Gcn5 ortholog
DEVELOPMENTAL BIOLOGY

Embryonic

Ada2 proteins are present in many Gcn5-containing complexes. To confirm that Drosophila Ada2b and Gcn5 are able to interact, a GST pull-down assay was performed. Bacterially expressed GST-dGcn5 efficiently binds to in vitro-translated dAda2b protein. Embryonic expression pattern of dAda2b was compared to that of dGcn5 by in situ hybridization. Both dAda2b and dGcn5 are maternally contributed; their transcripts can be detected at the precellular stage. This finding was confirmed by staining ovaries with the dAda2b and dGcn5 probes. Both transcripts are present in nurse cells of stage 10 egg chambers and are dumped in the oocyte when the nurse cells degenerate at late oogenesis (Qi, 2004).

In wild-type embryos undergoing cellularization, no maternal dAda2b mRNA remains, whereas that of dGcn5 persists. In advanced-stage embryos, zygotic expression of both dAda2b and dGcn5 is strongest in the central nervous system (CNS) and is present at lower levels in most other tissues. These results suggest that dAda2b and dGcn5 are found in many of the same cells during embryo development. As expected, in dAda2b homozygous embryos, no dAda2b mRNA could be detected after stage 5, when the maternal message is no longer present. In contrast, the maternal protein persists until late in stage 15, at which point reduced dAda2b protein levels are found in homozygous mutant embryos. In order to investigate the contribution of dAda2b to early embryo development, attempts were made to remove maternal dAda2b in germ line clones. However, dAda2b homozygous germ cells arrest at an early stage of oogenesis, and for this reason it was not possible to assess dAda2b function in early embryo development (Qi, 2004).

Effects of Mutation or Deletion

The insertion sequences of several P elements that cytologically mapped close to the 69C4 position of Gcn5 on polytene chromosomes were determined, as well as which ones were lethal over the Df(3L)iro-2 deficiency. The 1479/10 P element was found inserted 521 bp upstream of the Gcn5 transcription start site, in the coding sequence of the CG14121 gene. This insertion was used to generate deletions via transposase-mediated P-element excision and the lethal deficiency sex204 was recovered that removes a part of the CG14121 coding sequence, the Gcn5 and CG10686 genes, and the transcription start site of the citron gene. To obtain Gcn5-specific alleles, a screen was performed for EMS-induced Gcn5 mutants by noncomplementation of the sex204 deficiency. Eleven mutants were obtained that failed to complement the sex204 deficiency for viability. Five mutants were rescued by a PL transgenic construct encompassing the Gcn5 and CG14121 genes. Four of these alleles were fully rescued by a UAS-Gcn5 cDNA transgene expressed under the control of the ubiquitous da-GAL4 driver, indicating that these alleles specifically impair Gcn5 function. The remaining allele is a mutation of the CG14121 gene. The Gcn5 mutations were sequenced and it was found that they all lie within the first dGcn5 exon. Two of them are localized in the Pcaf homology domain and result in a cysteine-to-tyrosine substitution (C137T) or a small deletion from tyrosine 280 to phenylalanine 285 (DeltaT280-F285). The two other mutations generate stop codons (Q186st and E333st). Gcn5E333st mutants were outcrossed and then rescued as homozygous by the PL transgene, indicating that they do not bear other irrelevant EMS-induced lethal mutations (Carré, 2005).

To determine the lethal phase associated with the Gcn5 loss of function, the Gcn5E333st allele was maintained over a GFP-expressing balancer chromosome. About 26% of eggs laid by females from this stock did not reach the first larval instar, a number attributable to embryonic lethality from the homozygous balancer chromosome. Nonfluorescent homozygous Gcn5E333st larvae developed similarly to their GFP-expressing fluorescent heterozygous siblings until the end of the third larval instar, despite the absence of detectable Gcn5 protein in mutant tissues at this stage. However, Gcn5E333st mutant larvae did not form their puparium at the normal time and continued to wander for 4 to 5 additional days. They eventually stopped moving but failed to evert spiracles, formed abnormally long prepupae, and died with partial separation of the posterior part of the prepupa from the pupal case. Similar developmental arrest and defects were observed with the heteroallelic null combination Gcn5Q186st/Gcn5E333st (Carré, 2005).

