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.
When incubated with a 9,000 x g rat-liver supernatant, benzo(a)pyrene 7,8-diol and benz(a)anthracene 8,9-diol were more active than the parent hydrocarbons in inducing his+ revertant colonies of S. typhimurium TA 100. Benzo(a) pyrene 9,10-diol was less active than benzo(a)pyrene; the K-region diols, benz(a)anthracene 5,6-diol and benzo(a)pyrene 4,5-diol, were inactive. None of the diols was active when the cofactors for the microsomal mono-oxygenase were omitted. The diol-epoxides benzo(a)pyrene 7,8-diol 9,10-oxide, benz(a)anthracene 8,9-diol 10,11-oxide and 7-methylbenz(a)anthracene 8,9-diol 10,11-oxide and the K-region epoxides, benzo(a)pyrene 4,5-oxide and benz(a)anthracene 5,6-oxide, were mutagenic without further metabolism.  相似文献   

2.
The syntheses of 7,12-dimethylbenz[a]anthracene 5,6-oxide, 7-acetoxymethyl-12-methylbenz[a]anthracene 5,6-oxide and a product that appears to be mainly 7-hydroxymethyl-12-methylbenz[a]anthracene 5,6-oxide are described. The compounds readily rearranged to phenols in the presence of mineral acid, and 7,12-dimethylbenz[a]anthracene 5,6-oxide and its 7-hydroxymethyl derivative reacted slowly with water to yield trans-5,6-dihydro-5,6-dihydroxy-7,12-dimethylbenz[a] anthracene and trans-5,6-dihydro-5,6-dihydroxy-7-hydroxymethyl-12-methylbenz [a]anthracene respectively. Both epoxides were converted enzymically by rat liver microsomal fractions and homogenates into the related trans-dihydrodiols. The epoxides reacted chemically with GSH to form conjugates that were identical with the conjugates formed when the epoxides were incubated with rat liver homogenates. The GSH conjugates were more stable to acid than conjugates derived from other arene oxides. In the alkylation of 4-(p-nitrobenzyl)pyridine, 7,12-dimethyl-benz[a]anthracene 5,6-oxide was more active than the 5,6-oxides of 7-methylbenz[a]-anthracene and benz[a]anthracene.  相似文献   

3.
To facilitate predictions of the transport and fate of contaminants at future coal conversion facilities, rates of microbial transformation of polycyclic aromatic hydrocarbons were measured in stream water and sediment samples collected in the vicinity of a coal-coking treated wastewater discharge from November 1977 through August 1979. Six radiolabeled polycyclic aromatic hydrocarbons were incubated with sediment and water samples; 14CO2, cell-bound 14C, and polar transformation products were isolated and quantified. Whereas 14CO2 and bound 14C were major transformation products in sediment assays, soluble polar 14C dominated transformation in water samples. Mean rate constants (measured at 20°C) in sediments collected downstream from the effluent outfall were 7.8 × 10−2 h−1 (naphthalene), 1.6 × 10−2 h−1 (anthracene), and 3.3 × 10−3 h−1 [benz(a)anthracene], which corresponded to turnover times of 13, 62, and 300 h, respectively. No unequivocal evidence for transformation of benzo(a)pyrene or dibenz(a,h)anthracene was obtained. Only naphthalene and anthracene transformations were observed in water samples; rate constants were consistently 5- and 20-fold lower, respectively, than in the corresponding sediment samples. The measured rate constants for anthracene transformation in July 1978 sediment samples were not related to total heterotroph numbers. In late July 1978, the effluent was diverted from the primary study area; however, no differences were observed either in transformation rate constants or in the downstream/upstream sediment rate constant ratio. These results are consistent with the hypothesis that continuous inputs of polycyclic aromatic hydrocarbons result in an increased ability within a microbial community to utilize certain polycyclic aromatic hydrocarbons. However, because transformation rates remained elevated for more than 1 year after removal of the polycyclic aromatic hydrocarbon source, microbial communities may shift only slowly in response to changes in polycyclic aromatic hydrocarbon concentrations.  相似文献   

