Histone H4


DEVELOPMENTAL BIOLOGY

Embryonic

Histone H4 methyltransferase activity is cell cycle-regulated, consistent with increased H4 Lys 20 methylation at mitosis. This increase closely follows the cell cycle-regulated expression of the H4 Lys 20 methyltransferase, PR-Set7. Localization of PR-Set7 to mitotic chromosomes and subsequent increase in H4 Lys 20 methylation were inversely correlated to transient H4 Lys 16 acetylation in early S-phase. To expand the observation made in mammalian cells that H4 Lys 20 methylation decreases in S-phase and peaks at mitosis, immunofluorescence studies were performed in Drosophila embryos. A recent report shows that the H4 Lys 20-methyl modification is present in Drosophila and is essential for Drosophila development and viability (Nishioka, 2002). In addition, the embryos provide an excellent model to study H4 Lys 20 methylation during the cell cycle since they rapidly and repeatedly shift from S-phase to mitosis -- this can be easily determined by DAPI staining. Consistent with the above findings, H4 Lys 20 methylation is clearly detected on chromosomes during both metaphase and anaphase, whereas staining during S-phase results in a faint signal even upon overexposure. It is hypothesized that the faint signal during S-phase most likely reflects a combination of dilution of the modification by histone deposition as well as the decrease in chromatin condensation, which could contribute to a dispersion of the signal resulting in a decreased ability to detect the modification. Regardless, these data confirm that H4 Lys 20 methylation is decreased during S-phase and increased specifically during mitosis (Rice, 2002).

Larval

Using indirect immunofluorescence with site-specific antisera, histone H4 isoforms acetylated at lysines 5, 8, 12, or 16 have been shown to have distinct patterns of distribution in interphase, polytene chromosomes from Drosophila larvae. H4 molecules acetylated at lysines 5 or 8 are distributed in overlapping, but nonidentical, islands throughout the euchromatic chromosome arms. beta-Heterochromatin in the chromocenter is depleted in these isoforms, but relatively enriched in H4 acetylated at lysine 12. H4 acetylated at lysine 16 is found at numerous sites along the transcriptionally hyperactive X chromosome in male larvae, but not in male autosomes or any chromosome in female cells. These findings support the hypothesis that H4 molecules acetylated at particular sites mediate unique and specific effects on chromatin function (Turner, 1992).

Dosage compensation in Drosophila occurs by an increase in transcription of genes on the X chromosome in males. This elevated expression requires the function of at least four loci, known collectively as the male-specific lethal (msl) genes. The proteins encoded by two of these genes, maleless (mle) and male-specific lethal-1 (msl-1), (see Sex lethal and MSL-2) are found associated with the X chromosome in males, suggesting that they act as positive regulators of dosage compensation. A specific acetylated isoform of histone H4, H4Ac16, is also detected predominantly on the male X chromosome. MLE and MSL-1 bind to the X chromosome in an identical pattern; the pattern of H4Ac16 on the X is largely coincident with that of MLE/MSL-1. No H4Ac16 is found on the X chromosome in homozygous msl mutant males, correlating with the lack of dosage compensation in these mutants. Conversely, in Sex lethal mutants, H4Ac16 is detected on the female X chromosomes, coincident with an inappropriate increase in X chromosome transcription. These data suggest that synthesis or localization of H4Ac16 is controlled by the dosage compensation regulatory hierarchy. Dosage compensation may involve H4Ac16 function, potentially through interaction with the product of the msl genes (Bone, 1994).

MLE, MSL-1 and histone H4 acetylated at lysine 16 (H4Ac16) are located almost exclusively on the male X chromosome in interphase (polytene) cells. In neuroblasts from third instar Drosophila larva antisera to H4Ac16, MLE and MSL-1 uniquely label the distal, euchromatic region of the male X chromosome through mitosis. The centromere-proximal, heterochromatic region of the male X is not labelled with these antisera, nor are male autosomes or any chromosomes in female cells. That the association of H4Ac16 with the male X chromosomepersists, even when the chromosome is maximally compacted and transcriptionally quiescent, argues that this modified histone is an integral component of the dosage compensation pathway. In the nuclei of interphase neuroblasts from male (but never female) larva, antibodies to H4Ac16 reveal a small, brightly labelled patch against a background of generally weak nuclear staining. In double-labelling experiments, this patch was also labelled, albeit comparatively weakly, with antibodies to MSL-1. These results strongly suggest that the distal, euchromatic region of the X chromosome in male cells occupies a limited and relatively compact nuclear domain (Lavender, 1994).

To gain insights into the function of H4-K20 methylation, the highly specific methyl H4-K20-specific antibodies were used to analyze the distribution of methylated H4-K20 on Drosophila polytene chromosomes and mouse embryonic fibroblasts (MEFs). Methylation of histone H4-K20 on polytene chromosomes coincides with condensed chromosomal regions, including chromocentric heterochromatin and numerous bands on the euchromatic arms. Competition experiments were performed to verify that the observed staining pattern was specific for methyl H4-K20. In these experiments, the staining was completely removed with peptides that contained the H4 tail methylated at lysine 20, but not with unmodified peptides or by peptides that contained the H3 tail methylated at lysine 4 or lysine 9. Moreover, costaining of polytene chromosomes with antibodies raised in different organisms, a rabbit polyclonal and a mouse monoclonal antibody to methyl H4-K20, shows complete overlap (Nishioka, 2002).

Having a monoclonal antibody allowed for a direct comparison of the distribution of methylated H4-K20 with that of other modifications that occur on the H3 and H4 tails, using existing rabbit polyclonal antibodies. Comparison of the distribution of methyl H4-K20 and methyl H3-K9 on polytene chromosomes established that the H4-K20 methylation pattern is distinct from the predominantly chromocentric pattern of methylated H3-K9, a modification that has been associated with constitutive heterochromatin in various species (Nishioka, 2002).

The distribution of the methyl H4-K20 was analyzed with respect to transcriptionally active or competent genes. To accomplish this, the staining pattern on Drosophila polytene chromosomes obtained with the polyclonal antibody to methyl H4-K20 was compared to that observed with monoclonal antibody to the transcription-engaged form of RNA polymerase II. This analysis demonstrated nonoverlapping patterns for each of the antibodies in the entire chromosomes, except at regions that were not fully spread; this led to the conclusion that methylated H4-K20 was very low or absent from transcriptionally competent regions. Similar results were obtained when the staining pattern of the antibody to methyl H4-K20 was compared to that of the transcriptionally active form of RNA polymerase II at heat shock loci under heat shock (transcriptionally permissive) conditions (Nishioka, 2002).

To further analyze the association of the methyl H4-K20 mark with transcriptionally silent chromatin, the methyl H4-K20 staining pattern was compared to that obtained with antibodies specific to methyl H3-K4, a mark that has been correlated with transcriptionally competent genes in higher eukaryotes. Consistent with the results with RNAPII staining, nonoverlapping patterns for methyl H3-K4 and methyl H4-K20 were observed in the entire polytene chromosomes, except at regions that were not fully spread. Thus, it is concluded that the methylated H4-K20 modification marks transcriptionally silent chromatin (Nishioka, 2002).