Puparium formation is triggered by a pulse of 20-hydroxyecdysone at the end of the third larval instar. This involves the coordinated induction of a small number of early puff genes, whose products in turn regulate the expression of a larger set of late puff genes. Homozygous Gcn5E333st mutants were taken at various times during their extended wandering stage, and polytene chromosomes from their salivary glands were squashed. A strong reduction was observed of the size of the early puffs 2B, 74EF, and 75B in these animals, indicating a failure of the ecdysone-controlled genetic program in Gcn5 mutants (Carré, 2005).

Heteroallelic Gcn5C137T/Gcn5E333st and Gcn5E333st/Gcn5DeltaT280-F285 mutant larvae form their puparium normally, indicating that the Gcn5C137T and Gcn5DeltaT280-F285 variant proteins retain partial function. However, these mutants fail to elongate their metathoracic legs correctly, have rough eyes, and display a partial to complete absence of abdominal cuticle deposition. Both heteroallelic combinations give rise to rare adult escapers with misshapen wings, rough eyes, and crooked and twisted metathoracic legs. These escapers died a few hours after eclosion. Altogether, these data point to an essential function of Gcn5 at the onset of and during metamorphosis (Carré, 2005).

Since an important stock of Gcn5 protein is detected in oocytes and presyncytial embryos, this maternal contribution of Gcn5 may be sufficient to allow embryonic and larval development. To generate embryos lacking a maternal contribution, advantage was taken of the absence of expression of the pUAST-derived construct UAS-Gcn5 in the female germ line. UAS-Gcn5/+; Gcn5E333st/sex204 da-GAL4 rescued adults were generated, and oogenesis in females was found to be arrested at stages 5 and 6. This effect was due to the lack of UAS-Gcn5 expression in the germ line, since control females rescued by the Gcn5 genomic construct (PL/+; Gcn5E333st/sex204) were fertile. With Gcn5 being required for oogenesis, its contribution to embryonic development was not examined (Carré, 2005).

Imaginal discs from homozygous Gcn5E333st mutant third-instar larvae are misshapen and severely reduced in size, suggesting that a Gcn5 loss of function impairs the proliferation of imaginal cells during larval instars. In a previous report, use was made of the inverted-repeat transgenic construct UAS-IR[Gcn5] to target RNA interference (RNAi) against Gcn5 (Roignant, 2003). To further analyze the role of Gcn5 in the cell proliferation of imaginal tissues, a UAS-IR[gcn5] transgenic line was crossed with en-GAL4 UAS-GFP individuals expressing the GAL4 activator. The en-GAL4 driver induced both GFP expression and specific dGcn5 depletion in the posterior (P) compartments of imaginal discs of third-larval-instar progeny. In a control experiment, silencing of the unrelated CBP protein was not observed. The Gcn5-depleted P compartments often had a reduced size and appeared flatter than anterior compartments, suggesting that cell proliferation is slowed down in these compartments. Although Gcn5 RNAi in the P compartments of imaginal discs induced strong lethality during late pupal development, few animals survived until the adult stage. The posterior part of their wings was reduced in size and displayed abnormal veins and cross veins. In the most dramatic cases, a bubble indicative of abnormal adhesion between the dorsal and ventral wing epithelial sheets was observed. The cell density was not significantly changed in the silenced compartment, indicating that the size reduction is due to a reduction in the cell number. A vg-GAL4 driver triggered Gcn5 RNAi in the wing margin and induced a strong reduction of this structure in silenced adults, while the induction of Gcn5 RNAi using a dll-GAL4 driver led to a reduction in the size of the distal part of the tarsus and wings. Finally, the induction of Gcn5 RNAi by an esg-GAL4 driver, which is strongly expressed in imaginal histoblasts, resulted in lethality of pharate adults, with a partial to complete absence of the abdominal cuticle. Collectively, these data strongly suggested that Gcn5 RNAi limits cell proliferation (Carré, 2005).

