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 46 毫秒
1.
p53 is a cellular-encoded transformation-related protein. It is synthesized at elevated levels in tumor cells but has also been detected at low concentrations in several types of nontransformed cells. The p53 of tumor cells is immunogenic and elicits specific antibody production. The antigenic determinants of the p53 protein were studied by specific binding to anti-p53 monoclonal antibodies obtained from the RA3-2C2, PAb122, and PAb421 established hybridoma cell lines, and their conservation was followed in various animal species. We found that whereas mouse p53 efficiently immunoprecipitated with all three anti-p53 monoclonal antibodies, human and rat p53 bound PAb122 and PAb421 but lacked a determinant binding RA3-2C2. The hamster p53 molecule represented a third category, which immunoprecipitated with polyclonal anti-p53 antibodies but failed to bind all three monoclonal antibodies analyzed here. Using these monoclonal antibodies, we detected no variations between p53 found in transformed and p53 found in nontransformed cells, within a given species. The results also showed that RA3-2C2, which recognizes a mouse-specific determinant, binds a site located at a proteolytic digestion fragment of the p53 molecule that differs from that containing PAb122 and PAb421 recognition site(s). p53 is a single protein that can be immunoprecipitated through different antigenic determinants that vary between species.  相似文献   

2.
We have examined the effect that microinjection of a monoclonal antibody directed against human DNA polymerase-alpha (SJK-287) has on DNA synthesis in exponentially growing human, mouse, and hamster cell lines. We show that the SJK-287 antibody, when microinjected directly into the nuclei of cells is capable of inhibiting DNA synthesis in all three cell lines tested. Moreover, the effectiveness with which this antibody can inhibit ongoing DNA synthesis by the microinjection assay is closely correlated with the ability of the antibody to neutralize DNA polymerase-alpha activity fractionated from each cell line in vitro. Two other monoclonal antibodies of the same class, one directed against the cellular p53 protein (PAb122), and one directed against the c-myc protein (PM-8) were also tested for their ability to inhibit ongoing DNA synthesis by direct microinjection and in lysolecithin permeabilized cells. Both monoclonal antibodies failed to inhibit ongoing DNA synthesis in exponentially growing cells by these assays.  相似文献   

3.
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   

4.
Transfection of a cloned p53 gene into a p53 nonproducer Abelson murine leukemia virus-transformed cell line, L12, reconstituted p53 expression. The protein expressed in these cells was indistinguishable from that naturally expressed in p53 producer tumor cells. Conversely, p53 protein expressed in L12-derived clones that were established by transfection with a full-length p53 cDNA clone (pM8) exhibited a discrete immunological form. Immunoprecipitation of p53 with a panel of monoclonal anti-p53 antibodies showed that L12-derived clones that were transfected with the genomic p53 clone contained the same antigenic determinants as those found in the p53 protein expressed in tumor cells. These p53 proteins bound all monoclonal antibody types as well as the polyclonal anti-p53 tested. However, L12-derived clones established by transfection of the p53 cDNA clone (pM8) expressed a p53 protein that bound the RA3-2C2 and PAb200.47 anti-p53 monoclonal antibodies as well as polyclonal anti-p53 serum but totally lacked the antigenic receptor for the PAb122 and PAb421 monoclonal antibodies. The p53 proteins expressed by either genomic or cDNA p53 clones exhibited the same apparent molecular sizes and identical partial peptide maps. We suggest that transfection of the p53 gene induced expression of the entire group of the possible mRNA species, whereas cloned p53 cDNA (pM8) represented a single mRNA molecule that codes for a discrete species of p53 protein.  相似文献   