Oogenesis and Spermatogenesis

There are two distinct phases of histone gene expression during oogenesis. In the first phase, which occurs during early to middle oogenesis (stages 5-10A), there is a mosaic pattern of histone mRNA in the 15 nurse cells of the egg chamber: some cells have very high levels of mRNA, while others have little or no mRNA. This analysis suggests that there is a cyclic accumulation and subsequent degradation of histone mRNA in the egg chamber and that very little histone mRNA is transported into the growing oocyte. Moreover, since the endomitotic replication cycles of the nurse cells are asynchronous during this period, the mosaic distribution of histone message would suggest that the expression of the histone genes in each nurse cell nucleus is probably coupled to DNA replication as in most somatic cells. The second phase begins at stage 10B. During this period, histone gene expression appears to be "induced" in all 15 nurse cells of the egg chamber; instead of a mosaic pattern, high levels of histone mRNA are found in all cells. Unlike the earlier phase, this expression is apparently uncoupled from the endomitotic replication of the nurse cells (which are completed by the end of stage 10A). Moreover, much of the newly synthesized histone mRNA is transported from the nurse cells into the oocyte where it accumulates and is stored for use during early embryogenesis. Tightly clustered grains are observed within nurse cell nuclei in non-denatured tissue sections. As was the case with cytoplasmic histone mRNA, there is a mosaic distribution of nuclear grains from stages 5 to 10A, while at stage 10B, virtually all nurse cell nuclei have grain clusters. These grain clusters appear to be due to the hybridization of nurse cell histone gene DNA to the histone mRNA probe, and are localized in specific regions of the nucleus (Ambrosio, 1985).

Relatively low levels of histone mRNAs are present in egg chambers prior to stage 10, during the period of nurse and follicle cell polyploidization. Histone mRNAs accumulate rapidly and selectively after stage 10, coinciding with the onset of nurse cell degeneration and well after DNA synthesis and actin mRNA accumulation have ceased. A large proportion of the histone mRNAs is associated with polysomes at all times, indicating that expression of histone genes is not strictly coupled to DNA synthesis. Another burst of histone mRNA accumulation near the end of oogenesis may provide a store of maternal histone mRNA to support the rapid cleavages that occur during early embryogenesis (Ruddell, 1985).

A genomic fragment was cloned from a DNA library constructed from a Drosophila enhancer trap line in which reporter gene expression was observed at the anterior-most tip of the ovaries and testes. This genomic clone was identified as the L-repeat of the Drosophila melanogaster histone gene cluster. Northern blotting and in situ hybridization to RNA in tissues with individual cDNAs and PCR-generated probes for each histone confirm that gene expression is greatest at the anterior portion of each ovariole, in the germarium, and is also elevated in a few individual nurse cells and somatic follicle cells within the egg chamber during early developmental stages. Histone H1 and each of the core histones have a similar expression pattern that is correlated to cell division. Maternal stores of histone transcripts are also transported to the mature oocyte from the nurse cells at a later stage of oogenesis (stage 10), when virtually all the nurse cells contain high levels of histone transcripts. The results are consistent with expression of the somatic histone gene cluster during oogenesis as a coordinate unit. There does not seem to be a reduced level of somatic type H1 in the germ-line, as is observed in some other species. The relationship between the P[lacZ] expression pattern in the germarium and the overall expression of the histone cluster suggests there are specific regulatory elements for germ-line expression (Walker, 1998).

maleless (mle) is essential in Drosophila melanogaster males both in somatic cells and in germ cells. In somatic cells mle is necessary for X-chromosome dosage compensation. The role of mle in the germline is unknown. The expression pattern and localization of MLE, the other MSLs and acetylated isoforms of histone H4 in male germ cells have been examined to address whether dosage compensation and/or X inactivation occur in the Drosophila germline. MLE is the only MSL expressed in the male germ cells and is not localized to the X chromosome, nor with any other chromosomal cluster. Weak Mle expression is detectable in spermatogonia and early spermatocytes. Mel is very abundant in the nuclei of fully developed primary spermatocytes; it continues to be detectable in round stage spermatids and elongated spermatides. Mle protein is not detectable during later stages of spermatid development. Studies using a mle temperature sensitive allele reveal that, genetically, the amount of mle activity required for fertility is higher than that required for viability. The analysis of mle mutant gonads reveals that complete loss of mle in the germline and in the soma does not affect the development of male germ cells. Spermatogeneisis can proceed to the final stages of differentiation in the apparent absence of Mle protein. In male germ cells, loss of mle has no detectable effect on the expression or localization of Histone H4 acetylated on amino acid 16. The lack of specific X chromosome localization of H4Ac16 in the transcriptionally active stages of spermatogenesis argues against a role of H4Ac16 in dosage compensation and favors a more general role in transcriptional activation. It is concludedthat in the germline mle is not involved in chromosomal dosage compensation but, in its requirement for male fertility, may be involved inpost-transcriptional gene regulation, perhaps mediated by the Mle helicase function. The acetylation pattern of histone H4 is verydynamic during spermatogenesis. While the pattern is not compatible with dosage compensation or Xinactivation, it is consistent with all premeiotic chromosomes being in an active configuration that isreplaced in post-meiotic stages with an inactive chromatin constitution (Rastelli, 1998).

Transition from a nucleosome-based to a protamine-based chromatin configuration during spermiogenesis in Drosophila

In higher organisms, the chromatin of sperm is organised in a highly condensed protamine-based structure. In pre-meiotic stages and shortly after meiosis, histones carry multiple modifications. This study focused on post-meiotic stages and shows that also after meiosis, histone H3 shows a high overall methylation of K9 and K27; it was hypothesised that these modifications ensure maintenance of transcriptional silencing in the haploid genome. Furthermore, histones are lost during the early canoe stage, and just before this stage, hyper-acetylation of histone H4 and mono-ubiquitylation of histone H2A occurs. It is believed that these histone modifications within the histone-based chromatin architecture may lead to better access of enzymes and chromatin remodellers. This notion is supported by the presence of the architectural protein CTCF, numerous DNA breaks, SUMO, UbcD6 and high content of ubiquitin, as well as testes-specific nuclear proteasomes at this time. Moreover, the first transition protein-like chromosomal protein to be found in Drosophila, Tpl94D, is reported. It is proposed that Tpl94D (an HMG box protein) and the numerous DNA breaks facilitate chromatin unwinding as a prelude to protamine and Mst77F deposition. Finally, it is showm that histone modifications and removal are independent of protamine synthesis (Rathke, 2007).

