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.
Metabolism of phenanthrene by Phanerochaete chrysosporium.   总被引:10,自引:8,他引:2       下载免费PDF全文
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.
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   

3.
Three filamentous fungi were examined for the ability to biotransform phenanthrene to oxidative (phase I) and conjugative (phase II) metabolites. Phenanthrene metabolites were purified by high-performance liquid chromatography (HPLC) and identified by UV/visible absorption, mass, and1H NMR spectra.Aspergillus niger ATCC 6275,Syncephalastrum racemosum UT-70, andCunninghamella elegans ATCC 9245 initially transformed [9-14C]phenanthrene to produce metabolites at the 9,10-, 1,2-, and 3,4- positions. Subsequently, sulfate conjugates of phase I metabolites were formed byA. niger, S. racemosum, andC. elegans. Minor glucuronide conjugates of 9-phenanthrol and phenanthrenetrans-9,10-dihydrodiol were formed byS. racemosum andA. niger, respectively. In addition,C. elegans produced the glucose conjugates 1-phenanthryl -d-glucopyranoside and 2-hydroxy-1-phenanthryl -d-glucopyranoside, a novel metabolite. [9-14C]Phenanthrene metabolites were not detected in organic extracts from biotransformation experiments with the yeasts,Candida lipolytica 37-1,Candida tropicalis ATCC 32113, andCandida maltosa R-42.  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 fungal metabolism of 7-methylbenz[a]anthracene (7-MBA) and 7-hydroxymethylbenz[a]anthracene (7-OHMBA) was studied. 7-MBA was metabolized by Cunninghamella elegans to form 7-OHMBA-trans-8,9-dihydrodiol and 7-OHMBA-trans-3,4-dihydrodiol as the predominant metabolites. Other metabolites were identified as 7-OHMBA, 7-MBA-trans-8,9-dihydrodiol and 7-MBA-trans-3,4-dihydrodiol, and 7-MBA-8,9,10,11-tetraol. Incubation of 7-OHMBA with C. elegans cells indicated that 7-OHMBA-trans-8,9-dihydrodiol and 7-OHMBA-trans-3,4-dihydrodiol were major metabolites. The metabolism of 7-MBA by rat liver microsomes from 3-methylcholanthrene-treated rats showed that the metabolites were qualitatively similar to those formed by C. elegans, except additional dihydrodiol metabolites were formed at the 5,6 and 10,11 positions. The metabolites formed were isolated by high-performance liquid chromatography and identified by comparing their chromatographic, UV-visible absorption and mass spectral properties with those of reference compounds.  相似文献   

6.
For the first time the dependence of completeness of pyrene degradation by the white-rot fungus Pleurotus ostreatus D1 on cultivation conditions was found. In Kirk’s medium about 65.6 ± 0.9% of the initial pyrene was metabolized after 3 weeks, with pyrene-4,5-dihydrodiol accumulating. This process was accompanied by laccase production only. In basidiomycetes rich medium, P. ostreatus D1 metabolized up to 89.8 ± 2.3% of pyrene within 3 weeks without pyrene-4,5-dihydrodiol accumulation throughout the time of cultivation. Phenanthrene and phthalic acid were identified as the metabolites produced from pyrene degradation under these conditions. Accumulation of phenanthrene with its subsequent disappearance was observed. One more metabolite probably was the product of phenanthrene degradation. Pyrene metabolism in basidiomycetes rich medium was accompanied first by laccase and tyrosinase production and later by versatile peroxidase production. The cell-associated activities of laccase, tyrosinase, and versatile peroxidase were found. The data obtained indicate that both enzymes (laccase and versatile peroxidase) are necessary for complete degradation of pyrene. Furthermore, both cell-associated and extracellular laccases can catalyse the first stages of pyrene degradation, and versatile peroxidase can be necessary for oxidation of the resulting metabolites.  相似文献   

7.
Estuarine sediment slurries and microorganisms were examined for the ability to methylate inorganic tin. Under controlled redox conditions, tin was methylated only in oxygen-free sediment slurries. Monomethyltin usually comprised greater than 90% of the alkyltin products formed, although dimethyltin was also produced. Autoclaved anoxic sediments did not produce organotins. Several bacterial cultures, most notably sulfate-reducing bacteria isolated from anoxic estuarine sediments, formed monoand dimethyltin from inorganic tin in the absence of sediment. The results suggest that inorganic tin methylation in estuarine environments is an anaerobic process catalyzed primarily by sulfate-reducing microorganisms.  相似文献   

