Skip to main content
Nucleus logoLink to Nucleus
. 2011 May-Jun;2(3):182–188. doi: 10.4161/nucl.2.3.15876

Where splicing joins chromatin

Jarmila Hnilicová 1, David Staněk 1,
PMCID: PMC3149878  PMID: 21818411

Abstract

There are numerous data suggesting that two key steps in gene expression—transcription and splicing influence each other closely. For a long time it was known that chromatin modifications regulate transcription, but only recently it was shown that chromatin and histone modifications play a significant role in pre-mRNA splicing. Here we summarize interactions between splicing machinery and chromatin and discuss their potential functional significance. We focus mainly on histone acetylation and methylation and potential mechanisms of their role in splicing. It seems that whereas histone acetylation acts mainly by alterating the transcription rate, histone methylation can also influence splicing directly by recruiting various splicing components.

Key words: chromatin, exon, alternative splicing, transcription, snRNP, histone methylation, histone acetylation, nucleosome

Pre-mRNA Splicing Regulation

The vast majority of our genes has a puzzle-like structure where short coding sequences (exons) must be correctly identified and joined together while surrounding much longer non-coding sequences (introns) have to be removed before a mature mRNA is exported to the cytoplasm and translated. A typical human gene contains 8 introns of average length ∼3.4 kb but much larger examples were found (e.g., the last intron of glypican 5 gene is 721 kb in length).1 In contrast to long introns, exons are short having an average length of 145 bp and represent only a small fraction of primary pre-mRNA transcripts. An extreme example is the largest human gene CNTNAP2 encoding Caspr2 protein which spans over 2.3 Mb (!) but its spliced mRNA is only 10 kb long because 99.6% of the CNTNAP2 gene sequence corresponds to introns.2

To further complicate matters, exons and introns are not always recognized identically and, for example, a particular exon can be skipped in a fraction of mature mRNAs. This process, called alternative splicing, produces different mRNA isoforms from one gene which significantly increases the coding potential of our genome. It is estimated that in human cells almost 95% of genes are alternatively spliced3,4 and that there are on average 7 different alternative splicing events per single gene leading to different mRNA variants.4 The complexity of the process puts pressure on cells to regulate splicing precisely. Then how is alternative splicing regulated?

Introns contain consensus splice sites at both ends. However, compared to yeasts, most splice site sequences in higher metazoans are degenerate and additional regulatory sequences in pre-mRNA are needed to help basal splicing machinery (the spliceosome) to recognize correct splice sites.5 These regulatory sequences are bound by many different splicing regulatory proteins which are able to interact with the splicing complex. The combinatorial interplay among splicing factors results in the usage (or skipping) of individual splice sites. Some of the splicing regulatory proteins are widely expressed in different tissues (such as PTB or SR proteins) while the others are highly tissue specific (for example, Nova proteins are expressed almost exclusively in neurons). The differences in the expression of splicing regulatory proteins are believed to be responsible for tissue-specific alternative splicing.6

The splicing code—the sum of all splicing related features in a pre-mRNA sequence—can explain a majority of differences in alternative splicing between individual tissues (e.g., 74% out of 97 tested alternative splicing events specific for the central nervous system and muscle tissue were properly predicted based on a pre-mRNA sequence).7 Although the splicing code model was effective to estimate whether the alternative exon was included or skipped it was less precise in the prediction of the level of inclusion/exclusion when tested by Barash et al. which indicates a presence of additional regulatory signals.7

Splicing is Cotranscriptional and Occurs in the Vicinity of Chromatin

Intron recognition and removal likely occur in the cell nucleus very shortly after intron transcription. Splicing complexes associate with pre-mRNA immediately after the target sequences are synthesized and splicing of many introns is completed before pre-mRNA transcription termination.815 This was shown not only for long human genes but also for many yeast genes, which are much shorter than human genes and are spliced cotranscriptionally as well.1621 Moreover, splicing can induce pausing of RNA polymerase II during transcription21 and RNA polymerase II was shown to pause in terminal exons, which increases the time window for cotranscriptional splicing.16

This suggests that transcription and splicing are coupled not only in time, but also functionally. It was shown that the promoter composition and the speed of transcription can influence splicing.2233 The speed of transcription affects the dynamics of presentation of individual splice and regulatory sites to the splicing machinery, which then results in using different splice sites not only in transfected minigenes but also in endogenous genes.3443 The carboxyterminal domain of RNA polymerase II itself can also contribute to splicing factors recruitment which then might influence alternative splicing.44,45

Average eukaryotic RNA polymerase II elongation rate vary from ∼1.5–4.5 kb/min,8,14,4649 which means that an average intron is transcribed in 1–3 minutes. An average splicing event takes 30 seconds to accomplish19,20,50,51 and therefore it is very likely that recognition of splice sites as well as intron excision occur when pre-mRNA is still in close proximity to the chromatin of its own gene, and this makes it possible for chromatin modifications to influence splicing.

