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.
Chloroethylnitrosoureas (CENUs), which are bifunctional alkylating agents widely used in the clinical treatment of cancer, exert anticancer activity by inducing crosslink within a guanine-cytosine DNA base pair. However, the formation of dG-dC crosslinks can be prevented by O6-alkylguanine-DNA alkyltransferase (AGT), ultimately leading to drug resistance. Therefore, the level of AGT expression is related to the formation of dG-dC crosslinks and the sensitivity of cells to CENUs. In this work, we determined the CENU-induced dG-dC crosslink in mouse L1210 leukemia cells and in human glioblastoma cells (SF-763, SF-767 and SF-126) containing different levels of AGT using high-performance liquid chromatography coupled with electrospray ionization tandem mass spectrometry. The results indicate that nimustine (ACNU) induced more dG-dC crosslinks in L1210 leukemia cells than those induced by carmustine (BCNU), lomustine (CCNU) and fotemustine (FTMS). This result was consistent with a previously reported cohort study, which demonstrated that ACNU had a better survival gain than BCNU, CCNU and FTMS for patients with high-grade glioma. Moreover, we compared the crosslinking levels and the cytotoxicity in SF-763, SF-767 and SF-126 cells with different AGT expression levels after exposure to ACNU. The levels of dG-dC crosslink in SF-126 cells (low AGT expression) were significantly higher than those in SF-767 (medium AGT expression) and SF-763 (high AGT expression) cells at each time point. Correspondingly, the cytotoxicity of SF-126 was the highest followed by SF-767 and SF-763. The results obtained in this work provided unequivocal evidence for drug resistance to CENUs induced by AGT-mediated repair of DNA ICLs. We postulate that the level of dG-dC crosslink has the potential to be employed as a biomarker for estimating drug resistance and anticancer efficiencies of novel CENU chemotherapies.  相似文献   

2.
3.
《Life sciences》1996,58(19):PL303-PL308
O6-Alkylguanine derivatives sensitize tumor cells to chloroethylnitrosourea (CENU) chemotherapy by inactivation of O6-methylguanine-DNA methyltransferase (MGMT), which repairs CENU-induced O6-alkylguanines in DNA by accepting the alkyl group at a cysteine moiety. To test the biological significance of synthesized O6-fluorobenzylguanine derivatives, we measured their ability of inactivation of MGMT activity and their effects on the cytotoxicity of 1-(4-amino-2-methyl-5-pyrimidinyl)methyl-3-(2-chloroethyl)-3-nitrosourea hydrochloride (ACNU) in comparison with the effects of O6-benzylguanine and O6-phenylguanine. The O6-(4-and 3-fluorobenzyl)guanines considerably reduced the MGMT activity of HeLa S3 cell-free extract as did O6-benzylguanine. In contrast, O6-(2-fluorobenzyl)guanine and O6-phenylguanine had less of an effect on the activity. Two-hour pretreatment of O6-(4-and 3-fluorobenzyl)guanines potentiated ACNU cytotoxicity in HeLa S3 cells to a greater extent than did O6-(2-fluorobenzyl)guanine and O6-phenylguanine. The enhancement effects were consistent with the depletion of MGMT activity after the pretreatment of O6-fluorobenzylguanine derivatives. O6-Fluorobenzylguanines with a fluoro-substitution at the 4- or 3-position of the benzyl group were comparable to O6-benzylguanine and were powerful MGMT inactivators. The chemical features of the O6-benzyl group are a biologically important determinant in the reaction evolution with MGMT.  相似文献   

