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 515 毫秒
1.
Metabolism of phenanthrene by Phanerochaete chrysosporium.   总被引:9,自引:0,他引:9  
The white rot fungus Phanerochaete chrysosporium metabolized phenanthrene when it was grown for 7 days at 37 degrees C in a medium containing malt extract, D-glucose, D-maltose, yeast extract, and Tween 80. After cultures were grown with [9-14C]phenanthrene, radioactive metabolites were extracted from the medium with ethyl acetate, separated by high-performance liquid chromatography, and detected by liquid scintillation counting. Metabolites from cultures grown with unlabeled phenanthrene were identified as phenanthrene trans-9,10-dihydrodiol, phenanthrene trans-3,4-dihydrodiol, 9-phenanthrol, 3-phenanthrol, 4-phenanthrol, and the novel conjugate 9-phenanthryl beta-D-glucopyranoside. Identification of the compounds was based on their UV absorption, mass, and nuclear magnetic resonance spectra. Since lignin peroxidase was not detected in the culture medium, these results suggest the involvement of monooxygenase and epoxide hydrolase activity in the initial oxidation and hydration of phenanthrene by P. chrysosporium.  相似文献   

2.
Under photoautotrophic growth conditions, the marine cyanobacterium Agmenellum quadruplicatum PR-6 metabolized phenanthrene to form trans-9,10-dihydroxy-9,10-dihydrophenanthrene (phenanthrene trans-9,10-dihydrodiol) and 1-methoxyphenanthrene as the major ethyl acetate-extractable metabolites. Small amounts of phenanthrols were also formed. The metabolites were purified by high-pressure liquid chromatography and identified from their UV, infrared, mass, and proton magnetic resonance spectral properties. A. quadruplicatum PR-6 formed phenanthrene trans-9,10-dihydrodiol with a 22% enantiomeric excess of the (-)-9S,10S-enantiomer. Incorporation experiments with 18O2 showed that one atom of oxygen from O2 was incorporated into the dihydrodiol. Toxicity studies, using an algal lawn bioassay, indicated that 9-phenanthrol and 9,10-phenanthrenequinone inhibit the growth of A. quadruplicatum PR-6.  相似文献   

3.
Under photoautotrophic growth conditions, the marine cyanobacterium Agmenellum quadruplicatum PR-6 metabolized phenanthrene to form trans-9,10-dihydroxy-9,10-dihydrophenanthrene (phenanthrene trans-9,10-dihydrodiol) and 1-methoxyphenanthrene as the major ethyl acetate-extractable metabolites. Small amounts of phenanthrols were also formed. The metabolites were purified by high-pressure liquid chromatography and identified from their UV, infrared, mass, and proton magnetic resonance spectral properties. A. quadruplicatum PR-6 formed phenanthrene trans-9,10-dihydrodiol with a 22% enantiomeric excess of the (-)-9S,10S-enantiomer. Incorporation experiments with 18O2 showed that one atom of oxygen from O2 was incorporated into the dihydrodiol. Toxicity studies, using an algal lawn bioassay, indicated that 9-phenanthrol and 9,10-phenanthrenequinone inhibit the growth of A. quadruplicatum PR-6.  相似文献   

4.
The white rot fungus Pleurotus ostreatus, grown for 11 days in basidiomycetes rich medium containing [14C] phenanthrene, metabolized 94% of the phenanthrene added. Of the total radioactivity, 3% was oxidized to CO2. Approximately 52% of phenanthrene was metabolized to trans-9,10-dihydroxy-9,10-dihydrophenanthrene (phenanthrene trans-9,10-dihydrodiol) (28%), 2,2'-diphenic acid (17%), and unidentified metabolites (7%). Nonextractable metabolites accounted for 35% of the total radioactivity. The metabolites were extracted with ethyl acetate, separated by reversed-phase high-performance liquid chromatography, and characterized by 1H nuclear magnetic resonance, mass spectrometry, and UV spectroscopy analyses. 18O2-labeling experiments indicated that one atom of oxygen was incorporated into the phenanthrene trans-9,10-dihydrodiol. Circular dichroism spectra of the phenanthrene trans-9,10-dihydrodiol indicated that the absolute configuration of the predominant enantiomer was 9R,10R, which is different from that of the principal enantiomer produced by Phanerochaete chrysosporium. Significantly less phenanthrene trans-9,10-dihydrodiol was observed in incubations with the cytochrome P-450 inhibitor SKF 525-A (77% decrease), 1-aminobenzotriazole (83% decrease), or fluoxetine (63% decrease). These experiments with cytochrome P-450 inhibitors and 18O2 labeling and the formation of phenanthrene trans-9R,10R-dihydrodiol as the predominant metabolite suggest that P. ostreatus initially oxidizes phenanthrene stereoselectively by a cytochrome P-450 monoxygenase and that this is followed by epoxide hydrolase-catalyzed hydration reactions.  相似文献   

