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 15 毫秒
1.
The inhibitory effects of diesel exhaust components and flavonoids on 20α-hydroxysteroid dehydrogenase (20α-HSD) activity were examined in cytosolic fractions from the liver, kidney and lung of male mice. 9,10-Phenanthrenequinone (9,10-PQ) and 1,2-naphthoquinone (1,2-NQ), which are contained in diesel exhaust particles (DEPs), potently inhibited 20α-HSD activity in liver cytosol. 9,10-PQ also inhibited the enzyme activity in lung cytosol. However, 20α-HSD activity in kidney cytosol was little inhibited by 9,10-PQ or 1,2-NQ. Flavonoids such as quercetin, fisetin and kaempferol exhibited high inhibitory potencies for 20α-HSD activity in liver cytosol, whereas these flavonoids were poor inhibitors for the enzyme activity in kidney cytosol. It is likely that several diesel exhaust components and flavonoids augment the signaling of progesterone in the liver cells, by potently inhibiting 20α-HSD activity in mouse liver cytosol. The possibility that there are distinct enzymes catalyzing 20α-HSD activity in the non-reproductive tissues of male mice is also discussed.  相似文献   

2.
Shimada H  Miura K  Imamura Y 《Life sciences》2006,78(25):2931-2936
Progesterone was stereoselectively reduced to a metabolite 20alpha-hydroxy-4-pregnen-3-one in the cytosolic fraction from the liver of male mice, indicating that the reduction of progesterone is catalyzed by 20alpha-hydroxysteroid dehydrogenase (20alpha-HSD). The cytosolic 20alpha-HSD activity was observed not only in the liver, but also in the kidney and lung. In liver cytosol, both NADPH and NADH were effective as cofactors for 20alpha-HSD activity, although NADPH was better than NADH for the enzyme activity. On the other hand, 20alpha-HSD activity in kidney cytosol required only NADPH as a cofactor. No significant sex-related difference of 20alpha-HSD activity was observed in liver and kidney cytosols. Flavonoids have been reported to inhibit the biosynthesis and metabolism of steroids. However, little is known about inhibitory effects of flavonoids on 20alpha-HSD activity. Thus, the effects of 16 flavonoids on 20alpha-HSD activity were examined, using liver cytosol of male mice. Among flavonoids tested, fisetin, apigenin, naringenin, luteolin, quercetin and kaempferol exhibited high inhibitory potencies for the 20alpha-HSD activity. We propose the possibility that these flavonoids augment progesterone signaling by inhibiting potently 20alpha-HSD activity in non-reproductive tissues.  相似文献   

3.
The effects of flavonoids and quinones on NADPH- and NADH-dependent 20alpha-hydroxysteroid dehydrogenase (20alpha-HSD) activities were examined in cytosolic fractions from the liver and kidney of mice. Judging from the data for the inhibition of NADPH- and NADH-dependent 20alpha-HSD activities by flavonoids and quinones, enzyme catalyzing renal NADPH-dependent 20alpha-HSD activity was found to be distinct from enzyme(s) catalyzing hepatic NADPH- and NADH-dependent 20alpha-HSD activities. Sulfobromophthalein (SBP) had little ability to inhibit hepatic NADPH-dependent 20alpha-HSD activity and bromophenol blue (BPB) exhibited a weak activation against the enzyme activity, whereas SBP and BPB were potent and moderate inhibitors, respectively, of hepatic NADH-dependent 20alpha-HSD activity. Thus, enzyme catalyzing hepatic NADPH-dependent 20alpha-HSD activity appeared to be distinct from enzyme catalyzing hepatic NADH-dependent 20alpha-HSD activity. The data for the pH profiles of hepatic NADPH- and NADH-dependent 20alpha-HSD activities also led us to the conclusion. Based on these results, we propose the possibility that several distinct enzymes catalyze NADPH- and NADH-dependent 20alpha-HSD activities in cytosolic fractions from the liver and kidney of mice.  相似文献   

