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 46 毫秒
1.
Camptothecins are a new class of anticancer drugs that target DNA topoisomerase I; current efforts are directed toward elucidating optimal combinations of these drugs with other antineoplastic agents. A rationale for the use of sequential therapy involving the combination of camptothecins with topoisomerase II-targeting drugs, such as etoposide, has arisen from observations of increased topoisomerase II protein levels in cell lines resistant to camptothecin. In an effort to understand potential mechanisms of resistance to this strategy, we developed a U-937 cell subline, denoted RERC, that is capable of surviving exposure to sequential topoisomerase poisoning. The RERC cells are 200-fold resistant to camptothecin, 8-fold resistant to etoposide, and 10-fold hypersensitive to cisplatin compared to the parental U-937 cells. Biochemical analyses indicate that the resistant phenotype involves alterations in both topoisomerase I and topoisomerase IIalpha. Topoisomerase I catalytic activity in the resistant cells is similar to that of the parental line but is resistant to camptothecin. Moreover, the resistant cells express a single mRNA species of topoisomerase I that codes for a mutation in codon 533. In addition, topoisomerase IIalpha protein levels are decreased 10-fold in the resistant line, coincident with a two-fold decrease in the expression of topoisomerase IIalpha mRNA. Collectively, these results indicate that resistance to sequential topoisomerase poisoning may involve a reduction in total cellular topoisomerase activity.  相似文献   

2.
The efficacy of all chemotherapeutic agents is limited by the occurrence of drug resistance. For etoposide (VP-16), increased expression of MDR-1 or MRP and alterations in topoisomerase IIalpha have been shown to confer tolerance. To further understand resistance to VP-16, three sublines, designated MCF-7-VP17, ZR-75B-VP13, and MDA-MB-231-VP7, were initially isolated as single clones from parental cells by exposure to VP-16. Subsequently, a population of cells from each subline was exposed to 3-fold higher drug concentrations, allowing stable sublines to be established at higher extracellular drug concentrations. Characterization of the resistant sublines demonstrates the adaptation that occurs with advancing drug concentrations during in vitro selections. Reduced topoisomerase II mRNA and protein levels were observed in the initial isolates. This reduction was accompanied by a decrease in topoisomerase II activity and cellular growth rate and was associated with 6-314-fold resistance to topoisomerase II poisons. With advancing resistance, MRP expression increased and VP-16 accumulation decreased. This adaptation allowed for partial restoration of topoisomerase II activity as a result of increased expression (MCF-7-VP17 and ZR-75B-VP13) or hyperphosphorylation (MDA-MB-231-VP7), with a resultant increase in growth rate. In MDA-MB-231-VP7 cells, hyperphosphorylation coincided with increased casein kinase II mRNA and protein levels, suggesting a role for this kinase in the acquired hyperphosphorylation. In this cell line, hyperphosphorylation mediated the increased activity despite a fall in topoisomerase IIalpha protein levels secondary to an acquired 600-bp deletion in one topoisomerase IIalpha allele, which resulted in reduced protein levels. In all three sublines, high levels of resistance were attained as a result of synergism between the reduced topoisomerase IIalpha levels and MRP overexpression. These studies demonstrate how cellular adaptation to increasing drug pressure occurs and how more than one mechanism can contribute to the resistant phenotype when increasing selecting pressure is applied. Reduced expression of topoisomerase II is sufficient to confer substantial resistance early in the selection process, with synergy from MRP overexpression helping to confer high levels of resistance.  相似文献   

3.
In this report we examine biochemical and genetic alterations in DNA topoisomerase II (topoisomerase II) in K562 cells selected for resistance in the presence of etoposide (VP-16). Previously, we have demonstrated that the 30-fold VP-16-resistant K/VP.5 cell line exhibits decreased stability of drug-induced topoisomerase II/DNA covalent complexes, requires greater ATP concentrations to stimulate VP-16-induced topoisomerase II/DNA complex formation, and contains reduced mRNA and protein levels of the M(r) 170,000 isoform of topoisomerase II, compared with parental K562 cells. K/VP.5 cells grown in the absence of VP-16 for 2 years maintained resistance to VP-16, decreased levels of topoisomerase II, and attenuated ATP stimulation of VP-16-induced topoisomerase II/DNA binding, compared with K562 cells. Sequencing of cDNA coding for two consensus ATP binding sites and the active site tyrosine in the K/VP.5 topoisomerase II gene indicated that no mutations were present in these domains. In addition, single-strand conformational polymorphism analysis of restriction fragments encompassing the entire topoisomerase II cDNA revealed no evidence of mutations in the gene for this enzyme in K/VP.5 cells. Nuclear extracts from K562 (but not K/VP.5) cells contained a heat-labile factor that potentiated VP-16-induced topoisomerase II/DNA covalent complex formation in isolated nuclei from K/VP.5 cells. Immunoprecipitated topoisomerase II from K/VP.5 cells was 2.5-fold less phosphorylated, compared with enzyme from K562 cells. Collectively, our data suggest that acquired VP-16 resistance is mediated, at least in part, by altered levels or activity of a kinase that regulates topoisomerase II phosphorylation and hence drug-induced topoisomerase II/DNA covalent complex formation and stability.  相似文献   

