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.
2.
3.
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   

4.
The concept that the tumor suppressor p53 is a latent DNA-binding protein that must become activated for sequence-specific DNA binding recently has been challenged, although the "activation" phenomenon has been well established in in vitro DNA binding assays. Using electrophoretic mobility shift assays and fluorescence correlation spectroscopy, we analyzed the binding of "latent" and "activated" p53 to double-stranded DNA oligonucleotides containing or not containing a p53 consensus binding site (DNAspec or DNAunspec, respectively). In the absence of competitor DNA, latent p53 bound DNAspec and DNAunspec with high affinity in a sequence-independent manner. Activation of p53 by the addition of the C-terminal antibody PAb421 significantly decreased the binding affinity for DNAunspec and concomitantly increased the binding affinity for DNAspec. The net result of this dual effect is a significant difference in the affinity of activated p53 for DNAspec and DNAunspec, which explains the activation of p53. High affinity nonspecific DNA binding of latent p53 required both the p53 core domain and the p53 C terminus, whereas high affinity sequence-specific DNA binding of activated p53 was mediated by the p53 core domain alone. The data suggest that high affinity nonspecific DNA binding of latent and high affinity sequence-specific binding of activated p53 to double-stranded DNA differ in their requirement for the C terminus and involve different structural features of the core domain. Because high affinity nonspecific DNA binding of latent p53 is restricted to wild type p53, we propose that it relates to its tumor suppressor functions.  相似文献   

5.
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   

6.
Activation of the cryptic DNA binding function of mutant forms of p53.   总被引:16,自引:3,他引:13       下载免费PDF全文
Wild type p53 assembles into a latent multiprotein complex which can be activated for sequence-specific DNA binding in vitro by proteins targeting the carboxy-terminal domain. Using an optimized system coupling the post-translational modification of wild type p53 to activation of sequence specific DNA binding, we examined the affects of common mutations on the cryptic DNA binding function of p53. Two mutant forms of p53 were shown to be efficiently converted from the latent state by PAb421 and DnaK, but were defective in activation by casein kinase II, indicating that mutant p53 may not be receptive to allosteric regulation by casein kinase II phosphorylation. A reactive sulfhydryl group is absolutely required for DNA binding by wild type and mutant forms of p53 once converted to the activated state. Together, these data show that some mutant forms of p53 harbour the wild-type machinery required to engage in sequence-specific DNA binding and define a signalling pathway whose inactivation may directly result in a loss of p53 function.  相似文献   

7.
P53 acetylation requires p300-docking to two contiguous sites in the activation domain that in turn mediates DNA-dependent acetylation of the tetramer. In an attempt to further define the mechanism of DNA-dependent acetylation of p53, an in vitro system has been reconstituted with distinct p53 isoforms and has been used to reveal conformational constraints on p53 acetylation. Two native p53 tetrameric isoforms purified from Sf9 cells differing by the extent of phosphorylation within the C-terminal acetylation site are both acetylated in a sequence-specific DNA-dependent manner. By contrast, p53 purified from an Escherichia coli expression system is in a largely denatured conformation and its acetylation is DNA-independent. Heating native p53 to destroy the folded structure restores DNA-independent acetylation similar to that seen with bacterially expressed p53. There are at least two sites of conformational flexibility in the p53 tetramer: the first in the flexible S10 beta-sheet within the MDM2 ubiquitination sequence and the second in the C-terminal regulatory domain. We analysed therefore whether DNA-dependent acetylation correlated with conformational changes in either of these two regions. DNA-dependent acetylation of p53 is maintained in a dose-dependent manner by low concentrations of consensus site DNA under conditions where flexibility in the S10 beta-sheet region is maintained. Oligonucleotide DNAs that promote acetylation stimulate the binding of monoclonal antibodies PAb421 and ICA-9; two antibodies whose contiguous epitopes overlap the C-terminal acetylation motif. By contrast, bent oligonucleotide DNAs that conceal both the S10 beta-sheet from binding of the monoclonal antibody DO-12 and attenuate binding of the monoclonal antibody PAb421 can preclude acetylation. These data suggest that, in the absence of DNA, the acetylation motif of p53 is in a cryptic state, but after DNA binding, allosteric effects mediate an exposure of the acetylation motif to allow DNA-dependent acetylation of the tetramer.  相似文献   

