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.
Alterations in the amino acid composition, phosphorylation pattern, or intracellular levels of topoisomerase II have been associated with resistance to antineoplastic agents whose effects are mediated through interactions with this enzyme. To develop a model system with which to investigate the determinants of topoisomerase II sensitivity or resistance to antineoplastic agents that target this enzyme, a cDNA encoding the wild-type Drosophila melanogaster topoisomerase II was ligated into a mammalian expression vector containing a glucocorticoid-inducible mouse mammary tumor virus promoter and transfected into an epipodophyllotoxin-resistant Chinese hamster ovary cell line (VPM(r)-5). In two transfectants carrying an intact, full-length Drosophila topoisomerase II cDNA, exposure to the inducing agent, dexamethasone (10 microM), resulted in complementation of the endogenous mutant topoisomerase II and phenotypic reversion to etoposide sensitivity. In the presence of glucocorticoid, etoposide-induced cytotoxicity increased 20-fold, despite the fact that Drosophila topoisomerase II mRNA expression was only 0.1% of that of the endogenous mammalian topoisomerase II. Induced cells demonstrated a marked increase in DNA single strand breaks compared with uninduced resistant cells, thereby providing biochemical evidence supporting increased DNA strand cleavage due to activation of the Drosophila enzyme. These observations demonstrate the ability of a wild-type Drosophila topoisomerase II to complement a mutant mammalian enzyme and suggest that transfectants capable of conditional topoisomerase II expression represent a useful model for studies of the biochemical pharmacology and structure-function relationships of normal and mutant enzymes.  相似文献   

2.
DNA topoisomerase VI from the hyperthermophilic archaeon Sulfolobus shibatae is the prototype of a novel family of type II DNA topoisomerases that share little sequence similarity with other type II enzymes, including bacterial and eukaryal type II DNA topoisomerases and archaeal DNA gyrases. DNA topoisomerase VI relaxes both negatively and positively supercoiled DNA in the presence of ATP and has no DNA supercoiling activity. The native enzyme is a heterotetramer composed of two subunits, A and B, with apparent molecular masses of 47 and 60 kDa, respectively. Here wereport the overexpression in Escherichia coli and the purification of each subunit. The A subunit exhibits clusters of arginines encoded by rare codons in E.coli . The expression of this protein thus requires the co-expression of the minor E.coli arginyl tRNA which reads AGG and AGA codons. The A subunit expressed in E.coli was obtained from inclusion bodies after denaturation and renaturation. The B subunit was overexpressed in E.coli and purified in soluble form. When purified B subunit was added to the renatured A subunit, ATP-dependent relaxation and decatenation activities of the hyperthermophilic DNA topoisomerase were reconstituted. The reconstituted recombinant enzyme exhibits a specific activity similar to the enzyme purified from S.shibatae . It catalyzes transient double-strand cleavage of DNA and becomes covalently attached to the ends of the cleaved DNA. This cleavage is detected only in the presence of both subunits and in the presence of ATP or its non-hydrolyzable analog AMPPNP.  相似文献   

3.
A novel, heat-resistant and Pronase-sensitive, inhibitor of eukaryotic DNA topoisomerase I has been purified from Xenopus laevis ovaries. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) of the most purified fraction revealed three bands with apparent molecular masses of 25, 28.5, and 33.5 kDa. The 25- and 33.5-kDa peptides recovered from an SDS-PAGE gel inhibited X. laevis DNA topoisomerase I. The purified inhibitor was specific to DNA topoisomerase I and did not inhibit other DNA enzymes tested. The inhibitor blocked the catalytic activity of DNA topoisomerase I by interacting with the enzyme, rather than by competing for binding sites on substrate DNA. Binding of DNA topoisomerase I to substrate DNA was blocked by the inhibitor, as was the cleavage reaction catalyzed by DNA topoisomerase I. Inhibition of DNA topoisomerase I was relieved by divalent cations Ca2+, Mg2+, or Mn2+.  相似文献   

