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.
Products that appeared to be mainly benzo[a]pyrene 7,8-oxide and benzo[a]pyrene 9,10-oxide were synthesized and their chemical and biochemical properties were investigated. The oxides were unstable and readily rearranged to phenols. They were converted by rat liver homogenates and microsomal preparations into phenols and dihydrodiols, but glutathione conjugates were not formed in appreciable amounts. The dihydrodiols formed from benzo[a]pyrene 7,8- and 9,10-oxide by rat liver microsomal preparations were identical in their chromatographic and spectrographic properties with dihydrodiols formed when benzo[a]pyrene was metabolized by rat liver homogenates. 9,10-Dihydrobenzo[a]pyrene 7,8-oxide and 7,8-dihydrobenzo[a]pyrene 9,10-oxide were also synthesized. They were converted by rat liver homogenates and microsomal preparations into the related cis- and trans-dihydroxy compounds. Glutathione conjugates were formed from the oxides by rat liver homogenates. Both 7,8- and 9,10-dihydrobenzo[a]pyrene were metabolized by rat liver homogenates to mainly the trans-isomers of the related dihydroxy compounds. In experiments with boiled homogenates, the benzo[a]pyrene oxides were converted into phenols, whereas the dihydrobenzo[a]pyrene oxides yielded small amounts of the related dihydroxy compounds.  相似文献   

2.
Oxidation of benzo[a]pyrene by the filamentous fungus Cunninghamella elegans.   总被引:18,自引:0,他引:18  
Cunninghamella elegans oxidized benzo[a]pyrene to several metabolic products. Compounds that were isolated and identified were: trans-9,10-dihydroxy-9,10-dihydrobenzo[a]pyrene, trans-7,8-dihydroxy-7,8-dihydrobenzo[a]pyrene, benzo[a]pyrene 1,6-quinone, benzo[a]pyrene 3,6-quinone, 9-hydroxybenz[a]pyrene, and 3-hydroxybenzo[a]pyrene. In addition, an unidentified dihydroxybenzo[a]pyrene metabolite was also formed. Experiments with [14C]benzo[a]pyrene showed that over a 96-h period, 18.4% of the hydrocarbon was converted to metabolic products. Most of the metabolites were sulfate conjugates as demonstrated by the formation of benzo[a]pyrene quinones and phenols after treatment with aryl sulfatase. Glucuronide and sulfate conjugates were also detected as water-soluble metabolites. The results show that benzo[a]pyrene is metabolized by a filamentous fungus in a manner that is remarkably similar to that observed in higher organisms.  相似文献   

3.
(±)-7β,8α-Dihydroxy-9β,10β-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene (diol epoxide-1) and (±)-7β,8α-dihydroxy-9α,10α-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene (diol epoxide-2) are highly mutagenic diol epoxide diastereomers that are formed during metabolism of the carcinogen (±)-trans-7,8-dihydroxy-7,8-dihydrobenzo[a]pyrene. Remarkable stereoselectivity has been observed on metabolism of the optically pure (+)- and (?)-enantiomers of the dihydrodiol which are obtained by separation of the diastereomeric diesters with (?)-α-methoxy-α-trifluoromethylphenylacetic acid. The high stereoselectivity in the formation of diol epoxide-1 relative to diol epoxide-2 was observed with liver microsomes from 3-methylcholanthrene-treated rats and with a purified cytochrome P-448-containing monoxygenase system where the (?)-enantiomer produced a diol epoxide-2 to diol epoxide-1 ratio of 6 : 1 and the (+)-enantiomer produced a ratio of 1 : 22. Microsomes from control and phenobarbital-treated rats were less stereospecific in the metabolism of enantiomers of BP 7,8-dihydrodiol. The ratio of diol epoxide-2 to diol epoxide-1 formed from the (?)- and (+)-enantiomers with microsomes from control rats was 2 : 1 and 1 : 6, respectively. Both enantiomers of BP 7,8-dihydrodiol were also metabolized to a phenolic derivative, tentatively identified as 6,7,8-trihydroxy-7,8-dihydrobenzo[a]pyrene, which accounted for ~30% of the total metabolites formed by microsomes from control and phenobarbital-pretreated rats whereas this metabolite represents ~5% of the total metabolites with microsomes from 3-methylcholanthrene-treated rats. With benzo[a]pyrene as substrate, liver microsomes produced the 4,5-, 7,8- and 9,10-dihydrodiol with high optical purity (>85%), and diol epoxides were also formed. Most of the optical activity in the BP 7,8-dihydrodiol was due to metabolism by the monoxygenase system rather than by epoxide hydrase, since hydration of (±)-benzo[a]pyrene 7,8-oxide by liver microsomes produced dihydrodiol which was only 8% optically pure. Thus, the stereospecificity of both the monoxygenase system and, to a lesser extent, epoxide hydrase plays important roles in the metabolic activation of benzo[a]pyrene to carcinogens and mutagens.  相似文献   

