51
|
Buganim Y, Rotter V. p53: Balancing tumour suppression and implications for the clinic. Eur J Cancer 2009; 45 Suppl 1:217-34. [DOI: 10.1016/s0959-8049(09)70037-1] [Citation(s) in RCA: 6] [Impact Index Per Article: 0.4] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 12/29/2022]
|
52
|
Schneider WM, Inouye M, Montelione GT, Roth MJ. Independently inducible system of gene expression for condensed single protein production (cSPP) suitable for high efficiency isotope enrichment. ACTA ACUST UNITED AC 2009; 10:219-25. [PMID: 19642019 DOI: 10.1007/s10969-009-9067-x] [Citation(s) in RCA: 13] [Impact Index Per Article: 0.8] [Reference Citation Analysis] [Abstract] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 03/25/2009] [Accepted: 07/04/2009] [Indexed: 11/29/2022]
Abstract
The ability to produce isotope-enriched proteins is fundamental to the success of modern protein NMR, and is particularly essential for NMR activities in structural genomics projects. Conventional methods of protein production often prove to be cost prohibitive for obtaining samples, particularly perdeuterated and site-specifically labeled proteins. The condensed single protein production system (cSPP), providing protein expression following condensation of cells 10-40 fold, allows for the production of such samples at a fraction of the cost. The previously described cSPP system is a two plasmid system where both the MazF toxin and ACA-less target gene are coinduced with IPTG. Coinduction results in 10-20% of the target protein produced without isotopic enrichment. Though the unlabeled protein is generally not visible in isotope-filtered NMR experiments, it results in an effective reduction in yield of the observable sample. By altering the cSPP system and separating the induction of the MazF toxin, required to convert cells into a semiquiescent state prior to condensation, from the expression of the target gene, we are now able to eliminate the unlabeled protein fraction and improve the isotope incorporation. Here we describe a series of pCold(tet) vectors with various features that can be used in the dual inducible cSPP(tet) system to obtain high-quality isotopically enriched protein at as little as 2.5% the cost of traditional methods.
Collapse
Affiliation(s)
- William M Schneider
- Department of Biochemistry, University of Medicine and Dentistry of New Jersey-Robert Wood Johnson Medical School, Piscataway, NJ 08854, USA
| | | | | | | |
Collapse
|
53
|
Pan Y, Nussinov R. Cooperativity dominates the genomic organization of p53-response elements: a mechanistic view. PLoS Comput Biol 2009; 5:e1000448. [PMID: 19629163 PMCID: PMC2705680 DOI: 10.1371/journal.pcbi.1000448] [Citation(s) in RCA: 17] [Impact Index Per Article: 1.1] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Download PDF] [Figures] [Journal Information] [Subscribe] [Scholar Register] [Received: 12/17/2008] [Accepted: 06/24/2009] [Indexed: 11/18/2022] Open
Abstract
p53-response elements (p53-REs) are organized as two repeats of a palindromic DNA segment spaced by 0 to 20 base pairs (bp). Several experiments indicate that in the vast majority of the human p53-REs there are no spacers between the two repeats; those with spacers, particularly with sizes beyond two nucleotides, are rare. This raises the question of what it indicates about the factors determining the p53-RE genomic organization. Clearly, given the double helical DNA conformation, the orientation of two p53 core domain dimers with respect to each other will vary depending on the spacer size: a small spacer of 0 to 2 bps will lead to the closest p53 dimer-dimer orientation; a 10-bp spacer will locate the p53 dimers on the same DNA face but necessitate DNA looping; while a 5-bp spacer will position the p53 dimers on opposite DNA faces. Here, via conformational analysis we show that when there are 0-2 bp spacers, p53-DNA binding is cooperative; however, cooperativity is greatly diminished when there are spacers with sizes beyond 2 bp. Cooperative binding is broadly recognized to be crucial for biological processes, including transcriptional regulation. Our results clearly indicate that cooperativity of the p53-DNA association dominates the genomic organization of the p53-REs, raising questions of the structural organization and functional roles of p53-REs with larger spacers. We further propose that a dynamic landscape scenario of p53 and p53-REs can better explain the selectivity of the degenerate p53-REs. Our conclusions bear on the evolutionary preference of the p53-RE organization and as such, are expected to have broad implications to other multimeric transcription factor response element organization.
Collapse
Affiliation(s)
- Yongping Pan
- Basic Research Program, SAIC-Frederick, Inc., Center for Cancer Research Nanobiology Program, NCI-Frederick, Frederick, Maryland, United States of America
| | - Ruth Nussinov
- Basic Research Program, SAIC-Frederick, Inc., Center for Cancer Research Nanobiology Program, NCI-Frederick, Frederick, Maryland, United States of America
- Sackler Institute of Molecular Medicine, Department of Human Genetics and Molecular Medicine, Sackler School of Medicine, Tel Aviv University, Tel Aviv, Israel
- * E-mail:
| |
Collapse
|
54
|
Abstract
The p53 gene is crucial for effective tumor suppression in humans as supported by its universal inactivation in cancer cells either through mutations affecting the p53 locus directly or through aberration of its normal regulation. The p53 tumor repressor is regulated through a negative feedback loop involving its transcriptional target MDM2. MDMX is also an essential negative regulator of p53. Several computational models have been proposed to simulate the dynamics of the p53-MDM2 loop, but they do not include MDMX, only account for some basic interactions between p53 and MDM2 and cannot capture the intrinsic noise in the loop. In this article, we present a comprehensive model for the p53-MDM2/MDMX loop that accounts for most known interactions among p53, MDM2 and MDMX. Our model is characterized by a set of molecular reactions, which enables us to employ stochastic simulation to investigate the dynamics of the loop. In agreement with experiments, our results show that p53 and MDM2 undergo oscillations after DNA damage in the presence of noise, and the variation in oscillation amplitudes is much higher than that in oscillation periods. Our simulations predict that intrinsic noise contributes to 60%-70% of the total variation in oscillation amplitudes and periods. The protein levels of p53, MDM2, and MDMX after treatment with Nutlin in our simulations are also consistent with experimental results. Our simulation results further predict that p53 levels increase dramatically after MDM2 is knocked out, but increase with a much less amount after MDMX is knocked out. This may partially explain why MDM2-null and MDMX-null mouse embryos die in different developmental stages. Our stochastic model and simulation provide insights into the variability of the behavior of the p53 pathway and can be used to predict the dynamics of the pathway after certain interventions.
Collapse
Affiliation(s)
- Xiaodong Cai
- Department of Electrical and Computer Engineering, University of Miami, Coral Gables, Florida 33146, USA.
| | | |
Collapse
|
55
|
Nomura T, Kamada R, Ito I, Chuman Y, Shimohigashi Y, Sakaguchi K. Oxidation of methionine residue at hydrophobic core destabilizes p53 tetrameric structure. Biopolymers 2009; 91:78-84. [PMID: 18781628 DOI: 10.1002/bip.21084] [Citation(s) in RCA: 19] [Impact Index Per Article: 1.2] [Reference Citation Analysis] [Abstract] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/05/2023]
Abstract
The tumor suppressor protein p53 is a tetrameric phosphoprotein that induces cell cycle, development, and differentiation by regulating the expression of target genes. The tetramerization of p53 is essential for its tumor suppressor functions. It has been known that oxidation of proteins affects their structure and function. A methionine residue (Met340) is located at the hydrophobic core in p53 tetramerization domain. Here, we demonstrated that Met340 residue can be oxidized to methionine sulfoxide under oxidative conditions and investigated effects of the oxidation of p53 tetramerization domain on its stability and oligomerization state by CD measurement and gel filtration. The oxidation of Met340 drastically induced destabilization of the p53 tetramer by 22.8 kJ/mol of DeltaDeltaG(Tm), while retaining the identical conformation as that of the wild-type peptide. Trypsin digestion experiments also showed that oxidation of Met340 allowed the peptide to form locally loose structure and become more sensitive to enzyme degradation. The tetrameric structure may be destabilized because the oxidation of Met340 induces charge repulsion and/or steric hindrance between the sulfoxide groups. These results taken together suggested that oxidation of methionine residues in the p53 protein might be one of the inactivation mechanisms of p53 transcriptional function under conditions of oxidative stress.
Collapse
Affiliation(s)
- Takao Nomura
- Department of Chemistry, Faculty of Science, Hokkaido University, Sapporo 060-0810, Japan
| | | | | | | | | | | |
Collapse
|
56
|
Fukuchi S, Homma K, Minezaki Y, Gojobori T, Nishikawa K. Development of an accurate classification system of proteins into structured and unstructured regions that uncovers novel structural domains: its application to human transcription factors. BMC STRUCTURAL BIOLOGY 2009; 9:26. [PMID: 19402914 PMCID: PMC2687452 DOI: 10.1186/1472-6807-9-26] [Citation(s) in RCA: 33] [Impact Index Per Article: 2.1] [Reference Citation Analysis] [Abstract] [Track Full Text] [Download PDF] [Figures] [Subscribe] [Scholar Register] [Received: 01/05/2009] [Accepted: 04/30/2009] [Indexed: 12/26/2022]
Abstract
BACKGROUND In addition to structural domains, most eukaryotic proteins possess intrinsically disordered (ID) regions. Although ID regions often play important functional roles, their accurate identification is difficult. As human transcription factors (TFs) constitute a typical group of proteins with long ID regions, we regarded them as a model of all proteins and attempted to accurately classify TFs into structural domains and ID regions. Although an extremely high fraction of ID regions besides DNA binding and/or other domains was detected in human TFs in our previous investigation, 20% of the residues were left unassigned. In this report, we exploit the generally higher sequence divergence in ID regions than in structural regions to completely divide proteins into structural domains and ID regions. RESULTS The new dichotomic system first identifies domains of known structures, followed by assignment of structural domains and ID regions with a combination of pre-existing tools and a newly developed program based on sequence divergence, taking un-aligned regions into consideration. The system was found to be highly accurate: its application to a set of proteins with experimentally verified ID regions had an error rate as low as 2%. Application of this system to human TFs (401 proteins) showed that 38% of the residues were in structural domains, while 62% were in ID regions. The preponderance of ID regions makes a sharp contrast to TFs of Escherichia coli (229 proteins), in which only 5% fell in ID regions. The method also revealed that 4.0% and 11.8% of the total length in human and E. coli TFs, respectively, are comprised of structural domains whose structures have not been determined. CONCLUSION The present system verifies that sequence divergence including information of unaligned regions is a good indicator of ID regions. The system for the first time estimates the complete fractioning of structured/un-structured regions in human TFs, also revealing structural domains without homology to known structures. These predicted novel structural domains are good targets of structural genomics. When applied to other proteins, the system is expected to uncover more novel structural domains.
