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 15 毫秒
1.
Xu N  Shiraki T  Yamada T  Nakajima M  Gauthier JM  Pfeiffer CJ  Sato S 《Gene》2002,288(1-2):159-166
The cDNA (DNA complementary to RNA) of the p53 gene of the beluga whale (Delphinapterus leucas) was sequenced by the method of 5'- and 3'-rapid amplification of cDNA ends (RACE) with the cDNA made for the RNA obtained from fresh peripheral blood leukocytes isolated from two animals. Primers for the RACE method were synthesized based on the sequence of the DNA of beluga whale corresponding to exon 5 of the human p53 gene, which was determined after amplification of the DNA isolated from the liver from a beluga whale by using a pair of primers for the human sequence. The sequenced cDNA had a 2150-nucleotide length and contained the whole region corresponding to human exons 1 through 11. The reading frame was 1164 bp (base pair) long and began in exon 2 and ended in exon 11, coding for a 387-amino acid protein. The nucleotide sequence of the reading frame showed high similarity over 85% with pig, sheep, cow, and human genes. The similarities with the former two animals at the amino acid level were also more than 85%. Lower similarity of the beluga whale p53 gene was also found with those of lower tetrapods, fish and invertebrates.  相似文献   

2.
The p53-activated gene PAG608, which encodes a nuclear zinc finger protein, is a p53-inducible gene that contributes to p53-mediated apoptosis. However, the mechanisms by which PAG608 is involved in the apoptosis of neuronal cells are still obscure. In this study, we demonstrated that expression of p53 was induced by 100 microm 6-hydroxydopamine (6-OHDA), accompanied by increased PAG608 expression in PC12 cells. On the other hand, transient or permanent transfection of antisense PAG608 cDNA into PC12 cells significantly prevented apoptotic cell death induced by 100 microm 6-OHDA or 200 microm hydrogen peroxide but not by 250 microm 1-methyl-4-phenylpyridinium ion. The 6-OHDA-induced activation of caspase-3, DNA fragmentation, loss of mitochondrial membrane potential, and induction of p53 and Bax were also prevented in PC12 cells that stably expressed antisense PAG608 cDNA. These results suggest that PAG608 is associated with the apoptotic pathway induced by these oxidative stress-generating reagents, upstream of the collapse in the mitochondrial membrane potential in PC12 cells. Interestingly, transient transfection with PAG608 cDNA increased p53 expression in both PC12 cells and B65 cells, indicating that PAG608 induced by p53 is able to induce p53 expression in these cells inversely. Furthermore, transient transfection of a truncated mutant PAG608 cDNA, lacking the first zinc finger domain, inhibited 6-OHDA-induced cell death and altered the nuclear and nucleolar localization of wild-type PAG608 in PC12 cells. These results suggest that PAG608 may induce or regulate p53 expression and translocate to the nucleus and nucleolus using its first zinc finger domain during oxidative stress-induced apoptosis of catecholamine-containing cells.  相似文献   

3.
The expression of 1B6, a growth-regulated sequence isolated from a Syrian hamster fibroblast cDNA library, was studied in BALB/c 3T3 cells. The level of cytoplasmic 1B6 mRNA (1600 bases) was low in quiescent cells and plateaued in mid/late G1 after the cells were stimulated with 15% fetal calf serum (FCS). Protein synthesis was not required for the induction of 1B6 mRNA; therefore, the expression of 1B6 is a primary response to serum stimulation. The induction of 1B6 mRNA was also observed after stimulation with insulin, epidermal growth factor, and fibroblast growth factor but not with platelet-derived growth factor. When quiescent cells were serum-stimulated, the percentage of cells that became committed to enter DNA synthesis was proportional to the length of their incubation with serum. To determine if 1B6 expression was also correlated with the time of exposure to serum, quiescent cells were stimulated with a pulse of 15% FCS and the abundance level of 1B6 induced by that pulse was determined. The amount of 1B6 mRNA increased with increasing time of exposure to serum and paralleled the increase in the percentage of nuclei that were induced into DNA synthesis by the serum pulse. Comparison of the nucleotide sequence of the p1B6 cDNA to the GenBank database revealed a striking identity of 1B6 to the 3' end of p36, the heavy chain of calpactin I. The previous characterization of p36 as a substrate for tyrosine kinases suggests a possible role for 1B6/p36 in cell proliferation.  相似文献   