Chromatin Marks Affecting Splicing

In cells, DNA is packed together with histones in nucleosomes. Each nucleosome consists of an octamer composed of four core histones (H2A, H2B, H3 and H4) and 147 bp DNA, which is wrapped around this octamer. Histones can be posttranslationally modified on many sites, especially at their N-terminal ends that protrude out of the nucleosomal core. Regulation of gene activity via nucleosome positioning and histone modifications has been well established.52,53 However, it seems that nucleosome positions and modifications are also important for exon recognition. In the genome, nucleosomes are preferentially found at exons and exon-positioned nucleosomes carry a specific set of histone modifications (see Fig. 1),5462 although the latter statement is still a matter of a debate.63,64 Three histone marks were shown to affect splicing and/or to mediate interaction of histones with splicing factors: methylation, acetylation and phosphorylation (see Table 1). These modifications differ in the charge they introduce into the chromatin. While methylations are electrostatically neutral and their action is mediated via specific protein readers, it is supposed that histone acetylation, which adds a negative charge, can itself affect nucleosomal stability and make chromatin more accessible (reviewed in ref. 61). Histone phosphorylation, which brings a negative charge, can regulate the binding of SR proteins to the chromatin during the cell cycle but its role in splicing regulation is unclear.65 The modifications also differ substantially in their turnover: the addition and removal of acetylation is rapid and most acetyl residues are exchanged within minutes, phosphorylations can last up to several hours and methylations, the most stable ones, can be present even for days.66

Figure 1.

Figure 1

Different sets of chromatin modifications were identified at exons and at introns. The enrichment of a specific histone mark depends on the position of an exon within a gene (e.g., H2BK5me1 and H4K20me1 are found preferentially at exons toward the 5′ end of genes),61 or it can depend on gene expression (highlighted in bold), e.g., H3K27me3 is high at exons of low expressed genes, but it was reported to be elevated at introns of highly expressed genes.58

Table 1.

Interaction of chromatin and splicing machinery

Histone modification >Interacting protein >Link to splicing
H3K4me3 Chd1 Chd1 associates with SRp20 (SRSF3)119 and U2 snRNP (via SF3 subunits) and increases efficiency of pre-mRNA splicing86
H3K4me3 Sgf29 Sgf29 interacts with SF3B5 (SF3b10) and SF3B3 (SF3b130) subunits of U2 snRNP87
H3K9me3 PTB, hnRNP A1, hnRNP A/B, hnRNP A2/B1, hnRNP K, hnRNP L PTB and most of the hnRNP proteins are direct regulators of alternative splicing but it is not known whether the association of hnRNP proteins with H3K9me387 affects splicing
methylated H3K9 HP1 (HP1a) HP1 binds to Drosophila hnRNP proteins (PEP, DDP1, HRB87F)114
H3K36me3 MRG15 MRG15 recruits PTB; tethering of PTB to chromatin changes alternative splicing88
methylated H3K79 TP53BP1 TP53BP1 immunoprecipitates U1 and U2 snRNA (but also other small RNAs)120
histone H3 (not phosphorylated at serine 10) SRp20 (SRFS3), SF2/ASF (SRSF1) Both SR proteins participate in constitutive and alternative splicing, but the role of interaction with histone H3 in splicing is not known65
DNA methylation MeCP2 MeCP2 regulates alternative splicing121

Histone Acetylation Induces Changes in the Alternative Splicing

Histone acetylation can be easily increased in cells by histone deacetylase (HDAC) inhibitors and the treatment with HDAC inhibitors influences an alternative splicing pattern of transiently expressed minigene splicing reporters.36 Recently, we showed that HDAC inhibition changes alternative splicing of ∼700 endogenous human genes. In addition, HDAC1 depletion affected the splicing pattern of the fibronectin gene.38 The correlation between histone acetylation and alternative splicing was also observed with the NCAM gene, whose alternative splicing is regulated by depolarization of neuronal membranes.67 In this instance, it is worthy to note that many genes that changed alternative splicing after HDAC inhibition were encoding various ion channels and regulators of cell cycle or apoptosis.38 Thus it seems that acetylation preferentially modulates alternative splicing of genes that must react quickly to changing conditions.

