Protein Lysine Acetylation by p300/CBPClick to copy article linkArticle link copied!
SPECIAL ISSUE
This article is part of the
1 Introduction
2 Background Information
2.1 Epigenetics and Chromatin
2.2 Post-Translational Modifications of Proteins
Figure 1
Figure 1. Reversible acetylation of the epsilon amine group of lysine side chains.
Figure 2
Figure 2. Major metabolic processes that produce or consume acetyl-CoA. Processes occurring in the cytoplasm are indicated using purple font, and processes occurring in the mitochondrion are indicated using orange font. Note that PDH can also be nuclear. This figure was adapted in part from Albaugh et al. (30)
2.3 Case Study: Histone H4 Tail
2.4 Discovery of p300/CBP
Figure 3
Figure 3. Domain structure of p300/CBP. Exon–intron gene diagrams are shown for p300 and CBP (top). Below are example protein structures for the bromodomain (PDB 3I3J, 2.33 Å), catalytic HAT domain (PDB 3BIY, 1.7 Å), ZZ zinc finger (PDB 1TOT), and TAZ2 domain (PDB 3IO2, 2.5 Å). All structures were produced using purified p300, except the ZZ zinc finger, which used purified CBP. p300/CBP proteins are colored with a rainbow, with blue at the N-terminus and red at the C-terminus, and residues included in the structure are listed below each. Zinc ions are black spheres. All structures are based on X-ray crystallography, except the ZZ zinc finger structures from solution NMR. The p300 bromodomain structure shown here is remarkably similar to an independently generated CBP bromodomain structure (not shown, PDB 3DWY, 1.98 Å). Below is a model for full-length p300/CBP with all domains shown, and is a compilation based on several recent analyses.: (94, 95) three cysteine/histidine-rich (C/H) domains are shown in turquoise, three zinc fingers are shown in yellow, and the catalytic acetyltransferase domain is shown in orange, with its autoacetylated regulatory loop drawn above, which corresponds to residues 1523–1554. A few examples of proteins that bind p300/CBP are listed below the protein model, with the particular domain involved in binding indicated with a black line. Below that, amino acid similarity is indicated, for comparing p300 and CBP sequences within either the catalytic BHC region (from the bromodomain to the C/H3 domain) or the entire protein. At the bottom, commonly purified active p300 variants are indicated, including p300 acetyltransferase/HAT domain, BHC enzyme (bromodomain-HAT-C/H3), and full-length protein. It should be noted that p300 HAT has a deletion in residues 1529–1560.
2.5 p300/CBP Function in Signaling Pathways
Figure 4
Figure 4. p300/CBP is central to many important signaling pathways. These include pathways that respond to intracellular signals (turquoise), extracellular signals (purple), and intercellular signals (blue). These pathways control the key cellular functions via altering expression of target genes, through the action of p300/CBP in the nucleus.
Figure 5
Figure 5. p300/CBP functions as a scaffold, bridge, and acetyltransferase. The acetyltransferase reactions are illustrated by turquoise arrows, indicating acetylation of histone and nonhistone substrates (in yellow), as well as autoacetylation of the p300/CBP acetyltransferase domain. The bridge function is illustrated by turquoise squares, representing DNA-binding proteins that bring DNA elements into proximity with p300/CBP through their interactions. The scaffold function is illustrated by orange squares, representing a protein complex being recruited by p300/CBP. These functions together allow for gene expression.
3 p300 Acetyltransferase Activity
3.1 p300 Acetyltransferase Domain Structure
Figure 6
Figure 6. p300 acetyltransferase domain structure bound to Lys-CoA. (A) Secondary structures of p300 acetyltransferase domain. (B) L1 loop and an acidic surface. (C) Parts of Lys-CoA bisubstrate analog (gray) and four p300 residues of interest (green). Generated in PyMol based on Protein Databank entry 3BIY, published by Liu et al. (143)
3.2 p300 Acetyltransferase Reaction Chemistry
Figure 7
Figure 7. Acetyl transfer catalysis by p300. (A) The p300 active site is drawn in green, and histone H4 substrate in blue, with important residues indicated. CoA is drawn in black, and binds in a specific tunnel. (B) Four steps in a proposed p300 mechanism. acetyl-CoA binds, then peptidyl-lysine binds. The hydrophobic indole of W1436 promotes an uncharged lysine and positions it for attack. The lysine attacks the carbonyl of acetyl-CoA, while Y1467 acts as a general acid to protonate the leaving group. Acetyl-lysine-containing product leaves quickly, then CoASH departs slowly.
4 Inhibitors of p300 Acetyltransferase Activity
4.1 Bisubstrate Acetyltransferase Inhibitors
Figure 8
Figure 8. Bisubstrate inhibitors of acetyltransferases.
4.2 Natural Product Acetyltransferase Inhibitors
Figure 9
Figure 9. Natural products implicated as modulators of acetyltransferases.
4.3 Synthetic Acetyltransferase Inhibitors
Figure 10
Figure 10. Synthetic inhibitors of acetyltransferases.
Figure 11
Figure 11. C646 modeled in the acetyltransferase active site. (A) C646 is shown in magenta, computationally docked in the crystal structure of the acetyltransferase active site, which was generated as a cocrystal with Lys-CoA. Several residues that coordinate CoA binding are predicted to similarly coordinate C646 binding, as shown in aqua stick representations of the side chains. (B) The structure of C646, shown in an orientation similar to that in the docked model above.
5 Protein Ligands of p300/CBP
5.1 Structures of p300/CBP with Bound Protein Ligands
Figure 12
Figure 12. Structures of CBP bromodomain bound to ligands. The purified CBP bromodomain (residues 1081–1197, shown in a rainbow blue to red) is shown bound to (A) histone H4 residues 14–28 acetylated at K20 (PDB 2RNY); (B) p53 residues 367–386 acetylated at K382 (PDB 1JSP); (C) the compound ischemin (PDB 2L84); and (D) the compound dimethylisoxazole (PDB 3SVH). Peptide ligands are shown in gray (A,B) or stick models colored by atom (C,D and acetyl-lysines in A,B). All structures are based on solution NMR except for (D), which is from X-ray crystallography (1.8 Å).
Figure 13
Figure 13. Structures of p300/CBP TAZ2 domain bound to ligands. The TAZ2 domain is shown colored in a rainbow (blue to red, residues included listed below each) bound to various ligands: STAT1 (A, PDB 2KA6); E1A (B, PDB 2KJE); p53 (C, PDB 2K8F); MEF2-DNA complex (D, PDB 3P57), and C/EBPε (PDB 3T92). All structures are based on solution NMR except for two from X-ray crystallography: that in (D) (2.192 Å) and that in (E) (1.5 Å). All structures were produced using purified p300, except the (A) and (B), which used purified CBP. Zinc ions are black spheres, protein ligands are gray, and DNA is yellow. The crystal structure with MEF2 revealed binding in three possible conformations with TAZ2, and one example is shown here.
5.2 Flanking Domains on p300/CBP Acetyltransferase Activity
Figure 14
Figure 14. Models for targeting influenced by p300-ligand binding. In these models, p300 is shown in green, the histone octamer is shown in yellow, DNA is shown with a red strand, and p300 ligands are indicated with an “L”. In (A), a ligand targets p300 to a gene or other DNA element due to the DNA binding affinity of the ligand. In (B), a ligand targets p300 to a protein complex due to the protein binding affinity of the ligand. In (C), two ligands bridged by p300 allow for chromatin (bound by the purple ligand) to come into proximity with a chromatin-modifying enzyme (the orange ligand). In (D), two ligands compete for the same site within p300, and the one ligand could be seen as a competitive inhibitor for the p300 association with the other.
Figure 15
Figure 15. Mastermind-Notch-CSL-DNA core complex. The complex formed by DNA, CSL, Notch (ANK repeats and RAM region purified separately), and Mastermind N-terminal helix is shown with two different view angles. The X-ray crystal structure was generated at 3.85 Å, and this figure was produced in PyMol using Protein Databank entry 3V79. Proteins are shown as ribbons, with the surfaces at 70% transparency. DNA is shown as stick models colored by atom.
Figure 16
Figure 16. Model for Mastermind activation of p300. In this model, p300 (green) initially has inhibited acetyltransferase activity due to an autoinhibitory loop (orange, left). This is relieved upon recruitment by the Notch-Mastermind-CSL complex. Mastermind (purple) binds to the p300 C/H3 domain, and also to Notch intracellular domain (magenta) and CSL (red). p300 autoacetylation, Mastermind acetylation, and histone acetylation are then catalyzed by p300 (turquoise ▲).
5.3 Compounds That Block or Mimic p300/CBP Ligand Binding
Figure 17
Figure 17. Inhibitors of the p300/CBP bromodomain.
Figure 18
Figure 18. Inhibitors of the other p300/CBP domains.
6 Potential Applications for Inhibitors and Activators of p300/CBP
6.1 Loss of p300/CBP Activity in Human Disease
6.2 Overexpression, Inappropriate Activation, or Mistargeting of p300/CBP
Figure 19
Figure 19. Diseases of potential therapeutic application for a p300/CBP inhibitor. No p300/CBP inhibitor has yet made it into clinical trials, but the biology of p300/CBP action and documented effects of p300/CBP disruption lead us to hypothesize a beneficial therapeutic potential for a p300/CBP inhibitor in many diseases.
6.3 Cell-Based Assays of p300/CBP Activity and Inhibition
6.4 Overlap between Genetic and Pharmacologic Disruption of p300/CBP
7 Future Directions
8 Conclusion
Supporting Information
Table of 98 protein acetylation substrates of p300/CBP and table of 411 protein binding partners of p300/CBP. This material is available free of charge via the Internet at http://pubs.acs.org.
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.
Biographies
Beverley M. Dancy
Beverley M. Dancy, originally from Cape Town, was introduced to research in the laboratory of Greg J. Hermann at Lewis & Clark College. Her graduate studies with Philip A. Cole, centered on p300 acetyltransferase inhibition, earned her a Ph.D. in Pharmacology and Molecular Sciences from the Johns Hopkins University School of Medicine in 2013. As a postdoctoral fellow in the Morgan-Sedensky laboratory, she investigated enzymes and pathways involved in mitochondrial disease, and in 2014 she will begin a fellowship at the National Institutes of Health Laboratory of Cardiac Energetics, with Robert S. Balaban. Dr. Dancy is interested in developing molecular tools and imaging technologies for understanding the consequences of disease at the molecular, cellular, and organismal levels. In addition to research, she is passionate about bridging the gap between science research and a broader audience through communication and engagement.
Philip A. Cole
Philip A. Cole was born in Paterson, NJ and graduated from Yale University with a B.S. in Chemistry in 1984 and then spent a year as a Churchill Scholar at the University of Cambridge, England. Cole went on to obtain M.D. and Ph.D. degrees from Johns Hopkins University where he pursued research in bioorganic chemistry in 1991. Cole then entered postdoctoral fellowship training at Harvard Medical School prior to joining the Rockefeller University in 1996 as a junior lab head. In 1999, Cole moved back to Johns Hopkins as the Marshall-Maren professor and director of pharmacology. His research interests are in the area of protein post-translational modifications and chemical biology.
Acknowledgment
We thank the National Institutes of Health (NIH) and Flight Attendant Medical Research Institute (FAMRI) Foundation for funding.
Ac | acetyl |
CBP | CREB (cAMP response element-binding) binding protein |
CoA | coenzyme A |
ED50 | half-maximal effective dose in a cell, tissue, or organism |
HAT | histone acetyltransferase |
HDAC | histone deacetylase |
IC50 | half-maximal inhibitory concentration in solution with a purified enzyme |
K | lysine |
NAD | nicotinamide adenine dinucleotide |
p300 | E1A binding protein p300 |
PTM | post-translational modification |
TSA | trichostatin A |
References
This article references 499 other publications.
- 1Waddington, C. H. Endeavour 1942, 18Google ScholarThere is no corresponding record for this reference.
- 2Klar, A. J. Trends Genet. 1998, 14, 299Google Scholar2Propagating epigenetic states through meiosis: where Mendel's gene is more than a DNA moietyKlar, Amar J. S.Trends in Genetics (1998), 14 (8), 299-301CODEN: TRGEE2; ISSN:0168-9525. (Elsevier Science Ltd.)A review, with 21 refs., on epigenetic inheritance as illustrated by Drosophila and Schizosaccharomyces pombe.
- 3Redon, C.; Pilch, D.; Rogakou, E.; Sedelnikova, O.; Newrock, K.; Bonner, W. Curr. Opin. Genet. Dev. 2002, 12, 162Google ScholarThere is no corresponding record for this reference.
- 4Dawson, M. A.; Kouzarides, T. Cell 2012, 150, 12Google Scholar4Cancer Epigenetics: From Mechanism to TherapyDawson, Mark A.; Kouzarides, TonyCell (Cambridge, MA, United States) (2012), 150 (1), 12-27CODEN: CELLB5; ISSN:0092-8674. (Cell Press)A review. The epigenetic regulation of DNA-templated processes has been intensely studied over the last 15 years. DNA methylation, histone modification, nucleosome remodeling, and RNA-mediated targeting regulate many biol. processes that are fundamental to the genesis of cancer. Here, we present the basic principles behind these epigenetic pathways and highlight the evidence suggesting that their misregulation can culminate in cancer. This information, along with the promising clin. and preclin. results seen with epigenetic drugs against chromatin regulators, signifies that it is time to embrace the central role of epigenetics in cancer.
- 5Mukherjee, S.; Hao, Y. H.; Orth, K. Trends Biochem. Sci. 2007, 32, 210Google ScholarThere is no corresponding record for this reference.
- 6Moellering, R. E.; Cravatt, B. F. Science 2013, 341, 549Google ScholarThere is no corresponding record for this reference.
- 7Cohen, I.; Poręba, E.; Kamieniarz, K.; Schneider, R. Genes Cancer 2011, 2, 631Google Scholar7Histone modifiers in cancer: friends or foes?Cohen, Idan; Poreba, Elzbieta; Kamieniarz, Kinga; Schneider, RobertGenes & Cancer (2011), 2 (6), 631-647CODEN: GCEAAY; ISSN:1947-6019. (Sage Publications)A review. Covalent modifications of histories can regulate all DNA-dependent processes. In the last few years, it has become more and more evident that histone modifications are key players in the regulation of chromatin states and dynamics as well as in gene expression. Therefore, histone modifications and the enzymic machineries that set them are crucial regulators that can control cellular proliferation, differentiation, plasticity, and malignancy processes. This review discusses the biol. and biochem. of covalent histone posttranslational modifications (PTMs) and evaluates the dual role of their modifiers in cancer: as oncogenes that can initiate and amplify tumorigenesis or as tumor suppressors.
- 8Xing, S.; Poirier, Y. Trends Plant Sci. 2012, 17, 423Google Scholar8The protein acetylome and the regulation of metabolismXing, Shufan; Poirier, YvesTrends in Plant Science (2012), 17 (7), 423-430CODEN: TPSCF9; ISSN:1360-1385. (Elsevier Ltd.)A review. Acetyl-CoA (CoA) is a central metabolite involved in numerous anabolic and catabolic pathways, as well as in protein acetylation. Beyond histones, a large no. of metabolic enzymes are acetylated in both animal and bacteria, and the protein acetylome is now emerging in plants. Protein acetylation is influenced by the cellular level of both acetyl-CoA and NAD+, and regulates the activity of several enzymes. Acetyl-CoA is thus ideally placed to act as a key mol. linking the energy balance of the cell to the regulation of gene expression and metabolic pathways via the control of protein acetylation. Better knowledge over how to influence acetyl-CoA levels and the acetylation process promises to be an invaluable tool to control metabolic pathways.
- 9Daugherty, M.; Polanuyer, B.; Farrell, M.; Scholle, M.; Lykidis, A.; de Crécy-Lagard, V.; Osterman, A. J. Biol. Chem. 2002, 277, 21431Google ScholarThere is no corresponding record for this reference.
- 10Robinson, P. J.; An, W.; Routh, A.; Martino, F.; Chapman, L.; Roeder, R. G.; Rhodes, D. J. Mol. Biol. 2008, 381, 816Google Scholar1030 nm Chromatin Fibre Decompaction Requires both H4-K16 Acetylation and Linker Histone EvictionRobinson, Philip J. J.; An, Woojin; Routh, Andrew; Martino, Fabrizio; Chapman, Lynda; Roeder, Robert G.; Rhodes, DanielaJournal of Molecular Biology (2008), 381 (4), 816-825CODEN: JMOBAK; ISSN:0022-2836. (Elsevier Ltd.)The mechanism by which chromatin is decondensed to permit access to DNA is largely unknown. Here, using a model nucleosome array reconstituted from recombinant histone octamers, we have defined the relative contribution of the individual histone octamer N-terminal tails as well as the effect of a targeted histone tail acetylation on the compaction state of the 30 nm chromatin fiber. This study goes beyond previous studies as it is based on a nucleosome array that is very long (61 nucleosomes) and contains a stoichiometric concn. of bound linker histone, which is essential for the formation of the 30 nm chromatin fiber. We find that compaction is regulated in two steps: Introduction of H4 acetylated to 30% on K16 inhibits compaction to a greater degree than deletion of the H4 N-terminal tail. Further decompaction is achieved by removal of the linker histone.
- 11Neumann, H.; Peak-Chew, S. Y.; Chin, J. W. Nat. Chem. Biol. 2008, 4, 232Google Scholar11Genetically encoding Nε-acetyllysine in recombinant proteinsNeumann, Heinz; Peak-Chew, Sew Y.; Chin, Jason W.Nature Chemical Biology (2008), 4 (4), 232-234CODEN: NCBABT; ISSN:1552-4450. (Nature Publishing Group)Nε-acetylation of lysine is a reversible post-translational modification with a regulatory role that rivals that of phosphorylation in eukaryotes. No general methods exist to synthesize proteins contg. Nε-acetyllysine at defined sites. The site-specific incorporation of Nε-acetyllysine in recombinant proteins produced in Escherichia coli was achieved via the evolution of an orthogonal Nε-acetyllysyl-tRNA synthetase/tRNACUA pair. This strategy should find wide applications in defining the cellular role of this modification.
- 12Guo, J.; Wang, J.; Lee, J. S.; Schultz, P. G. Angew. Chem., Int. Ed. 2008, 47, 6399Google ScholarThere is no corresponding record for this reference.
- 13Muir, T. W.; Sondhi, D.; Cole, P. A. Proc. Natl. Acad. Sci. U.S.A. 1998, 95, 6705Google Scholar13Expressed protein ligation: A general method for protein engineeringMuir, Tom W.; Sondhi, Dolan; Cole, Philip A.Proceedings of the National Academy of Sciences of the United States of America (1998), 95 (12), 6705-6710CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)A protein semisynthesis method-expressed protein ligation-is described that involves the chemoselective addn. of a peptide to a recombinant protein. This method was used to ligate a phosphotyrosine peptide to the C terminus of the protein tyrosine kinase C-terminal Src kinase (Csk). By intercepting a thioester generated in the recombinant protein with an N-terminal cysteine contg. synthetic peptide, near quant. chem. ligation of the peptide to the protein was achieved. The semisynthetic tail-phosphorylated Csk showed evidence of an intramol. phosphotyrosine-Src homol. 2 interaction and an unexpected increase in catalytic phosphoryl transfer efficiency toward a physiol. relevant substrate compared with the non-tail-phosphorylated control. This work illustrates that expressed protein ligation is a simple and powerful new method in protein engineering to introduce sequences of unnatural amino acids, posttranslational modifications, and biophys. probes into proteins of any size.
- 14Karukurichi, K. R.; Wang, L.; Uzasci, L.; Manlandro, C. M.; Wang, Q.; Cole, P. A. J. Am. Chem. Soc. 2010, 132, 1222Google ScholarThere is no corresponding record for this reference.
- 15Huang, R.; Holbert, M. A.; Tarrant, M. K.; Curtet, S.; Colquhoun, D. R.; Dancy, B. M.; Dancy, B. C.; Hwang, Y.; Tang, Y.; Meeth, K.; Marmorstein, R.; Cole, R. N.; Khochbin, S.; Cole, P. A. J. Am. Chem. Soc. 2010, 132, 9986Google ScholarThere is no corresponding record for this reference.
- 16Johns, E. W.; Phillips, D. M.; Simson, P.; Butler, J. A. Biochem. J. 1961, 80, 189Google ScholarThere is no corresponding record for this reference.
- 17Phillips, D. M. Biochem. J. 1963, 87, 258Google ScholarThere is no corresponding record for this reference.
- 18Cohen, I.; Poręba, E.; Kamieniarz, K.; Schneider, R. Genes Cancer 2011, 2, 631Google Scholar18Histone modifiers in cancer: friends or foes?Cohen, Idan; Poreba, Elzbieta; Kamieniarz, Kinga; Schneider, RobertGenes & Cancer (2011), 2 (6), 631-647CODEN: GCEAAY; ISSN:1947-6019. (Sage Publications)A review. Covalent modifications of histories can regulate all DNA-dependent processes. In the last few years, it has become more and more evident that histone modifications are key players in the regulation of chromatin states and dynamics as well as in gene expression. Therefore, histone modifications and the enzymic machineries that set them are crucial regulators that can control cellular proliferation, differentiation, plasticity, and malignancy processes. This review discusses the biol. and biochem. of covalent histone posttranslational modifications (PTMs) and evaluates the dual role of their modifiers in cancer: as oncogenes that can initiate and amplify tumorigenesis or as tumor suppressors.
- 19Close, P.; Creppe, C.; Gillard, M.; Ladang, A.; Chapelle, J. P.; Nguyen, L.; Chariot, A. Cell. Mol. Life Sci. 2010, 67, 1255Google Scholar19The emerging role of lysine acetylation of non-nuclear proteinsClose, Pierre; Creppe, Catherine; Gillard, Magali; Ladang, Aurelie; Chapelle, Jean-Paul; Nguyen, Laurent; Chariot, AlainCellular and Molecular Life Sciences (2010), 67 (8), 1255-1264CODEN: CMLSFI; ISSN:1420-682X. (Birkhaeuser Verlag)A review. The acetylation of protein Lys residues is a post-translational modification that critically regulates gene transcription by targeting histones as well as a variety of transcription factors in the nucleus. More recent reports have also demonstrated that numerous proteins located outside the nucleus are also acetylated and that this modification has profound consequences on their functions. This review describes the latest findings on the protein substrates acetylated outside the nucleus and on the acetylases and deacetylates that catalyze these modifications. Protein acetylation is emerging as a major mechanism by which key proteins are regulated in many physiol. processes such as migration, metab., and aging as well as in pathol. circumstances such as cancer and neurodegenerative disorders.
- 20Glozak, M. A.; Sengupta, N.; Zhang, X.; Seto, E. Gene 2005, 363, 15Google ScholarThere is no corresponding record for this reference.
- 21Norris, K. L.; Lee, J. Y.; Yao, T. P. Sci. Signaling 2009, 2, pe76Google Scholar21Acetylation goes global: the emergence of acetylation biologyNorris Kristi L; Lee Joo-Yong; Yao Tso-PangScience signaling (2009), 2 (97), pe76 ISSN:.For the first 30 years since its discovery, reversible protein acetylation has been studied and understood almost exclusively in the context of histone modification and gene transcription. With the discovery of non-histone acetylated proteins and acetylation-modifying enzymes in cellular compartments outside the nucleus, the regulatory potential of reversible acetylation has slowly been recognized in the last decade. However, the scope of protein acetylation involvement in complex biological processes remains uncertain. The recent development of new technology has enabled, for the first time, the identification and quantification of the acetylome, acetylation events at the whole-proteome level. These efforts have uncovered a stunning complexity of the acetylome that potentially rivals that of the phosphoproteome. The remarkably ubiquitous and conserved nature of protein acetylation revealed by these new studies suggests the regulatory power of this dynamic modification. The establishment of comprehensive acetylomes will change the landscape of protein acetylation, where an exciting research frontier awaits.
- 22Choudhary, C.; Kumar, C.; Gnad, F.; Nielsen, M. L.; Rehman, M.; Walther, T. C.; Olsen, J. V.; Mann, M. Science 2009, 325, 834Google Scholar22Lysine Acetylation Targets Protein Complexes and Co-Regulates Major Cellular FunctionsChoudhary, Chunaram; Kumar, Chanchal; Gnad, Florian; Nielsen, Michael L.; Rehman, Michael; Walther, Tobias C.; Olsen, Jesper V.; Mann, MatthiasScience (Washington, DC, United States) (2009), 325 (5942), 834-840CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)Lysine acetylation is a reversible posttranslational modification of proteins and plays a key role in regulating gene expression. Technol. limitations have so far prevented a global anal. of lysine acetylation's cellular roles. We used high-resoln. mass spectrometry to identify 3600 lysine acetylation sites on 1750 proteins and quantified acetylation changes in response to the deacetylase inhibitors suberoylanilide hydroxamic acid and MS-275. Lysine acetylation preferentially targets large macromol. complexes involved in diverse cellular processes, such as chromatin remodeling, cell cycle, splicing, nuclear transport, and actin nucleation. Acetylation impaired phosphorylation-dependent interactions of 14-3-3 and regulated the yeast cyclin-dependent kinase Cdc28. Our data demonstrate that the regulatory scope of lysine acetylation is broad and comparable with that of other major posttranslational modifications.
- 23Chen, Y.; Zhao, W.; Yang, J. S.; Cheng, Z.; Luo, H.; Lu, Z.; Tan, M.; Gu, W.; Zhao, Y. Mol. Cell. Proteomics 2012, 11, 1048Google Scholar23Quantitative acetylome analysis reveals the roles of SIRT1 in regulating diverse substrates and cellular pathwaysChen, Yue; Zhao, Wenhui; Yang, Jeong Soo; Cheng, Zhongyi; Luo, Hao; Lu, Zhike; Tan, Minjia; Gu, Wei; Zhao, YingmingMolecular and Cellular Proteomics (2012), 11 (10), 1048-1062CODEN: MCPOBS; ISSN:1535-9484. (American Society for Biochemistry and Molecular Biology)Despite the progress in identifying many Lys acetylation (Kac) proteins, Kac substrates for Kac-regulatory enzymes remain largely unknown, presenting a major knowledge gap in Kac biol. Here, we identified and quantified 4623 Kac sites in 1800 Kac proteins in SIRT1+/+ and SIRT1-/- MEF cells, representing the first study to reveal an enzyme-regulated Kac subproteome and the largest Lys acetylome reported to date from a single study. Four hundred eighty-five Kac sites were enhanced by more than 100% after SIRT1 knockout. Our results indicate that SIRT1 regulates the Kac states of diverse cellular pathways. Interestingly, we found that a no. of acetyltransferases and major acetyltransferase complexes are targeted by SIRT1. Moreover, we showed that the activities of the acetyltransferases are regulated by SIRT1-mediated deacetylation. Taken together, our results reveal the Lys acetylome in response to SIRT1, provide new insights into mechanisms of SIRT1 function, and offer biomarker candidates for the clin. evaluation of SIRT1-activator compds.
- 24Zhao, S.; Xu, W.; Jiang, W.; Yu, W.; Lin, Y.; Zhang, T.; Yao, J.; Zhou, L.; Zeng, Y.; Li, H.; Li, Y.; Shi, J.; An, W.; Hancock, S. M.; He, F.; Qin, L.; Chin, J.; Yang, P.; Chen, X.; Lei, Q.; Xiong, Y.; Guan, K. L. Science 2010, 327, 1000Google ScholarThere is no corresponding record for this reference.
- 25Albaugh, B. N.; Arnold, K. M.; Denu, J. M. ChemBioChem 2011, 12, 290Google ScholarThere is no corresponding record for this reference.
- 26Sutendra, G.; Kinnaird, A.; Dromparis, P.; Paulin, R.; Stenson, T. H.; Haromy, A.; Hashimoto, K.; Zhang, N.; Flaim, E.; Michelakis, E. D. Cell 2014, 158, 84Google ScholarThere is no corresponding record for this reference.
- 27Begley, T. P.; Kinsland, C.; Strauss, E. Vitam. Horm. 2001, 61, 157Google Scholar27The biosynthesis of coenzyme A in bacteriaBegley, Tadhg P.; Kinsland, Cynthia; Strauss, ErickVitamins and Hormones (San Diego, CA, United States) (2001), 61 (), 157-171CODEN: VIHOAQ; ISSN:0083-6729. (Academic Press)A review with 50 refs. CoA (I) and enzyme-bound phosphopantetheine (II) function as acyl carriers and as carbonyl activating groups for Claisen reactions as well as for amide-, ester-, and thioester-forming reactions in the cell. These cofactors play a key role in the biosynthesis and breakdown of fatty acids and in the biosynthesis of polyketides and nonribosomal peptides. CoA is biosynthesized in bacteria in nine steps. The biosynthesis begins with the decarboxylation of aspartate to give β-alanine. Pantoic acid is formed by the hydroxymethylation of α-ketoisovalerate followed by redn. These intermediates are then condensed to give pantothenic acid. Phosphorylation of pantothenic acid followed by condensation with cysteine and decarboxylation gives 4'-phosphopantetheine. Adenylation and phosphorylation of 4'-phosphopantetheine completes the biosynthesis of CoA. This review will focus on the mechanistic enzymol. of CoA biosynthesis in bacteria. (c) 2001 Academic Press.
- 28Xing, S.; Poirier, Y. Trends Plant Sci. 2012, 17, 423Google Scholar28The protein acetylome and the regulation of metabolismXing, Shufan; Poirier, YvesTrends in Plant Science (2012), 17 (7), 423-430CODEN: TPSCF9; ISSN:1360-1385. (Elsevier Ltd.)A review. Acetyl-CoA (CoA) is a central metabolite involved in numerous anabolic and catabolic pathways, as well as in protein acetylation. Beyond histones, a large no. of metabolic enzymes are acetylated in both animal and bacteria, and the protein acetylome is now emerging in plants. Protein acetylation is influenced by the cellular level of both acetyl-CoA and NAD+, and regulates the activity of several enzymes. Acetyl-CoA is thus ideally placed to act as a key mol. linking the energy balance of the cell to the regulation of gene expression and metabolic pathways via the control of protein acetylation. Better knowledge over how to influence acetyl-CoA levels and the acetylation process promises to be an invaluable tool to control metabolic pathways.
- 29Wellen, K. E.; Thompson, C. B. Nat. Rev. Mol. Cell Biol. 2012, 13, 270Google Scholar29A two-way street: reciprocal regulation of metabolism and signallingWellen, Kathryn E.; Thompson, Craig B.Nature Reviews Molecular Cell Biology (2012), 13 (4), 270-276CODEN: NRMCBP; ISSN:1471-0072. (Nature Publishing Group)A review. It is becoming increasingly clear that cellular signaling and metab. are not just sep. entities but rather are tightly linked. Although nutrient metab. is known to be regulated by signal transduction, an emerging paradigm is that signaling and transcriptional networks can be modulated by nutrient-sensitive protein modifications, such as acetylation and glycosylation, which depend on the availability of acetyl-CoA and sugar donors such as UDP-N-acetylglucosamine (UDP-GlcNAc), resp. The integration of metabolic and signaling cues allows cells to modulate activities such as metab., cell survival and proliferation according to their intracellular metabolic resources.
- 30Albaugh, B. N.; Arnold, K. M.; Denu, J. M. ChemBioChem 2011, 12, 290Google ScholarThere is no corresponding record for this reference.
- 31Allfrey, V. G.; Faulkner, R.; Mirsky, A. E. Proc. Natl. Acad. Sci. U.S.A. 1964, 51, 786Google Scholar31Acetylation and methylation of histones and their possible role in the regulation of ribonucleic acid (RNA) synthesisAllfrey, V. G.; Faulkner, R.; Mirsky, A. E.Proceedings of the National Academy of Sciences of the United States of America (1964), 51 (5), 786-94CODEN: PNASA6; ISSN:0027-8424.Isolated calf thymus nuclei was incubated with 14C-labeled Na acetate. Evidence that acetate was incorporated as acetyl groups attached to histones included: (1) following chromatography the label was eluted together with basic proteins; (2) acetyl-labeled histones were sepd. by electrophoresis; (3) acetate-14C was not removed from histones by org. solvents; (4) over 75% of the label remained after histones were treated with trichloroacetic acid. Most extensive acetate uptake occurred with arginine-rich histones. Puromycin did not inhibit the acetylation, indicating histones were modified by acetylation after completion of the polypeptide chain. Acetylation lowered the histone effectiveness as inhibitor of the RNA polymerase reaction, and a dynamic and reversible mechanism was suggested for activation as well as repression of RNA synthesis.
- 32Hebbes, T. R.; Thorne, A. W.; Crane-Robinson, C. EMBO J. 1988, 7, 1395Google ScholarThere is no corresponding record for this reference.
- 33Hebbes, T. R.; Clayton, A. L.; Thorne, A. W.; Crane-Robinson, C. EMBO J. 1994, 13, 1823Google ScholarThere is no corresponding record for this reference.
- 34Braunstein, M.; Rose, A. B.; Holmes, S. G.; Allis, C. D.; Broach, J. R. Genes Dev. 1993, 7, 592Google Scholar34Transcriptional silencing in yeast is associated with reduced nucleosome acetylationBraunstein, Miriam; Rose, Alan B.; Holmes, Scott G.; Allis, C. David; Broach, James R.Genes & Development (1993), 7 (4), 592-604CODEN: GEDEEP; ISSN:0890-9369.Two classes of sequences in the yeast Saccharomyces cerevisiae are subject to transcription silencing: the silent mating-type cassettes and telomers. The silencing of these regions is shown to be strictly assocd. with acetylation of the ε-amino groups of lysines in the N-terminal domains of 3 of the 4 core histones. Both the silent mating-type cassettes and the Y domains of telomeres are packaged in nucleosomes in vivo that are hypoacetylated relative to those packaging active genes. This difference in acetylation is eliminated by genetic inactivation of silencing:. The silent cassettes from sir2, sir3, or sir4 cells show the same level of acetylation as other active genes. The correspondence of silencing and hypoacetylation of the mating-type cassettes is obsd. even for an allele lacking a promoter, indicating that silencing per se, rather than the absence of transcription, is correlated with hypoacetylation. Finally, overexpression of Sir2p, a protein required for transcriptional silencing in yeast, yields substantial histone deacetylation in vivo. These studies fortify the hypothesis that silencing in yeast results from heterochromatin formation and argue that the silencing proteins participate in this formation.
- 35Turner, B. M.; Birley, A. J.; Lavender, J. Cell 1992, 69, 375Google ScholarThere is no corresponding record for this reference.
- 36Allfrey, V. G.; Faulkner, R.; Mirsky, A. E. Proc. Natl. Acad. Sci. U.S.A. 1964, 51, 786Google Scholar36Acetylation and methylation of histones and their possible role in the regulation of ribonucleic acid (RNA) synthesisAllfrey, V. G.; Faulkner, R.; Mirsky, A. E.Proceedings of the National Academy of Sciences of the United States of America (1964), 51 (5), 786-94CODEN: PNASA6; ISSN:0027-8424.Isolated calf thymus nuclei was incubated with 14C-labeled Na acetate. Evidence that acetate was incorporated as acetyl groups attached to histones included: (1) following chromatography the label was eluted together with basic proteins; (2) acetyl-labeled histones were sepd. by electrophoresis; (3) acetate-14C was not removed from histones by org. solvents; (4) over 75% of the label remained after histones were treated with trichloroacetic acid. Most extensive acetate uptake occurred with arginine-rich histones. Puromycin did not inhibit the acetylation, indicating histones were modified by acetylation after completion of the polypeptide chain. Acetylation lowered the histone effectiveness as inhibitor of the RNA polymerase reaction, and a dynamic and reversible mechanism was suggested for activation as well as repression of RNA synthesis.
- 37Mutskov, V.; Gerber, D.; Angelov, D.; Ausio, J.; Workman, J.; Dimitrov, S. Mol. Cell. Biol. 1998, 18, 6293Google Scholar37Persistent interactions of core histone tails with nucleosomal DNA following acetylation and transcription factor bindingMutskov, Vesco; Gerber, Delphine; Angelov, Dimitri; Ausio, Juan; Workman, Jerry; Dimitrov, StefanMolecular and Cellular Biology (1998), 18 (11), 6293-6304CODEN: MCEBD4; ISSN:0270-7306. (American Society for Microbiology)In this study, we examd. the effect of acetylation of the NH2 tails of core histones on their binding to nucleosomal DNA in the absence or presence of bound transcription factors. To do this, we used a novel UV laser-induced protein-DNA crosslinking technique, combined with immunochem. and mol. biol. approaches. Nucleosomes contg. one or five GAL4 binding sites were reconstituted with hypo-acetylated or hyper-acetylated core histones. Within these reconstituted particles, UV laser-induced histone-DNA crosslinking was found to occur only via the non-structured histone tails and thus presented a unique tool for studying histone tail interactions with nucleosomal DNA. Importantly, these studies demonstrated that the NH2 tails were not released from nucleosomal DNA upon histone acetylation, although some weakening of their interactions was obsd. at elevated ionic strengths. Moreover, the binding of up to five GAL4-AH dimers to nucleosomes occupying the central 90 bp occurred without displacement of the histone NH2 tails from DNA. GAL4-AH binding perturbed the interaction of each histone tail with nucleosomal DNA to different degrees. However, in all cases, greater than 50% of the interactions between the histone tails and DNA was retained upon GAL4-AH binding, even if the tails were highly acetylated. These data illustrate an interaction of acetylated or non-acetylated histone tails with DNA that persists in the presence of simultaneously bound transcription factors.
- 38Zeng, L.; Zhou, M. M. FEBS Lett. 2002, 513, 124Google ScholarThere is no corresponding record for this reference.
- 39Taverna, S. D.; Li, H.; Ruthenburg, A. J.; Allis, C. D.; Patel, D. J. Nat. Struct. Mol. Biol. 2007, 14, 1025Google Scholar39How chromatin-binding modules interpret histone modifications: lessons from professional pocket pickersTaverna, Sean D.; Li, Haitao; Ruthenburg, Alexander J.; Allis, C. David; Patel, Dinshaw J.Nature Structural & Molecular Biology (2007), 14 (11), 1025-1040CODEN: NSMBCU; ISSN:1545-9993. (Nature Publishing Group)A review. Histones comprise the major protein component of chromatin, the scaffold in which the eukaryotic genome is packaged, and are subject to many types of post-translational modifications (PTMs), esp. on their flexible tails. These modifications may constitute a 'histone code' and could be used to manage epigenetic information that helps extend the genetic message beyond DNA sequences. This proposed code, read in part by histone PTM-binding 'effector' modules and their assocd. complexes, has been predicted to define unique functional states of chromatin and/or regulate various chromatin-templated processes. A wealth of structural and functional data show how chromatin effector modules target their cognate covalent histone modifications. Here, the authors summarize key features in the mol. recognition of histone PTMs by a diverse family of 'reader pockets' (including bromodomains, Royal superfamily modules, PHD fingers, WD40 protein repeats, 14-3-3 proteins, and BRCT domains), highlighting specific readout mechanisms for individual marks, common themes, and insights into the downstream functional consequences of the interactions. Changes in these interactions may have far-reaching implications for human biol. and disease, notably cancer.
- 40Botuyan, M. V.; Lee, J.; Ward, I. M.; Kim, J. E.; Thompson, J. R.; Chen, J.; Mer, G. Cell 2006, 127, 1361Google ScholarThere is no corresponding record for this reference.
- 41Lee, J.; Thompson, J. R.; Botuyan, M. V.; Mer, G. Nat. Struct. Mol. Biol. 2008, 15, 109Google Scholar41Distinct binding modes specify the recognition of methylated histones H3K4 and H4K20 by JMJD2A-tudorLee, Joseph; Thompson, James R.; Botuyan, Maria Victoria; Mer, GeorgesNature Structural & Molecular Biology (2008), 15 (1), 109-111CODEN: NSMBCU; ISSN:1545-9993. (Nature Publishing Group)The lysine demethylase JMJD2A has the unique property of binding trimethylated peptides from two different histone sequences (H3K4me3 and H4K20me3) through its tudor domains. Here we show using X-ray crystallog. and calorimetry that H3K4me3 and H4K20me3, which are recognized with similar affinities by JMJD2A, adopt radically different binding modes, to the extent that we were able to design single point mutations in JMJD2A that inhibited the recognition of H3K4me3 but not H4K20me3 and vice versa.
- 42Guo, Y.; Nady, N.; Qi, C.; Allali-Hassani, A.; Zhu, H.; Pan, P.; Adams-Cioaba, M. A.; Amaya, M. F.; Dong, A.; Vedadi, M.; Schapira, M.; Read, R. J.; Arrowsmith, C. H.; Min, J. Nucleic Acids Res. 2009, 37, 2204Google ScholarThere is no corresponding record for this reference.
- 43Li, H.; Fischle, W.; Wang, W.; Duncan, E. M.; Liang, L.; Murakami-Ishibe, S.; Allis, C. D.; Patel, D. J. Mol. Cell 2007, 28, 677Google ScholarThere is no corresponding record for this reference.
- 44Li, H.; Ilin, S.; Wang, W.; Duncan, E. M.; Wysocka, J.; Allis, C. D.; Patel, D. J. Nature 2006, 442, 91Google ScholarThere is no corresponding record for this reference.
- 45Jenuwein, T.; Allis, C. D. Science 2001, 293, 1074Google Scholar45Translating the histone codeJenuwein, Thomas; Allis, C. DavidScience (Washington, DC, United States) (2001), 293 (5532), 1074-1080CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)The review. Chromatin, the physiol. template of all eukaryotic genetic information, is subject to a diverse array of posttranslational modifications that largely impinge on histone amino termini, thereby regulating access to the underlying DNA. Distinct histone amino-terminal modifications can generate synergistic or antagonistic interaction affinities for chromatin-assocd. proteins, which in turn dictate dynamic transitions between transcriptionally active or transcriptionally silent chromatin states. The combinatorial nature of histone amino-terminal modifications thus reveals a "histone code" that considerably extends the information potential of the genetic code. We propose that this epigenetic marking system represents a fundamental regulatory mechanism that has an impact on most, if not all; chromatin-templated processes, with far-reaching consequences for cell fate decisions and both normal and pathol. development.
- 46Bernstein, E.; Hake, S. B. Biochem. Cell Biol. 2006, 84, 505Google Scholar46The nucleosome: a little variation goes a long wayBernstein, Emily; Hake, Sandra B.Biochemistry and Cell Biology (2006), 84 (4), 505-517CODEN: BCBIEQ; ISSN:0829-8211. (National Research Council of Canada)A review. Changes in the overall structure of chromatin are essential for the proper regulation of cellular processes, including gene activation and silencing, DNA repair, chromosome segregation during mitosis and meiosis, X chromosome inactivation in female mammals, and chromatin compaction during apoptosis. Such alterations of the chromatin template occur through at least 3 interrelated mechanisms: post-translational modifications of histones, ATP-dependent chromatin remodeling, and the incorporation (or replacement) of specialized histone variants into chromatin. Of these mechanisms, the exchange of variants into and out of chromatin is the least well understood. However, the exchange of conventional histones for variant histones has distinct and profound consequences within the cell. This review focuses on the growing no. of mammalian histone variants, their particular biol. functions and unique features, and how they may affect the structure of the nucleosome. The authors propose that a given nucleosome might not consist of heterotypic variants, but rather, that only specific histone variants come together to form a homotypic nucleosome, a hypothesis that the authors refer to as the nucleosome code. Such nucleosomes might in turn participate in marking specific chromatin domains that may contribute to epigenetic inheritance.
- 47Schreiber, S. L.; Bernstein, B. E. Cell 2002, 111, 771Google ScholarThere is no corresponding record for this reference.
- 48Ernst, J.; Kellis, M. Nat. Biotechnol. 2010, 28, 817Google ScholarThere is no corresponding record for this reference.
- 49Jenuwein, T.; Allis, C. D. Science 2001, 293, 1074Google Scholar49Translating the histone codeJenuwein, Thomas; Allis, C. DavidScience (Washington, DC, United States) (2001), 293 (5532), 1074-1080CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)The review. Chromatin, the physiol. template of all eukaryotic genetic information, is subject to a diverse array of posttranslational modifications that largely impinge on histone amino termini, thereby regulating access to the underlying DNA. Distinct histone amino-terminal modifications can generate synergistic or antagonistic interaction affinities for chromatin-assocd. proteins, which in turn dictate dynamic transitions between transcriptionally active or transcriptionally silent chromatin states. The combinatorial nature of histone amino-terminal modifications thus reveals a "histone code" that considerably extends the information potential of the genetic code. We propose that this epigenetic marking system represents a fundamental regulatory mechanism that has an impact on most, if not all; chromatin-templated processes, with far-reaching consequences for cell fate decisions and both normal and pathol. development.
- 50Feng, Y.; Wang, J.; Asher, S.; Hoang, L.; Guardiani, C.; Ivanov, I.; Zheng, Y. G. J. Biol. Chem. 2011, 286, 20323Google ScholarThere is no corresponding record for this reference.
- 51Nathan, D.; Ingvarsdottir, K.; Sterner, D. E.; Bylebyl, G. R.; Dokmanovic, M.; Dorsey, J. A.; Whelan, K. A.; Krsmanovic, M.; Lane, W. S.; Meluh, P. B.; Johnson, E. S.; Berger, S. L. Genes Dev. 2006, 20, 966Google ScholarThere is no corresponding record for this reference.
- 52Messner, S.; Altmeyer, M.; Zhao, H.; Pozivil, A.; Roschitzki, B.; Gehrig, P.; Rutishauser, D.; Huang, D.; Caflisch, A.; Hottiger, M. O. Nucleic Acids Res. 2010, 38, 6350Google ScholarThere is no corresponding record for this reference.
- 53Dhalluin, C.; Carlson, J. E.; Zeng, L.; He, C.; Aggarwal, A. K.; Zhou, M. M. Nature 1999, 399, 491Google ScholarThere is no corresponding record for this reference.
- 54Thorne, A. W.; Kmiciek, D.; Mitchelson, K.; Sautiere, P.; Crane-Robinson, C. Eur. J. Biochem. 1990, 193, 701Google Scholar54Patterns of histone acetylationThorne, Alan W.; Kmiciek, Daniel; Mitchelson, Keith; Sautiere, Pierre; Crane-Robinson, ColynEuropean Journal of Biochemistry (1990), 193 (3), 701-13CODEN: EJBCAI; ISSN:0014-2956.The N-terminal domains of all four core histones are subject to reversible acetylation at certain lysine residues. This modification has been functionally linked to transcription, histone deposition at replication, and to histone removal during spermatogenesis. To increase understanding of the significance of this modification the specificity of site utilization in the monoacetyl, diacetyl, and triacetyl forms of histones H3, H4, and H2B (histone H2A has only a single modification site) was studied using pig thymus and HeLa cells as the source. The HeLa histones were extd. from cells grown both with and without butyrate treatment. For histone H3 there is a fairly strict order of site occupancy: Lys14, followed by Lys23, followed by Lys18 in both pig and HeLa histones. Since the order and specificity is the same when butyrate is added to the HeLa cell cultures, it is concluded that addn. of the fatty acid does not scramble the specificity of site utilization, but merely allows more of the natural forms of modified histone to accumulate. For histone H4, the monoacetyl form is exclusively modified at Lys16, but further addn. of acetyl groups is less specific and progresses through sites 12, 8 and 5 in an N-terminal direction. Similar results were obtained for H4 from both pig thymus and butyrate-treated HeLa cells. Histone H2B could be studied in detail only from butyrate-treated HeLa cells and a much lower level of site specificity was found: sites 12 and 15 were preferred to the more N- and C-terminal sites at Lys5 and Lys20. The data reinforces the view that lysine acetylation in core histones is a very specific phenomenon that plays several functionally distinct roles. The high degree of site specificity makes it unlikely that the structural effects of acetylation are mediated merely by a generalized redn. of charge in the histone N-terminal domains.
- 55Zhang, K.; Williams, K. E.; Huang, L.; Yau, P.; Siino, J. S.; Bradbury, E. M.; Jones, P. R.; Minch, M. J.; Burlingame, A. L. Mol. Cell. Proteomics 2002, 1, 500Google Scholar55Histone acetylation and deacetylation: identification of acetylation and methylation sites of HeLa histone H4 by mass spectrometryZhang, Kangling; Williams, Katherine E.; Huang, Lan; Yau, Peter; Siino, Joseph S.; Bradbury, E. Morton; Jones, Patrick R.; Minch, Michael J.; Burlingame, Alma L.Molecular and Cellular Proteomics (2002), 1 (7), 500-508CODEN: MCPOBS; ISSN:1535-9476. (American Society for Biochemistry and Molecular Biology, Inc.)The acetylation isoforms of histone H4 from butyrate-treated HeLa cells were sepd. by C4 reverse-phase high pressure liq. chromatog. and by PAGE. Histone H4 bands were excised and digested in-gel with the endoprotease trypsin. Matrix-assisted laser desorption ionization time-of-flight mass spectrometry was used to characterize the level of acetylation, and nanoelectrospray tandem mass spectrometric anal. of the acetylated peptides was used to det. the exact sites of acetylation. Although there are 15 acetylation sites possible, only four acetylated peptide sequences were actually obsd. The tetra-acetylated form is modified at lysines 5, 8, 12, and 16, the tri-acetylated form is modified at lysines 8, 12, and 16, and the di-acetylated form is modified at lysines 12 and 16. The only significant amt. of the mono-acetylated form was found at position 16. These results are consistent with the hypothesis of a "zip" model whereby acetylation of histone H4 proceeds in the direction of from Lys-16 to Lys-5, and deacetylation proceeds in the reverse direction. Histone acetylation and deacetylation are coordinated processes leading to a non-random distribution of isoforms. Our results also revealed that lysine 20 is dimethylated in all modified isoforms, as well as the nonacetylated isoform of H4.
- 56Clayton, A. L.; Hazzalin, C. A.; Mahadevan, L. C. Mol. Cell 2006, 23, 289Google ScholarThere is no corresponding record for this reference.
- 57Waterborg, J. H. Biochem. Cell Biol. 2002, 80, 363Google Scholar57Dynamics of histone acetylation in vivo. A function for acetylation turnover?Waterborg, Jakob H.Biochemistry and Cell Biology (2002), 80 (3), 363-378CODEN: BCBIEQ; ISSN:0829-8211. (National Research Council of Canada)A review. Histone acetylation discovered more than 40 yr ago is a reversible modification of lysines within the amino-terminal domain of core histones. Amino-terminal histone domains contribute to the compaction of genes into repressed chromatin fibers. It is thought that their acetylation causes localized relaxation of chromatin as a necessary but not sufficient condition for processes that repackage DNA such as transcription, replication, repair, recombination, and sperm formation. While increased histone acetylation enhances gene transcription and loss of acetylation represses and silences genes, the function of the rapid continuous or repetitive acetylation and deacetylation reactions with half-lives of just a few minutes remains unknown. Thirty years of in vivo measurements of acetylation turnover and rates of change in histone modification levels have been reviewed to identify common chromatin characteristics measured by distinct protocols. It has now become possible to look across a wider spectrum of organisms than ever before and identify common features. The rapid turnover rates in transcriptionally active and competent chromatin are one such feature. While ubiquitously obsd., we still do not know whether turnover itself is linked to chromatin transcription beyond its contribution to rapid changes towards hyper- or hypoacetylation of nucleosomes. However, recent expts. suggest that turnover may be linked directly to steps in gene transcription, interacting with nucleosome remodeling complexes.
- 58Katan-Khaykovich, Y.; Struhl, K. Genes Dev. 2002, 16, 743Google ScholarThere is no corresponding record for this reference.
- 59Annunziato, A. T.; Eason, M. B.; Perry, C. A. Biochemistry 1995, 34, 2916Google ScholarThere is no corresponding record for this reference.
- 60Hendzel, M. J.; Davie, J. R. Biochem. J. 1991, 273, 753Google ScholarThere is no corresponding record for this reference.
- 61Hazzalin, C. A.; Mahadevan, L. C. PLoS Biol. 2005, 3, e393Google ScholarThere is no corresponding record for this reference.
- 62Crump, N. T.; Hazzalin, C. A.; Bowers, E. M.; Alani, R. M.; Cole, P. A.; Mahadevan, L. C. Proc. Natl. Acad. Sci. U.S.A. 2011, 108, 7814Google ScholarThere is no corresponding record for this reference.
- 63Taunton, J.; Hassig, C. A.; Schreiber, S. L. Science 1996, 272, 408Google Scholar63A mammalian histone deacetylase related to the yeast transcriptional regulator Rpd3pTaunton, Jack; Hassig, Christian A.; Schreiber, Stuart L.Science (Washington, D. C.) (1996), 272 (5260), 408-11CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)Trapoxin is a microbially derived cyclotetrapeptide that inhibits histone deacetylation in vivo and causes mammalian cells to arrest in the cell cycle. A trapoxin affinity matrix was used to isolate two nuclear proteins that copurified with histone deacetylase activity. Both proteins were identified by peptide microsequencing, and a complementary DNA encoding the histone deacetylase catalytic subunit (HD1) was cloned from a human Jurkat T cell library. As the predicted protein is very similar to the yeast transcriptional regulator Rpd3p, these results support a role for histone deacetylase as a key regulator of eukaryotic transcription.
- 64Schiltz, R. L.; Mizzen, C. A.; Vassilev, A.; Cook, R. G.; Allis, C. D.; Nakatani, Y. J. Biol. Chem. 1999, 274, 1189Google ScholarThere is no corresponding record for this reference.
- 65Kimura, A.; Horikoshi, M. Genes Cells 1998, 3, 789Google Scholar65Tip60 acetylates six lysines of a specific class in core histones in vitroKimura, Akatsuki; Horikoshi, MasamiGenes to Cells (1998), 3 (12), 789-800CODEN: GECEFL; ISSN:1356-9597. (Blackwell Science Ltd.)Tip60, an HIV-1-Tat interactive protein, is a nuclear histone acetyltransferase (HAT) with unique histone substrate specificity. Since the acetylation of core histones at particular lysines mediates distinct effects on chromatin assembly and gene regulation, the identification of Lys site specificity of the HAT activity of Tip60 is an initial step in the anal. of its mol. function. Tip60 significantly acetylates N-terminal tail peptides of histones H2A, H3 and H4, but not H2B, consistent with substrate preference on intact histones. Preferred acetylation sites for Tip60 are Lys-5 of histone H2A, Lys-14 of histone H3, and Lys-5, Lys-8, Lys-12, Lys-16 of histone H4, detd. by a method which combined matrix-assisted laser desorption/ionization mass spectrometry (MALDI-MS) measurements and Lys-C endopeptidase digestion, or a method detecting the incorporation of radiolabeled acetate into synthetic peptides. The Lys site specificity of Tip60 in histone N-terminal tail peptides in vitro was characterized by an assay measuring the mol. wt. of endopeptidase-digested peptides, or a previously described assay. These results agree well with the authors' proposed classification of Lys residues in core histones. The classification may be useful for an anal. of the relations between HATs and the substrates of other uncharacterized HATs.
- 66Thompson, P. R.; Kurooka, H.; Nakatani, Y.; Cole, P. A. J. Biol. Chem. 2001, 276, 33721Google ScholarThere is no corresponding record for this reference.
- 67Ogryzko, V. V.; Schiltz, R. L.; Russanova, V.; Howard, B. H.; Nakatani, Y. Cell 1996, 87, 953Google ScholarThere is no corresponding record for this reference.
- 68Schiltz, R. L.; Mizzen, C. A.; Vassilev, A.; Cook, R. G.; Allis, C. D.; Nakatani, Y. J. Biol. Chem. 1999, 274, 1189Google ScholarThere is no corresponding record for this reference.
- 69McManus, K. J.; Hendzel, M. J. Mol. Cell. Biol. 2003, 23, 7611Google Scholar69Quantitative analysis of CBP- and P300-induced histone acetylations in vivo using native chromatinMcManus, Kirk J.; Hendzel, Michael J.Molecular and Cellular Biology (2003), 23 (21), 7611-7627CODEN: MCEBD4; ISSN:0270-7306. (American Society for Microbiology)In vivo, histone tails are involved in numerous interactions, including those with DNA, adjacent histones, and other, nonhistone proteins. The N-termini are also the substrates for a no. of enzymes, including histone acetyltransferases (HATs), histone deacetylases, and histone methyltransferases. Traditional biochem. approaches defining the substrate specificity profiles of HATs have been performed using purified histone tails, recombinant histones, or purified mononucleosomes as substrates. It is clear that the in vivo presentation of the substrate cannot be accurately represented by using these in vitro approaches. Because of the difficulty in translating in vitro results into in vivo situations, the authors developed a novel single-cell HAT assay that provides quant. measurements of endogenous HAT activity. The HAT assay is performed under in vivo conditions by using the native chromatin structure as the physiol. substrate. The assay combines the spatial resolving power of laser scanning confocal microscopy with simple statistical analyses to characterize CREB binding protein (CBP)- and P300-induced changes in global histone acetylation levels at specific lysine residues. Here the authors show that CBP and P300 exhibit unique substrate specificity profiles, consistent with the developmental and functional differences between the two HATs.
- 70Kasper, L. H.; Lerach, S.; Wang, J.; Wu, S.; Jeevan, T.; Brindle, P. K. EMBO J. 2010, 29, 3660Google ScholarThere is no corresponding record for this reference.
- 71Dawson, M. A.; Kouzarides, T. Cell 2012, 150, 12Google Scholar71Cancer Epigenetics: From Mechanism to TherapyDawson, Mark A.; Kouzarides, TonyCell (Cambridge, MA, United States) (2012), 150 (1), 12-27CODEN: CELLB5; ISSN:0092-8674. (Cell Press)A review. The epigenetic regulation of DNA-templated processes has been intensely studied over the last 15 years. DNA methylation, histone modification, nucleosome remodeling, and RNA-mediated targeting regulate many biol. processes that are fundamental to the genesis of cancer. Here, we present the basic principles behind these epigenetic pathways and highlight the evidence suggesting that their misregulation can culminate in cancer. This information, along with the promising clin. and preclin. results seen with epigenetic drugs against chromatin regulators, signifies that it is time to embrace the central role of epigenetics in cancer.
- 72Yee, S. P.; Branton, P. E. Virology 1985, 147, 142Google ScholarThere is no corresponding record for this reference.
- 73Whyte, P.; Williamson, N. M.; Harlow, E. Cell 1989, 56, 67Google ScholarThere is no corresponding record for this reference.
- 74Eckner, R.; Ewen, M. E.; Newsome, D.; Gerdes, M.; DeCaprio, J. A.; Lawrence, J. B.; Livingston, D. M. Genes Dev. 1994, 8, 869Google ScholarThere is no corresponding record for this reference.
- 75Chrivia, J. C.; Kwok, R. P.; Lamb, N.; Hagiwara, M.; Montminy, M. R.; Goodman, R. H. Nature 1993, 365, 855Google ScholarThere is no corresponding record for this reference.
- 76Arany, Z.; Sellers, W. R.; Livingston, D. M.; Eckner, R. Cell 1994, 77, 799Google ScholarThere is no corresponding record for this reference.
- 77Heery, D. M.; Fischer, P. M. Drug Discovery Today 2007, 12, 88Google ScholarThere is no corresponding record for this reference.
- 78Brownell, J. E.; Zhou, J.; Ranalli, T.; Kobayashi, R.; Edmondson, D. G.; Roth, S. Y.; Allis, C. D. Cell 1996, 84, 843Google Scholar78Tetrahymena histone acetyltransferase A: a homolog to yeast Gcn5p linking histone acetylation to gene activationBrownell, James E.; Zhou, Jianxin; Ranalli, Tamara; Kobayashi, Ryuji; Edmondson, Diane G.; Roth, Sharon Y.; Allis, C. DavidCell (Cambridge, Massachusetts) (1996), 84 (6), 843-51CODEN: CELLB5; ISSN:0092-8674. (Cell Press)We report the cloning of a transcription-assocd. histone acetyltransferase type A (HAT A). This Tetrahymena enzyme is strikingly homologous to the yeast protein Gcn5, a putative transcriptional adaptor, and we demonstrate that recombinant Gcn5p possesses HAT activity. Both the ciliate enzyme and Gcn5p contain potential active site residues found in other acetyltransferases and a highly conserved bromodomain. The presence of this domain in nuclear A-type HATs, but not in cytoplasmic B-type HATs, suggests a mechanism whereby HAT A is directed to chromatin to facilitate transcriptional activation. These findings shed light on the biochem. function of the evolutionarily conserved Gcn5p-Ada complex, directly linking histone acetylation to gene activation, and indicate that histone acetylation is a targeted phenomenon.
- 79Brownell, J. E.; Allis, C. D. Proc. Natl. Acad. Sci. U.S.A. 1995, 92, 6364Google ScholarThere is no corresponding record for this reference.
- 80Kleff, S.; Andrulis, E. D.; Anderson, C. W.; Sternglanz, R. J. Biol. Chem. 1995, 270, 24674Google ScholarThere is no corresponding record for this reference.
- 81Bannister, A. J.; Kouzarides, T. Nature 1996, 384, 641Google ScholarThere is no corresponding record for this reference.
- 82Ogryzko, V. V.; Schiltz, R. L.; Russanova, V.; Howard, B. H.; Nakatani, Y. Cell 1996, 87, 953Google ScholarThere is no corresponding record for this reference.
- 83Eckner, R. Biol. Chem. 1996, 377, 685Google ScholarThere is no corresponding record for this reference.
- 84Taddei, A.; Gasser, S. M. Genetics 2012, 192, 107Google ScholarThere is no corresponding record for this reference.
- 85Rando, O. J.; Chang, H. Y. Annu. Rev. Biochem. 2009, 78, 245Google Scholar85Genome-wide views of chromatin structureRando, Oliver J.; Chang, Howard Y.Annual Review of Biochemistry (2009), 78 (), 245-271CODEN: ARBOAW; ISSN:0066-4154. (Annual Reviews Inc.)A review. Eukaryotic genomes are packaged into a nucleoprotein complex known as chromatin, which affects most processes that occur on DNA. Along with genetic and biochem. studies of resident chromatin proteins and their modifying enzymes, mapping of chromatin structure in vivo is one of the main pillars in our understanding of how chromatin relates to cellular processes. In this review, we discuss the use of genomic technologies to characterize chromatin structure in vivo, with a focus on data from budding yeast and humans. The picture emerging from these studies is the detailed chromatin structure of a typical gene, where the typical behavior gives insight into the mechanisms and deep rules that establish chromatin structure. Important deviation from the archetype is also obsd., usually as a consequence of unique regulatory mechanisms at special genomic loci. Chromatin structure shows substantial conservation from yeast to humans, but mammalian chromatin has addnl. layers of complexity that likely relate to the requirements of multicellularity such as the need to establish faithful gene regulatory mechanisms for cell differentiation.
- 86Bordoli, L.; Netsch, M.; Lüthi, U.; Lutz, W.; Eckner, R. Nucleic Acids Res. 2001, 29, 589Google ScholarThere is no corresponding record for this reference.
- 87Tang, Y.; Holbert, M. A.; Wurtele, H.; Meeth, K.; Rocha, W.; Gharib, M.; Jiang, E.; Thibault, P.; Verreault, A.; Cole, P. A.; Marmorstein, R. Nat. Struct. Mol. Biol. 2008, 15, 998Google Scholar87Fungal Rtt109 histone acetyltransferase is an unexpected structural homolog of metazoan p300/CBP. [Erratum to document cited in CA149:217962]Tang, Yong; Holbert, Marc A.; Wurtele, Hugo; Meeth, Katrina; Rocha, Walter; Gharib, Marlene; Jiang, Eva; Thibault, Pierre; Verreault, Alain; Cole, Philip A.; Marmorstein, RonenNature Structural & Molecular Biology (2008), 15 (9), 998CODEN: NSMBCU; ISSN:1545-9993. (Nature Publishing Group)On page 738, the author name "Verreault" was mistakenly spelled as "Verrault". The error has been cor. in the HTML and PDF versions of the article.
- 88Bordoli, L.; Netsch, M.; Lüthi, U.; Lutz, W.; Eckner, R. Nucleic Acids Res. 2001, 29, 589Google ScholarThere is no corresponding record for this reference.
- 89Giles, R. H.; Dauwerse, H. G.; van Ommen, G. J.; Breuning, M. H. Am. J. Hum. Genet. 1998, 63, 1240Google Scholar89Do human chromosomal bands 16p13 and 22q11-13 share ancestral origins?Giles R H; Dauwerse H G; van Ommen G J; Breuning M HAmerican journal of human genetics (1998), 63 (4), 1240-2 ISSN:0002-9297.There is no expanded citation for this reference.
- 90Chan, H. M.; La Thangue, N. B. J. Cell Sci. 2001, 114, 2363Google Scholar90p300/CBP proteins: HATs for transcriptional bridges and scaffoldsChan, Ho Man; La Thangue, Nicholas B.Journal of Cell Science (2001), 114 (13), 2363-2373CODEN: JNCSAI; ISSN:0021-9533. (Company of Biologists Ltd.)A review with 139 refs. P300/CBP transcriptional co-activator proteins play a central role in co-ordinating and integrating multiple signal-dependent events with the transcription app., allowing the appropriate level of gene activity to occur in response to diverse physiol. cues that influence, for example, proliferation, differentiation and apoptosis. P300/CBP activity can be under aberrant control in human disease, particularly in cancer, which may inactivate a p300/CBP tumor-suppressor-like activity. The transcription regulating-properties of p300 and CBP appear to be exerted through multiple mechanisms. They act as protein bridges, thereby connecting different sequence-specific transcription factors to the transcription app. Providing a protein scaffold upon which to build a multicomponent transcriptional regulatory complex is likely to be an important feature of p300/CBP control. Another key property is the presence of histone acetyltransferase (HAT) activity, which endows p300/CBP with the capacity to influence chromatin activity by modulating nucleosomal histones. Other proteins, including the p53 tumor suppressor, are targets for acetylation by p300/CBP. With the current intense level of research activity, p300/CBP will continue to be in the limelight and, we can be confident, yield new and important information on fundamental processes involved in transcriptional control.
- 91Ogryzko, V. V.; Schiltz, R. L.; Russanova, V.; Howard, B. H.; Nakatani, Y. Cell 1996, 87, 953Google ScholarThere is no corresponding record for this reference.
- 92Yang, X. J.; Seto, E. Mol. Cell 2008, 31, 449Google ScholarThere is no corresponding record for this reference.
- 93Bedford, D. C.; Brindle, Ph. Aging (N. Y.) 2012, 4, 247Google Scholar93Is histone acetylation the most important physiological function for CBP and p300?Bedford David C; Brindle Paul KAging (2012), 4 (4), 247-55 ISSN:.Protein lysine acetyltransferases (HATs or PATs) acetylate histones and other proteins, and are principally modeled as transcriptional coactivators. CREB binding protein (CBP, CREBBP) and its paralog p300 (EP300) constitute the KAT3 family of HATs in mammals, which has mostly unique sequence identity compared to other HAT families. Although studies in yeast show that many histone mutations cause modest or specific phenotypes, similar studies are impractical in mammals and it remains uncertain if histone acetylation is the primary physiological function for CBP/p300. Nonetheless, CBP and p300 mutations in humans and mice show that these coactivators have important roles in development, physiology, and disease, possibly because CBP and p300 act as network "hubs" with more than 400 described protein interaction partners. Analysis of CBP and p300 mutant mouse fibroblasts reveals CBP/p300 are together chiefly responsible for the global acetylation of histone H3 residues K18 and K27, and contribute to other locus-specific histone acetylation events. CBP/p300 can also be important for transcription, but the recruitment of CBP/p300 and their associated histone acetylation marks do not absolutely correlate with a requirement for gene activation. Rather, it appears that target gene context (e.g. DNA sequence) influences the extent to which CBP and p300 are necessary for transcription.
- 94Yang, X. J.; Seto, E. Mol. Cell 2008, 31, 449Google ScholarThere is no corresponding record for this reference.
- 95Bedford, D. C.; Brindle, P. K. Aging (N. Y.) 2012, 4, 247Google Scholar95Is histone acetylation the most important physiological function for CBP and p300?Bedford David C; Brindle Paul KAging (2012), 4 (4), 247-55 ISSN:.Protein lysine acetyltransferases (HATs or PATs) acetylate histones and other proteins, and are principally modeled as transcriptional coactivators. CREB binding protein (CBP, CREBBP) and its paralog p300 (EP300) constitute the KAT3 family of HATs in mammals, which has mostly unique sequence identity compared to other HAT families. Although studies in yeast show that many histone mutations cause modest or specific phenotypes, similar studies are impractical in mammals and it remains uncertain if histone acetylation is the primary physiological function for CBP/p300. Nonetheless, CBP and p300 mutations in humans and mice show that these coactivators have important roles in development, physiology, and disease, possibly because CBP and p300 act as network "hubs" with more than 400 described protein interaction partners. Analysis of CBP and p300 mutant mouse fibroblasts reveals CBP/p300 are together chiefly responsible for the global acetylation of histone H3 residues K18 and K27, and contribute to other locus-specific histone acetylation events. CBP/p300 can also be important for transcription, but the recruitment of CBP/p300 and their associated histone acetylation marks do not absolutely correlate with a requirement for gene activation. Rather, it appears that target gene context (e.g. DNA sequence) influences the extent to which CBP and p300 are necessary for transcription.
- 96Arany, Z.; Newsome, D.; Oldread, E.; Livingston, D. M.; Eckner, R. Nature 1995, 374, 81Google ScholarThere is no corresponding record for this reference.
- 97Ferreon, J. C.; Martinez-Yamout, M. A.; Dyson, H. J.; Wright, P. E. Proc. Natl. Acad. Sci. U.S.A. 2009, 106, 13260Google ScholarThere is no corresponding record for this reference.
- 98Lundblad, J. R.; Kwok, R. P.; Laurance, M. E.; Harter, M. L.; Goodman, R. H. Nature 1995, 374, 85Google ScholarThere is no corresponding record for this reference.
- 99Radhakrishnan, I.; Pérez-Alvarado, G. C.; Parker, D.; Dyson, H. J.; Montminy, M. R.; Wright, P. E. Cell 1997, 91, 741Google ScholarThere is no corresponding record for this reference.
- 100Kalkhoven, E. Biochem. Pharmacol. 2004, 68, 1145Google ScholarThere is no corresponding record for this reference.
- 101Nam, Y.; Sliz, P.; Song, L.; Aster, J. C.; Blacklow, S. C. Cell 2006, 124, 973Google ScholarThere is no corresponding record for this reference.
- 102Mayr, B.; Montminy, M. Nat. Rev. Mol. Cell Biol. 2001, 2, 599Google Scholar102Transcriptional regulation by the phosphorylation-dependent factor CREBMayr, Bernhard; Montminy, MarcNature Reviews Molecular Cell Biology (2001), 2 (8), 599-609CODEN: NRMCBP; ISSN:1471-0072. (Nature Publishing Group)A review. The transcription factor CREB - for 'cAMP response element-binding protein' - functions in glucose homeostasis, growth-factor-dependent cell survival, and has been implicated in learning and memory. CREB is phosphorylated in response to various signals, but how is specificity achieved in these signaling pathways.
- 103Bourdeau, V.; Deschênes, J.; Métivier, R.; Nagai, Y.; Nguyen, D.; Bretschneider, N.; Gannon, F.; White, J. H.; Mader, S. Mol. Endocrinol. 2004, 18, 1411Google Scholar103Genome-wide identification of high-affinity estrogen response elements in human and mouseBourdeau, Veronique; Deschenes, Julie; Metivier, Raphael; Nagai, Yoshihiko; Nguyen, Denis; Bretschneider, Nancy; Gannon, Frank; White, John H.; Mader, SylvieMolecular Endocrinology (2004), 18 (6), 1411-1427CODEN: MOENEN; ISSN:0888-8809. (Endocrine Society)Although estrogen receptors (ERs) recognize 15-bp palindromic estrogen response elements (EREs) with maximal affinity in vitro, few near-consensus sequences have been characterized in estrogen target genes. Here we report the design of a genome-wide screen for high-affinity EREs and the identification of approx. 70,000 motifs in the human and mouse genomes. EREs are enriched in regions proximal to the transcriptional start sites, and approx. 1% of elements appear conserved in the flanking regions (-10 kb to +5 kb) of orthologous human and mouse genes. Conserved and nonconserved elements were also found, often in multiple occurrences, in more than 230 estrogen-stimulated human genes previously identified from expression studies. In genes contg. known EREs, we also identified addnl. distal elements, sometimes with higher in vitro binding affinity and/or better conservation between the species considered. Chromatin immunopptn. expts. in breast cancer cell lines indicate that most novel elements present in responsive genes bind ERα in vivo, including some EREs located up to approx. 10 kb from transcriptional start sites. Our results demonstrate that near-consensus EREs occur frequently in both genomes and that whereas chromatin structure likely modulates access to binding sites, far upstream elements can be evolutionarily conserved and bind ERs in vivo.
- 104Saito, H.; Posas, F. Genetics 2012, 192, 289Google ScholarThere is no corresponding record for this reference.
- 105Sheikh-Hamad, D.; Di Mari, J.; Suki, W. N.; Safirstein, R.; Watts, B. A., III; Rouse, D. J. Biol. Chem. 1998, 273, 1832Google ScholarThere is no corresponding record for this reference.
- 106Chen, Y. J.; Wang, Y. N.; Chang, W. C. J. Biol. Chem. 2007, 282, 27215Google ScholarThere is no corresponding record for this reference.
- 107Jang, E. R.; Choi, J. D.; Lee, J. S. FEBS Lett. 2011, 585, 47Google ScholarThere is no corresponding record for this reference.
- 108Sakaguchi, K.; Herrera, J. E.; Saito, S.; Miki, T.; Bustin, M.; Vassilev, A.; Anderson, C. W.; Appella, E. Genes Dev. 1998, 12, 2831Google ScholarThere is no corresponding record for this reference.
- 109Iyer, N. G.; Chin, S. F.; Ozdag, H.; Daigo, Y.; Hu, D. E.; Cariati, M.; Brindle, K.; Aparicio, S.; Caldas, C. Proc. Natl. Acad. Sci. U.S.A. 2004, 101, 7386Google ScholarThere is no corresponding record for this reference.
- 110Yuan, L. W.; Gambee, J. E. J. Biol. Chem. 2000, 275, 40946Google ScholarThere is no corresponding record for this reference.
- 111Yang, X. J.; Seto, E. Mol. Cell 2008, 31, 449Google ScholarThere is no corresponding record for this reference.
- 112Legube, G.; Trouche, D. EMBO Rep. 2003, 4, 944Google ScholarThere is no corresponding record for this reference.
- 113Schwartz, C.; Beck, K.; Mink, S.; Schmolke, M.; Budde, B.; Wenning, D.; Klempnauer, K. H. EMBO J. 2003, 22, 882Google ScholarThere is no corresponding record for this reference.
- 114Poizat, C.; Puri, P. L.; Bai, Y.; Kedes, L. Mol. Cell. Biol. 2005, 25, 2673Google Scholar114Phosphorylation-dependent degradation of p300 by doxorubicin-activated p38 mitogen-activated protein kinase in cardiac cellsPoizat, Coralie; Puri, Pier Lorenzo; Bai, Yan; Kedes, LarryMolecular and Cellular Biology (2005), 25 (7), 2673-2687CODEN: MCEBD4; ISSN:0270-7306. (American Society for Microbiology)P300 and CBP are general transcriptional coactivators implicated in different cellular processes, including regulation of the cell cycle, differentiation, tumorigenesis, and apoptosis. Posttranslational modifications such as phosphorylation are predicted to select a specific function of p300/CBP in these processes; however, the identification of the kinases that regulate p300/CBP activity in response to individual stimuli and the physiol. significance of p300 phosphorylation have not been elucidated. Here we demonstrate that the cardiotoxic anticancer agent doxorubicin (adriamycin) induces the phosphorylation of p300 in primary neonatal cardiomyocytes. Hyperphosphorylation precedes the degrdn. of p300 and parallels apoptosis in response to doxorubicin. Doxorubicin-activated p38 kinases α and β assoc. with p300 and are implicated in the phosphorylation-mediated degrdn. of p300, as pharmacol. blockade of p38 prevents p300 degrdn. P38 phosphorylates p300 in vitro at both the N and C termini of the protein, and enforced activation of p38 by the constitutively active form of its upstream kinase (MKK6EE) triggers p300 degrdn. These data support the conclusion that p38 mitogen-activated protein kinase regulates p300 protein stability and function in cardiomyocytes undergoing apoptosis in response to doxorubicin.
- 115Xu, W.; Chen, H.; Du, K.; Asahara, H.; Tini, M.; Emerson, B. M.; Montminy, M.; Evans, R. M. Science 2001, 294, 2507Google ScholarThere is no corresponding record for this reference.
- 116Ceschin, D. G.; Walia, M.; Wenk, S. S.; Duboé, C.; Gaudon, C.; Xiao, Y.; Fauquier, L.; Sankar, M.; Vandel, L.; Gronemeyer, H. Genes Dev. 2011, 25, 1132Google ScholarThere is no corresponding record for this reference.
- 117Ryan, C. M.; Kindle, K. B.; Collins, H. M.; Heery, D. M. Biochem. Biophys. Res. Commun. 2010, 391, 1136Google ScholarThere is no corresponding record for this reference.
- 118Kuo, H. Y.; Chang, C. C.; Jeng, J. C.; Hu, H. M.; Lin, D. Y.; Maul, G. G.; Kwok, R. P.; Shih, H. M. Proc. Natl. Acad. Sci. U.S.A. 2005, 102, 16973Google ScholarThere is no corresponding record for this reference.
- 119Karanam, B.; Jiang, L.; Wang, L.; Kelleher, N. L.; Cole, P. A. J. Biol. Chem. 2006, 281, 40292Google ScholarThere is no corresponding record for this reference.
- 120Karukurichi, K. R.; Wang, L.; Uzasci, L.; Manlandro, C. M.; Wang, Q.; Cole, P. A. J. Am. Chem. Soc. 2010, 132, 1222Google ScholarThere is no corresponding record for this reference.
- 121Thompson, P. R.; Wang, D.; Wang, L.; Fulco, M.; Pediconi, N.; Zhang, D.; An, W.; Ge, Q.; Roeder, R. G.; Wong, J.; Levrero, M.; Sartorelli, V.; Cotter, R. J.; Cole, P. A. Nat. Struct. Mol. Biol. 2004, 11, 308Google Scholar121Regulation of the p300 HAT domain via a novel activation loopThompson, Paul R.; Wang, Dongxia; Wang, Ling; Fulco, Marcella; Pediconi, Natalia; Zhang, Dianzheng; An, Woojin; Ge, Qingyuan; Roeder, Robert G.; Wong, Jiemin; Levrero, Massimo; Sartorelli, Vittorio; Cotter, Robert J.; Cole, Philip A.Nature Structural & Molecular Biology (2004), 11 (4), 308-315CODEN: NSMBCU; ISSN:1545-9993. (Nature Publishing Group)The transcriptional coactivator p300 is a histone acetyltransferase (HAT) whose function is crit. for regulating gene expression in mammalian cells. However, the mol. events that regulate p300 HAT activity are poorly understood. We evaluated autoacetylation of the p300 HAT protein domain to det. its function. Using expressed protein ligation, the p300 HAT protein domain was generated in hypoacetylated form and found to have reduced catalytic activity. This basal catalytic rate was stimulated by autoacetylation of several key lysine sites within an apparent activation loop motif. This post-translational modification and catalytic regulation of p300 HAT activity is conceptually analogous to the activation of most protein kinases by autophosphorylation. We therefore propose that this autoregulatory loop could influence the impact of p300 on a wide variety of signaling and transcriptional events.
- 122Karanam, B.; Wang, L.; Wang, D.; Liu, X.; Marmorstein, R.; Cotter, R.; Cole, P. A. Biochemistry 2007, 46, 8207Google ScholarThere is no corresponding record for this reference.
- 123Kasper, L. H.; Fukuyama, T.; Biesen, M. A.; Boussouar, F.; Tong, C.; de Pauw, A.; Murray, P. J.; van Deursen, J. M.; Brindle, P. K. Mol. Cell. Biol. 2006, 26, 789Google Scholar123Conditional knockout mice reveal distinct functions for the global transcriptional coactivators CBP and p300 in T-cell developmentKasper, Lawryn H.; Fukuyama, Tomofusa; Biesen, Michelle A.; Boussouar, Faycal; Tong, Caili; de Pauw, Antoine; Murray, Peter J.; van Deursen, Jan M. A.; Brindle, Paul K.Molecular and Cellular Biology (2006), 26 (3), 789-809CODEN: MCEBD4; ISSN:0270-7306. (American Society for Microbiology)The global transcriptional coactivators CREB-binding protein (CBP) and the closely related p300 interact with over 312 proteins, making them among the most heavily connected hubs in the known mammalian protein-protein interactome. It is largely uncertain, however, if these interactions are important in specific cell lineages of adult animals, as homozygous null mutations in either CBP or p300 result in early embryonic lethality in mice. Here we describe a Cre/LoxP conditional p300 null allele (p300flox) that allows for the temporal and tissue-specific inactivation of p300. We used mice carrying p300flax and a CBP conditional knockout allele (CBPflax) in conjunction with an Lck-Cre transgene to delete CBP and p300 starting at the CD4- CD8- double-neg. thymocyte stage of T-cell development. Loss of either p300 or CBP led to a decrease in CD4+ CD8+ double-pos. thymocytes, but an increase in the percentage of CD8+ single-pos. thymocytes seen in CBP mutant mice was not obsd. in p300 mutants. T cells completely lacking both CBP and p300 did not develop normally and were nonexistent or very rare in the periphery, however. T cells lacking CBP or p300 had reduced tumor necrosis factor alpha gene expression in response to phorbol ester and ionophore, while signal-responsive gene expression in CBP- or p300-deficient macrophages was largely intact. Thus, CBP and p300 each supply a surprising degree of redundant coactivation capacity in T cells and macrophages, although each gene has also unique properties in thymocyte development.
- 124Personal communication with Paul Brindle, see: http://www.stjude.org/stjude/v/index.jsp?vgnextoid=c30215f204294210VgnVCM1000001e0215acRCRD.Google ScholarThere is no corresponding record for this reference.
- 125Yan, G.; Eller, M. S.; Elm, C.; Larocca, C. A.; Ryu, B.; Panova, I. P.; Dancy, B. M.; Bowers, E. M.; Meyers, D.; Lareau, L.; Cole, P. A.; Taverna, S. D.; Alani, R. M. J. Invest. Dermatol. 2013, 133, 2444Google ScholarThere is no corresponding record for this reference.
- 126Cohen, I.; Poręba, E.; Kamieniarz, K.; Schneider, R. Genes Cancer 2011, 2, 631Google Scholar126Histone modifiers in cancer: friends or foes?Cohen, Idan; Poreba, Elzbieta; Kamieniarz, Kinga; Schneider, RobertGenes & Cancer (2011), 2 (6), 631-647CODEN: GCEAAY; ISSN:1947-6019. (Sage Publications)A review. Covalent modifications of histories can regulate all DNA-dependent processes. In the last few years, it has become more and more evident that histone modifications are key players in the regulation of chromatin states and dynamics as well as in gene expression. Therefore, histone modifications and the enzymic machineries that set them are crucial regulators that can control cellular proliferation, differentiation, plasticity, and malignancy processes. This review discusses the biol. and biochem. of covalent histone posttranslational modifications (PTMs) and evaluates the dual role of their modifiers in cancer: as oncogenes that can initiate and amplify tumorigenesis or as tumor suppressors.
- 127Chan, H. M.; La Thangue, N. B. J. Cell Sci. 2001, 114, 2363Google Scholar127p300/CBP proteins: HATs for transcriptional bridges and scaffoldsChan, Ho Man; La Thangue, Nicholas B.Journal of Cell Science (2001), 114 (13), 2363-2373CODEN: JNCSAI; ISSN:0021-9533. (Company of Biologists Ltd.)A review with 139 refs. P300/CBP transcriptional co-activator proteins play a central role in co-ordinating and integrating multiple signal-dependent events with the transcription app., allowing the appropriate level of gene activity to occur in response to diverse physiol. cues that influence, for example, proliferation, differentiation and apoptosis. P300/CBP activity can be under aberrant control in human disease, particularly in cancer, which may inactivate a p300/CBP tumor-suppressor-like activity. The transcription regulating-properties of p300 and CBP appear to be exerted through multiple mechanisms. They act as protein bridges, thereby connecting different sequence-specific transcription factors to the transcription app. Providing a protein scaffold upon which to build a multicomponent transcriptional regulatory complex is likely to be an important feature of p300/CBP control. Another key property is the presence of histone acetyltransferase (HAT) activity, which endows p300/CBP with the capacity to influence chromatin activity by modulating nucleosomal histones. Other proteins, including the p53 tumor suppressor, are targets for acetylation by p300/CBP. With the current intense level of research activity, p300/CBP will continue to be in the limelight and, we can be confident, yield new and important information on fundamental processes involved in transcriptional control.
- 128Pan, C. Q.; Sudol, M.; Sheetz, M.; Low, B. C. Cell. Signalling 2012, 24, 2143Google Scholar128Modularity and functional plasticity of scaffold proteins as p(l)acemakers in cell signalingPan, Catherine Qiurong; Sudol, Marius; Sheetz, Michael; Low, Boon ChuanCellular Signalling (2012), 24 (11), 2143-2165CODEN: CESIEY; ISSN:0898-6568. (Elsevier Inc.)A review. Cells coordinate and integrate various functional modules that control their dynamics, intracellular trafficking, metab. and gene expression. Such capacity is mediated by specific scaffold proteins that tether multiple components of signaling pathways at plasma membrane, Golgi app., mitochondria, endoplasmic reticulum, nucleus and in more specialized subcellular structures such as focal adhesions, cell-cell junctions, endosomes, vesicles and synapses. Scaffold proteins act as "pacemakers" as well as "placemakers" that regulate the temporal, spatial and kinetic aspects of protein complex assembly by modulating the local concns., proximity, subcellular dispositions and biochem. properties of the target proteins through the intricate use of their modular protein domains. These regulatory mechanisms allow them to gate the specificity, integration and crosstalk of different signaling modules. In addn. to acting as phys. platforms for protein assembly, many professional scaffold proteins can also directly modify the properties of their targets while they themselves can be regulated by post-translational modifications and/or mech. forces. Furthermore, multiple scaffold proteins can form alliances of higher-order regulatory networks. Here, we highlight the emerging themes of scaffold proteins by analyzing their common and distinctive mechanisms of action and regulation, which underlie their functional plasticity in cell signaling. Understanding these mechanisms in the context of space, time and force should have ramifications for human physiol. and for developing new therapeutic approaches to control pathol. states and diseases.
- 129Tanaka, Y.; Naruse, I.; Maekawa, T.; Masuya, H.; Shiroishi, T.; Ishii, S. Proc. Natl. Acad. Sci. U.S.A. 1997, 94, 10215Google ScholarThere is no corresponding record for this reference.
- 130Kamei, Y.; Xu, L.; Heinzel, T.; Torchia, J.; Kurokawa, R.; Gloss, B.; Lin, S. C.; Heyman, R. A.; Rose, D. W.; Glass, C. K.; Rosenfeld, M. G. Cell 1996, 85, 403Google ScholarThere is no corresponding record for this reference.
- 131Hottiger, M. O.; Felzien, L. K.; Nabel, G. J. EMBO J. 1998, 17, 3124Google ScholarThere is no corresponding record for this reference.
- 132Yin, X.; Warner, D. R.; Roberts, E. A.; Pisano, M. M.; Greene, R. M. Biochem. Biophys. Res. Commun. 2005, 329, 1010Google ScholarThere is no corresponding record for this reference.
- 133Avantaggiati, M. L.; Ogryzko, V.; Gardner, K.; Giordano, A.; Levine, A. S.; Kelly, K. Cell 1997, 89, 1175Google ScholarThere is no corresponding record for this reference.
- 134Kamei, Y.; Xu, L.; Heinzel, T.; Torchia, J.; Kurokawa, R.; Gloss, B.; Lin, S. C.; Heyman, R. A.; Rose, D. W.; Glass, C. K.; Rosenfeld, M. G. Cell 1996, 85, 403Google ScholarThere is no corresponding record for this reference.
- 135Goodman, R. H.; Smolik, S. Genes Dev. 2000, 14, 1553Google Scholar135CBP/p300 in cell growth, transformation, and developmentGoodman, Richard H.; Smolik, SarahGenes & Development (2000), 14 (13), 1553-1577CODEN: GEDEEP; ISSN:0890-9369. (Cold Spring Harbor Laboratory Press)A review with several refs. This review discusses the involvement of CBP/p300 in cell growth and transformation, chromosomal translocations, involvement of cellular oncogenes, assocn. with viral oncoproteins, involvement with DNA and RNA tumor viruses, the regulation of CBP/p300 by phosphorylation, and usage in development of model systems.
- 136Lee, C. W.; Sørensen, T. S.; Shikama, N.; La Thangue, N. B. Oncogene 1998, 16, 2695Google ScholarThere is no corresponding record for this reference.
- 137Nakajima, T.; Fukamizu, A.; Takahashi, J.; Gage, F. H.; Fisher, T.; Blenis, J.; Montminy, M. R. Cell 1996, 86, 465Google ScholarThere is no corresponding record for this reference.
- 138Thompson, P. R.; Wang, D.; Wang, L.; Fulco, M.; Pediconi, N.; Zhang, D.; An, W.; Ge, Q.; Roeder, R. G.; Wong, J.; Levrero, M.; Sartorelli, V.; Cotter, R. J.; Cole, P. A. Nat. Struct. Mol. Biol. 2004, 11, 308Google Scholar138Regulation of the p300 HAT domain via a novel activation loopThompson, Paul R.; Wang, Dongxia; Wang, Ling; Fulco, Marcella; Pediconi, Natalia; Zhang, Dianzheng; An, Woojin; Ge, Qingyuan; Roeder, Robert G.; Wong, Jiemin; Levrero, Massimo; Sartorelli, Vittorio; Cotter, Robert J.; Cole, Philip A.Nature Structural & Molecular Biology (2004), 11 (4), 308-315CODEN: NSMBCU; ISSN:1545-9993. (Nature Publishing Group)The transcriptional coactivator p300 is a histone acetyltransferase (HAT) whose function is crit. for regulating gene expression in mammalian cells. However, the mol. events that regulate p300 HAT activity are poorly understood. We evaluated autoacetylation of the p300 HAT protein domain to det. its function. Using expressed protein ligation, the p300 HAT protein domain was generated in hypoacetylated form and found to have reduced catalytic activity. This basal catalytic rate was stimulated by autoacetylation of several key lysine sites within an apparent activation loop motif. This post-translational modification and catalytic regulation of p300 HAT activity is conceptually analogous to the activation of most protein kinases by autophosphorylation. We therefore propose that this autoregulatory loop could influence the impact of p300 on a wide variety of signaling and transcriptional events.