4.
The syntheses of 10,11-dihydrobenz[a]anthracene 8,9-oxide, benz[a]anthracene 8,9-oxide and 9-hydroxybenz[a]anthracene are described, together with those of a number of related compounds. The epoxides react both chemically and enzymically with water to yield the corresponding dihydrodiols and with reduced glutathione to form glutathione conjugates, and they react chemically with N-acetylcysteine to yield the corresponding mercapturic acids. 8,9-Dihydro-8,9-dihydroxybenz[a]anthracene, formed enzymically from benz[a]anthracene 8,9-oxide, was identical with a dihydrodiol formed when benz[a]anthracene was metabolized by rat liver homogenates. Similarly 10,11-dihydrobenz[a]anthracene 8,9-oxide yielded a dihydrodiol identical with the product formed when 10,11-dihydrobenz[a]anthracene was metabolized.  相似文献   

5.
The introduction of a polycyclic hydrocarbon such as benzo(alpha)pyrene (BP) into normal golden hamster embryo cell cultures results, in addition to cytotoxicity, in malignant cell transformation. Studies on the effect of different doses of BP on the normal cells showed that the frequency of transformed colonies was directly related to the dose of the carcinogen. Analysis of this dose-response curve suggests a one-event ("one-hit") response for transformation by this carcinogen. The one-event response for transformation by carcinogenic polycyclic hydrocarbons and the fact that these carcinogens bind to DNA in susceptible cells suggests that transformation can involve a single alteration in the genetic constitution of the treated cells. Carcinogens may, therefore, produce somatic mutations, some of which may involve the genes that control malignancy. Recently, considerable progress has been made in developing models for the study of chemical mutagenesis in mammalian cells. Using resistance to 8-azaguanine as a marker, positive correlations between mutagenicity and transformation were obtained with chemically reactive carcinogens such as N-acetoxy-N-2-fluorenyl-acetamide, N-methyl-N'-nitro-N-nitrosoguanidine and K-region epoxides of polycyclic hydrocarbons. However, no such correlations were obtained with the carcinogenic polycyclic hydrocarbons themselves, since the cell lines used in chemical mutagenesis do not metabolize these carcinogens. In order to obtain better correlations, we have developed a cell-mediated mutagenic assay with carcinogenic hydrocarbons in which Chinese hamster cells, which are susceptible for mutagenesis, were co-cultivated with lethally irradiated rodent cells that can metabolize these compounds. Using this cell mediated assay, we obtained mutagenesis with the carcinogenic hydrocarbons 7,12-dimethylbenz(alpha)anthracene (DMBA), BP, 3-methylcholanthrene and 7-methylbenz(alpha)anthracene; the most potent carcinogen, DMBA, gave the highest frequency of mutations. The polycyclic hydrocarbons, pyrene and benz(alpha)anthracene, which are not carcinogenic were also not mutagenic. We have therefore demonstrated a relationship between the carcinogenecity of polycyclic hydrocarbons and their mutagenicity in mammalian cells, without having to isolate their reative metabolic intermediates. It should be possible to use in this system human cells from different organs and individuals to screen for environmental chemicals hazardous to humans which have to be metabolically activated.  相似文献   

6.
7.
To determine rates of microbial transformation of polycyclic aromatic hydrocarbons (PAH) in freshwater sediments, 14C-labeled PAH were incubated with samples from both pristine and petroleum-contaminated streams. Evolved 14CO2 was trapped in KOH, unaltered PAH and polar metabolic intermediate fractions were quantitated after sediment extraction and column chromatography, and bound cellular 14C was measured in sediment residues. Large fractions of 14C were incorporated into microbial cellular material; therefore, measurement of rates of 14CO2 evolution alone would seriously underestimate transformation rates of [14C]naphthalene and [14C]anthracene. PAH compound turnover times in petroleum-contaminated sediment increased from 7.1 h for naphthalene to 400 h for anthracene, 10,000 h for benz(a)anthracene, and more than 30,000 h for benz(a)pyrene. Turnover times in uncontaminated stream sediment were 10 to 400 times greater than in contaminated samples, while absolute rates of PAH transformation (micrograms of PAH per gram of sediment per hour) were 3,000 to 125,000 times greater in contaminated sediment. The data indicate that four- and five-ring PAH compounds, several of which are carcinogenic, may persist even in sediments that have received chronic PAH inputs and that support microbial populations capable of transforming two- and three-ring PAH compounds.  相似文献   