Collapse
Affiliation(s)
- Satoshi Fukuchi
- Center for Information Biology & DNA Data Bank of Japan, National Institute of Genetics, Mishima, Shizuoka, Japan.
| | | | | | | | | |
Collapse
|
57
|
Yi F, Saha A, Murakami M, Kumar P, Knight JS, Cai Q, Choudhuri T, Robertson ES. Epstein-Barr virus nuclear antigen 3C targets p53 and modulates its transcriptional and apoptotic activities. Virology 2009; 388:236-47. [PMID: 19394062 DOI: 10.1016/j.virol.2009.03.027] [Citation(s) in RCA: 81] [Impact Index Per Article: 5.1] [Reference Citation Analysis] [Abstract] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 11/04/2008] [Revised: 03/09/2009] [Accepted: 03/21/2009] [Indexed: 01/17/2023]
Abstract
The p53 tumor suppressor gene is one of the most commonly mutated genes in human cancers and the corresponding encoded protein induces apoptosis or cell-cycle arrest at the G1/S checkpoint in response to DNA damage. To date, previous studies have shown that antigens encoded by human tumor viruses such as SV40 large T antigen, adenovirus E1A and HPV E6 interact with p53 and disrupt its functional activity. In a similar fashion, we now show that EBNA3C, one of the EBV latent antigens essential for the B-cell immortalization in vitro, interacts directly with p53. Additionally, we mapped the interaction of EBNA3C with p53 to the C-terminal DNA-binding and the tetramerization domain of p53, and the region of EBNA3C responsible for binding to p53 was mapped to the N-terminal domain of EBNA3C (residues 130-190), previously shown to interact with a number of important cell-cycle components, specifically SCF(Skp2), cyclin A, and cMyc. Furthermore, we demonstrate that EBNA3C substantially represses the transcriptional activity of p53 in luciferase based reporter assays, and rescues apoptosis induced by ectopic p53 expression in SAOS-2 (p53(-/-)) cells. Interestingly, we also show that the DNA-binding ability of p53 is diminished in the presence of EBNA3C. Thus, the interaction between the p53 and EBNA3C provides new insights into the mechanism(s) by which the EBNA3C oncoprotein can alter cellular gene expression in EBV associated human cancers.
Collapse
Affiliation(s)
- Fuming Yi
- Department of Microbiology and Tumor Virology Program, Abramson Comprehensive Cancer Center, University of Pennsylvania Medical School, 201E Johnson Pavilion, 3610 Hamilton Walk, PA 19104, USA
| | | | | | | | | | | | | | | |
Collapse
|
58
|
Itahana Y, Ke H, Zhang Y. p53 Oligomerization is essential for its C-terminal lysine acetylation. J Biol Chem 2009; 284:5158-64. [PMID: 19106109 PMCID: PMC2643511 DOI: 10.1074/jbc.m805696200] [Citation(s) in RCA: 57] [Impact Index Per Article: 3.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 07/24/2008] [Revised: 12/03/2008] [Indexed: 11/06/2022] Open
Abstract
Acetylation of multiple lysine residues in the p53 plays critical roles in the protein stability and transcriptional activity of p53. To better understand how p53 acetylation is regulated, we generated a number of p53 mutants and examined acetylation of each mutant in transfected cells. We found that p53 mutants that are defective in tetramer formation are also defective in C-terminal lysine residue acetylation. Consistently, we found that several cancer-derived p53 mutants that bear mutations in the tetramerization domain cannot form oligomers and are defective in C-terminal lysine acetylation, and these mutants are inactive in p21 transactivation. We demonstrated that the acetyltransferase p300 interacts with and promotes acetylation of wild-type p53 but not with any of the artificially generated or human cancer-derived p53 mutants that are defective in oligomerization. These results, combined with a computer-aided crystal structure analysis, suggest a model in which p53 oligomerization precedes its acetylation by providing docking sites for acetyltransferases.
Collapse
Affiliation(s)
- Yoko Itahana
- Department of Radiation Oncology, University of North Carolina, Chapel Hill, North Carolina 27599-7512, USA
| | | | | |
Collapse
|
59
|
Cai BH, Chen JY, Lu MH, Chang LT, Lin HC, Chang YM, Chao CF. Functional four-base A/T gap core sequence CATTAG of P53 response elements specifically bound tetrameric P53 differently than two-base A/T gap core sequence CATG bound both dimeric and tetrameric P53. Nucleic Acids Res 2009; 37:1984-90. [PMID: 19208646 PMCID: PMC2665222 DOI: 10.1093/nar/gkp033] [Citation(s) in RCA: 19] [Impact Index Per Article: 1.2] [Reference Citation Analysis] [Abstract] [Track Full Text] [Download PDF] [Figures] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/14/2022] Open
Abstract
The consensus sequence of p53 is repeated half sites of PuPuPuC(A/T)(A/T)GPyPyPy. GtAGCAttAGCCCAGACATGTCC is a 14-3-3sigma promoter p53 regulation site; the first core sequence is CAttAG, and the second is CATG. Both mutants GtAGgAttAGCCCAGACATGTCC and GtAGCAttAGCCCAGACATcTCC can be activated by p53 as a 1.5-fold half site. The original p53 regulated site on the 14-3-3sigma promoter is a whole site, and CATTAG is a functional core sequence. The p53-binding affinity and the activity of CATTAG were lower than for the mutant CATATG core sequence. Wild-type p53 acts as a tetramer to bind to the whole site; however, it also can bind to a half site by one of its dimers. Wild-type p53 can only bind to a half site with core sequence CATG but not to CATATG. The 1.5-fold half site or whole site with core sequence CATATG can be bound by wild-type p53. A p53 mutant, A344, forms dimeric p53; it can only bind to CATG, and not to CATATG. Therefore, tetrameric and dimeric p53 can bind to a two-base A/T gap core sequence, but only tetrameric p53 can bind to a four-base A/T gap core sequence.
Collapse
Affiliation(s)
- Bi-He Cai
- National Defense Medical Center, Institute of Life Sciences, Taipei, Taiwan, ROC
| | | | | | | | | | | | | |
Collapse
|
60
|
Gabizon R, Mor M, Rosenberg MM, Britan L, Hayouka Z, Kotler M, Shalev DE, Friedler A. Using peptides to study the interaction between the p53 tetramerization domain and HIV-1 Tat. Biopolymers 2008; 90:105-16. [PMID: 18189286 DOI: 10.1002/bip.20919] [Citation(s) in RCA: 21] [Impact Index Per Article: 1.2] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/07/2022]
Abstract
Peptides are valuable tools for studying protein-protein interactions, especially in cases of isolated protein domains and natively unfolded proteins. Here, we used peptides to quantitatively characterize the interaction between the natively unfolded HIV-1 Tat protein and the tetramerization domain of the cellular tumor suppressor protein p53. We used peptide mapping, fluorescence anisotropy, and NMR spectroscopy to perform a detailed structural and biophysical characterization of the interaction between the two proteins and elucidate its molecular mechanism, which have so far been studied using cell-based methods. We show that the p53 tetramerization domain, p53(326-355), binds directly to residues 1-35 and 47-57 in Tat. We have characterized the interaction between p53(326-355) and Tat(47-57) in detail. The p53 residues that are mainly involved in binding to Tat(47-57) are E343 and E349, which bind to the positively charged arginine-rich motif of Tat by a partly electrostatic mechanism. All oligomerization states of p53(326-355) bind Tat(47-57) without inhibiting p53 tetramerization, since the residues in p53(326-355) that bind Tat(47-57) face away from the tetramerization interface. We conclude that p53 is able to bind Tat as a transcriptionally active tetramer.
Collapse
Affiliation(s)
- Ronen Gabizon
- Institute of Chemistry, The Hebrew University of Jerusalem, Safra Campus, Givat Ram, Jerusalem 91904, Israel
| | | | | | | | | | | | | | | |
Collapse
|
61
|
Pan Y, Nussinov R. p53-Induced DNA bending: the interplay between p53-DNA and p53-p53 interactions. J Phys Chem B 2008; 112:6716-24. [PMID: 18461991 DOI: 10.1021/jp800680w] [Citation(s) in RCA: 27] [Impact Index Per Article: 1.6] [Reference Citation Analysis] [Abstract] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/19/2023]
Abstract
Specific p53 binding-induced DNA bending and its underlying responsible forces are crucial for the understanding of selective transcription activation. Diverse p53-response elements exist in the genome; however, it is not known what determines the DNA bending and to what extent. In order to gain knowledge of the forces that govern the DNA bending, molecular dynamics simulations were performed on a series of p53 core domain tetramer-DNA complexes in which each p53 core domain was bound to a DNA quarter site specifically. By varying the sequence of the central 4-base pairs of each half-site, different DNA bending extents were observed. The analysis showed that the dimer-dimer interactions in p53 were similar for the complexes; on the other hand, the specific interactions between the p53 and DNA, including the interactions of Arg280, Lys120, and Arg248 with the DNA, varied more significantly. In particular, the Arg280 interactions were better maintained in the complex with the CATG-containing DNA sequence and were mostly lost in the complex with the CTAG-containing DNA sequence. Structural analysis shows that the base pairings for the CATG sequence were stable throughout the simulation trajectory, whereas those for the CTAG sequence were partially dissociated in part of the trajectory, which affected the stability of the nearby Arg280-Gua base interactions. Thus, DNA bending depends on the balance between the p53 dimer-dimer interactions and p53-DNA interactions, which is in turn related to the DNA sequence and DNA flexibility.
Collapse
Affiliation(s)
- Yongping Pan
- Center for Cancer Research Nanobiology Program, SAIC-Frederick, Inc. NCI-Frederick, Frederick, Maryland 21702, USA
| | | |
Collapse
|
62
|
De Grandis V, Bizzarri AR, Cannistraro S. Docking study and free energy simulation of the complex between p53 DNA-binding domain and azurin. J Mol Recognit 2007; 20:215-26. [PMID: 17703463 DOI: 10.1002/jmr.840] [Citation(s) in RCA: 50] [Impact Index Per Article: 2.8] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/23/2023]
Abstract
Molecular interaction between p53 tumor suppressor and the copper protein azurin (AZ) has been demonstrated to enhance p53 stability and hence antitumoral function, opening new perspectives in cancer treatment. While some experimental work has provided evidence for AZ binding to p53, no crystal structure for the p53-AZ complex was solved thus far. In this work the association between AZ and the p53 DNA-binding domain (DBD) was investigated by computational methods. Using a combination of rigid-body protein docking, experimental mutagenesis information, and cluster analysis 10 main p53 DBD-AZ binding modes were generated. The resulting structures were further characterized by molecular dynamics (MD) simulations and free energy calculations. We found that the highest scored docking conformation for the p53 DBD-AZ complex also yielded the most favorable free energy value. This best three-dimensional model for the complex was validated by using a computational mutagenesis strategy. In this structure AZ binds to the flexible L(1) and s(7)-s(8) loops of the p53 DBD and stabilizes them through protein-protein tight packing interactions, resulting in high degree of both surface matching and electrostatic complementarity.