4.
Studies of yeast DNA topoisomerase II with various alanine-substitution mutations provide strong biochemical support of a recent hypothesis that the type IA and IIA DNA topoisomerases act similarly in their cleavage and rejoining of DNA. DNA breakage and rejoining by either a type IA or a type IIA enzyme are shown to involve cooperation between a DNA-binding domain containing the active-site tyrosine and a Rossmann fold containing several highly conserved acidic residues. For a homodimeric type IIA enzyme, cooperation occurs in trans: the active-site tyrosine in the DNA-binding domain of one protomer cooperates with several residues in the Rossmann fold as well as other regions of the other protomer.  相似文献   

5.
Site-directed mutagenesis was carried out at 10 highly conserved polar residues within the C-terminal half of yeast DNA topoisomerase II, which corresponds to the A subunit of bacterial DNA gyrase, to identify amino acid side chains that augment the active site tyrosine Tyr-782 in the breakage and rejoining of DNA strands. Complementation tests show that alanine substitution at Arg-690, Asp-697, Lys-700, Arg-704, or Arg-781, but not at His-735, His-736, Glu-738, Gln-750, or Asn-828, inactivates the enzyme in vivo. Measurements of DNA relaxation and cleavage by purified mutant enzymes show that these activities are abolished in the R690A mutant and are much reduced in the mutants D697A, K700A, R704A, and R781A. When a Y782F polypeptide with a phenylalanine substituting for the active site tyrosine was expressed in cells that also express the R690A polypeptide, the resulting heterodimeric yeast DNA topoisomerase II was found to nick plasmid DNA. Thus in a dimeric wild-type enzyme, Tyr-782 in one protomer and Arg-690 in the other cooperate in trans in the catalysis of DNA cleavage. For the residues D697A, K700A, R704A, and R781A, their locations in the crystal structures of type II DNA topoisomerase fragments suggest that Arg-781 and Lys-700 might be involved in anchoring the 5' and 3' sides of the broken DNA, respectively, and the roles of Asp-697 and Arg-704 are probably less direct.  相似文献   

6.
Several experimental data support the notion that the recognition of DNA crossovers play an important role in the multiple functions of topoisomerase II. Here, a theoretical analysis of the possible modes of assembly of yeast topoisomerase II with right and left-handed tight DNA crossovers is performed, using the crystal coordinates of the docking partners. The DNA crossovers are assumed to be clamped into the central hole of the enzyme. Taking into account the rules for building symmetric ternary complexes and the structural constraints imposed by DNA-DNA and protein-DNA interactions, this analysis shows that two geometric solutions could exist, depending on the chirality of the DNA crossovers. In the first one, the two DNA segments are symmetrically recognized by the enzyme while each single double helix binds asymmetrically the protein dimer. In the second one, each double helix is symmetrically recognized by the protein around its dyad axis, while the two DNA segments have their own binding modes. The finding of potential DNA-binding domains which could interact with the crossovers provides structural supports for each model. The structural similarity of a loop containing a cluster of conserved basic residues pointing into the central hole of topoisomerase II and the second DNA-binding site of histone H5 which binds DNA crossover is of particular interest. Each solution, which is consistent with different sets of experimental data found in the literature, could either correspond to different functions of the enzyme or different steps of the reaction. This work provides structural insights for better understanding the role of chirality and symmetry in topoisomerase II-DNA crossover recognition, suggests testable experiments to further elucidate the structure of ternary complexes, and raises new questions about the relationships between the mechanism of strand-passage and strand-exchange catalyzed by the enzyme.  相似文献   

7.
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   

8.
The interaction of topoisomerase II with its DNA cleavage site is critical to the physiological functions of the enzyme. Despite this importance, the specific enzyme-DNA interactions that drive topoisomerase II-mediated DNA cleavage and religation are poorly understood. Therefore, to dissect interactions between the enzyme and its cleavage site, abasic DNA lesions were incorporated into a bilaterally symmetrical and identical cleavage site. Results indicate that topoisomerase II has unique interactions with each position of the 4-base overhang generated by enzyme-mediated DNA cleavage. Lesions located 2 bases 3' to the point of scission stimulated cleavage the most, whereas those 3 bases from the point of scission stimulated cleavage the least. Moreover, an additive and in some cases synergistic cleavage enhancement was observed in oligonucleotides that contained multiple DNA lesions, with levels reaching >60-fold higher than the wild-type substrate. Finally, topoisomerase II efficiently cleaved and religated a DNA substrate in which apyrimidinic sites were simultaneously incorporated at every position on one strand of the 4-base overhang. Therefore, unlike classical DNA ligases in which base pairing is the driving force behind closure of the DNA break, it appears that for topoisomerase II, the enzyme is responsible for the spatial orientation of the DNA termini for ligation.  相似文献   

