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.
Abstract

The DNA-binding properties of the anti-cancer drug amsacrine and a 9-aminoacridine analogue substituted at the 4 position with a 4-methanesulphonanilido-group, have been examined by means of unwinding, melting and equilibrium binding experiments. These find that the latter compound is at least as effective as a DNA-binder and intercalator as amsacrine itself. Molecular modelling and energetic calculations have confirmed this, and have produced plausible intercalation geometries. These show that there are subtle differences in the low-energy minor groove arrangements adopted by the substituents of the two drugs. Speculation is advanced that these differences may be relevant to the marked differences in cytotoxicity shown by the two compounds.  相似文献   

2.
The effect of the antitumour acridine derivative amsacrine [4'-(9-acridinylamino)methanesulphon-m-anisidide] on the fluorescence lifetime of DNA-bound ethidium has been investigated using a synchronously pumped cavity dumped dye laser producing picosecond pulses for sample excitation and a time-correlated single photon counting detection system. As the proportion of DNA-bound amsacrine on the synthetic DNA polymer poly[deoxyadenylic-thymidylic acid] is increased, the fluorescence decay curve of ethidium can be accurately resolved into two exponential components. The short lifetime component, whose proportion increases with increasing proportions of DNA-bound amsacrine, has a lifetime of between 3 and 4 ns, significantly longer than that of ethidium in aqueous solution (1.63 ns). The magnitude of the long lifetime component decreases from 25.4 to 14 ns with increasing proportions of bound amsacrine. It is concluded that a new fluorescence state of ethidium (lifetime 3-4 ns) is present, probably resulting from reversible electron transfer between ethidium and amsacrine. The ability of various 9-anilinoacridine derivatives to quench the fluorescence of DNA-bound ethidium appears to be related to the electron donor properties of the substituents on the anilino ring, as well as to experimental antitumour activity. The electron donor properties of DNA-bound amsacrine may therefore be relevant to its antitumour action.  相似文献   

3.
The mutagenicity of CI-921, the 4-methyl-5-(N-methyl)carboxamide derivative of the clinical antileukaemia agent, amsacrine, has been assessed using both bacterial and mammalian cells. CI-921 is distinguished from amsacrine in its high activity against some experimental tumours and is currently undergoing phase I clinical trial. Like 9-aminoacridine and amsacrine, CI-921 is mutagenic to the Salmonella typhimurium frameshift tester strain TA1537, but shows no sign of inducing base pair changes in strain TA100. In Chinese hamster cell culture, however, it differs from 9-aminoacridine in causing extensive chromosomal aberrations and an increase in mutations at the hypoxanthine-guanine phosphoribosyltransferase locus. It induces the formation of tightly packed and multilayered colonies in treated cultures of C3H/10T1/2 cells, but its action differs from that of benzo[a]pyrene, which induces type III fibroblastic multilayered colonies. Side-by-side comparison of the mutagenic properties of CI-921 and amsacrine showed no substantial differences at similar toxicity, suggesting that the increased lipophilicity and DNA-binding affinity of CI-921, which are thought to contribute to its increased antitumour activity, do not concomitantly increase the efficiency of in vitro mutagenesis or cell transformation.  相似文献   

4.
5.
Amsacrine is a DNA intercalating agent which is active against a number of tumours in mice and is used for the treatment of leukaemia in humans. In its DNA-bound form, amsacrine efficiently quenches the fluorescence of ethidium. Fluorescence lifetime studies demonstrate two populations of DNA-bound ethidium. The first, whose fluorescence lifetime is constant at approx. 3 ns and whose proportion increases with increasing amsacrine binding ratio, may comprise molecules bound in close proximity to amsacrine. The second, whose fluorescence lifetime is longer and variable (10-24 ns) and whose proportion decreases with increasing amsacrine binding ratio, may comprise molecules three or more base-pairs away from ethidium. Studies with a number of derivatives of 9-anilinoacridine containing different anilino substituents suggest that the observed wide variation in quenching capacity is correlated with the magnitude of the substituent dipole moment in a particular direction. Consideration of the geometry of the DNA-binding complex indicates that the negative pole of a dipole established in the anilino ring is directed towards a positively charged site on the ethidium molecule. Quenching of ethidium fluorescence may therefore occur where an electron-transfer complex has formed between ethidium and amsacrine molecules. To ascertain whether electron-transfer complex formation is biologically important in the amsacrine series, ethidium quenching has been quantitated and compared with activity against a transplantable neoplasm in mice, the Lewis lung carcinoma. Compounds which strongly quench ethidium fluorescence are in general highly active antitumour agents. The results are discussed in terms of a model where amsacrine has both a DNA-binding and a protein-binding domain, the latter possibly interacting by formation of an electron-transfer complex. The most likely protein-binding domain is on the enzyme topoisomerase II, the target for its cytotoxic activity.  相似文献   