4.
5.
We isolated and sequenced mouse lipocortin I cDNA clones from a lambda gt10 cDNA library prepared from Swiss 3T3 mRNA. The homology with human lipocortin I at the amino acid level is 86%. When confluent layers of Swiss 3T3 cells were stimulated with 10% fetal calf serum, expression of lipocortin I was strongly stimulated. In parallel, DNA synthesis was induced with a peak at 24 hours after glucocorticoid treatment indicating induction of cell proliferation. In the absence of serum glucocorticoid treatment provoked neither induction of DNA synthesis nor expression of lipocortin I. We conclude that serum contains an unidentified factor, which acts synergistically with glucocorticoids on cell proliferation and lipocortin I expression.  相似文献   

6.

Background

PPP1R13L gene has been found to be over-expressed in variety of cancers and its expression in p53 wild-type background is sufficient to promote tumor growth in vivo. However, in the non-transformed cells it acts as a tumor suppressor which suggests that the role of PPP1R13L is multifaceted.

Methods

We have used siRNA optimized for inhibition of p53, PPP1R13L, BAX and GADD45 alpha expression and investigated the role of those gene products for PPP1R13L expression and induction in a variety of mouse and human cells with different p53 status. In addition we have applied Western Blot, Q-PCR and proteasome inhibition analysis to further ascertain the link between PPP1R13L induction and p53 status.

Results

We show that the pattern and extent of the PPP1R13L expression depend on the presence of active p53. Downregulation of p53 target genes BAX and/or GADD45 alpha led to decreased in PPP1R13L activation after adriamycin and/or etoposide treatments. Treatment of the cells with the proteasome inhibitor MG-132 resulted in the accumulation of both p53 and PPP1R13L proteins.

Conclusions

We have provided evidence that endogenous PPP1R13L acts as a negative regulator of p53 function, presumably by direct binding. p53 accumulation and activity after DNA damage is compromised by PPP1R13L expression. We suggest that PPP1R13L and p53 form a negative feedback loop which regulates their amount and activity.

General significance

The profound modulatory effect of the PPP1R13L protein on the ability of p53 to cause cellular apoptosis has important implications in cancer and presents new therapeutic possibilities.  相似文献   

7.
We have investigated the role of the cellular p53 protein in the induction of growth in size and cell DNA replication in human peripheral blood lymphocytes (PBL) and in monocyte/macrophage-depleted lymphocyte (MDL) cultures stimulated with phytohemagglutinin (PHA). Our results show that in human lymphocytes exposed to PHA, the induction of p53 protein synthesis and accumulation correlates with the extent of cellular DNA replication, rather than with growth in size. Moreover, the induction of p53 is dependent on the presence of the T-cell mitogen, Interleukin-2. A monoclonal antibody to Interleukin-2 receptors (anti-Tac) inhibits PHA-stimulated cellular DNA synthesis, and this inhibition is correlated with a reduction in the percentage of p53-positive cells. We conclude from this work that the p53 protein is a cell cycle-dependent gene whose expression can be regulated by different mitogens in different cell types.  相似文献   

