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.
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   

2.
Polar, ethyl acetate soluble metabolites formed in incubations of dibenz[a,c]-anthracene (DB[a,c]A), dibenz[a,h]anthracene (DB[a,h]A) and the related DB[a,h]A 3,4-diol and dibenz[a,j]anthracene (DB[a,j]A) with 3-methylcholanthrene (3-MC)-induced rat liver microsomal preparations have been separated by HPLC and examined using fluorescence, UV and NMR spectroscopy. Metabolites with spectral properties consistant with their identification as the 3,4:8,9-bis-diol of DB[a,j]A and a 1,2,3,4,12,13-hexol derived from DB[a,c]A were found. DB[a,h]A was metabolized to three polar products identified as the 3,4:10,11-bis-diol and the related 1,2,3,4,8,9- and 1,2,3,4,10,11-hexols, which were also formed, together with the related 1,2,3,4-tetrol, from the DB[a,h]A 3,4-diol. The possible role of bis-diols in the metabolic activation of these three dibenzanthracenes is discussed.  相似文献   

3.
Mouse skin and human skin have been treated in vivo or in short-term organ culture with dibenz[a,h]anthracene (DB[a,h]A), the related 3,4- or 5,6-diols or the anti- or syn-3,4-diol 1,2-oxides. DNA hydrolysates have been 32P-postlabelled and the adducts present examined by HPLC using a phenyl-modified reverse phase column and, for comparison, by PEI-cellulose TLC and autoradiography. The adducts formed when the diol-epoxides were reacted with salmon sperm DNA were also examined. The results show that in mouse skin treated in vivo, the major adducts formed from DB[a,h]A and the 3,4-diol were the same and that two of them were more polar than those formed in skin or in DNA that had been treated with the related anti- or syn-diol epoxides. Human skin treated with DB[a,h]A in culture yielded an adduct profile that was qualitatively similar to the profiles obtained with mouse skin.  相似文献   

4.
The formation of trans-dihydrodiols from dibenz[a,c]anthracene, dibenz[a,h]anthracene and chrysene by chemical oxidation in an ascorbic acid-ferrous sulphate-EDTA system and by rat-liver microsomal fractions has been studied using a combination of thin-layer (TLC) and high pressure liquid chromatography (HPLC) to separate the mixtures of isomeric dihydrodiols. The 1,2- and 3,4-dihydrodiols of dibenz[a,c]anthracene, the 1,2-,3,4- and 5,6-dihydrodiols of dibenz[a,h]anthracene and the 1,2-, 3,4- and 5,6-dihydrodiols of chrysene were formed in chemical oxidations. These dihydrodiols were also formed when the three parent hydrocarbons were metabolized by rat-liver microsomal fractions and, in addition, dibenz[a,c]anthracene yielded the 10,11-dihydrodiol. The 1,2- and 3,4-dihydrodiols of dibenz[a,c]anthracene have not been reported previously either as metabolites of the hydrocarbon or as products of chemical syntheses and the 5,6-dihydrodiol of chrysene was not detected in earlier metabolic studies.  相似文献   

5.
According to epidemiologic studies, exposure of women to fumes from cooking oils appears to be an important risk factor for lung cancer. Fume samples from three different commercial cooking oils frequently used in Taiwan were collected and analyzed for mutagenicity in the Salmonella/microsome assay. Polycyclic aromatic hydrocarbons were extracted from the samples and identified by HPLC chromatography. Extracts from three cooking oil fumes were found to be mutagenic in the presence of S9 mix. All samples contained dibenz[a,h]anthracene (DB[a,h]A) and benz[a]anthracene (B[a]A). Concentration of DB[a,h]A and B[a]A were 1.9 and 2.2 μg/m3 in fumes from lard oil, 2.1 and 2.3 μg/m3 in soybean oil, 1.8 and 1.3 μg/m3 in peanut oil, respectively. Benzo[a]pyrene (B[a]P) was identified in fume samples of soybean and peanut oil, in concentrations of 19.6 and 18.3 μg/m3, in this order. These results provide experimental evidence and support the findings of epidemiologic observations, in which women exposed to the emitted fumes of cooking oils are at increased risk of contracting lung cancer.  相似文献   