Collapse
Affiliation(s)
- Valentina De Grandis
- Biophysics and Nanoscience Centre, CNISM, Facoltà di Scienze, Università della Tuscia, Largo dell'Università-I-01100 Viterbo, Italy
| | | | | |
Collapse
|
63
|
Enhancement of oligomeric stability by covalent linkage and its application to the human p53tet domain: thermodynamics and biological implications. Biochem Soc Trans 2007; 35:1574-8. [DOI: 10.1042/bst0351574] [Citation(s) in RCA: 13] [Impact Index Per Article: 0.7] [Reference Citation Analysis] [Abstract] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/17/2022]
Abstract
The formation of oligomeric proteins proceeds at a major cost of reducing the translational and rotational entropy for their subunits in order to form the stabilizing interactions found in the oligomeric state. Unlike site-directed mutations, covalent linkage of subunits represents a generically applicable strategy for enhancing oligomeric stability by reducing the entropic driving force for dissociation. Although this can be realized by introducing de novo disulfide cross-links between subunits, issues with irreversible aggregation limit the utility of this approach. In contrast, tandem linkage of subunits in a single polypeptide chain offers a universal method of pre-paying the entropic cost of oligomer formation. In the present paper, thermodynamic, structural and experimental aspects of designing and characterizing tandem-linked oligomers are discussed with reference to engineering a stabilized tetramer of the oligomerization domain of the human p53 tumour-suppressor protein by tandem dimerization.
Collapse
|
64
|
Lwin TZ, Durant JJ, Bashford D. A fluid salt-bridging cluster and the stabilization of p53. J Mol Biol 2007; 373:1334-47. [PMID: 17900613 PMCID: PMC2104543 DOI: 10.1016/j.jmb.2007.07.080] [Citation(s) in RCA: 17] [Impact Index Per Article: 0.9] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 05/09/2007] [Accepted: 07/31/2007] [Indexed: 11/22/2022]
Abstract
p53 is a homotetrameric tumor suppressor protein that is found to be mutated in most human cancers. Some of these mutations, particularly mutations to R337, fall in the tetramerization domain and cause defects in tetramer formation leading to loss of function. Mutation to His at this site has been found to destabilize the tetramer in a pH-dependent fashion. In structures of the tetramerization domain determined by crystallography, R337 from one monomer makes a salt bridge with D352 from another monomer, apparently helping to stabilize the tetramer. Here we present molecular dynamics simulations of wild-type p53 and the R337His mutant at several different pH and salt conditions. We find that the 337-352 salt bridge is joined by two other charged side chains, R333 and E349. These four residues do not settle into a fixed pattern of salt bridging, but continue to exchange salt-bridging partners on the nanosecond time scale throughout the simulation. This unusual system of fluid salt bridging may explain the previous finding from alanine scanning experiments that R333 contributes significantly to protein stability, even though in the crystal structure it is extended outward into solvent. This extended conformation of R333 appears to be the result of a specific crystal contact and, this contact being absent in the simulation, R333 turns inward to join its interaction partners. When R337 is mutated to His but remains positively charged, it maintains the original interaction with D352, but the newly observed interaction with E349 is weakened, accounting for the reduced stability of R337H even under mildly acidic conditions. When this His is deprotonated, the interaction with D352 is also lost, accounting for the further destabilization observed under mildly alkaline conditions. Simulations were carried out using both explicit and implicit solvent models, and both displayed similar behavior of the fluid salt-bridging cluster, suggesting that implicit solvent models can capture at least the qualitative features of this phenomenon as well as explicit solvent. Simulations under strongly acidic conditions in implicit solvent displayed the beginnings of the unfolding process, a destabilization of the hydrophobic dimer-dimer interface. Computational alanine scanning using the molecular mechanics Poisson-Boltzmann surface area method showed significant correlation to experimental unfolding data for charged and polar residues, but much weaker correlation for hydrophobic residues.
Collapse
Affiliation(s)
- Thu Zar Lwin
- Hartwell Center for Bioinformatics and Biotechnology, Saint Jude Children's Research Hospital, 322 N. Lauderdale St., Mail Stop 312, Memphis, TN 38105, USA.
| | | | | |
Collapse
|
65
|
Baryshnikova EN, Balobanov VA, Katina NS, Melnik BS, Dolgikh DA, Semisotnov GV, Bychkova VE. Equilibrium unfolding of mutant apomyoglobins carrying substitutions of conserved nonfunctional residues with alanine. Mol Biol 2007. [DOI: 10.1134/s0026893307040139] [Citation(s) in RCA: 3] [Impact Index Per Article: 0.2] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/22/2022]
|
66
|
Wahl GM. Mouse bites dogma: how mouse models are changing our views of how P53 is regulated in vivo. Cell Death Differ 2007; 13:973-83. [PMID: 16575406 DOI: 10.1038/sj.cdd.4401911] [Citation(s) in RCA: 26] [Impact Index Per Article: 1.4] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 12/16/2022] Open
Abstract
P53 is a transcription factor that can cause cells to be eliminated by apoptosis or senescent-like arrest upon its activation by irreparable genetic damage, excessively expressed oncogenes, or a broad spectrum of other stresses. As P53 executes life and death decisions, its activity must be stringently regulated, which implies that it is not likely to be controlled by a simple regulatory mechanism involving a binary on-off switch. This brief review will summarize a subset of the new information presented at the 10th P53 workshop in Dunedin, New Zealand in November 2004 as well as very recent publications that provide new insights into the molecular regulators of P53. Data emerging from mouse models provide a fundamentally different view of how P53 is regulated than suggested by more traditional in vitro approaches. The differences between cell culture and mouse models demonstrate the importance of preserving stoichiometric relationships between P53 and its various regulators to obtain an accurate view of the relevant molecular mechanisms that control P53 activity.
Collapse
Affiliation(s)
- G M Wahl
- Salk Institute for Biological Studies, Gene Expression Laboratory, 10010 N. Torrey Pines Road, La Jolla, CA 92037, USA.
| |
Collapse
|
67
|
Bu W, Carroll KD, Palmeri D, Lukac DM. Kaposi's sarcoma-associated herpesvirus/human herpesvirus 8 ORF50/Rta lytic switch protein functions as a tetramer. J Virol 2007; 81:5788-806. [PMID: 17392367 PMCID: PMC1900300 DOI: 10.1128/jvi.00140-07] [Citation(s) in RCA: 31] [Impact Index Per Article: 1.7] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 12/22/2022] Open
Abstract
The Kaposi's sarcoma-associated herpesvirus open reading frame 50 (ORF50) protein (called Rta), is necessary and sufficient for reactivation of the virus from latency. We previously demonstrated that a truncated mutant of ORF50 lacking its C-terminal transcriptional activation domain, called ORF50DeltaSTAD, formed mixed multimers with wild-type (WT) ORF50 and functioned as a dominant negative inhibitor of reactivation. For this report, we investigated the requirements for multimerization of ORF50/Rta in transactivation and viral reactivation. We analyzed multimerization of WT, mutant, and chimeric ORF50 proteins, using Blue Native polyacrylamide gel electrophoresis and size exclusion chromatography. WT and mutant ORF50 proteins form tetramers and higher-order multimers, but not monomers, in solution. The proline-rich, N-terminal leucine heptapeptide repeat (LR) of ORF50 (amino acids [aa] 244 to 275) is necessary but not sufficient for oligomer formation and functions in concert with the central portion of ORF50/Rta (aa 245 to 414). The dominant negative mutant ORF50DeltaSTAD requires the LR to form mixed multimers with WT ORF50 and inhibit its function. In the context of the WT ORF50/Rta protein, mutagenesis of the LR, or replacement of the LR by heterologous multimerization domains from the GCN4 or p53 proteins, demonstrates that tetramers of Rta are sufficient for transactivation and viral reactivation. Mutants of Rta that are unable to form tetramers but retain the ability to form higher-order multimers are reduced in function or are nonfunctional. We concluded that the proline content, but not the leucine content, of the LR is critical for determining the oligomeric state of Rta.
Collapse
Affiliation(s)
- Wei Bu
- University of Medicine and Dentistry of New Jersey/New Jersey Medical School, Department of Microbiology and Molecular Genetics and Graduate School of Biomedical Sciences, Newark, NJ 07103, USA
| | | | | | | |
Collapse
|
68
|
Abstract
Specific p53 binding-induced DNA bending has important biological implications such as transcription activation. However, the detailed structures of the bent DNA and the p53-DNA complex are still unavailable, hampering our understanding of the mechanism for p53-induced DNA bending and its consequent biological significance. To gain insight into the p53 binding-induced DNA bending, we performed molecular dynamics simulations on DNA segments with the consensus sequence for p53-specific binding, half site DNA-p53 complexes, and full site DNA-p53 complexes. We show that each DNA-bound p53 core domain caused a local DNA conformational change within the quarter site; upon the binding of the p53 dimer, there was an apparent DNA bending at the center of the half site; when bound with two p53 dimers, the full site DNAs with two different sequences bent 20 and 35 degrees, respectively. These results are in agreement with experimental observations. Our simulations demonstrate that the two p53 dimers favored a staggered conformation in which they make favorable interactions at the interface. This dimer-dimer interface organization necessitated conformational changes in the DNA, leading to the bending at the center of the full site, which in turn is dependent on the DNA sequence. Overall, our results provide the detailed atomic model for the DNA-p53 tetramer complex and delineate the roles of DNA-p53, p53 dimer-dimer interactions, and DNA sequence in specific p53 binding-induced DNA conformational changes.
Collapse
Affiliation(s)
- Yongping Pan
- Center for Cancer Research Nanobiology Program, SAIC-Frederick, Inc., NCI-Frederick, National Institutes of Health, Frederick, Maryland 21702, USA
| | | |
Collapse
|
69
|
Wagner J, Ma L, Rice JJ, Hu W, Levine AJ, Stolovitzky GA. p53-Mdm2 loop controlled by a balance of its feedback strength and effective dampening using ATM and delayed feedback. ACTA ACUST UNITED AC 2006; 152:109-18. [PMID: 16986275 DOI: 10.1049/ip-syb:20050025] [Citation(s) in RCA: 59] [Impact Index Per Article: 3.1] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/19/2022]
Abstract
When the genomic integrity of a cell is challenged, its fate is determined in part by signals conveyed by the p53 tumour suppressor protein. It was observed recently that such signals are not simple gradations of p53 concentration, but rather a counter-intuitive limit-cycle behaviour. Based on a careful mathematical interpretation of the experimental body of knowledge, we propose a model for the p53 signalling network and characterise the p53 stability and oscillatory dynamics. In our model, ATM, a protein that senses DNA damage, activates p53 by phosphorylation. In its active state, p53 has a decreased degradation rate and an enhanced transactivation of Mdm2, a gene whose protein product Mdm2 tags p53 for degradation. Thus the p53-Mdm2 system forms a negative feedback loop. However, the feedback in this loop is delayed, as the pool of Mdm2 molecules being induced by p53 at a given time will mark for degradation the pool of p53 molecules at some later time, after the Mdm2 molecules have been transcribed, exported out of the nucleus, translated and transported back into the nucleus. The analysis of our model demonstrates how this time lag combines with the ATM-controlled feedback strength and effective dampening of the negative feedback loop to produce limit-cycle oscillations. The picture that emerges is that ATM, once activated by DNA damage, makes the p53-Mdm2 oscillator undergo a supercritical Hopf bifurcation. This approach yields an improved understanding of the global dynamics and bifurcation structure of our time-delayed, negative feedback model and allows for predictions of the behaviour of the p53 system under different perturbations.
