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 15 毫秒
1.
Mutations induced by polycyclic aromatic hydrocarbons (PAH) are expected to be produced when error-prone DNA replication occurs across unrepaired DNA lesions formed by reactive PAH metabolites such as diol epoxides. The mutagenicity of the two PAH-diol epoxides (+)-anti-7,8-dihydroxy-9,10-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene (BPDE) and (+/-)-anti-11,12-dihydroxy-13,14-epoxy-11,12,13,14-tetrahydrodibenzo[a,l]pyrene (DBPDE) was compared in nucleotide excision repair (NER) proficient and deficient hamster cell lines. We applied the (32)P-postlabelling assay to analyze adduct levels and the hprt gene mutation assay for monitoring mutations. It was found that the mutagenicity per target dose was 4 times higher for DBPDE compared to BPDE in NER proficient cells while in NER deficient cells, the mutagenicity per target dose was 1.4 times higher for BPDE. In order to investigate to what extent the mutagenicity of the different adducts in NER proficient cells was influenced by repair or replication bypass, we measured the overall NER incision rate, the rate of adduct removal, the rate of replication bypass and the frequency of induced recombination in the hprt gene. The results suggest that NER of BPDE lesions are 5 times more efficient than for DBPDE lesions, in NER proficient cells. However, DBPDE adducts block replication more efficiently and also induce 6 times more recombination events in the hprt gene than adducts of BPDE, suggesting that DBPDE adducts are, to a larger extent, bypassed by homologous recombination. The results obtained here indicate that the mutagenicity of PAH is influenced not only by NER, but also by replication bypass fidelity. This has been postulated earlier based on results using in vitro enzyme assays, but is now also being recognized in terms of forward mutations in intact mammalian cells.  相似文献   

2.
Hu W  Feng Z  Tang MS 《Biochemistry》2004,43(44):14282-14289
Chromium(VI) [Cr(VI)], a ubiquitous environmental contaminant, is a well-known carcinogen to both humans and experimental animals, although it is a weak mutagen by itself. Occupational exposure to Cr(VI) is strongly associated with a high incidence of lung cancer, but the underlying mechanisms remain unclear. Tobacco smoking is the major cause of lung cancer, and polycyclic aromatic hydrocarbons (PAHs) in tobacco smoke are the major etiological agents. Since humans are frequently exposed to both Cr(VI) and PAHs, it is possible that Cr(VI) and PAHs have a synergistic effect on mutagenecity and cytotoxicity that contributes to the high incidence of lung cancer associated with exposure to both agents. In this study, we tested this possibility by determining the effect of Cr(VI) exposure on (+/-)-anti-7beta,8alpha-dihydroxy-9alpha,10alpha-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene (BPDE, an active metabolite of PAHs) induced cytotoxicity, mutagenicity, and DNA adduct formation in Chinese hamster ovary (CHO) cells. Using the adenine phosphoribosyltransferase (APRT(+)) --> APRT(-) forward mutation assay, we found that while Cr(VI) alone induced low mutation frequency, it greatly enhanced BPDE-induced mutations in nucleotide excision repair (NER)-proficient CHO cells. Cr(VI) exposure also greatly enhanced BPDE-induced killing in NER-proficient cells. It is known that the cytotoxicity and mutagenicity of BPDE are mainly caused by the formation of DNA adduct, which are removed by NER. To test the possibility that the enhancement of cytotoxicity and mutagenicity by Cr(VI) is caused by the inhibition of NER, NER-deficient cells were used, and the enhancement effects of Cr(VI) were not observed in those cells. We further found that while Cr(VI) exposure does not change the total BPDE-DNA adduct formation, it significantly inhibited the repair of BPDE-DNA adducts from genomic DNA in NER-proficient cells. Using a host cell reactivation assay, we found that the repair of BPDE-DNA adduct in a luciferase reporter gene is greatly inhibited after Cr(VI) exposure in NER-proficient cells while not in NER-deficient cells. Together these results clearly demonstrate that Cr(VI) exposure can greatly enhance the mutagenicity and cytotoxicity of PAHs by inhibiting the cellular NER pathway, and this may constitute an important mechanism for Cr(VI)-induced human carcinogenesis.  相似文献   