6.
Microbiological analysis of soils from a polycyclic aromatic hydrocarbon (PAH)-contaminated site resulted in the enrichment of five microbial communities capable of utilizing pyrene as a sole carbon and energy source. Communities 4 and 5 rapidly degraded a number of different PAH compounds. Three pure cultures were isolated from community 5 using a spray plate method with pyrene as the sole carbon source. The cultures were identified as strains of Burkholderia ( Pseudomonas ) cepacia on the basis of biochemical and growth tests. The pure cultures (VUN 10 001, VUN 10 002 and VUN 10 003) were capable of degrading fluorene, phenanthrene and pyrene (100 mg l−1) to undetectable levels within 7–10 d in standard serum bottle cultures. Pyrene degradation was observed at concentrations up to 1000 mg l−1. The three isolates were also able to degrade other PAHs including fluoranthene, benz[ a ]anthracene and dibenz[ a , h ]anthracene as sole carbon and energy sources. Stimulation of dibenz[ a , h ]anthracene and benzo[ a ]pyrene degradation was achieved by the addition of small quantities of phenanthrene to cultures containing these compounds. Substrate utilization tests revealed that these micro-organisms could also grow on n -alkanes, chlorinated- and nitro-aromatic compounds.  相似文献   

7.
Studies were carried out on the incidence of sister-chromatid exchanges induced in Chinese hamster ovary cells by in vitro treatment with the polycyclic aromatic hydrocarbons 7-methylbenz[a]anthracene and benzo[a]pyrene and with related K-region and non-K-region dihydrodiols. Appreciable increases in the incidence of sister-chromatid exchanges were apparent in cells treated with non-K-region dihydrodiols: the most active compounds were 3,4-dihydro-3,4-dihydroxy-7-methylbenz[a]anthracene and 7,8-dihydro-7,8-dihydroxybenzo[a]pyrene and the effects were dose-dependent. The parent hydrocarbons and the related K-region dihydrodiols induced some sister-chromatid exchanges but they were considerably less active than these two non-K-region diols. The results suggest that this system may usefully be applied to studies aimed at determining which dihydrodiols are important in the metabolic activation of the carcinogenic polycyclic hydrocarbons. These and other results also infer that Chinese hamster ovary cells possess some intrinsic ability to metabolize such compounds in the absence of exogenous activation systems.  相似文献   

8.
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   

9.
Early carcinogenicity tests found no evidence of activity for picene but found considerable initiating and carcinogenic activity for dibenz[a,h]anthracene (DBA). More recent investigation suggested that both pentacyclics were complete carcinogens when administered as single sc injections in NMRI mice, despite findings that picene acted as neither an initiating nor promoting agent. To investigate this contradiction, the complete carcinogenicities of both isomers were compared by sc injection in female Sprague-Dawley rats. The results demonstrate that 1 micromol of DBA, administered three times weekly for 20 doses, induces sarcomas in all test animals by 33 weeks (100%). Similar treatment with picene did not induce sarcoma in any test animals by 37 weeks (0%). The present results agree with the earlier studies. It follows from these results that the predictions of the unified theory for the appearance of carcinogenic properties following administration of picene and dibenz[a,h]anthracene to test animals have been confirmed.  相似文献   

10.
Large inocula of Stenotrophomonas maltophilia VUN 10,003 were used to investigate bacterial degradation of benzo[a]pyrene and dibenz[a,h]anthracene. Although strain VUN 10,003 was capable of degrading 10–15 mg l−1 of the five-ring compounds in the presence of pyrene after 63 days, further addition of pyrene after degradation of the five-ring polycyclic aromatic hydrocarbons (PAHs) ceased did not stimulate significant decreases in the concentration of benzo[a]pyrene or dibenz[a,h]anthracene. However, pyrene was degraded to undetectable levels 21 days after its addition. The amount of benzo[a]pyrene and dibenz[a,h]anthracene degraded by strain VUN 10,003 was not affected by the initial concentration of the compounds when tested at 25–100 mg l−1, by the accumulation of by-products from pyrene catabolism or a loss of ability by the cells to catabolise benzo[a]pyrene or dibenz[a,h]anthracene. Metabolite or by-product repression was suspected to be responsible for the inhibition: By-products from the degradation of the five-ring compounds inhibited their further degradation. Journal of Industrial Microbiology & Biotechnology (2002) 28, 88–96 DOI: 10.1038/sj/jim/7000216 Received 30 January 2001/ Accepted in revised form 10 October 2001  相似文献   