8.
The white rot fungus, Pleurotus ostreatus, metabolized four soil adsorbed polycyclic aromatic hydrocarbons: 50% of pyrene (0.1 mg g–1 dry soil), 68% of anthracene and 63% of phenanthrene were mineralized after 21 d. Biodegradation was increased to 75%, 80% and 75%, respectively of the initial concentration when 0.15% Tween 40 was added. Biodegradation of pyrene in the presence of surfactant and H2O2 (1.0 mM) was 90%. Benz[a]pyrene was also oxidized by Pleurotus ostreatus but it is not mineralized.  相似文献   

9.
1. 7- and 12-Methylbenz[a]anthracene were converted by rat-liver homogenates into the corresponding hydroxymethyl derivatives, products that are probably the 8,9-dihydro-8,9-dihydroxy and the 5,6-dihydro-5,6-dihydroxy derivatives, and a number of phenolic products. 2. Both hydrocarbons were converted into glutathione conjugates; that from 7-methylbenz[a]anthracene was also formed, together with 5,6-dihydro-5,6-dihydroxy- and 5-hydroxy-benz[a]anthracene, from 5,6-epoxy-5,6-dihydro-7-methylbenz[a]anthracene. 3. 7- and 12-Hydroxymethyl-benz[a]anthracene were converted into products that are probably 8,9-dihydro-8,9-dihydroxy derivatives, and into phenols. 4. The preparation of a number of derivatives of the hydrocarbons is described. 5. The oxidation of the hydrocarbons with lead tetra-acetate was investigated.  相似文献   

10.
Detoxification of polycyclic aromatic hydrocarbons by fungi   总被引:8,自引:0,他引:8  
Summary The polycyclic aromatic hydrocarbons (PAHs) are a group of hazardous environmental pollutants, many of which are acutely toxic, mutagenic, or carcinogenic. A diverse group of fungi, includingAspergillus ochraceus, Cunninghamella elegans, Phanerochaete chrysosporium, Saccharomyces cerevisiae, andSyncephalastrum racemosum, have the ability to oxidize PAHs. The PAHs anthracene, benz[a]anthracene, benzo[a]pyrene, fluoranthene, fluorene, naphthalene, phenanthrene, and pyrene, as well as several methyl-, nitro-, and fluoro-substituted PAHs, are metabolized by one or more of these fungi. Unsubstituted PAHs are oxidized initially to arene oxides,trans-dihydrodiols, phenols, quinones, and tetralones. Phenols andtrans-dihydrodiols may be further metabolized, and thus detoxified, by conjugation with sulfate, glucuronic acid, glucose, or xylose. Although dihydrodiol epoxides and other mutagenic and carcinogenic compounds have been detected as minor fungal metabolites of a few PAHs, most transformations performed by fungi reduce the mutagenicity and thus detoxify the PAHs.  相似文献   

