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.
We have studied several features of the interactions of 7r,8t-dihydroxy-9t,10t-oxy-7,8,9,10-tetrahydrobenzo[a]pyrene (BPDE-I) with a DNA repair-proficient derivative of Chinese hamster ovary cells (CHO), AT3-2, and with a UV-light sensitive mutant, UVL-10, derived from AT3-2. Methods were developed for quantitating the amount of unhydrolysed BPDE-I associated with cells and for purifying DNA from cells under conditions where artificial labeling during preparation is minimized. In both cell types, about 30% of the BPDE-I added to a cell culture is rapidly taken up by the cells and is maintained in a cellular compartment in which the half-life of BPDE-I is about 10-fold longer than in aqueous medium. The kinetics of covalent binding to DNA were measured in both cell types and found to be described well by a single exponential process with a half-life of about 60 min. This is virtually identical to the half-life for intracellular hydrolysis of BPDE-I (57 min), consistent with the suggestion that this intracellular, relatively stable BPDE-I is responsible for binding.  相似文献   

2.
DNA covalent binding studies with enantiomers of trans-7,8-dihydroxy- anti-9,10-epoxy-7,8,9,10-tetrahydro-benzo[a]pyrene (anti-BPDE) have been carried out by means of spectroscopic techniques (UV, CD, and fluorescence). Synthetic polynucleotides are employed to investigate binding differences between the G.C and A.T base pairs and to elucidate the bases for the stereoselective covalent binding of DNA toward anti-BPDE. The results indicate that of all the polynucleotides studied, only poly(dA-dT).poly(dA-dT) exhibits predominant intercalative covalent binding towards (+)-anti-BPDE and suffers the least covalent modification. Only minor intercalative covalent contributions are found in alternating polymer poly(dA-dC).poly(dG-dT). These observations parallel the DNA physical binding results of anti-BPDE and its hydrolysis products. They support the hypothesis that intercalative covalent adducts derive from intercalative physical binding while the external covalent adducts derive from external bimolecular associations. In contrast to the A.T polymers, the guanine containing polymers exhibit pronounced reduction in covalent modification by (-)-anti-BPDE. The intercalative covalent binding mode becomes relatively more important in the adducts formed by the (-) enantiomer as a consequence of decreased external guanine binding. These findings are consistent with the guanine specificity, stereoselective covalent binding at dG, the absence of stereoselectivity at dA for anti-BPDE, and the enhanced binding heterogeneity for the (-) enantiomer as found in the native DNA studies. The possible sequence and/or conformational dependence of such stereoselective covalent binding is indicated by the opposite pyrenyl CD sign exhibited by (+)-anti-BPDE bound to polynucleotides with pyrimidine on one strand and purine on another vs. that bound to polymers containing alternating purine-pyrimidine sequences.  相似文献   

3.
Abstract

DNA covalent binding studies with enantiomers of trans-7,8-dihydroxy- anti-9,10-epoxy- 7,8,9,10-tetrahydro-benzo [a] pyrene (anti-BPDE) have been carried out by means of spectroscopic techniques (UV, CD, and fluorescence). Synthetic polynucleotides are employed to investigate binding differences between the G · C and A · T base pairs and to elucidate the bases for the stereoselective covalent binding of DNA toward anti-BPDE. The results indicate that of all the polynucleotides studied, only poly(dA-dT) · poly(dA-dT) exhibits predominant intercalative covalent binding towards (+)-anti-BPDE and suffers the least covalent modification. Only minor intercalative covalent contributions are found in alternating polymer poly(dA-dC) · poly(dG-dT). These observations parallel the DNA physical binding results of anti-BPDE and its hydrolysis products. They support the hypothesis that intercalative covalent adducts derive from intercalative physical binding while the external covalent adducts derive from external bimolecular associations. In contrast to the A · T polymers, the guanine containing polymers exhibit pronounced reduction in covalent modification by (-)-anti-BPDE. The intercalative covalent binding mode becomes relatively more important in the adducts formed by the (-) enantiomer as a consequence of decreased external guanine binding. These findings are consistent with the guanine specificity, stereoselective covalent binding at dG, the absence of stereoselectivity at dA for anti-BPDE, and the enhanced binding heterogeneity for the (-) enantiomer as found in the native DNA studies. The possible sequence and/or conformational dependence of such stereoselective covalent binding is indicated by the opposite pyrenyl CD sign exhibited by (+)-anti-BPDE bound to polynucleotides with pyrimidine on one strand and purine on another vs. that bound to polymers containing alternating purine-pyrimidine sequences.  相似文献   

