151
|
Uberti D, Belloni M, Grilli M, Spano P, Memo M. Induction of tumour-suppressor phosphoprotein p53 in the apoptosis of cultured rat cerebellar neurones triggered by excitatory amino acids. Eur J Neurosci 1998; 10:246-54. [PMID: 9753133 DOI: 10.1046/j.1460-9568.1998.00042.x] [Citation(s) in RCA: 80] [Impact Index Per Article: 3.0] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/07/2023]
Abstract
We found that primary cultures of rat cerebellar granule cells, although definitely postmitotic and terminally differentiated, express the tumour-suppressor phosphoprotein p53. In particular, granule cells both expressed significant levels of p53 mRNA and positively reacted to an anti-p53 antibody, from the first day of culturing. During neurone differentiation, p53 mRNA content did not significantly change, at least up to 12 days in vitro, while p53 immunoreactivity increased gradually. p53 expression appeared to be further modulable being upregulated after stimulation of glutamate ionotropic receptors by glutamate or kainate. Although qualitatively similar, p53 induction by glutamate and kainate differed in terms of intensity and time-course. The glutamate increase of p53 immunoreactivity appeared within 30 min after the treatment and lasted for at least 2 h. Kainate-induced increase of p53 immunoreactivity was delayed, becoming apparent within 2 h and lasting for at least 8 h. Both kainate- and glutamate-induced increases of p53 immunoreactivity were prevented by the non-competitive NMDA receptor antagonist MK 801. As shown by the electrophoretic mobility shift analysis, both glutamate and kainate induced increases of p53 DNA binding activity. Blockade of p53 induction by a specific p53 antisense oligonucleotide resulted in a partial reduction of excitotoxicity with a complete inhibition of the excitatory amino acids induced apoptosis. Our data suggest that stimulation of ionotropic glutamate receptors in neurones results in a p53-dependent apoptosis.
Collapse
Affiliation(s)
- D Uberti
- Department of Biomedical Sciences and Biotechnologies, School of Medicine, University of Brescia, Italy
| | | | | | | | | |
Collapse
|
152
|
Campomenosi P, Fronza G, Ottaggio L, Roncella S, Inga A, Bogliolo M, Monti P, Assereto P, Moro F, Cutrona G, Bozzo S, Chiorazzi N, Abbondandolo A, Ferrarini M. Heterogeneous p53 mutations in a Burkitt lymphoma from an AIDS patient with monoclonal c-myc and VDJ rearrangements. Int J Cancer 1997; 73:816-21. [PMID: 9399658 DOI: 10.1002/(sici)1097-0215(19971210)73:6<816::aid-ijc8>3.0.co;2-5] [Citation(s) in RCA: 5] [Impact Index Per Article: 0.2] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/05/2023]
Abstract
This study investigates the timing of p53 mutations detected in the malignant cells of a Burkitt's lymphoma cell line (BRG-P) with respect to other maturation or transforming events. The BRG-P cell line, derived from an AIDS patient, was of special value since it displayed subclones that had undergone an isotype switch from IgM to IgA1 (BRG-M and BRG-A cells). BRG-M and BRG-A cells were characterized by the same monoclonal c-myc and VDJ rearrangements and by the expression of Ig receptors with specificity for a 45 kDa protein of human breast cells. Analysis of p53 mutations in the different BRG subclones showed that 1) BRG-M cells displayed 2 different p53 mutations in trans; since the original BL cells also showed the same mutations, this finding indicated that both occurred in vivo; 2) one of the p53 alleles of BRG-A cells was lost, while the other showed a mutation different from those seen in BRG-M cells; and 3) all 3 mutations observed in BRG-M or BRG-A cells resulted in the functional inactivation of the transcriptional activation function of p53. Together, our data demonstrate that p53 mutations were relatively late events during lymphomagenesis. Moreover, in view of the role of p53 in cell apoptosis, it is conceivable that BRG cells were subjected to a strong selective pressure that favored p53 inactivation. Such inactivation was possibly required to counterbalance other potentially apoptotic events, including the presence of a deregulated c-myc oncogene and signals delivered by the host environment in situ.
Collapse
Affiliation(s)
- P Campomenosi
- CSTA-Mutagenesis Laboratory, National Institute for Cancer Research (IST), Genoa, Italy
| | | | | | | | | | | | | | | | | | | | | | | | | | | |
Collapse
|
153
|
Chène P, Mittl P, Grütter M. In vitro structure-function analysis of the beta-strand 326-333 of human p53. J Mol Biol 1997; 273:873-81. [PMID: 9367778 DOI: 10.1006/jmbi.1997.1360] [Citation(s) in RCA: 28] [Impact Index Per Article: 1.0] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/05/2023]
Abstract
The beta-strand 326-333 is a key structural element in the formation of p53 tetramers. To investigate the contribution of its amino acid residues, an alanine scan was performed. The oligomerisation and DNA-binding properties of the mutant proteins were compared with those of wild-type proteins in vitro and analysed on the basis of the crystal structure of the p53 tetramerisation domain at 1.5 A resolution. Two categories of mutant proteins were identified. Phe328Ala, Leu330Ala and Ile332Ala mutant proteins are inactive for DNA binding and oligomerisation, while the Glu326Ala, Tyr327Ala, Thr329Ala, Gln331Ala and Arg333Ala mutant proteins have properties similar to those of wild-type proteins. These results suggest that single mutations within the p53 tetramerisation domain destabilise the structure of the whole protein, inhibiting its DNA-binding activity. Furthermore, the mutation of leucine 330 to alanine within the tetramerisation domain of the Arg175His protein abolishes the dominant negative effect of this mutant. This shows that the beta-strand 326-333 is a key structural element that mediates the dominant negative effect of p53 mutants.
Collapse
Affiliation(s)
- P Chène
- Oncology Department and Core Drug Doscovery Technologies, Novartis CH-4002 Basel, Switzerland
| | | | | |
Collapse
|
154
|
McCoy M, Stavridi ES, Waterman JL, Wieczorek AM, Opella SJ, Halazonetis TD. Hydrophobic side-chain size is a determinant of the three-dimensional structure of the p53 oligomerization domain. EMBO J 1997; 16:6230-6. [PMID: 9321402 PMCID: PMC1326307 DOI: 10.1093/emboj/16.20.6230] [Citation(s) in RCA: 36] [Impact Index Per Article: 1.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/05/2023] Open
Abstract
The p53 tumor suppressor oligomerization domain, a dimer of two primary dimers, is an independently folding domain whose subunits consist of a beta-strand, a tight turn and an alpha-helix. To evaluate the effect of hydrophobic side-chains on three-dimensional structure, we substituted residues Phe341 and Leu344 in the alpha-helix with other hydrophobic amino acids. Substitutions that resulted in residue 341 having a smaller side-chain than residue 344 switched the stoichiometry of the domain from tetrameric to dimeric. The three-dimensional structure of one such dimer was determined by multidimensional NMR spectroscopy. When compared with the primary dimer of the wild-type p53 oligomerization domain, the mutant dimer showed a switch in alpha-helical packing from anti-parallel to parallel and rotation of the alpha-helices relative to the beta-strands. Hydrophobic side-chain size is therefore an important determinant of a protein fold.
Collapse
Affiliation(s)
- M McCoy
- Departments of Molecular Genetics and Structural Biology, The Wistar Institute, Philadelphia, PA 19104-4268, USA
| | | | | | | | | | | |
Collapse
|
155
|
Okorokov AL, Ponchel F, Milner J. Induced N- and C-terminal cleavage of p53: a core fragment of p53, generated by interaction with damaged DNA, promotes cleavage of the N-terminus of full-length p53, whereas ssDNA induces C-terminal cleavage of p53. EMBO J 1997; 16:6008-17. [PMID: 9312058 PMCID: PMC1170231 DOI: 10.1093/emboj/16.19.6008] [Citation(s) in RCA: 50] [Impact Index Per Article: 1.8] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/05/2023] Open
Abstract
p53 is able to recognize and bind sites of DNA damage and, in some way, damage to cellular DNA activates a p53 response leading to G1 arrest or apoptosis. We have previously shown that 'damaged DNA' induces N-terminal cleavage of p53 to generate p40(DeltaN) and p35 (core) protein products. We now show that the p35 product has protease activity and is able to cleave between residues 23 and 24 of full-length p53 to generate a novel product, p50(DeltaN23). This activity was inhibited by bestatin, an aminopeptidase inhibitor. Residues 23 and 24 lie within the mdm-2 binding domain of p53 and the possibility that p50(DeltaN23) may be resistant to feedback regulation by mdm-2 is discussed. Unexpectedly, interaction with ssDNA induced two further cleavage products of p53, generated by C-terminal cleavage and designated p50(DeltaC) and p40(DeltaC). In vivo generation of a C-terminal cleavage product of endogenous p53 similar in size to p50(DeltaC) correlated with up-regulation of p21 expression in ML-1 cells exposed to either adriamycin or cisplatin. The possible significance of the various p53 cleavage products in relation to the cellular response to DNA damage is discussed.