The switch between a nucleosome-based chromatin configuration and a protamine-based structure is a specialised form of chromatin remodelling in the male germline. The mammalian zinc finger protein CTCF is involved in many epigenetic processes. Furthermore, paralogous variant of CTCF which is testis-specifically expressed, called BORIS, is exclusively expressed in the mammalian male germline. The function of BORIS in this context is still not clear. Drosophila, in contrast to mammals, contains only one CTCF gene. It was therefore asked whether Drosophila CTCF is also expressed in the testes, and immunostaining and anti-histone staining was performed on testes of transgenic flies expressing protamine-eGFP. CTCF expression was observed during pre-meiotic and meiotic stages at the chromosomes as has been shown for mitotic cell division in mammalian cell culture. Shortly after meiosis, CTCF is visible in young elongating nuclei, where it co-localises with the chromatin as indicated by the histone distribution. CTCF is also present in the early and late canoe stage spermatid heads. At the early canoe stage, CTCF is very diffusely distributed in comparison to histones. CTCF does not co-localise with the chromatin which starts to condense at one side of the nucleus. This diffuse distribution is still visible at the late canoe stage when protamine-eGFP starts to be deposited to the chromatin. CTCF is no longer detectable after the canoe stage. The earlier chromatin-associated CTCF localisation might indicate a very early role in chromatin reorganisation at the switch between the nebenkern and canoe stage. Furthermore, CTCF might be associated primarily with the chromatin, which is not yet condensing during these stages. The late canoe stage is the only post-meiotic stage where distinct regions of RNA polymerase II are found with an antibody directed against a phosphorylated subunit of active polymerase, indicative of transcription. At this precise stage, only a very small set of genes is thought to be transcribed. Also CTCF expression during chromatin reorganisation in the nucleus was detected in D. hydei (Rathke, 2007).

Sperm morphogenesis is characterised by an impressive degree of changes in cell architecture based on stored, translationally repressed mRNAs that are recruited at the appropriate time to the polysomes. Among these are mRNAs that encode Tpl94D and protamines. A dramatic switch in structure from the nucleosomal- to the protamine-based structure of chromatin takes place, and this remarkable chromatin reorganisation of the complete genome is a typical feature depending on stored mRNAs, e.g. for protamine synthesis. This process ultimately leads to an extremely condensed state of the haploid genome in the sperm, which is essential for male fertility in mammals. This study focused on the switch between a nucleosomal- and a protamine-based chromatin reorganisation. The major steps in chromatin organisation take place in the canoe stage of spermatid development. A candidate for a transition protein in Drosophila was identified. The corresponding gene tpl94D (CG31281) encodes a predicted basic high mobility group (HMG) protein of 18.8 kDa. In transgenic flies, Tpl94D-eGFP fusion proteins are expressed solely during the switch between histones and protamines, as is typical for mammalian transition proteins. Since a highly similar chain of events to those reported in mammals is observed, the Drosophila system is considered an excellent choice to study the mechanism of chromatin remodelling during male germ cell development (Rathke, 2007).

Generally, the bulk of histones, including their diverse modifications in the N-terminal tail, appear to be removed during the canoe stage. Furthermore, the nucleus accumulates ubiquitin at the early canoe stage, when mono-ubiquitylation of histone H2A is no longer detectable. Therefore, taking into account the known presence of proteasomes in the nucleus at this stage of chromatin reorganisation and the overlap of expression shown in this study, it is hypothesised that this ubiquitylation is targeting histones for degradation. This study investigated several mutants having mutations in ubiquitin-conjugating enzymes or ubiquitin ligases, exhibiting arrested spermiogenesis during spermatid development and that are male sterile. However, in all investigated mutants, histone removal is indistinguishable from that of wild-type flies (Rathke, 2007).

Many histone modifications were found after meiosis and were categorised into three classes (Rathke, 2007).

  1. Histone modifications that persist from pre-meiotic stages and keep the genome silent.

    The vast majority of the genome is transcriptionally silent in post-meiotic stages. This is accompanied by multiple histone modifications that persist from pre-meiotic stages and indicate silencing such as H3K9 and H3K27 methylation. These modifications do not change significantly during post-meiotic stages, which is in agreement with the hypothesis that these modifications predominantly play a role in maintaining transcriptional silencing. Previously, phosphorylation of histones have been analysed during spermatogenesis. Phosphorylated histone H4S1 and H3S10 are present during meiotic divisions. H3S10 phosphorylation is hardly detectable after meiosis, whereas phosphorylation of H4S1 persists until chromatin compaction starts.

  2. Histone modifications that persist from pre-meiotic stages and characterise transcriptionally active chromatin.

    The primary spermatocyte phase is characterised by a high level of transcriptional activity of housekeeping genes. In addition, genes are transcribed that are needed for the subsequent steps in spermatogenesis, as the majority of transcription ceases once meiotic division starts. H4 acetylation and H3K4 and H4R3 methylation of histones were investigated. These histone modifications, which are indicative of transcriptional activity, persist until histone degradation.

  3. Increasing or de novo appearance of histone modifications that decrease the affinity between histones and DNA as a prelude to histone removal.

    It might be that H4 hyper-acetylation, as postulated for mammals and/or other secondary modifications of histones are the first step towards histone removal. The fact that these modifications are conserved between mammals and flies adds support to this hypothesis. Indeed, histone H4 acetylation is very pronounced at the canoe stage and de novo mono-ubiquitylation of histone H2A is seen in round spermatids. Both types of histone modifications are proposed to be necessary for opening the chromatin and decreasing the contact between DNA and histones. The fact that histone H2A mono-ubiquitylation vanishes before the early canoe stage, thus before the hyper-acetylation of histone H4, leads to thinking about a stepwise remodelling of the chromatin. This study proposes that these histone modifications open the chromatin, so that enzymes and regulators have access to histone-based chromatin and can induce and prepare the reorganisation of the genome in the male germline.

It remains to be clarified whether and how these histone modifications influence the topology of the chromatin as a prelude to histone removal as well as for Tpl94D, Mst77F and protamine deposition. A functional approach based on analysis of mutants of histone-modifying enzymes is difficult, as all characterised histone-modifying enzymes are already active during Drosophila development or at least in spermatogonia and spermatocytes. Therefore a tissue-specific knock-out mutant would most probably exhibit arrest of spermatogenesis before meiosis, rendering it useless for experimental purposes (Rathke, 2007).

At the first glance, it might seem surprising that histones and all their modifications are removed. Instead of specifically reverting the differentially modified histones to their unmodified state, they are removed together with all histones. This might allow the paternal genome to form nucleosomes with unmodified histones after fertilisation and before zygote formation. Thus, the paternal genome starts embryogenesis with a nucleosomal chromatin lacking histone modifications (Rathke, 2007).

The data show that most of the histones are removed between the early and late canoe stage; such a process requires a loosening of contact between the histones and DNA, which in turn requires an unwinding of the chromatin structure. It is proposed that this unwinding process is facilitated by DNA nicks as they were widespread at this stage of chromatin reorganisation. Finally, Tpl94D, UbcD6 and SUMO were also observed to accumulate in the chromatin during this process. DNA breaks, Tpl94D, UbcD6 and SUMO were no longer detectable when protamines were fully expressed. Thus, it is proposed that all these proteins and the DNA breaks act together in an unknown manner to allow chromatin remodelling (Rathke, 2007).

The CTCF protein is present during pre-meiotic stages in the nucleus and stays associated with the chromosomes during meiosis. After meiosis, however, strong localisation to the nucleus is detected during the transition from round spermatid nuclei to the early canoe stage of spermiogenesis. It is speculated that CTCF might set borders in the chromatin for the histone modifications, which are characteristic of the canoe stage, such as acetylation and ubiquitylation. CTCF is visible for longer than histones and disappears together with active RNA polymerase II. CTCF might maintain chromatin accessibility to RNA polymerase II since a few genes are known to be transcribed at this time. In addition, transient occurrence of RNA polymerase II at the late canoe stage might require CTCF to insulate active genes from inactive ones. This idea needs to be tested in tissue-specific CTCF loss-of-function mutants; such mutants are, however, currently unavailable (Rathke, 2007).