3.
A benzo[a]pyrene(BP)-Gua adduct was extracted in the urine of rats treated with BP. Some (0.15%) of the administered dose of BP was excreted as BP-Gua within 48 h. A double labelling experiment demonstrated that the excreted product contained both a BP and a Gua moiety. Partially hepatectomized rats treated with [14C]Gua during the regenerative phase were injected with [3H]BP and the urine collected and processed by chromatographic procedures. The adduct had similar chromatographic properties to the adduct released from human PLC/5 cells treated with 7,8-dihydroxy-9,10-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene (BPDE) and co-chromatographed with 7-BPDE-Gua released from BPDE-adducted DNA under aqueous conditions. Detection and quantitation of BP-Gua offers an alternative, non-invasive method of monitoring individuals exposed to carcinogenic polycyclic aromatic hydrocarbons (PAHs).  相似文献   

4.
The effects of separate lipoproteins or of serum with high or low lipoprotein concentrations on formation of lipophilic carcinogen adducts with DNA and on mutagenicity of the carcinogen was investigated using V79 Chinese hamster lung cells. Binding of r-7,t-8-dihydroxy-t-9,10-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene (BPDE) to DNA and BPDE induction of 6-thioguanine (6-TG)-resistant mutants in V79 cells was significantly lower after 1 or 4 h when the medium was supplemented with purified HDL, and was lower after 1 h but not 4 h when the medium was supplemented with serum containing a high concentration of mixed lipoproteins (LP). Cells grown in medium without serum or LP supplementation exhibited the highest levels of both BPDE-DNA adduct formation and mutagenesis after 1 h. At 1 h, cells exposed to BPDE in LDL-supplemented medium showed decreased adduct formation and mutagenesis when compared to cells treated with BPDE in PBS-supplemented medium. After 4 h, cells treated with BPDE in LDL-supplemented medium gave the highest levels of adduct formation and the highest mutation frequency. These results suggest that both LDL and HDL effectively decrease the concentration of BPDE available to V79 cells exposed to the mutagen for short periods of time, resulting in decreased interaction of BPDE with DNA and decreased BPDE-associated mutagenesis, but that both BPDE-DNA adduct formation and mutagenesis increased as a function of increased exposure time in the presence of LDL. The results suggest that LDL, but not HDL, uptake by adsorptive endocytosis may be associated with potentiated entry of BPDE into V79 cells as a function of time.  相似文献   

5.
When the benzo(a)pyrene diol epoxide (±)-7β,8α-dihydroxy-9α,10α-epoxy-7,8,9,10-tetrahydrobenzo(a)pyrene (BPDE) is mixed into a DNA solution, a 10nm red shift in the absorption maximum of BPDE appears at 354nm which is due to a non-covalent intercalation complex. The major reaction pathway at this intercalation site is the hydrolysis of BPDE to its tetraol which is accompanied by a decrease in the absorbance and a shift from 354 to 353nm (the latter is due to intercalated tetraol). The non-covalent binding constants are approximately 8200M?1 for BPDE and 3300M?1 for the tetraol at 25°C, pH 7.0. Covalent adduct formation between BPDE and DNA occurs either at another, external binding site, or after some rearrangement of the intercalated BPDE, since covalent adducts display a 345nm absorption maximum (2nm red shift only).  相似文献   