11.
When benz[a] anthracene was oxidised in a reaction mixture containing ascorbic acid, ferrous sulphate and EDTA, the non-K-region dihydrodiols, trans-1,2-dihydro-1,2-dihydroxybenz[a] anthracene and trans-3,4-dihydro-3,4-dihydroxybenz[a] anthracene together with small amounts of the 8,9- and 10,11-dihydrodiols were formed. When oxidised in a similar system, 7,12-dimethylbenz[a] anthracene yielded the K-region dihydrodiol, trans-5,6-dihydro-5,6-dihydroxy-7,12-dimethylbenz[a] anthracene and the non-K-region dihydrodiols, trans-3,4-dihydro-3,4-dihydroxy-7,12-dimethylbenz[a] anthracene, trans-8,9-dihydro-8,9-dihydroxy-7,12-dimethylbenz[a] anthracene, trans-10,11-dihydro-10,11-dihydroxy-7,12-dimethylbenz[a] anthracene and a trace of the 1,2-dihydrodiol. The structures and sterochemistry of the dihydrodiols were established by comparisons of their UV spectra and chromatographic characteristics using HPLC with those of authentic compounds or, when no authentic compounds were available, by UV, NMR and mass spectral analysis. An examination by HPLC of the dihydrodiols formed in the metabolism, by rat-liver microsomal fractions, of benz[a] anthracene and 7,12-dimethylbenz[a] anthracene was carried out. The metabolic dihydriols were identified by comparisons of their chromatographic and UV or fluorescence spectral characteristics with compounds of known structures. The principle metabolic dihydriols formed from both benz[a] anthracene and 7,12-dimethylbenz[a] anthracene were the trans-5,6- and trans-8,9-dihydrodiols. The 1,2- and 10,11-dihydrodiols were identified as minor products of the metabolism of benz [a] anthracene and the tentative identification of the trans-3,4-dihydriol as a metabolite was made from fluorescence and chromatographic data. The minor metabolic dihydriols formed from 7,12-dimethylbenz[a] anthracene were the trans-3,4-dihydrodiol and the trans-10,11-dihydriol but the trans-1,2-dihydrodiol was not detected in the present study.  相似文献   

12.
We have examined the selectivity of rat liver microsomal epoxide hydrolase (EC 3.3.2.3) toward all of the possible positional isomers of benzo-ring diol epoxides and tetrahydroepoxides of benz[a]anthracene, as well as the 1,2-diol 3,4-epoxides of triphenylene. This set includes compounds with no bay region in the vicinity of the benzo-ring, a bay-region diol group, a bay-region epoxide group, and (for the triphenylene derivatives) both a bay-region diol and a bay-region epoxide. In all cases where both the tetrahydroepoxides and the corresponding diol epoxides were examined, there is a large retarding effect of hydroxyl substitution on the rate of the enzyme-catalyzed hydration. When the tetrahydroepoxides are fair or poor substrates (epoxide group in the 1,2-, 8,9-, or 10,11-position), the additional retardation introduced by adjacent hydroxyl groups causes the enzyme-catalyzed hydrolysis of the corresponding diol epoxides to be insignificantly slow or nonexistent. In contrast, a benz[a]anthracene derivative with an epoxide group in the 3,4-position, (-)-tetrahydrobenz[a]anthracene (3R,4S)-epoxide, has been identified as the best substrate known for epoxide hydrolase, with a Vmax at 37 degrees C and pH 8.4 of 6800 nmol/min/mg of protein, and the two diastereomeric (+/-)-benz[a]anthracene 1,2-diol 3,4-epoxides, unlike all the other diol epoxides examined to date, are moderately good substrates for epoxide hydrolase. This novel observation is accounted for by the fact that the very high reactivity of the tetrahydrobenz[a]anthracene 3,4-epoxide system towards epoxide hydrolase is large enough to overcome a kinetically unfavorable effect of hydroxyl substitution. The enantioselectivity and positional selectivity of the enzyme have been determined for the tetrahydro-1,2- and -3,4-epoxides of benz[a]anthracene as well as the 1,2-diol 3,4-epoxides. When the epoxide is located in the 3,4-position, the benzylic carbon is the preferred site of attack, whereas for the enantiomers of the bay-region tetrahydro-1,2-epoxides, the chemically less reactive non-benzylic carbon is preferred. The regio- and enantioselectivity of epoxide hydrolase are discussed in terms of a possible model for the hydrophobic binding site of this enzyme.  相似文献   