Surprisingly, however, a greater proportion of cells in S phase as well as a significantly larger number of cells undergoing mitosis was observed in the P compartments of wing discs silenced by UAS-IR[Gcn5] upon activation by en-GAL4. Notably, a high level of apoptosis was detected in imaginal discs from third-instar Gcn5 mutant larvae as well as in response to Gcn5 RNAi triggered by either en-GAL4 or ptc-GAL4. Together, these data suggest that the net effect of the Gcn5 loss of function on the compartment size results from a deregulation of the cell cycle coupled to cell death (Carré, 2005).

The biological function of the WD40 repeat-containing protein p55/Caf1 in Drosophila

The p55 family WD40 repeat-containing histone chaperone proteins are components of several chromatin regulatory complexes (such as PRC2, NURF and CAF-1) and interact with histone H4, yet their functional relevance in vivo is unclear. This study used Drosophila as a genetic model to dissect the function of p55/Caf1 during development. p55 was found essential for Drosophila development and is required for cell proliferation and viability. However, the data further demonstrate that histone H3K27 di-/tri-methylation and PRC2-mediated gene silencing still occur normally when p55 is missing. p55 is also implicated in bridging chromatin regulatory complexes to the chromatin by binding to histone H4, but a transgene of p55 whose binding pocket is disrupted could still functionally substitute the wild-type p55 for the survival. These studies suggest that p55 is not crucial for PRC2-mediated gene silencing in vivo, and the vital function of p55 is probably not dependent on its interaction with histone H4 (Wen, 2012).


REFERENCES

Reference names in red indicate recommended papers.

Agalioti, T., Chen, G. and Thanos, D. (2002). Deciphering the transcriptional histone acetylation code for a human gene. Cell 111: 381-392. 12419248

Bertrand, C., et al. (2003). Arabidopsis histone acetyltransferase AtGCN5 regulates the floral meristem activity through the WUSCHEL/AGAMOUS pathway. J. Biol. Chem. 278: 28246-28251. 12740375

Biswas, D., Yu, Y., Mitra, D. and Stillman, D. (2005). Genetic interactions between Nhp6 and Gcn5 with Mot1 and the Ccr4-Not Complex that regulate TBP binding in Saccharomyces cerevisiae. Genetics [Epub ahead of print]. 16272410

Blanco, J. C., et al. (1998). The histone acetylase PCAF is a nuclear receptor coactivator. Genes Dev. 12: 1638-1651.

Bouchard, C., et al, (2001). Regulation of cyclin D2 gene expression by the Myc/Max/Mad network: Myc-dependent TRRAP recruitment and histone acetylation at the cyclin D2 promoter. Genes Dev. 15: 2042-2047. 11511535

Boyer, L. A., et al. (2002). Essential role for the SANT domain in the functioning of multiple chromatin remodeling enzymes. Mol. Cell 10: 935-942. 12419236

Brand, M., Yamamoto, K., Staub, A. and Tora, L. (1999). Identification of TATA-binding protein-free TAFII-containing complex subunits suggests a role in nucleosome acetylation and signal transduction. J. Biol Chem. 274: 18285-18289. 10373431

Brown, C. E., et al. (2001). Recruitment of HAT complexes by direct activator interactions with the ATM-related Tra1 subunit. Science 292: 2333-2337. 11423663

Brownell, J. E., et al. (1996). Tetrahymena histone acetyltransferase A: a homolog to yeast Gcn5p linking histone acetylation to gene activation. Cell 84: 843-851. 8601308

