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.
Under anaerobic conditions and with proper electron donors, NADPH-cytochrome P-450 reductase (EC 1.6.2.4) and xanthine oxidase (EC 1.2.3.2) similarly reductively metabolized mitomycin C. Reversed phase high performance liquid chromatography was used to separate, detect, and isolate several metabolites. Three metabolites were identified by mass spectrometry and thin layer chromatography as 1,2-cis- and trans-2,7-diamino-1-hydroxymitosene and 2,7-diaminomitosene. Three metabolites were phosphate-dependent, and two of them were identified to be 1,2-cis- and trans-2,7-diaminomitosene 1-phosphate. The amounts of the five identified metabolites generated during the reduction of mitomycin C varied with pH and nucleophile concentration. At pH 6.5, 2,7-diaminomitosene was essentially the only metabolite formed, whereas from pH 6.8 to 8.0, trans- and cis-2,7-diamino-1-hydroxymitosene increased in quantity as 2,7-diaminomitosene decreased. The disappearance of mitomycin C and the production of metabolites were enzyme and mitomycin C concentration-dependent. Substrate saturation was not reached for either enzyme up to 5 mM mitomycin C. Electron paramagnetic resonance studies demonstrated the formation of mitomycin C radical anion as an intermediate during enzymatic activation. Our results indicate that either enzyme catalyzed the initial activation of mitomycin C to a radical anion intermediate. Subsequent spontaneous reactions, including the elimination of methanol and the opening of the aziridine ring, generate one active center at C-1 which facilitates nucleophilic attack. Simultaneous generation of two reactive centers was not observed. All five primary metabolites were metabolized further by either flavoenzyme. The secondary metabolites exhibited similar changes in their absorbance spectra and were unlike the primary metabolites, suggesting that a second alkylating center other than C-1 was generated during secondary activation. We propose that secondary activation of monofunctionally bound mitomycin C is probably a main route for the bifunctional binding of mitomycin C to macromolecules and that the cytotoxic actions of mitomycin C result from multiple metabolic activations and reactions.  相似文献   

2.
This article describes a metabolite profiling method for evaluating the effect of oxygen exposure on human liver microsomal metabolism of mitomycin C (MC) in the presence of glutathione (GSH) and NADPH under hypoxic (100% helium), limitedly and fully aerobic, and hyperoxic (100% oxygen) conditions. MC and its metabolite(s) were characterized and the relative percentages of these components were estimated at different incubation times using liquid chromatography and quadrupole time-of-flight mass spectrometry. The MC metabolite profiles were confirmed using purified human cytochrome P450 reductase, acidic activation, and UV-Vis detection at 550 nm. In hypoxia, MC was exclusively metabolized into 2,7-diaminomitosene-10-glutathione-S-conjugate (2,7-DAM-10-SG) within 30 min, whereas approximately 5% of this conjugate, 16% of 2,7-diaminomitosene (2,7-DAM), and 77% of MC were observed under a fully aerobic condition at 90 min. Under limitedly aerobic conditions, the relative percentages of the two metabolites in incubations varied greatly depending on the volume ratio of air to liquid. In hyperoxia, 2% of 2,7-DAM-10-SG, 9% of 2,7-DAM, and 86% of MC were obtained at 90 min. The results indicate that oxygen strongly inhibits the in vitro metabolism of MC. These data suggest that GSH may serve a dual function in facilitating the formation of a leucoaziridinomitosene followed by electron rearrangement giving intermediate metabolite 2,7-DAM, and then trapping this intermediate giving rise to 2,7-DAM-10-SG. These findings provide direct evidence for understanding the fate of oxygen-sensitive metabolic deactivation of MC by GSH.  相似文献   

