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.
We have identified the phosphorylation sites in monkey p53 as well as specific changes in the phosphorylation state of free and complexed forms of simian virus 40 (SV40) large T antigen (T) and monkey p53 isolate from SV40 lytically infected CV1 cells. Phosphopeptide analyses of free T and p53 (To and p53o) and complexed T and p53 (T+ and p53+) fractions indicated several quantitative increases in the specific phosphorylation of complexed forms of both proteins. The N terminus of monkey p53+ is phosphorylated at Ser-9, Ser-15, Ser-20, either Ser-33 or Ser-37, and at least one of Ser-90 to Ser-99. The C-terminal sites are Ser-315 and Ser-392. On comparing p53+ with p53o, we found that labeling of the two N-terminal phosphotryptic peptides encompassing residues 1 to 20 and 33 to 101 was increased fivefold and that Ser-315 was sevenfold more labeled than was Ser-392. When T+ was compared with To, the N-terminal peptide containing phosphorylation sites Ser-106 through Thr-124 was twofold more labeled, the peptide containing Ser-657 through Ser-679 was sixfold more labeled and contained up to four phosphorylated serine residues, and Ser-639 and Thr-701 appeared unchanged. Overall, T+ molecules appeared to contain 3.5 mol more of labeled phosphate than did To, with the N-terminal peptide appearing fully phosphorylated. The phosphopeptide patterns obtained for lytic T+ and To fractions were nearly identical to those found for wild-type SV40 T (stably complexed with mouse p53) and mutant 5080 T (defective for p53 binding) expressed in transformed C3H10T1/2 cells (L. Tack, C. Cartwright, J. Wright, A. Srinivasan, W. Eckhart, K. Peden, and J. Pipas, J. Virol. 63:3362-3367, 1989). These results indicate that increases in specific phosphorylation sites in both T+ and p53+ correlate with the association of T with p53. The enhanced phosphorylation state may be a consequence of complex formation between T and p53 or reflect an increased affinity of p53 for highly phosphorylated forms of T.  相似文献   

2.
A 13 kDa protein, covalently linked to a small RNA from the cytoplasm of mouse cells, was studied. Sequence analysis of its tryptic peptides revealed that the RNA-linked protein is identical to prothymosin alpha. Very similar RNA-protein complexes were identified in human, bovine and yeast cells. Tryptic peptide maps of 125I-labelled RNA-linked proteins of diverse origin demonstrated their marked similarity, thus indicating high evolutionary conservation of prothymosin alpha from yeast to man.  相似文献   

3.
Simian virus 40 (SV40) large-T antigen and the cellular protein p53 were phosphorylated in vivo by growing cells in the presence of 32Pi. The large-T/p53 complex was isolated by immunoprecipitation and used as a substrate for protein phosphatase 2A (PP2A) consisting of the catalytic subunit (C) and the two regulatory subunits, A and B. Three different purified forms of PP2A, including free C, the AC form, and the ABC form, could readily dephosphorylate both proteins. With both large-T and p53, the C subunit was most active, followed by the AC form, which was more active than the ABC form. The activity of all three forms of PP2A toward these proteins was strongly stimulated by manganese ions and to a lesser extent by magnesium ions. The presence of complexed p53 did not affect the dephosphorylation of large-T antigen by PP2A. The dephosphorylation of individual phosphorylation sites of large-T and p53 were determined by two-dimensional peptide mapping. Individual sites within large-T and p53 were dephosphorylated at different rates by all three forms of PP2A. The phosphates at Ser-120 and Ser-123 of large-T, which affect binding to the origin of SV40 DNA, were removed most rapidly. Three of the six major phosphopeptides of p53 were readily dephosphorylated, while the remaining three were relatively resistant to PP2A. Dephosphorylation of most of the sites in large-T and p53 by the AC form was inhibited by SV40 small-t antigen. The inhibition was most apparent for those sites which were preferentially dephosphorylated. Inhibition was specific for the AC form; no effect was observed on the dephosphorylation of either protein by the free C subunit or the ABC form. The inhibitory effect of small-t on dephosphorylation by PP2A could explain its role in transformation.  相似文献   

