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.
7-hydroxystaurosporine (UCN-01) is a more selective protein kinase C inhibitor than staurosporine. UCN-01 exhibits antitumor activity in experimental tumor models and is presently in clinical trials. Our study reveals that human myeloblastic leukemia HL60 and K562 and colon carcinoma HT29 cells undergo internucleosomal DNA fragmentation and morphological changes characteristic of apoptosis after UCN-01 treatment. These three cell lines lack functional p53, and K562 and HT29 cells are usually resistant to apoptosis. DNA fragmentation in HT29 and K562 cells occurred after 1 day of treatment while it took less than 4 h in HL60 cells. Cycloheximide prevented UCN-01-induced DNA fragmentation in HT-29 cells, but not in HL60 and K562 cells, suggesting that macromolecular synthesis is selectively required for apoptotic DNA fragmentation in HT29 cells. UCN-01-induced DNA fragmentation was preceded by activation of cyclin B1/cdc2 kinase. Further studies in HL60 cells showed that UCN-01-induced apoptosis was associated with degradation of CPP32, PARP, and lamin B and that the inhibitor of caspases (ICE/CED-3 cysteine proteases), Z-VAD-FMK, and the serine protease inhibitor, DCI, protected HL60 cells from UCN-01-induced DNA fragmentation. However, only DCI and TPCK, but not Z-VAD-FMK, inhibited DNA fragmentation in the HL60 cell-free system, suggesting that serine protease(s) may play a role in the execution phase of apoptosis in HL60 cells treated with UCN-01. Z-VAD-FMK and DCI also inhibited apoptosis in HT29 cells. These data demonstrate that the protein kinase C inhibitor and antitumor agent, UCN-01 is a potent apoptosis inducer in cell lines that are usually resistant to apoptosis and lack p53 and that caspases and probably serine proteases are activated during UCN-01-induced apoptosis.  相似文献   

2.
Discrete cleavages within 28S rRNA divergent domains have previously been found to coincide with DNA fragmentation during apoptosis. Here we show that rRNA and DNA cleavages can occur independently in apoptotic cells, i.e. that the previously observed correlation is likely to be coincidental. In HL-60 cells, apoptosis with massive DNA fragmentation could be induced without any signs of rRNA cleavage. The opposite situation; rRNA cleavage without concomitant internucleosomal DNA fragmentation, was found in okadaic acid-treated Molt-4 cells. Other leukemia cell lines underwent apoptosis either without (K562 and Molt-3) or with (U937) both forms of polynucleotide cleavage. In K562 cells transfected with a temperature-sensitive p53 mutant, internucleosomal DNA fragmentation but not 28S rRNA cleavage was inducible by wild-type p53 expression. The absence of apoptotic rRNA cleavage in some cell types suggests that this phenomenon is tightly regulated and unrelated to DNA fragmentation or a presumed scheme for general macromolecular degradation in apoptotic cells.  相似文献   

3.
Kim JM  Yamada M  Masai H 《Mutation research》2003,532(1-2):29-40
Cdc7 kinase plays an essential role in firing of replication origins by phosphorylating components of the replication complexes. Cdc7 kinase has also been implicated in S phase checkpoint signaling downstream of the ATR and Chk1 kinases. Inactivation of Cdc7 in yeast results in arrest of cell growth with 1C DNA content after completion of the ongoing DNA replication. In contrast, conditional inactivation of Cdc7 in undifferentiated mouse embryonic stem (ES) cells leads to growth arrest with rapid cessation of DNA synthesis, suggesting requirement of Cdc7 functions for continuation of ongoing DNA synthesis. Furthermore, loss of Cdc7 function induces recombinational repair (nuclear Rad51 foci) and G2/M checkpoint responses (inhibition of Cdc2 kinase). Eventually, p53 becomes highly activated and the cells undergo massive p53-dependent apoptosis. Thus, defective origin activation in mammalian cells can generate DNA replication checkpoint signals. Efficient removal of those cells in which replication has been perturbed, through cell death, may be beneficial to maintain the highest level of genetic integrity in totipotent stem cells. Partial, rather than total, loss of Cdc7 kinase expression results in retarded growth at both cellular and whole body levels, with especially profound impairment of germ cell development.  相似文献   