4.
5.
Unique functions of mammalian DNA-topoisomerases IIalpha and -beta are suggested by their distinct cellular distribution and chromatin binding at mitosis. Here, we studied H69-VP cells that, due to a homozygous mutation, express topoisomerase IIalpha mostly outside the nucleus. In these cells topoisomerase IIbeta showed a normal nuclear localization. However, at mitosis it diffused away from the chromatin despite the nuclear lack of the alpha-isoform. 80% of these cells performed chromosome condensation and disjunction with the aid of cytosolic topoisomerase IIalpha, which bound to the mitotic chromatin with low affinity. However, the genotype of these cells was highly polyploid indicating an increased rate of non-disjunction. In 20% of the mutant cells neither topoisomerase II isoform was bound to the mitotic chromatin, which appeared as an unstructured DNA spheroid unable to undergo disjunction and cytokinesis. Parental H69 cells expressing topoisomerase IIalpha inside the nucleus exhibited high affinity binding of the enzyme to the mitotic chromatin. Their genotype was mostly diploid and stable. We conclude (i) that high affinity chromatin binding of topoisomerase IIalpha is essential for chromosome condensation/disjunction and (ii) that topoisomerase IIbeta does not adopt these functions.  相似文献   

6.
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   

7.
A Chinese hamster ovary (CHO) cell line highly resistant to the non-cleavable complex-forming topoisomerase II inhibitor dexrazoxane (ICRF-187, Zinecard) was selected. The resistant cell line (DZR) was 1500-fold resistant (IC50 = 2800 vs 1.8 microM) to continuous dexrazoxane exposure. DZR cells were also cross-resistant (8- to 500-fold) to other bisdioxopiperazines (ICRF-193, ICRF-154, and ICRF-186), and somewhat cross-resistant (4- to 14-fold) to anthracyclines (daunorubicin, doxorubicin, epirubicin, and idarubicin) and etoposide (8.5-fold), but not to the other non-cleavable complex-forming topoisomerase II inhibitors suramin and merbarone. The cytotoxicity of dexrazoxane to both cell lines was unchanged in the presence of the membrane-active agent verapamil. DZR cells were 9-fold resistant to dexrazoxane-mediated inhibition of topoisomerase II DNA decatenation activity compared with CHO cells (IC50 = 400 vs 45 microM), but were only 1.4-fold (IC50 = 110 vs 83 microM) resistant to etoposide. DZR cells contained one-half the level of topoisomerase II protein compared with parental CHO cells. However, the specific activity for decatenation using nuclear extract topoisomerase II was unchanged. Etoposide (100 microM)-induced topoisomerase II-DNA complexes in DZR cells and isolated nuclei were similarly one-half the level found in CHO cells and in isolated nuclei. However, the ability of 500 microM dexrazoxane to inhibit etoposide (100 microM)-induced topoisomerase II-DNA covalent complexes was reduced 4- to 6-fold in both DZR cells and nuclei compared with CHO cells and nuclei. In contrast, there was no differential ability of aclarubicin or merbarone to inhibit etoposide-induced topoisomerase II-DNA complexes in CHO compared with DZR cells and isolated nuclei. It was concluded that the DZR cell line acquired its resistance to dexrazoxane mainly through an alteration in the topoisomerase II target.  相似文献   

