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.
The nuclear enzymes DNA topoisomerases I and II appeared as cellular targets for several antitumor drugs: campthotecin derivatives interacting with topoisomerase I, and actinomycin D, anthracycline derivatives, elliptinium acetate, mitoxantrone, epipodophyllotoxine derivatives, amsacrine and a new olivacine derivative, NSC-6596871 (S 16020-2), which interact with topoisomerase II. The functions of these enzymes are numerous and important since they are critical for DNA functions and cell survival. Despite the fact that they share the same target, topoisomerase II inhibitors have different mechanisms of action. Two principle types of induced alterations are involved in cellular resistance to topoisomerase II drugs: qualitative or quantitative alteration of the enzyme and/or increased drug efflux due to overexpression of P-glycoprotein. S 16020-2, a new olivacine derivative with a high antitumor activity against solid tumors, shows a potent cytotoxic effect against tumor cells expressing P-glycoprotein. This observation suggests that the comprehension of the respective effects of topoisomerase inhibitors and the precise knowledge of their mechanisms of resistance would improve the use of this therapeutic class in the clinic within rational chemotherapeutic combinations.  相似文献   

2.
DNA topoisomerase I (topo I) is the molecular target for the camptothecin group of anticancer drugs. These drugs are showing activity against a wide array of human tumors. Many data have indicated that the sensitivity of a tumor cell to the camptothecins is dependent on tumor topo I levels. Drug-sensitive cells have high levels of topo I. Unfortunately, there is still a relative lack of information on topo I levels in human malignancies. Because of this, we investigated topo I activity and immunoprotein levels in a variety of normal murine and human tissues, as well as tissues obtained from several carcinomas, lymphomas, and sarcomas. Flow cytometric analysis was also performed on the neoplastic specimens to determine the percentage of cycling cells. Topo I catalytic activity was detected in all normal tissues at a fairly constant level. The average topo I catalytic activity in normal mammalian tissues was 2.7 +/- 1.3 x 10(4) units/mg protein (range 1.1 to 5.0 x 10(4)). Topo I catalytic activity was much more variable in human malignancies and ranged from a low of 1.4 x 10(4) units/mg protein in a rhabdomyosarcoma to a high of 160 x 10(4) units/mg protein in a poorly differentiated ovarian carcinoma. Western blot analysis with either a mouse monoclonal antibody or scleroderma antibodies directed against topo I revealed that the elevated topo I catalytic activity levels in the malignant tissues are due to elevated amounts of topo I immunoprotein. It is possible that the high topo I levels that characterize several different types of human malignancies might indicate that these tumors would be sensitive to many of the new drugs that target topo I.  相似文献   

3.
The budding yeast Saccharomyces cerevisiae has been exploited to investigate the cytotoxic mechanisms of drugs that target DNA topoisomerases. This model organism has been used to establish eukaryotic DNA topoisomerase I or II as the cellular target of specific antineoplastic agents, to define mutations in these enzymes that confer drug resistance and to elucidate the cellular factors that modulate cell sensitivity to DNA topoisomerase-targeted drugs. These findings have provided valuable insights into the critical activities of these enzymes and how perturbing their functions produces DNA damage and cell death.  相似文献   

4.
New anticancer drugs that target DNA topoisomerase I (topo I) are showing activity against a wide variety of solid human neoplasms. These drugs work by a novel mechanism of action and cause topo I-mediated DNA breaks. These DNA breaks become lethal in cycling cells when they interact with the replication fork. Because of the challenges in treating metastatic malignant melanoma, we performed an immunohistochemical study of this group of neoplasms to search for the presence of molecular markers that might indicate tumor response to topo I active drugs. Using a new immunohistochemical stain for topo I, we found elevation of this protein in 10 of 24 cases (41.6%) of metastatic malignant melanoma. The metastatic tumors that showed increased expression of topo I (2+ or 3+) had statistically significant higher proliferation indices, measured by immunohistochemical staining for DNA topo II-alpha, than did metastatic lesions with no detectable topo I expression. The average topo II-alpha index of metastatic melanomas with 2+ topo I expression was 45.1 (SD = 17.9) and with 3+ topo I expression was 52.3 (SD = 32.5). These values were found to be statistically different (P = .05) than the average topo II-alpha index of 18.9 (SD = 17.7) found for metastatic melanomas without detectable topo I immunostaining. Immunohistochemical staining for p53 suggested abnormal p53 function in 6 of the 10 melanomas (60%), which showed elevations of topo I (2 to 3+ topo I immunostaining) but normal p53 function in all 14 metastatic lesions that showed normal topo I expression.  相似文献   