6.
Tretyakova N  Matter B  Jones R  Shallop A 《Biochemistry》2002,41(30):9535-9544
The mutagenicity of a prominent tobacco carcinogen, benzo[a]pyrene (B[a]P), is believed to result from chemical reactions between its diol epoxide metabolite, (+)-anti-7r,8t-dihydroxy-c9,10-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene (BPDE), and DNA, producing promutagenic lesions, e.g., (+)-trans-anti-7R,8S,9S-trihydroxy-10S-(N(2)-deoxyguanosyl)-7,8,9,10-tetrahydrobenzo[a]pyrene (N(2)-BPDE-dG). Previous studies used the DNA repair enzyme UvrABC endonuclease in combination with ligation-mediated PCR (LMPCR) to demonstrate an increased reactivity of BPDE toward guanine nucleobases within codons 157, 248, and 273 of the p53 tumor suppressor gene (Denissenko, M. F., Pao, A., Tang, M., and Pfeifer, G. P. Science 274, 430-432). These sites are also "hot spots" for mutations observed in lung tumors of smokers, suggesting an involvement of B[a]P in the initiation of lung cancer. However, the LMPCR approach relies on the ability of the repair enzyme to excise BPDE-induced lesions, and thus the slowly repaired lesions may escape detection. Furthermore, BPDE-DNA adduct structure and stereochemistry cannot be determined. In the present work, we performed a direct quantitative analysis of N(2)-BPDE-dG originating from specific guanine nucleobases within p53- and K-ras-derived DNA sequences by using a stable isotope labeling-mass spectrometry approach recently developed in our laboratory. (15)N-labeled dG was placed at defined positions within DNA sequences derived from the K-ras proto-oncogene and p53 tumor suppressor gene, the two genes most frequently mutated in smoking-induced lung cancer. (15)N-labeled DNA was annealed to the complementary strands, followed by BPDE treatment and liquid chromatography-electrospray ionization tandem mass spectrometry analysis (HPLC-ESI-MS/MS) of N(2)-BPDE-dG lesions. The extent of adduct formation at (15)N-labeled guanine was determined directly from the HPLC-ESI-MS/MS peak area ratios of (15)N-N(2)-BPDE-dG and N(2)-BPDE-dG. BPDE-induced guanine adducts were produced nonrandomly along K-ras and p53 gene-derived DNA sequences, with over 5-fold differences in adduct formation depending on sequence context. N(2)-BPDE-dG yield was enhanced by the presence of 5-Me substituent at the cytosine base-paired with the target guanine nucleobase, an endogenous DNA modification characteristic for CpG dinucleotides within the p53 gene. In the K-ras-derived DNA sequence, the majority of N(2)-BPDE-dG adducts originated from the first position of the codon 12 (GGT), consistent with the large number of G --> T transversions observed at this nucleotide in smoking-induced lung cancer. On the contrary, the pattern of N(2)-BPDE-dG formation within the p53 exon 5 sequences did not correlate with the mutational spectrum in lung cancer, suggesting that factors other than N(2)-BPDE-dG formation are responsible for these mutations. The stable isotope labeling HPLC-ESI-MS/MS approach described in this work is universally applicable to studies of modifications to isolated DNA by other carcinogens and alkylating drugs.  相似文献   

7.
The solution structure of the adduct derived from the covalent bonding of the fjord region (+)-(11S, 12R, 13R, 14S) stereoisomer of anti -11,12-dihydroxy-13,14-epoxy-11,12,13, 14-tetrahydrobenzo[g]chrysene, (+)- anti -B[g]CDE, to the exocyclic N(6)amino group of the adenine residue dA6, (designated (+)- trans-anti -(B[g]C)dA6), positioned opposite a thymine residue dT17 in the DNA sequence context d(C1-T2-C3-T4-C5-(B[g]C)A6-C7-T8-T9-C10-C11). d(G12-G13-A14-A15-G16-T17-G18-A19-G20++ +-A21-G22) (designated (B[g]C)dA. dT 11-mer duplex), has been studied using structural information derived from NMR data in combination with molecular dynamics (MD) calculations. The solution structure of the (+)- trans-anti -(B[g]C)dA.dT 11-mer duplex has been determined using an MD protocol where both interproton distance and dihedral angle restraints deduced from NOESY and COSY spectra are used during the refinement process, followed by additional relaxation matrix refinement to the observed NOESY intensities to account for spin diffusion effects. The results established that the covalently attached benzo[g]chrysene ring intercalates into the DNA helix directed towards the 5'-side of the modified strand and stacks predominantly with dT17 when intercalated between dC5.dG18 and (B[g]C)dA6.dT17 base-pairs. All base-pairs, including the modified (B[g]C)dA6.dT17 base-pair, are aligned through Watson-Crick pairing as in normal B -DNA. In addition, the potential strain associated with the highly sterically hindered fjord region of the aromatic portion of the benzo[g]chrysenyl ring is relieved through the adoption of a non-planar, propeller-like geometry within the chrysenyl ring system. This conformation shares common structural features with the related (+)- trans-anti -(B[c]Ph)dA adduct in the identical base sequence context, derived from the fjord region (+)-(1S,2R,3R,4S)-3, 4-dihydroxy-1,2-epoxy-1,2,3,4-tetrahydrobenzo[c]phenanthrene stereoisomer, in which intercalation is also observed towards the 5'-side of the modified dA6.dT17 base-pair.  相似文献   

