence of a substrate such as p-nitrobenzaldehyde, DL-glyceraldehyde, or 9,10-phenanthrenequinone, suggesting that PGH2 9,11-endoperoxide and other substrates are reduced at different active site(s). The reaction product formed from PGH2 by this enzyme was identified as PGF2 alpha by gas chromatography/mass spectrometry. These results suggest that aldehyde reductase is not exactly identical to PGF synthetase in terms of its immunological property and substrate specificity for PGs, but that this enzyme is also involved in the direct conversion of PGH2 to PGF2 alpha similar to PGF synthetase.  相似文献   

7.
Chinese hamster monomeric carbonyl reductases (CHCRs) belong to the short-chain dehydrogenase/reductase (SDR) superfamily, which is a family of enzymes that metabolize many endogenous and xenobiotic compounds. We previously cloned three carbonyl reductase cDNAs-Chcr1, Chcr2, and Chcr3. By performing spectrophotometric analyses, we indicated that the enzymes CHCR1, CHCR2, and CHCR3 had similar specificities toward steroids; only CHCR3 did not show any reactivity with prostaglandins (PGs). In the present study, we investigated the characteristics of CHCRs in detail, that is, the differences in their expression patterns, physicochemical properties, and enzymatic activities. CHCR1 exhibited sex-dependent expression patterns. CHCRs showed multiple surface potentials in the zeta potential analysis and CHCR3 exhibited an isatin reductase activity with a high K(m) value. By the present HPLC-analysis, all the three enzymes exhibited PGF(2alpha) dehydrogenase activity and could oxidize PGF(2alpha) to PGE(2) and 15-keto-PGF(2alpha), i.e., the three enzymes exhibited 9- and 15-hydroxy PG dehydrogenase activities. Moreover, 15-keto-PGE(2) was detected in a comparatively higher amount in the dehydrogenase reaction products of CHCR2 than in those of CHCR1 and CHCR3, suggesting that CHCR2 can oxidize PGE(2) and/or 15-keto-PGF(2alpha) to 15-keto-PGE(2); however, these two PGs did not seem to be efficient substrates of CHCR1. Despite the differences in the dehydrogenase activities between CHCR1 and CHCR2, PGE(2) reductase activities of the two enzymes were similar, and PGF(2alpha) was predominantly produced from PGE(2) as a result of the PG 9-keto reductase activity. On the other hand, CHCR3 exhibited a reduced PGE(2) reductase activity. In conclusion, although the CHCRs share a high degree of sequence identity (>70%), they clearly differed in their enzymatic characteristics.  相似文献   

8.
A homogeneous preparation of sepiapterin reductase, an enzyme involved in the biosynthesis of tetrahydrobiopterin, from rat erythrocytes was found to be responsible for the reduction with NADPH of various carbonyl compounds of non-pteridine derivatives including some vicinal dicarbonyl compounds which were reported in the previous paper (Katoh, S. and Sueoka, T. (1984) Biochem. Biophys. Res. Commun. 118, 859–866) in addition to the general substrate, sepiapterin (2-amino-4-hydroxy-6-lactoyl-7,8-dihydropteridine). The compounds sensitive as substrates of the enzyme were quinones, e.g., p-quinone and menadione; other vicinal dicarbonyls, e.g., methylglyoxal and phenylglyoxal; monoaldehydes, e.g., p-nitrobenzaldehyde; and monoketones, e.g., acetophenone, acetoin, propiophenone and benzylacetone. Rutin, dicoumarol, indomethacin, and ethacrynic acid inhibited the enzyme activity toward either a carbonyl compound of a non-pteridine derivative or sepiapterin as substrate. Sepiapterin reductase is quite similar to general aldo-keto reductases, especially to carbonyl reductase.  相似文献   