8.
In this study we further characterized the 3'-5' exonuclease activity intrinsic to wild-type p53. We showed that this activity, like sequence-specific DNA binding, is mediated by the p53 core domain. Truncation of the C-terminal 30 amino acids of the p53 molecule enhanced the p53 exonuclease activity by at least 10-fold, indicating that this activity, like sequence-specific DNA binding, is negatively regulated by the C-terminal basic regulatory domain of p53. However, treatments which activated sequence-specific DNA binding of p53, like binding of the monoclonal antibody PAb421, which recognizes a C-terminal epitope on p53, or a higher phosphorylation status, strongly inhibited the p53 exonuclease activity. This suggests that at least on full-length p53, sequence-specific DNA binding and exonuclease activities are subject to different and seemingly opposing regulatory mechanisms. Following up the recent discovery in our laboratory that p53 recognizes and binds with high affinity to three-stranded DNA substrates mimicking early recombination intermediates (C. Dudenhoeffer, G. Rohaly, K. Will, W. Deppert, and L. Wiesmueller, Mol. Cell. Biol. 18:5332-5342), we asked whether such substrates might be degraded by the p53 exonuclease. Addition of Mg2+ ions to the binding assay indeed started the p53 exonuclease and promoted rapid degradation of the bound, but not of the unbound, substrate, indicating that specifically recognized targets can be subjected to exonucleolytic degradation by p53 under defined conditions.  相似文献   

9.
The mouse p53 gene generates two alternative splice products encoding p53 protein and a naturally occurring protein (p53as) with changes at the C-terminus. In p53as the negative regulatory region for DNA binding and PAb421 antibody binding site are replaced, and p53as is constitutively active for sequence-specific DNA binding. Using the technique of randomized synthetic oligonucleotide in cyclic amplification and selection of targets, we have found that p53as and p53 proteins have the same DNA binding specificities but that these specificities frequently diverge from the consensus of two copies of PuPuPuCATGPyPyPy. The importance of tetranucleotide CATG was confirmed but there was a less rigorous requirement for patterns of flanking or intervening sequences. In particular, the three purines upstream and three pyrimidines downstream of CATG are not required for p53 or p53as binding, 29 or more intervening nucleotides are tolerated, and one CATG is sufficient where adjacent nucleotides contain a region of homology with certain previously reported non-consensus p53 binding sequences. These results suggested further definition of the non-consensus motifs, and database searches with these uncovered additional candidate genes for p53 protein binding. We conclude that p53as and perhaps other activated forms of p53 exert their effects on the same genes and that differential activities of p53 protein forms are not due to inherently different sequence selectivities of DNA binding.  相似文献   

10.
We have investigated the effect of PARP gene inactivation on the expression of wild-type (wt) p53 protein. Using immortalized fibroblasts from control and PARP knock-out mice we have found by immunoblotting with the PAb421 antibody a profound decrease of the p53 expression to a barely detectable level in PARP knock-out cells. Surprisingly, longer exposure of immunoblots revealed an immunoreactive band at about 75 kD which was stronger in PARP-deficient cells than in wt cells and was not affected upon doxorubicin treatment. The size of the PAb421 immunoreactive protein and the lack of its inducibility in response to DNA damage resembled those of p73, the first described p53 homologue. Therefore, we examined the reactivity of anti-p53 antibodies with in vitro translated p73 protein. Interestingly, p73 was efficiently immunoprecipitated with distinct antibodies recognizing the carboxy-terminus of p53. In Northern blots we observed p73 signals of comparable intensity in controls and PARP-deficient cells. We conclude that elevated expression of p73 may compensate the reduced level of p53 in PARP-deficient cells.  相似文献   

11.
A temperature-sensitive mutant of human p53.   总被引:11,自引:5,他引:11       下载免费PDF全文
W Zhang  X Y Guo  G Y Hu  W B Liu  J W Shay    A B Deisseroth 《The EMBO journal》1994,13(11):2535-2544
  相似文献   

12.
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   

13.
14.
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   

15.
We analysed by analytical ultracentrifugation and fluorescence anisotropy the binding of p53 truncation mutants to sequence-specific DNA. The synthetic 30 base-pair DNA oligomers contained the 20 base-pair recognition elements for p53, consisting of four sites of five base-pairs per p53 monomer. We found that the binding at low ionic strengths was obscured by artifacts of non-specific binding and so made measurements at higher ionic strengths. Analytical ultracentrifugation of the construct p53CT (residues 94-360, containing the DNA-binding core and tetramerization domains) gave a dissociation constant of approximately 3 microM for its dimer-tetramer equilibrium, similar to that of full-length protein. Analytical ultracentrifugation and fluorescence anisotropy showed that p53CT formed a complex with the DNA constructs with 2:1 stoichiometry (dimer:DNA). The binding of p53CT (1-100 nm range) to DNA was highly cooperative, with a Hill coefficient of 1.8 (dimer:DNA). The dimeric L344A mutant of p53CT has impaired tetramerization. It bound to full-length DNA p53 recognition sequence, but with sixfold less affinity than wild-type protein. It did not form a detectable complex with a 30-mer DNA construct containing two specific five base-pair sites and two random sites, emphasizing the high co-operativity of the binding. The fundamental active unit of p53 appears to be the tetramer, which is induced by DNA binding, although it is a dimer at low concentrations.  相似文献   