Collapse
Affiliation(s)
- A L Okorokov
- Yorkshire Cancer Research Campaign P53 Research Group, Department of Biology, University of York, York YO1 5DD, UK
| | | | | |
Collapse
|
156
|
Abstract
Many tumour therapies act by inducing a cellular damage response pathway mediated by the tumour suppressor protein p53. Alternative outcomes of p53 induction include apoptosis or transient cell-cycle arrest, both thought to require the transcriptional activity of wild-type p53. Current research highlights the action of a p53-activated gene, p21Cip1/WAF1/Sdi1, which encodes a cyclin-kinase inhibitor important in mediating p53-dependent cell-cycle arrest, while programmed cell death in response to DNA damage requires transcriptionally active p53 but not activation of p21Cip1/WAF1/Sdi1. This review examines the roles of p53 and p21Cip1/WAF1/Sdi1 in controlling cell proliferation, in the light of a new study on expression of p53 and p21Cip1/WAF1/Sdi1 in squamous cell carcinoma of the larynx.
Collapse
Affiliation(s)
- L S Cox
- Department of Biochemistry, University of Oxford, U.K.
| |
Collapse
|
157
|
Wagner P. Phosphorylation mutants of p53 show differential complex formation with putative dehydrogenase Tms1 of fission yeast. EUROPEAN JOURNAL OF BIOCHEMISTRY 1997; 248:441-4. [PMID: 9346300 DOI: 10.1111/j.1432-1033.1997.t01-1-00441.x] [Citation(s) in RCA: 3] [Impact Index Per Article: 0.1] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 02/05/2023]
Abstract
The yeast tms1 gene was originally identified as a multi-copy suppressor of a lethal growth arrest caused by expression of a tumour mutant cDNA of p53 in fission yeast. The tms1 gene product (Tms1) was found to form stable complexes with p53 in yeast and in vitro; using purified recombinant proteins, the interaction was mapped to the C-terminal region of p53. This part is known to be modified by several protein kinases resulting in a transition of p53 from a latent to an activated state capable of transactivating various cellular genes involved in growth suppression or apoptosis. Since there is evidence for an evolutionary conservation of a Tms1-related protein in mammals, the effect of the phosphorylation status of the C-terminus of p53 on Tms1/p53 complex formation in vitro has been investigated. Whereas mutants changing the cdc2 phosphorylation site at position 315 of human p53 had only little effect on Tms1/p53 complex formation, we found that mutants involving the protein kinase CK2 site at position 392 showed a significantly decreased relative affinity for the Tms1 protein. The same result was obtained by using a C-terminal fragment of p53 which was phosphorylated by purified protein kinase CK2, suggesting that the complex formation of p53 with cellular C-terminal binding proteins like Tms1 impairs regulation by phosphorylation.
Collapse
Affiliation(s)
- P Wagner
- Medical Biochemistry, University of Saarland, Homburg/Saar, Germany.
| |
Collapse
|
158
|
Fourie AM, Hupp TR, Lane DP, Sang BC, Barbosa MS, Sambrook JF, Gething MJ. HSP70 binding sites in the tumor suppressor protein p53. J Biol Chem 1997; 272:19471-9. [PMID: 9235949 DOI: 10.1074/jbc.272.31.19471] [Citation(s) in RCA: 72] [Impact Index Per Article: 2.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/04/2023] Open
Abstract
Mutations within conserved regions of the tumor suppressor protein, p53, result in oncogenic forms of the protein with altered tertiary structures. In most cases, the mutant p53 proteins are selectively recognized and bound by members of the HSP70 family of molecular chaperones, but the binding site(s) in p53 for these chaperones have not been clearly defined. We have screened a library of overlapping biotinylated peptides, spanning the entire human p53 sequence, for binding to the HSP70 proteins, Hsc70 and DnaK. We show that most of the high affinity binding sites for these proteins map to secondary structure elements, particularly beta-strands, in the hydrophobic core of the central DNA binding domain, where the majority of oncogenic p53 mutations are found. Although peptides corresponding to the C-terminal region of p53 also contain potential binding sites, p53 proteins with C-terminal deletions are capable of binding to Hsc70, indicating that this region is not required for complex formation. We propose that mutations in the p53 protein alter the tertiary structure of the central DNA binding domain, thus exposing high affinity HSP70 binding sites that are cryptic in the wild-type molecule.
Collapse
Affiliation(s)
- A M Fourie
- R. W. Johnson Pharmaceutical Research Institute, San Diego, California 92121, USA
| | | | | | | | | | | | | |
Collapse
|
159
|
Walters KJ, Matsuo H, Wagner G. A Simple Method to Distinguish Intermonomer Nuclear Overhauser Effects in Homodimeric Proteins with C2 Symmetry. J Am Chem Soc 1997. [DOI: 10.1021/ja963309k] [Citation(s) in RCA: 44] [Impact Index Per Article: 1.6] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/28/2022]
Affiliation(s)
- Kylie J. Walters
- Program for Higher Degrees in Biophysics Department of Biological Chemistry and Molecular Pharmacology Harvard Medical School, Boston, Massachusetts 02115
| | - Hiroshi Matsuo
- Program for Higher Degrees in Biophysics Department of Biological Chemistry and Molecular Pharmacology Harvard Medical School, Boston, Massachusetts 02115
| | - Gerhard Wagner
- Program for Higher Degrees in Biophysics Department of Biological Chemistry and Molecular Pharmacology Harvard Medical School, Boston, Massachusetts 02115
| |
Collapse
|
160
|
Fontoura BM, Atienza CA, Sorokina EA, Morimoto T, Carroll RB. Cytoplasmic p53 polypeptide is associated with ribosomes. Mol Cell Biol 1997; 17:3146-54. [PMID: 9154813 PMCID: PMC232167 DOI: 10.1128/mcb.17.6.3146] [Citation(s) in RCA: 53] [Impact Index Per Article: 1.9] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/04/2023] Open
Abstract
Our previous finding that the tumor suppressor p53 is covalently linked to 5.8S rRNA suggested functional association of p53 polypeptide with ribosomes. p53 polypeptide is expressed at low basal levels in the cytoplasm of normal growing cells in the G1 phase of the cell cycle. We report here that cytoplasmic wild-type p53 polypeptide from both rat embryo fibroblasts and MCF7 cells and the A135V transforming mutant p53 polypeptide were found associated with ribosomes to various extents. Treatment of cytoplasmic extracts with RNase or puromycin in the presence of high salt, both of which are known to disrupt ribosomal function, dissociated p53 polypeptide from the ribosomes. In immunoprecipitates of p53 polypeptide-associated ribosomes, 5.8S rRNA was detectable only after proteinase K treatment, indicating all of the 5.8S rRNA in p53-associated ribosomes is covalently linked to protein. While 5.8S rRNA linked to protein was found in the immunoprecipitates of either wild-type or A135V mutant p53 polypeptide associated with ribosomes, little 5.8S rRNA was found in the immunoprecipitates of the slowly sedimenting p53 polypeptide, which was not associated with ribosomes. In contrast, 5.8S rRNA was liberated from bulk ribosomes by 1% sodium dodecyl sulfate, without digestion with proteinase K, indicating that these ribosomes contain 5.8S rRNA, which is not linked to protein. Immunoprecipitation of p53 polypeptide coprecipitated a small fraction of ribosomes. p53 mRNA immunoprecipitated with cytoplasmic p53 polypeptide, while GAPDH mRNA did not. These results show that cytoplasmic p53 polypeptide is associated with a subset of ribosomes, having covalently modified 5.8S rRNA.
Collapse
Affiliation(s)
- B M Fontoura
- Department of Pathology, New York University School of Medicine, New York 10016, USA
| | | | | | | | | |
Collapse
|
161
|
Delphin C, Huang KP, Scotto C, Chapel A, Vincon M, Chambaz E, Garin J, Baudier J. The in vitro phosphorylation of p53 by calcium-dependent protein kinase C--characterization of a protein-kinase-C-binding site on p53. EUROPEAN JOURNAL OF BIOCHEMISTRY 1997; 245:684-92. [PMID: 9183006 DOI: 10.1111/j.1432-1033.1997.t01-1-00684.x] [Citation(s) in RCA: 31] [Impact Index Per Article: 1.1] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 02/04/2023]
Abstract
We show that, in vitro, Ca2+-dependent protein kinase C (PKC) phosphorylates recombinant murine p53 protein on several residues contained within a conserved basic region of 25 amino acids, located in the C-terminal part of the protein. Accordingly, synthetic p53-(357-381)-peptide is phosphorylated by PKC at multiple Ser and Thr residues, including Ser360, Thr365, Ser370 and Thr377. We also establish that p53-(357-381)-peptide at micromolar concentrations has the ability to stimulate sequence-specific DNA binding by p53. That stimulation is lost upon phosphorylation by PKC. To further characterise the mechanisms that regulate PKC-dependent phosphorylation of p53-(357-381)-peptide, the phosphorylation of recombinant p53 and p53-(357-381)-peptide by PKC were compared. The results suggest that phosphorylation of full-length p53 on the C-terminal PKC sites is highly dependent on the accessibility of the phosphorylation sites and that a domain on p53 distinct from p53-(357-381)-peptide is involved in binding PKC. Accordingly, we have identified a conserved 27-amino-acid peptide, p53-(320-346)-peptide, within the C-terminal region of p53 and adjacent to residues 357-381 that interacts with PKC in vitro. The interaction between p53-(320-346)-peptide and PKC inhibits PKC autophosphorylation and the phosphorylation of substrates, including p53-(357-381)-peptide, neurogranin and histone H1. Conventional Ca2+-dependent PKC alpha, beta and gamma and the catalytic fragment of PKC (PKM) were nearly equally susceptible to inhibition by p53-(320-346)-peptide. The Ca2+-independent PKC delta was much less sensitive to inhibition. The significance of these findings for understanding the in vivo phosphorylation of p53 by PKC are discussed.