4.
9,10-Phenanthrenequinone (9,10-PQ), a major quinone in diesel exhaust particles, induces apoptosis via the generation of reactive oxygen species (ROS) because of 9,10-PQ redox cycling. We have found that intratracheal infusion of 9,10-PQ facilitates the secretion of surfactant into rat alveolus. In the cultured rat lung, treatment with 9,10-PQ results in an increase in a lower-density surfactant by ROS generation through redox cycling of the quinone. The surfactant contains aldo-keto reductase (AKR) 1C15, which reduces 9,10-PQ and the enzyme level in the surfactant increases on treatment with 9,10-PQ suggesting an involvement of AKR1C15 in the redox cycling of the quinone. In six human cell types (A549, MKN45, Caco2, Hela, Molt4 and U937) only type II epithelial A549 cells secrete three human AKR1C subfamily members (AKR1C1, AKR1C2 and AKR1C3) with the surfactant into the medium; this secretion is highly increased by 9,10-PQ treatment. Using in vitro enzyme inhibition analysis, we have identified AKR1C3 as the most abundantly secreted AKR1C member. The AKR1C enzymes in the medium efficiently reduce 9,10-PQ and initiate its redox cycling accompanied by ROS production. The exposure of A549 cells to 9,10-PQ provokes viability loss, which is significantly protected by the addition of the AKR1C3 inhibitor and antioxidant enzyme and by the removal of the surfactants from the culture medium. Thus, the AKR1C enzymes secreted in pulmonary surfactants probably participate in the toxic mechanism triggered by 9,10-PQ.  相似文献   

5.
9,10-Phenanthrenequinone (9,10-PQ), a major quinone found in diesel exhaust particles, is considered to generate reactive oxygen species (ROS) through its redox cycling. Here, we show that 9,10-PQ evokes apoptosis in human aortic endothelial cells (HAECs) and its apoptotic signaling includes ROS generation and caspase activation. The 9,10-PQ-induced cytotoxicity was inhibited by ROS scavengers, indicating that intracellular ROS generation is responsible for the 9,10-PQ-induced apoptosis. Comparison of mRNA expression levels and kinetic constants in the 9,10-PQ reduction among 10 human reductases suggests that aldo-keto reductase 1C3 (AKR1C3) is a 9,10-PQ reductase in HAECs. In in vitro 9,10-PQ reduction by AKR1C3, the reduced product 9,10-dihydroxyphenanthrene and superoxide anions were formed, suggesting the enzymatic two-electron reduction of 9,10-PQ that thereby causes oxidative stress through its redox cycling. In addition, the participation of AKR1C3 in 9,10-PQ-redox cycling was confirmed by the data that AKR1C3 overexpression in endothelial cells augmented the ROS generation and cytotoxicity by 9,10-PQ, and the ROS scavengers inhibited the toxic effects. Pretreatment of the overexpressing cells with AKR1C3 inhibitors, flufenamic acid and indomethacin, suppressed the 9,10-PQ-induced GSH depletion. These results suggest that AKR1C3 is a key enzyme in the initial step of 9,10-PQ-induced cytotoxicity in HAECs.  相似文献   

6.
9,10-Phenanthrenequinone (9,10-PQ), a major component in diesel exhaust particles, is suggested to generate reactive oxygen species (ROS) through its redox cycling, leading to cell toxicity. l-Xylulose reductase (XR), a NADPH-dependent enzyme in the uronate pathway, strongly reduces alpha-dicarbonyl compounds and was thought to act as a detoxification enzyme against reactive carbonyl compounds. Here, we have investigated the role of intracellular ROS generation in apoptotic signaling in human acute T-lymphoblastic leukemia MOLT-4 cells treated with 9,10-PQ and the role of XR in the generation of ROS. Treatment with 9,10-PQ elicited not only apoptotic signaling, including mitochondrial membrane dysfunction and activation of caspases and poly(ADP-ribose) polymerase, but also intracellular ROS generation and consequent glutathione depletion. The apoptotic effects of 9,10-PQ were drastically mitigated by pretreatment with intracellular ROS scavengers, such as N-acetyl-l-cysteine, glutathione monoethyl ester, and polyethylene glycol-conjugated catalase, indicating that intracellular ROS generation is responsible for the 9,10-PQ-evoked apoptosis. Surprisingly, the ROS generation and cytotoxicity by 9,10-PQ were augmented in an XR-transformed cell line. XR indeed reduced 9,10-PQ and produced superoxide anion through redox cycling. In addition, the expression levels of XR and its mRNA in the T lymphoma cells were markedly enhanced after the exposure to 9,10-PQ, and the induction was completely abolished by the ROS scavengers. Moreover, the 9,10-PQ-induced apoptosis was partially inhibited by the pretreatment with XR-specific inhibitors. These results suggest that initially produced ROS induce XR, which accelerates the generation of ROS.  相似文献   