The mechanism by which acetylation modulates splicing has been connected to RNA polymerase II processivity. Increased RNA polymerase II processivity, which is induced by histone acetylation, modulates the association of splicing regulators with the nascent RNA as was shown in the case of the fibronectin gene and the splicing regulator SRp40 (SRSF5).38 However, in yeast the deletion of histone deacetylases or the histone acetyltransferase Gcn5 influenced cotranscriptional recruitment of spliceosomal proteins without obvious changes in RNA polymerase II distribution along the gene.68,69 It is then possible that histone acetylation also affects splicing directly via the recruitment of spliceosomal subunits, without altering transcription elongation.

Mutual Influence between Splicing Machinery and Chromatin

The spliceosome assembles from five small ribonucleoprotein particles (snRNPs) that assemble on the pre-mRNA in a stepwise manner.17,18,50,7075 Each snRNP consists of several proteins and one small nuclear RNA (snRNA) according to which the snRNPs are named.76 First, the U1 snRNP binds to the premRNA and recognizes the 5′ splice site. Then, the U2 snRNP binds to the vicinity of the 3′ splice site, closely followed by the U4/U6•U5 tri-snRNP. After tri-snRNP incorporation the spliceosome undergoes extensive rearrangements that result in the formation of an active complex that catalyzes both steps of the splicing reaction. The U1 and U2 snRNP binding is important for the accuracy of splicing as this defines the intron. U1 and U2 snRNPs associate with pre-mRNA quickly and they are already present when RNA polymerase transcribes only hundreds of nucleotides downstream of splice sites.11,17,74,75 This implies that snRNPs are in close vicinity to chromatin and their association with pre-mRNA could be influenced by chromatin and vice versa, snRNPs themselves might influence the chromatin state. One example of plausible mutual influence is discussed below.

5′ Splice Site Influences Transcription and Chromatin

U1 snRNA interacts with chromatin77 and it has been shown that the presence of functional 5′ splice site in the gene can affect chromatin state.78 However, U1 snRNA—chromatin interaction might reflect only the role of U1 snRNA in transcription. The U1 snRNA associates with proteins from two general transcription factor complexes, TFIID and TFIIH.77,79 In the TFIIH complex U1 snRNA enhances transcription presumably through the interaction with cyclin H and regulation of the cyclin H associated kinase CDK7,79,80 which phosphorylates carboxyterminal domain of RNA polymerase II. It was shown that the functional 5′ splice site near the promoter enhances the assembly of general transcription factors at this promoter.78 If the presence of U1 snRNA (and U1 snRNP) stimulates transcription then introncontaining genes should be transcribed more than intron-less genes. This is true in Saccharomyces cerevisiae where almost half of all introns found in the genome are located in highly expressed genes coding ribosomal proteins (95 out of 132 of these genes contain introns).70,81 Moreover, deletion of introns from two nonribosomal yeast genes decreases the expression of these genes.82 In mammalian cells, the presence of intron with functional splice sites also increases gene expression.78,8285

Interestingly, the presence of a functional 5′ splice site can increase gene specific H3K9 acetylation and H3K4me3,78 which are chromatin marks associated with active transcription. In another study, the level of H3K4me3 at the minigene decreased when both splice sites were mutated compared to the minigene containing intron with functional splice sites, however the reported difference was quite small.83 It is not known whether these changes in the chromatin state are coupled generally with splicing or whether they are specifically related to U1 snRNP recruitment. Moreover, it is difficult to distinguish between the cause and the consequence in these observations. Chromatin modifications might reflect only changes in transcription and the increase of H3K4 trimethylation and histone H3 acetylation occurs independently of splicing. An example of mutual dependence is H3K4me3: the level of H3K4me3 correlates with transcription activity and is enhanced by the presence of a functional 5′ splice site within the gene but in turn this chromatin modification affects splicing (see below).