8.
Etoposides block cell division by interfering with the action of topoisomerase II, leaving enzyme-DNA double-strand breaks. We found that certain components of the trimeric DNA-dependent protein kinase influence cell survival following etoposide damage. Interestingly, either Ku70- or Ku80-deficient cell lines, but not mutant cell lines of the DNA-PK catalytic sub-unit (DNA-PKcs), were found to be hypersensitive to the effects of etoposide VP16. Ku70- and Ku80-deficient cells can be complemented to an etoposide resistant phenotype by introducing wildtype Ku70 or Ku80 cDNAs. Mutational analysis of introduced Ku70 cDNAs into murine embryonic stem cells deleted for Ku70 (-/-) showed that mutants where heterodimerization and DNA binding functions of Ku were disrupted, also blocked the restoration of etoposide resistance. In contrast with the differential etoposide sensitivity of DNA-PK mutants, both Ku- and DNA-PKcs-deficient cell lines showed G2 ionizing radiation-induced delays, a cell cycle phase where topoisomerase II function is critical. Thus, the topoisomerase II cleaved complexes may be an example of DNA lesions requiring the Ku heterodimer, but not DNA-PK for DNA repair.  相似文献   

9.
DNA topoisomerase IIalpha was monitored with the monoclonal antibody Ki-S1 in human fibroblasts after irradiation of cells with gamma rays from a 137Cs source or treatment with the DNA topoisomerase II inhibitor doxorubicin. DNA topoisomerase IIalpha was localized immunohistochemically as bright fluorescent dots in the karyoplasm. The fibroblasts investigated originated from normal human donors and a xeroderma pigmentosum (XP) patient (XP12BE). All cell lines examined showed a time- and dose-dependent increase in DNA topoisomerase IIalpha abundance after irradiation or treatment with doxorubicin. No principal difference in response was seen between normal and XP fibroblasts towards either treatment alone. After irradiation with 9 Gy, the effect was detectable after as little as 30 min and lasted for at least 6 h. After doxorubicin treatment, topoisomerase II overexpression occurred within less than 2 h. It passed through a maximum and began to decrease after approximately 6 h. In principle, the increase in DNA topoisomerase IIalpha may result from (i) architectural changes of interphase chromatin leading to enhanced accessibility of preformed enzyme to the antibody, (ii) enhanced gene expression, or (iii) enhanced stabilization of mRNA or protein molecules. The increase in enzyme levels may be part of the well-known DNA damage responses that operate in cell-protective or DNA-reparative pathways. Thus, the action of DNA topoisomerase II would serve to catalyze preparatory steps in DNA repair. We also found overexpression of the Bax protein and p16 predominantly in treated XP cells, suggesting that the DNA-damaging protocols elicited signals for apoptosis and cell-cycle arrest. From the simultaneous increase in DNA topoisomerase IIalpha and Bax, one may conclude that DNA topoisomerase IIalpha also plays role in apoptosis.  相似文献   

10.
In mammalian cells, DNA topoisomerase II is the product of two distinct genes encoding the alpha and beta isoforms of the enzyme. Besides homodimeric topoisomerase IIalpha and IIbeta, we have recently shown that alpha/beta heterodimers constitute a third population of topoisomerase II in HeLa cells. We found that topoisomerase II heterodimers are not restricted to HeLa cells but exist in different mammalian cell types, and up to 25% of the total topoisomerase IIbeta population is involved in heterodimer formation. Studies of topoisomerase II phosphorylation in HeLa cells show that heterodimers are phosphorylated in vivo to a significantly lower level compared to homodimeric alpha enzymes, but in contrast to the latter neither heterodimers nor topoisomerase IIbeta homodimers coprecipitate together with a kinase activity that is able to mediate their phosphorylation. However, both enzymes can still be phosphorylated by exogenously added casein kinase II. The differential phosphorylation of topoisomerase II heterodimers suggests an alternative regulation of this topoisomerase II subclass compared to the homodimeric topoisomerase IIalpha counterparts.  相似文献   

