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 31 毫秒
1.
Induction of microsomal 1-acyl-glycerophosphocholine (GPC) acyltransferase in rat tissues by four peroxisome proliferators, clofibric acid, tiadenol, DEHP and PFOA, was examined. Among the nine tissues examined, kidney, liver and intestinal mucosa responded to the challenges by the peroxisome proliferators to induce the enzyme. The treatment of rats with various dose of clofibric acid, tiadenol, DEHP or PFOA resulted in an induction of kidney microsomal 1-acyl-GPC acyltransferase in a dose-dependent manner. Despite the structural dissimilarity of peroxisome proliferators, the induction of microsomal 1-acyl-GPC acyltransferase was highly correlated with the induction of peroxisomal beta-oxidation. The activity of microsomal 1-acyl-GPC acyltransferase was not affected by changes in hormonal (adrenalectomy, diabetes, hyperthyroidism and hypothyroidism) and nutritional (starvation, starvation-refeeding, fat-free-diet feeding and high-fat-diet feeding) states. The induction of renal microsomal 1-acyl-GPC acyltransferase was seen in mice subsequent to the administration of clofibric acid and tiadenol and in guinea pigs subsequent to the administration of tiadenol. These results may indicate that kidney microsomal 1-acyl-GPC acyltransferase is a highly specific parameter responsive to the challenges by peroxisome proliferators and may suggest that the possibility that the inductions by peroxisome proliferators of microsomal 1-acyl-GPC acyltransferase and peroxisomal beta-oxidation in kidney are co-regulated.  相似文献   

2.
The influence of the peroxisomal proliferators perfluorooctanoic acid (PFOA), perfluorooctane sulphonic acid (PFOSA) and clofibric acid on lipid metabolism in rats was studied. Dietary treatment of male Wistar rats with these three compounds resulted in rapid and pronounced reduction in both cholesterol and triacylglycerols in serum. The concentration of liver triacylglycerols was increased by about 300% by PFOSA. Free cholesterol was increased by both perfluoro compounds. Cholesteryl ester was reduced to 50% by PFOSA as well by clofibrate. In hepatocytes from fed rats, all the compounds resulted in reduced cholesterol synthesis from acetate, pyruvate and hydroxymethyl glutarate, but there was no reduction of synthesis from mevalonic acid. The oxidation of palmitate was also increased in all groups. The perfluoro compounds, but not clofibrate, caused some reduction in fatty acid synthesis. The activity of liver HMG-CoA reductase was reduced to 50% or less in all treatment groups and all three compounds led to lower activity of acyl-CoA:cholesterol acyltransferase (ACAT). Changes in other enzymes related to lipid metabolism were inconsistent. The present data suggest that the hypolipemic effect of these compounds may, at least partly, be mediated via a common mechanism; impaired production of lipoprotein particles due to reduced synthesis and esterification of cholesterol together with enhanced oxidation of fatty acids in the liver.  相似文献   

3.
Administration of clofibric acid, 2,2'-(decamethylenedithio)diethanol, di(2-ethylhexyl)phthalate or perfluorooctanoic acid to male rates increased markedly microsomal 1-acylglycerophosphocholine (a-acyl-GPC) acyltransferase in a dose-dependent manner in liver. Simultaneous administration of actinomycin D or cycloheximide completely abolished the increase in the enzyme activity. The treatment of rats with clofibric acid did not affect the rate of decay of 1-acyl-GPC acyltransferase. Regardless of a great difference in the chemical structures of the peroxisome proliferators, high correlation was observed between the induced activities of microsomal 1-acyl-GPC acyltransferase and peroxisomal beta-oxidation. Stearoyl-CoA desaturase was induced by peroxisome proliferators in a dose-dependent manner; nevertheless, high correlation was not seen between the induced activities of desaturase and peroxisomal beta-oxidation. Hormonal (adrenalectomy, diabetes, hyperthyroidism and hypothyroidism) and nutritional (starvation, starvation-refeeding, fat-free diet feeding and high-fat diet feeding) alterations hardly affected the activity of 1-acyl-GPC acyltransferase. The present results indicate that microsomal 1-acyl-GPC acyltransferase is a useful parameter responsive to the challenges by peroxisome proliferators and suggest that a similar regulatory mechanism operates for the inductions of microsomal 1-acyl-GPC acyltransferase and peroxisomal beta-oxidation.  相似文献   