4.
A dual-label HPLC assay to measure femtomole quantities of ethyl acetate-extractable [3H]benzo[a]pyrene metabolites was developed. 14C-labeled metabolites of benzo[a]pyrene formed by rat liver 9000g supernatant were used as both internal standards and chromatographic markers. The percentage deviation between assays was determined to be between 11 and 13% for 9,10-dihydro-9,10-dihydroxybenzo[a]pyrene, 7,8-dihydro-7,8-dihydroxybenzo[a]pyrene, benzo[a]pyrene-3,6-quinone, benzo[a]pyrene-1,6-quinone, and 9-hydroxybenzo[a]pyrene, 22% for 4,5-dihydro-4,5-dihydroxybenzo[a]pyrene, and less than 5% for 3-hydroxybenzo[a]pyrene. The detection limit of this assay was between 3 and 10 fmol per metabolite. The application of this technique to the metabolism of [3H]benzo[a]pyrene by microsomes of hamster and human oral cavity tissue is described.  相似文献   

5.
The metabolism of benzo[a]pyrene in randomly proliferating and confluent cultures of human skin fibroblast cells was compared with cell cultures in early S phase of the cell cycle after a G1 block. When each cell population was exposed to [G-3H]benzo[a]pyrene for 24 hours and the organic soluble metabolites in the extracellular medium and intracellular components were analyzed by HPLC, a quantitative increase in metabolism was observed in the confluent cell populations. The amount of organic soluble metabolites in the extracellular medium of the confluent dense cultures was 2.7 times the amount found in randomly proliferating cultures and 1.5 times that of the synchronized cultures. The trans-7,8- and 9,10 dihydrodiols and 3-hydroxy benzo[a]pyrene were the major metabolites formed. Small amounts of the sulphate conjugate, 9-hydroxy-benzo[a]pyrene and the tetrols were also detected. Cytoplasmic as well as nuclear extracts from the confluent cell cultures also contained higher amounts of metabolites compared to those from the randomly proliferating and S-phase cells. The levels of DNA modification by metabolically activated benzo[a]pyrene did not differ among the randomly proliferating, confluent and S-phase cells. However, the S-phase cells exhibited approximately 50-fold increase in the frequency of transformation compared to the randomly proliferating cells. Confluent cells were not transformed by benzo[a]pyrene. These data suggest that factors other than random modification of DNA by the carcinogen might have a significant role in the expression of a transformed phenotype and that metabolism and transformation are not directly related. Furthermore, confluent dense cultures with a heightened capability for metabolism of benzo[a]pyrene were more active in the detoxification of benzo[a]pyrene than in the production of the metabolites associated with cellular transformation.Abbreviations BaP benzo[a]pyrene - BaP-4,5-diol trans-4,5 dihydroxy-4,5-dihydrobenzo[a]pyrene - BaP-7,8-diol trans-7,8-dihydroxy-7,8-dihydrobenzo[a]pyrene - Bap-9,10-diol trans-9,10-dihydroxy-9,10 dihydrobenzo[a]pyrene - CM complete medium - HNF human neonatal foreskin - HPLC high pressure liquid chromatography - PAH polycyclic aromatic hydrocarbon - PDL population doubling - RP randomly proliferating  相似文献   