Collapse
Affiliation(s)
- C Delphin
- Département de Biologie Moléculaire et Structurale du CEA, BMCC INSERM Unité 309, Grenoble, France
| | | | | | | | | | | | | | | |
Collapse
|
162
|
Ishioka C, Shimodaira H, Englert C, Shimada A, Osada M, Jia LQ, Suzuki T, Gamo M, Kanamaru R. Oligomerization is not essential for growth suppression by p53 in p53-deficient osteosarcoma Saos-2 cells. Biochem Biophys Res Commun 1997; 232:54-60. [PMID: 9125151 DOI: 10.1006/bbrc.1997.6216] [Citation(s) in RCA: 17] [Impact Index Per Article: 0.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/04/2023]
Abstract
The carboxy-terminal portion of the p53 protein contains the tetramerization domain, and the introduction of multiple missense mutations in this domain disrupts the formation of p53 tetramers, resulting in the production of dimeric or monomeric forms of p53. It has recently been shown that a single missense or nonsense mutation in this domain affects the functional properties of p53 both in yeast and in mammalian cells. In this study, we tested the oligomerization of p53 with mutations in the oligomerization domain, when expressed in a human osteosarcoma cell line, Saos-2, in vivo. We found that single point mutations, including two missense and two nonsense mutations, in the alpha-helix of the oligomerization domain disrupted the oligomerization of p53, but that p53 still retained its ability to inhibit colony formation of cells to some degree. These results suggest that oligomerization and the carboxy-terminal basic domain are not prerequisite for p53-dependent tumor suppression, and this may explain why few of the tumor-derived p53 mutations that have been examined so far are carboxy-terminal mutations.
Collapse
Affiliation(s)
- C Ishioka
- Department of Clinical Oncology, Tohoku University, Sendai, Japan.
| | | | | | | | | | | | | | | | | |
Collapse
|
163
|
Almog N, Li R, Peled A, Schwartz D, Wolkowicz R, Goldfinger N, Pei H, Rotter V. The murine C'-terminally alternatively spliced form of p53 induces attenuated apoptosis in myeloid cells. Mol Cell Biol 1997; 17:713-22. [PMID: 9001225 PMCID: PMC231797 DOI: 10.1128/mcb.17.2.713] [Citation(s) in RCA: 17] [Impact Index Per Article: 0.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/03/2023] Open
Abstract
The onset of p53-dependent apoptosis results from the accumulation of damaged DNA. Recently, it was shown that the C' terminus of the p53 protein plays a central role in sensing damaged DNA. In our present study, we examined the role of the C' terminus in the induction of apoptosis. A temperature-sensitive (ts) mutant of the alternatively spliced form of p53 (p53AS-ts) and the ts mutant of the regularly spliced form (p53RS-ts) were used to generate series of stable clones with increasing amounts of p53 protein. Apoptotic patterns induced by either the regularly spliced p53 product (p53RS) or a C'-terminally alternatively spliced p53 product (p53AS) were compared. We found that although both forms of p53 induced apoptosis following expression of the wild-type protein conformation, the kinetics were different. Apoptosis induced by the p53AS protein was attenuated compared to that induced by p53RS. The delay in the manifestation of the apoptotic features following p53AS expression was in agreement with a delay in the regulation of the expression of apoptosis-related genes. The observation that p53 with an altered C' terminus is still capable of inducing apoptosis suggests that the actual onset of the apoptotic process most probably involves structural domains other than the C' terminus of the p53 molecule. However, the fact that the apoptotic activity mediated by the p53AS product was slower than that mediated by the p53RS product suggests that the C' terminus indeed exerts a certain control on the apoptotic activity of the p53 molecule.
Collapse
Affiliation(s)
- N Almog
- Department of Molecular Cell Biology, Weizmann Institute of Science, Rehovot, Israel
| | | | | | | | | | | | | | | |
Collapse
|
164
|
Affiliation(s)
- W D Kohn
- Department of Biochemistry, MRC Group in Protein Structure and Function, University of Alberta, Edmonton, Alberta T6G 2H7, Canada
| | | | | |
Collapse
|
165
|
Affiliation(s)
- L Grasso
- Department of Microbiology, Jefferson Medical College, Thomas Jefferson University, Philadelphia, Pennsylvania 19107, USA
| | | |
Collapse
|
166
|
Affiliation(s)
- J M Varley
- CRC Department of Cancer Genetics, Paterson Institute for Cancer Research, Manchester, UK
| | | | | |
Collapse
|
167
|
Miller M, Lubkowski J, Rao JK, Danishefsky AT, Omichinski JG, Sakaguchi K, Sakamoto H, Appella E, Gronenborn AM, Clore GM. The oligomerization domain of p53: crystal structure of the trigonal form. FEBS Lett 1996; 399:166-70. [PMID: 8980144 DOI: 10.1016/s0014-5793(96)01231-8] [Citation(s) in RCA: 32] [Impact Index Per Article: 1.1] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/03/2023]
Abstract
The structure of the oligomerization domain of the p53 tumor suppressor protein was determined in the trigonal crystal form, using a refined NMR structure as a model. A synthetic peptide comprising residues 319-360 of human p53 crystallized in the space group P3(1)21. There is one biologically relevant tetrameric domain in the crystallographic asymmetric unit. The structure was refined jointly with NMR data, only the third such case (the previous examples being IL-1beta (Shaanan, B., Gronenborn, A.M., Cohen, G.H., Gilliland, G.L., Veerapandian, B., Davies, D.R. and Clore, G.M. (1992) Science 257, 961-964 [1]) and BPTI (Schiffer, C., Huber, R., Wuthrich, K. and Van Gunsteren, W.F. (1994) J. Mol. Biol. 241, 588-599 [21)), to 2.5 A resolution with an R factor of 0.207. The distribution of tumor-derived mutations in the oligomerization region together with structural and biological data suggest a strategy for the design of antitumor therapeutics.
Collapse
Affiliation(s)
- M Miller
- Macromolecular Structure Laboratory, NCI-Frederick Cancer Research and Development Center, ABL-Basic Research Program, Frederick, MD 21702, USA
| | | | | | | | | | | | | | | | | | | |
Collapse
|
168
|
Kim KK, Daud AI, Wong SC, Pajak L, Tsai SC, Wang H, Henzel WJ, Field LJ. Mouse RAD50 has limited epitopic homology to p53 and is expressed in the adult myocardium. J Biol Chem 1996; 271:29255-64. [PMID: 8910585 DOI: 10.1074/jbc.271.46.29255] [Citation(s) in RCA: 16] [Impact Index Per Article: 0.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/03/2023] Open
Abstract
Previous studies have identified a 180-kDa mouse cardiomyocyte phosphoprotein with limited epitopic homology to p53. In this study, the protein was purified and partially sequenced. Oligonucleotide probes based on the available amino acid sequence data were used to isolate cDNA clones. Sequence analyses revealed that the clones encoded a protein with regional homology to the yeast RAD50 gene product. Expression of the mouse cDNA rescued the methyl methanesulfonate-sensitive phenotype in rad50 mutant yeast, indicating that the cardiomyocyte phosphoprotein is the mammalian homologue of the yeast RAD50 gene product. Fluorescence in situ hybridization analyses localized the mouse RAD50 gene to the A5-B1 region of chromosome 11. Northern blot analyses demonstrated a complex pattern of RAD50 expression during mouse development which was further complicated by the presence of several alternatively spliced transcripts. High levels of RAD50 expression was evident in the adult myocardium, a somewhat surprising observation given the absence of DNA synthesis in adult cardiomyocytes.