13.
Bioaugmentation of polycyclic aromatic hydrocarbon (PAH)-contaminated soil was investigated using a mixed bacterial culture (community five) isolated from an abandoned industrial site. Community five was inoculated into contaminated soil containing a total PAH (two- to five-ring compounds) concentration of approximately 820 mg/kg soil. PAH degradation by the indigenous microbial population was restricted to the lower molecular weight compounds (naphthalene, acenaphthene, fluorene and phenanthrene) even with yeast extract addition: these compounds decreased by 14 to 37%, in soil hydrated to 50% water capacity, following 91 days of incubation at 24°C. Inoculation of community five into this PAH-contaminated soil resulted in significant decreases in the concentration of all PAHs over the incubation period: greater than 86% of naphthalene, acenaphthene, fluorene, and phenanthrene were degraded after 91 days, while anthracene, fluoranthene, and pyrene were degraded to lesser extents (51.7 to 57.6%). A lag period of 48 to 63 days was observed before the onset of benz[a]anthracene, benzo[a]pyrene, and dibenz[a,h]anthracene removal. However, significant decreases in the concentration of these compounds (32.6, 25.2, and 18.5%, respectively) were observed after 91 days. No significant decrease in the mutagenic potential of organic soil extracts (as measured by the Ames Test) was observed after incubation of the soil with the indigenous microflora; however, the Microtox toxicity of aqueous soil extracts was reduced sevenfold. In contrast, extracts from contaminated soil inoculated with community five underwent a 43% decrease in mutagenic potential and the toxicity was reduced 170-fold after 91 days incubation. These observations suggest that community five could be utilised for the detoxification of PAH-contaminated soil.  相似文献   

14.
Addition of arachidonic acid to ram seminal vesicle microsomes oxidizes 3,4-dihydroxy-3,4-dihydrobenzo[a]anthracene (BA-3,4-diol) to five more polar products. Four of the products are identified by chromatographic and spectroscopic analysis as tetrahydrotetraols, which are solvolysis products of dihydrodiolepoxides. The fifth product is a 10-methyl ether formed by methanolysis of the anti-diolepoxide. Quantitation of the individual products indicates that anti-diolepoxides predominate over syn-diolepoxides by approximately 2:1. Identical product profiles are detected from the reaction of BA-3,4-diol with hematin and 13-hydroperoxy-octadecadienoic acid in the presence of Tween 20. No other products are detected in either system, which indicates that peroxyl radicals oxidize BA-3,4-diol exclusively by epoxidation of the 1,2-double bond. The stereochemical and regiochemical differences between oxidation of BA-3,4-diol by peroxyl radicals and cytochrome P-450 are dramatic and suggest that BA-3,4-diol is uniquely suited as a probe to quantitate peroxyl radical-dependent epoxidation in vitro and in vivo.  相似文献   

15.
Previous studies by other investigators have established that L-region methyl derivatives of dibenz[a,h]anthracene (DBA) were more carcinogenic than the parent hydrocarbon. The bioalkylation of DBA was investigated by incubating the hydrocarbon with rat liver cytosol fortified with S-adenosyl-L-methionine (SAM) in 0.1 M phosphate buffer (pH 7.4) for 1 h at 37 degrees C in air. The reaction was stopped by the addition of cold acetone and the mixture extracted with ethyl acetate and washed with water. The organic phase was evaporated and the residue dissolved in methylene chloride for analysis by reverse phase high performance liquid chromatography (HPLC) and gas chromatography/mass spectroscopy GC/MS. Products were found that were indistinguishable from 7-methyl-DBA and 7,14-dimethyl-DBA, 7-hydroxymethyl-DBA, 7-hydroxymethyl-14-methyl-DBA, and 7,14-dihydroxymethyl-DBA. The results suggest that unsubstituted carcinogenic hydrocarbons are preprocarcinogens that react with SAM in liver cytosol preparations, to form alkyl substituted procarcinogens, which are more potent than the corresponding preprocarcinogens.  相似文献   