6.
Human placental microsomes were incubated with [3H]benzo[a]pyrene (BP) and Salmon sperm DNA and the resulting metabolite-nucleoside complexes resolved by Sephadex LH-20 chromatography. The metabolite pattern was analyzed by high-pressure liquid chromatography (HPLC). The incubates were also co-chromatographed with extracts obtained from incubates with rat liver microsomes and [14C]BP. Phenols, quinones and 7,8-dihydrodiol were detected in the placental incubates. Both 9,10- and 4,5-dihydrodiols were very low as compared with control rat liver samples. Placental microsomes catalyzed the binding of BP metabolites to DNA in vitro, giving rise to two main complexes which co-chromatographed with rat liver-produced peaks attributable to 7,8-diol-9,10-epoxide and 7,8-oxide and/or quinones when metabolized further. The nucleoside metabolite peaks attributable to 4,5-oxide and 9-phenol-4,5-oxide were lacking when compared with the binding pattern catalyzed by rat liver. Both the total binding and specific metabolite-nucleoside adducts in the placenta correlated with fluorometrically measured aryl hydrocarbon hydroxylase (AHH) activity and with the amount of dihydrodiol formed. The results demonstrate that both the metabolite pattern and the nucleoside-metabolite complexes formed by the placental microsomes in vitro differed greatly from those produced by rat liver microsomes. These studies also suggest that it is not possible to predict specific patterns of DNA binding from AHH measurements or even from BP metabolite patterns, especially when comparing different tissues and species.  相似文献   

7.
The 7-methylbenzo[a]pyrene (7-MBaP) was incubated with liver microsomes of rats pretreated with polychlorinated biphenyls (Aroclor 1254) (PCBs). Metabolites of 7-MBaP were isolated by both reversed-phase and normal-phase high performance liquid chromatography (HPLC) and were characterized by nuclear magnetic resonance, UV-visible and mass spectral analyses. The predominant metabolite of 7-MBaP was found to be 3-hydroxy-7-methylbenzo[a]pyrene (3-hydroxy-7-MBaP). Other identified metabolites include 7-MBaP 4,5-, 7,8-, and 9,10-trans-dihydrodiols, 7-hydroxymethyl-BaP, 7-hydroxymethyl-BaP trans-9,10-dihydrodiol, 9-hydroxy-7-MBaP, 3-hydroxy-7-hydroxymethyl-BaP, 7-MBaP 1,6- and 3,6- quinones, and a hydroquinone which is also formed by further metabolism of the 3-hydroxy-7-MBaP. Comparative metabolic studies of 7-MBaP and BaP indicated that, relative to that of BaP, the methyl substituent of 7-MBaP slightly increases the formation of 3-hydroxy-7-MBaP and decreases the metabolism at other regions of the 7-MBaP molecule. The finding that a 7,8-dihydrodiol is a metabolite indicates that, like BaP, 7-MBaP may also be activated to the potentially reactive 7,8-dihydrodiol 9,10-epoxides although their formations are significantly reduced.  相似文献   

8.
1. Liver microsomes from rats were considerably more active in metabolizing benzo[f]quinoline (B f Q) than those from brown bullheads (Ictalurus nebulosus). 2. The main B f Q metabolites formed by both rat and brown bullhead liver microsomes were qualitatively similar and included B f Q-7,8-dihydrodiol, B f Q-9,10-dihydrodiol, B f Q-N-oxide, 7-hydroxy B f Q, and 9-hydroxy B f Q. 3. The liver microsomes from control brown bullheads and rats metabolized B f Q primarily at the 7,8-and 9,10-positions, respectively, whereas in the case of microsomes from 3-methylcholanthrene (3-MC)-treated rats or brown bullheads, the major site of metabolic attack was the 7,8-position. 4. A 3-MC-type of cytochrome P-450 appears to be primarily responsible for the oxidation of B f Q by control brown bullhead liver microsomes, whereas a phenobarbital-inducible type of cytochrome P-450 seems to be involved in the metabolism of B f Q by control rat liver microsomes.  相似文献   