6.
Because of the critical role of the DNA topoisomerases in the synthesis and conformation of DNA, and the well-known observation that radiation inhibits replicative DNA synthesis, we have examined the possibility that inhibitors of these enzymes might influence radiation lethality. In particular, using protocols involving the administration of either fresh or conditioned medium, we examined the ability of intercalative and nonintercalative inhibitors to affect the expression of potentially lethal damage and/or sublethal damage. The inhibitors examined were amsacrine, teniposide, etoposide, and novobiocin; only the latter compound was clearly effective in a selective way at nontoxic concentrations, and this was observed specifically in reference to the repair of potentially lethal damage effected by incubation in conditioned medium. These results are another example of differences between the repair of sublethal versus potentially lethal damage that further support distinctions between the two. At a mechanistic level, these and other data suggest that the property of novobiocin that is relevant in the foregoing is its metabolic inhibition of replicative DNA synthesis, a process which may be more important in the repair of potentially lethal damage as opposed to sublethal damage.  相似文献   

7.
Amsacrine is an acridine derivative drug applied in haematological malignancies. It targets topoisomerase II enhancing the formation of a cleavable DNA-enzyme complex and leading to DNA fragmentation in dividing cancer cells. Little is known about other modes of the interaction of amsacrine with DNA, by which it could affect also normal cells. Using the alkaline comet assay, we showed that amsacrine at concentrations from the range 0.01 to 10 microM induced DNA damage in normal human lymphocytes, human promyelocytic leukemia HL-60 cells lacking the p53 gene and murine pro-B lymphoid cells BaF3 expressing BCR/ABL oncogene measured as the increase in percentage tail DNA. The effect was dose-dependent. Treated cells were able to recover within a 120-min incubation. Amifostine at 14 mM decreased the level of DNA damage in normal lymphocytes, had no effect on the HL-60 cells and potentiated the DNA-damaging effect of the drug in BCR/ABL-transformed cells. Vitamin C at 10 and 50 microM diminished the extent of DNA damage in normal lymphocytes, but had no effect in cancer cells. Pre-treatment of the cells with the nitrone spin trap, N-tert-butyl-alpha-phenylnitrone or ebselen, which mimics glutathione peroxidase, reduced the extent of DNA damage evoked by amsacrine in all types of cells. The cells exposed to amsacrine and treated with endonuclease III and 3-methyladenine-DNA glycosylase II, the enzymes recognizing oxidized and alkylated bases, respectively, displayed greater extent of DNA damage than those not treated with these enzymes. The results obtained suggest that free radicals may be involved in the formation of DNA lesions induced by amsacrine. The drug can also methylate DNA bases. Our results indicate that the induction of secondary malignancies should be taken into account as diverse side effects of amsacrine. Amifostine may potentate DNA-damage effect of amsacrine in cancer cells and decrease this effect in normal cells and Vitamin C can be considered as a protective agent against DNA damage in normal cells.  相似文献   

8.
DNase I footprinting has been used to probe the sequence selectivity of binding of a series of intercalating amsacrine-4-carboxamides and a related 9-aminoacridine-4-carboxamide to three DNA restriction fragments. These ligands have good experimental antileukemic activity, and for those members of the series that gave evaluable footprints, our principal finding is that they bind preferentially to GC-rich regions in agreement with the conclusion of equilibrium and kinetic measurements. The highest affinity sites generally occur in clusters of GC base pairs with runs of AT pairs being excluded from binding. It is important to appreciate that the 9-aminoacridine- and amsacrine-4-carboxamides exhibit a very high degree of selectivity for GC sites which, to our knowledge, has not been previously matched by acridine derivatives in footprinting experiments. The principal determinant of specificity appears to be the 4-carboxamide group itself since neither variations in the terminal funtionality of the 4-carboxamide sidechain nor the presence of the 9-anilino substituent modifies sequence preferences. The molecular origins of selectivity may be discerned in terms of potential hydrogen bonding interactions between the 4-carboxamide moiety and carbonyl oxygen and amino groups of GC base pairs in the DNA minor groove at CG dinucleotide sites. The related therapeutic agent amsacrine failed to inhibit cleavage by DNase I, so no conclusion can be drawn concerning its binding selectivity, save to note that amsacrine does not possess the 4-carboxamide group which appears to be the crucial determinant of GC specificity. Whether selectivity for binding to GC-rich sequences is an important element in the antitumor activity of both the 9-aminoacridine- and amsacrine-4-carboxamides remains to be determined.  相似文献   