3.
The anti-cancer drug mitomycin C is metabolically activated to bind and cross-link DNA. The cross-linking contributes significantly to the cytotoxicity. The complex chemical structure of mitomycin C allows its metabolism by several known (cytosolic NAD(P)H:quinone oxidoreductase and microsomal NADPH:cytochrome P450 reductase) and unknown enzymes. The identification of new enzymes/proteins that metabolize mitomycin C and like drugs is an area of significant research interest since these studies have direct implications in drug development and clinical usage. In the present studies, we have investigated a role of cytosolic glucose regulatory protein GRP58 in mitomycin C-induced DNA cross-linking and cytotoxicity. The control and GRP58 siRNA were transfected in human colon carcinoma HCT116 cells in culture. The transfection of GRP58 siRNA but not control siRNA significantly inhibited GRP58 in human colon carcinoma HCT116 cells. The inhibition of GRP58 led to decrease in mitomycin C-induced DNA cross-linking and cytotoxicity. These results establish a role of GRP58 in mitomycin C-induced DNA cross-linking and cytotoxicity. Site-directed mutagenesis of cysteines to serines in thioredoxin domains of GRP58 and cross-linking assays revealed that both N- and C-terminal thioredoxin domains are required for GRP58-mediated mitomycin C-induced DNA cross-linking. These results suggest that GRP58 might be an important target enzyme for further studies on mitomycin C and similar drug therapy.  相似文献   

4.
A new covalent mitomycin C-DNA adduct (4) was isolated from DNA exposed to reductively activated mitomycin C (MC) in vitro. The MC-treated DNA was hydrolyzed enzymatically under certain conditions, and the new adduct was isolated from the hydrolysate by HPLC. Its structure was determined by ultraviolet and circular dichroism spectroscopy and chemical and enzymatic transformations conducted on microscale. In the structure, a single 2" beta, 7"-diaminomitosene residue is linked bifunctionally to two guanines in the dinucleoside phosphate d(GpG). The guanines are linked at their N2 atoms to the C1" and C10" positions of the mitosene, respectively. A key to the structure was a finding that removal of the mitosene from the adduct by hot piperidine yielded d(GpG); another was that the adduct was slowly converted to the known interstrand cross-link adduct 3 by snake venom diesterase and alkaline phosphatase. Adduct 4 represents an intrastrand cross-link in DNA formed by MC. Of the two possible strand-polarity isomers of 4, 4a in which the mitosene 1"-position is linked to the 3'-guanine of d(GpG) is designated as the proper structure, on the basis of the mechanism of the cross-linking reaction. The same adduct 4 was isolated from poly(dG).poly(dC), synthetic oligonucleotides containing the GpG sequence, and Micrococcus luteus and calf thymus DNAs. The relative yields of interstrand and intrastrand cross-links (3 and 4) were determined under first-order kinetic conditions; an average 3.6-fold preference for the formation of 3 over that of 4 was observed. An explanation for this preference is proposed.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
The reductive activation of mitomycin C in aqueous bicarbonate buffer resulted in the formation of a previously unknown compound, characterized as an oxazolidinone derivative of cis-1-hydroxy-2,7-diaminomitosene. This compound is the result of a cyclization reaction of bicarbonate with the aziridine ring of aziridinomitosene, and was observed at bicarbonate concentrations close to those present in physiological plasma.  相似文献   

6.
Malondialdehyde (MDA), an in vivo metabolite of lipid peroxidation and prostaglandin biosynthesis, is mutagenic in Salmonella typhimurium. It is a reactive electrophile that can form interstrand cross-links in DNA. To explore the possibility that MDA-induced interstrand cross-links are the pre-mutagenic lesion, we have quantitated the ability of highly purified preparations of MDA to form interstrand cross-links when reacted with linear plasmid DNA. At physiological temperature and pH, MDA did not form DNA cross-links as determined by DNA denaturation followed by agarose gel electrophoresis. DNA cross-links were formed, however, when incubations with MDA were carried out at either pH 4.2 or temperatures exceeding 60 degrees. alpha-Methylmalondialdehyde (CH3MDA) was found to cross-link DNA more efficiently than MDA, but was not mutagenic in any tester strain of Salmonella. MDA polymers, formed by acid incubation of MDA, also were capable of inducing cross-links. However, an inverse relationship was observed between mutagenicity and extent of polymerization. The pattern of mutagenic response for MDA in different strains of Salmonella was compared with mitomycin C, an established mutagenic cross-linking agent. Error-prone repair and a UvrB+ phenotype, which are needed for the induction of mutations by mitomycin C, were not required for MDA mutagenesis. These findings, taken together, dissociate the mutagenicity of MDA from its ability to form interstrand cross-links with DNA.  相似文献   