11.
Although it has been observed that many epoxides are ultimate mutagens, surprisingly little is known about epoxides to which man may be extensively exposed, e.g., physiological compounds, drugs, drug metabolites and pesticides. We have now investigated 35 such and related epoxides for mutagenicity, using reversion of his?Salmonella typhimurium TA98 and TA100 as biological end-point. None of the tested steroids (12 compounds), vitamin K epoxides (3 compounds) and pesticides (dieldrin, endrin, HEOM (1,2,3,4,9,9-hexachloro-6,7-epoxy-1,4,4a5,6,7,8,8a-octahydro-1,4-methanonaphthalene), heptachlor epoxide) showed any mutagenic activity. Negative results were also obtained with the antibiotics oleandomycin, anti-capsin and asperlin, the cardiotonic drug resibufogenin, the widely used parasympatholytic drugs butylscopolamine and scopolamine, the sedatives valtratum, didovaltratum and acevaltratum, the tranquilizer oxanamide as well as with the drug metabolites carbamazepine 10,11-oxide and diethylstilbestrol α,β-oxide. Three barbiturate epoxides, formed by metabolism of allobarbital, alphenal and secobarbital, caused weak but reproducible mutagenic effects at high concentrations. The cytostatic agent ethoglucide was the only drug having substantial mutagenic activity. Its mutagenic potency was similar to those of the control epoxides styrene 7,8-oxide, p-bromostyrene 7,8-oxide and m-bromostyrene 7,8-oxide, but much lower than those of benzo[a]pyrene 4,5-oxide, benzo[e]pyrene 4,5-oxide and 7,12-dimethylbenz[a]-anthracene 5,6-oxide.Some epoxides were also tested in other Salmonella typhimurium strains or in the presence of rat-liver S9 mix. Positive results were only obtained with compounds that had already been detected as mutagens in the direct test with strain TA100.  相似文献   

12.
An SV40-based shuttle vector system was used to identify the types of mutational changes and the sites of mutation within the supF DNA sequence generated by the four stereoisomers of benzo[c]phenanthrene 3,4-dihydrodiol 1,2-epoxide (B[c]PhDE), by racemic mixtures of bay or fjord region dihydrodiol epoxides (DE) of 5-methylchrysene, of 5,6-dimethylchrysene, of benzo[g]chrysene and of 7-methylbenz[a]anthracene and by two direct acting polycyclic aromatic hydrocarbon carcinogens, 7-bromomethylbenz[a]anthracene (7-BrMeBA) and 7-bromomethyl-12-methylbenz[a]anthracene (7-BrMe-12-MeBA). The results of these studies demonstrated that the predominant type of mutation induced by these compounds is the base substitution. The chemical preference for reaction at deoxyadenosine (dAdo) or deoxyguanosine (dGuo) residues in DNA, which is in general correlated with the spatial structure (planar or non-planar) of the reactive polycyclic aromatic hydrocarbon, is reflected in the preference for mutation at AT or GC pairs. In addition, if the ability to react with DNA in vivo is taken into account, the relative mutagenic potencies of the B[c]PhDE stereoisomers are consistent with the higher tumorigenic activity associated with non-planar polycyclic aromatic hydrocarbons and their extensive reaction with dAdo residues in DNA. Comparison of the types of mutations generated by polycyclic aromatic hydrocarbons and other bulky carcinogens in this shuttle vector system suggests that all bulky lesions may be processed by a similar mechanism related to that involved in replication past apurinic sites. However, inspection of the distribution of mutations over the target gene induced by the different compounds demonstrated that individual polycyclic aromatic hydrocarbons induce unique patterns of mutational hotspots within the target gene. A polymerase arrest assay was used to determine the sequence specificity of the interaction of reactive polycyclic aromatic hydrocarbons with the shuttle vector DNA. The results of these assays revealed a divergence between mutational hotspots and polymerase arrest sites for all compounds investigated, i.e., sites of mutational hotspots do not correspond to sites where high levels of adduct formation occur, and suggested that some association between specific adducts and sequence context may be required to constitute a premutagenic lesion. A site-specific mutagenesis system employing a single-stranded vector (M13mp7L2) was used to investigate the mutational events a single benzo[a]pyrene or benzo[c]phenanthrene dihydrodiol epoxide–DNA adduct elicits within specific sequence contexts. These studies showed that sequence context can cause striking differences in mutagenic frequencies for given adducts. In addition, these sequence context effects do not originate only from nucleotides immediately adjacent to the adduct, but are also modulated by more distal nucleotides. The implications of these results for mechanisms of polycyclic aromatic hydrocarbon-induced mutagenesis and carcinogenesis are discussed.  相似文献   