4.
Insights into the mechanisms of chemical carcinogenesis can sometimes be gained by comparing the effects of closely related chemicals which differ in carcinogenic potency. We have treated Chinese hamster ovary (CHO) cells with a non-carcinogenic metabolite of benzo[a]pyrene, 9r,10t-dihydroxy-7c,8c-oxy-7,8,9,10-tetrahydrobenzo[a]pyrene (BPDE-III), and measured the formation and persistence of DNA adducts. We have correlated this binding data with cytotoxicity and mutagenicity in a DNA-repair-proficient CHO cell line (AT3-2) and in two derived lines, UVL-1 and UVL-10, which are unable to repair bulky DNA adducts. These data are compared with similar studies of the effects of the carcinogenic metabolite, 7r,8t-dihydroxy-9t,10t-oxy-7,8,9,10-tetrahydrobenzo[a]pyrene (BPDE-I). Synchronous fluorescence spectroscopy was used to measure the levels of BPDE-III-DNA adducts in treated cells. Adduct levels increased linearly with dose, but the absolute binding levels were about 30-fold lower than in comparable incubations with BPDE-I. Measurements of the removal of adducts derived from these two diol epoxides indicated no significant difference in the rate of repair measured 24 h post-treatment. When cells were treated with increasing doses of BPDE-III, survival curves were obtained which exhibited a shoulder region at low doses and an exponential decrease in plating efficiency at higher doses. By comparison of the D0's, the DNA-repair-deficient cell lines were found to be 4-5-fold more sensitive to the killing effects of BPDE-III than were the repair-proficient AT3-2 cells.  相似文献   

5.
Linear dichroism and absorption methods are used to study the orientations of transition moments of absorption bands of polycyclic aromatic epoxide derivatives which overlap with those of the DNA band in the 240-300 nm region. Both the short and long axes of the pyrene residues of 1-oxiranylpyrene (1-OP) and the (+) and (-) enantiomers of trans-7,8-dihydroxy-anti-9,10-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene (BPDE) noncovalently bound to double-stranded native DNA are oriented approximately perpendicular to the axis of the DNA helix, consistent with intercalative modes of binding. The covalent binding of these three epoxide derivatives to DNA is accompanied by reorientations of both the short and long axes of the pyrene residues. Covalent adducts derived from the highly mutagenic (+)-anti-BPDE are characterized by tilts of the short axis within 35 degrees or less, and of the long axis by more than 60-80 degrees, with respect to the planes of the DNA bases. In the adducts derived from the binding of the less mutagenic (-)-anti-BPDE and 1-OP epoxide derivatives to DNA, the long axes of the pyrenyl rings are predominantly oriented within 25 degrees of the planes of the DNA bases; however, in the case of the (-) enantiomer of BPDE, there is significant heterogeneity of conformations. In the case of the 1-OP covalent DNA adducts, the short axis of the pyrene ring system is tilted away from the planes of the DNA bases, and the pyrene ring system is not intercalated between DNA base-pairs as in the noncovalent complexes. The stereochemical properties of the saturated 7,8,9,10-ring in BPDE, or the lack of the 7 and 8 carbon atoms in 1-OP, do not seem to affect noncovalent intercalative complex formation which, most likely, is influenced mainly by the flat pyrenyl residues. These structural features, however, strongly influence the conformations of the covalent adducts, which in turn may be responsible for the differences in the mutagenic activities of these molecules.  相似文献   