5.
The metabolism of phenanthrene, a polycyclic aromatic hydrocarbon (PAH), by Streptomyces flavovirens was investigated. When grown for 72 h in tryptone yeast extract broth saturated with phenanthrene, the actinomycete oxidized 21.3% of the hydrocarbon at the K-region to form trans-9,10-dihydroxy-9,10-dihydrophenanthrene (phenanthrene trans-9,10-dihydrodiol). A trace of 9-phenanthrol was also detected. Metabolites isolated by thin-layer and high performance liquid chromatography were identified by comparing chromatographic, mass spectral, and nuclear magnetic resonance properties with those of authentic compounds. Experiments using [9-14C]phenanthrene showed that the trans-9,10-dihydrodiol had 62.8% of the radioactivity found in the metabolites. Circular dichroism spectra of the phenanthrene trans-9,10-dihydrodiol indicated that the absolute configuration of the predominant enantiomer was (–)-9S,10S, the same as that of the principal enantiomer produced by mammalian enzymes. Incubation of S. flavovirens with phenanthrene is an atmosphere of 18O2, followed by gas chromatographic/mass spectral analysis of the metabolites, indicated that one atom from molecular oxygen was incorporated into each molecule of the phenanthrene trans-9,10-dihydrodiol. Cytochrome P-450 was detected in 105,000×g supernatants prepared from cell extracts of S. flavovirens. The results show that the oxidation of phenanthrene by S. flavovirens was both regio- and stereospecific.Abbreviations CD circular dichroism - DMF N,N-dimethyl-formamide - GC/MS gas chromatography/mass spectrometry - HPLC high performance liquid chromatography - NMR nuclear magnetic resonance - ODS octadecylsilane - PAH polycyclic aromatic hydrocarbon - TLC thin-layer chromatography - TMS tetramethylsilane - UV ultraviolet  相似文献   

6.
THE RELATIVE ROLE OF EUKARYOTIC VERSUS PROKARYOTIC MICROORGANISMS IN PHENANTHRENE TRANSFORMATION WAS MEASURED IN SLURRIES OF COASTAL SEDIMENT BY TWO DIFFERENT APPROACHES: detection of marker metabolites and use of selective inhibitors on phenanthrene biotransformation. Phenanthrene biotransformation was measured by polar metabolite formation and CO(2) evolution from [9-C]phenanthrene. Radiolabeled metabolites were tentatively identified by high-performance liquid chromatography (HPLC) separation combined with UV/visible spectral analysis of HPLC peaks and comparison to authentic standards. Both yeasts and bacteria transformed phenanthrene in slurries of coastal sediment. Two products of phenanthrene oxidation by fungi, phenanthrene trans-3,4-dihydrodiol and 3-phenanthrol, were produced in yeast-inoculated sterile sediment. However, only products of phenanthrene oxidation typical of bacterial transformation, 1-hydroxy-2-naphthoic acid and phenanthrene cis-3,4-dihydrodiol, were isolated from slurries of coastal sediment with natural microbial populations. Phenanthrene trans-dihydrodiols or other products of fungal oxidation of phenanthrene were not detected in the slurry containing a natural microbial population. A predominant role for bacterial transformation of phenanthrene was also suggested from selective inhibitor experiments. Addition of streptomycin to slurries, at a concentration which suppressed bacterial viable counts and rates of [methyl-H]thymidine uptake, completely inhibited phenanthrene transformation. Treatment with colchicine, at a concentration which suppressed yeast viable counts, depressed phenanthrene transformation by 40%, and this was likely due to nontarget inhibition of bacterial activity. The relative contribution of eukaryotic microorganisms to phenanthrene transformation in inoculated sterile sediment was estimated to be less than 3% of the total activity. We conclude that the predominant degraders of phenanthrene in muddy coastal sediments are bacteria and not eukaryotic microorganisms.  相似文献   