8.
A new method was developed to determine the mutagenic efficacy of a suspected mutagen by employing green fluorescent protein (GFP) as a direct biosensor for mutation detection. Alterations in target gene (AcGFP1) expression after mutagen [(±)-7p,8a-dihydroxy-9a,10a-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene (BPDE)] treatment were measured to detect the mutagenic efficacy of the carcinogen. In contrast to mutagen treatment of the entire plasmid or cell culture, the target AcGFP1gene devoid of the plasmid backbone was exposed to BPDE (10–500 μM) to eliminate the need for an additional fusion gene. Shuttle vectors (pAcGFP-N1) were religated to the AcGFP1 gene with BPDE adducts (0–8.59 μM) and replicated in the eukaryotic host. This approach eliminated false-negative errors in target gene expression that arose from BPDE adduct formation in the residual plasmid backbone rather than in the AcGFP1 gene. Determination of the BPDE–AcGFP1 adducts allowed the quantitative mutagenic effect of the BPDE adducts on AcGFP1 gene expression to be monitored. The results obtained with flow cytometry and confocal microscopy validate our method and demonstrate efficient and direct use of GFP as a biosensor for mutation detection.  相似文献   

9.
The structural origin underlying differential nucleotide excision repair (NER) susceptibilities of bulky DNA lesions remains a challenging problem. We investigated the 10S (+)-trans-anti-[BP]-N(2)-2'-deoxyguanosine (G*) adduct in double-stranded DNA. This adduct arises from the reaction, in vitro and in vivo, of a major genotoxic metabolite of benzo[a]pyrene (BP), (+)-(7R,8S,9S,10R)-7,8-dihydroxy-9,10-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene, with the exocyclic amino group of guanine. Removal of this lesion by the NER apparatus in cell-free extracts has been found to depend on the base sequence context in which the lesion is embedded, providing an excellent opportunity for elucidating the properties of the damaged DNA duplexes that favor NER. While the BP ring system is in the B-DNA minor groove, 5' directed along the modified strand, there are orientational distinctions that are sequence dependent and are governed by flanking amino groups [Nucleic Acids Res.35 (2007), 1555-1568]. To elucidate sequence-governed NER susceptibility, we conducted molecular dynamics simulations for the 5'-...CG*GC..., 5'-...CGG*C..., and 5'-...TCG*CT... adduct-containing duplexes. We also investigated the 5'-...CG*IC... and 5'-...CIG*C... sequences, which contain "I" (2'-deoxyinosine), with hydrogen replacing the amino group in 2'-deoxyguanosine, to further characterize the structural and dynamic roles of the flanking amino groups in the damaged duplexes. Our results pinpoint explicit roles for the amino groups in tandem GG sequences on the efficiency of NER and suggest a hierarchy of destabilizing structural features that differentially facilitate NER of the BP lesion in the sequence contexts investigated. Furthermore, combinations of several locally destabilizing features in the hierarchy, consistent with a multipartite model, may provide a relatively strong recognition signal.  相似文献   

10.
In a previous study, we showed that naturally occurring organosulfur compounds (OSCs) from garlic and onion modulated the activation of carcinogen via the alteration of cytochromes P450. The present study was undertaken to determine the incidence of the in vivo induction of phase II enzymes by individual OSCs on the genotoxicity of several carcinogens. Diallyl sulfide (DAS), diallyl disulfide (DADS), dipropyl sulfide (DPS) and dipropyl disulfide (DPDS), were administered by gavage (1mmol/kg) to male SPF Wistar rats for 4 consecutive days. The effects of treatments on phase II enzymes and on the genotoxicity of carcinogens were evaluated with hepatic cytosols and microsomes from OSCs-treated rats. DADS strongly increased all the phase II enzymes activities examined, i.e. total glutathione S-transferase (GST) activity, mu GST activity, quinone reductase (QR) activity and epoxide hydrolase (EH) activity. In addition, DADS strongly increased the protein level of rGSTP1. QR activity, total and mu GST activities were also increased by DAS and DPDS whereas DPS increased only mu GST activity and QR activity. To assess the repercussions of these inductions on the genotoxicity of carcinogens, the effects of cytosols or microsomes from OSCs-treated rats on the mutagenicity of (+)-anti-7beta,8alpha-dihydroxy-9alpha,10alpha-oxy-7,8,9,10-tetrahydrobenzo[a]pyrene (BPDE), styrene oxide (SO) and 4-nitroquinoline 1-oxide (4-NQO) were measured in the Ames test. DADS showed a very effective antimutagenic activity against BPDE, SO and 4-NQO. DAS reduced the mutagenicity of BPDE and SO. In contrast, DPS and DPDS showed little efficient antimutagenic activity since they only reduced the mutagenicity of BPDE and 4-NQO, respectively. Interestingly, DADS appeared to be as effective as ethoxyquin, a model inducer of phase II enzymes, in both inducing phase II enzymes and inhibiting the mutagenicity of carcinogens. This study demonstrated that the antimutagenic activities of OSCs against several ultimate carcinogens were closely related to their ability to induce phase II enzymes.  相似文献   

