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 30 毫秒
1.
We recently observed an interaction between poly(ADP-ribose) polymerase-1 (PARP-1) and the tumor suppressor p53 protein. However, more extensive studies on both proteins, especially those on characterization of their domains involved in the interaction were difficult due to very low expression levels of p53 in mammalian cells. Therefore, we generated recombinant proteins for such studies. To clarify which domains of human PARP-1 and of human wild-type (wt) p53 were involved in this protein-protein interaction, we generated baculoviral constructs encoding full length or distinct functional domains of both proteins. Full length PARP-1 was simultaneously coexpressed in insect cells with full length wt p53 protein or its distinct truncated fragments and vice versa. Reciprocal immunoprecipitation of Sf9 cell lysates revealed that the central and carboxy-terminal fragments of p53 each were sufficient to confer binding to PARP-1, whereas the amino-terminal part harbouring the transactivation functional domain was dispensable. On the other hand, the amino-terminal and central fragments of PARP-1 were both necessary for complex formation with p53 protein. Since the most important features of p53 protein are regulated by phosphorylation, we addressed the question whether its phosphorylation is essential for the binding between the two proteins. Baculovirally expressed wt p53 was post-translationally modified. At least six distinct p53 isomers were resolved by immunoblotting following two-dimensional separation of baculovirally expressed wt p53 protein. Using specific phospho-serine antibodies, we identified phosphorylation of baculovirally expressed p53 protein at five distinct sites. To define the role of p53 phosphorylation, pull-down assays using untreated and dephosphorylated p53 protein were performed. Dephosphorylated p53 failed to bind PARP-1, indicating that complex formation between the two proteins was regulated by phosphorylation of p53. The marked phosphorylation of p53 at Ser392 observed in unstressed cells suggests that the phosphorylated carboxy-terminal part of p53 undergoes complex formation with PARP-1 resulting in masking of the NES and thereby preventing its export.  相似文献   

2.
It has been previously described by different groups that poly(ADP-ribose) polymerase-1 (PARP-1) and the product of the tumor suppressor gene p53 form tight complexes. We investigated which domains of human PARP-1 and of human wild-type p53 were involved in this protein-protein interaction. We generated baculoviral constructs encoding full length protein or distinct functional domains of both proteins. Baculovirally expressed wild-type p53 was posttranslationally modified. Full length PARP-1 was simultaneously coexpressed in insect cells with full length wt p53 protein or its distinct truncated fragments and vice versa. Reciprocal immunoprecipitation of Sf9 cell lysates revealed that the central and carboxy-terminal fragments of p53 were sufficient to confer binding to PARP-1. The amino-terminal part harboring the transactivation functional domain of p53 was dispensable. On the other hand, the amino-terminal and central fragments of PARP-1 were necessary for complex formation with p53 protein. Finally, we explored the functional significance of the interaction between both proteins. Inactivation of PARP-1 resulted in the reduction of p53 steady-state levels. Inhibition of nuclear export by leptomycin B prevented accelerated degradation of p53 in PARP-1 KO cells and led to accumulation of p53 protein. Considering the fact that the accelerated p53 nuclear export in the absence of PARP-1 contributes to enhanced p53 degradation, we conclude that PARP-1 may mask the NES of p53 through complex formation with its carboxy-terminal part, thereby preventing the export.  相似文献   

3.
4.
The constitutive activation of the Janus kinase 2 (JAK2) and mutation of the p53 tumor suppressor are both detected in human cancer. We examined the potential regulation of JAK2 phosphorylation by wild-type (wt) p53 in human ovarian cancer cell lines, Caov-3 and MDAH2774, which harbor mutant form of p53 tumor suppressor gene and high levels of phosphorylated JAK2. The wt p53 gene was re-introduced into the cells using an adenovirus vector. In addition to wt p53, mutant p53 22/23, mutant p53-175, and NCV (negative control virus) were introduced into the cells in the control groups. Expression of wt p53, but not that of p53-175 mutant, diminished JAK2 tyrosine phosphorylation in MDAH2774 and Caov-3 cell lines. Expression of wt p53 or p53 22/23 mutant did not cause a reduction in the phosphorylation of unrelated protein kinases, ERK1 and ERK2 (ERK1/2). The inhibition of JAK2 tyrosine phosphorylation can be reversed by tyrosine phosphatase inhibitor, sodium orthovanadate. Protein tyrosine phosphatase 1-B levels increased with introduction of wt p53 and may be involved in the dephosphorylation of JAK2. These findings present a possible p53-dependent cellular process of modulating JAK2 tyrosine phosphorylation in ovarian cancer cell lines.  相似文献   