Collapse
Affiliation(s)
- J Wagner
- IBM Computational Biology Center, IBM TJ Watson Research Center, Yorktown Heights, NY 10598, USA
| | | | | | | | | | | |
Collapse
|
70
|
Wilder PT, Lin J, Bair CL, Charpentier TH, Yang D, Liriano M, Varney KM, Lee A, Oppenheim AB, Adhya S, Carrier F, Weber DJ. Recognition of the tumor suppressor protein p53 and other protein targets by the calcium-binding protein S100B. BIOCHIMICA ET BIOPHYSICA ACTA-MOLECULAR CELL RESEARCH 2006; 1763:1284-97. [PMID: 17010455 DOI: 10.1016/j.bbamcr.2006.08.024] [Citation(s) in RCA: 72] [Impact Index Per Article: 3.8] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Subscribe] [Scholar Register] [Received: 07/15/2006] [Revised: 08/16/2006] [Accepted: 08/17/2006] [Indexed: 01/10/2023]
Abstract
S100B is an EF-hand containing calcium-binding protein of the S100 protein family that exerts its biological effect by binding and affecting various target proteins. A consensus sequence for S100B target proteins was published as (K/R)(L/I)xWxxIL and matches a region in the actin capping protein CapZ (V.V. Ivanenkov, G.A. Jamieson, Jr., E. Gruenstein, R.V. Dimlich, Characterization of S-100b binding epitopes. Identification of a novel target, the actin capping protein, CapZ, J. Biol. Chem. 270 (1995) 14651-14658). Several additional S100B targets are known including p53, a nuclear Dbf2 related (NDR) kinase, the RAGE receptor, neuromodulin, protein kinase C, and others. Examining the binding sites of such targets and new protein sequence searches provided additional potential target proteins for S100B including Hdm2 and Hdm4, which were both found to bind S100B in a calcium-dependent manner. The interaction between S100B and the Hdm2 and/or the Hdm4 proteins may be important physiologically in light of evidence that like Hdm2, S100B also contributes to lowering protein levels of the tumor suppressor protein, p53. For the S100B-p53 interaction, it was found that phosphorylation of specific serine and/or threonine residues reduces the affinity of the S100B-p53 interaction by as much as an order of magnitude, and is important for protecting p53 from S100B-dependent down-regulation, a scenario that is similar to what is found for the Hdm2-p53 complex.
Collapse
Affiliation(s)
- Paul T Wilder
- Department of Biochemistry and Molecular Biology, The University of Maryland School of Medicine, 108 N. Greene Street, Baltimore, MD 21201, USA
| | | | | | | | | | | | | | | | | | | | | | | |
Collapse
|
71
|
Hussain SP, Harris CC. p53 biological network: at the crossroads of the cellular-stress response pathway and molecular carcinogenesis. J NIPPON MED SCH 2006; 73:54-64. [PMID: 16641528 DOI: 10.1272/jnms.73.54] [Citation(s) in RCA: 74] [Impact Index Per Article: 3.9] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/10/2023]
Abstract
p53 as a key molecular node in the stress response pathway, including inflammation. p53 is involved in several critical pathways including cell cycle arrest, apoptosis, DNA repair, and cellular senescence, which are essential for normal cellular homeostasis and maintaining genome integrity. The alteration of the TP53 gene or posttranslational modification in the p53 protein can alter its response to cellular stress. The molecular archaeology of the TP53 mutation spectrum generates hypotheses concerning the etiology and molecular pathogenesis of human cancer. The spectrum of somatic mutations in the TP53 gene implicates environmental carcinogens, and both endogenous agents and processes in the etiology of human cancer.
Collapse
Affiliation(s)
- S Perwez Hussain
- Laboratory of Human Carcinogenesis, Center for Cancer Research, National Cancer Institute, Bethesda, MD 20892-4255, USA
| | | |
Collapse
|
72
|
Kantarci N, Doruker P, Haliloglu T. Cooperative fluctuations point to the dimerization interface of p53 core domain. Biophys J 2006; 91:421-32. [PMID: 16807229 PMCID: PMC1483080 DOI: 10.1529/biophysj.106.077800] [Citation(s) in RCA: 11] [Impact Index Per Article: 0.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 11/21/2005] [Accepted: 03/21/2006] [Indexed: 11/18/2022] Open
Abstract
Elastic network models are used for investigation of the p53 core domain functional dynamics. Global modes of motion indicate high positive correlations for residue fluctuations across the A-B interface, which are not observed at the B-C interface. Major hinge formation is observed at the A-B interface upon dimerization indicating stability of the A-B dimer. These findings imply A-B as the native dimerization interface, whereas B-C is the crystal interface. The A-B dimer exhibits an opening-closing motion about DNA, supporting the previously suggested clamp-like model of nonspecific DNA binding followed by diffusion. Monomer A has limited positive correlations with DNA, while monomer B exhibits high positive correlations with DNA in the functionally significant slow modes. Thus, monomer B might seem to maintain the stability of the dimer-DNA complex by forming the relatively fixed arm of the dimer clamp, whereas the other arm of the clamp, monomer A, might allow sliding via continuous association/dissociation mechanisms.
Collapse
Affiliation(s)
- Nigar Kantarci
- Department of Chemical Engineering and Polymer Research Center, Bogazici University, Istanbul, Turkey
| | | | | |
Collapse
|
73
|
Hinow P, Rogers CE, Barbieri CE, Pietenpol JA, Kenworthy AK, DiBenedetto E. The DNA binding activity of p53 displays reaction-diffusion kinetics. Biophys J 2006; 91:330-42. [PMID: 16603489 PMCID: PMC1479054 DOI: 10.1529/biophysj.105.078303] [Citation(s) in RCA: 63] [Impact Index Per Article: 3.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/18/2022] Open
Abstract
The tumor suppressor protein p53 plays a key role in maintaining the genomic stability of mammalian cells and preventing malignant transformation. In this study, we investigated the intracellular diffusion of a p53-GFP fusion protein using confocal fluorescence recovery after photobleaching. We show that the diffusion of p53-GFP within the nucleus is well described by a mathematical model for diffusion of particles that bind temporarily to a spatially homogeneous immobile structure with binding and release rates k1 and k2, respectively. The diffusion constant of p53-GFP was estimated to be Dp53-GFP=15.4 microm2 s-1, significantly slower than that of GFP alone, DGFP=41.6 microm2 s-1. The reaction rates of the binding and unbinding of p53-GFP were estimated as k1=0.3 s-1 and k2=0.4 s-1, respectively, values suggestive of nonspecific binding. Consistent with this finding, the diffusional mobilities of tumor-derived sequence-specific DNA binding mutants of p53 were indistinguishable from that of the wild-type protein. These data are consistent with a model in which, under steady-state conditions, p53 is latent and continuously scans DNA, requiring activation for sequence-specific DNA binding.
Collapse
Affiliation(s)
- Peter Hinow
- Department of Mathematics, Vanderbilt University, and Department of Biochemistry, Center in Molecular Toxicology, Vanderbilt-Ingram Cancer Center, Nashville, Tennessee, USA
| | | | | | | | | | | |
Collapse
|
74
|
Riley KJL, Cassiday LA, Kumar A, Maher LJ. Recognition of RNA by the p53 tumor suppressor protein in the yeast three-hybrid system. RNA (NEW YORK, N.Y.) 2006; 12:620-30. [PMID: 16581806 PMCID: PMC1421098 DOI: 10.1261/rna.2286706] [Citation(s) in RCA: 17] [Impact Index Per Article: 0.9] [Reference Citation Analysis] [Abstract] [Key Words] [MESH Headings] [Grants] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 05/08/2023]
Abstract
The p53 tumor suppressor protein is a homotetrameric transcription factor whose gene is mutated in nearly half of all human cancers. In an unrelated screen of RNA/protein interactions using the yeast three-hybrid system, we inadvertently detected p53 interactions with several different RNAs. A literature review revealed previous reports of both sequence-specific and -non-specific interactions between p53 and RNA. Using yeast three-hybrid selections to identify preferred RNA partners for p53, we failed to identify primary RNA sequences or obvious secondary structures required for p53 binding. The cationic p53 C-terminus was shown to be required for RNA binding in yeast. We show that while p53 strongly discriminates between certain RNAs in the yeast three-hybrid assay, the same RNAs are bound equally by p53 in vitro. We further show that the p53 RNA-binding preferences in yeast are mirrored almost exactly by a recombinant tetrameric form of the HIV-1 nucleocapsid (NC) protein thought to be a sequence-nonspecific RNA-binding protein. However, the possibility of specific RNA binding by p53 could not be ruled out because p53 and HIV-1 NC displayed certain differences in RNA-binding preference. We conclude that (1) p53 binds RNA in vivo, (2) RNA binding by p53 is largely sequence-nonspecific in the yeast nucleus, (3) some structure-specific RNA binding by p53 cannot be ruled out, and (4) caution is required when interpreting results of RNA screens in the yeast three-hybrid system because sequence-dependent differences in RNA folding and display can masquerade as sequence-dependent differences in protein recognition.