H3K4me3 Associates with U2 snRNP

Association of the U2 snRNP with chromatin is mediated by trimethylated lysine 4 histone H3.86,87 In addition, the functional significance of H3K4me3 for splicing was shown.86,88 Several complexes that methylate lysine 4 of histone H3 in human cells share the same structural subunits.89 The depletion of one of these subunits, Ash2, leads to the decrease of overall H3K4me3.90,91 Ash2 depletion causes a decrease of splicing efficiency immediately after induction of transcription86 and Ash2 overexpression affects alternative splicing.88

U2 snRNP binds to chromatin via the chromatin remodeling protein Chd1 that binds H3K4me3 and enhances pre-mRNA splicing efficiency within a short time window after activation of gene transcription.86 U2 snRNP subunits also interact with another H3K4me3 reader, Sgf29 protein.87 Considering that H3K4me3 is frequently elevated around transcription start sites,9297 it is surprising that this particular histone modification interacts with U2 snRNP. However, the level of H3K4me3 positively correlates with gene expression indicating that U2 snRNP might be recruited preferentially to highly expressed genes.93,94,97

As mentioned before, U2 snRNP recruitment to the transcription unit is also regulated by the histone acetyltransferase Gcn5.68 Although Gcn5 participates in a number of protein complexes the most important for U2 snRNP recruitment seems to be the SAGA complex.68,98 Gcn5 in the SAGA complex acetylates predominantly histones at promoters68 that subsequently serve as an anchor stabilizing the binding of the whole SAGA complex to the promoter.99 Interestingly, yeast Msl1 (a homologue of mammalian U2B”, one of the U2 snRNP proteins) was also found at promoters.68 Sgf29 and Chd1, both of the proteins interacting with U2 snRNP in human cells, are conserved in yeasts and they are components of the yeast SAGA complex.98 It is possible that Gcn5 plays a role in the recruitment of the SAGA complex to promoters and that SAGA complex subunits are bringing the U2 snRNP. In human cells, the STAGA complex (homologue of the yeast SAGA complex) interacts with the U2 snRNP specific protein SF3B3 (SF3b130).100

MRG15: Chromatin Binding Protein that Recruits Splicing Regulator PTB

PTB is an example of a splicing regulator that was shown to associate with chromatin and, more importantly, to modulate through this association alternative splicing of a number of genes.88 PTB binds chromatin via MRG15 protein and this interaction preferentially regulates splicing of pre-mRNAs with weak PTB binding sites. MRG15 contains chromodomain that is able to bind H3K36me3 101 and H3K36me3 level on a particular gene correlates with PTB recruitment to pre-mRNAs transcribed from this gene.88 Interestingly, MRG15 does not only bind H3K36me3, but also interacts with H3K4me3.87 A simple MRG15 knock-down influences alternative splicing of more than 180 genes independently of PTB.88 This suggests that there might be an alternative mechanism that does not depend on PTB, although many of these alternative splicing changes might be due to the indirect effects caused by altered expression of splicing regulators after MRG15 knockdown. MRG15 is a component of several chromatin modifying complexes. These complexes contain histone acetylases Tip60 102,103 and hMOF,104 histone deacetylases such as HDAC1 or HDAC2 105,106 and histone demethylases.107,108 Eaf3, a yeast orthologue of MRG15, associates with RNA polymerase II109 and is important for histone deacetylation during transcription elongation that inhibits the internal transcription initiation inside coding sequences of genes.110,111 In mammalian cells, a similar complex that consist of both HDAC1 and MRG15 was identified recently and it was shown to affect progression of RNA polymerase II through transcribed regions.106 Thus, MRG15 significantly participates in regulation of chromatin acetylation and transcription, processes that were shown to modulate splicing outcome (see above) and it would be interesting to know whether MRG15 alone is able to affect alternative splicing directly through modulation of chromatin.

hnRNP Proteins and H3K9 Methylation

H3K9 methylation, which recruits the HP1 protein, is important for the formation and maintenance of heterochromatin. Although average levels of H3K9me3 and H3K9me2 correlate with gene silencing, both histone modifications can be also detected in some actively transcribed genes.61,93,112,113 Recently, PTB, together with other splicing regulators hnRNPs (hnRNP A1, hnRNP L, hnRNP K, hnRNP A2/B1, hnRNP A/B), were identified as proteins able to recognize H3K9me3.87 In Drosophila, several other hnRNP proteins interact with HP1 and heterochromatin.114 What can be the reason for such an association? One of the H3K9me3 recognizing proteins, hnRNP K, binds to large intergenic noncoding RNAs (linc-RNAs) and the complex of linc-RNA and hnRNP K is an effector of p53 signaling and represses transcription of hundreds of genes.115 When PTB-RNA interactions were analyzed genome-wide by CLIP-seq method, a substantial fraction of tags (almost 30%) was mapped to intergenic regions, implying that PTB may also bind to noncoding RNAs.116 PTB was also detected at the promoter of the HMGA2 gene containing a very specific ∼60 bp polypyrimidine/polypurine element but it is unclear whether binding to this promoter is mediated by chromatin.117,118 It is intriguing to speculate that the association of hnRNP proteins with the chromatin might regulate transcription. In addition, hnRNP proteins binding to HP1 are involved in heterochromatin formation in Drosophila.114