11.
Gu Y  Xiao B  Wargo HL  Bucher MH  Singh SV  Ji X 《Biochemistry》2003,42(4):917-921
Murine class alpha glutathione S-transferase subunit types A2 (mGSTA2-2) and A1 (mGSTA1-1) have high catalytic efficiency for glutathione (GSH) conjugation of the ultimate carcinogenic metabolite of benzo[a]pyrene, (+)-anti-7,8-dihydroxy-9,10-oxy-7,8,9,10-tetrahydrobenzo[a]pyrene, [(+)-anti-BPDE]. Only 10 residues differ between the sequences of mGSTA1-1 and 2-2. However, the catalytic efficiency of mGSTA1-1 for GSH conjugation of (+)-anti-BPDE is >3-fold higher as compared with mGSTA2-2. The crystal structure of mGSTA1-1 in complex with the GSH conjugate of (+)-anti-7,8-dihydroxy-9,10-oxy-7,8,9,10-tetrahydrobenzo[a]pyrene (GSBpd) reveals that R216 and I221 in the last helix play important roles in catalysis [Gu, Y., Singh, S. V., and Ji, X. (2000) Biochemistry 39, 12552-12557]. The crystal structure of mGSTA2-2 in complex with GSBpd has been determined, which reveals a different binding mode of GSBpd. Comparison of the two structures suggests that residues 207 and 221 are responsible for the different binding mode of GSBpd and therefore contribute to the distinct catalytic efficiency of the two isozymes.  相似文献   

12.
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   

13.
The UvrABC nuclease system from Escherichia coli removes DNA damages induced by a wide range of chemical carcinogens with variable efficiencies. The interactions with UvrABC proteins of the following three lesions site-specifically positioned in DNA, and of known conformations, were investigated: (i) adducts derived from the binding of the (-)-(7S,8R,9R,10S) enantiomer of 7,8-dihydroxy-9, 10-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene [(-)-anti-BPDE] by cis-covalent addition to N(2)-2'-deoxyguanosine [(-)-cis-anti-BP-N(2)-dG], (ii) an adduct derived from the binding of the (+)-(1R,2S,3S,4R) enantiomer of 1,2-dihydroxy-3,4-epoxy-1,2,3, 4-tetrahydro-5-methylchrysene [(+)-anti-5-MeCDE] by trans addition to N(2)-2'-deoxyguanosine [(+)-trans-anti-MC-N(2)-dG], and (iii) a C8-2'-deoxyguanosine adduct (C8-AP-dG) formed by reductively activated 1-nitropyrene (1-NP). The influence of these three different adducts on UvrA binding affinities, formation of UvrB-DNA complexes by quantitative gel mobility shift analyses, and the rates of UvrABC incision were investigated. The binding affinities of UvrA varied among the three adducts. UvrA bound to the DNA adduct (+)-trans-anti-MC-N(2)-dG with the highest affinity (K(d) = 17 +/- 2 nM) and to the DNA containing C8-AP-dG with the least affinity (K(d) = 28 +/- 1 nM). The extent of complex formation with UvrB was also the lowest with the C8-AP-dG adduct. 5' Incisions occurred at the eighth phosphate from the modified guanine. The major 3' incision site corresponded to the fifth phosphodiester bond for all three adducts. However, additional 3' incisions were observed at the fourth and sixth phosphates in the case of the C8-AP-dG adduct, whereas in the case of the (-)-cis-anti-BP-N(2)-dG and (+)-trans-anti-MC-N(2)-dG lesions additional 3' cleavage occurred at the sixth and seventh phosphodiester bonds. Both the initial rate and the extent of 5' and 3' incisions revealed that C8-AP-dG was repaired less efficiently in comparison to the (-)-cis-anti-BP-N(2)-dG and (+)-trans-anti-MC-N(2)-dG containing DNA adducts. Our study showed that UvrA recognizes conformational changes induced by structurally different lesions and that in certain cases the binding affinities of UvrA and UvrB can be correlated with the incision rates. The size of the bubble formed around the damaged site with mismatched bases also appears to influence the incision rates. A particularly noteworthy finding in this study is that UvrABC repair of a substrate with no base opposite C8-AP-dG was quite inefficient as compared to the same adduct with a C opposite it. These findings are discussed in terms of the available NMR solution structures.  相似文献   

