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 484 毫秒
1.
2.
3.
Computational models reveal the structural origins of cooperativity in the association of the DNA binding domains (DBD) of p53 (and its two homologues p63 and p73) with consensus DNA. In agreement with experiments they show that cooperativity, as defined by sequential binding of monomers to DNA is strong for p53 and weak for homologues p63 and p73. Computations also suggest that cooperativity can arise from the dimerization of the DBD prior to binding the DNA for all 3 family members. Dimerization between the DBDs is driven by packing interactions originating in residues of helix H1 and loop L3, while DNA binding itself is dominated by local and global electrostatics. Calculations further suggest that low affinity oligomerization of the p53 DBD can precede the oligomerization of the tetramerization domain (TD). During synthesis of multiple chains on the polysome, this may increase fidelity by reducing the possibility of the highly hydrophobic TD from nonspecific aggregation. Mutations have been suggested to test these findings.  相似文献   

4.
The p53 protein is the major tumor suppressor in mammals. The discovery of the p53 homologs p63 and p73 defined a family of p53 members with distinct roles in tumor suppression, differentiation, and development. Here, we describe the biochemical characterization of the core DNA-binding domain of a human isoform of p63, p63-delta, and particularly novel features in comparison with p53. In contrast to p53, the free p63 core domain did not show specific binding to p53 DNA consensus sites. However, glutathione S-transferase-fused and thus dimerized p63 and p53 core domains had similar affinity and specificity for the p53 consensus sites p21, gadd45, cyclin G, and bax. Furthermore, the fold of p63 core was remarkably stable compared with p53 as judged by differential scanning calorimetry (T(m) = 61 degrees C versus 44 degrees C for p53) and equilibrium unfolding ([urea](50%) = 5.2 m versus 3.1 m for p53). A homology model of p63 core highlights differences at a segment near the H1 helix hypothetically involved in the formation of the dimerization interface in p53, which might reduce cooperativity of p63 core DNA binding compared with p53. The model also shows differences in the electrostatic and hydrophobic potentials of the domains relevant to folding stability.  相似文献   

5.
Wyka IM  Dhar K  Binz SK  Wold MS 《Biochemistry》2003,42(44):12909-12918
Human replication protein A (RPA) is a heterotrimeric (70, 32, and 14 kDa subunits), eukaryotic single-stranded DNA (ssDNA) binding protein required for DNA recombination, repair, and replication. The three subunits of human RPA are composed of six conserved DNA binding domains (DBDs). Deletion and mutational studies have identified a high-affinity DNA binding core in the central region of the 70 kDa subunit, composed of DBDs A and B. To define the roles of each DBD in DNA binding, monomeric and tandem DBD A and B domain chimeras were created and characterized. Individually, DBDs A and B have a very low intrinsic affinity for ssDNA. In contrast, tandem DBDs (AA, AB, BA, and BB) bind ssDNA with moderate to high affinity. The AA chimera had a much higher affinity for ssDNA than did the other tandem DBDs, demonstrating that DBD A has a higher intrinsic affinity for ssDNA than DBD B. The RPA-DNA interface is similar in both DBD A and DBD B. Mutational analysis was carried out to probe the relative contributions of the two domains to DNA binding. Mutation of polar residues in either core DBD resulted in a significant decrease in the affinity of the RPA complex for ssDNA. RPA complexes with pairs of mutated polar residues had lower affinities than those with single mutations. The decrease in affinity observed when polar mutations were combined suggests that multiple polar interactions contribute to the affinity of the RPA core for DNA. These results indicate that RPA-ssDNA interactions are the result of binding of multiple nonequivalent domains. Our data are consistent with a sequential binding model for RPA, in which DBD A is responsible for positioning and initial binding of the RPA complex while DBD A together with DBD B direct stable, high-affinity binding to ssDNA.  相似文献   

