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 62 毫秒
1.
The adenovirus type 5 243R E1A protein induces p53-dependent apoptosis in the absence of the 19- and 55-kDa E1B polypeptides. This effect appears to result from an accumulation of p53 protein and is unrelated to expression of E1B products. We now report that in the presence of the E1B 55-kDa polypeptide, the 289R E1A protein does not induce such p53 accumulation and, in fact, is able to block that induced by E1A 243R. This inhibition also requires the 289R-dependent transactivation of E4orf6 expression. E4orf6 is known to form complexes with the E1B 55-kDa protein and to function both in the transport and stabilization of viral mRNA and in shutoff of host cell protein synthesis. We demonstrated that the block in p53 accumulation is not due to the generalized shutoff of host cell metabolism. Rather, it appears to result from a mechanism targeted specifically to p53, most likely involving a decrease in the stability of p53 protein. The E1B 55-kDa protein is known to interact with both E4orf6 and p53, and as demonstrated recently by others, we showed that E4orf6 also binds directly to p53. Thus, multiple interactions between all three proteins may regulate p53 stability, resulting in the maintenance of low levels of p53 following virus infection.  相似文献   

2.
Bap31 is a polytopic integral membrane protein of the endoplasmic reticulum and forms a complex with Bcl-2/Bcl-XL and procaspase-8 (Ng, F. W. H., Nguyen, M., Kwan, T., Branton, P. E., Nicholson, W. D., Cromlish, J. A., and Shore, G. C. (1997) J. Cell Biol. 139, 327-338). In co-transfected human cells, procaspase-8 is capable of interacting with Ced-4, an important adaptor molecule in Caenorhabditis elegans that binds to and activates the C. elegans procaspase, proCed-3. Here, we show that the predicted death effector homology domain within the cytosolic region of Bap31 interacts with Ced-4 and contributes to recruitment of procaspase-8. Bcl-XL, which binds directly but weakly to the polytopic transmembrane region of Bap31, indirectly and cooperatively associates with the Bap31 cytosolic domain, dependent on the presence of procaspase-8 and Ced-4. Ced-4Deltac does not interact with Bcl-XL but rather displaces it from Bap31, suggesting that an endogenous Ced-4-like adaptor is a normal constituent of the Bap31 complex and is required for stable association of Bcl-XL with Bap31 in vivo. These findings indicate that Bap31 is capable of recruiting essential components of a core death regulatory machinery.  相似文献   

3.
4.
The product of open reading frame 14 (orf14) of herpesvirus saimiri (HVS) exhibits significant homology with mouse mammary tumor virus superantigen. orf14 encodes a 50-kDa secreted glycoprotein, as shown previously (Z. Yao, E. Maraskovsky, M. K. Spriggs, J. I. Cohen, R. J. Armitage, and M. R. Alderson, J. Immunol. 156:3260-3266, 1996). orf14 expressed from recombinant baculovirus powerfully induces proliferation of CD4-positive cells originating from several different species. To study the role of orf14 in transformation, a mutant form of HVS (HVS Deltaorf14) was constructed with a deletion in the orf14 gene. The transforming potential of HVS Deltaorf14 was tested in cell culture and in common marmosets. Parental HVS subgroup C strain 488 immortalized common marmoset T lymphocytes in vitro to interleukin-2-independent growth, while the HVS Deltaorf14 mutant did not produce such a growth transformation. In addition, HVS Deltaorf14 was nononcogenic in common marmosets. In contrast to other nononcogenic HVS mutant viruses which were repeatedly isolated from peripheral blood mononuclear cells of infected marmosets for more than 5 months, HVS Deltaorf14 did not persist at a high level in vivo. These results demonstrate that orf14 of HVS is not required for replication but is required for transformation and for high-level persistence in vivo.  相似文献   

5.
From previous studies on the induction of DNA synthesis in quiescent primary baby rat kidney cells by adenovirus type 5 (Ad5) E1A deletion mutants, we concluded that induction is prevented only when cellular proteins p300 and pRb are both uncomplexed with E1A (J.A. Howe, J.S. Mymryk, C. Egan, P.E. Branton, and S.T. Bayley, Proc. Natl. Acad. Sci. USA 87:5883-5887, 1990). We have now examined induction by these same mutants in virus lacking the E1B region, so that cellular p53 was no longer complexed to the E1B 55-kDa protein. E1A mutants that fail to bind pRb induced DNA synthesis at a significantly lower level in Ad5 lacking E1B than in Ad5 containing E1B. Apparently, therefore, uncomplexed p53 can partially replace p300 in cooperating with pRb to suppress DNA synthesis in baby rat kidney cells.  相似文献   