6.
The formation of DNA adducts by the ultimate carcinogen 7r,8t-dihydroxy-9t,10t-oxy-7,8,9,10-tetrahydrobenzo[alpha]pyrene (BPDE-I) has been implicated in the process of carcinogenesis. In a line of Chinese hamster ovary (CHO) cells designated AT3-2 and in two derivative mutant lines, UVL-1 and UVL-10, originally selected for hypersensitivity to UV-irradiation, we have measured the formation of BPDE-I: DNA adducts and the production of biological damage. The quantity and quality of BPDE-I: DNA adducts formed initially in the 3 cell lines are identical over a wide range of BPDE-I doses. However, the UVL lines are unable to remove adducts from their DNA, while the AT3-2 cells remove about 50% of the BPDE-I: DNA adducts in a 24-h incubation. Correlated with this, the UVL lines are more sensitive to the lethal effects of BPDE-I than are the AT3-2 cells. Mutant frequencies were measured at the aprt, hprt and oua loci and were found to increase linearly with BPDE-I: DNA adduct formation at doses which gave greater than 50% survival. At the hprt and oua loci, the efficiency of mutation induction was similar for AT3-2 and UVL-10 cells. UVL-1 cells showed slightly higher (within a factor of 2-3) mutant frequencies in response to BPDE-I compared to AT3-2 at these two loci. However, at the aprt locus the repair-deficient cells were much more highly mutable (9-15-fold) than the repair-proficient AT3-2 cells. Based on the measured average level of adduct formation, it is calculated that 15% of the BPDE-I: DNA adducts in the aprt gene are converted into mutations. However, the possibility exists that the aprt locus is subject to higher levels of modification by BPDE-I than is the bulk DNA, which would lead to an artifactually high apparent conversion frequency.  相似文献   

7.
The initiation of carcinogenesis by carcinogens such as 7r,8t-dihydroxy-9,10t-oxy-7,8,9,10-tetrahydrobenzo[a]pyrene (BPDE-I) is thought to involve the formation of DNA adducts. However, the diastereomeric diol epoxide, 7r,8t-dihydroxy-9,10c-oxy-7,8,9,10-tetrahydrobenzo[a]pyrene (BPDE-II), also forms DNA adducts but is inactive in standard carcinogenesis models. We have measured the formation and loss of DNA adducts derived from BPDE-II in a DNA-repair-proficient line of Chinese hamster ovary (CHO) cells, AT3-2, and in two derived mutant cell lines, UVL-1 and UVL-10, which are unable to repair bulky DNA adducts. BPDE-II adducts were lost from cellular DNA in AT3-2 cells with a half-life of 13.8 h; this was about twice the rate found for BPDE-I adducts. BPDE-II adducts were also lost from DNA in UVL-1 and UVL-10 cells, but at a much slower rate. When purified DNA was modified in vitro with BPDE-II and then held at 37 degrees C, DNA adducts were removed at a rate identical to that seen in UVL-1 and UVL-10 cells, suggesting that the loss in these cells was not due to enzymatic DNA-repair processes but to chemical lability of the adducts. Mutant frequencies at the APRT and HPRT loci were measured at BPDE-II doses that resulted in greater than 20% survival, and were found to increase linearly with dose. In the DNA-repair-deficient cells, the HPRT locus was moderately hypermutable compared with AT3-2 cells (about 5-fold); the APRT locus was extremely hypermutable, giving about 25-fold higher mutant fractions in UVL-1 and UVL-10 than in AT3-2 cells at equal initial levels of binding. When we compared the mutational efficiency of BPDE-II at both loci in AT3-2 cells (the mutant frequency in mutants/10(6) survivors at a dose that resulted in one adduct per 10(6) base pairs) with our previous studies of BPDE-1, we found that BPDE-II was 4-5 times less efficient as a mutagen than BPDE-I. This difference in mutational efficiency could be explained in part by the increased rate of loss of BPDE-II adducts from the cellular DNA, part of which was due to an increased rate of enzymatic removal of these lesions compared with the removal of BPDE-I adducts.  相似文献   