4.
Two new agents based upon the structure of the clinically active prodrug laromustine were synthesized. These agents, 2-(2-chloroethyl)-N-methyl-1,2-bis(methylsulfonyl)-N-nitrosohydrazinecarboxamide (1) and N-(2-chloroethyl)-2-methyl-1,2-bis(methylsulfonyl)-N-nitrosohydrazinecarboxamide (2), were designed to retain the potent chloroethylating and DNA cross-linking functions of laromustine, and gain the ability to methylate DNA at the O-6 position of guanine, while lacking the carbamoylating activity of laromustine. The methylating arm was introduced with the intent of depleting the DNA repair protein O6-alkylguanine-DNA alkyltransferase (AGT). Compound 1 is markedly more cytotoxic than laromustine in both AGT minus EMT6 mouse mammary carcinoma cells and high AGT expressing DU145 human prostate carcinoma cells. DNA cross-linking studies indicated that its cross-linking efficiency is nearly identical to its predicted active decomposition product, 1,2-bis(methylsulfonyl)-1-(2-chloroethyl)hydrazine (90CE), which is also produced by laromustine. AGT ablation studies in DU145 cells demonstrated that 1 can efficiently deplete AGT. Studies assaying methanol and 2-chloroethanol production as a consequence of the methylation and chloroethylation of water by 1 and 2 confirmed their ability to function as methylating and chloroethylating agents and provided insights into the superior activity of 1.  相似文献   

5.
The DNA repair protein O6-alkylguanine alkyltransferase (AGT) is responsible for removing promutagenic alkyl lesions from exocyclic oxygens located in the major groove of DNA, i.e. the O6 and O4 positions of guanine and thymine. The protein carries out this repair reaction by transferring the alkyl group to an active site cysteine and in doing so directly repairs the premutagenic lesion in a reaction that inactivates the protein. In order to trap a covalent AGT–DNA complex, oligodeoxyribonucleotides containing the novel nucleoside N1,O6-ethanoxanthosine (eX) have been prepared. The eX nucleoside was prepared by deamination of 3′,5′-protected O6-hydroxyethyl-2′-deoxyguanosine followed by cyclization to produce 3′,5′-protected N1,O6-ethano-2′-deoxyxanthosine, which was converted to the nucleoside phosphoramidite and used in the preparation of oligodeoxyribonucleotides. Incubation of human AGT with a DNA duplex containing eX resulted in the formation of a covalent protein–DNA complex. Formation of this complex was dependent on both active human AGT and eX and could be prevented by chemical inactivation of the AGT with O6-benzylguanine. The crosslinking of AGT to DNA using eX occurs with high yield and the resulting complex appears to be well suited for further biochemical and biophysical characterization.  相似文献   

6.
M F Hacques  C Marion 《Biopolymers》1986,25(12):2281-2293
CD and uv spectroscopy reveal that the synthetic polynucleotides poly(dG–dC) · poly(dG–dC) and poly(dG–m5dC) · poly(dG–m5dC) undergo a transition induced by small amounts of Ni++ cation from a right-handed B-form to left-handed Z-type conformations. We describe the application of steady-state and transient electric birefringence to the characterization of the transition observed at very low ionic strength (10 mM Tris HCl, pH 7.4). Dialysis experiments show that the changes in spectroscopic and electro-optic properties upon addition of Ni++ are completely reversible. The differences in shape of the inverted CD spectra suggest the existence of a family of left-handed conformations, depending on the polymer used and on the amounts of cation added. The stoichiometry required for inducing the Z-conformation of poly(dG–m5dC) is 1 cation/4 nucleotide phosphates. The transition is accompanied by a decrease in birefringence, an increase in length, and the more important contribution of a permanent or slowly induced dipole moment to the orientation mechanism. High concentrations of Ni++ promote the Z → Z* transition. Upon increasing the Ni++ concentration, poly(dG–dC) undergoes a biphasic transition, first to one intermediate conformation that is neither B- nor Z-like and then to a left-handed form that is probably different from Z*. These conversions are accompanied by regular decreases both in birefringence and in chain length, but no transient appears in the field-reversal experiments.  相似文献   