9.
The reversible DNA physical binding of a series of non-reactive metabolites and metabolite model compounds derived from benzo[a]pyrene (BP) has been examined in UV absorption and in fluorescence emission and fluorescence lifetime studies. Members of this series have steric and pi electronic properties similar to the highly carcinogenic metabolite trans-7,8-dihydroxy-anti-9,10-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene (BPDE) and the less potent metabolite 4,5-epoxy-4,5-dihydrobenzo(a)pyrene (4,5-BPE). The molecules examined are trans-7,8-dihydroxy-7,8-dihydrobenzo[a]-pyrene (7,8-di(OH)H2BP), 7,8,9,10-tetrahydroxytetrahydrobenzo[a]pyrene (tetrol) 7,8,9,10-tetrahydrobenzo[a]pyrene (7,8,9,10-H4BP), pyrene, trans-4,5-dihydroxy-4,5-dihydrobenzo[a]pyrene (4,5-di(OH)H2BP) and 4,5-dihydrobenzo[a]pyrene (4,5-H2BP). In 15% methanol at 23 degrees C the intercalation binding constants of the molecules studied lie in the range 0.79-6.1 X 10(3) M-1. Of all the molecules examined the proximate carcinogen 7,8-di(OH)-H2BP is the best intercalating agent. The proximate carcinogen has a binding constant which in UV absorption studies is found to be 2.8-6.0 times greater than that of the other hydroxylated metabolites. Intercalation is the major mode of binding for 7,8-di(OH)H2BP and accounts for more than 95% of the total binding. Details concerning the specific role of physical bonding in BP carcinogenesis remain to be elucidated. However, the present studies demonstrate that the reversible binding constants for BP metabolites are of the same magnitude as reversible binding constants which arise from naturally occurring base-base hydrogen bonding and pi stacking interactions in DNA. Furthermore, previous autoradiographic studies indicate that in human skin fibroblasts incubated in BP, pooling of the unmetabolized hydrocarbons occurs at the nucleus. The high affinity of 7,8-di(OH)H2BP for DNA may play a role in similarly elevating in vivo nuclear concentrations of the non-reactive proximate carcinogen.  相似文献   

10.
When [3H]benzo[a]pyrene is incubated in vitro together with DNA, NADPH and rat lung microsomes, covalent binding of benzo[a]pyrene (BP) metabolites to DNA occurs. These metabolite-nucleoside complexes can be resolved into several distinct peaks by elution of a Sephadex LH-20 column with a water-methanol gradient. 3-Methylcholanthrene (MC) pretreatment of animals induces the total covalent binding in vitro several-fold and increases the amounts of at least five metabolite-nucleoside complexes associated with the 7,8-diol-9,10-epoxidcs, the 7,8-oxide or quinones oxygenated further, the 4,5-oxide and phenols oxygenated further. These increases correspond well with the increases in the production of both non-K-region and K-region metabolites of BP by lung microsomes, as determined by highpressure liquid chromatography (HPLC). On the other hand, when [3H]BP is metabolized in isolated perfused rat lung, only the peak representing the 7,8-diol-9,10-epoxide bound to nucleoside(s) is readily detectable and then only in lungs from MC-treated animals. The extent of binding of BP metabolites to lung DNA is very low, about 0.0004% of the total dose applied to the perfusion medium; more than 60% of this can be accounted for by the binding of the 7,8-diol-9,10-epoxides to nucleoside(s). It is suggested that the further metabolism leading to metabolites not available to covalent binding, (e.g. conjugation) of primary BP metabolites in the intact tissue is responsible for the differences in the metabolite-nucleoside patterns observed in vivo, as compared with microsomal metabolism in vitro.  相似文献   

11.
The stereoselectivity of the oxidation of 7,8-dihydrobenzo[a]pyrene (H2BP) to 9,10-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene (H4BP-epoxide) by prostaglandin H (PGH) synthase and cytochrome P-450 has been studied using microsomal preparations from ram seminal vesicles and rat liver. Incubations were performed in the presence of polyguanylic acid and the adducts formed between H4BP-epoxide and guanosine were isolated following the recovery and hydrolysis of the poly(G). When (+/-)-H4BP-epoxide was reacted with poly(G), four diastereomeric adducts were formed by the cis and trans addition of the exocyclic amino group of guanine to the benzylic carbon of the epoxide enantiomers. Each diastereomer was identified by a combination of ultraviolet, nuclear magnetic resonance, circular dichroism, and mass spectroscopy. Under comparable conditions, ram seminal vesicle microsomes in the presence of arachidonic acid triggered the binding of H2BP to poly(G) to a greater extent than rat liver microsomes from untreated and phenobarbital- and methylcholanthrene pretreated animals in the presence of NADPH. Quantitation of the (-)-cis- and (+)-cis-guanosine adducts revealed the degree of stereoselectivity of epoxidation. The ratio of (-)/(+) adducts was 54:46 for PGH synthase and 89:11 (control), 62:38 (phenobarbital), and 69:31 (methylcholanthrene) for cytochrome P-450-catalyzed reactions. PGH synthase catalyzed the epoxidation of H2BP with little or no stereoselectivity in contrast to cytochrome P-450. The utility of the poly(G) binding technique for the elucidation of the stereoselective generation of chiral electrophiles is discussed along with the mechanistic implications of the results.  相似文献   