8.
DNA repair and replication were examined in diploid human fibroblasts after treatment with (+/-)-r-7,t-8-dihydroxy-t-9,10-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene (BPDE-I). Unscheduled DNA synthesis exhibited a linear response to BPDE-I concentrations up to 1.5 microM and a saturation plateau after higher concentrations. Maximal unscheduled DNA synthesis was observed in the first hour after treatment with synthesis diminishing progressively thereafter. Half-maximal unscheduled DNA synthesis was seen within 4-6 h after treatment with 0.7 microM BPDE-I. DNA replication was inhibited by BPDE-I in a dose- and time-dependent fashion. The mechanisms of this inhibition were characterized by velocity sedimentation of pulse-labeled nascent DNA in alkaline sucrose gradients. Very low concentrations of BPDE-I (0.03 and 0.07 microM) were found to inhibit replicon initiation by up to 50% within 30-60 min after treatment. Recovery of initiation following these low concentrations was evident within 3 h after treatment. Higher concentrations of carcinogen inhibited DNA synthesis in active replicons. This effect was manifested by a reduction in incorporation of precursor into replication intermediates of greater than 1 X 10(7) Da with the concurrent production of abnormally small nascent DNA. When viewed 45 min after treatment with 0.17 microM BPDE-I the combination of these two effects partially masked the inhibition of replicon initiation. However, even after treatment with 0.33 microM BPDE-I an effect on initiation was evident. These results reveal a pattern of response to BPDE-I that is quite similar to that produced by 254 nm radiation.  相似文献   

9.
Abstract

A theoretical model is proposed for the covalent binding of (+) 7 β,8α-dihydroxy-9α, 10α- epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene denoted by BPDE I(+), to N2 on guanine. The DNA must kink a minimum of 39° to allow proper hybrid configurations about the C10 and N2 atoms involved in bond formation and to allow stacking of the pyrene moiety with the non-bonded adjacent base pair. Conservative (same sugar puckers and glycosidic angles as in B-DNA) and non-conservative (alternating sugar puckers as in intercalation sites) conformations are found and they are proposed structures in pathways connecting B-DNA, an intercalation site, and a kink site in the formation of a covalently intercalative bound adduct of BPDE I(+) to N2 on guanine. Stereographic projections are presented for (3′) and (5′) binding in the DNA. Experimental data for bending of DNA by BPDE, orientation of BPDE in DNA and unwinding of superhelical DNA is explained. The structure of a covalent intercalative complex is predicted to result from the reaction. Also, an anti ? syn transition of guanine results in a structure which allows the DNA to resume its overall B-form. The only change is that guanine has been rotated by 200° about its glycosidic bond so that the BPDE I(+) is bound in the major groove. The latter step may allow the DNA to be stored with an adduct which may produce an error in the genetic code.  相似文献   

10.
Kinetic flow dichroism studies indicate that the (+) enantiomer of 7 beta, 8 alpha-dihydroxy-9 alpha, 10 alpha-epoxy-7,8,9,10-tetrahydrobenzo(a)pyrene physically bound at intercalative-type sites in double-stranded DNA undergoes covalent binding reactions to form adducts at external binding sites. The conformation of the non-covalent complex derived from the (-) stereoisomer is also intercalative in nature, but the conformations of the covalent adducts are heterogeneous and are characterized by both intercalative-type and external conformations. It is suggested that the distinctly higher biological activity of the (+) enantiomer relative to the activity of the (-) enantiomer may be related to the preponderance of 7,8,9-triol benzo(a)pyrene residues covalently linked to deoxyguanine and located at external binding sites in the DNA adducts.  相似文献   