The question of whether histone removal is dependent on a signal that monitors the start of protamine and Mst77F mRNA translation was addressed. Both histone modification and degradation are indistinguishable from the wild-type in loss-of-function mutants of Mst35Ba and Mst35Bb, the genes encoding protamine A and B, respectively. Also in nc3 mutants of Mst77F, histone removal is not disturbed. It is concluded that N-terminal tail modification of histones and histone degradation, on the one hand, and protamine deposition, on the other, are controlled by different pathways in the cell (Rathke, 2007).

In mammals, it is well known that after meiosis the nucleosomal conformation is lost. This is accompanied by the appearance of testis-specific linker histones. So far, no linker histone variants have been identified in Drosophila, but variants of H2A (H2AvD) and H3 (H3.3) are known. In mammals, histones are hyper-acetylated before being displaced from the DNA, and phosphorylation and ubiquitylation have also been proposed to occur. For Drosophila, H2A mono-ubiquitylation and a strong increase in H4 acetylation occur shortly before histone removal and degradation. In mammals, histones are replaced first by transition proteins (major types: TP1 and TP2). This study identified the high mobility group protein Tpl94D, a first probable candidate for a functional homologue of mammalian transition proteins. In mammals, transition proteins are subsequently replaced by protamines leading to chromatin with a doughnut structure. In Drosophila, it has recently been shown that the sperm nucleus also contains protamines. Protamines A and B are encoded by two closely related protamine genes, Mst35Ba and Mst35Bb. In addition, the identification of Mst77F shows that sperm nuclei contain at least one further abundant chromatin component. Moreover, in human sperm several new putative protamines have been identified by 2D gel electrophoresis and protein sequencing. In mammals, this chromatin reorganisation is essential for male fertility. Male flies carrying the deletion protDelta38.1, where both protamines as well as three additional ORFs are removed, show severely reduced fertility (Rathke, 2007).

In summary, a step-by-step scheme is proposed for chromatin reorganisation: (1) histone modifications lead to subsequent histone removal and degradation; (2) the exposed chromatin becomes nicked, resulting in DNA breaks; (3) Tpl94D deposition constitutes an intermediate stage that triggers subsequent protamine-based chromatin organisation (Rathke, 2007).

Since many features concerning spermiogenesis are conserved between Drosophila and mammals, it is proposed that Drosophila is an ideal system to gain further insight into the mechanism of chromatin reorganisation in spermatid nuclei, a process that is crucial for male fertility (Rathke, 2007).


REFERENCES

Adenot, P. G., et al. (1997). Differential H4 acetylation of paternal and maternal chromatin precedes DNA replication and differential transcriptional activity in pronuclei of 1-cell mouse embryos. Development 124(22): 4615-4625. PubMed Citation: 9409678

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

Agarwal, S., and Behe, M. J. (1996). Non-conservative mutations are well tolerated in the globular region of yeast histone H4. Cell. Mol. Biol. 255: 401-411. PubMed Citation: 8568885

Aggarwal, B. D. and Calvi, B. R. (2004). Chromatin regulates origin activity in Drosophila follicle cells. Nature 430(6997): 372-6. 15254542

Akhmanova, A., Miedema, K. and Hennig, W. (1996). Identification and characterization of the Drosophila histone H4 replacement gene. FEBS Lett. 388: 219-222. PubMed Citation: 8690091

Akhtar, A. and Becker, P. B., 2000). Activation of transcription through histone H4 acetylation by Mof, an acetyltransferase essential for dosage compensation in Drosophila. Molec. Cell 5: 367-375. PubMed Citation: 10882077

Alberts, A. S., Geneste, O. and Treisman, R. (1998). Activation of SRF-regulated chromosomal templates by Rho-family GTPases requires a signal that also induces H4 hyperacetylation. Cell 92: 475-487. PubMed Citation: 9491889

Alland, L., et al. (1997). Role for N-CoR and histone deacetylase in Sin3-mediated transcriptional repression. Nature 387 (6628): 49-55. PubMed Citation: 9139821

Ambrosio, L. and Schedl, P. (1985). Two discrete modes of histone gene expression during oogenesis in Drosophila melanogaster. Dev. Biol. 111: 220-31. PubMed Citation: 3928419

An, W., et al. (2002). Selective requirements for histone H3 and H4 N termini in p300-dependent transcriptional activation from chromatin. Molec. Cell 9: 811-821. 11983172

Arents, G., and Moudrianakis, E. N. (1995). The histone fold: a ubiquitous architectural motif utilized in DNA compaction and protein dimerization. Proc. Nat. Acad. Sci. 92: 11170-74. PubMed Citation: 7479959

Badenhorst, P., Voas, M., Rebay, I. and Wu, C. (2002). Biological functions of the ISWI chromatin remodeling complex NURF. Genes Dev. 16: 3186-3198. 12502740

Baxevanis, A. D. and Landsman, D. (1996). Histone Sequence Database: a compilation of highly-conserved nucleoprotein sequences. Nucleic Acids Res. 24: 245-247. PubMed Citation: 8594591

Beisel, C., Imhof, A., Greene, J., Kremmer, E. and Sauer, F. (2002). Histone methylation by the Drosophila epigenetic transcriptional regulator Ash1. Nature 419(6909): 857-62. 12397363

Beisel, C., et al. (2002). Histone methylation by the Drosophila epigenetic transcriptional regulator Ash1. Nature 419: 857-862. 12397363

Belyaev, N. D., Keohane, A. M. and Turner, B. M. (1996). Histone H4 acetylation and replication timing in Chinese hamster chromosomes. Exp. Cell Res. 225: 277-285. PubMed Citation: 8660915

Berloco, M., et al. (2001). The maternal effect gene, abnormal oocyte (abo), of Drosophila melanogaster encodes a specific negative regulator of histones. Proc. Natl. Acad. Sci. 98: 12126-12131. 11593026

Bhadra, U., Pal-Bhadra, M. and Birchler, J. A. (1999). Role of the male specific lethal (msl) genes in modifying the effects of sex chromosomal dosage in Drosophila. Genetics 152(1): 249-268. PubMed Citation: 10224258

Bird, A. W., et al. (2002). Acetylation of histone H4 by Esa1 is required for DNA double-strand break repair. Nature 419: 411-415. 12353039

Bone, J. R., et al. (1994). Acetylated histone H4 on the male X chromosome is associated with dosage compensation in Drosophila. Genes Dev. 8: 96-104. PubMed Citation: 8288132

Brown, E. J. and Bachtrog, D. (2014). The chromatin landscape of Drosophila: comparisons between species, sexes, and chromosomes. Genome Res 24(7):1125-37. PubMed ID: 24840603

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

Braunstein, M., et al. (1996). Efficient transcriptional silencing in Saccharomyces cerevisiae requires a heterochromatin histone acetylation pattern. Mol. Cell. Biol. 16(8): 4349-56. PubMed Citation: 8754835

Cavalli, G. and Paro, R. (1999). Epigenetic inheritance of active chromatin after removal of the main transactivator. Science 286(5441): 955-8. PubMed Citation: 10542150