7.
Mycobacterium strain S1, originally described as Rhodococcus strain S1 by chemotaxonomic criteria, was isolated by growth on anthracene, and is unable to use any of nine other polycyclic aromatic compounds as carbon source. Metabolism of phenanthrene during growth on anthracene as sole carbon source results in the accumulation of traces of a dihydrodiol metabolite in the growth medium, which, by comparison with authentic standards, has been tentatively identified as phenanthrene trans-9,10-dihydrodiol. Anthracene metabolites were ruled out on the basis of comparisons with authentic anthracene dihydrodiols from Pseudomonas fluorescens D1 and chemically synthesized anthrols. The original source of phenanthrene for dihydrodiol production was phenanthrene present as a < 1% contaminant in the anthracene used as carbon source. However, addition of further phenanthrene to the anthracene growth medium increased the level of phenanthrene trans-9,10-dihydrodiol formed. Mycobacterium strain S1 also produced phenanthrene trans-9,10-dihydrodiol when grown in a glucose-salts medium in the presence of phenanthrene. This dihydrodiol is a dead-end metabolite, and neither it nor its parent hydrocarbon are able to support the growth of Mycobacterium strain S1. Studies with metyrapone and ancimidol, which did not inhibit growth on anthracene but did inhibit formation of phenanthrene trans-9,10-dihydrodiol, suggest it is likely the product of a cytochrome P450 monooxygenase-like activity.  相似文献   

8.
The metabolism of the carcinogen benz[a]anthracene (BA), a tetracyclic aromatic hydrocarbon, by Cunninghamella elegans was investigated. C. elegans grown on Sabouraud dextrose broth transformed [14C]BA to labeled BA trans-8,9-dihydrodiol (90%), BA trans-10,11-dihydrodiol (6%), and BA trans-3,4-dihydrodiol (4%), but not to BA trans-5,6-dihydrodiol. These metabolites were separated by thin-layer chromatography and reversed-phase high-performance liquid chromatography and were identified by UV and mass spectral techniques. A BA tetraol, 8 beta,9 alpha,10 alpha,11 beta-tetrahydroxy-8 alpha, 9 beta,10 beta,11 alpha-tetrahydro-BA, was also identified as a metabolite and may have arisen as an additional oxidation product of either BA 8,9- or 10,11-dihydrodiol. This is the first study in which a biologically produced BA tetraol has been identified. Our results suggest that the transformation of BA to trans-dihydrodiols by C. elegans is similar to the transformation of BA found in mammals, except that BA 5,6-dihydrodiol is not produced.  相似文献   

9.
The enzymatic mechanisms involved in the degradation of phenanthrene by the white rot fungus Pleurotus ostreatus were examined. Phase I metabolism (cytochrome P-450 monooxygenase and epoxide hydrolase) and phase II conjugation (glutathione S-transferase, aryl sulfotransferase, UDP-glucuronosyltransferase, and UDP-glucosyltransferase) enzyme activities were determined for mycelial extracts of P. ostreatus. Cytochrome P-450 was detected in both cytosolic and microsomal fractions at 0.16 and 0.38 nmol min(sup-1) mg of protein(sup1), respectively. Both fractions oxidized [9,10-(sup14)C]phenanthrene to phenanthrene trans-9,10-dihydrodiol. The cytochrome P-450 inhibitors 1-aminobenzotriazole (0.1 mM), SKF-525A (proadifen, 0.1 mM), and carbon monoxide inhibited the cytosolic and microsomal P-450s differently. Cytosolic and microsomal epoxide hydrolase activities, with phenanthrene 9,10-oxide as the substrate, were similar, with specific activities of 0.50 and 0.41 nmol min(sup-1) mg of protein(sup-1), respectively. The epoxide hydrolase inhibitor cyclohexene oxide (5 mM) significantly inhibited the formation of phenanthrene trans-9,10-dihydrodiol in both fractions. The phase II enzyme 1-chloro-2,4-dinitrobenzene glutathione S-transferase was detected in the cytosolic fraction (4.16 nmol min(sup-1) mg of protein(sup-1)), whereas aryl adenosine-3(prm1)-phosphate-5(prm1)-phosphosulfate sulfotransferase (aryl PAPS sulfotransferase) UDP-glucuronosyltransferase, and UDP-glucosyltransferase had microsomal activities of 2.14, 4.25, and 4.21 nmol min(sup-1) mg of protein(sup-1), respectively, with low activity in the cytosolic fraction. However, when P. ostreatus culture broth incubated with phenanthrene was screened for phase II metabolites, no sulfate, glutathione, glucoside, or glucuronide conjugates of phenanthrene metabolites were detected. These experiments indicate the involvement of cytochrome P-450 monooxygenase and epoxide hydrolase in the initial phase I oxidation of phenanthrene to form phenanthrene trans-9,10-dihydrodiol. Laccase and manganese-independent peroxidase were not involved in the initial oxidation of phenanthrene. Although P. ostreatus had phase II xenobiotic metabolizing enzymes, conjugation reactions were not important for the elimination of hydroxylated phenanthrene.  相似文献   