11.
R Xu  B Mao  J Xu  B Li  S Birke  C E Swenberg    N E Geacintov 《Nucleic acids research》1995,23(12):2314-2319
The apparent persistence length of enzymatically linearized pIBI30 plasmid DNA molecules approximately 2300 bp long, as measured by a hydrodynamic linear flow dichroism method, is markedly decreased after covalent binding of the highly tumorigenic benzo[a]pyrene metabolite 7R,8S-dihydroxy-9S,10R-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene [(+)-anti-BPDE]. In striking contrast, the binding of the non-tumorigenic, mirror-image 7S,8R,9R,10S enantiomer [(-)-anti-BPDE] to DNA has no measurable effect on its alignment in hydrodynamic flow gradients (< or = 2.2% of the DNA bases modified). In order to relate this effect to BPDE-nucleotide lesions of defined stereochemistry, the bending induced by site-specifically placed and stereochemically defined (+)- and (-)-anti-BPDE-N2-dG lesions in an 11mer deoxyoligonucleotide duplex was studied by ligation and gel electrophoresis methods. Out of the four stereochemically isomeric anti-BPDE-N2-deoxyguanosyl (dG) adducts with either (+)-trans, (-)-trans, (+)-cis, and (-)-cis adduct stereochemistry, only the (+)-trans adduct gives rise to prominent bends or flexible hinge joints in the modified oligonucleotide duplexes. Since both anti-BPDE enantiomers are known to bind preferentially to dG (> or = 85%), these observations can account for the differences in persistence lengths of DNA modified with either (+)-anti-BPDE or the chiral (-)-anti-BPDE isomer.  相似文献   

12.
DNA repair and replication were examined in diploid human fibroblasts after treatment with (±)-r-7,t-8-dihydroxy-t-9,10-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene (BPDE-I). Unscheduled DNA synthesis exhibited a linear response to BPDE-I concentrations up to 1.5 μM and a saturation plateau after higher concentrations. Maximal unscheduled DNA synthesis was observed in the first hour after treatment with synthesis diminishing progressively thereafter. Half-maximal unscheduled DNA synthesis was seen within 4–6 h after treatment with 0.7 μM BPDE-I. DNA replication was inhibited by BPDE-I in a dose- and time-dependent fashion. The mechanisms of this inhibition were characterized by velocity sedimentation of pulse-labeled nascent DNA in alkaline sucrose gradients. Very low concentrations of BPDE-I (0.03 and 0.07 μM) were found to inhibit replicon initiation by up to 50% within 30–60 min after treatment. Recovery of initiation following these low concentrations was evident within 3 h after treatment. Higher concentrations of carcinogen inhibited DNA synthesis in active replicons. This effect was manifested by a reduction in incorporation of precursor into replication intermediates of greater than 1·107 Da with the concurrent production of abnormally small nascent DNA. When viewed 45 min after treatment with 0.17 μM BPDE-I the combination of these two effects partially masked the inhibition of replicon initiation. However, even after treatment with 0.33 μM BPDE-I an effect on initiation was evident. These results reveal a pattern of response to BPDE-I that is quite similar to that produced by 254 nm radiation.  相似文献   

13.
The primary mode of non-covalent interaction of the strong carcinogen, benzo(a)pyrene diol epoxide, with DNA is through intercalation. It has variously been suggested that intercalative complexes may be prerequisite for either covalent binding or DNA-catalysed hydrolysis of the epoxide or both. Geacintov [Geacintov, N. E. (1986). Carcinogenesis 7, 589.] has recently argued that intercalation is important in covalent binding and presented theoretical constructs consistent with this proposal. A more general theoretical model is presented here which includes the possibilities that either catalysis of hydrolysis or covalent binding of benzo(a)pyrene diol epoxide DNA can occur (a) in an intercalation complex, or (b) without formation of a detectable, physically bound complex. It is shown that a variety of possible mechanisms formulated under this general theory lead to equations for overall reaction rates and covalent binding fractions which are all of the same form with respect to DNA concentration dependence. A consequence of this is that experimental studies of the dependence of hydrolysis rates and covalent binding fractions on DNA concentration do not distinguish between the various possible mechanisms. These findings are discussed in relation to the interactions of benzo(a)pyrene diol epoxide with chromatin in cells.  相似文献   