7.
1,3-Bis(2-chloroethyl)-1-nitrosourea (BCNU) and related 2-haloethylnitrosoureas covalently cross-link DNA under physiological conditions. The rate of the cross-linking increases with increasing pH in the range 4–10 and with the (G + C) content of natural DNAs. The reaction leads to stable interstrand cross-links by a two-step process and is strongly dependent on the 2-halogen in the nitrosourea where Cl ? Br > F ? I. Only one 2-haloethyl group is necessary for cross-linking, which is not observed when the halogen is replaced by -OH or -OCH3. Promoting the acidity of the N3H group by appropriate aryl substitution increases the rate of cross-linking. The position of the halogen is critical since, while 1-(2-chloroethyl)-1-nitrosourea cross-links DNA efficiently, 1-(3-chloropropyl)-1-nitrosourea shows no reactivity. N4-(2-Chloroethyl)-1-methylcytosine hydrochloride, very similar to a suggested intermediate in the cross-linking process, alkylates PM2-CCC-DNA readily. The modes of aqueous decomposition of nitrosoureas as they apply to alkylation and cross-linking are discussed. The results are in accord with formation of a haloethonium ion which forms a nitrogen half mustard intermediate with a DNA base then completes the cross-link. A correlation exists between the extent of DNA cross-linking and activity of the nitrosoureas against L1210 leukemia. Based on the results of this work, a new nitrosourea is designed and synthesized which shows more efficient cross-linking.  相似文献   

8.
O6-Methyl[8-3H]deoxyguanosine in a synthetic DNA polymer, poly(dC, dG, m6dG), is demethylated by cell-free extracts of EscherichiacoliBr adapted by exposure to N-methyl-N′-nitro-N-nitrosoguanidine, as shown by the appearance of 3H-labeled deoxyguanosine in hydrolysates of the recovered DNA. The demethylating activity could not be detected in extracts of nonadapted E. coli. These results provide direct evidence that a previously described inducible repair activity in E. coli acts by demethylating O6-methylguanine at the DNA level.  相似文献   

9.
C Colombier  B Lippert    M Leng 《Nucleic acids research》1996,24(22):4519-4524
Our aim was to determine whether a single transplatin monofunctional adduct, either trans-[Pt(NH3)2(dC)Cl]+ or trans-[Pt(NH3)2(dG)Cl]+ within a homopyrimidine oligonucleotide, could further react and form an interstrand cross-link once the platinated oligonucleotide was bound to the complementary duplex. The single monofunctional adduct was located at either the 5' end or in the middle of the platinated oligonucleotide. In all the triplexes, specific interstrand cross-links were formed between the platinated Hoogsteen strand and the complementary purine-rich strand. No interstrand cross-links were detected between the platinated oligonucleotides and non-complementary DNA. The yield and the rate of the cross-linking reaction depend upon the nature and location of the monofunctional adducts. Half-lives of the monofunctional adducts within the triplexes were in the range 2-6 h. The potential use of the platinated oligonucleotides to modulate gene expression is discussed.  相似文献   