4.
Cell-free synthesis of simian virus 40 T-antigens.   总被引:27,自引:18,他引:9       下载免费PDF全文
  相似文献   

5.
Myxobacteria have been shown to contain a large number of branched RNA-linked single-stranded DNA (multicopy single-stranded DNA (msDNA] molecules. In addition, we found that Myxococcus xanthus contains another smaller msDNA-like molecule, designated mrDNA, consisting of a 65-base single-stranded DNA covalently linked by a 2',5'-phosphodiester linkage to a 49-base branched RNA. In spite of their different primary sequences, the RNA-linked mrDNA is remarkably similar in secondary structure to msDNA, sharing similar stem-loop folding as well as the unique 2',5'-phosphodiester linkage. These results indicate that these novel molecules are synthesized by common molecular mechanisms.  相似文献   

6.
The structurally related ATM (ataxia-telangiectasia-mutated) and ATR (ATM-Rad3-related) protein kinases fulfill overlapping yet non-redundant functions as key regulators of cellular DNA damage responses. We recently showed that ATM phosphorylates the cyclic AMP response element-binding protein, CREB, following exposure to ionizing radiation (IR) and other DNA-damaging stimuli. Here, we show that a phospho-specific antibody recognizing the major ATM phosphorylation site in CREB cross-reacts with SV40 large tumor antigen (LTag), a multifunctional oncoprotein required for replication of the SV40 minichromosome. The relevant IR-induced phosphorylation site in LTag recognized by phospho-CREB antibody was mapped to Ser-120. IR strongly induced the phosphorylation of Ser-120 in an ATM-dependent manner in mouse embryo fibroblasts. Infection of African green monkey CV1 cells with SV40 resulted in the activation of ATM and phosphorylation of LTag and endogenous ATM substrates. Infection-induced LTag phosphorylation correlated with the onset of DNA replication, was ATM-dependent, and peaked when viral DNA levels reached their maximum. SV40 replication in CV1 cells required an intact LTag Ser-120 phosphorylation site and was inhibited following transfection with ATM small interfering RNA suggesting that ATM is required for optimal SV40 replication in primate cells. Our findings uncover a direct link between ATM and SV40 LTag that may have implications for understanding the replication cycle of oncogenic polyoma viruses.  相似文献   

7.
The primary transforming functions of simian virus 40 large T antigen (SV40 LT) are conferred primarily through the binding and inactivation of p53 and the retinoblastoma family members. Normal p53 function requires an association with the CREB binding protein (CBP)/p300 coactivators, and a ternary complex containing SV40 LT, p53, and CBP/p300 has been identified previously. In this report, we have evaluated a secondary function of p53 bound to the SV40 LT complex in mediating the binding of human CBP/p300. We demonstrate that p53 associated with SV40 LT was posttranslationally modified in a manner consistent with the binding of CBP/p300. Furthermore, expression of SV40 LT induced the proportion of p53 phosphorylated on S15. An essential function for p53 in bridging the interaction between SV40 LT and CBP/p300 was identified through the reconstitution of the SV40 LT-CBP/p300 complex upon p53 reexpression in p53-null cells. In addition, the SV40 LT-CBP/p300 complex was disrupted through RNA interference-mediated depletion of endogenous p53. We also demonstrate that SV40 LT was acetylated in a p300- and p53-dependent manner, at least in part through the CH3 domain of p300. Therefore, the binding of p53 serves to modify SV40 LT by targeting CBP and p300 binding to direct the acetylation of SV40 LT.  相似文献   