14.
A comparison of the binding of (–)- and (+)-[3H]nicotine to rat brain membranes revealed that only the (–)-enantiomer showed high affinity binding; while the (+)-enantiomer was at least 1/10 as effective as the (–)-enantiomer when in competition with (–)-[3H]nicotine as the ligand. Positive cooperativity, which is observed with (–)-[3H]nicotine as the presence of low concentrations of (+)-nicotine, may account for the seeming paradox.  相似文献   

15.
Benzo[a]pyrene (BP) is activated within tissues in both a regio- and a stereoselective manner and, since human skin is sensitive to tumour induction by polycyclic aromatic hydrocarbons (PAH), the steroselective metabolism of BP in this tissue has been investigated. Samples of skin from eleven individuals were treated with [3H]BP in short-term organ culture. Two samples were also treated with mixtures of [14C](+)- and (-)-trans-7,8-dihydro-7,8-dihydroxybenzo[a]pyrene (BP-7,8-dihydrodiol) in varying proportions. Following application of [3H] BP, more 7,8-dihydrodiol was recovered from the skin itself than from the culture fluid in ten cases; no 7.8-dihydrodiol was detected in extracts from the eleventh. The 7,8-dihydrodiol metabolite was extracted predominantly (range 74-greater than 99%) as the (-)-enantiomer in nine of these ten patients, although proportionately more (+)-enantiomer was recovered from the culture fluid than from the skin in each case. The relative proportions of [3H]BP tetrols derived from syn- and anti-7,8-dihydroxy-9,10-oxy-7,8,9,10-tetrahydroxybenzo[a]pyrene (BPDE) detected in these extracts was more variable. When skin samples were treated with [14C]BP-7,8-dihydrodiol, more anti- than syn-BPDE-derived tetrols were extracted, irrespective of the optical purity of the dihydrodiol applied. These findings provide evidence for interindividual variations in the stereoselective metabolism of BP, which may be of some importance in determining individual susceptibility to PAH-induced skin carcinogenesis.  相似文献   

16.
(±)-7β,8α-Dihydroxy-9β,10β-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene (diol epoxide-1) and (±)-7β,8α-dihydroxy-9α,10α-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene (diol epoxide-2) are highly mutagenic diol epoxide diastereomers that are formed during metabolism of the carcinogen (±)-trans-7,8-dihydroxy-7,8-dihydrobenzo[a]pyrene. Remarkable stereoselectivity has been observed on metabolism of the optically pure (+)- and (?)-enantiomers of the dihydrodiol which are obtained by separation of the diastereomeric diesters with (?)-α-methoxy-α-trifluoromethylphenylacetic acid. The high stereoselectivity in the formation of diol epoxide-1 relative to diol epoxide-2 was observed with liver microsomes from 3-methylcholanthrene-treated rats and with a purified cytochrome P-448-containing monoxygenase system where the (?)-enantiomer produced a diol epoxide-2 to diol epoxide-1 ratio of 6 : 1 and the (+)-enantiomer produced a ratio of 1 : 22. Microsomes from control and phenobarbital-treated rats were less stereospecific in the metabolism of enantiomers of BP 7,8-dihydrodiol. The ratio of diol epoxide-2 to diol epoxide-1 formed from the (?)- and (+)-enantiomers with microsomes from control rats was 2 : 1 and 1 : 6, respectively. Both enantiomers of BP 7,8-dihydrodiol were also metabolized to a phenolic derivative, tentatively identified as 6,7,8-trihydroxy-7,8-dihydrobenzo[a]pyrene, which accounted for ~30% of the total metabolites formed by microsomes from control and phenobarbital-pretreated rats whereas this metabolite represents ~5% of the total metabolites with microsomes from 3-methylcholanthrene-treated rats. With benzo[a]pyrene as substrate, liver microsomes produced the 4,5-, 7,8- and 9,10-dihydrodiol with high optical purity (>85%), and diol epoxides were also formed. Most of the optical activity in the BP 7,8-dihydrodiol was due to metabolism by the monoxygenase system rather than by epoxide hydrase, since hydration of (±)-benzo[a]pyrene 7,8-oxide by liver microsomes produced dihydrodiol which was only 8% optically pure. Thus, the stereospecificity of both the monoxygenase system and, to a lesser extent, epoxide hydrase plays important roles in the metabolic activation of benzo[a]pyrene to carcinogens and mutagens.  相似文献   