- 139Muir, T. W.; Sondhi, D.; Cole, P. A. Proc. Natl. Acad. Sci. U.S.A. 1998, 95, 6705Google Scholar139Expressed protein ligation: A general method for protein engineeringMuir, Tom W.; Sondhi, Dolan; Cole, Philip A.Proceedings of the National Academy of Sciences of the United States of America (1998), 95 (12), 6705-6710CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)A protein semisynthesis method-expressed protein ligation-is described that involves the chemoselective addn. of a peptide to a recombinant protein. This method was used to ligate a phosphotyrosine peptide to the C terminus of the protein tyrosine kinase C-terminal Src kinase (Csk). By intercepting a thioester generated in the recombinant protein with an N-terminal cysteine contg. synthetic peptide, near quant. chem. ligation of the peptide to the protein was achieved. The semisynthetic tail-phosphorylated Csk showed evidence of an intramol. phosphotyrosine-Src homol. 2 interaction and an unexpected increase in catalytic phosphoryl transfer efficiency toward a physiol. relevant substrate compared with the non-tail-phosphorylated control. This work illustrates that expressed protein ligation is a simple and powerful new method in protein engineering to introduce sequences of unnatural amino acids, posttranslational modifications, and biophys. probes into proteins of any size.
- 140Dancy, B. M.; Crump, N. T.; Peterson, D. J.; Mukherjee, C.; Bowers, E. M.; Ahn, Y. H.; Yoshida, M.; Zhang, J.; Mahadevan, L. C.; Meyers, D. J.; Boeke, J. D.; Cole, P. A. ChemBioChem 2012, 13, 2113Google ScholarThere is no corresponding record for this reference.
- 141Liu, X.; Wang, L.; Zhao, K.; Thompson, P. R.; Hwang, Y.; Marmorstein, R.; Cole, P. A. Nature 2008, 451, 846Google ScholarThere is no corresponding record for this reference.
- 142Hwang, Y.; Thompson, P. R.; Wang, L.; Jiang, L.; Kelleher, N. L.; Cole, P. A. Angew. Chem., Int. Ed. 2007, 46, 7621Google Scholar142A selective chemical probe for coenzyme A-requiring enzymesHwang, Yousang; Thompson, Paul R.; Wang, Ling; Jiang, Lihua; Kelleher, Neil L.; Cole, Philip A.Angewandte Chemie, International Edition (2007), 46 (40), 7621-7624CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A CoA-based affinity probe with a sulfoxycarbamate functionality can selectively identify several acetyltransferases relative to other enzymes and proteins. It leaves behind a desthiobiotin tag that can be used for western blotting and mass spectrometric characterization.
- 143Liu, X.; Wang, L.; Zhao, K.; Thompson, P. R.; Hwang, Y.; Marmorstein, R.; Cole, P. A. Nature 2008, 451, 846Google ScholarThere is no corresponding record for this reference.
- 144Tang, Y.; Holbert, M. A.; Wurtele, H.; Meeth, K.; Rocha, W.; Gharib, M.; Jiang, E.; Thibault, P.; Verreault, A.; Cole, P. A.; Marmorstein, R. Nat. Struct. Mol. Biol. 2008, 15, 998Google Scholar144Fungal Rtt109 histone acetyltransferase is an unexpected structural homolog of metazoan p300/CBP. [Erratum to document cited in CA149:217962]Tang, Yong; Holbert, Marc A.; Wurtele, Hugo; Meeth, Katrina; Rocha, Walter; Gharib, Marlene; Jiang, Eva; Thibault, Pierre; Verreault, Alain; Cole, Philip A.; Marmorstein, RonenNature Structural & Molecular Biology (2008), 15 (9), 998CODEN: NSMBCU; ISSN:1545-9993. (Nature Publishing Group)On page 738, the author name "Verreault" was mistakenly spelled as "Verrault". The error has been cor. in the HTML and PDF versions of the article.
- 145Wang, L.; Tang, Y.; Cole, P. A.; Marmorstein, R. Curr. Opin. Struct. Biol. 2008, 18, 741Google Scholar145Structure and chemistry of the p300/CBP and Rtt109 histone acetyltransferases: implications for histone acetyltransferase evolution and functionWang, Ling; Tang, Yong; Cole, Philip A.; Marmorstein, RonenCurrent Opinion in Structural Biology (2008), 18 (6), 741-747CODEN: COSBEF; ISSN:0959-440X. (Elsevier B.V.)A review. The recent crystal structure and assocd. biochem. studies of the metazoan-specific p300/CBP and fungal-specific Rtt109 histone acetyltransferases (HATs) have provided new insights into the ancestral relation between HATs and their functions. These studies point to a common HAT ancestor that has evolved around a common structural framework to form HATs with divergent catalytic and substrate-binding properties. These studies also point to the importance of regulatory loops within HATs and autoacetylation in HAT function. Implications for future studies are discussed.
- 146Poux, A. N.; Cebrat, M.; Kim, C. M.; Cole, P. A.; Marmorstein, R. Proc. Natl. Acad. Sci. U.S.A. 2009, 99, 14065Google ScholarThere is no corresponding record for this reference.
- 147Yuan, H.; Rossetto, D.; Mellert, H.; Dang, W.; Srinivasan, M.; Johnson, J.; Hodawadekar, S.; Ding, E. C.; Speicher, K.; Abshiru, N.; Perry, R.; Wu, J.; Yang, C.; Zheng, Y. G.; Speicher, D. W.; Thibault, P.; Verreault, A.; Johnson, F. B.; Berger, S. L.; Sternglanz, R.; McMahon, S. B.; Côté, J.; Marmorstein, R. EMBO J. 2012, 31, 58Google ScholarThere is no corresponding record for this reference.
- 148Karukurichi, K. R.; Cole, P. A. Bioorg. Chem. 2011, 39, 42Google Scholar148Probing the reaction coordinate of the p300/CBP histone acetyltransferase with bisubstrate analogsKarukurichi, Kannan R.; Cole, Philip A.Bioorganic Chemistry (2011), 39 (1), 42-47CODEN: BOCMBM; ISSN:0045-2068. (Elsevier B.V.)Histone and protein acetylation catalyzed by p300/CBP transcriptional coactivator regulates a variety of key biol. pathways. This study investigates the proposed Theorell-Chance or "hit-and-run" catalytic mechanism of p300/CBP histone acetyltransferase (HAT) using bisubstrate analogs. A range of histone peptide tail peptide-CoA conjugates with different length linkers were synthesized and evaluated as inhibitors of p300 HAT. We show that longer linkers between the histone tail peptide and the CoA substrate moieties appear to allow for dual engagement of the two binding surfaces. Results with D1625R/D1628R double mutant p300 HAT further confirm the requirement for a neg. charged surface on the enzyme to interact with the histone tail.
- 149Cleland, W. W. Biochim. Biophys. Acta 1963, 67, 104Google ScholarThere is no corresponding record for this reference.
- 150Cleland, W. W. Biochim. Biophys. Acta 1963, 67, 104Google ScholarThere is no corresponding record for this reference.
- 151Segel, I. Enzyme Kinetics; Wiley Interscience: New York, 1975.Google ScholarThere is no corresponding record for this reference.
- 152Zheng, Y.; Thompson, P. R.; Cebrat, M.; Wang, L.; Devlin, M. K.; Alani, R. M.; Cole, P. A. Methods Enzymol. 2004, 376, 188Google ScholarThere is no corresponding record for this reference.
- 153Bowers, E. M.; Yan, G.; Mukherjee, C.; Orry, A.; Wang, L.; Holbert, M. A.; Crump, N. T.; Hazzalin, C. A.; Liszczak, G.; Yuan, H.; Larocca, C.; Saldanha, S. A.; Abagyan, R.; Sun, Y.; Meyers, D. J.; Marmorstein, R.; Mahadevan, L. C.; Alani, R. M.; Cole, P. A. Chem. Biol. 2010, 17, 471Google Scholar153Virtual Ligand Screening of the p300/CBP Histone Acetyltransferase: Identification of a Selective Small Molecule InhibitorBowers, Erin M.; Yan, Gai; Mukherjee, Chandrani; Orry, Andrew; Wang, Ling; Holbert, Marc A.; Crump, Nicholas T.; Hazzalin, Catherine A.; Liszczak, Glen; Yuan, Hua; Larocca, Cecilia; Saldanha, S. Adrian; Abagyan, Ruben; Sun, Yan; Meyers, David J.; Marmorstein, Ronen; Mahadevan, Louis C.; Alani, Rhoda M.; Cole, Philip A.Chemistry & Biology (Cambridge, MA, United States) (2010), 17 (5), 471-482CODEN: CBOLE2; ISSN:1074-5521. (Cell Press)Summary: The histone acetyltransferase (HAT) p300/CBP is a transcriptional coactivator implicated in many gene regulatory pathways and protein acetylation events. Although p300 inhibitors have been reported, a potent, selective, and readily available active-site-directed small mol. inhibitor is not yet known. Here the authors use a structure-based, in silico screening approach to identify a com. available pyrazolone-contg. small mol. p300 HAT inhibitor, C646. C646 is a competitive p300 inhibitor with a Ki of 400 nM and is selective vs. other acetyltransferases. Studies on site-directed p300 HAT mutants and synthetic modifications of C646 confirm the importance of predicted interactions in conferring potency. Inhibition of histone acetylation and cell growth by C646 in cells validate its utility as a pharmacol. probe and suggest that p300/CBP HAT is a worthy anticancer target.
- 154Lau, O. D.; Kundu, T. K.; Soccio, R. E.; Ait-Si-Ali, S.; Khalil, E. M.; Vassilev, A.; Wolffe, A. P.; Nakatani, Y.; Roeder, R. G.; Cole, P. A. Mol. Cell 2000, 5, 589Google ScholarThere is no corresponding record for this reference.
- 155Sagar, V.; Zheng, W.; Thompson, P. R.; Cole, P. A. Bioorg. Med. Chem. 2004, 12, 3383Google ScholarThere is no corresponding record for this reference.
- 156Karukurichi, K. R.; Cole, P. A. Bioorg. Chem. 2011, 39, 42Google Scholar156Probing the reaction coordinate of the p300/CBP histone acetyltransferase with bisubstrate analogsKarukurichi, Kannan R.; Cole, Philip A.Bioorganic Chemistry (2011), 39 (1), 42-47CODEN: BOCMBM; ISSN:0045-2068. (Elsevier B.V.)Histone and protein acetylation catalyzed by p300/CBP transcriptional coactivator regulates a variety of key biol. pathways. This study investigates the proposed Theorell-Chance or "hit-and-run" catalytic mechanism of p300/CBP histone acetyltransferase (HAT) using bisubstrate analogs. A range of histone peptide tail peptide-CoA conjugates with different length linkers were synthesized and evaluated as inhibitors of p300 HAT. We show that longer linkers between the histone tail peptide and the CoA substrate moieties appear to allow for dual engagement of the two binding surfaces. Results with D1625R/D1628R double mutant p300 HAT further confirm the requirement for a neg. charged surface on the enzyme to interact with the histone tail.
- 157Cebrat, M.; Kim, C. M.; Thompson, P. R.; Daugherty, M.; Cole, P. A. Bioorg. Med. Chem. 2003, 11, 3307Google ScholarThere is no corresponding record for this reference.
- 158Lau, O. D.; Kundu, T. K.; Soccio, R. E.; Ait-Si-Ali, S.; Khalil, E. M.; Vassilev, A.; Wolffe, A. P.; Nakatani, Y.; Roeder, R. G.; Cole, P. A. Mol. Cell 2000, 5, 589Google ScholarThere is no corresponding record for this reference.
- 159Chase, J. F.; Tubbs, P. K. Biochem. J. 1969, 111, 225Google ScholarThere is no corresponding record for this reference.
- 160Williams, J. W.; Northrop, D. B. J. Antibiot. 1979, 32, 1147Google ScholarThere is no corresponding record for this reference.
- 161Cullis, P. M.; Wolfenden, R.; Cousens, L. S.; Alberts, B. M. J. Biol. Chem. 1982, 257, 12165Google ScholarThere is no corresponding record for this reference.
- 162Erwin, B. G.; Persson, L.; Pegg, A. E. Biochemistry 1984, 23, 4250Google ScholarThere is no corresponding record for this reference.
- 163Khalil, E. M.; Cole, P. A. J. Am. Chem. Soc. 1998, 120, 6195Google ScholarThere is no corresponding record for this reference.
- 164Khalil, E. M.; De Angelis, J.; Ishii, M.; Cole, P. A. Proc. Natl. Acad. Sci. U.S.A. 1999, 96, 12418Google ScholarThere is no corresponding record for this reference.
- 165Kim, C. M.; Cole, P. A. J. Med. Chem. 2001, 44, 2479Google ScholarThere is no corresponding record for this reference.
- 166Yu, M.; Magalhães, M. L.; Cook, P. F.; Blanchard, J. S. Biochemistry 2006, 45, 14788Google ScholarThere is no corresponding record for this reference.
- 167Gao, F.; Yan, X.; Baettig, O. M.; Berghuis, A. M.; Auclair, K. Angew. Chem., Int. Ed. 2005, 44, 6859Google ScholarThere is no corresponding record for this reference.
- 168Gao, F.; Yan, X.; Shakya, T.; Baettig, O. M.; Ait-Mohand-Brunet, S.; Berghuis, A. M.; Wright, G. D.; Auclair, K. J. Med. Chem. 2006, 49, 5273Google ScholarThere is no corresponding record for this reference.
- 169Gao, F.; Yan, X.; Zahr, O.; Larsen, A.; Vong, K.; Auclair, K. Bioorg. Med. Chem. Lett. 2008, 18, 5518Google ScholarThere is no corresponding record for this reference.
- 170Magalhães, M. L.; Vetting, M. W.; Gao, F.; Freiburger, L.; Auclair, K.; Blanchard, J. S. Biochemistry 2008, 47, 579Google ScholarThere is no corresponding record for this reference.
- 171Barnett, B. P.; Hwang, Y.; Taylor, M. S.; Kirchner, H.; Pfluger, P. T.; Bernard, V.; Lin, Y. Y.; Bowers, E. M.; Mukherjee, C.; Song, W. J.; Longo, P. A.; Leahy, D. J.; Hussain, M. A.; Tschöp, M. H.; Boeke, J. D.; Cole, P. A. Science 2010, 330, 1689Google ScholarThere is no corresponding record for this reference.
- 172Taylor, M. S.; Hwang, Y.; Hsiao, P. Y.; Boeke, J. D.; Cole, P. A. Methods Enzymol. 2012, 514, 205Google ScholarThere is no corresponding record for this reference.
- 173Parang, K.; Till, J. H.; Ablooglu, A. J.; Kohanski, R. A.; Hubbard, S. R.; Cole, P. A. Nat. Struct. Biol. 2001, 8, 37Google ScholarThere is no corresponding record for this reference.
- 174Shen, K.; Hines, A. C.; Schwarzer, D.; Pickin, K. A.; Cole, P. A. Biochim. Biophys. Acta 2005, 1754, 65Google ScholarThere is no corresponding record for this reference.
- 175Lau, O. D.; Kundu, T. K.; Soccio, R. E.; Ait-Si-Ali, S.; Khalil, E. M.; Vassilev, A.; Wolffe, A. P.; Nakatani, Y.; Roeder, R. G.; Cole, P. A. Mol. Cell 2000, 5, 589Google ScholarThere is no corresponding record for this reference.
- 176Wu, J.; Xie, N.; Wu, Z.; Zhang, Y.; Zheng, Y. G. Bioorg. Med. Chem. 2009, 17, 1381Google ScholarThere is no corresponding record for this reference.
- 177Zheng, Y.; Thompson, P. R.; Cebrat, M.; Wang, L.; Devlin, M. K.; Alani, R. M.; Cole, P. A. Methods Enzymol. 2004, 376, 188Google ScholarThere is no corresponding record for this reference.
- 178De Angelis, J.; Gastel, J.; Klein, D. C.; Cole, P. A. J. Biol. Chem. 1998, 273, 3045Google ScholarThere is no corresponding record for this reference.
- 179Lau, O. D.; Courtney, A. D.; Vassilev, A.; Marzilli, L. A.; Cotter, R. J.; Nakatani, Y.; Cole, P. A. J. Biol. Chem. 2000, 275, 21953Google ScholarThere is no corresponding record for this reference.
- 180Tanner, K. G.; Langer, M. R.; Denu, J. M. Biochemistry 2000, 39, 15652Google ScholarThere is no corresponding record for this reference.
- 181Tanner, K. G.; Langer, M. R.; Kim, Y.; Denu, J. M. J. Biol. Chem. 2000, 275, 22048Google ScholarThere is no corresponding record for this reference.
- 182Yan, Y.; Barlev, N. A.; Haley, R. H.; Berger, S. L.; Marmorstein, R. Mol. Cell 2000, 6, 1195Google ScholarThere is no corresponding record for this reference.
- 183Wu, J.; Xie, N.; Wu, Z.; Zhang, Y.; Zheng, Y. G. Bioorg. Med. Chem. 2009, 17, 1381Google ScholarThere is no corresponding record for this reference.
- 184Lau, O. D.; Kundu, T. K.; Soccio, R. E.; Ait-Si-Ali, S.; Khalil, E. M.; Vassilev, A.; Wolffe, A. P.; Nakatani, Y.; Roeder, R. G.; Cole, P. A. Mol. Cell 2000, 5, 589Google ScholarThere is no corresponding record for this reference.
- 185Lau, O. D.; Courtney, A. D.; Vassilev, A.; Marzilli, L. A.; Cotter, R. J.; Nakatani, Y.; Cole, P. A. J. Biol. Chem. 2000, 275, 21953Google ScholarThere is no corresponding record for this reference.
- 186Poux, A. N.; Cebrat, M.; Kim, C. M.; Cole, P. A.; Marmorstein, R. Proc. Natl. Acad. Sci. U.S.A. 2002, 99, 14065Google ScholarThere is no corresponding record for this reference.
- 187Wu, J.; Xie, N.; Wu, Z.; Zhang, Y.; Zheng, Y. G. Bioorg. Med. Chem. 2009, 17, 1381Google ScholarThere is no corresponding record for this reference.
- 188Victor, M.; Bei, Y.; Gay, F.; Calvo, D.; Mello, C.; Shi, Y. EMBO Rep. 2002, 3, 50Google ScholarThere is no corresponding record for this reference.
- 189Huang, Z. Q.; Li, J.; Sachs, L. M.; Cole, P. A.; Wong, J. EMBO J. 2003, 22, 2146Google ScholarThere is no corresponding record for this reference.
- 190Costanzo, A.; Merlo, P.; Pediconi, N.; Fulco, M.; Sartorelli, V.; Cole, P. A.; Fontemaggi, G.; Fanciulli, M.; Schiltz, L.; Blandino, G.; Balsano, C.; Levrero, M. Mol. Cell 2002, 9, 175Google ScholarThere is no corresponding record for this reference.
- 191Kaehlcke, K.; Dorr, A.; Hetzer-Egger, C.; Kiermer, V.; Henklein, P.; Schnoelzer, M.; Loret, E.; Cole, P. A.; Verdin, E.; Ott, M. Mol. Cell 2003, 12, 167Google ScholarThere is no corresponding record for this reference.
- 192Polesskaya, A.; Naguibneva, I.; Fritsch, L.; Duquet, A.; Ait-Si-Ali, S.; Robin, P.; Vervisch, A.; Pritchard, L. L.; Cole, P.; Harel-Bellan, A. EMBO J. 2001, 20, 6816Google ScholarThere is no corresponding record for this reference.
- 193Bandyopadhyay, D.; Okan, N. A.; Bales, E.; Nascimento, L.; Cole, P. A.; Medrano, E. E. Cancer Res. 2002, 62, 6231Google ScholarThere is no corresponding record for this reference.
- 194Cebrat, M.; Kim, C. M.; Thompson, P. R.; Daugherty, M.; Cole, P. A. Bioorg. Med. Chem. 2003, 11, 3307Google ScholarThere is no corresponding record for this reference.
- 195Zheng, Y.; Balasubramanyam, K.; Cebrat, M.; Buck, D.; Guidez, F.; Zelent, A.; Alani, R. M.; Cole, P. A. J. Am. Chem. Soc. 2005, 127, 17182Google ScholarThere is no corresponding record for this reference.
- 196Guidez, F.; Howell, L.; Isalan, M.; Cebrat, M.; Alani, R. M.; Ivins, S.; Hormaeche, I.; McConnell, M. J.; Pierce, S.; Cole, P. A.; Licht, J.; Zelent, A. Mol. Cell. Biol. 2005, 25, 5552Google Scholar196Histone acetyltransferase activity of p300 is required for transcriptional repression by the promyelocytic leukemia zinc finger proteinGuidez, Fabien; Howell, Louise; Isalan, Mark; Cebrat, Marek; Alani, Rhoda M.; Ivins, Sarah; Hormaeche, Itsaso; McConnell, Melanie J.; Pierce, Sarah; Cole, Philip A.; Licht, Jonathan; Zelent, ArthurMolecular and Cellular Biology (2005), 25 (13), 5552-5566CODEN: MCEBD4; ISSN:0270-7306. (American Society for Microbiology)Histone acetyltransferase (HAT) activities of proteins such as p300, CBP, and P/CAF play important roles in activation of gene expression. We now show that the HAT activity of p300 can also be required for down-regulation of transcription by a DNA binding repressor protein. Promyelocytic leukemia zinc finger (PLZF), originally identified as a fusion with retinoic acid receptor alpha in rare cases of all-trans-retinoic acid-resistant acute promyelocytic leukemia, is a transcriptional repressor that recruits histone deacetylase-contg. corepressor complexes to specific DNA binding sites. PLZF assocs. with p300 in vivo, and its ability to repress transcription is specifically dependent on HAT activity of p300 and acetylation of lysines in its C-terminal C2-H2 zinc finger motif. An acetylation site mutant of PLZF does not repress transcription and is functionally deficient in a colony suppression assay despite retaining its abilities to interact with corepressor/histone deacetylase complexes. This is due to the fact that acetylation of PLZF activates its ability to bind specific DNA sequences both in vitro and in vivo. Taken together, our results indicate that a histone deacetylase-dependent transcriptional repressor can be pos. regulated through acetylation and point to an unexpected role of a coactivator protein in transcriptional repression.
- 197Balasubramanyam, K.; Swaminathan, V.; Ranganathan, A.; Kundu, T. K. J. Biol. Chem. 2003, 278, 19134Google ScholarThere is no corresponding record for this reference.
- 198Rosen, T.; Fordice, D. B. South Med. J. 1994, 87, 543Google Scholar198Cashew nut dermatitisRosen T; Fordice D BSouthern medical journal (1994), 87 (4), 543-6 ISSN:0038-4348.The urushiol dermatitis caused by plants of the Anacardiaceae family is the most common cause of acute allergic contact dermatitis. We have reported a case of cashew nut urushiol dermatitis due to ingestion of homemade cashew nut butter contaminated by cashew nut shell oil. With the precautions taken today to avoid contamination of food products with cashew urushiols, it is rare to find a case of cashew nut dermatitis in the United States. We have found no other report of contact dermatitis due to cashew nut butter. Moreover, though hinted at in the literature, there has been no previous detailed report of perianal contact dermatitis due to cashew ingestion. The fact that our patient was ill enough to require treatment with 3 weeks of systemic steroid therapy highlights the potential public health hazard of consumption of improperly prepared cashew products. However, the risk of cashew nut dermatitis today remains small, and this should not discourage cashew lovers from enjoying their treats. A final lesson to be learned from this case is that perianal eruptions may be due to materials deliberately applied to the anogenital region or to ingested antigens that remain sufficiently intact within the feces to affect perianal skin.
- 199Balasubramanyam, K.; Swaminathan, V.; Ranganathan, A.; Kundu, T. K. J. Biol. Chem. 2003, 278, 19134Google ScholarThere is no corresponding record for this reference.
- 200Ghizzoni, M.; Wu, J.; Gao, T.; Haisma, H. J.; Dekker, F. J.; George Zheng, Y. Eur. J. Med. Chem. 2012, 47, 337Google ScholarThere is no corresponding record for this reference.
- 201Wu, J.; Xie, N.; Wu, Z.; Zhang, Y.; Zheng, Y. G. Bioorg. Med. Chem. 2009, 17, 1381Google ScholarThere is no corresponding record for this reference.
- 202Chandregowda, V.; Kush, A.; Reddy, G. C. Eur. J. Med. Chem. 2009, 44, 2711Google ScholarThere is no corresponding record for this reference.
- 203Balasubramanyam, K.; Swaminathan, V.; Ranganathan, A.; Kundu, T. K. J. Biol. Chem. 2003, 278, 19134Google ScholarThere is no corresponding record for this reference.
- 204Devipriya, B.; Parameswari, A. R.; Rajalakshmi, G.; Palvannan, T.; Kumaradhas, P. Indian J. Biochem. Biophys. 2010, 47, 364Google Scholar204Exploring the binding affinities of p300 enzyme activators CTPB and CTB using docking methodDevipriya, B.; Parameswari, A. Renuga; Rajalakshmi, G.; Palvannan, T.; Kumaradhas, P.Indian Journal of Biochemistry & Biophysics (2010), 47 (6), 364-369CODEN: IJBBBQ; ISSN:0301-1208. (National Institute of Science Communication and Information Resources)CREB binding protein (CBP) and E1A binding protein p300, also known as p300 are functionally related transcriptional co-activators (CoAs) and histone acetyltransferases (HATs). Some small mols., which target HATs can activate or inhibit the p300 enzyme potently. Here, the binding affinities are reported for two small mols. CTPB [N-(4-chloro-3-trifluoromethyl-phenyl)-2-ethoxy-6-pentadecyl-benzamide] and CTB [N-(4-chloro-3-trifluoromethyl-phenyl)-2-ethoxybenzamide] with p300 using docking method to obtain the insight of their interaction with p300. These small mols. bind to the enzyme, subsequently causing a structural change in the enzyme, which is responsible for the HAT activation. CTB exhibits higher binding affinity than CTPB, and their lowest docked energies are -7.72, -1.18 kcal/mol, resp. In CTPB mol., phenolic hydroxyl of Tyr1397 interacts with the non-polar atoms C(5E) and C(5F), and forms polar-non polar interactions. Similar interactions have also been obsd. in CTB. The residues Tyr1446 and Cys1438 interact with the non-pentadecyl atoms. Further, the docking study predicts a N-H···O hydrogen bonding interaction between CTB and Leu1398, in which the H···O contact distance is 2.06 Å. The long pentadecyl chain of CTPB reduces the formation of hydrogen bond with the p300. The H-bond interaction could be the key factor for the better activation of CTB.
- 205Mantelingu, K.; Kishore, A. H.; Balasubramanyam, K.; Kumar, G. V.; Altaf, M.; Swamy, S. N.; Selvi, R.; Das, C.; Narayana, C.; Rangappa, K. S.; Kundu, T. K. J. Phys. Chem. B 2007, 111, 4527Google ScholarThere is no corresponding record for this reference.
- 206Varier, R. A.; Swaminathan, V.; Balasubramanyam, K.; Kundu, T. K. Biochem. Pharmacol. 2004, 68, 1215Google ScholarThere is no corresponding record for this reference.
- 207Balasubramanyam, K.; Varier, R. A.; Altaf, M.; Swaminathan, V.; Siddappa, N. B.; Ranga, U.; Kundu, T. K. J. Biol. Chem. 2004, 279, 51163Google ScholarThere is no corresponding record for this reference.
- 208Goel, A.; Kunnumakkara, A. B.; Aggarwal, B. B. Biochem. Pharmacol. 2008, 75, 787Google ScholarThere is no corresponding record for this reference.
- 209Maheshwari, R. K.; Singh, A. K.; Gaddipati, J.; Srimal, R. C. Life Sci. 2006, 78, 2081Google ScholarThere is no corresponding record for this reference.
- 210Aggarwal, B. B.; Shishodia, S. Biochem. Pharmacol. 2006, 71, 1397Google Scholar210Molecular targets of dietary agents for prevention and therapy of cancerAggarwal, Bharat B.; Shishodia, ShishirBiochemical Pharmacology (2006), 71 (10), 1397-1421CODEN: BCPCA6; ISSN:0006-2952. (Elsevier B.V.)A review. While fruits and vegetables are recommended for prevention of cancer and other diseases, their active ingredients (at the mol. level) and their mechanisms of action less well understood. Extensive research during the last half century has identified various mol. targets that can potentially be used not only for the prevention of cancer but also for treatment. However, lack of success with targeted monotherapy resulting from bypass mechanisms has forced researchers to employ either combination therapy or agents that interfere with multiple cell-signaling pathways. In this review, we present evidence that numerous agents identified from fruits and vegetables can interfere with several cell-signaling pathways. The agents include curcumin (turmeric), resveratrol (red grapes, peanuts, and berries), genistein (soybean), diallyl sulfide (allium), S-allyl cysteine (allium), allicin (garlic), lycopene (tomato), capsaicin (red chilli), diosgenin (fenugreek), 6-gingerol (ginger), ellagic acid (pomegranate), ursolic acid (apples, pears, prunes), silymarin (milk thistle), anethol (anise, camphor, and fennel), catechins (green tea), eugenol (cloves), indole-3-carbinol (cruciferous vegetables), limonene (citrus fruits), β-carotene (carrots), and dietary fiber. For instance, the cell-signaling pathways inhibited by curcumin alone include NF-κB, AP-1, STAT3, Akt, Bcl-2, Bcl-XL, caspases, PARP, IKK, EGFR, HER2, JNK, MAPK, COX2, and 5-LOX. The active principle identified in fruit and vegetables and the mol. targets modulated may be the basis for how these dietary agents not only prevent but also treat cancer and other diseases. This work reaffirms what Hippocrates said 25 centuries ago, let food be thy medicine and medicine be thy food.
- 211Zhou, H.; Beevers, C. S.; Huang, S. Curr. Drug Targets 2011, 12, 332Google Scholar211The targets of curcuminZhou, Hongyu; Beevers, Christopher S.; Huang, ShileCurrent Drug Targets (2011), 12 (3), 332-347CODEN: CDTUAU; ISSN:1389-4501. (Bentham Science Publishers Ltd.)A review. Curcumin (diferuloylmethane), an orange-yellow component of turmeric or curry powder, is a polyphenol natural product isolated from the rhizome of the plant Curcuma longa. For centuries, curcumin has been used in some medicinal prepn. or used as a food-coloring agent. In recent years, extensive in vitro and in vivo studies suggested curcumin has anticancer, antiviral, antiarthritic, anti-amyloid, antioxidant, and anti-inflammatory properties. The underlying mechanisms of these effects are diverse and appear to involve the regulation of various mol. targets, including transcription factors (such as nuclear factor-κB), growth factors (such as vascular endothelial cell growth factor), inflammatory cytokines (such as tumor necrosis factor, interleukin 1 and interleukin 6), protein kinases (such as mammalian target of rapamycin, mitogen-activated protein kinases, and Akt) and other enzymes (such as cyclooxygenase 2 and 5 lipoxygenase). Thus, due to its efficacy and regulation of multiple targets, as well as its safety for human use, curcumin has received considerable interest as a potential therapeutic agent for the prevention and/or treatment of various malignant diseases, arthritis, allergies, Alzheimer's disease, and other inflammatory illnesses. This review summarizes various in vitro and in vivo pharmacol. aspects of curcumin as well as the underlying action mechanisms. The recently identified mol. targets and signaling pathways modulated by curcumin are also discussed here.
- 212Han, X.; Xu, B.; Beevers, C. S.; Odaka, Y.; Chen, L.; Liu, L.; Luo, Y.; Zhou, H.; Chen, W.; Shen, T.; Huang, S. Carcinogenesis 2012, 33, 868Google ScholarThere is no corresponding record for this reference.
- 213Gupta, S. C.; Prasad, S.; Kim, J. H.; Patchva, S.; Webb, L. J.; Priyadarsini, I. K.; Aggarwal, B. B. Nat. Prod. Rep. 2011, 28, 1937Google Scholar213Multitargeting by curcumin as revealed by molecular interaction studiesGupta, Subash C.; Prasad, Sahdeo; Kim, Ji Hye; Patchva, Sridevi; Webb, Lauren J.; Priyadarsini, Indira K.; Aggarwal, Bharat B.Natural Product Reports (2011), 28 (12), 1937-1955CODEN: NPRRDF; ISSN:0265-0568. (Royal Society of Chemistry)A review. Curcumin (diferuloylmethane), the active ingredient in turmeric (Curcuma longa), is a highly pleiotropic mol. with anti-inflammatory, anti-oxidant, chemopreventive, chemosensitization, and radiosensitization activities. The pleiotropic activities attributed to curcumin come from its complex mol. structure and chem., as well as its ability to influence multiple signaling mols. Curcumin has been shown to bind by multiple forces directly to numerous signaling mols., such as inflammatory mols., cell survival proteins, protein kinases, protein reductases, histone acetyltransferase, histone deacetylase, glyoxalase I, xanthine oxidase, proteasome, HIV1 integrase, HIV1 protease, sarco (endo) plasmic reticulum Ca2+ ATPase, DNA methyltransferases 1, FtsZ protofilaments, carrier proteins, and metal ions. Curcumin can also bind directly to DNA and RNA. Owing to its β-diketone moiety, curcumin undergoes keto-enol tautomerism that has been reported as a favorable state for direct binding. The functional groups on curcumin found suitable for interaction with other macromols. include the α, β-unsatd. β-diketone moiety, carbonyl and enolic groups of the β-diketone moiety, methoxy and phenolic hydroxyl groups, and the Ph rings. Various biophys. tools have been used to monitor direct interaction of curcumin with other proteins, including absorption, fluorescence, Fourier transform IR (FTIR) and CD (CD) spectroscopy, surface plasmon resonance, competitive ligand binding, Forster type fluorescence resonance energy transfer (FRET), radiolabeling, site-directed mutagenesis, matrix-assisted laser desorption/ionization time-of-flight mass spectrometry (MALDI-TOF MS), immunopptn., phage display biopanning, electron microscopy, 1-anilino-8-naphthalene-sulfonate (ANS) displacement, and co-localization. Mol. docking, the most commonly employed computational tool for calcg. binding affinities and predicting binding sites, has also been used to further characterize curcumin's binding sites. Furthermore, the ability of curcumin to bind directly to carrier proteins improves its soly. and bioavailability. In this review, we focus on how curcumin directly targets signaling mols., as well as the different forces that bind the curcumin-protein complex and how this interaction affects the biol. properties of proteins. We will also discuss various analogs of curcumin designed to bind selective targets with increased affinity.
- 214Heery, D. M.; Fischer, P. M. Drug Discovery Today 2007, 12, 88Google ScholarThere is no corresponding record for this reference.
- 215Marcu, M. G.; Jung, Y. J.; Lee, S.; Chung, E. J.; Lee, M. J.; Trepel, J.; Neckers, L. Med. Chem. 2006, 2, 169Google Scholar215Curcumin is an inhibitor of p300 histone acetyltransferaseMarcu, Monica G.; Jung, Yun-Jin; Lee, Sunmin; Chung, Eun-Joo; Lee, Min-Jung; Trepel, Jane; Neckers, LenMedicinal Chemistry (2006), 2 (2), 169-174CODEN: MCEHAJ; ISSN:1573-4064. (Bentham Science Publishers Ltd.)Histone acetyltransferases (HATs), and p300/CBP in particular, have been implicated in cancer cell growth and survival, and as such, HATs represent novel, therapeutically relevant mol. targets for drug development. In this study, we demonstrate that the small mol. natural product curcumin, whose medicinal properties have long been recognized in India and Southeast Asia, is a selective HAT inhibitor. Furthermore the data indicate that α, β unsatd. carbonyl groups in the curcumin side chain function as Michael reaction sites and that the Michael reaction acceptor functionality of curcumin is required for its HAT-inhibitory activity. In cells, curcumin promoted proteasome-dependent degrdn. of p300 and the closely related CBP protein without affecting the HATs PCAF or GCN5. In addn. to inducing p300 degrdn. curcumin inhibited the acetyltransferase activity of purified p300 as assessed using either histone H3 or p53 as substrate. Radiolabeled curcumin formed a covalent assocn. with p300, and tetrahydrocurcumin displayed no p300 inhibitory activity, consistent with a Michael reaction-dependent mechanism. Finally, curcumin was able to effectively block histone hyperacetylation in both PC3-M prostate cancer cells and peripheral blood lymphocytes induced by the histone deacetylase inhibitor MS-275. These data thus identify the medicinal natural product curcumin as a novel lead compd. for development of possibly therapeutic, p300/CBP-specific HAT inhibitors.
- 216Balasubramanyam, K.; Varier, R. A.; Altaf, M.; Swaminathan, V.; Siddappa, N. B.; Ranga, U.; Kundu, T. K. J. Biol. Chem. 2004, 279, 51163Google ScholarThere is no corresponding record for this reference.
- 217Anand, P.; Kunnumakkara, A. B.; Newman, R. A.; Aggarwal, B. B. Mol. Pharmaceutics 2007, 4, 807Google Scholar217Bioavailability of Curcumin: Problems and PromisesAnand, Preetha; Kunnumakkara, Ajaikumar B.; Newman, Robert A.; Aggarwal, Bharat B.Molecular Pharmaceutics (2007), 4 (6), 807-818CODEN: MPOHBP; ISSN:1543-8384. (American Chemical Society)A review. Curcumin, a polyphenolic compd. derived from dietary spice turmeric, possesses diverse pharmacol. effects including anti-inflammatory, antioxidant, antiproliferative and antiangiogenic activities. Phase I clin. trials have shown that curcumin is safe even at high doses (12 g/day) in humans but exhibit poor bioavailability. Major reasons contributing to the low plasma and tissue levels of curcumin appear to be due to poor absorption, rapid metab., and rapid systemic elimination. To improve the bioavailability of curcumin, numerous approaches have been undertaken. These approaches involve, first, the use of adjuvant like piperine that interferes with glucuronidation; second, the use of liposomal curcumin; third, curcumin nanoparticles; fourth, the use of curcumin phospholipid complex; and fifth, the use of structural analogs of curcumin (e.g., EF-24). The latter has been reported to have a rapid absorption with a peak plasma half-life. Despite the lower bioavailability, therapeutic efficacy of curcumin against various human diseases, including cancer, cardiovascular diseases, diabetes, arthritis, neurol. diseases and Crohn's disease, has been documented. Enhanced bioavailability of curcumin in the near future is likely to bring this promising natural product to the forefront of therapeutic agents for treatment of human disease.
- 218Sharma, R. A.; Euden, S. A.; Platton, S. L.; Cooke, D. N.; Shafayat, A.; Hewitt, H. R.; Marczylo, T. H.; Morgan, B.; Hemingway, D.; Plummer, S. M.; Pirmohamed, M.; Gescher, A. J.; Steward, W. P. Clin. Cancer Res. 2004, 10, 6847Google Scholar218Phase I Clinical Trial of Oral Curcumin: Biomarkers of Systemic Activity and ComplianceSharma, Ricky A.; Euden, Stephanie A.; Platton, Sharon L.; Cooke, Darren N.; Shafayat, Aisha; Hewitt, Heather R.; Marczylo, Timothy H.; Morgan, Bruno; Hemingway, David; Plummer, Simon M.; Pirmohamed, Munir; Gescher, Andreas J.; Steward, William P.Clinical Cancer Research (2004), 10 (20), 6847-6854CODEN: CCREF4; ISSN:1078-0432. (American Association for Cancer Research)Curcumin, a polyphenolic antioxidant derived from a dietary spice, exhibits anticancer activity in rodents and in humans. Its efficacy appears to be related to induction of glutathione S-transferase enzymes, inhibition of prostaglandin E2 (PGE2) prodn., or suppression of oxidative DNA adduct (M1G) formation. We designed a dose-escalation study to explore the pharmacol. of curcumin in humans. Fifteen patients with advanced colorectal cancer refractory to std. chemotherapies consumed capsules compatible with curcumin doses between 0.45 and 3.6 g daily for up to 4 mo. Levels of curcumin and its metabolites in plasma, urine, and feces were analyzed by high-pressure liq. chromatog. and mass spectrometry. Three biomarkers of the potential activity of curcumin were translated from preclin. models and measured in patient blood leukocytes: glutathione S-transferase activity, levels of M1G, and PGE2 prodn. induced ex vivo. Dose-limiting toxicity was not obsd. Curcumin and its glucuronide and sulfate metabolites were detected in plasma in the 10 nmol/L range and in urine. A daily dose of 3.6 g curcumin engendered 62% and 57% decreases in inducible PGE2 prodn. in blood samples taken 1 h after dose on days 1 and 29, resp., of treatment compared with levels obsd. immediately predose (P < 0.05). A daily oral dose of 3.6 g of curcumin is advocated for Phase II evaluation in the prevention or treatment of cancers outside the gastrointestinal tract. PGE2 prodn. in blood and target tissue may indicate biol. activity. Levels of curcumin and its metabolites in the urine can be used to assess general compliance.
- 219Mishra, S.; Kapoor, N.; Mubarak Ali, A.; Pardhasaradhi, B. V.; Kumari, A. L.; Khar, A.; Misra, K. Free Radical Biol. Med. 2005, 38, 1353Google Scholar219Differential apoptotic and redox regulatory activities of curcumin and its derivativesMishra, Satyendra; Kapoor, Neha; Mubarak Ali, A.; Pardhasaradhi, B. V. V.; Kumari, A. Leela; Khar, Ashok; Misra, KrishnaFree Radical Biology & Medicine (2005), 38 (10), 1353-1360CODEN: FRBMEH; ISSN:0891-5849. (Elsevier)We have synthesized different bioconjugates of curcumin, which were tested for their pro- and antioxidant properties. In the present study five representative derivs. of curcumin, i.e., 4,4'-di-(O-acetyl) curcumin, 4,4'-di-(O-glycinoyl) curcumin, 4,4'-di-(O-glycinoyl-di-N-piperoyl) curcumin, 4,4'-di-(O-piperoyl) curcumin, and 4,4'-(O,O-cystinoyl)-3,3'-dimethoxydiphenyl-1,6-heptadiene-3,5-dione, were used for testing their apoptotic potential on tumor cells. Dipiperoyl and diglycinoyl derivs. showed higher apoptotic activity at lower concns., whereas diacetyl curcumin had slightly lower apoptotic activity on tumor cells. On the other hand, diglycinoyl-dipiperoyl and cystinoyl heptadiene derivs. had lost their apoptotic potential significantly. The apoptotic activity of these derivs. correlated very well with the generation of ROS by the tumor cells, whereas GSH levels remained unaltered. Our studies also indicate down-regulation of Bcl-2 and participation of caspase-3 in the apoptotic death of tumor cells.
- 220Weber, W. M.; Hunsaker, L. A.; Abcouwer, S. F.; Deck, L. M.; Vander Jagt, D. L. Bioorg. Med. Chem. 2005, 13, 3811Google ScholarThere is no corresponding record for this reference.
- 221Varier, R. A.; Swaminathan, V.; Balasubramanyam, K.; Kundu, T. K. Biochem. Pharmacol. 2004, 68, 1215Google ScholarThere is no corresponding record for this reference.
- 222Balasubramanyam, K.; Altaf, M.; Varier, R. A.; Swaminathan, V.; Ravindran, A.; Sadhale, P. P.; Kundu, T. K. J. Biol. Chem. 2004, 279, 33716Google ScholarThere is no corresponding record for this reference.
- 223Balasubramanyam, K.; Altaf, M.; Varier, R. A.; Swaminathan, V.; Ravindran, A.; Sadhale, P. P.; Kundu, T. K. J. Biol. Chem. 2004, 279, 33716Google ScholarThere is no corresponding record for this reference.
- 224Mantelingu, K.; Reddy, B. A.; Swaminathan, V.; Kishore, A. H.; Siddappa, N. B.; Kumar, G. V.; Nagashankar, G.; Natesh, N.; Roy, S.; Sadhale, P. P.; Ranga, U.; Narayana, C.; Kundu, T. K. Chem. Biol. 2007, 14, 645Google Scholar224Specific Inhibition of p300-HAT Alters Global Gene Expression and Represses HIV ReplicationMantelingu, K.; Reddy, B. A. Ashok; Swaminathan, V.; Kishore, A. Hari; Siddappa, Nagadenahalli B.; Kumar, G. V. Pavan; Nagashankar, G.; Natesh, Nagashayana; Roy, Siddhartha; Sadhale, Parag P.; Ranga, Udaykumar; Narayana, Chandrabhas; Kundu, Tapas K.Chemistry & Biology (Cambridge, MA, United States) (2007), 14 (6), 645-657CODEN: CBOLE2; ISSN:1074-5521. (Cell Press)Summary: Reversible acetylation of histone and nonhistone proteins plays pivotal role in cellular homeostasis. Dysfunction of histone acetyltransferases (HATs) leads to several diseases including cancer, neurodegenaration, asthma, diabetes, AIDS, and cardiac hypertrophy. We describe the synthesis and characterization of a set of p300-HAT-specific small-mol. inhibitors from a natural nonspecific HAT inhibitor, garcinol, which is highly toxic to cells. We show that the specific inhibitor selectively represses the p300-mediated acetylation of p53 in vivo. Furthermore, inhibition of p300-HAT down regulates several genes but significantly a few important genes are also upregulated. Remarkably, these inhibitors were found to be nontoxic to T cells, inhibit histone acetylation of HIV infected cells, and consequently inhibit the multiplication of HIV.
- 225Arif, M.; Pradhan, S. K.; Thanuja, G. R.; Vedamurthy, B. M.; Agrawal, S.; Dasgupta, D.; Kundu, T. K. J. Med. Chem. 2009, 52, 267Google ScholarThere is no corresponding record for this reference.
- 226Baggett, S.; Protiva, P.; Mazzola, E. P.; Yang, H.; Ressler, E. T.; Basile, M. J.; Weinstein, I. B.; Kennelly, E. J. J. Nat. Prod. 2005, 68, 354Google ScholarThere is no corresponding record for this reference.
- 227Gartner, M.; Müller, T.; Simon, J. C.; Giannis, A.; Sleeman, J. P. ChemBioChem 2005, 6, 171Google ScholarThere is no corresponding record for this reference.
- 228Dal Piaz, F.; Tosco, A.; Eletto, D.; Piccinelli, A. L.; Moltedo, O.; Franceschelli, S.; Sbardella, G.; Remondelli, P.; Rastrelli, L.; Vesci, L.; Pisano, C.; De Tommasi, N. ChemBioChem 2010, 11, 818Google ScholarThere is no corresponding record for this reference.
- 229Biel, M.; Kretsovali, A.; Karatzali, E.; Papamatheakis, J.; Giannis, A. Angew. Chem., Int. Ed. 2004, 43, 3974Google ScholarThere is no corresponding record for this reference.
- 230Heery, D. M.; Fischer, P. M. Drug Discovery Today 2007, 12, 88Google ScholarThere is no corresponding record for this reference.
- 231Ravindra, K. C.; Selvi, B. R.; Arif, M.; Reddy, B. A.; Thanuja, G. R.; Agrawal, S.; Pradhan, S. K.; Nagashayana, N.; Dasgupta, D.; Kundu, T. K. J. Biol. Chem. 2009, 284, 24453Google ScholarThere is no corresponding record for this reference.
- 232Sandur, S. K.; Ichikawa, H.; Sethi, G.; Ahn, K. S.; Aggarwal, B. B. J. Biol. Chem. 2006, 281, 17023Google ScholarThere is no corresponding record for this reference.
- 233Padhye, S.; Dandawate, P.; Yusufi, M.; Ahmad, A.; Sarkar, F. H. Med. Res. Rev. 2012, 32, 1131Google Scholar233Perspectives on medicinal properties of plumbagin and its analogsPadhye, Subhash; Dandawate, Prasad; Yusufi, Mujahid; Ahmad, Aamir; Sarkar, Fazlul H.Medicinal Research Reviews (2012), 32 (6), 1131-1158CODEN: MRREDD; ISSN:0198-6325. (John Wiley & Sons, Inc.)A review. Plumbagin is one of the simplest plant secondary metabolite of three major phylogenic families viz. Plumbaginaceae, Droseraceae, and Ebenceae, and exhibits highly potent biol. activities, including antioxidant, antiinflammatory, anticancer, antibacterial, and antifungal activities. Recent investigations indicate that these activities arise mainly out of its ability to undergo redox cycling, generating reactive oxygen species and chelating trace metals in biol. system. The compd. is endowed with a property to inhibit the drug efflux mechanism in drug-resistant bacteria, thereby allowing intracellular accumulation of the potent drug mols. An interesting bioactivity exhibited by this compd. is the elimination of stringent, conjugative, multidrug-resistant plasmids from several bacterial strains including opportunistic bacteria, such as Acinetobacter baumannii. Moreover, plumbagin effectively induces apoptosis and causes cell cycle arrest, which is, in part, due to the inactivation of NF-κB in cancer cells. Therefore, it has been suggested that designing "hybrid drug mols." of plumbagin by combining it with other appropriate anticancer agents may lead to the generation of novel and potent anticancer drugs with pleiotropic action against human cancers. This comprehensive review is an attempt to understand the chem. of plumbagin and catalog its biol. activities reported to date.
- 234Ravindra, K. C.; Selvi, B. R.; Arif, M.; Reddy, B. A.; Thanuja, G. R.; Agrawal, S.; Pradhan, S. K.; Nagashayana, N.; Dasgupta, D.; Kundu, T. K. J. Biol. Chem. 2009, 284, 24453Google ScholarThere is no corresponding record for this reference.
- 235Choi, K. C.; Jung, M. G.; Lee, Y. H.; Yoon, J. C.; Kwon, S. H.; Kang, H. B.; Kim, M. J.; Cha, J. H.; Kim, Y. J.; Jun, W. J.; Lee, J. M.; Yoon, H. G. Cancer Res. 2009, 69, 583Google ScholarThere is no corresponding record for this reference.
- 236Neukam, K.; Pastor, N.; Cortés, F. Mutat. Res. 2008, 654, 8Google Scholar236Tea flavanols inhibit cell growth and DNA topoisomerase II activity and induce endoreduplication in cultured Chinese hamster cellsNeukam, Karin; Pastor, Nuria; Cortes, FelipeMutation Research, Genetic Toxicology and Environmental Mutagenesis (2008), 654 (1), 8-12CODEN: MRGMFI; ISSN:1383-5718. (Elsevier B.V.)Tea polyphenols are promising chemopreventive anticancer agents, the properties of which have been studied both in vitro and in vivo, providing evidence that - within this group of compds. - the tea flavanols are able to inhibit carcinogenesis, an effect that in some cases could be correlated with increased cell apoptosis and decreased cell proliferation. Of four main tea flavanols, namely (-)-epigallocatechin-3-gallate (EGCG), (-)-epigallocatechin (EGC), (+)-catechin (CA) and (-)-epicatechin (EC), it was found that EGCG was the most potent to inhibit dose dependently the topoisomerase II (TOPO II) catalytic activity isolated from hamster ovary AA8 cells. In the range of concns. that caused TOPO II inhibition, a high level of endoreduplication, a rare phenomenon that consists in two successive rounds of DNA replication without intervening mitosis, was obsd., while neither micronuclei nor DNA strand breaks (Comet assay) were detected at the same doses. We propose that the anticarcinogenic effect of tea flavanols can be partly explained by their potency and effectiveness to induce endoreduplication. Concerning such an induction, max. effect seems to require a pyrogallol structure at the B-ring. Addnl. substitution with a galloylic residue at the C3 hydroxyl group leads to further augmentation of the effect. Thus, we suggest that the chemopreventive properties of tea flavanols can be at least partly due to their ability to interfere with the cell cycle and block cell proliferation at early stages of mitosis.
- 237Berger, S. J.; Gupta, S.; Belfi, C. A.; Gosky, D. M.; Mukhtar, H. Biochem. Biophys. Res. Commun. 2001, 288, 101Google ScholarThere is no corresponding record for this reference.
- 238Suzuki, K.; Yahara, S.; Hashimoto, F.; Uyeda, M. Biol. Pharm. Bull. 2001, 24, 1088Google Scholar238Inhibitory activities of (-)-epigallocatechin-3-O-gallate against topoisomerases I and IISuzuki, Keitarou; Yahara, Shoji; Hashimoto, Furnio; Uyeda, MasaruBiological & Pharmaceutical Bulletin (2001), 24 (9), 1088-1090CODEN: BPBLEO; ISSN:0918-6158. (Pharmaceutical Society of Japan)The substitution of gallic acid at the 3 position of (-)-epigallocatechin-3-O-gallate (EGCG) increased the inhibition against topoisomerase I from calf thymus gland and topoisomerase II from human placenta, and the substitution of a hydroxyl group at the 3' position increased the inhibition against the topoisomerase I. These results suggested that the 3 and 3' positions of the EGCG mol. play important roles in the process of inhibition of topoisomerases I and II. EGCG showed strong inhibition against topoisomerases I from wheat germ, calf thymus gland, and Vero cells, and showed weak or no inhibition against topoisomerases I from carcinoma cells such as A549, HeLaand COLO 201 cells. EGCG differentially inhibited the topoisomerases I from different sources.
- 239Bandele, O. J.; Osheroff, N. Chem. Res. Toxicol. 2008, 21, 936Google ScholarThere is no corresponding record for this reference.
- 240Golden, E. B.; Lam, P. Y.; Kardosh, A.; Gaffney, K. J.; Cadenas, E.; Louie, S. G.; Petasis, N. A.; Chen, T. C.; Schönthal, A. H. Blood 2009, 113, 5927Google ScholarThere is no corresponding record for this reference.
- 241Ge, J.; Tan, B. X.; Chen, Y.; Yang, L.; Peng, X. C.; Li, H. Z.; Lin, H. J.; Zhao, Y.; Wei, M.; Cheng, K.; Li, L. H.; Dong, H.; Gao, F.; He, J. P.; Wu, Y.; Qiu, M.; Zhao, Y. L.; Su, J. M.; Hou, J. M.; Liu, J. Y. J. Mol. Med. 2011, 89, 595Google Scholar241Interaction of green tea polyphenol epigallocatechin-3-gallate with sunitinib: potential risk of diminished sunitinib bioavailabilityGe, Jun; Tan, Ben-Xu; Chen, Ye; Yang, Li; Peng, Xing-Chen; Li, Hong-Ze; Lin, Hong-Jun; Zhao, Yu; Wei, Meng; Cheng, Ke; Li, Long-Hao; Dong, Hang; Gao, Feng; He, Jian-Ping; Wu, Yang; Qiu, Meng; Zhao, Ying-Lan; Su, Jing-Mei; Hou, Jian-Mei; Liu, Ji-YanJournal of Molecular Medicine (Heidelberg, Germany) (2011), 89 (6), 595-602CODEN: JMLME8; ISSN:0946-2716. (Springer)Sunitinib, a novel oral multi-targeted tyrosine kinase inhibitor for patients with metastatic renal cell carcinoma (mRCC) and advanced gastrointestinal stromal tumor, has a good prospect for clin. application and is being investigated for the potential therapy of other tumors. We obsd. the phenomenon that drinking tea interfered with symptom control in an mRCC patient treated with sunitinib and speculated that green tea or its components might interact with sunitinib. This study was performed to investigate whether epigallocatechin-3-gallate (EGCG), the major constituent of green tea, interacted with sunitinib. The interaction between EGCG and sunitinib was examd. in vitro and in vivo. 1H NMR (1H-NMR) spectroscopy and mass spectrometry (MS) were used to analyze the interaction between these two mols. and whether a new compd. was formed. Solns. of sunitinib and EGCG were intragastrically administered to rats to investigate whether the plasma concns. of sunitinib were affected by EGCG. In this study, we noticed that a ppt. was formed when the solns. of sunitinib and EGCG were mixed under both neutral and acidic conditions. 1H-NMR spectra indicated an interaction between EGCG and sunitinib, but no new compd. was obsd. by MS. Sticky semisolid contents were found in the stomachs of sunitinib and EGCG co-administrated mice. The and C max of plasma sunitinib were markedly reduced by co-administration of EGCG to rats. Our study firstly showed that EGCG interacted with sunitinib and reduced the bioavailability of sunitinib. This finding has significant practical implications for tea-drinking habit during sunitinib administration.
- 242Strick, R.; Strissel, P. L.; Borgers, S.; Smith, S. L.; Rowley, J. D. Proc. Natl. Acad. Sci. U.S.A. 2000, 97, 4790Google ScholarThere is no corresponding record for this reference.
- 243Ruiz, P. A.; Braune, A.; Hölzlwimmer, G.; Quintanilla-Fend, L.; Haller, D. J. Nutr. 2007, 137, 1208Google Scholar243Quercetin inhibits TNF-induced NF-κB transcription factor recruitment to proinflammatory gene promoters in murine intestinal epithelial cellsRuiz, Pedro A.; Braune, Annett; Hoelzlwimmer, Gabriele; Quintanilla-Fend, Leticia; Haller, DirkJournal of Nutrition (2007), 137 (5), 1208-1215CODEN: JONUAI; ISSN:0022-3166. (American Society for Nutrition)Flavonoids may play an important role for adjunct nutritional therapy of chronic intestinal inflammation. In this study, we characterized the mol. mechanisms by which quercetin and its enteric bacterial metabolites, taxifolin, alphitonin, and 3, 4-dihydroxy-phenylacetic acid, inhibit tumor necrosis factor α (TNF)-induced proinflammatory gene expression in the murine small intestinal epithelial cell (IEC) line Mode-K as well as in heterozygous TNFΔARE/WT mice, a murine model of exptl. ileitis. Quercetin inhibited TNF-induced interferon-γ-inducible protein 10 (IP-10) and macrophage inflammatory protein 2 (MIP-2) gene expression in Mode-K cells with effective inhibitory concn. of 40 and 44 μmol/L, resp. Interestingly, taxifolin, alphitonin, and 3,4-dihydroxy-phenylacetic acid did not inhibit TNF responses in IEC, suggesting that microbial transformation of quercetin completely abolished its anti-inflammatory effect. At the mol. level, quercetin inhibited Akt phosphorylation but did not inhibit TNF-induced RelA/I-κB phosphorylation and IκB degrdn. or TNF-α-induced nuclear factor-κB transcriptional activity. Most important for understanding the mechanism involved, chromatin immunopptn. anal. revealed inhibitory effects of quercetin on phospho-RelA recruitment to the IP-10 and MIP-2 gene promoters. In addn., and consistent with the lack of cAMP response element binding protein (CBP)/p300 recruitment and phosphorylation/acetylation of histone 3 at the promoter binding site, quercetin inhibited histone acetyl transferase activity. The oral application of quercetin to heterozygous TNFΔARE/WT mice [10 mg/(d × kg body wt)] significantly inhibited IP-10 and MIP-2 gene expression in primary ileal epithelial cells but did not affect tissue pathol. These studies support an anti-inflammatory effect of quercetin in epithelial cells through mechanisms that inhibit cofactor recruitment at the chromatin of proinflammatory genes.
- 244Chen, J.; Han, J.; Wang, J. Toxicol. Ind. Health 2013, 29, 360Google Scholar244Prevention of cytotoxicity of nickel by quercetin: the role of reactive oxygen species and histone acetylationChen, Jie; Han, Jia; Wang, JianminToxicology and Industrial Health (2013), 29 (4), 360-366CODEN: TIHEEC; ISSN:0748-2337. (Sage Publications)Excessive exposure to nickel may cause health effects on the blood, lung, nose, kidney, reproductive system, skin and the unborn child. In the present study, we found that Ni2+ exposure led to a time- and dosedependent proliferation arrest and death in human leukemia HL-60 cells. In the presence of 1 μM Ni2+, reactive oxygen species (ROS) generation (indicated by the level of malondialdehyde) increased to 323% and histone acetylation decreased to 32%. Interestingly, quercetin (QU) dose dependently prevented Ni2+-induced cell proliferation arrest and death from 0 to 80 μM but showed similar activity of scavenging ROS at the concns. of 20, 40 and 80 μM. When the effect of QU on histone acetylation was studied, QU significantly prevented Ni2+-induced histone hypoacetylation at 40 or 80 μM. Moreover, increase in histone acetylation by trichostatin A could also significantly enhance the protection effect of QU at 10 or 20 μM but not at higher concns. Thus, our results further confirmed the crit. role of ROS and histone hypoacetylation in the cytotoxicity of Ni2+ exposure and proved that QU is a potentially useful native dietary compd. to efficiently prevent Ni2+-caused cytotoxicity through both diminishing ROS generation and increasing histone acetylation.
- 245Si, D.; Wang, Y.; Zhou, Y. H.; Guo, Y.; Wang, J.; Zhou, H.; Li, Z. S.; Fawcett, J. P. Drug Metab. Dispos. 2009, 37, 629Google ScholarThere is no corresponding record for this reference.
- 246Saaby, L.; Rasmussen, H. B.; Jäger, A. K. J. Ethnopharmacol. 2009, 121, 178Google ScholarThere is no corresponding record for this reference.
- 247Hilliard, J. J.; Krause, H. M.; Bernstein, J. I.; Fernandez, J. A.; Nguyen, V.; Ohemeng, K. A.; Barrett, J. F. Adv. Exp. Med. Biol. 1995, 390, 59Google ScholarThere is no corresponding record for this reference.
- 248Xiao, X.; Shi, D.; Liu, L.; Wang, J.; Xie, X.; Kang, T.; Deng, W. PLoS One 2011, 6, e22934Google ScholarThere is no corresponding record for this reference.
- 249O’Brien, E.; Dietrich, D. R. Crit. Rev. Toxicol. 2005, 35, 33Google Scholar249Ochratoxin A: The Continuing EnigmaO'Brien, Evelyn; Dietrich, DanielCritical Reviews in Toxicology (2005), 35 (1), 33-60CODEN: CRTXB2; ISSN:1040-8444. (Taylor & Francis, Inc.)A review. The mycotoxin ochratoxin A (OTA) has been linked to the genesis of several disease states in both animals and humans. It has been described as nephrotoxic, carcinogenic, teratogenic, immunotoxic, and hepatotoxic in lab. and domestic animals, as well as being thought to be the probable causal agent in the development of nephropathies (Balkan Endemic Nephropathy, BEN and Chronic Interstitial Nephropathy, CIN) and urothelial tumors in humans. As a result, several international agencies are currently attempting to define safe legal limits for OTA concn. in foodstuffs (e.g., grain, meat, wine, and coffee), in processed foods, and in animal fodder. In order to achieve this goal, an accurate risk assessment of OTA toxicity including mechanistic and epidemiol. studies must be carried out. Ochratoxin has been suggested by various researchers to mediate its toxic effects via induction of apoptosis, disruption of mitochondrial respiration and/or the cytoskeleton, or, indeed, via the generation of DNA adducts. Thus, it is still unclear if the predominant mechanism is of a genotoxic or an epigenetic nature. One aspect that is clear, however, is that the toxicity of OTA is subject to and characterized by large species- and sex-specific differences, as well as an apparently strict structure-activity relationship. These considerations could be crucial in the investigation of OTA-mediated toxicity. Furthermore, the use of appropriate in vivo and in vitro model systems appears to be vital in the generation of relevant exptl. data. The intention of this review is to collate and discuss the currently available data on OTA-mediated toxicity with particular focus on their relevance for the in vivo situation, and also to suggest possible future strategies for unlocking the secrets of ochratoxin A.
- 250Czakai, K.; Müller, K.; Mosesso, P.; Pepe, G.; Schulze, M.; Gohla, A.; Patnaik, D.; Dekant, W.; Higgins, J. M.; Mally, A. Toxicol. Sci. 2011, 122, 317Google Scholar250Perturbation of Mitosis through Inhibition of Histone Acetyltransferases: The Key to Ochratoxin A Toxicity and Carcinogenicity?Czakai, Kristin; Mueller, Katja; Mosesso, Pasquale; Pepe, Gaetano; Schulze, Markus; Gohla, Antje; Patnaik, Debasis; Dekant, Wolfgang; Higgins, Jonathan M. G.; Mally, AngelaToxicological Sciences (2011), 122 (2), 317-329CODEN: TOSCF2; ISSN:1096-0929. (Oxford University Press)Ochratoxin A (OTA) is one of the most potent rodent renal carcinogens studied to date. Although controversial results regarding OTA genotoxicity have been published, it is now widely accepted that OTA is not a mutagenic, DNA-reactive carcinogen. Instead, increasing evidence from both in vivo and in vitro studies suggests that OTA may promote genomic instability and tumorigenesis through interference with cell division. The aim of the present study was to provide further support for disruption of mitosis as a key event in OTA toxicity and to understand how OTA mediates these effects. Immortalized human kidney epithelial cells (IHKE) were treated with OTA and monitored by differential interference contrast microscopy for 15 h. Image anal. confirmed that OTA at concns. ≥ 5 μM, which correlate with plasma concns. in rats under conditions of carcinogenesis, causes sustained mitotic arrest and exit from mitosis without nuclear or cellular division. Mitotic chromosomes were characterized by aberrant condensation and premature sister chromatid sepn. assocd. with altered phosphorylation and acetylation of core histones. To test if OTA directly interferes with histone acetyltransferases (HATs) which regulate lysine acetylation of histones and nonhistone proteins, a cell-free HAT activity assay was conducted using total nuclear exts. of IHKE cells. In this assay, OTA significantly blocked HAT activity in a concn.-dependent manner Overall, results from this study provide further support for a mechanism of OTA carcinogenicity involving interference with the mitotic machinery and suggest HATs as a primary cellular target of OTA.
- 251Ravindra, K. C.; Narayan, V.; Lushington, G. H.; Peterson, B. R.; Prabhu, K. S. Chem. Res. Toxicol. 2012, 25, 337Google ScholarThere is no corresponding record for this reference.
- 252Reginato, M. J.; Krakow, S. L.; Bailey, S. T.; Lazar, M. A. J. Biol. Chem. 1998, 273, 1855Google ScholarThere is no corresponding record for this reference.
- 253Ravindra, K. C.; Narayan, V.; Lushington, G. H.; Peterson, B. R.; Prabhu, K. S. Chem. Res. Toxicol. 2012, 25, 337Google ScholarThere is no corresponding record for this reference.
- 254Stimson, L.; Rowlands, M. G.; Newbatt, Y. M.; Smith, N. F.; Raynaud, F. I.; Rogers, P.; Bavetsias, V.; Gorsuch, S.; Jarman, M.; Bannister, A.; Kouzarides, T.; McDonald, E.; Workman, P.; Aherne, G. W. Mol. Cancer Ther. 2005, 4, 1521Google Scholar254Isothiazolones as inhibitors of PCAF and p300 histone acetyltransferase activityStimson, Lindsay; Rowlands, Martin G.; Newbatt, Yvette M.; Smith, Nicola F.; Raynaud, Florence I.; Rogers, Paul; Bavetsias, Vassilios; Gorsuch, Stephen; Jarman, Michael; Bannister, Andrew; Kouzarides, Tony; McDonald, Edward; Workman, Paul; Aherne, G. WynneMolecular Cancer Therapeutics (2005), 4 (10), 1521-1532CODEN: MCTOCF; ISSN:1535-7163. (American Association for Cancer Research)Histone acetylation plays an important role in regulating the chromatin structure and is tightly regulated by two classes of enzyme, histone acetyltransferases (HAT) and histone deacetylases (HDAC). Deregulated HAT and HDAC activity plays a role in the development of a range of cancers. Consequently, inhibitors of these enzymes have potential as anticancer agents. Several HDAC inhibitors have been described; however, few inhibitors of HATs have been disclosed. Following a FlashPlate high-throughput screen, we identified a series of isothiazolone-based HAT inhibitors. Thirty-five N-substituted analogs inhibited both p300/cAMP-responsive element binding protein-binding protein-assocd. factor (PCAF) and p300 (1 to > 50 μmol/L, resp.) and the growth of a panel of human tumor cell lines (50% growth inhibition, 0.8 to > 50 μmol/L). CCT077791 and CCT077792 decreased cellular acetylation in a time-dependent manner (2-48 h of exposure) and a concn.-dependent manner (one to five times, 72 h, 50% growth inhibition) in HCT116 and HT29 human colon tumor cell lines. CCT077791 reduced total acetylation of histones H3 and H4, levels of specific acetylated lysine marks, and acetylation of α-tubulin. Four and 24 h of exposure to the compds. produced the same extent of growth inhibition as 72 h of continuous exposure, suggesting that growth arrest was an early event. Chem. reactivity of these compds., as measured by covalent protein binding and loss of HAT inhibition in the presence of DTT, indicated that reaction with thiol groups might be important in their mechanism of action. As one of the first series of small-mol. inhibitors of HAT activity, further analog synthesis is being pursued to examine the potential scope for reducing chem. reactivity while maintaining HAT inhibition.
- 255Dekker, F. J.; Ghizzoni, M.; van der Meer, N.; Wisastra, R.; Haisma, H. J. Bioorg. Med. Chem. 2009, 17, 460Google ScholarThere is no corresponding record for this reference.
- 256Gorsuch, S.; Bavetsias, V.; Rowlands, M. G.; Aherne, G. W.; Workman, P.; Jarman, M.; McDonald, E. Bioorg. Med. Chem. 2009, 17, 467Google ScholarThere is no corresponding record for this reference.
- 257Gorsuch, S.; Bavetsias, V.; Rowlands, M. G.; Aherne, G. W.; Workman, P.; Jarman, M.; McDonald, E. Bioorg. Med. Chem. 2009, 17, 467Google ScholarThere is no corresponding record for this reference.
- 258Wisastra, R.; Ghizzoni, M.; Maarsingh, H.; Minnaard, A. J.; Haisma, H. J.; Dekker, F. J. Org. Biomol. Chem. 2011, 9, 1817Google ScholarThere is no corresponding record for this reference.
- 259Dekker, F. J.; Ghizzoni, M.; van der Meer, N.; Wisastra, R.; Haisma, H. J. Bioorg. Med. Chem. 2009, 17, 460Google ScholarThere is no corresponding record for this reference.
- 260Ornaghi, P.; Rotili, D.; Sbardella, G.; Mai, A.; Filetici, P. Biochem. Pharmacol. 2005, 70, 911Google ScholarThere is no corresponding record for this reference.
- 261Smith, A. T.; Livingston, M. R.; Mai, A.; Filetici, P.; Queener, S. F.; Sullivan, W. J., Jr. Antimicrob. Agents Chemother. 2007, 51, 1109Google ScholarThere is no corresponding record for this reference.
- 262Mai, A.; Rotili, D.; Tarantino, D.; Ornaghi, P.; Tosi, F.; Vicidomini, C.; Sbardella, G.; Nebbioso, A.; Miceli, M.; Altucci, L.; Filetici, P. J. Med. Chem. 2006, 49, 6897Google ScholarThere is no corresponding record for this reference.
- 263Mai, A.; Rotili, D.; Tarantino, D.; Nebbioso, A.; Castellano, S.; Sbardella, G.; Tini, M.; Altucci, L. Bioorg. Med. Chem. Lett. 2009, 19, 1132Google ScholarThere is no corresponding record for this reference.
- 264Rahim, R.; Strobl, J. S. Anticancer Drugs 2009, 20, 736Google ScholarThere is no corresponding record for this reference.
- 265Tahan, F.; Jazrawi, E.; Moodley, T.; Rovati, G. E.; Adcock, I. M. Clin. Exp. Allergy 2008, 38, 805Google Scholar265Montelukast inhibits tumour necrosis factor-α-mediated interleukin-8 expression through inhibition of nuclear factor-κB p65-associated histone acetyltransferase activityTahan, F.; Jazrawi, E.; Moodley, T.; Rovati, G. E.; Adcock, I. M.Clinical and Experimental Allergy (2008), 38 (5), 805-811CODEN: CLEAEN; ISSN:0954-7894. (Blackwell Publishing Ltd.)Background: Montelukast is a potent cysteinyl leukotriene-1 receptor antagonist possessing some anti-inflammatory effects although the mol. mechanism of these anti-inflammatory effects is unknown. In this study, we aimed to investigate the effect of montelukast on nuclear factor (NF)-κB-assocd. histone acetylation activity in phorbol myristate acetate (PMA)-differentiated U937 cells. Methods: We examd. the inhibitory effects of montelukast on TNF-α-induced IL-8 prodn. in PMA-differentiated U-937 cells. U-937 cells were exposed to PMA (50 ng/mL) for 48 h to allow differentiation to macrophages. Macrophages were then exposed to TNF-α (10 ng/mL) in the presence or absence of montelukast (0.01-10 μM) for 24 h. After this time, the concn. of IL-8 in the culture supernatant was measured by sandwich-type ELISA kit. The effect of signalling pathways on TNF-α-induced IL-8 release was examd. pharmacol. using selective NF-κB/IKK2 (AS602868, 3 μM), (PD98059, 10 μM) and p38 mitogen activated protein kinase (MAPK) (SB203580, 1 μM) inhibitors. NF-κB DNA binding activity was measured by a DNA-binding ELISA-based assay. NF-κB-p65-assocd. histone acetyltransferase (HAT) activity was measured by immunopptn. linked to com. fluorescent HAT. Results: TNF-α-induced IL-8 release was suppressed by an NF-κB inhibitor but not by MEK or p38 MAPK inhibitors. Montelukast induced a concn.-dependent inhibition of TNF-α-induced IL-8 release and mRNA expression that reached a plateau at 0.1 μM without affecting cell viability. Montelukast did not affect NF-κB p65 activation as measured by DNA binding but suppressed NF-κB p65-assocd. HAT activity. Conclusion: Montelukast inhibits TNF-α-stimulated IL-8 expression through changes in NF-κB p65-assocd. HAT activity. Drugs targeting these enzymes may enhance the anti-inflammatory actions of montelukast.
- 266Chimenti, F.; Bizzarri, B.; Maccioni, E.; Secci, D.; Bolasco, A.; Chimenti, P.; Fioravanti, R.; Granese, A.; Carradori, S.; Tosi, F.; Ballario, P.; Vernarecci, S.; Filetici, P. J. Med. Chem. 2009, 52, 530Google ScholarThere is no corresponding record for this reference.
- 267Wu, J.; Wang, J.; Li, M.; Yang, Y.; Wang, B.; Zheng, Y. G. Bioorg. Chem. 2011, 39, 53Google Scholar267Small molecule inhibitors of histone acetyltransferase Tip60Wu, Jiang; Wang, Ju-Xian; Li, Min-Yong; Yang, Yu-Tao; Wang, Bing-He; Zheng, Y. GeorgeBioorganic Chemistry (2011), 39 (1), 53-58CODEN: BOCMBM; ISSN:0045-2068. (Elsevier B.V.)Tip60 is a key member of the MYST family of histone acetyltransferases and involved in a broad spectrum of cellular pathways and disease conditions. So far, small mol. inhibitors of Tip60 and other members of MYST HATs are rarely reported. To discover new small mol. inhibitors of Tip60 as mechanistic tools for functional study and as chem. leads for therapeutic development, we performed virtual screening using the crystal structure of Esa1 (the yeast homolog of Tip60) on a small mol. library database. Radioactive acetylation assays were carried out to further evaluate the virtual screen hits. Several compds. with new structural scaffolds were identified with micromolar inhibition potency for Tip60 from the biochem. studies. Further, computer modeling and kinetic assays suggest that these mols. target the acetyl-CoA binding site in Tip60. These new inhibitors provide valuable chem. hits to develop further potent inhibitors for the MYST HATs.
- 268Whitty, A. Future Med. Chem. 2011, 3, 797Google Scholar268Growing PAINS in academic drug discoveryWhitty, AdrianFuture Medicinal Chemistry (2011), 3 (7), 797-801CODEN: FMCUA7; ISSN:1756-8919. (Future Science Ltd.)In a recent article it was argued that compds. published as drug leads by academic labs. commonly contain functionality that identifies them as nonspecific pan-assay interference compds.'' (PAINS). The article raises broad questions about why best practices for hit and lead qualification that are well known in industry are not more widely employed in academia, as well as about the role of journals in publishing manuscripts that report drug leads of little potential value. Barriers to adoption of best practices for some academic drug-discovery researchers include knowledge gaps and infrastructure deficiencies, but they also arise from fundamental differences in how academic research is structured and how success is measured. Academic drug discovery should not seek to become identical to com. pharmaceutical research, but we can do a better job of assessing and communicating the true potential of the drug leads we publish, thereby reducing the wastage of resources on nonviable compds.
- 269Baell, J. B.; Holloway, G. A. J. Med. Chem. 2010, 53, 2719Google Scholar269New Substructure Filters for Removal of Pan Assay Interference Compounds (PAINS) from Screening Libraries and for Their Exclusion in BioassaysBaell, Jonathan B.; Holloway, Georgina A.Journal of Medicinal Chemistry (2010), 53 (7), 2719-2740CODEN: JMCMAR; ISSN:0022-2623. (American Chemical Society)This report describes a no. of substructural features which can help to identify compds. that appear as frequent hitters (promiscuous compds.) in many biochem. high throughput screens. The compds. identified by such substructural features are not recognized by filters commonly used to identify reactive compds. Even though these substructural features were identified using only one assay detection technol., such compds. have been reported to be active from many different assays. In fact, these compds. are increasingly prevalent in the literature as potential starting points for further exploration, whereas they may not be.
- 270Mujtaba, S.; Zeng, L.; Zhou, M. M. Oncogene 2007, 26, 5521Google ScholarThere is no corresponding record for this reference.
- 271Manning, E. T.; Ikehara, T.; Ito, T.; Kadonaga, J. T.; Kraus, W. L. Mol. Cell. Biol. 2001, 21, 3876Google Scholar271p300 forms a stable, template-committed complex with chromatin: role for the bromodomainManning, E. Tory; Ikehara, Tsuyoshi; Ito, Takashi; Kadonaga, James T.; Kraus, W. LeeMolecular and Cellular Biology (2001), 21 (12), 3876-3887CODEN: MCEBD4; ISSN:0270-7306. (American Society for Microbiology)The nature of the interaction of coactivator proteins with transcriptionally active promoters in chromatin is a fundamental question in transcriptional regulation by RNA polymerase II. In this study, we used a biochem. approach to examine the functional assocn. of the coactivator p300 with chromatin templates. Using in vitro transcription template competition assays, we obsd. that p300 forms a stable, template-committed complex with chromatin during the transcription process. The template commitment is dependent on the time of incubation of p300 with the chromatin template and occurs independently of the presence of a transcriptional activator protein. In studies examg. interactions between p300 and chromatin, we found that p300 binds directly to chromatin and that the binding requires the p300 bromodomain, a conserved 110-amino-acid sequence found in many chromatin-assocd. proteins. Furthermore, we obsd. that the isolated p300 bromodomain binds directly to histones, preferentially to histone H3. However, the isolated p300 bromodomain does not bind to nucleosomal histones under the same assay conditions, suggesting that free histones and nucleosomal histones are not equiv. as binding substrates. Collectively, our results suggest that the stable assocn. of p300 with chromatin is mediated, at least in part, by the bromodomain and is critically important for p300 function. Furthermore, our results suggest a model for p300 function that involves distinct activator-dependent targeting and activator-independent chromatin binding activities.
- 272Wei, L.; Jamonnak, N.; Choy, J.; Wang, Z.; Zheng, W. Biochem. Biophys. Res. Commun. 2008, 368, 279Google ScholarThere is no corresponding record for this reference.
- 273Polesskaya, A.; Naguibneva, I.; Duquet, A.; Bengal, E.; Robin, P.; Harel-Bellan, A. Mol. Cell. Biol. 2001, 21, 5312Google Scholar273Interaction between acetylated MyoD and the bromodomain of CBP and/or p300Polesskaya, Anna; Naguibneva, Irina; Duquet, Arnaud; Bengal, Eyal; Robin, Philippe; Harel-Bellan, AnnickMolecular and Cellular Biology (2001), 21 (16), 5312-5320CODEN: MCEBD4; ISSN:0270-7306. (American Society for Microbiology)Acetylation is emerging as a posttranslational modification of nuclear proteins that is essential to the regulation of transcription and that modifies transcription factor affinity for binding sites on DNA, stability, and/or nuclear localization. Here, we present both in vitro and in vivo evidence that acetylation increases the affinity of myogenic factor MyoD for acetyltransferases CBP and p300. In myogenic cells, the fraction of endogenous MyoD that is acetylated was found assocd. with CBP or p300. In vitro, the interaction between MyoD and CBP was more resistant to high salt concns. and was detected with lower doses of MyoD when MyoD was acetylated. Interestingly, an anal. of CBP mutants revealed that the interaction with acetylated MyoD involves the bromodomain of CBP. In live cells, MyoD mutants that cannot be acetylated did not assoc. with CBP or p300 and were strongly impaired in their ability to cooperate with CBP for transcriptional activation of a muscle creatine kinase-luciferase construct. Taken together, our data suggest a new mechanism for activation of protein function by acetylation and demonstrate for the first time an acetylation-dependent interaction between the bromodomain of CBP and a nonhistone protein.
- 274Hou, T.; Ray, S.; Lee, C.; Brasier, A. R. J. Biol. Chem. 2008, 283, 30725Google ScholarThere is no corresponding record for this reference.
- 275Mujtaba, S.; He, Y.; Zeng, L.; Yan, S.; Plotnikova, O.; Sachchidanand; Sanchez, R.; Zeleznik-Le, N. J.; Ronai, Z.; Zhou, M. M. Mol. Cell 2004, 13, 251Google ScholarThere is no corresponding record for this reference.
- 276Kim, J. H.; Cho, E. J.; Kim, S. T.; Youn, H. D. Nat. Struct. Mol. Biol. 2005, 12, 423Google Scholar276CtBP represses p300-mediated transcriptional activation by direct association with its bromodomainKim, Jae-Hwan; Cho, Eun-Jung; Kim, Seong-Tae; Youn, Hong-DukNature Structural & Molecular Biology (2005), 12 (5), 423-428CODEN: NSMBCU; ISSN:1545-9993. (Nature Publishing Group)Histone acetyltransferase coactivators bind to acetylated histones through their bromodomains and catalyze the acetylation of histone H3 and H4 tails for transcriptional activation. C-terminal binding protein (CtBP) serves as a transcriptional corepressor by recruiting histone deacetylases. However, the precise mechanism by which CtBP represses transcription has not been detd. In this study authors found that CtBP1 directly assocs. with p300 by binding to the PXDLS motif in the bromodomain of p300. Moreover, CtBP1 blocks the accessibility of p300 to histones in an NADH-sensitive manner and thus represses p300-mediated histone acetylation and transcriptional activation. In addn., an NADH-nonresponsive, monomeric mutant, CtBP1 (G183V), was found to strongly repress p300-mediated transcriptional activation. Thus, the dissocn. of NADH from CtBP1 leads to the repression of p300-driven general transcriptional activity by CtBP1. These results suggest a novel mechanism whereby CtBP1 serves as an energy-sensing repressor of histone acetyltransferase(s) and thus affects general transcription.
- 277Reynoird, N.; Schwartz, B. E.; Delvecchio, M.; Sadoul, K.; Meyers, D.; Mukherjee, C.; Caron, C.; Kimura, H.; Rousseaux, S.; Cole, P. A.; Panne, D.; French, C. A.; Khochbin, S. EMBO J. 2010, 29, 2943Google ScholarThere is no corresponding record for this reference.
- 278Chan, H. M.; La Thangue, N. B. J. Cell Sci. 2001, 114, 2363Google Scholar278p300/CBP proteins: HATs for transcriptional bridges and scaffoldsChan, Ho Man; La Thangue, Nicholas B.Journal of Cell Science (2001), 114 (13), 2363-2373CODEN: JNCSAI; ISSN:0021-9533. (Company of Biologists Ltd.)A review with 139 refs. P300/CBP transcriptional co-activator proteins play a central role in co-ordinating and integrating multiple signal-dependent events with the transcription app., allowing the appropriate level of gene activity to occur in response to diverse physiol. cues that influence, for example, proliferation, differentiation and apoptosis. P300/CBP activity can be under aberrant control in human disease, particularly in cancer, which may inactivate a p300/CBP tumor-suppressor-like activity. The transcription regulating-properties of p300 and CBP appear to be exerted through multiple mechanisms. They act as protein bridges, thereby connecting different sequence-specific transcription factors to the transcription app. Providing a protein scaffold upon which to build a multicomponent transcriptional regulatory complex is likely to be an important feature of p300/CBP control. Another key property is the presence of histone acetyltransferase (HAT) activity, which endows p300/CBP with the capacity to influence chromatin activity by modulating nucleosomal histones. Other proteins, including the p53 tumor suppressor, are targets for acetylation by p300/CBP. With the current intense level of research activity, p300/CBP will continue to be in the limelight and, we can be confident, yield new and important information on fundamental processes involved in transcriptional control.
- 279Wang, F.; Marshall, C. B.; Ikura, M. Cell. Mol. Life Sci. 2013, 70, 3989Google Scholar279Transcriptional/epigenetic regulator CBP/p300 in tumorigenesis: structural and functional versatility in target recognitionWang, Feng; Marshall, Christopher B.; Ikura, MitsuhikoCellular and Molecular Life Sciences (2013), 70 (21), 3989-4008CODEN: CMLSFI; ISSN:1420-682X. (Birkhaeuser Basel)A review. In eukaryotic cells, gene transcription is regulated by sequence-specific DNA-binding transcription factors that recognize promoter and enhancer elements near the transcriptional start site. Some coactivators promote transcription by connecting transcription factors to the basal transcriptional machinery. The highly conserved coactivators CREB-binding protein (CBP) and its paralog, E1A-binding protein (p300), each have four sep. transactivation domains (TADs) that interact with the TADs of a no. of DNA-binding transcription activators as well as general transcription factors (GTFs), thus mediating recruitment of basal transcription machinery to the promoter. Most promoters comprise multiple activator-binding sites, and many activators contain tandem TADs, thus multivalent interactions may stabilize CBP/p300 at the promoter, and intrinsically disordered regions in CBP/p300 and many activators may confer adaptability to these multivalent complexes. CBP/p300 contains a catalytic histone acetyltransferase (HAT) domain, which remodels chromatin to 'relax' its superstructure and enables transcription of proximal genes. The HAT activity of CBP/p300 also acetylates some transcription factors (e.g., p53), hence modulating the function of key transcriptional regulators. Through these numerous interactions, CBP/p300 has been implicated in complex physiol. and pathol. processes, and, in response to different signals, can drive cells towards proliferation or apoptosis. Dysregulation of the transcriptional and epigenetic functions of CBP/p300 is assocd. with leukemia and other types of cancer, thus it has been recognized as a potential anti-cancer drug target. In this review, we focus on recent exciting findings in the structural mechanisms of CBP/p300 involving multivalent and dynamic interactions with binding partners, which may pave new avenues for anti-cancer drug development.
- 280Ragvin, A.; Valvatne, H.; Erdal, S.; Arskog, V.; Tufteland, K. R.; Breen, K.; ØYan, A. M.; Eberharter, A.; Gibson, T. J.; Becker, P. B.; Aasland, R. J. Mol. Biol. 2004, 337, 773Google ScholarThere is no corresponding record for this reference.
- 281He, J.; Ye, J.; Cai, Y.; Riquelme, C.; Liu, J. O.; Liu, X.; Han, A.; Chen, L. Nucleic Acids Res. 2011, 39, 4464Google ScholarThere is no corresponding record for this reference.
- 282Zeng, L.; Zhang, Q.; Gerona-Navarro, G.; Moshkina, N.; Zhou, M. M. Structure 2008, 16, 643Google ScholarThere is no corresponding record for this reference.
- 283Mujtaba, S.; He, Y.; Zeng, L.; Yan, S.; Plotnikova, O.; Sachchidanand; Sanchez, R.; Zeleznik-Le, N. J.; Ronai, Z.; Zhou, M. M. Mol. Cell 2004, 13, 251Google ScholarThere is no corresponding record for this reference.
- 284Ferreon, J. C.; Martinez-Yamout, M. A.; Dyson, H. J.; Wright, P. E. Proc. Natl. Acad. Sci. U.S.A. 2009, 106, 13260Google ScholarThere is no corresponding record for this reference.
- 285Feng, H.; Jenkins, L. M.; Durell, S. R.; Hayashi, R.; Mazur, S. J.; Cherry, S.; Tropea, J. E.; Miller, M.; Wlodawer, A.; Appella, E.; Bai, Y. Structure 2009, 17, 202Google ScholarThere is no corresponding record for this reference.
- 286Wojciak, J. M.; Martinez-Yamout, M. A.; Dyson, H. J.; Wright, P. E. EMBO J. 2009, 28, 948Google ScholarThere is no corresponding record for this reference.
- 287Das, C.; Roy, S.; Namjoshi, S.; Malarkey, C. S.; Jones, D. N.; Kutateladze, T. G.; Churchill, M. E.; Tyler, J. K. Proc. Natl. Acad. Sci. U.S.A. 2014, 111, E1072Google ScholarThere is no corresponding record for this reference.
- 288Delvecchio, M.; Gaucher, J.; Aguilar-Gurrieri, C.; Ortega, E.; Panne, D. Nat. Struct. Mol. Biol. 2013, 20, 1040Google Scholar288Structure of the p300 catalytic core and implications for chromatin targeting and HAT regulationDelvecchio, Manuela; Gaucher, Jonathan; Aguilar-Gurrieri, Carmen; Ortega, Esther; Panne, DanielNature Structural & Molecular Biology (2013), 20 (9), 1040-1046CODEN: NSMBCU; ISSN:1545-9993. (Nature Publishing Group)CBP and p300 are histone acetyltransferases (HATs) that assoc. with and acetylate transcriptional regulators and chromatin. Mutations in their catalytic 'cores' are linked to genetic disorders, including cancer. Here we present the 2.8-Å crystal structure of the catalytic core of human p300 contg. its bromodomain, CH2 region and HAT domain. The structure reveals that the CH2 region contains a discontinuous PHD domain interrupted by a RING domain. The bromodomain, PHD, RING and HAT domains adopt an assembled configuration with the RING domain positioned over the HAT substrate-binding pocket. Disease mutations that disrupt RING attachment led to upregulation of HAT activity, thus revealing an inhibitory role for this domain. The structure provides a starting point for understanding how chromatin-substrate targeting and HAT regulation are coupled and why mutations in the p300 core lead to dysregulation.
- 289He, J.; Ye, J.; Cai, Y.; Riquelme, C.; Liu, J. O.; Liu, X.; Han, A.; Chen, L. Nucleic Acids Res. 2011, 39, 4464Google ScholarThere is no corresponding record for this reference.