13.
The synthesis of dibenz[a,c]anthracene 10,11-oxide is described. The oxide was unstable and was rapidly decomposed with cold mineral acid into a mixture of 10- and 11- hydroxydibenz[a,c]anthracene. The oxide was converted by rat liver microsomal preparations and homogenates into a product that is probably 10,11-dihydro-10,11-dihydroxydibenz[a,c]anthracene and which was identical with the metabolite formed when dibenz[a,c]anthracene was metabolized by rat liver homogenates. The oxide did not react either chemically or enzymically with GSH. 10,11-Dihydrodibenz[a,c]anthracene and 10,11-dihydrodibenz[a,c]anthracene 12,13-oxide were both metabolized by rat liver preparations into trans-10,11,12,13-tetrahydro-10,11-dihydroxydibenz[a,c] anthracene and the oxide was converted chemically into this dihydroxy compound, and it reacted chemically but not enzymically with GSH. In the alkylation of 4-(p-nitrobenzyl)pyridine, the ;K-region' epoxide, dibenz[a,h]anthracene 5,6-oxide, was more active than either dibenz[a,c]anthracene 10,11-oxide or 10,11-dihydrobenz[a,c]anthracene 12,13-oxide.  相似文献   

14.
MANY mixed function1 oxygenases are membrane-bound multicomponent enzyme systems which require NADPH and molecular oxygen for the oxidative metabolism of polycyclic hydrocarbons, drugs and insecticides, as well as numerous lipophilic endogenous substrates2–4. Any day to day fluctuations in these enzyme activities may therefore influence the intensity and duration of drug action and also the rate of metabolism of chemical carcinogens, insecticides and many normal body substrates.  相似文献   

15.
The effect of a series of bromomethylated polycyclic hydrocarbons on in vitro DNA and RNA synthesis has been studied by measurement of the incorporation of [3H]-dTMP or [14C]-AMP into new chains. The inhibition of RNA synthesis was less than 12% for 9-bromomethylanthracene, 9-methyl-10-bromomethylanthracene and 12-bromomethylbenzo(a)acridine, and more than 37% for 7-bromomethylbenzo(a)anthracene, 7,12-dibromomethylbenzo(a)anthracene and 7-bromomethylbenzo(c)acridine. Analogous results were found for the inhibition of DNA synthesis, except for 7-bromomethylbenzo(c)acridine which had little effect. Apart from this exception a good correlation was found between the inhibitory action of the bromo derivatives and the carcinogenicity of the non-halogenated parent hydrocarbons.  相似文献   

16.
Carbon partitioning and residue formation during microbial degradation of polycyclic aromatic hydrocarbons (PAH) in soil and soil-compost mixtures were examined by using [14C]anthracenes labeled at different positions. In native soil 43.8% of [9-14C]anthracene was mineralized by the autochthonous microflora and 45.4% was transformed into bound residues within 176 days. Addition of compost increased the metabolism (67.2% of the anthracene was mineralized) and decreased the residue formation (20.7% of the anthracene was transformed). Thus, the higher organic carbon content after compost was added did not increase the level of residue formation. [14C]anthracene labeled at position 1,2,3,4,4a,5a was metabolized more rapidly and resulted in formation of higher levels of residues (28.5%) by the soil-compost mixture than [14C]anthracene radiolabeled at position C-9 (20.7%). Two phases of residue formation were observed in the experiments. In the first phase the original compound was sequestered in the soil, as indicated by its limited extractability. In the second phase metabolites were incorporated into humic substances after microbial degradation of the PAH (biogenic residue formation). PAH metabolites undergo oxidative coupling to phenolic compounds to form nonhydrolyzable humic substance-like macromolecules. We found indications that monomeric educts are coupled by C-C- or either bonds. Hydrolyzable ester bonds or sorption of the parent compounds plays a minor role in residue formation. Moreover, experiments performed with 14CO2 revealed that residues may arise from CO2 in the soil in amounts typical for anthracene biodegradation. The extent of residue formation depends on the metabolic capacity of the soil microflora and the characteristics of the soil. The position of the 14C label is another important factor which controls mineralization and residue formation from metabolized compounds.  相似文献   