17.
The effect of the enantiomers of a novel 5-HT2 receptor antagonist, (+/-)-(1R,3S)-1-[2-[4-[3-(p-fluorophenyl)-1-indanyl]-piperazinyl] ethyl]-2-imidazolidinone, was studied on serotonin (5-HT), noradrenaline (NA), potassium (K+), and calcium (Ca2+)-induced contractions in isolated rat thoracic aorta. The enantiomers shifted the 5-HT, NA, K+, and Ca2+ concentration-response curves to the right in a concentration-dependent manner and depressed the maximal contractile responses. The (+)-enantiomer was a far more potent inhibitor of 5-HT-induced contractions than the (-)-enantiomer. The (+)-enantiomer and phentolamine, both at 10(-6) M, had equal inhibitory effects on NA-evoked contractions. The (+)-enantiomer was again more potent inhibiting NA-induced contractions than the (-)-enantiomer. Both enantiomers had an equieffective inhibitory effect on K+ and Ca2(+)-induced contractions. The results show that the 5-HT and alpha-adrenoceptor antagonism of the two enantiomers is stereoselective, the (+)-enantiomer being more potent than the (-)-enantiomer. In contrast the enantiomers had equal, nonstereoselective inhibitory effects on K+ and Ca2(+)-evoked contractions.  相似文献   

18.
Here we show that several cell signaling inhibitors have effect on cyp1a1 expression and the metabolism of benzo[a]pyrene (B[a]P) in Hepa1c1c7 cells. The CYP1A1 inhibitor alpha-naphthoflavone (alpha-NF), the p53 inhibitor pifithrin-alpha (PFT-alpha), the ERK inhibitors PD98059 and U0126, and the p38 MAPK inhibitors SB202190 and PD169316 induced the expression and level of cyp1a1 protein. On the other hand, during the first h the inhibitors appeared to reduce the metabolism of B[a]P as measured by the generation of tetrols and by covalent binding of B[a]P to macromolecules. In contrast, the phosphatidylinositol-3 (PI-3) kinase inhibitor wortmannin, had neither an effect on the cyp1a1 expression nor the B[a]P-metabolism. In order to avoid these unspecific effects, we characterized the mechanisms involved in the apoptotic effects of B[a]P-metabolites. B[a]P and the B[a]P-metabolites B[a]P-7,8-DHD and BPDE-I induced apoptosis, whereas B[a]P-4,5-DHD had no effect. B[a]P, B[a]P-7,8-DHD and BPDE-I induced an accumulation and phosphorylation of p53, while the Bcl-2 proteins Bcl-xl, Bad and Bid were down-regulated. Interestingly, the levels of anti-apoptotic phospho-Bad were up-regulated in response to B[a]P as well as to B[a]P-7,8-DHD and BPDE-I. Both p38 MAPK and JNK were activated, but the p38 MAPK inhibitors were not able to inhibit BPDE-I-induced apoptosis. PFT-alpha reduced the BPDE-I-induced apoptosis, while both the PI-3 kinase inhibitor and the ERK inhibitors increased the apoptosis in combination with BPDE-I. BPDE-I also triggered apoptosis in primary cultures of rat lung cells. In conclusion, often used cell signaling inhibitors both enhanced the expression and the level of cyp1a1 and more directly acted as inhibitors of cyp1a1 metabolism of B[a]P. However, studies with the B[a]P-metabolite BPDE-I supported the previous suggestion that p53 has a role in the pro-apoptotic signaling pathway induced by B[a]P. Furthermore, these studies also show that the reactive metabolites of B[a]P induce the anti-apoptotic signals, Akt and ERK. Neither the induction nor the activity of p38 MAPK and JNK seems to be of major importance for the B[a]P-induced apoptosis.  相似文献   