Collapse
Affiliation(s)
- Kasandra J-L Riley
- Department of Biochemistry and Molecular Biology, Guggenheim 16, Mayo Clinic College of Medicine, 200 First Street SW, Rochester, Minnesota 55905, USA
| | | | | | | |
Collapse
|
75
|
Nabuurs SB, Spronk CAEM, Vuister GW, Vriend G. Traditional biomolecular structure determination by NMR spectroscopy allows for major errors. PLoS Comput Biol 2006; 2:e9. [PMID: 16462939 PMCID: PMC1359070 DOI: 10.1371/journal.pcbi.0020009] [Citation(s) in RCA: 71] [Impact Index Per Article: 3.7] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Download PDF] [Figures] [Journal Information] [Subscribe] [Scholar Register] [Received: 09/28/2005] [Accepted: 12/29/2005] [Indexed: 12/04/2022] Open
Abstract
One of the major goals of structural genomics projects is to determine the three-dimensional structure of representative members of as many different fold families as possible. Comparative modeling is expected to fill the remaining gaps by providing structural models of homologs of the experimentally determined proteins. However, for such an approach to be successful it is essential that the quality of the experimentally determined structures is adequate. In an attempt to build a homology model for the protein dynein light chain 2A (DLC2A) we found two potential templates, both experimentally determined nuclear magnetic resonance (NMR) structures originating from structural genomics efforts. Despite their high sequence identity (96%), the folds of the two structures are markedly different. This urged us to perform in-depth analyses of both structure ensembles and the deposited experimental data, the results of which clearly identify one of the two models as largely incorrect. Next, we analyzed the quality of a large set of recent NMR-derived structure ensembles originating from both structural genomics projects and individual structure determination groups. Unfortunately, a visual inspection of structures exhibiting lower quality scores than DLC2A reveals that the seriously flawed DLC2A structure is not an isolated incident. Overall, our results illustrate that the quality of NMR structures cannot be reliably evaluated using only traditional experimental input data and overall quality indicators as a reference and clearly demonstrate the urgent need for a tight integration of more sophisticated structure validation tools in NMR structure determination projects. In contrast to common methodologies where structures are typically evaluated as a whole, such tools should preferentially operate on a per-residue basis. Three-dimensional biomolecular structures provide an invaluable source of biologically relevant information. To be able to learn the most of the wealth of information that these structures can provide us, it is of great importance that the quality and accuracy of the protein structure models deposited in the Protein Data Bank are as high as possible. In this work, the authors describe an analysis that illustrates that this is unfortunately not the case for many protein structures solved using nuclear magnetic resonance spectroscopy. They present an example in which two strikingly different models describing the same protein are analyzed using commonly available structure validation tools, and the results of this analysis show one of the two models to be incorrect. Subsequently, using a large set of recently determined structures, the authors demonstrate that unfortunately this example does not stand on its own. The analyses and examples clearly illustrate that relying solely on the experimental data to evaluate structural quality can provide a false sense of correctness and the combination of multiple sophisticated structure validation tools is required to detect the presence of errors in protein nuclear magnetic resonance structures.
Collapse
Affiliation(s)
- Sander B Nabuurs
- Center for Molecular and Biomolecular Informatics, Nijmegen Center for Molecular Life Sciences, Radboud University, Nijmegen, Netherlands
| | - Chris A. E. M Spronk
- Center for Molecular and Biomolecular Informatics, Nijmegen Center for Molecular Life Sciences, Radboud University, Nijmegen, Netherlands
| | - Geerten W Vuister
- Department of Biophysical Chemistry, Institute for Molecules and Materials, Radboud University, Nijmegen, Netherlands
- * To whom correspondence should be addressed. E-mail: (GWV); (GV)
| | - Gert Vriend
- Center for Molecular and Biomolecular Informatics, Nijmegen Center for Molecular Life Sciences, Radboud University, Nijmegen, Netherlands
- * To whom correspondence should be addressed. E-mail: (GWV); (GV)
| |
Collapse
|
76
|
Galic V, Willner J, Wollan M, Garg R, Garcia R, Goff BA, Gray HJ, Swisher EM. Common polymorphisms inTP53 andMDM2 and the relationship toTP53 mutations and clinical outcomes in women with ovarian and peritoneal carcinomas. Genes Chromosomes Cancer 2006; 46:239-47. [PMID: 17171684 DOI: 10.1002/gcc.20407] [Citation(s) in RCA: 38] [Impact Index Per Article: 2.0] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/09/2022] Open
Abstract
The importance of somatic TP53 mutations and germline TP53 codon 72 genotype in the survival of women with epithelial ovarian cancer is controversial. Recent data suggest that a promoter polymorphism in the MDM2 gene may influence age of cancer onset in a gender-specific fashion. We sought to determine the relationship between somatic TP53 mutations, germline genotypes at TP53 codon 72 and MDM2 SNP309, and overall survival and response to chemotherapy in a large series of patients with ovarian and peritoneal carcinomas. Of the 188 cancers, 103 (54.8%) had a TP53 mutation, of which 71% were missense mutations and 29% were null mutations. TP53 mutation status and mutation type (null vs. missense) did not influence response to therapy or overall survival. Women with the codon 72 Pro/Pro had a decreased overall survival (median, 29 months) compared with women with one or two arginine alleles (median, 49 months; P=0.04). Somatic mutation or deletion was equally common for either codon 72 allele. Age of diagnosis was not influenced by codon 72 but showed a trend for younger age in women with somatic TP53 mutations and the MDM2 G/G genotype.
Collapse
Affiliation(s)
- Vijaya Galic
- Division of Gynecologic Oncology, Department of Obstetrics and Gynecology, University of Washington, Seattle, WA 98195, USA
| | | | | | | | | | | | | | | |
Collapse
|
77
|
Chi SW, Lee SH, Kim DH, Ahn MJ, Kim JS, Woo JY, Torizawa T, Kainosho M, Han KH. Structural Details on mdm2-p53 Interaction. J Biol Chem 2005; 280:38795-802. [PMID: 16159876 DOI: 10.1074/jbc.m508578200] [Citation(s) in RCA: 115] [Impact Index Per Article: 5.8] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/01/2023] Open
Abstract
Mdm2 is a cellular antagonist of p53 that keeps a balanced cellular level of p53. The two proteins are linked by a negative regulatory feedback loop and physically bind to each other via a putative helix formed by residues 18-26 of p53 transactivation domain (TAD) and its binding pocket located within the N-terminal 100-residue domain of mdm2 (Kussie, P. H., Gorina, S., Marechal, V., Elenbaas, B., Moreau, J., Levine, A. J., and Pavletich, N. P. (1996) Science 274, 948-953). In a previous report we demonstrated that p53 TAD in the mdm2-freee state is mostly unstructured but contains two nascent turns in addition to a "preformed" helix that is the same as the putative helix mediating p53-mdm2 binding. Here, using heteronuclear multidimensional NMR methods, we show that the two nascent turn motifs in p53 TAD, turn I (residues 40-45) and turn II (residues 49-54), are also capable of binding to mdm2. In particular, the turn II motif has a higher mdm2 binding affinity ( approximately 20 mum) than the turn I and targets the same site in mdm2 as the helix. Upon mdm2 binding this motif becomes a well defined full helix turn whose hydrophobic face formed by the side chains of Ile-50, Trp-53, and Phe-54 inserts deeply into the helix binding pocket. Our results suggest that p53-mdm2 binding is subtler than previously thought and involves global contacts such as multiple "non-contiguous" minimally structured motifs instead of being localized to one small helix mini-domain in p53 TAD.
Collapse
Affiliation(s)
- Seung-Wook Chi
- Protein Analysis and Design Laboratory, Division of Drug Discovery, Korea Research Institute of Bioscience and Biotechnology, Yusong P. O. Box 115, Daejon 305-600, Korea
| | | | | | | | | | | | | | | | | |
Collapse
|
78
|
Datta AB, Panjikar S, Weiss MS, Chakrabarti P, Parrack P. Structure of lambda CII: implications for recognition of direct-repeat DNA by an unusual tetrameric organization. Proc Natl Acad Sci U S A 2005; 102:11242-7. [PMID: 16061804 PMCID: PMC1183575 DOI: 10.1073/pnas.0504535102] [Citation(s) in RCA: 14] [Impact Index Per Article: 0.7] [Reference Citation Analysis] [Abstract] [Key Words] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 02/21/2005] [Indexed: 11/18/2022] Open
Abstract
The temperate coliphage lambda, after infecting its host bacterium Escherichia coli, can develop either along the lytic or the lysogenic pathway. Crucial to the lysis/lysogeny decision is the homotetrameric transcription-activator protein CII (4 x 11 kDa) of the phage that binds to a unique direct-repeat sequence T-T-G-C-N6-T-T-G-C at each of the three phage promoters it activates: p(E), p(I), and p(aQ). Several regions of CII have been identified for its various functions (DNA binding, oligomerization, and susceptibility to host protease), but the crystal structure of the protein long remained elusive. Here, we present the three-dimensional structure of CII at 2.6-angstroms resolution. The CII monomer is comprised of four alpha helices and a disordered C terminus. The first three helices (alpha1-alpha3) form a compact domain, whereas the fourth helix (alpha4) protrudes in different orientations in each subunit. A four-helix bundle, formed by alpha4 from each subunit, holds the tetramer. The quaternary structure can be described as a dimer of dimers, but the tetramer does not exhibit a closed symmetry. This unusual quaternary arrangement allows the placement of the helix-turn-helix motifs of two of the four CII subunits for interaction with successive major grooves of B-DNA, from one face of DNA. This structure provides a simple explanation for how a homotetrameric protein may recognize a direct-repeat DNA sequence rather than the inverted-repeat sequences of most prokaryotic activators.
Collapse
Affiliation(s)
- Ajit B Datta
- Department of Biochemistry, Bose Institute, P1/12 CIT Scheme VIIM, Calcutta 700 054, India
| | | | | | | | | |
Collapse
|
79
|
Ziebell MR, Prestwich GD. Interactions of peptide mimics of hyaluronic acid with the receptor for hyaluronan mediated motility (RHAMM). J Comput Aided Mol Des 2005; 18:597-614. [PMID: 15849992 DOI: 10.1007/s10822-004-5433-8] [Citation(s) in RCA: 16] [Impact Index Per Article: 0.8] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/26/2022]
Abstract
Using the hyaluronic acid (HA) binding region of the receptor for hyaluronan-mediated motility (RHAMM) as a model, a molecular perspective for peptide mimicry of the natural ligand was established by comparing the interaction sites of HA and unnatural peptide-ligands to RHAMM. This was accomplished by obtaining a series of octapeptide-ligands through screening experiments that bound to the HA binding domains of RHAMM (amino acids 517-576) and could be displaced by HA. These molecules were computationally docked onto a three-dimensional NMR based model of RHAMM. The NMR model showed that RHAMM(517-576) was a set of three helices, two of which contained the HA binding domains (HABDs) flanking a central groove. The structure was stabilized by hydrophobic interactions from four pairs of Val and Ile side chains extending into the groove. The presence of solvent exposed, positively charged side chains spaced 11 A apart matched the spacing of negative charges on HA. Docking experiments using flexible natural and artificial ligands demonstrated that HA and peptide-mimetics preferentially bound to the second helix that contains HABD-2. Three salt bridges between HA carboxylates and Lys548, Lys553 and Lys560 and two hydrophobic interactions involving Val538 and Val559 were predicted to stabilize the RHAMM-HA complex. The high affinity peptides and HA utilized the same charged residues, with additional contacts to other basic residues. However, hydrophobic contacts do not contribute to affinity for peptide ligand-RHAMM complexes. These results offer insight into how selectivity is achieved in the binding of HA to RHAMM, and how peptide competitors may compete for binding with HA on a single hyaladherin.