8.
Previous experiments have brought into question which amino acid sequence of the p53 oncogene product should be considered wild type and whether the normal protein is capable of cooperating with the ras oncogene to transform cells in culture. To address these questions, a series of p53 cDNA-genomic hybrid clones have been compared for the ability to cooperate with the ras oncogene in transformation assays. From these experiments, it has become clear that the amino acid alanine at position 135, in either the genomic clone or the cDNA clone, failed to produce a p53 protein that cooperated with the ras oncogene and transformed cells. Replacing alanine with valine at this position in either the genomic or the cDNA clone activated for transformation in this assay. Using restriction enzyme polymorphisms in the p53 gene, it was shown that normal mouse DNA encodes alanine at position 135 in the p53 protein. Thus, mutation is required to activate the p53 protein for cooperation with the ras oncogene. After cotransfection with the activated ras gene, the genomic p53 DNA clone always produced more transformed cell foci (1.7-fold) than similar cDNA clones and these foci were more readily cloned (3.6-fold) into permanent cell lines. A series of deletion mutants of the genomic p53 clone were employed to show that the presence of intron 4 in the p53 gene was sufficient to provide much enhanced clonability of transformed foci from culture dishes. The presence of introns in the p53 gene constructions also resulted in elevated levels of p53 protein in the p53-plus-ras-transformed cell lines. Thus, qualitative changes in the p53 protein are required to activate p53 for transformation with the oncogene ras. Quantitative improvements of transformation frequencies are associated with the higher expression levels of altered p53 protein that are provided by having one of the p53 introns in the transforming plasmid.  相似文献   

9.
The p53 null HL-60 cell line was transfected with plasmids coding for either the wild-type p53 or mutant p53 gene. The stable expression of wild-type p53 resulted in a significant increase in sensitivity to the topoisomerase II poisons etoposide and doxorubicin, but not to the topoisomerase II inhibitors razoxane and ADR-529. HL-60 cells expressing wild-type p53 demonstrated 8- to 10-fold more VP-16 induced DNA breaks by the alkaline elution assay. The effect of inducible expression of wild-type p53 was also studied in the p53 null erythroblastoid cell line K562 and in the human squamous carcinoma cell line SqCC. The inducible expression of wild-type p53 in the K562 cell line resulted in a 3-fold increase in sensitivity to VP-16. The quantity of topoisomerase IIalpha was not altered by the transfection as determined by immunoblotting, while the amount of the beta isoform was increased 2.5-fold in HL-60 cells. The topo II catalytic activity present in nuclear extracts was measured as the decatenation of kinetoplast DNA, and found to be unaltered by p53 expression. Immunostaining for topoisomerase IIalpha was substantially diminished in both stable and inducible wild-type p53 expressing cells when three different antibodies were used (two polyclonal and one monoclonal). However, the addition of VP-16 resulted in a rapid appearance of nuclear fluorescence for topoisomerase IIalpha. No changes in topoisomerase IIbeta immunostaining were observed. These results suggest that an epitope for topoisomerase IIalpha is concealed in cells expressing wild-type p53 and that a complex between topoisomerase IIalpha and p53 may be disrupted by the addition of antitumor drugs.  相似文献   

10.
M Oren  B Bienz  D Givol  G Rechavi    R Zakut 《The EMBO journal》1983,2(10):1633-1639
Three cDNA clones, corresponding to two non-overlapping regions of the mRNA coding for the mouse p53 cellular tumor antigen, were isolated and characterized. In hybridization-selection assays, these clones were capable of selectively binding p53 mRNA, as demonstrated by in vitro translation and immunoprecipitation with anti-p53 monoclonal antibodies. The p53 mRNA appeared to be the only messenger species specifically selected by these clones. The size of the p53 mRNA was found to be approximately 2 kb, and its levels to vary substantially among different types of transformed cells. Evidence was found for the existence of two distinct p53-specific genes in mouse genomic DNA. Two partially overlapping recombinant phage clones were obtained, both derived from the same p53-specific genomic DNA region. The orientation of the various cDNA clones relative to that of the p53 mRNA was established by S1 analysis and the relationship between the cDNA clones and the genomic ones was determined by comparative restriction enzyme mapping and nucleic acid hybridization.  相似文献   