16.
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   

17.
The cytochrome P450 isoforms responsible for the regio-selective metabolism of benz[a]anthracene (BA) are poorly defined but as with other polycyclic aromatic hydrocarbons (PAHs) may include members of the CYP2C sub-family. Since the expression of some of these is regulated in a gender-specific manner and may be altered by age, rat strain or by phenobarbital treatment, the effects of these variables on metabolism of BA to diols was investigated. These studies used hepatic, microsomal membranes from immature and adult Long-Evans rats and adult Hooded Lister rats. BA-diols were resolved by normal phase HPLC into three discrete peaks identified as benz[a]anthracene-5,6-diol (BA-5,6-diol), benz[a]anthracene-10, 11-diol (BA-10,11-diol) and a mixture of benz[a]anthracene-3,4- and -8,9-diols (BA-3,4-diol and BA-8,9-diol and termed Peak(3/8)). Significant gender-related differences were found in the rates of diol formation in adults of both the Long-Evans and Hooded Lister rat strains. Formation of BA-10,11-diol and to a lesser extent the components of Peak(3/8) were greater in the male compared to female animals by factors of at least 14 and two, respectively. An age-dependent effect is also observed in the Long-Evans rat since these differences are still apparent in prepubertal animals but to a lesser extent (gender ratio male:female BA-10,11-diol 9X; Peak(3/8) 1.4X). In contrast BA-5,6-diol was formed at similar rates by membranes from female and male rats whether mature (Long-Evans and Hooded Lister) or immature (Long-Evans). Phenobarbital treatment of the adult Long-Evans rats resulted in a moderate increase in the formation of each diol other than at the 10,11-position and the induction was not gender specific. The rate of formation of BA-10, 11-diol was decreased in phenobarbital-treated male rats suggesting modulation of a male specific isoform. Measurement of microsomal epoxide hydrolase revealed no gender or age differences and suggests that this enzyme is not rate limiting in BA-diol formation and thus is not responsible for the differences in BA-diol formation observed. The results suggest that CYP2C11 along with a male-specific isoenzyme not regulated by age are important in the formation of BA-10,11-diol and a component(s) of Peak(3/8) in males. CYPs 2B2 and/or 2C6 appear to be involved in formation of BA-5,6-diol in male and female. Identification of the CYPs involved in the regio-selective metabolism of BA may lead to an explanation of the lower carcinogenic potency of this PAH compared to dimethylbenz[a]anthracene and this study provides novel clues concerning the identities of the CYPs, which are important.  相似文献   

18.
The role of vicinal diol-epoxides in the metabolic activation of 7,12-dimethylbenz[a]anthracene to intermediates that react with nucleic acids was investigated using Sephadex LH-20 column chromatography and high pressure liquid chromatography. The results show that some of the hydrocarbon-DNA products formed in mouse skin treated in vivo with 7,12-dimethylbenz[a]anthracene arise from the reaction of DNA with 3,4-dihydro-3,4-dihydroxy-7,12-dimethylbenz[a]anthracene 1,2-oxides which, on the basis of this and other evidence, appears to be a biologically-active metabolite of 7,12-dimethylbenz[a]anthracene. However, since other nucleic acid-hydrocarbon adducts were also present that have not been identified as resulting from the reaction of the 3,4-diol 1,2-oxides with DNA, other mechanisms may also be involved in the metabolic activation of 7,12-dimethylbenz[a]anthracene in mouse skin.  相似文献   

19.
1. Benz[a]anthracene was hydroxylated by rat-liver homogenates on the 3,4-,5,6- or 8,9-bond to yield phenols and dihydrodihydroxy compounds. Metabolic action at the 7- and 12-positions was also detected. 5,6-Epoxy-5,6-dihydrobenzanthracene was converted into a phenol that is probably 5-hydroxybenzanthracene and 5,6-dihydro-5,6-dihydroxybenzanthracene. Both substrates yielded a product that is probably S-(5,6-dihydro-6-hydroxy-5-benzanthracenyl)glutathione. 2. Dibenz[a,h]anthracene was hydroxylated by rat-liver homogenates to yield products that are probably 3- and 4-hydroxydibenzanthracene, 1,2-dihydro-1,2-dihydroxydibenzanthracene, 3,4-dihydro-3,4-dihydroxydibenzanthracene and 5,6-dihydro-5,6-dihydroxydibenzanthracene. There was no evidence for metabolic action at the 7- and 14-positions. 5,6-Epoxy-5,6-dihydrodibenzanthracene was converted into a phenol that is probably 5-hydroxydibenzanthracene and 5,6-dihydro-5,6-dihydroxydibenzanthracene. Both substrates yielded a glutathione conjugate that is probably S-(5,6-dihydro-6-hydroxy-5-dibenzanthracenyl)glutathione. 3. The synthesis of 5,6-epoxy-5,6-dihydrodibenzanthracene is described and the reactions of this epoxide and 5,6-epoxy-5,6-dihydrobenzanthracene with water and thiols have been investigated. 4. The oxidation of dibenzanthracene in the ascorbic acid-Fe(2+) ion-oxygen model system is described.  相似文献   

20.
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   

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