Collapse
Affiliation(s)
- Michael R Ziebell
- Department of Physiology and Biophysics, State University of New York, Stony Brook, NY 11794, USA.
| | | |
Collapse
|
80
|
Friedler A, Veprintsev DB, Freund SMV, von Glos KI, Fersht AR. Modulation of Binding of DNA to the C-Terminal Domain of p53 by Acetylation. Structure 2005; 13:629-36. [PMID: 15837201 DOI: 10.1016/j.str.2005.01.020] [Citation(s) in RCA: 73] [Impact Index Per Article: 3.7] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 10/04/2004] [Revised: 12/15/2004] [Accepted: 01/26/2005] [Indexed: 11/23/2022]
Abstract
The binding of nonspecific DNA to the C-terminal negative regulatory domain (CTD) of p53 modulates its activity. The CTD is a natively unfolded region, which is subject to acetylation and phosphorylation at several residues as part of control. To measure the effect of covalent modification on binding to DNA, we synthesized a series of fluorescein-labeled CTD peptides with single and multiple acetylations at lysine residues that we had identified by NMR as making contact with DNA, and developed an analytical ultracentrifugation method to study their binding to DNA. Binding depended on ionic strength, indicating an electrostatic contribution. Monoacetylation weakened DNA binding at physiological ionic strength 2- to 3-fold, diacetylations resulted in further 2- to 3-fold decrease in the affinity, and tri- and tetraacetylations rendered DNA binding undetectable. Phosphorylation at S392 did not affect DNA binding. NMR spectroscopy showed binding to DNA did not induce significant structure into CTD, apart possibly from local helix formation.
Collapse
Affiliation(s)
- Assaf Friedler
- MRC Centre for Protein Engineering, Hills Road, Cambridge CB2 2QH, United Kingdom
| | | | | | | | | |
Collapse
|
81
|
Ma B, Pan Y, Gunasekaran K, Venkataraghavan RB, Levine AJ, Nussinov R. Comparison of the protein-protein interfaces in the p53-DNA crystal structures: towards elucidation of the biological interface. Proc Natl Acad Sci U S A 2005; 102:3988-93. [PMID: 15738397 PMCID: PMC554805 DOI: 10.1073/pnas.0500215102] [Citation(s) in RCA: 25] [Impact Index Per Article: 1.3] [Reference Citation Analysis] [Abstract] [Key Words] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 11/24/2004] [Indexed: 11/18/2022] Open
Abstract
p53, the tumor suppressor protein, functions as a dimer of dimers. However, how the tetramer binds to the DNA is still an open question. In the crystal structure, three copies of the p53 monomers (containing chains A, B, and C) were crystallized with the DNA-consensus element. Although the structure provides crucial data on the p53-DNA contacts, the active oligomeric state is unclear because the two dimeric (A-B and B-C) interfaces present in the crystal cannot both exist in the tetramer. Here, we address the question of which of these two dimeric interfaces may be more biologically relevant. We analyze the sequence and structural properties of the p53-p53 dimeric interfaces and carry out extensive molecular dynamics simulations of the crystal structures of the human and mouse p53 dimers. We find that the A-B interface residues are more conserved than those of the B-C. Molecular dynamics simulations show that the A-B interface can provide a stable DNA-binding motif in the dimeric state, unlike B-C. Our results indicate that the interface between chains A-B in the p53-DNA complex constitutes a better candidate for a stable biological interface, whereas the B-C interface is more likely to be due to crystal packing. Thus, they have significant implications toward our understanding of DNA binding by p53 as well as p53-mediated interactions with other proteins.
Collapse
Affiliation(s)
- Buyong Ma
- Basic Research Program, SAIC-Frederick, Inc., Laboratory of Experimental and Computational Biology, National Cancer Institute, Frederick, MD 21702, USA.
| | | | | | | | | | | |
Collapse
|
82
|
Huang YJ, Moseley HNB, Baran MC, Arrowsmith C, Powers R, Tejero R, Szyperski T, Montelione GT. An integrated platform for automated analysis of protein NMR structures. Methods Enzymol 2005; 394:111-41. [PMID: 15808219 DOI: 10.1016/s0076-6879(05)94005-6] [Citation(s) in RCA: 61] [Impact Index Per Article: 3.1] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 12/03/2022]
Abstract
Recent developments provide automated analysis of NMR assignments and three-dimensional (3D) structures of proteins. These approaches are generally applicable to proteins ranging from about 50 to 150 amino acids. In this chapter, we summarize progress by the Northeast Structural Genomics Consortium in standardizing the NMR data collection process for protein structure determination and in building an integrated platform for automated protein NMR structure analysis. Our integrated platform includes the following principal steps: (1) standardized NMR data collection, (2) standardized data processing (including spectral referencing and Fourier transformation), (3) automated peak picking and peak list editing, (4) automated analysis of resonance assignments, (5) automated analysis of NOESY data together with 3D structure determination, and (6) methods for protein structure validation. In particular, the software AutoStructure for automated NOESY data analysis is described in this chapter, together with a discussion of practical considerations for its use in high-throughput structure production efforts. The critical area of data quality assessment has evolved significantly over the past few years and involves evaluation of both intermediate and final peak lists, resonance assignments, and structural information derived from the NMR data. Methods for quality control of each of the major automated analysis steps in our platform are also discussed. Despite significant remaining challenges, when good quality data are available, automated analysis of protein NMR assignments and structures with this platform is both fast and reliable.
Collapse
Affiliation(s)
- Yuanpeng Janet Huang
- Center for Advanced Biotechnology and Medicine, Department of Molecular Biology and Biochemistry, Rutgers University, Piscataway, New Jersey 08854, USA
| | | | | | | | | | | | | | | |
Collapse
|
83
|
Friedler A, Veprintsev DB, Rutherford T, von Glos KI, Fersht AR. Binding of Rad51 and other peptide sequences to a promiscuous, highly electrostatic binding site in p53. J Biol Chem 2004; 280:8051-9. [PMID: 15611070 DOI: 10.1074/jbc.m411176200] [Citation(s) in RCA: 56] [Impact Index Per Article: 2.7] [Reference Citation Analysis] [Abstract] [MESH Headings] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 12/20/2022] Open
Abstract
Homologous recombination is repressed by the binding of p53 to Rad51. We identified by fluorescence and NMR spectroscopy that peptides corresponding to residues 179-190 of Rad51 bind to the core domain of p53 in a site that overlaps with its specific DNA binding site. The p53 site is quite promiscuous, since it also binds peptides derived from 53BP1, 53BP2, Hif-1alpha, and BCL-X(L) in overlapping regions. Binding is mediated mainly by a strong, nonspecific, electrostatic component and is fine tuned by specific interactions. Competition of the different proteins with each other and with specific DNA for a single site in p53 could be a factor in regulation of its activity.
Collapse
Affiliation(s)
- Assaf Friedler
- Cambridge Centre for Protein Engineering, Medical Research Council Centre, Hills Road, Cambridge CB2 2QH, United Kingdom
| | | | | | | | | |
Collapse
|
84
|
Weinberg RL, Veprintsev DB, Fersht AR. Cooperative binding of tetrameric p53 to DNA. J Mol Biol 2004; 341:1145-59. [PMID: 15321712 DOI: 10.1016/j.jmb.2004.06.071] [Citation(s) in RCA: 173] [Impact Index Per Article: 8.2] [Reference Citation Analysis] [Abstract] [MESH Headings] [Journal Information] [Subscribe] [Scholar Register] [Received: 03/30/2004] [Revised: 06/24/2004] [Accepted: 06/28/2004] [Indexed: 10/26/2022]
Abstract
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.
Collapse
Affiliation(s)
- Richard L Weinberg
- Cambridge University Chemical Laboratory and Cambridge Centre for Protein Engineering, Medical Research Council Centre, Hills Road, Cambridge, CB2 2QH, United Kingdom
| | | | | |
Collapse
|
85
|
Weinberg RL, Freund SMV, Veprintsev DB, Bycroft M, Fersht AR. Regulation of DNA binding of p53 by its C-terminal domain. J Mol Biol 2004; 342:801-11. [PMID: 15342238 DOI: 10.1016/j.jmb.2004.07.042] [Citation(s) in RCA: 79] [Impact Index Per Article: 3.8] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 03/01/2004] [Revised: 06/22/2004] [Accepted: 07/09/2004] [Indexed: 11/28/2022]
Abstract
The tumor suppressor p53 is a tetrameric multi-domain transcription factor. Its C-terminal domain is thought to regulate the binding of its core domain to specific recognition sequences in promoters. The mechanism of regulation by the C-terminal domain and the role of its post-translational modification are controversial. We have examined the binding of DNA in solution to a series of unmodified p53 constructs that lack various domains. The specific DNA sequences bind tightly to the core domain, irrespective of whether or not the C-terminal domain is part of the construct. Unmodified p53 is accordingly an active DNA binding protein. Non-specific DNA sequences do not inhibit directly the binding of the specific sequences to the core but bind to the C terminus and inhibit p53 via that binding mode. Using NMR, we identified the residues of the C terminus that interact with the non-specific DNA. They include residues that are known to be modified post-translationally. Our data provide direct support for the regulatory role of the C terminus in the activity of p53 and show that p53 containing the unmodified C terminus actively binds to short double-stranded DNA.
Collapse
Affiliation(s)
- Richard L Weinberg
- Cambridge University Chemical Laboratory and Cambridge Centre for Protein Engineering, Medical Research Council Centre, Hills Road, Cambridge CB2 2QH, UK
| | | | | | | | | |
Collapse
|
86
|
Voeltzel T, Morel AP, Rostan MC, Ji J, Chiodino C, Ponchel F, Vigouroux J, Caron de Fromentel C, Soussi T, Ozturk M. A New Set of Monoclonal Antibodies Directed to Proline-Rich and Central Regions of p53. ACTA ACUST UNITED AC 2004; 23:287-92. [PMID: 15672606 DOI: 10.1089/hyb.2004.23.287] [Citation(s) in RCA: 2] [Impact Index Per Article: 0.1] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/12/2022]
Abstract
The p53 protein can adopt several conformations in cells--"latent," "active," or mutant--depending on cellular stress or mutations of the TP53 gene. Today, only a few antibodies discriminating these conformations are available. We produced three new anti-p53 monoclonal antibodies (MAbs) directed against epitopes of human p53. The H53C1 MAb recognizes an epitope located at the N-terminal part of the central region of p53 and can discriminate mutant from wild-type conformation. The H53C2 and H53C3 MAbs are against different epitopes within the proline-rich region of p53. Moreover, the H53C2 epitope is located in the second negative regulatory domain of p53 between residues 80 and 93. These MAbs can be used as new tools to study and modulate the cellular functions of p53.