9.
Amsacrine is an anilinoacridine derivative anticancer drug, used to treat a wide variety of malignancies. In cells, amsacrine poisons topoisomerase 2 by stabilizing DNA-drug-enzyme ternary complex. Presence of amsacrine increases the steady-state concentration of these ternary complexes which in turn hampers DNA replication and results in subsequent cell death. Due to reversible binding and rapid slip-out of amsacrine from DNA duplex, structural data is not available on amsacrine-DNA complexes. In the present work, we designed five oligonucleotide duplexes, differing in their minor groove widths and hydration pattern, and examined their binding with amsacrine using attenuated total reflection Fourier transform infrared (ATR-FTIR) spectroscopy. Complexes of amsacrine with calf thymus DNA were also evaluated for a comparison. Our results demonstrate for the first time that amsacrine is not a simple intercalator; rather mixed type of DNA binding (intercalation and minor groove) takes place between amsacrine and DNA. Further, this binding is highly sensitive towards the geometries and hydration patterns of different minor grooves present in the DNA. This study shows that ligand binding to DNA could be very sensitive to DNA base composition and DNA groove structures. Results demonstrated here could have implication for understanding cytotoxic mechanism of aminoacridine based anticancer drugs and provide directions to modify these drugs for better efficacy and few side-effects.  相似文献   

10.
11.
It has been shown that inhibitors of topoisomerase II (topo II) etoposide and amsacrine results in accumulation of GM-130 and Hep-2 cells with 4c DNA amount. The differential analysis based on flow cytometry (Zenin et al., 2001) and enabled us to discriminate cells with 4c DNA--G2, M, including metaphase and anaphase cells and cells in pseudo-G1. 1 microM etoposide evoked cell accumulation in G2 phase, while 40 microM etoposide blocked cell proliferation, which was confirmed by a complete absence of both mitotic cells and 4c DNA cell accumulation. GM-130 and Hep-2 cells that were first blocked and then washed from nocodazole, and after that treated with 50 microM etoposide or 20 microM amsacrine, were shown to enter pseudo-G1 with 4c DNA amount per cell. In the presence of nocodazole, 4 and 40 microM amsacrine evoked transition of all mitotic cells to pseudo-G1 within 1 h. 15 or 30 minutes pulse treatments of GM-130 cells with 40 microM amsacrine in the presence of nocodazole, followed by incubation in drug-free medium, resulted in the similar transition of cells to pseudo-G1.  相似文献   

12.
Diploid human fibroblast strains were treated for 10 min with inhibitors of type I and type II DNA topoisomerases, and after removal of the inhibitors, the rate of initiation of DNA synthesis at replicon origins was determined. By alkaline elution chromatography, 4'-(9-acridinylamino)methanesulfon-m-anisidide (amsacrine), an inhibitor of DNA topoisomerase II, was shown to produce DNA strand breaks. These strand breaks are thought to reflect drug-induced stabilization of topoisomerase-DNA cleavable complexes. Removal of the drug led to a rapid resealing of the strand breaks by dissociation of the complexes. Velocity sedimentation analysis was used to quantify the effects of amsacrine treatment on DNA replication. It was demonstrated that transient exposure to low concentrations of amsacrine inhibited replicon initiation but did not substantially affect DNA chainelongation within operating replicons. Maximal inhibition of replicon initiation occurred 20 to 30 min after drug treatment, and the initiation rate recovered 30 to 90 min later. Ataxia telangiectasia cells displayed normal levels of amsacrine-induced DNA strand breaks during stabilization of cleavable complexes but failed to downregulate replicon initiation after exposure to the topoisomerase inhibitor. Thus, inhibition of replicon initiation in response to DNA damage appears to be an active process which requires a gene product which is defective or missing in ataxia telangiectasia cells. In normal human fibroblasts, the inhibition of DNA topoisomerase I by camptothecin produced reversible DNA strand breaks. Transient exposure to this drug also inhibited replicon initiation. These results suggest that the cellular response pathway which downregulates replicon initiation following genotoxic damage may respond to perturbations of chromatin structure which accompany stabilization of topoisomerase-DNA cleavable complexes.  相似文献   