4.
We compare the effects of Imatinib mesylate (Glivec) on chronic myeloid leukemia derived cell lines K562 and JURL-MK1. In both cell lines, the cell cycle arrests in G(1)/G(0) phase within 24 h after the addition of 1 microM Imatinib. This is followed by a decrease of Ki-67 expression and the induction of apoptosis. In JURL-MK1 cells, the apoptosis is faster in comparison with K562 cells: the caspase-3 activity reaches the peak value (20 to 30 fold of the control) after about 40 h and the apoptosis proceeds to its culmination point, the DNA fragmentation, within 48 h following 1 microM Imatinib addition. Unlike K562 cells, JURL-MK1 cells possess a probably functional p53 protein inducible by TPA (tetradecanoyl phorbol acetate) or UV-B irradiation. However, no increase in p53 expression was observed in Imatinib-treated JURL-MK1 cells indicating that the difference in the apoptosis rate between the two cell lines is not due to the lack of p53 in K562 cells. Imatinib also triggers erythroid differentiation both in JURL-MK1 and K562 cells. Glycophorin A expression occurred simultaneously with the apoptosis, even at the single cell level. In K562 cells, but not in JURL-MK1 cells, the differentiation process involved increased hemoglobin synthesis. However, during spontaneous evolution of JURL-MK1 cells in culture, the effects produced by Imatinib progressively changed from the fast apoptosis to the more complete erythroid differentiation. We suggest that the apoptosis and the erythroid differentiation are parallel effects of Imatinib and their relative contributions, kinetics and completeness are related to the differentiation stage of the treated cells.  相似文献   

5.
We studied the role of proteases in apoptosis using a cell-free system prepared from a human leukemia cell line. HL60 cells are p53 null and extremely sensitive to a variety of apoptotic stimuli including DNA damage induced by the topoisomerase I inhibitor, camptothecin. We measured DNA fragmentation induced in isolated nuclei by cytosolic extracts using a filter elution assay. Cytosol from camptothecin-treated HL60 cells induced internucleosomal DNA fragmentation in nuclei from untreated cells. This fragmentation was suppressed by serine protease inhibitors. Serine proteases (trypsin, endoproteinase Glu-C, chymotrypsin A, and proteinase K) and papain by themselves induced DNA fragmentation in naive nuclei. This effect was enhanced in the presence of cytosol from untreated cells. Cysteine protease inhibitors (E-64, leupeptin, Ac-YVAD-CHO [ICE inhibitor]) did not affect camptothecin-induced DNA fragmentation. The apopain/Yama inhibitor, Ac-DEVD-CHO, and the proteasome inhibitor, MG-132, were also inactive both in the cell-free system and in whole cells. Interleukin-1β converting enzyme (ICE) or human immunodeficiency virus protease failed to induce DNA fragmentation in naive nuclei. Together, these results suggest that DNA damage activates serine protease(s) which in turn activate(s) nuclear endonuclease(s) during apoptosis in HL60 cells.  相似文献   

6.
The wild-type human MDM2 protooncogene was tested for its ability to modulate apoptotic activity of the de novo expressed p53 tumor suppressor gene in K562 cells. We also studied the role of some cytokines in this phenomenon. K562, a human myeloid leukemia cell line, does not express p53 at the mRNA or protein level. In this study, we stably transfected K562 with eukaryotic vectors containing either normal p53 cDNA (pC53-SN3) or mutated p53 (143Val-->Ala) cDNA (pC53-SCX3). Transfectants expressing WT p53 or those expressing mutant p53 are called K562 SN and K562 SM respectively. Many leukemic cell lines undergo apoptosis when de novo WT p53 is expressed alone. In contrast, while the resulting clones (K562 SN and K562 SM) expressed p53, they did not undergo apoptosis. However, when treated with MDM2 mRNA antisense (MDM2 AS) oligodeoxynucleotides (ODNs), K562 SN demonstrated apoptotic features at both molecular and morphological levels. No change was observed when the other clones (K562 and K562 SM) were treated with MDM2 AS. Apoptosis induced in this manner was associated with a relatively small increase in intracellular calcium [Ca2+]i. Cells cultured in medium previously supplemented with recombinant human (rh) interleukin (IL)-3 and rh-erythropoietin (Epo) did not undergo apoptosis. Moreover, K562 SN cells were induced to differentiate. This differentiation was evaluated by measuring hemoglobin (Hb) level in cellular extracted proteins and by analyzing erythroid colony number and morphology. High Hb synthesis was obtained when K562 SN cells were cultured with cytokines (IL-3 + Epo) combined with MDM2 AS. Our results are consistent with the hypothesis that the function of the proto-oncogene MDM2 is to provide a 'feedback' mechanism for the p53-dependent pathway of apoptosis that could be shunted toward differentiation.  相似文献   

