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 234 毫秒
1.
Antineoplastic bis(dioxopiperazine)s, such as meso-2,3-bis(2,6-dioxopiperazin-4-yl)butane (ICRF-193), are widely believed to be only catalytic inhibitors of topoisomerase II. However, topoisomerase inhibitors have little or no antineoplastic activity unless they are topoisomerase poisons, a special subclass of topoisomerase-targeting drugs that stabilize topoisomerase-DNA strand passing intermediates and thus cause the topoisomerase to become a cytotoxic DNA-damaging agent. Here we report that ICRF-193 is a very significant topoisomerase II poison. Detection of topoisomerase II poisoning by ICRF-193 required the use of a chaotropic protein denaturant in the topoisomerase poisoning assays. ICRF-193 caused dose-dependent cross-linking of human topoisomerase IIbeta to DNA and stimulated topoisomerase IIbeta-mediated DNA cleavage at specific sites on (32)P-end-labeled DNA. Human topoisomerase IIalpha-mediated DNA cleavage was stimulated to a lesser extent by ICRF-193. In vivo experiments with MCF-7 cells also showed the requirement of a chaotropic protein denaturant in the assays and selectivity for the beta-isozyme of human topoisomerase II. Studies with two topoisomerase IIbeta-negative cell model systems confirmed significant topoisomerase II poisoning by ICRF-193 in the wild type cells and were consistent with beta-isozyme selectivity. Common use of only the detergent, SDS, in assays may have led to failure to detect topoisomerase II poisoning by ICRF-193 in earlier studies.  相似文献   

2.
p14ARF (ARF) and topoisomerase I play central roles in cancer and have recently been shown to interact. The interaction activates topoisomerase I, an important target for camptothecin-like chemotherapeutic drugs, but the regulation of the interaction is poorly understood. We have used the H358 and H23 lung cancer cell lines and purified recombinant human topoisomerase I to demonstrate that the ARF/topoisomerase I interaction is regulated by topoisomerase I serine phosphorylation, a modification that regulates topoisomerase I activity. Both cell lines express wild-type ARF and topoisomerase I proteins at equivalent levels, but H23 topoisomerase I, unlike that of H358 cells, is largely devoid of serine phosphorylation, has low activity, and complexes poorly with ARF. The ability of H23 topoisomerase I to complex with ARF can be restored by treatment with the serine kinase, casein kinase II. Consistent with these observations, we show that the response of H23 cells to camptothecin treatment is unaffected by changes in intracellular levels of ARF. However, in H358 and PC-3 cells, which express a serine phosphorylated topoisomerase I that complexes with ARF, ectopic overexpression of ARF causes sensitization to camptothecin, and siRNA-mediated down-regulation of endogenous ARF causes desensitization to camptothecin. These biological responses correlate with increased and decreased levels, respectively, of ARF/topoisomerase I complex and DNA-bound topoisomerase I. Thus, ARF is a serine phosphorylation-dependent coregulator of topoisomerase I in vivo, and it regulates cellular sensitivity to camptothecin by interacting with topoisomerase I. Certain cancer associated defects affecting ARF/topoisomerase I complex formation could contribute to cellular resistance to camptothecin.  相似文献   

3.
The physiological role of topoisomerase III is unclear for any organism. We show here that the removal of topoisomerase III in temperature sensitive topoisomerase IV mutants in Escherichia coli results in inviability at the permissive temperature. The removal of topoisomerase III has no effect on the accumulation of catenated intermediates of DNA replication, even when topoisomerase IV activity is removed. Either recQ or recA null mutations, but not helD null or lexA3, partially rescued the synthetic lethality of the double topoisomerase III/IV mutant, indicating a role for topoisomerase III in recombination. We find a bias against deleting the gene encoding topoisomerase III in ruvC53 or DeltaruvABC backgrounds compared with the isogenic wild-type strains. The topoisomerase III RuvC double mutants that can be constructed are five- to 10-fold more sensitive to UV irradiation and mitomycin C treatment and are twofold less efficient in transduction efficiency than ruvC53 mutants. The overexpression of ruvABC allows the construction of the topoisomerase III/IV double mutant. These data are consistent with a role for topoisomerase III in disentangling recombination intermediates as an alternative to RuvABC to maintain the stability of the genome.  相似文献   