13.
CHO-AA8 cells were used as a model system to study the role of DNA topoisomerase II in the resistance of non-cycling cells to amsacrine. Plateau-phase AA8 cells have previously been shown to be resistant to amsacrine and to contain fewer DNA breaks than log-phase cells after drug treatment (Robbie, M.A., Baguley, B.C., Denny, W.A., Gavin, J.R. and Wilson, W.R. (1988) Cancer Res., in press). The phage P4-unknotting activity of nuclear extracts decreased 2-fold when AA8 cells entered into the non-cycling state, but there was no difference in sensitivity to amsacrine between log- and plateau-phase nuclear extracts. Drug stimulation of protein-DNA complex formation was similar in whole cells, isolated nuclei and nuclear extracts from either log- or plateau-phase cells. However, stimulation of complex formation in cells, nuclei or nuclear extracts was approx. 4-fold lower in plateau-phase than in log-phase. The data presented suggested that drug-enzyme interaction was altered in plateau-phase cells.  相似文献   

14.
Okada Y  Tosaka A  Nimura Y  Kikuchi A  Yoshida S  Suzuki M 《Gene》2001,272(1-2):141-148
In human cells, atypical drug resistance was previously identified with reduced catalytic activity or nuclear localization efficiency of DNA topoisomerase II alpha (TOP2 alpha). We have shown two etoposide resistant hTOP2 alpha mutants, K798L and K798P confer resistance to etoposide. In this work, we showed these mutants are also resistant against doxorubicin and mAMSA in vivo in the yeast strain ISE2, rad52, top2-4 at the non-permissive temperature. We purified these mutants to characterize the drug resistant mechanism. Purified recombinant proteins were 8- to 12-fold more resistant to etoposide and doxorubicin than wild type TOP2 alpha, and 2-fold more resistant to amsacrine, as measured by accumulation of cleavable DNA. These data show that K798L and K798P may be intrinsically resistant against these drugs in vitro and that this character may confer atypical multidrug resistant phenotype in vivo in yeast.  相似文献   

15.
We have investigated the biochemical basis for the hypersensitivity to intercalating agents and epipodophyllotoxins of a Chinese hamster cell mutant, ADR-1. More topoisomerase II-induced DNA strand breaks are accumulated by ADR-1 than by parental CHO-K1 cells following exposure to the intercalating agent amsacrine. Levels of induced DNA strand breaks correlate with cell killing. Topoisomerase II activity is elevated in ADR-1 cells as a consequence of an increased cellular level of topoisomerase II protein. We have studied the phenotype of cell hybrids generated by fusing parental and mutant cells. The hybrid ADR-1/CHO-K1 exhibits normal levels of resistance to amsacrine and expresses the lower, parental level of topoisomerase II. These results provide additional evidence that topoisomerase II mediates the cytotoxic action of intercalating agents and epipodophyllotoxins and suggest that the intracellular level of topoisomerase II is an important determinant of cellular sensitivity to these drugs. This has implications for antitumor therapy. ADR-1 cells provide a model system for studying the effects of topoisomerase II overproduction on cell proliferation and chromosome organization.  相似文献   

16.
CHO-AA8 cells were used as a model system to study the role of DNA topoisomerase II in the resistance of non-cycling cells to amsacrine. Plateau-phase AA8 cells have previously been shown to be resistant to amsacrine and to contain fewer DNA breaks than log-phase cells after drug treatment (Robbie, M.A., Baguley, B.C., Denny, W.A., Gavin, J.R. and Wilson, W.R. (1988) Cancer Res., in press). The phage P4-unknotting activity of nuclear extracts decreased 2-fold when AA8 cells entered into the non-cycling state, but there was no difference in sensitivity to amsacrine between log- and plateau-phase nuclear extracts. Drug stimulation of protein-DNA complex formation was similar in whole cells, isolated nuclei and nuclear extracts from either log- or plateau-phase cells. However, stimulation of complex formation in cells, nuclei or nuclear extracts was approx. 4-fold lower in plateau-phase than in log-phase. The data presented suggested that drug-enzyme interaction was altered in plateau-phase cells.  相似文献   