10.
《Inorganica chimica acta》1986,121(2):167-174
The reaction of 2,3-tri with CrCl3·6H2O1, dehydrated in boiling DMF, results in the formation of mer-CrCl3(2,3-tri) and anation of hydrolysed solutions of mer-MCl3(2,3,-tri) (M=Co, Cr) with 6 M HCl containing HClO4, forms trans-dichloro- mer-[MCl2(2,3-tri)(OH2)]ClO4·H2O (M=Cr, Co; I, II). trans-Dinitro-mer-[Co(NO2)2(NH3)(2,3-tri)] ClO4 crystallises from the reaction between mer-Co(NO2)3(2,3-tri) and aqueous 7 M ammonia, on addition of NaClO4·H2O, and trans-dichloro-mer-[CoCl2(NH3)(2,3-tri)]ClO4 (III) can be isolated by treatment of the dinitro with 12 M HCl. Reaction of mer-CoCl3(2,3-tri) with C2O42, followed by addition of aqueous NH3 and NaClO4·H2O results in the isolation of racemic mer-[Co(ox)(NH3)(2,3-tri)]ClO4· H2O. This complex was resolved into its enantiomeric forms and treatment of these with SOCl2/MeOH/ HClO4 gave the chiral forms of trans-dichloro-mer- [CoCl2(NH3)(2,3-tri)]ClO4 (R or S at the see-NH center). The rates of loss of the first chloro ligand from these dichloro complexes have been measured spectrophotometrically in 0.1 M HNO3 over a 15 K temperature range to give the following kinetic parameters; (I) kH(298)=7.25 × 10−5 s−1, Ea=78.5 kJ mol−1, δS298#=69 J K−1 mol−1; (II) kH(298)=4.00 × 10−3 s−1, Ea=89.9, δS298#= +87.5; (III) kH(298)=3.09 × 10−4 s−1, Ea=103, δS298#=+27. Treatment of the dichloro cations with Hg2+/HNO3 results in the generation of mer- M(2,3-tri)(OH2)33+ (M=Cr, Co; IV, V) and trans- diaqua-mer-Co(NH3)(2,3-tri)(OH2)23+ (VI). The Co(III) cations isomerise to the fac configuration with (V) Kisom(298) μ=1.0 M)=2.97 × 10−5 s−1, Ea=115, δS298#=+46. (VI) Kisom(298) (μ=1.0 M)=4.13 × 10−5 s−1, Ea=113, δS298#=+52.  相似文献   

11.
《Inorganica chimica acta》2006,359(9):2896-2909
[RuCl3(NO)(P–P)], [P–P = R2P(CH2)nPR2 (n = 1–3) and R2P(CH2)POR2, PR2–CHCH–PR2, R = Ph and (C6H11)2P-(CH2)2-P(C6H11)2] were obtained and characterized by 31P {1H} NMR, IR spectroscopies and cyclic voltammetry. The structures of fac-[RuCl3(NO)(P–P)], P–P = dppm (1), dppe (2), c-dppen (3) and dppp (4), mer-[RuCl3(NO)(dcpe)] (6a) and mer-[RuCl3(NO)(dppmO)] (7) have been determined by X-ray diffraction. Photochemical isomerization of fac- to mer-[RuCl3(NO)(P–P)] was observed under white light in a CH2Cl2 solution and in solid state. The isomerization processes were followed by IR and 31P {1H} spectra. The mer-[RuCl3(15NO)(dppb)] isomer was used for the definition of the phosphorus atoms in the structure of the complex in solution. The electrochemical study shows that the oxidation/reduction processes observed in these complexes are dependent on both the isomer (fac or mer) and the solvent. In CH2Cl2, the NO+ reduction potentials are less negative for the mer-isomers than for the fac ones, while in CH3CN solvent these potentials are, in general, very close for both isomers.  相似文献   

12.
Chloroethylnitrosureas (CNUs) are powerful DNA-reactive alkylating agents used in cancer therapy. Here, we analyzed cyto- and genotoxicity of nimustine (ACNU), a representative of CNUs, in synchronized cells and in cells deficient in repair proteins involved in homologous recombination (HR) or nonhomologous end-joining (NHEJ). We show that HR mutants are extremely sensitive to ACNU, as measured by colony formation, induction of apoptosis and chromosomal aberrations. The NHEJ mutants differed in their sensitivity, with Ku80 mutants being moderately sensitive and DNA-PKcs mutated cells being resistant. HR mutated cells displayed a sustained high level of γH2AX foci and displayed co-staining with Rad51 and 53BP1, indicating DNA double-strand breaks (DSB) to be formed. Using synchronized cells, we analyzed whether DSB formation after ACNU treatment was replication-dependent. We show that γH2AX foci were not induced in G1 but increased significantly in S phase and remained at a high level in G2, where a fraction of cells became arrested and underwent, with a delay of > 12 h, cell death by apoptosis and necrosis. Rad51, ATM, MDC-1 and RPA-2 foci were also formed and shown to co-localize with γH2AX foci induced in S phase, indicating that the DNA damage response was activated. All effects observed were abrogated by MGMT, which repairs O6-chloroethylguanine that is converted into DNA cross-links. We deduce that the major genotoxic and killing lesion induced by CNUs are O6-chloroethylguanine-triggered cross-links, which give rise to DSBs in the treatment cell cycle, and that HR, but not NHEJ, is the major route of protection against this group of anticancer drugs. Base excision repair had no significant impact on ACNU-induced cytotoxicity.  相似文献   