5.
The topoisomerase II alpha (topo II alpha) enzyme is the target for several chemotherapeutic agents, including etoposide, teniposide, mitoxantrone, and doxorubicin (topo II poisons). The enzyme also is a marker of cell proliferation. Most cases of Hodgkin's disease (HD) are responsive to combination chemotherapy regimes that include topo II poisons such as doxorubicin. Immunoperoxidase methods for detection of the topo II alpha isoenzyme are now available for use in formalin-fixed, paraffin-embedded tissues, which may provide information about the proliferative capacity and possible sensitivity of tumors to drugs that target topo II. We used a specific antibody to analyze subsets of HD for topo II alpha staining patterns. Formalin-fixed blocks from 49 cases of HD, including 20 nodular sclerosis (NS), 14 mixed-cellularity (MC), and 15 lymphocyte-predominant (LP) subtypes, were analyzed by dual staining for topo II in combination with monoclonal antibodies against Reed-Sternberg (RS) cells consisting of CD15 for the NS and MC subtypes and CD20 for LP lymphocytic and histiocytic (L & H) cells. The number of morphologically appropriate cells coexpressing the RS or L & H marker and topo II alpha was quantitated. Positive nuclear staining for topo II alpha in RS or L & H cells was seen in 100% of cases, irrespective of subtype. Coexpression of CD15 and topo II alpha was seen in 58.4% of the RS cells or mononuclear variants in NSHD cases and 68.4% in MCHD cases. No significant difference in the percentage of neoplastic cells expressing topo II alpha was found between NS and MC subtypes. Cases of LPHD showed coexpression of CD20 and topo II alpha in 84.4% of the L & H cells, a significant increase over the level of tumor cell coexpression seen in NSHD and MCHD (P < .001). Only one case was found to have a low (< 25% of tumor cell coexpression) level of topo II alpha expression. Immunohistochemical detection of a high level of topo II alpha expression in HD, irrespective of subtype, suggests a molecular explanation for the excellent response of most HD to standard combination chemotherapy, which can include topo II poisons. The LP subtype has a higher expression of topo II alpha in the neoplastic cell population than do NS or MC subtypes, perhaps indicating increased sensitivity of these tumors to topo II poisons. It may be possible to identify subsets of HD that are more or less sensitive to conventional chemotherapeutic regimes, which would help in the selection of appropriate treatment.  相似文献   

6.
We have shown that both DNA topoisomerase (topo) IIalpha and beta are in vivo targets for etoposide using a new assay which directly measures topo IIalpha and beta cleavable complexes in individual cells after treatment with topo II targeting drugs. CCRF-CEM human leukemic cells were exposed to etoposide for 2 hr, then embedded in agarose on microscope slides before cell lysis. DNA from each cell remained trapped in the agarose and covalently bound topo II molecules from drug-stabilized cleavable complexes remained associated with the DNA. The covalently bound topo II was detected in situ by immunofluorescence. Isoform-specific covalent complexes were detected with antisera specific for either the alpha or beta isoform of topo II followed by a fluorescein isothiocyanate-conjugated second antibody. DNA was detected using the fluorescent stain Hoechst 33258. A cooled slow scan charged coupled device camera was used to capture images. A dose-dependent increase in green immunofluorescence was observed when using antisera to either the alpha or beta isoforms of topo II, indicating that both isoforms are targets for etoposide. We have called this the TARDIS method, for trapped in agarose DNA immunostaining. Two key advantages of the TARDIS method are that it is isoform-specific and that it requires small numbers of cells, making it suitable for analysis of samples from patients being treated with topo II-targeting drugs. The isoform specificity will enable us to extend our understanding of the mechanism of interaction between topo II-targeting agents and their target, the two human isoforms.  相似文献   

7.
Aclarubicin and doxorubicin are DNA binding anthracycline antibiotics of related chemical structure but differing cytotoxic action. Although doxorubicin mediates its cytotoxicity by poisoning the enzyme topoisomerase II, aclarubicin has been hypothesized to inhibit the catalytic action of topoisomerase II. We show here that aclarubicin, in contrast to doxorubicin, is highly effective in inhibiting the action of topoisomerase I. Aclarubicin not only inhibits this enzyme in a cell-free assay but also markedly inhibits DNA-protein cross-linking in H460 human lung adenocarcinoma cells as measured by the K(+)-SDS precipitation technique. It also displaces topoisomerase I from DNA as measured by Western blotting. Aclarubicin reverses the cytotoxicity of both amsacrine and camptothecin in clonogenic survival assays, consistent with the hypothesis that it is a dual topoisomerase I/II inhibitor. We suggest that the self-inhibition of topoisomerase I in short-term assays may mask the underlying activity of aclarubicin as a topoisomerase I poison. In short-term (1-H) drug exposure assays, aclarubicin kills both exponential and plateau phase cells by a non-cell cycle-selective mechanism apparently not involving G2 phase arrest. This may be a consequence of simultaneous inhibition of topoisomerases I and II.  相似文献   