10.
Aspergillus niger, isolated from hydrocarbon-contaminated soil, was examined for its potential to degrade phenanthrene and pyrene. Two novel metabolites, 1-methoxyphenanthrene and 1-methoxypyrene, were identified by conventional chemical techniques. Minor metabolites identified were 1- and 2-phenanthrol and 1-pyrenol. No 14CO2 evolution was observed in either [14C]phenanthrene or [14C]pyrene cultures.  相似文献   

11.
The presence of cytochrome P450 and P450-mediated phenanthrene oxidation in the white rot fungus Phanerochaete chrysosporium under ligninolytic condition was first demonstrated in this study. The carbon monoxide difference spectra indicated induction of P450 (130 pmol mg−1 in the microsomal fraction) by phenanthrene. The microsomal P450 degraded phenanthrene with a NADPH-dependent activity of 0.44 ± 0.02 min−1. One of major detectable metabolites of phenanthrene in the ligninolytic cultures and microsomal fractions was identified as phenanthrene trans-9,10-dihydrodiol. Piperonyl butoxide, a P450 inhibitor which had no effect on manganese peroxidase activity, significantly inhibited phenanthrene degradation and the trans-9,10-dihydrodiol formation in both intact cultures and microsomal fractions. Furthermore, phenanthrene was also efficiently degraded by the extracellular fraction with high manganese peroxidase activity. These results indicate important roles of both manganese peroxidase and cytochrome P450 in phenanthrene metabolism by ligninolytic P. chrysosporium.  相似文献   

12.
Phenanthrene degradation by Polyporus sp. S133, a new phenanthrene-degrading strain, was investigated in this work. The analysis of degradation was performed by calculation of the remaining phenanthrene by gas chromatography-mass spectrometry. When cells were grown in phenanthrene culture after 92 h, all but 200 and 250 mg/l of the phenanthrene had been degraded. New metabolic pathways of phenanthrene and a better understanding of the phenoloxidases and dioxygenase mechanism involved in degradation of phenanthrene were explored in this research. The mechanism of degradation was determined through identification of the several metabolites; 9,10-phenanthrenequinone, 2,2'-diphenic acid, salicylic acid, and catechol. 9,10-Oxidation and ring cleavage to give 9,10-phenanthrenequinone is the major fate of phenanthrene in ligninolytic Polyporus sp. S133. The identification of 2,2'-diphenic acid in culture extracts indicates that phenanthrene was initially attacked through dioxigenation at C9 and C10 to give cis-9,10-dihydrodiol. Dehydrogenation of phenanthrene-cis-9,10-dihydrodiol to produce the corresponding diol, followed by ortho-cleavage of the oxygenated ring, produced 2,2'-diphenic acid. Several enzymes (manganese peroxidase, lignin peroxidase, laccase, 1,2-dioxygenase, and 2,3-dioxygenase) produced by Polyporus sp. S133 was detected during the incubation. The highest level of activity was shown at 92 h of culture.  相似文献   