7.
Progesterone, 17alpha-hydroxyprogesterone, cortisone and cortisol, which are C(21)-steroids with a ketone group at the 20-position, potently inhibited the activity of enzyme acetohexamide reductase (AHR) responsible for the reductive metabolism of acetohexamide in kidney microsomes of male rats. Furthermore, progesterone was a competitive inhibitor of AHR. In the case of progesterone usage as the substrate, 20beta-hydroxysteroid dehydrogenase (20beta-HSD) activity was much higher than 20alpha-hydroxysteroid dehydrogenase (20alpha-HSD) activity in kidney microsomes of male rats. These results indicate that AHR present in kidney microsomes of male rats, functions as 20beta-HSD with carbonyl reductase-like activity. In male rats, both testectomy and hypophysectomy decreased the renal microsomal 20beta-HSD activity, but the decreased enzyme activities were increased by the treatment with testosterone propionate (TP). We propose the possibility that TP treatment regulates the renal microsomal 20beta-HSD activity by acting directly on the kidney of male rats. This is supported from the fact that when TP was given to ovariectomized and hypophysectomized female rats, the male-specific 20beta-HSD activity was detected in their kidney microsomes.  相似文献   

8.
9,10-Phenanthrenequinone (9,10-PQ) is one of the most abundant quinones among diesel exhaust particulates. Recent data have suggested that quinones induce apoptosis in immune, epithelial and tumor cells, leading to respirator illness; however, the mechanisms by which quinones induce apoptosis and the structure required for this remain unknown. We studied the antitumor activity of 9,10-PQ analogs against two human tumor cell lines, HCT-116 colon tumor cells and HL-60 promyelocytic leukemia cells. The loss of the cis-orthoquinone unit in 9,10-PQ abrogated its ability to induce apoptosis in the two tumor cell lines, and the LC50 values of these analogs were indicated over 10 μM. An analog of 9,10-PQ in which the biaryl unit had been deleted displayed a reduced ability to induce tumor cell apoptosis, while the analogs 1,10-phenanthroline-5,6-dione (9) and pyrene-4,5-dione (10), which also had modified biaryl units, exhibited increased tumor cell apoptotic activity. The cis-orthoquinone unit in 9,10-PQ was identified as essential for its ability to induce apoptosis in tumor cells, and its biaryl unit is also considered to influence orthoquinone-mediated apoptotic activity.  相似文献   

9.
Interaction of camel lens zeta-crystallin, an NADPH:quinone oxidoreductase, with several quinone derivatives was examined by fluorescence spectroscopy and activity measurements. Fluorescence of zeta-crystallin was quenched to different levels by the different quinones:juglone (5-OH, 1,4 naphthoquinone), 1,4 naphthoquinone (1,4-NQ), and 1,2 naphthoquinone (1,2-NQ) considerably quenched the fluorescence of zeta-crystallin, where as the commonly used substrate, 9,10-phenanthrenequinone (PQ) did not induce significant quenching. Activity measurements showed only PQ served as a substrate for camel lens zeta-crystallin, while juglone, 1,4-NQ, and 1,2-NQ were inhibitors. Thus quinones that interacted with zeta-crystallin directly inhibited the enzyme, whereas the substrate had very low affinity for the enzyme in the absence of NADPH. Another substrate, dichlorophenol indophenol (DCIP), conformed to the same pattern; DCIP did not quench the fluorescence of the enzyme significantly, but served as a substrate. This pattern is consistent with an ordered mechanism of catalysis with quinone being the second substrate. All three naphthoquinones were uncompetitive inhibitors with respect to NADPH and noncompetitive with respect to PQ. These kinetics are similar to those exhibited by cysteine- and/or lysine-modifying agents. Juglone, 1,4-NQ, and 1,2-NQ interacted with and quenched the fluorescence of camel lens alpha-crystallin, but to lesser extent than that of zeta-crystallin.  相似文献   