6.
We show that changes in the nucleotide sequence alter the DNA conformation in the crystal structures of p63 DNA-binding domain (p63DBD) bound to its response element. The conformation of a 22-bp canonical response element containing an AT spacer between the two half-sites is unaltered compared with that containing a TA spacer, exhibiting superhelical trajectory. In contrast, a GC spacers abolishes the DNA superhelical trajectory and exhibits less bent DNA, suggesting that increased GC content accompanies increased double helix rigidity. A 19-bp DNA, representing an AT-rich response element with overlapping half-sites, maintains superhelical trajectory and reveals two interacting p63DBD dimers crossing one another at 120°. p63DBD binding assays to response elements of increasing length complement the structural studies. We propose that DNA deformation may affect promoter activity, that the ability of p63DBD to bind to superhelical DNA suggests that it is capable of binding to nucleosomes, and that overlapping response elements may provide a mechanism to distinguish between p63 and p53 promoters.  相似文献   

7.
8.
Butler JS  Loh SN 《Biochemistry》2003,42(8):2396-2403
The p53 DNA binding domain (DBD) contains a single bound zinc ion that is essential for activity. Zinc remains bound to wild-type DBD at temperatures below 30 degrees C; however, it rapidly dissociates at physiological temperature. The resulting zinc-free protein (apoDBD) is folded and stable. NMR spectra reveal that the DNA binding surface is altered in the absence of Zn(2+). Fluorescence anisotropy studies show that Zn(2+) removal abolishes site-specific DNA binding activity, although full nonspecific DNA binding affinity is retained. Surprisingly, the majority of tumorigenic mutations that destabilize DBD do not appreciably destabilize apoDBD. The R175H mutation instead substantially accelerates the rate of Zn(2+) loss. A considerable fraction of cellular p53 may therefore exist in the folded zinc-free form, especially when tumorigenic mutations are present. ApoDBD appears to promote aggregation of zinc-bound DBD via a nucleation-growth process. These data provide an explanation for the dominant negative phenotype exhibited by many mutations. Through a combination of induced p53 aggregation and diminished site-specific DNA binding activity, Zn(2+) loss may represent a significant inactivation pathway for p53 in the cell.  相似文献   

9.
10.
p53 is a conformationally flexible sequence-specific DNA binding protein mutated in many human tumors. To understand why the mutant p53 proteins associated with human tumors fail to bind DNA, we mapped the DNA binding domain of wild-type p53 and examined its regulation by changes in the protein conformation. Using site-directed mutagenesis, residues 90-286 of mouse p53 were shown to form the sequence-specific DNA binding domain. Two highly conserved regions within this domain, regions IV and V, were implicated in contacting DNA. Wild-type p53 bound DNA as a tetramer, each subunit recognizing five nucleotides of the 20 nucleotide-long DNA site. Conformational shifts of the oligomerization domain propagated to the tetrameric DNA binding domain, regulating DNA binding activity, but did not affect the subunit stoichiometry of wild-type p53 oligomers. Interestingly, conformational shifts could also be propagated within certain p53 mutants, rescuing DNA binding. One of these mutants was the mouse equivalent of human histidine 273, which is frequently associated with human tumors.  相似文献   

11.
12.
13.
Although p53 and p73 share considerable homology in their DNA-binding domains, there have been few studies examining their relative interactions with DNA as purified proteins. Comparing p53 and p73β proteins, our data show that zinc chelation by EDTA is significantly more detrimental to the ability of p73β than of p53 to bind DNA, most likely due to the greater effect that the loss of zinc has on the conformation of the DNA-binding domain of p73. Furthermore, prebinding to DNA strongly protects p73β but not p53 from chelation by EDTA suggesting that DNA renders the core domain of p73 less accessible to its environment. Further exploring these biochemical differences, a five-base sub-sequence was identified in the p53 consensus binding site that confers a greater DNA-binding stability on p73β than on full-length p53 in vitro. Surprisingly, p53 lacking its C-terminal non-specific DNA-binding domain (p53Δ30) demonstrates the same sequence discrimination as does p73β. In vivo, both p53 and p73β exhibit higher transactivation of a reporter with a binding site containing this sub-sequence, suggesting that lower in vitro dissociation translates to higher in vivo transactivation of sub-sequence-containing sites.  相似文献   