4.
Perfluorinated fatty acids alter hepatic lipid metabolism and are potent peroxisome proliferators in rodents. Two such perfluorinated acids, perfluorodecanoic acid (PFDA) and perfluorooctanoic acid (PFOA), were examined to determine if they covalently bind cellular proteins. PFDA and PFOA were found to covalently bind proteins when administered to rats in vivo. The liver, plasma and testes of male rats treated with [1-14C]PFDA or PFOA (9.4 mumol/kg) contained detectable levels of covalently bound 14C (0.1-0.5% of the tissue 14C content). Characterization of PFDA covalent binding to albumin in vitro showed that cysteine significantly decreased binding with no effect of methionine, suggesting protein sulfhydryl groups are involved. In cytosolic and microsomal incubation there was no effect of the addition of CoA, ATP or NADPH on the magnitude of the covalent binding of PFDA. Therefore PFDA need not be metabolically activated to form covalent adducts. Despite demonstration of covalent binding of PFDA and PFOA to proteins both in vivo and in vitro, the role of this macromolecular binding in perfluorinated fatty acid toxicity is not known.  相似文献   

5.
6.
The peroxisome proliferators perfluorooctanoic acid (PFOA; 0.02% w/w), perfluorodecanoic acid (PFDA; 0.02%, w/w), nafenopin (0.125%, w/w), clofibrate (0.5%, w/w), and acetylsalicylic acid (ASA; 1%, w/w) were administered to male C57 BL/6 mice in their diet for two weeks. Parameters for Fe3+ ADP, NADPH or ascorbic acid-initiated lipid peroxidation in vitro were measured. Approximately a twofold increase in susceptibility to lipid peroxidation was obtained for all the peroxisome proliferators tested. Cotreatment of mice with the peroxisome proliferator ASA (1%, w/w) and a catalase inhibitor, 3-amino-1,2,4-triazole (AT; 0.4%, w/w) for 7 days resulted in little inhibition of peroxisome proliferation, an elevated level of H2O2 in vivo, and total inhibition of the increased susceptibility to lipid peroxidation in vitro. No increase in lipid peroxidation in vivo was observed. Certain antioxidant enzymes (DT-diaphorase, superoxide dismutase, glutathione transferase, glutathione peroxidase, and glutathione reductase) and components (ubiquinone and α-tocopherol) were also measured. The results showed that there was some induction of these antioxidant enzymes and components by ASA or aminotriazole, except for glutathione peroxidase and superoxide dismutase, which were inhibited. The possible involvement of oxidative stress in the carcinogenicity of peroxisome proliferators is discussed.  相似文献   