9.
Pyridoxal reductase (PL reductase), which catalyzes reduction of PL by NADPH to form pyridoxine and NADP(+), was purified from Schizosaccharomyces pombe. The purified enzyme was very unstable but was stabilized by low concentrations of various detergents such as Tween 40. The enzyme was a monomeric protein with the native molecular weight of 41,000 +/- 1,600. The enzyme showed a single absorption peak at 280 nm (E(1%) = 10.0). PL and 2-nitrobenzaldehyde were excellent substrates, and no measurable activity was observed with short chain aliphatic aldehydes; substrate specificity of PL reductase was obviously different from those of yeast aldo-keto reductases (AKRs) so far purified. The peptide sequences of PL reductase were identical with those in a hypothetical 333-amino acid protein from S. pombe (the DDBJ/EMBL/GenBank(TM) accession number D89205). The gene corresponding to this protein was expressed in Escherichia coli, and the purified protein was found to have PL reductase activity. The recombinant PL reductase showed the same properties as those of native PL reductase. PL reductase showed only low sequence identities with members of AKR superfamily established to date; it shows the highest identity (18.5%) with human Shaker-related voltage-gated K(+) channel beta2 subunit. The elements of secondary structure of PL reductase, however, distributed similarly to those demonstrated in the three-dimensional structure of human aldose reductase except that loop A region is lost, and loop B region is extended. Amino acid residues involved in substrate binding or catalysis are also conserved. Conservation of these features, together with the major modifications, establish PL reductase as the first member of a new AKR family, AKR8.  相似文献   

10.
An NAD(P)-dependent 3 alpha-hydroxysteroid dehydrogenase (EC 1.1.1.50) was purified to homogeneity from rat liver cytosol, where it is responsible for most if not all of the capacity for the oxidation of androsterone, 1-acenaphthenol and benzenedihydrodiol (trans-1,2-dihydroxycyclohexa-3,5-diene). The dehydrogenase has many properties (substrate specificity, pI, Mr, amino acid composition) in common with the dihydrodiol dehydrogenase (EC 1.3.1.20) purified from the same source [Vogel, Bentley, Platt & Oesch (1980) J. Biol. Chem. 255, 9621-9625]. Since 3 alpha-hydroxysteroids are by far the most efficient substrates, the enzyme is more appropriately designated a 3 alpha-hydroxysteroid dehydrogenase. It also promotes the NAD(P)H-dependent reductions of quinones (e.g. 9,10-phenanthrenequinone, 1,4-benzoquinone), aromatic aldehydes (4-nitrobenzaldehyde) and aromatic ketones (4-nitroacetophenone). The dehydrogenase is not inhibited by dicoumarol, disulfiram, hexobarbital or pyrazole. The mechanism of the powerful inhibition of this enzyme by both non-steroidal and steroidal anti-inflammatory drugs [Penning & Talalay (1983) Proc. Natl. Acad. Sci. U.S.A. 80, 4504-4508] was examined with several substrates. Most non-steroidal anti-inflammatory drugs are competitive inhibitors (e.g. Ki for indomethacin, 0.20 microM for 9,10-phenanthrenequinone reduction at pH 6.0, and 0.835 microM for androsterone oxidation at pH 7.0), except for salicylates, which act non-competitively (e.g. Ki for aspirin, 650 microM for androsterone oxidation). The inhibitory potency of these agents falls sharply as the pH is increased from 6 to 9. Most anti-inflammatory steroids are likewise competitive inhibitors, except for the most potent (betamethasone and dexamethasone), which act non-competitively. The enzyme is inhibited competitively by arachidonic acid and various prostaglandins.  相似文献   