Chadwick, B. P. and Willard, H. F. (2004). Multiple spatially distinct types of facultative heterochromatin on the human inactive X chromosome. Proc. Natl. Acad. Sci. 101(50): 17450-5. 15574503

Chen, Z. H., Zhu, M., Yang, J., Liang, H., He, J., He, S., Wang, P., Kang, X., McNutt, M. A., Yin, Y. and Shen, W. H. (2014). PTEN interacts with histone H1 and controls chromatin condensation. Cell Rep 8: 2003-2014. PubMed ID: 25199838

Choi, B. I., et al. (1996). Isolation and characterization of a cDNA encoding a novel human transcription factor TFIID subunit containing similarities with histones H2B and H3. Gene 169: 263-267. PubMed Citation: 8647459

Chou, R. H., et al. (2014). EGFR modulates DNA synthesis and repair through Tyr phosphorylation of histone H4. Dev Cell 30: 224-237. PubMed ID: 25073158

Cirillo, L. A., et al. (2002). Opening of compacted chromatin by early developmental transcription factors HNF3 (FoxA) and GATA-4. Molec. Cell 9: 279-289. 11864602

Clapier, C. R., et al. (2001). Critical role for the histone H4 N terminus in nucleosome remodeling by ISWI. Mol. Cell. Biol. 21(3): 875-83. 11154274

Collas, P., et al. (1999). Active transgenes in zebrafish are enriched in acetylated histone H4 and dynamically associate with RNA pol II and splicing complexes. J. Cell Sci. 112 (Pt 7): 1045-54. PubMed Citation: 10198286

Copur, O., Gorchakov, A., Finkl, K., Kuroda, M. I. and Muller, J. (2018). Sex-specific phenotypes of histone H4 point mutants establish dosage compensation as the critical function of H4K16 acetylation in Drosophila. Proc Natl Acad Sci U S A. PubMed ID: 30530664

Corona, D. F., et al. (2002). Modulation of ISWI function by site-specific histone acetylation. EMBO Rep. 3(3): 242-7. 11882543

Couture, J.-F., et, al. (2005). Structural and functional analysis of SET8, a histone H4 Lys-20 methyltransferase. Genes Dev. 19: 1455-1465. 15933070

Dalal, Y., Wang, H., Lindsay, S. and Henikoff, S. (2007). Tetrameric structure of centromeric nucleosomes in interphase Drosophila cells. PLoS Biol. 5(8): e218. Medline abstract: 17676993

Dimitrov, S., et al. (1993). Chromatin transitions during early Xenopus embryogenesis: changes in histone H4 acetylation and in linker histone type. Dev. Biol. 160: 214-27. PubMed Citation: 8224538

Dominski, Z., Yang, X.C., Purdy, M. and Marzluff, W. F. (2003). Cloning and characterization of the Drosophila U7 small nuclear RNA. Proc. Natl. Acad. Sci. 100(16): 9422-7. 12872004

Dominski, Z., Yang, X. C., Purdy, M. and Marzluff, W. F. (2005). Differences and similarities between Drosophila and mammalian 3' end processing of histone pre-mRNAs. RNA 11(12): 1835-47. 16251385

Dou, Y., et al. (2005). Physical association and coordinate function of the H3 K4 methyltransferase MLL1 and the H4 K16 acetyltransferase MOF. Cell 121: 873-885. 15960975

Du, H. N., Fingerman, I. M. and Briggs, S. D. (2008). Histone H3 K36 methylation is mediated by a trans-histone methylation pathway involving an interaction between Set2 and histone H4. Genes Dev. 22(20): 2786-98. PubMed Citation: 18923077

Eberharter, A., et al. (2001). Acf1, the largest subunit of CHRAC, regulates ISWI-induced nucleosome remodelling. EMBO J. 20: 3781-3788. 11447119

Edmondson, D. G., Smith, M. M. and Roth, S. Y. (1996). Repression domain of the yeast global repressor Tup1 interacts directly with histones H3 and H4. Genes Dev. 10: 1247-1259. PubMed Citation: 8675011

English, C. M., et al. (2006). Structural basis for the histone chaperone activity of Asf1. Cell 127(3): 495-508. Medline abstract: 17081973

Fanti, L., Berloco, M. and Pimpinelli, S. (1994). Carnitine suppression of position-effect variegation in Drosophila melanogaster. Mol. Gen. Genet. 244: 588-595. PubMed Citation: 7969027

Frank, S. R., et al. (2001). Binding of c-Myc to chromatin mediates mitogen-induced acetylation of histone H4 and gene activation. Genes Dev. 15: 2069-2082. 11511539

Furuyama, T., Dalal, Y. and Henikoff, S. (2006a). Chaperone-mediated assembly of centromeric chromatin in vitro. Proc. Natl. Acad. Sci. 103(16): 6172-7. Medline abstract: 16601098

Furuyama, T. and Henikoff, S. (2006b). Biotin-tag affinity purification of a centromeric nucleosome assembly complex. Cell Cycle 5(12): 1269-74. Medline abstract: 16775420

Georgel, P. T., Tsukiyama, T. and Wu, C. (1997). Role of histone tails in nucleosome remodeling by Drosophila NURF. EMBO J. 16: 4717-4726. 9303316

Godfrey, A. C., et al. (2006). U7 snRNA mutations in Drosophila block histone pre-mRNA processing and disrupt oogenesis. RNA 12: 396-409. 16495235

Govin, J., et al. (2010). Systematic screen reveals new functional dynamics of histones H3 and H4 during gametogenesis. Genes Dev. 24(16): 1772-86. PubMed Citation: 20713519

Guertin, M. J., Martins, A. L., Siepel, A. and Lis, J. T. (2012). Accurate prediction of inducible transcription factor binding intensities in vivo. PLoS Genet 8: e1002610. Pubmed: 22479205

Guglielmi, B., La Rochelle, N. and Tjian, R. (2013). Gene-specific transcriptional mechanisms at the histone gene cluster revealed by single-cell imaging. Mol Cell 51: 480-492. PubMed ID: 23973376

Guo, B., et al. (1997). ATF1 and CREB trans-activate a cell cycle regulated histone H4 gene at a distal nuclear matrix associated promoter element. Biochemistry 36(47): 14447-14455. PubMed ID: 9398163

Gu, W., Szauter, P. and Lucchesi, J. C. (1998). Targeting of MOF, a putative histone acetyl transferase, to the X chromosome of Drosophila melanogaster. Dev. Genet. 22(1): 56-64. 9816055

Gupta, S., et al. (2003). Tumor suppressor pRB functions as a co-repressor of the CCAAT displacement protein (CDP/cut) to regulate cell cycle controlled histone H4 transcription. J. Cell Physiol. 196(3): 541-56. 12891711

Hamiche, A., Kang, J.-G., Dennis, C., Xiao, H. and Wu, C. (2001). Histone tails modulate nucleosome mobility and regulate ATP-dependent nucleosome sliding by NURF. Proc. Natl. Acad. Sci. 98: 14316-14321. 11724935

Hassig, C. A., et al. (1997). Histone deacetylase activity is required for full transcriptional repression by mSin3A. Cell 89 (3): 341-347

Hilfiker, A., et al. (1997). mof, a putative acetyl transferase gene related to the Tip60 and MOZ human genes and to the SAS genes of yeast, is required for dosage compensation in Drosophila. EMBO J. 16(8): 2054-2060.