7.
D M Peterson  J Fisher 《Biochemistry》1986,25(14):4077-4084
Mitomycin c in the presence of NADPH and brewers' yeast NADPH: (acceptor) oxidoreductase (Old Yellow enzyme, EC 1.6.99.1) is transformed, at pH 8.0 and with anaerobicity, to two major mitosene products (the cis- and trans-1-hydroxy-2,7-diaminomitosenes; respective yields of 45 and 30%). These arise by covalent trapping by solvent of a quinone methide intermediate, obtained by rearrangement of the mitomycin c hydroquinone. At lower pH (6.5), the major product of this reaction is 2,7-diaminomitosene, which arises by covalent trapping of the quinone methide by H+. In the former instance the quinone methide acts as an electrophile and in the latter as a nucleophile. A detailed kinetic analysis indicates that the role of the NADPH and Old Yellow enzyme is to initiate an autocatalytic reaction, propagated by mitomycin c reduction by the mitosene hydroquinones (arising by the electrophilic pathway). The evidence supporting this conclusion is the formation of oxidized mitosene products, under the rigorously anaerobic reaction circumstance, the nonstoichiometric participation of NADPH, a dependence of the velocity on the total mitomycin c concentration, the pH dependence of the reaction, and the accurate simulation of the overall kinetic course with a mathematical model of the autocatalytic pathway. These observations indicate that the autocatalytic pathway may be dominant during in vitro mitomycin c anaerobic reductive activation by other reducing agents and that (as with anthracycline reductive activation) oxidation of the mitosene hydroquinone obtained from nucleophile addition to the quinone methide may be a necessary event for the formation of stable covalent adducts in vivo.  相似文献   

8.
DT-diaphorase (DTD) mediated reduction of a series of 2,5-bis-substituted-3,6-diaziridinyl-1,4-benzoquinones was found to increase the level of DNA interstrand cross-linking (ISC) formed at neutral pH with an enhancement observed as the pH was decreased to 5.8. The analogues used were symmetrically alkyl-substituted carbamoyl ester analogues of AZQ (D1-D7), 3,6-diaziridinyl-1,4-benzoquinone (DZQ), the 2,5-dimethyl derivative (MeDZQ), and a 2,5-bis[(2-hydroxyethyl)amino] analogue (BZQ). At pH 5.8, the level of DNA ISC induced by enzymatic reduction was as follows: DZQ greater than MeDZQ much greater than D1 (methyl) greater than D3 (n-propyl) greater than D2 (AZQ; ethyl) greater than D5 (n-butyl) greater than D7 (sec-butyl) greater than D4 (isopropyl) D6 greater than (isobutyl). A similar trend was observed at pH 7.2. The level of DNA ISC induced by BZQ, which is not a substrate for DTD, was not increased by enzymatic reduction. Dicumarol, a known inhibitor of DTD, was capable of inhibiting the DNA ISC induced by these quinones upon enzymatic reduction. MeDZQ and DZQ reacted with guanines, as measured by Maxam and Gilbert sequencing, with a sequence selectivity similar to that of the nitrogen mustard class of antitumor agents. Enzymatic reduction of DZQ and MeDZQ by DTD was found to alter their sequence-selective alkylation. Reduced DZQ showed enhanced guanine alkylation in 5'-GC-3' sequences and new sites of adenine alkylation in 5'-(A/T)AA-3' sequences. Reduced MeDZQ only showed new sites of adenine alkylation at 5'-(A/T)AA-3' sequences but no enhancement of guanine alkylation. The new sites of adenine alkylation were found to be inhibited in the presence of magnesium and rapidly converted into apurinic sites.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
Nonerythroid α-spectrin (αIISp) is a structural protein involved in repair of DNA interstrand cross-links and is deficient in cells from patients with Fanconi anemia (FA), which are defective in ability to repair cross-links. In order to further demonstrate the importance of the role that αIISp plays in normal human cells and in the repair defect in FA, αIISp was knocked down in normal cells using siRNA. Depletion of αIISp in normal cells by siRNA resulted in chromosomal instability and cellular hypersensitivity to DNA interstrand cross-linking agents. An increased number of chromosomal aberrations were observed and, following treatment with a DNA interstrand cross-linking agent, mitomycin C, cells showed decreased cell growth and survival and decreased formation of damage-induced αIISp and XPF nuclear foci. Thus depletion of αIISp in normal cells leads to a number of defects observed in FA cells, such as chromosome instability and a deficiency in cross-link repair.  相似文献   