Collapse
Affiliation(s)
- K K Kim
- Krannert Institute of Cardiology, Indiana University School of Medicine, Indianapolis, Indiana 46202-4800, USA
| | | | | | | | | | | | | | | |
Collapse
|
169
|
Sakamoto H, Kodama H, Higashimoto Y, Kondo M, Lewis MS, Anderson CW, Appella E, Sakaguchi K. Chemical synthesis of phosphorylated peptides of the carboxy-terminal domain of human p53 by a segment condensation method. INTERNATIONAL JOURNAL OF PEPTIDE AND PROTEIN RESEARCH 1996; 48:429-42. [PMID: 8956076 DOI: 10.1111/j.1399-3011.1996.tb00861.x] [Citation(s) in RCA: 17] [Impact Index Per Article: 0.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 02/03/2023]
Abstract
A segment condensation method was developed for the chemical synthesis of large (> 90 amino acid) phosphopeptides and was used to produce phosphorylated and non-phosphorylated derivatives of the C-terminal tetramerization and regulatory domains of human p53 (residues 303-393). Efficient condensation synthesis of the 91 residue p53 domain was achieved in two steps. The non-phosphorylated N-terminal segment p53(303-334) (1) and its derivative phosphorylated at serine 315 (1P315), and the non-phosphorylated middle segment p53(335-360) (2), were synthesized as partially protected peptide thioesters in the solid phase using Boc chemistry. The C-terminal segment p53(361-393) (3) and its derivative phosphorylated at serine 392 (3P392) were synthesized as partially protected peptides in the solid phase using Fmoc chemistry. Phosphoamino acid was incorporated into the N-terminal segment (1P315) at the residue corresponding to p53 serine 315 as Boc-Ser(PO3(Bzl)2)-OH during synthesis. Serine 392 in the C-terminal segment was selectively phosphorylated after synthesis by phosphitylation followed by oxidation. A derivative phosphorylated at serine 378 was synthesized in a one-step condensation of the unphosphorylated N-terminal segment (1) and the phosphorylated long C-terminal segment p53(335-393) (2-3P378). Yields of the ligated peptides after removal of the protecting groups and HPLC purification averaged 60% for the first condensation and 35% for the second condensation. All five p53 peptides exhibited monomer-tetramer association as determined by analytical ultracentrifugation. Circular dichroism spectroscopy revealed that phosphorylation at Ser315 increased the alpha-helical content, which was abolished when Ser392 also was phosphorylated, suggesting an interaction between N-terminal and C-terminal residues of the C-terminal domain of p53.
Collapse
Affiliation(s)
- H Sakamoto
- Laboratory of Cell Biology, National Cancer Institute, National Institutes of Health, Bethesda, Maryland, USA
| | | | | | | | | | | | | | | |
Collapse
|
170
|
Runnebaum IB, Köhler T, Stickeler E, Kieback HR, Kreienberg R. p53 mutation is associated with high S-phase fraction in primary fallopian tube adenocarcinoma. Br J Cancer 1996; 74:1157-60. [PMID: 8883398 PMCID: PMC2075926 DOI: 10.1038/bjc.1996.510] [Citation(s) in RCA: 10] [Impact Index Per Article: 0.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Download PDF] [Figures] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/02/2023] Open
Abstract
Fallopian tube carcinoma (FTC) is a rare but lethal gynaecological malignancy. Four out of seven FTCs were identified with three point missense mutations, one single base deletion and one silent point mutation in the p53 gene. Genital-type HPV sequences were not detected. The S-phase fraction of tumours with mutant and wild-type p53 was 25.74% (median) and 12.55% (median) respectively.
Collapse
Affiliation(s)
- I B Runnebaum
- Department of Obstetrics and Gynaecology, University of Ulm, Germany.
| | | | | | | | | |
Collapse
|
171
|
Lin SR, Yang YC, Jung JH, Tsai JH. A significant decrease of the transcriptional activity of p53 mutants deriving from human functional adrenal tumors. DNA Cell Biol 1996; 15:793-803. [PMID: 8892752 DOI: 10.1089/dna.1996.15.793] [Citation(s) in RCA: 6] [Impact Index Per Article: 0.2] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/02/2023] Open
Abstract
Recently, our laboratory has found a high incidence (77%) of p53 gene mutations in human functional adrenal tumors. Furthermore, the majority of mutant sites were assembled at codons 100, 102, and 249. These mutation sites are not common, and there have been no studies addressing whether or not these mutants points or mutant styles cause the p53 protein to lose function. It has been well known that p53 is a transcription factor. To examine the transcriptional activities of these mutant p53 genes from patients with functional adrenal tumors, we constructed p53 expression plasmids from tumors and paired adjacent normal adrenal gland tissues, using a transient co-transfection assay with a reporter gene in H358 cells. Wild-type p53 from normal adrenal gland tissues specifically trans-activates the expression of a chloramphenicol acetyltransferase (CAT) reporter gene in H358 cells. Three mutant p53 proteins (at codons 100, 102, and 249, respectively) from tumors showed a >90% loss of transcriptional activity. One mutant at codon 68, other than at hot spots, remained at approximately 65% transcriptional activity. An immunoprecipitation assay showed that the mutant proteins of codon 68 and codon 102 could respond to the three monoclonal antibodies (PAbDO-1, PAb1620, and PAb421), indicating that there were no obvious changes in the antigenicity of the proteins. However, the mutant protein of codon 249 could not respond to the carboxy-terminus-specific antibody PAb421 and conformation-specific antibody PAb1620, indicating that there were some obvious changes in the conformation of the mutant proteins. The mutant protein of codon 100 could not be detected by immunoprecipitation assay but could be analyzed by Western blot. In a further study using a DNA-binding assay, it was shown that the loss of transcriptional activity was caused by the loss of DNA-binding ability. These results show that the p53 mutants, derived from functional adrenal tumors, actually lost DNA-binding ability and decreased the transcriptional activity. However, the role of the mutant protein in the tumorigenesis of functional adrenal tumors requires further investigation.
Collapse
Affiliation(s)
- S R Lin
- Department of Clinical Pathology, Kaohsiung Medical College, Taiwan
| | | | | | | |
Collapse
|
172
|
Oberleithner H, Schneider S, Bustamante JO. Atomic force microscopy visualizes ATP-dependent dissociation of multimeric TATA-binding protein before translocation into the cell nucleus. Pflugers Arch 1996; 432:839-44. [PMID: 8772134 DOI: 10.1007/s004240050206] [Citation(s) in RCA: 23] [Impact Index Per Article: 0.8] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/02/2023]
Abstract
The TATA-binding protein (TBP) is a universal transcription factor which plays an essential role in eukaryotic gene expression. As a karyophilic molecule, this cytosolic protein reaches its DNA-binding site through the transport channel of the nuclear pore complex. As occurs with other major cellular proteins, TBP forms multimers in solution, which is a limiting factor for nuclear translocation. While studying the nuclear translocation of TBP, we detected ATP-dependent multimerization of TBP with atomic force microscopy. In physiological solutions containing ATP, 14-molecule multimers dissociated into four-molecule multimers with a half-maximum dissociation constant of 10 microM. Electrophysiological experiments using isolated cell nuclei of cultured kidney cells revealed that TBP translocates into the cell nucleus only in the presence of ATP. When ATP was replaced with its slowly hydrolysing analogue, ATP[gamma-S] [i.e. adenosine 5'-o-(3-thiotriphosphate)], the aggregates remained intact and nuclear translocation was not possible. Taken together, our investigations suggest that TBP exhibits ATPase activity similar to that observed in relation to molecular chaperons. This activity secures physiological translocation of the transcription factor into the nucleus.
Collapse
Affiliation(s)
- H Oberleithner
- Department of Physiology, University of Würzburg School of Medicine, Röntgenring 9, D-97070 Würzburg, Germany
| | | | | |
Collapse
|
173
|
Bristow RG, Benchimol S, Hill RP. The p53 gene as a modifier of intrinsic radiosensitivity: implications for radiotherapy. Radiother Oncol 1996; 40:197-223. [PMID: 8940748 DOI: 10.1016/0167-8140(96)01806-3] [Citation(s) in RCA: 139] [Impact Index Per Article: 4.8] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/03/2023]
Abstract
BACKGROUND AND PURPOSE Experimental studies have implicated the normal or "wild type' p53 protein (i.e. WTp53) in the cellular response to ionizing radiation and other DNA damaging agents. Whether altered WTp53 protein function can lead to changes in cellular radiosensitivity and/or clinical radiocurability remains an area of ongoing study. In this review, we describe the potential implications of altered WTp53 protein function in normal and tumour cells as it relates to clinical radiotherapy, and describe novel treatment strategies designed to re-institute WTp53 protein function as a means of sensitizing cells to ionizing radiation. METHODS AND MATERIALS A number of experimental and clinical studies are critically reviewed with respect to the role of the p53 protein as a determinant of cellular oncogenesis, genomic stability, apoptosis, DNA repair and radioresponse in normal and transformed mammalian cells. RESULTS In normal fibroblasts, exposure to ionizing radiation leads to a G1 cell cycle delay (i.e. a "G1 checkpoint') as a result of WTp53 mediated inhibition of G1-cyclin-kinase and retinoblastoma (pRb) protein function. The G1 checkpoint response is absent in tumour cells which express a mutant form of the p53 protein (i.e. MTp53), leading to acquired radioresistance in vitro. Depending on the cell type studied, this increase in cellular radiation survival can be mediated through decreased radiation-induced apoptosis, or altered kinetics of the radiation-induced G1 checkpoint. Recent biochemical studies support an indirect role for the p53 protein in both nucleotide excision and recombinational DNA repair pathways. However, based on clinicopathologic data, it remains unclear as to whether WTp53 protein function can predict for human tumour radiocurability and normal tissue radioresponse. CONCLUSIONS Alterations in cell cycle control secondary to aberrant WTp53 protein function may be clinically significant if they lead to the acquisition of mutant cellular phenotypes, including the radioresistant phenotype. Pre-clinical studies suggest that these phenotypes may be reversed using adenovirus-mediated gene therapy or pharmacologic strategies designed to re-institute WTp53 protein function. Our analysis of the published data strongly argues for the use of functional assays for the determination of WTp53 protein function in studies which attempt to correlate normal and tumour tissue radioresponse with p53 genotype, or p53 protein expression.