8.
The simian virus 40 large tumor antigen (SV40 Tag) has been ascribed many functions critical to viral propagation, including binding to the mammalian tumor suppressor p53. Recent studies have demonstrated that SV40-transformed murine cells have functional p53. The status of p53 in SV40-immortalized human cells, however, has not been characterized. We have found that in response to ionizing radiation, p53-dependent p21 transactivation activity is present, albeit reduced, in SV40-immortalized cells and that this activity can be further reduced with either dominant negative p53 expression or higher SV40 Tag expression. Furthermore, overexpression of p53 in SV40-immortalized ataxia-telangiectasia (A-T) cells restores p53-dependent p21 induction to typical A-T levels. All SV40-immortalized cell lines exhibited an absence of G1 arrest. Moreover, all SV40-immortalized cell lines exhibited increased apoptosis relative to primary cells in response to ionizing radiation, suggesting that SV40 immortalization results in a unique phenotype with regard to DNA damage responses.  相似文献   

9.
In marked contrast to simian virus 40 (SV40), polyoma virus (PyV) has been reported to replicate discontinuously on both arms of replication forks. In an effort to clarify the relationship between the mechanisms of DNA replication in these closely related viruses, the distribution of RNA-primed DNA chains at replication forks was examined concurrently in PyV and SV40 replicating DNA purified from virus-infected cells. About one third of PyV DNA chains contained 7 to 9 ribonucleotides covalently linked to their 5'-end. A similar fraction of DNA chains from replicating SV40 DNA contained an oligoribonucleotide that was 6 to 9 residues long and began with either (p)ppA or (p)ppG. Greater than 80% of PyV or SV40 RNA-primed DNA chains hybridized specifically to the retrograde template. Moreover, at least 95% of the RNA-primed DNA chains from either PyV or SV40 whose initiation sites could be mapped to unique nucleotide locations originated from the retrograde template. Therefore, PyV and SV40 DNA replication forks are essentially the same; DNA synthesis is discontinuous predominantly, if not exclusively, on the retrograde template.  相似文献   

10.
A transient decaribonucleotide (iRNA) is covalently linked to nascent eukaryotic DNA chains at their 5' end. Searching for the putative iRNA polymerase (primase), we detected in extracts from SV40-infected cells a DNA-dependent incorporation of UMP residues from UTP into free and DNA linked deca- or similarly sized ribonucleotides. Denatured salmon sperm DNA served as the standard template in this reaction. SV40 FIII DNA was also an effective template, SV40 FII DNA was ineffective while FI yielded mainly free decaribonucleotides. The incorporation depended on the other rNTPs and was resistant to high concentrations of alpha-amanitin and rifamycin AF/013, drugs inhibitory to RNA polymerases I, II and III. The results implicate the decaribonucleotide polymerase in the priming of nascent DNA chains and suggest that the unique size of iRNA is encoded within its primase.  相似文献   

11.
We have analyzed in detail the phosphorylation of p53 from normal (3T3) and simian virus 40 (SV40)-transformed (SV3T3) BALB/c mouse cells and from normal (F111) and SV40-transformed [FR(wt648)] rat cells by two-dimensional tryptic peptide mapping and phosphoamino acid analyses. To accommodate the different half-lives of p53 in normal (half-life, 15 min) and transformed (half-life, 20 h) cells and possible differences in the rates of turnover of phosphate at specific sites, cells were labeled for 2 h (short-term labeling) or 18 h (long-term labeling). Depending on the labeling conditions, either close similarities or marked differences were observed in the phosphorylation patterns of p53 from normal and transformed cells. After the 2-h labeling, the phosphorylation patterns of p53 from normal and transformed mouse cells were quite similar. In contrast, p53 from normal and transformed rat cells exhibited dramatic quantitative and qualitative differences under these labeling conditions. The reverse was found after an 18-h label leading to steady-state phosphorylation of p53 in transformed cells: while p53 in transformed mouse cells revealed a marked quantitative increase in phosphorylation compared with p53 from normal cells, the corresponding patterns of p53 from normal and transformed rat cells were similar. Our data thus indicate species-specific differences in the phosphorylation of mouse and rat p53 in SV40-transformed cells, reflected by (i) different turnover rates at specific sites in mouse and rat p53 and (ii) phosphorylation of nonhomologous serine and threonine residues in rat p53, as revealed by indirect assignment of phosphorylation sites to the phosphopeptides of rat p53. Analyses of p53 from the SV40 tsA58 mutant-transformed F111 cell lines FR(tsA58)A (N type) and FR(tsA58)57 (A type) yielded no conclusive evidence for a direct correlation between phosphorylation of p53, the metabolic stabilization of p53, and expression of the transformed phenotype.  相似文献   