5.
The high level and intranuclear location of the cellular phosphoprotein p53 are usually regarded as invariant features of SV40-transformed fibroblast lines. During the development of improved methods for immunocytochemical detection of p53 using SVA31 E7 mouse fibroblasts, we have observed unexpected and marked variations in its distribution. Cells grown on plastic coverslips were fixed in acetone and the content and distribution of p53 and SV40 large T-antigen analysed by an indirect immunoperoxidase procedure using monoclonal antibodies PAb122 and PAb 248 for p53, and PAb416 for large T. First, we observed in all cultures an apparent reversal of intracellular compartmentalization, with strong cytoplasmic and absent nuclear/chromatin positivity in mitotic cells and young daughter cells. More importantly, for a short period, between 18-24 h after trypsinization and cell passage, we observed a marked overall reduction in detectable nuclear p53 content in all cells with both PAb122 and PAb248 antibodies. The first observation also held for SV40 large T-antigen, the second only for p53. These variations have important practical implications for the immunocytochemical analysis of cellular content and intracellular compartmentalization of p53. The biological implications of our findings are also discussed.  相似文献   

6.
Lepik D  Ustav M 《Journal of virology》2000,74(10):4688-4697
Small DNA tumor viruses like human papillomaviruses, simian virus 40, and adenoviruses modulate the activity of cellular tumor suppressor proteins p53 and/or pRB. These viruses replicate as nuclear multicopy extrachromosomal elements during the S phase of the cell cycle, and it has been suggested that inactivation of p53 and pRb is necessary for directing the cells to the S phase. Mouse polyomavirus (Py), however, modulates only the pRB protein activity without any obvious interference with the action of p53. We show here that Py replication was not suppressed by the p53 protein indeed in all tested different mouse cell lines. In addition, E1- and E2-dependent papillomavirus origin replication was insensitive to the action of p53 in mouse cells. We show that in hamster (Chinese hamster ovary) or human (osteosarcoma 143) cell lines the replication of both Py and papillomavirus origins was efficiently blocked by p53. The block of Py replication in human and hamster cells is not caused by the downregulation of large T-antigen expression. The deletion analysis of the p53 protein shows that the RPA binding, proline-rich regulatory, DNA-binding, and oligomerization domains are necessary for p53 action in both replication systems. These results indicate that in mouse cells the p53 protein could be inactive for the suppression of papovavirus replication.  相似文献   

7.
SV40 large T antigen associates with a cellular phosphoprotein, p53, in virus-transformed cells. We have raised three new monoclonal antibodies, PAb1101, PAb1102 and PAb1103, to this cellular protein, derived from SV40-transformed human fibroblasts. These define at least two non-overlapping determinants on human p53 that are in different areas of the molecule from those recognised by previously available antibodies. Unlike those antibodies, PAb1102 and PAb1103 do not react with rodent p53. PAb1101 reacts far more weakly with rodent p53 than with primate p53. All three antibodies show a preference for binding to the large T-associated form of p53, an effect that is particularly marked with PAb1102. The novel specificity of these antibodies allows further probing of the nature and function of the large T/p53 complex in human cells.  相似文献   

8.
J V Gannon  R Greaves  R Iggo    D P Lane 《The EMBO journal》1990,9(5):1595-1602
Point mutations in the p53 gene are the most frequently identified genetic change in human cancer. They convert murine p53 from a tumour suppressor gene into a dominant transforming oncogene able to immortalize primary cells and bring about full transformation in combination with an activated ras gene. In both the human and murine systems the mutations lie in regions of p53 conserved from man to Xenopus. We have developed a monoclonal antibody to p53 designated PAb240 which does not immunoprecipitate wild type p53. A series of different p53 mutants all react more strongly with PAb240 than with PAb246. The PAb240 reactive form of p53 cannot bind to SV40 large T antigen but does bind to HSP70. In contrast, the PAb246 form binds to T antigen but not to HSP70. PAb240 recognizes all forms of p53 when they are denatured. It reacts with all mammalian p53 and chicken p53 in immunoblots. We propose that immunoprecipitation of p53 by PAb240 is diagnostic of mutation in both murine and human systems and suggest that the different point mutations which convert p53 from a recessive to a dominant oncogene exert a common conformational effect on the protein. This conformational change abolishes T antigen binding and promotes self-oligomerization. These results are consistent with a dominant negative model where mutant p53 protein binds to and neutralizes the activity of p53 in the wild type conformation.  相似文献   