Bu, P., Evrard, Y. A., Lozano, G. and Dent, S. Y. (2007). Loss of Gcn5 acetyltransferase activity leads to neural tube closure defects and exencephaly in mouse embryos. Mol. Cell Biol. 27: 3405-3416. PubMed Citation: 17325035

Candau, R., and Berger, S. L. (1996). Structural and functional analysis of yeast putative adaptors. Evidence for an adaptor complex in vivo. J. Biol. Chem. 271: 5237-5245. 8617808

Candau, R., Zhou, J. X., Allis, C. D. and Berger, S. L. (1997). Histone acetyltransferase activity and interaction with ADA2 are critical for GCN5 function in vivo. EMBO J. 16: 555-565. 9034338

Carré, C., Szymczak, D., Pidoux, J. and Antoniewski, C. (2005). The Histone H3 acetylase dGcn5 is a key player in Drosophila melanogaster metamorphosis. Molec. Cell. Biol. 25: 8228-8238. 16135811

Dhasarathy, A. and Kladde, M. P. (2005). Promoter occupancy is a major determinant of chromatin remodeling enzyme requirements. Mol. Cell. Biol. 25(7): 2698-707. 15767675

Faiola, F., et al. (2005). Dual regulation of c-Myc by p300 via acetylation-dependent control of Myc protein turnover and coactivation of Myc-induced transcription. Mol. Cell. Biol. 25(23): 10220-34. 16287840

Georgieva, S., et al. (2000). Two novel Drosophila TAF(II)s have homology with human TAF(II)30 and are differentially regulated during development. Mol. Cell. Biol. 20: 1639-1648. 10669741

Grant, P. A., L. Duggan, J. Cote, S. M. Roberts, J. E. Brownell, R. Candau, R. Ohba, T. Owen-Hughes, C. D. Allis, F. Winston, S. L. Berger, and J. L. Workman. (1997). Yeast Gcn5 functions in two multisubunit complexes to acetylate nucleosomal histones: characterization of an Ada complex and the SAGA (Spt/Ada) complex. Genes Dev. 11: 1640-1650. 9224714

Grant, P. A., et al. (1999). Expanded lysine acetylation specificity of Gcn5 in native complexes. J. Biol. Chem. 274: 5895-5900. 10026213

Guelman, S., et al. (2006). Host cell factor and an uncharacterized SANT domain protein are stable components of ATAC, a novel dAda2A/dGcn5-containing histone acetyltransferase complex in Drosophila. Mol. Cell. Biol. 26(3): 871-82. 16428443

Hassan, A. H., et al. (2002). Function and selectivity of bromodomains in anchoring chromatin-modifying complexes to promoter nucleosomes. Cell 111: 369-379. 12419247

Herrera, J. E., et al. (1997). The histone acetyltransferase activity of human GCN5 and PCAF is stabilized by coenzymes. J. Biol. Chem. 272(43): 27253-27258. 9341171

Hudson, B. P., et al. (2000). Solution structure and acetyl-lysine binding activity of the GCN5 bromodomain. J. Mol. Biol. 304: 355-370. 11090279

Kanno, T., et al. (2004). Selective recognition of acetylated histones by bromodomain proteins visualized in living cells. Mol. Cell 13:33-43. 14731392

Kikuchi, H., Takami, Y. and Nakayama, T. (2005). GCN5: a supervisor in all-inclusive control of vertebrate cell cycle progression through transcription regulation of various cell cycle-related genes. Gene 347(1): 83-97. 15715965

Koutelou, E., Hirsch, C. L. and Dent, S. Y. (2010). Multiple faces of the SAGA complex. Curr. Opin. Cell Biol. 22: 374-382. PubMed Citation: 20363118

Krebs, J. E., et al. (2000). Global role for chromatin remodeling enzymes in mitotic gene expression. Cell 102: 587-598. 11007477

