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 0 毫秒
1.
An S phase-related nuclear 55K antigen, also designated PSL, has been characterized in various mammalian cells, using a human serum from a patient with autoimmune disorders (Barque et al., EMBO j 2 (1983) 743). In this report, we show by immunoelectron microscopy that the p55 protein associates in situ with the chromatin of rat hepatocytes. This association is a transient one, as indirect immunofluorescence studies show that PSL does not bind to individualized metaphase chromosomes. Furthermore, immunoprecipitation tests indicate that the majority of PSL is in the non-chromosomal cell fraction. These results suggest that this nuclear antigen is directly involved in the DNA replication process.  相似文献   

2.
PSL(p55) is a nuclear 55kD antigen present in various mammalian cell systems, which has been first identified by use of human autoimmune antibodies (Barque et al. 1983, EMBO J. 2, 743). It has been shown to be associated with interphase chromatine and to be synthesized in during the S phase of the cell cycle. In this work, we have analysed the status of PSL in promyelocytic HL-60 human cells in exponential or stationary growth, or undergoing granulocytic differentiation in presence of Retinoic acid. By use of 2-dimensional electrophoresis, PSL was found to be composed of two acidic proteins designated p55A and p55B. Unexpectedly, estimated 10-20 fold higher amounts of each species were found in cells treated for 5 days with 10(-6)M Retinoic acid, than in asynchronously growing cells or resting cells. Moreover, the p55A protein was phosphorylated during the process. On the basis of these results, PSL appears to be involved in some steps of the granulocytic differentiation process.  相似文献   

3.
Salt-extracted proteins of taxol-stabilized microtubules from Chinese hamster ovary cells arrested at mitosis were used to immunize mice for hybridoma production. From a group of related monoclonal antibodies (MAbs), one, C9, recognized an epitope on antigens localized by immunofluorescence microscopy to interphase centrosomes and nuclei. The availability of the nuclear antigen was cell cycle-dependent; however, permeabilization of cells before fixation revealed that the antigen was present throughout the cell cycle. The nuclear antigen was exposed during prophase and was released from the nucleus upon nuclear envelope breakdown filling the cytoplasm of the mitotic cell. Antigenic material re-accumulated at daughter nuclei and was concealed during G1 phase. Detergent extraction of the cytoplasmic antigen from mitotic cells enabled localization of antigens to centrosomes, kinetochores, and the furrowing region/midbody. Immunoblot analysis of cells of a variety of species of origin identified an approximate 250 kD polypeptide as corresponding to the nuclear antigen, whereas polypeptides of 107/117 kD as well as approximately 250 kD accounted for the mitotic cytoplasmic antigens. No polypeptides could be associated with antigens at centrosomes, kinetochores, or midbodies. This MAb joins the antibody preparations previously reported that describe nuclear antigens, or epitopes on antigens, enhanced at mitosis.  相似文献   

4.
The basic carboxy terminus of p53 plays an important role in directing the protein into the nuclear compartment. The C terminus of the p53 molecule contains a cluster of several nuclear localization signals (NLSs) that mediate the migration of the protein into the cell nucleus. NLSI, the most active domain, is highly conserved in genetically diverged species and shares perfect homology with consensus NLS sequences found in other nuclear proteins. The other two NLSs, II and III, appear to be less effective and less conserved. Although nuclear localization is dictated primarily by the NLSs inherent in the primary amino acid sequence, the actual nuclear homing can be modified by interactions with other proteins expressed in the cell. Comparison between wild-type p53 and naturally occurring mutant p53 showed that both protein categories could migrate into the nucleus of rat primary embryonic fibroblasts by essentially similar mechanisms. Nuclear localization of both proteins was totally dependent on the existence of functional NLS domains. In COS cells, however, we found that NLS-deprived wild-type p53 molecules could migrate into the nucleus by complexing with another nuclear protein, simian virus 40 large-T antigen. Wild-type and mutant p53 proteins differentially complexed with viral or cellular proteins, which may significantly affect the ultimate compartmentalization of p53 in the cell; this finding suggests that the actual subcellular compartmentalization of proteins may differ in various cell type milieux and may largely be affected by the ability of these proteins to complex with other proteins expressed in the cell. Experiments designed to test the physiological significance of p53 subcellular localization indicated that nuclear localization of mutant p53 is essential for this protein to enhance the process of malignant transformation of partially transformed cells, suggesting that p53 functions within the cell nucleus.  相似文献   