9.
The 11-4 p53 cDNA clone failed to transform primary rat fibroblasts when cotransfected with the ras oncogene. Two linker insertion mutations at amino acid 158 or 215 (of 390 amino acids) activated this p53 cDNA for transformation with ras. These mutant cDNAs produced a p53 protein that lacked an epitope, recognized by monoclonal antibody PAb246 (localized at amino acids 88 to 110 in the protein) and preferentially bound to a heat shock protein, hsc70. In rat cells transformed by a genomic p53 clone plus ras, two populations of p53 proteins were detected, PAb246+ and PAb246-, which did or did not bind to this monoclonal antibody, respectively. The PAb246- p53 preferentially associated with hsc70, and this protein had a half-life 4- to 20-fold longer than free p53 (PAb246+). These data suggest a possible functional role for hsc70 in the transformation process. cDNAs for p53 derived from methylcholanthrene-transformed cells transform rat cells in cooperation with the ras oncogene and produce a protein that bound with the heat shock proteins. Recombinant clones produced between a Meth A cDNA and 11-4 were tested for the ability to transform rat cells. A single amino acid substitution at residue 132 was sufficient to activate the 11-4 p53 cDNA for transformation. These studies have identified a region between amino acids 132 and 215 in the p53 protein which, when mutated, can activate the p53 cDNA. These results also call into question what the correct p53 wild-type sequence is and whether a wild-type p53 gene can transform cells in culture.  相似文献   

10.
The detailed analysis of mRNA structure coding for p53 protein from the intact and KonA stimulated lymphocytes from mouse spleen has shown both matrices to be identical. mRNAs have been analyzed by S1 mapping. Both mRNAs directed in vitro the synthesis of p53 protein reacting with monoclonal antibodies PAb421 and PAb248 both of which are specific for one of p53 forms synthesized by lymphocytes in vivo. Thus, the phenomenon of epitopes exclusion on p53 surface from intact or activated B-lymphocytes might be explained by peculiarities of posttranslational step of protein synthesis.  相似文献   

11.
Irradiation with ultraviolet (UV) triggers programmed cell death (apoptosis) in keratinocytes. This process is believed to protect against skin carcinogenesis since the cells with damaged DNA are selectively removed, limiting the likelihood of the development of a malignant keratinocyte clone. The p53 protein is able to detect mutation-bearing DNA fragments and is thus indispensable for the UV-induced apoptosis in the epidermis. Since age is a risk factor for the development of skin tumors we investigated whether ultraviolet induces apoptosis and p53 activation in senescent keratinocytes. Cultured senescent keratinocytes were irradiated with broad-band ultraviolet, apoptosis was assessed using TUNEL (terminal deoxynucleotidyl transferase-mediated dUTP nick end labelling) technique and the p53 activation pattern was determined with Western blotting and immunofluorescent staining with a panel of anti-p53 antibodies recognising different conformational forms of the protein (PAb 122, PAb 240, DO-7). In senescent keratinocytes arrested in the G1 phase of cell cycle, ultraviolet irradiation (100-2000 J/m2) caused accumulation and nuclear translocation of p53. However, in contrast to young cells where UV induces apoptotic cell death in G1, apoptosis was not detected in senescent cells. There were subtle differences in the p53 activation pattern between senescent keratinocytes and known patterns in young keratinocytes and other cell types. In senescent keratinocytes a constitutional nuclear expression of p53 (conformational form recognized by PAb 240) was present and the p53 induction in response to ultraviolet radiation was rapid. Suppression of apoptosis in senescent keratinocytes may be an important mechanism responsible for enhanced skin carcinogenesis in old age.  相似文献   