13.
The fungus Cunninghamella elegans oxidized anthracene and phenanthrene to form predominately trans-dihydrodiols. The metabolites were isolated by reversed-phase high-pressure liquid chromatography for structural and conformational analyses. Comparison of the circular dichroism spectrum of the fungal trans-1,2-dihydroxy-1,2-dihydroanthracene to that formed by rat liver microsomes indicated that the major enantiomer of the trans-1,2-dihydroxy-1,2-dihydroanthracene formed by C. elegans had an S,S absolute stereochemistry, which is opposite to the predominately 1R,2R dihydrodiol formed by rat liver microsomes. C. elegans oxidized phenanthrene primarily in the 1,2-positions to form trans-1,2-dihydroxy-1,2-dihydrophenanthrene. In addition, a minor amount of trans-3,4-dihydroxy-3,4-dihydrophenanthrene was detected. Metabolism at the K-region (9,10-positions) of phenanthrene was not detected. Comparison of the circular dichroism spectra of the phenanthrene trans-1,2- and trans-3,4-dihydrodiols formed by C. elegans to those formed by mammalian enzymes indicated that each of the dihydrodiols formed by C. elegans had an S,S absolute configuration. The results indicate that there are differences in both the regio- and stereoselective metabolism of anthracene and phenanthrene between the fungus C. elegans and rat liver microsomes.  相似文献   

14.
Fungal metabolism and detoxification of fluoranthene.   总被引:1,自引:0,他引:1  
Five metabolites produced by Cunninghamella elegans from fluoranthene (FA) in biotransformation studies were investigated for mutagenic activity towards Salmonella typhimurium TA100 and TA104. Whereas FA displayed positive, dose-related mutagenic responses in both tester strains in the presence of a rat liver homogenate fraction, 3-FA-beta-glucopyranoside, 3-(8-hydroxy-FA)-beta-glucopyranoside, FA trans-2,3-dihydrodiol, and 8-hydroxy-FA trans-2,3-dihydrodiol were negative. 9-Hydroxy-FA trans-2,3-dihydrodiol showed a weak positive response in S. typhimurium TA100. Mutagenicity assays performed with samples extracted at 24-h intervals during incubation of C. elegans with FA for 120 h showed that mutagenic activity decreased with time. Comparative studies with rat liver microsomes indicated that FA trans-2,3-dihydrodiol, the previously identified proximal mutagenic metabolite of FA, was the major metabolite. The circular dichroism spectrum of the rat liver microsomal FA trans-2,3-dihydrodiol indicated that it was optically active. In contrast, the circular dichroism spectrum of the fungal FA trans-2,3-dihydrodiol showed no optical activity. These results indicate that C. elegans has the potential to detoxify FA and that the stereochemistry of its trans-2,3-dihydrodiol metabolite reduces its mutagenic potential.  相似文献   

15.
The K-region 5,6-epoxides, formed in the metabolism of benzo[c]phenanthrene (BcPh) in the presence of an epoxide hydrolase inhibitor 3,3,3-trichloropropylene 1,2-oxide (TCPO) by liver microsomes from untreated, phenobarbital-treated, 3-methylcholanthrene-treated, and polychlorinated biphenyls (Aroclor 1254)-treated rats of the Sprague-Dawley and the Long-Evans strains, were found by chiral stationary phase high-performance liquid chromatography analyses to be enriched (58-72%) in the 5S, 6R enantiomer. In the absence of TCPO, the metabolically formed BcPh trans-5,6-dihydrodiol was enriched (78-86%) in the 5S,6S enantiomer. The major enantiomer of the BcPh 3,4-epoxide metabolite was found to be enriched in the 3S,4R enantiomer which undergoes racemization under the experimental conditions. The major enantiomer of the 5,6-dihydrodiol metabolite was elucidated by the exciton chirality circular dichroism (CD) method to have a 5S,6S absolute stereochemistry. Absolute configurations of enantiomeric methoxylation products derived from each of the two BcPh 5,6-epoxide enantiomers. Optically pure BcPh 5S,6R-epoxide was enzymatically hydrated exclusively at the C6 position to form an optically pure BcPh 5S,6S-dihydrodiol. However, optically pure BcPh 5R,6S-epoxide was hydrated at both C5 and C6 positions to form a BcPh trans-5,6-dihydrodiol with a (5S,6S):(5R,6R) enantiomer ratio of 32:68.  相似文献   