14.
Braithwaite E  Wu X  Wang Z 《Mutation research》1999,424(1-2):207-219
DNA is frequently damaged by endogenous agents inside the cells. Some exogenous agents such as polycyclic aromatic hydrocarbons (PAHs) are ubiquitous in the environment and may thus contribute to the 'background' DNA damage in humans. DNA lesions are normally removed by various repair mechanisms. The major repair mechanisms for various DNA lesions are summarized. In contrast to the extensively studied repair mechanisms, much less is known about the relative repair efficiencies of various DNA lesions. Since DNA repair is a crucial defense against carcinogenesis, it may constitute an important factor affecting the carcinogenicity of DNA damaging agents. We have adopted a human cell-free system for measuring relative DNA repair efficiencies based on the concept of repair competition between acetylaminofluorene adducts and other DNA lesions of interest. Using this in vitro system, we determined the relative repair efficiencies of PAH adducts induced by: anti-(+/-)-benzo[a]pyrene-trans-7,8-dihydrodiol-9,10-epoxide (BPDE), anti-(+/-)-benz[a]anthracene-trans-3,4-dihydrodiol-1,2-epoxide (BADE-I), anti-(+/-)-benz[a]anthracene-trans-8,9-dihydrodiol-10, 11-epoxide (BADE-II), anti-(+/-)-benzo[b]fluoranthene-trans-9, 10-dihydrodiol-11,12-epoxide (BFDE), anti-(+/-)-chrysene-trans-1, 2-dihydrodiol-3,4-epoxide (CDE), and anti-(+/-)-dibenzo[a, l]pyrene-trans-11,12-dihydrodiol-13,14-epoxide (DBPDE). While damage by BPDE, DBPDE, CDE, and BFDE were repaired by nucleotide excision repair as efficiently as AAF adducts, the repair of BADE-I and BADE-II adducts were significantly slower in human cell extracts. Damage by DBPDE at 3 microM in vitro yielded approximately 5-fold higher DNA adducts than BPDE as determined by quantitative PCR. This potent DNA reactivity may account in part for the potent carcinogenicity of dibenzo[a,l]pyrene. The correlation of these results to the carcinogenic properties of the PAH compounds is discussed. Furthermore, we show that NER plays a role in AP site repair in vivo in the eukaryotic model organism yeast.  相似文献   

15.
The reversible DNA physical binding of a series of non-reactive metabolites and metabolite model compounds derived from benzo[a]pyrene (BP) has been examined in UV absorption and in fluorescence emission and fluorescence lifetime studies. Members of this series have steric and pi electronic properties similar to the highly carcinogenic metabolite trans-7,8-dihydroxy-anti-9,10-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene (BPDE) and the less potent metabolite 4,5-epoxy-4,5-dihydrobenzo(a)pyrene (4,5-BPE). The molecules examined are trans-7,8-dihydroxy-7,8-dihydrobenzo[a]-pyrene (7,8-di(OH)H2BP), 7,8,9,10-tetrahydroxytetrahydrobenzo[a]pyrene (tetrol) 7,8,9,10-tetrahydrobenzo[a]pyrene (7,8,9,10-H4BP), pyrene, trans-4,5-dihydroxy-4,5-dihydrobenzo[a]pyrene (4,5-di(OH)H2BP) and 4,5-dihydrobenzo[a]pyrene (4,5-H2BP). In 15% methanol at 23 degrees C the intercalation binding constants of the molecules studied lie in the range 0.79-6.1 X 10(3) M-1. Of all the molecules examined the proximate carcinogen 7,8-di(OH)-H2BP is the best intercalating agent. The proximate carcinogen has a binding constant which in UV absorption studies is found to be 2.8-6.0 times greater than that of the other hydroxylated metabolites. Intercalation is the major mode of binding for 7,8-di(OH)H2BP and accounts for more than 95% of the total binding. Details concerning the specific role of physical bonding in BP carcinogenesis remain to be elucidated. However, the present studies demonstrate that the reversible binding constants for BP metabolites are of the same magnitude as reversible binding constants which arise from naturally occurring base-base hydrogen bonding and pi stacking interactions in DNA. Furthermore, previous autoradiographic studies indicate that in human skin fibroblasts incubated in BP, pooling of the unmetabolized hydrocarbons occurs at the nucleus. The high affinity of 7,8-di(OH)H2BP for DNA may play a role in similarly elevating in vivo nuclear concentrations of the non-reactive proximate carcinogen.  相似文献   