10.
Homogeneous 3 alpha-hydroxysteroid dehydrogenase (3 alpha-HSD) from rat liver cytosol displays 9, 11, and 15-hydroxyprostaglandin dehydrogenase activity. Using [14C]-PGF2 alpha as substrate the products of this reaction were separated by TLC and identified by autoradiography as PGE2 and PGB2. The purified enzyme catalyzes this reaction at a rate 200 times faster than cytosol. This corresponds to the rate enhancement observed when the enzyme is purified from cytosol using androsterone (a 3 alpha-hydroxysteroid) as substrate and suggests that it may represent a major 9-hydroxyprostaglandin dehydrogenase in this tissue. Although the 3 alpha-HSD has many properties in common with the 9-hydroxyprostaglandin dehydrogenase of rat kidney, rat kidney contains no protein that is immunodetectable with polyclonal antibody raised against the purified 3 alpha-HSD.  相似文献   

11.
1,2-Naphthoquinone (1,2-NQ) is electrophilic, and forms covalent bonds with protein thiols, but its two-electron reduction product 1,2-dihydroxynaphthalene (1,2-NQH(2)) is not, so enzymes catalyzing the reduction with reduced pyridine nucleotides as cofactors could protect cells from electrophile-based chemical insults. To assess this possibility, we examined proteins isolated from the 9000g supernatant from mouse liver for 1,2-NQ reductase activity using an HPLC assay procedure for the hydroquinone of 1,2-NQ and Cibacron Blue 3GA column chromatography and Western blot analysis with specific antibody to determine 1,2-NQ-bound proteins. Among the proteins with high affinities for pyridine nucleotides that also inhibited 1,2-NQ-protein adduct formation in the presence of NADH, a 37-kDa protein was found and identified as glyceraldehyde-3-phosphate dehydrogenase (GAPDH). Using recombinant human GAPDH, we found that this glycolytic enzyme indeed catalyzes the two-electron reduction of 1,2-NQ accompanied by extensive NADH consumption under 20% oxygen conditions. When either 1,2-NQH(2) or 1,2-NQ was incubated with GAPDH in the presence of NADH, minimal covalent bonding to the enzyme occurred compared to that in its absence. These results indicate that GAPDH can inhibit 1,2-NQ-based electrophilic protein modification by conversion to the nonelectrophilic 1,2-NQH(2) via an NADH-dependent process.  相似文献   

12.
9,10-Phenanthrenequinone (9,10-PQ), a redox-active quinone in diesel exhausts, triggers cellular apoptosis via reactive oxygen species (ROS) generation in its redox cycling. This study found that induction of CCAAT/enhancer-binding protein-homologous protein (CHOP), a pro-apoptotic factor derived from endoplasmic reticulum stress, participates in the mechanism of rat endothelial cell damage. The 9,10-PQ-mediated CHOP induction was strengthened by a proteasome inhibitor (MG132) and the MG132-induced cell sensitization to the 9,10-PQ toxicity was abolished by a ROS inhibitor, suggesting that ROS generation and consequent proteasomal dysfunction are responsible for the CHOP up-regulation caused by 9,10-PQ. Aldo-keto reductase (AKR) 1C15 expressed in rat endothelial cells reduced 9,10-PQ into 9,10-dihydroxyphenanthrene concomitantly with superoxide anion formation, implying its participation in evoking the 9,10-PQ-redox cycling. The 9,10-PQ-induced damage was augmented by AKR1C15 over-expression. 9,10-PQ also provoked the AKR1C15 up-regulation, which sensitized against the quinone toxicity. These results suggest the presence of a negative feedback loop exacerbating the quinone toxicity in rat endothelial cells.  相似文献   