5.
NSP1 is an essential nuclear pore protein in yeast. We observed that anti-NSP1 antibodies label mammalian nuclear pore complexes and recognize nucleoporin p62. Also peptide antibodies raised against the NSP1 carboxy-terminal end cross-react with p62, a conserved component of the nuclear pore complex in higher eukaryotes. To further analyze the structural and functional similarity between NSP1 and mammalian nucleoporins, we cloned and sequenced nucleoporin p62 from a HeLa cDNA library. Human p62 consists of a carboxy-terminal domain homologous to the essential yeast NSP1 carboxy-terminal domain and an amino-terminal half resembling the repetitive middle domain of NSP1. The full-length p62 and a fusion protein consisting of cytosolic mouse dihydrofolate reductase (DHFR) and the p62 carboxy-terminal domain were expressed in transfected HeLa cells. Only overexpressed full-length p62, but not the DHFR-C-p62 fusion protein, binds wheat germ agglutinin (WGA). This suggests that modification by N-acetylglucosamine is mainly restricted to the repetitive amino-terminal half of p62 and implies a role of this type of repetitive sequences in nuclear transport. In the transfected HeLa cells, the DHFR-C-p62 fusion protein forms patchy aggregates that accumulate at the nuclear periphery but are also scattered through the cytoplasm. It is suggested that nucleoporin p62 may be targeted and anchored to the pore complex via its carboxy-terminal domain which reveals a hydrophobic heptad repeat organization similar to that found in lamins and other intermediate filament proteins.  相似文献   

6.
7.
8.
Azurin, a bacterial protein, can be internalized in cancer cells and induce apoptosis. Such anticancer effect is coupled to the formation of a complex with the tumour‐suppressor p53. The mechanism by which azurin stabilizes p53 and the binding sites of their complex are still under investigation. It is also known that the predominant mechanism for p53 down‐regulation implies its association to Mdm2, the main ubiquitin ligase affecting its stability. However, the p53/Mdm2 interaction, occurring at the level of both their N‐terminal domains, has been characterized so far by experiments involving only partial domains of these proteins. The relevance of the p53/Mdm2 complex as a possible target of the anticancer therapies requires a deeper study of this complex as made up of the two entire proteins. Moreover, the apparent antagonist action of azurin against Mdm2, with respect of p53 regulation, might suggest the possibility that azurin binds p53 at the same site of Mdm2, preventing in such a way p53 and Mdm2 from association and thus p53 from degradation. By following the interaction of the two entire proteins by atomic force spectroscopy, we have assessed the formation of a specific complex between p53 and Mdm2. We found for it a binding strength and a dissociation rate constant typical of dynamical protein–protein interactions and we observed that azurin, even if capable to bind p53, does not compete with Mdm2 for the same binding site on p53. The formation of the p53/Mdm2/azurin ternary complex might suggest an alternative anti‐cancer mechanism adopted by azurin. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

9.
Although wild‐type p53 protein is overexpressed in first trimester trophoblast, it is inactive towards its target genes Metalloproteinase 2 and 9. This seems to be due to a complex mechanism of inactivation and stabilization of p53 relying on the formation of protein complexes involving the N‐terminus of p53. To detect the proteins associated with this sequence, we incubated biotinylated p53 N‐terminal peptide in cytotrophoblastic cell medium 24 h before lysis of cells. We purified the proteins retained on biotinylated peptide using a neutravidin affinity column. Proteins were then identified by peptide mass finger printing followed or not by peptide fragmentation sequencing. Among these proteins, we identified glucose‐regulated protein 78 (GRP78) and verified its interaction with p53 in trophoblastic cells by immunoprecipitation and Western blot analysis. Moreover, the decreased expression of GRP78 induced by GRP78siRNA or versipelostatin decreased the formation of high molecular weight p53 complexes and p53 monomer and increased trophoblastic invasion. These results suggest that GRP78 is involved in inactivation and stabilization of p53 and in the regulation of trophoblastic invasion.  相似文献   