12.
13.
D W Meek  S Simon  U Kikkawa    W Eckhart 《The EMBO journal》1990,9(10):3253-3260
The entire coding sequence of wild-type mouse p53 was expressed in Escherichia coli under control of the PL promoter of bacteriophage lambda. The bacterial p53 protein had identical mobility to p53 from SV3T3 cells on SDS polyacrylamide gels and was recognized in bacterial lysates by three p53-specific monoclonal antibodies, including PAb246 which is specific for wild-type mouse p53. Immunoprecipitates of the bacterial p53 were phosphorylated by a highly purified preparation of rat casein kinase II; the stoichiometry of incorporation was approximately 1 mol of phosphate per mol of p53. The phosphorylated residue was identified by phosphopeptide mapping as serine 389, which is a major site of p53 phosphorylation in vivo. p53 (serine 389) kinase activity was detected on lysates of SV3T3 cells; this activity co-purified with casein kinase II on phosphocellulose and Mono Q columns and was inhibited by heparin. Immunoprecipitates of the p53-T antigen complex from SV3T3 cells also had associated serine 389 kinase activity. Phosphorylation of serine 389 by this kinase was potently inhibited by heparin and quenched by excess unlabelled GTP. The data indicate that p53 is a physiological substrate of casein kinase II, which is stimulated in response to mitogens, phosphorylates nuclear oncoproteins, and may play a role in the transduction of extracellular signals to the nucleus.  相似文献   

14.
HeLa cells infected with the nondefective adenovirus 2 (Ad2)-simian virus 40 (SV40) hybrid viruses (Ad2(+)ND1, Ad2(+)ND2, Ad2(+)ND4, and Ad2(+)ND5) synthesize SV40-specific proteins ranging in size from 28,000 to 100,000 daltons. By analysis of their methionine-containing tryptic peptides, we demonstrated that all these proteins shared common amino acid sequences. Most methionine-containing tryptic peptides derived from proteins of smaller size were contained within the proteins of larger size. Seventeen of the 21 methionine-containing tryptic peptides of the largest SV40-specific protein (100,000 daltons) from Ad2(+)ND4-infected cells were identical to methionine-containing peptides of SV40 T-antigen immunoprecipitated from extracts of SV40-infected cells. All of the methionine-containing tryptic peptides of the Ad2(+)ND4 100,000-dalton protein were found in SV40 T-antigen immunoprecipitated from SV40-transformed cells. All SV40-specific proteins observed in vivo could be synthesized in vitro using the wheat germ cell-free system and SV40-specific RNA from hybrid virus-infected cells that was purified by hybridization to SV40 DNA. As proof of identity, the in vitro products were shown to have methionine-containing tryptic peptides identical to those of their in vivo counterparts. Based on the extensive overlap in amino acid sequence between the SV40-specific proteins from hybrid virus-infected cells and SV40 T-antigen from SV40-infected and -transformed cells, we conclude that at least the major portion of the SV40-specific proteins cannot be Ad2 coded. From the in vitro synthesis experiments with SV40-selected RNA, we further conclude that the SV40-specific proteins must be SV40 coded and not host coded. Since SV40 T-antigen is related to the SV40-specific proteins, it must also be SV40 coded.  相似文献   