13.
The enzyme 20alpha-hydroxysteroid dehydrogenase (20alpha-HSD) catalyzes the conversion of progesterone into its inactive form, 20alpha-hydroxyprogesterone. To gain information about the exact sites of 20alpha-HSD mRNA expression, we performed in situ hybridization using a (35)S-labeled cRNA probe in tissues of adult mice of both sexes. 20alpha-HSD mRNA was expressed in both male and female gonads. In the ovary, high expression was found in luteal cells of corpora lutea, while much lower expression could be detected in granulosa cells of growing follicles. In the testis, a specific hybridization signal was detected only in Leydig cells. In the female reproductive tract, 20alpha-HSD mRNA was found in the epithelial cells of the uterine cervix. In the adrenal cortex, only the zona reticularis exhibited specific radiolabeling, the expression being very high in the female and very low in the male. In the skin, specific labeling was restricted to sebaceous glands, the hybridization signal being much higher in the female than in the male. In the liver, 20alpha-HSD mRNA was found in hepatocytes, with a higher degree of expression in the female. In the kidney, specific labeling was observed in the epithelial cells of distal convoluted tubules, the signal being also much more striking in the female than in the male. In non-reproductive tissues, it clearly appears that the expression of 20alpha-HSD mRNA is higher in the female than in the male, suggesting that 20alpha-HSD may play an important role in reducing the intracellular concentration of progesterone originating from the circulation at a much higher level in the female.  相似文献   

14.
《Free radical research》2013,47(7):848-857
Abstract

9,10-Phenanthrenequinone (9,10-PQ), a redox-active quinone in diesel exhausts, triggers cellular apoptosis via reactive oxygen species (ROS) generation in its redox cycling. This study found that induction of CCAAT/enhancer-binding protein-homologous protein (CHOP), a pro-apoptotic factor derived from endoplasmic reticulum stress, participates in the mechanism of rat endothelial cell damage. The 9,10-PQ-mediated CHOP induction was strengthened by a proteasome inhibitor (MG132) and the MG132-induced cell sensitization to the 9,10-PQ toxicity was abolished by a ROS inhibitor, suggesting that ROS generation and consequent proteasomal dysfunction are responsible for the CHOP up-regulation caused by 9,10-PQ. Aldo-keto reductase (AKR) 1C15 expressed in rat endothelial cells reduced 9,10-PQ into 9,10-dihydroxyphenanthrene concomitantly with superoxide anion formation, implying its participation in evoking the 9,10-PQ-redox cycling. The 9,10-PQ-induced damage was augmented by AKR1C15 over-expression. 9,10-PQ also provoked the AKR1C15 up-regulation, which sensitized against the quinone toxicity. These results suggest the presence of a negative feedback loop exacerbating the quinone toxicity in rat endothelial cells.  相似文献   

15.
The first aim of this study is to characterize the reduction of progesterone in rat liver. Progesterone was mainly reduced to 20alpha-hydroxyprogesterone in the cytosolic fraction of rat liver. The amount of 20alpha-hydroxyprogesterone formed from progesterone in the cytosolic fraction was significantly larger in the males than in the females and this enzyme reaction proceeded not only in the presence of NADPH, but also in the presence of NADH. Furthermore, we attempted to evaluate the inhibitory effects of 15 flavonoids on the NADPH-dependent reduction of progesterone to 20alpha-hydroxyprogesterone in liver cytosol of male rats. The order of the inhibitory potencies was luteolin>apigenin>quercetin>myricetin=fisetin=kaempferol. Other flavonoids exhibited lower inhibitory potencies. Energy-minimized molecular models demonstrated that a planar benzopyrone ring (A and C rings) with a coplanar phenyl ring (B ring) is a structural characteristic determining the inhibitory effects of flavonoids other than isoflavones.  相似文献   

16.
Glyceraldehyde-3-phosphate dehydrogenase (GAPDH) catalyzes the oxidative phosphorylation of glyceraldehyde 3-phosphate to 1,3-diphosphoglycerate, one of the precursors for glycolytic ATP biosynthesis. The enzyme contains an active site cysteine thiolate, which is critical for its catalytic function. As part of a continuing study of the interactions of quinones with biological systems, we have examined the susceptibility of GAPDH to inactivation by 9,10-phenanthrenequinone (9,10-PQ). In a previous study of quinone toxicity, this quinone, whose actions have been exclusively attributed to reactive oxygen species (ROS) generation, caused a reduction in the glycolytic activity of GAPDH under aerobic and anaerobic conditions, indicating indirect and possible direct actions on this enzyme. In this study, the effects of 9,10-PQ on GAPDH were examined in detail under aerobic and anaerobic conditions so that the role of oxygen could be distinguished from the direct effects of the quinone. The results indicate that, in the presence of the reducing agent DTT, GAPDH inhibition by 9,10-PQ under aerobic conditions was mostly indirect and comparable to the direct actions of exogenously-added H2O2 on this enzyme. GAPDH was also inhibited by 9,10-PQ anaerobically, but in a somewhat more complex manner. This quinone, which is not considered an electrophile, inhibited GAPDH in a time-dependent manner, consistent with irreversible modification and comparable to the electrophilic actions of 1,4-benzoquinone (1,4-BQ). Analysis of the anaerobic inactivation kinetics for the two quinones revealed comparable inactivation rate constants (k(inac)), but a much lower inhibitor binding constant (K(i)) for 1,4-BQ. Protection and thiol titration studies suggest that these quinones bind to the NAD+ binding site and modify the catalytic thiol from this site. Thus, 9,10-PQ inhibits GAPDH by two distinct mechanisms: through ROS generation that results in the oxidization of GAPDH thiols, and by an oxygen-independent mechanism that results in the modification of GAPDH catalytic thiols.  相似文献   