16.
Polycyclic aromatic hydrocarbons (PAHs) are environmental pollutants, formed during incomplete burning of coal, oil and gas. Several PAHs have carcinogenic and mutagenic potencies, but these compounds must be activated in order to exert their mutagenic effects. One of the principal pathways proposed for metabolic activation of PAHs involves the cytochrome P450 enzymes. The DNA damaging potential of cytochrome P450-activated PAHs is generally associated with their bay and fjord regions, and the DNA repair response of PAHs containing such regions has been thoroughly studied. However, little is known about the repair of DNA damage resulting from metabolites from PAHs without bay and fjord regions. We have investigated the six-ringed PAH anthanthrene (dibenzo[def,mno]chrysene), which does not posses bay or fjord motifs. We analyzed the repair profile of human cell extracts and of cell cultures in response to DNA damage induced by cytochrome P450-activated anthanthrene. In cell extracts, functional nucleotide excision repair (NER) and mismatch repair (MMR) activities were necessary to trigger a response to anthanthrene metabolite-induced DNA damage. In cell cultures, NER was responsible for the repair of anthanthrene metabolite-induced DNA damage. However, when the NER pathway was inactivated, a residual repair pathway performed the DNA repair.  相似文献   

17.
The formation and removal of covalent adducts of racemic 7 beta, 8 alpha-dihydroxy-9 alpha,10 alpha-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene (BPDE I) was studied in nucleosomal DNA of confluent cultures of normal human fibroblasts (NF). For this purpose NF were prelabeled in their DNA with [14C]-thymidine and treated with [3H]BPDE I. The adducts were composed of 77% (7R)-N2-(7 beta, 8 alpha, 9 alpha-trihydroxy-7,8,9,10-tetrahydrobenzo[a]pyrene-10-yl)deoxyguanosine, 12% of the corresponding 7S-enantiomer and of minor amounts of adducts to cytosine and adenine. The adduct composition did not change significantly in 24-h post treatment incubation. Bulk mononucleosomes were prepared from micrococcal nuclease digested nuclei and their DNA analyzed by gel electrophoresis. The adduct concentrations were determined in 145 base pair (b.p.) nucleosomal core-DNA, 165 b.p. chromatosomal DNA and in total nuclear DNA. From these data the concentration in nucleosomal linker-DNA was calculated. The initial adduct distribution was non-random and 6.3 times higher in 47 b.p. linker-DNA relative to 145 b.p. core-DNA and 9.2 times higher in 27 b.p. linker-DNA relative to 165 b.p. chromatosomal DNA. Adduct removal was very rapid during the first 8 h and more efficient from linker-DNA than from core-DNA. After this early phase the adducts located in 145 b.p. core-DNA became refractory to further excision and represent a major fraction of the adducts persisting in DNA of NF over a prolonged period. In contrast, further adduct removal was observed from nucleosomal linker-DNA.  相似文献   

18.
The function of the human nucleotide excision repair (NER) apparatus is to remove bulky adducts from damaged DNA. In an effort to gain insights into the molecular mechanisms involved in the recognition and excision of bulky lesions, we investigated a series of site specifically modified oligonucleotides containing single, well-defined polycyclic aromatic hydrocarbon (PAH) diol epoxide-adenine adducts. Covalent adducts derived from the bay region PAH, benzo[a]pyrene, are removed by human NER enzymes in vitro. In contrast, the stereochemically analogous N(6)-dA adducts derived from the topologically different fjord region PAH, benzo[c]phenanthrene, are resistant to repair. The evasion of DNA repair may play a role in the observed higher tumorigenicity of the fjord region PAH diol epoxides. We are elucidating the structural and thermodynamic features of these adducts that may underlie their marked distinction in biologic function, employing high-resolution nuclear magnetic resonance studies, measurements of thermal stabilities of the PAH diol epoxide-modified oligonucleotide duplexes, and molecular dynamics simulations with free energy calculations. Our combined findings suggest that differences in the thermodynamic properties and thermal stabilities are associated with differences in distortions to the DNA induced by the lesions. These structural effects correlate with the differential NER susceptibilities and stem from the intrinsically distinct shapes of the fjord and bay region PAH diol epoxide-N(6)-adenine adducts.  相似文献   