11.
zeta-Crystallin is a major protein in the lens of certain mammals. In guinea pigs it comprises 10% of the total lens protein, and it has been shown that a mutation in the zeta-crystallin gene is associated with autosomal dominant congenital cataract. As with several other lens crystallins of limited phylogenetic distribution, zeta-crystallin has been characterized as an "enzyme/crystallin" based on its ability to reduce catalytically the electron acceptor 2,6-dichlorophenolindophenol. We report here that certain naturally occurring quinones are good substrates for the enzymatic activity of zeta-crystallin. Among the various quinones tested, the orthoquinones 1,2-naphthoquinone and 9,10-phenanthrenequinone were the best substrates whereas menadione, ubiquinone, 9,10-anthraquinone, vitamins K1 and K2 were inactive as substrates. This quinone reductase activity was NADPH specific and exhibited typical Michaelis-Menten kinetics. Activity was sensitive to heat and sulfhydryl reagents but was very stable on freezing. Dicumarol (Ki = 1.3 x 10(-5) M) and nitrofurantoin (Ki = 1.4 x 10(-5) M) inhibited the activity competitively with respect to the electron acceptor, quinone. NADPH protected the enzyme against inactivation caused by heat, N-ethylmaleimide, or H2O2. Electron paramagnetic resonance spectroscopy of the reaction products showed formation of a semiquinone radical. The enzyme activity was associated with O2 consumption, generation of O2- and H2O2, and reduction of ferricytochrome c. These properties indicate that the enzyme acts through a one-electron transfer process. The substrate specificity, reaction characteristics, and physicochemical properties of zeta-crystallin demonstrate that it is an active NADPH:quinone oxidoreductase distinct from quinone reductases described previously.  相似文献   

12.
FQR1 is a novel primary auxin-response gene that codes for a flavin mononucleotide-binding flavodoxin-like quinone reductase. Accumulation of FQR1 mRNA begins within 10 min of indole-3-acetic acid application and reaches a maximum of approximately 10-fold induction 30 min after treatment. This increase in FQR1 mRNA abundance is not diminished by the protein synthesis inhibitor cycloheximide, demonstrating that FQR1 is a primary auxin-response gene. Sequence analysis reveals that FQR1 belongs to a family of flavin mononucleotide-binding quinone reductases. Partially purified His-tagged FQR1 isolated from Escherichia coli catalyzes the transfer of electrons from NADH and NADPH to several substrates and exhibits in vitro quinone reductase activity. Overexpression of FQR1 in plants leads to increased levels of FQR1 protein and quinone reductase activity, indicating that FQR1 functions as a quinone reductase in vivo. In mammalian systems, glutathione S-transferases and quinone reductases are classified as phase II detoxification enzymes. We hypothesize that the auxin-inducible glutathione S-transferases and quinone reductases found in plants also act as detoxification enzymes, possibly to protect against auxin-induced oxidative stress.  相似文献   

13.
The microsomes of placenta and uterus from pregnant rabbits have been found to catalyze the omega-hydroxylation of PGE1, PGE2, PGF2 alpha, and PGA1 as well as the omega- and (omega-1)-hydroxylation of palmitate and myristate in the presence of NADPH. These activities were greatly inhibited by carbon monoxide, indicating the involvement of cytochrome P-450. The apparent Km for PGE1 was 2.38 microM and 2.1 microM with the placental and uterus microsomes, respectively. Cytochrome P-450 has been solubilized with 1% cholate from the placental microsomes, and partially purified by chromatography on 6-amino-n-hexyl Sepharose 4B, DEAE-Sephadex A-50 and hydroxylapatite columns. The partially purified cytochrome P-450 efficiently catalyzed the omega-hydroxylation of various prostaglandins such as PGE1, PGE2, PGF2 alpha, PGD2, and PGA1 in a reconstituted system containing NADPH-cytochrome P-450 reductase, cytochrome b5, and phosphatidylcholine. The reconstituted system also hydroxylated palmitate and myristate at the omega- and (omega-1)-position, but could not hydroxylate laurate. These catalytic properties resemble those of a new form of cytochrome P-450 highly purified from the lung microsomes of progesterone-treated rabbits (Yamamoto, S., Kusunose, E., Ogita, K., Kaku, M., Ichihara, K., and Kusunose, M. (1984) J. Biochem. 96, 593-603). This type of cytochrome P-450, viz., cytochrome P-450 with high prostaglandin omega-hydroxylase activity may play a role in the regulation of prostaglandin levels in pregnancy.  相似文献   