10.
DT-diaphorase (DTD) activity has been related to bioactivation and cytotoxicity of antitumor quinones. A pair of human colon adenocarcinoma cell lines, HT29 and BE, were used in this study to examine the role of DTD in antitumor quinone induced apoptosis. HT29 cells have elevated levels of DTD whereas BE cells lack functional DTD due to a point mutation which results in a complete lack of DTD activity. MeDZQ, a quinone that is efficiently bioactivated by DTD, induced apoptosis both in HT29 and BE cells, but with a much higher incidence in HT29, as assessed by morphological criteria and the formation of oligonucleosomal fragments of DNA. Two other quinone compounds which are also substrates for DTD, i.e. streptonigrin and mitomycin C, also preferentially induced apoptosis in HT29 cells, which could be inhibited by dicoumarol. Our data suggest that bioreductive activation of antitumor quinones by DTD results in induction of apoptosis in human colon carcinoma cells.  相似文献   

11.
The processing of damaged DNA was altered in a mitomycin C-sensitive mutant (mtcA) of Micrococcus radiodurans. Even though the mutant retained resistance to 254-nm UV radiation, it did not, in contrast to the wild-type strain, show any excessive DNA degradation or cell death when incubated with chloramphenicol after sublethal doses of either UV light or mitomycin C. The results suggest the constitutive synthesis of an enzyme system responsible for wild-type proficiency in the repair of mitomycin C-induced damage. An alternative system able to repair damage caused by mitomycin C was demonstrated in the mtcA background. In this strain, additional damage inflicted upon the cellular DNA effected a massive rescue of cells previously inactivated by mitomycin C. Rescue was provoked by ionizing radiation, by UV light, or by simple alkylating agents. Cells treated with psoralen plus near-UV radiation could be rescued only when inactivation was due primarily to psoralen-DNA interstrand cross-links rather than to monoadducts. The rescue of inactivated cells was prevented in the presence of chloramphenicol. These results can be interpreted most readily in terms of an alternative repair system able to overcome DNA interstrand cross-links produced by mitomycin C or psoralen plus near-UV light, but induced only by the more abundant number of damages produced by radiation or simple alkylating agents.  相似文献   