- 290Feng, H.; Jenkins, L. M.; Durell, S. R.; Hayashi, R.; Mazur, S. J.; Cherry, S.; Tropea, J. E.; Miller, M.; Wlodawer, A.; Appella, E.; Bai, Y. Structure 2009, 17, 202Google ScholarThere is no corresponding record for this reference.
- 291Ferreon, J. C.; Martinez-Yamout, M. A.; Dyson, H. J.; Wright, P. E. Proc. Natl. Acad. Sci. U.S.A. 2009, 106, 13260Google ScholarThere is no corresponding record for this reference.
- 292Wojciak, J. M.; Martinez-Yamout, M. A.; Dyson, H. J.; Wright, P. E. EMBO J. 2009, 28, 948Google ScholarThere is no corresponding record for this reference.
- 293Chan, H. M.; La Thangue, N. B. J. Cell Sci. 2001, 114, 2363Google Scholar293p300/CBP proteins: HATs for transcriptional bridges and scaffoldsChan, Ho Man; La Thangue, Nicholas B.Journal of Cell Science (2001), 114 (13), 2363-2373CODEN: JNCSAI; ISSN:0021-9533. (Company of Biologists Ltd.)A review with 139 refs. P300/CBP transcriptional co-activator proteins play a central role in co-ordinating and integrating multiple signal-dependent events with the transcription app., allowing the appropriate level of gene activity to occur in response to diverse physiol. cues that influence, for example, proliferation, differentiation and apoptosis. P300/CBP activity can be under aberrant control in human disease, particularly in cancer, which may inactivate a p300/CBP tumor-suppressor-like activity. The transcription regulating-properties of p300 and CBP appear to be exerted through multiple mechanisms. They act as protein bridges, thereby connecting different sequence-specific transcription factors to the transcription app. Providing a protein scaffold upon which to build a multicomponent transcriptional regulatory complex is likely to be an important feature of p300/CBP control. Another key property is the presence of histone acetyltransferase (HAT) activity, which endows p300/CBP with the capacity to influence chromatin activity by modulating nucleosomal histones. Other proteins, including the p53 tumor suppressor, are targets for acetylation by p300/CBP. With the current intense level of research activity, p300/CBP will continue to be in the limelight and, we can be confident, yield new and important information on fundamental processes involved in transcriptional control.
- 294Simone, C.; Stiegler, P.; Forcales, S. V.; Bagella, L.; De Luca, A.; Sartorelli, V.; Giordano, A.; Puri, P. L. Oncogene 2004, 23, 2177Google ScholarThere is no corresponding record for this reference.
- 295Saint Just Ribeiro, M.; Hansson, M. L.; Wallberg, A. E. Biochem. J. 2007, 404, 289Google ScholarThere is no corresponding record for this reference.
- 296Tax, F. E.; Thomas, J. H.; Ferguson, E. L.; Horvitz, H. R. Genetics 1997, 147, 1675Google ScholarThere is no corresponding record for this reference.
- 297Petcherski, A. G.; Kimble, J. Nature 2000, 405, 364Google ScholarThere is no corresponding record for this reference.
- 298Petcherski, A. G.; Kimble, J. Curr. Biol. 2000, 10, R471Google ScholarThere is no corresponding record for this reference.
- 299Kitagawa, M.; Oyama, T.; Kawashima, T.; Yedvobnick, B.; Kumar, A.; Matsuno, K.; Harigaya, K. Mol. Cell. Biol. 2001, 21, 4337Google Scholar299A human protein with sequence similarity to Drosophila mastermind coordinates the nuclear form of Notch and a CSL protein to build a transcriptional activator complex on target promotersKitagawa, Motoo; Oyama, Toshinao; Kawashima, Taichi; Yedvobnick, Barry; Kumar, Anumeha; Matsuno, Kenji; Harigaya, KenichiMolecular and Cellular Biology (2001), 21 (13), 4337-4346CODEN: MCEBD4; ISSN:0270-7306. (American Society for Microbiology)Mastermind (Mam) has been implicated as an important pos. regulator of the Notch signaling pathway by genetic studies using Drosophila melanogaster. Here we describe a biochem. mechanism of action of Mam within the Notch signaling pathway. Expression of a human sequence related to Drosophila Mam (hMam-1) in mammalian cells augments induction of Hairy Enhancer of split (HES) promoters by Notch signaling. HMam-1 stabilizes and participates in the DNA binding complex of the intracellular domain of human Notch1 and a CSL protein. Truncated versions of hMam-1 that can maintain an assocn. with the complex behave in a dominant neg. fashion and depress transactivation. Furthermore, Drosophila Mam forms a similar complex with the intracellular domain of Drosophila Notch and Drosophila CSL protein during activation of Enhancer of split, the Drosophila counterpart of HES. These results indicate that Mam is an essential component of the transcriptional app. of Notch signaling.
- 300Wu, L.; Aster, J. C.; Blacklow, S. C.; Lake, R.; Artavanis-Tsakonas, S.; Griffin, J. D. Nat. Genet. 2000, 26, 484Google Scholar300MAML1, a human homologue of Drosophila mastermind, is a transcriptional co-activator for NOTCH receptorsWu, Lizi; Aster, Jon C.; Blacklow, Stephen C.; Lake, Robert; Artavanis-Tsakonas, Spyros; Griffin, James D.Nature Genetics (2000), 26 (4), 484-489CODEN: NGENEC; ISSN:1061-4036. (Nature America Inc.)Notch receptors are involved in cell-fate detn. in organisms as diverse as flies, frogs and humans. In Drosophila melanogaster, loss-of-function mutations of Notch produce a "neurogenic" phenotype in which cells destined to become epidermis switch fate and differentiate to neural cells. Upon ligand activation, the intracellular domain of Notch (ICN) translocates to the nucleus, and interacts directly with the DNA-binding protein Suppressor of hairless (Su(H)) in flies, or recombination signal binding protein Jκ (RBP-Jκ) in mammals, to activate gene transcription. But the precise mechanisms of Notch-induced gene expression are not completely understood. The gene mastermind has been identified in multiple genetic screens for modifiers of Notch mutations in Drosophila. Here we clone MAML1, a human homolog of the Drosophila gene Mastermind, and show that it encodes a protein of 130 kD localizing to nuclear bodies. MAML1 binds to the ankyrin repeat domain of all four mammalian NOTCH receptors, forms a DNA-binding complex with ICN and RBP-Jκ, and amplifies NOTCH-induced transcription of HES1. These studies provide a mol. mechanism to explain the genetic links between mastermind and Notch in Drosophila and indicate that MAML1 functions as a transcriptional co-activator for NOTCH signaling.
- 301Wilson, J. J.; Kovall, R. A. Cell 2006, 124, 985Google ScholarThere is no corresponding record for this reference.
- 302Arnett, K. L.; Hass, M.; McArthur, D. G.; Ilagan, M. X.; Aster, J. C.; Kopan, R.; Blacklow, S. C. Nat. Struct. Mol. Biol. 2010, 17, 1312Google Scholar302Structural and mechanistic insights into cooperative assembly of dimeric Notch transcription complexesArnett, Kelly L.; Hass, Matthew; McArthur, Debbie G.; Ilagan, Ma Xenia G.; Aster, Jon C.; Kopan, Raphael; Blacklow, Stephen C.Nature Structural & Molecular Biology (2010), 17 (11), 1312-1317CODEN: NSMBCU; ISSN:1545-9993. (Nature Publishing Group)Ligand-induced proteolysis of Notch produces an intracellular effector domain that transduces essential signals by regulating the transcription of target genes. This function relies on the formation of transcriptional activation complexes that include intracellular Notch, a Mastermind co-activator and the transcription factor CSL bound to cognate DNA. These complexes form higher-order assemblies on paired, head-to-head CSL recognition sites. Here we report the X-ray structure of a dimeric human Notch1 transcription complex loaded on the paired site from the human HES1 promoter. The small interface between the Notch ankyrin domains could accommodate DNA bending and untwisting to allow a range of spacer lengths between the two sites. Cooperative dimerization occurred on the human and mouse Hes5 promoters at a sequence that diverged from the CSL-binding consensus at one of the sites. These studies reveal how promoter organizational features control cooperativity and, thus, the responsiveness of different promoters to Notch signaling.
- 303Nam, Y.; Sliz, P.; Song, L.; Aster, J. C.; Blacklow, S. C. Cell 2006, 124, 973Google ScholarThere is no corresponding record for this reference.
- 304Choi, S. H.; Wales, T. E.; Nam, Y.; O’Donovan, D. J.; Sliz, P.; Engen, J. R.; Blacklow, S. C. Structure 2012, 20, 340Google ScholarThere is no corresponding record for this reference.
- 305Hansson, M. L.; Popko-Scibor, A. E.; Saint Just Ribeiro, M.; Dancy, B. M.; Lindberg, M. J.; Cole, P. A.; Wallberg, A. E. Nucleic Acids Res. 2009, 37, 2996Google ScholarThere is no corresponding record for this reference.
- 306Fryer, C. J.; Lamar, E.; Turbachova, I.; Kintner, C.; Jones, K. A. Genes Dev. 2002, 16, 1397Google ScholarThere is no corresponding record for this reference.
- 307Saint Just Ribeiro, M.; Hansson, M. L.; Wallberg, A. E. Biochem. J. 2007, 404, 289Google ScholarThere is no corresponding record for this reference.
- 308Black, J. C.; Choi, J. E.; Lombardo, S. R.; Carey, M. Mol. Cell 2006, 23, 809Google ScholarThere is no corresponding record for this reference.
- 309Kim, J. H.; Cho, E. J.; Kim, S. T.; Youn, H. D. Nat. Struct. Mol. Biol. 2005, 12, 423Google Scholar309CtBP represses p300-mediated transcriptional activation by direct association with its bromodomainKim, Jae-Hwan; Cho, Eun-Jung; Kim, Seong-Tae; Youn, Hong-DukNature Structural & Molecular Biology (2005), 12 (5), 423-428CODEN: NSMBCU; ISSN:1545-9993. (Nature Publishing Group)Histone acetyltransferase coactivators bind to acetylated histones through their bromodomains and catalyze the acetylation of histone H3 and H4 tails for transcriptional activation. C-terminal binding protein (CtBP) serves as a transcriptional corepressor by recruiting histone deacetylases. However, the precise mechanism by which CtBP represses transcription has not been detd. In this study authors found that CtBP1 directly assocs. with p300 by binding to the PXDLS motif in the bromodomain of p300. Moreover, CtBP1 blocks the accessibility of p300 to histones in an NADH-sensitive manner and thus represses p300-mediated histone acetylation and transcriptional activation. In addn., an NADH-nonresponsive, monomeric mutant, CtBP1 (G183V), was found to strongly repress p300-mediated transcriptional activation. Thus, the dissocn. of NADH from CtBP1 leads to the repression of p300-driven general transcriptional activity by CtBP1. These results suggest a novel mechanism whereby CtBP1 serves as an energy-sensing repressor of histone acetyltransferase(s) and thus affects general transcription.
- 310Kraus, W. L.; Manning, E. T.; Kadonaga, J. T. Mol. Cell. Biol. 1999, 19, 8123Google Scholar310Biochemical analysis of distinct activation functions in p300 that enhance transcription initiation with chromatin templatesKraus, W. Lee; Manning, E. Tory; Kadonaga, James T.Molecular and Cellular Biology (1999), 19 (12), 8123-8135CODEN: MCEBD4; ISSN:0270-7306. (American Society for Microbiology)To investigate the mechanisms of transcriptional enhancement by the p300 coactivator, we analyzed wild-type and mutant versions of p300 with a chromatin transcription system in vitro. Estrogen receptor, NF-κB p65 plus Sp1, and Gal4-VP16 were used as different sequence-specific activators. The CH3 domain (or E1A-binding region) was found to be essential for the function of each of the activators tested. The bromodomain was also obsd. to be generally important for p300 coactivator activity, though to a lesser extent than the CH3 domain/E1A-binding region. The acetyltransferase activity and the C-terminal region (contg. the steroid receptor coactivator/p160-binding region and the glutamine-rich region) were each found to be important for activation by estrogen receptor but not for that by Gal4-VP16. The N-terminal region of p300, which had been previously found to interact with nuclear hormone receptors, was not seen to be required for any of the activators, including estrogen receptor. Single-round transcription expts. revealed that the functionally important subregions of p300 contribute to its ability to promote the assembly of transcription initiation complexes. In addn., the acetyltransferase activity of p300 was obsd. to be distinct from the broadly essential activation function of the CH3 domain/E1A-binding region. These results indicate that specific regions of p300 possess distinct activation functions that are differentially required to enhance the assembly of transcription initiation complexes. Interestingly, with the estrogen receptor, four distinct regions of p300 each have an essential role in the transcription activation process. These data exemplify a situation in which a network of multiple activation functions is required to achieve gene transcription.
- 311Delvecchio, M.; Gaucher, J.; Aguilar-Gurrieri, C.; Ortega, E.; Panne, D. Nat. Struct. Mol. Biol. 2013, 20, 1040Google Scholar311Structure of the p300 catalytic core and implications for chromatin targeting and HAT regulationDelvecchio, Manuela; Gaucher, Jonathan; Aguilar-Gurrieri, Carmen; Ortega, Esther; Panne, DanielNature Structural & Molecular Biology (2013), 20 (9), 1040-1046CODEN: NSMBCU; ISSN:1545-9993. (Nature Publishing Group)CBP and p300 are histone acetyltransferases (HATs) that assoc. with and acetylate transcriptional regulators and chromatin. Mutations in their catalytic 'cores' are linked to genetic disorders, including cancer. Here we present the 2.8-Å crystal structure of the catalytic core of human p300 contg. its bromodomain, CH2 region and HAT domain. The structure reveals that the CH2 region contains a discontinuous PHD domain interrupted by a RING domain. The bromodomain, PHD, RING and HAT domains adopt an assembled configuration with the RING domain positioned over the HAT substrate-binding pocket. Disease mutations that disrupt RING attachment led to upregulation of HAT activity, thus revealing an inhibitory role for this domain. The structure provides a starting point for understanding how chromatin-substrate targeting and HAT regulation are coupled and why mutations in the p300 core lead to dysregulation.
- 312Filippakopoulos, P.; Knapp, S. Nat. Rev. Drug Discovery 2014, 13, 337Google Scholar312Targeting bromodomains: epigenetic readers of lysine acetylationFilippakopoulos, Panagis; Knapp, StefanNature Reviews Drug Discovery (2014), 13 (5), 337-356CODEN: NRDDAG; ISSN:1474-1776. (Nature Publishing Group)A review. Lysine acetylation is a key mechanism that regulates chromatin structure; aberrant acetylation levels have been linked to the development of several diseases. Acetyl-lysine modifications create docking sites for bromodomains, which are small interaction modules found on diverse proteins, some of which have a key role in the acetylation-dependent assembly of transcriptional regulator complexes. These complexes can then initiate transcriptional programs that result in phenotypic changes. The recent discovery of potent and highly specific inhibitors for the BET (bromodomain and extra-terminal) family of bromodomains has stimulated intensive research activity in diverse therapeutic areas, particularly in oncol., where BET proteins regulate the expression of key oncogenes and anti-apoptotic proteins. In addn., targeting BET bromodomains could hold potential for the treatment of inflammation and viral infection. Here, we highlight recent progress in the development of bromodomain inhibitors, and their potential applications in drug discovery.
- 313Nicodeme, E.; Jeffrey, K. L.; Schaefer, U.; Beinke, S.; Dewell, S.; Chung, C. W.; Chandwani, R.; Marazzi, I.; Wilson, P.; Coste, H.; White, J.; Kirilovsky, J.; Rice, C. M.; Lora, J. M.; Prinjha, R. K.; Lee, K.; Tarakhovsky, A. Nature 2010, 468, 1119Google ScholarThere is no corresponding record for this reference.
- 314Mirguet, O.; Lamotte, Y.; Donche, F.; Toum, J.; Gellibert, F.; Bouillot, A.; Gosmini, R.; Nguyen, V. L.; Delannée, D.; Seal, J.; Blandel, F.; Boullay, A. B.; Boursier, E.; Martin, S.; Brusq, J. M.; Krysa, G.; Riou, A.; Tellier, R.; Costaz, A.; Huet, P.; Dudit, Y.; Trottet, L.; Kirilovsky, J.; Nicodeme, E. Bioorg. Med. Chem. Lett. 2012, 22, 2963Google ScholarThere is no corresponding record for this reference.
- 315Dawson, M. A.; Prinjha, R. K.; Dittmann, A.; Giotopoulos, G.; Bantscheff, M.; Chan, W. I.; Robson, S. C.; Chung, C. W.; Hopf, C.; Savitski, M. M.; Huthmacher, C.; Gudgin, E.; Lugo, D.; Beinke, S.; Chapman, T. D.; Roberts, E. J.; Soden, P. E.; Auger, K. R.; Mirguet, O.; Doehner, K.; Delwel, R.; Burnett, A. K.; Jeffrey, P.; Drewes, G.; Lee, K.; Huntly, B. J.; Kouzarides, T. Nature 2011, 478, 529Google ScholarThere is no corresponding record for this reference.
- 316Filippakopoulos, P.; Qi, J.; Picaud, S.; Shen, Y.; Smith, W. B.; Fedorov, O.; Morse, E. M.; Keates, T.; Hickman, T. T.; Felletar, I.; Philpott, M.; Munro, S.; McKeown, M. R.; Wang, Y.; Christie, A. L.; West, N.; Cameron, M. J.; Schwartz, B.; Heightman, T. D.; La Thangue, N.; French, C. A.; Wiest, O.; Kung, A. L.; Knapp, S.; Bradner, J. E. Nature 2010, 468, 1067Google Scholar316Selective inhibition of BET bromodomainsFilippakopoulos, Panagis; Qi, Jun; Picaud, Sarah; Shen, Yao; Smith, William B.; Fedorov, Oleg; Morse, Elizabeth M.; Keates, Tracey; Hickman, Tyler T.; Felletar, Ildiko; Philpott, Martin; Munro, Shongah; McKeown, Michael R.; Wang, Yuchuan; Christie, Amanda L.; West, Nathan; Cameron, Michael J.; Schwartz, Brian; Heightman, Tom D.; La Thangue, Nicholas; French, Christopher; Wiest, Olaf; Kung, Andrew L.; Knapp, Stefan; Bradner, James E.Nature (London, United Kingdom) (2010), 468 (7327), 1067-1073CODEN: NATUAS; ISSN:0028-0836. (Nature Publishing Group)Epigenetic proteins are intently pursued targets in ligand discovery. So far, successful efforts have been limited to chromatin modifying enzymes, or so-called epigenetic 'writers' and 'erasers'. Potent inhibitors of histone binding modules have not yet been described. Here the authors report a cell-permeable small mol. (I,JQ1) that binds competitively to acetyl-lysine recognition motifs, or bromodomains. High potency and specificity towards a subset of human bromodomains is explained by co-crystal structures with bromodomain and extra-terminal (BET) family member BRD4, revealing excellent shape complementarity with the acetyl-lysine binding cavity. Recurrent translocation of BRD4 is obsd. in a genetically-defined, incurable subtype of human squamous carcinoma. Competitive binding by JQ1 displaces the BRD4 fusion oncoprotein from chromatin, prompting squamous differentiation and specific antiproliferative effects in BRD4-dependent cell lines and patient-derived xenograft models. These data establish proof-of-concept for targeting protein-protein interactions of epigenetic 'readers', and provide a versatile chem. scaffold for the development of chem. probes more broadly throughout the bromodomain family.
- 317Fish, P. V.; Filippakopoulos, P.; Bish, G.; Brennan, P. E.; Bunnage, M. E.; Cook, A. S.; Federov, O.; Gerstenberger, B. S.; Jones, H.; Knapp, S.; Marsden, B.; Nocka, K.; Owen, D. R.; Philpott, M.; Picaud, S.; Primiano, M. J.; Ralph, M. J.; Sciammetta, N.; Trzupek, J. D. J. Med. Chem. 2012, 55, 9831Google ScholarThere is no corresponding record for this reference.
- 318Picaud, S.; Da Costa, D.; Thanasopoulou, A.; Filippakopoulos, P.; Fish, P. V.; Philpott, M.; Fedorov, O.; Brennan, P.; Bunnage, M. E.; Owen, D. R.; Bradner, J. E.; Taniere, P.; O’Sullivan, B.; Müller, S.; Schwaller, J.; Stankovic, T.; Knapp, S. Cancer Res. 2013, 73, 3336Google ScholarThere is no corresponding record for this reference.
- 319Hewings, D. S.; Wang, M.; Philpott, M.; Fedorov, O.; Uttarkar, S.; Filippakopoulos, P.; Picaud, S.; Vuppusetty, C.; Marsden, B.; Knapp, S.; Conway, S. J.; Heightman, T. D. J. Med. Chem. 2011, 54, 6761Google ScholarThere is no corresponding record for this reference.
- 320Hay, D.; Fedorov, O.; Filippakopoulos, P.; Martin, S.; Philpott, M.; Picaud, S.; Hewings, D. S.; Uttakar, S.; Heightman, T. D.; Conway, S. J.; Knapp, S.; Brennan, P. E. Med. Chem. Commun. 2013, 4, 140Google ScholarThere is no corresponding record for this reference.
- 321Bamborough, P.; Diallo, H.; Goodacre, J. D.; Gordon, L.; Lewis, A.; Seal, J. T.; Wilson, D. M.; Woodrow, M. D.; Chung, C. W. J. Med. Chem. 2012, 55, 587Google ScholarThere is no corresponding record for this reference.
- 322Boehm, D.; Calvanese, V.; Dar, R. D.; Xing, S.; Schroeder, S.; Martins, L.; Aull, K.; Li, P. C.; Planelles, V.; Bradner, J. E.; Zhou, M. M.; Siliciano, R. F.; Weinberger, L.; Verdin, E.; Ott, M. Cell Cycle 2013, 12, 452Google ScholarThere is no corresponding record for this reference.
- 323Li, Z.; Guo, J.; Wu, Y.; Zhou, Q. Nucleic Acids Res. 2013, 41, 277Google ScholarThere is no corresponding record for this reference.
- 324Zhu, J.; Gaiha, G. D.; John, S. P.; Pertel, T.; Chin, C. R.; Gao, G.; Qu, H.; Walker, B. D.; Elledge, S. J.; Brass, A. L. Cell Rep. 2012, 2, 807Google Scholar324Reactivation of latent HIV-1 by inhibition of BRD4Zhu, Jian; Gaiha, Gaurav D.; John, Sinu P.; Pertel, Thomas; Chin, Christopher R.; Gao, Geng; Qu, Hongjing; Walker, Bruce D.; Elledge, Stephen J.; Brass, Abraham L.Cell Reports (2012), 2 (4), 807-816CODEN: CREED8; ISSN:2211-1247. (Cell Press)HIV-1 depends on many host factors for propagation. Other host factors, however, antagonize HIV-1 and may have profound effects on viral activation. Curing HIV-1 requires the redn. of latent viral reservoirs that remain in the face of antiretroviral therapy. Using orthologous genetic screens, we identified bromodomain contg. 4 (BRD4) as a neg. regulator of HIV-1 replication. Antagonism of BRD4, via RNA interference or with a small mol. inhibitor, JQ1, both increased proviral transcriptional elongation and alleviated HIV-1 latency in cell-line models. In multiple instances, JQ1, when used in combination with the NF-κB activators Prostratin or PHA, enhanced the in vitro reactivation of latent HIV-1 in primary T cells. These data are consistent with a model wherein BRD4 competes with the virus for HIV-1 dependency factors (HDFs) and suggests that combinatorial therapies that activate HDFs and antagonize HIV-1 competitive factors may be useful for curing HIV-1 infection.
- 325Banerjee, C.; Archin, N.; Michaels, D.; Belkina, A. C.; Denis, G. V.; Bradner, J.; Sebastiani, P.; Margolis, D. M.; Montano, M. J. Leukocyte Biol. 2012, 92, 1147Google Scholar325BET bromodomain inhibition as a novel strategy for reactivation of HIV-1Banerjee, Camellia; Archin, Nancie; Michaels, Daniel; Belkina, Anna C.; Denis, Gerald V.; Bradner, James; Sebastiani, Paola; Margolis, David M.; Montano, MontyJournal of Leukocyte Biology (2012), 92 (6), 1147-1154CODEN: JLBIE7; ISSN:0741-5400. (Society for Leukocyte Biology)The persistence of latent HIV-1 remains a major challenge in therapeutic efforts to eradicate infection. We report the capacity for HIV reactivation by a selective small mol. inhibitor of BET family bromodomains, JQ1, a promising therapeutic agent with antioncogenic properties. JQ1 reactivated HIV transcription in models of latent T cell infection and latent monocyte infection. We also tested the effect of exposure to JQ1 to allow recovery of replication-competent HIV from pools of resting CD4+ T cells isolated from HIV-infected, ART-treated patients. In one of three patients, JQ1 allowed recovery of virus at a frequency above unstimulated conditions. JQ1 potently suppressed T cell proliferation with minimal cytotoxic effect. Transcriptional profiling of T cells with JQ1 showed potent down-regulation of T cell activation genes, including CD3, CD28, and CXCR4, similar to HDAC inhibitors, but JQ1 also showed potent up-regulation of chromatin modification genes, including SIRT1, HDAC6, and multiple lysine demethylases (KDMs). Thus, JQ1 reactivates HIV-1 while suppressing T cell activation genes and up-regulating histone modification genes predicted to favor increased Tat activity. Thus, JQ1 may be useful in studies of potentially novel mechanisms for transcriptional control as well as in translational efforts to identify therapeutic mols. to achieve viral eradication.
- 326Wang, X.; Li, J.; Schowalter, R. M.; Jiao, J.; Buck, C. B.; You, J. PLoS Pathog. 2012, 8, e1003021Google ScholarThere is no corresponding record for this reference.
- 327Goupille, O.; Penglong, T.; Lefèvre, C.; Granger, M.; Kadri, Z.; Fucharoen, S.; Maouche-Chrétien, L.; Leboulch, P.; Chrétien, S. Biochem. Biophys. Res. Commun. 2012, 429, 1Google ScholarThere is no corresponding record for this reference.
- 328Ott, C. J.; Kopp, N.; Bird, L.; Paranal, R. M.; Qi, J.; Bowman, T.; Rodig, S. J.; Kung, A. L.; Bradner, J. E.; Weinstock, D. M. Blood 2012, 120, 2843Google ScholarThere is no corresponding record for this reference.
- 329Picaud, S.; Da Costa, D.; Thanasopoulou, A.; Filippakopoulos, P.; Fish, P. V.; Philpott, M.; Fedorov, O.; Brennan, P.; Bunnage, M. E.; Owen, D. R.; Bradner, J. E.; Taniere, P.; O’Sullivan, B.; Müller, S.; Schwaller, J.; Stankovic, T.; Knapp, S. Cancer Res. 2013, 73, 3336Google ScholarThere is no corresponding record for this reference.
- 330Dawson, M. A.; Prinjha, R. K.; Dittmann, A.; Giotopoulos, G.; Bantscheff, M.; Chan, W. I.; Robson, S. C.; Chung, C. W.; Hopf, C.; Savitski, M. M.; Huthmacher, C.; Gudgin, E.; Lugo, D.; Beinke, S.; Chapman, T. D.; Roberts, E. J.; Soden, P. E.; Auger, K. R.; Mirguet, O.; Doehner, K.; Delwel, R.; Burnett, A. K.; Jeffrey, P.; Drewes, G.; Lee, K.; Huntly, B. J.; Kouzarides, T. Nature 2011, 478, 529Google ScholarThere is no corresponding record for this reference.
- 331Zuber, J.; Shi, J.; Wang, E.; Rappaport, A. R.; Herrmann, H.; Sison, E. A.; Magoon, D.; Qi, J.; Blatt, K.; Wunderlich, M.; Taylor, M. J.; Johns, C.; Chicas, A.; Mulloy, J. C.; Kogan, S. C.; Brown, P.; Valent, P.; Bradner, J. E.; Lowe, S. W.; Vakoc, C. R. Nature 2011, 478, 524Google ScholarThere is no corresponding record for this reference.
- 332Mertz, J. A.; Conery, A. R.; Bryant, B. M.; Sandy, P.; Balasubramanian, S.; Mele, D. A.; Bergeron, L.; Sims, R. J., III. Proc. Natl. Acad. Sci. U.S.A. 2011, 108, 16669Google ScholarThere is no corresponding record for this reference.
- 333Delmore, J. E.; Issa, G. C.; Lemieux, M. E.; Rahl, P. B.; Shi, J.; Jacobs, H. M.; Kastritis, E.; Gilpatrick, T.; Paranal, R. M.; Qi, J.; Chesi, M.; Schinzel, A. C.; McKeown, M. R.; Heffernan, T. P.; Vakoc, C. R.; Bergsagel, P. L.; Ghobrial, I. M.; Richardson, P. G.; Young, R. A.; Hahn, W. C.; Anderson, K. C.; Kung, A. L.; Bradner, J. E.; Mitsiades, C. S. Cell 2011, 146, 904Google ScholarThere is no corresponding record for this reference.
- 334Lockwood, W. W.; Zejnullahu, K.; Bradner, J. E.; Varmus, H. Proc. Natl. Acad. Sci. U.S.A. 2012, 109, 19408Google ScholarThere is no corresponding record for this reference.
- 335Bandukwala, H. S.; Gagnon, J.; Togher, S.; Greenbaum, J. A.; Lamperti, E. D.; Parr, N. J.; Molesworth, A. M.; Smithers, N.; Lee, K.; Witherington, J.; Tough, D. F.; Prinjha, R. K.; Peters, B.; Rao, A. Proc. Natl. Acad. Sci. U.S.A. 2012, 109, 14532Google ScholarThere is no corresponding record for this reference.
- 336Mirguet, O.; Lamotte, Y.; Donche, F.; Toum, J.; Gellibert, F.; Bouillot, A.; Gosmini, R.; Nguyen, V. L.; Delannée, D.; Seal, J.; Blandel, F.; Boullay, A. B.; Boursier, E.; Martin, S.; Brusq, J. M.; Krysa, G.; Riou, A.; Tellier, R.; Costaz, A.; Huet, P.; Dudit, Y.; Trottet, L.; Kirilovsky, J.; Nicodeme, E. Bioorg. Med. Chem. Lett. 2012, 22, 2963Google ScholarThere is no corresponding record for this reference.
- 337Bamborough, P.; Diallo, H.; Goodacre, J. D.; Gordon, L.; Lewis, A.; Seal, J. T.; Wilson, D. M.; Woodrow, M. D.; Chung, C. W. J. Med. Chem. 2012, 55, 587Google ScholarThere is no corresponding record for this reference.
- 338Zhang, G.; Liu, R.; Zhong, Y.; Plotnikov, A. N.; Zhang, W.; Zeng, L.; Rusinova, E.; Gerona-Nevarro, G.; Moshkina, N.; Joshua, J.; Chuang, P. Y.; Ohlmeyer, M.; He, J. C.; Zhou, M. M. J. Biol. Chem. 2012, 287, 28840– 51Google ScholarThere is no corresponding record for this reference.
- 339Matzuk, M. M.; McKeown, M. R.; Filippakopoulos, P.; Li, Q.; Ma, L.; Agno, J. E.; Lemieux, M. E.; Picaud, S.; Yu, R. N.; Qi, J.; Knapp, S.; Bradner, J. E. Cell 2012, 150, 673Google ScholarThere is no corresponding record for this reference.
- 340Ferguson, F. M.; Fedorov, O.; Chaikuad, A.; Philpott, M.; Muniz, J. R.; Felletar, I.; von Delft, F.; Heightman, T.; Knapp, S.; Abell, C.; Ciulli, A. J. Med. Chem. 2013, 56, 10183Google ScholarThere is no corresponding record for this reference.
- 341Zhang, W.; Prakash, C.; Sum, C.; Gong, Y.; Li, Y.; Kwok, J. J.; Thiessen, N.; Pettersson, S.; Jones, S. J.; Knapp, S.; Yang, H.; Chin, K. C. J. Biol. Chem. 2012, 287, 43137Google ScholarThere is no corresponding record for this reference.
- 342Bartholomeeusen, K.; Xiang, Y.; Fujinaga, K.; Peterlin, B. M. J. Biol. Chem. 2012, 287, 36609Google ScholarThere is no corresponding record for this reference.
- 343Devaiah, B. N.; Lewis, B. A.; Cherman, N.; Hewitt, M. C.; Albrecht, B. K.; Robey, P. G.; Ozato, K.; Sims, R. J., III; Singer, D. S. Proc. Natl. Acad. Sci. U.S.A. 2012, 109, 6927Google ScholarThere is no corresponding record for this reference.
- 344Palermo, R. D.; Webb, H. M.; West, M. J. PLoS Pathog. 2011, 7, e1002334Google ScholarThere is no corresponding record for this reference.
- 345Delmore, J. E.; Issa, G. C.; Lemieux, M. E.; Rahl, P. B.; Shi, J.; Jacobs, H. M.; Kastritis, E.; Gilpatrick, T.; Paranal, R. M.; Qi, J.; Chesi, M.; Schinzel, A. C.; McKeown, M. R.; Heffernan, T. P.; Vakoc, C. R.; Bergsagel, P. L.; Ghobrial, I. M.; Richardson, P. G.; Young, R. A.; Hahn, W. C.; Anderson, K. C.; Kung, A. L.; Bradner, J. E.; Mitsiades, C. S. Cell 2011, 146, 904Google ScholarThere is no corresponding record for this reference.
- 346Sachchidanand; Resnick-Silverman, L.; Yan, S.; Mutjaba, S.; Liu, W. J.; Zeng, L.; Manfredi, J. J.; Zhou, M. M. Chem. Biol. 2006, 13, 81Google Scholar346Target Structure-Based Discovery of Small Molecules that Block Human p53 and CREB Binding Protein AssociationSachchidanand; Resnick-Silverman, Lois; Yan, Sherry; Mutjaba, Shiraz; Liu, Wen-jun; Zeng, Lei; Manfredi, James J.; Zhou, Ming-MingChemistry & Biology (Cambridge, MA, United States) (2006), 13 (1), 81-90CODEN: CBOLE2; ISSN:1074-5521. (Cell Press)Lysine acetylation of human tumor suppressor p53 in response to cellular stress signals is required for its function as a transcription factor that regulates cell cycle arrest, senescence, or apoptosis. Here, we report small mols. that block lysine 382-acetylated p53 assocn. with the bromodomain of the coactivator CBP, an interaction essential for p53-induced transcription of the cell cycle inhibitor p21 in response to DNA damage. These chems. were discovered in target structure-guided NMR spectroscopy screening of a focused chem. library constructed based on the structural knowledge of CBP bromodomain/p53-AcK382 binding. Structural characterization shows that these chems. inhibit CBP/p53 assocn. by binding to the acetyl-lysine binding site of the bromodomain. Cell-based functional assays demonstrate that the lead chems. can modulate p53 stability and function in response to doxorubicin-induced DNA damage.
- 347Borah, J. C.; Mujtaba, S.; Karakikes, I.; Zeng, L.; Muller, M.; Patel, J.; Moshkina, N.; Morohashi, K.; Zhang, W.; Gerona-Navarro, G.; Hajjar, R. J.; Zhou, M. M. Chem. Biol. 2011, 18, 531Google Scholar347A Small Molecule Binding to the Coactivator CREB-Binding Protein Blocks Apoptosis in CardiomyocytesBorah, Jagat C.; Mujtaba, Shiraz; Karakikes, Ioannis; Zeng, Lei; Muller, Michaela; Patel, Jigneshkumar; Moshkina, Natasha; Morohashi, Keita; Zhang, Weijia; Gerona-Navarro, Guillermo; Hajjar, Roger J.; Zhou, Ming-MingChemistry & Biology (Cambridge, MA, United States) (2011), 18 (4), 531-541CODEN: CBOLE2; ISSN:1074-5521. (Cell Press)Summary: As a master transcription factor in cellular responses to external stress, tumor suppressor p53 is tightly regulated. Excessive p53 activity during myocardial ischemia causes irreversible cellular injury and cardiomyocyte death. P53 activation is dependent on lysine acetylation by the lysine acetyltransferase and transcriptional coactivator CREB-binding protein (CBP) and on acetylation-directed CBP recruitment for p53 target gene expression. Here, we report a small mol. ischemin, developed with a structure-guided approach to inhibit the acetyl-lysine binding activity of the bromodomain of CBP. We show that ischemin alters post-translational modifications on p53 and histones, inhibits p53 interaction with CBP and transcriptional activity in cells, and prevents apoptosis in ischemic cardiomyocytes. Our study suggests small mol. modulation of acetylation-mediated interactions in gene transcription as a new approach to therapeutic interventions of human disorders such as myocardial ischemia.
- 348Gerona-Navarro, G.; Yoel-Rodríguez; Mujtaba, S.; Frasca, A.; Patel, J.; Zeng, L.; Plotnikov, A. N.; Osman, R.; Zhou, M. M. J. Am. Chem. Soc. 2011, 133, 2040Google ScholarThere is no corresponding record for this reference.
- 349Rooney, T. P.; Filippakopoulos, P.; Fedorov, O.; Picaud, S.; Cortopassi, W. A.; Hay, D. A.; Martin, S.; Tumber, A.; Rogers, C. M.; Philpott, M.; Wang, M.; Thompson, A. L.; Heightman, T. D.; Pryde, D. C.; Cook, A.; Paton, R. S.; Müller, S.; Knapp, S.; Brennan, P. E.; Conway, S. J. Angew. Chem., Int. Ed. 2014, 53, 6126Google ScholarThere is no corresponding record for this reference.
- 350Fedorov, O.; Lingard, H.; Wells, C.; Monteiro, O. P.; Picaud, S.; Keates, T.; Yapp, C.; Philpott, M.; Martin, S. J.; Felletar, I.; Marsden, B. D.; Filippakopoulos, P.; Müller, S.; Knapp, S.; Brennan, P. E. J. Med. Chem. 2014, 57, 462Google ScholarThere is no corresponding record for this reference.
- 351Chung, C. W.; Dean, A. W.; Woolven, J. M.; Bamborough, P. J. Med. Chem. 2012, 55, 576Google ScholarThere is no corresponding record for this reference.
- 352Hewings, D. S.; Wang, M.; Philpott, M.; Fedorov, O.; Uttarkar, S.; Filippakopoulos, P.; Picaud, S.; Vuppusetty, C.; Marsden, B.; Knapp, S.; Conway, S. J.; Heightman, T. D. J. Med. Chem. 2011, 54, 6761Google ScholarThere is no corresponding record for this reference.
- 353Hay, D. A.; Fedorov, O.; Martin, S.; Singleton, D. C.; Tallant, C.; Wells, C.; Picaud, S.; Philpott, M.; Monteiro, O. P.; Rogers, C. M.; Conway, S. J.; Rooney, T. P.; Tumber, A.; Yapp, C.; Filippakopoulos, P.; Bunnage, M. E.; Müller, S.; Knapp, S.; Schofield, C. J.; Brennan, P. E. J. Am. Chem. Soc. 2014, 136, 9308Google ScholarThere is no corresponding record for this reference.
- 354Dey, A.; Wong, E. T.; Cheok, C. F.; Tergaonkar, V.; Lane, D. P. Cell Death Differ. 2008, 15, 263Google Scholar354R-Roscovitine simultaneously targets both the p53 and NF-κB pathways and causes potentiation of apoptosis: implications in cancer therapyDey, A.; Wong, E. T.; Cheok, C. F.; Tergaonkar, V.; Lane, D. P.Cell Death and Differentiation (2008), 15 (2), 263-273CODEN: CDDIEK; ISSN:1350-9047. (Nature Publishing Group)Seliciclib (CYC202, R-Roscovitine) is a 2, 6, 9-substituted purine analog that is currently in phase II clin. trials as an anticancer agent. We show in this study that R-Roscovitine can downregulate nuclear factor-kappa B (NF-κB) activation in response to tumor necrosis factor (TNF)α and interleukin 1. Activation of p53-dependent transcription is not compromised when R-Roscovitine is combined with TNFα. We characterize the mol. mechanism governing NF-κB repression and show that R-Roscovitine inhibits the IκB kinase (IKK) kinase activity, which leads to defective IκBα phosphorylation, degrdn. and hence nuclear function of NF-κB. We further show that the downregulation of the NF-κB pathway is also at the level of p65 modification and that the phosphorylation of p65 at Ser 536 is repressed by R-Roscovitine. Consistent with repression of canonical IKK signaling pathway, the induction of NF-κB target genes monocyte chemoattractant protein, intercellular adhesion mol.-1, cyclooxygenase-2 and IL-8 is also inhibited by R-Roscovitine. We further show that treatment of cells with TNFα and R-Roscovitine causes potentiation of cell death. Based on these results, we suggest the potential use of R-Roscovitine as a bitargeted anticancer drug that functions by simultaneously causing p53 activation and NF-κB suppression. This study also provides mechanistic insight into the mol. mechanism of action of R-Roscovitine, thereby possibly explaining its anti-inflammatory properties. Cell Death and Differentiation (2008) 15, 263-273; doi:10.1038/sj.cdd.4402257; published online 2 Nov. 2007.
- 355Li, B. X.; Xiao, X. ChemBioChem 2009, 10, 2721Google ScholarThere is no corresponding record for this reference.
- 356Best, J. L.; Amezcua, C. A.; Mayr, B.; Flechner, L.; Murawsky, C. M.; Emerson, B.; Zor, T.; Gardner, K. H.; Montminy, M. Proc. Natl. Acad. Sci. U.S.A. 2004, 101, 17622Google ScholarThere is no corresponding record for this reference.
- 357Minter, A. R.; Brennan, B. B.; Mapp, A. K. J. Am. Chem. Soc. 2004, 126, 10504Google ScholarThere is no corresponding record for this reference.
- 358Buhrlage, S. J.; Bates, C. A.; Rowe, S. P.; Minter, A. R.; Brennan, B. B.; Majmudar, C. Y.; Wemmer, D. E.; Al-Hashimi, H.; Mapp, A. K. ACS Chem. Biol. 2009, 4, 335Google ScholarThere is no corresponding record for this reference.
- 359Bates, C. A.; Pomerantz, W. C.; Mapp, A. K. Biopolymers 2011, 95, 17Google ScholarThere is no corresponding record for this reference.
- 360Rowe, S. P.; Casey, R. J.; Brennan, B. B.; Buhrlage, S. J.; Mapp, A. K. J. Am. Chem. Soc. 2007, 129, 10654Google ScholarThere is no corresponding record for this reference.
- 361Henderson, W. R., Jr.; Chi, E. Y.; Ye, X.; Nguyen, C.; Tien, Y. T.; Zhou, B.; Borok, Z.; Knight, D. A.; Kahn, M. Proc. Natl. Acad. Sci. U.S.A. 2010, 107, 14309Google ScholarThere is no corresponding record for this reference.
- 362Emami, K. H.; Nguyen, C.; Ma, H.; Kim, D. H.; Jeong, K. W.; Eguchi, M.; Moon, R. T.; Teo, J. L.; Kim, H. Y.; Moon, S. H.; Ha, J. R.; Kahn, M. Proc. Natl. Acad. Sci. U.S.A. 2004, 101, 12682Google ScholarThere is no corresponding record for this reference.
- 363Sasaki, T.; Hwang, H.; Nguyen, C.; Kloner, R. A.; Kahn, M. PLoS One 2013, 8, e75010Google ScholarThere is no corresponding record for this reference.
- 364Hao, S.; He, W.; Li, Y.; Ding, H.; Hou, Y.; Nie, J.; Hou, F. F.; Kahn, M.; Liu, Y. J. Am. Soc. Nephrol. 2011, 22, 1642Google Scholar364Targeted inhibition of β-catenin/CBP signaling ameliorates renal interstitial fibrosisHao, Sha; He, Weichun; Li, Yingjian; Ding, Hong; Hou, Yayi; Nie, Jing; Hou, Fan Fan; Kahn, Michael; Liu, YouhuaJournal of the American Society of Nephrology (2011), 22 (9), 1642-1653CODEN: JASNEU; ISSN:1046-6673. (American Society of Nephrology)Because fibrotic kidneys exhibit aberrant activation of β-catenin signaling, this pathway may be a potential target for antifibrotic therapy. In this study, we examd. the effects of β-catenin activation on tubular epithelial-mesenchymal transition (EMT) in vitro and evaluated the therapeutic efficacy of the peptidomimetic small mol. ICG-001, which specifically disrupts β-catenin-mediated gene transcription, in obstructive nephropathy. In vitro, ectopic expression of stabilized β-catenin in tubular epithelial (HKC-8) cells suppressed E-cadherin and induced Snail1, fibronectin, and plasminogen activator inhibitor-1 (PAI-1) expression. ICG-001 suppressed β-catenin-driven gene transcription in a dose-dependent manner and abolished TGF-β1-induced expression of Snail1, PAI-1, collagen I, fibronectin, and α-smooth muscle actin (α-SMA). This antifibrotic effect of ICG-001 did not involve disruption of Smad signaling. In the unilateral ureteral obstruction model, ICG-001 ameliorated renal interstitial fibrosis and suppressed renal expression of fibronectin, collagen I, collagen III, α-SMA, PAI-1, fibroblast-specific protein-1, Snail1, and Snail2. Late administration of ICG-001 also effectively attenuated fibrotic lesions in obstructive nephropathy. In conclusion, inhibiting β-catenin signaling may be an effective approach to the treatment of fibrotic kidney diseases.
- 365Majmudar, C. Y.; Højfeldt, J. W.; Arevang, C. J.; Pomerantz, W. C.; Gagnon, J. K.; Schultz, P. J.; Cesa, L. C.; Doss, C. H.; Rowe, S. P.; Vásquez, V.; Tamayo-Castillo, G.; Cierpicki, T.; Brooks, C. L., III; Sherman, D. H.; Mapp, A. K. Angew. Chem., Int. Ed. 2012, 51, 11258Google ScholarThere is no corresponding record for this reference.
- 366Kung, A. L.; Zabludoff, S. D.; France, D. S.; Freedman, S. J.; Tanner, E. A.; Vieira, A.; Cornell-Kennon, S.; Lee, J.; Wang, B.; Wang, J.; Memmert, K.; Naegeli, H. U.; Petersen, F.; Eck, M. J.; Bair, K. W.; Wood, A. W.; Livingston, D. M. Cancer Cell 2004, 6, 33Google Scholar366Small molecule blockade of transcriptional coactivation of the hypoxia-inducible factor pathwayKung, Andrew L.; Zabludoff, Sonya D.; France, Dennis S.; Freedman, Steven J.; Tanner, Elizabeth A.; Vieira, Annelisa; Cornell-Kennon, Susan; Lee, Jennifer; Wang, Beqing; Wang, Jamin; Memmert, Klaus; Naegeli, Hans-Ulrich; Petersen, Frank; Eck, Michael J.; Bair, Kenneth W.; Wood, Alexander W.; Livingston, David M.Cancer Cell (2004), 6 (1), 33-43CODEN: CCAECI; ISSN:1535-6108. (Cell Press)Homeostasis under hypoxic conditions is maintained through a coordinated transcriptional response mediated by the hypoxia-inducible factor (HIF) pathway and requires coactivation by the CBP and p300 transcriptional coactivators. Through a target-based high-throughput screen, the authors identified chetomin as a disrupter of HIF binding to p300. At a mol. level, chetomin disrupts the structure of the CH1 domain of p300 and precludes its interaction with HIF, thereby attenuating hypoxia-inducible transcription. Systemic administration of chetomin inhibited hypoxia-inducible transcription within tumors and inhibited tumor growth. These results demonstrate a therapeutic window for pharmacol. attenuation of HIF activity and further establish the feasibility of disrupting a signal transduction pathway by targeting the function of a transcriptional coactivator with a small mol.
- 367Block, K. M.; Wang, H.; Szabó, L. Z.; Polaske, N. W.; Henchey, L. K.; Dubey, R.; Kushal, S.; László, C. F.; Makhoul, J.; Song, Z.; Meuillet, E. J.; Olenyuk, B. Z. J. Am. Chem. Soc. 2009, 131, 18078Google ScholarThere is no corresponding record for this reference.
- 368Yin, S.; Kaluz, S.; Devi, N. S.; Jabbar, A. A.; de Noronha, R. G.; Mun, J.; Zhang, Z.; Boreddy, P. R.; Wang, W.; Wang, Z.; Abbruscato, T.; Chen, Z.; Olson, J. J.; Zhang, R.; Goodman, M. M.; Nicolaou, K. C.; Van Meir, E. G. Clin. Cancer Res. 2012, 18, 6623Google Scholar368Arylsulfonamide KCN1 Inhibits In Vivo Glioma Growth and Interferes with HIF Signaling by Disrupting HIF-1α Interaction with Cofactors p300/CBPYin, Shaoman; Kaluz, Stefan; Devi, Narra S.; Jabbar, Adnan A.; de Noronha, Rita G.; Mun, Jiyoung; Zhang, Zhaobin; Boreddy, Purushotham R.; Wang, Wei; Wang, Zhibo; Abbruscato, Thomas; Chen, Zhengjia; Olson, Jeffrey J.; Zhang, Ruiwen; Goodman, Mark M.; Nicolaou, K. C.; Van Meir, Erwin G.Clinical Cancer Research (2012), 18 (24), 6623-6633CODEN: CCREF4; ISSN:1078-0432. (American Association for Cancer Research)Purpose: The hypoxia-inducible factor-1 (HIF-1) plays a crit. role in tumor adaptation to hypoxia, and its elevated expression correlates with poor prognosis and treatment failure in patients with cancer. In this study, we detd. whether 3,4-dimethoxy-N-[(2,2-dimethyl-2H-chromen-6-yl)methyl]-N-phenylbenzenesulfonamide, KCN1, the lead inhibitor in a novel class of arylsulfonamide inhibitors of the HIF-1 pathway, had antitumorigenic properties in vivo and further defined its mechanism of action. Exptl. Design: We studied the inhibitory effect of systemic KCN1 delivery on the growth of human brain tumors in mice. To define mechanisms of KCN1 anti-HIF activities, we examd. its influence on the assembly of a functional HIF-1α/HIF-1β/p300 transcription complex. Results: KCN1 specifically inhibited HIF reporter gene activity in several glioma cell lines at the nanomolar level. KCN1 also downregulated transcription of endogenous HIF-1 target genes, such as VEGF, Glut-1, and carbonic anhydrase 9, in a hypoxia-responsive element (HRE)-dependent manner. KCN1 potently inhibited the growth of s.c. malignant glioma tumor xenografts with minimal adverse effects on the host. It also induced a temporary survival benefit in an intracranial model of glioma but had no effect in a model of melanoma metastasis to the brain. Mechanistically, KCN1 did not downregulate the levels of HIF-1α or other components of the HIF transcriptional complex; rather, it antagonized hypoxia-inducible transcription by disrupting the interaction of HIF-1α with transcriptional coactivators p300/CBP. Conclusions: Our results suggest that the new HIF pathway inhibitor KCN1 has antitumor activity in mouse models, supporting its further translation for the treatment of human tumors displaying hypoxia or HIF overexpression.
- 369Tanaka, Y.; Naruse, I.; Maekawa, T.; Masuya, H.; Shiroishi, T.; Ishii, S. Proc. Natl. Acad. Sci. U.S.A. 1997, 94, 10215Google ScholarThere is no corresponding record for this reference.
- 370Yao, T. P.; Oh, S. P.; Fuchs, M.; Zhou, N. D.; Ch’ng, L. E.; Newsome, D.; Bronson, R. T.; Li, E.; Livingston, D. M.; Eckner, R. Cell 1998, 93, 361Google ScholarThere is no corresponding record for this reference.
- 371Roelfsema, J. H.; Peters, D. J. Expert Rev. Mol. Med. 2007, 9, 1Google Scholar371Rubinstein-Taybi syndrome: clinical and molecular overviewRoelfsema Jeroen H; Peters Dorien J MExpert reviews in molecular medicine (2007), 9 (23), 1-16 ISSN:.Rubinstein-Taybi syndrome is characterised by mental retardation, growth retardation and a particular dysmorphology. The syndrome is rare, with a frequency of approximately one affected individual in 100,000 newborns. Mutations in two genes - CREBBP and EP300 - have been identified to cause the syndrome. These two genes show strong homology and encode histone acetyltransferases (HATs), which are transcriptional co-activators involved in many signalling pathways. Loss of HAT activity is sufficient to account for the phenomena seen in Rubinstein-Taybi patients. Although some mutations found in CREBBP are translocations, inversions and large deletions, most are point mutations or small deletions and insertions. Mutations in EP300 are comparatively rare. Extensive screening of patients has revealed mutations in CREBBP and EP300 in around 50% of cases. The cause of the syndrome in the remaining patients remains to be identified, but other genes could also be involved. Here, we describe the clinical presentation of Rubinstein-Taybi syndrome, review the mutation spectrum and discuss the current understanding of causative molecular mechanisms.
- 372Qaksen, H.; Bartsch, O.; Okur, M.; Temel, H.; Açikgoz, M.; Yilmaz, C. Genet. Couns. 2009, 20, 255Google ScholarThere is no corresponding record for this reference.
- 373Viosca, J.; Lopez-Atalaya, J. P.; Olivares, R.; Eckner, R.; Barco, A. Neurobiol. Dis. 2010, 37, 186Google Scholar373Syndromic features and mild cognitive impairment in mice with genetic reduction on p300 activity: Differential contribution of p300 and CBP to Rubinstein-Taybi syndrome etiologyViosca Jose; Lopez-Atalaya Jose P; Olivares Roman; Eckner Richard; Barco AngelNeurobiology of disease (2010), 37 (1), 186-94 ISSN:.Rubinstein-Taybi syndrome (RSTS) is a complex autosomal-dominant disease characterized by mental and growth retardation and skeletal abnormalities. A majority of the individuals diagnosed with RSTS carry heterozygous mutation in the gene CREBBP, but a small percentage of cases are caused by mutations in EP300. To investigate the contribution of p300 to RSTS pathoetiology, we carried out a comprehensive and multidisciplinary characterization of p300(+/-) mice. These mice exhibited facial abnormalities and impaired growth, two traits associated to RSTS in humans. We also observed abnormal gait, reduced swimming speed, enhanced anxiety in the elevated plus maze, and mild cognitive impairment during the transfer task in the water maze. These analyses demonstrate that p300(+/-) mice exhibit phenotypes that are reminiscent of neurological traits observed in RSTS patients, but their comparison with previous studies on CBP deficient strains also indicates that, in agreement with the most recent findings in human patients, the activity of p300 in cognition is likely less relevant or more susceptible to compensation than the activity of CBP.
- 374Roelfsema, J. H.; Peters, D. J. Expert Rev. Mol. Med. 2007, 9, 1Google Scholar374Rubinstein-Taybi syndrome: clinical and molecular overviewRoelfsema Jeroen H; Peters Dorien J MExpert reviews in molecular medicine (2007), 9 (23), 1-16 ISSN:.Rubinstein-Taybi syndrome is characterised by mental retardation, growth retardation and a particular dysmorphology. The syndrome is rare, with a frequency of approximately one affected individual in 100,000 newborns. Mutations in two genes - CREBBP and EP300 - have been identified to cause the syndrome. These two genes show strong homology and encode histone acetyltransferases (HATs), which are transcriptional co-activators involved in many signalling pathways. Loss of HAT activity is sufficient to account for the phenomena seen in Rubinstein-Taybi patients. Although some mutations found in CREBBP are translocations, inversions and large deletions, most are point mutations or small deletions and insertions. Mutations in EP300 are comparatively rare. Extensive screening of patients has revealed mutations in CREBBP and EP300 in around 50% of cases. The cause of the syndrome in the remaining patients remains to be identified, but other genes could also be involved. Here, we describe the clinical presentation of Rubinstein-Taybi syndrome, review the mutation spectrum and discuss the current understanding of causative molecular mechanisms.
- 375Gervasini, C.; Mottadelli, F.; Ciccone, R.; Castronovo, P.; Milani, D.; Scarano, G.; Bedeschi, M. F.; Belli, S.; Pilotta, A.; Selicorni, A.; Zuffardi, O.; Larizza, L. Eur. J. Hum. Genet. 2010, 18, 768Google Scholar375High frequency of copy number imbalances in Rubinstein-Taybi patients negative to CREBBP mutational analysisGervasini, Cristina; Mottadelli, Federica; Ciccone, Roberto; Castronovo, Paola; Milani, Donatella; Scarano, Gioacchino; Bedeschi, Maria Francesca; Belli, Serena; Pilotta, Alba; Selicorni, Angelo; Zuffardi, Orsetta; Larizza, LidiaEuropean Journal of Human Genetics (2010), 18 (7), 768-775CODEN: EJHGEU; ISSN:1018-4813. (Nature Publishing Group)Rubinstein-Taybi syndrome (RSTS) is a rare autosomal dominant disorder characterized by facial dysmorphisms, growth and psychomotor development delay, and skeletal defects. The known genetic causes are point mutations or deletions of the CREBBP (50-60%) and EP300 (5%) genes. To detect chromosomal rearrangements indicating novel positional candidate RSTS genes, we used a-CGH to study 26 patients fulfilling the diagnostic criteria for RSTS who were neg. at fluorescence in situ hybridization analyses of the CREBBP and EP300 regions, and direct sequencing analyses of the CREBBP gene. We found seven imbalances (27%): Four de novo and three inherited rearrangements not reported among the copy no. variants. A de novo 7p21.1 deletion of 500 kb included the TWIST1 gene, a suggested candidate for RSTS that is responsible for the Saethre-Chotzen syndrome, an entity that enters in differential diagnosis with RSTS. A similar issue of differential diagnosis was raised by a large 4.3 Mb 2q22.3q23.1 deletion encompassing ZEB2, the gene responsible for the Mowat-Wilson syndrome, whose signs may overlap with RSTS. Positional candidate genes could not be sought in the remaining pathogenetic imbalances, because of the size of the involved region (a 9 Mb 2q24.3q31.1 deletion) and/or the relative paucity of suitable genes (a 5 Mb 3p13p12.3 duplication). One of the inherited rearrangements, the 17q11.2 379Kb duplication, represents the reciprocal event of the deletion underlying an overgrowth syndrome, both being mediated by the NF1-REP-P1 and REP-P2 sub-duplicons. The contribution of this and the other detected CNVs to the clin. RSTS phenotype is difficult to assess.
- 376Foley, P.; Bunyan, D.; Stratton, J.; Dillon, M.; Lynch, S. A. Am. J. Med. Genet., Part A 2009, 149A, 997Google Scholar376Further case of Rubinstein-Taybi syndrome due to a deletion in EP300Foley, Patricia; Bunyan, David; Stratton, John; Dillon, Michelle; Lynch, Sally AnnAmerican Journal of Medical Genetics, Part A (2009), 149A (5), 997-1000CODEN: AJMGB8; ISSN:1552-4825. (Wiley-Liss, Inc.)Rubinstein-Taybi syndrome (RSTS) is a heterogeneous disorder with approx. 45-55% of patients showing mutations in the CREB binding protein and a further 3% of patients having mutations in EP300. We report a male child with a deletion of exons 3-8 of the EP300 gene who has RSTS. He has a milder skeletal phenotype, a finding that has been described in other cases with EP300 mutations. The mother suffered from preeclampsia and HELLP syndrome in the pregnancy. She subsequently developed a mullerian tumor of her cervix 6 years after the birth of her son.
- 377Bartholdi, D.; Roelfsema, J. H.; Papadia, F.; Breuning, M. H.; Niedrist, D.; Hennekam, R. C.; Schinzel, A.; Peters, D. J. J. Med. Genet. 2007, 44, 327Google Scholar377Genetic heterogeneity in Rubinstein-Taybi syndrome: delineation of the phenotype of the first patients carrying mutations in EP300Bartholdi, Deborah; Roelfsema, Jeroen H.; Papadia, Francesco; Breuning, Martijn H.; Niedrist, Dunja; Hennekam, Raoul C.; Schinzel, Albert; Peters, Dorien J. M.Journal of Medical Genetics (2007), 44 (5), 327-333CODEN: JMDGAE; ISSN:0022-2593. (BMJ Publishing Group)Rubinstein-Taybi syndrome (RSTS) is a congenital disorder characterised by growth retardation, facial dysmorphisms, skeletal abnormalities and mental retardation. Broad thumbs and halluces are the hallmarks of the syndrome. RSTS is assocd. with chromosomal rearrangements and mutations in the CREB-binding protein gene (CREBBP), also termed CBP, encoding the CREB-binding protein. Recently, it was shown that mutations in EP300, coding for the p300 protein, also cause RSTS. CBP and EP300 are highly homologous genes, which play important roles as global transcriptional coactivators. To report the phenotype of the presently known patients with RSTS (n = 4) carrying germline mutations of EP300. The patients with EP300 mutations displayed the typical facial gestalt and malformation pattern compatible with the diagnosis of RSTS. However, three patients exhibited much milder skeletal findings on the hands and feet than typically obsd. in patients with RSTS. Part of the clin. variability in RSTS is explained by genetic heterogeneity. The diagnosis of RSTS must be expanded to include patients without broad thumbs or halluces.
- 378Tsai, A. C.; Dossett, C. J.; Walton, C. S.; Cramer, A. E.; Eng, P. A.; Nowakowska, B. A.; Pursley, A. N.; Stankiewicz, P.; Wiszniewska, J.; Cheung, S. W. Eur. J. Hum. Genet. 2011, 19, 43Google Scholar378Exon deletions of the EP300 and CREBBP genes in two children with Rubinstein-Taybi syndrome detected by aCGHTsai Anne Chun-Hui; Dossett Cherilyn J; Walton Carol S; Cramer Andrea E; Eng Patti A; Nowakowska Beata A; Pursley Amber N; Stankiewicz Pawel; Wiszniewska Joanna; Cheung Sau WaiEuropean journal of human genetics : EJHG (2011), 19 (1), 43-9 ISSN:.We demonstrate the utility of an exon coverage microarray platform in detecting intragenic deletions: one in exons 24-27 of the EP300 gene and another in exons 27 and 28 of the CREBBP gene in two patients with Rubinstein-Taybi syndrome (RSTS). RSTS is a heterogeneous disorder in which approximately 45-55% of cases result from deletion or mutations in the CREBBP gene and an unknown portion of cases result from gene changes in EP300. The first case is a 3-year-old female with an exonic deletion of the EP300 gene who has classic facial features of RSTS without the thumb and great toe anomalies, consistent with the milder skeletal phenotype that has been described in other RSTS cases with EP300 mutations. In addition, the mother of this patient also had preeclampsia during pregnancy, which has been infrequently reported. The second case is a newborn male who has the classical features of RSTS. Our results illustrate that exon-targeted array comparative genomic hybridization (aCGH) is a powerful tool for detecting clinically significant intragenic rearrangements that would be otherwise missed by aCGH platforms lacking sufficient exonic coverage or sequencing of the gene of interest.
- 379Zimmermann, N.; Acosta, A. M.; Kohlhase, J.; Bartsch, O. Eur. J. Hum. Genet. 2007, 15, 837Google Scholar379Confirmation of EP300 gene mutations as a rare cause of Rubinstein-Taybi syndromeZimmermann, Nicole; Acosta, Ana Maria Bravo Ferrer; Kohlhase, Juergen; Bartsch, OliverEuropean Journal of Human Genetics (2007), 15 (8), 837-842CODEN: EJHGEU; ISSN:1018-4813. (Nature Publishing Group)The Rubinstein-Taybi syndrome (RSTS, MIM 180849), a dominant Mendelian disorder with typical face, short stature, skeletal abnormalities, and mental retardation, is usually caused by heterozygous mutations of the CREBBP gene, but recently, EP300 gene mutations were reported in three individuals. Using quant. PCR (for the CREBBP and EP300 genes) and genomic sequencing (for the EP300 gene), we studied here 13 patients who had shown no mutation after genomic sequencing of the CREBBP gene in a previous investigation. Two new disease-causing mutations were identified, including a partial deletion of CREBBP and a 1-bp deletion in EP300, c.7100delC (p.P2366fsX2401). The 1-bp deletion represents the fourth EP300 mutation reported to date and was identified in a patient with non-classical RSTS. Based on the very similar structure of the CREBBP and EP300 genes and the higher rate of single-nucleotide polymorphisms in EP300 (2.23 per individual) as compared to CREBBP (0.71 per individual) (P>0.001, Wilcoxon test), it may be assumed that EP300 gene mutations should be as frequent as CREBBP gene mutations. Based on the location of the EP300 gene mutations identified so far (outside the histone acetyl transferase domain) and the obsd. (although not very striking) phenotypical differences with the EP300 mutations, we propose that most EP300 mutations could be assocd. with other phenotypes, not classical RSTS.
- 380Bartholdi, D.; Roelfsema, J. H.; Papadia, F.; Breuning, M. H.; Niedrist, D.; Hennekam, R. C.; Schinzel, A.; Peters, D. J. J. Med. Genet. 2007, 44, 327Google Scholar380Genetic heterogeneity in Rubinstein-Taybi syndrome: delineation of the phenotype of the first patients carrying mutations in EP300Bartholdi, Deborah; Roelfsema, Jeroen H.; Papadia, Francesco; Breuning, Martijn H.; Niedrist, Dunja; Hennekam, Raoul C.; Schinzel, Albert; Peters, Dorien J. M.Journal of Medical Genetics (2007), 44 (5), 327-333CODEN: JMDGAE; ISSN:0022-2593. (BMJ Publishing Group)Rubinstein-Taybi syndrome (RSTS) is a congenital disorder characterised by growth retardation, facial dysmorphisms, skeletal abnormalities and mental retardation. Broad thumbs and halluces are the hallmarks of the syndrome. RSTS is assocd. with chromosomal rearrangements and mutations in the CREB-binding protein gene (CREBBP), also termed CBP, encoding the CREB-binding protein. Recently, it was shown that mutations in EP300, coding for the p300 protein, also cause RSTS. CBP and EP300 are highly homologous genes, which play important roles as global transcriptional coactivators. To report the phenotype of the presently known patients with RSTS (n = 4) carrying germline mutations of EP300. The patients with EP300 mutations displayed the typical facial gestalt and malformation pattern compatible with the diagnosis of RSTS. However, three patients exhibited much milder skeletal findings on the hands and feet than typically obsd. in patients with RSTS. Part of the clin. variability in RSTS is explained by genetic heterogeneity. The diagnosis of RSTS must be expanded to include patients without broad thumbs or halluces.
- 381Lopez-Atalaya, J. P.; Gervasini, C.; Mottadelli, F.; Spena, S.; Piccione, M.; Scarano, G.; Selicorni, A.; Barco, A.; Larizza, L. J. Med. Genet. 2012, 49, 66Google Scholar381Histone acetylation deficits in lymphoblastoid cell lines from patients with Rubinstein-Taybi syndromeLopez-Atalaya, J. P.; Gervasini, C.; Mottadelli, F.; Spena, S.; Piccione, M.; Scarano, G.; Selicorni, A.; Barco, A.; Larizza, L.Journal of Medical Genetics (2012), 49 (1), 66-74CODEN: JMDGAE; ISSN:0022-2593. (BMJ Publishing Group)Background: Rubinstein-Taybi syndrome (RSTS) is a congenital neurodevelopmental disorder defined by postnatal growth deficiency, characteristic skeletal abnormalities and mental retardation and caused by mutations in the genes encoding for the transcriptional co-activators with intrinsic lysine acetyltransferase (KAT) activity CBP and p300. Previous studies have shown that neuronal histone acetylation is reduced in mouse models of RSTS. Methods: The authors identified different mutations at the CREBBP locus and generated lymphoblastoid cell lines derived from nine patients with RSTS carrying distinct CREBBP mutations that illustrate different grades of the clin. severity in the spectrum of the syndrome. They next assessed whether histone acetylation levels were altered in these cell lines. Results: The comparison of CREBBP-mutated RSTS cell lines with cell lines derived from patients with an unrelated mental retardation syndrome or healthy controls revealed significant deficits in histone acetylation, affecting primarily histone H2B and histone H2A. The most severe defects were obsd. in the lines carrying the whole deletion of the CREBBP gene and the truncating mutation, both leading to a haploinsufficiency state. Interestingly, this deficit was rescued by treatment with an inhibitor of histone deacetylases (HDACi). Conclusions: The authors' results extend to humans the seminal observations in RSTS mouse models and point to histone acetylation defects, mainly involving H2B and H2A, as relevant mol. markers of the disease.
- 382Miller, R. W.; Rubinstein, J. H. Am. J. Med. Genet. 1995, 56, 112Google Scholar382Tumors in Rubinstein-Taybi syndromeMiller R W; Rubinstein J HAmerican journal of medical genetics (1995), 56 (1), 112-5 ISSN:0148-7299.The 14 tumors reported in Rubinstein-Taybi syndrome since 1989, when added to the 22 previously reported, are beginning to show a pattern of neural and developmental tumors, especially of the head, which is malformed in the syndrome. Among the neoplasms were 12 of the nervous system: 2 each of oligodendroglioma, medulloblastoma, neuroblastoma, and benign meningioma, a pheochromocytoma, and 3 other benign tumors; 2 of nasopharyngeal rhabdomyosarcoma; and 1 each of leiomyosarcoma, seminoma, and embryonal carcinoma. Among the other benign tumors were an odontoma, a choristoma, a dermoid cyst, and 2 pilomatrixomas.
- 383Iyer, N. G.; Ozdag, H.; Caldas, C. Oncogene 2004, 23, 4225Google Scholar383p300/CBP and cancerIyer, Narayanan Gopalakrishna; Oezdag, Hilal; Caldas, CarlosOncogene (2004), 23 (24), 4225-4231CODEN: ONCNES; ISSN:0950-9232. (Nature Publishing Group)A review. P300 and cAMP response element-binding protein (CBP) are adenoviral E1A-binding proteins involved in multiple cellular processes, and function as transcriptional co-factors and histone acetyltransferases. Germline mutation of CBP results in Rubinstein-Taybi syndrome, which is characterized by an increased predisposition to childhood malignancies. Furthermore, somatic mutations of p300 and CBP occur in a no. of malignancies. Chromosome translocations target CBP and, less commonly, p300 in acute myeloid leukemia and treatment-related hematol. disorders. P300 mutations in solid tumors result in truncated p300 protein products or amino-acid substitutions in crit. protein domains, and these are often assocd. with inactivation of the second allele. A mouse model confirms that p300 and CBP function as suppressors of hematol. tumor formation. The involvement of these proteins in crit. tumorigenic pathways (including TGF-β, p53 and Rb) provides a mechanistic route as to how their inactivation could result in cancer.
- 384Tillinghast, G. W.; Partee, J.; Albert, P.; Kelley, J. M.; Burtow, K. H.; Kelly, K. Genes, Chromosomes Cancer 2003, 37, 121Google Scholar384Analysis of genetic stability at the EP300 and CREBBP loci in a panel of cancer cell linesTillinghast, Guy W.; Partee, Jason; Albert, Paul; Kelley, Jenny M.; Burtow, Kenneth H.; Kelly, KathleenGenes, Chromosomes & Cancer (2003), 37 (2), 121-131CODEN: GCCAES; ISSN:1045-2257. (Wiley-Liss, Inc.)EP300 (p300) and CREBBP (CBP) are highly related transcriptional co-activators possessing histone acetyltransferase activity. These proteins have been implicated in coordinating numerous transcriptional responses that are important in the processes of proliferation and differentiation. A role for EP300 and CREBBP as tumor suppressors in cancer has been suggested by the fact that they are targeted by viral oncogenes; there is an increased incidence of hematol. malignancies in mice monoallelic for CREBBP; and loss, albeit at a low frequency, of both EP300 alleles in epithelial cancers has been obsd. Because the level of EP300/CREBBP appears to have a crit. effect on integrating certain transcriptional processes, we sought to det. whether a loss in the combined gene dosage of EP300 and CREBBP might play a role in cancer. Accordingly, we screened a panel of 103 cell lines for loss of heterozygosity and found 35 and 51% LOH for the CREBBP and EP300 loci, resp. Concordant loss of CREBBP and EP300 was not assocd. with mutations in important regions of the remaining EP300 or CREBBP genes. In addn., there did not appear to be a statistically significant selection in cancer cells, stratified by various criteria, for the concordant loss of EP300 and CREBBP. We conclude that EP300 and CREBBP rarely act as classical tumor suppressors in human cancer.
- 385Bryan, E. J.; Jokubaitis, V. J.; Chamberlain, N. L.; Baxter, S. W.; Dawson, E.; Choong, D. Y.; Campbell, I. G. Int. J. Cancer 2002, 102, 137Google Scholar385Mutation analysis of EP300 in colon, breast and ovarian carcinomasBryan, Emma J.; Jokubaitis, Venta J.; Chamberlain, Narelle L.; Baxter, Simon W.; Dawson, Elisabeth; Choong, David Y. H.; Campbell, Ian G.International Journal of Cancer (2002), 102 (2), 137-141CODEN: IJCNAW; ISSN:0020-7136. (Wiley-Liss, Inc.)The putative tumor suppressor gene EP300 is located on chromosome 22q13 which is a region showing frequent loss of heterozygosity (LOH) in colon, breast and ovarian cancers. The authors analyzed 203 human breast, colon and ovarian primary tumors and cell lines for somatic mutations in EP300. LOH across the EP300 locus was detected in 38% of colon, 36% of breast, and 49% of ovarian primary tumors but no somatic mutations in EP300 were identified in any primary tumor. Anal. of 17 colon, 11 breast, and 11 ovarian cancer cell lines identified truncating mutations in 4 colon cancer cell lines (HCT116, HT29, LIM2405 and LIM2412). The authors confirmed the presence of a previously reported frameshift mutation in HCT116 at codon 1699 and identified a second frameshift mutation at codon 1468. Bi-allelic inactivation of EP300 was also detected in LIM2405 that harbors an insC mutation at codon 927 as well an insA mutation at codon 1468. An insA mutation at codon 1468 was identified in HT29 and a CGA>TGA mutation at codon 86 was identified in LIM2412. Both these lines were heterozygous across the EP300 locus and western blot anal. confirmed the presence of an apparently wild-type protein. The authors' study has established that genetic inactivation of EP300 is rare in primary colorectal, breast and ovarian cancers. In contrast, mutations are common among colorectal cancer cell lines with 4/17 harboring homozygous or heterozygous mutations. The rarity of EP300 mutations among these tumor types that show a high frequency of LOH across 22q13 may indicate that another gene is the target of the loss. It is possible that bi-allelic inactivation of EP300 is not necessary and that haploinsufficiency is sufficient to promote tumorigenesis. Alternatively, silencing of EP300 may be achieved by epigenetic mechanisms such as promoter methylation.
- 386Muraoka, M.; Konishi, M.; Kikuchi-Yanoshita, R.; Tanaka, K.; Shitara, N.; Chong, J. M.; Iwama, T.; Miyaki, M. Oncogene 1996, 12, 1565Google ScholarThere is no corresponding record for this reference.
- 387Iyer, N. G.; Ozdag, H.; Caldas, C. Oncogene 2004, 23, 4225Google Scholar387p300/CBP and cancerIyer, Narayanan Gopalakrishna; Oezdag, Hilal; Caldas, CarlosOncogene (2004), 23 (24), 4225-4231CODEN: ONCNES; ISSN:0950-9232. (Nature Publishing Group)A review. P300 and cAMP response element-binding protein (CBP) are adenoviral E1A-binding proteins involved in multiple cellular processes, and function as transcriptional co-factors and histone acetyltransferases. Germline mutation of CBP results in Rubinstein-Taybi syndrome, which is characterized by an increased predisposition to childhood malignancies. Furthermore, somatic mutations of p300 and CBP occur in a no. of malignancies. Chromosome translocations target CBP and, less commonly, p300 in acute myeloid leukemia and treatment-related hematol. disorders. P300 mutations in solid tumors result in truncated p300 protein products or amino-acid substitutions in crit. protein domains, and these are often assocd. with inactivation of the second allele. A mouse model confirms that p300 and CBP function as suppressors of hematol. tumor formation. The involvement of these proteins in crit. tumorigenic pathways (including TGF-β, p53 and Rb) provides a mechanistic route as to how their inactivation could result in cancer.
- 388Pasqualucci, L.; Dominguez-Sola, D.; Chiarenza, A.; Fabbri, G.; Grunn, A.; Trifonov, V.; Kasper, L. H.; Lerach, S.; Tang, H.; Ma, J.; Rossi, D.; Chadburn, A.; Murty, V. V.; Mullighan, C. G.; Gaidano, G.; Rabadan, R. Nature 2011, 471, 189Google ScholarThere is no corresponding record for this reference.
- 389Kishimoto, M.; Kohno, T.; Okudela, K.; Otsuka, A.; Sasaki, H.; Tanabe, C.; Sakiyama, T.; Hirama, C.; Kitabayashi, I.; Minna, J. D.; Takenoshita, S.; Yokota, J. Clin. Cancer Res. 2005, 11, 512Google Scholar389Mutations and deletions of the CBP gene in human lung cancerKishimoto, Masahiro; Kohno, Takashi; Okudela, Koji; Otsuka, Ayaka; Sasaki, Hiroki; Tanabe, Chikako; Sakiyama, Tokuki; Hirama, Chie; Kitabayashi, Issay; Minna, John D.; Takenoshita, Seiichi; Yokota, JunClinical Cancer Research (2005), 11 (2, Pt. 1), 512-519CODEN: CCREF4; ISSN:1078-0432. (American Association for Cancer Research)Microarray-based comparative genomic hybridization anal. led us to detect a homozygous deletion at the cAMP response element binding protein-binding protein (CBP) locus in a lung cancer cell line. Oncogenic roles of CBP had been suggested by functional and genetic studies; thus, involvement of CBP gene alterations in lung carcinogenesis was investigated by undertaking comprehensive anal. of genetic CBP alterations in human lung cancer. Fifty-nine cell lines and 95 surgical specimens of lung cancer were analyzed for mutations, homozygous and hemizygous deletions, and expression of the CBP gene. Homozygous CBP deletions, including two intragenic deletions, were detected in three (5.1%) lung cancer cell lines. CBP mutations, including missense, nonsense, and frame-shift mutations, were detected in six (10.2 %) cell lines and five (5.3%) surgical specimens of lung cancer. The wild-type CBP allele was retained in 9 of 11 cases with CBP mutations, and both the wild-type and mutant alleles were expressed in all the six cases with heterozygous CBP mutations examd. Three mutations with amino acid substitutions in the histone acetyltransferase domain caused significant redn. in transcription activation activity of CBP protein in vivo. A fraction of lung cancers carried mutations and/or deletions of the CBP gene, suggesting that genetic CBP alterations are involved in the genesis and/or progression of a subset of lung cancers.
- 390Suganuma, T.; Kawabata, M.; Ohshima, T.; Ikeda, M. A. Proc. Natl. Acad. Sci. U.S.A. 2002, 99, 13073Google ScholarThere is no corresponding record for this reference.
- 391Mullighan, C. G.; Zhang, J.; Kasper, L. H.; Lerach, S.; Payne-Turner, D.; Phillips, L. A.; Heatley, S. L.; Holmfeldt, L.; Collins-Underwood, J. R.; Ma, J.; Buetow, K. H.; Pui, C. H.; Baker, S. D. Nature 2011, 471, 235Google ScholarThere is no corresponding record for this reference.
- 392Sánchez-Molina, S.; Oliva, J. L.; García-Vargas, S.; Valls, E.; Rojas, J. M.; Martínez-Balbás, M. A. Biochem. J. 2006, 398, 215Google ScholarThere is no corresponding record for this reference.
- 393Iyer, N. G.; Ozdag, H.; Caldas, C. Oncogene 2004, 23, 4225Google Scholar393p300/CBP and cancerIyer, Narayanan Gopalakrishna; Oezdag, Hilal; Caldas, CarlosOncogene (2004), 23 (24), 4225-4231CODEN: ONCNES; ISSN:0950-9232. (Nature Publishing Group)A review. P300 and cAMP response element-binding protein (CBP) are adenoviral E1A-binding proteins involved in multiple cellular processes, and function as transcriptional co-factors and histone acetyltransferases. Germline mutation of CBP results in Rubinstein-Taybi syndrome, which is characterized by an increased predisposition to childhood malignancies. Furthermore, somatic mutations of p300 and CBP occur in a no. of malignancies. Chromosome translocations target CBP and, less commonly, p300 in acute myeloid leukemia and treatment-related hematol. disorders. P300 mutations in solid tumors result in truncated p300 protein products or amino-acid substitutions in crit. protein domains, and these are often assocd. with inactivation of the second allele. A mouse model confirms that p300 and CBP function as suppressors of hematol. tumor formation. The involvement of these proteins in crit. tumorigenic pathways (including TGF-β, p53 and Rb) provides a mechanistic route as to how their inactivation could result in cancer.
- 394Fan, S.; Ma, Y. X.; Wang, C.; Yuan, R. Q.; Meng, Q.; Wang, J. A.; Erdos, M.; Goldberg, I. D.; Webb, P.; Kushner, P. J.; Pestell, R. G.; Rosen, E. M. Cancer Res. 2002, 62, 141Google ScholarThere is no corresponding record for this reference.
- 395Liang, J.; Prouty, L.; Williams, B. J.; Dayton, M. A.; Blanchard, K. L. Blood 1998, 92, 2118Google ScholarThere is no corresponding record for this reference.
- 396Carapeti, M.; Aguiar, R. C.; Goldman, J. M.; Cross, N. C. Blood 1998, 91, 3127Google ScholarThere is no corresponding record for this reference.
- 397Kindle, K. B.; Troke, P. J.; Collins, H. M.; Matsuda, S.; Bossi, D.; Bellodi, C.; Kalkhoven, E.; Salomoni, P.; Pelicci, P. G.; Minucci, S.; Heery, D. M. Mol. Cell. Biol. 2005, 25, 988Google Scholar397MOZ-TIF2 inhibits transcription by nuclear receptors and p53 by impairment of CBP functionKindle, Karin B.; Troke, Philip J. F.; Collins, Hilary M.; Matsuda, Sachiko; Bossi, Daniela; Bellodi, Cristian; Kalkhoven, Eric; Salomoni, Paolo; Pelicci, Pier Giuseppe; Minucci, Saverio; Heery, David M.Molecular and Cellular Biology (2005), 25 (3), 988-1002CODEN: MCEBD4; ISSN:0270-7306. (American Society for Microbiology)Chromosomal rearrangements assocd. with acute myeloid leukemia (AML) include fusions of the genes encoding the acetyltransferase MOZ or MORF with genes encoding the nuclear receptor coactivator TIF2, p300, or CBP. Here we show that MOZ-TIF2 acts as a dominant inhibitor of the transcriptional activities of CBP-dependent activators such as nuclear receptors and p53. The dominant neg. property of MOZ-TIF2 requires the CBP-binding domain (activation domain 1 [AD1]), and coimmunopptn. and fluorescent resonance energy transfer expts. show that MOZ-TIF2 interacts with CBP directly in vivo. The CBP-binding domain is also required for the ability of MOZ-TIF2 to extend the proliferative potential of murine bone marrow lineage-neg. cells in vitro. We show that MOZ-TIF2 displays an aberrant nuclear distribution and that cells expressing this protein have reduced levels of cellular CBP, leading to depletion of CBP from PML bodies. In summary, our results indicate that disruption of the normal function of CBP and CBP-dependent activators is an important feature of MOZ-TIF2 action in AML.
- 398Ito, Y.; Miyazono, K. Curr. Opin. Genet. Dev. 2003, 13, 43Google ScholarThere is no corresponding record for this reference.
- 399Grossman, S. R. Eur. J. Biochem. 2001, 268, 2773Google Scholar399p300/CBP/p53 interaction and regulation of the p53 responseGrossman, Steven R.European Journal of Biochemistry (2001), 268 (10), 2773-2778CODEN: EJBCAI; ISSN:0014-2956. (Blackwell Science Ltd.)A review with 50 refs. Substantial evidence points to a crit. role for the p300/CREB binding protein (CBP) co-activators in p53 responses to DNA damage. The p300/CBP and the assocd. protein P/CAF bind to and acetylate p53 during the DNA damage response, and are needed for full p53 transactivation as well as downstream p53 effects of growth arrest and/or apoptosis. Beyond this simplistic model, p300/CBP appear to be complex integrators of signals that regulate p53, and biochem., the multipartite p53/p300/CBP interaction is equally complex. Through phys. interaction with p53, p300/CBP can both pos. and neg. regulate p53 transactivation, as well as p53 protein turnover depending on cellular context and environmental stimuli, such as DNA damage.
- 400Iyer, N. G.; Xian, J.; Chin, S. F.; Bannister, A. J.; Daigo, Y.; Aparicio, S.; Kouzarides, T.; Caldas, C. Oncogene 2007, 26, 21Google ScholarThere is no corresponding record for this reference.
- 401Galbiati, L.; Mendoza-Maldonado, R.; Gutierrez, M. I.; Giacca, M. Cell Cycle 2005, 4, 930Google ScholarThere is no corresponding record for this reference.
- 402Oike, Y.; Hata, A.; Mamiya, T.; Kaname, T.; Noda, Y.; Suzuki, M.; Yasue, H.; Nabeshima, T.; Araki, K.; Yamamura, K. Hum. Mol. Genet. 1999, 8, 387Google Scholar402Truncated CBP protein leads to classical Rubinstein-Taybi syndrome phenotypes in mice: implications for a dominant-negative mechanismOike, Yuichi; Hata, Akira; Mamiya, Takayoshi; Kaname, Tadashi; Noda, Yukihiro; Suzuki, Misao; Yasue, Hirofumi; Nabeshima, Toshitaka; Araki, Kimi; Yamamura, Ken-IchiHuman Molecular Genetics (1999), 8 (3), 387-396CODEN: HMGEE5; ISSN:0964-6906. (Oxford University Press)A mouse model of Rubinstein-Taybi syndrome (RTS) was generated by an insertional mutation into the cAMP response element-binding protein (CREB)-binding protein (CBP) gene. Heterozygous CBP-deficient mice, which had truncated CBP protein (residues 1-1084) contg. the CREB-binding domain (residues 462-661), showed clin. features of RTS, such as growth retardation (100%), retarded osseous maturation (100%), hypoplastic maxilla with narrow palate (100%), cardiac anomalies (15%) and skeletal abnormalities (7%). Truncated CBP is considered to have been acting during development as a dominant-neg. inhibitor to lead to the phenotypes of RTS in mice. Our studies with step-through-type passive avoidance tests and with fear conditioning test showed that mice were deficient in long-term memory (LTM). In contrast, short-term memory (STM) appeared to be normal. These results implicate a crucial role for CBP in mammalian LTM. Our CBP+/- mice would be an excellent model for the study of the role of CBP in development and memory storage mechanisms.
- 403Viosca, J.; Lopez-Atalaya, J. P.; Olivares, R.; Eckner, R.; Barco, A. Neurobiol. Dis. 2010, 37, 186Google Scholar403Syndromic features and mild cognitive impairment in mice with genetic reduction on p300 activity: Differential contribution of p300 and CBP to Rubinstein-Taybi syndrome etiologyViosca Jose; Lopez-Atalaya Jose P; Olivares Roman; Eckner Richard; Barco AngelNeurobiology of disease (2010), 37 (1), 186-94 ISSN:.Rubinstein-Taybi syndrome (RSTS) is a complex autosomal-dominant disease characterized by mental and growth retardation and skeletal abnormalities. A majority of the individuals diagnosed with RSTS carry heterozygous mutation in the gene CREBBP, but a small percentage of cases are caused by mutations in EP300. To investigate the contribution of p300 to RSTS pathoetiology, we carried out a comprehensive and multidisciplinary characterization of p300(+/-) mice. These mice exhibited facial abnormalities and impaired growth, two traits associated to RSTS in humans. We also observed abnormal gait, reduced swimming speed, enhanced anxiety in the elevated plus maze, and mild cognitive impairment during the transfer task in the water maze. These analyses demonstrate that p300(+/-) mice exhibit phenotypes that are reminiscent of neurological traits observed in RSTS patients, but their comparison with previous studies on CBP deficient strains also indicates that, in agreement with the most recent findings in human patients, the activity of p300 in cognition is likely less relevant or more susceptible to compensation than the activity of CBP.
- 404Kung, A. L.; Rebel, V. I.; Bronson, R. T.; Ch’ng, L. E.; Sieff, C. A.; Livingston, D. M.; Yao, T. P. Genes Dev. 2000, 14, 272Google ScholarThere is no corresponding record for this reference.
- 405Kasper, L. H.; Fukuyama, T.; Biesen, M. A.; Boussouar, F.; Tong, C.; de Pauw, A.; Murray, P. J.; van Deursen, J. M.; Brindle, P. K. Mol. Cell. Biol. 2006, 26, 789Google Scholar405Conditional knockout mice reveal distinct functions for the global transcriptional coactivators CBP and p300 in T-cell developmentKasper, Lawryn H.; Fukuyama, Tomofusa; Biesen, Michelle A.; Boussouar, Faycal; Tong, Caili; de Pauw, Antoine; Murray, Peter J.; van Deursen, Jan M. A.; Brindle, Paul K.Molecular and Cellular Biology (2006), 26 (3), 789-809CODEN: MCEBD4; ISSN:0270-7306. (American Society for Microbiology)The global transcriptional coactivators CREB-binding protein (CBP) and the closely related p300 interact with over 312 proteins, making them among the most heavily connected hubs in the known mammalian protein-protein interactome. It is largely uncertain, however, if these interactions are important in specific cell lineages of adult animals, as homozygous null mutations in either CBP or p300 result in early embryonic lethality in mice. Here we describe a Cre/LoxP conditional p300 null allele (p300flox) that allows for the temporal and tissue-specific inactivation of p300. We used mice carrying p300flax and a CBP conditional knockout allele (CBPflax) in conjunction with an Lck-Cre transgene to delete CBP and p300 starting at the CD4- CD8- double-neg. thymocyte stage of T-cell development. Loss of either p300 or CBP led to a decrease in CD4+ CD8+ double-pos. thymocytes, but an increase in the percentage of CD8+ single-pos. thymocytes seen in CBP mutant mice was not obsd. in p300 mutants. T cells completely lacking both CBP and p300 did not develop normally and were nonexistent or very rare in the periphery, however. T cells lacking CBP or p300 had reduced tumor necrosis factor alpha gene expression in response to phorbol ester and ionophore, while signal-responsive gene expression in CBP- or p300-deficient macrophages was largely intact. Thus, CBP and p300 each supply a surprising degree of redundant coactivation capacity in T cells and macrophages, although each gene has also unique properties in thymocyte development.
- 406Rebel, V. I.; Kung, A. L.; Tanner, E. A.; Yang, H.; Bronson, R. T.; Livingston, D. M. Proc. Natl. Acad. Sci. U.S.A. 2002, 99, 14789Google ScholarThere is no corresponding record for this reference.