16.
The degradation of 7,12-dimethylbenz[a]anthracene (DMBA), a carcinogenic polycyclic aromatic hydrocarbon, by cultures of Mycobacterium vanbaalenii PYR-1 was studied. When M. vanbaalenii PYR-1 was grown in the presence of DMBA for 136 h, high-pressure liquid chromatography (HPLC) analysis showed the presence of four ethyl acetate-extractable compounds and unutilized substrate. Characterization of the metabolites by mass and nuclear magnetic resonance spectrometry indicated initial attack at the C-5 and C-6 positions and on the methyl group attached to C-7 of DMBA. The metabolites were identified as cis-5,6-dihydro-5,6-dihydroxy-7,12-dimethylbenz[a]anthracene (DMBA cis-5,6-dihydrodiol), trans-5,6-dihydro-5,6-dihydroxy-7,12-dimethylbenz[a]anthracene (DMBA trans-5,6-dihydrodiol), and 7-hydroxymethyl-12-methylbenz[a]anthracene, suggesting dioxygenation and monooxygenation reactions. Chiral stationary-phase HPLC analysis of the dihydrodiols showed that DMBA cis-5,6-dihydrodiol had 95% 5S,6R and 5% 5R,6S absolute stereochemistry. On the other hand, the DMBA trans-5,6-dihydrodiol was a 100% 5S,6S enantiomer. A minor photooxidation product, 7,12-epidioxy-7,12-dimethylbenz[a]anthracene, was also formed. The results demonstrate that M. vanbaalenii PYR-1 is highly regio- and stereoselective in the degradation of DMBA.  相似文献   

17.
Fungal metabolism and detoxification of fluoranthene.   总被引:6,自引:4,他引:2       下载免费PDF全文
Five metabolites produced by Cunninghamella elegans from fluoranthene (FA) in biotransformation studies were investigated for mutagenic activity towards Salmonella typhimurium TA100 and TA104. Whereas FA displayed positive, dose-related mutagenic responses in both tester strains in the presence of a rat liver homogenate fraction, 3-FA-beta-glucopyranoside, 3-(8-hydroxy-FA)-beta-glucopyranoside, FA trans-2,3-dihydrodiol, and 8-hydroxy-FA trans-2,3-dihydrodiol were negative. 9-Hydroxy-FA trans-2,3-dihydrodiol showed a weak positive response in S. typhimurium TA100. Mutagenicity assays performed with samples extracted at 24-h intervals during incubation of C. elegans with FA for 120 h showed that mutagenic activity decreased with time. Comparative studies with rat liver microsomes indicated that FA trans-2,3-dihydrodiol, the previously identified proximal mutagenic metabolite of FA, was the major metabolite. The circular dichroism spectrum of the rat liver microsomal FA trans-2,3-dihydrodiol indicated that it was optically active. In contrast, the circular dichroism spectrum of the fungal FA trans-2,3-dihydrodiol showed no optical activity. These results indicate that C. elegans has the potential to detoxify FA and that the stereochemistry of its trans-2,3-dihydrodiol metabolite reduces its mutagenic potential.  相似文献   