9.
A covalently cross-linked dimer of yeast DNA topoisomerase II was created by fusing the enzyme with the GCN4 leucine zipper followed by two glycines and a cysteine. Upon oxidation of the chimeric protein, a disulfide bond forms between the two carboxyl termini, covalently and intradimerically cross-linking the two protomers. In addition, all nine of the cysteines naturally occurring in topoisomerase II have been changed to alanines in this construct. This cross-linked, cysteine-less topoisomerase II is catalytically active in DNA duplex passage as indicated by ATP-dependent DNA supercoil relaxation and kinetoplast DNA decatenation assays. However, these experiments do not directly distinguish between a "one-gate" and a "two-gate" mechanism for the enzyme.  相似文献   

10.
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   

11.
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   

12.
A 30-kDa DNA topoisomerase has been purified to near homogeneity from the purple nonsulfur photosynthetic bacterium Rhodobacter capsulatus. The enzyme is recognized by an antibody against a 16-mer peptide sequence from human DNA topoisomerase I. The purified enzyme is a type I topoisomerase. Consistent with the properties of other prokaryotic type I DNA topoisomerases, the isolated enzyme is unable to relax positively supercoiled DNA and absolutely requires divalent cations for its relaxation activity. However, regardless of the Mg+2 concentrations, ATP concentrations above 5 mM completely inhibit the relaxing activity. The enzyme is sensitive to high salt concentrations and the optimal activity occurs at salt concentrations between 3 and 30 mM for monovalent cations. Single-stranded M13 DNA is a strong inhibitor of this relaxing activity. The enzyme is inhibited by ethidium bromide, confirming that this DNA topoisomerase is incapable of relaxing positive supercoils. Topoisomerase I-specific inhibitors like Hoechst 32258 and actinomycin D inhibit the enzymatic activity while the enzyme is resistant to type II topoisomerase inhibitors such as norfloxacin, nalidixic acid, and novobiocin. From these enzymatic characteristics, we conclude that the R. capsulatus DNA topoisomerase is a prokaryotic type I DNA topoisomerase.  相似文献   

13.
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   

14.
Apurinic sites are position-specific poisons of topoisomerase II and stimulate DNA scission approximately 10-18-fold when they are located within the 4-base overhang generated by enzyme-mediated cleavage (Kingma, P. S., and Osheroff, N. (1997) J. Biol. Chem. 272, 1148-1155). To determine whether other major forms of spontaneous DNA damage also act as topoisomerase II poisons, the effects of position-specific apyrimidinic sites and deaminated cytosines (i.e. uracil:guanine mismatches) on the type II enzyme were determined. Both of these lesions stimulated topoisomerase II-mediated DNA scission with the same positional specificity as apurinic sites but were less efficacious. Moreover, apurinic sites dominated the effects of apyrimidinic sites in substrates that contained multiple lesions. The differential ability of spontaneous lesions to enhance DNA cleavage did not correlate with either a decreased stability of the double helix or the size of the gap formed by base loss. Rather, it appears to be due (at least in part) to increased rates of religation for substrates containing apyrimidinic sites or deaminated cytosines. These results suggest that several forms of spontaneous DNA damage are capable of acting as endogenous poisons of topoisomerase II.  相似文献   