Collapse
Affiliation(s)
- T Voeltzel
- INSERM U590, Centre Léon Bérard, Lyon, France.
| | | | | | | | | | | | | | | | | | | |
Collapse
|
87
|
Elgavish A, Wood PA, Pinkert CA, Eltoum IE, Cartee T, Wilbanks J, Mentor-Marcel R, Tian L, Scroggins SE. Transgenic mouse with human mutant p53 expression in the prostate epithelium. Prostate 2004; 61:26-34. [PMID: 15287091 DOI: 10.1002/pros.20071] [Citation(s) in RCA: 15] [Impact Index Per Article: 0.7] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Submit a Manuscript] [Subscribe] [Scholar Register] [Indexed: 11/07/2022]
Abstract
BACKGROUND Apoptosis is disrupted in prostate tumor cells, conferring a survival advantage. p53 is a nuclear protein believed to regulate cancer progression, in part by inducing apoptosis. To test this possibility in future studies, the objective of the present study was to generate a transgenic mouse model expressing mutant p53 in the prostate (PR). METHODS Transgene incorporation was tested using Southern analysis. Expression of mutant p53 protein was examined using immunofluorescence microscopy. Apoptosis in the PR was evaluated using the Tunnel method. RESULTS A construct, consisting of the rat probasin promoter and a mutant human p53 fragment, was prepared and used to generate transgenic mice. rPB-mutant p53 transgene incorporation, as well as nuclear accumulation of mutant human p53 protein, was demonstrated. Prostatic intraepithelial neoplasia (PIN) III and IV were found in PR of 52-week old transgenic mice, whereas no pathological changes were found in the other organs examined. PR ability to undergo apoptosis following castration was reduced in rPB-mutant p53 mice as compared to non transgenic littermates. CONCLUSIONS Transgenic rPB-mutant p53 mice accumulate mutant p53 protein in PR, resulting in neoplastic lesions and reduced apoptotic potential in the PR. Breeding rPB-mutant p53 mice with mice expressing an oncogene in their PR will be useful in examining interactions of multiple genes that result in progression of slow growing prostate tumors expressing oncogenes alone to metastatic cancer.
Collapse
Affiliation(s)
- Ada Elgavish
- Department of Genetics, Kaul Building 624, University of Alabama at Birmingham, Birmingham, AL 35294, USA.
| | | | | | | | | | | | | | | | | |
Collapse
|
88
|
la Cour T, Kiemer L, Mølgaard A, Gupta R, Skriver K, Brunak S. Analysis and prediction of leucine-rich nuclear export signals. Protein Eng Des Sel 2004; 17:527-36. [PMID: 15314210 DOI: 10.1093/protein/gzh062] [Citation(s) in RCA: 622] [Impact Index Per Article: 29.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 12/31/2022] Open
Abstract
We present a thorough analysis of nuclear export signals and a prediction server, which we have made publicly available. The machine learning prediction method is a significant improvement over the generally used consensus patterns. Nuclear export signals (NESs) are extremely important regulators of the subcellular location of proteins. This regulation has an impact on transcription and other nuclear processes, which are fundamental to the viability of the cell. NESs are studied in relation to cancer, the cell cycle, cell differentiation and other important aspects of molecular biology. Our conclusion from this analysis is that the most important properties of NESs are accessibility and flexibility allowing relevant proteins to interact with the signal. Furthermore, we show that not only the known hydrophobic residues are important in defining a nuclear export signals. We employ both neural networks and hidden Markov models in the prediction algorithm and verify the method on the most recently discovered NESs. The NES predictor (NetNES) is made available for general use at http://www.cbs.dtu.dk/.
Collapse
Affiliation(s)
- Tanja la Cour
- Center for Biological Sequence Analysis, Biocentrum-DTU, Technical University of Denmark, Building 208, DK-2800 Lyngby, Denmark
| | | | | | | | | | | |
Collapse
|
89
|
Wood MJ, Storz G, Tjandra N. Structural basis for redox regulation of Yap1 transcription factor localization. Nature 2004; 430:917-21. [PMID: 15318225 DOI: 10.1038/nature02790] [Citation(s) in RCA: 128] [Impact Index Per Article: 6.1] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 03/24/2004] [Accepted: 06/28/2004] [Indexed: 01/28/2023]
Abstract
The ability of organisms to alter their gene expression patterns in response to environmental changes is essential for viability. A central regulator of the response to oxidative stress in Saccharomyces cerevisiae is the Yap1 transcription factor. Upon activation by increased levels of reactive oxygen species, Yap1 rapidly redistributes to the nucleus where it regulates the expression of up to 70 genes. Here we identify a redox-regulated domain of Yap1 and determine its high-resolution solution structure. In the active oxidized form, a nuclear export signal (NES) in the carboxy-terminal cysteine-rich domain is masked by disulphide-bond-mediated interactions with a conserved amino-terminal alpha-helix. Point mutations that weaken the hydrophobic interactions between the N-terminal alpha-helix and the C-terminal NES-containing domain abolished redox-regulated changes in subcellular localization of Yap1. Upon reduction of the disulphide bonds, Yap1 undergoes a change to an unstructured conformation that exposes the NES and allows redistribution to the cytoplasm. These results reveal the structural basis of redox-dependent Yap1 localization and provide a previously unknown mechanism of transcription factor regulation by reversible intramolecular disulphide bond formation.
Collapse
Affiliation(s)
- Matthew J Wood
- Cell Biology and Metabolism Branch, National Institute of Child Health and Human Development, National Institutes of Health, Bethesda, Maryland 20892-5430, USA
| | | | | |
Collapse
|
90
|
Yoshida Y, Izumi H, Torigoe T, Ishiguchi H, Yoshida T, Itoh H, Kohno K. Binding of RNA to p53 regulates its oligomerization and DNA-binding activity. Oncogene 2004; 23:4371-9. [PMID: 15064727 DOI: 10.1038/sj.onc.1207583] [Citation(s) in RCA: 30] [Impact Index Per Article: 1.4] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 12/11/2022]
Abstract
The C-terminus of p53 is responsible for maintaining the latent, non-DNA-binding form of p53. However, the mechanism by which the C-terminus regulates DNA binding is not yet fully understood. We show here that p53 interacts with RNA via its C-terminal domain and that disruption of this interaction, by RNase A treatment, truncation or phosphorylation of the C-terminus, restores DNA-binding activity. Furthermore, the oligomerization of p53 is significantly enhanced by disrupting the interaction between p53 and RNA. These findings suggest that binding of RNA to p53 is involved in the mechanism of p53 latency.
Collapse
Affiliation(s)
- Yoichiro Yoshida
- Department of Molecular Biology, School of Medicine, University of Occupational and Environmental Health, 1-1 Iseigaoka, Yahatanishi-ku, Kitakyushu, Fukuoka 807-8555, Japan
| | | | | | | | | | | | | |
Collapse
|
91
|
Kato S, Han SY, Liu W, Otsuka K, Shibata H, Kanamaru R, Ishioka C. Understanding the function-structure and function-mutation relationships of p53 tumor suppressor protein by high-resolution missense mutation analysis. Proc Natl Acad Sci U S A 2003; 100:8424-9. [PMID: 12826609 PMCID: PMC166245 DOI: 10.1073/pnas.1431692100] [Citation(s) in RCA: 649] [Impact Index Per Article: 29.5] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/01/2023] Open
Abstract
Inactivation of the tumor suppressor p53 by missense mutations is the most frequent genetic alteration in human cancers. The common missense mutations in the TP53 gene disrupt the ability of p53 to bind to DNA and consequently to transactivate downstream genes. However, it is still not fully understood how a large number of the remaining mutations affect p53 structure and function. Here, we used a comprehensive site-directed mutagenesis technique and a yeast-based functional assay to construct, express, and evaluate 2,314 p53 mutants representing all possible amino acid substitutions caused by a point mutation throughout the protein (5.9 substitutions per residue), and correlated p53 function with structure- and tumor-derived mutations. This high-resolution mutation analysis allows evaluation of previous predictions and hypotheses through interrelation of function, structure and mutation.
Collapse
|
92
|
Zhu X, Yu QS, Cutler RG, Culmsee CW, Holloway HW, Lahiri DK, Mattson MP, Greig NH. Novel p53 inactivators with neuroprotective action: syntheses and pharmacological evaluation of 2-imino-2,3,4,5,6,7-hexahydrobenzothiazole and 2-imino-2,3,4,5,6,7-hexahydrobenzoxazole derivatives. J Med Chem 2002; 45:5090-7. [PMID: 12408720 DOI: 10.1021/jm020044d] [Citation(s) in RCA: 87] [Impact Index Per Article: 3.8] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/29/2022]
Abstract
Tumor suppressor protein, p53, is an intracellular protein that is critical within the biochemical cascade that leads to cell death via apoptosis. Recent studies identified the tetrahydrobenzothiazole analogue, pifithrin-alpha (2), as a p53 inhibitor that was effective in protecting neuronal cells against a variety of lethal insults and reducing the side effects of anticancer drugs. As up-regulation of p53 has been described as a common feature of several neurodegenerative disorders, including Alzheimer's disease, 2 and novel analogues (3-16) were synthesized to (i) assess the value of tetrahydrobenzothiazole analogues as neuroprotective agents and (ii) define the structural requirements for p53 inactivation. Not only did 2 exhibit neuroprotective activity in both tissue culture and in vivo stroke models but also compounds 6, 7, 10, 13, 15, and 16 proved to be highly potent in protecting PC12 cells and compounds 3, 4, and 6 were highly potent in protecting primary hippocampal cells against death induced by the DNA-damaging agent, camptothecin.
Collapse
Affiliation(s)
- Xiaoxiang Zhu
- Laboratory of Neurosciences, Gerontology Research Center, National Institute on Aging, Intramural Research Program, National Institutes of Health, 5600 Nathan Shock Drive, Baltimore, Maryland 21224, USA
| | | | | | | | | | | | | | | |
Collapse
|
93
|
Miyaki M, Iijima T, Yasuno M, Kita Y, Hishima T, Kuroki T, Mori T. High incidence of protein-truncating mutations of the p53 gene in liver metastases of colorectal carcinomas. Oncogene 2002; 21:6689-93. [PMID: 12242668 DOI: 10.1038/sj.onc.1205887] [Citation(s) in RCA: 15] [Impact Index Per Article: 0.7] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 03/24/2002] [Revised: 07/05/2002] [Accepted: 07/18/2002] [Indexed: 11/10/2022]
Abstract
To clarify the significance of p53 mutations in liver metastasis of colorectal carcinogenesis, the characteristics of p53 mutations from 51 liver metastases and 76 primary invasive carcinomas without liver metastasis (Dukes' A, B and C) were compared. The frequency of tumors with p53 mutations was 61% (31 out of 51) in the liver metastases, and 51% (39 out of 76) in the primary carcinomas without liver metastasis. Approximately 90% of the informative cases having p53 mutation showed 17pLOH. Mutations detected within exons 4-10 of the p53 gene included missense, nonsense, frameshift, inframe deletion, and inframe insertion mutations. Out of the tumors with p53 mutations, we found that the percentage of tumors with protein-truncating mutations (nonsense and frameshift mutations) was extremely higher in liver metastases (16 out of 31, 52%) than in primary carcinomas without liver metastasis (5 out of 39, 13%) (P=0.0005). The present results suggest that protein-truncating mutations of the p53 gene are more relevant than missense mutations as one of the prognostic factors in liver metastasis of colorectal carcinomas.