15.
We have previously reported that EBNA-5, one of the Epstein–Barr virus-encoded proteins, accumulates in the nuclear bodies containing PML, the promyelocytic leukemia associated protein. In this study, we examine the intranuclear distribution of SV40 large T-antigen (SV40T), the p53 tumor suppressor protein (p53), and PML in a conditionally immortalized cell line, IDH4. In IDH4 cells, the expression of SV40T is regulated by a dexamethasone (Dex)-driven promoter. Withdrawal of Dex results in down-regulation of SV40T and growth arrest, whereas addition of Dex to the growth-arrested cells results in up-regulation of SV40T and proliferation. In proliferating IDH4 cells, SV40T is concentrated in nuclear dots that are also positive for p53. Many of these dots are juxtaposed to PML positive structures but do not colocalize with them. After removal of Dex, SV40T–p53 dots gradually disappear, while the PML structures remain. Induction of SV40T in nonproliferating IDH4 cells causes a coordinated redistribution of SV40T and p53. The immunostaining for SV40T and p53 is first weak, then strong with a homogeneous distribution, and 3–4 days later becomes dot-like again. This reappearance of SV40T–p53 dots coincides with the recovery of proliferation in restimulated IDH4 cells. Also, the p53 pattern correlates with the SV40T pattern with regard to both morphology and intensity during both suppression and induction of SV40T. Taken together, our data suggest that (i) the level of p53 is coregulated with the level of SV40T in a dose-dependent fashion; (ii) the formation of SV40T–p53 nuclear dots correlates with the transformed phenotype; (iii) the SV40T–p53 dots localize preferentially to the neighborhood of PML bodies which are already present in normal cells.  相似文献   

16.
Ca2+/phospholipid-dependent protein kinase (protein kinase C) and trypsin-activated protein kinase C (protein kinase M) phosphorylated the synthetic peptide R1-A13 (Arg-Arg-Leu-Ser-Ser-Leu-Arg-Ala-Ser-Thr-Ser-Lys-Ala) which contains both cAMP- and insulin-regulated phosphorylation sites in rat liver ribosomal protein S6 [Wettenhall, R. E. H. & Morgan, F. J. (1984) J. Biol. Chem. 259, 2084-2091]. Both enzymes showed essentially the same kinetic properties; V and apparent Km were determined to be 0.16 mumol min-1 mg-1 and 30 microM, respectively. At first, tryptic phosphopeptides were prepared at the early stage of phosphorylation and purified by high-performance liquid chromatography (HPLC). Through these analyses, four radioactive peptides were isolated. When protein kinase C was employed, phosphorylation was observed on all four peptides in a Ca2+/phospholipid-dependent manner. Irrespective of the protein kinase employed, phosphate incorporation into these peptides increased linearly with time; the peptide concentration did not affect the ratio of phosphate distribution into these four peptides. Analysis of amino acid composition and phosphoamino acid of radioactive peptides obtained after extensive phosphorylation showed that phosphates were incorporated into Ser-4, Ser-5, Ser-9 and Ser-11. The latter three serine residues were major phosphorylated sites. When rat liver 40-S ribosomal subunits were employed as substrate for protein kinases C and M, a radioactive protein with Mr,app = 31,000, which corresponded to S6 protein, was detected on an autoradiogram of a sodium dodecyl sulfate/polyacrylamide slab gel. The rate of phosphorylation with protein kinase M was twice as fast as that with protein kinase C. The elution profile of radioactive tryptic peptides in HPLC suggest that phosphorylation occurred on the sites in S6 protein corresponding to Ser-5, Ser-9 and Ser-11 as major sites and Ser-4 as the minor one. These results indicate that protein kinase C has an ability to recognize at least four sites derived from hormone-dependent phosphorylation sites in ribosomal protein S6 irrespective of the mode of activation of this enzyme.  相似文献   