14.
On nitroaryl reductase activities in several Clostridia   总被引:5,自引:0,他引:5  
Crude extracts of Clostridium kluyveri, Clostridium spec. La 1, Clostridium sporogenes and Clostridium pasteurianum catalyse the NADH-dependent reduction of the nitro group of p-nitrobenzoate. The former three Clostridia also use pyruvate as electron donor for this reduction. The NADH-dependent reductases have been partially purified and characterized from Clostridium kluyveri. Nitroalkyl compounds as well as nitrite, sulfite, sulfate and hydroxylamine are no substrates. Based on chromatographic behavior, separation pattern, yields, stability, pH optima, molecular masses and EPR studies the three NADH-dependent nitroaryl group reducing enzymes in Clostridium kluyveri (three activities in Clostridium spec. La 1 and two activities in Clostridium sporogenes) are different from alcohol dehydrogenase, aldehyde dehydrogenase, 3-hydroxy-butyryl-CoA dehydrogenase, butyryrl-CoA dehydrogenase, 2-enoate reductase, ferredoxin-NAD and ferredoxin-NADP reductase. The physiological roles of the nitroaryl reductases are not known. The reductase activities show losses of 80-90% during classical protein purification procedures. One of the three nitroaryl reductases exhibits a pH optimum of 10.5. The crude extract reveals a pH optimum at 11.5. The first step of the reduction reaction leads to the nitroradical anion (1 electron transfer). The electron transfer to p-nitrobenzoate is also catalysed by ferrodoxin-NAD reductase from NADH and by ferredoxin-NADP reductase from NADP. Partially purified 2-oxo-acid synthases from Clostridium sporogenes catalyse with low rates the reduction of p-nitrobenzoate as well as 2-nitroethanol in the presence and absence of ferredoxin using pyruvate or 2-oxo-4-methylpentanoate as electron donors, respectively. The NADH-dependent reduction of p-nitro-benzoate accounts for at least 70% and the 2-oxo acid-dependent reduction for about 5% of the total nitroaryl reductase activity in the Clostridia. It seems that the pyridine nucleotide-dependent nitroaryl reductases are enzymes so far unknown in Clostridia.  相似文献   

15.
Testicular interstitial cells were utilized to study the effects of prostaglandins (PG) on in vitro testosterone production and to examine the role of cyclic adenosine-3',5'-monophosphate (cAMP) in this process. Testosterone production was assessed after 3 hour incubations while cAMP accumulation was examined after a 0.5 hour incubation period. Testosterone and cAMP were measured by radioimmunoassay. None of the PGs tested (PGA, PGA2, PGB1, PGE1, PGE2, PGF1alpha PGF2alpha) altered basal testosterone production when present in incubates at concentrations of 1.3 X 10(-8) M to 1.3 X 10(-4). However, at concentrations of 1.3 X 10(-4) M all of these PGs were capable of decreasing Luteinizing Hormone (LH; 100ng)-induced testosterone production. The inhibition of LH-induced testosterone production by the B, E and F series PGs was less pronounced than that for the A series. PGA1 and PGA2 exhibited 80% and 95% inhibition, respectively, at 1.3 X 10(4) M. The inhibitory action of 4 X 10(5) M PGA1 or PGA2 was evident within 30 minutes. Preincubation of interstitial cells with indomethacin (10(-5) or 10(-6) M) for 30 minutes did not alter subsequent basal or LH (100ng)-induced testosterone production. Accumulation of cAMP was stimulated by LH (10 microgram) or by PGs (1.3 X 10(-4) M PGA1, PGA2, PGB1, PGE1 or PGF2alpha). The PG-induced cAMP accumulation thus occurred at concentrations where LH-stimulated testosterone production was inhibited. Furthermore, PGA1 and PGA2 (1.3 X 10(-4) M) inhibited testosterone production induced by either 3-isobutyl-1-methyl xanthine (MIX; 10(-4) M or 10(-3) M) or dibutyryl cAMP (dbcAMP; 10(-4) M or 10(-3) M). These results indicate that PGs can block testosterone production by a direct effect on testicular interstitial cells and suggest that PGs exert their inhibitory action distal to stimulation of cAMP formation. PGs do not appear to play a role in the mechanism of LH action.  相似文献   