12.
A human liver cytochrome P-450 (P-450) IIIA4 cDNA clone was inserted behind an alcohol dehydrogenase promoter in the plasmid vector pAAH5 and expressed in Saccharomyces cerevisiae (D12 and AH22 strains). A cytochrome P-450 with typical spectral properties was expressed at a level of approximately 8 x 10(5) molecules/cell in either strain of yeast. The expressed P-450 IIIA4 had the same apparent monomeric Mr as the corresponding protein in human liver microsomes (P-450NF) and could be isolated from yeast microsomes. Catalytic activity of the yeast microsomes toward putative P-450 IIIA4 substrates was seen in the reactions supported by cumene hydroperoxide but was often lower and variable when supported by the physiological donor NADPH. The catalytic activity of purified P-450 IIIA4 was also poor in some systems reconstituted with rabbit liver NADPH-P-450 reductase and best when both the detergent 3-[(3-cholamidopropyl)dimethylammonio]-1-propanesulfonate and a lipid extract (from liver or yeast microsomes) or L-alpha-1,2-dilauroyl-sn-glycero-3-phosphocholine were present. Under these conditions the expressed P-450 IIIA4 was an efficient catalyst for nifedipine oxidation, 6 beta-hydroxylation of testosterone and cortisol, 2-hydroxylation of 17 beta-estradiol and 17 alpha-ethynylestradiol, N-oxygenation and 3-hydroxylation of quinidine, 16 alpha-hydroxylation of dehydroepiandrosterone 3-sulfate, erythromycin N-demethylation, the 10-hydroxylation of (R)-warfarin, the formation of 9,10-dehydrowarfarin from (S)-warfarin, and the activation of aflatoxins B1 and G1, sterigmatocystin, 7,8-dihydroxy-7,8-dihydrobenzo[a]pyrene (both + and - diastereomers), 3,4-dihydroxy-3,4-dihydrobenz[a]anthracene, 3,4-dihydroxy-3,4-dihydro-7, 12-dimethylbenz[a]anthracene, 9,10-dihydroxy-9,10-dihydrobenzo[b]fluoranthene, 6-aminochrysene, and tris(2,3-dibromopropyl) phosphate to products genotoxic in a Salmonella typhimurium TA1535/pSK1002 system where a chimeric umuC' 'lacZ plasmid is responsive to DNA alkylation. Reaction rates were stimulated by 7,8-benzoflavone and inhibited by rabbit anti-P-450 IIIA (anti-P-450NF), troleandomycin, gestodene, and cimetidine. Evidence was obtained that rates of reduction of ferric P-450 IIIA4 in yeast microsomes and the reconstituted systems are slow and at least partially responsible for the lower rates of catalysis seen in these systems (relative to liver microsomes). The results of these studies with a defined protein clearly demonstrate the ability of P-450 IIIA4 to catalyze regio- and stereoselective oxidations with a diverse group of substrates, and this enzyme appears to be one of the most versatile catalysts in the P-450 family.  相似文献   

13.
Incubation of 15-hydroperoxy-5,8,11,13-eicosatetraenoic acid with ram seminal vesicle microsomes (RSVM) triggers the oxygenation of trans-7,8-dihydroxy-7,8-dihydrobenzo[a]pyrene (BP-7,8-diol). The principal oxidation products are 7,8,9,10-tetrahydroxy-7,8,9,10-tetrahydrobenzo[a]pyrenes which are non-enzymatic hydrolysis products of r-7,t-8-dihydroxy-t-9,10-oxy-7,8,9,10-tetrahydrobenzo[a]pyrene. At short incubation times, an additional product is isolated which is identified as r-7,t-8,t-9-trihydroxy-c-10-methoxy-7,8,9,10-tetrahydrobenzo[a]pyrene. This product appears to arise by solvolysis of the extracted diolepoxide during high performance liquid chromatography using methanol-water solvent systems. The incubation of 18O-labeled 15-hydroperoxy-5,8,11,13-eicosatetraenoic acid with BP-7,8-diol and RSVM leads to very little incorporation of 18O into the stable solvolysis products (analyzed by gc-ms of their peracetates). Parallel incubations conducted with 16O-labeled hydroperoxide under an 18O2 atmosphere indicate that the principle source of the epoxide oxygen is molecular oxygen.  相似文献   