17.
Several classes of antitumor drugs are known to stabilize topoisomerase complexes in which the enzyme is covalently bound to a terminus of a DNA strand break. The DNA cleavage sites generally are different for each class of drugs. We have determined the DNA sequence locations of a large number of drug-stimulated cleavage sites of topoisomerase II, and find that the results provide a clue to the possible structure of the complexes and the origin of the drug-specific differences. Cleavage enhancements by VM-26 and amsacrine (m-AMSA), which are representative of different classes of topoisomerase II inhibitors, have strong dependence on bases directly at the sites of cleavage. The preferred bases were C at the 3' terminus for VM-26 and A at the 5' terminus for m-AMSA. Also, a region of dyad symmetry of 12 to 16 base pairs was detected about the enzyme cleavage positions. These results are consistent with those obtained with doxorubicin, although in the case of doxorubicin, cleavage requires the presence of an A at the 3' terminus of at least one the pair of breaks that constitute a double-strand cleavage (Capranico et al., Nucleic Acids Res., 1990, 18: 6611). These findings suggest that topoisomerase II inhibitors may stack with one or the other base pair flanking the enzyme cleavage sites.  相似文献   

18.
A number of topoisomerase II-targeted anticancer drugs, including amsacrine, utilize an acridine or related aromatic core as a scaffold. Therefore, to further explore the potential of acridine-related compounds to act as topoisomerase II poisons, we synthesized a series of novel trifluoromethylated 9-amino-3,4-dihydroacridin-1(2H)-one derivatives and examined their ability to enhance DNA cleavage mediated by human topoisomerase IIα. Derivatives containing a H, Cl, F, and Br at C7 enhanced enzyme-mediated double-stranded DNA cleavage ~5.5- to 8.5-fold over baseline, but were less potent than amsacrine. The inclusion of an amino group at C9 was critical for activity. The compounds lost their activity against topoisomerase IIα in the presence of a reducing agent, displayed no activity against the catalytic core of topoisomerase IIα, and inhibited DNA cleavage when incubated with the enzyme prior to the addition of DNA. These findings strongly suggest that the compounds act as covalent, rather than interfacial, topoisomerase II poisons.  相似文献   

19.
Werner's syndrome (WS) is a rare autosomal recessive human disorder and the patients exhibit many symptoms of accelerated ageing in their early adulthood. The gene (WRN) responsible for WS has been biochemically characterised as a 3'-5' helicase and is homologous to a number of RecQ superfamily of helicases. The yeast SGS1 helicase is considered as a human WRN homologue and SGS1 physically interacts with topoisomerases II and III. In view of this, it has been hypothesised that the WRN gene may also interact with topoisomerases II and III. The purpose of this study is to determine whether the loss of function of WRN protein alters the sensitivity of WS cells to agents that block the action of topoisomerase II. This study deals with the comparison of the chromosomal damage induced by the two anti-topoisomerase II drugs, VP-16 and amsacrine, in both G1 and G2 phases of the cell cycle, in lymphoblastoid cells from WS patients and from a healthy donor. Our results show that the WS cell lines are hypersensitive to chromosome damage induced by VP-16 and amsacrine only in the G2 phase of the cell cycle. No difference either in the yield of the induced aberrations or SCEs was found after treatment of cells at G1 stage. These data might suggest that in WS cells, because of the mutation of the WRN protein, the inhibition of topoisomerase II activity results in a higher rate of misrepair, probably due to some compromised G2 phase processes involving the WRN protein.  相似文献   

20.
The frameshift mutagenicity of the DNA intercalating drug proflavine is known to be enhanced by photoirradiation of bacterial cultures. To determine whether this phenomenon was also present in acridine-derived antitumour drugs, cultures of Salmonella typhimurium were exposed to the antileukaemia agent amsacrine and the experimental agent N-[2-(dimethylamino)ethyl]acridine-4-carboxamide dihydrochloride (acridine carboxamide) in the presence or absence of visible light. A small increase in mutagenicity was observed with amsacrine but not with acridine carboxamide. A series of analogues of amsacrine were then tested, and a striking relationship was found between the minimum drug concentration for mutagenicity and DNA binding affinity. In each case, photoirradiation was associated with a small increase in mutagenicity. Each of the compounds showing the photo-enhancement effect was capable of reversible one-electron oxidation. It is suggested that this oxidation occurs in bacteria, and that the DNA binding constant of the resulting acridine radical species will increase because of the extra positive charge. This increased DNA binding would be sufficient to explain the photo-enhancement of mutagenicity of these drugs.  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