Hiratani, I., et al. (2008). Global reorganization of replication domains during embryonic stem cell differentiation. PLoS Biol. 6: e245. PubMed Citation: 18842067

Hori, T., Shang, W. H., Toyoda, A., Misu, S., Monma, N., Ikeo, K., Molina, O., Vargiu, G., Fujiyama, A., Kimura, H., Earnshaw, W. C. and Fukagawa, T. (2014). Histone H4 Lys 20 monomethylation of the CENP-A nucleosome is essential for kinetochore assembly. Dev Cell 29: 740-749. PubMed ID: 24960696

Hu, H., et al. (2011). Structure of a CENP-A-histone H4 heterodimer in complex with chaperone HJURP. Genes Dev. 25(9): 901-6. PubMed Citation: 21478274

Huang, L., Zhang, W. and Roth, S. Y. (1997). Amino termini of histones H3 and H4 are required for a1-alpha2 repression in yeast. Mol. Cell. Biol. 17(11): 6555-6562

Huang, S., Litt, M. and Felsenfeld, G. (2005). Methylation of histone H4 by arginine methyltransferase PRMT1 is essential in vivo for many subsequent histone modifications. Genes Dev. 19: 1885-1893. 16103216

Hug, C. B., Grimaldi, A. G., Kruse, K. and Vaquerizas, J. M. (2017). Chromatin architecture emerges during zygotic genome activation independent of transcription. Cell 169(2): 216-228 e219. PubMed ID: 28388407

Ito, S., Fujiyama-Nakamura, S., Kimura, S., Lim, J., Kamoshida, Y., Shiozaki-Sato, Y., Sawatsubashi, S., Suzuki, E., Tanabe, M., Ueda, T., Murata, T., Kato, H., Ohtake, F., Fujiki, R., Miki, T., Kouzmenko, A., Takeyama, K., Kato, S. (2012). Epigenetic Silencing of Core Histone Genes by HERS in Drosophila. Mol. Cell 45(4): 494-504. PubMed Citation: 22365829

Kanno, T., et al. (2004). Selective recognition of acetylated histones by bromodomain proteins visualized in living cells. Mol. Cell 13(1): 33-43. Medline abstract: 14731392

Karachentsev, D., Sarma, K., Reinberg, D. and Steward, R. (2005). PR-Set7-dependent methylation of histone H4 Lys 20 functions in repression of gene expression and is essential for mitosis. Genes Dev. 19(4): 431-5. 15681608

Kida, S., Umehara, T. and Horikoshi, M. (1995). Three distinct regions in a rat TFIID subunit containing histone H4 homology. Gene 161: 303-304

Keener, J., et al. (1997). Histones H3 and H4 are components of upstream activation factor required for the high-level transcription of yeast rDNA by RNA polymerase I. Proc. Natl. Acad. Sci. 94(25): 13458-13462

Keohane, A., et al. (1996). X-inactivation and histone H4 acetylation in embryonic stem cells. Dev. Biol. 180: 618-30

Kind, J. and Akhtar, A. (2007). Cotranscriptional recruitment of the dosage compensation complex to X-linked target genes. Genes Dev. 21: 2030-2040. PubMed Citation: 17699750

Kind, J., et al. (2008) Genome-wide analysis reveals MOF as a key regulator of dosage compensation and gene expression in Drosophila. Cell 133: 813-828. PubMed Citation: 18510926

Klymenko, T., Papp, B., Fischle, W., Köcher, T., Schelder, M., Fritsch, C., Wild, B., Wilm, M. and Müller, J. (2006). A Polycomb group protein complex with sequence-specific DNA-binding and selective methyl-lysine-binding activities. Genes Dev. 20: 1110-1122. Medline abstract: 16618800

Kotake, Y., et al. (2007). pRB family proteins are required for H3K27 trimethylation and Polycomb repression complexes binding to and silencing p16INK4a tumor suppressor gene. Genes Dev. 21: 49-54. Medline abstract: 17210787

Krishnamoorthy, T., et al. (2006). Phosphorylation of histone H4 Ser1 regulates sporulation in yeast and is conserved in fly and mouse spermatogenesis. Genes Dev. 20(18): 2580-92. Medline abstract: 16980586

Kruger, W., et al. (1995). Amino acid substitutions in the structured domains of histones H3 and H4 partially relieve the requirement of the yeast SWI/SNF complex for transcription. Genes Dev. 9: 2770-2779

Larschan, E., et al. (2011). X chromosome dosage compensation via enhanced transcriptional elongation in Drosophila. Nature 471(7336): 115-8. PubMed Citation: 21368835

Lavender, L. S., et a. (1994). Histone H4 acetylated at lysine 16 and proteins of the Drosophila dosage compensation pathway co-localize on the male X chromosome through mitosis. Chromosome Resc 2: 398-404

Lee, D. Y., et al. (1993). A positive role for histone acetylation in transcription factor access to nucleosomal DNA. Cell 72: 73-84

Lee, M. C., Toh, L. L., Yaw, L. P. and Luo, Y. (2010). Drosophila octamer elements and Pdm-1 dictate the coordinated transcription of core histone genes. J. Biol. Chem. 285(12): 9041-53. PubMed Citation: 20097756

Lifton, R. P., et al. (1977). The organization of histone genes in Drosophila melanogaster: Functional and evolutionary implications. Cold Spring Harbor Symposia on Quantitative Biology XLII: 1047-51

Ling, X., et al. (1996). Yeast histone H3 and H4 amino termini are important for nucleosome assembly in vivo and in vitro: redundant and position-independent functions in assembly but not in gene regulation. Genes Dev. 10: 686-699

Loyola, A., et al. (2001). Reconstitution of recombinant chromatin establishes a requirement for histone-tail modifications during chromatin assembly and transcription. Genes Dev. 15: 2837-2851. 11691835

Lucas, I., et al. (2007). High-throughput mapping of origins of replication in human cells. EMBO Rep. 8: 770-777. PubMed Citation: 17668008

Lv, W. W., Wei, H. M., Wang, D. L., Ni, J. Q. and Sun, F. L. (2012). Depletion of histone deacetylase 3 antagonizes PI3K-mediated overgrowth through the acetylation of histone H4 at lysine 16. J. Cell Sci. [Epub ahead of print]. PubMed Citation: 22956542

Ma, X. J., Lu, Q. and Grunstein, M. (1996). A search for proteins that interact genetically with histone H3 and H4 amino termini uncovers novel regulators of the Swe1 kinase in Saccharomyces cerevisiae. Genes Dev. 10: 1327-1340

Ma, X. J., et al. (1998). Deposition-related sites K5/K12 in histone H4 are not required for nucleosome deposition in yeast. Proc. Natl. Acad. Sci. 95(12): 6693-6698

Matangkasombut, O. and Buratowski, S. (2003). Different sensitivities of Bromodomain factors 1 and 2 to Histone H4 acetylation. Mol. Cell 11: 353-363. 12620224