12.
Differential activation of p53 by the various adducts of mitomycin C   总被引:1,自引:0,他引:1  
Mitomycin C (MC) is a cytotoxic chemotherapeutic agent that causes DNA damage in the form of DNA cross-links as well as a variety of DNA monoadducts and is known to induce p53. The various DNA adducts formed upon treatment of mouse mammary tumor cells with MC as well as 10-decarbamoyl MC (DMC) and 2,7-diaminomitosene (2,7-DAM), the major MC metabolite, have been elucidated. The cytotoxicity of DMC parallels closely that of MC in a number of rodent cell lines tested, whereas 2,7-DAM is relatively noncytotoxic. In this study, we investigate the ability of MC, DMC, and 2,7-DAM to activate p53 at equidose concentrations by treating tissue culture cell lines with the three mitomycins. Whereas MC and DMC induced p53 protein levels and increased the levels of p21 and Gadd45 mRNA, 2,7-DAM did not. Furthermore, MC and DMC, but not 2,7-DAM, were able to induce apoptosis efficiently in ML-1 cells. Therefore the 2,7-DAM monoadducts were unable to activate the p53 pathway. Interestingly, DMC was able to initiate apoptosis via a p53-independent pathway whereas MC was not. This is the first finding that adducts of a multiadduct type DNA-damaging agent are differentially recognized by DNA damage sensor pathways.  相似文献   

13.
It has been known for several years that Adriamycin forms adducts and interstrand cross-links when reacted for long periods of time with bacterial and mammalian DNA in vitro, with the cross-link being restricted to 2 bp elements containing GpC sequences. The self-complementary 20mer deoxyoligonucleotide TA4T4GCA4T4A has been used in this study as a model of the apparent G-G cross-linking site at GpC sequences. The rate of formation of cross-links, as well as the dependence on both Adriamycin and Fe(III) concentration, were similar with this oligonucleotide as compared with calf thymus DNA. The cross-linking was demonstrated on both denaturing and non-denaturing sequencing gels. The half-life of the G-G cross-link was 40 h, consistent with that implied with high molecular weight, heterogeneous sequence DNA. Exonuclease III digests of adducts formed with 20mer deoxyoligonucleotides containing single, central G-G, G-I and I-I potential cross-links revealed that a guanine residue is required at both ends of the cross-link. No cross-linking was observed with a similar oligonucleotide containing only a single central (G.C) bp.  相似文献   

14.
The cytotoxic action of the antitumor antibiotic mitomycin C occurs primarily at the level of DNA. Using highly sensitive fluorescence assays which depend on the enhancement of ethidium fluorescence only when it intercalates duplex regions of DNA, three aspects of mitomycin C action on DNA have been studied: (a) cross-linking events, (b) alkylation without necessarily cross-linking, and (c) strand breakage. Cross-linking of DNA is determined by the return of fluorescence after a heat denaturation step at alkaline pH's. Under these conditions denatured DNA gives no fluorescence. The cross-linking was independently confirmed by S1-endonuclease (EC 3.1.4.-) digestion. At relatively high concentrations of mitomycin the suppression of ethidium fluorescence enhancement was shown not to be due to depurination but rather to alkylation, as a result of losses in potential intercalation sites. A linear relationship exists between binding ratio for mitomycin and loss of fluorescence. The proportional decrease in fluorescence with pH strongly suggests that the alkylation is due to the aziridine moiety of the antibiotic under these conditions. A parallel increase in the rate and overall efficiency of covalent cross-linking of DNA with lower pH suggests that the cross-linking event, to which the primary cytotoxic action has been linked, occurs sequentially with alkylation by aziridine and then by carbamate. Mitomycin C, reduced chemically, was shown to induce single strand cleavage as well as monoaklylation and covalent cross-linking in PM2 covalently closed circular DNA. The inhibition of this cleavage by superoxide dismutase (EC 1.15.1.1) and catalase (EC 1.11.1.6), and by free radical scavengers suggests that the degradation of DNA observed to accompany the cytotoxic action of mitomycin C is largely due to the free radical O2. In contrast to the behavior of the antibiotic streptonigrin, mitomycin C does not inactivate the protective enzymes superoxide dismutase or catalase. Lastly, mitomycin C is able to cross-link DNA in the absence of reduction at pH 4. This is consistent with the postulated cross-linking mechansims.  相似文献   