6.
Paclitaxel induces a cell cycle block at G2-M phase by preventing the depolymerization of microtubules and induces p53-independent apoptosis in many cancer cells. We observed that gastric cancer cells treated with paclitaxel have shown a cyclin-dependent kinase (CDK)4 down-regulation. This paclitaxel-induced CDK4 down-regulation resulted in a cell cycle arrest at G1-S phase. To confirm this observation, we prepared stable transfectants that overexpressed CDK4 and analyzed the cell cycle progression. Ectopic expression of CDK4 in SNU cells resulted in a release of paclitaxel-induced G1 arrest. The release of G1 arrest by enforced expression of CDK4 seems to make the cells more sensitive to paclitaxel-induced apoptosis. From this finding, we could then suggest that paclitaxel treatment induces both G1-S and G2-M blocks in the cell cycle progression of gastric cancer cells.  相似文献   

7.
The 34-kDa early-region 4 open reading frame 6 (E4orf6) product of human adenovirus type 5 forms complexes with both the cellular tumor suppressor p53 and the viral E1B 55-kDa protein (E1B-55kDa). E4orf6 can inhibit p53 transactivation activity, as can E1B-55kDa, and in combination these viral proteins cause the rapid turnover of p53. In addition, E4orf6-55kDa complexes play a critical role at later times in the regulation of viral mRNA transport and shutoff of host cell protein synthesis. In the present study, we have further characterized some of the biological properties of E4orf6. Analysis of extracts from infected cells by Western blotting indicated that E4orf6, like E1A and E1B products, is present at high levels until very late times, suggesting that it is available to act throughout the infectious cycle. This pattern is similar to that of E4orf4 but differs markedly from that of another E4 product, E4orf6/7, which is present only transiently. Synthesis of E4orf6 is maximal at early stages but ceases completely with the onset of shutoff of host protein synthesis; however, it was found that unlike E4orf6/7, E4orf6 is very stable, thus allowing high levels to be maintained even at late times. E4orf6 was shown to be phosphorylated at low levels. Coimmunoprecipitation studies in cells lacking p53 indicated that E4orf6 interacts with a number of other proteins. Five of these were shown to be viral or virally induced proteins ranging in size from 102 to 27 kDa, including E1B-55kDa. One such species, of 72 kDa, was shown not to represent the E2 DNA-binding protein and thus remains to be identified. Another appeared to be the L4 100-kDa nonstructural adenovirus late product, but it appeared to be present nonspecifically and not as part of an E4orf6 complex. Apart from p53, three additional cellular proteins, of 84, 19, and 14 kDa were detected by using an adenovirus vector that expresses only E4orf6. The 19-kDa species and a 16-kDa cellular protein were also shown to interact with E4orf6/7. It is possible that complex formation with these viral and cellular proteins plays a role in one or more of the biological activities associated with E4orf6 and E4orf6/7.  相似文献   

8.
9.
Recently we have shown that in fibroblasts (NIH 3T3 and Rat-1 cells) inhibition of protein geranylgeranylation leads to a G0/G1 arrest, whereas inhibition of protein farnesylation does not affect cell cycle distribution. Here we demonstrate that in human tumor cells the geranylgeranyltransferase-I (GGTase-I) inhibitor GGTI-298 blocked cells in G0/G1, whereas the farnesyltransferase (FTase) inhibitor FTI-277 showed a differential effect depending on the cell line. FTI-277 accumulated Calu-1 and A-549 lung carcinoma and Colo 357 pancreatic carcinoma cells in G2/M, T-24 bladder carcinoma, and HT-1080 fibrosarcoma cells in G0/G1, but had no effect on cell cycle distribution of pancreatic (Panc-1), breast (SKBr 3 and MDAMB-231), and head and neck (A-253) carcinoma cells. Furthermore, treatment of Calu-1, Panc-1, Colo 357, T-24, A-253, SKBr 3, and MDAMB-231 cells with GGTI-298, but not FTI-277, induced the protein expression levels of the cyclin-dependent kinase inhibitor p21WAF. HT-1080 and A-549 cells had a high basal level of p21WAF, and GGTI-298 did not further increase these levels. Furthermore, GGTI-298 also induces the accumulation of large amounts of p21WAF mRNA in Calu-1 cells, a cell line that lacks the tumor suppressor gene p53. There was little effect of GGTI-298 on the cellular levels of another cyclin- dependent kinase inhibitor p27KIP as well as cyclin E and cyclin D1. These results demonstrate that GGTase-I inhibitors arrest cells in G0/G1 and induce accumulation of p21WAF in a p53-independent manner and that FTase inhibitors can interfere with cell cycle events by a mechanism that involves neither p21WAF nor p27KIP. The results also point to the potential of GGTase-I inhibitors as agents capable of restoring growth arrest in cells lacking functional p53.  相似文献   