17.
L J Askonas  T M Penning 《Biochemistry》1991,30(49):11553-11560
Nonsteroidal anti-inflammatory drugs (NSAIDs) exert their effect by inhibiting the target enzyme cyclooxygenase (prostaglandin H2 synthase); however, little is known about the peptides comprising its NSAID binding site. Hydroxyprostaglandin dehydrogenases also bind NSAIDs, but their NSAID binding sites have not been well characterized. Using existing synthetic strategies, we have incorporated the bromoacetoxy affinity labeling moiety around the perimeter of two potent NSAIDs, indomethacin and mefenamate, a N-phenylanthranilate. The compounds synthesized were 1-(4-(bromoacetamido)benzyl)-5-methoxy-2-methylindole-3-acetic acid (1), 3-(2-(2-bromoacetoxy)ethyl)-1-(4-chlorobenzyl)-5-methoxy-2-methylindole (2), 4-(bromoacetamido)-N-(2,3-dimethylphenyl)anthranilic acid (3), N-(3-(bromoacetamido)phenyl)-anthranilic acid (4), and N-(4-(bromoacetamido)phenyl)anthranilic acid (5). To access whether these compounds have general utility in labeling NSAID binding sites, the compounds were evaluated as affinity labeling agents for 3 alpha-hydroxysteroid dehydrogenase (3 alpha-HSD) from rat liver cytosol. This enzyme displays 9-, 11-, and 15-hydroxyprostaglandin dehydrogenase activity, is inhibited potently by NSAIDs, and is homologous to bovine lung prostaglandin F synthase. Compounds 1-5 were shown to affinity label the NSAID binding site of 3 alpha-HSD. They inactivated 3 alpha-HSD through an E.I complex in a time- and concentration-dependent manner with t1/2 values ranging from seconds to hours. Ligands that compete for the active site of 3 alpha-HSD (NAD+ and indomethacin) afforded protection against inactivation, and the inactivators could demonstrate competitive kinetics against 3 alpha-hydroxysteroid substrates by forming an E.NAD+.I complex.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
Steroidogenic enzymes belonging to the aldo-keto reductase family (AKR) possess highly homologous sequences while having different activities. To gain further knowledge about the function as well as the regulation of these enzymes in the monkey, we have isolated cDNA sequences encoding monkey type 5 17beta-hydroxysteroid dehydrogenase, 20alpha-hydroxysteroid dehydrogenase and 3alpha-hydroxysteroid dehydrogenase, and characterized their enzymatic activity and mRNA tissue distribution. Sequence analysis indicates that these enzymes share approximately 94 and 76% amino acid identity with human and mouse homologs, respectively. Monkey type 5 17beta-HSD possesses 95.9% amino acid sequence identity with human type 5 17beta-HSD. It catalyzes the transformation of 4-androstenedione into testosterone, but it lacks 20alpha-hydroxysteroid dehydrogenase activity that is present in the human enzyme. This activity seems to be specific to human, since mouse type 5 17beta-HSD does not show significant 20alpha-HSD activity. In addition, monkey and mouse 20alpha-HSD possess relatively high 20alpha-, 3alpha-, and 17beta-HSD activities, while their human counterpart is confined to 20alpha-HSD activity. The monkey 3alpha-HSD possesses relatively high 3alpha-, 17beta-, and 20alpha-HSD activities; human type 1 3alpha-HSD exerts 3alpha- and 20alpha-HSD activities; the mouse 3alpha-HSD displays a unique 3alpha-HSD activity. Quantification of mRNA expression shows that the monkey 3alpha-HSD is exclusively expressed in the liver, while the type 5 17beta-HSD is predominately found in the kidney, with lower levels observed in the stomach, liver, and colon. Monkey 20alpha-HSD mRNA is highly expressed in the kidney, stomach, and liver. Our study provides the basis for future investigations on the regulation and function of these enzymes in the monkey.  相似文献   