Miotto, B., Sagnier, T., Berenger, H., Bohmann, D., Pradel, J. and Graba, Y. (2005). Chameau HAT and DRpd3 HDAC function as antagonistic cofactors of JNK/AP-1-dependent transcription during Drosophila metamorphosis. Genes Dev. 20(1): 101-12. 16391236

Mizzen, C. A., et al. (1996). The TAFII250 subunit of TFIID has histone acetyltransferase activity. Cell 87: 1261-70. Medline abstract: 8980232

Mizuguchi, G., et al. (2007). Nonhistone Scm3 and histones CenH3-H4 assemble the core of centromere-specific nucleosomes. Cell 129: 1153-1164. Medline abstract: 17574026

Nagy, L., et al. (1997). Nuclear receptor repression mediated by a complex containing SMRT, mSin3A, and histone deacetylase. Cell 89 (3): 373-380

Nakatani, Y., Bagby, S. and Ikura, M. (1996). The histone folds in transcription factor TFIID. J. Biol. Chem. 271: 6575-6578

Ng, K. W., et al. (1997). The binding of a Fos/Jun heterodimer can completely disrupt the structure of a nucleosome. EMBO J. 16: 2072-85

Nightingale, K. P., et al. (1998). Histone acetylation facilitates RNA polymerase II transcription of the Drosophila hsp26 gene in chromatin. EMBO J. 17: 2865-2876

Ogryzko, V. V., et al. (1996). The transcriptional coactivators p300 and CBP are histone acetyltransferases. Cell 87: 953-959

Nishioka, K., Rice, J. C., Sarma, K., Erdjument-Bromage, H., Werner, J., Wang, Y., Chuikov, S., Valenzuela, P., Tempst, P. and Steward, R., et al. (2002). PR-Set7 is a nucleosome-specific methyltransferase that modifies lysine 20 of histone H4 and is associated with silent chromatin. Mol. Cell 9: 1201-1213. 12086618

Nowak, S. J. and Corces, V. G. (2000). Phosphorylation of histone H3 correlates with transcriptionally active loci. Genes Dev. 14: 3003-3013. 11114889

O'Neill, L. P., et al. (1999). A developmental switch in H4 acetylation upstream of Xist plays a role in X chromosome inactivation. EMBO J. 18(10): 2897-2907

Papp, B. and Muller, J. (2006). Histone trimethylation and the maintenance of transcriptional ON and OFF states by trxG and PcG proteins. Genes Dev. 20(15): 2041-54. Medline abstract: 16882982

Parthun, M. R., Widom, J. and Gottschling, D. E. (1996). the major cytoplasmic histone acetyltransferase in yeast: Links to chromatin replication and histone metabolism. Cell 87: 85-94

Pinnola, A., Naumova, N., Shah, M. and Tulin, A. V. (2007). Nucleosomal core histones mediate dynamic regulation of Poly(ADP-ribose) polymerase 1 protein binding to chromatin and induction of its enzymatic activity. J. Biol. Chem. 282(44): 32511-32519. PubMed Citation: 17827147

Prieto-Sanchez, S., Moreno-Castro, C., Hernandez-Munain, C. and Sune, C. (2020). Drosophila Prp40 localizes to the histone locus body and regulates gene transcription and development. J Cell Sci. PubMed ID: 32094262

Rastelli, L. and Kuroda, M. I. (1998). An analysis of maleless and histone H4 acetylation in Drosophila melanogaster spermatogenesis. Mech. Dev. 71(1-2): 107-117

Rathke, C., et al. (2007). Transition from a nucleosome-based to a protamine-based chromatin configuration during spermiogenesis in Drosophila. J. Cell Sci. 120(Pt 9): 1689-700. Medline abstract: 17452629

Reid, J. L., Moqtaderi, Z.. and Struhl, K. (2004). Eaf3 regulates the global pattern of histone acetylation in Saccharomyces cerevisiae. Mol. Cell. Biol. 24(2): 757-64. 14701747

Rice, J. C., et al. (2002). Mitotic-specific methylation of histone H4 Lys 20 follows increased PR-Set7 expression and its localization to mitotic chromosomes. Genes Dev. 16: 2225-2230. 12208845

Roth, S. Y. and Allis, C. D. (1996). Histone acetylation and Chromatin assembly: A single escort, multiple dances. Cell 87: 5-8. PubMed ID: 8858142

Ruddell, A., and Jacobs-Lorena, M. (1985). Biphasic pattern of histone gene expression during Drosophila oogenesis. Proc. Natl. Acad. Sci. 82: 3316-9. PubMed ID: 3923476

Sakaguchi, A., Joyce, E., Aoki, T., Schedl, P. and Steward, R. (2012). The histone H4 lysine 20 monomethyl mark, set by PR-Set7 and stabilized by L(3)mbt, is necessary for proper interphase chromatin organization. PLoS One 7: e45321. PubMed ID:23024815

Sahashi, R., Crevel, G., Pasko, J., Suyari, O., Nagai, R., Saura, M. M., Yamaguchi, M. and Cotterill, S. (2014). DNA polymerase alpha interacts with PrSet7 and mediates H4K20 monomethylation in Drosophila. J Cell Sci 127(Pt 14):3066-78. PubMed ID: 24806961

Samata, M., Alexiadis, A., Richard, G., Georgiev, P., Nuebler, J., Kulkarni, T., Renschler, G., Basilicata, M. F., Zenk, F. L., Shvedunova, M., Semplicio, G., Mirny, L., Iovino, N. and Akhtar, A. (2020). Intergenerationally maintained histone H4 lysine 16 acetylation is instructive for future gene activation. Cell. PubMed ID: 32502394

Sanders, S. L., et al. (2004). Methylation of histone H4 lysine 20 controls recruitment of Crb2 to sites of DNA damage. Cell 119: 603-614. 15550243

Santisteban, M. S., et al. (1997). Histone octamer function in vivo: mutations in the dimer-tetramer interfaces disrupt both gene activation and repression. EMBO J 16 (9): 2493-2506

Schotta, G., et al. (2004). A silencing pathway to induce H3-K9 and H4-K20 trimethylation at constitutive heterochromatin. Genes Dev. 18: 1251-1262. 15145825

Schotta, G., et al. (2008). A chromatin-wide transition to H4K20 monomethylation impairs genome integrity and programmed DNA rearrangements in the mouse. Genes Dev. 22: 2048-2061. PubMed Citation: 18676810

Schübeler, D., et al. (2004). The histone modification pattern of active genes revealed through genome-wide chromatin analysis of a higher eukaryote. Genes Dev. 18: 1263-1271. 15175259

Schwaiger, M., et al. (2009). Chromatin state marks cell-type- and gender-specific replication of the Drosophila genome. Genes Dev. 23(5): 589-601. PubMed Citation: 19270159

Shogren-Knaak, M., Ishii, H., Sun, J. M, Pazin, M. J., Davie, J. R. and Peterson, C. L. (2006). Histone H4-K16 acetylation controls chromatin structure and protein interactions. Science 311(5762): 844-7. 16469925

Smith, E. R., et al. (2000). The Drosophila MSL complex acetylates histone H4 at lysine 16, a chromatin modification linked to dosage compensation. Molec. Cell. Biol. 20: 312-318