16.
A novel aldo-keto reductase (AKR) from Escherichia coli has been cloned, expressed and purified. This protein, YghZ, is distantly related (<40%) to mammalian aflatoxin dialdehyde reductases of the aldo-keto reductase AKR7 family and to potassium channel beta-subunits in the AKR6 family. The enzyme has been placed in a new AKR family (AKR14), with the designation AKR14A1. Sequences encoding putative homologues of this enzyme exist in many other bacteria. The enzyme can reduce several aldehyde and diketone substrates, including the toxic metabolite methylglyoxal. The K(m) for the model substrate 4-nitrobenzaldehyde is 1.06 mM and for the endogenous dicarbonyl methylglyoxal it is 3.4 mM. Overexpression of the recombinant enzyme in E. coli leads to increased resistance to methylglyoxal. It is possible that this enzyme plays a role in the metabolism of methylglyoxal, and can influence its levels in vivo.  相似文献   

17.
Detailed analysis of the action of prostaglandins (PGs) on the corpus luteum in primate species is very limited. In this study we examined the response of the adenylate cyclase system to PGs in homogenates prepared from the corpus luteum of rhesus monkeys at midluteal phase of the menstrual cycle. The conversion of [alpha 32p] ATP to [32p] cyclic AMP (cAMP) was assessed in the absence (control activity; 50 microM GTP) and presence of various concentrations of seven PGs and arachidonic acid, either alone or in combination with 250 nM hCG. Cyclic AMP production increased up to three-fold in the presence of PGD2, PGE2, PGI2 or PGF2 alpha; however PGA2, PGB2, 13, 14-dihydro-15-keto PGE2 and arachidonic acid alone did not alter cAMP levels. In dose-response studies, adenylate cyclase was 10 and 100-fold more sensitive to PGD2 (Vmax at 1 X 10(-5) M) than to PGE2 or to PGI2 and PGF2 alpha, respectively. Activity in the presence of hCG plus either PGD2, PGE2, PGI2 or PGF2 alpha did not differ from that for hCG (or the PG) alone. In contrast, addition of PGA2 or arachidonate inhibited (p less than 0.05) hCG-stimulated cAMP production by 50 and 100 percent. We conclude that the gonadotropin-sensitive adenylate cyclase of the macaque corpus luteum is also modulated by several PGs. These factors may either mimic (e.g., PGD2, PGE2, PGI2) or suppress (PGA2) gonadotropin-stimulated cAMP production and possibly cAMP-mediated events in luteal cells.  相似文献   

18.
The yakC gene in Schizosaccharomyces pombe, which encodes yakC protein (YakC), a potential member of an aldo-keto reductase (AKR) family, was cloned and expressed in Escherichia coli cells. The recombinant YakC purified to homogeneity catalyzed the reduction of 2-nitrobenzaldehyde (k(cat), 44.1 s(-1), K(m), 0.185 +/- 0.018 mM), 2-phthalaldehyde (19.8, 0.333 +/- 0.032), and pyridine-2-aldehyde (7.64, 0.302 +/- 0.028). Neither pyridoxal nor other compounds examined acted as substrates. NADPH, but not NADH, was a hydrogen donor. The enzyme is a monomer with a molecular weight of 38,900 +/- 6,600 (SDS-PAGE). The amino acid sequence deduced from yakC showed the highest (34%) identity with that of pyridoxal reductase (AKR8A1) among the identified AKRs. Twenty-one function-unknown proteins showed 40% or higher identity to the deduced amino acid sequence: DR2261 protein of Deionococcus radiodurans showed the highest (50%) identity. The predicted secondary structure of YakC is similar to that of human aldose reductase, a representative AKR. The results establish YakC as the first member of a new AKR family, AKR13. The yeast cells contained enzyme(s) other than YakC and pyridoxal reductase with the ability to reduce 2-nitrobenzaldehyde: total (100%) activity in the crude extract consisted of about 23% YakC, about 44% pyridoxal reductase, and about 33% other enzyme(s).  相似文献   