7.
The abilities of the hepatic peroxisome proliferators (HPPs) clofibrate, di(2-ethylhexyl)phthalate (DEHP), mono(2-ethylhexyl)- phthalate (MEHP), 2,4-dichlorophenoxy acetic acid (2,4-D), 2,4,5-trichlorophenoxy acetic acid (2,4,5-T) and tiadenol to induce morphological transformation and to increase the catalase activity of Syrian hamster embryo (SHE) cells were studied. DEHP, MEHP, clofibrate and tiadenol induced morphological transformation of SHE cells and increased the catalase activity. DEHP was more potent than clofibrate and tiadenol in both inducing catalase and morphological transformation, while MEHP seemed more potent than DEHP in inducing catalase, but not morphological transformation, 2,4,5-T and 2,4-D did not induce morphological transformation, but 2,4,5-T was more potent than clofibrate in increasing the catalase activity. These results show that several HPPs induce morphological transformation of SHE cells and an increase in the catalase activity. There is, however, no direct connection between these two parameters, as seen from the results of 2,4,5-T. The tumor promoter TPA, and the metal salt nickel sulphate, induced morphological transformation of SHE cells without any appreciable increase in the catalase activity. These results further corroborate the dissociation between induction of morphological transformation and the increase in catalase activity.Abbreviations Clofibrate ethyl-2-(p-chlorophenox) isobutyrate - 2,4-D 2,4-dichlorophenoxy acetic acid - DEHP di(2-ethylhexyl)phthalate - HPP hepatic peroxisome proliferator - MEHP mono(2-ethylhexyl)phthalate - SHE Syrian hamster embryo - 2,4,5-T 2,4,5-trichlorophenoxy acetic acid - tiadenol di(hydroxyethylthio)-1,10-decane  相似文献   

8.
It has been suggested that peroxisome proliferators stimulate Kupffer cells, an effect which may be involved in their mechanism of action. To evaluate this hypothesis, this study was designed to investigate the effect of stimulating Kupffer cells on basal as well as induced peroxisomal enzyme activity. Twenty four hours following treatment of male Sprague-Dawley rats with the peroxisome proliferating agent perfluorooctanoic acid (PFOA), in corn oil or with corn oil alone, hepatic peroxisomal -oxidation was 4.6 ± 0.2 and 1.8 ± 0.1 U/g liver, respectively. As expected, PFOA did not influence the catalase activity. Stimulating Kupffer cells in vivo by zymosan A (25 mg/kg, iv) prior to treatment with corn oil or PFOA diminished basal as well as PFOA-induced peroxisomal b-oxidation by 20-35%. Activation of Kupffer cells by zymosan A also diminished catalase activity by over 60%. Furthermore, PFOA reduced blood colloidal carbon clearance by 35% within 2 h of its administration. The data suggest that activation of Kupffer cells exerts a negative effect on basal as well as PFOA-induced peroxisomal enzyme activities. Data also suggest that PFOA inhibits Kupffer cells. Activated Kupffer cells may indeed produce factors which interfere with normal hepatic peroxisomal functions and responses.  相似文献   

9.
The occurrence and activation of glutathione S-transferase (GST) and the GST activities in biofilms in cold sulfidic spring waters were compared to the occurrence and activation of GST and the GST activities of the aquatic fungal strains EH5 and EH7 of Mucor hiemalis isolated for the first time from such waters. Using fluorescently labeled polyclonal anti-GST antibodies and GST activity measurements, we demonstrated that a high level of GST occurred in situ in natural biofilms and pure cultures of strain EH5. Measurement of microsomal and cytosolic soluble GST activities using different xenobiotic substrates, including 1-chloro-2,4-dinitrobenzene (CDNB), 1,2-dichloro-4-nitrobenzene, 1,2-epoxy-3-(4-nitrophenoxy)propane, 1-iodo-2,4-dinitrobenzene, and fluorodifen, showed that the overall biotransforming abilities of biofilms were at least sixfold greater than that of strain EH5 alone. Increasing the level of sodium thiosulfate (STS) in the medium stimulated the microsomal and cytosolic GST activities with CDNB of strain EH5 about 44- and 94-fold, respectively, compared to the activities in the control. The induction of microsomal GST activity with fluorodifen by STS was strongly linear, but the initial strong linear increase in cytosolic GST activity with fluorodifen showed saturation-like effects at STS concentrations higher than approximately 1 mM. Using laser scanning confocal and conventional fluorescence microscopy, abundant fluorescently labeled GST proteins were identified in germinating sporangiospores of strain EH5 after activation by STS. High-performance size exclusion chromatography and sodium dodecyl sulfate-polyacrylamide gel electrophoresis revealed the presence of at least two main GSTs (~27.8- and ~25.6-kDa subunits) in the cytosol of EH5, whereas the major 27.8-kDa subunit was the only GST in microsomes. We suggest that differential cellular GST expression takes place in strain EH5 depending on spore and hyphal development. Our results may contribute to our understanding of induction of GST by sulfurous compounds, as well as to the immunofluorescence visualization of GST in aquatic fungus and fungus-bacterium biofilms.  相似文献   