4.
Mammalian cells express two genetically distinct isoforms of DNA topoisomerase II, designated topoisomerase IIalphaand topoisomerase IIbeta. We have recently shown that mouse topoisomerase IIalpha can substitute for the yeast topoisomerase II enzyme and complement yeast top2 mutations. This functional complementation allowed functional analysis of the C-terminal domain (CTD) of mammalian topoisomerase II, where the amino acid sequences are divergent and species-specific, in contrast to the highly conserved N-terminal and central domains. Several C-terminal deletion mutants of mouse topoisomerase IIalpha were constructed and expressed in yeast top2 cells. We found that the CTD of topoisomerase IIalphais dispensable for enzymatic activity in vitro but is required for nuclear localization in vivo. Interestingly, the CTD of topoisomerase IIbetawas also able to function as a signal for nuclear targeting. We therefore examined whether the CTD alone is sufficient for nuclear localization in vivo . The C-terminal region was fused to GFP (green fluorescent protein) and expressed under the GAL1 promoter in yeast cells. As expected, GFP signal was exclusively detected in the nucleus, irrespective of the CTD derived from either topoisomerase IIalphaor IIbeta. Surprisingly, when the upstream sequence of each CTD was added nuclear localization of the GFP signal was found to be cell cycle dependent: topoisomerase IIalpha-GFP was seen in the mitotic nucleus but was absent from the interphase nucleus, while topoisomerase IIbeta-GFP was detected predominantly in the interphase nucleus and less in the mitotic nucleus. Our results suggest that the catalytically dispensable CTD of topoisomerase II is sufficient as a signal for nuclear localization and that yeast cells can distinguish between the two isoforms of mammalian topoisomerase II, localizing each protein properly.  相似文献   

5.
Two forms of topoisomerase I can be purified from Xenopus laevis. A protein with a molecular mass of 165 kDa has been identified as topoisomerase I in ovaries (Richard and Bogenhagen, 1989. J. Biol. Chem. 264, 4704-4709). When a similar purification is performed using liver tissue, topoisomerase I is purified as a 110-kDa protein. Separate rabbit antisera were raised against oocyte and liver topoisomerase I polypeptides. Each antiserum reacts in immunoblotting or immunoprecipitation procedures only with the tissue-specific topoisomerase I polypeptide against which it was generated. The failure of the antiserum raised against liver topoisomerase I to cross-react with the oocyte enzyme suggests that the smaller topoisomerase I is not derived from the 165-kDa oocyte enzyme by proteolysis. X. laevis tissue culture cells lysed and processed in the presence of SDS contain the 110-kDa form of topoisomerase I. The 165-kDa form of topoisomerase I disappears during oocyte maturation in vitro.  相似文献   

6.
It has been known for some time that ATP inhibits the DNA relaxation activity of human DNA topoisomerase I. However, the underlying mechanism of this inhibitory effect remains largely unknown. Using filter binding assays, the binding of human DNA topoisomerase I to DNA was decreased in the presence of ATP. This result suggests that the inhibition of DNA relaxation activity of human DNA topoisomerase I by ATP is at the binding step rather than at the nicking or resealing step. DNA topoisomerase I cleavage assay further supports this notion. ATP-agarose binding and UV cross-linking assays also demonstrate that ATP directly and specifically binds human DNA topoisomerase I. To address whether the ATP binding results in conformational changes in human DNA topoisomerase I, various proteases were employed for detecting potential protein conformational changes. Our results indicated that the proteolytic susceptibilities of trypsin and chymotrypsin were altered in the presence of ATP. The result suggests that the conformation of human DNA topoisomerase I was altered upon ATP binding. In addition, the binding between ATP and human DNA topoisomerase I was also reduced by increasing concentrations of DNA. Our data suggests that human DNA topoisomerase I exhibits at least two incompatible conformations. One conformation is in the form of a topoisomerase I-ATP complex, which inhibits DNA relaxation activity of human DNA topoisomerase I, and the other, a topoisomerase I-DNA complex, which exerts DNA relaxation activity. Our studies identify the role of ATP in the regulation of human DNA topoisomerase I and provide a substantial implication of how human DNA topoisomerase I compromises its versatile functions.  相似文献   