11.
Anthracyclines are among the most clinically useful topoisomerase II poisons. A complete understanding of their molecular mechanism is thus fundamental for a rational design of novel agents. We evaluated four anthracycline analogues with respect to human topoisomerase IIalpha-dependent DNA cleaving activity, efficiency in killing yeast cells, and uptake and retention in yeast and compared the yeast system to tumor cell line models. The yeast JN394top2-4 strain was used because it has a topoisomerase II ts gene mutation: enzyme activity is much less at 30 degrees C than at 25 degrees C and is completely lost at 35 degrees C. Untransformed JN394top2-4 cells were 33-fold more sensitive to idarubicin at 25 degrees C than at 30 degrees C, showing that topoisomerase II is the primary drug target. Overexpression of human topoisomerase IIalpha was toxic to yeast cells when the yeast enzyme was inactivated. Drug-dependent killing of yeast cells expressing low levels of the human alpha isoenzyme at 35 degrees C showed that the analogues spanned a 3-log range of cytotoxic potency in yeast, as they did in tumor cells. However, the compounds were much less active against the yeast strain than mammalian tumor cell lines. Drug uptake was determined and found to be altered in yeast with respect to tumor cells. Although DNA cleavage stimulated by anthracyclines roughly correlated with cytotoxicity, the cleavage level:cytotoxicity ratios were different for the studied drugs. Thus, the results suggest that other drug-dependent molecular factors contribute to drug activity in addition to the cellular content of topoisomerase IIalpha and drug uptake.  相似文献   

12.
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   

13.
Drug resistance to anti-tumour agents often coincides with mutations in the gene encoding DNA topoisomerase II alpha. To examine how inactive forms of topoisomerase II can influence resistance to the chemotherapeutic agent VP-16 (etoposide) in the presence of a wild-type allele, we have expressed point mutations and carboxy-terminal truncations of yeast topoisomerase II from a plasmid in budding yeast. Truncations that terminate the coding region of topoisomerase II at amino acid (aa) 750, aa 951 and aa 1044 are localised to both the cytosol and the nucleus and fail to complement a temperature-sensitive top2-1 allele at non-permissive temperature. In contrast, the plasmid-borne wild-type TOP2 allele and a truncation at aa 1236 are nuclear localised and complement the top2-1 mutation. At low levels of expression, truncated forms of topoisomerase II render yeast resistant to levels of etoposide 2- and 3-fold above that tolerated by cells expressing the full-length enzyme. Maximal resistance is conferred by the full-length enzyme carrying a mutated active site (Y783F) or a truncation at aa 1044. The level of phosphorylation of topoisomerase II was previously shown to correlate with drug resistance in cultured cells, hence we tested mutants in the major casein kinase II acceptor sites in the C-terminal domain of yeast topoisomerase II for changes in drug sensitivity. Neither ectopic expression of the C-terminal domain alone nor phosphoacceptor site mutants significantly alter the host cell's sensitivity to etoposide.  相似文献   

14.
Merbarone is a catalytic inhibitor of topoisomerase II that is in clinical trials as an anticancer agent. Despite the potential therapeutic value of this drug, the mechanism by which it blocks topoisomerase II activity has not been delineated. Therefore, to determine the mechanistic basis for the inhibitory action of merbarone, the effects of this drug on individual steps of the catalytic cycle of human topoisomerase IIalpha were assessed. Concentrations of merbarone that inhibited catalytic activity >/=80% had no effect on either enzyme.DNA binding or ATP hydrolysis. In contrast, the drug was a potent inhibitor of enzyme-mediated DNA scission (in the absence or presence of ATP), and the inhibitory profiles of merbarone for DNA cleavage and relaxation were similar. These data indicate that merbarone acts primarily by blocking topoisomerase II-mediated DNA cleavage. Merbarone inhibited DNA scission in a global (rather than site-specific) fashion but did not appear to intercalate into DNA or bind in the minor groove. Since the drug competed with etoposide (a cleavage-enhancing agent that binds directly to topoisomerase II), it is proposed that merbarone exerts its inhibitory effects through interactions with the enzyme and that the drug shares an interaction domain on topoisomerase II with cleavage-enhancing agents.  相似文献   