5.
To determine functional subcellular loci of p53, a cellular protein associated with cellular transformation, we analyzed the nucleoplasmic, chromatin, and nuclear matrix fractions from normal mouse 3T3 cells, from methylcholanthren-transformed mouse (MethA) cells, and from various simian virus 40 (SV40)-transformed cells for the presence of p53. In 3T3 and MethA cells, p53 was present in all nuclear subfractions, suggesting an association of p53 with different structural components of the nucleus. In 3T3 cells, p53 was rapidly turned over, whereas in MethA cells, p53 was metabolically stable. In SV40-transformed cells, p53 complexed to large tumor antigen (large T) was found in the nucleoplasmic and nuclear matrix fractions, as described previously (M. Staufenbiel and W. Deppert, Cell 33:173-181, 1983). In addition, however, metabolically stable p53 not complexed to large T (free p53) was also found in the chromatin and nuclear matrix fractions of these cells. This free p53 did not arise by dissociation of large T-p53 complexes, suggesting that stabilization of p53 in SV40-transformed cells can also occur by means other than formation of a complex with large T.  相似文献   

6.
Five monoclonal antibodies which are reactive to early antigens of Epstein-Barr virus have been produced by using somatic cell hybridization techniques. The specificity of the monoclonal antibodies to early antigens was demonstrated by indirect immunofluorescence, which showed that the antigens were localized to the nucleus of early antigen-induced Raji cells. Additional indirect immunofluorescence studies showed that like patient antisera to diffuse-staining early antigen, the monoclonal antibodies gave positive staining reactions after methanol fixation. One of the antibodies, 1150-4, was positive by the anti-complement immunofluorescence technique but differed with Epstein-Barr virus-associated nuclear antigen-positive patient sera in that it only stained induced cells. Different fixation methods were found to alter dramatically the appearance of the nuclear staining reactions produced by the monoclonal antibodies. Immunoprecipitation and immunoblot experiments revealed that monoclonal antibodies 1108-1 and 1129-1 recognized two polypeptides of 55,000 and 50,000 daltons (p55;50), 1173-6 and 1180-2 recognized just p50, and 1150-4 identified a 65,000-dalton nuclear protein. Immunobiochemical characterization of these viral antigens showed that p55 is a phosphoprotein, and p55;50 has strong DNA-binding activity preferentially to single-stranded DNA. Elucidation of the role of these nuclear proteins in Epstein-Barr virus infection and the events associated with Epstein-Barr virus-directed lymphocyte transformation may provide significant information on the pathogenicity of this important human virus.  相似文献   

7.
A new radioimmunoassay for p53, employing an anti-peptide antibody directed against conserved Domain V, exhibited specificity for a relatively dephosphorylated form of p53. This form, correlated with the monoclonal antibody PAB421+ conformation, appeared transiently in the cytosol of cycloheximide-treated T cells undergoing activation by concanavalin-A/serum. Concurrently, there were decreased levels of p53 in the nucleus that correlated with increased phosphorylation of p53. After 90 min nuclear levels of p53 increased steadily above levels of unstimulated cells. Anti-p53 antibodies introduced into cells prior to stimulation enhanced cell proliferation in response to mitogens.  相似文献   

8.
9.
We have characterized a new antibody specificity in a panel of sera from dogs developing systemic lupus erythematosus (SLE) or clinically related autoimmune disorders. This antibody stains in a speckled fashion the nucleus of cells of different mammalian origins. The target antigen is a basic (pI 9.2) nuclear polypeptide with an apparent molecular weight of 43 kDa (p43) which is detected in various mammalian cell nuclei. p43, as studied in HeLa cells, appears to be cell cycle-independent. It is released from nuclei by salts (0.5 M NaCl or 0.25 M ammonium sulfate). Upon subfractionation of nuclear components, p43 is found in the fraction containing HnRNPs and is recovered in immunoprecipitates obtained with 4F4 monoclonal antibody to HnRNP C proteins. Immunoelectron microscopy revealed that p43 is concentrated over the dense chromatin periphery and interchromatin granule clusters. Another important feature of p43 is its ability to specifically bind wheat germ agglutinin lectin but not concanavalin A nor Ulex europaeus I, supporting the notion that p43 is a glycoprotein bearing an N-acetyl-glucosamine moiety. Consistent with this result, a radio-active p43 band is specifically immunoprecipitated by canine anti-p43 autoantibodies from HeLa cells metabolically labeled with [14C]glucosamine. Finally, anti-p43 antibodies do not immunoprecipitate SnRNA, indicating that p43 has no apparent association with SnRNPs.  相似文献   