14.
BackgroundMutations within the DNA binding domain (DBD) of the tumor suppressor p53 are found in > 50% of human cancers and may significantly modify p53 secondary structure impairing its function. p28, an amphipathic cell-penetrating peptide, binds to the DBD through hydrophobic interaction and induces a posttranslational increase in wildtype and mutant p53 restoring functionality. We use mutation analyses to explore which elements of secondary structure may be critical to p28 binding.MethodsMolecular modeling, Raman spectroscopy, Atomic Force Spectroscopy (AFS) and Surface Plasmon Resonance (SPR) were used to identify which secondary structure of site-directed and naturally occurring mutant DBDs are potentially altered by discrete changes in hydrophobicity and the molecular interaction with p28.ResultsWe show that specific point mutations that alter hydrophobicity within non-mutable and mutable regions of the p53 DBD alter specific secondary structures. The affinity of p28 was positively correlated with the β-sheet content of a mutant DBD, and reduced by an increase in unstructured or random coil that resulted from a loss in hydrophobicity and redistribution of surface charge.ConclusionsThese results help refine our knowledge of how mutations within p53-DBD alter secondary structure and provide insight on how potential structural alterations in p28 or similar molecules improve their ability to restore p53 function.General significanceRaman spectroscopy, AFS, SPR and computational modeling are useful approaches to characterize how mutations within the p53DBD potentially affect secondary structure and identify those structural elements prone to influence the binding affinity of agents designed to increase the functionality of p53.  相似文献   

15.
The pharmacological suppression of the DNA damage response and DNA repair can increase the therapeutic indices of conventional chemotherapeutics. Replication Protein A (RPA), the major single-stranded DNA binding protein in eukaryotes, is required for DNA replication, DNA repair, DNA recombination, and DNA damage response signaling. Through the use of high-throughput screening of 1500 compounds, we have identified a small molecule inhibitor, 15-carboxy-13-isopropylatis-13-ene-17,18-dioic acid (NSC15520), that inhibited both the binding of Rad9-GST and p53-GST fusion proteins to the RPA N-terminal DNA binding domain (DBD), interactions that are essential for robust DNA damage signaling. NSC15520 competitively inhibited the binding of p53-GST peptide with an IC(50) of 10 μM. NSC15520 also inhibited helix destabilization of a duplex DNA (dsDNA) oligonucleotide, an activity dependent on the N-terminal domain of RPA70. NSC15520 did not inhibit RPA from binding single-stranded oligonucleotides, suggesting that the action of this inhibitor is specific for the N-terminal DBD of RPA, and does not bind to DBDs essential for single-strand DNA binding. Computer modeling implicates direct competition between NSC15520 and Rad9 for the same binding surface on RPA. Inhibitors of protein-protein interactions within the N-terminus of RPA are predicted to act synergistically with DNA damaging agents and inhibitors of DNA repair. Novel compounds such as NSC15520 have the potential to serve as chemosensitizing agents.  相似文献   

16.
S Mader  J Y Chen  Z Chen  J White  P Chambon    H Gronemeyer 《The EMBO journal》1993,12(13):5029-5041
We show here that, in addition to generating an increase in DNA binding efficiency, heterodimerization of retinoid X receptor (RXR) with either retinoic acid receptor (RAR) or thyroid hormone receptor (TR) alters the binding site repertoires of RAR, RXR and TR homodimers. The binding site specificities of both homo- and heterodimers appear to be largely determined by their DNA binding domains (DBDs), and are dictated by (i) homocooperative DNA binding of the RXR DBD, (ii) heterocooperative DNA binding of RXR/RAR and RXR/TR DBDs, and (iii) steric hindrance. No homodimerization domain exists in the DBDs of TR and RAR. The dimerization function which is located in the ligand binding domain further stabilizes, but in general does not change, the repertoire dictated by the corresponding DBD(s). The binding repertoire can be further modified by the actual sequence of the binding site. We also provide evidence supporting the view that the cooperative binding of the RXR/RAR and RXR/TR DBDs to directly repeated elements is anisotropic, with interactions between the dimerization interfaces occurring only with RXR bound to the 5' located motif. This polarity, which appears to be maintained in the full-length receptor heterodimers, may constitute a novel parameter in promoter-specific transactivation.  相似文献   