8.
Phenanthrene metabolites (phenols and dihydrodiols) and 1-hydroxypyrene excreted in the 24-h urine of smokers, non-smokers and lung cancer patients, who after heavy smoking became light smokers, were determined and compared. In contrast to 1- hydroxypyrene, no significant differences of the absolute amounts of phenanthrene metabolites were found between smokers and non-smokers. A ratio phenanthrene metabolites/l-hydroxypyrene of 10.4 was observed for non-smokers and 9.9 for lung cancer patients, but 4.2 for smokers. Significantly different ratios for the regiospecific oxidation of phenanthrene were found for smokers when compared with non-smokers (1,2-oxidation vs 3,4-oxidation was 1.45 in the case of smokers, but 2.34 in the case of non-smokers) indicating a cigarette smoke - but not PAH - caused induction of CYP 1A2 in smokers. As a consequence of the degree of PAH exposure the ratio dihydrodiols/phenols depends on the total amount of metabolites excreted. Phenols predominate, equally in smokers and non-smokers after low exposure, while dihydrodiols become more prominent in highly exposed persons (coke plant workers). Both (i) the regiospecific oxidation of PAH and (ii) the ratio of dihydrodiol vs phenol formation may be recognized from the urinary phenanthrene metabolite profile. This pattern mirrors the enzymatic status (balance of the CYP isoforms and epoxide hydrolase) in individuals. Accordingly, more detailed information may be obtained from the urinary metabolite pattern than from 1- hydroxypyrene, commonly used in PAH biomonitoring.  相似文献   

9.
Phenanthrene metabolites (phenols and dihydrodiols) and 1-hydroxypyrene excreted in the 24-h urine of smokers, non-smokers and lung cancer patients, who after heavy smoking became light smokers, were determined and compared. In contrast to 1- hydroxypyrene, no significant differences of the absolute amounts of phenanthrene metabolites were found between smokers and non-smokers. A ratio phenanthrene metabolites/l-hydroxypyrene of 10.4 was observed for non-smokers and 9.9 for lung cancer patients, but 4.2 for smokers. Significantly different ratios for the regiospecific oxidation of phenanthrene were found for smokers when compared with non-smokers (1,2-oxidation vs 3,4-oxidation was 1.45 in the case of smokers, but 2.34 in the case of non-smokers) indicating a cigarette smoke - but not PAH - caused induction of CYP 1A2 in smokers. As a consequence of the degree of PAH exposure the ratio dihydrodiols/phenols depends on the total amount of metabolites excreted. Phenols predominate, equally in smokers and non-smokers after low exposure, while dihydrodiols become more prominent in highly exposed persons (coke plant workers). Both (i) the regiospecific oxidation of PAH and (ii) the ratio of dihydrodiol vs phenol formation may be recognized from the urinary phenanthrene metabolite profile. This pattern mirrors the enzymatic status (balance of the CYP isoforms and epoxide hydrolase) in individuals. Accordingly, more detailed information may be obtained from the urinary metabolite pattern than from 1- hydroxypyrene, commonly used in PAH biomonitoring.  相似文献   

10.
Oxidative metabolism of the carcinogen 6-fluorobenzo[c]phenanthrene (6-FB[c]Ph) was compared with that of benzo[c]phenanthrene (B[c]Ph) to elucidate the enhancement of carcinogenicity of B[c]Ph by the 6-fluoro substituent. Liver microsomes from untreated (control), phenobarbital-treated, and 3-methylcholanthrene-treated rats metabolized 6-FB[c]Ph at rates of 3.5, 1.5, and 7.7 nmol of products/nmol of cytochrome P-450/min, respectively. The rates of metabolism of B[c]Ph by the same microsomes were 2.9, 1.6, and 5.5 nmol of products/nmol of cytochrome P-450/min, respectively. Whereas the K-region 5,6-dihydrodiol was the major metabolite of B[c]Ph, the major metabolite of 6-FB[c]Ph was the K-region 7,8-oxide, which underwent slow rearrangement to an oxepin. Thus, the 6-fluoro substituent blocks oxidation at the 5,6-double bond and inhibits hydration of the K-region 7,8-oxide by epoxide hydrolase. Substitution with fluorine at C-6 caused an almost 2.5-fold increase in the percentages of the putative proximate carcinogens, i.e. benzo-ring dihydrodiols with bay-region double bonds, when liver microsomes from 3-methylcholanthrene-treated rats were used. Little or no increase was observed in their formation by liver microsomes from control or phenobarbital-treated rats. Interestingly, liver microsomes from control rats formed almost 3-fold as much 3,4-dihydrodiol as isosteric 9,10-dihydrodiol. The R,R-enantiomers of the 3,4- and 9,10-dihydrodiols and the S,S-enantiomer of the 7,8-dihydrodiol were predominantly formed by all three microsomal preparations.  相似文献   