18.
The homogeneous 3 alpha-hydroxysteroid dehydrogenase (EC 1.1.1.50) of rat liver cytosol is indistinguishable from dihydrodiol dehydrogenase (trans-1,2-dihydrobenzene-1,2-diol dehydrogenase EC 1.3.1.20), Penning, T. M., Mukharji, I., Barrows, S., and Talalay, P. (1984) Biochem. J. 222, 601-611). Examination of the substrate specificity of the purified dehydrogenase for trans-dihydrodiol metabolites of polycyclic aromatic hydrocarbons indicates that the enzyme will catalyze the NAD(P)-dependent oxidation of trans-dihydrodiols of benzene, naphthalene, phenanthrene, chrysene, 5-methylchrysene, and benzo[a]pyrene under physiological conditions. Comparison of the utilization ratios Vmax/Km indicates that benzenedihydrodiol and the trans-1,2- and trans-7,8-dihydrodiols of 5-methylchrysene were most efficiently oxidized by the purified dehydrogenase, followed by the trans-7,8-dihydrodiol of benzo[a]pyrene and the trans-1,2-dihydrodiols of phenanthrene, chrysene, and naphthalene. The purified enzyme appears to display rigid regio-selectivity, since it will readily oxidize non-K-region trans-dihydrodiols but will not oxidize the K-region trans-dihydrodiols of phenanthrene and benzo[a]pyrene. The stereochemical course of enzymatic dehydrogenation was investigated by circular dichroism spectrometry. For the trans-1,2-dihydrodiols of benzene, naphthalene, phenanthrene, chrysene, and 5-methylchrysene, the dehydrogenase preferentially oxidized the (+)-[S,S]-isomer. Apparent inversion of this stereochemical preference occurred with the trans-7,8-dihydrodiol of 5-methylchrysene, as the (-)-enantiomer was preferentially oxidized. No change in the sign of the Cotton Effect was observed following oxidation of the racemic trans-7,8-dihydrodiol of benzo[a]pyrene, suggesting that both stereoisomers of this compound were substrates. Large-scale incubation of the [3H]-(+/-)-trans-7,8-dihydrodiol of benzo[a]pyrene with the purified dehydrogenase resulted in greater than 90% utilization of this potent proximate carcinogen, suggesting that the enzyme utilizes both the (-)-[R,R] and the (+)-[S,S]-stereoisomers, which confirms the circular dichroism result. These data show that dihydrodiol dehydrogenase displays the appropriate regio- and stereospecificity to catalyze the oxidation of both the major and minor non-K-region trans-dihydrodiols that arise from the microsomal metabolism of benzo[a]pyrene in vivo.  相似文献   

19.
Fungal transformation of fluoranthene.   总被引:8,自引:8,他引:0       下载免费PDF全文
The fungus Cunninghamella elegans ATCC 36112 metabolized approximately 80% of the 3-14C-labeled fluoranthene (FA) added within 72 h of incubation. C. elegans metabolized FA to trans-2,3-dihydroxy-2,3-dihydrofluoranthene (trans-2,3-dihydrodiol), 8- and 9-hydroxyfluoranthene trans-2,3-dihydrodiol, 3-fluoranthene-beta-glucopyranoside, and 3-(8-hydroxyfluoranthene)-beta-glucopyranoside. These metabolites were separated by thin-layer and reversed-phase high-performance liquid chromatography and identified by 1H nuclear magnetic resonance, UV, and mass spectral techniques. The major pathway involved hydroxylation to form a glucoside conjugate of 3-hydroxyfluoranthene and a glucoside conjugate of 3,8-dihydroxyfluoranthene which together accounted for 52% of the total ethyl acetate-soluble metabolites. C. elegans initially metabolized FA in the 2,3 position to form fluoranthene trans-2,3-dihydrodiol, which has previously been shown to be a biologically active compound in mammalian and bacterial genotoxicity tests. However, C. elegans formed predominantly glucoside conjugates of the phenolic derivatives of FA, which suggests that this fungus has the potential to detoxify FA.  相似文献   

20.
The fungus Cunninghamella elegans ATCC 36112 metabolized approximately 80% of the 3-14C-labeled fluoranthene (FA) added within 72 h of incubation. C. elegans metabolized FA to trans-2,3-dihydroxy-2,3-dihydrofluoranthene (trans-2,3-dihydrodiol), 8- and 9-hydroxyfluoranthene trans-2,3-dihydrodiol, 3-fluoranthene-beta-glucopyranoside, and 3-(8-hydroxyfluoranthene)-beta-glucopyranoside. These metabolites were separated by thin-layer and reversed-phase high-performance liquid chromatography and identified by 1H nuclear magnetic resonance, UV, and mass spectral techniques. The major pathway involved hydroxylation to form a glucoside conjugate of 3-hydroxyfluoranthene and a glucoside conjugate of 3,8-dihydroxyfluoranthene which together accounted for 52% of the total ethyl acetate-soluble metabolites. C. elegans initially metabolized FA in the 2,3 position to form fluoranthene trans-2,3-dihydrodiol, which has previously been shown to be a biologically active compound in mammalian and bacterial genotoxicity tests. However, C. elegans formed predominantly glucoside conjugates of the phenolic derivatives of FA, which suggests that this fungus has the potential to detoxify FA.  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