17.
The isoenzymes of human and rat lung glutathione S-transferase (GST) differ among themselves in their activities towards the epoxides of benzo(a)pyrene (BP). The Ya' and Yc-type subunits of rat lung GST exhibit maximum activities towards BP-4,5-oxide and BP-7,8-oxide suggesting that these two subunits are preferentially involved in the detoxification of highly reactive epoxides and diol-epoxides of polycyclic aromatic hydrocarbons (PAH). The studies with human lung GST isoenzymes indicate that BP-4,5-oxide, and BP-7,8-oxide are preferred substrates for the cationic (pI 8.3) form of the enzyme. Identification of compounds which can selectively induce these isoenzymes of GST could prove useful as inhibitors of PAH induced neoplasia.  相似文献   

18.
Enantiomers of nine K-region and one non-K-region epoxides of polycyclic aromatic hydrocarbons have been resolved by high-performance liquid chromatography with chiral stationary phases either ionically or covalently bonded to gamma-aminopropylsilanized silica. Resolution of enantiomers was confirmed by ultraviolet-visible absorption, circular dichroism, and mass spectral analyses. This method has been applied to the determination of optical purity and absolute configuration of the K-region epoxides formed in the metabolism of 1-methylbenz[a]anthracene, 7-methylbenz[a]anthracene, and 12-methylbenz[a]anthracene by rat liver microsomes.  相似文献   

19.
Experiments were performed to investigate the effects of 3 polycyclic aromatic hydrocarbons, benz[a]anthracene, dibenz[a,c]anthracene and dibenz[a,h]anthracene and K-regio epoxides and some of their related dihydrodiols on the chromosomes of Chinese hamster ovary cells in vitro. Of the 3 hydrocarbons only benz[a]anthracene showed any activity in inducing sister-chromatid exchanges. The K-region epoxide and the 3,4-dihydrodiol have been found to be more active than the corresponding K-region or the other non K-region dihydrodiols derived from benz[a]anthracene. Athough dibenz[a,c]anthracene was almost inactive, the K-region 5,6-epoxide and all 3 possible dihydrodiols, the 1,2-, 3,4- and 10,11-diols were active in inducing increased numbers of sister-chromatid exchanges in the chromosomes of these cells. The 3,4-dihydrodiol of dibenz[a,h]anthrecene was also active in inducing sister-chromatid exchanges whereas the 1,2- and 5,6-dihydrodiols were only weakly active. This study provides some support for the suggestiion that the activation of these 3 hydrocarbons proceeds by the metabolic conversion of non K-region dihydrodiols into vicinal diol-epoxides.  相似文献   

20.
Quinones and other oxygenated polycyclic aromatic hydrocarbons (oxy-PAHs) are toxic and/or genotoxic compounds observed to be cocontaminants at PAH-contaminated sites, but their formation and fate in contaminated environmental systems have not been well studied. Anthracene-9,10-dione (anthraquinone) has been found in most PAH-contaminated soils and sediments that have been analyzed for oxy-PAHs. However, little is known about the biodegradation of oxy-PAHs, and no bacterial isolates have been described that are capable of growing on or degrading anthraquinone. PAH-degrading Mycobacterium spp. are the only organisms that have been investigated to date for metabolism of a PAH quinone, 4,5-pyrenequinone. We utilized DNA-based stable-isotope probing (SIP) with [U-13C]anthraquinone to identify bacteria associated with anthraquinone degradation in PAH-contaminated soil from a former manufactured-gas plant site both before and after treatment in a laboratory-scale bioreactor. SIP with [U-13C]anthracene was also performed to assess whether bacteria capable of growing on anthracene are the same as those identified to grow on anthraquinone. Organisms closely related to Sphingomonas were the most predominant among the organisms associated with anthraquinone degradation in bioreactor-treated soil, while organisms in the genus Phenylobacterium comprised the majority of anthraquinone degraders in the untreated soil. Bacteria associated with anthracene degradation differed from those responsible for anthraquinone degradation. These results suggest that Sphingomonas and Phenylobacterium species are associated with anthraquinone degradation and that anthracene-degrading organisms may not possess mechanisms to grow on anthraquinone.  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