- 407Kasper, L. H.; Fukuyama, T.; Biesen, M. A.; Boussouar, F.; Tong, C.; de Pauw, A.; Murray, P. J.; van Deursen, J. M.; Brindle, P. K. Mol. Cell. Biol. 2006, 26, 789Google Scholar407Conditional knockout mice reveal distinct functions for the global transcriptional coactivators CBP and p300 in T-cell developmentKasper, Lawryn H.; Fukuyama, Tomofusa; Biesen, Michelle A.; Boussouar, Faycal; Tong, Caili; de Pauw, Antoine; Murray, Peter J.; van Deursen, Jan M. A.; Brindle, Paul K.Molecular and Cellular Biology (2006), 26 (3), 789-809CODEN: MCEBD4; ISSN:0270-7306. (American Society for Microbiology)The global transcriptional coactivators CREB-binding protein (CBP) and the closely related p300 interact with over 312 proteins, making them among the most heavily connected hubs in the known mammalian protein-protein interactome. It is largely uncertain, however, if these interactions are important in specific cell lineages of adult animals, as homozygous null mutations in either CBP or p300 result in early embryonic lethality in mice. Here we describe a Cre/LoxP conditional p300 null allele (p300flox) that allows for the temporal and tissue-specific inactivation of p300. We used mice carrying p300flax and a CBP conditional knockout allele (CBPflax) in conjunction with an Lck-Cre transgene to delete CBP and p300 starting at the CD4- CD8- double-neg. thymocyte stage of T-cell development. Loss of either p300 or CBP led to a decrease in CD4+ CD8+ double-pos. thymocytes, but an increase in the percentage of CD8+ single-pos. thymocytes seen in CBP mutant mice was not obsd. in p300 mutants. T cells completely lacking both CBP and p300 did not develop normally and were nonexistent or very rare in the periphery, however. T cells lacking CBP or p300 had reduced tumor necrosis factor alpha gene expression in response to phorbol ester and ionophore, while signal-responsive gene expression in CBP- or p300-deficient macrophages was largely intact. Thus, CBP and p300 each supply a surprising degree of redundant coactivation capacity in T cells and macrophages, although each gene has also unique properties in thymocyte development.
- 408Qiao, Y.; Molina, H.; Pandey, A.; Zhang, J.; Cole, P. A. Science 2006, 311, 1293Google ScholarThere is no corresponding record for this reference.
- 409Liao, Z. W.; Zhou, T. C.; Tan, X. J.; Song, X. L.; Liu, Y.; Shi, X. Y.; Huang, W. J.; Du, L. L.; Tu, B. J.; Lin, X. D. J. Transl. Med. 2012, 10, 110Google Scholar409High expression of p300 is linked to aggressive features and poor prognosis of nasopharyngeal carcinomaLiao, Zhi-Wei; Zhou, Tong-Chong; Tan, Xiao-Jun; Song, Xian-Lu; Liu, Yuan; Shi, Xing-Yuan; Huang, Wen-Jin; Du, Li-Li; Tu, Bo-Jun; Lin, Xiao-DanJournal of Translational Medicine (2012), 10 (), 110CODEN: JTMOBV; ISSN:1479-5876. (BioMed Central Ltd.)Background: Increased expression of transcriptional coactivator p300 has been obsd. in a variety of human cancers. However, the expression status of p300 protein/mRNA in nasopharyngeal carcinoma (NPC) tissues and its clinicopathol./prognostic implication are poorly understood. Methods: In our study, mRNA and protein expression levels of p300 was explored by reverse transcription-polymerase chain reaction (RT-PCR), Western blotting (WB) and immunohistochem. (IHC) in nasopharyngeal mucosal and NPC tissues. The data were analyzed by receiver operating characteristic (ROC) curve anal., spearman's rank correlation, Kaplan-Meier plots and Cox proportional hazards regression model. Results: Up-regulated expression of p300 mRNA/p300 protein was detected in NPC tissues by RT-PCR and WB, when compared to nasopharyngeal mucosal tissues. Based on ROC curve anal., the cutoff score for p300 high expression was defined when more than 35% of the tumor cells were pos. stained. High expression of p300 was obsd. in 127/209 (60.7%) of NPCs. In NPCs, high expression of p300 was pos. assocd. with later T classification, later N classification, distant metastasis and later clin. stage (P < 0.05). In univariate survival anal., overexpression of p300 was found to be an indicator of progression-free (P = 0.002) and overall survival (P = 0.001) in NPCs. More importantly, p300 expression was evaluated as an independent prognostic factor for NPC in multivariate anal. (P = 0.036). Conclusions: Our findings support that high expression of p300 protein might be important in conferring a more aggressive behavior, and is an independent mol. marker for shortened survival time of patients with NPC.
- 410Li, M.; Luo, R. Z.; Chen, J. W.; Cao, Y.; Lu, J. B.; He, J. H.; Wu, Q. L.; Cai, M. Y. J. Transl. Med. 2011, 9, 5Google Scholar410High expression of transcriptional coactivator p300 correlates with aggressive features and poor prognosis of hepatocellular carcinomaLi, Mei; Luo, Rong-Zhen; Chen, Jie-Wei; Cao, Yun; Lu, Jia-Bin; He, Jie-Hua; Wu, Qiu-Liang; Cai, Mu-YanJournal of Translational Medicine (2011), 9 (), 5CODEN: JTMOBV; ISSN:1479-5876. (BioMed Central Ltd.)Background: It has been suggested that p300 participates in the regulation of a wide range of cell biol. processes and mutation of p300 has been identified in certain types of human cancers. However, the expression dynamics of p300 in hepatocellular carcinoma (HCC) and its clin./prognostic significance are unclear. Methods: In this study, the methods of reverse transcription-polymerase chain reaction (RT-PCR), Western blotting and immunohistochem. (IHC) were utilized to investigate protein/mRNA expression of p300 in HCCs. Receiver operating characteristic (ROC) curve anal., spearman's rank correlation, Kaplan-Meier plots and Cox proportional hazards regression model were used to analyze the data. Results: Up-regulated expression of p300 mRNA and protein was obsd. in the majority of HCCs by RT-PCR and Western blotting, when compared with their adjacent non-malignant liver tissues. According to the ROC curves, the cutoff score for p300 high expression was defined when more than 60% of the tumor cells were pos. stained. High expression of p300 was examd. in 60/123 (48.8%) of HCCs and in 8/123 (6.5%) of adjacent non-malignant liver tissues. High expression of p300 was correlated with higher AFP level, larger tumor size, multiplicity, poorer differentiation and later stage (P < 0.05). In univariate survival anal., a significant assocn. between overexpression of p300 and shortened patients' survival was found (P = 0.001). In different subsets of HCC patients, p300 expression was also a prognostic indicator in patients with stage II (P = 0.007) and stage III (P = 0.011). Importantly, p300 expression was evaluated as an independent prognostic factor in multivariate anal. (P = 0.021). Consequently, a new clinicopathol. prognostic model with three poor prognostic factors (p300 expression, AFP level and vascular invasion) was constructed. The model could significantly stratify risk (low, intermediate and high) for overall survival (P < 0.0001). Conclusions: Our findings provide a basis for the concept that high expression of p300 in HCC may be important in the acquisition of an aggressive phenotype, suggesting that p300 overexpression, as examd. by IHC, is an independent biomarker for poor prognosis of patients with HCC. The combined clinicopathol. prognostic model may become a useful tool for identifying HCC patients with different clin. outcomes.
- 411Vleugel, M. M.; Shvarts, D.; van der Wall, E.; van Diest, P. J. Hum. Pathol. 2006, 37, 1085Google Scholar411p300 and p53 levels determine activation of HIF-1 downstream targets in invasive breast cancerVleugel, Marije M.; Shvarts, David; van der Wall, Elsken; van Diest, Paul J.Human Pathology (2006), 37 (8), 1085-1092CODEN: HPCQA4; ISSN:0046-8177. (Elsevier Inc.)In previous studies, we noted that overexpression of hypoxia-inducible factor (HIF)-1α in breast cancer, esp. the diffuse form, does not always lead to functional activation of its downstream genes. Transcriptional activity of HIF-1 may be repressed by p53 through competition for transcriptional coactivators such as p300. The aim of this study was therefore to explore the role of p53 and p300 in relation to overexpression of HIF-1α and activation of HIF-1 downstream genes in invasive breast cancer. p300 immunohistochem. was performed in a group of 183 early-stage invasive breast cancers, and related to p53 accumulation, overexpression of HIF-1α, and several HIF-1 downstream genes. p300 was expressed in varying degrees in 84% of invasive breast cancers. p300 staining intensity correlated pos. with HIF-1α expression (P = .04), p53 accumulation (P = .001), and overexpression of glucose transporter 1 (GLUT-1) (P < .001), a glucose transporter downstream target gene of HIF-1. GLUT-1 levels were significantly assocd. with p300 in HIF-1α pos. patients (P = .02). p53 accumulation significantly pos. correlated with carbonic anhydrase IX (CAIX)/GLUT-1 coexpression in HIF-1α-pos. patients (P = .007). p53 accumulation/high p300 levels, the most favorable situation for HIF-1 downstream activation, were significantly assocd. with GLUT-1 overexpression (P = .01) and coexpression of CAIX/GLUT-1 (P = .03), compared with low p53/low p300 levels, the most unfavorable situation for HIF-1 downstream activation. p300 is a cofactor highly assocd. with p53 accumulation and HIF-1α levels in invasive breast cancer. Furthermore, low levels of p300 may explain absence of downstream effects in HIF-1α-overexpressing cancers, an effect that seems to be enhanced by wild-type levels of p53. This underlines the importance of p300 levels and p53 accumulation in the HIF-1-regulated response toward hypoxia.
- 412Hudelist, G.; Czerwenka, K.; Kubista, E.; Marton, E.; Pischinger, K.; Singer, C. F. Breast Cancer Res. Treat. 2003, 78, 193Google Scholar412Expression of Sex Steroid Receptors and their Co-Factors in Normal and Malignant Breast Tissue: AIB1 is a Carcinoma-Specific Co-ActivatorHudelist, Gernot; Czerwenka, Klaus; Kubista, Ernst; Marton, Erika; Pischinger, Kerstin; Singer, Christian F.Breast Cancer Research and Treatment (2003), 78 (2), 193-204CODEN: BCTRD6; ISSN:0167-6806. (Kluwer Academic Publishers)The differential expression pattern of estrogen receptor alpha (ER-α), estrogen receptor beta (ER-β) and their co-activator/co-repressor proteins is thought to modulate estrogenic action and to be present already during the early stages of tumorigenesis. It has therefore been postulated that certain co-activator and co-repressor proteins contribute to the development of breast cancer. There are some reports providing information on gene amplification and mRNA over-expression of certain co-factors in breast cancer, but to date there is only limited knowledge about their resp. protein expressions. The aim of this study was to examine the expression of four steroid receptor co-activators (steroid receptor co-activator 1 (SRC-1), transcription intermediary factor 2 (TIF 2), protein 300 kDa/CREB binding protein (p300/CBP), amplified in breast cancer 1 (AIB1)), and of the co-repressor nuclear receptor co-repressor (NCoR), in malignant breast tissues and in matching normal breast biopsies of the same individuals. Protein expression was analyzed by immunohistochem. and was compared to prognostic parameters such as lymph node involvement, tumor grading and receptor status. All members of the co-regulatory protein family were detected in both, benign and matching malignant tissue samples, except for AIB1, which was found to be expressed exclusively in malignant epithelium. AIB1 was preferentially present in carcinomas with high tumor grade (r = 0.48, p = 0.014), and was co-expressed with p300/CBP (r = 0.54, p = 0.006). TIF 2 correlated significantly to nodal status (r = 0.46, p = 0.025). Furthermore, protein levels of ER-β, p300/CBP and AIB1 were higher in invasive ductal carcinomas than in normal mammary tissue. The tumoral ER-α protein expression was significantly correlated with that of PgR (r = 0.61, p = 0.001) and NCoR (r = 0.4, p = 0.043), whereas ER-β expression was assocd. with SRC-1 (r = 0.68, p ≤ 0.001), TIF 2 (r = 0.64, p = 0.001) and NCoR (r = 0.48, p = 0.014) protein levels in malignant specimens. In our hands, 20% of ER-β pos. tumors did not express ER-α protein, thereby suggesting that a substantial fraction of ER-beta pos. tumors is falsely considered to be estrogen receptor neg.' if only ER-α specific antibodies are employed in the histol. assessment of the ER status.
- 413Debes, J. D.; Sebo, T. J.; Lohse, C. M.; Murphy, L. M.; Haugen, D. A.; Tindall, D. J. Cancer Res. 2003, 63, 7638Google ScholarThere is no corresponding record for this reference.
- 414Isharwal, S.; Miller, M. C.; Marlow, C.; Makarov, D. V.; Partin, A. W.; Veltri, R. W. Prostate 2008, 68, 1097Google ScholarThere is no corresponding record for this reference.
- 415Heery, D. M.; Fischer, P. M. Drug Discovery Today 2007, 12, 88Google ScholarThere is no corresponding record for this reference.
- 416Lin, W. M.; Baker, A. C.; Beroukhim, R.; Winckler, W.; Feng, W.; Marmion, J. M.; Laine, E.; Greulich, H.; Tseng, H.; Gates, C.; Hodi, F. S.; Dranoff, G.; Sellers, W. R.; Thomas, R. K.; Meyerson, M.; Golub, T. R.; Dummer, R.; Herlyn, M.; Getz, G.; Garraway, L. A. Cancer Res. 2008, 68, 664Google ScholarThere is no corresponding record for this reference.
- 417Rotte, A.; Bhandaru, M.; Cheng, Y.; Sjoestroem, C.; Martinka, M.; Li, G. PLoS One 2013, 8, e75405Google ScholarThere is no corresponding record for this reference.
- 418Bhandaru, M.; Ardekani, G. S.; Zhang, G.; Martinka, M.; McElwee, K. J.; Li, G.; Rotte, A. BMC Cancer 2014, 14– 398Google ScholarThere is no corresponding record for this reference.
- 419Bandyopadhyay, D.; Okan, N. A.; Bales, E.; Nascimento, L.; Cole, P. A.; Medrano, E. E. Cancer Res. 2002, 62, 6231Google ScholarThere is no corresponding record for this reference.
- 420Yan, G.; Eller, M. S.; Elm, C.; Larocca, C. A.; Ryu, B.; Panova, I. P.; Dancy, B. M.; Bowers, E. M.; Meyers, D.; Lareau, L.; Cole, P. A.; Taverna, S. D.; Alani, R. M. J. Invest. Dermatol. 2013, 133, 2444Google ScholarThere is no corresponding record for this reference.
- 421Iyer, N. G.; Chin, S. F.; Ozdag, H.; Daigo, Y.; Hu, D. E.; Cariati, M.; Brindle, K.; Aparicio, S.; Caldas, C. Proc. Natl. Acad. Sci. U.S.A. 2004, 101, 7386Google ScholarThere is no corresponding record for this reference.
- 422Giles, R. H.; Petrij, F.; Dauwerse, H. G.; den Hollander, A. I.; Lushnikova, T.; van Ommen, G. J.; Goodman, R. H.; Deaven, L. L.; Doggett, N. A.; Peters, D. J.; Breuning, M. H. Genomics 1997, 42, 96Google ScholarThere is no corresponding record for this reference.
- 423Rozman, M.; Camós, M.; Colomer, D.; Villamor, N.; Esteve, J.; Costa, D.; Carrió, A.; Aymerich, M.; Aguilar, J. L.; Domingo, A.; Solé, F.; Gomis, F.; Florensa, L.; Montserrat, E.; Campo, E. Genes, Chromosomes Cancer 2004, 40, 140Google Scholar423Type I MOZ/CBP (MYST3/CREBBP) is the most common chimeric transcript in acute myeloid leukemia with t(8;16)(p11;p13) translocationRozman, Maria; Camos, Mireia; Colomer, Dolors; Villamor, Neus; Esteve, Jordi; Costa, Dolors; Carrio, Ana; Aymerich, Marta; Aguilar, Josep Lluis; Domingo, Alicia; Sole, Francesc; Gomis, Federico; Florensa, Lourdes; Montserrat, Emili; Campo, EliasGenes, Chromosomes & Cancer (2004), 40 (2), 140-145CODEN: GCCAES; ISSN:1045-2257. (Wiley-Liss, Inc.)The t(8;16)(p11;p13) fuses the MOZ (MYST3) gene at 8p11 with CBP (CREBBP) at 16p13 and is assocd. with an infrequent but well-defined type of acute myeloid leukemia (AML) that has unique morphocytochem. findings (monocytoid blast morphol. with erythrophagocytosis and simultaneously pos. for myeloperoxidase and nonspecific esterases). RT-PCR amplification of MOZ/CBP (MYST3/CREBBP) chimera has proved difficult, with four different transcripts found in four reported cases. The authors studied 7 AML-t(8;16) patients, 5 with cytogenetically demonstrated t(8;16) and 2 with similar morphocytochem. and immunophenotypical characteristics. Clin., 3 cases presented as therapy-related leukemia. Extramedullar involvement was obsd. at presentation in 2 patients and coagulopathy in 4. The clinicobiol. findings confirmed the distinctiveness of this entity. Of note is the erythrophagocytosis in 5 of 7 cases and the immunol. negativity for CD34 and CD117 and positivity for CD56. Using a new RT-PCR strategy, the authors were able to amplify a specific band of 212 bp in six cases in which sequence anal. confirmed the presence of the previously described MOZ/CBP fusion transcript type 1. This is the largest molecularly studied AML-t(8;16) series, which demonstrates that MOZ/CBP breakpoints are usually clustered in intron 16 of MOZ and intron 2 of CBP. The newly designed single-round PCR provides a simple tool for the mol. confirmation of MOZ/CBP rearrangement.
- 424Borrow, J.; Stanton, V. P., Jr.; Andresen, J. M.; Becher, R.; Behm, F. G.; Chaganti, R. S.; Civin, C. I.; Disteche, C.; Dubé, I.; Frischauf, A. M.; Horsman, D.; Mitelman, F.; Volinia, S.; Watmore, A. E.; Housman, D. E. Nat. Genet. 1996, 14, 33Google Scholar424The translocation t(8;16)(p11;p13) of acute myeloid leukemia fuses a putative acetyltransferase to the CREB-binding proteinBorrow, Julian; Stanton, Vincent P., Jr.; Andresen, J. Michael; Becher, Reinhard; Behm, Frederick G.; Chagaanti, Raju S. K.; Civin, Curt I.; Disteche, Christine; Dube, Ian; et al.Nature Genetics (1996), 14 (1), 33-41CODEN: NGENEC; ISSN:1061-4036. (Nature Publishing Co.)The recurrent translocation t(8;16)(p11;p13) is a cytogenetic hallmark for the M4/M5 subtype of acute myeloid leukemia. Here the authors identify the breakpoint-assocd. genes. Positional cloning on chromosome 16 implicates the CREB-binding protein (CBP), a transcriptional adaptor/co-activator protein. At the chromosome 8 breakpoint the authors identify a novel gene, MOZ, which encodes a 2,004-amino-acid protein characterized by two C4HC3 zinc fingers and a single C2HC zinc finger in conjunction with a putative acetyltransferase signature. In-frame MOZ-CBP fusion transcripts combine the MOZ finger motifs and putative acetyltransferase domain with a largely intact CBP. The authors suggest that MOZ may represent a chromatin-assocd. acetyltransferase, and raise the possibility that a dominant MOZ-CBP fusion protein could mediate leukemogenesis via aberrant chromatin acetylation.
- 425Chaffanet, M.; Gressin, L.; Preudhomme, C.; Soenen-Cornu, V.; Birnbaum, D.; Pébusque, M. J. Genes, Chromosomes Cancer 2000, 28, 138Google Scholar425MOZ is fused to p300 in an acute monocytic leukemia with t(8;22)Chaffanet, Max; Gressin, Laetitia; Preudhomme, Claude; Soenen-Cornu, Valerie; Birnbaum, Daniel; Pebusque, Marie-JosepheGenes, Chromosomes & Cancer (2000), 28 (2), 138-144CODEN: GCCAES; ISSN:1045-2257. (Wiley-Liss, Inc.)The authors report on the fusion of the monocytic leukemia zinc finger protein (MOZ) gene to the adenoviral EIA-assocd. protein p300 gene in acute monocytic leukemia MS assocd. with a t(8;22)(p11;q13) translocation. The authors studied two patients with double-color fluorescence in situ hybridization (FISH) using the yeast artificial chromosome 176C9 and the bacterial artificial chromosome clone H59D10 specific to the MOZ and p300 genes, resp. Both probes were split in the patients' chromosome metaphase cells, and the two deriv. chromosomes were each labeled with both probes. The authors showed by Southern blot the rearrangement of the MOZ gene, and cloned the fusion transcripts in one patient carrying the t(8;22) by reverse transcription-polymerase chain reaction using MOZ- and p300-specific primers. Both fusion transcripts were expressed. This result defines a novel reciprocal translocation involving two acetyltransferases, MOZ and p300, resulting in an abnormal transcriptional co-activator that could play a crit. role in leukemogenesis.
- 426Kitabayashi, I.; Aikawa, Y.; Yokoyama, A.; Hosoda, F.; Nagai, M.; Kakazu, N.; Abe, T.; Ohki, M. Leukemia 2001, 15, 89Google ScholarThere is no corresponding record for this reference.
- 427Panagopoulos, I.; Fioretos, T.; Isaksson, M.; Samuelsson, U.; Billström, R.; Strömbeck, B.; Mitelman, F.; Johansson, B. Hum. Mol. Genet. 2001, 10, 395Google Scholar427Fusion of the MORF and CBP genes in acute myeloid leukemia with the t(10;16)(q22;p13)Panagopoulos, Ioannis; Fioretos, Thoas; Isaksson, Margareth; Samuelsson, Ulf; Billstrom, Rolf; Strombeck, Bodil; Mitelman, Felix; Johansson, BertilHuman Molecular Genetics (2001), 10 (4), 395-404CODEN: HMGEE5; ISSN:0964-6906. (Oxford University Press)The CBP gene at 16p13 fuses to MOZ and MLL as a result of the t(8;16)(p11;p13) in acute (myelo)monocytic leukemias (AML M4/M5) and the t(11;16)(q23;p13) in treatment-related AML, resp. Here the authors show that a novel t(10;16)(q22;p13) in a childhood AML M5a leads to a MORF-CBP chimera. RT-PCR using MORF forward and CBP reverse primers amplified a MORF-CBP fusion in which nucleotide 3103 of MORF was fused in-frame with nucleotide 284 of CBP. Nested RT-PCR with CBP forward and MORF reverse primers generated a CBP-MORF transcript in which nucleotide 283 of CBP was fused in-frame with nucleotide 3104 of MORF. Genomic analyses revealed that the breaks were close to Alu elements in intron 16 of MORF and intron 2 of CBP and that duplications had occurred near the breakpoints. A database search using MORF cDNA enabled us to construct an exon-intron map of the MORF gene. The MORF-CBP protein retains the zinc fingers, two nuclear localization signals, the histone acetyltransferase (HAT) domain, a portion of the acidic domain of MORF and the CBP protein downstream of codon 29. Thus, the part of CBP encoding the RARA-binding domain, the CREB-binding domain, the three Cys/His-rich regions, the bromodomain, the HAT domain and the Glu-rich domains is present. In the reciprocal CBP-MORF, part of the acidic domain and the C-terminal Ser- and Met-rich regions of MORF are likely to be driven by the CBP promoter. Since both fusion transcripts were present, their exact role in the leukemogenic process remains to be elucidated.
- 428Vizmanos, J. L.; Larráyoz, M. J.; Lahortiga, I.; Floristán, F.; Alvarez, C.; Odero, M. D.; Novo, F. J.; Calasanz, M. J. Genes, Chromosomes Cancer 2003, 36, 402Google Scholar428t(10;16)(q22;p13) and MORF-CREBBP fusion is a recurrent event in acute myeloid leukemiaVizmanos, Jose L.; Larrayoz, Maria J.; Lahortiga, Idoya; Floristan, Filomena; Alvarez, Carmen; Odero, Maria D.; Novo, Francisco J.; Calasanz, Maria J.Genes, Chromosomes & Cancer (2003), 36 (4), 402-405CODEN: GCCAES; ISSN:1045-2257. (Wiley-Liss, Inc.)Recently, it was shown that t(10;16)(q22;p13) fuses the MORF and CREBBP genes in a case of childhood acute myeloid leukemia (AML) M5a, with a complex karyotype contg. other rearrangements. Here, we report a new case with the MORF-CREBBP fusion in an 84-yr-old patient diagnosed with AML M5b, in which the t(10;16)(q22;p13) was the only cytogenetic aberration. This supports that this is a recurrent pathogenic translocation in AML.
- 429Uppal, G. K.; Leighton, J.; Da Costa, D.; Czulewicz, A.; Palazzo, I. E. Hematol. Rep. 2011, 3, e23Google ScholarThere is no corresponding record for this reference.
- 430Kojima, K.; Kaneda, K.; Yoshida, C.; Dansako, H.; Fujii, N.; Yano, T.; Shinagawa, K.; Yasukawa, M.; Fujita, S.; Tanimoto, M. Br. J. Haematol. 2003, 120, 271Google Scholar430A novel fusion variant of the MORF and CBP genes detected in therapy-related myelodysplastic syndrome with t(10;16)(q22;p13)Kojima, Kensuke; Kaneda, Kinuyo; Yoshida, Chikamasa; Dansako, Hirokata; Fujii, Nobuharu; Yano, Tomofumi; Shinagawa, Katsuji; Yasukawa, Masaki; Fujita, Shigeru; Tanimoto, MitsuneBritish Journal of Haematology (2003), 120 (2), 271-273CODEN: BJHEAL; ISSN:0007-1048. (Blackwell Publishing Ltd.)We report a case of therapy-related myelodysplastic syndrome (t-MDS) with t(10;16)(q22;p13), in which novel fusion transcripts of the MORF and CBP genes were detected. In one MORF-CBP fusion transcript, exon 15 of the MORF gene was fused in frame with exon 5 of the CBP gene. In a reciprocal CBP-MORF transcript, exon 4 of the CBP gene was fused in frame with exon 16 of the MORF gene. This is the first reported case of t-MDS assocd. with t(10;16), and provides mol. evidence that the novel MORF-CBP and/or CBP-MORF fusion protein(s) might play an important role in the development of t-MDS.
- 431Taki, T.; Sako, M.; Tsuchida, M.; Hayashi, Y. Blood 1997, 89, 3945Google ScholarThere is no corresponding record for this reference.
- 432Sobulo, O. M.; Borrow, J.; Tomek, R.; Reshmi, S.; Harden, A.; Schlegelberger, B.; Housman, D.; Doggett, N. A.; Rowley, J. D.; Zeleznik-Le, N. J. Proc. Natl. Acad. Sci. U.S.A. 1997, 94, 8732Google ScholarThere is no corresponding record for this reference.
- 433Satake, N.; Ishida, Y.; Otoh, Y.; Hinohara, S.; Kobayashi, H.; Sakashita, A.; Maseki, N.; Kaneko, Y. Genes, Chromosomes Cancer 1997, 20, 60Google Scholar433Novel MLL-CBP fusion transcript in therapy-related chronic myelomonocytic leukemia with a t(11;16) (q23;p13) chromosome translocationSatake, Noriko; Ishida, Yasushi; Otoh, Yoshiko; Hinohara, Shin-Ichi; Kobayashi, Hirofumi; Sakashita, Akiko; Maseki, Nobuo; Kaneko, YasuhikoGenes, Chromosomes & Cancer (1997), 20 (1), 60-63CODEN: GCCAES; ISSN:1045-2257. (Wiley-Liss)CBP, which is located on 16p13 and encodes a transcriptional adaptor/coactivator protein, has been shown to fuse by the t(8;16)(p11;p13) translocation to MOZ on 8p11 in acute myeloid leukemia. The authors found a t(11;16)(q23;p13) in a child with therapy-related chronic myelomonocytic leukemia. Subsequent reverse transcriptase-polymerase chain reaction and direct sequencing analyses revealed the MLL-CBP fusion transcript in CMML cells. Because 11q23 translocations involving MLL and t(8;16) involving MOZ and CBP have been reported in therapy-related leukemias, both the MLL and CBP genes may be targets for topoisomerase II inhibitors. Accordingly, the authors believe that most t(11;16)-assocd. leukemias may develop in patients who have been treated with cytotoxic chemotherapy for primary malignant diseases.
- 434Ida, K.; Kitabayashi, I.; Taki, T.; Taniwaki, M.; Noro, K.; Yamamoto, M.; Ohki, M.; Hayashi, Y. Blood 1997, 90, 4699Google ScholarThere is no corresponding record for this reference.
- 435Wang, J.; Iwasaki, H.; Krivtsov, A.; Febbo, P. G.; Thorner, A. R.; Ernst, P.; Anastasiadou, E.; Kutok, J. L.; Kogan, S. C.; Zinkel, S. S.; Fisher, J. K.; Hess, J. L.; Golub, T. R.; Armstrong, S. A.; Akashi, K.; Korsmeyer, S. J. EMBO J. 2005, 24, 368Google ScholarThere is no corresponding record for this reference.
- 436Wang, L.; Gural, A.; Sun, X. J.; Zhao, X.; Perna, F.; Huang, G.; Hatlen, M. A.; Vu, L.; Liu, F.; Xu, H.; Asai, T.; Xu, H.; Deblasio, T.; Menendez, S.; Voza, F.; Jiang, Y.; Cole, P. A.; Zhang, J.; Melnick, A.; Roeder, R. G.; Nimer, S. D. Science 2011, 333, 765Google ScholarThere is no corresponding record for this reference.
- 437Reynoird, N.; Schwartz, B. E.; Delvecchio, M.; Sadoul, K.; Meyers, D.; Mukherjee, C.; Caron, C.; Kimura, H.; Rousseaux, S.; Cole, P. A.; Panne, D.; French, C. A.; Khochbin, S. EMBO J. 2010, 29, 2943Google ScholarThere is no corresponding record for this reference.
- 438Crump, N. T.; Hazzalin, C. A.; Bowers, E. M.; Alani, R. M.; Cole, P. A.; Mahadevan, L. C. Proc. Natl. Acad. Sci. U.S.A. 2011, 108, 7814Google ScholarThere is no corresponding record for this reference.
- 439Pattabiraman, D. R.; Sun, J.; Dowhan, D. H.; Ishii, S.; Gonda, T. J. Mol. Cancer Res. 2009, 7, 1477Google Scholar439Mutations in Multiple Domains of c-Myb Disrupt Interaction with CBP/p300 and Abrogate Myeloid Transforming AbilityPattabiraman, Diwakar R.; Sun, Jane; Dowhan, Dennis H.; Ishii, Shunsuke; Gonda, Thomas J.Molecular Cancer Research (2009), 7 (9), 1477-1486CODEN: MCROC5; ISSN:1541-7786. (American Association for Cancer Research)The c-myb proto-oncogene is a key regulator of hematopoietic cell proliferation and differentiation. MYB mRNA is expressed at high levels in, and is required for the proliferation of, most human myeloid and acute lymphoid leukemias. Recently, chromosomal translocation and genomic duplications of c-MYB have been identified in human T-cell acute leukemia. The present work focuses on the effects of mutations in different domains of the murine c-Myb protein on its transforming ability as defined by suppression of myelomonocytic differentiation and continued proliferation. Using both a novel myeloid cell line-based assay and a primary hematopoietic cell assay, we have shown that mutation of single residues in the transactivation domain important for CBP/p300 binding leads to complete loss of transforming ability. We also simultaneously mutated residues in the DNA-binding domain and the neg. regulatory domain of the protein. These double mutants, but not the corresponding single mutants, show a complete loss of transforming activity. Surprisingly, these double mutants show severely impaired transactivation and are also defective for CBP/p300 binding. Our results imply that multiple Myb domains influence its interaction with CBP/p300, highlight the importance of this interaction for myeloid transformation, and suggest an approach for mol. targeting of Myb in leukemia. (Mol Cancer Res 2009;7(9):1477-86).
- 440Iyer, N. G.; Ozdag, H.; Caldas, C. Oncogene 2004, 23, 4225Google Scholar440p300/CBP and cancerIyer, Narayanan Gopalakrishna; Oezdag, Hilal; Caldas, CarlosOncogene (2004), 23 (24), 4225-4231CODEN: ONCNES; ISSN:0950-9232. (Nature Publishing Group)A review. P300 and cAMP response element-binding protein (CBP) are adenoviral E1A-binding proteins involved in multiple cellular processes, and function as transcriptional co-factors and histone acetyltransferases. Germline mutation of CBP results in Rubinstein-Taybi syndrome, which is characterized by an increased predisposition to childhood malignancies. Furthermore, somatic mutations of p300 and CBP occur in a no. of malignancies. Chromosome translocations target CBP and, less commonly, p300 in acute myeloid leukemia and treatment-related hematol. disorders. P300 mutations in solid tumors result in truncated p300 protein products or amino-acid substitutions in crit. protein domains, and these are often assocd. with inactivation of the second allele. A mouse model confirms that p300 and CBP function as suppressors of hematol. tumor formation. The involvement of these proteins in crit. tumorigenic pathways (including TGF-β, p53 and Rb) provides a mechanistic route as to how their inactivation could result in cancer.
- 441Varier, R. A.; Kundu, T. K. Curr. Pharm. Des. 2006, 12, 1975Google Scholar441Chromatin modifications (acetylation/ deacetylation/ methylation) as new targets for HIV therapyVarier, Radhika A.; Kundu, Tapas K.Current Pharmaceutical Design (2006), 12 (16), 1975-1993CODEN: CPDEFP; ISSN:1381-6128. (Bentham Science Publishers Ltd.)A review. Human immunodeficiency virus (HIV) is one of the most deadly threats to the human race. Though the developed countries have been able to control the epidemic by utilizing the discovery of very expensive diagnostics, the situation is dangerously alarming in developing and underdeveloped countries. The development of highly active antiretroviral drugs has improved the survival and quality of life, but prolonged treatment results in viral load rebound to pretherapy levels. Recent advances in the understanding of eukaryotic and genome-integrated viral gene expression showed that regulation of chromatin function is closely linked to the multiplication of HIV. Therefore, a new therapeutic approach has been initiated targeting the chromatin-modifying enzymes mainly histone acetyltransferases and deacetylases which may lead to a better and economical anti- HIV combinatorial therapeutics. In this review, the authors have discussed the mechanisms of HIV gene expression in the chromatin context and its potentiality to be exploited as a new therapeutic target.
- 442Mujtaba, S.; Zhou, M. M. Methods 2011, 53, 97Google ScholarThere is no corresponding record for this reference.
- 443Sakane, N.; Kwon, H. S.; Pagans, S.; Kaehlcke, K.; Mizusawa, Y.; Kamada, M.; Lassen, K. G.; Chan, J.; Greene, W. C.; Schnoelzer, M.; Ott, M. PLoS Pathog. 2011, 7, e1002184Google ScholarThere is no corresponding record for this reference.
- 444Mujtaba, S.; Zhou, M. M. Methods 2011, 53, 97Google ScholarThere is no corresponding record for this reference.
- 445Allouch, A.; Di Primio, C.; Alpi, E.; Lusic, M.; Arosio, D.; Giacca, M.; Cereseto, A. Cell Host Microbe 2011, 9, 484Google Scholar445The TRIM Family Protein KAP1 Inhibits HIV-1 IntegrationAllouch, Awatef; Di Primio, Cristina; Alpi, Emanuele; Lusic, Marina; Arosio, Daniele; Giacca, Mauro; Cereseto, AnnaCell Host & Microbe (2011), 9 (6), 484-495CODEN: CHMECB; ISSN:1931-3128. (Cell Press)Summary: The integration of viral cDNA into the host genome is a crit. step in the life cycle of HIV-1. This step is catalyzed by integrase (IN), a viral enzyme that is pos. regulated by acetylation via the cellular histone acetyl transferase (HAT) p300. To investigate the relevance of IN acetylation, we searched for cellular proteins that selectively bind acetylated IN and identified KAP1, a protein belonging to the TRIM family of antiviral proteins. KAP1 binds acetylated IN and induces its deacetylation through the formation of a protein complex which includes the deacetylase HDAC1. Modulation of intracellular KAP1 levels in different cell types including T cells, the primary HIV-1 target, revealed that KAP1 curtails viral infectivity by selectively affecting HIV-1 integration. This study identifies KAP1 as a cellular factor restricting HIV-1 infection and underscores the relevance of IN acetylation as a crucial step in the viral infectious cycle.
- 446Terreni, M.; Valentini, P.; Liverani, V.; Gutierrez, M. I.; Di Primio, C.; Di Fenza, A.; Tozzini, V.; Allouch, A.; Albanese, A.; Giacca, M.; Cereseto, A. Retrovirology 2010, 7, 18Google ScholarThere is no corresponding record for this reference.
- 447Zou, W.; Wang, Z.; Liu, Y.; Fan, Y.; Zhou, B. Y.; Yang, X. F.; He, J. J. Glia 2010, 58, 1640Google ScholarThere is no corresponding record for this reference.
- 448Kauppi, M.; Murphy, J. M.; de Graaf, C. A.; Hyland, C. D.; Greig, K. T.; Metcalf, D.; Hilton, A. A.; Nicola, N. A.; Kile, B. T.; Hilton, D. J.; Alexander, W. S. Blood 2008, 112, 3148Google ScholarThere is no corresponding record for this reference.
- 449Hilton, D. J.; Kile, B. T.; Alexander, W. S. Blood 2009, 113, 5599Google ScholarThere is no corresponding record for this reference.
- 450Yanazume, T.; Morimoto, T.; Wada, H.; Kawamura, T.; Hasegawa, K. Mol. Cell. Biochem. 2003, 248, 115Google ScholarThere is no corresponding record for this reference.
- 451Gusterson, R. J.; Jazrawi, E.; Adcock, I. M.; Latchman, D. S. J. Biol. Chem. 2003, 278, 6838Google ScholarThere is no corresponding record for this reference.
- 452Zhou, X. Y.; Shibusawa, N.; Naik, K.; Porras, D.; Temple, K.; Ou, H.; Kaihara, K.; Roe, M. W.; Brady, M. J.; Wondisford, F. E. Nat. Med. 2004, 10, 633Google Scholar452Insulin regulation of hepatic gluconeogenesis through phosphorylation of CREB-binding proteinZhou, Xiao Yan; Shibusawa, Nobuyuki; Naik, Karuna; Porras, Delia; Temple, Karla; Ou, Hesheng; Kaihara, Kelly; Roe, Michael W.; Brady, Matthew J.; Wondisford, Fredric E.Nature Medicine (New York, NY, United States) (2004), 10 (6), 633-637CODEN: NAMEFI; ISSN:1078-8956. (Nature Publishing Group)Hepatic gluconeogenesis is essential for maintenance of normal blood glucose concns. and is regulated by opposing stimulatory (cAMP) and inhibitory (insulin) signaling pathways. The cAMP signaling pathway leads to phosphorylation of cAMP response element-binding (CREB) protein, resulting in recruitment of the coactivators CREB-binding protein (CBP) and p300 and subsequent activation of gluconeogenesis. Insulin signaling leads to phosphorylation of CBP at serine 436, a residue near its CREB-interacting domain, but it is unknown whether this event modulates cAMP signaling. Here, the authors show in vitro and in 'knock-in' mice that a mutant CBP (S436A) is aberrantly recruited to CREB protein, resulting in inappropriate activation of gluconeogenesis in the fed state and glucose intolerance resulting from increased hepatic glucose prodn. The authors propose that insulin signaling may directly regulate many cAMP signaling pathways at the transcriptional level by controlling CBP recruitment.
- 453Cha-Molstad, H.; Saxena, G.; Chen, J.; Shalev, A. J. Biol. Chem. 2009, 284, 16898Google ScholarThere is no corresponding record for this reference.
- 454Bricambert, J.; Miranda, J.; Benhamed, F.; Girard, J.; Postic, C.; Dentin, R. J. Clin. Invest. 2010, 120, 4316Google ScholarThere is no corresponding record for this reference.
- 455Schnell, U.; Dijk, F.; Sjollema, K. A.; Giepmans, B. N. Nat. Methods 2012, 9, 152Google Scholar455Immunolabeling artifacts and the need for live-cell imagingSchnell, Ulrike; Dijk, Freark; Sjollema, Klaas A.; Giepmans, Ben N. G.Nature Methods (2012), 9 (2), 152-158CODEN: NMAEA3; ISSN:1548-7091. (Nature Publishing Group)A review. Fluorescent fusion proteins have revolutionized examn. of proteins in living cells. Still, studies using these proteins are met with criticism because proteins are modified and ectopically expressed, in contrast to immunofluorescence studies. However, introducing immunoreagents inside cells can cause protein extn. or relocalization, not reflecting the in vivo situation. Here we discuss pitfalls of immunofluorescence labeling that often receive little attention and argue that immunostaining expts. in dead, permeabilized cells should be complemented with live-cell imaging when scrutinizing protein localization.
- 456McNeil, P. L. J. Cell Sci. 1987, 88, 669Google Scholar456Glass beads load macromolecules into living cellsMcNeil P L; Warder EJournal of cell science (1987), 88 ( Pt 5) (), 669-78 ISSN:0021-9533.We describe and characterize an exceptionally rapid and simple new technique for loading large numbers of cultured cells with large macromolecules. The culture medium of the cell monolayer is replaced by a small volume of the macromolecule to be loaded. Glass beads (75-500 micron diameter) are then sprinkled onto the cells, the cells are washed free of beads and exogenous macromolecules, and 'bead-loading' is completed. The conditions for bead-loading can readily be modified to accommodate cell type and loading objectives: for example, the amount of loading per cell increases if bead size is increased or if beads are agitated after sprinkling onto the monolayer, but at the expense of increased cell loss. As many as 97% of a population of bovine aortic endothelial (BAE) cells were loaded with a 10,000 Mr dextran; and 79% with a 150,000 Mr dextran using bead-loading. Various cell lines have been loaded using glass beads. Moreover, bead-loading has the advantage of producing loaded cells that remain adherent and well-spread, thus minimizing recovery time and permitting immediate microscopic examination.
- 457Hayashi-Takanaka, Y.; Yamagata, K.; Wakayama, T.; Stasevich, T. J.; Kainuma, T.; Tsurimoto, T.; Tachibana, M.; Shinkai, Y.; Kurumizaka, H.; Nozaki, N.; Kimura, H. Nucleic Acids Res. 2011, 39, 6475Google ScholarThere is no corresponding record for this reference.
- 458Hayashi-Takanaka, Y.; Yamagata, K.; Nozaki, N.; Kimura, H. J. Cell Biol. 2009, 187, 781Google ScholarThere is no corresponding record for this reference.
- 459Johansen, K. M.; Johansen, J. Chromosome Res. 2006, 14, 393Google Scholar459Regulation of chromatin structure by histone H3S10 phosphorylationJohansen, Kristen M.; Johansen, JorgenChromosome Research (2006), 14 (4), 393-404CODEN: CRRSEE; ISSN:0967-3849. (Springer)A review. The epigenetic phospho-serine 10 modification of histone H3 has been a puzzle due to its assocn. with two apparently opposed chromatin states. It is found at elevated levels on the highly condensed, transcriptionally inactive mitotic chromosomes yet is also correlated with the more extended chromatin configuration of active genes, euchromatic interband regions, and activated heat shock puffs of Drosophila polytene chromosomes. In addn., phosphorylation of histone H3S10 is up-regulated on the hypertranscribed male X chromosome. Here we review the cellular effects of histone H3S10 phosphorylation and discuss a model for its involvement in regulating chromatin organization and heterochromatization that would be applicable to both interphase and mitotic chromosomes.
- 460Stasevich, T. J.; Hayashi-Takanaka, Y.; Sato, Y.; Maehara, K.; Ohkawa, Y.; Sakata-Sogawa, K.; Tokunaga, M.; Nagase, T.; Nozaki, N.; McNally, J. G.; Kimura, H. Nature 2014, 516, 272Google ScholarThere is no corresponding record for this reference.
- 461Rothbauer, U.; Zolghadr, K.; Tillib, S.; Nowak, D.; Schermelleh, L.; Gahl, A.; Backmann, N.; Conrath, K.; Muyldermans, S.; Cardoso, M. C.; Leonhardt, H. Nat. Methods. 2006, 3, 887Google Scholar461Targeting and tracing antigens in live cells with fluorescent nanobodiesRothbauer, Ulrich; Zolghadr, Kourosh; Tillib, Sergei; Nowak, Danny; Schermelleh, Lothar; Gahl, Anja; Backmann, Natalija; Conrath, Katja; Muyldermans, Serge; Cardoso, M. Cristina; Leonhardt, HeinrichNature Methods (2006), 3 (11), 887-889CODEN: NMAEA3; ISSN:1548-7091. (Nature Publishing Group)The authors fused the epitope-recognizing fragment of heavy-chain antibodies from Camelidae sp. with fluorescent proteins to generate fluorescent, antigen-binding nanobodies (chromobodies) that can be expressed in living cells. The authors demonstrate that chromobodies can recognize and trace antigens in different subcellular compartments throughout S phase and mitosis. Chromobodies should enable new functional studies, as potentially any antigenic structure can be targeted and traced in living cells in this fashion.
- 462Sato, Y.; Mukai, M.; Ueda, J.; Muraki, M.; Stasevich, T. J.; Horikoshi, N.; Kujirai, T.; Kita, H.; Kimura, T.; Hira, S.; Okada, Y.; Hayashi-Takanaka, Y.; Obuse, C.; Kurumizaka, H.; Kawahara, A.; Yamagata, K.; Nozaki, N.; Kimura, H. Sci. Rep. 2013, 3, 2436Google Scholar462Genetically encoded system to track histone modification in vivoSato Yuko; Mukai Masanori; Ueda Jun; Muraki Michiko; Stasevich Timothy J; Horikoshi Naoki; Kujirai Tomoya; Kita Hiroaki; Kimura Taisuke; Hira Seiji; Okada Yasushi; Hayashi-Takanaka Yoko; Obuse Chikashi; Kurumizaka Hitoshi; Kawahara Atsuo; Yamagata Kazuo; Nozaki Naohito; Kimura HiroshiScientific reports (2013), 3 (), 2436 ISSN:.Post-translational histone modifications play key roles in gene regulation, development, and differentiation, but their dynamics in living organisms remain almost completely unknown. To address this problem, we developed a genetically encoded system for tracking histone modifications by generating fluorescent modification-specific intracellular antibodies (mintbodies) that can be expressed in vivo. To demonstrate, an H3 lysine 9 acetylation specific mintbody (H3K9ac-mintbody) was engineered and stably expressed in human cells. In good agreement with the localization of its target acetylation, H3K9ac-mintbody was enriched in euchromatin, and its kinetics measurably changed upon treatment with a histone deacetylase inhibitor. We also generated transgenic fruit fly and zebrafish stably expressing H3K9ac-mintbody for in vivo tracking. Dramatic changes in H3K9ac-mintbody localization during Drosophila embryogenesis could highlight enhanced acetylation at the start of zygotic transcription around mitotic cycle 7. Together, this work demonstrates the broad potential of mintbody and lays the foundation for epigenetic analysis in vivo.
- 463Vincenz, C.; Kerppola, T. K. Proc. Natl. Acad. Sci. U.S.A. 2008, 105, 16572Google ScholarThere is no corresponding record for this reference.
- 464Forster, T. H. Ann. Phys. 1948, 437, 55Google ScholarThere is no corresponding record for this reference.
- 465Helms, V. Fluorescence Resonance Energy Transfer. Principles of Computational Cell Biology; Wiley-VCH: Weinheim, 2008; p 202, ISBN 978-3-527-31555-0.Google ScholarThere is no corresponding record for this reference.
- 466Kanno, T.; Kanno, Y.; Siegel, R. M.; Jang, M. K.; Lenardo, M. J.; Ozato, K. Mol. Cell 2004, 13, 33Google ScholarThere is no corresponding record for this reference.
- 467Sasaki, K.; Ito, T.; Nishino, N.; Khochbin, S.; Yoshida, M. Proc. Natl. Acad. Sci. U.S.A. 2009, 106, 16257Google ScholarThere is no corresponding record for this reference.
- 468Ito, T.; Umehara, T.; Sasaki, K.; Nakamura, Y.; Nishino, N.; Terada, T.; Shirouzu, M.; Padmanabhan, B.; Yokoyama, S.; Ito, A.; Yoshida, M. Chem. Biol. 2011, 18, 495Google Scholar468Real-Time Imaging of Histone H4K12-Specific Acetylation Determines the Modes of Action of Histone Deacetylase and Bromodomain InhibitorsIto, Tamaki; Umehara, Takashi; Sasaki, Kazuki; Nakamura, Yoshihiro; Nishino, Norikazu; Terada, Takaho; Shirouzu, Mikako; Padmanabhan, Balasundaram; Yokoyama, Shigeyuki; Ito, Akihiro; Yoshida, MinoruChemistry & Biology (Cambridge, MA, United States) (2011), 18 (4), 495-507CODEN: CBOLE2; ISSN:1074-5521. (Cell Press)Histone acetylation constitutes an epigenetic mark for transcriptional regulation. Here we developed a fluorescent probe to visualize acetylation of histone H4 Lys12 (H4K12) in living cells using fluorescence resonance energy transfer (FRET) and the binding of the BRD2 bromodomain to acetylated H4K12. Using this probe designated as Histac-K12, we demonstrated that histone H4K12 acetylation is retained in mitosis and that some histone deacetylase (HDAC) inhibitors continue to inhibit cellular HDAC activity even after their removal from the culture. In addn., a small mol. that interferes with ability of the bromodomain to bind to acetylated H4K12 could be assessed using Histac-K12 in cells. Thus, Histac-K12 will serve as a powerful tool not only to understand the dynamics of H4K12-specific acetylation but also to characterize small mols. that modulate the acetylation or interaction status of histones.
- 469Carrillo, L. D.; Krishnamoorthy, L.; Mahal, L. K. J. Am. Chem. Soc. 2006, 128, 14768Google ScholarThere is no corresponding record for this reference.
- 470Newman, R. H.; Zhang, J. Mol. BioSyst. 2008, 4, 496Google Scholar470Visualization of phosphatase activity in living cells with a FRET-based calcineurin activity sensorNewman, Robert H.; Zhang, JinMolecular BioSystems (2008), 4 (6), 496-501CODEN: MBOIBW; ISSN:1742-206X. (Royal Society of Chemistry)Protein kinases and phosphatases are organized into complex intracellular signaling networks designed to coordinate their activities in both space and time. In order to better understand the mol. mechanisms underlying the regulation of signal transduction networks, it is important to define the spatiotemporal dynamics of both protein kinases and phosphatases within their endogenous environment. Herein, we report the development of a genetically-encoded protein biosensor designed to specifically probe the activity of the Ca2+/calmodulin-dependent protein phosphatase, calcineurin. Our reporter design utilizes a phosphatase activity-dependent mol. switch based on the N-terminal regulatory domain of the nuclear factor of activated T-cells as a specific substrate of calcineurin, sandwiched between cyan fluorescent protein and yellow fluorescent protein. Using this reporter, calcineurin activity can be monitored as dephosphorylation-induced increases in fluorescence resonance energy transfer and can be simultaneously imaged with intracellular calcium dynamics. The successful design of a prototype phosphatase activity sensor lays a foundation for studying targeting and compartmentation of phosphatases.
- 471Nagai, Y.; Miyazaki, M.; Aoki, R.; Zama, T.; Inouye, S.; Hirose, K.; Iino, M.; Hagiwara, M. Nat. Biotechnol. 2000, 18, 313Google ScholarThere is no corresponding record for this reference.
- 472Ting, A. Y.; Kain, K. H.; Klemke, R. L.; Tsien, R. Y. Proc. Natl. Acad. Sci. U.S.A. 2001, 98, 15003Google ScholarThere is no corresponding record for this reference.
- 473Kurokawa, K.; Mochizuki, N.; Ohba, Y.; Mizuno, H.; Miyawaki, A.; Matsuda, M. J. Biol. Chem. 2001, 276, 31305Google ScholarThere is no corresponding record for this reference.
- 474Lin, C. W.; Jao, C. Y.; Ting, A. Y. J. Am. Chem. Soc. 2004, 126, 5982Google ScholarThere is no corresponding record for this reference.
- 475Lin, C. W.; Ting, A. Y. Angew. Chem., Int. Ed. 2004, 43, 2940Google ScholarThere is no corresponding record for this reference.
- 476Yoshida, M.; Kijima, M.; Akita, M.; Beppu, T. J. Biol. Chem. 1990, 265, 17174Google ScholarThere is no corresponding record for this reference.
- 477Hanahan, D.; Weinberg, R. A. Cell 2000, 100, 57Google Scholar477The hallmarks of cancerHanahan, Douglas; Weinberg, Robert A.Cell (Cambridge, Massachusetts) (2000), 100 (1), 57-70CODEN: CELLB5; ISSN:0092-8674. (Cell Press)A review is given with many refs. on principles governing the transformation of normal human cells into malignant cancers. The authors suggest that research over the past decades has revealed a small no. of mol., biochem., and cellular traits (acquired capabilities) shared by most and perhaps all types of cancer. Topics included are the acquired capabilities self-sufficiency in growth signals, insensitivity to antigrowth signals, evading apoptosis, limitless replicative potential, sustained angiogenesis, and tissue invasion and metastasis followed by genome instability as an enabling characteristic and alternative pathways to cancer.
- 478Hanahan, D.; Weinberg, R. A. Cell 2011, 144, 646Google Scholar478Hallmarks of cancer: the next generationHanahan, Douglas; Weinberg, Robert A.Cell (Cambridge, MA, United States) (2011), 144 (5), 646-674CODEN: CELLB5; ISSN:0092-8674. (Cell Press)A review. The hallmarks of cancer comprise six biol. capabilities acquired during the multistep development of human tumors. The hallmarks constitute an organizing principle for rationalizing the complexities of neoplastic disease. They include sustaining proliferative signaling, evading growth suppressors, resisting cell death, enabling replicative immortality, inducing angiogenesis, and activating invasion and metastasis. Underlying these hallmarks are genome instability, which generates the genetic diversity that expedites their acquisition, and inflammation, which fosters multiple hallmark functions. Conceptual progress in the last decade has added two emerging hallmarks of potential generality to this list-reprogramming of energy metab. and evading immune destruction. In addn. to cancer cells, tumors exhibit another dimension of complexity: they contain a repertoire of recruited, ostensibly normal cells that contribute to the acquisition of hallmark traits by creating the "tumor microenvironment.". Recognition of the widespread applicability of these concepts will increasingly affect the development of new means to treat human cancer.
- 479Reynoird, N.; Schwartz, B. E.; Delvecchio, M.; Sadoul, K.; Meyers, D.; Mukherjee, C.; Caron, C.; Kimura, H.; Rousseaux, S.; Cole, P. A.; Panne, D.; French, C. A.; Khochbin, S. EMBO J. 2010, 29, 2943Google ScholarThere is no corresponding record for this reference.
- 480Crump, N. T.; Hazzalin, C. A.; Bowers, E. M.; Alani, R. M.; Cole, P. A.; Mahadevan, L. C. Proc. Natl. Acad. Sci. U.S.A. 2011, 108, 7814Google ScholarThere is no corresponding record for this reference.
- 481Santer, F. R.; Höschele, P. P.; Oh, S. J.; Erb, H. H.; Bouchal, J.; Cavarretta, I. T.; Parson, W.; Meyers, D. J.; Cole, P. A.; Culig, Z. Mol. Cancer Ther. 2011, 10, 1644Google Scholar481Inhibition of the Acetyltransferases p300 and CBP Reveals a Targetable Function for p300 in the Survival and Invasion Pathways of Prostate Cancer Cell LinesSanter, Frederic R.; Hoeschele, Philipp P. S.; Oh, Su Jung; Erb, Holger H. H.; Bouchal, Jan; Cavarretta, Ilaria T.; Parson, Walther; Meyers, David J.; Cole, Philip A.; Culig, ZoranMolecular Cancer Therapeutics (2011), 10 (9), 1644-1655CODEN: MCTOCF; ISSN:1535-7163. (American Association for Cancer Research)Inhibitors of histone deacetylases have been approved for clin. application in cancer treatment. On the other hand, histone acetyltransferase (HAT) inhibitors have been less extensively investigated for their potential use in cancer therapy. In prostate cancer, the HATs and coactivators p300 and CBP are upregulated and may induce transcription of androgen receptor (AR)-responsive genes, even in the absence or presence of low levels of AR. To discover a potential anticancer effect of p300/CBP inhibition, we used two different approaches: (i) downregulation of p300 and CBP by specific short interfering RNA (siRNA) and (ii) chem. inhibition of the acetyltransferase activity by a newly developed small mol., C646. Knockdown of p300 by specific siRNA, but surprisingly not of CBP, led to an increase of caspase-dependent apoptosis involving both extrinsic and intrinsic cell death pathways in androgen-dependent and castration-resistant prostate cancer cells. Induction of apoptosis was mediated by several pathways including inhibition of AR function and decrease of the nuclear factor kappa B (NF-κB) subunit p65. Furthermore, cell invasion was decreased upon p300, but not CBP, depletion and was accompanied by lower matrix metalloproteinase (MMP)-2 and MMP-9 transcriptions. Thus, p300 and CBP have differential roles in the processes of survival and invasion of prostate cancer cells. Induction of apoptosis in prostate cancer cells was confirmed by the use of C646. This was substantiated by a decrease of AR function and downregulation of p65 impairing several NF-κB target genes. Taken together, these results suggest that p300 inhibition may be a promising approach for the development of new anticancer therapies. Mol Cancer Ther; 10(9); 1644-55.
- 482Wang, Y.; Toh, H. C.; Chow, P.; Chung, A. Y.; Meyers, D. J.; Cole, P. A.; Ooi, L. L.; Lee, C. G. FASEB J. 2012, 26, 3032Google ScholarThere is no corresponding record for this reference.
- 483Yan, G.; Eller, M. S.; Elm, C.; Larocca, C. A.; Ryu, B.; Panova, I. P.; Dancy, B. M.; Bowers, E. M.; Meyers, D.; Lareau, L.; Cole, P. A.; Taverna, S. D.; Alani, R. M. J. Invest. Dermatol. 2013, 133, 2444Google ScholarThere is no corresponding record for this reference.
- 484Santer, F. R.; Höschele, P. P.; Oh, S. J.; Erb, H. H.; Bouchal, J.; Cavarretta, I. T.; Parson, W.; Meyers, D. J.; Cole, P. A.; Culig, Z. Mol. Cancer Ther. 2011, 10, 1644Google Scholar484Inhibition of the Acetyltransferases p300 and CBP Reveals a Targetable Function for p300 in the Survival and Invasion Pathways of Prostate Cancer Cell LinesSanter, Frederic R.; Hoeschele, Philipp P. S.; Oh, Su Jung; Erb, Holger H. H.; Bouchal, Jan; Cavarretta, Ilaria T.; Parson, Walther; Meyers, David J.; Cole, Philip A.; Culig, ZoranMolecular Cancer Therapeutics (2011), 10 (9), 1644-1655CODEN: MCTOCF; ISSN:1535-7163. (American Association for Cancer Research)Inhibitors of histone deacetylases have been approved for clin. application in cancer treatment. On the other hand, histone acetyltransferase (HAT) inhibitors have been less extensively investigated for their potential use in cancer therapy. In prostate cancer, the HATs and coactivators p300 and CBP are upregulated and may induce transcription of androgen receptor (AR)-responsive genes, even in the absence or presence of low levels of AR. To discover a potential anticancer effect of p300/CBP inhibition, we used two different approaches: (i) downregulation of p300 and CBP by specific short interfering RNA (siRNA) and (ii) chem. inhibition of the acetyltransferase activity by a newly developed small mol., C646. Knockdown of p300 by specific siRNA, but surprisingly not of CBP, led to an increase of caspase-dependent apoptosis involving both extrinsic and intrinsic cell death pathways in androgen-dependent and castration-resistant prostate cancer cells. Induction of apoptosis was mediated by several pathways including inhibition of AR function and decrease of the nuclear factor kappa B (NF-κB) subunit p65. Furthermore, cell invasion was decreased upon p300, but not CBP, depletion and was accompanied by lower matrix metalloproteinase (MMP)-2 and MMP-9 transcriptions. Thus, p300 and CBP have differential roles in the processes of survival and invasion of prostate cancer cells. Induction of apoptosis in prostate cancer cells was confirmed by the use of C646. This was substantiated by a decrease of AR function and downregulation of p65 impairing several NF-κB target genes. Taken together, these results suggest that p300 inhibition may be a promising approach for the development of new anticancer therapies. Mol Cancer Ther; 10(9); 1644-55.
- 485Wang, L.; Gural, A.; Sun, X. J.; Zhao, X.; Perna, F.; Huang, G.; Hatlen, M. A.; Vu, L.; Liu, F.; Xu, H.; Asai, T.; Xu, H.; Deblasio, T.; Menendez, S.; Voza, F.; Jiang, Y.; Cole, P. A.; Zhang, J.; Melnick, A.; Roeder, R. G.; Nimer, S. D. Science 2011, 333, 765Google ScholarThere is no corresponding record for this reference.
- 486Liu, Y.; Wang, L.; Predina, J.; Han, R.; Beier, U. H.; Wang, L. C.; Kapoor, V.; Bhatti, T. R.; Akimova, T.; Singhal, S.; SBrindle, P. K.; Cole, P. A.; Albelda, S. M.; Hancock, W. W. Nat. Med. 2013, 19, 1173Google Scholar486Inhibition of p300 impairs Foxp3+ T regulatory cell function and promotes antitumor immunityLiu, Yujie; Wang, Liqing; Predina, Jarrod; Han, Rongxiang; Beier, Ulf H.; Wang, Liang-Chuan S.; Kapoor, Veena; Bhatti, Tricia R.; Akimova, Tatiana; Singhal, Sunil; Brindle, Paul K.; Cole, Philip A.; Albelda, Steven M.; Hancock, Wayne W.Nature Medicine (New York, NY, United States) (2013), 19 (9), 1173-1177CODEN: NAMEFI; ISSN:1078-8956. (Nature Publishing Group)Forkhead box P3 (Foxp3)+ T regulatory (Treg) cells maintain immune homeostasis and limit autoimmunity but can also curtail host immune responses to various types of tumors. Foxp3+ Treg cells are therefore considered promising targets to enhance antitumor immunity, and approaches for their therapeutic modulation are being developed. However, although studies showing that exptl. depleting Foxp3+ Treg cells can enhance antitumor responses provide proof of principle, these studies lack clear translational potential and have various shortcomings. Histone/protein acetyltransferases (HATs) promote chromatin accessibility, gene transcription and the function of multiple transcription factors and nonhistone proteins. We now report that conditional deletion or pharmacol. inhibition of one HAT, p300 (also known as Ep300 or KAT3B), in Foxp3+ Treg cells increased T cell receptor-induced apoptosis in Treg cells, impaired Treg cell suppressive function and peripheral Treg cell induction, and limited tumor growth in immunocompetent but not in immunodeficient mice. Our data thereby demonstrate that p300 is important for Foxp3+ Treg cell function and homeostasis in vivo and in vitro, and identify mechanisms by which appropriate small-mol. inhibitors can diminish Treg cell function without overtly impairing T effector cell responses or inducing autoimmunity. Collectively, these data suggest a new approach for cancer immunotherapy.
- 487Min, S. W.; Cho, S. H.; Zhou, Y.; Schroeder, S.; Haroutunian, V.; Seeley, W. W.; Huang, E. J.; Shen, Y.; Masliah, E.; Mukherjee, C.; Meyers, D.; Cole, P. A.; Ott, M.; Gan, L. Neuron 2010, 67, 953
Erratum in: Neuron2010, 68, 801
Google ScholarThere is no corresponding record for this reference. - 488Mali, P.; Chou, B. K.; Yen, J.; Ye, Z.; Zou, J.; Dowey, S.; Brodsky, R. A.; Ohm, J. E.; Yu, W.; Baylin, S. B.; Yusa, K.; Bradley, A.; Meyers, D. J.; Mukherjee, C.; Cole, P. A.; Cheng, L. Stem Cells 2010, 28, 713Google ScholarThere is no corresponding record for this reference.
- 489Xu, C. R.; Cole, P. A.; Meyers, D. J.; Kormish, J.; Dent, S.; Zaret, K. S. Science 2011, 332, 963Google ScholarThere is no corresponding record for this reference.
- 490Marek, R.; Coelho, C. M.; Sullivan, R. K.; Baker-Andresen, D.; Li, X.; Ratnu, V.; Dudley, K. J.; Meyers, D.; Mukherjee, C.; Cole, P. A.; Sah, P.; Bredy, T. W. J. Neurosci. 2011, 31, 7486Google Scholar490Paradoxical enhancement of fear extinction memory and synaptic plasticity by inhibition of the histone acetyltransferase p300Marek, Roger; Coelho, Carlos M.; Sullivan, Robert K. P.; Baker-Andresen, Danay; Li, Xiang; Ratnu, Vikram; Dudley, Kevin J.; Meyers, David; Mukherjee, Chandrani; Cole, Philip A.; Sah, Pankaj; Bredy, Timothy W.Journal of Neuroscience (2011), 31 (20), 7486-7491CODEN: JNRSDS; ISSN:0270-6474. (Society for Neuroscience)It is well established that the coordinated regulation of activity-dependent gene expression by the histone acetyltransferase (HAT) family of transcriptional coactivators is crucial for the formation of contextual fear and spatial memory, and for hippocampal synaptic plasticity. However, no studies have examd. the role of this epigenetic mechanism within the infralimbic prefrontal cortex (ILPFC), an area of the brain that is essential for the formation and consolidation of fear extinction memory. Here, the authors report that a postextinction training infusion of a combined p300/CBP inhibitor (Lys-CoA-Tat), directly into the ILPFC, enhances fear extinction memory in mice. These results also demonstrated that the HAT p300 was highly expressed within pyramidal neurons of the ILPFC and that the small-mol. p300-specific inhibitor (C 646) infused into the ILPFC immediately after weak extinction training enhanced the consolidation of fear extinction memory. C 646 infused 6 h after extinction had no effect on fear extinction memory, nor did an immediate postextinction training infusion into the prelimbic prefrontal cortex. Consistent with the behavioral findings, inhibition of p300 activity within the ILPFC facilitated long-term potentiation (LTP) under stimulation conditions that did not evoke long-lasting LTP. These data suggested that one function of p300 activity within the ILPFC is to constrain synaptic plasticity, and that a redn. in the function of this HAT is required for the formation of fear extinction memory.
- 491Knight, Z. A.; Shokat, K. M. Cell 2007, 128, 425Google ScholarThere is no corresponding record for this reference.
- 492Kasper, L. H.; Fukuyama, T.; Biesen, M. A.; Boussouar, F.; Tong, C.; de Pauw, A.; Murray, P. J.; van Deursen, J. M.; Brindle, P. K. Mol. Cell. Biol. 2006, 26, 789Google Scholar492Conditional knockout mice reveal distinct functions for the global transcriptional coactivators CBP and p300 in T-cell developmentKasper, Lawryn H.; Fukuyama, Tomofusa; Biesen, Michelle A.; Boussouar, Faycal; Tong, Caili; de Pauw, Antoine; Murray, Peter J.; van Deursen, Jan M. A.; Brindle, Paul K.Molecular and Cellular Biology (2006), 26 (3), 789-809CODEN: MCEBD4; ISSN:0270-7306. (American Society for Microbiology)The global transcriptional coactivators CREB-binding protein (CBP) and the closely related p300 interact with over 312 proteins, making them among the most heavily connected hubs in the known mammalian protein-protein interactome. It is largely uncertain, however, if these interactions are important in specific cell lineages of adult animals, as homozygous null mutations in either CBP or p300 result in early embryonic lethality in mice. Here we describe a Cre/LoxP conditional p300 null allele (p300flox) that allows for the temporal and tissue-specific inactivation of p300. We used mice carrying p300flax and a CBP conditional knockout allele (CBPflax) in conjunction with an Lck-Cre transgene to delete CBP and p300 starting at the CD4- CD8- double-neg. thymocyte stage of T-cell development. Loss of either p300 or CBP led to a decrease in CD4+ CD8+ double-pos. thymocytes, but an increase in the percentage of CD8+ single-pos. thymocytes seen in CBP mutant mice was not obsd. in p300 mutants. T cells completely lacking both CBP and p300 did not develop normally and were nonexistent or very rare in the periphery, however. T cells lacking CBP or p300 had reduced tumor necrosis factor alpha gene expression in response to phorbol ester and ionophore, while signal-responsive gene expression in CBP- or p300-deficient macrophages was largely intact. Thus, CBP and p300 each supply a surprising degree of redundant coactivation capacity in T cells and macrophages, although each gene has also unique properties in thymocyte development.
- 493Kasper, L. H.; Lerach, S.; Wang, J.; Wu, S.; Jeevan, T.; Brindle, P. K. EMBO J. 2010, 29, 3660Google ScholarThere is no corresponding record for this reference.
- 494Phan, H. M.; Xu, A. W.; Coco, C.; Srajer, G.; Wyszomierski, S.; Evrard, Y. A.; Eckner, R.; Dent, S. Y. Dev. Dyn. 2005, 233, 1337Google ScholarThere is no corresponding record for this reference.
- 495Huang, Y.; Vasilatos, S. N.; Boric, L.; Shaw, P. G.; Davidson, N. E. Breast Cancer Res. Treat. 2012, 131, 777– 89Google Scholar495Inhibitors of histone demethylation and histone deacetylation cooperate in regulating gene expression and inhibiting growth in human breast cancer cellsHuang, Yi; Vasilatos, Shauna N.; Boric, Lamia; Shaw, Patrick G.; Davidson, Nancy E.Breast Cancer Research and Treatment (2012), 131 (3), 777-789CODEN: BCTRD6; ISSN:0167-6806. (Springer)Abnormal activities of histone lysine demethylases (KDMs) and lysine deacetylases (HDACs) are assocd. with aberrant gene expression in breast cancer development. However, the precise mol. mechanisms underlying the crosstalk between KDMs and HDACs in chromatin remodeling and regulation of gene transcription are still elusive. In this study, we showed that treatment of human breast cancer cells with inhibitors targeting the zinc cofactor dependent class I/II HDAC, but not NAD+ dependent class III HDAC, led to significant increase of H3K4me2 which is a specific substrate of histone lysine-specific demethylase 1 (LSD1) and a key chromatin mark promoting transcriptional activation. We also demonstrated that inhibition of LSD1 activity by a pharmacol. inhibitor, pargyline, or siRNA resulted in increased acetylation of H3K9 (AcH3K9). However, siRNA knockdown of LSD2, a homolog of LSD1, failed to alter the level of AcH3K9, suggesting that LSD2 activity may not be functionally connected with HDAC activity. Combined treatment with LSD1 and HDAC inhibitors resulted in enhanced levels of H3K4me2 and AcH3K9, and exhibited synergistic growth inhibition of breast cancer cells. Finally, microarray screening identified a unique subset of genes whose expression was significantly changed by combination treatment with inhibitors of LSD1 and HDAC. Our study suggests that LSD1 intimately interacts with histone deacetylases in human breast cancer cells. Inhibition of histone demethylation and deacetylation exhibits cooperation and synergy in regulating gene expression and growth inhibition, and may represent a promising and novel approach for epigenetic therapy of breast cancer.
- 496Han, H.; Yang, X.; Pandiyan, K.; Liang, G. PLoS One 2013, 8, e75136Google ScholarThere is no corresponding record for this reference.
- 497Prusevich, P.; Kalin, J. H.; Ming, S. A.; Basso, M.; Givens, J.; Li, X.; Hu, J.; Taylor, M. S.; Cieniewicz, A. M.; Hsiao, P. Y.; Huang, R.; Roberson, H.; Adejola, N.; Avery, L. B.; Casero, R. A., Jr.; Taverna, S. D.; Qian, J.; Tackett, A. J.; Ratan, R. R.; McDonald, O. G.; Feinberg, A. P.; Cole, P. A. ACS Chem. Biol. 2014, 9, 1284Google ScholarThere is no corresponding record for this reference.
- 498Yan, G.; Eller, M. S.; Elm, C.; Larocca, C. A.; Ryu, B.; Panova, I. P.; Dancy, B. M.; Bowers, E. M.; Meyers, D.; Lareau, L.; Cole, P. A.; Taverna, S. D.; Alani, R. M. J. Invest. Dermatol. 2013, 133, 2444Google ScholarThere is no corresponding record for this reference.
- 499Das, N.; Dhanawat, M.; Dash, B.; Nagarwal, R. C.; Shrivastava, S. K. Eur. J. Pharm. Sci. 2010, 41, 571Google Scholar499Codrug: An efficient approach for drug optimizationDas, N.; Dhanawat, M.; Dash, B.; Nagarwal, R. C.; Shrivastava, S. K.European Journal of Pharmaceutical Sciences (2010), 41 (5), 571-588CODEN: EPSCED; ISSN:0928-0987. (Elsevier B.V.)A review. Codrug or mutual prodrug is an approach where various effective drugs, which are assocd. with some drawbacks, can be modified by attaching with other drugs of same or different categories directly or via a linkage. More appropriately one can say combining two different pharmacophores with similar or different pharmacol. activities elicit synergistic action or help to target the parent drug to specific site/organ/cells resp. This approach is commonly used to improve physicochem., biopharmaceutical and drug delivery properties of therapeutic agents.
Cited By
Smart citations by scite.ai include citation statements extracted from the full text of the citing article. The number of the statements may be higher than the number of citations provided by ACS Publications if one paper cites another multiple times or lower if scite has not yet processed some of the citing articles.
This article is cited by 416 publications.