10.
We have earlier shown that wild-type (wt) p53 expressed from a temperature-sensitive construct (ts p53) triggers apoptosis in the v-myc retrovirus-induced, p53-negative T-cell lymphoma line J3D (Y. Wang et al., Cell Growth & Differ., 4: 467-473, 1993). We also found that constitutive bcl-2 expression inhibits wt p53-triggered apoptosis in these cells (Y. Wang et al., Oncogene, 8: 3427-3431, 1993). Here we demonstrate that more than 90% of the ts p53-transfected J3D cells were arrested in G1 at 18 h after induction of wt p53 expression by temperature shift to 32 degrees C. At this time, at least 80% of the cells remained viable. After 30 h at 32 degrees C, around 50% of the cells had died by apoptosis, while most of the remaining cells were still alive in G1, indicating that p53-induced apoptosis occurred following G1 arrest. The G1 cell cycle arrest at 18 h after temperature shift to 32 degrees C was reversible, as shown by the fact that the cells readily resumed exponential growth following temperature shift back to 37 degrees C, although viability dropped from around 80 to 65%. Expression of both WAF1 and bax mRNA was induced by wt p53 in both the ts p53 and ts p53/bcl-2 transfected cells. The kinetics of G1 cell cycle arrest at 32 degrees C was similar in both the ts p53 and the ts p53/bcl-2 double transfectants.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
The caspase-3 has been shown to be involved in mediating apoptosis induced by different stimuli. However, it is still unclear whether p53 is required for the ionizing radiation (IR)-induced caspase-3 activation. In the present study, we examined IR-induced apoptosis in three closely related human lymphoblast cell lines that differ in p53 status. Irradiation of TK6 cells (wild-type p53) with 4 Gy gamma-rays resulted in rapid apoptosis, whereas the apoptotic response was delayed and reduced in WTK1 cells (mutant p53) and the TK6 derivative line expressing HPV16 E6 (abrogated p53). The differential apoptotic responses in these cell lines correlated with caspase-3 activation. IR induced an early as well as a late phase of caspase-3 activation in TK6 but only a delayed onset in WTK1 and TK6-E6-5E cells. The early phase of caspase-3 activation coincided with an elevation of p53 and bax protein levels. Pretreatment of all three cell lines with a caspases inhibitor z-VAD-FMK inhibited apoptosis. These results suggest that IR-induced apoptosis is mediated by a mechanism involving the caspase-3 cascade, which is shared by both p53-dependent and -independent pathways. The activation of caspase-3 by IR may thus engage at least two separate mechanisms, one through the regulation of the bcl-2 family members by p53, whereas the other yet-to-be-identified one involves neither p53 nor bax.  相似文献   

12.
Previous observations that the adenovirus type 5 (Ad5) E4orf6 and E4orf3 gene products have redundant effects in viral lytic infection together with the recent findings that E4orf6 possesses transforming potential prompted us to investigate the effect of E4orf3 expression on the transformation of primary rat cells in combination with adenovirus E1 oncogene products. Our results demonstrate for the first time that E4orf3 can cooperate with adenovirus E1A and E1A plus E1B proteins to transform primary baby rat kidney cells, acting synergistically with E4orf6 in the presence of E1B gene products. Transformed rat cells expressing E4orf3 exhibit morphological alterations, higher growth rates and saturation densities, and increased tumorigenicity compared with transformants expressing E1 proteins only. Consistent with previous results for adenovirus-infected cells, the E4orf3 protein is immunologically restricted to discrete nuclear structures known as PML oncogenic domains (PODs) in transformed rat cells. As opposed to E4orf6, the ability of E4orf3 to promote oncogenic cell growth is probably not linked to a modulation of p53 functions and stability. Instead, our results indicate that the transforming activities of E4orf3 are due to combinatorial effects that involve the binding to the adenovirus 55-kDa E1B protein and the colocalization with PODs independent from interactions with the PML gene product. These data fit well with a model in which the reorganization of PODs may trigger a cascade of processes that cause uncontrolled cell proliferation and neoplastic growth. In sum, our results provide strong evidence for the idea that interactions with PODs by viral proteins are linked to oncogenic transformation.  相似文献   

13.
14.
In hematopoietic cells, gamma-irradiation causes a p53-dependent transient G1 phase cell cycle arrest. Various extracellular growth inhibitory signals elicit G1 arrest by targeting CDK4. Here we show that in a myeloid cell line, 32D cl 3, enforced expression of CDK4, but not cyclins D2 nor D3, overrides the gamma-irradiation-induced G1 arrest. CDK4 does not confer resistance to the radiation-induced G2 block observed in parental cells. Ectopic expression of CDK4 overcomes the ionizing radiation-induced inhibition of CDK4 and CDK2 kinase activity. The levels of CDK4 protein do not change after exposure to ionizing radiation in either parental cells or those overexpressing CDK4. Ionizing radiation induces the expression of both p53 and p21, and in cells constitutively synthesizing exogenous CDK4, the return of p53 protein levels to baseline is prolonged. Increased levels of p21 are found associated with CDK4, and not CDK2, in the lines overexpressing CDK4, compared to the parental line, after exposure to ionizing radiation. Enforced expression of CDK4 may therefore overcome a gamma-irradiation-induced G1 arrest through the titration of the CDK inhibitor p21 allowing both CDK4 and CDK2 to remain active.  相似文献   