10.
J Laffin  D Fogleman  J M Lehman 《Cytometry》1989,10(2):205-213
Human diploid fibroblasts (HDF) have a finite life span in cell culture which can be extended when transformed with simian virus 40 (SV40). Flow cytometric analysis of SV40-HDF transformation allowed DNA content changes to be correlated with the appearance, quantity, and distribution of T antigen, p53, and V antigen, three proteins associated with this process. These studies demonstrated a shift in the DNA content to tetraploidy, which was correlated with the age of the SV40-HDF but not the time of infection. A significant increase of the epitope recognized by PAb122 to host p53 and the epitope PAb101 to SV40 T antigen occurred at the same time the tetraploid population appeared. However, an antigen reactive with SV40 V antibody was present at high levels in most of the population early after infection, but the levels declined with time. The percentage of PAb101-T antigen-positive cells increased more rapidly in cells infected at a late passage, and this was concomitant with the shift in DNA content to tetraploid. Analysis of the mean fluorescence of total, gated populations (G1, G2, and greater than G2) demonstrated that a threshold level of p53 and T antigen was reached in each compartment of the cell cycle. As the transformed phenotype appeared, a population of cells was continually released into the supernatant, and although these cells had a DNA pattern similar to the monolayer cells, the T antigen and p53 levels were 3-5 times higher in the tetraploid G2 cells. These studies correlated the expression of proteins associated with viral transformation in HDF which vary with time and shift in DNA content.  相似文献   

11.
p53 regulates apoptosis and the cell cycle through actions in the nucleus and cytoplasm. Altering the subcellular localization of p53 can alter its biological function. Therefore, small molecules that change the localization of p53 would be useful chemical probes to understand the influence of subcellular localization on the function of p53. To identify such molecules, a high-content screen for compounds that increased the localization of p53 to the nucleus or cytoplasm was developed, automated, and conducted. With this image-based assay, we identified ellipticine that increased the nuclear localization of GFP-mutant p53 protein but not GFP alone in Saos-2 osteosarcoma cells. In addition, ellipticine increased the nuclear localization of endogenous p53 in HCT116 colon cancer cells with a resultant increase in the transactivation of the p21 promoter. Increased nuclear p53 after ellipticine treatment was not associated with an increase in DNA double stranded breaks, indicating that ellipticine shifts p53 to the nucleus through a mechanism independent of DNA damage. Thus, a chemical biology approach has identified a molecule that shifts the localization of p53 and enhances its nuclear activity. Electronic supplementary material The online version of this article (doi:) contains supplementary material, which is available to authorized users. G. Wei Xu and Imtiaz A. Mawji have contributed equally to this work.  相似文献   

12.
The intracellular location and kinetics at which the simian virus 40 T antigen and the cellular protein p53 associate with one another were determined for simian virus 40-transformed mouse (215) and rat (14B) cells. Cells were labeled under pulse-chase conditions and fractionated into nuclear and cytoplasmic components, and the proteins were immunoprecipitated with monoclonal antibodies (pAb 416, 101, and 122). We found that newly made T antigen and p53 migrated to the nucleus of these cells independently; that is, in uncomplexed form. Newly made p53 was transported to the nucleus more rapidly than T antigen in both cell lines and formed a complex with a mature form of T antigen recognizable by pAb 101. This association was very rapid in both cell lines (t 1/2, 5 to 15 min). In contrast, the time course of complex formation between newly made T antigen and the p53 in the nucleus varied with the ratio of T antigen to p53 of the cell line studied. In 215 cells, where the ratio was 3.6, the kinetics were quite slow (t 1/2, 30 min), whereas in 14B cells, where the ratio was 1.7, they were quite rapid (t 1/2, 5 min). We suggest that a competition between newly made and uncomplexed T antigen for the p53 in the nucleus is the major determinant of the rate of complex formation for newly made T antigen. Our studies indicate that this macromolecular interaction is extremely dynamic.  相似文献   