14.
The ability of rat liver microsomes to catalyze the formation of benzo(a)pyrene 7,8-diol-9,10-epoxide — DNA nucleoside adduct was increased threefold by feeding 0.5% ethoxyquin to the animals. Microsomal epoxide hydratase activity was enhanced i parallel by a factor of 3 while aryl hydrocarbon hydroxylase activity was not induced. Liver microsomes from rat pretreated with 3-methylcholanthrene produced an increased proportion of diol epoxide — DNA adduct when ethoxyquin had been fed to the animals. The main chromatographic peak formed by microsomes from 3-methylcholanthrene treated rats which contains DNA adducts of secondary benzo(a)pyrene phenol metabolites is reduced when the animals had received ethoxyquin.  相似文献   

15.
Neuromedin B: a novel bombesin-like peptide identified in porcine spinal cord   总被引:19,自引:0,他引:19  
Metabolism of 1-nitrobenzo(a)pyrene (1-nitro-BaP) by rat liver microsomes yielded 1-nitro-BaP trans-7,8-dihydrodiol, 1-nitro-BaP trans-9,10-dihydrodiol and 1-nitro-BaP 7,8,9,10-tetrahydrotetrol. Formation of these metabolites suggests that a vicinal 7,8,9,10-dihydrodiol-epoxide is a metabolite of 1-nitro-BaP.  相似文献   

16.
The syn- and anti-diastereoisomers of trans-7,8-dihydroxy-9,10-oxy-7,8,9,10-tetrahydrobenzo[a]pyrene (BPDE) were further metabolized by rat liver microsomes obtained from 3-methylcholanthrene(MC)-pretreated rats and NADPH to reactive intermediates, presumably 1,7,8- and 3,7,8-trihydroxy-9,10-oxy-7,8,9,10-tetrahydrobenzo[a]pyrenes (triol-epoxides), that bound to macromolecules or decomposed to products consistent with pentahydroxy derivatives of benzo[a]pyrene (BP-pentols). Three major metabolites of syn-BPDE and four major metabolites of anti-BPDE were isolated by high performance liquid chromatography and characterized by spectroscopic techniques. When fluorescence spectroscopy was employed all metabolites exhibited very similar spectral properties and showed substantial shifts in excitation and emission maxima to longer wavelengths when measured under alkaline conditions, consistent with the presence of a phenolic hydroxyl group. Furthermore, the spectral properties of the metabolites from syn- and anti-BPDE were similar to those of 1-hydroxypyrene. Previous data from this laboratory together with the data presented in this study thus strongly suggest that further metabolism of BPDE involves hydroxylation at the 1- and 3-positions to yield the corresponding triol-epoxides and various BP-pentols. The pentols could also be formed by incubating tetrols derived from syn- and anti-BPDE with microsomes and NADPH. However, the rate of formation of pentols from the BP-tetrols was much slower than the rate of further metabolism of BPDE. Accordingly, the major route of BP-pentol formation is likely to be via the intermediate formation of triol-epoxides. Isolated liver nuclei from MC-pretreated rats were also found to catalyze the activation of anti-BPDE in presence of NADPH to reactive intermediates. This resulted in a substantial increase in binding to histone and non-histone proteins, with a concomitant decrease in binding to DNA. No qualitative change in the distribution of DNA-bound products of anti-BPDE could be demonstrated as a result of the further metabolism of anti-BPDE.  相似文献   

17.
The effects of three aryl acetylenes, 1-ethynylpyrene (EP), 2-ethynylnaphthalene (EN) and 3-ethynylperylene (EPE), upon the metabolism of benzo[a]pyrene (BaP) by microsomes isolated from rat liver were investigated. These aryl acetylenes all inhibited the total metabolism of BaP. Formation of BaP 7,8-dihydrodiol and BaP tetrol products by microsomal preparations from rats that had been pretreated with 3-methylcholanthrene (3MC) were preferentially inhibited. The effects of EP upon the metabolism of BaP 7,8-dihydrodiol by microsomes from rat liver were also studied. This aryl acetylene strongly inhibited the formation of BaP tetrols from BaP 7,8-dihydrodiol by liver microsomes both from untreated rats and from rats pretreated with 3MC, but enhanced the conversion of the BaP dihydrodiol into other metabolites.  相似文献   