Kristjuhan, A., et al. (2002). Transcriptional inhibition of genes with severe Histone H3 hypoacetylation in the coding region. Mol. Cell 10: 925-933. 12419235

Kuo, M. H., et al. (1996). Transcription-linked acetylation by Gcn5p of histones H3 and H4 at specific lysines. Nature 383: 269-272. 8805705

Kusch, T., Guelman, S., Abmayr, S. M. and Workman. J. L. (2003). Two Drosophila Ada2 homologues function in different multiprotein complexes. Mol. Cell. Biol. 23: 3305-3319. 12697829

Lebedeva, L. A., et al. (2005). Occupancy of the Drosophila hsp70 promoter by a subset of basal transcription factors diminishes upon transcriptional activation. Proc. Natl. Acad. Sci. [Epub ahead of print] . 16330756

Liu, Y., et al. (2005). Histone H3 Ser10 phosphorylation-independent function of snf1 and reg1 proteins rescues a gcn5- mutant in HIS3 expression. Mol. Cell. Biol. 25(23): 10566-79. 16287868

Lo, W. S., et al. (2005). Histone H3 phosphorylation can promote TBP recruitment through distinct promoter-specific mechanisms. EMBO J. 24: 997-1008. 15719021

Luciano, R. L. and Wilson, A. C. (2000). N-terminal transcriptional activation domain of LZIP comprises two LxxLL motifs and the host cell factor-1 binding motif. Proc. Natl. Acad. Sci. 97: 10757-10762. 10984507

Luciano, R. L., and Wilson, A. C. (2003). HCF-1 functions as a coactivator for the zinc finger protein Krox20. J. Biol. Chem. 278: 51116-51124. 14532282

Marcus, G. A., et al. (1994). Functional similarity and physical association between GCN5 and ADA2: putative transcriptional adaptors. EMBO J. 13: 4807-4815. 7957049

Martinez, E., et al. (2001). Human STAGA complex is a chromatin-acetylating transcription coactivator that interacts with pre-mRNA splicing and DNA damage- binding-factors in vivo. Mol. Cell. Biol. 21: 6782-6795. 11564863

Martinez-Balbas, M. A., (2000). Regulation of E2F1 activity by acetylation. EMBO J. 19: 662-671. 10675335

Muratoglu, S., et al. (2003). Two different Drosophila ADA2 homologues are present in distinct GCN5 histone acetyltransferase-containing complexes. Mol. Cell. Biol. 23: 306-321. 12482983

Ornaghi, P., et al. (1999). The bromodomain of Gcn5p interacts in vitro with specific residues in the N terminus of histone H4. J. Mol. Biol. 287: 1-7. 10074402

Owen, D. J., et al. (2000). The structural basis for the recognition of acetylated histone H4 by the bromodomain of histone acetyltransferase Gcn5p. EMBO J. 19: 6141-6149. 11080160

Pankotai, T., et al. (2005). The homologous Drosophila transcriptional adaptors ADA2a and ADA2b are both required for normal development but have different functions. Mol. Cell. Biol. 25(18): 8215-27. 16135810

Papai, G., et al. (2005). Intimate relationship between the genes of two transcriptional coactivators, ADA2a and PIMT, of Drosophila. Gene 348: 13-23. 15777699

Patel, J. H., et al. (2005). The c-MYC oncoprotein is a substrate of the acetyltransferases hGCN5/PCAF and TIP60. Mol. Cell. Biol. 24(24): 10826-34. 15572685

Phan, H. M., et al. (2005). GCN5 and p300 share essential functions during early embryogenesis. Dev. Dyn. 233(4): 1337-47. 15937931

Pray-Grant, M. G., et al. (2005). Chd1 chromodomain links histone H3 methylation with SAGA- and SLIK-dependent acetylation. Nature 433: 434-438. 15647753

Qi, D., Larsson, J. and Mannervik. M. (2004). Drosophila Ada2b is required for viability and normal histone H3 acetylation. Mol. Cell. Biol. 24: 8080-8089. 15340070