7.
Human topoisomerase II plays a crucial role in DNA replication and repair. It exists in two isoforms: topoisomerase II alpha (alpha) and topoisomerase II beta (beta). The alpha isoform is localized predominantly in the nucleus, while the beta isoform exhibits a reticular pattern of distribution both in the cytosol and in the nucleus. We show that both isoforms of topoisomerase II are phosphorylated in HIV infected cells and also by purified viral lysate. An analysis of the phosphorylation of topoisomerase II isoforms showed that extracts of HIV infected cells at 8 and 32 h. post-infection (p.i.) contain maximal phosphorylated topoisomerase II alpha, whereas infected cell extracts at 4 and 64 h p.i. contain maximum levels of phosphorylated topoisomerase II beta. In concurrent to phosphorylated topoisomerase II isoforms, we have also observed increased topoisomerase II alpha kinase activity after 8h p.i and topoisomerase beta kinase activity at 4 and 64 h p.i. These findings suggest that both topoisomerase II alpha and beta kinase activities play an important role in early as well as late stages of HIV-1 replication. Further analysis of purified virus showed that HIV-1 virion contained topoisomerase II isoform-specific kinase activities, which were partially isolated. One of the kinase activities of higher hydrophobicity can phosphorylate both topoisomerase II alpha and beta, while lower hydrophobic kinase could predominantly phosphorylate topoisomerase II alpha. The phosphorylation status was correlated with catalytic activity of the enzyme. Western blot analysis using phosphoamino-specific antibodies shows that both the kinase activities catalyze the phosphorylation at serine residues of topoisomerase II alpha and beta. The catalytic inhibitions by serine kinase inhibitors further suggest that the alpha and beta kinase activities associated with virus are distinctly different.  相似文献   

8.
It is known that topoisomerase IIalpha is phosphorylated by several kinases. To elucidate the role of phosphorylation of topoisomerase IIalpha in the cell cycle, we have examined the cell cycle behavior of phosphorylated topoisomerase IIalpha in HeLa cells using antibodies against several phospho-oligopeptides of this enzyme. Here we demonstrate that serine1212 in topoisomerase IIalpha is phosphorylated only in the mitotic phase. Using an antibody against an oligopeptide containing phosphoserine-1212 in topoisomerase IIalpha (PS1212), subcellular localization of topoisomerase IIalpha phosphorylated at serine1212 was examined by indirect immunofluorescence staining, and compared with that of overall topoisomerase IIalpha. Serine1212-phosphorylated topoisomerase IIalpha was localized specifically on mitotic chromosomes, but not on interphase chromosomes; this result contrasts with overall topoisomerase IIalpha which was observed on chomosomes in both interphase and mitosis. Serine1212-phosphorylated topoisomerase lIalpha first appeared on chromosome arms in prophase, became concentrated on the centromeres in metaphase, and disappeared in early telophase. In addition, ICRF-193, a catalytic inhibitor of topoisomerase II, prevented accumulation of serine1212-phosphorylated topoisomerase IIalpha at the centromeres. These results indicate that serine1212 of topoisomerase IIalpha is phosphorylated specifically during mitosis, and suggest that the serine1212-phosphorylated topoisomerase IIalpha acts on resolving topological constraint progressively from the chromosome arm to the centromere during metaphase chromosome condensation.  相似文献   

9.
10.
Poly(ADP-ribosylation) of DNA topoisomerase I from calf thymus   总被引:13,自引:0,他引:13  
We demonstrate that the activity of the major DNA topoisomerase I from calf thymus is severely inhibited after modification by purified poly(ADP-ribose) synthetase. Polymeric chains of poly(ADP-ribose) are covalently attached to DNA topoisomerase I. These observations with highly purified enzymes suggest that poly(ADP-ribosylation) may be a cellular mechanism for modulating DNA topoisomerase I activity in response to the state of DNA in the nucleus. Although extensive poly(ADP-ribosylation) of the Mr = 100,000 DNA topoisomerase I from calf thymus resulted in greater than 90% enzyme inhibition, exogenous poly(ADP-ribose) does not, by itself, inhibit topoisomerase activity. After modification, the apparent molecular weight of both the topoisomerase enzyme protein and of the topoisomerase enzyme activity was increased. In vitro, the extent of modification of DNA topoisomerase I could be controlled either by changing the ratio of topoisomerase to the synthetase or by varying the reaction time. More than 40 residues of ADP ribose per topoisomerase molecule could be added by the synthetase. Analysis of a poly(ADP-ribosylated) topoisomerase preparation that was about 50% inhibited revealed an average polymer chain length of 7.4, with 1-2 chains per enzyme molecule.  相似文献   