10.
The effects of the anti-wetting agent perfluoro-n-decanoic acid (PFDA) on various glutathione S-transferase (GST) enzyme activities were studied in vitro and in vivo. In addition the effects of PFDA treatment on the amount of some glutathione S-transferase subunits and their corresponding translatable mRNA were studied in vivo. PFDA like some other peroxisome proliferators was a non-competitive inhibitor of several GST enzyme activities in vitro. In vivo PFDA reduced the enzyme activity towards substrates which are indicative for the Ya, Yb1 and Yb2 subunits of GSTs to a larger extent than the enzyme activity towards the substrate indicative for the Yc subunit. Whereas the reduction of GST enzyme activities by other peroxisome proliferators was shown to be caused by an inhibition of the relevant enzymes in vivo, PFDA was found to decrease the GST enzyme activities at least in part by lowering the amount of the various GST subunits in vivo due to a lowered concentration of translatable mRNA coding for these enzymes. In addition PFDA abolished the inducibility of GST mRNAs by phenobarbital. Thus PFDA might be an interesting tool for mechanistic studies of the control of GST expression in the liver.  相似文献   

11.
Immunocytochemical localization of delta 3, delta 2-enoyl-CoA isomerase (isomerase) was investigated in rat liver. Livers of di-(2-ethylhexyl)phthalate (DEHP)-treated or untreated rats were perfusion-fixed and embedded in Epon or Lowicryl K4M. By light microscopy, reaction deposits for the enzyme were present in the cytoplasmic granules of hepatocytes and interlobular bile duct epithelium. Weak staining was noted in sinus-lining cells. After administration of DEHP, the granular staining of the hepatocytes was markedly enhanced, whereas the staining reaction of the sinus-lining cells decreased. The isomerase staining pattern was quite similar to that of long-chain acyl-CoA dehydrogenase (a mitochondrial marker), but different from that of catalase (a peroxisomal marker). Under electron microscopy, gold particles for isomerase were seen to be confined mainly to mitochondria of the hepatocytes, the bile duct epithelial cells and sinus-lining cells. Peroxisomes were weakly labeled. After DEHP administration, the peroxisomes were markedly induced, but the mitochondria were not. Quantitative analysis showed that the induction of the peroxisomal isomerase was only 2-fold whereas the mitochondrial isomerase was enhanced about 5-fold, 40 times as high as the peroxisomal enzyme. The results show that the mitochondria are the main intracellular site for isomerase and the peroxisomes a minor site. The mitochondrial isomerase of the rat liver is markedly induced by peroxisome proliferators, DEHP and clofibrate.  相似文献   

12.
Various antihyperlipemic peroxisome proliferators are known to be carcinogenic in rodents but not in human, other primates and guinea pig, which species lost their ability to synthesize ascorbate due to mutations in the gulonolactone oxidase gene. Ascorbate synthesis is accompanied by H2O2 production, consequently its induction can be potentially harmful; therefore, the in vivo effect of the peroxisome proliferator clofibrate was investigated on gulonolactone oxidase expression in mouse liver. Liver weights and peroxisomal protein contents were increased upon clofibrate treatment. Elevated plasma ascorbate concentrations were found in clofibrate-treated mice due to the higher microsomal gulonolactone oxidase activities. Remarkable gulonolactone oxidase activity appeared in the peroxisomal fraction upon the treatment. Increased activity of the enzyme was associated with an elevation of its mRNA level. According to the present results the evolutionary loss of gulonolactone oxidase may contribute to the explanation of the missing carcinogenic effect of peroxisome proliferators in humans.  相似文献   