7.
BACKGROUND: Some forms of chemoresistance in leukemia may start from failure of tumour cells to successfully undergo apoptosis and Bcl-2 may play a role in this defect. Therefore, we evaluated the Bcl-2 content and synthesis in relation with the apoptotic potential in leukemic cell lines after anthracycline treatment. METHODS: U937, HL60, and K562 cells and their drug resistant (DR) variants were treated with varying concentrations of Idarubicin (IDA). Apoptosis was evaluated by fluorescence microscopy after acridine orange staining. Bcl-2 and Bax content were evaluated either by flow cytometry after indirect immunolabelling or by Western blot. RESULTS: High Bcl-2 contents were not related to a poor ability to undergo apoptosis in U937, HL60, K562 and their DR variants. IDA induced a concentration-dependent increase in Bcl-2 content in all cell lines as long as they do not perform apoptosis. Enhanced Bcl-2 expression was inhibited by cycloheximide, actinomycin D, or antisense oligonucleotide directed against bcl-2 mRNA. Bcl-2 expression was also increased in the resistant U937 variant after serum deprivation or C2-ceramide treatment. The synthesis of Bcl-2 led to an increased Bcl-2/Bax ratio solely in the cells with an apoptosis-resistance phenotype. CONCLUSIONS: These data suggest that exposure to IDA induces Bcl-2 expression in leukemic cell lines, and that this mechanism could contribute to apoptosis resistance and participate in the acquisition of chemoresistance. They also confirm that the evolution of the Bcl-2/Bax ratio reflects apoptotic ability better than the steady state level of Bcl-2 expression.  相似文献   

8.
9.
Luteolin is 3',4',5,7-tetrahydroxyflavone found in celery, green pepper, and perilla leaf that inhibits tumorigenesis in animal models. We examined luteolin-mediated regulation of cell cycle progression and apoptosis in the HT-29 human colon cancer cell line. Luteolin decreased DNA synthesis and viable HT-29 cell numbers in a concentration-dependent manner. It inhibited cyclin-dependent kinase (CDK)4 and CDK2 activity, resulting in G1 arrest with a concomitant decrease of phosphorylation of retinoblastoma protein. Activities of CDK4 and CDK2 decreased within 2 h after luteolin treatment, with a 38% decrease in CDK2 activity (P < 0.05) observed in cells treated with 40 micromol/l luteolin. Luteolin inhibited CDK2 activity in a cell-free system, suggesting that it directly inhibits CDK2. Cyclin D1 levels decreased after luteolin treatment, although no changes in expression of cyclin A, cyclin E, CDK4, or CDK2 were detected. Luteolin also promoted G2/M arrest at 24 h posttreatment by downregulating cyclin B1 expression and inhibiting cell division cycle (CDC)2 activity. Luteolin promoted apoptosis with increased activation of caspases 3, 7, and 9 and enhanced poly(ADP-ribose) polymerase cleavage and decreased expression of p21(CIP1/WAF1), survivin, Mcl-1, Bcl-x(L), and Mdm-2. Decreased expression of these key antiapoptotic proteins could contribute to the increase in p53-independent apoptosis that was observed in HT-29 cells. We demonstrate that luteolin promotes both cell cycle arrest and apoptosis in the HT-29 colon cancer cell line, providing insight about the mechanisms underlying its antitumorigenic activities.  相似文献   

10.
In the present study, we compared the abilities of ricin and diphtheria toxin to induce apoptosis in Vero cells. The cytolysis and DNA fragmentation by ricin paralleled its protein synthesis inhibitory activity. However, unlike ricin, diphtheria toxin could induce neither cytolysis nor DNA fragmentation in Vero cells up to very high concentration, in spite of the fact that Vero cells were even more sensitive to protein synthesis inhibition by diphtheria toxin than ricin. Interestingly, coexistence of brefeldin A (BFA) and okadaic acid (OA) significantly enhanced diphtheria toxin-mediated cytolysis and DNA fragmentation without affecting the activity of protein synthesis inhibition. Ammonium chloride almost completely abolished the ability of diphtheria toxin to induce apoptosis in the presence of BFA and OA as well as the protein synthesis inhibitory activity. The mutant CRM 197, which does not catalyze the ADP ribosylation of elongation factor-2 (EF-2), failed to induce apoptosis in Vero cells even in the presence of BFA and OA. Thus, translocation of diphtheria toxin into the cytosol and subsequent enzymatic inactivation of EF-2 may be necessary steps to induce apoptosis. Taken together our results suggest that protein synthesis inhibition by toxins is not sufficient to induce apoptosis, and underlying mechanisms of apoptosis induction may be distinct between ricin and diphtheria toxin. Since a morphological change in the Golgi complex was observed in Vero cells treated with BFA and OA, modulation of the Golgi complex by these reagents may be partly responsible for enhanced apoptosis induction by diphtheria toxin.  相似文献   