11.
HL-60/AMSA is a human leukemia cell line that is 100 times more resistant to the cytotoxic actions of the antineoplastic, topoisomerase II-reactive DNA intercalating acridine derivative amsacrine (m-AMSA) than is its parent HL-60 line. HL-60/AMSA cells are minimally resistant to etoposide, a topoisomerase II-reactive drug that does not intercalate. Previously we showed that HL-60 topoisomerase II activity in cells, nuclei, or nuclear extracts was sensitive to m-AMSA and etoposide, while HL-60/AMSA topoisomerase II was resistant to m-AMSA but sensitive to etoposide. Now we show that purified topoisomerase II from the two cell lines exhibits the same drug sensitivity or resistance as that in the nuclear extracts although the magnitude of the m-AMSA resistance of HL-60/AMSA topoisomerase II in vitro is not as great as the resistance of the intact HL-60/AMSA cells. In addition HL-60/AMSA cells are cross-resistant to topoisomerase II-reactive intercalators from the anthracycline and ellipticine families and the pattern of sensitivity or resistance to the cytotoxic actions of the various topoisomerase II-reactive drugs is paralleled by topoisomerase II-reactive drug-induced DNA cleavage and protein cross-link production in cells and the production of drug-induced, topoisomerase II-mediated DNA cleavage and protein cross-linking in isolated biochemical systems. In addition to its lowered sensitivity to intercalators, HL-60/AMSA differed from HL-60 in 1) the susceptibility of its topoisomerase II to stimulation of DNA topoisomerase II complex formation by ATP, 2) the catalytic activity of its topoisomerase II in an ionic environment chosen to reproduce the environment found within the living cell, and 3) the observed restriction enzyme pattern on a Southern blot probed with a cDNA for human topoisomerase II. These data indicate that an m-AMSA-resistant form of topoisomerase II contributes to the resistance of HL-60/AMSA to m-AMSA and to other topoisomerase II-reactive DNA intercalating agents. The drug resistance is associated with additional biochemical and molecular alterations that may be important determinants of cellular sensitivity or resistance to topoisomerase II-reactive drugs.  相似文献   

12.
Characterization of a potent catenation activity of HeLa cell nuclei   总被引:1,自引:0,他引:1  
Using an assay which measures catenation of a supercoiled DNA template, we have characterized and quantitated a potent activity identified in crude fractions of HeLa cell nuclei. Catenation requires Mg-ATP and a DNA-condensing agent, polyvinyl alcohol. A filter-binding or agarose gel assay can be used to quantitate activity. In this reaction, DNA topoisomerase I relaxes the input supercoiled DNA to provide DNA topoisomerase II, a strongly favored template for catenation. DNA topoisomerase II preferentially catenates relaxed DNA over supercoiled DNA by a factor of 100. One molecule of DNA topoisomerase II is able to catenate about 20 circles of relaxed DNA/min at 30 degrees C but only 0.16 circle of supercoiled DNA/min at 30 degrees C. The purified HeLa topoisomerase I can also catenate DNA under these assay conditions, yet in an ATP-independent fashion. It is much less efficient than topoisomerase II; one molecule of topoisomerase I catenates only about 3.8 X 10(-3) molecules of supercoiled DNA/min at 30 degrees C with a DNA template containing 5% nicked circles. This remarkable difference between the two enzymes allows quantitation of DNA topoisomerase II activity seen in the presence of excess topoisomerase I. Unlike Escherichia coli topoisomerase I (omega), catenation by the HeLa topoisomerase I is not stimulated by gapped circles.  相似文献   

13.
TAS-103 is a novel antineoplastic agent that is active against in vivo tumor models [Utsugi, T., et al. (1997) Jpn. J. Cancer Res. 88, 992-1002]. This drug is believed to be a dual topoisomerase I/II-targeted agent, because it enhances both topoisomerase I- and topoisomerase II-mediated DNA cleavage in treated cells. However, the relative importance of these two enzymes for the cytotoxic actions of TAS-103 is not known. Therefore, the primary cellular target of the drug and its mode of action were determined. TAS-103 stimulated DNA cleavage mediated by mammalian topoisomerase I and human topoisomerase IIalpha and beta in vitro. The drug was less active than camptothecin against the type I enzyme but was equipotent to etoposide against topoisomerase IIalpha. A yeast genetic system that allowed manipulation of topoisomerase activity and drug sensitivity was used to determine the contributions of topoisomerase I and II to drug cytotoxicity. Results indicate that topoisomerase II is the primary cellular target of TAS-103. In addition, TAS-103 binds to human topoisomerase IIalpha in the absence of DNA, suggesting that enzyme-drug interactions play a role in formation of the ternary topoisomerase II.drug.DNA complex. TAS-103 induced topoisomerase II-mediated DNA cleavage at sites similar to those observed in the presence of etoposide. Like etoposide, it enhanced cleavage primarily by inhibiting the religation reaction of the enzyme. Based on these findings, it is suggested that TAS-103 be classified as a topoisomerase II-targeted drug.  相似文献   