13.
This report describes the preliminary characterization of a novel antigen reactive with a murine monoclonal antibody designated B1N produced in our laboratory. This antibody (IgM) reacts in IFI with mammals and also insect cells, by staining in a speckled fashion the nucleus of these cells. Immunoblotting analysis of Hela and murine D55 nuclear extracts revealed a polypeptide with an apparent molecular weight of 120kD (p120). In this work we demonstrated that: 1. this polypeptide appeared in human peripheral blood lymphocytes only when they were induced to proliferate in vitro after phytohemagglutinin stimulation; 2. this polypeptide was no longer detected in D55 resting cells, following serum deprivation; 3. the MAb B1N specifically revealed the nucleus of proliferating cells on frozen sections of uterine tissue. These data strongly suggest that the p120 nuclear antigen expression is associated with the proliferation state of cells.  相似文献   

14.
15.
O M?rkve  J H?stmark 《Cytometry》1991,12(7):622-627
In a series of bronchial and bladder carcinomas, p53 protein expression was examined. Samples from formalin-fixed, paraffin-embedded tissue (routine-treated) were compared with parallel samples of fresh tissue and tissue fixed in paraformaldehyde and ethanol. The expression of p53 was measured by immunofluorescence staining and dual parameter flow cytometry, with simultaneous monitoring of DNA content. For each tumor, p53 fluorescence with different fixatives was expressed relative to fresh tissue. The p53 fluorescence signals were on average brighter from routine-treated tissue than from fresh tissue. The tissue fixed in paraformaldehyde showed no difference from fresh tissue. In the ethanol-fixed tissue, however, fluorescence signals were reduced by nearly 70%, and the fraction of detectable p53 positive cells in tumor tissue was reduced by more than 50%. This loss of fluorescence was probably due to a leakage of the antigen from nucleus to cytoplasm. Pepsin treatment did not influence p53 fluorescence. Within the same tumor, the S-phase fraction in p53 positive cells was significantly higher than in p53 negative cells (13.1 +/- 2.0% vs. 6.5 +/- 0.8%). This pattern was not influenced by formalin fixation or pepsin treatment. Our study demonstrates that in measuring a nuclear antigen, tissue handling may influence the results, and care should be taken to optimize the preparation procedure. Using the antibody PAb 1801, p53 expression measured in archival material is not reduced as compared to fresh tissue.  相似文献   

16.
The cellular phosphoprotein p53 binds tightly and specifically to simian virus 40 T antigen and the 58,000-molecular-weight adenovirus E1b protein. Many human and murine tumor cell lines contain elevated levels of the p53 protein even in the absence of these associated viral proteins. Recently the cloned p53 gene, linked to strong viral promoters, has been shown to complement activated ras genes in transformation of primary rodent cell cultures. Overexpression of the p53 gene alone rescues some primary rodent cell cultures from senescence. We isolated three new monoclonal antibodies to the p53 protein, designated PAb242, PAb246, and PAb248, and mapped the epitopes they recognized on p53 in comparison with other previously isolated antibodies. At least five sterically separate epitopes were defined on murine p53. One of the antibodies, PAb246, recognizes an epitope on p53 that is unstable in the absence of bound simian virus 40 T antigen. This effect is demonstrable in vivo and in newly developed in vitro assays of T-p53 complex formation. Using the panel of anti-p53 antibodies and sensitive immunocytochemical methods, we found that p53 has a predominantly nuclear location in established but not transformed cells as well as in the vast majority of transformed cell lines. Several monoclonal antibodies to p53 showed cross-reactions with non-p53 components in immunocytochemical staining.  相似文献   