19.
Aldehyde reductase (ALR1) and aldose reductase (ALR2) were purified from human placenta by a rapid and efficient scheme that included rapid extraction of both reductases from 100,000 x g supernatant material with Red Sepharose followed by purification by chromatofocusing on Pharmacia PBE 94 and then chromatography on a hydroxylapatite high performance liquid chromatography column. Expression of ALR1 and ALR2 in placenta is variable with ALR1/ALR2 ratios ranging from 1:4 to 4:1. ALR1 and ALR2 are immunochemically distinct. ALR1 shows broad specificity for aldehydes but does not efficiently catalyze the reduction of glucose due to poor binding (Km = 2.5 M). ALR1 exhibits substrate inhibition with many substrates. ALR2 also shows broad specificity for aldehydes. Although glucose is a poor substrate for ALR2 compared with other substrates, the affinity of ALR2 for glucose (Km = 70 mM) suggests that glucose can be a substrate under hyperglycemic conditions. ALR2 shows normal hyperbolic kinetics with most substrates except with glyceraldehyde, which exhibits substrate activation. Treatment of ALR2 with dithiothreitol converted it into a form that exhibited hyperbolic kinetics with glyceraldehyde. Dithiothreitol treatment of ALR2 did not alter its properties toward other substrates or affect its inhibition by aldose reductase inhibitors such as sorbinil (2,4-dihydro-6-fluorospiro-[4H-1-benzopyran-4,4'-imidazolidine]-2' ,5'- dione), tolrestat (N-[[6-methoxy-5-(trifluoromethyl)-1-naphthalenyl]thioxomethyl]-N- methylglycine), or statil (3-[(4-bromo-2-fluorophenyl)methyl]-3,4-dihydro-4-oxo-1-phthalazineac etic acid).  相似文献   

20.
Three kinds of NADPH-dependent aldehyde reducing enzymes were present in the dog kidney. Aldose reductase was located in the inner medulla region and aldehyde reductase in all regions of the renal cortex, outer medulla and inner medulla. In addition, a new reductase designated tentatively as high-Km aldose reductase, which was converted into an aldose reductase-like enzyme, was present in the inner medulla region of the kidney. Aldose reductase, aldehyde reductase and high-Km aldose reductase were purified to homogeneity from each region of the dog kidney. The molecular weight of aldose reductase was estimated to be 38,500 by SDS-polyacrylamide gel electrophoresis and the isoelectric point was found to be 5.7 by chromatofocusing. Aldose reductase had activity for aldo-sugars such as D-xylose, D-glucose and D-galactose as substrates and utilized both NADPH and NADH as coenzymes. Sulfate ions resulted in over 2-fold activation of aldose reductase. All aldehyde reductases from the three regions had the same properties. The molecular weights and isoelectric points of aldehyde reductases were 40,000 and 6.1, respectively. The aldehyde reductases were inactive for D-hexose, utilized only NADPH as coenzyme and were not affected by sulfate ions. High-Km aldose reductase had a molecular weight of 38,500 and an isoelectric point of 5.4. It had activity for aldo-sugars, but showed much higher Km and lower kcat/Km values than aldose reductase. Sulfate ions inhibited high-Km aldose reductase. It was converted into an aldose reductase-like enzyme by incubation in phosphate buffer at pH 7.0. The three kinds of enzymes were strongly inhibited by the known aldose reductase inhibitors. However, aldehyde reductase and high-Km aldose reductase were, in general, less susceptible than aldose reductase.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号