14.
Screening of a human B-cell cDNA library with a topoisomerase II beta gene-specific probe revealed the presence of two distinct forms of topoisomerase II beta cDNA. One form (designated topoisomerase II beta-1), representing the majority of the clones, would encode the topoisomerase II beta amino acid sequence reported recently [Jenkins, J.R. et al. (1992) Nucleic Acids Res., 20, 5587-5592]. The second form (designated topoisomerase II beta-2) would encode a protein containing an additional 5 amino acids inserted after Valine-23 of the topoisomerase II beta-1 protein sequence. The topoisomerase II beta-1 and beta-2 mRNAs were both widely expressed in human cell lines and tissues. Topoisomerase II beta-2 mRNA was expressed at a lower level than that of the beta-1 form, but the relative expression of the two forms varied in different cell types. Analysis of genomic DNA clones revealed that the two forms of topoisomerase II beta mRNA arose via differential splicing. These data indicate that in addition to the closely related topoisomerase II alpha and beta isozymes, there are two forms of topoisomerase II beta mRNA widely expressed in human cells.  相似文献   

15.
16.
17.
Cellular topoisomerase I has been reported to be present in retroviral particles and to enhance viral cDNA synthesis; however, the mechanisms involved remain unknown. In the present study, it has been demonstrated that human topoisomerase I combines with a stem-loop RNA and that the bound topoisomerase I can be dissociated from RNA substrates in the presence of ATP. In addition, in vitro cleaved synthetic RNA bound by topoisomerase I is subsequently relegated when the topoisomerase I is dissociated by ATP. A mechanism is proposed in which human topoisomerase I is carried into virions and regulates the repair of genomic RNA by its ligation activity.  相似文献   

18.
The induction by interleukin-2 of DNA topoisomerase I and DNA topoisomerase II activities in the human T cell line HuT 78 was investigated. HuT 78 cells were treated with 1000 U of interleukin-2/ml, and extracts of the HuT 78 nuclei were prepared over a 24 h period. The extracts were assayed quantitatively for the activities of DNA topoisomerase I and DNA topoisomerase II. Three concomitant, transient increases of 3- to 11-fold in the specific activities of both DNA topoisomerase I and DNA topoisomerase II were observed following treatment with IL-2 at 0.5, 4, and 10 h after treatment with interleukin-2. The specific activities of both enzymes returned to base-line values after each of these transient increases. These results reveal that the activities of DNA topoisomerase I and DNA topoisomerase II are highly regulated in HuT 78 cells upon treatment with IL-2.  相似文献   

19.
Kowalska-Loth  B.  Bubko  I.  Komorowska  B.  Szumiel  I.  Staron  K. 《Molecular biology reports》1998,25(1):21-26
An in vitro system composed of nicked pBR322 DNA and purified topoisomerase I was employed to study the efficiency of the topoisomerase I-driven single-strand to double-strand DNA breaks conversion. At 1.4 × 105 topoisomerase I activity units per mg DNA about 20% single-strand nicks were converted into double-strand breaks during 30 min due to topoisomerase I action. Camptothecin inhibited the conversion. The conversion was also inhibited when the relaxing activity of the used topoisomerase I was increased by phosphorylation of the enzyme with casein kinase 2. The presented data suggest that topoisomerase I may be involved in production of double-stranded breaks in irradiated cells and that this process positively depends on the amount of topoisomerase I but not on its phosphorylation state.  相似文献   

20.
DNA topoisomerases I and II are the two major nuclear enzymes capable of relieving torsional strain in DNA. Of these enzymes, topoisomerase I plays the dominant role in relieving torsional strain during chromatin assembly in cell extracts from oocytes, eggs, and early embryos. We tested if the topoisomerases are used differentially during chromatin assembly in Saccharomyces cerevisiae by a combined biochemical and pharmacological approach. As measured by plasmid supercoiling, nucleosome deposition is severely impaired in assembly extracts from a yeast mutant with no topoisomerase I and a temperature-sensitive form of topoisomerase II (strain top1-top2). Expression of wild-type topoisomerase II in strain top1-top2 fully restored assembly-driven supercoiling, and assembly was equally efficient in extracts from strains expressing either topoisomerase I or II alone. Supercoiling in top1-top2 extract was rescued by adding back either purified topoisomerase I or II. Using the topoisomerase II poison VP-16, we show that topoisomerase II activity during chromatin assembly is the same in the presence and absence of topoisomerase I. We conclude that both topoisomerases I and II can provide the DNA relaxation activity required for efficient chromatin assembly in mitotically cycling yeast 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