17.
Structural evolution of C-terminal domains in the p53 family   总被引:1,自引:0,他引:1  
Ou HD  Löhr F  Vogel V  Mäntele W  Dötsch V 《The EMBO journal》2007,26(14):3463-3473
The tetrameric state of p53, p63, and p73 has been considered one of the hallmarks of this protein family. While the DNA binding domain (DBD) is highly conserved among vertebrates and invertebrates, sequences C-terminal to the DBD are highly divergent. In particular, the oligomerization domain (OD) of the p53 forms of the model organisms Caenorhabditis elegans and Drosophila cannot be identified by sequence analysis. Here, we present the solution structures of their ODs and show that they both differ significantly from each other as well as from human p53. CEP-1 contains a composite domain of an OD and a sterile alpha motif (SAM) domain, and forms dimers instead of tetramers. The Dmp53 structure is characterized by an additional N-terminal beta-strand and a C-terminal helix. Truncation analysis in both domains reveals that the additional structural elements are necessary to stabilize the structure of the OD, suggesting a new function for the SAM domain. Furthermore, these structures show a potential path of evolution from an ancestral dimeric form over a tetrameric form, with additional stabilization elements, to the tetramerization domain of mammalian p53.  相似文献   

18.
The E1 and E2 proteins from bovine papillomavirus bind cooperatively to the viral origin of DNA replication (ori), forming a complex which is essential for initiation of DNA replication. Cooperative binding has two components, in which (i) the DNA binding domains (DBDs) of the two proteins interact with each other and (ii) the E2 transactivation domain interacts with the helicase domain of E1. By generating specific point mutations in the DBD of E2, we have defined two patches of amino acids that are involved in the interaction with the E1 DBD. These same mutations, when introduced into the viral genome, result in severely reduced replication of the viral genome, as well as failure to transform mouse cells in tissue culture. Thus, the interaction between the E1 and E2 DBDs is important for the establishment of the viral genome as an episome and most likely contributes to the formation of a preinitiation complex on the viral ori.  相似文献   

19.
The tumor suppressor protein, p53, selectively binds to supercoiled (sc) DNA lacking the specific p53 consensus binding sequence (p53CON). Using p53 deletion mutants, we have previously shown that the p53 C-terminal DNA-binding site (CTDBS) is critical for this binding. Here we studied supercoil-selective binding of bacterially expressed full-length p53 using modulation of activity of the p53 DNA-binding domains by oxidation of cysteine residues (to preclude binding within the p53 core domain) and/or by antibodies mapping to epitopes at the protein C-terminus (to block binding within the CTDBS). In the absence of antibody, reduced p53 preferentially bound scDNA lacking p53CON in the presence of 3 kb linear plasmid DNAs or 20 mer oligonucleotides, both containing and lacking the p53CON. Blocking the CTDBS with antibody caused reduced p53 to bind equally to sc and linear or relaxed circular DNA lacking p53CON, but with a high preference for the p53CON. The same immune complex of oxidized p53 failed to bind DNA, while oxidized p53 in the absence of antibody restored selective scDNA binding. Antibodies mapping outside the CTDBS did not prevent p53 supercoil-selective (SCS) binding. These data indicate that the CTDBS is primarily responsible for p53 SCS binding. In the absence of the SCS binding, p53 binds sc or linear (relaxed) DNA via the p53 core domain and exhibits strong sequence-specific binding. Our results support a hypothesis that alterations to DNA topology may be a component of the complex cellular regulatory mechanisms that control the switch between latent and active p53 following cellular stress.  相似文献   

20.
Autoinhibition of p53 binding to MDMX requires two short-linear motifs (SLiMs) containing adjacent tryptophan (WW) and tryptophan-phenylalanine (WF) residues. NMR spectroscopy was used to show the WW and WF motifs directly compete for the p53 binding site on MDMX and circular dichroism spectroscopy was used to show the WW motif becomes helical when it is bound to the p53 binding domain (p53BD) of MDMX. Binding studies using isothermal titration calorimetry showed the WW motif is a stronger inhibitor of p53 binding than the WF motif when they are both tethered to p53BD by the natural disordered linker. We also investigated how the WW and WF motifs interact with the DNA binding domain (DBD) of p53. Both motifs bind independently to similar sites on DBD that overlap the DNA binding site. Taken together our work defines a model for complex formation between MDMX and p53 where a pair of disordered SLiMs bind overlapping sites on both proteins.  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