16.
The wild-type form of p53 contains an intrinsic 3′–5′-exonuclease activity. As p53 forms a complex with DNA polymerase α-primase (pol-prim) in vivo this finding suggests that p53 might cooperate with pol-prim to stabilize the genetic information of living cells. To test this hypothesis, exonuclease-free DNA pol-prim was expressed alone or together with p53 for purification. Pol-prim formed a complex with p53, which was purified by ion exchange and immunoaffinity chromatography from baculovirus-infected insect cells. The p53-containing pol-prim fractions removed a 3′-unpaired nucleotide with a 1.5–2-fold higher rate than a paired nucleotide, whereas the four subunit pol-prim did not have any exonuclase activity. Therefore, only p53/pol-prim was able to elongate a primer-template that contained a 3′-unpaired primer end in vitro. To achieve this, the 3′–5′-exonuclease activity of p53 excised the unpaired nucleotide at the 3′-end of the primer and created a paired 3′-end, which pol-prim was able to elongate. The exonuclease activity of p53 as well as the elongation of a primer with a mispaired 3′-end was inhibited specifically by the anti-p53 monoclonal antibodies PAb240 and PAb421.  相似文献   

17.
The existence of multiple affinity states for the opiate receptor in neuroblastoma x glioma NG108-15 hybrid cells has been demonstrated by competition binding studies with tritiated diprenorphine and [D-Ala2, D-Leu5]enkephalin (DADLE). In the presence of 10 mM Mg2+, all receptors exist in a high affinity state with Kd = 1.88 +/- 0.16 nM. Addition of 10 microM guanyl-5'-yl imidodiphosphate (Gpp(NH)p) decreased the affinity of DADLE to Kd = 8.08 +/- 0.93 nM. However, in the presence of 100 mM Na+, which is required for opiate inhibition of adenylate cyclase activity, analysis of competition binding data revealed three sites: the first, consisting of 17.5% of total receptor population has a Kd = 0.38 +/- 0.18 nM; the second, 50.6% of the population, has a Kd = 6.8 +/- 2.2 nM; and the third, 31.9% of the population, has a Kd of 410 +/- 110 nM. Thus, in the presence of sodium, a high affinity complex between receptor (R), GTP binding component (Ni), and ligand (L) was formed which was different from that formed in the absence of sodium. These multiple affinity states of receptor in the hybrid cells are agonist-specific, and the percentage of total opiate receptor in high affinity state is relatively constant in various concentrations of Na+. Multiple affinity states of opiate receptor can be demonstrated further by Scatchard analysis of saturation binding studies with [3H]DADLE. In the presence of Mg2+, or Gpp(NH)p, analysis of [3H]DADLE binding demonstrates that opiate receptor can exist in a single affinity state, with apparent Kd values of [3H]DADLE in 10 mM Mg2+ = 1.75 +/- 0.28 nM and in 10 microM Gpp(NH)p = 0.85 +/- 0.12 nM. There is a reduction of Bmax value from 0.19 +/- 0.02 nM in the presence of Mg2+ to 0.14 +/- 0.03 nM in the presence of Gpp(NH)p. In the presence of 100 mM Na+, Scatchard analysis of saturation binding of [3H]DADLE reveals nonlinear plots; two-site analysis of the curves yields Kd = 0.43 +/- 0.09 and 7.9 +/- 3.2 nM. These Kd values are analogous to that obtained with competition binding studies. Again, this conversion of single site binding Scatchard plots to multiple sites binding plots in the presence of Na+ is restricted to 3H-agonist binding only.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