11.
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   

12.
Reductive dechlorination of 2,4-dichlorophenoxyacetic acid (2,4-D) and 2,4,5-trichlorophenoxyacetic acid (2,4,5-T) was investigated in anaerobic sediments by non-adapted microorganisms and by microorganisms adapted to either 2,4- or 3,4-dichlorophenol (DCP). The rate of dechlorination of 2,4-D was increased by adaptation of sediment microorganisms to 2,4-DCP while dechlorination by sediment microorganisms adapted to 3,4-DCP displayed a lag phase similar to non-adapted sediment slurries. Both 2,4- and 3,4-DCP-adapted microorganisms produced 4-chlorophenoxyacetic acid by ortho-chlorine removal. Lag phases prior to dechlorination of the initial addition of 2,4,5-T by DCP-adapted sediment microorganisms were comparable to those from non-adapted sediment slurries. However, the rates of dechlorination increased upon subsequent additions of 2,4,5-T. Biodegradation of 2,4,5-T by sediment microorganisms adapted to 2,4- and/ or 3,4-DCP produced 2,5-D as the initial intermediate followed by 3-chlorophenol and phenol indicating a para > ortho > meta order of dechlorination. Dechlorination of 2,4,5-T, by either adapted or non-adapted sediment microorganisms, progressed without detection of 2,4,5-trichlorophenol as an intermediate.  相似文献   

13.
A bacterial strain was isolated from a Mumbai coastal area. It was dosed with anthracene and phenanthrene, and, after 14 days of incubation, it had degraded 90% and 93% of the anthracene and phenanthrene, respectively. The metabolites were extracted and identified by ultraviolet (UV)-visible light absorption, high-performance liquid chromatography, mass spectrometry, and by comparing with actual compounds and data. Neutral extracts from anthracene showed four metabolites, viz 1,2-dihydroxyanthracene, 6,7-benzocoumarin, 1-methoxy-2-hydroxyanthracene, and 9,10 anthraquinone. When Pseudomonas were grown in the presence of phenanthrene, two metabolites, viz 9,10-dihydroxyphenanthrene and 3,4-dihydroxyphenanthrene were identified.  相似文献   

14.
Mycobacterium vanbaalenii PYR-1 is capable of degrading a number of polycyclic aromatic hydrocarbons (PAHs) to ring cleavage metabolites via multiple pathways. Genes for the large and small subunits of a pyrene dioxygenase, nidA and nidB, respectively, were previously identified in M. vanbaalenii PYR-1 [Appl. Environ. Microbiol. 67 (2001) 3577]. A library of the M. vanbaalenii PYR-1 genome was constructed in a fosmid vector to identify additional genes involved in PAH degradation. Twelve fosmid clones containing nidA were identified by Southern hybridization. Sequence analysis of one nidA-positive clone, pFOS608, revealed a number of additional genes involved in PAH degradation. At this locus, one putative operon contained genes involved in phthalate degradation, and another contained genes encoding a putative ABC transporter(s). A number of the genes found in this region are homologous to those involved in phenanthrene degradation via the phthalic acid pathway. The majority of phenanthrene degradation genes were located between putative transposase genes. In Escherichia coli, pFOS608 converted phenanthrene into phenanthrene cis-3,4-dihydrodiol, and converted 1-hydroxy-2-naphthoic acid into 2'-carboxybenzalpyruvate, 2-carboxybenzaldehyde, and phthalic acid. A subclone containing nidA and nidB converted phenanthrene into phenanthrene cis-3,4-dihydrodiol, suggesting that the NidAB dioxygenase is responsible for an initial attack on phenanthrene. This study is the first to identify genes responsible for the degradation of phenanthrene via the phthalic acid pathway in Mycobacterium species.  相似文献   

15.
Metabolism of (+)-, (-)-, and (+/-)-trans-3,4-dihydroxy-3, 4-dihydrobenzo[c]phenanthrenes by liver microsomes from rats and mice and by a purified monooxygenase system reconstituted with cytochrome P-450c has been examined. Bay-region 3,4-diol 1,2-epoxides are minor metabolites of both enantiomers of the 3,4-dihydrodiol with liver microsomes from 3-methylcholanthrene-treated rats or with the reconstituted system (less than 10% of total metabolites). Microsomes from control and phenobarbital-treated rats and from control mice form higher percentages of these diol epoxides (13-36% of total metabolites). Microsomes from 3-methylcholanthrene-treated rats and cytochrome P-450c in the reconstituted system form exclusively the diol expoxide-1 diastereomer, in which the benzylic hydroxyl group and oxirane oxygen are cis to each other, from the (+)-(3S,4S)-dihydrodiol. The same enzymes selectively form the diol expoxide-2 diastereomer, with its oxirane oxygen and benzylic hydroxyl groups trans to each other, from the (-)-(3R,4R)-dihydrodiol (77% of the total diol epoxides). Liver microsomes from control rats show similar stereoselectivity whereas liver microsomes from phenobarbital-treated rats and from control mice are less stereoselective. Three bis-dihydrodiols and three phenolic dihydrodiols are also formed from the enantiomeric 3,4-dihydrodiols of benzo[c]phenanthrene. A single diastereomer of one of these bis-dihydrodiols with the newly introduced dihydrodiol group at the 7,8-position accounts for 79-88% of the total metabolites of the (-)-(3R,4R)-dihydrodiol formed by liver microsomes from 3-methylcholanthrene-treated rats or by the reconstituted system containing epoxide hydrolase. In contrast, the (+)-(3S,4S)-dihydrodiol is metabolized to two diastereomers of this bis-dihydrodiol, a third bis-dihydrodiol, and two phenolic dihydrodiols.  相似文献   