12.
13.
In HPV-associated genital lesions, low or absent expression of p53 has been attributed to the rapid degradation of p53 through its binding with HPV E6 protein. In this study, we examined p53 protein expression with two antibodies (CM1 polyclonal and PAb 1801 monoclonal antibodies), and Ki-67 proliferation antigen (monoclonal antibody) using an immunohistochemical (IHC) double-staining technique in 77 HPV-positive cervical lesions (HPV6, HPV11, HPV16, HPV18, HPV31, and HPV33) and in 15 HPV-negative cases. p53 protein expression was detected in 36/92 (39.1%) of the specimens. of the p53-positive cases, 80.6% (29/36) were HPV-positive samples, including 10/23 (43.5%) of HPV16- and 3/10 (30%) of HPV18-positive biopsies. In 52.8% of the p53-positive samples, the expression was found in less than 5% of the basal cells which were also positive for Ki-67.
Ki-67 proliferation marker was found in 91/92 specimens, most intensely in those infected by HPV16. p53 was more abundant in progressive or persistent lesions, but no differences were found between HPV-positive and HPV-negative samples. the positive IHC double-staining of both p53 and Ki-67 proliferation antigen in the same basal (and parabasal) cells indicates that these two normal cell-cycle proteins are being expressed while the cells are entering from the G1 to the S phase of the cell cycle. Since the latter property is only attributed to the wild-type p53 (but not to mutated p53), the p53 protein detected in HPV lesions by IHC is likely to be the wild-type p53 rather than mutated p53, and the result was also confirmed by using p53 mutant specific antibody PAb 240. Accordingly, the concept of HPV inactivating the wild-type p53 protein should be re-examined, and other mechanisms for HPV-mediated carcinogenesis should be considered.  相似文献   

14.
We have developed quantitative radioimmunological solid phase assays for the host protein p53 from mouse cells and from human cells. The first assay, for mouse p53, depends on having two monoclonal antibodies reacting with different determinants on the p53 molecule. With this assay we have shown that SV40-transformed cells have approximately 100-fold more p53 than untransformed mouse cells and that other transformed cells have intermediate levels. Embryonal carcinoma cell lines have approximately 50-fold less p53 than SV40-transformed cells. This is in contrast to the high levels of incorporation of [35S]methionine into p53 in these cells and indicates that metabolic labelling is not a valid approach for measuring p53 levels. The second assay, for human p53, required a different approach and made use of the anti-p53 antibodies detected in the sera of some breast cancer patients. Human tumour cell lines contained amounts of p53 varying from the high level seen in SV40-transformed human fibroblasts down to less than one hundredth of this amount. Normal human cells showed low levels of p53. The data confirm that many, but not all, human tumour cell lines contain more p53 than normal cells.  相似文献   

15.
16.
17.
In human fibroblasts, growth arrest at the end of the normal proliferative life span (induction of senescence) is dependent on the activity of the tumor suppressor protein p53. In contrast, once senescence has been established, it is generally accepted that reinitiation of DNA synthesis requires loss of multiple suppressor pathways, for example, by expression of Simian virus 40 (SV40) large T antigen, and that even this will not induce complete cell cycle traverse. Here we have used microinjection of monoclonal antibodies to the N terminus of p53, PAb1801 and DO-1, to reinvestigate the effect of blocking p53 function in senescent human fibroblasts. Unexpectedly, we found that both antibodies induce senescent cells to reenter S phase almost as efficiently as SV40, accompanied by a reversion to the “young” morphology. Furthermore, this is followed by completion of the cell division cycle, as shown by the appearance of mitoses, and by a four- to fivefold increase in cell number 9 days after injection. Immunofluorescence analysis showed that expression of the p53-inducible cyclin/kinase inhibitor p21sdi1/WAF1 was greatly diminished by targeting p53 with either PAb1801 or DO-1 but remained high and, moreover, still p53 dependent in cells expressing SV40 T antigen. As previously observed for induction, the maintenance of fibroblast senescence therefore appears to be critically dependent on functional p53. We suggest that the previous failure to observe this by using SV40 T-antigen mutants to target p53 was most probably due to incomplete abrogation of p53 function.  相似文献   

18.
Cohen PA  Hupp TR  Lane DP  Daniels DA 《FEBS letters》1999,463(1-2):179-184
In this study, we expressed and purified the p53 mutant encoded by the His175 allele (p53His175) in a baculovirus expression system in order to study the folding and the DNA binding activity of the protein. A two-site ELISA revealed that purified p53His175 protein preferentially displayed a PAb1620 conformation, which appeared to be not sufficient to interact specifically with DNA. The cryptic DNA binding activity of this mutant was then investigated by electrophoretic mobility shift assay in the presence of anti-p53 antibodies, and shown to be refractory to significant activation by PAb421 (a potent allosteric activator of wild-type p53's DNA binding activity). Nevertheless, p53His175 DNA binding was regulated by antibodies targeting the N-terminal region of the protein. Furthermore, while the protein preferentially displayed a PAb1620 conformation, our data suggested the existence of an equilibrium between at least two folding states of the protein (PAb1620 and PAb240 conformations). A model rationalizing the conformation, antibody-interacting ability and DNA binding regulation potential of p53His175 is presented.  相似文献   