13.
Cytosolic glutathione S-transferase (GST) and microsomal epoxide hydrolase (EH) are important detoxification enzymes for many epoxide xenobiotics. We have developed a rapid, simple, and convenient HPLC assay which measures both of these enzyme activities toward benzo[a]pyrene-4,5-oxide (BaPO) in tissue homogenates. Tissue fractions were incubated at 37 degrees C in the presence of 5 mM glutathione. Reactions were initiated by addition of BaPO and terminated by the addition of ice-cold acetonitrile containing 2-methoxynaphthalene as an internal standard. Samples were analyzed directly on a 15-cm C18 reverse-phase column at room temperature, with a ternary solvent program which utilized 0.01% ammonium phosphate buffer (pH 3.5), acetonitrile, and water. The uv absorbance (260 nm) was monitored. Baseline resolution of BaPO, BaPO-GSH, and BaPO-diol and the internal standard was accomplished in 10 min. In rat hepatic S9, production of both BaPO-GSH and BaPO-diol was linear with time and protein up to 15 min and 500 micrograms/ml, respectively. Coefficients of variation for replicate analyses were 2.7 and 3.7% for GST and EH activities in S9, respectively. With fluorescence detection (ex, 241; em, 389 nm), this assay was sensitive enough to measure GST and EH activities in mononuclear leukocytes (MNL). GST and EH activities in 109 human MNL samples were 142 +/- 74 (mean +/- SD; range 21-435) pmol/mg/min and 19 +/- 9 (mean +/- SD; range 3-59) pmol/mg/min, respectively. These results demonstrate the simplicity, high sensitivity, and applicability of this assay for a broad range of tissues.  相似文献   

14.
15.
16.
Using dietary administration, mice were exposed to eight substances known to cause peroxisome proliferation (i.e. clofibrate clofibric acid, 2,4-dichlorophenoxyacetic acid, 2,4,5-trichlorophenoxyacetic acid, nafenopin, ICI-55.897, S-8527 and Wy-14.643) or the related substance p-chlorophenoxyacetic acid (group A). Other animals received di(2-ethylhexyl)phthalate, mono(2-ethylhexyl)phthalate, 2-ethylhexanoic acid, or one of 12 other metabolically and/or structurally related compounds (group B). The effects of these treatments on liver cytosolic and microsomal epoxide hydrolases, microsomal cytochrome P-450, cytosolic glutathione transferase activity, the liver-somatic index and the protein contents of the microsomal and cytosolic fractions prepared from liver were subsequently monitored. In general, peroxisome proliferation was accompanied by increases in cytosolic epoxide hydrolase activity. Many peroxisome proliferators also caused increases in microsomal epoxide hydrolase activity, although the correlation was poorer in this case. Immunochemical quantitation by radial immunodiffusion demonstrated that the increases observed in both of these enzyme activities reflected equivalent increases in enzyme protein, i.e. that induction truly occurred. Induction of total microsomal cytochrome P-450 was obtained after dietary exposure to clofibrate, clofibric acid, 2,4-dichlorophenoxyacetic acid, 2,4,5-trichlorophenoxyacetic acid, nafenopin, Wy-14.643, di(2-ethylhexyl)phthalate and di(2-ethylhexyl)phosphate. The most pronounced effects on cytosolic glutathione transferase activity were the decreases obtained after treatment with clofibrate, clofibric acid and Wy-14.643. Our results, together with those reported by others, suggest that the processes of peroxisome proliferation and induction of cytosolic epoxide hydrolase are intimately related. One possible explanation for this is presented.  相似文献   