- Yanrui Suo, Xu Qian, Zhaoping Xiong, Xiaohong Liu, Chao Wang, Baiyang Mu, Xinyuan Wu, Weiwei Lu, Meiying Cui, Jiaxiang Liu, Yujie Chen, Mingyue Zheng, Xiaojie Lu. Enhancing the Predictive Power of Machine Learning Models through a Chemical Space Complementary DEL Screening Strategy. Journal of Medicinal Chemistry 2024, 67
(21)
, 18969-18980. https://doi.org/10.1021/acs.jmedchem.4c01416
- Zonglong Chen, Hong Yang, Yan Zhang, Xiaodong Lyu, Qiongyu Shi, Cheng Zhang, Xingcan Wang, Zekun Wang, Ying Zhang, Yue Deng, Yujie Wang, Yuting Huang, Yong Xu, Xun Huang, Yingxia Li. Discovery of CZL-046 with an (S)-3-Fluoropyrrolidin-2-one Scaffold as a p300 Bromodomain Inhibitor for the Treatment of Multiple Myeloma. Journal of Medicinal Chemistry 2024, 67
(20)
, 18606-18628. https://doi.org/10.1021/acs.jmedchem.4c01984
- Mengjun Ma, Mengyao Li, Chengwei Zhang, Zixuan Yang, Xiaoyu Chen, Penghui Lu, Shuangshuang Nie, Siqi Zhang, Shumin Ma, Chong Qin. Discovery of a Highly Potent PROTAC Degrader of p300/CBP Proteins for the Treatment of Enzalutamide-Resistant Prostate Cancer. Journal of Medicinal Chemistry 2024, 67
(19)
, 17290-17318. https://doi.org/10.1021/acs.jmedchem.4c01100
- Iván Cheng-Sánchez, Katherine Gosselé, Leonardo Palaferri, Eleen Laul, Gionata Riccabella, Rajiv K. Bedi, Yaozong Li, Anna Müller, Ivan Corbeski, Amedeo Caflisch, Cristina Nevado. Structure-Based Design of CBP/EP300 Degraders: When Cooperativity Overcomes Affinity. JACS Au 2024, 4
(9)
, 3466-3474. https://doi.org/10.1021/jacsau.4c00292
- Li-Yun Chen, Soumya Jyoti Singha Roy, Appaso M. Jadhav, Wesley W. Wang, Pei-Hsin Chen, Timothy Bishop, Michael A. Erb, Christopher G. Parker. Functional Investigations of p53 Acetylation Enabled by Heterobifunctional Molecules. ACS Chemical Biology 2024, 19
(9)
, 1918-1929. https://doi.org/10.1021/acschembio.4c00438
- Xuemin Chen, McKenna C. Crawford, Ying Xiong, Anver Basha Shaik, Kiall F. Suazo, Ludwig G. Bauer, Manini S. Penikalapati, Joycelyn H. Williams, Kilian V. M. Huber, Thorkell Andressen, Rolf E. Swenson, Jordan L. Meier. Paralogue-Selective Degradation of the Lysine Acetyltransferase EP300. JACS Au 2024, 4
(8)
, 3094-3103. https://doi.org/10.1021/jacsau.4c00442
- Yanrui Suo, Kaige Li, Xing Ling, Kenian Yan, Weiwei Lu, Jinfeng Yue, Xiaohua Chen, Zhiqiang Duan, Xiaojie Lu. Discovery Small-Molecule p300 Inhibitors Derived from a Newly Developed Indazolone-Focused DNA-Encoded Library. Bioconjugate Chemistry 2024, 35
(8)
, 1251-1257. https://doi.org/10.1021/acs.bioconjchem.4c00307
- Raj V. Nithun, Yumi Minyi Yao, Omer Harel, Shaimaa Habiballah, Ariel Afek, Muhammad Jbara. Site-Specific Acetylation of the Transcription Factor Protein Max Modulates Its DNA Binding Activity. ACS Central Science 2024, 10
(6)
, 1295-1303. https://doi.org/10.1021/acscentsci.4c00686
- Tianbang Wu, Jiankang Hu, Xiaofan Zhao, Cheng Zhang, Ruibo Dong, Qingqing Hu, Hongrui Xu, Hui Shen, Xiaohan Zhang, Yan Zhang, Bin Lin, Xishan Wu, Qiuping Xiang, Yong Xu. Discovery of a Promising CBP/p300 Degrader XYD129 for the Treatment of Acute Myeloid Leukemia. Journal of Medicinal Chemistry 2024, 67
(11)
, 9194-9213. https://doi.org/10.1021/acs.jmedchem.4c00335
- Jiankang Hu, Hongrui Xu, Tianbang Wu, Cheng Zhang, Hui Shen, Ruibo Dong, Qingqing Hu, Qiuping Xiang, Shuang Chai, Guolong Luo, Xiaoshan Chen, Yumin Huang, Xiaofan Zhao, Chao Peng, Xishan Wu, Bin Lin, Yan Zhang, Yong Xu. Discovery of Highly Potent and Efficient CBP/p300 Degraders with Strong In Vivo Antitumor Activity. Journal of Medicinal Chemistry 2024, 67
(9)
, 6952-6986. https://doi.org/10.1021/acs.jmedchem.3c02195
- Zhixiang Chen, Mi Wang, Dimin Wu, Longchuan Bai, Tianfeng Xu, Hoda Metwally, Yu Wang, Donna McEachern, Lijie Zhao, Ruiting Li, John Takyi-Williams, Meilin Wang, Lu Wang, Qiuxia Li, Bo Wen, Duxin Sun, Shaomeng Wang. Discovery of CBPD-268 as an Exceptionally Potent and Orally Efficacious CBP/p300 PROTAC Degrader Capable of Achieving Tumor Regression. Journal of Medicinal Chemistry 2024, 67
(7)
, 5275-5304. https://doi.org/10.1021/acs.jmedchem.3c02124
- Iván Cheng-Sánchez, Katherine A. Gosselé, Leonardo Palaferri, Mariia S. Kirillova, Cristina Nevado. Discovery and Characterization of Active CBP/EP300 Degraders Targeting the HAT Domain. ACS Medicinal Chemistry Letters 2024, 15
(3)
, 355-361. https://doi.org/10.1021/acsmedchemlett.3c00490
- Qi Chang, Jiayi Li, Yue Deng, Ruilin Zhou, Bingwei Wang, Yujie Wang, Mingming Zhang, Xun Huang, Yingxia Li. Discovery of Novel PROTAC Degraders of p300/CBP as Potential Therapeutics for Hepatocellular Carcinoma. Journal of Medicinal Chemistry 2024, 67
(4)
, 2466-2486. https://doi.org/10.1021/acs.jmedchem.3c01468
- Aparajita Ghosh, Ambati Himaja, Swati Biswas, Onkar Kulkarni, Balaram Ghosh. Advances in the Delivery and Development of Epigenetic Therapeutics for the Treatment of Cancer. Molecular Pharmaceutics 2023, 20
(12)
, 5981-6009. https://doi.org/10.1021/acs.molpharmaceut.3c00610
- McKenna C. Crawford, Deepika R. Tripu, Samuel A. Barritt, Yihang Jing, Diamond Gallimore, Stephen C. Kales, Natarajan V. Bhanu, Ying Xiong, Yuhong Fang, Kamaria A. T. Butler, Christopher A. LeClair, Nathan P. Coussens, Anton Simeonov, Benjamin A. Garcia, Christian C. Dibble, Jordan L. Meier. Comparative Analysis of Drug-like EP300/CREBBP Acetyltransferase Inhibitors. ACS Chemical Biology 2023, 18
(10)
, 2249-2258. https://doi.org/10.1021/acschembio.3c00293
- Md Kabir, Ning Sun, Xiaoping Hu, Tiphaine C. Martin, Jingjie Yi, Yue Zhong, Yan Xiong, H. Ümit Kaniskan, Wei Gu, Ramon Parsons, Jian Jin. Acetylation Targeting Chimera Enables Acetylation of the Tumor Suppressor p53. Journal of the American Chemical Society 2023, 145
(27)
, 14932-14944. https://doi.org/10.1021/jacs.3c04640
- Matthew Jones, Philipp Grosche, Andreas Floersheimer, Jérome André, Raphael Gattlen, Dieter Oser, Juliette Tinchant, Roman Wille, Barbara Chie-Leon, Marc Gerspacher, Peter Ertl, Nils Ostermann, Eva Altmann, Eusebio Manchado, Thomas Vorherr, Patrick Chène. Design and Biochemical Characterization of Peptidic Inhibitors of the Myb/p300 Interaction. Biochemistry 2023, 62
(7)
, 1321-1329. https://doi.org/10.1021/acs.biochem.2c00690
- Zhipeng A. Wang, Jonathan W. Markert, Samuel D. Whedon, Maheeshi Yapa Abeywardana, Kwangwoon Lee, Hanjie Jiang, Carolay Suarez, Hening Lin, Lucas Farnung, Philip A. Cole. Structural Basis of Sirtuin 6-Catalyzed Nucleosome Deacetylation. Journal of the American Chemical Society 2023, 145
(12)
, 6811-6822. https://doi.org/10.1021/jacs.2c13512
- Ryutaro Kanada, Yoshiko Kagoshima, Takashi Suzuki, Akifumi Nakamura, Hideaki Funami, Jun Watanabe, Masayoshi Asano, Mizuki Takahashi, Osamu Ubukata, Kanae Suzuki, Tomoya Aikawa, Kazumi Sato, Megumi Goto, Genzui Setsu, Kentaro Ito, Kawori Kihara, Mutsumi Kuroha, Takashi Kohno, Hideaki Ogiwara, Takeshi Isoyama, Yuichi Tominaga, Saito Higuchi, Hiroyuki Naito. Discovery of DS-9300: A Highly Potent, Selective, and Once-Daily Oral EP300/CBP Histone Acetyltransferase Inhibitor. Journal of Medicinal Chemistry 2023, 66
(1)
, 695-715. https://doi.org/10.1021/acs.jmedchem.2c01641
- Xinrong Tian, Dominic Suarez, Douglas Thomson, William Li, Elizabeth A. King, Louis LaFrance, Jeffrey Boehm, Linda Barton, Christina Di Marco, Cuthbert Martyr, Reema Thalji, Jesus Medina, Steven Knight, Dirk Heerding, Enoch Gao, Eldridge Nartey, Ted Cecconie, Christopher Nixon, Guofeng Zhang, Thomas J. Berrodin, Christopher Phelps, Amish Patel, Xiaopeng Bai, Ken Lind, Ninad Prabhu, Jeffrey Messer, Zhengrong Zhu, Lisa Shewchuk, Rob Reid, Alan P. Graves, Charles McHugh, Biju Mangatt. Discovery of Proline-Based p300/CBP Inhibitors Using DNA-Encoded Library Technology in Combination with High-Throughput Screening. Journal of Medicinal Chemistry 2022, 65
(21)
, 14391-14408. https://doi.org/10.1021/acs.jmedchem.2c00670
- Liwen Hua, Qiuyue Zhang, Xinyue Zhu, Ruoning Wang, Qidong You, Lei Wang. Beyond Proteolysis-Targeting Chimeric Molecules: Designing Heterobifunctional Molecules Based on Functional Effectors. Journal of Medicinal Chemistry 2022, 65
(12)
, 8091-8112. https://doi.org/10.1021/acs.jmedchem.2c00316
- Mary E. Tarantino, Sarah Delaney. Kinetic Analysis of the Effect of N-Terminal Acetylation on Thymine DNA Glycosylase. Biochemistry 2022, 61
(10)
, 895-908. https://doi.org/10.1021/acs.biochem.1c00823
- Zhipeng A. Wang, Samuel D. Whedon, Mingxuan Wu, Siyu Wang, Edward A. Brown, Ananya Anmangandla, Liam Regan, Kwangwoon Lee, Jianfeng Du, Jun Young Hong, Louise Fairall, Taylor Kay, Hening Lin, Yingming Zhao, John W. R. Schwabe, Philip A. Cole. Histone H2B Deacylation Selectivity: Exploring Chromatin’s Dark Matter with an Engineered Sortase. Journal of the American Chemical Society 2022, 144
(8)
, 3360-3364. https://doi.org/10.1021/jacs.1c13555
- Wesley W. Wang, Li-Yun Chen, Jacob M. Wozniak, Appaso M. Jadhav, Hayden Anderson, Taylor E. Malone, Christopher G. Parker. Targeted Protein Acetylation in Cells Using Heterobifunctional Molecules. Journal of the American Chemical Society 2021, 143
(40)
, 16700-16708. https://doi.org/10.1021/jacs.1c07850
- Alex Muthengi, Virangika K. Wimalasena, Hailemichael O. Yosief, Melissa J. Bikowitz, Logan H. Sigua, Tingjian Wang, Deyao Li, Zied Gaieb, Gagan Dhawan, Shuai Liu, Jon Erickson, Rommie E. Amaro, Ernst Schönbrunn, Jun Qi, Wei Zhang. Development of Dimethylisoxazole-Attached Imidazo[1,2-a]pyridines as Potent and Selective CBP/P300 Inhibitors. Journal of Medicinal Chemistry 2021, 64
(9)
, 5787-5801. https://doi.org/10.1021/acs.jmedchem.0c02232
- Jonathan E. Wilson, Gaurav Patel, Chirag Patel, Francois Brucelle, Annissa Huhn, Anna S. Gardberg, Florence Poy, Nico Cantone, Archana Bommi-Reddy, Robert J. Sims, III, Richard T. Cummings, Julian R. Levell. Discovery of CPI-1612: A Potent, Selective, and Orally Bioavailable EP300/CBP Histone Acetyltransferase Inhibitor. ACS Medicinal Chemistry Letters 2020, 11
(6)
, 1324-1329. https://doi.org/10.1021/acsmedchemlett.0c00155
- Fangrui Wu, Yuanda Hua, Salma Kaochar, Shenyou Nie, Yi-Lun Lin, Yuan Yao, Jingyu Wu, Xiaowei Wu, Xiaoyong Fu, Rachel Schiff, Christel M. Davis, Matthew Robertson, Erik A. Ehli, Cristian Coarfa, Nicholas Mitsiades, Yongcheng Song. Discovery, Structure–Activity Relationship, and Biological Activity of Histone-Competitive Inhibitors of Histone Acetyltransferases P300/CBP. Journal of Medicinal Chemistry 2020, 63
(9)
, 4716-4731. https://doi.org/10.1021/acs.jmedchem.9b02164
- Robert E. Thompson, Tom W. Muir. Chemoenzymatic Semisynthesis of Proteins. Chemical Reviews 2020, 120
(6)
, 3051-3126. https://doi.org/10.1021/acs.chemrev.9b00450
- Yaxi Yang, Rukang Zhang, Zhaojun Li, Lianghe Mei, Shili Wan, Hong Ding, Zhifeng Chen, Jing Xing, Huijin Feng, Jie Han, Hualiang Jiang, Mingyue Zheng, Cheng Luo, Bing Zhou. Discovery of Highly Potent, Selective, and Orally Efficacious p300/CBP Histone Acetyltransferases Inhibitors. Journal of Medicinal Chemistry 2020, 63
(3)
, 1337-1360. https://doi.org/10.1021/acs.jmedchem.9b01721
- Beth E. Zucconi, Jessica L. Makofske, David J. Meyers, Yousang Hwang, Mingxuan Wu, Mitzi I. Kuroda, Philip A. Cole. Combination Targeting of the Bromodomain and Acetyltransferase Active Site of p300/CBP. Biochemistry 2019, 58
(16)
, 2133-2143. https://doi.org/10.1021/acs.biochem.9b00160
- Jonathan H. Shrimp, Carissa Grose, Stephanie R. T. Widmeyer, Abigail L. Thorpe, Ajit Jadhav, and Jordan L. Meier . Chemical Control of a CRISPR-Cas9 Acetyltransferase. ACS Chemical Biology 2018, 13
(2)
, 455-460. https://doi.org/10.1021/acschembio.7b00883
- Ibraheem Ali, Ryan J. Conrad, Eric Verdin, and Melanie Ott . Lysine Acetylation Goes Global: From Epigenetics to Metabolism and Therapeutics. Chemical Reviews 2018, 118
(3)
, 1216-1252. https://doi.org/10.1021/acs.chemrev.7b00181
- Michael R. Michaelides, Arthur Kluge, Michael Patane, John H. Van Drie, Ce Wang, T. Matthew Hansen, Roberto M. Risi, Robert Mantei, Carmen Hertel, Kannan Karukurichi, Alexandre Nesterov, David McElligott, Peter de Vries, J. William Langston, Philip A. Cole, Ronen Marmorstein, Hong Liu, Loren Lasko, Kenneth D. Bromberg, Albert Lai, and Edward A. Kesicki . Discovery of Spiro Oxazolidinediones as Selective, Orally Bioavailable Inhibitors of p300/CBP Histone Acetyltransferases. ACS Medicinal Chemistry Letters 2018, 9
(1)
, 28-33. https://doi.org/10.1021/acsmedchemlett.7b00395
- Sarah M. Bronner, Jeremy Murray, F. Anthony Romero, Kwong Wah Lai, Vickie Tsui, Patrick Cyr, Maureen H. Beresini, Gladys de leon Boenig, Zhongguo Chen, Edna F. Choo, Kevin R. Clark, Terry D. Crawford, Hariharan Jayaram, Susan Kaufman, Ruina Li, Yingjie Li, Jiangpeng Liao, Xiaorong Liang, Wenfeng Liu, Justin Ly, Jonathan Maher, John Wai, Fei Wang, Aijun Zheng, Xiaoyu Zhu, and Steven Magnuson . A Unique Approach to Design Potent and Selective Cyclic Adenosine Monophosphate Response Element Binding Protein, Binding Protein (CBP) Inhibitors. Journal of Medicinal Chemistry 2017, 60
(24)
, 10151-10171. https://doi.org/10.1021/acs.jmedchem.7b01372
- Tingjun Chen, Dan Gao, Roushu Zhang, Guihua Zeng, Hao Yan, Eunju Lim, and Fu-Sen Liang . Chemically Controlled Epigenome Editing through an Inducible dCas9 System. Journal of the American Chemical Society 2017, 139
(33)
, 11337-11340. https://doi.org/10.1021/jacs.7b06555
- Zhen Han, Chau-wen Chou, Xiangkun Yang, Michael G. Bartlett, and Y. George Zheng . Profiling Cellular Substrates of Lysine Acetyltransferases GCN5 and p300 with Orthogonal Labeling and Click Chemistry. ACS Chemical Biology 2017, 12
(6)
, 1547-1555. https://doi.org/10.1021/acschembio.7b00114
- Mostafa M. Hamed, Sarah S. Darwish, Jennifer Herrmann, Ashraf H. Abadi, and Matthias Engel . First Bispecific Inhibitors of the Epidermal Growth Factor Receptor Kinase and the NF-κB Activity As Novel Anticancer Agents. Journal of Medicinal Chemistry 2017, 60
(7)
, 2853-2868. https://doi.org/10.1021/acs.jmedchem.6b01774
- Beth E. Zucconi, Birgit Luef, Wei Xu, Ryan A. Henry, Ilana M. Nodelman, Gregory D. Bowman, Andrew J. Andrews, and Philip A. Cole . Modulation of p300/CBP Acetylation of Nucleosomes by Bromodomain Ligand I-CBP112. Biochemistry 2016, 55
(27)
, 3727-3734. https://doi.org/10.1021/acs.biochem.6b00480
- David C. Montgomery, Julie M. Garlick, Rhushikesh A. Kulkarni, Steven Kennedy, Abdellah Allali-Hassani, Yin-Ming Kuo, Andrew J. Andrews, Hong Wu, Masoud Vedadi, and Jordan L. Meier . Global Profiling of Acetyltransferase Feedback Regulation. Journal of the American Chemical Society 2016, 138
(20)
, 6388-6391. https://doi.org/10.1021/jacs.6b03036
- Cheryl E. McCullough and Ronen Marmorstein . Molecular Basis for Histone Acetyltransferase Regulation by Binding Partners, Associated Domains, and Autoacetylation. ACS Chemical Biology 2016, 11
(3)
, 632-642. https://doi.org/10.1021/acschembio.5b00841
- Roman P. Simon, Dina Robaa, Zayan Alhalabi, Wolfgang Sippl, and Manfred Jung . KATching-Up on Small Molecule Modulators of Lysine Acetyltransferases. Journal of Medicinal Chemistry 2016, 59
(4)
, 1249-1270. https://doi.org/10.1021/acs.jmedchem.5b01502
- Xueying Yuan, Jeffrey M. Rosen. Histone acetylation modulators in breast cancer. Breast Cancer Research 2025, 27
(1)
https://doi.org/10.1186/s13058-025-02006-9
- Lili Li, Huabiao Zhou, Miaomiao Li, Wenbo Liu, Yuxuan Li, Hanyang Xu, Jiemei Jiang, Yan Yang, Yongfang Gong. Salvianolic acid B ameliorates hepatic fibrosis via inhibiting p300/CBP. European Journal of Pharmacology 2025, 998 , 177495. https://doi.org/10.1016/j.ejphar.2025.177495
- Rahma Benhassoun, Anne-Pierre Morel, Victoria Jacquot, Alain Puisieux, Maria Ouzounova. The epipliancy journey: Tumor initiation at the mercy of identity crisis and epigenetic drift. Biochimica et Biophysica Acta (BBA) - Reviews on Cancer 2025, 1880
(3)
, 189307. https://doi.org/10.1016/j.bbcan.2025.189307
- Francesco Fiorentino, Filippo Spriano, Daniela Tomaselli, Giorgia Risi, Emanuele Fabbrizi, Valeria Pecci, Simona Nanni, Francesco Bertoni, Dante Rotili, Antonello Mai. Development of C646‐Based Proteolysis Targeting Chimeras Degraders of the Lysine Acetyltransferases CBP and p300. ChemMedChem 2025, 10 https://doi.org/10.1002/cmdc.202400792
- Jiasheng Geng, Weina Lu, Qinglong Kong, Jiao Lv, Yue Liu, Guowei Zu, Yanmei Chen, Caiying Jiang, Zhengying You, Zuoming Nie. Validation of selective catalytic
BmCBP
inhibitors that regulate the
Bm30K
‐24 protein expression in silkworm,
Bombyx mori. Insect Molecular Biology 2025, 34
(2)
, 322-334. https://doi.org/10.1111/imb.12974
- Schilin Wen, Javiera Santander, Daniel Barria, Luis A. Salazar, Cristian Sandoval, Consuelo Arias, Verónica Iturriaga. Epigenetic Biomarkers in Temporomandibular Joint Osteoarthritis: An Emerging Target in Treatment. International Journal of Molecular Sciences 2025, 26
(8)
, 3668. https://doi.org/10.3390/ijms26083668
- Keqiang Rao, Xinchao Zhang, Yi Luo, Qiang Xia, Yuting Jin, Jing He. Lactylation orchestrates ubiquitin-independent degradation of cGAS and promotes tumor growth. Cell Reports 2025, 44
(4)
, 115441. https://doi.org/10.1016/j.celrep.2025.115441
- Jiaxin Tian, Tongyu Zhang, Ruidan Zhang, Sijia Hao, Jingyu Dong, Yiyan Chen, Jinpeng Zhou, Yang Tian. Lactylation in CNS disorders: mechanisms, cellular function, and disease relevance. Frontiers in Cell and Developmental Biology 2025, 13 https://doi.org/10.3389/fcell.2025.1566921
- Chao Ma, Wangang Zhang, Lujuan Xing. Differences in protein lactylation between pale, soft and exudative and red, firm and non-exudative pork. Meat Science 2025, 221 , 109736. https://doi.org/10.1016/j.meatsci.2024.109736
- Joseph Adu‐Amankwaah, Qiang You, Xiaoer Liu, Jiayi Jiang, Dongqi Yang, Kuntao Liu, Jinxiang Yuan, Yanfang Wang, Qinghua Hu, Rubin Tan. Pulmonary Hypertension: Molecular Mechanisms and Clinical Studies. MedComm 2025, 6
(3)
https://doi.org/10.1002/mco2.70134
- Felix Osten, Alea K. Bodenschatz, Karina Ivaskevica, Simon Kröhn, Birgit Piep, Tim Holler, Jana Teske, Judith Montag, Bogdan Iorga, Natalie Weber, Robert Zweigerdt, Theresia Kraft, Joachim D. Meissner. Differential impact of substrates on myosin heavy and light chain expression in human stem cell-derived cardiomyocytes at single-cell level. Journal of Muscle Research and Cell Motility 2025, 20 https://doi.org/10.1007/s10974-025-09690-2
- Longxia Xu, Hongwen Xuan, Xiaobing Shi. Dysregulation of the p300/CBP histone acetyltransferases in human cancer. Epigenomics 2025, 17
(3)
, 193-208. https://doi.org/10.1080/17501911.2024.2447807
- Nana Ma, Lairong Wang, Meijuan Meng, Yan Wang, Ran Huo, Guangjun Chang, Xiangzhen Shen. D-sodium lactate promotes the activation of NF-κB signaling pathway induced by lipopolysaccharide via histone lactylation in bovine mammary epithelial cells. Microbial Pathogenesis 2025, 199 , 107198. https://doi.org/10.1016/j.micpath.2024.107198
- Ming-Ming Zhou, Philip A. Cole. Targeting lysine acetylation readers and writers. Nature Reviews Drug Discovery 2025, 24
(2)
, 112-133. https://doi.org/10.1038/s41573-024-01080-6
- Yilan Yang, Jiatao Liao, Zhe Pan, Jin Meng, Li Zhang, Wei Shi, Xiaofang Wang, Xiaomeng Zhang, Zhirui Zhou, Jurui Luo, Xingxing Chen, Zhaozhi Yang, Xin Mei, Jinli Ma, Zhen Zhang, Yi‐Zhou Jiang, Zhi‐Min Shao, Fei Xavier Chen, Xiaoli Yu, Xiaomao Guo. Dual Inhibition of CDK4/6 and CDK7 Suppresses Triple‐Negative Breast Cancer Progression via Epigenetic Modulation of SREBP1‐Regulated Cholesterol Metabolism. Advanced Science 2025, 12
(5)
https://doi.org/10.1002/advs.202413103
- Jian-kang Hu, Xin Tang, Guo-long Luo, Cheng Zhang, Tian-bang Wu, Chao Wang, Hui Shen, Xiao-fan Zhao, Xi-shan Wu, Jeff B. Smaill, Yong Xu, Yan Zhang, Qiu-ping Xiang. Discovery of 5-imidazole-3-methylbenz[d]isoxazole derivatives as potent and selective CBP/p300 bromodomain inhibitors for the treatment of acute myeloid leukemia. Acta Pharmacologica Sinica 2025, 85 https://doi.org/10.1038/s41401-025-01478-x
- Megan J Ritter, Izuki Amano, Anthony N Hollenberg. Transcriptional Cofactors for Thyroid Hormone Receptors. Endocrinology 2025, 166
(2)
https://doi.org/10.1210/endocr/bqae164
- Nolan D. Swain, Y. George Zheng. Design of a Bioluminescent Assay Platform for Quantitative Measurement of Histone Acetyltransferase Enzymatic Activity. ChemBioChem 2025, 26
(1)
https://doi.org/10.1002/cbic.202400692
- Yao Deng, Guanxi Wen, Yongtao Yin, Dandan Chen, Difei Li, Rongchang Chen. Pharmacological inhibition of P300 with C646 ameliorates LPS-induced acute lung injury by modulating CXCL1 in M1 alveolar macrophages. International Immunopharmacology 2025, 144 , 113674. https://doi.org/10.1016/j.intimp.2024.113674
- Graham P. Marsh, Mark S. Cooper, Sean Goggins, Stephen J. Reynolds, Dean F. Wheeler, Joel O. Cresser-Brown, Robert E. Arnold, Emily G. Babcock, Gareth Hughes, Darko Bosnakovski, Michael Kyba, Samuel Ojeda, Drew A. Harrison, Christopher J. Ott, Hannah J. Maple. Development of p300-targeting degraders with enhanced selectivity and onset of degradation. RSC Medicinal Chemistry 2025, 550 https://doi.org/10.1039/D4MD00969J
- Swarna Beesetti. Ubiquitin Ligases in Control: Regulating NLRP3 Inflammasome Activation. Frontiers in Bioscience-Landmark 2025, 30
(3)
https://doi.org/10.31083/FBL25970
- Xuan Zhang, Lei Sun. Activated PRDM1-CREBBP contributes to preeclampsia by regulating apoptosis and invasion of the human trophoblast cells. iScience 2024, 27
(12)
, 111484. https://doi.org/10.1016/j.isci.2024.111484
- Noha A. M. Shendy, Melissa Bikowitz, Logan H. Sigua, Yang Zhang, Audrey Mercier, Yousef Khashana, Stephanie Nance, Qi Liu, Ian M. Delahunty, Sarah Robinson, Vanshita Goel, Matthew G. Rees, Melissa A. Ronan, Tingjian Wang, Mustafa Kocak, Jennifer A. Roth, Yingzhe Wang, Burgess B. Freeman, Brent A. Orr, Brian J. Abraham, Martine F. Roussel, Ernst Schonbrunn, Jun Qi, Adam D. Durbin. Group 3 medulloblastoma transcriptional networks collapse under domain specific EP300/CBP inhibition. Nature Communications 2024, 15
(1)
https://doi.org/10.1038/s41467-024-47102-0
- Priya Mittal, Jacquelyn A. Myers, Raymond D. Carter, Sandi Radko-Juettner, Hayden A. Malone, Wojciech Rosikiewicz, Alexis N. Robertson, Zhexin Zhu, Ishwarya V. Narayanan, Baranda S. Hansen, Meadow Parrish, Natarajan V. Bhanu, Robert J. Mobley, Jerold E. Rehg, Beisi Xu, Yiannis Drosos, Shondra M. Pruett-Miller, Mats Ljungman, Benjamin A. Garcia, Gang Wu, Janet F. Partridge, Charles W. M. Roberts. PHF6 cooperates with SWI/SNF complexes to facilitate transcriptional progression. Nature Communications 2024, 15
(1)
https://doi.org/10.1038/s41467-024-51566-5
- Shuangshuang Sun, Zhe Xu, Liying He, Yihui Shen, Yuqing Yan, Xubing Lv, Xujing Zhu, Wei Li, Wei-Ya Tian, Yongjun Zheng, Sen Lin, Yadong Sun, Lei Li. Metabolic regulation of cytoskeleton functions by HDAC6-catalyzed α-tubulin lactylation. Nature Communications 2024, 15
(1)
https://doi.org/10.1038/s41467-024-52729-0
- Julien Van Gils, Slim Karkar, Aurélien Barre, Seyta Ley-Ngardigal, Sophie Nothof, Stéphane Claverol, Caroline Tokarski, Jean-Philippe Trani, Raphael Chevalier, Natacha Broucqsault, Claire El Yazidi, Didier Lacombe, Patricia Fergelot, Frédérique Magdinier. Transcriptome and acetylome profiling identify crucial steps of neuronal differentiation in Rubinstein-Taybi syndrome. Communications Biology 2024, 7
(1)
https://doi.org/10.1038/s42003-024-06939-3
- Ashraf K. Shahib, Mojgan Rastegar, Andre J. van Wijnen, James R. Davie. Neurodevelopmental functions and activities of the KAT3 class of lysine acetyltransferases. Biochemistry and Cell Biology 2024, 102
(6)
, 430-447. https://doi.org/10.1139/bcb-2024-0156
- Ruimin He, Yifan Liu, Weijie Fu, Xuan He, Shuang Liu, Desheng Xiao, Yongguang Tao. Mechanisms and cross-talk of regulated cell death and their epigenetic modifications in tumor progression. Molecular Cancer 2024, 23
(1)
https://doi.org/10.1186/s12943-024-02172-y
- Yao Wang, Ping Li, Yuan Xu, Linyu Feng, Yongkang Fang, Guini Song, Li Xu, Zhou Zhu, Wei Wang, Qi Mei, Minjie Xie. Lactate metabolism and histone lactylation in the central nervous system disorders: impacts and molecular mechanisms. Journal of Neuroinflammation 2024, 21
(1)
https://doi.org/10.1186/s12974-024-03303-4
- Qun Wei, Cailing Gan, Meng Sun, Yuting Xie, Hongyao Liu, Taixiong Xue, Conghui Deng, Chunheng Mo, Tinghong Ye. BRD4: an effective target for organ fibrosis. Biomarker Research 2024, 12
(1)
https://doi.org/10.1186/s40364-024-00641-6
- Simon Maria Kienle, Tobias Schneider, Christine Bernecker, Janina Bracker, Andreas Marx, Michael Kovermann, Martin Scheffner, Katrin Stuber. Biochemical and Structural Consequences of NEDD8 Acetylation. ChemBioChem 2024, 25
(22)
https://doi.org/10.1002/cbic.202400478
- Xueying Yuan, Xiaoxin Hao, Hilda L. Chan, Na Zhao, Diego A. Pedroza, Fengshuo Liu, Kang Le, Alex J. Smith, Sebastian J. Calderon, Nadia Lieu, Michael J. Soth, Philip Jones, Xiang H.F. Zhang, Jeffrey M. Rosen. CREB-binding protein/P300 bromodomain inhibition reduces neutrophil accumulation and activates antitumor immunity in triple-negative breast cancer. JCI Insight 2024, 9
(20)
https://doi.org/10.1172/jci.insight.182621
- Chloe P. Whitworth, Wen Y. Aw, Elizabeth L. Doherty, Chenchen Handler, Yogeshwari Ambekar, Aanya Sawhney, Giuliano Scarcelli, William J. Polacheck. P300 Modulates Endothelial Mechanotransduction of Fluid Shear Stress. Cellular and Molecular Bioengineering 2024, 17
(5)
, 507-523. https://doi.org/10.1007/s12195-024-00805-2
- Sharon Y.R. Dent. KAT tales: Functions of Gcn5 and PCAF lysine acetyltransferases in SAGA and ATAC. Journal of Biological Chemistry 2024, 300
(10)
, 107744. https://doi.org/10.1016/j.jbc.2024.107744
- Runhua Du, Yanmei Gao, Cong Yan, Xuelian Ren, Shankang Qi, Guobin Liu, Xinlong Guo, Xiaohan Song, Hanmin Wang, Jingxin Rao, Yi Zang, Mingyue Zheng, Jia Li, He Huang. Sirtuin 1/sirtuin 3 are robust lysine delactylases and sirtuin 1-mediated delactylation regulates glycolysis. iScience 2024, 27
(10)
, 110911. https://doi.org/10.1016/j.isci.2024.110911
- Qi Liu, Aimee L. Qi, Adam D. Durbin, Jun Qi. Chemical Targeting of Histone Acetyltransferases. 2024, 266-306. https://doi.org/10.1039/9781837674916-00266
- . Drug Discovery for Chromatin Readers, Writers, Erasers, Remodelers and Epitranscriptomic Targets. 2024https://doi.org/10.1039/9781837674916-part3
- Emiliano Bolesani, Dorothee Bornhorst, Lavanya M Iyer, Dorota Zawada, Nina Friese, Michael Morgan, Lucas Lange, David M Gonzalez, Nadine Schrode, Andreas Leffler, Julian Wunder, Annika Franke, Lika Drakhlis, Robert Sebra, Axel Schambach, Alexander Goedel, Nicole C Dubois, Gergana Dobreva, Alessandra Moretti, Laura C Zelaráyan, Salim Abdelilah-Seyfried, Robert Zweigerdt. Transient stabilization of human cardiovascular progenitor cells from human pluripotent stem cells
in vitro
reflects stage-specific heart development
in vivo. Cardiovascular Research 2024, 120
(11)
, 1295-1311. https://doi.org/10.1093/cvr/cvae118
- Ye Chen, Tian Lan. N-terminal domain of androgen receptor is a major therapeutic barrier and potential pharmacological target for treating castration resistant prostate cancer: a comprehensive review. Frontiers in Pharmacology 2024, 15 https://doi.org/10.3389/fphar.2024.1451957
- Shuhong Tian, Keyan Zhong, Zhaoxin Yang, Jian Fu, Yangbo Cai, Min Xiao. Investigating the mechanism of tricyclic decyl benzoxazole -induced apoptosis in liver Cancer cells through p300-mediated FOXO3 activation. Cellular Signalling 2024, 121 , 111280. https://doi.org/10.1016/j.cellsig.2024.111280
- Jason P. Smith, Robert Paxton, Silvia Medrano, Nathan C. Sheffield, Maria Luisa S. Sequeira-Lopez, R. Ariel Gomez. Inhibition of Renin Expression Is Regulated by an Epigenetic Switch From an Active to a Poised State. Hypertension 2024, 81
(9)
, 1869-1882. https://doi.org/10.1161/HYPERTENSIONAHA.124.22886
- Jingyi Cai, Yudi Deng, Ziyang Min, Chaoyuan Li, Zhihe Zhao, Dian Jing. Deciphering the dynamics: Exploring the impact of mechanical forces on histone acetylation. The FASEB Journal 2024, 38
(15)
https://doi.org/10.1096/fj.202400907RR
- Hitoshi Ohtani, Minmin Liu, Gangning Liang, H Josh Jang, Peter A Jones. Efficient activation of hundreds of LTR12C elements reveals
cis
-regulatory function determined by distinct epigenetic mechanisms. Nucleic Acids Research 2024, 52
(14)
, 8205-8217. https://doi.org/10.1093/nar/gkae498
- Zijian Yang, Yingqi Zheng, Qiang Gao. Lysine lactylation in the regulation of tumor biology. Trends in Endocrinology & Metabolism 2024, 35
(8)
, 720-731. https://doi.org/10.1016/j.tem.2024.01.011
- Aaron Waddell, Nicole Grbic, Kassidy Leibowitz, William Austin Wyant, Sabah Choudhury, Kihyun Park, Marianne Collard, Philip A. Cole, Rhoda M. Alani. p300 KAT Regulates SOX10 Stability and Function in Human Melanoma. Cancer Research Communications 2024, 4
(8)
, 1894-1907. https://doi.org/10.1158/2767-9764.CRC-24-0124
- Xiaorui Su, Yuxuan Li, Yupeng Ren, Mingbo Cao, Gaoyuan Yang, Jing Luo, Ziyi Hu, Haixia Deng, Meihai Deng, Bo Liu, Zhicheng Yao. A new strategy for overcoming drug resistance in liver cancer: Epigenetic regulation. Biomedicine & Pharmacotherapy 2024, 176 , 116902. https://doi.org/10.1016/j.biopha.2024.116902
- Dabo Pan, Yaxuan Huang, Dewen Jiang, Yonghao Zhang, Mingkai Wu, Minzhen Han, Xiaojie Jin. Discovery of an EP300 Inhibitor using Structure-based Virtual Screening and
Bioactivity Evaluation. Current Pharmaceutical Design 2024, 30
(25)
, 1985-1994. https://doi.org/10.2174/0113816128298051240529113313
- Didier Lacombe, Agnès Bloch-Zupan, Cecilie Bredrup, Edward B Cooper, Sofia Douzgou Houge, Sixto García-Miñaúr, Hülya Kayserili, Lidia Larizza, Vanesa Lopez Gonzalez, Leonie A Menke, Donatella Milani, Francesco Saettini, Cathy A Stevens, Lloyd Tooke, Jill A Van der Zee, Maria M Van Genderen, Julien Van-Gils, Jane Waite, Jean-Louis Adrien, Oliver Bartsch, Pierre Bitoun, Antonia H M Bouts, Anna M Cueto-González, Elena Dominguez-Garrido, Floor A Duijkers, Patricia Fergelot, Elizabeth Halstead, Sylvia A Huisman, Camilla Meossi, Jo Mullins, Sarah M Nikkel, Chris Oliver, Elisabetta Prada, Alessandra Rei, Ilka Riddle, Cristina Rodriguez-Fonseca, Rebecca Rodríguez Pena, Janet Russell, Alicia Saba, Fernando Santos-Simarro, Brittany N Simpson, David F Smith, Markus F Stevens, Katalin Szakszon, Emmanuelle Taupiac, Nadia Totaro, Irene Valenzuena Palafoll, Daniëlle C M Van Der Kaay, Michiel P Van Wijk, Klea Vyshka, Susan Wiley, Raoul C Hennekam. Diagnosis and management in Rubinstein-Taybi syndrome: first international consensus statement. Journal of Medical Genetics 2024, 61
(6)
, 503-519. https://doi.org/10.1136/jmg-2023-109438
- Mélanie Lavaud, Robel Tesfaye, Léa Lassous, Bénédicte Brounais, Marc Baud'huin, Franck Verrecchia, François Lamoureux, Steven Georges, Benjamin Ory. Super-enhancers: drivers of cells’ identities and cells’ debacles. Epigenomics 2024, 16
(9)
, 681-700. https://doi.org/10.2217/epi-2023-0409
- Qiuzi Dai, Zigao Yuan, Qinsheng Sun, Zhuolin Ao, Binsheng He, Yuyang Jiang. Discovery of novel nucleoside derivatives as selective lysine acetyltransferase p300 inhibitors for cancer therapy. Bioorganic & Medicinal Chemistry Letters 2024, 104 , 129742. https://doi.org/10.1016/j.bmcl.2024.129742
- Yue Min, Hong Yu, Qi Li. Transcriptional and post-translational regulation of MITF mediated by bHLH domain during the melanogenesis and melanocyte proliferation in Crassostrea gigas. International Journal of Biological Macromolecules 2024, 266 , 131138. https://doi.org/10.1016/j.ijbiomac.2024.131138
- Magdalena Strachowska, Agnieszka Robaszkiewicz. Characteristics of anticancer activity of CBP/p300 inhibitors – Features of their classes, intracellular targets and future perspectives of their application in cancer treatment. Pharmacology & Therapeutics 2024, 257 , 108636. https://doi.org/10.1016/j.pharmthera.2024.108636
- Dustin C. Becht, Akinori Kanai, Soumi Biswas, Mohamed Halawa, Lei Zeng, Khan L. Cox, Michael G. Poirier, Ming-Ming Zhou, Xiaobing Shi, Akihiko Yokoyama, Tatiana G. Kutateladze. The winged helix domain of MORF binds CpG islands and the TAZ2 domain of p300. iScience 2024, 27
(4)
, 109367. https://doi.org/10.1016/j.isci.2024.109367
- Tsuyoshi Konuma, Ming-Ming Zhou. Distinct Histone H3 Lysine 27 Modifications Dictate Different Outcomes of Gene Transcription. Journal of Molecular Biology 2024, 436
(7)
, 168376. https://doi.org/10.1016/j.jmb.2023.168376
- Azharul Islam, Anirban Chakraborty, Altaf H Sarker, Uma K Aryal, Lang Pan, Gulshan Sharma, Istvan Boldogh, Tapas Hazra. Site-specific acetylation of polynucleotide kinase 3′-phosphatase regulates its distinct role in DNA repair pathways. Nucleic Acids Research 2024, 52
(5)
, 2416-2433. https://doi.org/10.1093/nar/gkae002
- Yun-Bin Lee, Hyun-Woo Rhee. Spray-type modifications: an emerging paradigm in post-translational modifications. Trends in Biochemical Sciences 2024, 49
(3)
, 208-223. https://doi.org/10.1016/j.tibs.2024.01.008
- Panhong Gou, Wenchao Zhang. Protein lysine acetyltransferase CBP/p300: A promising target for small molecules in cancer treatment. Biomedicine & Pharmacotherapy 2024, 171 , 116130. https://doi.org/10.1016/j.biopha.2024.116130
- Ruizhi Lai, Zhiqian Lin, Chunyan Yang, Li Hai, Zhongzhen Yang, Li Guo, Ruifang Nie, Yong Wu. Novel berberine derivatives as p300 histone acetyltransferase inhibitors in combination treatment for breast cancer. European Journal of Medicinal Chemistry 2024, 266 , 116116. https://doi.org/10.1016/j.ejmech.2023.116116
- Ning Wang, Xiaoyu Su, David Sams, Nanduri R. Prabhakar, Jayasri Nanduri. P300/CBP Regulates HIF-1–Dependent Sympathetic Activation and Hypertension by Intermittent Hypoxia. American Journal of Respiratory Cell and Molecular Biology 2024, 70
(2)
, 110-118. https://doi.org/10.1165/rcmb.2022-0481OC
Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.
Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.
The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.
Recommended Articles
Abstract
Figure 1
Figure 1. Reversible acetylation of the epsilon amine group of lysine side chains.
Figure 2
Figure 2. Major metabolic processes that produce or consume acetyl-CoA. Processes occurring in the cytoplasm are indicated using purple font, and processes occurring in the mitochondrion are indicated using orange font. Note that PDH can also be nuclear. This figure was adapted in part from Albaugh et al. (30)
Figure 3
Figure 3. Domain structure of p300/CBP. Exon–intron gene diagrams are shown for p300 and CBP (top). Below are example protein structures for the bromodomain (PDB 3I3J, 2.33 Å), catalytic HAT domain (PDB 3BIY, 1.7 Å), ZZ zinc finger (PDB 1TOT), and TAZ2 domain (PDB 3IO2, 2.5 Å). All structures were produced using purified p300, except the ZZ zinc finger, which used purified CBP. p300/CBP proteins are colored with a rainbow, with blue at the N-terminus and red at the C-terminus, and residues included in the structure are listed below each. Zinc ions are black spheres. All structures are based on X-ray crystallography, except the ZZ zinc finger structures from solution NMR. The p300 bromodomain structure shown here is remarkably similar to an independently generated CBP bromodomain structure (not shown, PDB 3DWY, 1.98 Å). Below is a model for full-length p300/CBP with all domains shown, and is a compilation based on several recent analyses.: (94, 95) three cysteine/histidine-rich (C/H) domains are shown in turquoise, three zinc fingers are shown in yellow, and the catalytic acetyltransferase domain is shown in orange, with its autoacetylated regulatory loop drawn above, which corresponds to residues 1523–1554. A few examples of proteins that bind p300/CBP are listed below the protein model, with the particular domain involved in binding indicated with a black line. Below that, amino acid similarity is indicated, for comparing p300 and CBP sequences within either the catalytic BHC region (from the bromodomain to the C/H3 domain) or the entire protein. At the bottom, commonly purified active p300 variants are indicated, including p300 acetyltransferase/HAT domain, BHC enzyme (bromodomain-HAT-C/H3), and full-length protein. It should be noted that p300 HAT has a deletion in residues 1529–1560.
Figure 4
Figure 4. p300/CBP is central to many important signaling pathways. These include pathways that respond to intracellular signals (turquoise), extracellular signals (purple), and intercellular signals (blue). These pathways control the key cellular functions via altering expression of target genes, through the action of p300/CBP in the nucleus.
Figure 5
Figure 5. p300/CBP functions as a scaffold, bridge, and acetyltransferase. The acetyltransferase reactions are illustrated by turquoise arrows, indicating acetylation of histone and nonhistone substrates (in yellow), as well as autoacetylation of the p300/CBP acetyltransferase domain. The bridge function is illustrated by turquoise squares, representing DNA-binding proteins that bring DNA elements into proximity with p300/CBP through their interactions. The scaffold function is illustrated by orange squares, representing a protein complex being recruited by p300/CBP. These functions together allow for gene expression.
Figure 6
Figure 6. p300 acetyltransferase domain structure bound to Lys-CoA. (A) Secondary structures of p300 acetyltransferase domain. (B) L1 loop and an acidic surface. (C) Parts of Lys-CoA bisubstrate analog (gray) and four p300 residues of interest (green). Generated in PyMol based on Protein Databank entry 3BIY, published by Liu et al. (143)
Figure 7
Figure 7. Acetyl transfer catalysis by p300. (A) The p300 active site is drawn in green, and histone H4 substrate in blue, with important residues indicated. CoA is drawn in black, and binds in a specific tunnel. (B) Four steps in a proposed p300 mechanism. acetyl-CoA binds, then peptidyl-lysine binds. The hydrophobic indole of W1436 promotes an uncharged lysine and positions it for attack. The lysine attacks the carbonyl of acetyl-CoA, while Y1467 acts as a general acid to protonate the leaving group. Acetyl-lysine-containing product leaves quickly, then CoASH departs slowly.
Figure 8
Figure 8. Bisubstrate inhibitors of acetyltransferases.
Figure 9
Figure 9. Natural products implicated as modulators of acetyltransferases.
Figure 10
Figure 10. Synthetic inhibitors of acetyltransferases.
Figure 11
Figure 11. C646 modeled in the acetyltransferase active site. (A) C646 is shown in magenta, computationally docked in the crystal structure of the acetyltransferase active site, which was generated as a cocrystal with Lys-CoA. Several residues that coordinate CoA binding are predicted to similarly coordinate C646 binding, as shown in aqua stick representations of the side chains. (B) The structure of C646, shown in an orientation similar to that in the docked model above.
Figure 12
Figure 12. Structures of CBP bromodomain bound to ligands. The purified CBP bromodomain (residues 1081–1197, shown in a rainbow blue to red) is shown bound to (A) histone H4 residues 14–28 acetylated at K20 (PDB 2RNY); (B) p53 residues 367–386 acetylated at K382 (PDB 1JSP); (C) the compound ischemin (PDB 2L84); and (D) the compound dimethylisoxazole (PDB 3SVH). Peptide ligands are shown in gray (A,B) or stick models colored by atom (C,D and acetyl-lysines in A,B). All structures are based on solution NMR except for (D), which is from X-ray crystallography (1.8 Å).
Figure 13
Figure 13. Structures of p300/CBP TAZ2 domain bound to ligands. The TAZ2 domain is shown colored in a rainbow (blue to red, residues included listed below each) bound to various ligands: STAT1 (A, PDB 2KA6); E1A (B, PDB 2KJE); p53 (C, PDB 2K8F); MEF2-DNA complex (D, PDB 3P57), and C/EBPε (PDB 3T92). All structures are based on solution NMR except for two from X-ray crystallography: that in (D) (2.192 Å) and that in (E) (1.5 Å). All structures were produced using purified p300, except the (A) and (B), which used purified CBP. Zinc ions are black spheres, protein ligands are gray, and DNA is yellow. The crystal structure with MEF2 revealed binding in three possible conformations with TAZ2, and one example is shown here.
Figure 14
Figure 14. Models for targeting influenced by p300-ligand binding. In these models, p300 is shown in green, the histone octamer is shown in yellow, DNA is shown with a red strand, and p300 ligands are indicated with an “L”. In (A), a ligand targets p300 to a gene or other DNA element due to the DNA binding affinity of the ligand. In (B), a ligand targets p300 to a protein complex due to the protein binding affinity of the ligand. In (C), two ligands bridged by p300 allow for chromatin (bound by the purple ligand) to come into proximity with a chromatin-modifying enzyme (the orange ligand). In (D), two ligands compete for the same site within p300, and the one ligand could be seen as a competitive inhibitor for the p300 association with the other.
Figure 15
Figure 15. Mastermind-Notch-CSL-DNA core complex. The complex formed by DNA, CSL, Notch (ANK repeats and RAM region purified separately), and Mastermind N-terminal helix is shown with two different view angles. The X-ray crystal structure was generated at 3.85 Å, and this figure was produced in PyMol using Protein Databank entry 3V79. Proteins are shown as ribbons, with the surfaces at 70% transparency. DNA is shown as stick models colored by atom.
Figure 16
Figure 16. Model for Mastermind activation of p300. In this model, p300 (green) initially has inhibited acetyltransferase activity due to an autoinhibitory loop (orange, left). This is relieved upon recruitment by the Notch-Mastermind-CSL complex. Mastermind (purple) binds to the p300 C/H3 domain, and also to Notch intracellular domain (magenta) and CSL (red). p300 autoacetylation, Mastermind acetylation, and histone acetylation are then catalyzed by p300 (turquoise ▲).
Figure 17
Figure 17. Inhibitors of the p300/CBP bromodomain.
Figure 18
Figure 18. Inhibitors of the other p300/CBP domains.
Figure 19
Figure 19. Diseases of potential therapeutic application for a p300/CBP inhibitor. No p300/CBP inhibitor has yet made it into clinical trials, but the biology of p300/CBP action and documented effects of p300/CBP disruption lead us to hypothesize a beneficial therapeutic potential for a p300/CBP inhibitor in many diseases.
Beverley M. Dancy
Beverley M. Dancy, originally from Cape Town, was introduced to research in the laboratory of Greg J. Hermann at Lewis & Clark College. Her graduate studies with Philip A. Cole, centered on p300 acetyltransferase inhibition, earned her a Ph.D. in Pharmacology and Molecular Sciences from the Johns Hopkins University School of Medicine in 2013. As a postdoctoral fellow in the Morgan-Sedensky laboratory, she investigated enzymes and pathways involved in mitochondrial disease, and in 2014 she will begin a fellowship at the National Institutes of Health Laboratory of Cardiac Energetics, with Robert S. Balaban. Dr. Dancy is interested in developing molecular tools and imaging technologies for understanding the consequences of disease at the molecular, cellular, and organismal levels. In addition to research, she is passionate about bridging the gap between science research and a broader audience through communication and engagement.
Philip A. Cole
Philip A. Cole was born in Paterson, NJ and graduated from Yale University with a B.S. in Chemistry in 1984 and then spent a year as a Churchill Scholar at the University of Cambridge, England. Cole went on to obtain M.D. and Ph.D. degrees from Johns Hopkins University where he pursued research in bioorganic chemistry in 1991. Cole then entered postdoctoral fellowship training at Harvard Medical School prior to joining the Rockefeller University in 1996 as a junior lab head. In 1999, Cole moved back to Johns Hopkins as the Marshall-Maren professor and director of pharmacology. His research interests are in the area of protein post-translational modifications and chemical biology.
References
This article references 499 other publications.
- 1Waddington, C. H. Endeavour 1942, 18There is no corresponding record for this reference.
- 2Klar, A. J. Trends Genet. 1998, 14, 2992Propagating epigenetic states through meiosis: where Mendel's gene is more than a DNA moietyKlar, Amar J. S.Trends in Genetics (1998), 14 (8), 299-301CODEN: TRGEE2; ISSN:0168-9525. (Elsevier Science Ltd.)A review, with 21 refs., on epigenetic inheritance as illustrated by Drosophila and Schizosaccharomyces pombe.
- 3Redon, C.; Pilch, D.; Rogakou, E.; Sedelnikova, O.; Newrock, K.; Bonner, W. Curr. Opin. Genet. Dev. 2002, 12, 162There is no corresponding record for this reference.
- 4Dawson, M. A.; Kouzarides, T. Cell 2012, 150, 124Cancer Epigenetics: From Mechanism to TherapyDawson, Mark A.; Kouzarides, TonyCell (Cambridge, MA, United States) (2012), 150 (1), 12-27CODEN: CELLB5; ISSN:0092-8674. (Cell Press)A review. The epigenetic regulation of DNA-templated processes has been intensely studied over the last 15 years. DNA methylation, histone modification, nucleosome remodeling, and RNA-mediated targeting regulate many biol. processes that are fundamental to the genesis of cancer. Here, we present the basic principles behind these epigenetic pathways and highlight the evidence suggesting that their misregulation can culminate in cancer. This information, along with the promising clin. and preclin. results seen with epigenetic drugs against chromatin regulators, signifies that it is time to embrace the central role of epigenetics in cancer.
- 5Mukherjee, S.; Hao, Y. H.; Orth, K. Trends Biochem. Sci. 2007, 32, 210There is no corresponding record for this reference.
- 6Moellering, R. E.; Cravatt, B. F. Science 2013, 341, 549There is no corresponding record for this reference.
- 7Cohen, I.; Poręba, E.; Kamieniarz, K.; Schneider, R. Genes Cancer 2011, 2, 6317Histone modifiers in cancer: friends or foes?Cohen, Idan; Poreba, Elzbieta; Kamieniarz, Kinga; Schneider, RobertGenes & Cancer (2011), 2 (6), 631-647CODEN: GCEAAY; ISSN:1947-6019. (Sage Publications)A review. Covalent modifications of histories can regulate all DNA-dependent processes. In the last few years, it has become more and more evident that histone modifications are key players in the regulation of chromatin states and dynamics as well as in gene expression. Therefore, histone modifications and the enzymic machineries that set them are crucial regulators that can control cellular proliferation, differentiation, plasticity, and malignancy processes. This review discusses the biol. and biochem. of covalent histone posttranslational modifications (PTMs) and evaluates the dual role of their modifiers in cancer: as oncogenes that can initiate and amplify tumorigenesis or as tumor suppressors.
- 8Xing, S.; Poirier, Y. Trends Plant Sci. 2012, 17, 4238The protein acetylome and the regulation of metabolismXing, Shufan; Poirier, YvesTrends in Plant Science (2012), 17 (7), 423-430CODEN: TPSCF9; ISSN:1360-1385. (Elsevier Ltd.)A review. Acetyl-CoA (CoA) is a central metabolite involved in numerous anabolic and catabolic pathways, as well as in protein acetylation. Beyond histones, a large no. of metabolic enzymes are acetylated in both animal and bacteria, and the protein acetylome is now emerging in plants. Protein acetylation is influenced by the cellular level of both acetyl-CoA and NAD+, and regulates the activity of several enzymes. Acetyl-CoA is thus ideally placed to act as a key mol. linking the energy balance of the cell to the regulation of gene expression and metabolic pathways via the control of protein acetylation. Better knowledge over how to influence acetyl-CoA levels and the acetylation process promises to be an invaluable tool to control metabolic pathways.
- 9Daugherty, M.; Polanuyer, B.; Farrell, M.; Scholle, M.; Lykidis, A.; de Crécy-Lagard, V.; Osterman, A. J. Biol. Chem. 2002, 277, 21431There is no corresponding record for this reference.
- 10Robinson, P. J.; An, W.; Routh, A.; Martino, F.; Chapman, L.; Roeder, R. G.; Rhodes, D. J. Mol. Biol. 2008, 381, 8161030 nm Chromatin Fibre Decompaction Requires both H4-K16 Acetylation and Linker Histone EvictionRobinson, Philip J. J.; An, Woojin; Routh, Andrew; Martino, Fabrizio; Chapman, Lynda; Roeder, Robert G.; Rhodes, DanielaJournal of Molecular Biology (2008), 381 (4), 816-825CODEN: JMOBAK; ISSN:0022-2836. (Elsevier Ltd.)The mechanism by which chromatin is decondensed to permit access to DNA is largely unknown. Here, using a model nucleosome array reconstituted from recombinant histone octamers, we have defined the relative contribution of the individual histone octamer N-terminal tails as well as the effect of a targeted histone tail acetylation on the compaction state of the 30 nm chromatin fiber. This study goes beyond previous studies as it is based on a nucleosome array that is very long (61 nucleosomes) and contains a stoichiometric concn. of bound linker histone, which is essential for the formation of the 30 nm chromatin fiber. We find that compaction is regulated in two steps: Introduction of H4 acetylated to 30% on K16 inhibits compaction to a greater degree than deletion of the H4 N-terminal tail. Further decompaction is achieved by removal of the linker histone.
- 11Neumann, H.; Peak-Chew, S. Y.; Chin, J. W. Nat. Chem. Biol. 2008, 4, 23211Genetically encoding Nε-acetyllysine in recombinant proteinsNeumann, Heinz; Peak-Chew, Sew Y.; Chin, Jason W.Nature Chemical Biology (2008), 4 (4), 232-234CODEN: NCBABT; ISSN:1552-4450. (Nature Publishing Group)Nε-acetylation of lysine is a reversible post-translational modification with a regulatory role that rivals that of phosphorylation in eukaryotes. No general methods exist to synthesize proteins contg. Nε-acetyllysine at defined sites. The site-specific incorporation of Nε-acetyllysine in recombinant proteins produced in Escherichia coli was achieved via the evolution of an orthogonal Nε-acetyllysyl-tRNA synthetase/tRNACUA pair. This strategy should find wide applications in defining the cellular role of this modification.
- 12Guo, J.; Wang, J.; Lee, J. S.; Schultz, P. G. Angew. Chem., Int. Ed. 2008, 47, 6399There is no corresponding record for this reference.
- 13Muir, T. W.; Sondhi, D.; Cole, P. A. Proc. Natl. Acad. Sci. U.S.A. 1998, 95, 670513Expressed protein ligation: A general method for protein engineeringMuir, Tom W.; Sondhi, Dolan; Cole, Philip A.Proceedings of the National Academy of Sciences of the United States of America (1998), 95 (12), 6705-6710CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)A protein semisynthesis method-expressed protein ligation-is described that involves the chemoselective addn. of a peptide to a recombinant protein. This method was used to ligate a phosphotyrosine peptide to the C terminus of the protein tyrosine kinase C-terminal Src kinase (Csk). By intercepting a thioester generated in the recombinant protein with an N-terminal cysteine contg. synthetic peptide, near quant. chem. ligation of the peptide to the protein was achieved. The semisynthetic tail-phosphorylated Csk showed evidence of an intramol. phosphotyrosine-Src homol. 2 interaction and an unexpected increase in catalytic phosphoryl transfer efficiency toward a physiol. relevant substrate compared with the non-tail-phosphorylated control. This work illustrates that expressed protein ligation is a simple and powerful new method in protein engineering to introduce sequences of unnatural amino acids, posttranslational modifications, and biophys. probes into proteins of any size.
- 14Karukurichi, K. R.; Wang, L.; Uzasci, L.; Manlandro, C. M.; Wang, Q.; Cole, P. A. J. Am. Chem. Soc. 2010, 132, 1222There is no corresponding record for this reference.
- 15Huang, R.; Holbert, M. A.; Tarrant, M. K.; Curtet, S.; Colquhoun, D. R.; Dancy, B. M.; Dancy, B. C.; Hwang, Y.; Tang, Y.; Meeth, K.; Marmorstein, R.; Cole, R. N.; Khochbin, S.; Cole, P. A. J. Am. Chem. Soc. 2010, 132, 9986There is no corresponding record for this reference.
- 16Johns, E. W.; Phillips, D. M.; Simson, P.; Butler, J. A. Biochem. J. 1961, 80, 189There is no corresponding record for this reference.
- 17Phillips, D. M. Biochem. J. 1963, 87, 258There is no corresponding record for this reference.
- 18Cohen, I.; Poręba, E.; Kamieniarz, K.; Schneider, R. Genes Cancer 2011, 2, 63118Histone modifiers in cancer: friends or foes?Cohen, Idan; Poreba, Elzbieta; Kamieniarz, Kinga; Schneider, RobertGenes & Cancer (2011), 2 (6), 631-647CODEN: GCEAAY; ISSN:1947-6019. (Sage Publications)A review. Covalent modifications of histories can regulate all DNA-dependent processes. In the last few years, it has become more and more evident that histone modifications are key players in the regulation of chromatin states and dynamics as well as in gene expression. Therefore, histone modifications and the enzymic machineries that set them are crucial regulators that can control cellular proliferation, differentiation, plasticity, and malignancy processes. This review discusses the biol. and biochem. of covalent histone posttranslational modifications (PTMs) and evaluates the dual role of their modifiers in cancer: as oncogenes that can initiate and amplify tumorigenesis or as tumor suppressors.
- 19Close, P.; Creppe, C.; Gillard, M.; Ladang, A.; Chapelle, J. P.; Nguyen, L.; Chariot, A. Cell. Mol. Life Sci. 2010, 67, 125519The emerging role of lysine acetylation of non-nuclear proteinsClose, Pierre; Creppe, Catherine; Gillard, Magali; Ladang, Aurelie; Chapelle, Jean-Paul; Nguyen, Laurent; Chariot, AlainCellular and Molecular Life Sciences (2010), 67 (8), 1255-1264CODEN: CMLSFI; ISSN:1420-682X. (Birkhaeuser Verlag)A review. The acetylation of protein Lys residues is a post-translational modification that critically regulates gene transcription by targeting histones as well as a variety of transcription factors in the nucleus. More recent reports have also demonstrated that numerous proteins located outside the nucleus are also acetylated and that this modification has profound consequences on their functions. This review describes the latest findings on the protein substrates acetylated outside the nucleus and on the acetylases and deacetylates that catalyze these modifications. Protein acetylation is emerging as a major mechanism by which key proteins are regulated in many physiol. processes such as migration, metab., and aging as well as in pathol. circumstances such as cancer and neurodegenerative disorders.
- 20Glozak, M. A.; Sengupta, N.; Zhang, X.; Seto, E. Gene 2005, 363, 15There is no corresponding record for this reference.
- 21Norris, K. L.; Lee, J. Y.; Yao, T. P. Sci. Signaling 2009, 2, pe7621Acetylation goes global: the emergence of acetylation biologyNorris Kristi L; Lee Joo-Yong; Yao Tso-PangScience signaling (2009), 2 (97), pe76 ISSN:.For the first 30 years since its discovery, reversible protein acetylation has been studied and understood almost exclusively in the context of histone modification and gene transcription. With the discovery of non-histone acetylated proteins and acetylation-modifying enzymes in cellular compartments outside the nucleus, the regulatory potential of reversible acetylation has slowly been recognized in the last decade. However, the scope of protein acetylation involvement in complex biological processes remains uncertain. The recent development of new technology has enabled, for the first time, the identification and quantification of the acetylome, acetylation events at the whole-proteome level. These efforts have uncovered a stunning complexity of the acetylome that potentially rivals that of the phosphoproteome. The remarkably ubiquitous and conserved nature of protein acetylation revealed by these new studies suggests the regulatory power of this dynamic modification. The establishment of comprehensive acetylomes will change the landscape of protein acetylation, where an exciting research frontier awaits.
- 22Choudhary, C.; Kumar, C.; Gnad, F.; Nielsen, M. L.; Rehman, M.; Walther, T. C.; Olsen, J. V.; Mann, M. Science 2009, 325, 83422Lysine Acetylation Targets Protein Complexes and Co-Regulates Major Cellular FunctionsChoudhary, Chunaram; Kumar, Chanchal; Gnad, Florian; Nielsen, Michael L.; Rehman, Michael; Walther, Tobias C.; Olsen, Jesper V.; Mann, MatthiasScience (Washington, DC, United States) (2009), 325 (5942), 834-840CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)Lysine acetylation is a reversible posttranslational modification of proteins and plays a key role in regulating gene expression. Technol. limitations have so far prevented a global anal. of lysine acetylation's cellular roles. We used high-resoln. mass spectrometry to identify 3600 lysine acetylation sites on 1750 proteins and quantified acetylation changes in response to the deacetylase inhibitors suberoylanilide hydroxamic acid and MS-275. Lysine acetylation preferentially targets large macromol. complexes involved in diverse cellular processes, such as chromatin remodeling, cell cycle, splicing, nuclear transport, and actin nucleation. Acetylation impaired phosphorylation-dependent interactions of 14-3-3 and regulated the yeast cyclin-dependent kinase Cdc28. Our data demonstrate that the regulatory scope of lysine acetylation is broad and comparable with that of other major posttranslational modifications.
- 23Chen, Y.; Zhao, W.; Yang, J. S.; Cheng, Z.; Luo, H.; Lu, Z.; Tan, M.; Gu, W.; Zhao, Y. Mol. Cell. Proteomics 2012, 11, 104823Quantitative acetylome analysis reveals the roles of SIRT1 in regulating diverse substrates and cellular pathwaysChen, Yue; Zhao, Wenhui; Yang, Jeong Soo; Cheng, Zhongyi; Luo, Hao; Lu, Zhike; Tan, Minjia; Gu, Wei; Zhao, YingmingMolecular and Cellular Proteomics (2012), 11 (10), 1048-1062CODEN: MCPOBS; ISSN:1535-9484. (American Society for Biochemistry and Molecular Biology)Despite the progress in identifying many Lys acetylation (Kac) proteins, Kac substrates for Kac-regulatory enzymes remain largely unknown, presenting a major knowledge gap in Kac biol. Here, we identified and quantified 4623 Kac sites in 1800 Kac proteins in SIRT1+/+ and SIRT1-/- MEF cells, representing the first study to reveal an enzyme-regulated Kac subproteome and the largest Lys acetylome reported to date from a single study. Four hundred eighty-five Kac sites were enhanced by more than 100% after SIRT1 knockout. Our results indicate that SIRT1 regulates the Kac states of diverse cellular pathways. Interestingly, we found that a no. of acetyltransferases and major acetyltransferase complexes are targeted by SIRT1. Moreover, we showed that the activities of the acetyltransferases are regulated by SIRT1-mediated deacetylation. Taken together, our results reveal the Lys acetylome in response to SIRT1, provide new insights into mechanisms of SIRT1 function, and offer biomarker candidates for the clin. evaluation of SIRT1-activator compds.
- 24Zhao, S.; Xu, W.; Jiang, W.; Yu, W.; Lin, Y.; Zhang, T.; Yao, J.; Zhou, L.; Zeng, Y.; Li, H.; Li, Y.; Shi, J.; An, W.; Hancock, S. M.; He, F.; Qin, L.; Chin, J.; Yang, P.; Chen, X.; Lei, Q.; Xiong, Y.; Guan, K. L. Science 2010, 327, 1000There is no corresponding record for this reference.
- 25Albaugh, B. N.; Arnold, K. M.; Denu, J. M. ChemBioChem 2011, 12, 290There is no corresponding record for this reference.
- 26Sutendra, G.; Kinnaird, A.; Dromparis, P.; Paulin, R.; Stenson, T. H.; Haromy, A.; Hashimoto, K.; Zhang, N.; Flaim, E.; Michelakis, E. D. Cell 2014, 158, 84There is no corresponding record for this reference.
- 27Begley, T. P.; Kinsland, C.; Strauss, E. Vitam. Horm. 2001, 61, 15727The biosynthesis of coenzyme A in bacteriaBegley, Tadhg P.; Kinsland, Cynthia; Strauss, ErickVitamins and Hormones (San Diego, CA, United States) (2001), 61 (), 157-171CODEN: VIHOAQ; ISSN:0083-6729. (Academic Press)A review with 50 refs. CoA (I) and enzyme-bound phosphopantetheine (II) function as acyl carriers and as carbonyl activating groups for Claisen reactions as well as for amide-, ester-, and thioester-forming reactions in the cell. These cofactors play a key role in the biosynthesis and breakdown of fatty acids and in the biosynthesis of polyketides and nonribosomal peptides. CoA is biosynthesized in bacteria in nine steps. The biosynthesis begins with the decarboxylation of aspartate to give β-alanine. Pantoic acid is formed by the hydroxymethylation of α-ketoisovalerate followed by redn. These intermediates are then condensed to give pantothenic acid. Phosphorylation of pantothenic acid followed by condensation with cysteine and decarboxylation gives 4'-phosphopantetheine. Adenylation and phosphorylation of 4'-phosphopantetheine completes the biosynthesis of CoA. This review will focus on the mechanistic enzymol. of CoA biosynthesis in bacteria. (c) 2001 Academic Press.
- 28Xing, S.; Poirier, Y. Trends Plant Sci. 2012, 17, 42328The protein acetylome and the regulation of metabolismXing, Shufan; Poirier, YvesTrends in Plant Science (2012), 17 (7), 423-430CODEN: TPSCF9; ISSN:1360-1385. (Elsevier Ltd.)A review. Acetyl-CoA (CoA) is a central metabolite involved in numerous anabolic and catabolic pathways, as well as in protein acetylation. Beyond histones, a large no. of metabolic enzymes are acetylated in both animal and bacteria, and the protein acetylome is now emerging in plants. Protein acetylation is influenced by the cellular level of both acetyl-CoA and NAD+, and regulates the activity of several enzymes. Acetyl-CoA is thus ideally placed to act as a key mol. linking the energy balance of the cell to the regulation of gene expression and metabolic pathways via the control of protein acetylation. Better knowledge over how to influence acetyl-CoA levels and the acetylation process promises to be an invaluable tool to control metabolic pathways.
- 29Wellen, K. E.; Thompson, C. B. Nat. Rev. Mol. Cell Biol. 2012, 13, 27029A two-way street: reciprocal regulation of metabolism and signallingWellen, Kathryn E.; Thompson, Craig B.Nature Reviews Molecular Cell Biology (2012), 13 (4), 270-276CODEN: NRMCBP; ISSN:1471-0072. (Nature Publishing Group)A review. It is becoming increasingly clear that cellular signaling and metab. are not just sep. entities but rather are tightly linked. Although nutrient metab. is known to be regulated by signal transduction, an emerging paradigm is that signaling and transcriptional networks can be modulated by nutrient-sensitive protein modifications, such as acetylation and glycosylation, which depend on the availability of acetyl-CoA and sugar donors such as UDP-N-acetylglucosamine (UDP-GlcNAc), resp. The integration of metabolic and signaling cues allows cells to modulate activities such as metab., cell survival and proliferation according to their intracellular metabolic resources.
- 30Albaugh, B. N.; Arnold, K. M.; Denu, J. M. ChemBioChem 2011, 12, 290There is no corresponding record for this reference.
- 31Allfrey, V. G.; Faulkner, R.; Mirsky, A. E. Proc. Natl. Acad. Sci. U.S.A. 1964, 51, 78631Acetylation and methylation of histones and their possible role in the regulation of ribonucleic acid (RNA) synthesisAllfrey, V. G.; Faulkner, R.; Mirsky, A. E.Proceedings of the National Academy of Sciences of the United States of America (1964), 51 (5), 786-94CODEN: PNASA6; ISSN:0027-8424.Isolated calf thymus nuclei was incubated with 14C-labeled Na acetate. Evidence that acetate was incorporated as acetyl groups attached to histones included: (1) following chromatography the label was eluted together with basic proteins; (2) acetyl-labeled histones were sepd. by electrophoresis; (3) acetate-14C was not removed from histones by org. solvents; (4) over 75% of the label remained after histones were treated with trichloroacetic acid. Most extensive acetate uptake occurred with arginine-rich histones. Puromycin did not inhibit the acetylation, indicating histones were modified by acetylation after completion of the polypeptide chain. Acetylation lowered the histone effectiveness as inhibitor of the RNA polymerase reaction, and a dynamic and reversible mechanism was suggested for activation as well as repression of RNA synthesis.
- 32Hebbes, T. R.; Thorne, A. W.; Crane-Robinson, C. EMBO J. 1988, 7, 1395There is no corresponding record for this reference.
- 33Hebbes, T. R.; Clayton, A. L.; Thorne, A. W.; Crane-Robinson, C. EMBO J. 1994, 13, 1823There is no corresponding record for this reference.
- 34Braunstein, M.; Rose, A. B.; Holmes, S. G.; Allis, C. D.; Broach, J. R. Genes Dev. 1993, 7, 59234Transcriptional silencing in yeast is associated with reduced nucleosome acetylationBraunstein, Miriam; Rose, Alan B.; Holmes, Scott G.; Allis, C. David; Broach, James R.Genes & Development (1993), 7 (4), 592-604CODEN: GEDEEP; ISSN:0890-9369.Two classes of sequences in the yeast Saccharomyces cerevisiae are subject to transcription silencing: the silent mating-type cassettes and telomers. The silencing of these regions is shown to be strictly assocd. with acetylation of the ε-amino groups of lysines in the N-terminal domains of 3 of the 4 core histones. Both the silent mating-type cassettes and the Y domains of telomeres are packaged in nucleosomes in vivo that are hypoacetylated relative to those packaging active genes. This difference in acetylation is eliminated by genetic inactivation of silencing:. The silent cassettes from sir2, sir3, or sir4 cells show the same level of acetylation as other active genes. The correspondence of silencing and hypoacetylation of the mating-type cassettes is obsd. even for an allele lacking a promoter, indicating that silencing per se, rather than the absence of transcription, is correlated with hypoacetylation. Finally, overexpression of Sir2p, a protein required for transcriptional silencing in yeast, yields substantial histone deacetylation in vivo. These studies fortify the hypothesis that silencing in yeast results from heterochromatin formation and argue that the silencing proteins participate in this formation.
- 35Turner, B. M.; Birley, A. J.; Lavender, J. Cell 1992, 69, 375There is no corresponding record for this reference.
- 36Allfrey, V. G.; Faulkner, R.; Mirsky, A. E. Proc. Natl. Acad. Sci. U.S.A. 1964, 51, 78636Acetylation and methylation of histones and their possible role in the regulation of ribonucleic acid (RNA) synthesisAllfrey, V. G.; Faulkner, R.; Mirsky, A. E.Proceedings of the National Academy of Sciences of the United States of America (1964), 51 (5), 786-94CODEN: PNASA6; ISSN:0027-8424.Isolated calf thymus nuclei was incubated with 14C-labeled Na acetate. Evidence that acetate was incorporated as acetyl groups attached to histones included: (1) following chromatography the label was eluted together with basic proteins; (2) acetyl-labeled histones were sepd. by electrophoresis; (3) acetate-14C was not removed from histones by org. solvents; (4) over 75% of the label remained after histones were treated with trichloroacetic acid. Most extensive acetate uptake occurred with arginine-rich histones. Puromycin did not inhibit the acetylation, indicating histones were modified by acetylation after completion of the polypeptide chain. Acetylation lowered the histone effectiveness as inhibitor of the RNA polymerase reaction, and a dynamic and reversible mechanism was suggested for activation as well as repression of RNA synthesis.
- 37Mutskov, V.; Gerber, D.; Angelov, D.; Ausio, J.; Workman, J.; Dimitrov, S. Mol. Cell. Biol. 1998, 18, 629337Persistent interactions of core histone tails with nucleosomal DNA following acetylation and transcription factor bindingMutskov, Vesco; Gerber, Delphine; Angelov, Dimitri; Ausio, Juan; Workman, Jerry; Dimitrov, StefanMolecular and Cellular Biology (1998), 18 (11), 6293-6304CODEN: MCEBD4; ISSN:0270-7306. (American Society for Microbiology)In this study, we examd. the effect of acetylation of the NH2 tails of core histones on their binding to nucleosomal DNA in the absence or presence of bound transcription factors. To do this, we used a novel UV laser-induced protein-DNA crosslinking technique, combined with immunochem. and mol. biol. approaches. Nucleosomes contg. one or five GAL4 binding sites were reconstituted with hypo-acetylated or hyper-acetylated core histones. Within these reconstituted particles, UV laser-induced histone-DNA crosslinking was found to occur only via the non-structured histone tails and thus presented a unique tool for studying histone tail interactions with nucleosomal DNA. Importantly, these studies demonstrated that the NH2 tails were not released from nucleosomal DNA upon histone acetylation, although some weakening of their interactions was obsd. at elevated ionic strengths. Moreover, the binding of up to five GAL4-AH dimers to nucleosomes occupying the central 90 bp occurred without displacement of the histone NH2 tails from DNA. GAL4-AH binding perturbed the interaction of each histone tail with nucleosomal DNA to different degrees. However, in all cases, greater than 50% of the interactions between the histone tails and DNA was retained upon GAL4-AH binding, even if the tails were highly acetylated. These data illustrate an interaction of acetylated or non-acetylated histone tails with DNA that persists in the presence of simultaneously bound transcription factors.
- 38Zeng, L.; Zhou, M. M. FEBS Lett. 2002, 513, 124There is no corresponding record for this reference.
- 39Taverna, S. D.; Li, H.; Ruthenburg, A. J.; Allis, C. D.; Patel, D. J. Nat. Struct. Mol. Biol. 2007, 14, 102539How chromatin-binding modules interpret histone modifications: lessons from professional pocket pickersTaverna, Sean D.; Li, Haitao; Ruthenburg, Alexander J.; Allis, C. David; Patel, Dinshaw J.Nature Structural & Molecular Biology (2007), 14 (11), 1025-1040CODEN: NSMBCU; ISSN:1545-9993. (Nature Publishing Group)A review. Histones comprise the major protein component of chromatin, the scaffold in which the eukaryotic genome is packaged, and are subject to many types of post-translational modifications (PTMs), esp. on their flexible tails. These modifications may constitute a 'histone code' and could be used to manage epigenetic information that helps extend the genetic message beyond DNA sequences. This proposed code, read in part by histone PTM-binding 'effector' modules and their assocd. complexes, has been predicted to define unique functional states of chromatin and/or regulate various chromatin-templated processes. A wealth of structural and functional data show how chromatin effector modules target their cognate covalent histone modifications. Here, the authors summarize key features in the mol. recognition of histone PTMs by a diverse family of 'reader pockets' (including bromodomains, Royal superfamily modules, PHD fingers, WD40 protein repeats, 14-3-3 proteins, and BRCT domains), highlighting specific readout mechanisms for individual marks, common themes, and insights into the downstream functional consequences of the interactions. Changes in these interactions may have far-reaching implications for human biol. and disease, notably cancer.
- 40Botuyan, M. V.; Lee, J.; Ward, I. M.; Kim, J. E.; Thompson, J. R.; Chen, J.; Mer, G. Cell 2006, 127, 1361There is no corresponding record for this reference.
- 41Lee, J.; Thompson, J. R.; Botuyan, M. V.; Mer, G. Nat. Struct. Mol. Biol. 2008, 15, 10941Distinct binding modes specify the recognition of methylated histones H3K4 and H4K20 by JMJD2A-tudorLee, Joseph; Thompson, James R.; Botuyan, Maria Victoria; Mer, GeorgesNature Structural & Molecular Biology (2008), 15 (1), 109-111CODEN: NSMBCU; ISSN:1545-9993. (Nature Publishing Group)The lysine demethylase JMJD2A has the unique property of binding trimethylated peptides from two different histone sequences (H3K4me3 and H4K20me3) through its tudor domains. Here we show using X-ray crystallog. and calorimetry that H3K4me3 and H4K20me3, which are recognized with similar affinities by JMJD2A, adopt radically different binding modes, to the extent that we were able to design single point mutations in JMJD2A that inhibited the recognition of H3K4me3 but not H4K20me3 and vice versa.
- 42Guo, Y.; Nady, N.; Qi, C.; Allali-Hassani, A.; Zhu, H.; Pan, P.; Adams-Cioaba, M. A.; Amaya, M. F.; Dong, A.; Vedadi, M.; Schapira, M.; Read, R. J.; Arrowsmith, C. H.; Min, J. Nucleic Acids Res. 2009, 37, 2204There is no corresponding record for this reference.
- 43Li, H.; Fischle, W.; Wang, W.; Duncan, E. M.; Liang, L.; Murakami-Ishibe, S.; Allis, C. D.; Patel, D. J. Mol. Cell 2007, 28, 677There is no corresponding record for this reference.
- 44Li, H.; Ilin, S.; Wang, W.; Duncan, E. M.; Wysocka, J.; Allis, C. D.; Patel, D. J. Nature 2006, 442, 91There is no corresponding record for this reference.
- 45Jenuwein, T.; Allis, C. D. Science 2001, 293, 107445Translating the histone codeJenuwein, Thomas; Allis, C. DavidScience (Washington, DC, United States) (2001), 293 (5532), 1074-1080CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)The review. Chromatin, the physiol. template of all eukaryotic genetic information, is subject to a diverse array of posttranslational modifications that largely impinge on histone amino termini, thereby regulating access to the underlying DNA. Distinct histone amino-terminal modifications can generate synergistic or antagonistic interaction affinities for chromatin-assocd. proteins, which in turn dictate dynamic transitions between transcriptionally active or transcriptionally silent chromatin states. The combinatorial nature of histone amino-terminal modifications thus reveals a "histone code" that considerably extends the information potential of the genetic code. We propose that this epigenetic marking system represents a fundamental regulatory mechanism that has an impact on most, if not all; chromatin-templated processes, with far-reaching consequences for cell fate decisions and both normal and pathol. development.
- 46Bernstein, E.; Hake, S. B. Biochem. Cell Biol. 2006, 84, 50546The nucleosome: a little variation goes a long wayBernstein, Emily; Hake, Sandra B.Biochemistry and Cell Biology (2006), 84 (4), 505-517CODEN: BCBIEQ; ISSN:0829-8211. (National Research Council of Canada)A review. Changes in the overall structure of chromatin are essential for the proper regulation of cellular processes, including gene activation and silencing, DNA repair, chromosome segregation during mitosis and meiosis, X chromosome inactivation in female mammals, and chromatin compaction during apoptosis. Such alterations of the chromatin template occur through at least 3 interrelated mechanisms: post-translational modifications of histones, ATP-dependent chromatin remodeling, and the incorporation (or replacement) of specialized histone variants into chromatin. Of these mechanisms, the exchange of variants into and out of chromatin is the least well understood. However, the exchange of conventional histones for variant histones has distinct and profound consequences within the cell. This review focuses on the growing no. of mammalian histone variants, their particular biol. functions and unique features, and how they may affect the structure of the nucleosome. The authors propose that a given nucleosome might not consist of heterotypic variants, but rather, that only specific histone variants come together to form a homotypic nucleosome, a hypothesis that the authors refer to as the nucleosome code. Such nucleosomes might in turn participate in marking specific chromatin domains that may contribute to epigenetic inheritance.
- 47Schreiber, S. L.; Bernstein, B. E. Cell 2002, 111, 771There is no corresponding record for this reference.
- 48Ernst, J.; Kellis, M. Nat. Biotechnol. 2010, 28, 817There is no corresponding record for this reference.
- 49Jenuwein, T.; Allis, C. D. Science 2001, 293, 107449Translating the histone codeJenuwein, Thomas; Allis, C. DavidScience (Washington, DC, United States) (2001), 293 (5532), 1074-1080CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)The review. Chromatin, the physiol. template of all eukaryotic genetic information, is subject to a diverse array of posttranslational modifications that largely impinge on histone amino termini, thereby regulating access to the underlying DNA. Distinct histone amino-terminal modifications can generate synergistic or antagonistic interaction affinities for chromatin-assocd. proteins, which in turn dictate dynamic transitions between transcriptionally active or transcriptionally silent chromatin states. The combinatorial nature of histone amino-terminal modifications thus reveals a "histone code" that considerably extends the information potential of the genetic code. We propose that this epigenetic marking system represents a fundamental regulatory mechanism that has an impact on most, if not all; chromatin-templated processes, with far-reaching consequences for cell fate decisions and both normal and pathol. development.
- 50Feng, Y.; Wang, J.; Asher, S.; Hoang, L.; Guardiani, C.; Ivanov, I.; Zheng, Y. G. J. Biol. Chem. 2011, 286, 20323There is no corresponding record for this reference.
- 51Nathan, D.; Ingvarsdottir, K.; Sterner, D. E.; Bylebyl, G. R.; Dokmanovic, M.; Dorsey, J. A.; Whelan, K. A.; Krsmanovic, M.; Lane, W. S.; Meluh, P. B.; Johnson, E. S.; Berger, S. L. Genes Dev. 2006, 20, 966There is no corresponding record for this reference.
- 52Messner, S.; Altmeyer, M.; Zhao, H.; Pozivil, A.; Roschitzki, B.; Gehrig, P.; Rutishauser, D.; Huang, D.; Caflisch, A.; Hottiger, M. O. Nucleic Acids Res. 2010, 38, 6350There is no corresponding record for this reference.
- 53Dhalluin, C.; Carlson, J. E.; Zeng, L.; He, C.; Aggarwal, A. K.; Zhou, M. M. Nature 1999, 399, 491There is no corresponding record for this reference.
- 54Thorne, A. W.; Kmiciek, D.; Mitchelson, K.; Sautiere, P.; Crane-Robinson, C. Eur. J. Biochem. 1990, 193, 70154Patterns of histone acetylationThorne, Alan W.; Kmiciek, Daniel; Mitchelson, Keith; Sautiere, Pierre; Crane-Robinson, ColynEuropean Journal of Biochemistry (1990), 193 (3), 701-13CODEN: EJBCAI; ISSN:0014-2956.The N-terminal domains of all four core histones are subject to reversible acetylation at certain lysine residues. This modification has been functionally linked to transcription, histone deposition at replication, and to histone removal during spermatogenesis. To increase understanding of the significance of this modification the specificity of site utilization in the monoacetyl, diacetyl, and triacetyl forms of histones H3, H4, and H2B (histone H2A has only a single modification site) was studied using pig thymus and HeLa cells as the source. The HeLa histones were extd. from cells grown both with and without butyrate treatment. For histone H3 there is a fairly strict order of site occupancy: Lys14, followed by Lys23, followed by Lys18 in both pig and HeLa histones. Since the order and specificity is the same when butyrate is added to the HeLa cell cultures, it is concluded that addn. of the fatty acid does not scramble the specificity of site utilization, but merely allows more of the natural forms of modified histone to accumulate. For histone H4, the monoacetyl form is exclusively modified at Lys16, but further addn. of acetyl groups is less specific and progresses through sites 12, 8 and 5 in an N-terminal direction. Similar results were obtained for H4 from both pig thymus and butyrate-treated HeLa cells. Histone H2B could be studied in detail only from butyrate-treated HeLa cells and a much lower level of site specificity was found: sites 12 and 15 were preferred to the more N- and C-terminal sites at Lys5 and Lys20. The data reinforces the view that lysine acetylation in core histones is a very specific phenomenon that plays several functionally distinct roles. The high degree of site specificity makes it unlikely that the structural effects of acetylation are mediated merely by a generalized redn. of charge in the histone N-terminal domains.
- 55Zhang, K.; Williams, K. E.; Huang, L.; Yau, P.; Siino, J. S.; Bradbury, E. M.; Jones, P. R.; Minch, M. J.; Burlingame, A. L. Mol. Cell. Proteomics 2002, 1, 50055Histone acetylation and deacetylation: identification of acetylation and methylation sites of HeLa histone H4 by mass spectrometryZhang, Kangling; Williams, Katherine E.; Huang, Lan; Yau, Peter; Siino, Joseph S.; Bradbury, E. Morton; Jones, Patrick R.; Minch, Michael J.; Burlingame, Alma L.Molecular and Cellular Proteomics (2002), 1 (7), 500-508CODEN: MCPOBS; ISSN:1535-9476. (American Society for Biochemistry and Molecular Biology, Inc.)The acetylation isoforms of histone H4 from butyrate-treated HeLa cells were sepd. by C4 reverse-phase high pressure liq. chromatog. and by PAGE. Histone H4 bands were excised and digested in-gel with the endoprotease trypsin. Matrix-assisted laser desorption ionization time-of-flight mass spectrometry was used to characterize the level of acetylation, and nanoelectrospray tandem mass spectrometric anal. of the acetylated peptides was used to det. the exact sites of acetylation. Although there are 15 acetylation sites possible, only four acetylated peptide sequences were actually obsd. The tetra-acetylated form is modified at lysines 5, 8, 12, and 16, the tri-acetylated form is modified at lysines 8, 12, and 16, and the di-acetylated form is modified at lysines 12 and 16. The only significant amt. of the mono-acetylated form was found at position 16. These results are consistent with the hypothesis of a "zip" model whereby acetylation of histone H4 proceeds in the direction of from Lys-16 to Lys-5, and deacetylation proceeds in the reverse direction. Histone acetylation and deacetylation are coordinated processes leading to a non-random distribution of isoforms. Our results also revealed that lysine 20 is dimethylated in all modified isoforms, as well as the nonacetylated isoform of H4.
- 56Clayton, A. L.; Hazzalin, C. A.; Mahadevan, L. C. Mol. Cell 2006, 23, 289There is no corresponding record for this reference.
- 57Waterborg, J. H. Biochem. Cell Biol. 2002, 80, 36357Dynamics of histone acetylation in vivo. A function for acetylation turnover?Waterborg, Jakob H.Biochemistry and Cell Biology (2002), 80 (3), 363-378CODEN: BCBIEQ; ISSN:0829-8211. (National Research Council of Canada)A review. Histone acetylation discovered more than 40 yr ago is a reversible modification of lysines within the amino-terminal domain of core histones. Amino-terminal histone domains contribute to the compaction of genes into repressed chromatin fibers. It is thought that their acetylation causes localized relaxation of chromatin as a necessary but not sufficient condition for processes that repackage DNA such as transcription, replication, repair, recombination, and sperm formation. While increased histone acetylation enhances gene transcription and loss of acetylation represses and silences genes, the function of the rapid continuous or repetitive acetylation and deacetylation reactions with half-lives of just a few minutes remains unknown. Thirty years of in vivo measurements of acetylation turnover and rates of change in histone modification levels have been reviewed to identify common chromatin characteristics measured by distinct protocols. It has now become possible to look across a wider spectrum of organisms than ever before and identify common features. The rapid turnover rates in transcriptionally active and competent chromatin are one such feature. While ubiquitously obsd., we still do not know whether turnover itself is linked to chromatin transcription beyond its contribution to rapid changes towards hyper- or hypoacetylation of nucleosomes. However, recent expts. suggest that turnover may be linked directly to steps in gene transcription, interacting with nucleosome remodeling complexes.
- 58Katan-Khaykovich, Y.; Struhl, K. Genes Dev. 2002, 16, 743There is no corresponding record for this reference.
- 59Annunziato, A. T.; Eason, M. B.; Perry, C. A. Biochemistry 1995, 34, 2916There is no corresponding record for this reference.
- 60Hendzel, M. J.; Davie, J. R. Biochem. J. 1991, 273, 753There is no corresponding record for this reference.
- 61Hazzalin, C. A.; Mahadevan, L. C. PLoS Biol. 2005, 3, e393There is no corresponding record for this reference.
- 62Crump, N. T.; Hazzalin, C. A.; Bowers, E. M.; Alani, R. M.; Cole, P. A.; Mahadevan, L. C. Proc. Natl. Acad. Sci. U.S.A. 2011, 108, 7814There is no corresponding record for this reference.
- 63Taunton, J.; Hassig, C. A.; Schreiber, S. L. Science 1996, 272, 40863A mammalian histone deacetylase related to the yeast transcriptional regulator Rpd3pTaunton, Jack; Hassig, Christian A.; Schreiber, Stuart L.Science (Washington, D. C.) (1996), 272 (5260), 408-11CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)Trapoxin is a microbially derived cyclotetrapeptide that inhibits histone deacetylation in vivo and causes mammalian cells to arrest in the cell cycle. A trapoxin affinity matrix was used to isolate two nuclear proteins that copurified with histone deacetylase activity. Both proteins were identified by peptide microsequencing, and a complementary DNA encoding the histone deacetylase catalytic subunit (HD1) was cloned from a human Jurkat T cell library. As the predicted protein is very similar to the yeast transcriptional regulator Rpd3p, these results support a role for histone deacetylase as a key regulator of eukaryotic transcription.
- 64Schiltz, R. L.; Mizzen, C. A.; Vassilev, A.; Cook, R. G.; Allis, C. D.; Nakatani, Y. J. Biol. Chem. 1999, 274, 1189There is no corresponding record for this reference.
- 65Kimura, A.; Horikoshi, M. Genes Cells 1998, 3, 78965Tip60 acetylates six lysines of a specific class in core histones in vitroKimura, Akatsuki; Horikoshi, MasamiGenes to Cells (1998), 3 (12), 789-800CODEN: GECEFL; ISSN:1356-9597. (Blackwell Science Ltd.)Tip60, an HIV-1-Tat interactive protein, is a nuclear histone acetyltransferase (HAT) with unique histone substrate specificity. Since the acetylation of core histones at particular lysines mediates distinct effects on chromatin assembly and gene regulation, the identification of Lys site specificity of the HAT activity of Tip60 is an initial step in the anal. of its mol. function. Tip60 significantly acetylates N-terminal tail peptides of histones H2A, H3 and H4, but not H2B, consistent with substrate preference on intact histones. Preferred acetylation sites for Tip60 are Lys-5 of histone H2A, Lys-14 of histone H3, and Lys-5, Lys-8, Lys-12, Lys-16 of histone H4, detd. by a method which combined matrix-assisted laser desorption/ionization mass spectrometry (MALDI-MS) measurements and Lys-C endopeptidase digestion, or a method detecting the incorporation of radiolabeled acetate into synthetic peptides. The Lys site specificity of Tip60 in histone N-terminal tail peptides in vitro was characterized by an assay measuring the mol. wt. of endopeptidase-digested peptides, or a previously described assay. These results agree well with the authors' proposed classification of Lys residues in core histones. The classification may be useful for an anal. of the relations between HATs and the substrates of other uncharacterized HATs.
- 66Thompson, P. R.; Kurooka, H.; Nakatani, Y.; Cole, P. A. J. Biol. Chem. 2001, 276, 33721There is no corresponding record for this reference.
- 67Ogryzko, V. V.; Schiltz, R. L.; Russanova, V.; Howard, B. H.; Nakatani, Y. Cell 1996, 87, 953There is no corresponding record for this reference.
- 68Schiltz, R. L.; Mizzen, C. A.; Vassilev, A.; Cook, R. G.; Allis, C. D.; Nakatani, Y. J. Biol. Chem. 1999, 274, 1189There is no corresponding record for this reference.
- 69McManus, K. J.; Hendzel, M. J. Mol. Cell. Biol. 2003, 23, 761169Quantitative analysis of CBP- and P300-induced histone acetylations in vivo using native chromatinMcManus, Kirk J.; Hendzel, Michael J.Molecular and Cellular Biology (2003), 23 (21), 7611-7627CODEN: MCEBD4; ISSN:0270-7306. (American Society for Microbiology)In vivo, histone tails are involved in numerous interactions, including those with DNA, adjacent histones, and other, nonhistone proteins. The N-termini are also the substrates for a no. of enzymes, including histone acetyltransferases (HATs), histone deacetylases, and histone methyltransferases. Traditional biochem. approaches defining the substrate specificity profiles of HATs have been performed using purified histone tails, recombinant histones, or purified mononucleosomes as substrates. It is clear that the in vivo presentation of the substrate cannot be accurately represented by using these in vitro approaches. Because of the difficulty in translating in vitro results into in vivo situations, the authors developed a novel single-cell HAT assay that provides quant. measurements of endogenous HAT activity. The HAT assay is performed under in vivo conditions by using the native chromatin structure as the physiol. substrate. The assay combines the spatial resolving power of laser scanning confocal microscopy with simple statistical analyses to characterize CREB binding protein (CBP)- and P300-induced changes in global histone acetylation levels at specific lysine residues. Here the authors show that CBP and P300 exhibit unique substrate specificity profiles, consistent with the developmental and functional differences between the two HATs.
- 70Kasper, L. H.; Lerach, S.; Wang, J.; Wu, S.; Jeevan, T.; Brindle, P. K. EMBO J. 2010, 29, 3660There is no corresponding record for this reference.
- 71Dawson, M. A.; Kouzarides, T. Cell 2012, 150, 1271Cancer Epigenetics: From Mechanism to TherapyDawson, Mark A.; Kouzarides, TonyCell (Cambridge, MA, United States) (2012), 150 (1), 12-27CODEN: CELLB5; ISSN:0092-8674. (Cell Press)A review. The epigenetic regulation of DNA-templated processes has been intensely studied over the last 15 years. DNA methylation, histone modification, nucleosome remodeling, and RNA-mediated targeting regulate many biol. processes that are fundamental to the genesis of cancer. Here, we present the basic principles behind these epigenetic pathways and highlight the evidence suggesting that their misregulation can culminate in cancer. This information, along with the promising clin. and preclin. results seen with epigenetic drugs against chromatin regulators, signifies that it is time to embrace the central role of epigenetics in cancer.
- 72Yee, S. P.; Branton, P. E. Virology 1985, 147, 142There is no corresponding record for this reference.
- 73Whyte, P.; Williamson, N. M.; Harlow, E. Cell 1989, 56, 67There is no corresponding record for this reference.
- 74Eckner, R.; Ewen, M. E.; Newsome, D.; Gerdes, M.; DeCaprio, J. A.; Lawrence, J. B.; Livingston, D. M. Genes Dev. 1994, 8, 869There is no corresponding record for this reference.
- 75Chrivia, J. C.; Kwok, R. P.; Lamb, N.; Hagiwara, M.; Montminy, M. R.; Goodman, R. H. Nature 1993, 365, 855There is no corresponding record for this reference.
- 76Arany, Z.; Sellers, W. R.; Livingston, D. M.; Eckner, R. Cell 1994, 77, 799There is no corresponding record for this reference.
- 77Heery, D. M.; Fischer, P. M. Drug Discovery Today 2007, 12, 88There is no corresponding record for this reference.
- 78Brownell, J. E.; Zhou, J.; Ranalli, T.; Kobayashi, R.; Edmondson, D. G.; Roth, S. Y.; Allis, C. D. Cell 1996, 84, 84378Tetrahymena histone acetyltransferase A: a homolog to yeast Gcn5p linking histone acetylation to gene activationBrownell, James E.; Zhou, Jianxin; Ranalli, Tamara; Kobayashi, Ryuji; Edmondson, Diane G.; Roth, Sharon Y.; Allis, C. DavidCell (Cambridge, Massachusetts) (1996), 84 (6), 843-51CODEN: CELLB5; ISSN:0092-8674. (Cell Press)We report the cloning of a transcription-assocd. histone acetyltransferase type A (HAT A). This Tetrahymena enzyme is strikingly homologous to the yeast protein Gcn5, a putative transcriptional adaptor, and we demonstrate that recombinant Gcn5p possesses HAT activity. Both the ciliate enzyme and Gcn5p contain potential active site residues found in other acetyltransferases and a highly conserved bromodomain. The presence of this domain in nuclear A-type HATs, but not in cytoplasmic B-type HATs, suggests a mechanism whereby HAT A is directed to chromatin to facilitate transcriptional activation. These findings shed light on the biochem. function of the evolutionarily conserved Gcn5p-Ada complex, directly linking histone acetylation to gene activation, and indicate that histone acetylation is a targeted phenomenon.
- 79Brownell, J. E.; Allis, C. D. Proc. Natl. Acad. Sci. U.S.A. 1995, 92, 6364There is no corresponding record for this reference.
- 80Kleff, S.; Andrulis, E. D.; Anderson, C. W.; Sternglanz, R. J. Biol. Chem. 1995, 270, 24674There is no corresponding record for this reference.
- 81Bannister, A. J.; Kouzarides, T. Nature 1996, 384, 641There is no corresponding record for this reference.
- 82Ogryzko, V. V.; Schiltz, R. L.; Russanova, V.; Howard, B. H.; Nakatani, Y. Cell 1996, 87, 953There is no corresponding record for this reference.
- 83Eckner, R. Biol. Chem. 1996, 377, 685There is no corresponding record for this reference.
- 84Taddei, A.; Gasser, S. M. Genetics 2012, 192, 107There is no corresponding record for this reference.
- 85Rando, O. J.; Chang, H. Y. Annu. Rev. Biochem. 2009, 78, 24585Genome-wide views of chromatin structureRando, Oliver J.; Chang, Howard Y.Annual Review of Biochemistry (2009), 78 (), 245-271CODEN: ARBOAW; ISSN:0066-4154. (Annual Reviews Inc.)A review. Eukaryotic genomes are packaged into a nucleoprotein complex known as chromatin, which affects most processes that occur on DNA. Along with genetic and biochem. studies of resident chromatin proteins and their modifying enzymes, mapping of chromatin structure in vivo is one of the main pillars in our understanding of how chromatin relates to cellular processes. In this review, we discuss the use of genomic technologies to characterize chromatin structure in vivo, with a focus on data from budding yeast and humans. The picture emerging from these studies is the detailed chromatin structure of a typical gene, where the typical behavior gives insight into the mechanisms and deep rules that establish chromatin structure. Important deviation from the archetype is also obsd., usually as a consequence of unique regulatory mechanisms at special genomic loci. Chromatin structure shows substantial conservation from yeast to humans, but mammalian chromatin has addnl. layers of complexity that likely relate to the requirements of multicellularity such as the need to establish faithful gene regulatory mechanisms for cell differentiation.
- 86Bordoli, L.; Netsch, M.; Lüthi, U.; Lutz, W.; Eckner, R. Nucleic Acids Res. 2001, 29, 589There is no corresponding record for this reference.
- 87Tang, Y.; Holbert, M. A.; Wurtele, H.; Meeth, K.; Rocha, W.; Gharib, M.; Jiang, E.; Thibault, P.; Verreault, A.; Cole, P. A.; Marmorstein, R. Nat. Struct. Mol. Biol. 2008, 15, 99887Fungal Rtt109 histone acetyltransferase is an unexpected structural homolog of metazoan p300/CBP. [Erratum to document cited in CA149:217962]Tang, Yong; Holbert, Marc A.; Wurtele, Hugo; Meeth, Katrina; Rocha, Walter; Gharib, Marlene; Jiang, Eva; Thibault, Pierre; Verreault, Alain; Cole, Philip A.; Marmorstein, RonenNature Structural & Molecular Biology (2008), 15 (9), 998CODEN: NSMBCU; ISSN:1545-9993. (Nature Publishing Group)On page 738, the author name "Verreault" was mistakenly spelled as "Verrault". The error has been cor. in the HTML and PDF versions of the article.
- 88Bordoli, L.; Netsch, M.; Lüthi, U.; Lutz, W.; Eckner, R. Nucleic Acids Res. 2001, 29, 589There is no corresponding record for this reference.
- 89Giles, R. H.; Dauwerse, H. G.; van Ommen, G. J.; Breuning, M. H. Am. J. Hum. Genet. 1998, 63, 124089Do human chromosomal bands 16p13 and 22q11-13 share ancestral origins?Giles R H; Dauwerse H G; van Ommen G J; Breuning M HAmerican journal of human genetics (1998), 63 (4), 1240-2 ISSN:0002-9297.There is no expanded citation for this reference.
- 90Chan, H. M.; La Thangue, N. B. J. Cell Sci. 2001, 114, 236390p300/CBP proteins: HATs for transcriptional bridges and scaffoldsChan, Ho Man; La Thangue, Nicholas B.Journal of Cell Science (2001), 114 (13), 2363-2373CODEN: JNCSAI; ISSN:0021-9533. (Company of Biologists Ltd.)A review with 139 refs. P300/CBP transcriptional co-activator proteins play a central role in co-ordinating and integrating multiple signal-dependent events with the transcription app., allowing the appropriate level of gene activity to occur in response to diverse physiol. cues that influence, for example, proliferation, differentiation and apoptosis. P300/CBP activity can be under aberrant control in human disease, particularly in cancer, which may inactivate a p300/CBP tumor-suppressor-like activity. The transcription regulating-properties of p300 and CBP appear to be exerted through multiple mechanisms. They act as protein bridges, thereby connecting different sequence-specific transcription factors to the transcription app. Providing a protein scaffold upon which to build a multicomponent transcriptional regulatory complex is likely to be an important feature of p300/CBP control. Another key property is the presence of histone acetyltransferase (HAT) activity, which endows p300/CBP with the capacity to influence chromatin activity by modulating nucleosomal histones. Other proteins, including the p53 tumor suppressor, are targets for acetylation by p300/CBP. With the current intense level of research activity, p300/CBP will continue to be in the limelight and, we can be confident, yield new and important information on fundamental processes involved in transcriptional control.
- 91Ogryzko, V. V.; Schiltz, R. L.; Russanova, V.; Howard, B. H.; Nakatani, Y. Cell 1996, 87, 953There is no corresponding record for this reference.
- 92Yang, X. J.; Seto, E. Mol. Cell 2008, 31, 449There is no corresponding record for this reference.
- 93Bedford, D. C.; Brindle, Ph. Aging (N. Y.) 2012, 4, 24793Is histone acetylation the most important physiological function for CBP and p300?Bedford David C; Brindle Paul KAging (2012), 4 (4), 247-55 ISSN:.Protein lysine acetyltransferases (HATs or PATs) acetylate histones and other proteins, and are principally modeled as transcriptional coactivators. CREB binding protein (CBP, CREBBP) and its paralog p300 (EP300) constitute the KAT3 family of HATs in mammals, which has mostly unique sequence identity compared to other HAT families. Although studies in yeast show that many histone mutations cause modest or specific phenotypes, similar studies are impractical in mammals and it remains uncertain if histone acetylation is the primary physiological function for CBP/p300. Nonetheless, CBP and p300 mutations in humans and mice show that these coactivators have important roles in development, physiology, and disease, possibly because CBP and p300 act as network "hubs" with more than 400 described protein interaction partners. Analysis of CBP and p300 mutant mouse fibroblasts reveals CBP/p300 are together chiefly responsible for the global acetylation of histone H3 residues K18 and K27, and contribute to other locus-specific histone acetylation events. CBP/p300 can also be important for transcription, but the recruitment of CBP/p300 and their associated histone acetylation marks do not absolutely correlate with a requirement for gene activation. Rather, it appears that target gene context (e.g. DNA sequence) influences the extent to which CBP and p300 are necessary for transcription.