19.
In order to study oxidative stress in peripheral cells of Alzheimer's disease (AD) patients, immortalized lymphocytes derived from two peculiar cohorts of patients, referring to early onset AD (EOSAD) and subjects harboured AD related mutation (ADmut), were used. Oxidative stress was evaluated measuring i) the typical oxidative markers, such as HNE Michel adducts, 3 Nitro-Tyrosine residues and protein carbonyl on protein extracts, ii) and the antioxidant capacity, following the enzymatic kinetic of superoxide dismutase (SOD), glutathione peroxidase (GPx) and glutathione reductase (GRD). We found that the signs of oxidative stress, measured as oxidative marker levels, were evident only in ADmut but not in EOSAD patients. However, oxidative imbalance in EOSAD as well as ADmut lymphocytes was underlined by a reduced SOD activity and GRD activity in both pathological groups in comparison with cells derived from healthy subjects. Furthermore, a redox modulated p53 protein was found conformational altered in both EOSAD and ADmut B lymphocytes in comparison with control cells. This conformational altered p53 isoform, named "unfolded p53", was recognized by the use of two specific conformational anti-p53 antibodies. Immunoprecipitation experiments, performed with the monoclonal antibodies PAb1620 (that recognizes p53wt) and PAb240 (that is direct towards unfolded p53), and followed by the immunoblotting with anti-4-hydroxynonenal (HNE) and anti- 3-nitrotyrosine (3NT) antibodies, showed a preferential increase of nitrated tyrosine residues in unfolded p53 isoform comparing to p53 wt protein, in both ADmut and EOSAD. In addition, a correlation between unfolded p53 and SOD activity was further found. Thus this study suggests that ROS/RNS contributed to change of p53 tertiary structure and that unfolded p53 can be considered as an early marker of oxidative imbalance in these patients.  相似文献   

20.
An expression vector utilizing the enhancer and promoter region of the simian virus 40 (SV40) DNA regulating a murine p53 cDNA clone was constructed. The vector produced murine p53 protein in monkey cells identified by five different monoclonal antibodies, three of which were specific for the murine form of p53. The murine p53 produced in monkey cells formed an oligomeric protein complex with the SV40 large tumor antigen. A large number of deletion mutations, in-frame linker insertion mutations, and linker insertion mutations resulting in a frameshift mutation were constructed in the cDNA coding portion of the p53 protein expression vector. The wild-type and mutant p53 cDNA vectors were expressed in monkey cells producing the SV40 large T antigen. The conformation and levels of p53 protein and its ability to form protein complexes with the SV40 T antigen were determined by using five different monoclonal antibodies with quite distinct epitope recognition sites. Insertion mutations between amino acid residues 123 and 215 (of a total of 390 amino acids) eliminated the ability of murine p53 to bind to the SV40 large T antigen. Deletion (at amino acids 11 through 33) and insertion mutations (amino acids 222 through 344) located on either side of this T-antigen-binding protein domain produced a murine p53 protein that bound to the SV40 large T antigen. The same five insertion mutations that failed to bind with the SV40 large T antigen also failed to react with a specific monoclonal antibody, PAb246. In contrast, six additional deletion and insertion mutations that produced p53 protein that did bind with T antigen were each recognized by PAb246. The proposed epitope for PAb246 has been mapped adjacent (amino acids 88 through 109) to the T-antigen-binding domain (amino acids 123 through 215) localized by the mutations mapped in this study. Finally, some insertion mutations that produced a protein that failed to bind to the SV40 T antigen appeared to have an enhanced ability to complex with a 68-kilodalton cellular protein in monkey cells.  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