11.
We studied the modulating effect of protein tyrosine kinase inhibitors on the response of cells of the human chronic myelogenous leukemia cell line K562 to radiation. The radiosensitivity of the cells was increased by treatment with herbimycin A and decreased by treatment with genistein. This modulating effect of protein tyrosine kinase inhibitors on radiation sensitivity was associated with the alteration of the mode of radiation-induced cell death. After X irradiation, the cells arrested in the G(2) phase of the cell cycle, but these TP53(-/-) cells were unable to sustain cell cycle arrest. This G(2)-phase checkpoint deficit caused cell death. The morphological pattern of cell death was characterized by swelling of the cytoplasmic compartments, cytosolic vacuolation, disruption of the plasma membrane, less evident nuclear condensation, and faint DNA fragmentation, all of which were consistent with oncosis or cytoplasmic apoptosis. The nonreceptor protein tyrosine kinase inhibitor herbimycin A accelerated the induction of typical apoptosis by X irradiation, which was demonstrated by morphological assessments using nuclear staining and electron microscopy as well as oligonucleosomal fragmentation and caspase 3 activity. Herbimycin A is known to be a selective antagonist of the BCR/ABL kinase of Philadelphia chromosome-positive K562 cells; this kinase blocks the induction of apoptosis after X irradiation. Our results showed that the inhibition of protein tyrosine kinase by herbimycin A enhanced radiation-induced apoptosis in K562 cells. This effect was associated with the activation of caspase 3 and rapid abrogation of the G(2)-phase checkpoint with progression out of G(2) into G(1) phase. In contrast, the receptor-type protein tyrosine kinase inhibitor genistein protected K562 cells from all types of radiation-induced cell death through the inhibition of caspase 3 activity and prolonged maintenance of G(2)-phase arrest. Further investigations using this model may give valuable information about the mechanisms of radiation-induced apoptosis and about the radiosensitivity and radioresistance of chronic myelogenous leukemia cells having the Philadelphia chromosome.  相似文献   

12.
During endomitosis, megakaryocytes undergo several rounds of DNA synthesis without division leading to polyploidization. In primary megakaryocytes and in the megakaryocytic cell line CHRF, loss or knock-down of p53 enhances cell cycling and inhibits apoptosis, leading to increased polyploidization. To support the hypothesis that p53 suppresses megakaryocytic polyploidization, we show that stable expression of wild-type p53 in K562 cells (a p53-null cell line) attenuates the cells' ability to undergo polyploidization during megakaryocytic differentiation due to diminished DNA synthesis and greater apoptosis. This suggested that p53's effects during megakaryopoiesis are mediated through cell cycle- and apoptosis-related target genes, possibly by arresting DNA synthesis and promoting apoptosis. To identify candidate genes through which p53 mediates these effects, gene expression was compared between p53 knock-down (p53-KD) and control CHRF cells induced to undergo terminal megakaryocytic differentiation using microarray analysis. Among substantially downregulated p53 targets in p53-KD megakaryocytes were cell cycle regulators CDKN1A (p21) and PLK2, proapoptotic FAS, TNFRSF10B, CASP8, NOTCH1, TP53INP1, TP53I3, DRAM1, ZMAT3 and PHLDA3, DNA-damage-related RRM2B and SESN1, and actin component ACTA2, while antiapoptotic CKS1B, BCL2, GTSE1, and p53 family member TP63 were upregulated in p53-KD cells. Additionally, a number of cell cycle-related, proapoptotic, and cytoskeleton-related genes with known functions in megakaryocytes but not known to carry p53-responsive elements were differentially expressed between p53-KD and control CHRF cells. Our data support a model whereby p53 expression during megakaryopoiesis serves to control polyploidization and the transition from endomitosis to apoptosis by impeding cell cycling and promoting apoptosis. Furthermore, we identify a putative p53 regulon that is proposed to orchestrate these effects.  相似文献   