11.
12.
13.
Oxidative stress and certain environmental carcinogens, e.g. vinyl chloride and its metabolite chloroacetaldehyde (CAA), introduce promutagenic exocyclic adducts into DNA, among them 1,N(6)-ethenoadenine (epsilonA), 3,N(4)-ethenocytosine (epsilonC) and N(2),3-ethenoguanine (epsilonG). We studied sequence-specific interaction of the vinyl-chloride metabolite CAA with human p53 gene exons 5-8, using DNA Polymerase Fingerprint Analysis (DPFA), and identified sites of the highest sensitivity. CAA-induced DNA damage was more extensive in p53 regions which revealed secondary structure perturbations, and were localized in regions of mutation hot-spots. These perturbations inhibited DNA synthesis on undamaged template. We also studied the repair kinetics of CAA-induced DNA lesions in E. coli at nucleotide resolution level. A plasmid bearing full length cDNA of human p53 gene was modified in vitro with 360 mM CAA and transformed into E. coli DH5alpha strain, in which the adaptive response system had been induced by MMS treatment before the cells were made competent. Following transformation, plasmids were re-isolated from transformed cultures 35, 40, 50 min and 1-24 h after transformation, and further subjected to LM-PCR, using ANPG, MUG and Fpg glycosylases to identify the sites of DNA damage. In adaptive response-induced E. coli cells the majority of DNA lesions recognized by ANPG glycosylase were removed from plasmid DNA within 35 min, while MUG glycosylase excised base modifications only within 50 min, both in a sequence-dependent manner. In non-adapted cells resolution of plasmid topological forms was perturbed, suggesting inhibition of one or more bacterial topoisomerases by unrepaired epsilon-adducts. We also observed delayed consequences of DNA modification with CAA, manifesting as secondary DNA breaks, which appeared 3 h after transformation of damaged DNA into E. coli, and were repaired after 24 h.  相似文献   

14.
15.
16.
During endomitosis, megakaryocytes undergo several rounds of DNA synthesis without division leading to polyploidization. In primary megakaryocytes and in the megakaryocytic cell line CHRF, loss or knock-down of p53 enhances cell cycling and inhibits apoptosis, leading to increased polyploidization. To support the hypothesis that p53 suppresses megakaryocytic polyploidization, we show that stable expression of wild-type p53 in K562 cells (a p53-null cell line) attenuates the cells' ability to undergo polyploidization during megakaryocytic differentiation due to diminished DNA synthesis and greater apoptosis. This suggested that p53's effects during megakaryopoiesis are mediated through cell cycle- and apoptosis-related target genes, possibly by arresting DNA synthesis and promoting apoptosis. To identify candidate genes through which p53 mediates these effects, gene expression was compared between p53 knock-down (p53-KD) and control CHRF cells induced to undergo terminal megakaryocytic differentiation using microarray analysis. Among substantially downregulated p53 targets in p53-KD megakaryocytes were cell cycle regulators CDKN1A (p21) and PLK2, proapoptotic FAS, TNFRSF10B, CASP8, NOTCH1, TP53INP1, TP53I3, DRAM1, ZMAT3 and PHLDA3, DNA-damage-related RRM2B and SESN1, and actin component ACTA2, while antiapoptotic CKS1B, BCL2, GTSE1, and p53 family member TP63 were upregulated in p53-KD cells. Additionally, a number of cell cycle-related, proapoptotic, and cytoskeleton-related genes with known functions in megakaryocytes but not known to carry p53-responsive elements were differentially expressed between p53-KD and control CHRF cells. Our data support a model whereby p53 expression during megakaryopoiesis serves to control polyploidization and the transition from endomitosis to apoptosis by impeding cell cycling and promoting apoptosis. Furthermore, we identify a putative p53 regulon that is proposed to orchestrate these effects.  相似文献   