Collapse
Affiliation(s)
- R G Bristow
- Department of Radiation Oncology, University of Toronto, Canada.
| | | | | |
Collapse
|
174
|
Dahiya R, Deng G, Chen KM, Chui RM, Haughney PC, Narayan P. P53 tumour-suppressor gene mutations are mainly localised on exon 7 in human primary and metastatic prostate cancer. Br J Cancer 1996; 74:264-8. [PMID: 8688333 PMCID: PMC2074577 DOI: 10.1038/bjc.1996.349] [Citation(s) in RCA: 23] [Impact Index Per Article: 0.8] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Download PDF] [Figures] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/01/2023] Open
Abstract
Mutations in the p53 tumour-suppressor gene are among the most common genetic alterations in human cancers. In the present study we analysed the mutations in the p53 tumor-suppressor gene in 25 primary and 20 metastatic human prostate cancer specimens. DNA extracted from the paraffin-embedded sections was amplified by hot-start polymerase chain reaction, and p53 gene mutations in the conserved mid-region (exons 4-9) were examined using single-strand conformation polymorphism (SSCP) analysis and immunohistochemistry. In the present study, we used a novel hot-start PCR-SSCP technique using DNA Taq polymerase antibody, which eliminates primer-dimers and non-specific products. Because of this new technique, the results of PCR-SSCP showed very high resolution. Polymerase chain reaction products were sequenced directly for point mutations for the p53 gene. Mutations were found in 2 out of 25 primary prostate cancers (8%) and 4 out of 20 metastatic cancers (20%). Mutations were observed exclusively in exon 7 and not in exons 4, 5, 6, 8 or 9. Nuclear accumulation of p53 protein, determined by immunohistochemistry, correlated with the degree of metastasis in prostatic cancer.
Collapse
Affiliation(s)
- R Dahiya
- Department of Urology, University of California San Francisco, USA
| | | | | | | | | | | |
Collapse
|
175
|
Affiliation(s)
- Cecil Dybowski
- Department of Chemistry and Biochemistry, University of Delaware, Newark, Delaware 19716-2522
| | - Martha D. Bruch
- Department of Chemistry and Biochemistry, University of Delaware, Newark, Delaware 19716-2522
| |
Collapse
|
176
|
Stivers JT, Abeygunawardana C, Whitman CP, Mildvan AS. 4-Oxalocrotonate tautomerase, a 41-kDa homohexamer: backbone and side-chain resonance assignments, solution secondary structure, and location of active site residues by heteronuclear NMR spectroscopy. Protein Sci 1996; 5:729-41. [PMID: 8845763 PMCID: PMC2143398 DOI: 10.1002/pro.5560050418] [Citation(s) in RCA: 17] [Impact Index Per Article: 0.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/02/2023]
Abstract
4-Oxalocrotonate tautomerase (4-OT), a homohexamer consisting of 62 residues per subunit, catalyzes the isomerization of unsaturated alpha-keto acids using Pro-1 as a general base (Stivers et al., 1996a, 1996b). We report the backbone and side-chain 1H, 15N, and 13C NMR assignments and the solution secondary structure for 4-OT using 2D and 3D homonuclear and heteronuclear NMR methods. The subunit secondary structure consists of an alpha-helix (residues 13-30), two beta-strands (beta 1, residues 2-8; beta 2, residues 39-45), a beta-hairpin (residues 50-57), two loops (I, residues 9-12; II, 34-38), and two turns (I, residues 30-33; II, 47-50). The remaining residues form coils. The beta 1 strand is parallel to the beta 2 strand of the same subunit on the basis of cross stand NH(i)-NH(j) NOEs in a 2D 15N-edited 1H-NOESY spectrum of hexameric 4-OT containing two 15N-labeled subunits/hexamer. The beta 1 strand is also antiparallel to another beta 1 strand from an adjacent subunit forming a subunit interface. Because only three such pairwise interactions are possible, the hexamer is a trimer of dimers. The diffusion constant, determined by dynamic light scattering, and the rotational correlation time (14.5 ns) estimated from 15N T1/T2 measurements, are consistent with the hexameric molecular weight of 41 kDa. Residue Phe-50 is in the active site on the basis of transferred NOEs to the bound partial substrate 2-oxo-1,6-hexanedioate. Modification of the general base, Pro-1, with the active site-directed irreversible inhibitor, 3-bromopyruvate, significantly alters the amide 15N and NH chemical shifts of residues in the beta-hairpin and in loop II, providing evidence that these regions change conformation when the active site is occupied.
Collapse
Affiliation(s)
- J T Stivers
- Department of Biological Chemistry, Johns Hopkins University School of Medicine, Baltimore, Maryland 21205-2185, USA
| | | | | | | |
Collapse
|
177
|
Tsai CJ, Lin SL, Wolfson HJ, Nussinov R. Protein-protein interfaces: architectures and interactions in protein-protein interfaces and in protein cores. Their similarities and differences. Crit Rev Biochem Mol Biol 1996; 31:127-52. [PMID: 8740525 DOI: 10.3109/10409239609106582] [Citation(s) in RCA: 93] [Impact Index Per Article: 3.2] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/01/2023]
Abstract
Protein structures generally consist of favorable folding motifs formed by specific arrangements of secondary structure elements. Similar architectures can be adopted by different amino acids sequences, although the details of the structures vary. It has long been known that despite the sequence variability, there is a striking preferential conservation of the hydrophobic character of the amino acids at the buried positions of these folding motifs. Differences in the sizes of the side-chains are accommodated by movements of the secondary structure elements with respect to each other, leading to compact packing. Scanning protein-protein interfaces reveals that similar architectures are also observed at and around their interacting surfaces, with preservation of the hydrophobic character, although not to the same extent. The general forces that determine the origin of the native structures of proteins have been investigated intensively. The major non-bonded forces operating on a protein chain as it folds into a three-dimensional structure are likely to be packing, the hydrophobic effect, and electrostatic interactions. While the substantial hydrophobic forces lead to a compact conformation, they are also nonspecific and cannot serve as a guide to a conformationally unique structure. For the general folding problem, it thus appears that packing is a prime candidate for determining a particular fold. Specific hydrogen-bonding patterns and salt-bridges have also been proposed to play a role. Inspection of protein-protein interfaces reveals that the hallmarks governing single chain protein structures also determine their interactions, suggesting that similar principles underlie protein folding and protein-protein associations. This review focuses on some aspects of protein-protein interfaces, particularly on the architectures and their interactions. These are compared with those present in protein monomers. This task is facilitated by the recently compiled, non-redundant structural dataset of protein-protein interfaces derived from the crystallographic database. In particular, although current view holds that protein-protein interfaces and interactions are similar to those found in the conformations of single-chain proteins, this review brings forth the differences as well. Not only is it logical that such differences would exist, it is these differences that further illuminate protein folding on the one hand and protein-protein recognition on the other. These are also particularly important in considering inhibitor (ligand) design.
Collapse
Affiliation(s)
- C J Tsai
- Laboratory of Mathematical Biology, NCI-FCRDC, Frederick, MD 21702, USA
| | | | | | | |
Collapse
|
178
|
Affiliation(s)
- K S Matthews
- Department of Biochemistry and Cell Biology, Rice University, Houston, TX 77251, USA.
| |
Collapse
|
179
|
Hansen S, Hupp TR, Lane DP. Allosteric regulation of the thermostability and DNA binding activity of human p53 by specific interacting proteins. CRC Cell Transformation Group. J Biol Chem 1996; 271:3917-24. [PMID: 8632013 DOI: 10.1074/jbc.271.7.3917] [Citation(s) in RCA: 77] [Impact Index Per Article: 2.7] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/01/2023] Open
Abstract
Conformational stability is a prerequisite for the physiological activity of the tumor suppressor protein p53. p53 protein can be allosterically activated for DNA binding by phosphorylation or through noncovalent interaction with proteins such as DnaK, the Escherichia coli homologue of the heat shock protein Hsp70. We present in vitro evidence for a rapid temperature-dependent change in the conformation and tetrameric nature of wild-type p53 upon incubation at 37 degrees C, which correlates with a permanent loss in DNA binding activity. We show that p53 is allosterically regulated for stabilization of the wild-type conformation and DNA binding activity at 37 degrees C by binding of two classes of ligands to regulatory sites on the N and C terminus of the molecule through which an intrinsic instability of p53 is neutralized. Deletion of the domain conferring instability at the C terminus is sufficient to confer enhanced stability to the total protein. DnaK binding to the C terminus can profoundly protect p53 at 37 degrees C from a temperature-dependent loss of the DNA binding activity but does not renature or activate denatured p53. In contrast, another activator of the DNA binding activity of latent p53, the monoclonal antibody PAb421, which also interacts with the C terminus of the protein, is not able to protect p53 from thermal denaturation. Two monoclonal antibodies to the N terminus of p53, PAb1801 and DO-1, do not activate the latent DNA binding function of p53 but can protect the p53 wild-type conformation at 37 degrees C. Thus, activation of the DNA binding function of p53 is not synonymous with protection from thermal denaturation, and therefore, both of these pathways may be used in cells to control the physiological activity of p53. The protection of p53 conformation from heat denaturation by interacting proteins suggests a novel mechanism by which p53 function could be regulated in vivo.