16.
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   

17.
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   

18.
Cultures of Mycobacterium sp. strain PYR-1 were dosed with anthracene or phenanthrene and after 14 days of incubation had degraded 92 and 90% of the added anthracene and phenanthrene, respectively. The metabolites were extracted and identified by UV-visible light absorption, high-pressure liquid chromatography retention times, mass spectrometry, (1)H and (13)C nuclear magnetic resonance spectrometry, and comparison to authentic compounds and literature data. Neutral-pH ethyl acetate extracts from anthracene-incubated cells showed four metabolites, identified as cis-1,2-dihydroxy-1,2-dihydroanthracene, 6,7-benzocoumarin, 1-methoxy-2-hydroxyanthracene, and 9,10-anthraquinone. A novel anthracene ring fission product was isolated from acidified culture media and was identified as 3-(2-carboxyvinyl)naphthalene-2-carboxylic acid. 6,7-Benzocoumarin was also found in that extract. When Mycobacterium sp. strain PYR-1 was grown in the presence of phenanthrene, three neutral metabolites were identified as cis- and trans-9,10-dihydroxy-9,10-dihydrophenanthrene and cis-3,4-dihydroxy-3,4-dihydrophenanthrene. Phenanthrene ring fission products, isolated from acid extracts, were identified as 2,2'-diphenic acid, 1-hydroxynaphthoic acid, and phthalic acid. The data point to the existence, next to already known routes for both gram-negative and gram-positive bacteria, of alternative pathways that might be due to the presence of different dioxygenases or to a relaxed specificity of the same dioxygenase for initial attack on polycyclic aromatic hydrocarbons.  相似文献   

19.
Yeast abundance in the sediments of 13 coastal sites in Massachusetts was quantified, and the potential of yeast isolates to biotransform polycyclic aromatic hydrocarbons (PAHs) was determined. Plate counts of yeasts varied between 10(2) to 10(7) CFU g (dry weight) of sediment-1. The most abundant genera isolated and identified included Candida, Cryptococcus, Rhodotorula, Torulopsis, and Trichosporon. More than 50% of the isolates from heavily contaminated sites transformed phenanthrene, as determined by spray-plate screening. The plate counts of phenanthrene-transforming yeasts correlated significantly to the sediment concentrations of phenanthrene. Transformation of [9-14C]phenanthrene and [12-14C]benz[a]anthracene by individual isolates varied greatly, ranging from 0.15 to 8.15 mumol of PAH g-1 in 120-h incubations. Of the isolated yeasts, Trichosporon penicillatum exhibited the greatest capacity for phenanthrene transformation. The ability to transform PAHs appears to be widespread among yeasts in coastal sediments.  相似文献   

20.
Alfalfa (Medicago sativa L.) and other plants bearing an important root system have been shown to be effective in the removal of organic compounds, including polycyclic aromatic hydrocarbons (PAHs). Phenanthrene is one of the main contaminants arising from the petrochemical industry and is included in the USEPA's list of priority toxic pollutants. Hydroponic cultures of alfalfa were employed as a model system to evaluate their capability of removing phenanthrene and to study the plant-pollutant interaction without the interference of a soil matrix. The removal of phenanthrene was followed over a period of 30 days. The half-life of phenanthrene in hydroponics (initial concentration 50 mg L-1) was reduced 2.7 times when plants were present. Growth index, chlorophyll content of leaves, and peroxidase activity of the roots of plants exposed to phenanthrene were lower than the corresponding values of nonexposed plants. Phenanthrene produced an acute negative effect on the total bacterial counts but also caused an increase in degraders/total bacteria ratio. The Ames Salmonella plate incorporation assay was employed to screen for potential genotoxic metabolites, which could be generated by metabolic activation of the parent compound. None of the samples exhibited a positive response. While lack of a positive response to this test is not a definitive evidence of the absence of genotoxic substances, these results suggest that the plant-assisted removal of phenanthrene merits further investigation.  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