13.
14.
Park YC  Jeong JH  Park KJ  Choi HJ  Park YM  Jeong BK  Higuchi Y  Yoo YH 《Life sciences》2005,77(16):2059-2070
Sulindac is one of the most widely studied nonsteroidal anti-inflammatory drugs in the prevention of colon cancer. Thus, from the viewpoint of colon cancer chemotherapy it is important to reveal the mechanism of sulindac-induced cell death. This study was undertaken to dissect the molecular mechanism underlying sulindac-induced apoptosis in human colon cancer cell line HT-29 (mutant p53), focusing on nuclear translocation of AIF, DFF and endonuclease G. On induction of apoptosis by sulindac, it was associated with decreased mitochondrial membrane potential, nuclear expression of active caspase-3, cleavage of poly(ADP-ribose) polymerase, translocation of mitochondrial proteins to the nucleus, and morphological evidence of nuclear condensation. However, sulindac led to only disintegration of nuclear DNA into high molecular weight DNA fragments of about 100-300 kbp as determined by a pulse-field gel electrophoresis, suggesting a predominantly AIF-mediated cell death process. In summary, our findings indicate that sulindac induces large-scale DNA fragmentation without oligonucleosomal DNA fragmentation. This result suggests that nuclear translocation of DFF and endonuclease G are not sufficient for the induction of oligonucleosomal DNA fragmentation in HT-29 cells.  相似文献   

15.
Cdc7, originally discovered by Hartwell as a budding yeast mutant that arrests immediately before the onset of S phase, is conserved through evolution and plays essential roles in initiation of mitotic DNA replication. Inducible inactivation of Cdc7 in mouse embryonic stem cells leads to rapid cessation of DNA synthesis and the subsequent activation of checkpoint responses, resulting in p53 activation and eventually p53-mediated apoptosis. This indicates a requirement of Cdc7 kinase for ongoing replication of mammalian genomes, and loss of Cdc7 kinase presumably generates arrested replication fork signals. Cdc7-/- mice or embryonic fibroblast cells (MEFs) expressing a low level of transgene-encoded Cdc7 protein are viable but exhibit reduced body size with impaired germ cell development and decreased cell proliferation. Interestingly, these phenotypes are largely corrected by the presence of an additional copy of the transgene, resulting in increased level of Cdc7 expression. This indicates the requirement of a critical level of Cdc7 for normal cell proliferation and development of specific organs. These results from mammals will be discussed in conjunction with the pleiotropic effects of Cdc7 mutation observed in yeasts.  相似文献   

16.
Cdc7, originally discovered by Hartwell1 as a budding yeast mutant that arrests immediately before the onset of S phase, is conserved through evolution and plays essential roles in initiation of mitotic DNA replication. Inducible inactivation of Cdc7 in mouse embryonic stem cells leads to rapid cessation of DNA synthesis and the subsequent activation of checkpoint responses, resulting in p53 activation and eventually p53-mediated apoptosis. This indicates a requirement of Cdc7 kinase for ongoing replication of mammalian genomes, and loss of Cdc7 kinase presumably generates arrested replication fork signals. Cdc7-/- mice or embryonic fibroblast cells (MEFs) expressing a low level of transgene-encoded Cdc7 protein are viable but exhibit reduced body size with impaired germ cell development and decreased cell proliferation. Interestingly, these phenotypes are largely corrected by the presence of an additional copy of the transgene, resulting in increased level of Cdc7 expression. This indicates the requirement of a critical level of Cdc7 for normal cell proliferation and development of specific organs. These results from mammals will be discussed in conjunction with the pleiotropic effects of Cdc7 mutation observed in yeasts.  相似文献   

17.
An increasing number of methods are being described to detect apoptotic cells. However, attempts to detect apoptotic cells in clinical samples are rarely successful. A hypothesis is that apoptotic cells are cleared from the circulation by phagocytosis before they become detectable by conventional morphological or cytometric methods. Using LR73 adhering cells as phagocytes in a model ofin vitrophagocytosis, we found that phagocytosis of daunorubicin (DNR)-treated U937, HL60, or K562 leukemia cell lines occurred prior to phosphatidylserine externalization, DNA hydrolysis, chromatin condensation, nuclear fragmentation, or mitochondrial potential alteration. Moreover DNR-treated K562 cells were eliminated by phagocytes while apoptosis was never observed by any of the above methods. By contrast, using a fluorometric batch analysis assay to detect caspase activity in ceramide- or DNR-treated cells (fluorogenic substrate for caspase), we found that caspase activity increased in apoptosis-committed cells before they were detected by flow cytometry or recognized by phagocytes. Similarly a caspase activity increase was detected in circulating mononuclear cells of leukemic patients 15 h after the beginning of anthracyclin treatment. We suggest that recent findings on enzymatic events (caspase activation) occurring in the early events of apoptosis must now allow the development of new markers for apoptosis, irrespective of the morphological features or internucleosomal fragmentation which are late events in apoptosis.  相似文献   