10.
A complete set of mono-, di- and triphosphorylated peptides comprising amino acids 10-27, the Mdm2 and p300 binding site(s) of p53, with and without a fluorescein label at the N-terminus, was synthesized by step-by-step solid phase synthesis. Fluorescence polarization analysis revealed that phosphorylation at Thr18 decreased binding to recombinant Mdm2 protein compared with the unphosphorylated and the two other single phosphorylated analogues. Unlabelled multiply phosphorylated peptides corresponding to this amino-terminal transactivation domain proved to be powerful tools in analysing the phosphate specificity of existing anti-p53 monoclonal and polyclonal antibodies using direct ELISA. The tetramerization domain of human p53 protein was modelled with a 53 residue-long unlabelled unphosphorylated and Ser315-phosphorylated peptide pair. CD analysis showed similar alpha-helical structures for both peptides and no major difference in the secondary structure could be observed upon phosphorylation. Size-exclusion HPLC indicated that these synthetic oligomerization domain mimics underwent a pH-dependent tetramerization process, but the presence of a phosphate group at Ser315 did not modify the oligomeric state of the 308-360 p53 fragments. Nevertheless, the fluorescein-labelled Ser315 phosphorylated peptide bound to the downstream signalling ligand DNA topoisomerase I protein with slightly higher affinity than did the unphosphorylated analogue.  相似文献   

11.
12.
Depletion of intracellular zinc by N,N,N′,N′-tetrakis(2-pyridylmethyl) ethylenediamine (TPEN) induces p53-mediated protein synthesis-dependent apoptosis of mouse cortical neurons. Here, we examined the requirement for poly(ADP-ribose) polymerase (PARP)-1 as an upstream regulator of p53 in zinc depletion-induced neuronal apoptosis. First, we found that chemical inhibition or genetic deletion of PARP-1 markedly attenuated TPEN-induced apoptosis of cultured mouse cortical neurons. Poly(ADP-ribosyl)ation of p53 occurred starting 1 h after TPEN treatment. Suggesting the critical role of PARP-1, the TPEN-induced increase of stability and activity of p53 as well as poly(ADP-ribosyl)ation of p53 was almost completely blocked by PARP inhibition. Consistent with this, the induction of downstream proapoptotic proteins PUMA and NOXA was noticeably reduced by chemical inhibitors or genetic deletion of PARP-1. TPEN-induced cytochrome C release into the cytosol and caspase-3 activation were also blocked by inhibition of PARP-1. Taken together, these findings indicate that PARP-1 is essential for TPEN-induced neuronal apoptosis.  相似文献   

13.
14.
We investigated the interaction between poly(ADP-ribose) polymerase-1 (PARP-1) and the product of the tumor suppressor gene p53 using two different approaches. In the first approach, we used primary and immortalized cells derived from wt and PARP-1 -/- mice. We examined whether PARP-1 deficiency would affect the expression of the wild-type (wt) p53 protein. The inactivation of the PARP-1 gene markedly affected the constitutive expression of the wt p53 protein. Interestingly, only the regularly spliced form of wt p53 was reduced to a barely detectable level in consequence to an approximately 8-fold shortening of its half-life, whereas the level of alternatively spliced p53 remained unchanged. Moreover, reconstitution of cells lacking the PARP-1 gene with the human counterpart restored the normal stability of the regularly spliced p53 protein. In the second approach, we performed experiments with c-Ha-ras transformed primary rat cells overexpressing the p53135val mutant alone or in combination with PARP-1. The advantage of this temperature sensitive p53135val mutant is its oncogenic character at 37 degrees C, connected with cytoplasmic localization of p53, and its tumor suppressor activity at 32 degrees C, accompanied by p53 translocation into the nucleus. No noticeable differences in proliferation and G1 accumulationwere observed between cells expressing p53135val with or without PARP-1. On the other hand, a comparison of the recovery of G1 arrested cells after a shift up to 37 degrees C for both cell lines showed dramatic differences in the kinetics. While cells expressing p53135val rapidly reached the characteristic S-phase level after a shift up to basal temperature, cells additionally expressing PARP-1 rested in G1 despite the temperature elevation. This coincided with exclusively cytoplasmic p53 protein in cells expressing p53135val and predominantly nuclear localization of p53 in p53135val +PARP-1 cells, as evidenced by immunostaining. Determination of the p53 level during the maintenance of cells at 32 degrees C revealed a marked decrease in the level of p53 in cells expressing p53135val alone, whereas in cells coexpressing PARP-1, the level of p53 remained largely unaffected. This indicates that the stability of wild-type p53 greatly differed between both cell lines. Furthermore, the inhibition of PARP-1 activity in G1 arrested cells by 3-aminobenzamide abolished its stabilizing effect on the wild-type p53 protein. Taken together, our results indicate that PARP-1 regulates the stability of the wt p53 protein and that its enzymatic activity is necessary for this stabilizing action.  相似文献   