13.
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   

14.
A circular dichroism study of poly dG, poly dC, and poly dG:dC   总被引:22,自引:0,他引:22  
D M Gray 《Biopolymers》1974,13(10):2087-2102
We have measured the ultraviolet circular dichroism spectra of oligo d(pG)5, poly dN AcG, poly dI, poly dC, two samples of poly dG, and four samples containing double-stranded poly dG:dC. We find that oligo d(pG)5 and poly dG exist in self-complexed forms as well as in single-stranded forms. Unlike the self-complexed form of poly dG, the single-stranded form of poly dG can hydrogen-bond with single-stranded poly dC. We present spectral data for double-stranded poly dG:dC, which can be used to help characterize poly dG:dC preparations and which provide a basis for resolving discrepancies among other reported poly dG:dC spectra.  相似文献   

15.
The binding site and the geometry of Co(III)meso-tetrakis(N-methylpyridinium-4-yl)porphyrin (CoTMPyP) complexed with double helical poly(dA)·poly(dT) and poly(dG)·poly(dC), and with triple helical poly(dA)·[poly(dT)]2 and poly(dC)·poly(dG)·poly(dC)+ were investigated by circular and linear dichroism (CD and LD). The appearance of monomeric positive CD at a low [porphyrin]/[DNA] ratio and bisignate CD at a high ratio of the CoTMPyP-poly(dA)·poly(dT) complex is almost identical with its triplex counterpart. Similarity in the CD spectra was also observed for the CoTMPyP-poly(dG)·poly(dC) and -poly(dC)·poly(dG)·poly(dC)+ complex. This observation indicates that both monomeric binding and stacking of CoTMPyP to these polynucleotides occur at the minor groove. However, different binding geometry of CoTMPyP, when bind to AT- and GC-rich polynucleotide, was observed by LD spectrum. The difference in the binding geometry may be attributed to the difference in the interaction between polynucleotides and CoTMPyP: in the GC polynucleotide case, amine group protrude into the minor groove while it is not present in the AT polynucleotide.  相似文献   

16.
Abstract

Sequence-dependency of cellular uptake of oligonucleotides into Vero cells has been studied. Cellular uptake of 5′-[35S]-labelled homopolymers decreased in the order (dG)16 >> (dT)16> (dA)16 > (dC)16. The change of two base-pairs (dG → dA) in a dG-rich antisense oligonucleotide with good antiviral activity dramatically decreased cellular uptake and abolished antiviral activity.  相似文献   

17.
The impuritiy profiles of acetonitrile solutions of the four standard O‐cyanoethyl‐N,N‐diisopropyl‐phosphoramidites of 5′‐O‐dimethoxytrityl (DMT) protected deoxyribonucleosides (dGib, dAbz, dCbz, T) were analyzed by HPLC‐MS. The solution stability of the phosphoramidites decreases in the order T, dC>dA>dG. After five weeks storage under inert gas atmosphere the amidite purity was reduced by 2% (T, dC), 6% (dA), and 39% (dG), respectively. The main degradation pathways involve hydrolysis, elimination of acrylonitrile and autocatalytic acrylonitrile‐induced formation of cyanoethyl phosphonoamidates. Consequently, the rate of degradation is reduced by reducing the water concentration in solution with molecular sieves and by lowering the amidite concentration. Acid‐catalyzed hydrolysis could also be reduced by addition of small amounts of base.  相似文献   