Reid, J. L., et al. (1998). E1A directly binds and regulates the P/CAF acetyltransferase. EMBO J. 17: 4469-4477. 9687513

Roignant, J. Y., et al. (2003). Absence of transitive and systemic pathways allows cell-specific and isoform-specific RNAi in Drosophila. RNA 9: 299-308. 12592004

Schiltz, R. L. and Nakatani, Y. (2000). The PCAF acetylase complex as a potential tumor suppressor. Biochim. Biophys. Acta 1470: M37-M53. 10722926

Smith, E. R., et al. (1998). Cloning of Drosophila GCN5: conserved features among metazoan GCN5 family members. Nucleic Acids Res. 26: 2948-2954. 9611240

Sterner, D. E., et al. (1999). Functional organization of the yeast SAGA complex: distinct components involved in structural integrity, nucleosome acetylation, and TATA-binding protein interaction. Mol. Cell. Biol. 19: 86-98. 9858534

Sterner, D. E., et al. (2002). The SANT domain of Ada2 is required for normal acetylation of histones by the yeast SAGA complex. J. Biol. Chem. 277: 8178-8186. 11777910

Syntichaki, P., Topalidou, I. and Thireos, G. (2000). The Gcn5 bromodomain co-ordinates nucleosome remodelling. Nature 404: 414-417. 10746732

Vlachonasios, K. E., Thomashow, M. F. and Triezenberg. S. J. (2003). Disruption mutations of ADA2b and GCN5 transcriptional adaptor genes dramatically affect Arabidopsis growth, development, and gene expression. Plant Cell 15: 626-638. 12615937

Weake, V. M., et al. (2008). SAGA-mediated H2B deubiquitination controls the development of neuronal connectivity in the Drosophila visual system. EMBO J. 27(2): 394-405. PubMed Citation: 18188155

Weake, V. M., et al. (2009). A novel histone fold domain-containing protein that replaces TAF6 in Drosophila SAGA is required for SAGA-dependent gene expression. Genes Dev. 23: 2818-2823. PubMed Citation: 20008933

Weake, V. M., et al. (2011). Post-transcription initiation function of the ubiquitous SAGA complex in tissue-specific gene activation. Genes Dev. 25(14): 1499-509. PubMed Citation: 21764853

Wen, P., Quan, Z. and Xi, R. (2012). The biological function of the WD40 repeat-containing protein p55/Caf1 in Drosophila. Dev. Dyn. 241(3): 455-64. PubMed Citation: 22241697

Wittschieben, B. O., et al. (2000). Overlapping roles for the histone acetyltransferase activities of SAGA and elongator in vivo. EMBO J. 19: 3060-3068. 10856249

Wu, P. Y., Ruhlmann, C. Winston, F. and Schultz, P. (2004). Molecular architecture of the S. cerevisiae SAGA complex. Mol. Cell 15: 199-208. 15260971

Xu, W., et al. (2000). Loss of Gcn5l2 leads to increased apoptosis and mesodermal defects during mouse development. Nat. Genet. 26: 229-232. 11017084

Zheng, Y., et al. (2005). Fluorescence analysis of a dynamic loop in the PCAF/GCN5 histone acetyltransferase. Biochemistry 44(31): 10501-9. 16060659

Zsindely, N., et al. (2009). The loss of histone H3 lysine 9 acetylation due to dSAGA-specific dAda2b mutation influences the expression of only a small subset of genes. Nucleic Acids Res. 37: 6665-6680. PubMed Citation: 19740772


Gcn5 ortholog: Biological Overview | Evolutionary Homologs | Regulation | Developmental Biology | Effects of Mutation

date revised: 20 March 2012

Home page: The Interactive Fly © 2003 Thomas B. Brody, Ph.D.

The Interactive Fly resides on the
Society for Developmental Biology's Web server.