- 94Yang, X. J.; Seto, E. Mol. Cell 2008, 31, 449There is no corresponding record for this reference.
- 95Bedford, D. C.; Brindle, P. K. Aging (N. Y.) 2012, 4, 24795Is histone acetylation the most important physiological function for CBP and p300?Bedford David C; Brindle Paul KAging (2012), 4 (4), 247-55 ISSN:.Protein lysine acetyltransferases (HATs or PATs) acetylate histones and other proteins, and are principally modeled as transcriptional coactivators. CREB binding protein (CBP, CREBBP) and its paralog p300 (EP300) constitute the KAT3 family of HATs in mammals, which has mostly unique sequence identity compared to other HAT families. Although studies in yeast show that many histone mutations cause modest or specific phenotypes, similar studies are impractical in mammals and it remains uncertain if histone acetylation is the primary physiological function for CBP/p300. Nonetheless, CBP and p300 mutations in humans and mice show that these coactivators have important roles in development, physiology, and disease, possibly because CBP and p300 act as network "hubs" with more than 400 described protein interaction partners. Analysis of CBP and p300 mutant mouse fibroblasts reveals CBP/p300 are together chiefly responsible for the global acetylation of histone H3 residues K18 and K27, and contribute to other locus-specific histone acetylation events. CBP/p300 can also be important for transcription, but the recruitment of CBP/p300 and their associated histone acetylation marks do not absolutely correlate with a requirement for gene activation. Rather, it appears that target gene context (e.g. DNA sequence) influences the extent to which CBP and p300 are necessary for transcription.
- 96Arany, Z.; Newsome, D.; Oldread, E.; Livingston, D. M.; Eckner, R. Nature 1995, 374, 81There is no corresponding record for this reference.
- 97Ferreon, J. C.; Martinez-Yamout, M. A.; Dyson, H. J.; Wright, P. E. Proc. Natl. Acad. Sci. U.S.A. 2009, 106, 13260There is no corresponding record for this reference.
- 98Lundblad, J. R.; Kwok, R. P.; Laurance, M. E.; Harter, M. L.; Goodman, R. H. Nature 1995, 374, 85There is no corresponding record for this reference.
- 99Radhakrishnan, I.; Pérez-Alvarado, G. C.; Parker, D.; Dyson, H. J.; Montminy, M. R.; Wright, P. E. Cell 1997, 91, 741There is no corresponding record for this reference.
- 100Kalkhoven, E. Biochem. Pharmacol. 2004, 68, 1145There is no corresponding record for this reference.
- 101Nam, Y.; Sliz, P.; Song, L.; Aster, J. C.; Blacklow, S. C. Cell 2006, 124, 973There is no corresponding record for this reference.
- 102Mayr, B.; Montminy, M. Nat. Rev. Mol. Cell Biol. 2001, 2, 599102Transcriptional regulation by the phosphorylation-dependent factor CREBMayr, Bernhard; Montminy, MarcNature Reviews Molecular Cell Biology (2001), 2 (8), 599-609CODEN: NRMCBP; ISSN:1471-0072. (Nature Publishing Group)A review. The transcription factor CREB - for 'cAMP response element-binding protein' - functions in glucose homeostasis, growth-factor-dependent cell survival, and has been implicated in learning and memory. CREB is phosphorylated in response to various signals, but how is specificity achieved in these signaling pathways.
- 103Bourdeau, V.; Deschênes, J.; Métivier, R.; Nagai, Y.; Nguyen, D.; Bretschneider, N.; Gannon, F.; White, J. H.; Mader, S. Mol. Endocrinol. 2004, 18, 1411103Genome-wide identification of high-affinity estrogen response elements in human and mouseBourdeau, Veronique; Deschenes, Julie; Metivier, Raphael; Nagai, Yoshihiko; Nguyen, Denis; Bretschneider, Nancy; Gannon, Frank; White, John H.; Mader, SylvieMolecular Endocrinology (2004), 18 (6), 1411-1427CODEN: MOENEN; ISSN:0888-8809. (Endocrine Society)Although estrogen receptors (ERs) recognize 15-bp palindromic estrogen response elements (EREs) with maximal affinity in vitro, few near-consensus sequences have been characterized in estrogen target genes. Here we report the design of a genome-wide screen for high-affinity EREs and the identification of approx. 70,000 motifs in the human and mouse genomes. EREs are enriched in regions proximal to the transcriptional start sites, and approx. 1% of elements appear conserved in the flanking regions (-10 kb to +5 kb) of orthologous human and mouse genes. Conserved and nonconserved elements were also found, often in multiple occurrences, in more than 230 estrogen-stimulated human genes previously identified from expression studies. In genes contg. known EREs, we also identified addnl. distal elements, sometimes with higher in vitro binding affinity and/or better conservation between the species considered. Chromatin immunopptn. expts. in breast cancer cell lines indicate that most novel elements present in responsive genes bind ERα in vivo, including some EREs located up to approx. 10 kb from transcriptional start sites. Our results demonstrate that near-consensus EREs occur frequently in both genomes and that whereas chromatin structure likely modulates access to binding sites, far upstream elements can be evolutionarily conserved and bind ERs in vivo.
- 104Saito, H.; Posas, F. Genetics 2012, 192, 289There is no corresponding record for this reference.
- 105Sheikh-Hamad, D.; Di Mari, J.; Suki, W. N.; Safirstein, R.; Watts, B. A., III; Rouse, D. J. Biol. Chem. 1998, 273, 1832There is no corresponding record for this reference.
- 106Chen, Y. J.; Wang, Y. N.; Chang, W. C. J. Biol. Chem. 2007, 282, 27215There is no corresponding record for this reference.
- 107Jang, E. R.; Choi, J. D.; Lee, J. S. FEBS Lett. 2011, 585, 47There is no corresponding record for this reference.
- 108Sakaguchi, K.; Herrera, J. E.; Saito, S.; Miki, T.; Bustin, M.; Vassilev, A.; Anderson, C. W.; Appella, E. Genes Dev. 1998, 12, 2831There is no corresponding record for this reference.
- 109Iyer, N. G.; Chin, S. F.; Ozdag, H.; Daigo, Y.; Hu, D. E.; Cariati, M.; Brindle, K.; Aparicio, S.; Caldas, C. Proc. Natl. Acad. Sci. U.S.A. 2004, 101, 7386There is no corresponding record for this reference.
- 110Yuan, L. W.; Gambee, J. E. J. Biol. Chem. 2000, 275, 40946There is no corresponding record for this reference.
- 111Yang, X. J.; Seto, E. Mol. Cell 2008, 31, 449There is no corresponding record for this reference.
- 112Legube, G.; Trouche, D. EMBO Rep. 2003, 4, 944There is no corresponding record for this reference.
- 113Schwartz, C.; Beck, K.; Mink, S.; Schmolke, M.; Budde, B.; Wenning, D.; Klempnauer, K. H. EMBO J. 2003, 22, 882There is no corresponding record for this reference.
- 114Poizat, C.; Puri, P. L.; Bai, Y.; Kedes, L. Mol. Cell. Biol. 2005, 25, 2673114Phosphorylation-dependent degradation of p300 by doxorubicin-activated p38 mitogen-activated protein kinase in cardiac cellsPoizat, Coralie; Puri, Pier Lorenzo; Bai, Yan; Kedes, LarryMolecular and Cellular Biology (2005), 25 (7), 2673-2687CODEN: MCEBD4; ISSN:0270-7306. (American Society for Microbiology)P300 and CBP are general transcriptional coactivators implicated in different cellular processes, including regulation of the cell cycle, differentiation, tumorigenesis, and apoptosis. Posttranslational modifications such as phosphorylation are predicted to select a specific function of p300/CBP in these processes; however, the identification of the kinases that regulate p300/CBP activity in response to individual stimuli and the physiol. significance of p300 phosphorylation have not been elucidated. Here we demonstrate that the cardiotoxic anticancer agent doxorubicin (adriamycin) induces the phosphorylation of p300 in primary neonatal cardiomyocytes. Hyperphosphorylation precedes the degrdn. of p300 and parallels apoptosis in response to doxorubicin. Doxorubicin-activated p38 kinases α and β assoc. with p300 and are implicated in the phosphorylation-mediated degrdn. of p300, as pharmacol. blockade of p38 prevents p300 degrdn. P38 phosphorylates p300 in vitro at both the N and C termini of the protein, and enforced activation of p38 by the constitutively active form of its upstream kinase (MKK6EE) triggers p300 degrdn. These data support the conclusion that p38 mitogen-activated protein kinase regulates p300 protein stability and function in cardiomyocytes undergoing apoptosis in response to doxorubicin.
- 115Xu, W.; Chen, H.; Du, K.; Asahara, H.; Tini, M.; Emerson, B. M.; Montminy, M.; Evans, R. M. Science 2001, 294, 2507There is no corresponding record for this reference.
- 116Ceschin, D. G.; Walia, M.; Wenk, S. S.; Duboé, C.; Gaudon, C.; Xiao, Y.; Fauquier, L.; Sankar, M.; Vandel, L.; Gronemeyer, H. Genes Dev. 2011, 25, 1132There is no corresponding record for this reference.
- 117Ryan, C. M.; Kindle, K. B.; Collins, H. M.; Heery, D. M. Biochem. Biophys. Res. Commun. 2010, 391, 1136There is no corresponding record for this reference.
- 118Kuo, H. Y.; Chang, C. C.; Jeng, J. C.; Hu, H. M.; Lin, D. Y.; Maul, G. G.; Kwok, R. P.; Shih, H. M. Proc. Natl. Acad. Sci. U.S.A. 2005, 102, 16973There is no corresponding record for this reference.
- 119Karanam, B.; Jiang, L.; Wang, L.; Kelleher, N. L.; Cole, P. A. J. Biol. Chem. 2006, 281, 40292There is no corresponding record for this reference.
- 120Karukurichi, K. R.; Wang, L.; Uzasci, L.; Manlandro, C. M.; Wang, Q.; Cole, P. A. J. Am. Chem. Soc. 2010, 132, 1222There is no corresponding record for this reference.
- 121Thompson, P. R.; Wang, D.; Wang, L.; Fulco, M.; Pediconi, N.; Zhang, D.; An, W.; Ge, Q.; Roeder, R. G.; Wong, J.; Levrero, M.; Sartorelli, V.; Cotter, R. J.; Cole, P. A. Nat. Struct. Mol. Biol. 2004, 11, 308121Regulation of the p300 HAT domain via a novel activation loopThompson, Paul R.; Wang, Dongxia; Wang, Ling; Fulco, Marcella; Pediconi, Natalia; Zhang, Dianzheng; An, Woojin; Ge, Qingyuan; Roeder, Robert G.; Wong, Jiemin; Levrero, Massimo; Sartorelli, Vittorio; Cotter, Robert J.; Cole, Philip A.Nature Structural & Molecular Biology (2004), 11 (4), 308-315CODEN: NSMBCU; ISSN:1545-9993. (Nature Publishing Group)The transcriptional coactivator p300 is a histone acetyltransferase (HAT) whose function is crit. for regulating gene expression in mammalian cells. However, the mol. events that regulate p300 HAT activity are poorly understood. We evaluated autoacetylation of the p300 HAT protein domain to det. its function. Using expressed protein ligation, the p300 HAT protein domain was generated in hypoacetylated form and found to have reduced catalytic activity. This basal catalytic rate was stimulated by autoacetylation of several key lysine sites within an apparent activation loop motif. This post-translational modification and catalytic regulation of p300 HAT activity is conceptually analogous to the activation of most protein kinases by autophosphorylation. We therefore propose that this autoregulatory loop could influence the impact of p300 on a wide variety of signaling and transcriptional events.
- 122Karanam, B.; Wang, L.; Wang, D.; Liu, X.; Marmorstein, R.; Cotter, R.; Cole, P. A. Biochemistry 2007, 46, 8207There is no corresponding record for this reference.
- 123Kasper, L. H.; Fukuyama, T.; Biesen, M. A.; Boussouar, F.; Tong, C.; de Pauw, A.; Murray, P. J.; van Deursen, J. M.; Brindle, P. K. Mol. Cell. Biol. 2006, 26, 789123Conditional knockout mice reveal distinct functions for the global transcriptional coactivators CBP and p300 in T-cell developmentKasper, Lawryn H.; Fukuyama, Tomofusa; Biesen, Michelle A.; Boussouar, Faycal; Tong, Caili; de Pauw, Antoine; Murray, Peter J.; van Deursen, Jan M. A.; Brindle, Paul K.Molecular and Cellular Biology (2006), 26 (3), 789-809CODEN: MCEBD4; ISSN:0270-7306. (American Society for Microbiology)The global transcriptional coactivators CREB-binding protein (CBP) and the closely related p300 interact with over 312 proteins, making them among the most heavily connected hubs in the known mammalian protein-protein interactome. It is largely uncertain, however, if these interactions are important in specific cell lineages of adult animals, as homozygous null mutations in either CBP or p300 result in early embryonic lethality in mice. Here we describe a Cre/LoxP conditional p300 null allele (p300flox) that allows for the temporal and tissue-specific inactivation of p300. We used mice carrying p300flax and a CBP conditional knockout allele (CBPflax) in conjunction with an Lck-Cre transgene to delete CBP and p300 starting at the CD4- CD8- double-neg. thymocyte stage of T-cell development. Loss of either p300 or CBP led to a decrease in CD4+ CD8+ double-pos. thymocytes, but an increase in the percentage of CD8+ single-pos. thymocytes seen in CBP mutant mice was not obsd. in p300 mutants. T cells completely lacking both CBP and p300 did not develop normally and were nonexistent or very rare in the periphery, however. T cells lacking CBP or p300 had reduced tumor necrosis factor alpha gene expression in response to phorbol ester and ionophore, while signal-responsive gene expression in CBP- or p300-deficient macrophages was largely intact. Thus, CBP and p300 each supply a surprising degree of redundant coactivation capacity in T cells and macrophages, although each gene has also unique properties in thymocyte development.
- 124Personal communication with Paul Brindle, see: http://www.stjude.org/stjude/v/index.jsp?vgnextoid=c30215f204294210VgnVCM1000001e0215acRCRD.There is no corresponding record for this reference.
- 125Yan, G.; Eller, M. S.; Elm, C.; Larocca, C. A.; Ryu, B.; Panova, I. P.; Dancy, B. M.; Bowers, E. M.; Meyers, D.; Lareau, L.; Cole, P. A.; Taverna, S. D.; Alani, R. M. J. Invest. Dermatol. 2013, 133, 2444There is no corresponding record for this reference.
- 126Cohen, I.; Poręba, E.; Kamieniarz, K.; Schneider, R. Genes Cancer 2011, 2, 631126Histone modifiers in cancer: friends or foes?Cohen, Idan; Poreba, Elzbieta; Kamieniarz, Kinga; Schneider, RobertGenes & Cancer (2011), 2 (6), 631-647CODEN: GCEAAY; ISSN:1947-6019. (Sage Publications)A review. Covalent modifications of histories can regulate all DNA-dependent processes. In the last few years, it has become more and more evident that histone modifications are key players in the regulation of chromatin states and dynamics as well as in gene expression. Therefore, histone modifications and the enzymic machineries that set them are crucial regulators that can control cellular proliferation, differentiation, plasticity, and malignancy processes. This review discusses the biol. and biochem. of covalent histone posttranslational modifications (PTMs) and evaluates the dual role of their modifiers in cancer: as oncogenes that can initiate and amplify tumorigenesis or as tumor suppressors.
- 127Chan, H. M.; La Thangue, N. B. J. Cell Sci. 2001, 114, 2363127p300/CBP proteins: HATs for transcriptional bridges and scaffoldsChan, Ho Man; La Thangue, Nicholas B.Journal of Cell Science (2001), 114 (13), 2363-2373CODEN: JNCSAI; ISSN:0021-9533. (Company of Biologists Ltd.)A review with 139 refs. P300/CBP transcriptional co-activator proteins play a central role in co-ordinating and integrating multiple signal-dependent events with the transcription app., allowing the appropriate level of gene activity to occur in response to diverse physiol. cues that influence, for example, proliferation, differentiation and apoptosis. P300/CBP activity can be under aberrant control in human disease, particularly in cancer, which may inactivate a p300/CBP tumor-suppressor-like activity. The transcription regulating-properties of p300 and CBP appear to be exerted through multiple mechanisms. They act as protein bridges, thereby connecting different sequence-specific transcription factors to the transcription app. Providing a protein scaffold upon which to build a multicomponent transcriptional regulatory complex is likely to be an important feature of p300/CBP control. Another key property is the presence of histone acetyltransferase (HAT) activity, which endows p300/CBP with the capacity to influence chromatin activity by modulating nucleosomal histones. Other proteins, including the p53 tumor suppressor, are targets for acetylation by p300/CBP. With the current intense level of research activity, p300/CBP will continue to be in the limelight and, we can be confident, yield new and important information on fundamental processes involved in transcriptional control.
- 128Pan, C. Q.; Sudol, M.; Sheetz, M.; Low, B. C. Cell. Signalling 2012, 24, 2143128Modularity and functional plasticity of scaffold proteins as p(l)acemakers in cell signalingPan, Catherine Qiurong; Sudol, Marius; Sheetz, Michael; Low, Boon ChuanCellular Signalling (2012), 24 (11), 2143-2165CODEN: CESIEY; ISSN:0898-6568. (Elsevier Inc.)A review. Cells coordinate and integrate various functional modules that control their dynamics, intracellular trafficking, metab. and gene expression. Such capacity is mediated by specific scaffold proteins that tether multiple components of signaling pathways at plasma membrane, Golgi app., mitochondria, endoplasmic reticulum, nucleus and in more specialized subcellular structures such as focal adhesions, cell-cell junctions, endosomes, vesicles and synapses. Scaffold proteins act as "pacemakers" as well as "placemakers" that regulate the temporal, spatial and kinetic aspects of protein complex assembly by modulating the local concns., proximity, subcellular dispositions and biochem. properties of the target proteins through the intricate use of their modular protein domains. These regulatory mechanisms allow them to gate the specificity, integration and crosstalk of different signaling modules. In addn. to acting as phys. platforms for protein assembly, many professional scaffold proteins can also directly modify the properties of their targets while they themselves can be regulated by post-translational modifications and/or mech. forces. Furthermore, multiple scaffold proteins can form alliances of higher-order regulatory networks. Here, we highlight the emerging themes of scaffold proteins by analyzing their common and distinctive mechanisms of action and regulation, which underlie their functional plasticity in cell signaling. Understanding these mechanisms in the context of space, time and force should have ramifications for human physiol. and for developing new therapeutic approaches to control pathol. states and diseases.
- 129Tanaka, Y.; Naruse, I.; Maekawa, T.; Masuya, H.; Shiroishi, T.; Ishii, S. Proc. Natl. Acad. Sci. U.S.A. 1997, 94, 10215There is no corresponding record for this reference.
- 130Kamei, Y.; Xu, L.; Heinzel, T.; Torchia, J.; Kurokawa, R.; Gloss, B.; Lin, S. C.; Heyman, R. A.; Rose, D. W.; Glass, C. K.; Rosenfeld, M. G. Cell 1996, 85, 403There is no corresponding record for this reference.
- 131Hottiger, M. O.; Felzien, L. K.; Nabel, G. J. EMBO J. 1998, 17, 3124There is no corresponding record for this reference.
- 132Yin, X.; Warner, D. R.; Roberts, E. A.; Pisano, M. M.; Greene, R. M. Biochem. Biophys. Res. Commun. 2005, 329, 1010There is no corresponding record for this reference.
- 133Avantaggiati, M. L.; Ogryzko, V.; Gardner, K.; Giordano, A.; Levine, A. S.; Kelly, K. Cell 1997, 89, 1175There is no corresponding record for this reference.
- 134Kamei, Y.; Xu, L.; Heinzel, T.; Torchia, J.; Kurokawa, R.; Gloss, B.; Lin, S. C.; Heyman, R. A.; Rose, D. W.; Glass, C. K.; Rosenfeld, M. G. Cell 1996, 85, 403There is no corresponding record for this reference.
- 135Goodman, R. H.; Smolik, S. Genes Dev. 2000, 14, 1553135CBP/p300 in cell growth, transformation, and developmentGoodman, Richard H.; Smolik, SarahGenes & Development (2000), 14 (13), 1553-1577CODEN: GEDEEP; ISSN:0890-9369. (Cold Spring Harbor Laboratory Press)A review with several refs. This review discusses the involvement of CBP/p300 in cell growth and transformation, chromosomal translocations, involvement of cellular oncogenes, assocn. with viral oncoproteins, involvement with DNA and RNA tumor viruses, the regulation of CBP/p300 by phosphorylation, and usage in development of model systems.
- 136Lee, C. W.; Sørensen, T. S.; Shikama, N.; La Thangue, N. B. Oncogene 1998, 16, 2695There is no corresponding record for this reference.
- 137Nakajima, T.; Fukamizu, A.; Takahashi, J.; Gage, F. H.; Fisher, T.; Blenis, J.; Montminy, M. R. Cell 1996, 86, 465There is no corresponding record for this reference.
- 138Thompson, P. R.; Wang, D.; Wang, L.; Fulco, M.; Pediconi, N.; Zhang, D.; An, W.; Ge, Q.; Roeder, R. G.; Wong, J.; Levrero, M.; Sartorelli, V.; Cotter, R. J.; Cole, P. A. Nat. Struct. Mol. Biol. 2004, 11, 308138Regulation of the p300 HAT domain via a novel activation loopThompson, Paul R.; Wang, Dongxia; Wang, Ling; Fulco, Marcella; Pediconi, Natalia; Zhang, Dianzheng; An, Woojin; Ge, Qingyuan; Roeder, Robert G.; Wong, Jiemin; Levrero, Massimo; Sartorelli, Vittorio; Cotter, Robert J.; Cole, Philip A.Nature Structural & Molecular Biology (2004), 11 (4), 308-315CODEN: NSMBCU; ISSN:1545-9993. (Nature Publishing Group)The transcriptional coactivator p300 is a histone acetyltransferase (HAT) whose function is crit. for regulating gene expression in mammalian cells. However, the mol. events that regulate p300 HAT activity are poorly understood. We evaluated autoacetylation of the p300 HAT protein domain to det. its function. Using expressed protein ligation, the p300 HAT protein domain was generated in hypoacetylated form and found to have reduced catalytic activity. This basal catalytic rate was stimulated by autoacetylation of several key lysine sites within an apparent activation loop motif. This post-translational modification and catalytic regulation of p300 HAT activity is conceptually analogous to the activation of most protein kinases by autophosphorylation. We therefore propose that this autoregulatory loop could influence the impact of p300 on a wide variety of signaling and transcriptional events.
- 139Muir, T. W.; Sondhi, D.; Cole, P. A. Proc. Natl. Acad. Sci. U.S.A. 1998, 95, 6705139Expressed protein ligation: A general method for protein engineeringMuir, Tom W.; Sondhi, Dolan; Cole, Philip A.Proceedings of the National Academy of Sciences of the United States of America (1998), 95 (12), 6705-6710CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)A protein semisynthesis method-expressed protein ligation-is described that involves the chemoselective addn. of a peptide to a recombinant protein. This method was used to ligate a phosphotyrosine peptide to the C terminus of the protein tyrosine kinase C-terminal Src kinase (Csk). By intercepting a thioester generated in the recombinant protein with an N-terminal cysteine contg. synthetic peptide, near quant. chem. ligation of the peptide to the protein was achieved. The semisynthetic tail-phosphorylated Csk showed evidence of an intramol. phosphotyrosine-Src homol. 2 interaction and an unexpected increase in catalytic phosphoryl transfer efficiency toward a physiol. relevant substrate compared with the non-tail-phosphorylated control. This work illustrates that expressed protein ligation is a simple and powerful new method in protein engineering to introduce sequences of unnatural amino acids, posttranslational modifications, and biophys. probes into proteins of any size.
- 140Dancy, B. M.; Crump, N. T.; Peterson, D. J.; Mukherjee, C.; Bowers, E. M.; Ahn, Y. H.; Yoshida, M.; Zhang, J.; Mahadevan, L. C.; Meyers, D. J.; Boeke, J. D.; Cole, P. A. ChemBioChem 2012, 13, 2113There is no corresponding record for this reference.
- 141Liu, X.; Wang, L.; Zhao, K.; Thompson, P. R.; Hwang, Y.; Marmorstein, R.; Cole, P. A. Nature 2008, 451, 846There is no corresponding record for this reference.
- 142Hwang, Y.; Thompson, P. R.; Wang, L.; Jiang, L.; Kelleher, N. L.; Cole, P. A. Angew. Chem., Int. Ed. 2007, 46, 7621142A selective chemical probe for coenzyme A-requiring enzymesHwang, Yousang; Thompson, Paul R.; Wang, Ling; Jiang, Lihua; Kelleher, Neil L.; Cole, Philip A.Angewandte Chemie, International Edition (2007), 46 (40), 7621-7624CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A CoA-based affinity probe with a sulfoxycarbamate functionality can selectively identify several acetyltransferases relative to other enzymes and proteins. It leaves behind a desthiobiotin tag that can be used for western blotting and mass spectrometric characterization.
- 143Liu, X.; Wang, L.; Zhao, K.; Thompson, P. R.; Hwang, Y.; Marmorstein, R.; Cole, P. A. Nature 2008, 451, 846There is no corresponding record for this reference.
- 144Tang, Y.; Holbert, M. A.; Wurtele, H.; Meeth, K.; Rocha, W.; Gharib, M.; Jiang, E.; Thibault, P.; Verreault, A.; Cole, P. A.; Marmorstein, R. Nat. Struct. Mol. Biol. 2008, 15, 998144Fungal Rtt109 histone acetyltransferase is an unexpected structural homolog of metazoan p300/CBP. [Erratum to document cited in CA149:217962]Tang, Yong; Holbert, Marc A.; Wurtele, Hugo; Meeth, Katrina; Rocha, Walter; Gharib, Marlene; Jiang, Eva; Thibault, Pierre; Verreault, Alain; Cole, Philip A.; Marmorstein, RonenNature Structural & Molecular Biology (2008), 15 (9), 998CODEN: NSMBCU; ISSN:1545-9993. (Nature Publishing Group)On page 738, the author name "Verreault" was mistakenly spelled as "Verrault". The error has been cor. in the HTML and PDF versions of the article.
- 145Wang, L.; Tang, Y.; Cole, P. A.; Marmorstein, R. Curr. Opin. Struct. Biol. 2008, 18, 741145Structure and chemistry of the p300/CBP and Rtt109 histone acetyltransferases: implications for histone acetyltransferase evolution and functionWang, Ling; Tang, Yong; Cole, Philip A.; Marmorstein, RonenCurrent Opinion in Structural Biology (2008), 18 (6), 741-747CODEN: COSBEF; ISSN:0959-440X. (Elsevier B.V.)A review. The recent crystal structure and assocd. biochem. studies of the metazoan-specific p300/CBP and fungal-specific Rtt109 histone acetyltransferases (HATs) have provided new insights into the ancestral relation between HATs and their functions. These studies point to a common HAT ancestor that has evolved around a common structural framework to form HATs with divergent catalytic and substrate-binding properties. These studies also point to the importance of regulatory loops within HATs and autoacetylation in HAT function. Implications for future studies are discussed.
- 146Poux, A. N.; Cebrat, M.; Kim, C. M.; Cole, P. A.; Marmorstein, R. Proc. Natl. Acad. Sci. U.S.A. 2009, 99, 14065There is no corresponding record for this reference.
- 147Yuan, H.; Rossetto, D.; Mellert, H.; Dang, W.; Srinivasan, M.; Johnson, J.; Hodawadekar, S.; Ding, E. C.; Speicher, K.; Abshiru, N.; Perry, R.; Wu, J.; Yang, C.; Zheng, Y. G.; Speicher, D. W.; Thibault, P.; Verreault, A.; Johnson, F. B.; Berger, S. L.; Sternglanz, R.; McMahon, S. B.; Côté, J.; Marmorstein, R. EMBO J. 2012, 31, 58There is no corresponding record for this reference.
- 148Karukurichi, K. R.; Cole, P. A. Bioorg. Chem. 2011, 39, 42148Probing the reaction coordinate of the p300/CBP histone acetyltransferase with bisubstrate analogsKarukurichi, Kannan R.; Cole, Philip A.Bioorganic Chemistry (2011), 39 (1), 42-47CODEN: BOCMBM; ISSN:0045-2068. (Elsevier B.V.)Histone and protein acetylation catalyzed by p300/CBP transcriptional coactivator regulates a variety of key biol. pathways. This study investigates the proposed Theorell-Chance or "hit-and-run" catalytic mechanism of p300/CBP histone acetyltransferase (HAT) using bisubstrate analogs. A range of histone peptide tail peptide-CoA conjugates with different length linkers were synthesized and evaluated as inhibitors of p300 HAT. We show that longer linkers between the histone tail peptide and the CoA substrate moieties appear to allow for dual engagement of the two binding surfaces. Results with D1625R/D1628R double mutant p300 HAT further confirm the requirement for a neg. charged surface on the enzyme to interact with the histone tail.
- 149Cleland, W. W. Biochim. Biophys. Acta 1963, 67, 104There is no corresponding record for this reference.
- 150Cleland, W. W. Biochim. Biophys. Acta 1963, 67, 104There is no corresponding record for this reference.
- 151Segel, I. Enzyme Kinetics; Wiley Interscience: New York, 1975.There is no corresponding record for this reference.
- 152Zheng, Y.; Thompson, P. R.; Cebrat, M.; Wang, L.; Devlin, M. K.; Alani, R. M.; Cole, P. A. Methods Enzymol. 2004, 376, 188There is no corresponding record for this reference.
- 153Bowers, E. M.; Yan, G.; Mukherjee, C.; Orry, A.; Wang, L.; Holbert, M. A.; Crump, N. T.; Hazzalin, C. A.; Liszczak, G.; Yuan, H.; Larocca, C.; Saldanha, S. A.; Abagyan, R.; Sun, Y.; Meyers, D. J.; Marmorstein, R.; Mahadevan, L. C.; Alani, R. M.; Cole, P. A. Chem. Biol. 2010, 17, 471153Virtual Ligand Screening of the p300/CBP Histone Acetyltransferase: Identification of a Selective Small Molecule InhibitorBowers, Erin M.; Yan, Gai; Mukherjee, Chandrani; Orry, Andrew; Wang, Ling; Holbert, Marc A.; Crump, Nicholas T.; Hazzalin, Catherine A.; Liszczak, Glen; Yuan, Hua; Larocca, Cecilia; Saldanha, S. Adrian; Abagyan, Ruben; Sun, Yan; Meyers, David J.; Marmorstein, Ronen; Mahadevan, Louis C.; Alani, Rhoda M.; Cole, Philip A.Chemistry & Biology (Cambridge, MA, United States) (2010), 17 (5), 471-482CODEN: CBOLE2; ISSN:1074-5521. (Cell Press)Summary: The histone acetyltransferase (HAT) p300/CBP is a transcriptional coactivator implicated in many gene regulatory pathways and protein acetylation events. Although p300 inhibitors have been reported, a potent, selective, and readily available active-site-directed small mol. inhibitor is not yet known. Here the authors use a structure-based, in silico screening approach to identify a com. available pyrazolone-contg. small mol. p300 HAT inhibitor, C646. C646 is a competitive p300 inhibitor with a Ki of 400 nM and is selective vs. other acetyltransferases. Studies on site-directed p300 HAT mutants and synthetic modifications of C646 confirm the importance of predicted interactions in conferring potency. Inhibition of histone acetylation and cell growth by C646 in cells validate its utility as a pharmacol. probe and suggest that p300/CBP HAT is a worthy anticancer target.
- 154Lau, O. D.; Kundu, T. K.; Soccio, R. E.; Ait-Si-Ali, S.; Khalil, E. M.; Vassilev, A.; Wolffe, A. P.; Nakatani, Y.; Roeder, R. G.; Cole, P. A. Mol. Cell 2000, 5, 589There is no corresponding record for this reference.
- 155Sagar, V.; Zheng, W.; Thompson, P. R.; Cole, P. A. Bioorg. Med. Chem. 2004, 12, 3383There is no corresponding record for this reference.
- 156Karukurichi, K. R.; Cole, P. A. Bioorg. Chem. 2011, 39, 42156Probing the reaction coordinate of the p300/CBP histone acetyltransferase with bisubstrate analogsKarukurichi, Kannan R.; Cole, Philip A.Bioorganic Chemistry (2011), 39 (1), 42-47CODEN: BOCMBM; ISSN:0045-2068. (Elsevier B.V.)Histone and protein acetylation catalyzed by p300/CBP transcriptional coactivator regulates a variety of key biol. pathways. This study investigates the proposed Theorell-Chance or "hit-and-run" catalytic mechanism of p300/CBP histone acetyltransferase (HAT) using bisubstrate analogs. A range of histone peptide tail peptide-CoA conjugates with different length linkers were synthesized and evaluated as inhibitors of p300 HAT. We show that longer linkers between the histone tail peptide and the CoA substrate moieties appear to allow for dual engagement of the two binding surfaces. Results with D1625R/D1628R double mutant p300 HAT further confirm the requirement for a neg. charged surface on the enzyme to interact with the histone tail.
- 157Cebrat, M.; Kim, C. M.; Thompson, P. R.; Daugherty, M.; Cole, P. A. Bioorg. Med. Chem. 2003, 11, 3307There is no corresponding record for this reference.
- 158Lau, O. D.; Kundu, T. K.; Soccio, R. E.; Ait-Si-Ali, S.; Khalil, E. M.; Vassilev, A.; Wolffe, A. P.; Nakatani, Y.; Roeder, R. G.; Cole, P. A. Mol. Cell 2000, 5, 589There is no corresponding record for this reference.
- 159Chase, J. F.; Tubbs, P. K. Biochem. J. 1969, 111, 225There is no corresponding record for this reference.
- 160Williams, J. W.; Northrop, D. B. J. Antibiot. 1979, 32, 1147There is no corresponding record for this reference.
- 161Cullis, P. M.; Wolfenden, R.; Cousens, L. S.; Alberts, B. M. J. Biol. Chem. 1982, 257, 12165There is no corresponding record for this reference.
- 162Erwin, B. G.; Persson, L.; Pegg, A. E. Biochemistry 1984, 23, 4250There is no corresponding record for this reference.
- 163Khalil, E. M.; Cole, P. A. J. Am. Chem. Soc. 1998, 120, 6195There is no corresponding record for this reference.
- 164Khalil, E. M.; De Angelis, J.; Ishii, M.; Cole, P. A. Proc. Natl. Acad. Sci. U.S.A. 1999, 96, 12418There is no corresponding record for this reference.
- 165Kim, C. M.; Cole, P. A. J. Med. Chem. 2001, 44, 2479There is no corresponding record for this reference.
- 166Yu, M.; Magalhães, M. L.; Cook, P. F.; Blanchard, J. S. Biochemistry 2006, 45, 14788There is no corresponding record for this reference.
- 167Gao, F.; Yan, X.; Baettig, O. M.; Berghuis, A. M.; Auclair, K. Angew. Chem., Int. Ed. 2005, 44, 6859There is no corresponding record for this reference.
- 168Gao, F.; Yan, X.; Shakya, T.; Baettig, O. M.; Ait-Mohand-Brunet, S.; Berghuis, A. M.; Wright, G. D.; Auclair, K. J. Med. Chem. 2006, 49, 5273There is no corresponding record for this reference.
- 169Gao, F.; Yan, X.; Zahr, O.; Larsen, A.; Vong, K.; Auclair, K. Bioorg. Med. Chem. Lett. 2008, 18, 5518There is no corresponding record for this reference.
- 170Magalhães, M. L.; Vetting, M. W.; Gao, F.; Freiburger, L.; Auclair, K.; Blanchard, J. S. Biochemistry 2008, 47, 579There is no corresponding record for this reference.
- 171Barnett, B. P.; Hwang, Y.; Taylor, M. S.; Kirchner, H.; Pfluger, P. T.; Bernard, V.; Lin, Y. Y.; Bowers, E. M.; Mukherjee, C.; Song, W. J.; Longo, P. A.; Leahy, D. J.; Hussain, M. A.; Tschöp, M. H.; Boeke, J. D.; Cole, P. A. Science 2010, 330, 1689There is no corresponding record for this reference.
- 172Taylor, M. S.; Hwang, Y.; Hsiao, P. Y.; Boeke, J. D.; Cole, P. A. Methods Enzymol. 2012, 514, 205There is no corresponding record for this reference.
- 173Parang, K.; Till, J. H.; Ablooglu, A. J.; Kohanski, R. A.; Hubbard, S. R.; Cole, P. A. Nat. Struct. Biol. 2001, 8, 37There is no corresponding record for this reference.
- 174Shen, K.; Hines, A. C.; Schwarzer, D.; Pickin, K. A.; Cole, P. A. Biochim. Biophys. Acta 2005, 1754, 65There is no corresponding record for this reference.
- 175Lau, O. D.; Kundu, T. K.; Soccio, R. E.; Ait-Si-Ali, S.; Khalil, E. M.; Vassilev, A.; Wolffe, A. P.; Nakatani, Y.; Roeder, R. G.; Cole, P. A. Mol. Cell 2000, 5, 589There is no corresponding record for this reference.
- 176Wu, J.; Xie, N.; Wu, Z.; Zhang, Y.; Zheng, Y. G. Bioorg. Med. Chem. 2009, 17, 1381There is no corresponding record for this reference.
- 177Zheng, Y.; Thompson, P. R.; Cebrat, M.; Wang, L.; Devlin, M. K.; Alani, R. M.; Cole, P. A. Methods Enzymol. 2004, 376, 188There is no corresponding record for this reference.
- 178De Angelis, J.; Gastel, J.; Klein, D. C.; Cole, P. A. J. Biol. Chem. 1998, 273, 3045There is no corresponding record for this reference.
- 179Lau, O. D.; Courtney, A. D.; Vassilev, A.; Marzilli, L. A.; Cotter, R. J.; Nakatani, Y.; Cole, P. A. J. Biol. Chem. 2000, 275, 21953There is no corresponding record for this reference.
- 180Tanner, K. G.; Langer, M. R.; Denu, J. M. Biochemistry 2000, 39, 15652There is no corresponding record for this reference.
- 181Tanner, K. G.; Langer, M. R.; Kim, Y.; Denu, J. M. J. Biol. Chem. 2000, 275, 22048There is no corresponding record for this reference.
- 182Yan, Y.; Barlev, N. A.; Haley, R. H.; Berger, S. L.; Marmorstein, R. Mol. Cell 2000, 6, 1195There is no corresponding record for this reference.
- 183Wu, J.; Xie, N.; Wu, Z.; Zhang, Y.; Zheng, Y. G. Bioorg. Med. Chem. 2009, 17, 1381There is no corresponding record for this reference.
- 184Lau, O. D.; Kundu, T. K.; Soccio, R. E.; Ait-Si-Ali, S.; Khalil, E. M.; Vassilev, A.; Wolffe, A. P.; Nakatani, Y.; Roeder, R. G.; Cole, P. A. Mol. Cell 2000, 5, 589There is no corresponding record for this reference.
- 185Lau, O. D.; Courtney, A. D.; Vassilev, A.; Marzilli, L. A.; Cotter, R. J.; Nakatani, Y.; Cole, P. A. J. Biol. Chem. 2000, 275, 21953There is no corresponding record for this reference.
- 186Poux, A. N.; Cebrat, M.; Kim, C. M.; Cole, P. A.; Marmorstein, R. Proc. Natl. Acad. Sci. U.S.A. 2002, 99, 14065There is no corresponding record for this reference.
- 187Wu, J.; Xie, N.; Wu, Z.; Zhang, Y.; Zheng, Y. G. Bioorg. Med. Chem. 2009, 17, 1381There is no corresponding record for this reference.
- 188Victor, M.; Bei, Y.; Gay, F.; Calvo, D.; Mello, C.; Shi, Y. EMBO Rep. 2002, 3, 50There is no corresponding record for this reference.
- 189Huang, Z. Q.; Li, J.; Sachs, L. M.; Cole, P. A.; Wong, J. EMBO J. 2003, 22, 2146There is no corresponding record for this reference.
- 190Costanzo, A.; Merlo, P.; Pediconi, N.; Fulco, M.; Sartorelli, V.; Cole, P. A.; Fontemaggi, G.; Fanciulli, M.; Schiltz, L.; Blandino, G.; Balsano, C.; Levrero, M. Mol. Cell 2002, 9, 175There is no corresponding record for this reference.
- 191Kaehlcke, K.; Dorr, A.; Hetzer-Egger, C.; Kiermer, V.; Henklein, P.; Schnoelzer, M.; Loret, E.; Cole, P. A.; Verdin, E.; Ott, M. Mol. Cell 2003, 12, 167There is no corresponding record for this reference.
- 192Polesskaya, A.; Naguibneva, I.; Fritsch, L.; Duquet, A.; Ait-Si-Ali, S.; Robin, P.; Vervisch, A.; Pritchard, L. L.; Cole, P.; Harel-Bellan, A. EMBO J. 2001, 20, 6816There is no corresponding record for this reference.
- 193Bandyopadhyay, D.; Okan, N. A.; Bales, E.; Nascimento, L.; Cole, P. A.; Medrano, E. E. Cancer Res. 2002, 62, 6231There is no corresponding record for this reference.
- 194Cebrat, M.; Kim, C. M.; Thompson, P. R.; Daugherty, M.; Cole, P. A. Bioorg. Med. Chem. 2003, 11, 3307There is no corresponding record for this reference.
- 195Zheng, Y.; Balasubramanyam, K.; Cebrat, M.; Buck, D.; Guidez, F.; Zelent, A.; Alani, R. M.; Cole, P. A. J. Am. Chem. Soc. 2005, 127, 17182There is no corresponding record for this reference.
- 196Guidez, F.; Howell, L.; Isalan, M.; Cebrat, M.; Alani, R. M.; Ivins, S.; Hormaeche, I.; McConnell, M. J.; Pierce, S.; Cole, P. A.; Licht, J.; Zelent, A. Mol. Cell. Biol. 2005, 25, 5552196Histone acetyltransferase activity of p300 is required for transcriptional repression by the promyelocytic leukemia zinc finger proteinGuidez, Fabien; Howell, Louise; Isalan, Mark; Cebrat, Marek; Alani, Rhoda M.; Ivins, Sarah; Hormaeche, Itsaso; McConnell, Melanie J.; Pierce, Sarah; Cole, Philip A.; Licht, Jonathan; Zelent, ArthurMolecular and Cellular Biology (2005), 25 (13), 5552-5566CODEN: MCEBD4; ISSN:0270-7306. (American Society for Microbiology)Histone acetyltransferase (HAT) activities of proteins such as p300, CBP, and P/CAF play important roles in activation of gene expression. We now show that the HAT activity of p300 can also be required for down-regulation of transcription by a DNA binding repressor protein. Promyelocytic leukemia zinc finger (PLZF), originally identified as a fusion with retinoic acid receptor alpha in rare cases of all-trans-retinoic acid-resistant acute promyelocytic leukemia, is a transcriptional repressor that recruits histone deacetylase-contg. corepressor complexes to specific DNA binding sites. PLZF assocs. with p300 in vivo, and its ability to repress transcription is specifically dependent on HAT activity of p300 and acetylation of lysines in its C-terminal C2-H2 zinc finger motif. An acetylation site mutant of PLZF does not repress transcription and is functionally deficient in a colony suppression assay despite retaining its abilities to interact with corepressor/histone deacetylase complexes. This is due to the fact that acetylation of PLZF activates its ability to bind specific DNA sequences both in vitro and in vivo. Taken together, our results indicate that a histone deacetylase-dependent transcriptional repressor can be pos. regulated through acetylation and point to an unexpected role of a coactivator protein in transcriptional repression.
- 197Balasubramanyam, K.; Swaminathan, V.; Ranganathan, A.; Kundu, T. K. J. Biol. Chem. 2003, 278, 19134There is no corresponding record for this reference.
- 198Rosen, T.; Fordice, D. B. South Med. J. 1994, 87, 543198Cashew nut dermatitisRosen T; Fordice D BSouthern medical journal (1994), 87 (4), 543-6 ISSN:0038-4348.The urushiol dermatitis caused by plants of the Anacardiaceae family is the most common cause of acute allergic contact dermatitis. We have reported a case of cashew nut urushiol dermatitis due to ingestion of homemade cashew nut butter contaminated by cashew nut shell oil. With the precautions taken today to avoid contamination of food products with cashew urushiols, it is rare to find a case of cashew nut dermatitis in the United States. We have found no other report of contact dermatitis due to cashew nut butter. Moreover, though hinted at in the literature, there has been no previous detailed report of perianal contact dermatitis due to cashew ingestion. The fact that our patient was ill enough to require treatment with 3 weeks of systemic steroid therapy highlights the potential public health hazard of consumption of improperly prepared cashew products. However, the risk of cashew nut dermatitis today remains small, and this should not discourage cashew lovers from enjoying their treats. A final lesson to be learned from this case is that perianal eruptions may be due to materials deliberately applied to the anogenital region or to ingested antigens that remain sufficiently intact within the feces to affect perianal skin.
- 199Balasubramanyam, K.; Swaminathan, V.; Ranganathan, A.; Kundu, T. K. J. Biol. Chem. 2003, 278, 19134There is no corresponding record for this reference.
- 200Ghizzoni, M.; Wu, J.; Gao, T.; Haisma, H. J.; Dekker, F. J.; George Zheng, Y. Eur. J. Med. Chem. 2012, 47, 337There is no corresponding record for this reference.
- 201Wu, J.; Xie, N.; Wu, Z.; Zhang, Y.; Zheng, Y. G. Bioorg. Med. Chem. 2009, 17, 1381There is no corresponding record for this reference.
- 202Chandregowda, V.; Kush, A.; Reddy, G. C. Eur. J. Med. Chem. 2009, 44, 2711There is no corresponding record for this reference.
- 203Balasubramanyam, K.; Swaminathan, V.; Ranganathan, A.; Kundu, T. K. J. Biol. Chem. 2003, 278, 19134There is no corresponding record for this reference.
- 204Devipriya, B.; Parameswari, A. R.; Rajalakshmi, G.; Palvannan, T.; Kumaradhas, P. Indian J. Biochem. Biophys. 2010, 47, 364204Exploring the binding affinities of p300 enzyme activators CTPB and CTB using docking methodDevipriya, B.; Parameswari, A. Renuga; Rajalakshmi, G.; Palvannan, T.; Kumaradhas, P.Indian Journal of Biochemistry & Biophysics (2010), 47 (6), 364-369CODEN: IJBBBQ; ISSN:0301-1208. (National Institute of Science Communication and Information Resources)CREB binding protein (CBP) and E1A binding protein p300, also known as p300 are functionally related transcriptional co-activators (CoAs) and histone acetyltransferases (HATs). Some small mols., which target HATs can activate or inhibit the p300 enzyme potently. Here, the binding affinities are reported for two small mols. CTPB [N-(4-chloro-3-trifluoromethyl-phenyl)-2-ethoxy-6-pentadecyl-benzamide] and CTB [N-(4-chloro-3-trifluoromethyl-phenyl)-2-ethoxybenzamide] with p300 using docking method to obtain the insight of their interaction with p300. These small mols. bind to the enzyme, subsequently causing a structural change in the enzyme, which is responsible for the HAT activation. CTB exhibits higher binding affinity than CTPB, and their lowest docked energies are -7.72, -1.18 kcal/mol, resp. In CTPB mol., phenolic hydroxyl of Tyr1397 interacts with the non-polar atoms C(5E) and C(5F), and forms polar-non polar interactions. Similar interactions have also been obsd. in CTB. The residues Tyr1446 and Cys1438 interact with the non-pentadecyl atoms. Further, the docking study predicts a N-H···O hydrogen bonding interaction between CTB and Leu1398, in which the H···O contact distance is 2.06 Å. The long pentadecyl chain of CTPB reduces the formation of hydrogen bond with the p300. The H-bond interaction could be the key factor for the better activation of CTB.
- 205Mantelingu, K.; Kishore, A. H.; Balasubramanyam, K.; Kumar, G. V.; Altaf, M.; Swamy, S. N.; Selvi, R.; Das, C.; Narayana, C.; Rangappa, K. S.; Kundu, T. K. J. Phys. Chem. B 2007, 111, 4527There is no corresponding record for this reference.
- 206Varier, R. A.; Swaminathan, V.; Balasubramanyam, K.; Kundu, T. K. Biochem. Pharmacol. 2004, 68, 1215There is no corresponding record for this reference.
- 207Balasubramanyam, K.; Varier, R. A.; Altaf, M.; Swaminathan, V.; Siddappa, N. B.; Ranga, U.; Kundu, T. K. J. Biol. Chem. 2004, 279, 51163There is no corresponding record for this reference.
- 208Goel, A.; Kunnumakkara, A. B.; Aggarwal, B. B. Biochem. Pharmacol. 2008, 75, 787There is no corresponding record for this reference.
- 209Maheshwari, R. K.; Singh, A. K.; Gaddipati, J.; Srimal, R. C. Life Sci. 2006, 78, 2081There is no corresponding record for this reference.
- 210Aggarwal, B. B.; Shishodia, S. Biochem. Pharmacol. 2006, 71, 1397210Molecular targets of dietary agents for prevention and therapy of cancerAggarwal, Bharat B.; Shishodia, ShishirBiochemical Pharmacology (2006), 71 (10), 1397-1421CODEN: BCPCA6; ISSN:0006-2952. (Elsevier B.V.)A review. While fruits and vegetables are recommended for prevention of cancer and other diseases, their active ingredients (at the mol. level) and their mechanisms of action less well understood. Extensive research during the last half century has identified various mol. targets that can potentially be used not only for the prevention of cancer but also for treatment. However, lack of success with targeted monotherapy resulting from bypass mechanisms has forced researchers to employ either combination therapy or agents that interfere with multiple cell-signaling pathways. In this review, we present evidence that numerous agents identified from fruits and vegetables can interfere with several cell-signaling pathways. The agents include curcumin (turmeric), resveratrol (red grapes, peanuts, and berries), genistein (soybean), diallyl sulfide (allium), S-allyl cysteine (allium), allicin (garlic), lycopene (tomato), capsaicin (red chilli), diosgenin (fenugreek), 6-gingerol (ginger), ellagic acid (pomegranate), ursolic acid (apples, pears, prunes), silymarin (milk thistle), anethol (anise, camphor, and fennel), catechins (green tea), eugenol (cloves), indole-3-carbinol (cruciferous vegetables), limonene (citrus fruits), β-carotene (carrots), and dietary fiber. For instance, the cell-signaling pathways inhibited by curcumin alone include NF-κB, AP-1, STAT3, Akt, Bcl-2, Bcl-XL, caspases, PARP, IKK, EGFR, HER2, JNK, MAPK, COX2, and 5-LOX. The active principle identified in fruit and vegetables and the mol. targets modulated may be the basis for how these dietary agents not only prevent but also treat cancer and other diseases. This work reaffirms what Hippocrates said 25 centuries ago, let food be thy medicine and medicine be thy food.
- 211Zhou, H.; Beevers, C. S.; Huang, S. Curr. Drug Targets 2011, 12, 332211The targets of curcuminZhou, Hongyu; Beevers, Christopher S.; Huang, ShileCurrent Drug Targets (2011), 12 (3), 332-347CODEN: CDTUAU; ISSN:1389-4501. (Bentham Science Publishers Ltd.)A review. Curcumin (diferuloylmethane), an orange-yellow component of turmeric or curry powder, is a polyphenol natural product isolated from the rhizome of the plant Curcuma longa. For centuries, curcumin has been used in some medicinal prepn. or used as a food-coloring agent. In recent years, extensive in vitro and in vivo studies suggested curcumin has anticancer, antiviral, antiarthritic, anti-amyloid, antioxidant, and anti-inflammatory properties. The underlying mechanisms of these effects are diverse and appear to involve the regulation of various mol. targets, including transcription factors (such as nuclear factor-κB), growth factors (such as vascular endothelial cell growth factor), inflammatory cytokines (such as tumor necrosis factor, interleukin 1 and interleukin 6), protein kinases (such as mammalian target of rapamycin, mitogen-activated protein kinases, and Akt) and other enzymes (such as cyclooxygenase 2 and 5 lipoxygenase). Thus, due to its efficacy and regulation of multiple targets, as well as its safety for human use, curcumin has received considerable interest as a potential therapeutic agent for the prevention and/or treatment of various malignant diseases, arthritis, allergies, Alzheimer's disease, and other inflammatory illnesses. This review summarizes various in vitro and in vivo pharmacol. aspects of curcumin as well as the underlying action mechanisms. The recently identified mol. targets and signaling pathways modulated by curcumin are also discussed here.
- 212Han, X.; Xu, B.; Beevers, C. S.; Odaka, Y.; Chen, L.; Liu, L.; Luo, Y.; Zhou, H.; Chen, W.; Shen, T.; Huang, S. Carcinogenesis 2012, 33, 868There is no corresponding record for this reference.
- 213Gupta, S. C.; Prasad, S.; Kim, J. H.; Patchva, S.; Webb, L. J.; Priyadarsini, I. K.; Aggarwal, B. B. Nat. Prod. Rep. 2011, 28, 1937213Multitargeting by curcumin as revealed by molecular interaction studiesGupta, Subash C.; Prasad, Sahdeo; Kim, Ji Hye; Patchva, Sridevi; Webb, Lauren J.; Priyadarsini, Indira K.; Aggarwal, Bharat B.Natural Product Reports (2011), 28 (12), 1937-1955CODEN: NPRRDF; ISSN:0265-0568. (Royal Society of Chemistry)A review. Curcumin (diferuloylmethane), the active ingredient in turmeric (Curcuma longa), is a highly pleiotropic mol. with anti-inflammatory, anti-oxidant, chemopreventive, chemosensitization, and radiosensitization activities. The pleiotropic activities attributed to curcumin come from its complex mol. structure and chem., as well as its ability to influence multiple signaling mols. Curcumin has been shown to bind by multiple forces directly to numerous signaling mols., such as inflammatory mols., cell survival proteins, protein kinases, protein reductases, histone acetyltransferase, histone deacetylase, glyoxalase I, xanthine oxidase, proteasome, HIV1 integrase, HIV1 protease, sarco (endo) plasmic reticulum Ca2+ ATPase, DNA methyltransferases 1, FtsZ protofilaments, carrier proteins, and metal ions. Curcumin can also bind directly to DNA and RNA. Owing to its β-diketone moiety, curcumin undergoes keto-enol tautomerism that has been reported as a favorable state for direct binding. The functional groups on curcumin found suitable for interaction with other macromols. include the α, β-unsatd. β-diketone moiety, carbonyl and enolic groups of the β-diketone moiety, methoxy and phenolic hydroxyl groups, and the Ph rings. Various biophys. tools have been used to monitor direct interaction of curcumin with other proteins, including absorption, fluorescence, Fourier transform IR (FTIR) and CD (CD) spectroscopy, surface plasmon resonance, competitive ligand binding, Forster type fluorescence resonance energy transfer (FRET), radiolabeling, site-directed mutagenesis, matrix-assisted laser desorption/ionization time-of-flight mass spectrometry (MALDI-TOF MS), immunopptn., phage display biopanning, electron microscopy, 1-anilino-8-naphthalene-sulfonate (ANS) displacement, and co-localization. Mol. docking, the most commonly employed computational tool for calcg. binding affinities and predicting binding sites, has also been used to further characterize curcumin's binding sites. Furthermore, the ability of curcumin to bind directly to carrier proteins improves its soly. and bioavailability. In this review, we focus on how curcumin directly targets signaling mols., as well as the different forces that bind the curcumin-protein complex and how this interaction affects the biol. properties of proteins. We will also discuss various analogs of curcumin designed to bind selective targets with increased affinity.
- 214Heery, D. M.; Fischer, P. M. Drug Discovery Today 2007, 12, 88There is no corresponding record for this reference.
- 215Marcu, M. G.; Jung, Y. J.; Lee, S.; Chung, E. J.; Lee, M. J.; Trepel, J.; Neckers, L. Med. Chem. 2006, 2, 169215Curcumin is an inhibitor of p300 histone acetyltransferaseMarcu, Monica G.; Jung, Yun-Jin; Lee, Sunmin; Chung, Eun-Joo; Lee, Min-Jung; Trepel, Jane; Neckers, LenMedicinal Chemistry (2006), 2 (2), 169-174CODEN: MCEHAJ; ISSN:1573-4064. (Bentham Science Publishers Ltd.)Histone acetyltransferases (HATs), and p300/CBP in particular, have been implicated in cancer cell growth and survival, and as such, HATs represent novel, therapeutically relevant mol. targets for drug development. In this study, we demonstrate that the small mol. natural product curcumin, whose medicinal properties have long been recognized in India and Southeast Asia, is a selective HAT inhibitor. Furthermore the data indicate that α, β unsatd. carbonyl groups in the curcumin side chain function as Michael reaction sites and that the Michael reaction acceptor functionality of curcumin is required for its HAT-inhibitory activity. In cells, curcumin promoted proteasome-dependent degrdn. of p300 and the closely related CBP protein without affecting the HATs PCAF or GCN5. In addn. to inducing p300 degrdn. curcumin inhibited the acetyltransferase activity of purified p300 as assessed using either histone H3 or p53 as substrate. Radiolabeled curcumin formed a covalent assocn. with p300, and tetrahydrocurcumin displayed no p300 inhibitory activity, consistent with a Michael reaction-dependent mechanism. Finally, curcumin was able to effectively block histone hyperacetylation in both PC3-M prostate cancer cells and peripheral blood lymphocytes induced by the histone deacetylase inhibitor MS-275. These data thus identify the medicinal natural product curcumin as a novel lead compd. for development of possibly therapeutic, p300/CBP-specific HAT inhibitors.
- 216Balasubramanyam, K.; Varier, R. A.; Altaf, M.; Swaminathan, V.; Siddappa, N. B.; Ranga, U.; Kundu, T. K. J. Biol. Chem. 2004, 279, 51163There is no corresponding record for this reference.
- 217Anand, P.; Kunnumakkara, A. B.; Newman, R. A.; Aggarwal, B. B. Mol. Pharmaceutics 2007, 4, 807217Bioavailability of Curcumin: Problems and PromisesAnand, Preetha; Kunnumakkara, Ajaikumar B.; Newman, Robert A.; Aggarwal, Bharat B.Molecular Pharmaceutics (2007), 4 (6), 807-818CODEN: MPOHBP; ISSN:1543-8384. (American Chemical Society)A review. Curcumin, a polyphenolic compd. derived from dietary spice turmeric, possesses diverse pharmacol. effects including anti-inflammatory, antioxidant, antiproliferative and antiangiogenic activities. Phase I clin. trials have shown that curcumin is safe even at high doses (12 g/day) in humans but exhibit poor bioavailability. Major reasons contributing to the low plasma and tissue levels of curcumin appear to be due to poor absorption, rapid metab., and rapid systemic elimination. To improve the bioavailability of curcumin, numerous approaches have been undertaken. These approaches involve, first, the use of adjuvant like piperine that interferes with glucuronidation; second, the use of liposomal curcumin; third, curcumin nanoparticles; fourth, the use of curcumin phospholipid complex; and fifth, the use of structural analogs of curcumin (e.g., EF-24). The latter has been reported to have a rapid absorption with a peak plasma half-life. Despite the lower bioavailability, therapeutic efficacy of curcumin against various human diseases, including cancer, cardiovascular diseases, diabetes, arthritis, neurol. diseases and Crohn's disease, has been documented. Enhanced bioavailability of curcumin in the near future is likely to bring this promising natural product to the forefront of therapeutic agents for treatment of human disease.
- 218Sharma, R. A.; Euden, S. A.; Platton, S. L.; Cooke, D. N.; Shafayat, A.; Hewitt, H. R.; Marczylo, T. H.; Morgan, B.; Hemingway, D.; Plummer, S. M.; Pirmohamed, M.; Gescher, A. J.; Steward, W. P. Clin. Cancer Res. 2004, 10, 6847218Phase I Clinical Trial of Oral Curcumin: Biomarkers of Systemic Activity and ComplianceSharma, Ricky A.; Euden, Stephanie A.; Platton, Sharon L.; Cooke, Darren N.; Shafayat, Aisha; Hewitt, Heather R.; Marczylo, Timothy H.; Morgan, Bruno; Hemingway, David; Plummer, Simon M.; Pirmohamed, Munir; Gescher, Andreas J.; Steward, William P.Clinical Cancer Research (2004), 10 (20), 6847-6854CODEN: CCREF4; ISSN:1078-0432. (American Association for Cancer Research)Curcumin, a polyphenolic antioxidant derived from a dietary spice, exhibits anticancer activity in rodents and in humans. Its efficacy appears to be related to induction of glutathione S-transferase enzymes, inhibition of prostaglandin E2 (PGE2) prodn., or suppression of oxidative DNA adduct (M1G) formation. We designed a dose-escalation study to explore the pharmacol. of curcumin in humans. Fifteen patients with advanced colorectal cancer refractory to std. chemotherapies consumed capsules compatible with curcumin doses between 0.45 and 3.6 g daily for up to 4 mo. Levels of curcumin and its metabolites in plasma, urine, and feces were analyzed by high-pressure liq. chromatog. and mass spectrometry. Three biomarkers of the potential activity of curcumin were translated from preclin. models and measured in patient blood leukocytes: glutathione S-transferase activity, levels of M1G, and PGE2 prodn. induced ex vivo. Dose-limiting toxicity was not obsd. Curcumin and its glucuronide and sulfate metabolites were detected in plasma in the 10 nmol/L range and in urine. A daily dose of 3.6 g curcumin engendered 62% and 57% decreases in inducible PGE2 prodn. in blood samples taken 1 h after dose on days 1 and 29, resp., of treatment compared with levels obsd. immediately predose (P < 0.05). A daily oral dose of 3.6 g of curcumin is advocated for Phase II evaluation in the prevention or treatment of cancers outside the gastrointestinal tract. PGE2 prodn. in blood and target tissue may indicate biol. activity. Levels of curcumin and its metabolites in the urine can be used to assess general compliance.
- 219Mishra, S.; Kapoor, N.; Mubarak Ali, A.; Pardhasaradhi, B. V.; Kumari, A. L.; Khar, A.; Misra, K. Free Radical Biol. Med. 2005, 38, 1353219Differential apoptotic and redox regulatory activities of curcumin and its derivativesMishra, Satyendra; Kapoor, Neha; Mubarak Ali, A.; Pardhasaradhi, B. V. V.; Kumari, A. Leela; Khar, Ashok; Misra, KrishnaFree Radical Biology & Medicine (2005), 38 (10), 1353-1360CODEN: FRBMEH; ISSN:0891-5849. (Elsevier)We have synthesized different bioconjugates of curcumin, which were tested for their pro- and antioxidant properties. In the present study five representative derivs. of curcumin, i.e., 4,4'-di-(O-acetyl) curcumin, 4,4'-di-(O-glycinoyl) curcumin, 4,4'-di-(O-glycinoyl-di-N-piperoyl) curcumin, 4,4'-di-(O-piperoyl) curcumin, and 4,4'-(O,O-cystinoyl)-3,3'-dimethoxydiphenyl-1,6-heptadiene-3,5-dione, were used for testing their apoptotic potential on tumor cells. Dipiperoyl and diglycinoyl derivs. showed higher apoptotic activity at lower concns., whereas diacetyl curcumin had slightly lower apoptotic activity on tumor cells. On the other hand, diglycinoyl-dipiperoyl and cystinoyl heptadiene derivs. had lost their apoptotic potential significantly. The apoptotic activity of these derivs. correlated very well with the generation of ROS by the tumor cells, whereas GSH levels remained unaltered. Our studies also indicate down-regulation of Bcl-2 and participation of caspase-3 in the apoptotic death of tumor cells.
- 220Weber, W. M.; Hunsaker, L. A.; Abcouwer, S. F.; Deck, L. M.; Vander Jagt, D. L. Bioorg. Med. Chem. 2005, 13, 3811There is no corresponding record for this reference.
- 221Varier, R. A.; Swaminathan, V.; Balasubramanyam, K.; Kundu, T. K. Biochem. Pharmacol. 2004, 68, 1215There is no corresponding record for this reference.
- 222Balasubramanyam, K.; Altaf, M.; Varier, R. A.; Swaminathan, V.; Ravindran, A.; Sadhale, P. P.; Kundu, T. K. J. Biol. Chem. 2004, 279, 33716There is no corresponding record for this reference.
- 223Balasubramanyam, K.; Altaf, M.; Varier, R. A.; Swaminathan, V.; Ravindran, A.; Sadhale, P. P.; Kundu, T. K. J. Biol. Chem. 2004, 279, 33716There is no corresponding record for this reference.
- 224Mantelingu, K.; Reddy, B. A.; Swaminathan, V.; Kishore, A. H.; Siddappa, N. B.; Kumar, G. V.; Nagashankar, G.; Natesh, N.; Roy, S.; Sadhale, P. P.; Ranga, U.; Narayana, C.; Kundu, T. K. Chem. Biol. 2007, 14, 645224Specific Inhibition of p300-HAT Alters Global Gene Expression and Represses HIV ReplicationMantelingu, K.; Reddy, B. A. Ashok; Swaminathan, V.; Kishore, A. Hari; Siddappa, Nagadenahalli B.; Kumar, G. V. Pavan; Nagashankar, G.; Natesh, Nagashayana; Roy, Siddhartha; Sadhale, Parag P.; Ranga, Udaykumar; Narayana, Chandrabhas; Kundu, Tapas K.Chemistry & Biology (Cambridge, MA, United States) (2007), 14 (6), 645-657CODEN: CBOLE2; ISSN:1074-5521. (Cell Press)Summary: Reversible acetylation of histone and nonhistone proteins plays pivotal role in cellular homeostasis. Dysfunction of histone acetyltransferases (HATs) leads to several diseases including cancer, neurodegenaration, asthma, diabetes, AIDS, and cardiac hypertrophy. We describe the synthesis and characterization of a set of p300-HAT-specific small-mol. inhibitors from a natural nonspecific HAT inhibitor, garcinol, which is highly toxic to cells. We show that the specific inhibitor selectively represses the p300-mediated acetylation of p53 in vivo. Furthermore, inhibition of p300-HAT down regulates several genes but significantly a few important genes are also upregulated. Remarkably, these inhibitors were found to be nontoxic to T cells, inhibit histone acetylation of HIV infected cells, and consequently inhibit the multiplication of HIV.
- 225Arif, M.; Pradhan, S. K.; Thanuja, G. R.; Vedamurthy, B. M.; Agrawal, S.; Dasgupta, D.; Kundu, T. K. J. Med. Chem. 2009, 52, 267There is no corresponding record for this reference.
- 226Baggett, S.; Protiva, P.; Mazzola, E. P.; Yang, H.; Ressler, E. T.; Basile, M. J.; Weinstein, I. B.; Kennelly, E. J. J. Nat. Prod. 2005, 68, 354There is no corresponding record for this reference.
- 227Gartner, M.; Müller, T.; Simon, J. C.; Giannis, A.; Sleeman, J. P. ChemBioChem 2005, 6, 171There is no corresponding record for this reference.
- 228Dal Piaz, F.; Tosco, A.; Eletto, D.; Piccinelli, A. L.; Moltedo, O.; Franceschelli, S.; Sbardella, G.; Remondelli, P.; Rastrelli, L.; Vesci, L.; Pisano, C.; De Tommasi, N. ChemBioChem 2010, 11, 818There is no corresponding record for this reference.
- 229Biel, M.; Kretsovali, A.; Karatzali, E.; Papamatheakis, J.; Giannis, A. Angew. Chem., Int. Ed. 2004, 43, 3974There is no corresponding record for this reference.
- 230Heery, D. M.; Fischer, P. M. Drug Discovery Today 2007, 12, 88There is no corresponding record for this reference.
- 231Ravindra, K. C.; Selvi, B. R.; Arif, M.; Reddy, B. A.; Thanuja, G. R.; Agrawal, S.; Pradhan, S. K.; Nagashayana, N.; Dasgupta, D.; Kundu, T. K. J. Biol. Chem. 2009, 284, 24453There is no corresponding record for this reference.
- 232Sandur, S. K.; Ichikawa, H.; Sethi, G.; Ahn, K. S.; Aggarwal, B. B. J. Biol. Chem. 2006, 281, 17023There is no corresponding record for this reference.
- 233Padhye, S.; Dandawate, P.; Yusufi, M.; Ahmad, A.; Sarkar, F. H. Med. Res. Rev. 2012, 32, 1131233Perspectives on medicinal properties of plumbagin and its analogsPadhye, Subhash; Dandawate, Prasad; Yusufi, Mujahid; Ahmad, Aamir; Sarkar, Fazlul H.Medicinal Research Reviews (2012), 32 (6), 1131-1158CODEN: MRREDD; ISSN:0198-6325. (John Wiley & Sons, Inc.)A review. Plumbagin is one of the simplest plant secondary metabolite of three major phylogenic families viz. Plumbaginaceae, Droseraceae, and Ebenceae, and exhibits highly potent biol. activities, including antioxidant, antiinflammatory, anticancer, antibacterial, and antifungal activities. Recent investigations indicate that these activities arise mainly out of its ability to undergo redox cycling, generating reactive oxygen species and chelating trace metals in biol. system. The compd. is endowed with a property to inhibit the drug efflux mechanism in drug-resistant bacteria, thereby allowing intracellular accumulation of the potent drug mols. An interesting bioactivity exhibited by this compd. is the elimination of stringent, conjugative, multidrug-resistant plasmids from several bacterial strains including opportunistic bacteria, such as Acinetobacter baumannii. Moreover, plumbagin effectively induces apoptosis and causes cell cycle arrest, which is, in part, due to the inactivation of NF-κB in cancer cells. Therefore, it has been suggested that designing "hybrid drug mols." of plumbagin by combining it with other appropriate anticancer agents may lead to the generation of novel and potent anticancer drugs with pleiotropic action against human cancers. This comprehensive review is an attempt to understand the chem. of plumbagin and catalog its biol. activities reported to date.
- 234Ravindra, K. C.; Selvi, B. R.; Arif, M.; Reddy, B. A.; Thanuja, G. R.; Agrawal, S.; Pradhan, S. K.; Nagashayana, N.; Dasgupta, D.; Kundu, T. K. J. Biol. Chem. 2009, 284, 24453There is no corresponding record for this reference.
- 235Choi, K. C.; Jung, M. G.; Lee, Y. H.; Yoon, J. C.; Kwon, S. H.; Kang, H. B.; Kim, M. J.; Cha, J. H.; Kim, Y. J.; Jun, W. J.; Lee, J. M.; Yoon, H. G. Cancer Res. 2009, 69, 583There is no corresponding record for this reference.
- 236Neukam, K.; Pastor, N.; Cortés, F. Mutat. Res. 2008, 654, 8236Tea flavanols inhibit cell growth and DNA topoisomerase II activity and induce endoreduplication in cultured Chinese hamster cellsNeukam, Karin; Pastor, Nuria; Cortes, FelipeMutation Research, Genetic Toxicology and Environmental Mutagenesis (2008), 654 (1), 8-12CODEN: MRGMFI; ISSN:1383-5718. (Elsevier B.V.)Tea polyphenols are promising chemopreventive anticancer agents, the properties of which have been studied both in vitro and in vivo, providing evidence that - within this group of compds. - the tea flavanols are able to inhibit carcinogenesis, an effect that in some cases could be correlated with increased cell apoptosis and decreased cell proliferation. Of four main tea flavanols, namely (-)-epigallocatechin-3-gallate (EGCG), (-)-epigallocatechin (EGC), (+)-catechin (CA) and (-)-epicatechin (EC), it was found that EGCG was the most potent to inhibit dose dependently the topoisomerase II (TOPO II) catalytic activity isolated from hamster ovary AA8 cells. In the range of concns. that caused TOPO II inhibition, a high level of endoreduplication, a rare phenomenon that consists in two successive rounds of DNA replication without intervening mitosis, was obsd., while neither micronuclei nor DNA strand breaks (Comet assay) were detected at the same doses. We propose that the anticarcinogenic effect of tea flavanols can be partly explained by their potency and effectiveness to induce endoreduplication. Concerning such an induction, max. effect seems to require a pyrogallol structure at the B-ring. Addnl. substitution with a galloylic residue at the C3 hydroxyl group leads to further augmentation of the effect. Thus, we suggest that the chemopreventive properties of tea flavanols can be at least partly due to their ability to interfere with the cell cycle and block cell proliferation at early stages of mitosis.
- 237Berger, S. J.; Gupta, S.; Belfi, C. A.; Gosky, D. M.; Mukhtar, H. Biochem. Biophys. Res. Commun. 2001, 288, 101There is no corresponding record for this reference.
- 238Suzuki, K.; Yahara, S.; Hashimoto, F.; Uyeda, M. Biol. Pharm. Bull. 2001, 24, 1088238Inhibitory activities of (-)-epigallocatechin-3-O-gallate against topoisomerases I and IISuzuki, Keitarou; Yahara, Shoji; Hashimoto, Furnio; Uyeda, MasaruBiological & Pharmaceutical Bulletin (2001), 24 (9), 1088-1090CODEN: BPBLEO; ISSN:0918-6158. (Pharmaceutical Society of Japan)The substitution of gallic acid at the 3 position of (-)-epigallocatechin-3-O-gallate (EGCG) increased the inhibition against topoisomerase I from calf thymus gland and topoisomerase II from human placenta, and the substitution of a hydroxyl group at the 3' position increased the inhibition against the topoisomerase I. These results suggested that the 3 and 3' positions of the EGCG mol. play important roles in the process of inhibition of topoisomerases I and II. EGCG showed strong inhibition against topoisomerases I from wheat germ, calf thymus gland, and Vero cells, and showed weak or no inhibition against topoisomerases I from carcinoma cells such as A549, HeLaand COLO 201 cells. EGCG differentially inhibited the topoisomerases I from different sources.
- 239Bandele, O. J.; Osheroff, N. Chem. Res. Toxicol. 2008, 21, 936There is no corresponding record for this reference.
- 240Golden, E. B.; Lam, P. Y.; Kardosh, A.; Gaffney, K. J.; Cadenas, E.; Louie, S. G.; Petasis, N. A.; Chen, T. C.; Schönthal, A. H. Blood 2009, 113, 5927There is no corresponding record for this reference.
- 241Ge, J.; Tan, B. X.; Chen, Y.; Yang, L.; Peng, X. C.; Li, H. Z.; Lin, H. J.; Zhao, Y.; Wei, M.; Cheng, K.; Li, L. H.; Dong, H.; Gao, F.; He, J. P.; Wu, Y.; Qiu, M.; Zhao, Y. L.; Su, J. M.; Hou, J. M.; Liu, J. Y. J. Mol. Med. 2011, 89, 595241Interaction of green tea polyphenol epigallocatechin-3-gallate with sunitinib: potential risk of diminished sunitinib bioavailabilityGe, Jun; Tan, Ben-Xu; Chen, Ye; Yang, Li; Peng, Xing-Chen; Li, Hong-Ze; Lin, Hong-Jun; Zhao, Yu; Wei, Meng; Cheng, Ke; Li, Long-Hao; Dong, Hang; Gao, Feng; He, Jian-Ping; Wu, Yang; Qiu, Meng; Zhao, Ying-Lan; Su, Jing-Mei; Hou, Jian-Mei; Liu, Ji-YanJournal of Molecular Medicine (Heidelberg, Germany) (2011), 89 (6), 595-602CODEN: JMLME8; ISSN:0946-2716. (Springer)Sunitinib, a novel oral multi-targeted tyrosine kinase inhibitor for patients with metastatic renal cell carcinoma (mRCC) and advanced gastrointestinal stromal tumor, has a good prospect for clin. application and is being investigated for the potential therapy of other tumors. We obsd. the phenomenon that drinking tea interfered with symptom control in an mRCC patient treated with sunitinib and speculated that green tea or its components might interact with sunitinib. This study was performed to investigate whether epigallocatechin-3-gallate (EGCG), the major constituent of green tea, interacted with sunitinib. The interaction between EGCG and sunitinib was examd. in vitro and in vivo. 1H NMR (1H-NMR) spectroscopy and mass spectrometry (MS) were used to analyze the interaction between these two mols. and whether a new compd. was formed. Solns. of sunitinib and EGCG were intragastrically administered to rats to investigate whether the plasma concns. of sunitinib were affected by EGCG. In this study, we noticed that a ppt. was formed when the solns. of sunitinib and EGCG were mixed under both neutral and acidic conditions. 1H-NMR spectra indicated an interaction between EGCG and sunitinib, but no new compd. was obsd. by MS. Sticky semisolid contents were found in the stomachs of sunitinib and EGCG co-administrated mice. The and C max of plasma sunitinib were markedly reduced by co-administration of EGCG to rats. Our study firstly showed that EGCG interacted with sunitinib and reduced the bioavailability of sunitinib. This finding has significant practical implications for tea-drinking habit during sunitinib administration.
- 242Strick, R.; Strissel, P. L.; Borgers, S.; Smith, S. L.; Rowley, J. D. Proc. Natl. Acad. Sci. U.S.A. 2000, 97, 4790There is no corresponding record for this reference.
- 243Ruiz, P. A.; Braune, A.; Hölzlwimmer, G.; Quintanilla-Fend, L.; Haller, D. J. Nutr. 2007, 137, 1208243Quercetin inhibits TNF-induced NF-κB transcription factor recruitment to proinflammatory gene promoters in murine intestinal epithelial cellsRuiz, Pedro A.; Braune, Annett; Hoelzlwimmer, Gabriele; Quintanilla-Fend, Leticia; Haller, DirkJournal of Nutrition (2007), 137 (5), 1208-1215CODEN: JONUAI; ISSN:0022-3166. (American Society for Nutrition)Flavonoids may play an important role for adjunct nutritional therapy of chronic intestinal inflammation. In this study, we characterized the mol. mechanisms by which quercetin and its enteric bacterial metabolites, taxifolin, alphitonin, and 3, 4-dihydroxy-phenylacetic acid, inhibit tumor necrosis factor α (TNF)-induced proinflammatory gene expression in the murine small intestinal epithelial cell (IEC) line Mode-K as well as in heterozygous TNFΔARE/WT mice, a murine model of exptl. ileitis. Quercetin inhibited TNF-induced interferon-γ-inducible protein 10 (IP-10) and macrophage inflammatory protein 2 (MIP-2) gene expression in Mode-K cells with effective inhibitory concn. of 40 and 44 μmol/L, resp. Interestingly, taxifolin, alphitonin, and 3,4-dihydroxy-phenylacetic acid did not inhibit TNF responses in IEC, suggesting that microbial transformation of quercetin completely abolished its anti-inflammatory effect. At the mol. level, quercetin inhibited Akt phosphorylation but did not inhibit TNF-induced RelA/I-κB phosphorylation and IκB degrdn. or TNF-α-induced nuclear factor-κB transcriptional activity. Most important for understanding the mechanism involved, chromatin immunopptn. anal. revealed inhibitory effects of quercetin on phospho-RelA recruitment to the IP-10 and MIP-2 gene promoters. In addn., and consistent with the lack of cAMP response element binding protein (CBP)/p300 recruitment and phosphorylation/acetylation of histone 3 at the promoter binding site, quercetin inhibited histone acetyl transferase activity. The oral application of quercetin to heterozygous TNFΔARE/WT mice [10 mg/(d × kg body wt)] significantly inhibited IP-10 and MIP-2 gene expression in primary ileal epithelial cells but did not affect tissue pathol. These studies support an anti-inflammatory effect of quercetin in epithelial cells through mechanisms that inhibit cofactor recruitment at the chromatin of proinflammatory genes.
- 244Chen, J.; Han, J.; Wang, J. Toxicol. Ind. Health 2013, 29, 360244Prevention of cytotoxicity of nickel by quercetin: the role of reactive oxygen species and histone acetylationChen, Jie; Han, Jia; Wang, JianminToxicology and Industrial Health (2013), 29 (4), 360-366CODEN: TIHEEC; ISSN:0748-2337. (Sage Publications)Excessive exposure to nickel may cause health effects on the blood, lung, nose, kidney, reproductive system, skin and the unborn child. In the present study, we found that Ni2+ exposure led to a time- and dosedependent proliferation arrest and death in human leukemia HL-60 cells. In the presence of 1 μM Ni2+, reactive oxygen species (ROS) generation (indicated by the level of malondialdehyde) increased to 323% and histone acetylation decreased to 32%. Interestingly, quercetin (QU) dose dependently prevented Ni2+-induced cell proliferation arrest and death from 0 to 80 μM but showed similar activity of scavenging ROS at the concns. of 20, 40 and 80 μM. When the effect of QU on histone acetylation was studied, QU significantly prevented Ni2+-induced histone hypoacetylation at 40 or 80 μM. Moreover, increase in histone acetylation by trichostatin A could also significantly enhance the protection effect of QU at 10 or 20 μM but not at higher concns. Thus, our results further confirmed the crit. role of ROS and histone hypoacetylation in the cytotoxicity of Ni2+ exposure and proved that QU is a potentially useful native dietary compd. to efficiently prevent Ni2+-caused cytotoxicity through both diminishing ROS generation and increasing histone acetylation.
- 245Si, D.; Wang, Y.; Zhou, Y. H.; Guo, Y.; Wang, J.; Zhou, H.; Li, Z. S.; Fawcett, J. P. Drug Metab. Dispos. 2009, 37, 629There is no corresponding record for this reference.
- 246Saaby, L.; Rasmussen, H. B.; Jäger, A. K. J. Ethnopharmacol. 2009, 121, 178There is no corresponding record for this reference.
- 247Hilliard, J. J.; Krause, H. M.; Bernstein, J. I.; Fernandez, J. A.; Nguyen, V.; Ohemeng, K. A.; Barrett, J. F. Adv. Exp. Med. Biol. 1995, 390, 59There is no corresponding record for this reference.
- 248Xiao, X.; Shi, D.; Liu, L.; Wang, J.; Xie, X.; Kang, T.; Deng, W. PLoS One 2011, 6, e22934There is no corresponding record for this reference.
- 249O’Brien, E.; Dietrich, D. R. Crit. Rev. Toxicol. 2005, 35, 33249Ochratoxin A: The Continuing EnigmaO'Brien, Evelyn; Dietrich, DanielCritical Reviews in Toxicology (2005), 35 (1), 33-60CODEN: CRTXB2; ISSN:1040-8444. (Taylor & Francis, Inc.)A review. The mycotoxin ochratoxin A (OTA) has been linked to the genesis of several disease states in both animals and humans. It has been described as nephrotoxic, carcinogenic, teratogenic, immunotoxic, and hepatotoxic in lab. and domestic animals, as well as being thought to be the probable causal agent in the development of nephropathies (Balkan Endemic Nephropathy, BEN and Chronic Interstitial Nephropathy, CIN) and urothelial tumors in humans. As a result, several international agencies are currently attempting to define safe legal limits for OTA concn. in foodstuffs (e.g., grain, meat, wine, and coffee), in processed foods, and in animal fodder. In order to achieve this goal, an accurate risk assessment of OTA toxicity including mechanistic and epidemiol. studies must be carried out. Ochratoxin has been suggested by various researchers to mediate its toxic effects via induction of apoptosis, disruption of mitochondrial respiration and/or the cytoskeleton, or, indeed, via the generation of DNA adducts. Thus, it is still unclear if the predominant mechanism is of a genotoxic or an epigenetic nature. One aspect that is clear, however, is that the toxicity of OTA is subject to and characterized by large species- and sex-specific differences, as well as an apparently strict structure-activity relationship. These considerations could be crucial in the investigation of OTA-mediated toxicity. Furthermore, the use of appropriate in vivo and in vitro model systems appears to be vital in the generation of relevant exptl. data. The intention of this review is to collate and discuss the currently available data on OTA-mediated toxicity with particular focus on their relevance for the in vivo situation, and also to suggest possible future strategies for unlocking the secrets of ochratoxin A.
- 250Czakai, K.; Müller, K.; Mosesso, P.; Pepe, G.; Schulze, M.; Gohla, A.; Patnaik, D.; Dekant, W.; Higgins, J. M.; Mally, A. Toxicol. Sci. 2011, 122, 317250Perturbation of Mitosis through Inhibition of Histone Acetyltransferases: The Key to Ochratoxin A Toxicity and Carcinogenicity?Czakai, Kristin; Mueller, Katja; Mosesso, Pasquale; Pepe, Gaetano; Schulze, Markus; Gohla, Antje; Patnaik, Debasis; Dekant, Wolfgang; Higgins, Jonathan M. G.; Mally, AngelaToxicological Sciences (2011), 122 (2), 317-329CODEN: TOSCF2; ISSN:1096-0929. (Oxford University Press)Ochratoxin A (OTA) is one of the most potent rodent renal carcinogens studied to date. Although controversial results regarding OTA genotoxicity have been published, it is now widely accepted that OTA is not a mutagenic, DNA-reactive carcinogen. Instead, increasing evidence from both in vivo and in vitro studies suggests that OTA may promote genomic instability and tumorigenesis through interference with cell division. The aim of the present study was to provide further support for disruption of mitosis as a key event in OTA toxicity and to understand how OTA mediates these effects. Immortalized human kidney epithelial cells (IHKE) were treated with OTA and monitored by differential interference contrast microscopy for 15 h. Image anal. confirmed that OTA at concns. ≥ 5 μM, which correlate with plasma concns. in rats under conditions of carcinogenesis, causes sustained mitotic arrest and exit from mitosis without nuclear or cellular division. Mitotic chromosomes were characterized by aberrant condensation and premature sister chromatid sepn. assocd. with altered phosphorylation and acetylation of core histones. To test if OTA directly interferes with histone acetyltransferases (HATs) which regulate lysine acetylation of histones and nonhistone proteins, a cell-free HAT activity assay was conducted using total nuclear exts. of IHKE cells. In this assay, OTA significantly blocked HAT activity in a concn.-dependent manner Overall, results from this study provide further support for a mechanism of OTA carcinogenicity involving interference with the mitotic machinery and suggest HATs as a primary cellular target of OTA.
- 251Ravindra, K. C.; Narayan, V.; Lushington, G. H.; Peterson, B. R.; Prabhu, K. S. Chem. Res. Toxicol. 2012, 25, 337There is no corresponding record for this reference.
- 252Reginato, M. J.; Krakow, S. L.; Bailey, S. T.; Lazar, M. A. J. Biol. Chem. 1998, 273, 1855There is no corresponding record for this reference.
- 253Ravindra, K. C.; Narayan, V.; Lushington, G. H.; Peterson, B. R.; Prabhu, K. S. Chem. Res. Toxicol. 2012, 25, 337There is no corresponding record for this reference.
- 254Stimson, L.; Rowlands, M. G.; Newbatt, Y. M.; Smith, N. F.; Raynaud, F. I.; Rogers, P.; Bavetsias, V.; Gorsuch, S.; Jarman, M.; Bannister, A.; Kouzarides, T.; McDonald, E.; Workman, P.; Aherne, G. W. Mol. Cancer Ther. 2005, 4, 1521254Isothiazolones as inhibitors of PCAF and p300 histone acetyltransferase activityStimson, Lindsay; Rowlands, Martin G.; Newbatt, Yvette M.; Smith, Nicola F.; Raynaud, Florence I.; Rogers, Paul; Bavetsias, Vassilios; Gorsuch, Stephen; Jarman, Michael; Bannister, Andrew; Kouzarides, Tony; McDonald, Edward; Workman, Paul; Aherne, G. WynneMolecular Cancer Therapeutics (2005), 4 (10), 1521-1532CODEN: MCTOCF; ISSN:1535-7163. (American Association for Cancer Research)Histone acetylation plays an important role in regulating the chromatin structure and is tightly regulated by two classes of enzyme, histone acetyltransferases (HAT) and histone deacetylases (HDAC). Deregulated HAT and HDAC activity plays a role in the development of a range of cancers. Consequently, inhibitors of these enzymes have potential as anticancer agents. Several HDAC inhibitors have been described; however, few inhibitors of HATs have been disclosed. Following a FlashPlate high-throughput screen, we identified a series of isothiazolone-based HAT inhibitors. Thirty-five N-substituted analogs inhibited both p300/cAMP-responsive element binding protein-binding protein-assocd. factor (PCAF) and p300 (1 to > 50 μmol/L, resp.) and the growth of a panel of human tumor cell lines (50% growth inhibition, 0.8 to > 50 μmol/L). CCT077791 and CCT077792 decreased cellular acetylation in a time-dependent manner (2-48 h of exposure) and a concn.-dependent manner (one to five times, 72 h, 50% growth inhibition) in HCT116 and HT29 human colon tumor cell lines. CCT077791 reduced total acetylation of histones H3 and H4, levels of specific acetylated lysine marks, and acetylation of α-tubulin. Four and 24 h of exposure to the compds. produced the same extent of growth inhibition as 72 h of continuous exposure, suggesting that growth arrest was an early event. Chem. reactivity of these compds., as measured by covalent protein binding and loss of HAT inhibition in the presence of DTT, indicated that reaction with thiol groups might be important in their mechanism of action. As one of the first series of small-mol. inhibitors of HAT activity, further analog synthesis is being pursued to examine the potential scope for reducing chem. reactivity while maintaining HAT inhibition.
- 255Dekker, F. J.; Ghizzoni, M.; van der Meer, N.; Wisastra, R.; Haisma, H. J. Bioorg. Med. Chem. 2009, 17, 460There is no corresponding record for this reference.
- 256Gorsuch, S.; Bavetsias, V.; Rowlands, M. G.; Aherne, G. W.; Workman, P.; Jarman, M.; McDonald, E. Bioorg. Med. Chem. 2009, 17, 467There is no corresponding record for this reference.
- 257Gorsuch, S.; Bavetsias, V.; Rowlands, M. G.; Aherne, G. W.; Workman, P.; Jarman, M.; McDonald, E. Bioorg. Med. Chem. 2009, 17, 467There is no corresponding record for this reference.
- 258Wisastra, R.; Ghizzoni, M.; Maarsingh, H.; Minnaard, A. J.; Haisma, H. J.; Dekker, F. J. Org. Biomol. Chem. 2011, 9, 1817There is no corresponding record for this reference.