15.
Five cell lines selected for resistance to the cytotoxicity of inhibitors of DNA topoisomerase II have point mutations in the gene that codes for the M(r) 170,000 form of this enzyme. In each case, the mutation results in an amino acid change in or near an ATP binding sequence of the M(r) 170,000 isozyme of topoisomerase II. We used single-strand conformational polymorphism analysis to screen for similar mutations in other drug-resistant cell lines or in leukemic cells from patients previously treated with etoposide or teniposide. We also analyzed the region of the gene that codes for amino acids adjacent to the tyrosine at position 804 of topoisomerase II which binds covalently to DNA. CEM/VM-1, CEM/VM-1-5, and HL-60/AMSA human leukemic cell lines were used as controls; 3 of 3 known mutations were detected by migration differences of polymerase chain reaction products from the RNA extracted from these three lines. A previously unknown mutation was found in the tyrosine 804 region of the M(r) 170,000 topoisomerase II expressed by CEM/VM-1 and CEM/VM-1-5 cells. Sequence analysis showed that substitution of a T for a C at nucleotide 2404 resulted in an amino acid change of a serine for a proline at amino acid 802. No mutations in any of the ATP binding sequences or in the tyrosine 804 region were detected in polymerase chain reaction products from RNA extracted from human leukemia HL-60/MX2 or CEM/MX1 cells (both cell lines selected for resistance to mitoxantrone) or in human myeloma 8226/Dox1V cells (selected for resistance by simultaneous exposure to doxorubicin and verapamil). No mutations were detected in polymerase chain reaction products from RNA extracted from blasts of 15 patients with relapsed acute lymphocytic leukemia, previously treated with etoposide or teniposide. We conclude that: (a) single-strand conformational polymorphism analysis is useful for screening for mutations in topoisomerase II; (b) resistance to the cytotoxicity of inhibitors of DNA topoisomerase II is not always associated with mutations in ATP binding sequences or the active site tyrosine region of M(r) 170,000 topoisomerase II; and (c) mutations similar to those detected in drug resistant cells selected in culture have not been identified in blast cells from patients with relapsed acute lymphocytic leukemia, previously treated with etoposide or teniposide.  相似文献   

16.
Increased expression of DNA topoisomerase II alpha has been associated with resistance to certain DNA-damaging alkylating agents, but no causal relationship or mechanism has been established. To investigate this observation, we developed a model of topoisomerase II overexpression by transfecting a full-length Chinese hamster ovary topoisomerase II alpha into EMT6 mouse mammary carcinoma. Topoisomerase II alpha-transfected cell lines demonstrated continued topoisomerase II alpha mRNA and protein expression, which were undetectable in vector-only lines, in stationary phase (G0-G1). The topoisomerase II transfectants were approximately 5-10-fold resistant to the alkylating agents cisplatin and mechlorethamine. Upon release from G0-G1, the topoisomerase II transfectants demonstrated more rapid thymidine incorporation and shorter cell-doubling times than control cells. Purified topoisomerase II and nuclear extracts with topoisomerase II-decatenating activity bound to cisplatin-treated DNA with significantly greater affinity than to untreated DNA in a cisplatin concentration-dependent manner. These observations suggest that expression of topoisomerase II alpha may have a role in cellular resistance to antineoplastic alkylating agents. The mechanism for this may involve increased binding of topoisomerase II alpha to alkylating agent-damaged DNA.  相似文献   

17.
PURPOSE: Often the best method of integrating chemotherapeutic agents is unknown. Recently there has been interest in the use of combinations of the topoisomerase II inhibitors and the topoisomerase I inhibitors as these agents have shown individual activity in malignancies such as non-small-cell lung cancer. This study examined the interaction of the topoisomerase II inhibitor etoposide with the topoisomerase I inhibitor topotecan (Tpt) in V79 cells (hamster lung fibroblast cells) to determine the optimal method of delivering these agents. METHODS AND RESULTS: Cell survival was assessed by colony formation. Synergistic interactions were assessed by the median effect principle in which a combination index (CI) of less than one suggests a synergistic interaction. The V79 cells were exposed to sequential 24-h incubations with the two chemotherapeutic agents. Initially, equitoxic doses of the two agents were delivered (i.e. 0.0275 microg/ml of topotecan alone or 0.089 microg/ml of etoposide alone resulting in a surviving fraction of 70%; Tpt:etoposide ratio 1: 3.2). It was determined that a sequence-dependent synergistic interaction (CI < 1) resulted at a lower level of cytotoxicity if the etoposide exposure followed the Tpt exposure compared to the opposite sequence. This same effect was seen after treatment of cells with various concentration (microg/ml) ratios of Tpt: etoposide (1:4.0, 1:1, 2.5:1). CONCLUSIONS: These results suggest that maximum synergy occurs for the delivery of etoposide following Tpt exposure (compared to the opposite sequence) and these findings may have important clinical implications.  相似文献   