17.
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   

18.
DNA polymerase eta (PolH) is the product of the xeroderma pigmentosum variant (XPV) gene and a well-characterized Y-family DNA polymerase for translesion synthesis. Cells derived from XPV patients are unable to faithfully bypass UV photoproducts and DNA adducts and thus acquire genetic mutations. Here, we found that PolH can be up-regulated by DNA breaks induced by ionizing radiation or chemotherapeutic agents, and knockdown of PolH gives cells resistance to apoptosis induced by DNA breaks in multiple cell lines and cell types in a p53-dependent manner. To explore the underlying mechanism, we examined p53 activation upon DNA breaks and found that p53 activation is impaired in PolH knockdown cells and PolH-null primary fibroblasts. Importantly, reconstitution of PolH into PolH knockdown cells restores p53 activation. Moreover, we provide evidence that, upon DNA breaks, PolH is partially colocalized with phosphorylated ATM at gamma-H2AX foci and knockdown of PolH impairs ATM to phosphorylate Chk2 and p53. However, upon DNA damage by UV, PolH knockdown cells exhibit two opposing temporal responses: at the early stage, knockdown of PolH suppresses p53 activation and gives cells resistance to UV-induced apoptosis in a p53-dependent manner; at the late stage, knockdown of PolH suppresses DNA repair, leading to sustained activation of p53 and increased susceptibility to apoptosis in both a p53-dependent and a p53-independent manner. Taken together, we found that PolH has a novel role in the DNA damage checkpoint and that a p53 target can modulate the DNA damage response and subsequently regulate p53 activation.  相似文献   

19.
A 2.5-kb cDNA clone for human p53 tumor antigen has been isolated. This clone contains the entire coding region including 135 bp upstream of the first ATG. Comparison of the nucleotide sequence of human p53 and mouse p53 demonstrates that the first ATG in human p53 corresponds to the second ATG (codon No. 4) in mouse p53. The human p53 comprises 393 residues and is longer than the mouse p53 due to six additional codons present at the region corresponding to exon 4 of the mouse p53 gene. The DNA sequence homology between the coding regions of mouse and human p53 is 81% and the conservation of homology is not equally distributed along the molecule. When inserted into SV40-based expression vectors the human p53 cDNA successfully directs the production of a polypeptide with an apparent mol. wt. of 55 kd which can be precipitated by monoclonal antibodies to p53.  相似文献   

20.
p53 is an important mediator of the cellular stress response with roles in cell cycle control, DNA repair, and apoptosis. 53BP2, a p53-interacting protein, enhances p53 transactivation, impedes cell cycle progression, and promotes apoptosis through unknown mechanisms. We now demonstrate that endogenous 53BP2 levels increase following UV irradiation induced DNA damage in a p53-independent manner. In contrast, we found that the presence of a wild-type (but not mutant) p53 gene suppressed 53BP2 steady-state levels in cell lines with defined p53 genotypes. Likewise, expression of a tetracycline-regulated wild-type p53 cDNA in p53-null fibroblasts caused a reduction in 53BP2 protein levels. However, 53BP2 levels were not reduced if the tetracycline-regulated p53 cDNA was expressed after UV damage in these cells. This suggests that UV damage activates cellular factors that can relieve the p53-mediated suppression of 53BP2 protein. To address the physiologic significance of 53BP2 induction, we utilized stable cell lines with a ponasterone A-regulated 53BP2 cDNA. Conditional expression of 53BP2 cDNA lowered the apoptotic threshold and decreased clonogenic survival following UV irradiation. Conversely, attenuation of endogenous 53BP2 induction with an antisense oligonucleotide resulted in enhanced clonogenic survival following UV irradiation. These results demonstrate that 53BP2 is a DNA damage-inducible protein that promotes DNA damage-induced apoptosis. Furthermore, 53BP2 expression is highly regulated and involves both p53-dependent and p53-independent mechanisms. Our data provide new insight into 53BP2 function and open new avenues for investigation into the cellular response to genotoxic stress.  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