19.
Rats were generated by pronuclear injection of the transgene with a cDNA construct encoding rat regucalcin that is a regulatory protein of Ca2+ signaling. Transgenic (TG) founders were fertile, transmitted the transgene at the expected frequency, and bred to homozygote. Western analysis of the cytosol prepared from the tissue of TG female rats (5-week-old) showed a remarkable expression of regucalcin (3.3 kDa) protein in the liver, kidney cortex, heart, lung, stomach, brain, spleen, muscle, colon, and duodenum. Regucalcin expression of TG male rats was seen in the liver, kidney cortex, heart, and lung. In wild-type (wt) male and female rats, regucalcin was mainly present in the liver and kidney cortex. Regucalcin inhibited protein phosphatase activity in rat kidney cortex cytosol and activated Ca2+-ATPase activity in rat heart muscle microsomes. The suppressive effect of regucalcin on protein phosphatase activity was significantly enhanced in the cytosol of kidney cortex of TG male and female rats as compared with those of wt rats. Likewise, heart muscle microsomal Ca2+-ATPase activity was significantly enhanced in TG rats. The changes in their enzyme's activities in TG rats were completely abolished in the presence of anti-regucalcin monoclonal antibody (100 ng/ml) in the enzyme reaction mixture. Moreover, the body weight of TG female rats was significantly lowered as compared with that of wt rats. Serum inorganic phosphorus concentration was significantly increased in TG male and female rats, while serum calcium, glucose, triglyceride, free cholesterol, albumin, and urea nitrogen concentrations were not significantly altered in TG rats. Regucalcin TG rats should be a useful model to define a regulatory role of endogenous regucalcin in the tissues in vivo.  相似文献   

20.
In earlier studies, two distinct molecules, 20 alpha-HSD-I and 20 alpha-HSD-II, responsible for 20 alpha-HSD activity of pig adrenal cytosol were purified to homogeneity and characterized [S. Nakajin et al., J. Steroid Biochem. 33 (1989) 1181-1189]. We report here that the purified 20 alpha-HSD-I, which mainly catalyzes the reduction of 17 alpha-hydroxyprogesterone to 17 alpha,20 alpha-dihydroxy-4-pregnen-3-one, catalyzes 3 alpha-hydroxysteroid oxidoreductase activity for 5 alpha (or 5 beta)-androstanes (C19), 5 alpha (or 5 beta)-pregnanes (C21) in the presence of NADPH as the preferred cofactor. The purified enzyme has a preference for the 5 alpha (or 5 beta)-androstane substrates rather than 5 alpha (or 5 beta)-pregnane substrates, and the 5 beta-isomers rather than 5 alpha-isomers, respectively. Kinetic constants in the reduction for 5 alpha-androstanedione (Km; 3.3 microM, Vmax; 69.7 nmol/min/mg) and 5 beta-androstanedione (Km; 7.7 microM, Vmax; 135.7 nmol/min/mg) were demonstrated for comparison with those for 17 alpha-hydroxyprogesterone (Km; 26.2 microM, Vmax; 1.3 nmol/min/mg) which is a substrate for 20 alpha-HSD activity. Regarding oxidation, the apparent Km and Vmax values for 3 alpha-hydroxy-5 alpha-androstan-17-one were 1.7 microM and 43.2 nmol/min/mg, and 1.2 microM and 32.1 nmol/min/mg for 3 alpha-hydroxy-5 beta-androstan-17-one, respectively. 20 alpha-HSD activity in the reduction of 17 alpha-hydroxyprogesterone catalyzed by the purified enzyme was inhibited competitively by addition of 5 alpha-DHT with a Ki value of 2.0 microM. Furthermore, 17 alpha-hydroxyprogesterone inhibited competitively 3 alpha-HSD activity with a Ki value of 150 microM.  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