15.
Type II DNA topoisomerases are required for the segregation of genomic DNA at cell division in prokaryotic and eukaryotic cells, and inhibitors of these enzymes are potential cytotoxic agents in both prokaryotes and eukaryotes. The bacterial member of the topoisomerase II family, DNA gyrase, and the chemotherapeutic agents which target it are the subject of a recent review (Maxwell, A. et al., 1993, in Molecular Biology of DNA Topoisomerases, Andoh, T. et al., eds., pp. 21-30, CRC Press, Boca Raton). Here we present an overview of current knowledge of eukaryotic topoisomerase II and the anticancer agents which target this enzyme, focussing predominantly on new observations and recent reports and reviews.  相似文献   

16.
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   

17.
Doxorubicin is a therapeutically useful anticancer drug that exerts multiple biological effects. Its antitumor and cardiotoxic properties have been ascribed to anthracycline-mediated free radical damage to DNA and membranes. Evidence for this idea comes in part from the selection by doxorubicin from stationary phase yeast cells of mutants (petites) deficient in mitochondrial respiration and therefore defective in free radical generation. However, doxorubicin also binds to DNA topoisomerase II, converting the enzyme into a DNA damaging agent through the trapping of a covalent enzyme-DNA complex termed the 'cleavable complex.' We have used yeast to determine whether stabilization of cleavable complexes plays a role in doxorubicin action and cytotoxicity. A plasmid-borne yeast TOP2 gene was mutagenized with hydroxylamine and used to transform drug-permeable yeast strain JN394t2-4, which carries a temperature-sensitive top2-4 mutation in its chromosomal TOP2 gene. Selection in growth medium at the nonpermissive temperature of 35 degrees in the presence of doxorubicin resulted in the isolation of plasmid-borne top2 mutants specifying functional doxorubicin-resistant DNA topoisomerase II. Single-point changes of Gly748 to Glu or Ala642 to Ser in yeast topoisomerase II, which lie in and adjacent to the CAP-like DNA binding domain, respectively, were identified as responsible for resistance to doxorubicin, implicating these regions in drug action. None of the mutants selected in JN394t2-4, which has a rad52 defect in double-strand DNA break repair, was respiration-deficient. We conclude that topoisomerase II is an intracellular target for doxorubicin and that the genetic background and/or cell proliferation status can determine the relative importance of topoisomerase II- versus free radical-killing.  相似文献   

18.
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   

19.
Topoisomerase II is the cytotoxic target for a number of clinically relevant antitumor drugs. Berberrubine, a protoberberine alkaloid which exhibits antitumor activity in animal models, has been identified as a specific poison of topoisomerase II in vitro. Topoisomerase II-mediated DNA cleavage assays showed that berberrubine poisons the enzyme by stabilizing topoisomerase II-DNA cleavable complexes. Subsequent proteinase K treatments revealed that berberrubine-induced DNA cleavage was generated solely by topoisomerase II. Topoisomerase II-mediated DNA religation with elevated temperature revealed a substantial reduction in DNA cleavage induced by berberrubine, to the extent comparable to that of other prototypical topoisomerase II poison, etoposide, suggesting that DNA cleavage involves stabilization of the reversible enzyme-DNA cleavable complex. However, the step at which berberrubine induces cleavable complex may differ from that of etoposide as revealed by the difference in the formation of the intermediate product, nicked DNA. This suggests that berberrubine's primary mode of linear formation may involve trapping nicked molecules, formed at transition from linear to covalently closed circular DNA. Unwinding of the duplex DNA by berberrubine is consistent with an intercalative binding mode for this compound. In addition to the ability to induce the cleavable complex mediated with topoisomerase II, berberrubine at high concentrations was shown to specifically inhibit topoisomerase II catalytic activity. Berberrubine, however, did not inhibit topoisomerase I at concentrations up to 240 microM. Cleavage sites induced by topoisomerase II in the presence of berberrubine and etoposide were mapped in DNA. Berberrubine induces DNA cleavage in a site-specific and concentration-dependent manner. Comparison of the cleavage pattern of berberrubine with that of etoposide revealed that they share many common sites of cleavage. Taken together, these results indicate that berberrubine represents a new class of antitumor agent which exhibits the topoisomerase II poison activity as well as catalytic inhibition activity and may have a potential clinical value in cancer treatment.  相似文献   

20.
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   

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