19.
We report below on the solution structures of stereoisomeric "fjord" region trans-anti-benzo[c]phenanthrene-N2-guanine (designated (BPh)G) adducts positioned opposite cytosine within the (C-(BPh)G-C).(G-C-G) sequence context. We observe intercalation of the phenanthrenyl ring with stereoisomer-dependent directionality, without disruption of the modified (BPh)G.C base-pair. Intercalation occurs to the 5' side of the modified strand for the 1S stereoisomeric adduct and to the 3' side for the 1R stereoisomeric adduct, with the S and R-trans-isomers related to one another by inversion in a mirror plane at all four chiral carbon atoms on the benzylic ring. Intercalation of the fjord region BPh ring into the helix without disruption of the modified base-pair is achieved through buckling of the (BPh)G.C base-pair, displacement of the linkage bond from the plane of the (BPh)G base, adaptation of a chair pucker by the BPh benzylic ring and the propeller-like deviation from planarity of the BPh phenanthrenyl ring. It is noteworthy that intercalation without base-pair disruption occurs from the minor groove side for S and R-trans-anti BPh-N2-guanine adducts opposite C, in contrast to our previous demonstration of intercalation without modified base-pair disruption from the major groove side for S and R-trans-anti BPh-N6-adenine adducts opposite T. Further, these results on fjord region 1S and 1R-trans-anti (BPh)G adducts positioned opposite C are in striking contrast to earlier research with "bay" region benzo[a]pyrene-N2-guanine (designated (BP)G) adducts positioned opposite cytosine, where 10S and 10R-trans-anti stereoisomers were positioned with opposite directionality in the minor groove without modified base-pair disruption. They also are in contrast to the 10S and 10R-cis-anti stereoisomers of (BP)G adducts opposite C, where the pyrenyl ring is intercalated into the helix with directionality, but the modified base and its partner on the opposite strand are displaced out of the helix. These results are especially significant given the known greater tumorigenic potential of fjord region compared to bay region polycyclic aromatic hydrocarbons. The tumorigenic potential has been linked to repair efficiency such that bay region adducts can be readily repaired while their fjord region counterparts are refractory to repair. Our structural results propose a link between DNA adduct conformation and repair-dependent mutagenic activity, which could ultimately translate into structure-dependent differences in tumorigenic activities. We propose that the fjord region minor groove-linked BPh-N2-guanine and major groove-linked BPh-N6-adenine adducts are refractory to repair based on our observations that the phenanthrenyl ring intercalates into the helix without modified base-pair disruption. The helix is therefore minimally perturbed and the phenanthrenyl ring is not available for recognition by the repair machinery. By contrast, the bay region BP-N2-G adducts are susceptible to repair, since the repair machinery can recognize either the pyrenyl ring positioned in the minor groove for the trans-anti groove-aligned stereoisomers, or the disrupted modified base-pair for the cis-anti base-displaced intercalated stereoisomers.  相似文献   

20.
(7S,8S)--Dihydroxy--7,8--dihydrobenzo[a]pyrene ((+)-BP-7,8-diol) is epoxidized to (7S,8R)-dihydroxy-(9S,10R)-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene ((+)-syn-BPDE) by cytochrome P-450 isoenzymes and to (7S,8R)-dihydroxy-(9R,10S)-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene ((-)-anti-BPDE) by peroxyl free radicals. 32P postlabeling analysis of the diastereomeric BPDE-DNA adducts was used to investigate the pathways of (+)-BP-7,8-diol oxidation in mouse skin in vivo. The pattern of deoxynucleoside 3',5'-bisphosphate adducts in epidermal scrapings from female CD-1 mice indicated that cytochrome P-450 was the major oxidant. Similar results were obtained when the tumor-promoting phorbol ester tetradecanoylphorbolacetate (TPA) was coadministered with (+)-BP-7,8-diol. However, when animals were pretreated with TPA 24 h before coadministration of TPA and (+)-BP-7,8-diol, the pattern of BPDE-DNA adducts indicated that peroxyl radicals made a major contribution to (+)-BP-7,8-diol epoxidation. Peroxyl radical-dependent epoxidation was maximal when the time between the two TPA administrations was 24-72 h. No increase in (-)-anti-BPDE-DNA was observed when the non-tumor-promoting phorbol ester 4-O-methyl-TPA was substituted for TPA. The calcium ionophore A23187 stimulated peroxyl radical generation when substituted for the first, but not the second, TPA treatment. The antiinflammatory steroid fluocinolone acetonide inhibited (-)-anti-BPDE-DNA adduct formation when coadministered with the first but not the second TPA treatment. These findings demonstrate the existence of two independent pathways of metabolic activation of (+)-BP-7,8-diol in mouse epidermis, one dependent on cytochrome P-450 and the other dependent on peroxyl free radicals. The results also suggest that repetitive topical administration of tumor-promoting phorbol esters remodels epidermal metabolism leading to a significant increase in free radical generation.  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