Collapse
Affiliation(s)
- S Hansen
- Department of Biochemistry, Medical Sciences Institute, The University of Dundee, Dundee, Scotland
| | | | | |
Collapse
|
180
|
Harris CC. The 1995 Walter Hubert Lecture--molecular epidemiology of human cancer: insights from the mutational analysis of the p53 tumour-suppressor gene. Br J Cancer 1996; 73:261-9. [PMID: 8562328 PMCID: PMC2074430 DOI: 10.1038/bjc.1996.47] [Citation(s) in RCA: 82] [Impact Index Per Article: 2.8] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Download PDF] [Figures] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/31/2023] Open
Affiliation(s)
- C C Harris
- Laboratory of Human Carcinogenesis, National Cancer Institute, NIH, Bethesda, MD 20892-4255, USA
| |
Collapse
|
181
|
Rand A, Glenn K, Alvares C, White M, Thibodeau S, Karnes W. p53 functional loss in a colon cancer cell line with two missense mutations (218leu and 248trp) on separate alleles. Cancer Lett 1996. [DOI: 10.1016/s0304-3835(06)80030-3] [Citation(s) in RCA: 1] [Impact Index Per Article: 0.0] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/28/2022]
|
182
|
Wagner P, Fuchs A, Prowald A, Montenarh M, Nastainczyk W. Precise mapping of the tms1 binding site on p53. FEBS Lett 1995; 377:155-8. [PMID: 8543040 DOI: 10.1016/0014-5793(95)01329-6] [Citation(s) in RCA: 9] [Impact Index Per Article: 0.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/31/2023]
Abstract
Originally identified as multicopy suppressor of a lethal growth arrest caused by expression of a tumour mutant cDNA of p53 in fission yeast the tms1 gene product was found to form stable complexes with p53 in yeast. By using purified recombinant proteins multimeric complexes of tms1 and p53 could be demonstrated and recently the p53 binding site on the tms1 protein was established to the sequence YYITTEDFCT (aa 116-125) in the vicinity of a well conserved cell division motif. Here we report the precise mapping of the tms1 binding site on the p53 protein to the sequence LQIRGRERFE (aa 330-339) which defines a new functional domain on the p53 protein.
Collapse
Affiliation(s)
- P Wagner
- University of Saarland, Homburg/Saar, Germany
| | | | | | | | | |
Collapse
|
183
|
Abstract
p53 is a multifunctional protein which plays a role in modulating gene transcription, policing cell cycle checkpoints, activating apoptosis, controlling DNA replication and repair, maintaining genomic stability and responding to genetic insults. Mutation of the p53 gene confers the single greatest known selective advantage favoring cancer formation. Point mutations result not only in the loss of tumor suppressor functions, but also in the gain of tumor promotion functions. These dual circumstances may be unique to p53 and, in part, could explain the relatively powerful force behind this selection pressure. General mechanisms of gain of function by mutated p53 may include alteration in transcriptional modulation and newly acquired targets for transcriptional regulation and protein binding. Despite the direct significance of p53 mutations, loss of the remaining wild-type allele is usually required for the formation of tumors in the natural setting. Novel applications of the basic scientific knowledge of p53 could lead to an improvement in cancer treatment, hopefully in the not so distant future.
Collapse
Affiliation(s)
- R M Elledge
- Division of Medical Oncology, University of Texas Health Science Center at San Antonio 78284, USA
| | | |
Collapse
|
184
|
Abstract
Mutation of the p53 gene is among the most common lesions in a variety of human tumors, including those of the central nervous system. In most instances, mutation of one p53 allele is followed by loss of the remaining wild-type allele, resulting in cells with a complete absence of functional wild-type p53 protein. However, in some situations, such as at initiation of spontaneously arising gliomas or as the germline configuration of patients with the Li-Fraumeni syndrome, cells clearly carry both wild-type and mutant p53 alleles. These observations lead to the hypothesis that p53 mutations can give rise to loss of tumor suppressor functions as well as to gain of oncogenic transformation capabilities. In this review, we define the types of mutations that occur in the p53 gene in various glial tumors, contrast that with the spectra described in other human tumor types, and discuss the biochemistry and physiology of the p53 protein and its ability to regulate and be regulated by other gene products. We use this information to propose roles for p53 in the initiation and progression of human gliomas.
Collapse
Affiliation(s)
- O Bögler
- Ludwig Institute for Cancer Research, University of California San Diego, La Jolla 92093-0660, USA
| | | | | | | |
Collapse
|
185
|
Balagurumoorthy P, Sakamoto H, Lewis MS, Zambrano N, Clore GM, Gronenborn AM, Appella E, Harrington RE. Four p53 DNA-binding domain peptides bind natural p53-response elements and bend the DNA. Proc Natl Acad Sci U S A 1995; 92:8591-5. [PMID: 7567980 PMCID: PMC41012 DOI: 10.1073/pnas.92.19.8591] [Citation(s) in RCA: 106] [Impact Index Per Article: 3.5] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/26/2023] Open
Abstract
Recent structural studies of the minimal core DNA-binding domain of p53 (p53DBD) complexed to a single consensus pentamer sequence and of the isolated p53 tetramerization domain have provided valuable insights into their functions, but many questions about their interacting roles and synergism remain unanswered. To better understand these relationships, we have examined the binding of the p53DBD to two biologically important full-response elements (the WAF1 and ribosomal gene cluster sites) by using DNA circularization and analytical ultracentrifugation. We show that the p53DBD binds DNA strongly and cooperatively with p53DBD to DNA binding stoichiometries of 4:1. For the WAF1 element, the mean apparent Kd is (8.3 +/- 1.4) x 10(-8) M, and no intermediate species of lower stoichiometries can be detected. We show further that complex formation induces an axial bend of at least 60 degrees in both response elements. These results, taken collectively, demonstrate that p53DBD possesses the ability to direct the formation of a tight nucleoprotein complex having the same 4:1 DNA-binding stoichiometry as wild-type p53 which is accompanied by a substantial conformational change in the response-element DNA. This suggests that the p53DBD may play a role in the tetramerization function of p53. A possible role in this regard is proposed.
Collapse
Affiliation(s)
- P Balagurumoorthy
- Department of Biochemistry, University of Nevada Reno 89557-0014, USA
| | | | | | | | | | | | | | | |
Collapse
|
186
|
Kersten S, Kelleher D, Chambon P, Gronemeyer H, Noy N. Retinoid X receptor alpha forms tetramers in solution. Proc Natl Acad Sci U S A 1995; 92:8645-9. [PMID: 7567990 PMCID: PMC41023 DOI: 10.1073/pnas.92.19.8645] [Citation(s) in RCA: 76] [Impact Index Per Article: 2.5] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/26/2023] Open
Abstract
Protein-protein interactions allow the retinoid X receptor (RXR) to bind to cognate DNA as a homo- or a heterodimer and to participate in mediating the effects of a variety of hormones on gene transcription. Here we report a systematic study of the oligomeric state of RXR in the absence of a DNA template. We have used electrophoresis under nondenaturing conditions and chemical crosslinking to show that in solution, RXR alpha forms homodimers as well as homotetramers. The dissociation constants governing dimer and tetramer formation were estimated by fluorescence anisotropy studies. The results indicate that RXR tetramers are formed with a high affinity and that at protein concentrations higher than about 70 nM, tetramers will constitute the predominant species. Tetramer formation may provide an additional level of the regulation of gene transcription mediated by RXRs.