- 259Dekker, F. J.; Ghizzoni, M.; van der Meer, N.; Wisastra, R.; Haisma, H. J. Bioorg. Med. Chem. 2009, 17, 460There is no corresponding record for this reference.
- 260Ornaghi, P.; Rotili, D.; Sbardella, G.; Mai, A.; Filetici, P. Biochem. Pharmacol. 2005, 70, 911There is no corresponding record for this reference.
- 261Smith, A. T.; Livingston, M. R.; Mai, A.; Filetici, P.; Queener, S. F.; Sullivan, W. J., Jr. Antimicrob. Agents Chemother. 2007, 51, 1109There is no corresponding record for this reference.
- 262Mai, A.; Rotili, D.; Tarantino, D.; Ornaghi, P.; Tosi, F.; Vicidomini, C.; Sbardella, G.; Nebbioso, A.; Miceli, M.; Altucci, L.; Filetici, P. J. Med. Chem. 2006, 49, 6897There is no corresponding record for this reference.
- 263Mai, A.; Rotili, D.; Tarantino, D.; Nebbioso, A.; Castellano, S.; Sbardella, G.; Tini, M.; Altucci, L. Bioorg. Med. Chem. Lett. 2009, 19, 1132There is no corresponding record for this reference.
- 264Rahim, R.; Strobl, J. S. Anticancer Drugs 2009, 20, 736There is no corresponding record for this reference.
- 265Tahan, F.; Jazrawi, E.; Moodley, T.; Rovati, G. E.; Adcock, I. M. Clin. Exp. Allergy 2008, 38, 805265Montelukast inhibits tumour necrosis factor-α-mediated interleukin-8 expression through inhibition of nuclear factor-κB p65-associated histone acetyltransferase activityTahan, F.; Jazrawi, E.; Moodley, T.; Rovati, G. E.; Adcock, I. M.Clinical and Experimental Allergy (2008), 38 (5), 805-811CODEN: CLEAEN; ISSN:0954-7894. (Blackwell Publishing Ltd.)Background: Montelukast is a potent cysteinyl leukotriene-1 receptor antagonist possessing some anti-inflammatory effects although the mol. mechanism of these anti-inflammatory effects is unknown. In this study, we aimed to investigate the effect of montelukast on nuclear factor (NF)-κB-assocd. histone acetylation activity in phorbol myristate acetate (PMA)-differentiated U937 cells. Methods: We examd. the inhibitory effects of montelukast on TNF-α-induced IL-8 prodn. in PMA-differentiated U-937 cells. U-937 cells were exposed to PMA (50 ng/mL) for 48 h to allow differentiation to macrophages. Macrophages were then exposed to TNF-α (10 ng/mL) in the presence or absence of montelukast (0.01-10 μM) for 24 h. After this time, the concn. of IL-8 in the culture supernatant was measured by sandwich-type ELISA kit. The effect of signalling pathways on TNF-α-induced IL-8 release was examd. pharmacol. using selective NF-κB/IKK2 (AS602868, 3 μM), (PD98059, 10 μM) and p38 mitogen activated protein kinase (MAPK) (SB203580, 1 μM) inhibitors. NF-κB DNA binding activity was measured by a DNA-binding ELISA-based assay. NF-κB-p65-assocd. histone acetyltransferase (HAT) activity was measured by immunopptn. linked to com. fluorescent HAT. Results: TNF-α-induced IL-8 release was suppressed by an NF-κB inhibitor but not by MEK or p38 MAPK inhibitors. Montelukast induced a concn.-dependent inhibition of TNF-α-induced IL-8 release and mRNA expression that reached a plateau at 0.1 μM without affecting cell viability. Montelukast did not affect NF-κB p65 activation as measured by DNA binding but suppressed NF-κB p65-assocd. HAT activity. Conclusion: Montelukast inhibits TNF-α-stimulated IL-8 expression through changes in NF-κB p65-assocd. HAT activity. Drugs targeting these enzymes may enhance the anti-inflammatory actions of montelukast.
- 266Chimenti, F.; Bizzarri, B.; Maccioni, E.; Secci, D.; Bolasco, A.; Chimenti, P.; Fioravanti, R.; Granese, A.; Carradori, S.; Tosi, F.; Ballario, P.; Vernarecci, S.; Filetici, P. J. Med. Chem. 2009, 52, 530There is no corresponding record for this reference.
- 267Wu, J.; Wang, J.; Li, M.; Yang, Y.; Wang, B.; Zheng, Y. G. Bioorg. Chem. 2011, 39, 53267Small molecule inhibitors of histone acetyltransferase Tip60Wu, Jiang; Wang, Ju-Xian; Li, Min-Yong; Yang, Yu-Tao; Wang, Bing-He; Zheng, Y. GeorgeBioorganic Chemistry (2011), 39 (1), 53-58CODEN: BOCMBM; ISSN:0045-2068. (Elsevier B.V.)Tip60 is a key member of the MYST family of histone acetyltransferases and involved in a broad spectrum of cellular pathways and disease conditions. So far, small mol. inhibitors of Tip60 and other members of MYST HATs are rarely reported. To discover new small mol. inhibitors of Tip60 as mechanistic tools for functional study and as chem. leads for therapeutic development, we performed virtual screening using the crystal structure of Esa1 (the yeast homolog of Tip60) on a small mol. library database. Radioactive acetylation assays were carried out to further evaluate the virtual screen hits. Several compds. with new structural scaffolds were identified with micromolar inhibition potency for Tip60 from the biochem. studies. Further, computer modeling and kinetic assays suggest that these mols. target the acetyl-CoA binding site in Tip60. These new inhibitors provide valuable chem. hits to develop further potent inhibitors for the MYST HATs.
- 268Whitty, A. Future Med. Chem. 2011, 3, 797268Growing PAINS in academic drug discoveryWhitty, AdrianFuture Medicinal Chemistry (2011), 3 (7), 797-801CODEN: FMCUA7; ISSN:1756-8919. (Future Science Ltd.)In a recent article it was argued that compds. published as drug leads by academic labs. commonly contain functionality that identifies them as nonspecific pan-assay interference compds.'' (PAINS). The article raises broad questions about why best practices for hit and lead qualification that are well known in industry are not more widely employed in academia, as well as about the role of journals in publishing manuscripts that report drug leads of little potential value. Barriers to adoption of best practices for some academic drug-discovery researchers include knowledge gaps and infrastructure deficiencies, but they also arise from fundamental differences in how academic research is structured and how success is measured. Academic drug discovery should not seek to become identical to com. pharmaceutical research, but we can do a better job of assessing and communicating the true potential of the drug leads we publish, thereby reducing the wastage of resources on nonviable compds.
- 269Baell, J. B.; Holloway, G. A. J. Med. Chem. 2010, 53, 2719269New Substructure Filters for Removal of Pan Assay Interference Compounds (PAINS) from Screening Libraries and for Their Exclusion in BioassaysBaell, Jonathan B.; Holloway, Georgina A.Journal of Medicinal Chemistry (2010), 53 (7), 2719-2740CODEN: JMCMAR; ISSN:0022-2623. (American Chemical Society)This report describes a no. of substructural features which can help to identify compds. that appear as frequent hitters (promiscuous compds.) in many biochem. high throughput screens. The compds. identified by such substructural features are not recognized by filters commonly used to identify reactive compds. Even though these substructural features were identified using only one assay detection technol., such compds. have been reported to be active from many different assays. In fact, these compds. are increasingly prevalent in the literature as potential starting points for further exploration, whereas they may not be.
- 270Mujtaba, S.; Zeng, L.; Zhou, M. M. Oncogene 2007, 26, 5521There is no corresponding record for this reference.
- 271Manning, E. T.; Ikehara, T.; Ito, T.; Kadonaga, J. T.; Kraus, W. L. Mol. Cell. Biol. 2001, 21, 3876271p300 forms a stable, template-committed complex with chromatin: role for the bromodomainManning, E. Tory; Ikehara, Tsuyoshi; Ito, Takashi; Kadonaga, James T.; Kraus, W. LeeMolecular and Cellular Biology (2001), 21 (12), 3876-3887CODEN: MCEBD4; ISSN:0270-7306. (American Society for Microbiology)The nature of the interaction of coactivator proteins with transcriptionally active promoters in chromatin is a fundamental question in transcriptional regulation by RNA polymerase II. In this study, we used a biochem. approach to examine the functional assocn. of the coactivator p300 with chromatin templates. Using in vitro transcription template competition assays, we obsd. that p300 forms a stable, template-committed complex with chromatin during the transcription process. The template commitment is dependent on the time of incubation of p300 with the chromatin template and occurs independently of the presence of a transcriptional activator protein. In studies examg. interactions between p300 and chromatin, we found that p300 binds directly to chromatin and that the binding requires the p300 bromodomain, a conserved 110-amino-acid sequence found in many chromatin-assocd. proteins. Furthermore, we obsd. that the isolated p300 bromodomain binds directly to histones, preferentially to histone H3. However, the isolated p300 bromodomain does not bind to nucleosomal histones under the same assay conditions, suggesting that free histones and nucleosomal histones are not equiv. as binding substrates. Collectively, our results suggest that the stable assocn. of p300 with chromatin is mediated, at least in part, by the bromodomain and is critically important for p300 function. Furthermore, our results suggest a model for p300 function that involves distinct activator-dependent targeting and activator-independent chromatin binding activities.
- 272Wei, L.; Jamonnak, N.; Choy, J.; Wang, Z.; Zheng, W. Biochem. Biophys. Res. Commun. 2008, 368, 279There is no corresponding record for this reference.
- 273Polesskaya, A.; Naguibneva, I.; Duquet, A.; Bengal, E.; Robin, P.; Harel-Bellan, A. Mol. Cell. Biol. 2001, 21, 5312273Interaction between acetylated MyoD and the bromodomain of CBP and/or p300Polesskaya, Anna; Naguibneva, Irina; Duquet, Arnaud; Bengal, Eyal; Robin, Philippe; Harel-Bellan, AnnickMolecular and Cellular Biology (2001), 21 (16), 5312-5320CODEN: MCEBD4; ISSN:0270-7306. (American Society for Microbiology)Acetylation is emerging as a posttranslational modification of nuclear proteins that is essential to the regulation of transcription and that modifies transcription factor affinity for binding sites on DNA, stability, and/or nuclear localization. Here, we present both in vitro and in vivo evidence that acetylation increases the affinity of myogenic factor MyoD for acetyltransferases CBP and p300. In myogenic cells, the fraction of endogenous MyoD that is acetylated was found assocd. with CBP or p300. In vitro, the interaction between MyoD and CBP was more resistant to high salt concns. and was detected with lower doses of MyoD when MyoD was acetylated. Interestingly, an anal. of CBP mutants revealed that the interaction with acetylated MyoD involves the bromodomain of CBP. In live cells, MyoD mutants that cannot be acetylated did not assoc. with CBP or p300 and were strongly impaired in their ability to cooperate with CBP for transcriptional activation of a muscle creatine kinase-luciferase construct. Taken together, our data suggest a new mechanism for activation of protein function by acetylation and demonstrate for the first time an acetylation-dependent interaction between the bromodomain of CBP and a nonhistone protein.
- 274Hou, T.; Ray, S.; Lee, C.; Brasier, A. R. J. Biol. Chem. 2008, 283, 30725There is no corresponding record for this reference.
- 275Mujtaba, S.; He, Y.; Zeng, L.; Yan, S.; Plotnikova, O.; Sachchidanand; Sanchez, R.; Zeleznik-Le, N. J.; Ronai, Z.; Zhou, M. M. Mol. Cell 2004, 13, 251There is no corresponding record for this reference.
- 276Kim, J. H.; Cho, E. J.; Kim, S. T.; Youn, H. D. Nat. Struct. Mol. Biol. 2005, 12, 423276CtBP represses p300-mediated transcriptional activation by direct association with its bromodomainKim, Jae-Hwan; Cho, Eun-Jung; Kim, Seong-Tae; Youn, Hong-DukNature Structural & Molecular Biology (2005), 12 (5), 423-428CODEN: NSMBCU; ISSN:1545-9993. (Nature Publishing Group)Histone acetyltransferase coactivators bind to acetylated histones through their bromodomains and catalyze the acetylation of histone H3 and H4 tails for transcriptional activation. C-terminal binding protein (CtBP) serves as a transcriptional corepressor by recruiting histone deacetylases. However, the precise mechanism by which CtBP represses transcription has not been detd. In this study authors found that CtBP1 directly assocs. with p300 by binding to the PXDLS motif in the bromodomain of p300. Moreover, CtBP1 blocks the accessibility of p300 to histones in an NADH-sensitive manner and thus represses p300-mediated histone acetylation and transcriptional activation. In addn., an NADH-nonresponsive, monomeric mutant, CtBP1 (G183V), was found to strongly repress p300-mediated transcriptional activation. Thus, the dissocn. of NADH from CtBP1 leads to the repression of p300-driven general transcriptional activity by CtBP1. These results suggest a novel mechanism whereby CtBP1 serves as an energy-sensing repressor of histone acetyltransferase(s) and thus affects general transcription.
- 277Reynoird, N.; Schwartz, B. E.; Delvecchio, M.; Sadoul, K.; Meyers, D.; Mukherjee, C.; Caron, C.; Kimura, H.; Rousseaux, S.; Cole, P. A.; Panne, D.; French, C. A.; Khochbin, S. EMBO J. 2010, 29, 2943There is no corresponding record for this reference.
- 278Chan, H. M.; La Thangue, N. B. J. Cell Sci. 2001, 114, 2363278p300/CBP proteins: HATs for transcriptional bridges and scaffoldsChan, Ho Man; La Thangue, Nicholas B.Journal of Cell Science (2001), 114 (13), 2363-2373CODEN: JNCSAI; ISSN:0021-9533. (Company of Biologists Ltd.)A review with 139 refs. P300/CBP transcriptional co-activator proteins play a central role in co-ordinating and integrating multiple signal-dependent events with the transcription app., allowing the appropriate level of gene activity to occur in response to diverse physiol. cues that influence, for example, proliferation, differentiation and apoptosis. P300/CBP activity can be under aberrant control in human disease, particularly in cancer, which may inactivate a p300/CBP tumor-suppressor-like activity. The transcription regulating-properties of p300 and CBP appear to be exerted through multiple mechanisms. They act as protein bridges, thereby connecting different sequence-specific transcription factors to the transcription app. Providing a protein scaffold upon which to build a multicomponent transcriptional regulatory complex is likely to be an important feature of p300/CBP control. Another key property is the presence of histone acetyltransferase (HAT) activity, which endows p300/CBP with the capacity to influence chromatin activity by modulating nucleosomal histones. Other proteins, including the p53 tumor suppressor, are targets for acetylation by p300/CBP. With the current intense level of research activity, p300/CBP will continue to be in the limelight and, we can be confident, yield new and important information on fundamental processes involved in transcriptional control.
- 279Wang, F.; Marshall, C. B.; Ikura, M. Cell. Mol. Life Sci. 2013, 70, 3989279Transcriptional/epigenetic regulator CBP/p300 in tumorigenesis: structural and functional versatility in target recognitionWang, Feng; Marshall, Christopher B.; Ikura, MitsuhikoCellular and Molecular Life Sciences (2013), 70 (21), 3989-4008CODEN: CMLSFI; ISSN:1420-682X. (Birkhaeuser Basel)A review. In eukaryotic cells, gene transcription is regulated by sequence-specific DNA-binding transcription factors that recognize promoter and enhancer elements near the transcriptional start site. Some coactivators promote transcription by connecting transcription factors to the basal transcriptional machinery. The highly conserved coactivators CREB-binding protein (CBP) and its paralog, E1A-binding protein (p300), each have four sep. transactivation domains (TADs) that interact with the TADs of a no. of DNA-binding transcription activators as well as general transcription factors (GTFs), thus mediating recruitment of basal transcription machinery to the promoter. Most promoters comprise multiple activator-binding sites, and many activators contain tandem TADs, thus multivalent interactions may stabilize CBP/p300 at the promoter, and intrinsically disordered regions in CBP/p300 and many activators may confer adaptability to these multivalent complexes. CBP/p300 contains a catalytic histone acetyltransferase (HAT) domain, which remodels chromatin to 'relax' its superstructure and enables transcription of proximal genes. The HAT activity of CBP/p300 also acetylates some transcription factors (e.g., p53), hence modulating the function of key transcriptional regulators. Through these numerous interactions, CBP/p300 has been implicated in complex physiol. and pathol. processes, and, in response to different signals, can drive cells towards proliferation or apoptosis. Dysregulation of the transcriptional and epigenetic functions of CBP/p300 is assocd. with leukemia and other types of cancer, thus it has been recognized as a potential anti-cancer drug target. In this review, we focus on recent exciting findings in the structural mechanisms of CBP/p300 involving multivalent and dynamic interactions with binding partners, which may pave new avenues for anti-cancer drug development.
- 280Ragvin, A.; Valvatne, H.; Erdal, S.; Arskog, V.; Tufteland, K. R.; Breen, K.; ØYan, A. M.; Eberharter, A.; Gibson, T. J.; Becker, P. B.; Aasland, R. J. Mol. Biol. 2004, 337, 773There is no corresponding record for this reference.
- 281He, J.; Ye, J.; Cai, Y.; Riquelme, C.; Liu, J. O.; Liu, X.; Han, A.; Chen, L. Nucleic Acids Res. 2011, 39, 4464There is no corresponding record for this reference.
- 282Zeng, L.; Zhang, Q.; Gerona-Navarro, G.; Moshkina, N.; Zhou, M. M. Structure 2008, 16, 643There is no corresponding record for this reference.
- 283Mujtaba, S.; He, Y.; Zeng, L.; Yan, S.; Plotnikova, O.; Sachchidanand; Sanchez, R.; Zeleznik-Le, N. J.; Ronai, Z.; Zhou, M. M. Mol. Cell 2004, 13, 251There is no corresponding record for this reference.
- 284Ferreon, J. C.; Martinez-Yamout, M. A.; Dyson, H. J.; Wright, P. E. Proc. Natl. Acad. Sci. U.S.A. 2009, 106, 13260There is no corresponding record for this reference.
- 285Feng, H.; Jenkins, L. M.; Durell, S. R.; Hayashi, R.; Mazur, S. J.; Cherry, S.; Tropea, J. E.; Miller, M.; Wlodawer, A.; Appella, E.; Bai, Y. Structure 2009, 17, 202There is no corresponding record for this reference.
- 286Wojciak, J. M.; Martinez-Yamout, M. A.; Dyson, H. J.; Wright, P. E. EMBO J. 2009, 28, 948There is no corresponding record for this reference.
- 287Das, C.; Roy, S.; Namjoshi, S.; Malarkey, C. S.; Jones, D. N.; Kutateladze, T. G.; Churchill, M. E.; Tyler, J. K. Proc. Natl. Acad. Sci. U.S.A. 2014, 111, E1072There is no corresponding record for this reference.
- 288Delvecchio, M.; Gaucher, J.; Aguilar-Gurrieri, C.; Ortega, E.; Panne, D. Nat. Struct. Mol. Biol. 2013, 20, 1040288Structure of the p300 catalytic core and implications for chromatin targeting and HAT regulationDelvecchio, Manuela; Gaucher, Jonathan; Aguilar-Gurrieri, Carmen; Ortega, Esther; Panne, DanielNature Structural & Molecular Biology (2013), 20 (9), 1040-1046CODEN: NSMBCU; ISSN:1545-9993. (Nature Publishing Group)CBP and p300 are histone acetyltransferases (HATs) that assoc. with and acetylate transcriptional regulators and chromatin. Mutations in their catalytic 'cores' are linked to genetic disorders, including cancer. Here we present the 2.8-Å crystal structure of the catalytic core of human p300 contg. its bromodomain, CH2 region and HAT domain. The structure reveals that the CH2 region contains a discontinuous PHD domain interrupted by a RING domain. The bromodomain, PHD, RING and HAT domains adopt an assembled configuration with the RING domain positioned over the HAT substrate-binding pocket. Disease mutations that disrupt RING attachment led to upregulation of HAT activity, thus revealing an inhibitory role for this domain. The structure provides a starting point for understanding how chromatin-substrate targeting and HAT regulation are coupled and why mutations in the p300 core lead to dysregulation.
- 289He, J.; Ye, J.; Cai, Y.; Riquelme, C.; Liu, J. O.; Liu, X.; Han, A.; Chen, L. Nucleic Acids Res. 2011, 39, 4464There is no corresponding record for this reference.
- 290Feng, H.; Jenkins, L. M.; Durell, S. R.; Hayashi, R.; Mazur, S. J.; Cherry, S.; Tropea, J. E.; Miller, M.; Wlodawer, A.; Appella, E.; Bai, Y. Structure 2009, 17, 202There is no corresponding record for this reference.
- 291Ferreon, J. C.; Martinez-Yamout, M. A.; Dyson, H. J.; Wright, P. E. Proc. Natl. Acad. Sci. U.S.A. 2009, 106, 13260There is no corresponding record for this reference.
- 292Wojciak, J. M.; Martinez-Yamout, M. A.; Dyson, H. J.; Wright, P. E. EMBO J. 2009, 28, 948There is no corresponding record for this reference.
- 293Chan, H. M.; La Thangue, N. B. J. Cell Sci. 2001, 114, 2363293p300/CBP proteins: HATs for transcriptional bridges and scaffoldsChan, Ho Man; La Thangue, Nicholas B.Journal of Cell Science (2001), 114 (13), 2363-2373CODEN: JNCSAI; ISSN:0021-9533. (Company of Biologists Ltd.)A review with 139 refs. P300/CBP transcriptional co-activator proteins play a central role in co-ordinating and integrating multiple signal-dependent events with the transcription app., allowing the appropriate level of gene activity to occur in response to diverse physiol. cues that influence, for example, proliferation, differentiation and apoptosis. P300/CBP activity can be under aberrant control in human disease, particularly in cancer, which may inactivate a p300/CBP tumor-suppressor-like activity. The transcription regulating-properties of p300 and CBP appear to be exerted through multiple mechanisms. They act as protein bridges, thereby connecting different sequence-specific transcription factors to the transcription app. Providing a protein scaffold upon which to build a multicomponent transcriptional regulatory complex is likely to be an important feature of p300/CBP control. Another key property is the presence of histone acetyltransferase (HAT) activity, which endows p300/CBP with the capacity to influence chromatin activity by modulating nucleosomal histones. Other proteins, including the p53 tumor suppressor, are targets for acetylation by p300/CBP. With the current intense level of research activity, p300/CBP will continue to be in the limelight and, we can be confident, yield new and important information on fundamental processes involved in transcriptional control.
- 294Simone, C.; Stiegler, P.; Forcales, S. V.; Bagella, L.; De Luca, A.; Sartorelli, V.; Giordano, A.; Puri, P. L. Oncogene 2004, 23, 2177There is no corresponding record for this reference.
- 295Saint Just Ribeiro, M.; Hansson, M. L.; Wallberg, A. E. Biochem. J. 2007, 404, 289There is no corresponding record for this reference.
- 296Tax, F. E.; Thomas, J. H.; Ferguson, E. L.; Horvitz, H. R. Genetics 1997, 147, 1675There is no corresponding record for this reference.
- 297Petcherski, A. G.; Kimble, J. Nature 2000, 405, 364There is no corresponding record for this reference.
- 298Petcherski, A. G.; Kimble, J. Curr. Biol. 2000, 10, R471There is no corresponding record for this reference.
- 299Kitagawa, M.; Oyama, T.; Kawashima, T.; Yedvobnick, B.; Kumar, A.; Matsuno, K.; Harigaya, K. Mol. Cell. Biol. 2001, 21, 4337299A human protein with sequence similarity to Drosophila mastermind coordinates the nuclear form of Notch and a CSL protein to build a transcriptional activator complex on target promotersKitagawa, Motoo; Oyama, Toshinao; Kawashima, Taichi; Yedvobnick, Barry; Kumar, Anumeha; Matsuno, Kenji; Harigaya, KenichiMolecular and Cellular Biology (2001), 21 (13), 4337-4346CODEN: MCEBD4; ISSN:0270-7306. (American Society for Microbiology)Mastermind (Mam) has been implicated as an important pos. regulator of the Notch signaling pathway by genetic studies using Drosophila melanogaster. Here we describe a biochem. mechanism of action of Mam within the Notch signaling pathway. Expression of a human sequence related to Drosophila Mam (hMam-1) in mammalian cells augments induction of Hairy Enhancer of split (HES) promoters by Notch signaling. HMam-1 stabilizes and participates in the DNA binding complex of the intracellular domain of human Notch1 and a CSL protein. Truncated versions of hMam-1 that can maintain an assocn. with the complex behave in a dominant neg. fashion and depress transactivation. Furthermore, Drosophila Mam forms a similar complex with the intracellular domain of Drosophila Notch and Drosophila CSL protein during activation of Enhancer of split, the Drosophila counterpart of HES. These results indicate that Mam is an essential component of the transcriptional app. of Notch signaling.
- 300Wu, L.; Aster, J. C.; Blacklow, S. C.; Lake, R.; Artavanis-Tsakonas, S.; Griffin, J. D. Nat. Genet. 2000, 26, 484300MAML1, a human homologue of Drosophila mastermind, is a transcriptional co-activator for NOTCH receptorsWu, Lizi; Aster, Jon C.; Blacklow, Stephen C.; Lake, Robert; Artavanis-Tsakonas, Spyros; Griffin, James D.Nature Genetics (2000), 26 (4), 484-489CODEN: NGENEC; ISSN:1061-4036. (Nature America Inc.)Notch receptors are involved in cell-fate detn. in organisms as diverse as flies, frogs and humans. In Drosophila melanogaster, loss-of-function mutations of Notch produce a "neurogenic" phenotype in which cells destined to become epidermis switch fate and differentiate to neural cells. Upon ligand activation, the intracellular domain of Notch (ICN) translocates to the nucleus, and interacts directly with the DNA-binding protein Suppressor of hairless (Su(H)) in flies, or recombination signal binding protein Jκ (RBP-Jκ) in mammals, to activate gene transcription. But the precise mechanisms of Notch-induced gene expression are not completely understood. The gene mastermind has been identified in multiple genetic screens for modifiers of Notch mutations in Drosophila. Here we clone MAML1, a human homolog of the Drosophila gene Mastermind, and show that it encodes a protein of 130 kD localizing to nuclear bodies. MAML1 binds to the ankyrin repeat domain of all four mammalian NOTCH receptors, forms a DNA-binding complex with ICN and RBP-Jκ, and amplifies NOTCH-induced transcription of HES1. These studies provide a mol. mechanism to explain the genetic links between mastermind and Notch in Drosophila and indicate that MAML1 functions as a transcriptional co-activator for NOTCH signaling.
- 301Wilson, J. J.; Kovall, R. A. Cell 2006, 124, 985There is no corresponding record for this reference.
- 302Arnett, K. L.; Hass, M.; McArthur, D. G.; Ilagan, M. X.; Aster, J. C.; Kopan, R.; Blacklow, S. C. Nat. Struct. Mol. Biol. 2010, 17, 1312302Structural and mechanistic insights into cooperative assembly of dimeric Notch transcription complexesArnett, Kelly L.; Hass, Matthew; McArthur, Debbie G.; Ilagan, Ma Xenia G.; Aster, Jon C.; Kopan, Raphael; Blacklow, Stephen C.Nature Structural & Molecular Biology (2010), 17 (11), 1312-1317CODEN: NSMBCU; ISSN:1545-9993. (Nature Publishing Group)Ligand-induced proteolysis of Notch produces an intracellular effector domain that transduces essential signals by regulating the transcription of target genes. This function relies on the formation of transcriptional activation complexes that include intracellular Notch, a Mastermind co-activator and the transcription factor CSL bound to cognate DNA. These complexes form higher-order assemblies on paired, head-to-head CSL recognition sites. Here we report the X-ray structure of a dimeric human Notch1 transcription complex loaded on the paired site from the human HES1 promoter. The small interface between the Notch ankyrin domains could accommodate DNA bending and untwisting to allow a range of spacer lengths between the two sites. Cooperative dimerization occurred on the human and mouse Hes5 promoters at a sequence that diverged from the CSL-binding consensus at one of the sites. These studies reveal how promoter organizational features control cooperativity and, thus, the responsiveness of different promoters to Notch signaling.
- 303Nam, Y.; Sliz, P.; Song, L.; Aster, J. C.; Blacklow, S. C. Cell 2006, 124, 973There is no corresponding record for this reference.
- 304Choi, S. H.; Wales, T. E.; Nam, Y.; O’Donovan, D. J.; Sliz, P.; Engen, J. R.; Blacklow, S. C. Structure 2012, 20, 340There is no corresponding record for this reference.
- 305Hansson, M. L.; Popko-Scibor, A. E.; Saint Just Ribeiro, M.; Dancy, B. M.; Lindberg, M. J.; Cole, P. A.; Wallberg, A. E. Nucleic Acids Res. 2009, 37, 2996There is no corresponding record for this reference.
- 306Fryer, C. J.; Lamar, E.; Turbachova, I.; Kintner, C.; Jones, K. A. Genes Dev. 2002, 16, 1397There is no corresponding record for this reference.
- 307Saint Just Ribeiro, M.; Hansson, M. L.; Wallberg, A. E. Biochem. J. 2007, 404, 289There is no corresponding record for this reference.
- 308Black, J. C.; Choi, J. E.; Lombardo, S. R.; Carey, M. Mol. Cell 2006, 23, 809There is no corresponding record for this reference.
- 309Kim, J. H.; Cho, E. J.; Kim, S. T.; Youn, H. D. Nat. Struct. Mol. Biol. 2005, 12, 423309CtBP represses p300-mediated transcriptional activation by direct association with its bromodomainKim, Jae-Hwan; Cho, Eun-Jung; Kim, Seong-Tae; Youn, Hong-DukNature Structural & Molecular Biology (2005), 12 (5), 423-428CODEN: NSMBCU; ISSN:1545-9993. (Nature Publishing Group)Histone acetyltransferase coactivators bind to acetylated histones through their bromodomains and catalyze the acetylation of histone H3 and H4 tails for transcriptional activation. C-terminal binding protein (CtBP) serves as a transcriptional corepressor by recruiting histone deacetylases. However, the precise mechanism by which CtBP represses transcription has not been detd. In this study authors found that CtBP1 directly assocs. with p300 by binding to the PXDLS motif in the bromodomain of p300. Moreover, CtBP1 blocks the accessibility of p300 to histones in an NADH-sensitive manner and thus represses p300-mediated histone acetylation and transcriptional activation. In addn., an NADH-nonresponsive, monomeric mutant, CtBP1 (G183V), was found to strongly repress p300-mediated transcriptional activation. Thus, the dissocn. of NADH from CtBP1 leads to the repression of p300-driven general transcriptional activity by CtBP1. These results suggest a novel mechanism whereby CtBP1 serves as an energy-sensing repressor of histone acetyltransferase(s) and thus affects general transcription.
- 310Kraus, W. L.; Manning, E. T.; Kadonaga, J. T. Mol. Cell. Biol. 1999, 19, 8123310Biochemical analysis of distinct activation functions in p300 that enhance transcription initiation with chromatin templatesKraus, W. Lee; Manning, E. Tory; Kadonaga, James T.Molecular and Cellular Biology (1999), 19 (12), 8123-8135CODEN: MCEBD4; ISSN:0270-7306. (American Society for Microbiology)To investigate the mechanisms of transcriptional enhancement by the p300 coactivator, we analyzed wild-type and mutant versions of p300 with a chromatin transcription system in vitro. Estrogen receptor, NF-κB p65 plus Sp1, and Gal4-VP16 were used as different sequence-specific activators. The CH3 domain (or E1A-binding region) was found to be essential for the function of each of the activators tested. The bromodomain was also obsd. to be generally important for p300 coactivator activity, though to a lesser extent than the CH3 domain/E1A-binding region. The acetyltransferase activity and the C-terminal region (contg. the steroid receptor coactivator/p160-binding region and the glutamine-rich region) were each found to be important for activation by estrogen receptor but not for that by Gal4-VP16. The N-terminal region of p300, which had been previously found to interact with nuclear hormone receptors, was not seen to be required for any of the activators, including estrogen receptor. Single-round transcription expts. revealed that the functionally important subregions of p300 contribute to its ability to promote the assembly of transcription initiation complexes. In addn., the acetyltransferase activity of p300 was obsd. to be distinct from the broadly essential activation function of the CH3 domain/E1A-binding region. These results indicate that specific regions of p300 possess distinct activation functions that are differentially required to enhance the assembly of transcription initiation complexes. Interestingly, with the estrogen receptor, four distinct regions of p300 each have an essential role in the transcription activation process. These data exemplify a situation in which a network of multiple activation functions is required to achieve gene transcription.
- 311Delvecchio, M.; Gaucher, J.; Aguilar-Gurrieri, C.; Ortega, E.; Panne, D. Nat. Struct. Mol. Biol. 2013, 20, 1040311Structure of the p300 catalytic core and implications for chromatin targeting and HAT regulationDelvecchio, Manuela; Gaucher, Jonathan; Aguilar-Gurrieri, Carmen; Ortega, Esther; Panne, DanielNature Structural & Molecular Biology (2013), 20 (9), 1040-1046CODEN: NSMBCU; ISSN:1545-9993. (Nature Publishing Group)CBP and p300 are histone acetyltransferases (HATs) that assoc. with and acetylate transcriptional regulators and chromatin. Mutations in their catalytic 'cores' are linked to genetic disorders, including cancer. Here we present the 2.8-Å crystal structure of the catalytic core of human p300 contg. its bromodomain, CH2 region and HAT domain. The structure reveals that the CH2 region contains a discontinuous PHD domain interrupted by a RING domain. The bromodomain, PHD, RING and HAT domains adopt an assembled configuration with the RING domain positioned over the HAT substrate-binding pocket. Disease mutations that disrupt RING attachment led to upregulation of HAT activity, thus revealing an inhibitory role for this domain. The structure provides a starting point for understanding how chromatin-substrate targeting and HAT regulation are coupled and why mutations in the p300 core lead to dysregulation.
- 312Filippakopoulos, P.; Knapp, S. Nat. Rev. Drug Discovery 2014, 13, 337312Targeting bromodomains: epigenetic readers of lysine acetylationFilippakopoulos, Panagis; Knapp, StefanNature Reviews Drug Discovery (2014), 13 (5), 337-356CODEN: NRDDAG; ISSN:1474-1776. (Nature Publishing Group)A review. Lysine acetylation is a key mechanism that regulates chromatin structure; aberrant acetylation levels have been linked to the development of several diseases. Acetyl-lysine modifications create docking sites for bromodomains, which are small interaction modules found on diverse proteins, some of which have a key role in the acetylation-dependent assembly of transcriptional regulator complexes. These complexes can then initiate transcriptional programs that result in phenotypic changes. The recent discovery of potent and highly specific inhibitors for the BET (bromodomain and extra-terminal) family of bromodomains has stimulated intensive research activity in diverse therapeutic areas, particularly in oncol., where BET proteins regulate the expression of key oncogenes and anti-apoptotic proteins. In addn., targeting BET bromodomains could hold potential for the treatment of inflammation and viral infection. Here, we highlight recent progress in the development of bromodomain inhibitors, and their potential applications in drug discovery.
- 313Nicodeme, E.; Jeffrey, K. L.; Schaefer, U.; Beinke, S.; Dewell, S.; Chung, C. W.; Chandwani, R.; Marazzi, I.; Wilson, P.; Coste, H.; White, J.; Kirilovsky, J.; Rice, C. M.; Lora, J. M.; Prinjha, R. K.; Lee, K.; Tarakhovsky, A. Nature 2010, 468, 1119There is no corresponding record for this reference.
- 314Mirguet, O.; Lamotte, Y.; Donche, F.; Toum, J.; Gellibert, F.; Bouillot, A.; Gosmini, R.; Nguyen, V. L.; Delannée, D.; Seal, J.; Blandel, F.; Boullay, A. B.; Boursier, E.; Martin, S.; Brusq, J. M.; Krysa, G.; Riou, A.; Tellier, R.; Costaz, A.; Huet, P.; Dudit, Y.; Trottet, L.; Kirilovsky, J.; Nicodeme, E. Bioorg. Med. Chem. Lett. 2012, 22, 2963There is no corresponding record for this reference.
- 315Dawson, M. A.; Prinjha, R. K.; Dittmann, A.; Giotopoulos, G.; Bantscheff, M.; Chan, W. I.; Robson, S. C.; Chung, C. W.; Hopf, C.; Savitski, M. M.; Huthmacher, C.; Gudgin, E.; Lugo, D.; Beinke, S.; Chapman, T. D.; Roberts, E. J.; Soden, P. E.; Auger, K. R.; Mirguet, O.; Doehner, K.; Delwel, R.; Burnett, A. K.; Jeffrey, P.; Drewes, G.; Lee, K.; Huntly, B. J.; Kouzarides, T. Nature 2011, 478, 529There is no corresponding record for this reference.
- 316Filippakopoulos, P.; Qi, J.; Picaud, S.; Shen, Y.; Smith, W. B.; Fedorov, O.; Morse, E. M.; Keates, T.; Hickman, T. T.; Felletar, I.; Philpott, M.; Munro, S.; McKeown, M. R.; Wang, Y.; Christie, A. L.; West, N.; Cameron, M. J.; Schwartz, B.; Heightman, T. D.; La Thangue, N.; French, C. A.; Wiest, O.; Kung, A. L.; Knapp, S.; Bradner, J. E. Nature 2010, 468, 1067316Selective inhibition of BET bromodomainsFilippakopoulos, Panagis; Qi, Jun; Picaud, Sarah; Shen, Yao; Smith, William B.; Fedorov, Oleg; Morse, Elizabeth M.; Keates, Tracey; Hickman, Tyler T.; Felletar, Ildiko; Philpott, Martin; Munro, Shongah; McKeown, Michael R.; Wang, Yuchuan; Christie, Amanda L.; West, Nathan; Cameron, Michael J.; Schwartz, Brian; Heightman, Tom D.; La Thangue, Nicholas; French, Christopher; Wiest, Olaf; Kung, Andrew L.; Knapp, Stefan; Bradner, James E.Nature (London, United Kingdom) (2010), 468 (7327), 1067-1073CODEN: NATUAS; ISSN:0028-0836. (Nature Publishing Group)Epigenetic proteins are intently pursued targets in ligand discovery. So far, successful efforts have been limited to chromatin modifying enzymes, or so-called epigenetic 'writers' and 'erasers'. Potent inhibitors of histone binding modules have not yet been described. Here the authors report a cell-permeable small mol. (I,JQ1) that binds competitively to acetyl-lysine recognition motifs, or bromodomains. High potency and specificity towards a subset of human bromodomains is explained by co-crystal structures with bromodomain and extra-terminal (BET) family member BRD4, revealing excellent shape complementarity with the acetyl-lysine binding cavity. Recurrent translocation of BRD4 is obsd. in a genetically-defined, incurable subtype of human squamous carcinoma. Competitive binding by JQ1 displaces the BRD4 fusion oncoprotein from chromatin, prompting squamous differentiation and specific antiproliferative effects in BRD4-dependent cell lines and patient-derived xenograft models. These data establish proof-of-concept for targeting protein-protein interactions of epigenetic 'readers', and provide a versatile chem. scaffold for the development of chem. probes more broadly throughout the bromodomain family.
- 317Fish, P. V.; Filippakopoulos, P.; Bish, G.; Brennan, P. E.; Bunnage, M. E.; Cook, A. S.; Federov, O.; Gerstenberger, B. S.; Jones, H.; Knapp, S.; Marsden, B.; Nocka, K.; Owen, D. R.; Philpott, M.; Picaud, S.; Primiano, M. J.; Ralph, M. J.; Sciammetta, N.; Trzupek, J. D. J. Med. Chem. 2012, 55, 9831There is no corresponding record for this reference.
- 318Picaud, S.; Da Costa, D.; Thanasopoulou, A.; Filippakopoulos, P.; Fish, P. V.; Philpott, M.; Fedorov, O.; Brennan, P.; Bunnage, M. E.; Owen, D. R.; Bradner, J. E.; Taniere, P.; O’Sullivan, B.; Müller, S.; Schwaller, J.; Stankovic, T.; Knapp, S. Cancer Res. 2013, 73, 3336There is no corresponding record for this reference.
- 319Hewings, D. S.; Wang, M.; Philpott, M.; Fedorov, O.; Uttarkar, S.; Filippakopoulos, P.; Picaud, S.; Vuppusetty, C.; Marsden, B.; Knapp, S.; Conway, S. J.; Heightman, T. D. J. Med. Chem. 2011, 54, 6761There is no corresponding record for this reference.
- 320Hay, D.; Fedorov, O.; Filippakopoulos, P.; Martin, S.; Philpott, M.; Picaud, S.; Hewings, D. S.; Uttakar, S.; Heightman, T. D.; Conway, S. J.; Knapp, S.; Brennan, P. E. Med. Chem. Commun. 2013, 4, 140There is no corresponding record for this reference.
- 321Bamborough, P.; Diallo, H.; Goodacre, J. D.; Gordon, L.; Lewis, A.; Seal, J. T.; Wilson, D. M.; Woodrow, M. D.; Chung, C. W. J. Med. Chem. 2012, 55, 587There is no corresponding record for this reference.
- 322Boehm, D.; Calvanese, V.; Dar, R. D.; Xing, S.; Schroeder, S.; Martins, L.; Aull, K.; Li, P. C.; Planelles, V.; Bradner, J. E.; Zhou, M. M.; Siliciano, R. F.; Weinberger, L.; Verdin, E.; Ott, M. Cell Cycle 2013, 12, 452There is no corresponding record for this reference.
- 323Li, Z.; Guo, J.; Wu, Y.; Zhou, Q. Nucleic Acids Res. 2013, 41, 277There is no corresponding record for this reference.
- 324Zhu, J.; Gaiha, G. D.; John, S. P.; Pertel, T.; Chin, C. R.; Gao, G.; Qu, H.; Walker, B. D.; Elledge, S. J.; Brass, A. L. Cell Rep. 2012, 2, 807324Reactivation of latent HIV-1 by inhibition of BRD4Zhu, Jian; Gaiha, Gaurav D.; John, Sinu P.; Pertel, Thomas; Chin, Christopher R.; Gao, Geng; Qu, Hongjing; Walker, Bruce D.; Elledge, Stephen J.; Brass, Abraham L.Cell Reports (2012), 2 (4), 807-816CODEN: CREED8; ISSN:2211-1247. (Cell Press)HIV-1 depends on many host factors for propagation. Other host factors, however, antagonize HIV-1 and may have profound effects on viral activation. Curing HIV-1 requires the redn. of latent viral reservoirs that remain in the face of antiretroviral therapy. Using orthologous genetic screens, we identified bromodomain contg. 4 (BRD4) as a neg. regulator of HIV-1 replication. Antagonism of BRD4, via RNA interference or with a small mol. inhibitor, JQ1, both increased proviral transcriptional elongation and alleviated HIV-1 latency in cell-line models. In multiple instances, JQ1, when used in combination with the NF-κB activators Prostratin or PHA, enhanced the in vitro reactivation of latent HIV-1 in primary T cells. These data are consistent with a model wherein BRD4 competes with the virus for HIV-1 dependency factors (HDFs) and suggests that combinatorial therapies that activate HDFs and antagonize HIV-1 competitive factors may be useful for curing HIV-1 infection.
- 325Banerjee, C.; Archin, N.; Michaels, D.; Belkina, A. C.; Denis, G. V.; Bradner, J.; Sebastiani, P.; Margolis, D. M.; Montano, M. J. Leukocyte Biol. 2012, 92, 1147325BET bromodomain inhibition as a novel strategy for reactivation of HIV-1Banerjee, Camellia; Archin, Nancie; Michaels, Daniel; Belkina, Anna C.; Denis, Gerald V.; Bradner, James; Sebastiani, Paola; Margolis, David M.; Montano, MontyJournal of Leukocyte Biology (2012), 92 (6), 1147-1154CODEN: JLBIE7; ISSN:0741-5400. (Society for Leukocyte Biology)The persistence of latent HIV-1 remains a major challenge in therapeutic efforts to eradicate infection. We report the capacity for HIV reactivation by a selective small mol. inhibitor of BET family bromodomains, JQ1, a promising therapeutic agent with antioncogenic properties. JQ1 reactivated HIV transcription in models of latent T cell infection and latent monocyte infection. We also tested the effect of exposure to JQ1 to allow recovery of replication-competent HIV from pools of resting CD4+ T cells isolated from HIV-infected, ART-treated patients. In one of three patients, JQ1 allowed recovery of virus at a frequency above unstimulated conditions. JQ1 potently suppressed T cell proliferation with minimal cytotoxic effect. Transcriptional profiling of T cells with JQ1 showed potent down-regulation of T cell activation genes, including CD3, CD28, and CXCR4, similar to HDAC inhibitors, but JQ1 also showed potent up-regulation of chromatin modification genes, including SIRT1, HDAC6, and multiple lysine demethylases (KDMs). Thus, JQ1 reactivates HIV-1 while suppressing T cell activation genes and up-regulating histone modification genes predicted to favor increased Tat activity. Thus, JQ1 may be useful in studies of potentially novel mechanisms for transcriptional control as well as in translational efforts to identify therapeutic mols. to achieve viral eradication.
- 326Wang, X.; Li, J.; Schowalter, R. M.; Jiao, J.; Buck, C. B.; You, J. PLoS Pathog. 2012, 8, e1003021There is no corresponding record for this reference.
- 327Goupille, O.; Penglong, T.; Lefèvre, C.; Granger, M.; Kadri, Z.; Fucharoen, S.; Maouche-Chrétien, L.; Leboulch, P.; Chrétien, S. Biochem. Biophys. Res. Commun. 2012, 429, 1There is no corresponding record for this reference.
- 328Ott, C. J.; Kopp, N.; Bird, L.; Paranal, R. M.; Qi, J.; Bowman, T.; Rodig, S. J.; Kung, A. L.; Bradner, J. E.; Weinstock, D. M. Blood 2012, 120, 2843There is no corresponding record for this reference.
- 329Picaud, S.; Da Costa, D.; Thanasopoulou, A.; Filippakopoulos, P.; Fish, P. V.; Philpott, M.; Fedorov, O.; Brennan, P.; Bunnage, M. E.; Owen, D. R.; Bradner, J. E.; Taniere, P.; O’Sullivan, B.; Müller, S.; Schwaller, J.; Stankovic, T.; Knapp, S. Cancer Res. 2013, 73, 3336There is no corresponding record for this reference.
- 330Dawson, M. A.; Prinjha, R. K.; Dittmann, A.; Giotopoulos, G.; Bantscheff, M.; Chan, W. I.; Robson, S. C.; Chung, C. W.; Hopf, C.; Savitski, M. M.; Huthmacher, C.; Gudgin, E.; Lugo, D.; Beinke, S.; Chapman, T. D.; Roberts, E. J.; Soden, P. E.; Auger, K. R.; Mirguet, O.; Doehner, K.; Delwel, R.; Burnett, A. K.; Jeffrey, P.; Drewes, G.; Lee, K.; Huntly, B. J.; Kouzarides, T. Nature 2011, 478, 529There is no corresponding record for this reference.
- 331Zuber, J.; Shi, J.; Wang, E.; Rappaport, A. R.; Herrmann, H.; Sison, E. A.; Magoon, D.; Qi, J.; Blatt, K.; Wunderlich, M.; Taylor, M. J.; Johns, C.; Chicas, A.; Mulloy, J. C.; Kogan, S. C.; Brown, P.; Valent, P.; Bradner, J. E.; Lowe, S. W.; Vakoc, C. R. Nature 2011, 478, 524There is no corresponding record for this reference.
- 332Mertz, J. A.; Conery, A. R.; Bryant, B. M.; Sandy, P.; Balasubramanian, S.; Mele, D. A.; Bergeron, L.; Sims, R. J., III. Proc. Natl. Acad. Sci. U.S.A. 2011, 108, 16669There is no corresponding record for this reference.
- 333Delmore, J. E.; Issa, G. C.; Lemieux, M. E.; Rahl, P. B.; Shi, J.; Jacobs, H. M.; Kastritis, E.; Gilpatrick, T.; Paranal, R. M.; Qi, J.; Chesi, M.; Schinzel, A. C.; McKeown, M. R.; Heffernan, T. P.; Vakoc, C. R.; Bergsagel, P. L.; Ghobrial, I. M.; Richardson, P. G.; Young, R. A.; Hahn, W. C.; Anderson, K. C.; Kung, A. L.; Bradner, J. E.; Mitsiades, C. S. Cell 2011, 146, 904There is no corresponding record for this reference.
- 334Lockwood, W. W.; Zejnullahu, K.; Bradner, J. E.; Varmus, H. Proc. Natl. Acad. Sci. U.S.A. 2012, 109, 19408There is no corresponding record for this reference.
- 335Bandukwala, H. S.; Gagnon, J.; Togher, S.; Greenbaum, J. A.; Lamperti, E. D.; Parr, N. J.; Molesworth, A. M.; Smithers, N.; Lee, K.; Witherington, J.; Tough, D. F.; Prinjha, R. K.; Peters, B.; Rao, A. Proc. Natl. Acad. Sci. U.S.A. 2012, 109, 14532There is no corresponding record for this reference.
- 336Mirguet, O.; Lamotte, Y.; Donche, F.; Toum, J.; Gellibert, F.; Bouillot, A.; Gosmini, R.; Nguyen, V. L.; Delannée, D.; Seal, J.; Blandel, F.; Boullay, A. B.; Boursier, E.; Martin, S.; Brusq, J. M.; Krysa, G.; Riou, A.; Tellier, R.; Costaz, A.; Huet, P.; Dudit, Y.; Trottet, L.; Kirilovsky, J.; Nicodeme, E. Bioorg. Med. Chem. Lett. 2012, 22, 2963There is no corresponding record for this reference.
- 337Bamborough, P.; Diallo, H.; Goodacre, J. D.; Gordon, L.; Lewis, A.; Seal, J. T.; Wilson, D. M.; Woodrow, M. D.; Chung, C. W. J. Med. Chem. 2012, 55, 587There is no corresponding record for this reference.
- 338Zhang, G.; Liu, R.; Zhong, Y.; Plotnikov, A. N.; Zhang, W.; Zeng, L.; Rusinova, E.; Gerona-Nevarro, G.; Moshkina, N.; Joshua, J.; Chuang, P. Y.; Ohlmeyer, M.; He, J. C.; Zhou, M. M. J. Biol. Chem. 2012, 287, 28840– 51There is no corresponding record for this reference.
- 339Matzuk, M. M.; McKeown, M. R.; Filippakopoulos, P.; Li, Q.; Ma, L.; Agno, J. E.; Lemieux, M. E.; Picaud, S.; Yu, R. N.; Qi, J.; Knapp, S.; Bradner, J. E. Cell 2012, 150, 673There is no corresponding record for this reference.
- 340Ferguson, F. M.; Fedorov, O.; Chaikuad, A.; Philpott, M.; Muniz, J. R.; Felletar, I.; von Delft, F.; Heightman, T.; Knapp, S.; Abell, C.; Ciulli, A. J. Med. Chem. 2013, 56, 10183There is no corresponding record for this reference.
- 341Zhang, W.; Prakash, C.; Sum, C.; Gong, Y.; Li, Y.; Kwok, J. J.; Thiessen, N.; Pettersson, S.; Jones, S. J.; Knapp, S.; Yang, H.; Chin, K. C. J. Biol. Chem. 2012, 287, 43137There is no corresponding record for this reference.
- 342Bartholomeeusen, K.; Xiang, Y.; Fujinaga, K.; Peterlin, B. M. J. Biol. Chem. 2012, 287, 36609There is no corresponding record for this reference.
- 343Devaiah, B. N.; Lewis, B. A.; Cherman, N.; Hewitt, M. C.; Albrecht, B. K.; Robey, P. G.; Ozato, K.; Sims, R. J., III; Singer, D. S. Proc. Natl. Acad. Sci. U.S.A. 2012, 109, 6927There is no corresponding record for this reference.
- 344Palermo, R. D.; Webb, H. M.; West, M. J. PLoS Pathog. 2011, 7, e1002334There is no corresponding record for this reference.
- 345Delmore, J. E.; Issa, G. C.; Lemieux, M. E.; Rahl, P. B.; Shi, J.; Jacobs, H. M.; Kastritis, E.; Gilpatrick, T.; Paranal, R. M.; Qi, J.; Chesi, M.; Schinzel, A. C.; McKeown, M. R.; Heffernan, T. P.; Vakoc, C. R.; Bergsagel, P. L.; Ghobrial, I. M.; Richardson, P. G.; Young, R. A.; Hahn, W. C.; Anderson, K. C.; Kung, A. L.; Bradner, J. E.; Mitsiades, C. S. Cell 2011, 146, 904There is no corresponding record for this reference.
- 346Sachchidanand; Resnick-Silverman, L.; Yan, S.; Mutjaba, S.; Liu, W. J.; Zeng, L.; Manfredi, J. J.; Zhou, M. M. Chem. Biol. 2006, 13, 81346Target Structure-Based Discovery of Small Molecules that Block Human p53 and CREB Binding Protein AssociationSachchidanand; Resnick-Silverman, Lois; Yan, Sherry; Mutjaba, Shiraz; Liu, Wen-jun; Zeng, Lei; Manfredi, James J.; Zhou, Ming-MingChemistry & Biology (Cambridge, MA, United States) (2006), 13 (1), 81-90CODEN: CBOLE2; ISSN:1074-5521. (Cell Press)Lysine acetylation of human tumor suppressor p53 in response to cellular stress signals is required for its function as a transcription factor that regulates cell cycle arrest, senescence, or apoptosis. Here, we report small mols. that block lysine 382-acetylated p53 assocn. with the bromodomain of the coactivator CBP, an interaction essential for p53-induced transcription of the cell cycle inhibitor p21 in response to DNA damage. These chems. were discovered in target structure-guided NMR spectroscopy screening of a focused chem. library constructed based on the structural knowledge of CBP bromodomain/p53-AcK382 binding. Structural characterization shows that these chems. inhibit CBP/p53 assocn. by binding to the acetyl-lysine binding site of the bromodomain. Cell-based functional assays demonstrate that the lead chems. can modulate p53 stability and function in response to doxorubicin-induced DNA damage.
- 347Borah, J. C.; Mujtaba, S.; Karakikes, I.; Zeng, L.; Muller, M.; Patel, J.; Moshkina, N.; Morohashi, K.; Zhang, W.; Gerona-Navarro, G.; Hajjar, R. J.; Zhou, M. M. Chem. Biol. 2011, 18, 531347A Small Molecule Binding to the Coactivator CREB-Binding Protein Blocks Apoptosis in CardiomyocytesBorah, Jagat C.; Mujtaba, Shiraz; Karakikes, Ioannis; Zeng, Lei; Muller, Michaela; Patel, Jigneshkumar; Moshkina, Natasha; Morohashi, Keita; Zhang, Weijia; Gerona-Navarro, Guillermo; Hajjar, Roger J.; Zhou, Ming-MingChemistry & Biology (Cambridge, MA, United States) (2011), 18 (4), 531-541CODEN: CBOLE2; ISSN:1074-5521. (Cell Press)Summary: As a master transcription factor in cellular responses to external stress, tumor suppressor p53 is tightly regulated. Excessive p53 activity during myocardial ischemia causes irreversible cellular injury and cardiomyocyte death. P53 activation is dependent on lysine acetylation by the lysine acetyltransferase and transcriptional coactivator CREB-binding protein (CBP) and on acetylation-directed CBP recruitment for p53 target gene expression. Here, we report a small mol. ischemin, developed with a structure-guided approach to inhibit the acetyl-lysine binding activity of the bromodomain of CBP. We show that ischemin alters post-translational modifications on p53 and histones, inhibits p53 interaction with CBP and transcriptional activity in cells, and prevents apoptosis in ischemic cardiomyocytes. Our study suggests small mol. modulation of acetylation-mediated interactions in gene transcription as a new approach to therapeutic interventions of human disorders such as myocardial ischemia.
- 348Gerona-Navarro, G.; Yoel-Rodríguez; Mujtaba, S.; Frasca, A.; Patel, J.; Zeng, L.; Plotnikov, A. N.; Osman, R.; Zhou, M. M. J. Am. Chem. Soc. 2011, 133, 2040There is no corresponding record for this reference.
- 349Rooney, T. P.; Filippakopoulos, P.; Fedorov, O.; Picaud, S.; Cortopassi, W. A.; Hay, D. A.; Martin, S.; Tumber, A.; Rogers, C. M.; Philpott, M.; Wang, M.; Thompson, A. L.; Heightman, T. D.; Pryde, D. C.; Cook, A.; Paton, R. S.; Müller, S.; Knapp, S.; Brennan, P. E.; Conway, S. J. Angew. Chem., Int. Ed. 2014, 53, 6126There is no corresponding record for this reference.
- 350Fedorov, O.; Lingard, H.; Wells, C.; Monteiro, O. P.; Picaud, S.; Keates, T.; Yapp, C.; Philpott, M.; Martin, S. J.; Felletar, I.; Marsden, B. D.; Filippakopoulos, P.; Müller, S.; Knapp, S.; Brennan, P. E. J. Med. Chem. 2014, 57, 462There is no corresponding record for this reference.
- 351Chung, C. W.; Dean, A. W.; Woolven, J. M.; Bamborough, P. J. Med. Chem. 2012, 55, 576There is no corresponding record for this reference.
- 352Hewings, D. S.; Wang, M.; Philpott, M.; Fedorov, O.; Uttarkar, S.; Filippakopoulos, P.; Picaud, S.; Vuppusetty, C.; Marsden, B.; Knapp, S.; Conway, S. J.; Heightman, T. D. J. Med. Chem. 2011, 54, 6761There is no corresponding record for this reference.
- 353Hay, D. A.; Fedorov, O.; Martin, S.; Singleton, D. C.; Tallant, C.; Wells, C.; Picaud, S.; Philpott, M.; Monteiro, O. P.; Rogers, C. M.; Conway, S. J.; Rooney, T. P.; Tumber, A.; Yapp, C.; Filippakopoulos, P.; Bunnage, M. E.; Müller, S.; Knapp, S.; Schofield, C. J.; Brennan, P. E. J. Am. Chem. Soc. 2014, 136, 9308There is no corresponding record for this reference.
- 354Dey, A.; Wong, E. T.; Cheok, C. F.; Tergaonkar, V.; Lane, D. P. Cell Death Differ. 2008, 15, 263354R-Roscovitine simultaneously targets both the p53 and NF-κB pathways and causes potentiation of apoptosis: implications in cancer therapyDey, A.; Wong, E. T.; Cheok, C. F.; Tergaonkar, V.; Lane, D. P.Cell Death and Differentiation (2008), 15 (2), 263-273CODEN: CDDIEK; ISSN:1350-9047. (Nature Publishing Group)Seliciclib (CYC202, R-Roscovitine) is a 2, 6, 9-substituted purine analog that is currently in phase II clin. trials as an anticancer agent. We show in this study that R-Roscovitine can downregulate nuclear factor-kappa B (NF-κB) activation in response to tumor necrosis factor (TNF)α and interleukin 1. Activation of p53-dependent transcription is not compromised when R-Roscovitine is combined with TNFα. We characterize the mol. mechanism governing NF-κB repression and show that R-Roscovitine inhibits the IκB kinase (IKK) kinase activity, which leads to defective IκBα phosphorylation, degrdn. and hence nuclear function of NF-κB. We further show that the downregulation of the NF-κB pathway is also at the level of p65 modification and that the phosphorylation of p65 at Ser 536 is repressed by R-Roscovitine. Consistent with repression of canonical IKK signaling pathway, the induction of NF-κB target genes monocyte chemoattractant protein, intercellular adhesion mol.-1, cyclooxygenase-2 and IL-8 is also inhibited by R-Roscovitine. We further show that treatment of cells with TNFα and R-Roscovitine causes potentiation of cell death. Based on these results, we suggest the potential use of R-Roscovitine as a bitargeted anticancer drug that functions by simultaneously causing p53 activation and NF-κB suppression. This study also provides mechanistic insight into the mol. mechanism of action of R-Roscovitine, thereby possibly explaining its anti-inflammatory properties. Cell Death and Differentiation (2008) 15, 263-273; doi:10.1038/sj.cdd.4402257; published online 2 Nov. 2007.
- 355Li, B. X.; Xiao, X. ChemBioChem 2009, 10, 2721There is no corresponding record for this reference.
- 356Best, J. L.; Amezcua, C. A.; Mayr, B.; Flechner, L.; Murawsky, C. M.; Emerson, B.; Zor, T.; Gardner, K. H.; Montminy, M. Proc. Natl. Acad. Sci. U.S.A. 2004, 101, 17622There is no corresponding record for this reference.
- 357Minter, A. R.; Brennan, B. B.; Mapp, A. K. J. Am. Chem. Soc. 2004, 126, 10504There is no corresponding record for this reference.
- 358Buhrlage, S. J.; Bates, C. A.; Rowe, S. P.; Minter, A. R.; Brennan, B. B.; Majmudar, C. Y.; Wemmer, D. E.; Al-Hashimi, H.; Mapp, A. K. ACS Chem. Biol. 2009, 4, 335There is no corresponding record for this reference.
- 359Bates, C. A.; Pomerantz, W. C.; Mapp, A. K. Biopolymers 2011, 95, 17There is no corresponding record for this reference.
- 360Rowe, S. P.; Casey, R. J.; Brennan, B. B.; Buhrlage, S. J.; Mapp, A. K. J. Am. Chem. Soc. 2007, 129, 10654There is no corresponding record for this reference.
- 361Henderson, W. R., Jr.; Chi, E. Y.; Ye, X.; Nguyen, C.; Tien, Y. T.; Zhou, B.; Borok, Z.; Knight, D. A.; Kahn, M. Proc. Natl. Acad. Sci. U.S.A. 2010, 107, 14309There is no corresponding record for this reference.
- 362Emami, K. H.; Nguyen, C.; Ma, H.; Kim, D. H.; Jeong, K. W.; Eguchi, M.; Moon, R. T.; Teo, J. L.; Kim, H. Y.; Moon, S. H.; Ha, J. R.; Kahn, M. Proc. Natl. Acad. Sci. U.S.A. 2004, 101, 12682There is no corresponding record for this reference.
- 363Sasaki, T.; Hwang, H.; Nguyen, C.; Kloner, R. A.; Kahn, M. PLoS One 2013, 8, e75010There is no corresponding record for this reference.
- 364Hao, S.; He, W.; Li, Y.; Ding, H.; Hou, Y.; Nie, J.; Hou, F. F.; Kahn, M.; Liu, Y. J. Am. Soc. Nephrol. 2011, 22, 1642364Targeted inhibition of β-catenin/CBP signaling ameliorates renal interstitial fibrosisHao, Sha; He, Weichun; Li, Yingjian; Ding, Hong; Hou, Yayi; Nie, Jing; Hou, Fan Fan; Kahn, Michael; Liu, YouhuaJournal of the American Society of Nephrology (2011), 22 (9), 1642-1653CODEN: JASNEU; ISSN:1046-6673. (American Society of Nephrology)Because fibrotic kidneys exhibit aberrant activation of β-catenin signaling, this pathway may be a potential target for antifibrotic therapy. In this study, we examd. the effects of β-catenin activation on tubular epithelial-mesenchymal transition (EMT) in vitro and evaluated the therapeutic efficacy of the peptidomimetic small mol. ICG-001, which specifically disrupts β-catenin-mediated gene transcription, in obstructive nephropathy. In vitro, ectopic expression of stabilized β-catenin in tubular epithelial (HKC-8) cells suppressed E-cadherin and induced Snail1, fibronectin, and plasminogen activator inhibitor-1 (PAI-1) expression. ICG-001 suppressed β-catenin-driven gene transcription in a dose-dependent manner and abolished TGF-β1-induced expression of Snail1, PAI-1, collagen I, fibronectin, and α-smooth muscle actin (α-SMA). This antifibrotic effect of ICG-001 did not involve disruption of Smad signaling. In the unilateral ureteral obstruction model, ICG-001 ameliorated renal interstitial fibrosis and suppressed renal expression of fibronectin, collagen I, collagen III, α-SMA, PAI-1, fibroblast-specific protein-1, Snail1, and Snail2. Late administration of ICG-001 also effectively attenuated fibrotic lesions in obstructive nephropathy. In conclusion, inhibiting β-catenin signaling may be an effective approach to the treatment of fibrotic kidney diseases.
- 365Majmudar, C. Y.; Højfeldt, J. W.; Arevang, C. J.; Pomerantz, W. C.; Gagnon, J. K.; Schultz, P. J.; Cesa, L. C.; Doss, C. H.; Rowe, S. P.; Vásquez, V.; Tamayo-Castillo, G.; Cierpicki, T.; Brooks, C. L., III; Sherman, D. H.; Mapp, A. K. Angew. Chem., Int. Ed. 2012, 51, 11258There is no corresponding record for this reference.
- 366Kung, A. L.; Zabludoff, S. D.; France, D. S.; Freedman, S. J.; Tanner, E. A.; Vieira, A.; Cornell-Kennon, S.; Lee, J.; Wang, B.; Wang, J.; Memmert, K.; Naegeli, H. U.; Petersen, F.; Eck, M. J.; Bair, K. W.; Wood, A. W.; Livingston, D. M. Cancer Cell 2004, 6, 33366Small molecule blockade of transcriptional coactivation of the hypoxia-inducible factor pathwayKung, Andrew L.; Zabludoff, Sonya D.; France, Dennis S.; Freedman, Steven J.; Tanner, Elizabeth A.; Vieira, Annelisa; Cornell-Kennon, Susan; Lee, Jennifer; Wang, Beqing; Wang, Jamin; Memmert, Klaus; Naegeli, Hans-Ulrich; Petersen, Frank; Eck, Michael J.; Bair, Kenneth W.; Wood, Alexander W.; Livingston, David M.Cancer Cell (2004), 6 (1), 33-43CODEN: CCAECI; ISSN:1535-6108. (Cell Press)Homeostasis under hypoxic conditions is maintained through a coordinated transcriptional response mediated by the hypoxia-inducible factor (HIF) pathway and requires coactivation by the CBP and p300 transcriptional coactivators. Through a target-based high-throughput screen, the authors identified chetomin as a disrupter of HIF binding to p300. At a mol. level, chetomin disrupts the structure of the CH1 domain of p300 and precludes its interaction with HIF, thereby attenuating hypoxia-inducible transcription. Systemic administration of chetomin inhibited hypoxia-inducible transcription within tumors and inhibited tumor growth. These results demonstrate a therapeutic window for pharmacol. attenuation of HIF activity and further establish the feasibility of disrupting a signal transduction pathway by targeting the function of a transcriptional coactivator with a small mol.
- 367Block, K. M.; Wang, H.; Szabó, L. Z.; Polaske, N. W.; Henchey, L. K.; Dubey, R.; Kushal, S.; László, C. F.; Makhoul, J.; Song, Z.; Meuillet, E. J.; Olenyuk, B. Z. J. Am. Chem. Soc. 2009, 131, 18078There is no corresponding record for this reference.
- 368Yin, S.; Kaluz, S.; Devi, N. S.; Jabbar, A. A.; de Noronha, R. G.; Mun, J.; Zhang, Z.; Boreddy, P. R.; Wang, W.; Wang, Z.; Abbruscato, T.; Chen, Z.; Olson, J. J.; Zhang, R.; Goodman, M. M.; Nicolaou, K. C.; Van Meir, E. G. Clin. Cancer Res. 2012, 18, 6623368Arylsulfonamide KCN1 Inhibits In Vivo Glioma Growth and Interferes with HIF Signaling by Disrupting HIF-1α Interaction with Cofactors p300/CBPYin, Shaoman; Kaluz, Stefan; Devi, Narra S.; Jabbar, Adnan A.; de Noronha, Rita G.; Mun, Jiyoung; Zhang, Zhaobin; Boreddy, Purushotham R.; Wang, Wei; Wang, Zhibo; Abbruscato, Thomas; Chen, Zhengjia; Olson, Jeffrey J.; Zhang, Ruiwen; Goodman, Mark M.; Nicolaou, K. C.; Van Meir, Erwin G.Clinical Cancer Research (2012), 18 (24), 6623-6633CODEN: CCREF4; ISSN:1078-0432. (American Association for Cancer Research)Purpose: The hypoxia-inducible factor-1 (HIF-1) plays a crit. role in tumor adaptation to hypoxia, and its elevated expression correlates with poor prognosis and treatment failure in patients with cancer. In this study, we detd. whether 3,4-dimethoxy-N-[(2,2-dimethyl-2H-chromen-6-yl)methyl]-N-phenylbenzenesulfonamide, KCN1, the lead inhibitor in a novel class of arylsulfonamide inhibitors of the HIF-1 pathway, had antitumorigenic properties in vivo and further defined its mechanism of action. Exptl. Design: We studied the inhibitory effect of systemic KCN1 delivery on the growth of human brain tumors in mice. To define mechanisms of KCN1 anti-HIF activities, we examd. its influence on the assembly of a functional HIF-1α/HIF-1β/p300 transcription complex. Results: KCN1 specifically inhibited HIF reporter gene activity in several glioma cell lines at the nanomolar level. KCN1 also downregulated transcription of endogenous HIF-1 target genes, such as VEGF, Glut-1, and carbonic anhydrase 9, in a hypoxia-responsive element (HRE)-dependent manner. KCN1 potently inhibited the growth of s.c. malignant glioma tumor xenografts with minimal adverse effects on the host. It also induced a temporary survival benefit in an intracranial model of glioma but had no effect in a model of melanoma metastasis to the brain. Mechanistically, KCN1 did not downregulate the levels of HIF-1α or other components of the HIF transcriptional complex; rather, it antagonized hypoxia-inducible transcription by disrupting the interaction of HIF-1α with transcriptional coactivators p300/CBP. Conclusions: Our results suggest that the new HIF pathway inhibitor KCN1 has antitumor activity in mouse models, supporting its further translation for the treatment of human tumors displaying hypoxia or HIF overexpression.
- 369Tanaka, Y.; Naruse, I.; Maekawa, T.; Masuya, H.; Shiroishi, T.; Ishii, S. Proc. Natl. Acad. Sci. U.S.A. 1997, 94, 10215There is no corresponding record for this reference.
- 370Yao, T. P.; Oh, S. P.; Fuchs, M.; Zhou, N. D.; Ch’ng, L. E.; Newsome, D.; Bronson, R. T.; Li, E.; Livingston, D. M.; Eckner, R. Cell 1998, 93, 361There is no corresponding record for this reference.
- 371Roelfsema, J. H.; Peters, D. J. Expert Rev. Mol. Med. 2007, 9, 1371Rubinstein-Taybi syndrome: clinical and molecular overviewRoelfsema Jeroen H; Peters Dorien J MExpert reviews in molecular medicine (2007), 9 (23), 1-16 ISSN:.Rubinstein-Taybi syndrome is characterised by mental retardation, growth retardation and a particular dysmorphology. The syndrome is rare, with a frequency of approximately one affected individual in 100,000 newborns. Mutations in two genes - CREBBP and EP300 - have been identified to cause the syndrome. These two genes show strong homology and encode histone acetyltransferases (HATs), which are transcriptional co-activators involved in many signalling pathways. Loss of HAT activity is sufficient to account for the phenomena seen in Rubinstein-Taybi patients. Although some mutations found in CREBBP are translocations, inversions and large deletions, most are point mutations or small deletions and insertions. Mutations in EP300 are comparatively rare. Extensive screening of patients has revealed mutations in CREBBP and EP300 in around 50% of cases. The cause of the syndrome in the remaining patients remains to be identified, but other genes could also be involved. Here, we describe the clinical presentation of Rubinstein-Taybi syndrome, review the mutation spectrum and discuss the current understanding of causative molecular mechanisms.
- 372Qaksen, H.; Bartsch, O.; Okur, M.; Temel, H.; Açikgoz, M.; Yilmaz, C. Genet. Couns. 2009, 20, 255There is no corresponding record for this reference.
- 373Viosca, J.; Lopez-Atalaya, J. P.; Olivares, R.; Eckner, R.; Barco, A. Neurobiol. Dis. 2010, 37, 186373Syndromic features and mild cognitive impairment in mice with genetic reduction on p300 activity: Differential contribution of p300 and CBP to Rubinstein-Taybi syndrome etiologyViosca Jose; Lopez-Atalaya Jose P; Olivares Roman; Eckner Richard; Barco AngelNeurobiology of disease (2010), 37 (1), 186-94 ISSN:.Rubinstein-Taybi syndrome (RSTS) is a complex autosomal-dominant disease characterized by mental and growth retardation and skeletal abnormalities. A majority of the individuals diagnosed with RSTS carry heterozygous mutation in the gene CREBBP, but a small percentage of cases are caused by mutations in EP300. To investigate the contribution of p300 to RSTS pathoetiology, we carried out a comprehensive and multidisciplinary characterization of p300(+/-) mice. These mice exhibited facial abnormalities and impaired growth, two traits associated to RSTS in humans. We also observed abnormal gait, reduced swimming speed, enhanced anxiety in the elevated plus maze, and mild cognitive impairment during the transfer task in the water maze. These analyses demonstrate that p300(+/-) mice exhibit phenotypes that are reminiscent of neurological traits observed in RSTS patients, but their comparison with previous studies on CBP deficient strains also indicates that, in agreement with the most recent findings in human patients, the activity of p300 in cognition is likely less relevant or more susceptible to compensation than the activity of CBP.
- 374Roelfsema, J. H.; Peters, D. J. Expert Rev. Mol. Med. 2007, 9, 1374Rubinstein-Taybi syndrome: clinical and molecular overviewRoelfsema Jeroen H; Peters Dorien J MExpert reviews in molecular medicine (2007), 9 (23), 1-16 ISSN:.Rubinstein-Taybi syndrome is characterised by mental retardation, growth retardation and a particular dysmorphology. The syndrome is rare, with a frequency of approximately one affected individual in 100,000 newborns. Mutations in two genes - CREBBP and EP300 - have been identified to cause the syndrome. These two genes show strong homology and encode histone acetyltransferases (HATs), which are transcriptional co-activators involved in many signalling pathways. Loss of HAT activity is sufficient to account for the phenomena seen in Rubinstein-Taybi patients. Although some mutations found in CREBBP are translocations, inversions and large deletions, most are point mutations or small deletions and insertions. Mutations in EP300 are comparatively rare. Extensive screening of patients has revealed mutations in CREBBP and EP300 in around 50% of cases. The cause of the syndrome in the remaining patients remains to be identified, but other genes could also be involved. Here, we describe the clinical presentation of Rubinstein-Taybi syndrome, review the mutation spectrum and discuss the current understanding of causative molecular mechanisms.
- 375Gervasini, C.; Mottadelli, F.; Ciccone, R.; Castronovo, P.; Milani, D.; Scarano, G.; Bedeschi, M. F.; Belli, S.; Pilotta, A.; Selicorni, A.; Zuffardi, O.; Larizza, L. Eur. J. Hum. Genet. 2010, 18, 768375High frequency of copy number imbalances in Rubinstein-Taybi patients negative to CREBBP mutational analysisGervasini, Cristina; Mottadelli, Federica; Ciccone, Roberto; Castronovo, Paola; Milani, Donatella; Scarano, Gioacchino; Bedeschi, Maria Francesca; Belli, Serena; Pilotta, Alba; Selicorni, Angelo; Zuffardi, Orsetta; Larizza, LidiaEuropean Journal of Human Genetics (2010), 18 (7), 768-775CODEN: EJHGEU; ISSN:1018-4813. (Nature Publishing Group)Rubinstein-Taybi syndrome (RSTS) is a rare autosomal dominant disorder characterized by facial dysmorphisms, growth and psychomotor development delay, and skeletal defects. The known genetic causes are point mutations or deletions of the CREBBP (50-60%) and EP300 (5%) genes. To detect chromosomal rearrangements indicating novel positional candidate RSTS genes, we used a-CGH to study 26 patients fulfilling the diagnostic criteria for RSTS who were neg. at fluorescence in situ hybridization analyses of the CREBBP and EP300 regions, and direct sequencing analyses of the CREBBP gene. We found seven imbalances (27%): Four de novo and three inherited rearrangements not reported among the copy no. variants. A de novo 7p21.1 deletion of 500 kb included the TWIST1 gene, a suggested candidate for RSTS that is responsible for the Saethre-Chotzen syndrome, an entity that enters in differential diagnosis with RSTS. A similar issue of differential diagnosis was raised by a large 4.3 Mb 2q22.3q23.1 deletion encompassing ZEB2, the gene responsible for the Mowat-Wilson syndrome, whose signs may overlap with RSTS. Positional candidate genes could not be sought in the remaining pathogenetic imbalances, because of the size of the involved region (a 9 Mb 2q24.3q31.1 deletion) and/or the relative paucity of suitable genes (a 5 Mb 3p13p12.3 duplication). One of the inherited rearrangements, the 17q11.2 379Kb duplication, represents the reciprocal event of the deletion underlying an overgrowth syndrome, both being mediated by the NF1-REP-P1 and REP-P2 sub-duplicons. The contribution of this and the other detected CNVs to the clin. RSTS phenotype is difficult to assess.
- 376Foley, P.; Bunyan, D.; Stratton, J.; Dillon, M.; Lynch, S. A. Am. J. Med. Genet., Part A 2009, 149A, 997376Further case of Rubinstein-Taybi syndrome due to a deletion in EP300Foley, Patricia; Bunyan, David; Stratton, John; Dillon, Michelle; Lynch, Sally AnnAmerican Journal of Medical Genetics, Part A (2009), 149A (5), 997-1000CODEN: AJMGB8; ISSN:1552-4825. (Wiley-Liss, Inc.)Rubinstein-Taybi syndrome (RSTS) is a heterogeneous disorder with approx. 45-55% of patients showing mutations in the CREB binding protein and a further 3% of patients having mutations in EP300. We report a male child with a deletion of exons 3-8 of the EP300 gene who has RSTS. He has a milder skeletal phenotype, a finding that has been described in other cases with EP300 mutations. The mother suffered from preeclampsia and HELLP syndrome in the pregnancy. She subsequently developed a mullerian tumor of her cervix 6 years after the birth of her son.
- 377Bartholdi, D.; Roelfsema, J. H.; Papadia, F.; Breuning, M. H.; Niedrist, D.; Hennekam, R. C.; Schinzel, A.; Peters, D. J. J. Med. Genet. 2007, 44, 327377Genetic heterogeneity in Rubinstein-Taybi syndrome: delineation of the phenotype of the first patients carrying mutations in EP300Bartholdi, Deborah; Roelfsema, Jeroen H.; Papadia, Francesco; Breuning, Martijn H.; Niedrist, Dunja; Hennekam, Raoul C.; Schinzel, Albert; Peters, Dorien J. M.Journal of Medical Genetics (2007), 44 (5), 327-333CODEN: JMDGAE; ISSN:0022-2593. (BMJ Publishing Group)Rubinstein-Taybi syndrome (RSTS) is a congenital disorder characterised by growth retardation, facial dysmorphisms, skeletal abnormalities and mental retardation. Broad thumbs and halluces are the hallmarks of the syndrome. RSTS is assocd. with chromosomal rearrangements and mutations in the CREB-binding protein gene (CREBBP), also termed CBP, encoding the CREB-binding protein. Recently, it was shown that mutations in EP300, coding for the p300 protein, also cause RSTS. CBP and EP300 are highly homologous genes, which play important roles as global transcriptional coactivators. To report the phenotype of the presently known patients with RSTS (n = 4) carrying germline mutations of EP300. The patients with EP300 mutations displayed the typical facial gestalt and malformation pattern compatible with the diagnosis of RSTS. However, three patients exhibited much milder skeletal findings on the hands and feet than typically obsd. in patients with RSTS. Part of the clin. variability in RSTS is explained by genetic heterogeneity. The diagnosis of RSTS must be expanded to include patients without broad thumbs or halluces.
- 378Tsai, A. C.; Dossett, C. J.; Walton, C. S.; Cramer, A. E.; Eng, P. A.; Nowakowska, B. A.; Pursley, A. N.; Stankiewicz, P.; Wiszniewska, J.; Cheung, S. W. Eur. J. Hum. Genet. 2011, 19, 43378Exon deletions of the EP300 and CREBBP genes in two children with Rubinstein-Taybi syndrome detected by aCGHTsai Anne Chun-Hui; Dossett Cherilyn J; Walton Carol S; Cramer Andrea E; Eng Patti A; Nowakowska Beata A; Pursley Amber N; Stankiewicz Pawel; Wiszniewska Joanna; Cheung Sau WaiEuropean journal of human genetics : EJHG (2011), 19 (1), 43-9 ISSN:.We demonstrate the utility of an exon coverage microarray platform in detecting intragenic deletions: one in exons 24-27 of the EP300 gene and another in exons 27 and 28 of the CREBBP gene in two patients with Rubinstein-Taybi syndrome (RSTS). RSTS is a heterogeneous disorder in which approximately 45-55% of cases result from deletion or mutations in the CREBBP gene and an unknown portion of cases result from gene changes in EP300. The first case is a 3-year-old female with an exonic deletion of the EP300 gene who has classic facial features of RSTS without the thumb and great toe anomalies, consistent with the milder skeletal phenotype that has been described in other RSTS cases with EP300 mutations. In addition, the mother of this patient also had preeclampsia during pregnancy, which has been infrequently reported. The second case is a newborn male who has the classical features of RSTS. Our results illustrate that exon-targeted array comparative genomic hybridization (aCGH) is a powerful tool for detecting clinically significant intragenic rearrangements that would be otherwise missed by aCGH platforms lacking sufficient exonic coverage or sequencing of the gene of interest.
- 379Zimmermann, N.; Acosta, A. M.; Kohlhase, J.; Bartsch, O. Eur. J. Hum. Genet. 2007, 15, 837379Confirmation of EP300 gene mutations as a rare cause of Rubinstein-Taybi syndromeZimmermann, Nicole; Acosta, Ana Maria Bravo Ferrer; Kohlhase, Juergen; Bartsch, OliverEuropean Journal of Human Genetics (2007), 15 (8), 837-842CODEN: EJHGEU; ISSN:1018-4813. (Nature Publishing Group)The Rubinstein-Taybi syndrome (RSTS, MIM 180849), a dominant Mendelian disorder with typical face, short stature, skeletal abnormalities, and mental retardation, is usually caused by heterozygous mutations of the CREBBP gene, but recently, EP300 gene mutations were reported in three individuals. Using quant. PCR (for the CREBBP and EP300 genes) and genomic sequencing (for the EP300 gene), we studied here 13 patients who had shown no mutation after genomic sequencing of the CREBBP gene in a previous investigation. Two new disease-causing mutations were identified, including a partial deletion of CREBBP and a 1-bp deletion in EP300, c.7100delC (p.P2366fsX2401). The 1-bp deletion represents the fourth EP300 mutation reported to date and was identified in a patient with non-classical RSTS. Based on the very similar structure of the CREBBP and EP300 genes and the higher rate of single-nucleotide polymorphisms in EP300 (2.23 per individual) as compared to CREBBP (0.71 per individual) (P>0.001, Wilcoxon test), it may be assumed that EP300 gene mutations should be as frequent as CREBBP gene mutations. Based on the location of the EP300 gene mutations identified so far (outside the histone acetyl transferase domain) and the obsd. (although not very striking) phenotypical differences with the EP300 mutations, we propose that most EP300 mutations could be assocd. with other phenotypes, not classical RSTS.
- 380Bartholdi, D.; Roelfsema, J. H.; Papadia, F.; Breuning, M. H.; Niedrist, D.; Hennekam, R. C.; Schinzel, A.; Peters, D. J. J. Med. Genet. 2007, 44, 327380Genetic heterogeneity in Rubinstein-Taybi syndrome: delineation of the phenotype of the first patients carrying mutations in EP300Bartholdi, Deborah; Roelfsema, Jeroen H.; Papadia, Francesco; Breuning, Martijn H.; Niedrist, Dunja; Hennekam, Raoul C.; Schinzel, Albert; Peters, Dorien J. M.Journal of Medical Genetics (2007), 44 (5), 327-333CODEN: JMDGAE; ISSN:0022-2593. (BMJ Publishing Group)Rubinstein-Taybi syndrome (RSTS) is a congenital disorder characterised by growth retardation, facial dysmorphisms, skeletal abnormalities and mental retardation. Broad thumbs and halluces are the hallmarks of the syndrome. RSTS is assocd. with chromosomal rearrangements and mutations in the CREB-binding protein gene (CREBBP), also termed CBP, encoding the CREB-binding protein. Recently, it was shown that mutations in EP300, coding for the p300 protein, also cause RSTS. CBP and EP300 are highly homologous genes, which play important roles as global transcriptional coactivators. To report the phenotype of the presently known patients with RSTS (n = 4) carrying germline mutations of EP300. The patients with EP300 mutations displayed the typical facial gestalt and malformation pattern compatible with the diagnosis of RSTS. However, three patients exhibited much milder skeletal findings on the hands and feet than typically obsd. in patients with RSTS. Part of the clin. variability in RSTS is explained by genetic heterogeneity. The diagnosis of RSTS must be expanded to include patients without broad thumbs or halluces.
- 381Lopez-Atalaya, J. P.; Gervasini, C.; Mottadelli, F.; Spena, S.; Piccione, M.; Scarano, G.; Selicorni, A.; Barco, A.; Larizza, L. J. Med. Genet. 2012, 49, 66381Histone acetylation deficits in lymphoblastoid cell lines from patients with Rubinstein-Taybi syndromeLopez-Atalaya, J. P.; Gervasini, C.; Mottadelli, F.; Spena, S.; Piccione, M.; Scarano, G.; Selicorni, A.; Barco, A.; Larizza, L.Journal of Medical Genetics (2012), 49 (1), 66-74CODEN: JMDGAE; ISSN:0022-2593. (BMJ Publishing Group)Background: Rubinstein-Taybi syndrome (RSTS) is a congenital neurodevelopmental disorder defined by postnatal growth deficiency, characteristic skeletal abnormalities and mental retardation and caused by mutations in the genes encoding for the transcriptional co-activators with intrinsic lysine acetyltransferase (KAT) activity CBP and p300. Previous studies have shown that neuronal histone acetylation is reduced in mouse models of RSTS. Methods: The authors identified different mutations at the CREBBP locus and generated lymphoblastoid cell lines derived from nine patients with RSTS carrying distinct CREBBP mutations that illustrate different grades of the clin. severity in the spectrum of the syndrome. They next assessed whether histone acetylation levels were altered in these cell lines. Results: The comparison of CREBBP-mutated RSTS cell lines with cell lines derived from patients with an unrelated mental retardation syndrome or healthy controls revealed significant deficits in histone acetylation, affecting primarily histone H2B and histone H2A. The most severe defects were obsd. in the lines carrying the whole deletion of the CREBBP gene and the truncating mutation, both leading to a haploinsufficiency state. Interestingly, this deficit was rescued by treatment with an inhibitor of histone deacetylases (HDACi). Conclusions: The authors' results extend to humans the seminal observations in RSTS mouse models and point to histone acetylation defects, mainly involving H2B and H2A, as relevant mol. markers of the disease.
- 382Miller, R. W.; Rubinstein, J. H. Am. J. Med. Genet. 1995, 56, 112382Tumors in Rubinstein-Taybi syndromeMiller R W; Rubinstein J HAmerican journal of medical genetics (1995), 56 (1), 112-5 ISSN:0148-7299.The 14 tumors reported in Rubinstein-Taybi syndrome since 1989, when added to the 22 previously reported, are beginning to show a pattern of neural and developmental tumors, especially of the head, which is malformed in the syndrome. Among the neoplasms were 12 of the nervous system: 2 each of oligodendroglioma, medulloblastoma, neuroblastoma, and benign meningioma, a pheochromocytoma, and 3 other benign tumors; 2 of nasopharyngeal rhabdomyosarcoma; and 1 each of leiomyosarcoma, seminoma, and embryonal carcinoma. Among the other benign tumors were an odontoma, a choristoma, a dermoid cyst, and 2 pilomatrixomas.
- 383Iyer, N. G.; Ozdag, H.; Caldas, C. Oncogene 2004, 23, 4225383p300/CBP and cancerIyer, Narayanan Gopalakrishna; Oezdag, Hilal; Caldas, CarlosOncogene (2004), 23 (24), 4225-4231CODEN: ONCNES; ISSN:0950-9232. (Nature Publishing Group)A review. P300 and cAMP response element-binding protein (CBP) are adenoviral E1A-binding proteins involved in multiple cellular processes, and function as transcriptional co-factors and histone acetyltransferases. Germline mutation of CBP results in Rubinstein-Taybi syndrome, which is characterized by an increased predisposition to childhood malignancies. Furthermore, somatic mutations of p300 and CBP occur in a no. of malignancies. Chromosome translocations target CBP and, less commonly, p300 in acute myeloid leukemia and treatment-related hematol. disorders. P300 mutations in solid tumors result in truncated p300 protein products or amino-acid substitutions in crit. protein domains, and these are often assocd. with inactivation of the second allele. A mouse model confirms that p300 and CBP function as suppressors of hematol. tumor formation. The involvement of these proteins in crit. tumorigenic pathways (including TGF-β, p53 and Rb) provides a mechanistic route as to how their inactivation could result in cancer.
- 384Tillinghast, G. W.; Partee, J.; Albert, P.; Kelley, J. M.; Burtow, K. H.; Kelly, K. Genes, Chromosomes Cancer 2003, 37, 121384Analysis of genetic stability at the EP300 and CREBBP loci in a panel of cancer cell linesTillinghast, Guy W.; Partee, Jason; Albert, Paul; Kelley, Jenny M.; Burtow, Kenneth H.; Kelly, KathleenGenes, Chromosomes & Cancer (2003), 37 (2), 121-131CODEN: GCCAES; ISSN:1045-2257. (Wiley-Liss, Inc.)EP300 (p300) and CREBBP (CBP) are highly related transcriptional co-activators possessing histone acetyltransferase activity. These proteins have been implicated in coordinating numerous transcriptional responses that are important in the processes of proliferation and differentiation. A role for EP300 and CREBBP as tumor suppressors in cancer has been suggested by the fact that they are targeted by viral oncogenes; there is an increased incidence of hematol. malignancies in mice monoallelic for CREBBP; and loss, albeit at a low frequency, of both EP300 alleles in epithelial cancers has been obsd. Because the level of EP300/CREBBP appears to have a crit. effect on integrating certain transcriptional processes, we sought to det. whether a loss in the combined gene dosage of EP300 and CREBBP might play a role in cancer. Accordingly, we screened a panel of 103 cell lines for loss of heterozygosity and found 35 and 51% LOH for the CREBBP and EP300 loci, resp. Concordant loss of CREBBP and EP300 was not assocd. with mutations in important regions of the remaining EP300 or CREBBP genes. In addn., there did not appear to be a statistically significant selection in cancer cells, stratified by various criteria, for the concordant loss of EP300 and CREBBP. We conclude that EP300 and CREBBP rarely act as classical tumor suppressors in human cancer.