Collapse
Affiliation(s)
- Michiko Miyaki
- Hereditary Tumor Research Project, Tokyo Metropolitan Komagome Hospital, Tokyo 113-8677, Japan.
| | | | | | | | | | | | | |
Collapse
|
94
|
Smith PC, Firestein S, Hunt JF. The crystal structure of the olfactory marker protein at 2.3 A resolution. J Mol Biol 2002; 319:807-21. [PMID: 12054872 DOI: 10.1016/s0022-2836(02)00242-5] [Citation(s) in RCA: 28] [Impact Index Per Article: 1.2] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/27/2022]
Abstract
Olfactory marker protein (OMP) is a highly expressed and phylogenetically conserved cytoplasmic protein of unknown function found almost exclusively in mature olfactory sensory neurons. Electrophysiological studies of olfactory epithelia in OMP knock-out mice show strongly retarded recovery following odorant stimulation leading to an impaired response to pulsed odor stimulation. Although these studies show that OMP is a modulator of the olfactory signal-transduction cascade, its biochemical role is not established. In order to facilitate further studies on the molecular function of OMP, its crystal structure has been determined at 2.3 A resolution using multiwavelength anomalous diffraction experiments on selenium-labeled protein. OMP is observed to form a modified beta-clamshell structure with eight antiparallel beta-strands. While OMP has no significant sequence homology to proteins of known structure, it has a similar fold to a domain found in a variety of existing structures, including in a large family of viral capsid proteins. The surface of OMP is mostly convex and lacking obvious small molecule binding sites, suggesting that it is more likely to be involved in modulating protein-protein interaction than in interacting with small molecule ligands. Three highly conserved regions have been identified as leading candidates for protein-protein interaction sites in OMP. One of these sites represents a loop known to mediate ligand interactions in the structurally homologous EphB2 receptor ligand-binding domain. This site is partially buried in the crystal structure but fully exposed in the NMR solution structure of OMP due to a change in the orientation of an alpha-helix that projects outward from the structurally invariant beta-clamshell core. Gating of this conformational change by molecular interactions in the signal-transduction cascade could be used to control access to OMP's equivalent of the EphB2 ligand-interaction loop, thereby allowing OMP to function as a molecular switch.
Collapse
Affiliation(s)
- Paul C Smith
- Department of Biological Sciences, 702A Fairchild Center, MC 2434, Columbia University, New York, NY 10027, USA
| | | | | |
Collapse
|
95
|
Rippin TM, Freund SMV, Veprintsev DB, Fersht AR. Recognition of DNA by p53 core domain and location of intermolecular contacts of cooperative binding. J Mol Biol 2002; 319:351-8. [PMID: 12051912 DOI: 10.1016/s0022-2836(02)00326-1] [Citation(s) in RCA: 69] [Impact Index Per Article: 3.0] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 10/27/2022]
Abstract
We present an analysis by NMR of a 58 kDa complex of the core domain of the tumour suppressor p53 with DNA that complements and extends the crystal structure analysis. Binding of specific DNA caused significant chemical shifts of residues on the DNA-binding interface that translated into the beta-sheet of the protein. Binding of non-specific DNA caused weak but qualitatively the same shifts, corresponding to weaker binding interactions. The observed chemical shift differences correlate with frequency of cancer-inducing mutations, suggesting that the affected residues contribute to the stability of p53 core domain-DNA complex. We also identified two affected regions on the surface of the protein: helix 1 (residues V173-C182) plus G244 and residues L114-T118, which may represent a dimerisation interface.
Collapse
Affiliation(s)
- Thomas M Rippin
- Cambridge University Chemical Laboratory and Cambridge Centre for Protein Engineering, MRC Centre, Hills Road, Cambridge CB2 2QH, UK
| | | | | | | |
Collapse
|
96
|
Nicholls CD, McLure KG, Shields MA, Lee PWK. Biogenesis of p53 involves cotranslational dimerization of monomers and posttranslational dimerization of dimers. Implications on the dominant negative effect. J Biol Chem 2002; 277:12937-45. [PMID: 11805092 DOI: 10.1074/jbc.m108815200] [Citation(s) in RCA: 103] [Impact Index Per Article: 4.5] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/06/2022] Open
Abstract
Precisely how mutant p53 exerts a dominant negative effect over wild type p53 has been an enigma. To understand how wild type and mutant p53 form hetero-oligomers, we studied p53 biogenesis in vitro. We show here that p53 dimers are formed cotranslationally (on the polysome), whereas tetramers are formed posttranslationally (by the dimerization of dimers in solution). Coexpression of wild type and mutant p53 therefore results in 50% of the p53 generated being heterotetramers comprised of a single species: wild type dimer/mutant dimer. Using hot spot mutants of p53 and a variety of natural target sites, we show that all wild type/mutant heterotetramers manifest impaired DNA binding activity. This impairment is not due to the mutant dimeric subunit inhibiting association of the complex with DNA but rather due to the lack of significant contribution (positive cooperativity) from the mutant partner. For all heterotetramers, bias in binding is particularly pronounced against those sequences in genes responsible for apoptosis rather than cell growth arrest. These results explain the molecular basis of p53 dominant negative effect and suggest a functional role in the regulation of p53 tetramerization.
Collapse
Affiliation(s)
- Chris D Nicholls
- Department of Microbiology and Infectious Diseases, Cancer Biology Research Group, University of Calgary Health Sciences Centre, Calgary, Alberta, Canada, T2N 4N1
| | | | | | | |
Collapse
|
97
|
Cromsigt J, van Buuren B, Schleucher J, Wijmenga S. Resonance assignment and structure determination for RNA. Methods Enzymol 2002; 338:371-99. [PMID: 11460559 DOI: 10.1016/s0076-6879(02)38229-6] [Citation(s) in RCA: 52] [Impact Index Per Article: 2.3] [Reference Citation Analysis] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/20/2023]
Affiliation(s)
- J Cromsigt
- Department of Medical Biosciences-Medical Biophysics, Umea University, S-901 87 Umea, Sweden
| | | | | | | |
Collapse
|
98
|
Klein C, Planker E, Diercks T, Kessler H, Künkele KP, Lang K, Hansen S, Schwaiger M. NMR spectroscopy reveals the solution dimerization interface of p53 core domains bound to their consensus DNA. J Biol Chem 2001; 276:49020-7. [PMID: 11606582 DOI: 10.1074/jbc.m107516200] [Citation(s) in RCA: 68] [Impact Index Per Article: 2.8] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/06/2022] Open
Abstract
The p53 protein is a transcription factor that acts as the major tumor suppressor in mammals. The core DNA-binding domain is mutated in about 50% of all human tumors. The crystal structure of the core domain in complex with DNA illustrated how a single core domain specifically interacts with its DNA consensus site and how it is inactivated by mutation. However, no structural information for the tetrameric full-length p53-DNA complex is available. Here, we present novel experimental insight into the dimerization of two p53 core domains upon cooperative binding to consensus DNA in solution obtained by NMR. The NMR data show that the p53 core domain itself does not appear to undergo major conformational changes upon addition of DNA and elucidate the dimerization interface between two DNA-bound core domains, which includes the short H1 helix. A NMR-based model for the dimeric p53 core-DNA complex incorporates these data and allows the conclusion that the dimerization interface also forms the actual interface in the tetrameric p53-DNA complex. The significance of this interface is further corroborated by the finding that hot spot mutations map to the H1 helix, and by the binding of the putative p53 inhibitor 53BP2 to this region via one of its ankyrin repeats. Based on symmetry considerations it is proposed that tetrameric p53 might link non-contiguous DNA consensus sites in a sandwich-like manner generating DNA loops as observed for transcriptionally active p53 complexes.
Collapse
Affiliation(s)
- C Klein
- Pharma Research, Roche Diagnostics GmbH, D-82372 Penzberg, Germany
| | | | | | | | | | | | | | | |
Collapse
|
99
|
Kanovsky M, Raffo A, Drew L, Rosal R, Do T, Friedman FK, Rubinstein P, Visser J, Robinson R, Brandt-Rauf PW, Michl J, Fine RL, Pincus MR. Peptides from the amino terminal mdm-2-binding domain of p53, designed from conformational analysis, are selectively cytotoxic to transformed cells. Proc Natl Acad Sci U S A 2001; 98:12438-43. [PMID: 11606716 PMCID: PMC60072 DOI: 10.1073/pnas.211280698] [Citation(s) in RCA: 66] [Impact Index Per Article: 2.8] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 06/05/2001] [Indexed: 02/07/2023] Open
Abstract
We have synthesized three peptides from the mdm-2 binding domain of human p53, residues 12-26 (PPLSQETFSDLWKLL), residues 12-20, and 17-26. To enable transport of the peptides across the cell membrane and at the same time to maximize the active mdm-2 binding alpha-helical conformation for these peptides, each was attached at its carboxyl terminus to the penetratin sequence, KKWKMRRNQFWVKVQRG, that contains many positively charged residues that stabilize an alpha-helix when present on its carboxyl terminal end. All three peptides were cytotoxic to human cancer cells in culture, whereas a control, unrelated peptide attached to the same penetratin sequence had no effect on these cell lines. The same three cytotoxic peptides had no effect on the growth of normal cells, including human cord blood-derived stem cells. These peptides were as effective in causing cell death in p53-null cancer cells as in those having mutant or normal p53. Peptide-induced cell death is not accompanied by expression of apoptosis-associated proteins such as Bax and waf(p21). Based on these findings, we conclude that the antiproliferative effects of these p53-derived peptides are not completely dependent on p53 activity and may prove useful as general anticancer agents.
Collapse
Affiliation(s)
- M Kanovsky
- Department of Pathology and Laboratory Medicine, Harbor Veterans Affairs Medical Center, 800 Poly Place, Brooklyn, NY 11209, USA
| | | | | | | | | | | | | | | | | | | | | | | | | |
Collapse
|
100
|
Abstract
Although cell death once was viewed exclusively as the disordered, chaotic outcome of metabolic catastrophe, apoptosis now is recognized as a highly ordered, evolutionarily conserved, and genetically selected program that is essential for normal development. The death receptor pathway of apoptosis, cytotoxic T cells, prolife survival signals, Bcl-2 family of regulators, p53 and regulated cell death in cancer, and oncogenes are reviewed. Future prospects in this arena also are discussed.
Collapse
Affiliation(s)
- D E Fisher
- Division of Pediatric Hematology and Oncology, Children's Hospital, Dana-Farber Cancer Institute, Boston, Massachusetts, USA
| |
Collapse
|