15.
The p53 cofactor Strap (stress responsive activator of p300) is directly targeted by the DNA damage signalling pathway where phosphorylation by ATM (ataxia telangiectasia mutated) kinase facilitates nuclear accumulation. Here, we show that Strap regulation reflects the coordinated interplay between different DNA damage-activated protein kinases, ATM and Chk2 (Checkpoint kinase 2), where phosphorylation by each kinase provides a distinct functional consequence on the activity of Strap. ATM phosphorylation prompts nuclear accumulation, which we show occurs by impeding nuclear export, whereas Chk2 phosphorylation augments protein stability once Strap has attained a nuclear location. These results highlight the various functional roles undertaken by the DNA damage signalling kinases in Strap control and, more generally, shed light on the pathways that contribute to the regulation of the p53 response.  相似文献   

16.
17.
Small-molecule inhibitors of poly (ADP-ribose) polymerase (PARP) have shown considerable promise in the treatment of homologous recombination (HR)-defective tumors, such as BRCA1- and BRCA2-deficient breast and ovarian cancers. We previously reported that mantle cell lymphoma cells with deficiency in ataxia telangiectasia mutated (ATM) are sensitive to PARP-1 inhibitors in vitro and in vivo. Here, we report that PARP inhibitors can potentially target ATM deficiency arising in a solid malignancy. We show that ATM protein expression varies between gastric cancer cell lines, with NUGC4 having significantly reduced protein levels. Significant correlation was found between ATM protein expression and sensitivity to the PARP inhibitor olaparib, with NUGC4 being the most sensitive. Moreover, reducing ATM kinase activity using a small-molecule inhibitor (KU55933) or shRNA-mediated depletion of ATM protein enhanced olaparib sensitivity in gastric cancer cell lines with depletion or inactivation of p53. Our results demonstrate that ATM is a potential predictive biomarker for PARP-1 inhibitor activity in gastric cancer harboring disruption of p53, and that combined inhibition of ATM and PARP-1 is a rational strategy for expanding the utility of PARP-1 inhibitors to gastric cancer with p53 disruption.  相似文献   

18.
The M band of sarcomeric muscle is a highly complex structure which contributes to the maintenance of the regular lattice of thick filaments. We propose that the spatial coordination of this assembly is regulated by specific interactions of myosin filaments, the M band protein myomesin and the large carboxy-terminal region of titin. Corresponding binding sites between these proteins were identified. Myomesin binds myosin in the central region of light meromyosin (LMM, myosin residues 1506-1674) by its unique amino-terminal domain My1. A single titin immunoglobulin domain, m4, interacts with a myomesin fragment spanning domains My4-My6. This interaction is regulated by phosphorylation of Ser482 in the linker between myomesin domains My4 and My5. Myomesin phosphorylation at this site by cAMP-dependent kinase and similar or identical activities in muscle extracts block the association with titin. We propose that this demonstration of a phosphorylation-controlled interaction in the sarcomeric cytoskeleton is of potential relevance for sarcomere formation and/or turnover. It also reveals how binding affinities of modular proteins can be regulated by modifications of inter-domain linkers.  相似文献   

19.
Half-minilamins, representing amino- and carboxy-terminal fragments of human lamins A, B1 and B2 with a truncated central rod domain, were investigated for their ability to form distinct head-to-tail-type dimer complexes. This mode of interaction represents an essential step in the longitudinal assembly reaction exhibited by full-length lamin dimers. As determined by analytical ultracentrifugation, the amino-terminal fragments were soluble under low ionic strength conditions sedimenting with distinct profiles and s-values (1.6-1.8 S) indicating the formation of coiled-coil dimers. The smaller carboxy-terminal fragments were, except for lamin B2, largely insoluble under these conditions. However, after equimolar amounts of homotypic amino- and carboxy-terminal lamin fragments had been mixed in 4 M urea, upon subsequent renaturation the carboxy-terminal fragments were completely rescued from precipitation and distinct soluble complexes with higher s-values (2.3-2.7 S) were obtained. From this behavior, we conclude that the amino- and carboxy-terminal coiled-coil dimers interact to form distinct oligomers (i.e. tetramers). Furthermore, a corresponding interaction occurred also between heterotypic pairs of A- and B-type lamin fragments. Hence, A-type lamin dimers may interact with B-type lamin dimers head-to-tail to yield linear polymers. These findings indicate that a lamin dimer principally has the freedom for a “combinatorial” head-to-tail association with all types of lamins, a property that might be of significant importance for the assembly of the nuclear lamina. Furthermore, we suggest that the head-to-tail interaction of the rod end domains represents a principal step in the assembly of cytoplasmic intermediate filament proteins too.  相似文献   

20.
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