17.
Superoxide dismutase (SOD), glutathione peroxidase (GPX), glutathione S-transferase (GST) and glutathione reductase (GR) play crucial roles in balancing the production and decomposition of reactive oxygen species (ROS) in living organisms. These enzymes act cooperatively and synergistically to scavenge ROS, as not one of them can singlehandedly clear all forms of ROS. In order to imitate the synergy of the enzymes, we designed and generated a recombinant protein, which comprises of a Schistosoma japonicum GST (SjGST) and a bifunctional 35-mer peptide with SOD and GPX activities. The engineered protein demonstrated SOD, GPX and GST activities simultaneously. This trifunctional enzyme with SOD, GPX and GST activities is expected to be the best ROS scavenger.  相似文献   

18.
BACKGROUND: Peroxisome proliferators are considered rodent carcinogens that are putative human non-carcinogens based on the presumed absence of direct genetic toxicity in rodent and human cells and the resistance of human cells to the induction of peroxisomes by peroxisome proliferators. The highly sensitive lacZ plasmid-based transgenic mouse mutation assay was employed to investigate the mutagenicity of several peroxisome proliferators based on several lines of evidence suggesting that these agents may in fact exert a genotoxic effect. METHODS: Male and female lacZ-plasmid based transgenic mice were treated at 4 months of age with 6 doses of 2,333 mg di-2-ethylhexyl phthalate (DHEP), 200 mg Wyeth-14,643, or 90 mg clofibrate per kg of bodyweight, respectively, over a two-week period. Control animals were treated with the respective vehicles only (35% propyl glycol for DEHP and Wyeth-14,643 treatment controls and sterile water for clofibrate treatment controls).The mutant frequency in liver, kidney and spleen DNA was determined as the proportion of retrieved mutant and wild-type lacZ plasmids expressed in Escherichia Coli C host cells employing a positive selection system for mutant plasmids. RESULTS: Exposure to DEHP or Wyeth-14,643 significantly increased the mutant frequency in liver, but not in kidney or spleen, of both female and male mice. Treatment with clofibrate did not lead to an increased mutant frequency in any of the organs studied. CONCLUSION: The results indicate that some peroxisome proliferators display an organ-specific mutagenicity in lacZ plasmid-based transgenic mice consistent with historical observations of organ- and compound-specific carcinogenicity.  相似文献   

19.
20.
We have reported [1-3] in immature golden hamster testis that 5 beta-reductase is localized in the tubular nongerm cells, while 5 alpha-reductase is present in the interstitial tissue and that the 17 beta-hydroxy-dehydrogenase activity is found predominantly in the tubular nongerm cells. Hormonal regulation of these enzyme activities was examined in the present study. Male golden hamsters were hypophysectomized on day 22 after birth. The hypophysectomized hamsters in groups of 3-8 were injected daily with 10 micrograms NIH-LH-S19, 50 micrograms NIAMD-Rat-FSH-B-1, 8 or 16 micrograms NIAMD-oFSH-13, 8 micrograms NIAMD-oFSH-13 plus 5 or 10 micrograms NIH-LH-S19, 1 mg testosterone propionate or saline for 5 days starting from day 23. Testicular homogenates of the treated hamsters and intact hamsters on day 28 were incubated with [14C]4-androstene-3,17-dione and NADPH, and enzyme activity (nmol/testes/h) was estimated. The activities of 5 beta- and 5 alpha-reductases and 17 beta-hydroxy-dehydrogenase decreased significantly 6 days after hypophysectomy. In the hypophysectomized hamster testis, a distinct response to FSH but not to LH in the activities of 5 beta-reductase and 17 beta-hydroxy-dehydrogenase was found. The injection of LH in addition to FSH showed no significant additive effects on these enzyme activities. The 5 alpha-reductase activity was stimulated significantly by LH plus FSH but not by LH alone, FSH alone or androgen. These results show that 5 beta-reduction of 4-ene-3-ketosteroids takes place in the Sertoli cells under the influence of FSH while 5 alpha-reduction occurs in the interstitial cells under the influence of LH and FSH in immature hamster testis.  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