18.
The interaction of the Cu(II) drugs CuL(NO3) and CuL′(NO3) (HL is pyridine-2-carbaldehyde thiosemicarbazone and HL′ is pyridine-2-carbaldehyde 4N-methylthiosemicarbazone, in water named [CuL]+ and [CuL′]+) with [poly(dA–dT)]2, [poly(dG–dC)]2, and calf thymus (CT) DNA has been probed in aqueous solution at pH 6.0, I = 0.1 M, and T = 25 °C by absorbance, fluorescence, circular dichroism, and viscosity measurements. The results reveal that these drugs act as groove binders with [poly(dA–dT)]2, with a site size n = 6–7, whereas they act as external binders with [poly(dG–dC)]2 and/or CT-DNA, thus establishing overall electrostatic interaction with n = 1. The binding constants with [CuL′]+ were slightly larger than with [CuL]+. The title compounds display some cleavage activity in the presence of thiols, bringing about the rupture of the DNA strands by the reactive oxygen species formed by reoxidation of Cu(I) to Cu(II); this feature was not observed in the absence of thiols. Mutagenic assays performed both in the presence and in the absence of S9 mix, probed by the Ames test on TA 98, TA 100, and TA 102, were negative. Weak genotoxic activity was detected for [CuL]+ and [CuL′]+, with a significative dose–response effect for [CuL′]+, which was shown to be more cytotoxic in the Ames test and 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide cell proliferation assays. Methylation of the terminal NH2 group enhances the antiproliferative activity of the pyridine-2-carbaldehyde thiosemicarbazones.  相似文献   

19.
The [RhCl3(N-N)(DMSO)] complexes, the N-N being 2,2′-bipyridine (1), 1,10-phenanthroline (2), 4,7-diphenyl-1,10-phenanthroline (3), 4,4′-dimethyl-2,2′-bipyridine (4) and 1,10-phenanthroline-5,6-dione (5), have been synthesized and characterized with spectroscopic methods. The compounds 2-5 adopt mer- and complex 1fac-structure. The molecular and electronic structure studies of mer- and fac-complexes with bpy and phen ligands at the DFT B3LYP level with 3-21G∗∗ basis set showed that mer-isomers are more stable. The cytostatic activity of the [RhCl3(N-N)(DMSO)] complexes against Caco-2 and A549 tumor cells have been studied. Their antibacterial activity have also been investigated. It has been found that the very promising biological activity show complexes 2, 3 and 4.  相似文献   

20.
Although it is known that (i) O6-alkylguanine-DNA alkyltransferase (AGT) confers tumor cell resistance to guanine O6-targeting drugs such as cloretazine, carmustine, and temozolomide and that (ii) AGT levels in tumors are highly variable, measurement of AGT activity in tumors before treatment is not a routine clinical practice. This derives in part from the lack of a reliable clinical AGT assay; therefore, a simple AGT assay was devised based on transfer of radioactive benzyl residues from [benzene-3H]O6-benzylguanine ([3H]BG) to AGT. The assay involves incubation of intact cells or cell homogenates with [3H]BG and measurement of radioactivity in a 70% methanol precipitable fraction. Approximately 85% of AGT in intact cells was recovered in cell homogenates. Accuracy of the AGT assay was confirmed by examination of AGT levels by Western blot analysis with the exception of false-positive results in melanin-containing cells due to [3H]BG binding to melanin. Second-order kinetic constants for human and murine AGT were 1100 and 380 M−1 s−1, respectively. AGT levels in various human cell lines ranged from less than 500 molecules/cell (detection limit) to 45,000 molecules/cell. Rodent cell lines frequently lacked AGT expression, and AGT levels in rodent cells were much lower than in human cells.  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