18.
The p53 tumor suppressor gene is thought to be required for the induction of programmed cell death (apoptosis) initiated by DNA damage. We show here, however, that the human promyelocytic leukemia cell line HL-60, which is known to be deficient in p53 because of large deletions in the p53 gene, can be induced to undergo apoptosis following X-irradiation. We demonstrate that the decision to undergo apoptosis in this cell line appears to be made at a G2 checkpoint. In addition, we characterize an HL-60 variant, HCW-2, which is radioresistant. HCW-2 cells display DNA damage induction and repair capabilities identical to those of the parental HL-60 cell line. Thus, the difference between the two cell lines appears to be that X-irradiation induces apoptosis in HL-60, but not in HCW-2, cells. Paradoxically, HCW-2 cells display high levels of expression of bax, which enhances apoptosis, and no longer express bcl-2, which blocks apoptosis. HCW-2 cells' resistance to apoptosis may be due to the acquisition of expression of bcl-XL, a bcl-2-related inhibitor of apoptosis. In summary, apoptosis can be induced in X-irradiated HL-60 cells by a p53-independent mechanism at a G2 checkpoint, despite the presence of endogenous bcl-2. The resistance shown by HCW-2 cells suggests that bcl-XL can block this process.  相似文献   

19.
Treatment of HL60 and Jurkat leukaemic cell lines, both not expressing p53, with the new non-covalent DNA minor groove binder -bromoacryloyl-distamycin (PNU 151807), induces apoptosis as shown either morphologically or by DNA laddering formation. We evaluated the p53-independent mechanisms of activation of apoptosis in these cell systems, by determining the levels of different caspases at different times after treatment with PNU 151807. Through Western blotting analysis we could measure the cleavage of the 110 Kd form of PARP in both HL60 and Jurkat cells and correspondingly the activation of CPP32-caspase 3. The levels of caspase-1 did not change at any time after drug treatment. By using the tetrapeptidic sequence recognized by caspase-3 (DEVD-AMC) or by caspase-1 (YVAD-AMC) linked to fluorogenic substrate, we also demonstrated that only the DEVD sequence was recognized and cleaved after drug treatment, while no significant changes were found for YVAD peptides. PNU 151807-induced DNA fragmentation and DEVD-cleavage were both inhibited by concomitant treatment with the specific inhibitor DEVD-CHO, but not by YVAD-CHO, clearly demonstrating the direct activation of caspase-3-like caspases in the induction of programmed cell death in these cell lines after minor groove binder exposure.  相似文献   

20.
为了探讨人野生型p53(wt-p53)基因增强大肠癌细胞化疗敏感性的分子生物学机制,将携带wt p53基因的质粒分别转染两种p53基因突变的人大肠癌细胞系HT-29及SW620,分析细胞中p53及细胞周期蛋白D1(cyclin D1)蛋白的表达水平;将化疗药物5 氟尿嘧啶(5-fluorouracil,5-FU)以不同浓度、不同时间分别作用于HT-29及SW620细胞,另外将已转染wt-p53基因的大肠癌细胞用5-FU进行诱导,Western印迹分析上述干预条件下细胞中p53蛋白及细胞周期蛋白D1表达水平的变化;流式细胞术检测wt p53基因联合5-FU组及对照组中细胞凋亡的改变情况.结果表明,wt-p53基因能增加癌细胞中细胞周期蛋白D1的表达,与wt-p53基因呈剂量依赖性关系;5-FU则降低其蛋白表达,与5-FU呈时间和剂量依赖性关系,而5-FU所致的细胞周期蛋白D1表达水平的降低在细胞预先转染了wt- p53基因时会被抑制;wt-p53基因与5-FU联合使用能提高大肠癌细胞凋亡率.结果提示,wt-p53基因可提高大肠癌细胞中细胞周期蛋白D1的表达水平,并抑制5-FU所致的细胞周期蛋白D1降解,从而提高大肠癌细胞对化疗药物5-FU的敏感性.  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