15.
Abstract

A dual prodrug conjugate between the antimetabolite cytarabine monophosphate and the alkylating agent 2,7-diaminomitosene (derived from mitomycin C), cytaramycin, was synthesized and tested for antileukemic activity in sensitive and resistant tumors. The compound was active against parental L-1210, CCRF-CEM, HL-60 and K-562 leukemia cells but did not overcome resistance in sublines developed for (1) multidrug resistance (L-1210/MDR and K-562-R) or (2) for cytarabine resistance (CCRF-CEM/ARA-C and HL-60/ARA-C). Alkaline DNA elution tests demonstrate a predominance of strand breaking activity due to the cytarabine moiety, and a lesser degree of DNA crosslinking, due to the mitosene moiety. The conjugate was active in mice bearing P-388 leukemia (80% increased lifespan), but was not more effective than mitomycin C alone in mice bearing a cytarabine-resistant L-1210 cell line (38% to 31% increased lifespan). These findings suggest that mitomycin nucleotide conjugates do not overcome resistance to the parent antimetabolites.  相似文献   

16.
Activation of Adriamycin by formaldehyde leads to the formation of drug–DNA adducts in vitro and these adducts stabilise the DNA to such a degree that they function as virtual interstrand cross-links. The formation of these virtual interstrand cross-links by Adriamycin was investigated in MCF-7 cells using a gene-specific interstrand cross-linking assay. Cross-linking was measured in both the nuclear-encoded DHFR gene and in mitochondrial DNA (mtDNA). Cross-link formation increased linearly with Adriamycin concentration following a 4 h exposure to the drug. The rate of formation of Adriamycin cross-links in each of the genomes was similar, reaching maximal levels of 0.55 and 0.4 cross-links/10 kb in the DHFR gene and mtDNA respectively, following exposure to 20 µM Adriamycin for 8 h. The interstrand cross-link was short lived in both DNA compartments, with a half-life of 4.5 and 3.3 h in the DHFR gene and mtDNA respectively. The kinetics of total Adriamycin adduct formation, detected using [14C]Adriamycin, was similar to that of cross-link formation. Maximal adduct levels (30 lesions/10 kb) were observed following incubation at 20 µM drug for 8 h. The formation of such high levels of adducts and cross-links could therefore be expected to contribute to the mechanism of action of Adriamycin.  相似文献   

17.
We previously reported that some Deinococcus radiodurans mutants are sensitive to DNA interstrand cross-linking agents but resistant to UV and gamma-rays. We isolated DNA fragments from a D. radiodurans genomic library which complemented the mitomycin C sensitivity of one of these mutants. One 3.2kb-long fragment contains an open reading frame of approximately 700bp and the deduced amino acid sequence is very homologous to other prokaryotic RecR proteins. This open reading frame in the mitomycin C-sensitive mutant strain contains a frame shift mutation at its carboxyl terminal region. These data suggest that RecR protein plays an important role in the resistance to interstrand cross-links in this bacterium.  相似文献   

18.
The nucleotide sequence preferences of the DNA interstrand cross-linking agents dehydroretronecine diacetate (DHRA), 2,3-bis(acetoxymethyl)-1-methylpyrrole (BAMP), dehydromonocrotaline, and dehydroretrorsine were studied by using synthetic DNA duplex fragments and polyacrylamide gel electrophoresis (PAGE). These agents have structural features in common with the reductively activated aziridinomitosene of mitomycin C (MC). Like MC, they preferentially cross-linked DNA duplexes containing the duplex sequence 5'-CG. For DHRA and BAMP interstrand cross-linked DNA duplexes, PAGE analysis of iron(II)-EDTA fragmentation reactions revealed the interstrand cross-links to be deoxyguanosine to deoxyguanosine (dG-to-dG), again analogous to DNA cross-links caused by MC. Unlike MC, DHRA could be shown to dG-to-dG cross-link a 5'-GC sequence. Furthermore, the impact of flanking sequence on the efficiency of interstrand cross-linking at 5'-CG was reduced for BAMP, with 5'-TCGA and 5'-GCGC being equally efficiently cross-linked. Possible origins of the 5'-CG sequence recognition common to all of the agents are discussed. A model is presented in which the transition state for the conversion of monoadducts to cross-links more closely resembles ground-state DNA at 5'-CG sequences.  相似文献   