18.
The effect of PGE1 on low density lipoprotein (LDL) apo-B-receptor binding was examined in human, rat and swine liver. Autologous LDL (for humans and swines) and homologous LDL (for rats) were isolated by ultracentrifugation and labelled with 123I using Iodogen followed by purification with dialysis. LDL-concentrations of 0.1-6 micrograms protein/ml were used for direct binding assays investigating the specific binding of labelled LDL in presence of increasing PGE1-concentrations (100 pM to 100 microM). In separate experiments the effect of PGE1 on displacement of specifically bound 123I-LDL by unlabelled ones was studied. The binding capacities estimated by Scatchard analysis were similar for human and rat liver LDL-apo-B-receptor binding, however, swine liver exhibited a significantly (p less than 0.001) lower binding capacity for 123I-LDL. PGE1 significantly (p less than 0.01-0.001) increased the amount of 123I-LDL specifically bound to the liver apo-B-receptors and the binding affinity in all liver preparations of the 3 species in a dose-dependent manner. PGE1 also significantly increased competition of unlabelled LDL for 123I-LDL bound to its specific apo-B-receptors in a dose-dependent manner (p less than 0.01-0.001) with an ED50 of 123 +/- 64 nM for human liver, 901 +/- 102 nM for rat liver obtained during anaesthesia, 74 +/- 23 nM for rat liver obtained after decapitation and 941 +/- 121 nM for swine liver. In human liver iloprost (ED50 = 876 +/- 53 nM) and PGI2 (ED50 = 52 +/- 12 microM) were less effective than PGE1, PGE2 had no effect on LDL-induced competition. It is concluded that PGE1 renders LDL more sensitive for apo-B-receptor binding suggesting a potential hypolipidemic action of PGE1.  相似文献   

19.
Escherichia coli RNA polymerase associated with the sigma54 factor (RNAP*sigma54) is a holoenzyme form that transcribes a special class of promoters not recognized by the standard RNA polymerase*sigma70 com plex. Promoters for RNAP*sigma54 vary in their overall 'strength' and show differences in their response to the presence of DNA curvature between enhancer and promoter. In order to examine whether these effects are related to the promoter affinity, we have determined the equilibrium dissociation constant K(d) for the binding of RNAP*sigma54 to the three promoters glnAp2, nifH and nifL. Binding studies were conducted by monitoring the changes in fluorescence anisotropy upon titrating RNAP*sigma54 to carboxyrhodamine-labeled DNA duplexes. For the glnAp2 and nifH promoters similar values of K(d) = 0.94 +/- 0.55 nM and K(d) = 0.85 +/- 0.30 nM were determined at physiological ionic strength, while the nifL promoter displayed a significantly weaker affinity with K(d) = 8.5 +/- 1.9 nM. The logarithmic dependence of K(d) on the ionic strength I was -Deltalog(K(d))/Deltalog(I) = 6.1 +/- 0.5 for the glnAp2, 5.2 +/- 1.2 for the nifH and 2.1 +/- 0.1 for the nifL promoter. This suggests that the polymerase can form fewer ion pairs with the nifL promoter, which would account for its weaker binding affinity.  相似文献   

20.
A population of latent (cryptic) receptors for tumor necrosis factor-alpha (TNF) has been characterized in the rat liver plasma membrane (PM). 125I-TNF bound to high (Kd = 1.51 +/- 0.35 nM) and low (Kd = 13.58 +/- 1.45 nM) affinity receptors in PM. Solubilization of PM with 1% Triton X-100 prior to incubation with 125I-TNF increased both high affinity (from 0.33 +/- 0.04 to 1.67 +/- 0.05 pmol/mg of protein) and low affinity (from 1.92 +/- 0.16 to 7.57 +/- 0.50 pmol/mg of protein) TNF binding without affecting the affinities for TNF. Digestion of intact PM with chymotrypsin abolished most of the TNF binding capacity of PM. However, substantial binding activity was recovered by solubilization of chymotrypsin-treated PM with 1% Triton X-100, suggesting the presence of a large latent pool of TNF receptors. The affinities of the high and low affinity sites recovered from chymotrypsin-treated membranes were similar to those of intact PM. Affinity labeling of receptors whether from PM, solubilized PM, or membranes digested with chymotrypsin and then solubilized resulted in cross-linking of 125I-TNF into Mr 130,000, 90,000, and 66,000 complexes. Thus, the properties of the latent TNF receptors were similar to those initially accessible to TNF. To determine if exposure of latent receptors is regulated by TNF, 125I-TNF binding to control and TNF-pretreated membranes was assayed. Specific binding was increased by pretreatment with TNF (p less than 0.05), demonstrating that hepatic PM contains latent TNF receptors whose exposure is promoted by TNF. Homologous up-regulation of TNF receptors may, in part, be responsible for sustained hepatic responsiveness during chronic exposure to TNF.  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