8.
The nuclear enzyme DNA topoisomerase II (topo II) is the target of important antitumor agents such as etoposide. Recent work has classified topo II targeting drugs into either topo II poisons that act by stabilizing enzyme-DNA cleavable complexes leading to DNA breaks, or topo II catalytic inhibitors that act at stages in the catalytic cycle of the enzyme where both DNA strands are intact and, therefore, do not cause DNA breaks. Accordingly, catalytic inhibitors are known to abrogate DNA damage and cytotoxicity caused by topo II poisons. In this commentary, we have focused on the possibilities of enabling high-dose therapy with the topo II poison etoposide by protection of normal tissue with catalytic inhibitors, analogous to folinic acid rescue in high-dose methotrexate treatment. Thus, we have demonstrated recently that (+)-1,2-bis(3,5-dioxopiperazinyl-1-yl)propane (ICRF-187) enabled a 3- to 4-fold dose escalation of etoposide in mice. Two high-dose etoposide models are described, namely use of the weak base chloroquine in tumors with acidic extracellular pH and targeting of CNS tumors with protection of normal tissue by the bisdioxopiperazine ICRF-187. In conclusion, high supralethal doses of topo II poisons in combination with catalytic inhibitor protection form a new strategy to improve the antitumor selectivity of etoposide and other topo II poisons. Such an approach may be used to overcome problems with drug resistance and drug penetration.  相似文献   

9.
Previously, we have observed a strong restriction of the Moloney murine leukemia virus (MoMLV) replicative cycle in a cell line displaying resistance to topoisomerase II (topo II)-interactive drugs. Resistance towards these antitumoral inhibitors was associated with decreased expression and activity of topo II, suggesting that such a decrease may be responsible for MoMLV restriction. To more specifically assess the role of topo II during the retroviral cycle, we have used the antisense strategy to obtain a selective decrease of cellular topo II expression. The RNA antisense was isolated from a retroviral library expressing random fragments of human topo II (alpha form). This system allowed us to investigate the HIV-1 replicative cycle in two related human CEM cell lines expressing different levels of topo II. Expression of the enzyme is decreased four- to sixfold following formation of a sense-antisense RNA hybrid. Repression of the topo II enzyme results in an impairment of the HIV-1 replicative cycle. Using the polymerase chain reaction, we showed that the number of integration events was decreased in cells repressing the enzyme, although viral DNA synthesis and circularization were equivalent to those in the parent cells.  相似文献   

10.
To examine the role of human DNA topoisomerase IIalpha (topo IIalpha) in drug resistance, we selectively inhibited topo IIalpha gene expression in U937 human monocytic leukaemia cells stably transfected with a plasmid that allowed for Zn-mediated conditional expression of a human alpha-topo IIalpha antisense sequence. Expression of topo IIalpha mRNA was reduced to <30%, whereas no significant alteration of topo IIbeta mRNA expression was observed. Under these conditions, drug sensitivity to the topo-II-directed agents, etoposide and daunorubicin, was reduced to approximately 50%, whereas sensitivity to 4-hydroperoxy-cyclophosphamide (4-HC) was not altered. This suggests that a reduced amount of topo IIalpha mRNA may be sufficient for the resistance to topo II inhibitors in leukaemia cells.  相似文献   

11.
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   

12.
The in vitro inhibitory effects of sitafloxacin (DU-6859a) and its three stereoisomers on bacterial DNA gyrase from Escherichia coli, topoisomerase IV from Staphylococcus aureus, and topoisomerase II from human placenta were compared. No correlation was observed between the inhibitory activities of quinolones against bacterial type II topoisomerases and those against human topoisomerase II. Sitafloxacin showed the most potent inhibitory activities against bacterial type II topoisomerases and the lowest activity against human type II topoisomerase.  相似文献   