17.
Cell transformation in vivo seems to be a multistep process. In in vitro studies certain combinations of two oncogenes, a cytoplasmic gene product together with a nuclear gene product, are sufficient to transform primary rodent cells. Polyoma virus large T antigen can immortalize and, in cooperation with polyoma virus middle T antigen, transform primary cells. On the other hand mutant mouse p53 can also immortalize and, in cooperation with an activated Ha-ras oncogene, transform primary cells. In the present study we analyzed whether mutant p53 can replace polyoma virus large T antigen in a cell transformation assay with polyoma virus middle T antigen. Transfection of mutant p53 alone resulted in a cell line which had retained the actin cable network, grew poorly in medium with low concentration of serum, and failed to grow in semisolid agar. Cotransfection of mutant p53 together with polyoma virus middle T led to cells which grew in medium containing low serum concentration, grew well in semisolid agar, and displayed an altered morphology with the tendency to overgrow the normal monolayer. By these criteria these cells were considered fully transformed. The rate of p53 synthesis was similar in both cell lines. However, only p53 from the transformed cell line turned out to be stable. Cells transformed by mutant p53 and polyoma virus middle T expressed nearly the same amount of the c-src-encoded pp60c-src protein as cells transformed by the same p53 and cotransfected activated Ha-ras oncogene. However, only the polyoma virus middle T/p53-transformed cells exhibited an elevated level of pp60c-src-specific tyrosine kinase activity. Thus, despite different mechanisms leading to cell transformation, mutant p53 can replace polyoma virus large T antigen and polyoma virus middle T can replace the activated Ha-ras oncogene in cell transformation.  相似文献   

18.
The antigenic structure of simian virus 40 (SV40) large tumor antigen (T-ag) in the plasma membranes of SV40-transformed mouse cells and SV40-infected monkey cells was characterized as a step toward defining possible biological function(s). Wild-type SV40, as well as a deletion mutant of SV40 (dl1263) which codes for a truncated T-ag with an altered carboxy terminus, was used to infect permissive cells. Members of a series of monoclonal antibodies directed against antigenic determinants on either the amino or the carboxy terminus of the T-ag polypeptide were able to precipitate surface T-ag (as well as nuclear T-ag) from both SV40-transformed and SV40-infected cells. Cellular protein p53 was coprecipitated with T-ag by all T-ag-reactive reagents from the surface and nucleus of SV40-transformed cells. In contrast, T-ag, but not T-ag-p53 complex, was recovered from the surface of SV40-infected cells. These results confirm that nuclear T-ag and surface T-ag are highly related molecules and that a complex of SV40 T-ag and p53 is present at the surface of SV40-transformed cells. Detectable levels of such a complex do not appear to be present on SV40-infected cells. Both the carboxy and amino termini of T-ag are exposed on the surfaces of SV40-transformed and -infected cells. The possible relevance of the presence of a T-ag-p53 complex on the surface of SV40-transformed cells and its absence from SV40-infected cells is considered.  相似文献   

19.
20.
Disruption of cell cycle regulation is one mechanism proposed for how nuclear envelope protein mutation can cause disease. Thus far only a few nuclear envelope proteins have been tested/found to affect cell cycle progression: to identify others, 39 novel nuclear envelope transmembrane proteins were screened for their ability to alter flow cytometry cell cycle/DNA content profiles when exogenously expressed. Eight had notable effects with seven increasing and one decreasing the 4N∶2N ratio. We subsequently focused on NET4/Tmem53 that lost its effects in p53−/− cells and retinoblastoma protein-deficient cells. NET4/TMEM53 knockdown by siRNA altered flow cytometry cell cycle/DNA content profiles in a similar way as overexpression. NET4/TMEM53 knockdown did not affect total retinoblastoma protein levels, unlike nuclear envelope-associated proteins Lamin A and LAP2α. However, a decrease in phosphorylated retinoblastoma protein was observed along with a doubling of p53 levels and a 7-fold increase in p21. Consequently cells withdrew from the cell cycle, which was confirmed in MRC5 cells by a drop in the percentage of cells expressing Ki-67 antigen and an increase in the number of cells stained for ß-galactosidase. The ß-galactosidase upregulation suggests that cells become prematurely senescent. Finally, the changes in retinoblastoma protein, p53, and p21 resulting from loss of NET4/Tmem53 were dependent upon active p38 MAP kinase. The finding that roughly a fifth of nuclear envelope transmembrane proteins screened yielded alterations in flow cytometry cell cycle/DNA content profiles suggests a much greater influence of the nuclear envelope on the cell cycle than is widely held.  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