17.
Cell-free extracts of simian virus 40 (SV40)-infected CV-1 cells can initiate large tumor antigen dependent bidirectional replication in circular DNA molecules containing a functional SV40 origin of replication (ori). To determine whether or not DNA replication under these conditions involves RNA-primed DNA synthesis, replication was carried out in the presence of 5-mercuri-deoxycytidine triphosphate to label nascent DNA chains. Newly synthesized mercurated DNA was isolated by its affinity for thiol-agarose, and the 5'-ends of the isolated chains were radiolabeled to allow identification of RNA primers. At least 50% of the isolated chains contained 4 to 7 ribonucleotides covalently linked to their 5'-end; 80% of the oligoribonucleotides began with adenosine and 19% began with guanosine. About 60% of the nascent DNA chains annealed to the SV40 ori region, and about 80% of these chains were synthesized in the same direction as early mRNA. These results are consistent with the properties of SV40 DNA replication in vivo and support a model for initiation of SV40 DNA replication in which DNA primase initiates DNA synthesis on that strand of ori that encodes early mRNA.  相似文献   

18.
A new method for the detection and assay of RNA-linked nascent DNA pieces has been developed. The method relies on selective degradation by spleen exonuclease of radioactive 5′-OH terminated DNA produced from the pulse-labelled nascent pieces upon alkaline hydrolysis. Analysis with this method in wild type Escherichia coli has shown relatively high proportions of the RNA-linked molecules after shorter pulses and in the smaller pieces, supporting the transient nature of the RNA attachment to the nascent pieces. The RNA-linked nascent DNA pieces are accumulated by both E. coli polAex1 (defective in 5′ → 3′ exonuclease of DNA polymerase I) and E. coli polA12 and polA1 (defective in polymerase of DNA polymerase I), suggesting the requirement of the concerted action of both 5′ → 3′ exonuclease and polymerase of DNA polymerase I for the removal of the RNA attached to the nascent pieces. Most of the nascent DNA pieces accumulated by E. coli ligts7 (defective in DNA ligase) are not linked to RNA, as expected from the direct role of DNA ligase in joining of the pieces. The analysis also has shown that a large portion of the nascent DNA pieces present in the cell under the normal steady-state conditions are not linked to RNA and that the level of the RNA-free DNA pieces is also increased in polA mutants. These findings suggest that the removal of RNA from the nascent pieces is a relatively rapid process and the joining reaction is a rate-limiting step that requires the concurrent action of DNA polymerase and DNA ligase.  相似文献   

19.
We report here the isolation and identification of the RNA specifically immunoprecipitated and covalently linked to the tumor suppressor gene product p53. After treatment with proteinase K, the sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) band of p53 yields a single, discrete 157-nucleotide RNA, which was cloned, sequenced, and identified as 5.8S rRNA. 5.8S rRNA was obtained only after proteolysis of the p53 SDS-PAGE band. Free 5.8S rRNA did not comigrate with p53 in SDS-PAGE. This RNA was only immunoprecipitated from cells containing p53. Protein-free RNA obtained by proteolysis of the p53 band hybridized to the single-stranded DNA vector containing the antisense sequence of 5.8S rRNA. The covalence of the p53-5.8S rRNA linkage was demonstrated by the following findings: (i) p53 and the linked 5.8S rRNA comigrated in SDS-PAGE; (ii) only after treatment of the p53-RNA complex with proteinase K did the 5.8S rRNA migrate differently from p53-linked 5.8S rRNA; and (iii) this isolated RNA was found linked to phosphoserine, presumably at the 5' end. Covalent linkage to the single, specific RNA suggests that p53 may be involved in regulating the expression or function of 5.8S rRNA.  相似文献   

20.
Wild-type (wt) murine p53 has been tested for its ability to block and reverse the transforming effects of simian virus 40 (SV40) large T antigen. Established and precrisis mouse cells overexpressing exogenously introduced wt p53 became resistant to SV40 transformation. The introduction of excess wt p53 into SV40-transformed precrisis cells reverted their transformed phenotype. However, the phenotype of SV40-transformed established cells was not reverted by excess wt p53. We conclude that an antioncogenic action of wt p53 is exerted during SV40 transformation and that in precrisis cells, the antitransforming action of wt p53 can be exerted both at initiation and during the maintenance of transformation.  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