13.
The mechanisms of action of intoplicine (RP-60475), a 7H-benzo[e]pyrido[4,3-b]indole derivative that is presently in early clinical trials, have been investigated. Intoplicine induced both topoisomerase I- and II-mediated DNA strand breaks, using purified topoisomerases. The topoisomerase cleavage site patterns induced by intoplicine were unique, relative to those of camptothecin, 4'-(9-acridinylamino)methanesulfon-m-anisidide (m-AMSA), and other known topoisomerase inhibitors. Both topoisomerase I- and II-induced DNA breaks decreased at drug concentrations higher than 1 microM, which is consistent with the DNA-intercalating activity of intoplicine. DNA damage was investigated in KB cells in culture by using alkaline elution. Intoplicine induced single-strand breaks (SSB) in a bell-shaped manner with respect to drug concentration (maximum frequency at 1 microM approximately 220 rad-equivalents). SSB formation was fast, whereas reversal after drug removal was slow. Similar bell-shaped curves were obtained for DNA double-strand breaks (DSB) and DNA-protein cross-links. SSB and DNA-protein cross-link frequencies were approximately equal, and no protein-free breaks were detectable, indicating the protein concealment of the breaks, as expected for topoisomerase inhibition. Comparison of SSB and DSB frequencies indicated that intoplicine produced a significant amount of SSB not related to DSB, which is consistent with concomitant inhibition of both DNA topoisomerases I and II in cells. Data derived from resistant cell lines indicated that multidrug-resistant cells were cross-resistant to intoplicine but that m-AMSA- and camptothecin-resistant cells were sensitive to intoplicine. Hence, intoplicine might circumvent topoisomerase I-mediated and topoisomerase II-mediated resistance by poisoning both enzymes simultaneously.  相似文献   

14.
Quinolones are potent broad spectrum antibacterial drugs that target the bacterial type II DNA topoisomerases. Their cytotoxicity derives from their ability to shift the cleavage-religation equilibrium required for topoisomerase action toward cleavage, thereby effectively trapping the enzyme on the DNA. It has been proposed that these drugs act by binding to the enzyme-DNA complex. Using catalytically inactive and quinolone-resistant mutant topoisomerase IV proteins, nitrocellulose filter DNA binding assays, and KMnO4 probing of drug-DNA and drug-DNA-enzyme complexes, we show: (i) that norfloxacin binding to DNA induces a structural alteration, which probably corresponds to an unwinding of the helix, that is exacerbated by binding of the topoisomerase and by binding of the drug to the enzyme and (ii) that formation of this structural perturbation in the DNA precedes DNA cleavage by the topoisomerase in the ternary complex. We conclude that cleavage of the DNA and the resultant opening of the DNA gate during topoisomerization requires the induction of strain in the DNA that is bound to the enzyme. We suggest that quinolones may act to accelerate the rate of DNA cleavage by stimulating acquisition of this structural perturbation in the ternary complex.  相似文献   

15.
Bacterial and archeal type I topoisomerases, including topoisomerase I, topoisomerase III and reverse gyrase, have different potential roles in the control of DNA topology including regulation of supercoiling and maintenance of genetic stability. Analysis of their coding sequences in different organisms shows that they belong to the type IA family of DNA topoisomerases, but there is variability in organization of various enzymatic domains necessary for topoisomerase activity. The torus-like structure of the conserved transesterification domain with the active site tyrosine for DNA cleavage/rejoining suggests steps of enzyme conformational change driven by DNA substrate and Mg(II) cofactor binding, that are required for catalysis of change in DNA linking number.  相似文献   

16.
Beyond the known mutagenic properties of DNA lesions, recent evidence indicates that several forms of genomic damage dramatically influence the catalytic activities of DNA topoisomerases. Apurinic sites, apyrimidinic sites, base mismatches, and ultraviolet photoproducts all enhance topoisomerase I-mediated DNA cleavage when they are located in close proximity to the point of scission. Furthermore, when located between the points of scission of a topoisomerase II cleavage site, these same lesions (with the exception of ultraviolet photoproducts) greatly stimulate the cleavage activity of the type II enzyme. Thus, as found for anticancer drugs, lesions have the capacity to convert topoisomerases from essential cellular enzymes to potent DNA toxins. These findings raise exciting new questions regarding the mechanism of anticancer drugs, the physiological functions of topoisomerases, and the processing of DNA damage in the cell.  相似文献   