Collapse
Affiliation(s)
- S Kersten
- Cornell University, Division of Nutritional Sciences, Ithaca, NY 14853-6301, USA
| | | | | | | | | |
Collapse
|
187
|
Thukral SK, Lu Y, Blain GC, Harvey TS, Jacobsen VL. Discrimination of DNA binding sites by mutant p53 proteins. Mol Cell Biol 1995; 15:5196-202. [PMID: 7651437 PMCID: PMC230767 DOI: 10.1128/mcb.15.9.5196] [Citation(s) in RCA: 34] [Impact Index Per Article: 1.1] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/26/2023] Open
Abstract
Critical determinants of DNA recognition by p53 have been identified by a molecular genetic approach. The wild-type human p53 fragment containing amino acids 71 to 330 (p53(71-330)) was used for in vitro DNA binding assays, and full-length human p53 was used for transactivation assays with Saccharomyces cerevisiae. First, we defined the DNA binding specificity of the wild-type p53 fragment by using systematically altered forms of a known consensus DNA site. This refinement indicates that p53 binds with high affinity to two repeats of PuGPuCA.TGPyCPy, a further refinement of an earlier defined consensus half site PuPuPuC(A/T).(T/A) GPyPyPy. These results were further confirmed by transactivation assays of yeast by using full-length human p53 and systematically altered DNA sites. Dimers of the pentamer AGGCA oriented either head-to-head or tail-to-tail bound efficiently, but transactivation was facilitated only through head-to-head dimers. To determine the origins of specificity in DNA binding by p53, we identified mutations that lead to altered specificities of DNA binding. Single-amino-acid substitutions were made at several positions within the DNA binding domain of p53, and this set of p53 point mutants were tested with DNA site variants for DNA binding. DNA binding analyses showed that the mutants Lys-120 to Asn, Cys-277 to Gln or Arg, and Arg-283 to Gln bind to sites with noncanonical base pair changes at positions 2, 3, and 1 in the pentamer (PuGPuCA), respectively. Thus, we implicate these residues in amino acid-base pair contacts. Interestingly, mutant Cys-277 to Gln bound a consensus site as two and four monomers, as opposed to the wild-type p53 fragment, which invariably binds this site as four monomers.
Collapse
Affiliation(s)
- S K Thukral
- Department of Developmental Biology, Amgen, Thousand Oaks, California 91320, USA
| | | | | | | | | |
Collapse
|
188
|
Wolkowicz R, Peled A, Elkind NB, Rotter V. Augmented DNA-binding activity of p53 protein encoded by a carboxyl-terminal alternatively spliced mRNA is blocked by p53 protein encoded by the regularly spliced form. Proc Natl Acad Sci U S A 1995; 92:6842-6. [PMID: 7624329 PMCID: PMC41425 DOI: 10.1073/pnas.92.15.6842] [Citation(s) in RCA: 22] [Impact Index Per Article: 0.7] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/26/2023] Open
Abstract
DNA-binding activity of the wild-type p53 is central to its function in vivo. However, recombinant or in vitro translated wild-type p53 proteins, unless modified, are poor DNA binders. The fact that the in vitro produced protein gains DNA-binding activity upon modification at the C terminus raises the possibility that similar mechanisms may exist in the cell. Data presented here show that a C-terminal alternatively spliced wild-type p53 (ASp53) mRNA expressed by bacteria or transcribed in vitro codes for a p53 protein that efficiently binds DNA. Our results support the conclusion that the augmented DNA binding activity of an ASp53 protein is probably due to attenuation of the negative effect residing at the C terminus of the wild-type p53 protein encoded by the regularly spliced mRNA (RSp53) rather than acquisition of additional functionality by the alternatively spliced C' terminus. In addition, we found that ASp53 forms a complex with the non-DNA-binding RSp53, which in turn blocks the DNA-binding activity of ASp53. Interaction between these two wild-type p53 proteins may underline a mechanism that controls the activity of the wild-type p53 protein in the cell.
Collapse
Affiliation(s)
- R Wolkowicz
- Department of Cell Biology, Weizmann Institute of Science, Rehovot, Israel
| | | | | | | |
Collapse
|
189
|
Lee S, Elenbaas B, Levine A, Griffith J. p53 and its 14 kDa C-terminal domain recognize primary DNA damage in the form of insertion/deletion mismatches. Cell 1995; 81:1013-20. [PMID: 7600570 DOI: 10.1016/s0092-8674(05)80006-6] [Citation(s) in RCA: 280] [Impact Index Per Article: 9.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/26/2023]
Abstract
Insertion/deletion (IDL) mismatches in DNA are lesions consisting of extra bases on one strand. Here, the binding of p53 and its 14 kDa C-terminal domain to DNAs containing one or three 3-cytosine IDL mismatches was examined. Electron microscopy showed that both p53 forms bound predominantly as tetramers at the lesions while single-stranded binding proteins did not bind. Gel retardation assays showed that p53 formed highly stable complexes when the DNA contained the IDL mismatches, but only unstable complexes when the DNA lacked lesions (but did contain free ends). The highly stable complexes had a half-life of > 2 hr, suggesting that upon encountering lesions, p53 may recruit other proteins to the site, providing a signal for DNA damage.
Collapse
Affiliation(s)
- S Lee
- Lineberger Comprehensive Cancer Center, University of North Carolina at Chapel Hill 27599-7295, USA
| | | | | | | |
Collapse
|
190
|
Jayaraman J, Prives C. Activation of p53 sequence-specific DNA binding by short single strands of DNA requires the p53 C-terminus. Cell 1995; 81:1021-9. [PMID: 7600571 DOI: 10.1016/s0092-8674(05)80007-8] [Citation(s) in RCA: 270] [Impact Index Per Article: 9.0] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/26/2023]
Abstract
Upon cellular DNA damage, the p53 tumor suppressor protein transmits a signal to genes that control the cell cycle and apoptosis. One function of p53 that is important for its role in this pathway is its ability to function as a sequence-specific transcriptional activator. We demonstrate here that short single DNA strands can markedly stimulate the ability of human and murine p53 proteins to bind specifically to a p53 response element in supercoiled DNA. We also show that single-stranded DNA does not stimulate binding by a truncated p53 that lacks the C-terminal domain. Finally, we establish that a peptide spanning the p53 C-terminus has the ability in trans to stimulate sequence-specific DNA binding by p53 dramatically. These data taken together suggest a model in which the p53 C-terminus can recognize DNA structures resulting from damage-induced lesions, and this interaction can be propagated to regulate positively p53 sequence-specific DNA binding.
Collapse
Affiliation(s)
- J Jayaraman
- Department of Biological Sciences, Columbia University, New York, New York 10027, USA
| | | |
Collapse
|
191
|
Bayle JH, Elenbaas B, Levine AJ. The carboxyl-terminal domain of the p53 protein regulates sequence-specific DNA binding through its nonspecific nucleic acid-binding activity. Proc Natl Acad Sci U S A 1995; 92:5729-33. [PMID: 7777576 PMCID: PMC41770 DOI: 10.1073/pnas.92.12.5729] [Citation(s) in RCA: 82] [Impact Index Per Article: 2.7] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/27/2023] Open
Abstract
The murine p53 protein contains two nucleic acid-binding sites, a sequence-specific DNA-binding region localized between amino acid residues 102-290 and a nucleic acid-binding site without sequence specificity that has been localized to residues 364-390. Alternative splicing of mRNA generates two forms of this p53 protein. The normal, or majority, splice form (NSp53) retains its carboxyl-terminal sequence-nonspecific nucleic acid-binding site, which can negatively regulate the sequence-specific DNA-binding site. The alternative splice form of p53 (ASp53) replaces amino acid residues 364-390 with 17 different amino acids. This protein fails to bind nucleic acids nonspecifically and is constitutive for sequence-specific DNA binding. Thus, the binding of nucleic acids at the carboxyl terminus regulates sequence-specific DNA binding by p53. The implications of these findings for the activation of p53 transcriptional activity following DNA damage are discussed.
Collapse
Affiliation(s)
- J H Bayle
- Department of Molecular Biology, Lewis Thomas Laboratory, Princeton University, NJ 08544, USA
| | | | | |
Collapse
|
192
|
Abstract
One of the most commonly detected abnormalities in human cancer is mutation of the p53 tumour suppressor gene. Intrinsic to the function of p53 is its ability to induce apoptotic cell death and to cause cell cycle arrest. Moreover, p53 plays an important role in controlling the cellular response to DNA damaging agents such as ionizing radiation and cancer chemotherapeutic drugs. Loss of p53 function causes increased resistance to radiation and chemotherapeutic agents, and there is increasing evidence that p53 mutational status is an important determinant of clinical outcome in cancer. This review will focus on recent data describing the biochemistry of p53 function, its role in mediating apoptosis and cell cycle arrest and in the control of tumour growth and death.