Smith, M. M., et al. (1996). A novel histone H4 mutant defective in nuclear division and mitotic chromosome transmission. Mol. Cell Biol. 16: 1017-1026

Smith, E. R., Cayrou, C., Huang, R., Lane, W. S., Cote, J. and Lucchesi, J. C. (2005). A human protein complex homologous to the Drosophila MSL complex is responsible for the majority of histone H4 acetylation at lysine 16. Mol. Cell. Biol. 25(21): 9175-88. 16227571

Sobel, R. E. Cook, R. G. and Allis, C. D. (1994). Non-random acetylation of histone H4 by a cytoplasmic histone acetyltransferase as determined by novel methodology. J. Biol. Chem. 269: 18576-82

Sobel, R. E., et al. (1995). Conservation of deposition-related acetylation sites in newly synthesized histone H3 and H4. Proc. Natl. Acad. Sci 92: 1237-41

Song, J. J., Garlick, J. D. and Kingston, R. E. (2008). Structural basis of histone H4 recognition by p55. Genes Dev. 22: 1313-1318. PubMed Citation: 18443147

Spierer, A., Seum, C., Delattre, M. and Spierer, P. (2005). Loss of the modifiers of variegation Su(var)3-7 or HP1 impacts male X polytene chromosome morphology and dosage compensation. J. Cell Sci. 118: 5047-5057. PubMed Citation: 16234327

Spierer, A., Begeot, F., Spierer, P. and Delattre, M. (2008). SU(VAR)3-7 links heterochromatin and dosage compensation in Drosophila. PLoS Genet. 4(5): e1000066. PubMed Citation: 18451980

Spencer, T. E., et al. (1997). Steroid receptor coactivator-1 is a histone acetyltransferase. Nature 389(6647): 194-198

Swaminathan J., Baxter, E. M. and Corces, V. G. (2005). The role of histone H2Av variant replacement and histone H4 acetylation in the establishment of Drosophila heterochromatin. Genes Dev. 19(1): 65-76. 15630020

Taipale, M., Rea, S., Richter, K., Vilar, A., Lichter, P., Imhof, A. and Akhtar, A. (2005). hMOF histone acetyltransferase is required for histone H4 lysine 16 acetylation in mammalian cells. Mol. Cell. Biol. 25(15):6798-810. 16024812

Torres-Rosell, J., et al. (2007). Anaphase onset before complete DNA replication with intact checkpoint responses. Science 315: 1411-1415. PubMed Citation: 17347440

Tse, C., Fletcher, T. M. and Hansen, J. C. (1998). Enhanced transcription factor access to arrays of histone H3/H4 tetramer.DNA complexes in vitro: implications for replication and transcription. Proc. Natl. Acad. Sci. 95(21): 12169-73

Tsang, C. K., Bertram, P. G., Ai, W., Drenan, R. and Zheng, X. F. (2003). Chromatin-mediated regulation of nucleolar structure and RNA Pol I localization by TOR. EMBO J. 22(22): 6045-56. 14609951

Turner, B. M., Birley, A. J. and Lavender, J. (1992). Histone H4 isoforms acetylated at specific lysine residues define individual chromosomes and chromatin domains in Drosophila polytene nuclei. Cell 69(2): 375-384

Tursun, B., Patel, T., Kratsios, P. and Hobert, O. (2011). Direct conversion of C. elegans germ cells into specific neuron types. Science 331: 304-308. Pubmed: 21148348

Tyler, J. K., et al. (1996). The p55 subunit of Drosophila chromatin assembly factor 1 is homologous to a Histone deacetylase-associated protein. Mol. Cell Biol. 16:6149-6159

van der Heijden, G. W., et al. (2005). Asymmetry in histone H3 variants and lysine methylation between paternal and maternal chromatin of the early mouse zygote. Mech. Dev. 122(9): 1008-22. 15922569

van Wijnen, A. J., et al. (1996). CDP/cut is the DNA-binding subunit of histone gene transcription factor HiNF-D: A mechanism for gene regulation at the G1-D phase cell cycle transition point indiependent of transcription factor E2F. Proc. Natl. Acad. Sci. 93: 11516-21. PubMed Citation: 8876167

van Wijnen, A. J., et al. (1997). Cell cycle-dependent modifications in activities of pRb-related tumor suppressors and proliferation-specific CDP/cut homeodomain factors in murine hematopoietic progenitor cells. J. Cell. Biochem. 66(4): 512-523. PubMed Citation: 9282329

Vermaak, D., et al. (1999). Functional analysis of the SIN3-histone deacetylase RPD3-RbAp48-histone H4 connection in the Xenopus oocyte. Mol. Cell Biol. 19(9): 5847-60. PubMed Citation: 10454532

Verreault, A., et al. (1996). Nucleosome assembly by a complex of CAF-1 and acetylated histones H3 and H4. Cell 87: 95-104. PubMed Citation: 8858152

Vettese-Dadey, M., et al. (1996). Acetylation of histone H4 plays a primary role in enhancing transcription factor binding to nucleosomal DNA in vitro. EMBO J. 15: 2508-2518. PubMed Citation: 8665858

Wakefield, M. J., et al. (1997). Histone underacetylation is an ancient component of mammalian X chromosome inactivation. Proc. Natl. Acad. Sci. 94(18): 9665-9668. PubMed Citation: 9275180

Walker, J. and Bownes, M. (1998). The expression of histone genes during Drosophila melanogaster oogenesis. Dev. Genes Evol. 207(8): 535-41. PubMed Citation: 9510548

Wiekowski, M., et al. (1997). Changes in histone synthesis and modification at the beginning of mouse development correlate with the establishment of chromatin mediated repression of transcription. J. Cell Sci. 110:1147-1158. PubMed Citation: 9191039

Wright, K. L., et al. (1995). Bipartite structure of the proximal promoter of a human H4 histone gene. J. Cell. Biochem. 58: 372-379. PubMed Citation: 7593258

Xie, X., et al. (1996). Structural similarity between TAFs and the heterotetrameric core of the histone octamer. Nature 380: 316-322. PubMed Citation: 8598927

Xiao, B., et al. (2005). Specificity and mechanism of the histone methyltransferase Pr-Set7. Genes Dev. 19: 1444-1454. 15933069

Zhang, J., Tan, D., DeRose, E. F., Perera, L., Dominski, Z., Marzluff, W. F., Tong, L. and Hall, T. M. (2014). Molecular mechanisms for the regulation of histone mRNA stem-loop-binding protein by phosphorylation. Proc Natl Acad Sci U S A 111(29): E2937-46. PubMed ID: 25002523

Zhou, J., et al. (2009). Histone deacetylase Rpd3 antagonizes Sir2-dependent silent chromatin propagation. Nucleic Acids Res. 37(11): 3699-3713. PubMed Citation: 19372273

Zippo, A., et al. (2009). Histone crosstalk between H3S10ph and H4K16ac generates a histone code that mediates transcription elongation. Cell 138(6): 1122-36. PubMed Citation: 19766566


Histone H4: Biological Overview | Evolutionary Homologs | Regulation | Developmental Biology

date revised: 26 December 2020

Home page: The Interactive Fly © 1997 Thomas B. Brody, Ph.D.

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