15.
16.
Photodynamic therapy (PDT) is an effective local cancer treatment that induces cytotoxicity through the intracellular generation of reactive oxygen species. The current study investigated whether abrogation of wild-type p53 expression modified the sensitivity of tumor cells to PDT-mediated oxidative stress. In these experiments, human colon (LS513) and breast (MCF-7) carcinoma cells exhibiting a wild-type p53 phenotype were directly compared to LS513 and MCF-7 cells with abrogated p53 function induced by stable integration of the human papillomavirus type 16 E6 viral oncoprotein. The effectiveness of this viral oncoprotein to target p53 for degradation was confirmed using a p53 transactivation reporter gene assay. Western analysis also confirmed attenuated expression of p53 in E6-transfected cells. Photosensitivity of PDT-treated cells was measured by a clonogenic assay and found to be equivalent for parental and p53-abrogated cells. PDT-mediated oxidative stress resulted in a rapid shift of pRb from a hyperphosphorylated form to a predominantly underphosphorylated form in parental cells that was not preceded by increases in p53 or p21 expression. Hypophosphorylated pRb was also observed in PDT-treated LS513/E6 and MCF-7/E6 cells, further indicating that p53 was not involved in this process. Delayed expression of p53 and p21 proteins was seen in parental cells 24-48 h after photosensitization. Cell cycle analysis showed that the abrogation of p53 had minimal effects on an observed PDT-induced G1 block. Rapid induction of apoptosis was documented in PDT-treated LS513 cells, whereas LS513/E6 treated cells exhibited reduced apoptosis in response to PDT. The MCF-7 cell lines exhibited a minimal apoptotic response to PDT. These results indicate that p53 expression does not directly modulate tumor cell sensitivity to PDT in either apoptosis-responsive (LS513) or nonresponsive (MCF-7) cells.  相似文献   

17.
18.
Tumor suppressor p53 is a nuclear protein that is induced by DNA damage and is involved in G1 and G2 phase control of the cell cycle. p21WAF1/CIP1/SDI1 (p21), a cyclin-dependent kinase inhibitor, is a downstream target and effector of p53 to induce G1 arrest. Mimosine is a potent reversible late G1 phase blocker of the cell cycle. In this study, we showed that mimosine can increase both p21 mRNA and protein levels, indirectly inhibit cyclin E-associated kinase activity without affecting the cyclin E protein level, block human breast cancer cells (21PT) in the late G1 phase of the cell cycle, and induce a p53-independent p21 pathway in these cells. These results support the possibility of restoring a G1 checkpoint by use of mimosine. They also suggest that the mechanism of the effect of mimosine is complex and may have more than one target in the cell.  相似文献   

19.
20.
The open reading frame (ORF) 1b-encoded part of the equine arteritis virus (EAV) replicase is expressed by ribosomal frameshifting during genome translation, which results in the production of an ORF1ab fusion protein (345 kDa). Four ORF1b-encoded processing products, nsp9 (p80), nsp10 (p50), nsp11 (p26), and nsp12 (p12), have previously been identified in EAV-infected cells (L. C. van Dinten, A. L. M. Wassenaar, A. E. Gorbalenya, W. J. M. Spaan, and E. J. Snijder, J. Virol. 70:6625-6633, 1996). In the present study, the generation of these four nonstructural proteins was shown to be mediated by the nsp4 serine protease, which is the main viral protease (E. J. Snijder, A. L. M. Wassenaar, L. C. van Dinten, W. J. M. Spaan, and A. E. Gorbalenya, J. Biol. Chem. 271:4864-4871, 1996). Mutagenesis of candidate cleavage sites revealed that Glu-2370/Ser, Gln-2837/Ser, and Glu-3056/Gly are the probable nsp9/10, nsp10/11, and nsp11/12 junctions, respectively. Mutations which abolished ORF1b protein processing were introduced into a recently developed infectious cDNA clone (L. C. van Dinten, J. A. den Boon, A. L. M. Wassenaar, W. J. M. Spaan, and E. J. Snijder, Proc. Natl. Acad. Sci. USA 94:991-997, 1997). An analysis of these mutants showed that the selective blockage of ORF1b processing affected different stages of EAV reproduction. In particular, the mutant with the nsp10/11 cleavage site mutation Gln-2837-->Pro displayed an unusual phenotype, since it was still capable of RNA synthesis but was incapable of producing infectious virus.  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