19.
The presence of high levels, as well as tissue-specific forms, of cytochrome P450 enzymes in mammalian olfactory mucosa (OM) has important implications in the bioactivation and toxicity of xenobiotics entering the tissue. Previous studies have shown that coumarin, a known olfactory toxicant in rats, is bioactivated by OM microsomal P450s to a number of products, presumably via coumarin-3,4-epoxide and other epoxide intermediates. The aim of the current study was to obtain direct evidence for the formation of such reactive intermediates in rat OM through the detection of protein covalent binding and glutathione (GSH) adduct formation. Protein covalent binding experiments with [14C]coumarin (10 μM) displayed a 7–9-fold higher NADPH-dependent radioactivity binding in rat OM microsomes (2.5 nmol/mg/30 min) compared to those in rat and human liver microsomes; the binding value in rat OM microsomes was substantially but not completely reduced by the addition of GSH (5 mM). LC/MS analyses detected a number of GSH adducts in GSH-supplemented coumarin metabolism reaction in rat OM microsomes; 3-glutathionyl coumarin was found to be the major one, indicating 3,4-epoxidation as the main bioactivation pathway. Additional GSH adducts were identified, presumably forming via the same pathway or epoxidation on the benzene moiety. Our findings provide direct evidence for the formation of multiple coumarin reactive intermediates in rat OM, leading to protein covalent binding and GSH conjugation.  相似文献   

20.
The major DNA adduct (greater than 95% total) resulting from the bioactivation of ethylene dibromide by conjugation with GSH is S-(2-(N7-guanyl)ethyl)GSH. The mutagenic potential of this adduct has been uncertain, however, because the observed mutagenicity might be caused by other adducts present at much lower levels, e.g. S-(2-N1-adenyl)ethyl)GSH. To assess the formation of other potential adducts, S-(2-(N3-deoxycytidyl)ethyl)GSH, S-(2-(O6-deoxyguanosyl)ethyl)GSH, and S-(2-(N2-deoxyguanosyl)ethyl)GSH were prepared and used as standards in the analysis of calf thymus DNA modified by treatment with [1,2-14C]ethylene dibromide and GSH in the presence of rat liver cytosol; only minor amounts (less than 0.2%) were found. A forward mutation assay in (repair-deficient) Salmonella typhimurium TA100 and sequence analysis were utilized to determine the type, site, and frequency of mutations in a portion of the lacZ gene resulting from in vitro modification of bacteriophage M13mp18 DNA with S-(2-chloroethyl)GSH, an analog of the ethylene dibromide-GSH conjugate. An adduct level of approximately 8 nmol (mg DNA)-1 resulted in a 10-fold increase in mutation frequency relative to the spontaneous level. The spectrum of spontaneous mutations was quite varied, but the spectrum of S-(2-chloroethyl)GSH-induced mutations consisted primarily of base substitutions of which G:C to A:T transitions accounted for 75% (70% of the total mutations). All available evidence implicates S-(2-(N7-guanyl)ethyl)GSH as the cause of these mutations inasmuch as the levels of the minor adducts are not consistent with the mutation frequency observed in this system. The sequence selectivity of alkylation was determined by treatment of end-labeled lac DNA fragments with S-(2-chloroethyl)GSH, cleavage of the DNA at adduct sites, and electrophoretic analysis. Comparison of the sequence selectivity with the mutation spectrum revealed no obligate relationship between the extent of adduct formation and the number of mutations which resulted at different sites. We suggest that the mechanism of mutagenesis involves DNA sequence-dependent alterations in the interaction of the polymerase with the (modified) template and incoming nucleotide.  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