18.
The effect of various co- and anti-carcinogens of colon carcinogenesis on the metabolism of benzo(a)pyrene (BP) in cultured rat colon is reported. Rat colon enzymatically converted BP into metabolites which bind to cellular macromolecules i.e., DNA and protein. Activity of aryl hydrocarbon hydroxylase (AHH) activity and binding levels of BP to macromolecules were higher in the descending colon when compared to other segments. The major metabolites of BP, extractable with ethylacetate, were quinones, tetrols, 7,8-diol and a peak containing 9,10-dihydroxy-9,10-dihydrobenzo(a)pyrene and 7,8,9-trihydroxy-7,8-dihydrobenzo(a)pyrene. The binding levels of BP to DNA and protein in the explant was lowered by co-incubation with 7,8-benzoflavone (7,8-BF) (3.6 and 18.0 μM), a known inhibitor of AHH, and with disulfiram (100 μM), an anti-oxidant. The absence of vitamin A in the media also resulted in a lower level of BP binding to DNA and protein and in lower activity of AHH. Pretreatment with known inducers of AHH such as phenobarbital (PB) or benz(a)anthracene (BA), did not have any significant effect on the binding levels of BP to DNA or on the AHH activity. of the bile acids investigated only taurodeoxycholic acid significantly increased the binding level of BP to DNA.  相似文献   

19.
The metabolism of benzo[a]pyrene by halogenated biphenyl-induced rat hepatic microsomal monooxygenases was determined using a high pressure liquid chromatographic assay system. Incubation of benzo[a]pyrene with microsomes from rats pretreated with phenobarbitone or phenobarbitone-type inducers (2,2',4,4',5,5'-hexachlorobiphenyl, 2,2',4,4',6,6'-hexachlorobiphenyl, 2,2',5,5'-tetrachlorobiphenyl, 2,2',4,4',5,5'-hexabromobiphenyl, and 2,2',5,5'-tetrabromobiphenyl) resulted in increased overall metabolism of the hydrocarbon (less than fourfold) into phenolic, quinone, and diol metabolites, with the most striking increase observed in the formation of 4,5-dihydro-4,5-dihydroxybenzo[a]pyrene. In contrast, the metabolism of benzo[a]pyrene by microsomes from rats induced with 3-methylcholanthrene or 3,3',4,4'-tetrachlorobiphenyl resulted in a greater than 10-fold increase in overall benzo[a]pyrene metabolism, with the largest increases observed in the formation of the trans-7,8- and -9,10-dihydrodiol metabolites of benzo[a]pyrene. However, in comparison to control and phenobarbitone-induced microsomes, the oxidative conversion of benzo[a]pyrene by microsomes induced with 3-methylcholanthrene and 3,3',4,4'-tetrachlorobiphenyl into the 6,12-quinone was substantially inhibited. Previous reports have shown that the commercial halogenated biphenyl mixtures, fireMaster BP-6, and Aroclor 1254 are mixed-type inducers and that microsomes from rats pretreated with these mixtures markedly enhance the overall metabolism of benzo[a]pyrene. Not surprisingly, the metabolism of benzo[a]pyrene by microsomes from rats pretreated with the mixed-type inducers, 2,3,3',4,4'-penta-,2,3,3',4,4',5-hexa-, and 2',3,3',4,4',5-hexa- chlorobiphenyl was also increased and the metabolic profile was similar to that observed with fireMaster BP-6 and Aroclor 1254 induced microsomes.  相似文献   

20.
Benzo(a)pyrene derivatives (1-, 2-, 3-, 7-, and 9-hydroxy-benzo(a)pyrene and trans-9,10-dihydro-9,10-dihydroxy-, -4,5-dihydro-4,5-dihydroxy-, and -7,8-dihydro-7,8-dihydroxy-benzo(a)pyrene) were metabolized by liver microsomes isolated from 3-methylcholanthrene-treated rats in the presence of calf thymus DNA. The isolated DNA was then assayed by fluorescence for bound metabolic products. Only 2-hydroxy-benzo(a)pyrene, 9-hydroxy-benzo(a)pyrene and trans-7,8-dihydro-7,8-dihydroxy-benzo(a)pyrene yielded detectable amounts of DNA-bound products. In contrast to the product(s) from 9-hydroxy-benzo(a)pyrene, the metabolites of 2-hydroxy-benzo(a)pyrene and trans-7,8-dihydro-7,8-dihydroxy-benzo(a)pyrene, both strong carcinogens, had similar excitation spectra and gave considerably increased fluorescence intensities when the DNA was denatured. These data indicate structural similarities in the DNA complexes formed after metabolic activation of 2-hydroxy-benzo(a)pyrene and trans-7,8-dihydro-7,8-dihydroxy-benzo(a)pyrene.  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