17.
DNA methylation is deregulated during oncogenesis. Since several major anti-cancer drugs act on topoisomerases, we investigated the effects of cytosine methylation on topoisomerase cleavage activities. Both topoisomerase I and II cleavage patterns were modified by CpG methylation in c-myc gene DNA fragments. Topoisomerase II changes, mainly cleavage reduction, occurred for methylation sites within 7 base pairs from the topoisomerase II breaks and were different for VM-26 and azatoxin. For topoisomerase I, cleavage enhancement as well as suppression were observed. Using synthetic methylated oligonucleotides, we show that hemimethylation is sufficient to alter topoisomerase I activity. Cytosine methylation on the scissile strand within the topoisomerase I consensus sequence had strong effects. Cleavage was stimulated by methylation at position -4 and was strongly inhibited by methylation at position -3 (with position -1 being the enzyme-linked nucleotide). This inhibitory effect was attributed to the presence of a methyl group in the major groove, since the transition uracil to thymine also inhibited cleavage. Altogether these results suggest an interaction of topoisomerase I with the DNA major grove at positions -3 and -4. In addition, DNA methylation may have profound effects on the activity of topoisomerases and may alter the distribution of cleavage sites produced by anticancer drugs in chromatin.  相似文献   

18.
The efficacy of all chemotherapeutic agents is limited by the occurrence of drug resistance. To further understand resistance to topoisomerase (topo) II inhibitors, 50 sublines were isolated as single clones from parental cells by exposure to ETP or m-AMSA. Subsequently, a population of cells from each subline was exposed to three-fold higher drug concentrations allowing 16 stable sublines to be established at higher extracellular drug concentration. The frequency and nature of mutations in topo II in the drug selected cell lines have been evaluated. In order to screen a large number of cell lines, an RNase protection assay was developed. Fragments covering the entire coding sequence of topo II was isolated after PCR amplification and subcloned in pGEM3Z vector. Using this approach, mismatches was observed in 13.6% of resistant cell lines (12% of resistant cell lines exposed to lower drug concentrations and 18.8% of resistant cell lines exposed to higher drug concentrations). Our findings suggest that mutations of topo II gene seem to be an important and frequent mechanism of resistance to topo II inhibitors.  相似文献   

19.
DNA topoisomerase (topo) II alpha is a major target for many anticancer agents. However, progress towards understanding how these agents interact with this enzyme in human cells and how resistance to these agents arises is greatly impeded by difficulties in expressing this gene. Here, we report on achieving a high level of expression of a full-length human topo II alpha gene in human cells. We started with the topo II alpha cDNA driven by a strong cytomegalovirus promoter and transiently transfected HeLa cells. Although topo II alpha mRNA was consistently detected in transfected cells, no exogenous topo II alpha protein was detected. By contrast, when the same cDNA was fused to an enhanced green fluorescent protein (EGFP), we detected a high level of expression at both mRNA and protein levels. The exogenous topo II alpha was localized to cell nuclei as expected, indicating that the fusion protein is properly folded. Furthermore, overexpression of the EGFP-topo II alpha fusion protein increased the sensitivity of the transfected cells to teniposide, suggesting that it functions as the endogenous counterpart. Thus, in addition to being used as a gene tag, the GFP fusion approach may be generally applicable for expressing genes, such as topo II alpha, that are difficult to express by conventional methods.  相似文献   

20.
We determined the inhibitory activities of gatifloxacin against Staphylococcus aureus topoisomerase IV, Escherichia coli DNA gyrase, and HeLa cell topoisomerase II and compared them with those of several quinolones. The inhibitory activities of quinolones against these type II topoisomerases significantly correlated with their antibacterial activities or cytotoxicities (correlation coefficient [r] = 0.926 for S. aureus, r = 0.972 for E. coli, and r = 0.648 for HeLa cells). Gatifloxacin possessed potent inhibitory activities against bacterial type II topoisomerases (50% inhibitory concentration [IC50] = 13.8 microg/ml for S. aureus topoisomerase IV; IC50 = 0.109 microg/ml for E. coli DNA gyrase) but the lowest activity against HeLa cell topoisomerase II (IC50 = 265 microg/ml) among the quinolones tested. There was also a significant correlation between the inhibitory activities of quinolones against S. aureus topoisomerase IV and those against E. coli DNA gyrase (r = 0.969). However, the inhibitory activity against HeLa cell topoisomerase II did not correlate with that against either bacterial enzyme. The IC50 of gatifloxacin for HeLa cell topoisomerase II was 19 and was more than 2,400 times higher than that for S. aureus topoisomerase IV and that for E. coli DNA gyrase. These ratios were higher than those for other quinolones, indicating that gatifloxacin possesses a higher selectivity for bacterial type II topoisomerases.  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