- 385Bryan, E. J.; Jokubaitis, V. J.; Chamberlain, N. L.; Baxter, S. W.; Dawson, E.; Choong, D. Y.; Campbell, I. G. Int. J. Cancer 2002, 102, 137385Mutation analysis of EP300 in colon, breast and ovarian carcinomasBryan, Emma J.; Jokubaitis, Venta J.; Chamberlain, Narelle L.; Baxter, Simon W.; Dawson, Elisabeth; Choong, David Y. H.; Campbell, Ian G.International Journal of Cancer (2002), 102 (2), 137-141CODEN: IJCNAW; ISSN:0020-7136. (Wiley-Liss, Inc.)The putative tumor suppressor gene EP300 is located on chromosome 22q13 which is a region showing frequent loss of heterozygosity (LOH) in colon, breast and ovarian cancers. The authors analyzed 203 human breast, colon and ovarian primary tumors and cell lines for somatic mutations in EP300. LOH across the EP300 locus was detected in 38% of colon, 36% of breast, and 49% of ovarian primary tumors but no somatic mutations in EP300 were identified in any primary tumor. Anal. of 17 colon, 11 breast, and 11 ovarian cancer cell lines identified truncating mutations in 4 colon cancer cell lines (HCT116, HT29, LIM2405 and LIM2412). The authors confirmed the presence of a previously reported frameshift mutation in HCT116 at codon 1699 and identified a second frameshift mutation at codon 1468. Bi-allelic inactivation of EP300 was also detected in LIM2405 that harbors an insC mutation at codon 927 as well an insA mutation at codon 1468. An insA mutation at codon 1468 was identified in HT29 and a CGA>TGA mutation at codon 86 was identified in LIM2412. Both these lines were heterozygous across the EP300 locus and western blot anal. confirmed the presence of an apparently wild-type protein. The authors' study has established that genetic inactivation of EP300 is rare in primary colorectal, breast and ovarian cancers. In contrast, mutations are common among colorectal cancer cell lines with 4/17 harboring homozygous or heterozygous mutations. The rarity of EP300 mutations among these tumor types that show a high frequency of LOH across 22q13 may indicate that another gene is the target of the loss. It is possible that bi-allelic inactivation of EP300 is not necessary and that haploinsufficiency is sufficient to promote tumorigenesis. Alternatively, silencing of EP300 may be achieved by epigenetic mechanisms such as promoter methylation.
- 386Muraoka, M.; Konishi, M.; Kikuchi-Yanoshita, R.; Tanaka, K.; Shitara, N.; Chong, J. M.; Iwama, T.; Miyaki, M. Oncogene 1996, 12, 1565There is no corresponding record for this reference.
- 387Iyer, N. G.; Ozdag, H.; Caldas, C. Oncogene 2004, 23, 4225387p300/CBP and cancerIyer, Narayanan Gopalakrishna; Oezdag, Hilal; Caldas, CarlosOncogene (2004), 23 (24), 4225-4231CODEN: ONCNES; ISSN:0950-9232. (Nature Publishing Group)A review. P300 and cAMP response element-binding protein (CBP) are adenoviral E1A-binding proteins involved in multiple cellular processes, and function as transcriptional co-factors and histone acetyltransferases. Germline mutation of CBP results in Rubinstein-Taybi syndrome, which is characterized by an increased predisposition to childhood malignancies. Furthermore, somatic mutations of p300 and CBP occur in a no. of malignancies. Chromosome translocations target CBP and, less commonly, p300 in acute myeloid leukemia and treatment-related hematol. disorders. P300 mutations in solid tumors result in truncated p300 protein products or amino-acid substitutions in crit. protein domains, and these are often assocd. with inactivation of the second allele. A mouse model confirms that p300 and CBP function as suppressors of hematol. tumor formation. The involvement of these proteins in crit. tumorigenic pathways (including TGF-β, p53 and Rb) provides a mechanistic route as to how their inactivation could result in cancer.
- 388Pasqualucci, L.; Dominguez-Sola, D.; Chiarenza, A.; Fabbri, G.; Grunn, A.; Trifonov, V.; Kasper, L. H.; Lerach, S.; Tang, H.; Ma, J.; Rossi, D.; Chadburn, A.; Murty, V. V.; Mullighan, C. G.; Gaidano, G.; Rabadan, R. Nature 2011, 471, 189There is no corresponding record for this reference.
- 389Kishimoto, M.; Kohno, T.; Okudela, K.; Otsuka, A.; Sasaki, H.; Tanabe, C.; Sakiyama, T.; Hirama, C.; Kitabayashi, I.; Minna, J. D.; Takenoshita, S.; Yokota, J. Clin. Cancer Res. 2005, 11, 512389Mutations and deletions of the CBP gene in human lung cancerKishimoto, Masahiro; Kohno, Takashi; Okudela, Koji; Otsuka, Ayaka; Sasaki, Hiroki; Tanabe, Chikako; Sakiyama, Tokuki; Hirama, Chie; Kitabayashi, Issay; Minna, John D.; Takenoshita, Seiichi; Yokota, JunClinical Cancer Research (2005), 11 (2, Pt. 1), 512-519CODEN: CCREF4; ISSN:1078-0432. (American Association for Cancer Research)Microarray-based comparative genomic hybridization anal. led us to detect a homozygous deletion at the cAMP response element binding protein-binding protein (CBP) locus in a lung cancer cell line. Oncogenic roles of CBP had been suggested by functional and genetic studies; thus, involvement of CBP gene alterations in lung carcinogenesis was investigated by undertaking comprehensive anal. of genetic CBP alterations in human lung cancer. Fifty-nine cell lines and 95 surgical specimens of lung cancer were analyzed for mutations, homozygous and hemizygous deletions, and expression of the CBP gene. Homozygous CBP deletions, including two intragenic deletions, were detected in three (5.1%) lung cancer cell lines. CBP mutations, including missense, nonsense, and frame-shift mutations, were detected in six (10.2 %) cell lines and five (5.3%) surgical specimens of lung cancer. The wild-type CBP allele was retained in 9 of 11 cases with CBP mutations, and both the wild-type and mutant alleles were expressed in all the six cases with heterozygous CBP mutations examd. Three mutations with amino acid substitutions in the histone acetyltransferase domain caused significant redn. in transcription activation activity of CBP protein in vivo. A fraction of lung cancers carried mutations and/or deletions of the CBP gene, suggesting that genetic CBP alterations are involved in the genesis and/or progression of a subset of lung cancers.
- 390Suganuma, T.; Kawabata, M.; Ohshima, T.; Ikeda, M. A. Proc. Natl. Acad. Sci. U.S.A. 2002, 99, 13073There is no corresponding record for this reference.
- 391Mullighan, C. G.; Zhang, J.; Kasper, L. H.; Lerach, S.; Payne-Turner, D.; Phillips, L. A.; Heatley, S. L.; Holmfeldt, L.; Collins-Underwood, J. R.; Ma, J.; Buetow, K. H.; Pui, C. H.; Baker, S. D. Nature 2011, 471, 235There is no corresponding record for this reference.
- 392Sánchez-Molina, S.; Oliva, J. L.; García-Vargas, S.; Valls, E.; Rojas, J. M.; Martínez-Balbás, M. A. Biochem. J. 2006, 398, 215There is no corresponding record for this reference.
- 393Iyer, N. G.; Ozdag, H.; Caldas, C. Oncogene 2004, 23, 4225393p300/CBP and cancerIyer, Narayanan Gopalakrishna; Oezdag, Hilal; Caldas, CarlosOncogene (2004), 23 (24), 4225-4231CODEN: ONCNES; ISSN:0950-9232. (Nature Publishing Group)A review. P300 and cAMP response element-binding protein (CBP) are adenoviral E1A-binding proteins involved in multiple cellular processes, and function as transcriptional co-factors and histone acetyltransferases. Germline mutation of CBP results in Rubinstein-Taybi syndrome, which is characterized by an increased predisposition to childhood malignancies. Furthermore, somatic mutations of p300 and CBP occur in a no. of malignancies. Chromosome translocations target CBP and, less commonly, p300 in acute myeloid leukemia and treatment-related hematol. disorders. P300 mutations in solid tumors result in truncated p300 protein products or amino-acid substitutions in crit. protein domains, and these are often assocd. with inactivation of the second allele. A mouse model confirms that p300 and CBP function as suppressors of hematol. tumor formation. The involvement of these proteins in crit. tumorigenic pathways (including TGF-β, p53 and Rb) provides a mechanistic route as to how their inactivation could result in cancer.
- 394Fan, S.; Ma, Y. X.; Wang, C.; Yuan, R. Q.; Meng, Q.; Wang, J. A.; Erdos, M.; Goldberg, I. D.; Webb, P.; Kushner, P. J.; Pestell, R. G.; Rosen, E. M. Cancer Res. 2002, 62, 141There is no corresponding record for this reference.
- 395Liang, J.; Prouty, L.; Williams, B. J.; Dayton, M. A.; Blanchard, K. L. Blood 1998, 92, 2118There is no corresponding record for this reference.
- 396Carapeti, M.; Aguiar, R. C.; Goldman, J. M.; Cross, N. C. Blood 1998, 91, 3127There is no corresponding record for this reference.
- 397Kindle, K. B.; Troke, P. J.; Collins, H. M.; Matsuda, S.; Bossi, D.; Bellodi, C.; Kalkhoven, E.; Salomoni, P.; Pelicci, P. G.; Minucci, S.; Heery, D. M. Mol. Cell. Biol. 2005, 25, 988397MOZ-TIF2 inhibits transcription by nuclear receptors and p53 by impairment of CBP functionKindle, Karin B.; Troke, Philip J. F.; Collins, Hilary M.; Matsuda, Sachiko; Bossi, Daniela; Bellodi, Cristian; Kalkhoven, Eric; Salomoni, Paolo; Pelicci, Pier Giuseppe; Minucci, Saverio; Heery, David M.Molecular and Cellular Biology (2005), 25 (3), 988-1002CODEN: MCEBD4; ISSN:0270-7306. (American Society for Microbiology)Chromosomal rearrangements assocd. with acute myeloid leukemia (AML) include fusions of the genes encoding the acetyltransferase MOZ or MORF with genes encoding the nuclear receptor coactivator TIF2, p300, or CBP. Here we show that MOZ-TIF2 acts as a dominant inhibitor of the transcriptional activities of CBP-dependent activators such as nuclear receptors and p53. The dominant neg. property of MOZ-TIF2 requires the CBP-binding domain (activation domain 1 [AD1]), and coimmunopptn. and fluorescent resonance energy transfer expts. show that MOZ-TIF2 interacts with CBP directly in vivo. The CBP-binding domain is also required for the ability of MOZ-TIF2 to extend the proliferative potential of murine bone marrow lineage-neg. cells in vitro. We show that MOZ-TIF2 displays an aberrant nuclear distribution and that cells expressing this protein have reduced levels of cellular CBP, leading to depletion of CBP from PML bodies. In summary, our results indicate that disruption of the normal function of CBP and CBP-dependent activators is an important feature of MOZ-TIF2 action in AML.
- 398Ito, Y.; Miyazono, K. Curr. Opin. Genet. Dev. 2003, 13, 43There is no corresponding record for this reference.
- 399Grossman, S. R. Eur. J. Biochem. 2001, 268, 2773399p300/CBP/p53 interaction and regulation of the p53 responseGrossman, Steven R.European Journal of Biochemistry (2001), 268 (10), 2773-2778CODEN: EJBCAI; ISSN:0014-2956. (Blackwell Science Ltd.)A review with 50 refs. Substantial evidence points to a crit. role for the p300/CREB binding protein (CBP) co-activators in p53 responses to DNA damage. The p300/CBP and the assocd. protein P/CAF bind to and acetylate p53 during the DNA damage response, and are needed for full p53 transactivation as well as downstream p53 effects of growth arrest and/or apoptosis. Beyond this simplistic model, p300/CBP appear to be complex integrators of signals that regulate p53, and biochem., the multipartite p53/p300/CBP interaction is equally complex. Through phys. interaction with p53, p300/CBP can both pos. and neg. regulate p53 transactivation, as well as p53 protein turnover depending on cellular context and environmental stimuli, such as DNA damage.
- 400Iyer, N. G.; Xian, J.; Chin, S. F.; Bannister, A. J.; Daigo, Y.; Aparicio, S.; Kouzarides, T.; Caldas, C. Oncogene 2007, 26, 21There is no corresponding record for this reference.
- 401Galbiati, L.; Mendoza-Maldonado, R.; Gutierrez, M. I.; Giacca, M. Cell Cycle 2005, 4, 930There is no corresponding record for this reference.
- 402Oike, Y.; Hata, A.; Mamiya, T.; Kaname, T.; Noda, Y.; Suzuki, M.; Yasue, H.; Nabeshima, T.; Araki, K.; Yamamura, K. Hum. Mol. Genet. 1999, 8, 387402Truncated CBP protein leads to classical Rubinstein-Taybi syndrome phenotypes in mice: implications for a dominant-negative mechanismOike, Yuichi; Hata, Akira; Mamiya, Takayoshi; Kaname, Tadashi; Noda, Yukihiro; Suzuki, Misao; Yasue, Hirofumi; Nabeshima, Toshitaka; Araki, Kimi; Yamamura, Ken-IchiHuman Molecular Genetics (1999), 8 (3), 387-396CODEN: HMGEE5; ISSN:0964-6906. (Oxford University Press)A mouse model of Rubinstein-Taybi syndrome (RTS) was generated by an insertional mutation into the cAMP response element-binding protein (CREB)-binding protein (CBP) gene. Heterozygous CBP-deficient mice, which had truncated CBP protein (residues 1-1084) contg. the CREB-binding domain (residues 462-661), showed clin. features of RTS, such as growth retardation (100%), retarded osseous maturation (100%), hypoplastic maxilla with narrow palate (100%), cardiac anomalies (15%) and skeletal abnormalities (7%). Truncated CBP is considered to have been acting during development as a dominant-neg. inhibitor to lead to the phenotypes of RTS in mice. Our studies with step-through-type passive avoidance tests and with fear conditioning test showed that mice were deficient in long-term memory (LTM). In contrast, short-term memory (STM) appeared to be normal. These results implicate a crucial role for CBP in mammalian LTM. Our CBP+/- mice would be an excellent model for the study of the role of CBP in development and memory storage mechanisms.
- 403Viosca, J.; Lopez-Atalaya, J. P.; Olivares, R.; Eckner, R.; Barco, A. Neurobiol. Dis. 2010, 37, 186403Syndromic features and mild cognitive impairment in mice with genetic reduction on p300 activity: Differential contribution of p300 and CBP to Rubinstein-Taybi syndrome etiologyViosca Jose; Lopez-Atalaya Jose P; Olivares Roman; Eckner Richard; Barco AngelNeurobiology of disease (2010), 37 (1), 186-94 ISSN:.Rubinstein-Taybi syndrome (RSTS) is a complex autosomal-dominant disease characterized by mental and growth retardation and skeletal abnormalities. A majority of the individuals diagnosed with RSTS carry heterozygous mutation in the gene CREBBP, but a small percentage of cases are caused by mutations in EP300. To investigate the contribution of p300 to RSTS pathoetiology, we carried out a comprehensive and multidisciplinary characterization of p300(+/-) mice. These mice exhibited facial abnormalities and impaired growth, two traits associated to RSTS in humans. We also observed abnormal gait, reduced swimming speed, enhanced anxiety in the elevated plus maze, and mild cognitive impairment during the transfer task in the water maze. These analyses demonstrate that p300(+/-) mice exhibit phenotypes that are reminiscent of neurological traits observed in RSTS patients, but their comparison with previous studies on CBP deficient strains also indicates that, in agreement with the most recent findings in human patients, the activity of p300 in cognition is likely less relevant or more susceptible to compensation than the activity of CBP.
- 404Kung, A. L.; Rebel, V. I.; Bronson, R. T.; Ch’ng, L. E.; Sieff, C. A.; Livingston, D. M.; Yao, T. P. Genes Dev. 2000, 14, 272There is no corresponding record for this reference.
- 405Kasper, L. H.; Fukuyama, T.; Biesen, M. A.; Boussouar, F.; Tong, C.; de Pauw, A.; Murray, P. J.; van Deursen, J. M.; Brindle, P. K. Mol. Cell. Biol. 2006, 26, 789405Conditional knockout mice reveal distinct functions for the global transcriptional coactivators CBP and p300 in T-cell developmentKasper, Lawryn H.; Fukuyama, Tomofusa; Biesen, Michelle A.; Boussouar, Faycal; Tong, Caili; de Pauw, Antoine; Murray, Peter J.; van Deursen, Jan M. A.; Brindle, Paul K.Molecular and Cellular Biology (2006), 26 (3), 789-809CODEN: MCEBD4; ISSN:0270-7306. (American Society for Microbiology)The global transcriptional coactivators CREB-binding protein (CBP) and the closely related p300 interact with over 312 proteins, making them among the most heavily connected hubs in the known mammalian protein-protein interactome. It is largely uncertain, however, if these interactions are important in specific cell lineages of adult animals, as homozygous null mutations in either CBP or p300 result in early embryonic lethality in mice. Here we describe a Cre/LoxP conditional p300 null allele (p300flox) that allows for the temporal and tissue-specific inactivation of p300. We used mice carrying p300flax and a CBP conditional knockout allele (CBPflax) in conjunction with an Lck-Cre transgene to delete CBP and p300 starting at the CD4- CD8- double-neg. thymocyte stage of T-cell development. Loss of either p300 or CBP led to a decrease in CD4+ CD8+ double-pos. thymocytes, but an increase in the percentage of CD8+ single-pos. thymocytes seen in CBP mutant mice was not obsd. in p300 mutants. T cells completely lacking both CBP and p300 did not develop normally and were nonexistent or very rare in the periphery, however. T cells lacking CBP or p300 had reduced tumor necrosis factor alpha gene expression in response to phorbol ester and ionophore, while signal-responsive gene expression in CBP- or p300-deficient macrophages was largely intact. Thus, CBP and p300 each supply a surprising degree of redundant coactivation capacity in T cells and macrophages, although each gene has also unique properties in thymocyte development.
- 406Rebel, V. I.; Kung, A. L.; Tanner, E. A.; Yang, H.; Bronson, R. T.; Livingston, D. M. Proc. Natl. Acad. Sci. U.S.A. 2002, 99, 14789There is no corresponding record for this reference.
- 407Kasper, L. H.; Fukuyama, T.; Biesen, M. A.; Boussouar, F.; Tong, C.; de Pauw, A.; Murray, P. J.; van Deursen, J. M.; Brindle, P. K. Mol. Cell. Biol. 2006, 26, 789407Conditional knockout mice reveal distinct functions for the global transcriptional coactivators CBP and p300 in T-cell developmentKasper, Lawryn H.; Fukuyama, Tomofusa; Biesen, Michelle A.; Boussouar, Faycal; Tong, Caili; de Pauw, Antoine; Murray, Peter J.; van Deursen, Jan M. A.; Brindle, Paul K.Molecular and Cellular Biology (2006), 26 (3), 789-809CODEN: MCEBD4; ISSN:0270-7306. (American Society for Microbiology)The global transcriptional coactivators CREB-binding protein (CBP) and the closely related p300 interact with over 312 proteins, making them among the most heavily connected hubs in the known mammalian protein-protein interactome. It is largely uncertain, however, if these interactions are important in specific cell lineages of adult animals, as homozygous null mutations in either CBP or p300 result in early embryonic lethality in mice. Here we describe a Cre/LoxP conditional p300 null allele (p300flox) that allows for the temporal and tissue-specific inactivation of p300. We used mice carrying p300flax and a CBP conditional knockout allele (CBPflax) in conjunction with an Lck-Cre transgene to delete CBP and p300 starting at the CD4- CD8- double-neg. thymocyte stage of T-cell development. Loss of either p300 or CBP led to a decrease in CD4+ CD8+ double-pos. thymocytes, but an increase in the percentage of CD8+ single-pos. thymocytes seen in CBP mutant mice was not obsd. in p300 mutants. T cells completely lacking both CBP and p300 did not develop normally and were nonexistent or very rare in the periphery, however. T cells lacking CBP or p300 had reduced tumor necrosis factor alpha gene expression in response to phorbol ester and ionophore, while signal-responsive gene expression in CBP- or p300-deficient macrophages was largely intact. Thus, CBP and p300 each supply a surprising degree of redundant coactivation capacity in T cells and macrophages, although each gene has also unique properties in thymocyte development.
- 408Qiao, Y.; Molina, H.; Pandey, A.; Zhang, J.; Cole, P. A. Science 2006, 311, 1293There is no corresponding record for this reference.
- 409Liao, Z. W.; Zhou, T. C.; Tan, X. J.; Song, X. L.; Liu, Y.; Shi, X. Y.; Huang, W. J.; Du, L. L.; Tu, B. J.; Lin, X. D. J. Transl. Med. 2012, 10, 110409High expression of p300 is linked to aggressive features and poor prognosis of nasopharyngeal carcinomaLiao, Zhi-Wei; Zhou, Tong-Chong; Tan, Xiao-Jun; Song, Xian-Lu; Liu, Yuan; Shi, Xing-Yuan; Huang, Wen-Jin; Du, Li-Li; Tu, Bo-Jun; Lin, Xiao-DanJournal of Translational Medicine (2012), 10 (), 110CODEN: JTMOBV; ISSN:1479-5876. (BioMed Central Ltd.)Background: Increased expression of transcriptional coactivator p300 has been obsd. in a variety of human cancers. However, the expression status of p300 protein/mRNA in nasopharyngeal carcinoma (NPC) tissues and its clinicopathol./prognostic implication are poorly understood. Methods: In our study, mRNA and protein expression levels of p300 was explored by reverse transcription-polymerase chain reaction (RT-PCR), Western blotting (WB) and immunohistochem. (IHC) in nasopharyngeal mucosal and NPC tissues. The data were analyzed by receiver operating characteristic (ROC) curve anal., spearman's rank correlation, Kaplan-Meier plots and Cox proportional hazards regression model. Results: Up-regulated expression of p300 mRNA/p300 protein was detected in NPC tissues by RT-PCR and WB, when compared to nasopharyngeal mucosal tissues. Based on ROC curve anal., the cutoff score for p300 high expression was defined when more than 35% of the tumor cells were pos. stained. High expression of p300 was obsd. in 127/209 (60.7%) of NPCs. In NPCs, high expression of p300 was pos. assocd. with later T classification, later N classification, distant metastasis and later clin. stage (P < 0.05). In univariate survival anal., overexpression of p300 was found to be an indicator of progression-free (P = 0.002) and overall survival (P = 0.001) in NPCs. More importantly, p300 expression was evaluated as an independent prognostic factor for NPC in multivariate anal. (P = 0.036). Conclusions: Our findings support that high expression of p300 protein might be important in conferring a more aggressive behavior, and is an independent mol. marker for shortened survival time of patients with NPC.
- 410Li, M.; Luo, R. Z.; Chen, J. W.; Cao, Y.; Lu, J. B.; He, J. H.; Wu, Q. L.; Cai, M. Y. J. Transl. Med. 2011, 9, 5410High expression of transcriptional coactivator p300 correlates with aggressive features and poor prognosis of hepatocellular carcinomaLi, Mei; Luo, Rong-Zhen; Chen, Jie-Wei; Cao, Yun; Lu, Jia-Bin; He, Jie-Hua; Wu, Qiu-Liang; Cai, Mu-YanJournal of Translational Medicine (2011), 9 (), 5CODEN: JTMOBV; ISSN:1479-5876. (BioMed Central Ltd.)Background: It has been suggested that p300 participates in the regulation of a wide range of cell biol. processes and mutation of p300 has been identified in certain types of human cancers. However, the expression dynamics of p300 in hepatocellular carcinoma (HCC) and its clin./prognostic significance are unclear. Methods: In this study, the methods of reverse transcription-polymerase chain reaction (RT-PCR), Western blotting and immunohistochem. (IHC) were utilized to investigate protein/mRNA expression of p300 in HCCs. Receiver operating characteristic (ROC) curve anal., spearman's rank correlation, Kaplan-Meier plots and Cox proportional hazards regression model were used to analyze the data. Results: Up-regulated expression of p300 mRNA and protein was obsd. in the majority of HCCs by RT-PCR and Western blotting, when compared with their adjacent non-malignant liver tissues. According to the ROC curves, the cutoff score for p300 high expression was defined when more than 60% of the tumor cells were pos. stained. High expression of p300 was examd. in 60/123 (48.8%) of HCCs and in 8/123 (6.5%) of adjacent non-malignant liver tissues. High expression of p300 was correlated with higher AFP level, larger tumor size, multiplicity, poorer differentiation and later stage (P < 0.05). In univariate survival anal., a significant assocn. between overexpression of p300 and shortened patients' survival was found (P = 0.001). In different subsets of HCC patients, p300 expression was also a prognostic indicator in patients with stage II (P = 0.007) and stage III (P = 0.011). Importantly, p300 expression was evaluated as an independent prognostic factor in multivariate anal. (P = 0.021). Consequently, a new clinicopathol. prognostic model with three poor prognostic factors (p300 expression, AFP level and vascular invasion) was constructed. The model could significantly stratify risk (low, intermediate and high) for overall survival (P < 0.0001). Conclusions: Our findings provide a basis for the concept that high expression of p300 in HCC may be important in the acquisition of an aggressive phenotype, suggesting that p300 overexpression, as examd. by IHC, is an independent biomarker for poor prognosis of patients with HCC. The combined clinicopathol. prognostic model may become a useful tool for identifying HCC patients with different clin. outcomes.
- 411Vleugel, M. M.; Shvarts, D.; van der Wall, E.; van Diest, P. J. Hum. Pathol. 2006, 37, 1085411p300 and p53 levels determine activation of HIF-1 downstream targets in invasive breast cancerVleugel, Marije M.; Shvarts, David; van der Wall, Elsken; van Diest, Paul J.Human Pathology (2006), 37 (8), 1085-1092CODEN: HPCQA4; ISSN:0046-8177. (Elsevier Inc.)In previous studies, we noted that overexpression of hypoxia-inducible factor (HIF)-1α in breast cancer, esp. the diffuse form, does not always lead to functional activation of its downstream genes. Transcriptional activity of HIF-1 may be repressed by p53 through competition for transcriptional coactivators such as p300. The aim of this study was therefore to explore the role of p53 and p300 in relation to overexpression of HIF-1α and activation of HIF-1 downstream genes in invasive breast cancer. p300 immunohistochem. was performed in a group of 183 early-stage invasive breast cancers, and related to p53 accumulation, overexpression of HIF-1α, and several HIF-1 downstream genes. p300 was expressed in varying degrees in 84% of invasive breast cancers. p300 staining intensity correlated pos. with HIF-1α expression (P = .04), p53 accumulation (P = .001), and overexpression of glucose transporter 1 (GLUT-1) (P < .001), a glucose transporter downstream target gene of HIF-1. GLUT-1 levels were significantly assocd. with p300 in HIF-1α pos. patients (P = .02). p53 accumulation significantly pos. correlated with carbonic anhydrase IX (CAIX)/GLUT-1 coexpression in HIF-1α-pos. patients (P = .007). p53 accumulation/high p300 levels, the most favorable situation for HIF-1 downstream activation, were significantly assocd. with GLUT-1 overexpression (P = .01) and coexpression of CAIX/GLUT-1 (P = .03), compared with low p53/low p300 levels, the most unfavorable situation for HIF-1 downstream activation. p300 is a cofactor highly assocd. with p53 accumulation and HIF-1α levels in invasive breast cancer. Furthermore, low levels of p300 may explain absence of downstream effects in HIF-1α-overexpressing cancers, an effect that seems to be enhanced by wild-type levels of p53. This underlines the importance of p300 levels and p53 accumulation in the HIF-1-regulated response toward hypoxia.
- 412Hudelist, G.; Czerwenka, K.; Kubista, E.; Marton, E.; Pischinger, K.; Singer, C. F. Breast Cancer Res. Treat. 2003, 78, 193412Expression of Sex Steroid Receptors and their Co-Factors in Normal and Malignant Breast Tissue: AIB1 is a Carcinoma-Specific Co-ActivatorHudelist, Gernot; Czerwenka, Klaus; Kubista, Ernst; Marton, Erika; Pischinger, Kerstin; Singer, Christian F.Breast Cancer Research and Treatment (2003), 78 (2), 193-204CODEN: BCTRD6; ISSN:0167-6806. (Kluwer Academic Publishers)The differential expression pattern of estrogen receptor alpha (ER-α), estrogen receptor beta (ER-β) and their co-activator/co-repressor proteins is thought to modulate estrogenic action and to be present already during the early stages of tumorigenesis. It has therefore been postulated that certain co-activator and co-repressor proteins contribute to the development of breast cancer. There are some reports providing information on gene amplification and mRNA over-expression of certain co-factors in breast cancer, but to date there is only limited knowledge about their resp. protein expressions. The aim of this study was to examine the expression of four steroid receptor co-activators (steroid receptor co-activator 1 (SRC-1), transcription intermediary factor 2 (TIF 2), protein 300 kDa/CREB binding protein (p300/CBP), amplified in breast cancer 1 (AIB1)), and of the co-repressor nuclear receptor co-repressor (NCoR), in malignant breast tissues and in matching normal breast biopsies of the same individuals. Protein expression was analyzed by immunohistochem. and was compared to prognostic parameters such as lymph node involvement, tumor grading and receptor status. All members of the co-regulatory protein family were detected in both, benign and matching malignant tissue samples, except for AIB1, which was found to be expressed exclusively in malignant epithelium. AIB1 was preferentially present in carcinomas with high tumor grade (r = 0.48, p = 0.014), and was co-expressed with p300/CBP (r = 0.54, p = 0.006). TIF 2 correlated significantly to nodal status (r = 0.46, p = 0.025). Furthermore, protein levels of ER-β, p300/CBP and AIB1 were higher in invasive ductal carcinomas than in normal mammary tissue. The tumoral ER-α protein expression was significantly correlated with that of PgR (r = 0.61, p = 0.001) and NCoR (r = 0.4, p = 0.043), whereas ER-β expression was assocd. with SRC-1 (r = 0.68, p ≤ 0.001), TIF 2 (r = 0.64, p = 0.001) and NCoR (r = 0.48, p = 0.014) protein levels in malignant specimens. In our hands, 20% of ER-β pos. tumors did not express ER-α protein, thereby suggesting that a substantial fraction of ER-beta pos. tumors is falsely considered to be estrogen receptor neg.' if only ER-α specific antibodies are employed in the histol. assessment of the ER status.
- 413Debes, J. D.; Sebo, T. J.; Lohse, C. M.; Murphy, L. M.; Haugen, D. A.; Tindall, D. J. Cancer Res. 2003, 63, 7638There is no corresponding record for this reference.
- 414Isharwal, S.; Miller, M. C.; Marlow, C.; Makarov, D. V.; Partin, A. W.; Veltri, R. W. Prostate 2008, 68, 1097There is no corresponding record for this reference.
- 415Heery, D. M.; Fischer, P. M. Drug Discovery Today 2007, 12, 88There is no corresponding record for this reference.
- 416Lin, W. M.; Baker, A. C.; Beroukhim, R.; Winckler, W.; Feng, W.; Marmion, J. M.; Laine, E.; Greulich, H.; Tseng, H.; Gates, C.; Hodi, F. S.; Dranoff, G.; Sellers, W. R.; Thomas, R. K.; Meyerson, M.; Golub, T. R.; Dummer, R.; Herlyn, M.; Getz, G.; Garraway, L. A. Cancer Res. 2008, 68, 664There is no corresponding record for this reference.
- 417Rotte, A.; Bhandaru, M.; Cheng, Y.; Sjoestroem, C.; Martinka, M.; Li, G. PLoS One 2013, 8, e75405There is no corresponding record for this reference.
- 418Bhandaru, M.; Ardekani, G. S.; Zhang, G.; Martinka, M.; McElwee, K. J.; Li, G.; Rotte, A. BMC Cancer 2014, 14– 398There is no corresponding record for this reference.
- 419Bandyopadhyay, D.; Okan, N. A.; Bales, E.; Nascimento, L.; Cole, P. A.; Medrano, E. E. Cancer Res. 2002, 62, 6231There is no corresponding record for this reference.
- 420Yan, G.; Eller, M. S.; Elm, C.; Larocca, C. A.; Ryu, B.; Panova, I. P.; Dancy, B. M.; Bowers, E. M.; Meyers, D.; Lareau, L.; Cole, P. A.; Taverna, S. D.; Alani, R. M. J. Invest. Dermatol. 2013, 133, 2444There is no corresponding record for this reference.
- 421Iyer, N. G.; Chin, S. F.; Ozdag, H.; Daigo, Y.; Hu, D. E.; Cariati, M.; Brindle, K.; Aparicio, S.; Caldas, C. Proc. Natl. Acad. Sci. U.S.A. 2004, 101, 7386There is no corresponding record for this reference.
- 422Giles, R. H.; Petrij, F.; Dauwerse, H. G.; den Hollander, A. I.; Lushnikova, T.; van Ommen, G. J.; Goodman, R. H.; Deaven, L. L.; Doggett, N. A.; Peters, D. J.; Breuning, M. H. Genomics 1997, 42, 96There is no corresponding record for this reference.
- 423Rozman, M.; Camós, M.; Colomer, D.; Villamor, N.; Esteve, J.; Costa, D.; Carrió, A.; Aymerich, M.; Aguilar, J. L.; Domingo, A.; Solé, F.; Gomis, F.; Florensa, L.; Montserrat, E.; Campo, E. Genes, Chromosomes Cancer 2004, 40, 140423Type I MOZ/CBP (MYST3/CREBBP) is the most common chimeric transcript in acute myeloid leukemia with t(8;16)(p11;p13) translocationRozman, Maria; Camos, Mireia; Colomer, Dolors; Villamor, Neus; Esteve, Jordi; Costa, Dolors; Carrio, Ana; Aymerich, Marta; Aguilar, Josep Lluis; Domingo, Alicia; Sole, Francesc; Gomis, Federico; Florensa, Lourdes; Montserrat, Emili; Campo, EliasGenes, Chromosomes & Cancer (2004), 40 (2), 140-145CODEN: GCCAES; ISSN:1045-2257. (Wiley-Liss, Inc.)The t(8;16)(p11;p13) fuses the MOZ (MYST3) gene at 8p11 with CBP (CREBBP) at 16p13 and is assocd. with an infrequent but well-defined type of acute myeloid leukemia (AML) that has unique morphocytochem. findings (monocytoid blast morphol. with erythrophagocytosis and simultaneously pos. for myeloperoxidase and nonspecific esterases). RT-PCR amplification of MOZ/CBP (MYST3/CREBBP) chimera has proved difficult, with four different transcripts found in four reported cases. The authors studied 7 AML-t(8;16) patients, 5 with cytogenetically demonstrated t(8;16) and 2 with similar morphocytochem. and immunophenotypical characteristics. Clin., 3 cases presented as therapy-related leukemia. Extramedullar involvement was obsd. at presentation in 2 patients and coagulopathy in 4. The clinicobiol. findings confirmed the distinctiveness of this entity. Of note is the erythrophagocytosis in 5 of 7 cases and the immunol. negativity for CD34 and CD117 and positivity for CD56. Using a new RT-PCR strategy, the authors were able to amplify a specific band of 212 bp in six cases in which sequence anal. confirmed the presence of the previously described MOZ/CBP fusion transcript type 1. This is the largest molecularly studied AML-t(8;16) series, which demonstrates that MOZ/CBP breakpoints are usually clustered in intron 16 of MOZ and intron 2 of CBP. The newly designed single-round PCR provides a simple tool for the mol. confirmation of MOZ/CBP rearrangement.
- 424Borrow, J.; Stanton, V. P., Jr.; Andresen, J. M.; Becher, R.; Behm, F. G.; Chaganti, R. S.; Civin, C. I.; Disteche, C.; Dubé, I.; Frischauf, A. M.; Horsman, D.; Mitelman, F.; Volinia, S.; Watmore, A. E.; Housman, D. E. Nat. Genet. 1996, 14, 33424The translocation t(8;16)(p11;p13) of acute myeloid leukemia fuses a putative acetyltransferase to the CREB-binding proteinBorrow, Julian; Stanton, Vincent P., Jr.; Andresen, J. Michael; Becher, Reinhard; Behm, Frederick G.; Chagaanti, Raju S. K.; Civin, Curt I.; Disteche, Christine; Dube, Ian; et al.Nature Genetics (1996), 14 (1), 33-41CODEN: NGENEC; ISSN:1061-4036. (Nature Publishing Co.)The recurrent translocation t(8;16)(p11;p13) is a cytogenetic hallmark for the M4/M5 subtype of acute myeloid leukemia. Here the authors identify the breakpoint-assocd. genes. Positional cloning on chromosome 16 implicates the CREB-binding protein (CBP), a transcriptional adaptor/co-activator protein. At the chromosome 8 breakpoint the authors identify a novel gene, MOZ, which encodes a 2,004-amino-acid protein characterized by two C4HC3 zinc fingers and a single C2HC zinc finger in conjunction with a putative acetyltransferase signature. In-frame MOZ-CBP fusion transcripts combine the MOZ finger motifs and putative acetyltransferase domain with a largely intact CBP. The authors suggest that MOZ may represent a chromatin-assocd. acetyltransferase, and raise the possibility that a dominant MOZ-CBP fusion protein could mediate leukemogenesis via aberrant chromatin acetylation.
- 425Chaffanet, M.; Gressin, L.; Preudhomme, C.; Soenen-Cornu, V.; Birnbaum, D.; Pébusque, M. J. Genes, Chromosomes Cancer 2000, 28, 138425MOZ is fused to p300 in an acute monocytic leukemia with t(8;22)Chaffanet, Max; Gressin, Laetitia; Preudhomme, Claude; Soenen-Cornu, Valerie; Birnbaum, Daniel; Pebusque, Marie-JosepheGenes, Chromosomes & Cancer (2000), 28 (2), 138-144CODEN: GCCAES; ISSN:1045-2257. (Wiley-Liss, Inc.)The authors report on the fusion of the monocytic leukemia zinc finger protein (MOZ) gene to the adenoviral EIA-assocd. protein p300 gene in acute monocytic leukemia MS assocd. with a t(8;22)(p11;q13) translocation. The authors studied two patients with double-color fluorescence in situ hybridization (FISH) using the yeast artificial chromosome 176C9 and the bacterial artificial chromosome clone H59D10 specific to the MOZ and p300 genes, resp. Both probes were split in the patients' chromosome metaphase cells, and the two deriv. chromosomes were each labeled with both probes. The authors showed by Southern blot the rearrangement of the MOZ gene, and cloned the fusion transcripts in one patient carrying the t(8;22) by reverse transcription-polymerase chain reaction using MOZ- and p300-specific primers. Both fusion transcripts were expressed. This result defines a novel reciprocal translocation involving two acetyltransferases, MOZ and p300, resulting in an abnormal transcriptional co-activator that could play a crit. role in leukemogenesis.
- 426Kitabayashi, I.; Aikawa, Y.; Yokoyama, A.; Hosoda, F.; Nagai, M.; Kakazu, N.; Abe, T.; Ohki, M. Leukemia 2001, 15, 89There is no corresponding record for this reference.
- 427Panagopoulos, I.; Fioretos, T.; Isaksson, M.; Samuelsson, U.; Billström, R.; Strömbeck, B.; Mitelman, F.; Johansson, B. Hum. Mol. Genet. 2001, 10, 395427Fusion of the MORF and CBP genes in acute myeloid leukemia with the t(10;16)(q22;p13)Panagopoulos, Ioannis; Fioretos, Thoas; Isaksson, Margareth; Samuelsson, Ulf; Billstrom, Rolf; Strombeck, Bodil; Mitelman, Felix; Johansson, BertilHuman Molecular Genetics (2001), 10 (4), 395-404CODEN: HMGEE5; ISSN:0964-6906. (Oxford University Press)The CBP gene at 16p13 fuses to MOZ and MLL as a result of the t(8;16)(p11;p13) in acute (myelo)monocytic leukemias (AML M4/M5) and the t(11;16)(q23;p13) in treatment-related AML, resp. Here the authors show that a novel t(10;16)(q22;p13) in a childhood AML M5a leads to a MORF-CBP chimera. RT-PCR using MORF forward and CBP reverse primers amplified a MORF-CBP fusion in which nucleotide 3103 of MORF was fused in-frame with nucleotide 284 of CBP. Nested RT-PCR with CBP forward and MORF reverse primers generated a CBP-MORF transcript in which nucleotide 283 of CBP was fused in-frame with nucleotide 3104 of MORF. Genomic analyses revealed that the breaks were close to Alu elements in intron 16 of MORF and intron 2 of CBP and that duplications had occurred near the breakpoints. A database search using MORF cDNA enabled us to construct an exon-intron map of the MORF gene. The MORF-CBP protein retains the zinc fingers, two nuclear localization signals, the histone acetyltransferase (HAT) domain, a portion of the acidic domain of MORF and the CBP protein downstream of codon 29. Thus, the part of CBP encoding the RARA-binding domain, the CREB-binding domain, the three Cys/His-rich regions, the bromodomain, the HAT domain and the Glu-rich domains is present. In the reciprocal CBP-MORF, part of the acidic domain and the C-terminal Ser- and Met-rich regions of MORF are likely to be driven by the CBP promoter. Since both fusion transcripts were present, their exact role in the leukemogenic process remains to be elucidated.
- 428Vizmanos, J. L.; Larráyoz, M. J.; Lahortiga, I.; Floristán, F.; Alvarez, C.; Odero, M. D.; Novo, F. J.; Calasanz, M. J. Genes, Chromosomes Cancer 2003, 36, 402428t(10;16)(q22;p13) and MORF-CREBBP fusion is a recurrent event in acute myeloid leukemiaVizmanos, Jose L.; Larrayoz, Maria J.; Lahortiga, Idoya; Floristan, Filomena; Alvarez, Carmen; Odero, Maria D.; Novo, Francisco J.; Calasanz, Maria J.Genes, Chromosomes & Cancer (2003), 36 (4), 402-405CODEN: GCCAES; ISSN:1045-2257. (Wiley-Liss, Inc.)Recently, it was shown that t(10;16)(q22;p13) fuses the MORF and CREBBP genes in a case of childhood acute myeloid leukemia (AML) M5a, with a complex karyotype contg. other rearrangements. Here, we report a new case with the MORF-CREBBP fusion in an 84-yr-old patient diagnosed with AML M5b, in which the t(10;16)(q22;p13) was the only cytogenetic aberration. This supports that this is a recurrent pathogenic translocation in AML.
- 429Uppal, G. K.; Leighton, J.; Da Costa, D.; Czulewicz, A.; Palazzo, I. E. Hematol. Rep. 2011, 3, e23There is no corresponding record for this reference.
- 430Kojima, K.; Kaneda, K.; Yoshida, C.; Dansako, H.; Fujii, N.; Yano, T.; Shinagawa, K.; Yasukawa, M.; Fujita, S.; Tanimoto, M. Br. J. Haematol. 2003, 120, 271430A novel fusion variant of the MORF and CBP genes detected in therapy-related myelodysplastic syndrome with t(10;16)(q22;p13)Kojima, Kensuke; Kaneda, Kinuyo; Yoshida, Chikamasa; Dansako, Hirokata; Fujii, Nobuharu; Yano, Tomofumi; Shinagawa, Katsuji; Yasukawa, Masaki; Fujita, Shigeru; Tanimoto, MitsuneBritish Journal of Haematology (2003), 120 (2), 271-273CODEN: BJHEAL; ISSN:0007-1048. (Blackwell Publishing Ltd.)We report a case of therapy-related myelodysplastic syndrome (t-MDS) with t(10;16)(q22;p13), in which novel fusion transcripts of the MORF and CBP genes were detected. In one MORF-CBP fusion transcript, exon 15 of the MORF gene was fused in frame with exon 5 of the CBP gene. In a reciprocal CBP-MORF transcript, exon 4 of the CBP gene was fused in frame with exon 16 of the MORF gene. This is the first reported case of t-MDS assocd. with t(10;16), and provides mol. evidence that the novel MORF-CBP and/or CBP-MORF fusion protein(s) might play an important role in the development of t-MDS.
- 431Taki, T.; Sako, M.; Tsuchida, M.; Hayashi, Y. Blood 1997, 89, 3945There is no corresponding record for this reference.
- 432Sobulo, O. M.; Borrow, J.; Tomek, R.; Reshmi, S.; Harden, A.; Schlegelberger, B.; Housman, D.; Doggett, N. A.; Rowley, J. D.; Zeleznik-Le, N. J. Proc. Natl. Acad. Sci. U.S.A. 1997, 94, 8732There is no corresponding record for this reference.
- 433Satake, N.; Ishida, Y.; Otoh, Y.; Hinohara, S.; Kobayashi, H.; Sakashita, A.; Maseki, N.; Kaneko, Y. Genes, Chromosomes Cancer 1997, 20, 60433Novel MLL-CBP fusion transcript in therapy-related chronic myelomonocytic leukemia with a t(11;16) (q23;p13) chromosome translocationSatake, Noriko; Ishida, Yasushi; Otoh, Yoshiko; Hinohara, Shin-Ichi; Kobayashi, Hirofumi; Sakashita, Akiko; Maseki, Nobuo; Kaneko, YasuhikoGenes, Chromosomes & Cancer (1997), 20 (1), 60-63CODEN: GCCAES; ISSN:1045-2257. (Wiley-Liss)CBP, which is located on 16p13 and encodes a transcriptional adaptor/coactivator protein, has been shown to fuse by the t(8;16)(p11;p13) translocation to MOZ on 8p11 in acute myeloid leukemia. The authors found a t(11;16)(q23;p13) in a child with therapy-related chronic myelomonocytic leukemia. Subsequent reverse transcriptase-polymerase chain reaction and direct sequencing analyses revealed the MLL-CBP fusion transcript in CMML cells. Because 11q23 translocations involving MLL and t(8;16) involving MOZ and CBP have been reported in therapy-related leukemias, both the MLL and CBP genes may be targets for topoisomerase II inhibitors. Accordingly, the authors believe that most t(11;16)-assocd. leukemias may develop in patients who have been treated with cytotoxic chemotherapy for primary malignant diseases.
- 434Ida, K.; Kitabayashi, I.; Taki, T.; Taniwaki, M.; Noro, K.; Yamamoto, M.; Ohki, M.; Hayashi, Y. Blood 1997, 90, 4699There is no corresponding record for this reference.
- 435Wang, J.; Iwasaki, H.; Krivtsov, A.; Febbo, P. G.; Thorner, A. R.; Ernst, P.; Anastasiadou, E.; Kutok, J. L.; Kogan, S. C.; Zinkel, S. S.; Fisher, J. K.; Hess, J. L.; Golub, T. R.; Armstrong, S. A.; Akashi, K.; Korsmeyer, S. J. EMBO J. 2005, 24, 368There is no corresponding record for this reference.
- 436Wang, L.; Gural, A.; Sun, X. J.; Zhao, X.; Perna, F.; Huang, G.; Hatlen, M. A.; Vu, L.; Liu, F.; Xu, H.; Asai, T.; Xu, H.; Deblasio, T.; Menendez, S.; Voza, F.; Jiang, Y.; Cole, P. A.; Zhang, J.; Melnick, A.; Roeder, R. G.; Nimer, S. D. Science 2011, 333, 765There is no corresponding record for this reference.
- 437Reynoird, N.; Schwartz, B. E.; Delvecchio, M.; Sadoul, K.; Meyers, D.; Mukherjee, C.; Caron, C.; Kimura, H.; Rousseaux, S.; Cole, P. A.; Panne, D.; French, C. A.; Khochbin, S. EMBO J. 2010, 29, 2943There is no corresponding record for this reference.
- 438Crump, N. T.; Hazzalin, C. A.; Bowers, E. M.; Alani, R. M.; Cole, P. A.; Mahadevan, L. C. Proc. Natl. Acad. Sci. U.S.A. 2011, 108, 7814There is no corresponding record for this reference.
- 439Pattabiraman, D. R.; Sun, J.; Dowhan, D. H.; Ishii, S.; Gonda, T. J. Mol. Cancer Res. 2009, 7, 1477439Mutations in Multiple Domains of c-Myb Disrupt Interaction with CBP/p300 and Abrogate Myeloid Transforming AbilityPattabiraman, Diwakar R.; Sun, Jane; Dowhan, Dennis H.; Ishii, Shunsuke; Gonda, Thomas J.Molecular Cancer Research (2009), 7 (9), 1477-1486CODEN: MCROC5; ISSN:1541-7786. (American Association for Cancer Research)The c-myb proto-oncogene is a key regulator of hematopoietic cell proliferation and differentiation. MYB mRNA is expressed at high levels in, and is required for the proliferation of, most human myeloid and acute lymphoid leukemias. Recently, chromosomal translocation and genomic duplications of c-MYB have been identified in human T-cell acute leukemia. The present work focuses on the effects of mutations in different domains of the murine c-Myb protein on its transforming ability as defined by suppression of myelomonocytic differentiation and continued proliferation. Using both a novel myeloid cell line-based assay and a primary hematopoietic cell assay, we have shown that mutation of single residues in the transactivation domain important for CBP/p300 binding leads to complete loss of transforming ability. We also simultaneously mutated residues in the DNA-binding domain and the neg. regulatory domain of the protein. These double mutants, but not the corresponding single mutants, show a complete loss of transforming activity. Surprisingly, these double mutants show severely impaired transactivation and are also defective for CBP/p300 binding. Our results imply that multiple Myb domains influence its interaction with CBP/p300, highlight the importance of this interaction for myeloid transformation, and suggest an approach for mol. targeting of Myb in leukemia. (Mol Cancer Res 2009;7(9):1477-86).
- 440Iyer, N. G.; Ozdag, H.; Caldas, C. Oncogene 2004, 23, 4225440p300/CBP and cancerIyer, Narayanan Gopalakrishna; Oezdag, Hilal; Caldas, CarlosOncogene (2004), 23 (24), 4225-4231CODEN: ONCNES; ISSN:0950-9232. (Nature Publishing Group)A review. P300 and cAMP response element-binding protein (CBP) are adenoviral E1A-binding proteins involved in multiple cellular processes, and function as transcriptional co-factors and histone acetyltransferases. Germline mutation of CBP results in Rubinstein-Taybi syndrome, which is characterized by an increased predisposition to childhood malignancies. Furthermore, somatic mutations of p300 and CBP occur in a no. of malignancies. Chromosome translocations target CBP and, less commonly, p300 in acute myeloid leukemia and treatment-related hematol. disorders. P300 mutations in solid tumors result in truncated p300 protein products or amino-acid substitutions in crit. protein domains, and these are often assocd. with inactivation of the second allele. A mouse model confirms that p300 and CBP function as suppressors of hematol. tumor formation. The involvement of these proteins in crit. tumorigenic pathways (including TGF-β, p53 and Rb) provides a mechanistic route as to how their inactivation could result in cancer.
- 441Varier, R. A.; Kundu, T. K. Curr. Pharm. Des. 2006, 12, 1975441Chromatin modifications (acetylation/ deacetylation/ methylation) as new targets for HIV therapyVarier, Radhika A.; Kundu, Tapas K.Current Pharmaceutical Design (2006), 12 (16), 1975-1993CODEN: CPDEFP; ISSN:1381-6128. (Bentham Science Publishers Ltd.)A review. Human immunodeficiency virus (HIV) is one of the most deadly threats to the human race. Though the developed countries have been able to control the epidemic by utilizing the discovery of very expensive diagnostics, the situation is dangerously alarming in developing and underdeveloped countries. The development of highly active antiretroviral drugs has improved the survival and quality of life, but prolonged treatment results in viral load rebound to pretherapy levels. Recent advances in the understanding of eukaryotic and genome-integrated viral gene expression showed that regulation of chromatin function is closely linked to the multiplication of HIV. Therefore, a new therapeutic approach has been initiated targeting the chromatin-modifying enzymes mainly histone acetyltransferases and deacetylases which may lead to a better and economical anti- HIV combinatorial therapeutics. In this review, the authors have discussed the mechanisms of HIV gene expression in the chromatin context and its potentiality to be exploited as a new therapeutic target.
- 442Mujtaba, S.; Zhou, M. M. Methods 2011, 53, 97There is no corresponding record for this reference.
- 443Sakane, N.; Kwon, H. S.; Pagans, S.; Kaehlcke, K.; Mizusawa, Y.; Kamada, M.; Lassen, K. G.; Chan, J.; Greene, W. C.; Schnoelzer, M.; Ott, M. PLoS Pathog. 2011, 7, e1002184There is no corresponding record for this reference.
- 444Mujtaba, S.; Zhou, M. M. Methods 2011, 53, 97There is no corresponding record for this reference.
- 445Allouch, A.; Di Primio, C.; Alpi, E.; Lusic, M.; Arosio, D.; Giacca, M.; Cereseto, A. Cell Host Microbe 2011, 9, 484445The TRIM Family Protein KAP1 Inhibits HIV-1 IntegrationAllouch, Awatef; Di Primio, Cristina; Alpi, Emanuele; Lusic, Marina; Arosio, Daniele; Giacca, Mauro; Cereseto, AnnaCell Host & Microbe (2011), 9 (6), 484-495CODEN: CHMECB; ISSN:1931-3128. (Cell Press)Summary: The integration of viral cDNA into the host genome is a crit. step in the life cycle of HIV-1. This step is catalyzed by integrase (IN), a viral enzyme that is pos. regulated by acetylation via the cellular histone acetyl transferase (HAT) p300. To investigate the relevance of IN acetylation, we searched for cellular proteins that selectively bind acetylated IN and identified KAP1, a protein belonging to the TRIM family of antiviral proteins. KAP1 binds acetylated IN and induces its deacetylation through the formation of a protein complex which includes the deacetylase HDAC1. Modulation of intracellular KAP1 levels in different cell types including T cells, the primary HIV-1 target, revealed that KAP1 curtails viral infectivity by selectively affecting HIV-1 integration. This study identifies KAP1 as a cellular factor restricting HIV-1 infection and underscores the relevance of IN acetylation as a crucial step in the viral infectious cycle.
- 446Terreni, M.; Valentini, P.; Liverani, V.; Gutierrez, M. I.; Di Primio, C.; Di Fenza, A.; Tozzini, V.; Allouch, A.; Albanese, A.; Giacca, M.; Cereseto, A. Retrovirology 2010, 7, 18There is no corresponding record for this reference.
- 447Zou, W.; Wang, Z.; Liu, Y.; Fan, Y.; Zhou, B. Y.; Yang, X. F.; He, J. J. Glia 2010, 58, 1640There is no corresponding record for this reference.
- 448Kauppi, M.; Murphy, J. M.; de Graaf, C. A.; Hyland, C. D.; Greig, K. T.; Metcalf, D.; Hilton, A. A.; Nicola, N. A.; Kile, B. T.; Hilton, D. J.; Alexander, W. S. Blood 2008, 112, 3148There is no corresponding record for this reference.
- 449Hilton, D. J.; Kile, B. T.; Alexander, W. S. Blood 2009, 113, 5599There is no corresponding record for this reference.
- 450Yanazume, T.; Morimoto, T.; Wada, H.; Kawamura, T.; Hasegawa, K. Mol. Cell. Biochem. 2003, 248, 115There is no corresponding record for this reference.
- 451Gusterson, R. J.; Jazrawi, E.; Adcock, I. M.; Latchman, D. S. J. Biol. Chem. 2003, 278, 6838There is no corresponding record for this reference.
- 452Zhou, X. Y.; Shibusawa, N.; Naik, K.; Porras, D.; Temple, K.; Ou, H.; Kaihara, K.; Roe, M. W.; Brady, M. J.; Wondisford, F. E. Nat. Med. 2004, 10, 633452Insulin regulation of hepatic gluconeogenesis through phosphorylation of CREB-binding proteinZhou, Xiao Yan; Shibusawa, Nobuyuki; Naik, Karuna; Porras, Delia; Temple, Karla; Ou, Hesheng; Kaihara, Kelly; Roe, Michael W.; Brady, Matthew J.; Wondisford, Fredric E.Nature Medicine (New York, NY, United States) (2004), 10 (6), 633-637CODEN: NAMEFI; ISSN:1078-8956. (Nature Publishing Group)Hepatic gluconeogenesis is essential for maintenance of normal blood glucose concns. and is regulated by opposing stimulatory (cAMP) and inhibitory (insulin) signaling pathways. The cAMP signaling pathway leads to phosphorylation of cAMP response element-binding (CREB) protein, resulting in recruitment of the coactivators CREB-binding protein (CBP) and p300 and subsequent activation of gluconeogenesis. Insulin signaling leads to phosphorylation of CBP at serine 436, a residue near its CREB-interacting domain, but it is unknown whether this event modulates cAMP signaling. Here, the authors show in vitro and in 'knock-in' mice that a mutant CBP (S436A) is aberrantly recruited to CREB protein, resulting in inappropriate activation of gluconeogenesis in the fed state and glucose intolerance resulting from increased hepatic glucose prodn. The authors propose that insulin signaling may directly regulate many cAMP signaling pathways at the transcriptional level by controlling CBP recruitment.
- 453Cha-Molstad, H.; Saxena, G.; Chen, J.; Shalev, A. J. Biol. Chem. 2009, 284, 16898There is no corresponding record for this reference.
- 454Bricambert, J.; Miranda, J.; Benhamed, F.; Girard, J.; Postic, C.; Dentin, R. J. Clin. Invest. 2010, 120, 4316There is no corresponding record for this reference.
- 455Schnell, U.; Dijk, F.; Sjollema, K. A.; Giepmans, B. N. Nat. Methods 2012, 9, 152455Immunolabeling artifacts and the need for live-cell imagingSchnell, Ulrike; Dijk, Freark; Sjollema, Klaas A.; Giepmans, Ben N. G.Nature Methods (2012), 9 (2), 152-158CODEN: NMAEA3; ISSN:1548-7091. (Nature Publishing Group)A review. Fluorescent fusion proteins have revolutionized examn. of proteins in living cells. Still, studies using these proteins are met with criticism because proteins are modified and ectopically expressed, in contrast to immunofluorescence studies. However, introducing immunoreagents inside cells can cause protein extn. or relocalization, not reflecting the in vivo situation. Here we discuss pitfalls of immunofluorescence labeling that often receive little attention and argue that immunostaining expts. in dead, permeabilized cells should be complemented with live-cell imaging when scrutinizing protein localization.
- 456McNeil, P. L. J. Cell Sci. 1987, 88, 669456Glass beads load macromolecules into living cellsMcNeil P L; Warder EJournal of cell science (1987), 88 ( Pt 5) (), 669-78 ISSN:0021-9533.We describe and characterize an exceptionally rapid and simple new technique for loading large numbers of cultured cells with large macromolecules. The culture medium of the cell monolayer is replaced by a small volume of the macromolecule to be loaded. Glass beads (75-500 micron diameter) are then sprinkled onto the cells, the cells are washed free of beads and exogenous macromolecules, and 'bead-loading' is completed. The conditions for bead-loading can readily be modified to accommodate cell type and loading objectives: for example, the amount of loading per cell increases if bead size is increased or if beads are agitated after sprinkling onto the monolayer, but at the expense of increased cell loss. As many as 97% of a population of bovine aortic endothelial (BAE) cells were loaded with a 10,000 Mr dextran; and 79% with a 150,000 Mr dextran using bead-loading. Various cell lines have been loaded using glass beads. Moreover, bead-loading has the advantage of producing loaded cells that remain adherent and well-spread, thus minimizing recovery time and permitting immediate microscopic examination.
- 457Hayashi-Takanaka, Y.; Yamagata, K.; Wakayama, T.; Stasevich, T. J.; Kainuma, T.; Tsurimoto, T.; Tachibana, M.; Shinkai, Y.; Kurumizaka, H.; Nozaki, N.; Kimura, H. Nucleic Acids Res. 2011, 39, 6475There is no corresponding record for this reference.
- 458Hayashi-Takanaka, Y.; Yamagata, K.; Nozaki, N.; Kimura, H. J. Cell Biol. 2009, 187, 781There is no corresponding record for this reference.
- 459Johansen, K. M.; Johansen, J. Chromosome Res. 2006, 14, 393459Regulation of chromatin structure by histone H3S10 phosphorylationJohansen, Kristen M.; Johansen, JorgenChromosome Research (2006), 14 (4), 393-404CODEN: CRRSEE; ISSN:0967-3849. (Springer)A review. The epigenetic phospho-serine 10 modification of histone H3 has been a puzzle due to its assocn. with two apparently opposed chromatin states. It is found at elevated levels on the highly condensed, transcriptionally inactive mitotic chromosomes yet is also correlated with the more extended chromatin configuration of active genes, euchromatic interband regions, and activated heat shock puffs of Drosophila polytene chromosomes. In addn., phosphorylation of histone H3S10 is up-regulated on the hypertranscribed male X chromosome. Here we review the cellular effects of histone H3S10 phosphorylation and discuss a model for its involvement in regulating chromatin organization and heterochromatization that would be applicable to both interphase and mitotic chromosomes.
- 460Stasevich, T. J.; Hayashi-Takanaka, Y.; Sato, Y.; Maehara, K.; Ohkawa, Y.; Sakata-Sogawa, K.; Tokunaga, M.; Nagase, T.; Nozaki, N.; McNally, J. G.; Kimura, H. Nature 2014, 516, 272There is no corresponding record for this reference.
- 461Rothbauer, U.; Zolghadr, K.; Tillib, S.; Nowak, D.; Schermelleh, L.; Gahl, A.; Backmann, N.; Conrath, K.; Muyldermans, S.; Cardoso, M. C.; Leonhardt, H. Nat. Methods. 2006, 3, 887461Targeting and tracing antigens in live cells with fluorescent nanobodiesRothbauer, Ulrich; Zolghadr, Kourosh; Tillib, Sergei; Nowak, Danny; Schermelleh, Lothar; Gahl, Anja; Backmann, Natalija; Conrath, Katja; Muyldermans, Serge; Cardoso, M. Cristina; Leonhardt, HeinrichNature Methods (2006), 3 (11), 887-889CODEN: NMAEA3; ISSN:1548-7091. (Nature Publishing Group)The authors fused the epitope-recognizing fragment of heavy-chain antibodies from Camelidae sp. with fluorescent proteins to generate fluorescent, antigen-binding nanobodies (chromobodies) that can be expressed in living cells. The authors demonstrate that chromobodies can recognize and trace antigens in different subcellular compartments throughout S phase and mitosis. Chromobodies should enable new functional studies, as potentially any antigenic structure can be targeted and traced in living cells in this fashion.
- 462Sato, Y.; Mukai, M.; Ueda, J.; Muraki, M.; Stasevich, T. J.; Horikoshi, N.; Kujirai, T.; Kita, H.; Kimura, T.; Hira, S.; Okada, Y.; Hayashi-Takanaka, Y.; Obuse, C.; Kurumizaka, H.; Kawahara, A.; Yamagata, K.; Nozaki, N.; Kimura, H. Sci. Rep. 2013, 3, 2436462Genetically encoded system to track histone modification in vivoSato Yuko; Mukai Masanori; Ueda Jun; Muraki Michiko; Stasevich Timothy J; Horikoshi Naoki; Kujirai Tomoya; Kita Hiroaki; Kimura Taisuke; Hira Seiji; Okada Yasushi; Hayashi-Takanaka Yoko; Obuse Chikashi; Kurumizaka Hitoshi; Kawahara Atsuo; Yamagata Kazuo; Nozaki Naohito; Kimura HiroshiScientific reports (2013), 3 (), 2436 ISSN:.Post-translational histone modifications play key roles in gene regulation, development, and differentiation, but their dynamics in living organisms remain almost completely unknown. To address this problem, we developed a genetically encoded system for tracking histone modifications by generating fluorescent modification-specific intracellular antibodies (mintbodies) that can be expressed in vivo. To demonstrate, an H3 lysine 9 acetylation specific mintbody (H3K9ac-mintbody) was engineered and stably expressed in human cells. In good agreement with the localization of its target acetylation, H3K9ac-mintbody was enriched in euchromatin, and its kinetics measurably changed upon treatment with a histone deacetylase inhibitor. We also generated transgenic fruit fly and zebrafish stably expressing H3K9ac-mintbody for in vivo tracking. Dramatic changes in H3K9ac-mintbody localization during Drosophila embryogenesis could highlight enhanced acetylation at the start of zygotic transcription around mitotic cycle 7. Together, this work demonstrates the broad potential of mintbody and lays the foundation for epigenetic analysis in vivo.
- 463Vincenz, C.; Kerppola, T. K. Proc. Natl. Acad. Sci. U.S.A. 2008, 105, 16572There is no corresponding record for this reference.
- 464Forster, T. H. Ann. Phys. 1948, 437, 55There is no corresponding record for this reference.
- 465Helms, V. Fluorescence Resonance Energy Transfer. Principles of Computational Cell Biology; Wiley-VCH: Weinheim, 2008; p 202, ISBN 978-3-527-31555-0.There is no corresponding record for this reference.
- 466Kanno, T.; Kanno, Y.; Siegel, R. M.; Jang, M. K.; Lenardo, M. J.; Ozato, K. Mol. Cell 2004, 13, 33There is no corresponding record for this reference.
- 467Sasaki, K.; Ito, T.; Nishino, N.; Khochbin, S.; Yoshida, M. Proc. Natl. Acad. Sci. U.S.A. 2009, 106, 16257There is no corresponding record for this reference.
- 468Ito, T.; Umehara, T.; Sasaki, K.; Nakamura, Y.; Nishino, N.; Terada, T.; Shirouzu, M.; Padmanabhan, B.; Yokoyama, S.; Ito, A.; Yoshida, M. Chem. Biol. 2011, 18, 495468Real-Time Imaging of Histone H4K12-Specific Acetylation Determines the Modes of Action of Histone Deacetylase and Bromodomain InhibitorsIto, Tamaki; Umehara, Takashi; Sasaki, Kazuki; Nakamura, Yoshihiro; Nishino, Norikazu; Terada, Takaho; Shirouzu, Mikako; Padmanabhan, Balasundaram; Yokoyama, Shigeyuki; Ito, Akihiro; Yoshida, MinoruChemistry & Biology (Cambridge, MA, United States) (2011), 18 (4), 495-507CODEN: CBOLE2; ISSN:1074-5521. (Cell Press)Histone acetylation constitutes an epigenetic mark for transcriptional regulation. Here we developed a fluorescent probe to visualize acetylation of histone H4 Lys12 (H4K12) in living cells using fluorescence resonance energy transfer (FRET) and the binding of the BRD2 bromodomain to acetylated H4K12. Using this probe designated as Histac-K12, we demonstrated that histone H4K12 acetylation is retained in mitosis and that some histone deacetylase (HDAC) inhibitors continue to inhibit cellular HDAC activity even after their removal from the culture. In addn., a small mol. that interferes with ability of the bromodomain to bind to acetylated H4K12 could be assessed using Histac-K12 in cells. Thus, Histac-K12 will serve as a powerful tool not only to understand the dynamics of H4K12-specific acetylation but also to characterize small mols. that modulate the acetylation or interaction status of histones.
- 469Carrillo, L. D.; Krishnamoorthy, L.; Mahal, L. K. J. Am. Chem. Soc. 2006, 128, 14768There is no corresponding record for this reference.
- 470Newman, R. H.; Zhang, J. Mol. BioSyst. 2008, 4, 496470Visualization of phosphatase activity in living cells with a FRET-based calcineurin activity sensorNewman, Robert H.; Zhang, JinMolecular BioSystems (2008), 4 (6), 496-501CODEN: MBOIBW; ISSN:1742-206X. (Royal Society of Chemistry)Protein kinases and phosphatases are organized into complex intracellular signaling networks designed to coordinate their activities in both space and time. In order to better understand the mol. mechanisms underlying the regulation of signal transduction networks, it is important to define the spatiotemporal dynamics of both protein kinases and phosphatases within their endogenous environment. Herein, we report the development of a genetically-encoded protein biosensor designed to specifically probe the activity of the Ca2+/calmodulin-dependent protein phosphatase, calcineurin. Our reporter design utilizes a phosphatase activity-dependent mol. switch based on the N-terminal regulatory domain of the nuclear factor of activated T-cells as a specific substrate of calcineurin, sandwiched between cyan fluorescent protein and yellow fluorescent protein. Using this reporter, calcineurin activity can be monitored as dephosphorylation-induced increases in fluorescence resonance energy transfer and can be simultaneously imaged with intracellular calcium dynamics. The successful design of a prototype phosphatase activity sensor lays a foundation for studying targeting and compartmentation of phosphatases.
- 471Nagai, Y.; Miyazaki, M.; Aoki, R.; Zama, T.; Inouye, S.; Hirose, K.; Iino, M.; Hagiwara, M. Nat. Biotechnol. 2000, 18, 313There is no corresponding record for this reference.
- 472Ting, A. Y.; Kain, K. H.; Klemke, R. L.; Tsien, R. Y. Proc. Natl. Acad. Sci. U.S.A. 2001, 98, 15003There is no corresponding record for this reference.
- 473Kurokawa, K.; Mochizuki, N.; Ohba, Y.; Mizuno, H.; Miyawaki, A.; Matsuda, M. J. Biol. Chem. 2001, 276, 31305There is no corresponding record for this reference.
- 474Lin, C. W.; Jao, C. Y.; Ting, A. Y. J. Am. Chem. Soc. 2004, 126, 5982There is no corresponding record for this reference.
- 475Lin, C. W.; Ting, A. Y. Angew. Chem., Int. Ed. 2004, 43, 2940There is no corresponding record for this reference.
- 476Yoshida, M.; Kijima, M.; Akita, M.; Beppu, T. J. Biol. Chem. 1990, 265, 17174There is no corresponding record for this reference.
- 477Hanahan, D.; Weinberg, R. A. Cell 2000, 100, 57477The hallmarks of cancerHanahan, Douglas; Weinberg, Robert A.Cell (Cambridge, Massachusetts) (2000), 100 (1), 57-70CODEN: CELLB5; ISSN:0092-8674. (Cell Press)A review is given with many refs. on principles governing the transformation of normal human cells into malignant cancers. The authors suggest that research over the past decades has revealed a small no. of mol., biochem., and cellular traits (acquired capabilities) shared by most and perhaps all types of cancer. Topics included are the acquired capabilities self-sufficiency in growth signals, insensitivity to antigrowth signals, evading apoptosis, limitless replicative potential, sustained angiogenesis, and tissue invasion and metastasis followed by genome instability as an enabling characteristic and alternative pathways to cancer.
- 478Hanahan, D.; Weinberg, R. A. Cell 2011, 144, 646478Hallmarks of cancer: the next generationHanahan, Douglas; Weinberg, Robert A.Cell (Cambridge, MA, United States) (2011), 144 (5), 646-674CODEN: CELLB5; ISSN:0092-8674. (Cell Press)A review. The hallmarks of cancer comprise six biol. capabilities acquired during the multistep development of human tumors. The hallmarks constitute an organizing principle for rationalizing the complexities of neoplastic disease. They include sustaining proliferative signaling, evading growth suppressors, resisting cell death, enabling replicative immortality, inducing angiogenesis, and activating invasion and metastasis. Underlying these hallmarks are genome instability, which generates the genetic diversity that expedites their acquisition, and inflammation, which fosters multiple hallmark functions. Conceptual progress in the last decade has added two emerging hallmarks of potential generality to this list-reprogramming of energy metab. and evading immune destruction. In addn. to cancer cells, tumors exhibit another dimension of complexity: they contain a repertoire of recruited, ostensibly normal cells that contribute to the acquisition of hallmark traits by creating the "tumor microenvironment.". Recognition of the widespread applicability of these concepts will increasingly affect the development of new means to treat human cancer.
- 479Reynoird, N.; Schwartz, B. E.; Delvecchio, M.; Sadoul, K.; Meyers, D.; Mukherjee, C.; Caron, C.; Kimura, H.; Rousseaux, S.; Cole, P. A.; Panne, D.; French, C. A.; Khochbin, S. EMBO J. 2010, 29, 2943There is no corresponding record for this reference.
- 480Crump, N. T.; Hazzalin, C. A.; Bowers, E. M.; Alani, R. M.; Cole, P. A.; Mahadevan, L. C. Proc. Natl. Acad. Sci. U.S.A. 2011, 108, 7814There is no corresponding record for this reference.
- 481Santer, F. R.; Höschele, P. P.; Oh, S. J.; Erb, H. H.; Bouchal, J.; Cavarretta, I. T.; Parson, W.; Meyers, D. J.; Cole, P. A.; Culig, Z. Mol. Cancer Ther. 2011, 10, 1644481Inhibition of the Acetyltransferases p300 and CBP Reveals a Targetable Function for p300 in the Survival and Invasion Pathways of Prostate Cancer Cell LinesSanter, Frederic R.; Hoeschele, Philipp P. S.; Oh, Su Jung; Erb, Holger H. H.; Bouchal, Jan; Cavarretta, Ilaria T.; Parson, Walther; Meyers, David J.; Cole, Philip A.; Culig, ZoranMolecular Cancer Therapeutics (2011), 10 (9), 1644-1655CODEN: MCTOCF; ISSN:1535-7163. (American Association for Cancer Research)Inhibitors of histone deacetylases have been approved for clin. application in cancer treatment. On the other hand, histone acetyltransferase (HAT) inhibitors have been less extensively investigated for their potential use in cancer therapy. In prostate cancer, the HATs and coactivators p300 and CBP are upregulated and may induce transcription of androgen receptor (AR)-responsive genes, even in the absence or presence of low levels of AR. To discover a potential anticancer effect of p300/CBP inhibition, we used two different approaches: (i) downregulation of p300 and CBP by specific short interfering RNA (siRNA) and (ii) chem. inhibition of the acetyltransferase activity by a newly developed small mol., C646. Knockdown of p300 by specific siRNA, but surprisingly not of CBP, led to an increase of caspase-dependent apoptosis involving both extrinsic and intrinsic cell death pathways in androgen-dependent and castration-resistant prostate cancer cells. Induction of apoptosis was mediated by several pathways including inhibition of AR function and decrease of the nuclear factor kappa B (NF-κB) subunit p65. Furthermore, cell invasion was decreased upon p300, but not CBP, depletion and was accompanied by lower matrix metalloproteinase (MMP)-2 and MMP-9 transcriptions. Thus, p300 and CBP have differential roles in the processes of survival and invasion of prostate cancer cells. Induction of apoptosis in prostate cancer cells was confirmed by the use of C646. This was substantiated by a decrease of AR function and downregulation of p65 impairing several NF-κB target genes. Taken together, these results suggest that p300 inhibition may be a promising approach for the development of new anticancer therapies. Mol Cancer Ther; 10(9); 1644-55.
- 482Wang, Y.; Toh, H. C.; Chow, P.; Chung, A. Y.; Meyers, D. J.; Cole, P. A.; Ooi, L. L.; Lee, C. G. FASEB J. 2012, 26, 3032There is no corresponding record for this reference.
- 483Yan, G.; Eller, M. S.; Elm, C.; Larocca, C. A.; Ryu, B.; Panova, I. P.; Dancy, B. M.; Bowers, E. M.; Meyers, D.; Lareau, L.; Cole, P. A.; Taverna, S. D.; Alani, R. M. J. Invest. Dermatol. 2013, 133, 2444There is no corresponding record for this reference.
- 484Santer, F. R.; Höschele, P. P.; Oh, S. J.; Erb, H. H.; Bouchal, J.; Cavarretta, I. T.; Parson, W.; Meyers, D. J.; Cole, P. A.; Culig, Z. Mol. Cancer Ther. 2011, 10, 1644484Inhibition of the Acetyltransferases p300 and CBP Reveals a Targetable Function for p300 in the Survival and Invasion Pathways of Prostate Cancer Cell LinesSanter, Frederic R.; Hoeschele, Philipp P. S.; Oh, Su Jung; Erb, Holger H. H.; Bouchal, Jan; Cavarretta, Ilaria T.; Parson, Walther; Meyers, David J.; Cole, Philip A.; Culig, ZoranMolecular Cancer Therapeutics (2011), 10 (9), 1644-1655CODEN: MCTOCF; ISSN:1535-7163. (American Association for Cancer Research)Inhibitors of histone deacetylases have been approved for clin. application in cancer treatment. On the other hand, histone acetyltransferase (HAT) inhibitors have been less extensively investigated for their potential use in cancer therapy. In prostate cancer, the HATs and coactivators p300 and CBP are upregulated and may induce transcription of androgen receptor (AR)-responsive genes, even in the absence or presence of low levels of AR. To discover a potential anticancer effect of p300/CBP inhibition, we used two different approaches: (i) downregulation of p300 and CBP by specific short interfering RNA (siRNA) and (ii) chem. inhibition of the acetyltransferase activity by a newly developed small mol., C646. Knockdown of p300 by specific siRNA, but surprisingly not of CBP, led to an increase of caspase-dependent apoptosis involving both extrinsic and intrinsic cell death pathways in androgen-dependent and castration-resistant prostate cancer cells. Induction of apoptosis was mediated by several pathways including inhibition of AR function and decrease of the nuclear factor kappa B (NF-κB) subunit p65. Furthermore, cell invasion was decreased upon p300, but not CBP, depletion and was accompanied by lower matrix metalloproteinase (MMP)-2 and MMP-9 transcriptions. Thus, p300 and CBP have differential roles in the processes of survival and invasion of prostate cancer cells. Induction of apoptosis in prostate cancer cells was confirmed by the use of C646. This was substantiated by a decrease of AR function and downregulation of p65 impairing several NF-κB target genes. Taken together, these results suggest that p300 inhibition may be a promising approach for the development of new anticancer therapies. Mol Cancer Ther; 10(9); 1644-55.
- 485Wang, L.; Gural, A.; Sun, X. J.; Zhao, X.; Perna, F.; Huang, G.; Hatlen, M. A.; Vu, L.; Liu, F.; Xu, H.; Asai, T.; Xu, H.; Deblasio, T.; Menendez, S.; Voza, F.; Jiang, Y.; Cole, P. A.; Zhang, J.; Melnick, A.; Roeder, R. G.; Nimer, S. D. Science 2011, 333, 765There is no corresponding record for this reference.
- 486Liu, Y.; Wang, L.; Predina, J.; Han, R.; Beier, U. H.; Wang, L. C.; Kapoor, V.; Bhatti, T. R.; Akimova, T.; Singhal, S.; SBrindle, P. K.; Cole, P. A.; Albelda, S. M.; Hancock, W. W. Nat. Med. 2013, 19, 1173486Inhibition of p300 impairs Foxp3+ T regulatory cell function and promotes antitumor immunityLiu, Yujie; Wang, Liqing; Predina, Jarrod; Han, Rongxiang; Beier, Ulf H.; Wang, Liang-Chuan S.; Kapoor, Veena; Bhatti, Tricia R.; Akimova, Tatiana; Singhal, Sunil; Brindle, Paul K.; Cole, Philip A.; Albelda, Steven M.; Hancock, Wayne W.Nature Medicine (New York, NY, United States) (2013), 19 (9), 1173-1177CODEN: NAMEFI; ISSN:1078-8956. (Nature Publishing Group)Forkhead box P3 (Foxp3)+ T regulatory (Treg) cells maintain immune homeostasis and limit autoimmunity but can also curtail host immune responses to various types of tumors. Foxp3+ Treg cells are therefore considered promising targets to enhance antitumor immunity, and approaches for their therapeutic modulation are being developed. However, although studies showing that exptl. depleting Foxp3+ Treg cells can enhance antitumor responses provide proof of principle, these studies lack clear translational potential and have various shortcomings. Histone/protein acetyltransferases (HATs) promote chromatin accessibility, gene transcription and the function of multiple transcription factors and nonhistone proteins. We now report that conditional deletion or pharmacol. inhibition of one HAT, p300 (also known as Ep300 or KAT3B), in Foxp3+ Treg cells increased T cell receptor-induced apoptosis in Treg cells, impaired Treg cell suppressive function and peripheral Treg cell induction, and limited tumor growth in immunocompetent but not in immunodeficient mice. Our data thereby demonstrate that p300 is important for Foxp3+ Treg cell function and homeostasis in vivo and in vitro, and identify mechanisms by which appropriate small-mol. inhibitors can diminish Treg cell function without overtly impairing T effector cell responses or inducing autoimmunity. Collectively, these data suggest a new approach for cancer immunotherapy.
- 487Min, S. W.; Cho, S. H.; Zhou, Y.; Schroeder, S.; Haroutunian, V.; Seeley, W. W.; Huang, E. J.; Shen, Y.; Masliah, E.; Mukherjee, C.; Meyers, D.; Cole, P. A.; Ott, M.; Gan, L. Neuron 2010, 67, 953
Erratum in: Neuron2010, 68, 801
There is no corresponding record for this reference. - 488Mali, P.; Chou, B. K.; Yen, J.; Ye, Z.; Zou, J.; Dowey, S.; Brodsky, R. A.; Ohm, J. E.; Yu, W.; Baylin, S. B.; Yusa, K.; Bradley, A.; Meyers, D. J.; Mukherjee, C.; Cole, P. A.; Cheng, L. Stem Cells 2010, 28, 713There is no corresponding record for this reference.
- 489Xu, C. R.; Cole, P. A.; Meyers, D. J.; Kormish, J.; Dent, S.; Zaret, K. S. Science 2011, 332, 963There is no corresponding record for this reference.
- 490Marek, R.; Coelho, C. M.; Sullivan, R. K.; Baker-Andresen, D.; Li, X.; Ratnu, V.; Dudley, K. J.; Meyers, D.; Mukherjee, C.; Cole, P. A.; Sah, P.; Bredy, T. W. J. Neurosci. 2011, 31, 7486490Paradoxical enhancement of fear extinction memory and synaptic plasticity by inhibition of the histone acetyltransferase p300Marek, Roger; Coelho, Carlos M.; Sullivan, Robert K. P.; Baker-Andresen, Danay; Li, Xiang; Ratnu, Vikram; Dudley, Kevin J.; Meyers, David; Mukherjee, Chandrani; Cole, Philip A.; Sah, Pankaj; Bredy, Timothy W.Journal of Neuroscience (2011), 31 (20), 7486-7491CODEN: JNRSDS; ISSN:0270-6474. (Society for Neuroscience)It is well established that the coordinated regulation of activity-dependent gene expression by the histone acetyltransferase (HAT) family of transcriptional coactivators is crucial for the formation of contextual fear and spatial memory, and for hippocampal synaptic plasticity. However, no studies have examd. the role of this epigenetic mechanism within the infralimbic prefrontal cortex (ILPFC), an area of the brain that is essential for the formation and consolidation of fear extinction memory. Here, the authors report that a postextinction training infusion of a combined p300/CBP inhibitor (Lys-CoA-Tat), directly into the ILPFC, enhances fear extinction memory in mice. These results also demonstrated that the HAT p300 was highly expressed within pyramidal neurons of the ILPFC and that the small-mol. p300-specific inhibitor (C 646) infused into the ILPFC immediately after weak extinction training enhanced the consolidation of fear extinction memory. C 646 infused 6 h after extinction had no effect on fear extinction memory, nor did an immediate postextinction training infusion into the prelimbic prefrontal cortex. Consistent with the behavioral findings, inhibition of p300 activity within the ILPFC facilitated long-term potentiation (LTP) under stimulation conditions that did not evoke long-lasting LTP. These data suggested that one function of p300 activity within the ILPFC is to constrain synaptic plasticity, and that a redn. in the function of this HAT is required for the formation of fear extinction memory.
- 491Knight, Z. A.; Shokat, K. M. Cell 2007, 128, 425There is no corresponding record for this reference.
- 492Kasper, L. H.; Fukuyama, T.; Biesen, M. A.; Boussouar, F.; Tong, C.; de Pauw, A.; Murray, P. J.; van Deursen, J. M.; Brindle, P. K. Mol. Cell. Biol. 2006, 26, 789492Conditional knockout mice reveal distinct functions for the global transcriptional coactivators CBP and p300 in T-cell developmentKasper, Lawryn H.; Fukuyama, Tomofusa; Biesen, Michelle A.; Boussouar, Faycal; Tong, Caili; de Pauw, Antoine; Murray, Peter J.; van Deursen, Jan M. A.; Brindle, Paul K.Molecular and Cellular Biology (2006), 26 (3), 789-809CODEN: MCEBD4; ISSN:0270-7306. (American Society for Microbiology)The global transcriptional coactivators CREB-binding protein (CBP) and the closely related p300 interact with over 312 proteins, making them among the most heavily connected hubs in the known mammalian protein-protein interactome. It is largely uncertain, however, if these interactions are important in specific cell lineages of adult animals, as homozygous null mutations in either CBP or p300 result in early embryonic lethality in mice. Here we describe a Cre/LoxP conditional p300 null allele (p300flox) that allows for the temporal and tissue-specific inactivation of p300. We used mice carrying p300flax and a CBP conditional knockout allele (CBPflax) in conjunction with an Lck-Cre transgene to delete CBP and p300 starting at the CD4- CD8- double-neg. thymocyte stage of T-cell development. Loss of either p300 or CBP led to a decrease in CD4+ CD8+ double-pos. thymocytes, but an increase in the percentage of CD8+ single-pos. thymocytes seen in CBP mutant mice was not obsd. in p300 mutants. T cells completely lacking both CBP and p300 did not develop normally and were nonexistent or very rare in the periphery, however. T cells lacking CBP or p300 had reduced tumor necrosis factor alpha gene expression in response to phorbol ester and ionophore, while signal-responsive gene expression in CBP- or p300-deficient macrophages was largely intact. Thus, CBP and p300 each supply a surprising degree of redundant coactivation capacity in T cells and macrophages, although each gene has also unique properties in thymocyte development.
- 493Kasper, L. H.; Lerach, S.; Wang, J.; Wu, S.; Jeevan, T.; Brindle, P. K. EMBO J. 2010, 29, 3660There is no corresponding record for this reference.
- 494Phan, H. M.; Xu, A. W.; Coco, C.; Srajer, G.; Wyszomierski, S.; Evrard, Y. A.; Eckner, R.; Dent, S. Y. Dev. Dyn. 2005, 233, 1337There is no corresponding record for this reference.
- 495Huang, Y.; Vasilatos, S. N.; Boric, L.; Shaw, P. G.; Davidson, N. E. Breast Cancer Res. Treat. 2012, 131, 777– 89495Inhibitors of histone demethylation and histone deacetylation cooperate in regulating gene expression and inhibiting growth in human breast cancer cellsHuang, Yi; Vasilatos, Shauna N.; Boric, Lamia; Shaw, Patrick G.; Davidson, Nancy E.Breast Cancer Research and Treatment (2012), 131 (3), 777-789CODEN: BCTRD6; ISSN:0167-6806. (Springer)Abnormal activities of histone lysine demethylases (KDMs) and lysine deacetylases (HDACs) are assocd. with aberrant gene expression in breast cancer development. However, the precise mol. mechanisms underlying the crosstalk between KDMs and HDACs in chromatin remodeling and regulation of gene transcription are still elusive. In this study, we showed that treatment of human breast cancer cells with inhibitors targeting the zinc cofactor dependent class I/II HDAC, but not NAD+ dependent class III HDAC, led to significant increase of H3K4me2 which is a specific substrate of histone lysine-specific demethylase 1 (LSD1) and a key chromatin mark promoting transcriptional activation. We also demonstrated that inhibition of LSD1 activity by a pharmacol. inhibitor, pargyline, or siRNA resulted in increased acetylation of H3K9 (AcH3K9). However, siRNA knockdown of LSD2, a homolog of LSD1, failed to alter the level of AcH3K9, suggesting that LSD2 activity may not be functionally connected with HDAC activity. Combined treatment with LSD1 and HDAC inhibitors resulted in enhanced levels of H3K4me2 and AcH3K9, and exhibited synergistic growth inhibition of breast cancer cells. Finally, microarray screening identified a unique subset of genes whose expression was significantly changed by combination treatment with inhibitors of LSD1 and HDAC. Our study suggests that LSD1 intimately interacts with histone deacetylases in human breast cancer cells. Inhibition of histone demethylation and deacetylation exhibits cooperation and synergy in regulating gene expression and growth inhibition, and may represent a promising and novel approach for epigenetic therapy of breast cancer.
- 496Han, H.; Yang, X.; Pandiyan, K.; Liang, G. PLoS One 2013, 8, e75136There is no corresponding record for this reference.
- 497Prusevich, P.; Kalin, J. H.; Ming, S. A.; Basso, M.; Givens, J.; Li, X.; Hu, J.; Taylor, M. S.; Cieniewicz, A. M.; Hsiao, P. Y.; Huang, R.; Roberson, H.; Adejola, N.; Avery, L. B.; Casero, R. A., Jr.; Taverna, S. D.; Qian, J.; Tackett, A. J.; Ratan, R. R.; McDonald, O. G.; Feinberg, A. P.; Cole, P. A. ACS Chem. Biol. 2014, 9, 1284There is no corresponding record for this reference.
- 498Yan, G.; Eller, M. S.; Elm, C.; Larocca, C. A.; Ryu, B.; Panova, I. P.; Dancy, B. M.; Bowers, E. M.; Meyers, D.; Lareau, L.; Cole, P. A.; Taverna, S. D.; Alani, R. M. J. Invest. Dermatol. 2013, 133, 2444There is no corresponding record for this reference.
- 499Das, N.; Dhanawat, M.; Dash, B.; Nagarwal, R. C.; Shrivastava, S. K. Eur. J. Pharm. Sci. 2010, 41, 571499Codrug: An efficient approach for drug optimizationDas, N.; Dhanawat, M.; Dash, B.; Nagarwal, R. C.; Shrivastava, S. K.European Journal of Pharmaceutical Sciences (2010), 41 (5), 571-588CODEN: EPSCED; ISSN:0928-0987. (Elsevier B.V.)A review. Codrug or mutual prodrug is an approach where various effective drugs, which are assocd. with some drawbacks, can be modified by attaching with other drugs of same or different categories directly or via a linkage. More appropriately one can say combining two different pharmacophores with similar or different pharmacol. activities elicit synergistic action or help to target the parent drug to specific site/organ/cells resp. This approach is commonly used to improve physicochem., biopharmaceutical and drug delivery properties of therapeutic agents.
Supporting Information
Supporting Information
Table of 98 protein acetylation substrates of p300/CBP and table of 411 protein binding partners of p300/CBP. This material is available free of charge via the Internet at http://pubs.acs.org.
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.