18.
The novel anticancer glucosyl derivative of indolo-carbazole (NB-506), an inhibitor of DNA topoisomerase I, exhibited strong in vitro cytotoxicity against various human cancer cell lines. In order to elucidate its cytotoxic mechanisms, we established nine NB-506-resistant sublines with different resistance ratios from human small cell lung cancer cells (SBC-3/P) by stepwise and brief exposure (24 h) to NB-506. Among them, SBC-3/NB#9 was 454 times more resistant to NB-506 than the parent cell line. The SBC-3/NB#9 cells showed cross-resistance only to topoisomerase I inhibitors, such as 11,7-ethyl-10-[4-(1-piperidino)-1-piperidino] carbonyloxycamptothecia and 7-ethyl-10-hydroxy-camptothecin, and not to other anticancer drugs, such as vincristine, vinblastine, Adriamycin, etoposide, and teniposide. These results indicate that the difference on the effect of topoisomerase I was considered to be related to a resistance mechanism. The topoisomerase I activities of nuclear extracts eluted from SBC-3/NB#9 cells was only one-tenth of the parent cell activity. A Western blotting study indicated that this lower activity was due to a lower amount of DNA topoisomerase I. Furthermore, we found correlations between topoisomerase I activity and sensitivity to NB-506 in sublines with different degrees of resistance. Accumulation of 3H-labeled NB-506 by SBC-3/NB#9 cells was only one-fifth of that by the parent cells, whereas intracellular accumulation of 3H-labeled camptothecin by both cell lines did not differ. The reduction of accumulation was specific to NB-506, and this result may explain why the resistance ratio for NB-506 was higher than those for 11,7-ethyl-10-[4-(1-piperidino)-1-piperidino] carbonyloxycamptothecin and 7-ethyl-10-hydroxy-camptothecin.  相似文献   

19.
The functional domain structure of human DNA topoisomerase IIalpha and Saccharomyces cerevisiae DNA topoisomerase II was studied by investigating the abilities of insertion and deletion mutant enzymes to support mitotic growth and catalyze transitions in DNA topology in vitro. Alignment of the human topoisomerase IIalpha and S. cerevisiae topoisomerase II sequences defined 13 conserved regions separated by less conserved or differently spaced sequences. The spatial tolerance of the spacer regions was addressed by insertion of linkers. The importance of the conserved regions was assessed through deletion of individual domains. We found that the exact spacing between most of the conserved domains is noncritical, as insertions in the spacer regions were tolerated with no influence on complementation ability. All conserved domains, however, are essential for sustained mitotic growth of S. cerevisiae and for enzymatic activity in vitro. A series of topoisomerase II carboxy-terminal truncations were investigated with respect to the ability to support viability, cellular localization, and enzymatic properties. The analysis showed that the divergent carboxy-terminal region of human topoisomerase IIalpha is dispensable for catalytic activity but contains elements that specifically locate the protein to the nucleus.  相似文献   

20.
KB/7D cells represent a multidrug-resistant subclone of human nasopharyngeal carcinoma KB cells generated by continuous exposure to the topoisomerase II inhibitor VP-16 (etoposide). KB/7D cells also show cross-resistance to doxorubicin and vincristine. Phenotypic traits of the cell line include a 2-fold decrease in topoisomerase II levels and a decrease in the uptake of VP-16 without an increase in the rate of drug efflux or expression of P-glycoprotein, suggesting a novel mechanism associated with the uptake of anticancer drugs. This study demonstrated that the multidrug-resistance associated protein (MRP) is overexpressed in KB/7D cells, and that the loss of resistance in revertant cells correlates with the loss of MRP. The resistance to VP-16 and doxorubicin could be overcome, partially, and resistance to vincristine could be overcome completely, by the L-enantiomer of verapamil, but not by the D-enantiomer or by BIBW 22 (4-[N-(2-hydroxy-2-methyl-propyl)-ethanolamino]-2,7-bis[cis-2,6-++ +dimethylmorpholino)-6-phenylpteridin), an inhibitor of MDR-1. L-Verapamil was shown to be significantly more potent than D-verapamil in modulating the accumulation defect in KB/7D cells towards doxorubicin, as measured by flow cytometry and confocal microscopy, and towards VP-16, as measured by increases in protein-linked DNA strand breaks. This suggests that KB/7D cells are multidrug resistant due to decreases in topoisomerase II levels and the overexpression of MRP, that MRP leads to a decrease in drug accumulation, and that L-verapamil can modulate the MRP-associated accumulation defect and drug-resistance phenotype. This contrasts with previous studies that suggest that MRP causes multidrug resistance by exporting cytotoxic drugs out of the cell and that did not show modulation of MRP by verapamil.  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