19.
DNA damage caused by the binding of the tumorigen 7R,8S-diol 9S,10R-epoxide (B[a]PDE), a metabolite of bezo[a]pyrene, to guanine in CpG dinucleotide sequences could affect DNA methylation and, thus, represent a potential epigenetic mechanism of chemical carcinogenesis. In this work, we investigated the impact of stereoisomeric (+)- and (-)-trans-anti-B[a]P-N(2)-dG adducts (B(+) and B(-)) on DNA methylation by prokaryotic DNA methyltransferases M.SssI and M.HhaI. These two methyltransferases recognize CpG and GCGC sequences, respectively, and transfer a methyl group to the C5 atom of cytosine (C). A series of 18-mer unmethylated or hemimethylated oligodeoxynucleotide duplexes containing trans-anti-B[a]P-N(2)-dG adducts was generated. The B(+) or B(-) residues were introduced either 5' or 3' adjacent or opposite to the target 2'-deoxycytidines. The B[a]PDE lesions practically produced no effect on M.SssI binding to DNA but reduced M.HhaI binding by 1-2 orders of magnitude. In most cases, the benzo[a]pyrenyl residues decreased the methylation efficiency of hemimethylated and unmethylated DNA by M.SssI and M.HhaI. An absence of the methylation of hemimethylated duplexes was observed when either the (+)- or the (-)-trans-anti-B[a]P-N(2)-dG adduct was positioned 5' to the target dC. The effects observed may be related to the minor groove conformation of the bulky benzo[a]pyrenyl residue and to a perturbation of the normal contacts of the methyltransferase catalytic loop with the B[a]PDE-modified DNA. Our results indicate that a trans-anti-B[a]P-N(2)-dG lesion flanking a target dC in the CpG dinucleotide sequence on its 5'-side has a greater adverse impact on methylation than the same lesion when it is 3' adjacent or opposite to the target dC.  相似文献   

20.
Benzo[a]pyrene is a polycyclic aromatic hydrocarbon (PAH) associated with potent carcinogenic activity. Mutagenesis induced by benzo[a]pyrene DNA adducts is believed to involve error-prone translesion synthesis opposite the lesion. However, the DNA polymerase involved in this process has not been clearly defined in eukaryotes. Here, we provide biochemical evidence suggesting a role for DNA polymerase eta (Poleta) in mutagenesis induced by benzo[a]pyrene DNA adducts in cells. Purified human Poleta predominantly inserted an A opposite a template (+)- and (-)-trans-anti-BPDE-N2-dG, two important DNA adducts of benzo[a]pyrene. Both lesions also dramatically elevated G and T mis-insertion error rates of human Poleta. Error-prone nucleotide insertion by human Poleta was more efficient opposite the (+)-trans-anti-BPDE-N2-dG adduct than opposite the (-)-trans-anti-BPDE-N2-dG. However, translesion synthesis by human Poleta largely stopped opposite the lesion and at one nucleotide downstream of the lesion (+1 extension). The limited extension synthesis of human Poleta from opposite the lesion was strongly affected by the stereochemistry of the trans-anti-BPDE-N2-dG adducts, the nucleotide opposite the lesion, and the sequence context 5' to the lesion. By combining the nucleotide insertion activity of human Poleta and the extension synthesis activity of human Polkappa, effective error-prone lesion bypass was achieved in vitro in response to the (+)- and (-)-trans-anti-BPDE-N2-dG DNA adducts.  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