Collapse
Affiliation(s)
- J M Lee
- Division of Molecular and Developmental Biology, Samuel Lunenfeld Research Institute, Mount Sinai Hospital, Toronto, Ontario, Canada
| | | |
Collapse
|
193
|
LaCasse EC, Lefebvre YA. Nuclear localization signals overlap DNA- or RNA-binding domains in nucleic acid-binding proteins. Nucleic Acids Res 1995; 23:1647-56. [PMID: 7540284 PMCID: PMC306917 DOI: 10.1093/nar/23.10.1647] [Citation(s) in RCA: 178] [Impact Index Per Article: 5.9] [Reference Citation Analysis] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/25/2023] Open
Affiliation(s)
- E C LaCasse
- Department of Medicine, University of Ottawa, Loeb Institute for Medical Research, Ottawa Civic Hospital, Canada
| | | |
Collapse
|
194
|
Devon RS, Porteous DJ, Brookes AJ. Splinkerettes--improved vectorettes for greater efficiency in PCR walking. Nucleic Acids Res 1995; 23:1644-5. [PMID: 7784225 PMCID: PMC306912 DOI: 10.1093/nar/23.9.1644] [Citation(s) in RCA: 157] [Impact Index Per Article: 5.2] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/27/2023] Open
Affiliation(s)
- R S Devon
- MRC Human Genetics Unit, Western General Hospital, Edinburgh, UK
| | | | | |
Collapse
|
195
|
Clubb RT, Omichinski JG, Sakaguchi K, Appella E, Gronenborn AM, Clore GM. Backbone dynamics of the oligomerization domain of p53 determined from 15N NMR relaxation measurements. Protein Sci 1995; 4:855-62. [PMID: 7663341 PMCID: PMC2143111 DOI: 10.1002/pro.5560040505] [Citation(s) in RCA: 26] [Impact Index Per Article: 0.9] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/26/2023]
Abstract
The backbone dynamics of the tetrameric p53 oligomerization domain (residues 319-360) have been investigated by two-dimensional inverse detected heteronuclear 1H-15N NMR spectroscopy at 500 and 600 MHz. 15N T1, T2, and heteronuclear NOEs were measured for 39 of 40 non-proline backbone NH vectors at both field strengths. The overall correlation time for the tetramer, calculated from the T1/T2 ratios, was found to be 14.8 ns at 35 degrees C. The correlation times and amplitudes of the internal motions were extracted from the relaxation data using the model-free formalism (Lipari G, Szabo A, 1982, J Am Chem Soc 104:4546-4559). The internal dynamics of the structural core of the p53 oligomerization domain are uniform and fairly rigid, with residues 327-354 exhibiting an average generalized order parameter (S2) of 0.88 +/- 0.08. The N- and C-termini exhibit substantial mobility and are unstructured in the solution structure of p53. Residues located at the N- and C-termini, in the beta-sheet, in the turn between the alpha-helix and beta-sheet, and at the C-terminal end of the alpha-helix display two distinct internal motions that are faster than the overall correlation time. Fast internal motions (< or = 20 ps) are within the extreme narrowing limit and are of uniform amplitude. The slower motions (0.6-2.2 ns) are outside the extreme narrowing limit and vary in amplitude.(ABSTRACT TRUNCATED AT 250 WORDS)
Collapse
Affiliation(s)
- R T Clubb
- Laboratory of Chemical Physics, National Institute of Diabetes and Digestive and Kidney Diseases, National Institutes of Health, Bethesda, Maryland 20892, USA
| | | | | | | | | | | |
Collapse
|
196
|
Torres AM, Forbes BE, Aplin SE, Wallace JC, Francis GL, Norton RS. Solution structure of human insulin-like growth factor II. Relationship to receptor and binding protein interactions. J Mol Biol 1995; 248:385-401. [PMID: 7739048 DOI: 10.1016/s0022-2836(95)80058-1] [Citation(s) in RCA: 39] [Impact Index Per Article: 1.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/26/2023]
Abstract
The three-dimensional structure of human insulin-like growth factor (IGF) II in aqueous solution at pH 3.1 and 300 K has been determined from nuclear magnetic resonance data and restrained molecular dynamics calculations. Structural constraints consisting of 502 NOE-derived distance constraints, 11 dihedral angle restraints, and three disulfide bridges were used as input for distance geometry calculations in DIANA and X-PLOR, followed by simulated annealing refinement and energy minimization in X-PLOR. The resulting family of 20 structures was well defined in the regions of residues 5 to 28 and 41 to 62, with an average pairwise root-mean-square deviation of 1.24 A for the backbone heavy-atoms (N, C2, C) and 1.90 A for all heavy atoms. The poorly defined regions consist of the N and C termini, part of the B-domain, and the C-domain loop. Resonances from these regions of the protein gave stronger cross peaks in two dimensional NMR spectra, consistent with significant motional averaging. The main secondary structure elements in IGF-II are alpha-helices encompassing residues 11 to 21, 42 to 49 and 53 to 59. A small anti-parallel beta-sheet is formed by residues 59 to 61 and 25 to 27, while residues 26 to 28 appear to participate in intermolecular beta-sheet formation. The structure of IGF-II in the well-defined regions is very similar to those of the corresponding regions of insulin and IGF-I. Significant differences between IGF-II and IGF-I occur near the start of the third helix, in a region known to modulate affinity for the type 2 IGF receptor, and at the C terminus. The IGF II structure is discussed in relation to its binding sites for the insulin and IGF receptors and the IGF binding proteins.
Collapse
Affiliation(s)
- A M Torres
- NMR Laboratory Biomolecular Research Institute, Parkville, Australia
| | | | | | | | | | | |
Collapse
|
197
|
Clore GM, Ernst J, Clubb R, Omichinski JG, Kennedy WM, Sakaguchi K, Appella E, Gronenborn AM. Refined solution structure of the oligomerization domain of the tumour suppressor p53. NATURE STRUCTURAL BIOLOGY 1995; 2:321-33. [PMID: 7796267 DOI: 10.1038/nsb0495-321] [Citation(s) in RCA: 165] [Impact Index Per Article: 5.5] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 01/27/2023]
Abstract
The NMR solution structure of the oligomerization domain of the tumour suppressor p53 (residues 319-360) has been refined. The structure comprises a dimer of dimers, oriented in an approximately orthogonal manner. The present structure determination is based on 4,472 experimental NMR restraints which represents a three and half fold increase over our previous work in the number of NOE restraints at the tetramerization interface. A comparison with the recently solved 1.7 A resolution X-ray structure shows that the structures are very similar and that the average angular root-mean-square difference in the interhelical angles is about 1 degree. The results of recent extensive mutagenesis data and the possible effects of mutations which have been identified in human cancers are discussed in the light of the present structure.
Collapse
Affiliation(s)
- G M Clore
- Laboratory of Chemical Physics, National Institute of Diabetes and Digestive and Kidney Diseases, National Institutes of Health, Bethesda, MD 20892-0520, USA
| | | | | | | | | | | | | | | |
Collapse
|
198
|
Wang Y, Schwedes JF, Parks D, Mann K, Tegtmeyer P. Interaction of p53 with its consensus DNA-binding site. Mol Cell Biol 1995; 15:2157-65. [PMID: 7891710 PMCID: PMC230443 DOI: 10.1128/mcb.15.4.2157] [Citation(s) in RCA: 124] [Impact Index Per Article: 4.1] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/27/2023] Open
Abstract
We have analyzed the specific interaction of murine p53 with the consensus DNA-binding sequence 5'-AGACATGCCT-AGACATGCCT-3'. We used segments of p53 lacking the C-terminal, nonspecific DNA-binding domain because the presence of an autonomous nonspecific DNA-binding domain in wild-type p53 would complicate analysis of site-specific DNA binding. p53 amino acids 1 to 360 bind the consensus sequence as tetramers, and DNA binding promotes tetramer-tetramer interactions. p53 amino acids 80 to 290, lacking both the nonspecific DNA-binding and tetramerization domains, consistently bind consensus DNA as four monomers and only as four monomers. The virtual absence of stable binding by fewer than four monomers, even at low concentrations of p53, argues that binding by amino acids 80 to 290 is strongly cooperative. Because p53 tetramers and monomers do not simultaneously bind a single DNA consensus sequence, we conclude that a single tetramer of wild-type p53 engages the recognition sequences of the entire DNA consensus site. We further show that consensus DNA consists of two functional half-sites. Insertions, deletions, or rearrangements within the half-sites reduce DNA binding dramatically. In contrast, two half-sites separated by insertions bind p53 relatively efficiently. Insertions that place half-sites on opposite faces of the DNA helix reduce DNA binding more than insertions that place half-sites on the same face of the helix. Transcription studies, in vivo, strongly confirm the rotational specificity of the p53 interaction with consensus DNA. The ability of single p53 tetramers to bind separated DNA half-sites argues that p53 has a flexible tetramerization region.
Collapse
Affiliation(s)
- Y Wang
- Department of Molecular Genetics and Microbiology, State University of New York, Stony Brook 11794
| | | | | | | | | |
Collapse
|
199
|
|
200
|
Clore GM, Omichinski JG, Sakaguchi K, Zambrano N, Sakamoto H, Appella E, Gronenborn AM. Interhelical angles in the solution structure of the oligomerization domain of p53: correction. Science 1995; 267:1515-6. [PMID: 7878474 DOI: 10.1126/science.7878474] [Citation(s) in RCA: 63] [Impact Index Per Article: 2.1] [Reference Citation Analysis] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/27/2023]
|