H3K9me2 and H3K9me3 were also reported to be enriched at introns in contrast to exons.58,59 hnRNP proteins associated with chromatin then might help to define intronic sequences. In human cells, the increase of heterochromatin marks (H3K9me2, H3K27me3) together with the presence of HP1α leads to changes in alternative splicing of the fibronectin gene.40 Recently, it was shown that association of HP1γ with H3K9me3 modulates alternative splicing of several genes including CD44.113 Although the induced heterochromatin formation near the alternative exons reduces the processivity of RNA polymerase II which causes the increased inclusion of alternative exon, it is still possible that methylation of H3K9 influences splicing of certain genes directly via tethering of splicing factors. Interestingly, exons enriched in the H3K9me3 are preferentially excluded from mature mRNAs indicating again the role of H3K9me3 in the regulation of splicing.61

Perspectives

The major task in the near future will be to reveal molecular details and mechanism(s) through which chromatin modifications affect pre-mRNA splicing. This knowledge is important not only from a theoretical point of view but might be essential for practical medical usage. Many small molecules that influence chromatin modifications serve as therapeutical agents and thus there is a hope that these compounds might help to treat splicing related diseases as well. The better we understand the connection between splicing and chromatin the higher chance to identify such a drug.

Acknowledgments

We thank Stanek's lab members for helpful comments. Projects in David Stanek's lab are supported by grants from the Czech Science Foundation (P305/10/0424) and from the Academy of Sciences of the Czech Republic (KAN200520801, AV0Z50520514).

Abbreviations

HDAC

histone deacetylase

snRNP

small ribonucleoprotein particle

hnRNP

heterogeneous nuclear ribonucleoprotein

References

  • 1.Lander ES, Linton LM, Birren B, Nusbaum C, Zody MC, Baldwin J, et al. Initial sequencing and analysis of the human genome. Nature. 2001;409:860–921. doi: 10.1038/35057062. [DOI] [PubMed] [Google Scholar]
  • 2.Scherer S. A short guide to the human genome. Cold Spring Harbor NY: Cold Spring Harbor Laboratory Press; 2008. p. 34. [Google Scholar]
  • 3.Wang ET, Sandberg R, Luo S, Khrebtukova I, Zhang L, Mayr C. Alternative isoform regulation in human tissue transcriptomes. Nature. 2008;456:470–476. doi: 10.1038/nature07509. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Pan Q, Shai O, Lee LJ, Frey BJ, Blencowe BJ. Deep surveying of alternative splicing complexity in the human transcriptome by high-throughput sequencing. Nat Genet. 2008;40:1413–1415. doi: 10.1038/ng.259. [DOI] [PubMed] [Google Scholar]
  • 5.Wang Z, Burge CB. Splicing regulation: from a parts list of regulatory elements to an integrated splicing code. RNA. 2008;14:802–813. doi: 10.1261/rna.876308. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Black DL. Mechanisms of alternative pre-messenger RNA splicing. Annu Rev Biochem. 2003;72:291–336. doi: 10.1146/annurev.biochem.72.121801.161720. [DOI] [PubMed] [Google Scholar]
  • 7.Barash Y, Calarco JA, Gao W, Pan Q, Wang X, Shai O, et al. Deciphering the splicing code. Nature. 2010;465:53–59. doi: 10.1038/nature09000. [DOI] [PubMed] [Google Scholar]
  • 8.Singh J, Padgett RA. Rates of in situ transcription and splicing in large human genes. Nat Struct Mol Biol. 2009;16:1128–1133. doi: 10.1038/nsmb.1666. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Tennyson CN, Klamut HJ, Worton RG. The human dystrophin gene requires 16 hours to be transcribed and is cotranscriptionally spliced. Nat Genet. 1995;9:184–190. doi: 10.1038/ng0295-184. [DOI] [PubMed] [Google Scholar]
  • 10.Pandya-Jones A, Black DL. Co-transcriptional splicing of constitutive and alternative exons. RNA. 2009;15:1896–1908. doi: 10.1261/rna.1714509. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Listerman I, Sapra AK, Neugebauer KM. Cotranscriptional coupling of splicing factor recruitment and precursor messenger RNA splicing in mammalian cells. Nat Struct Mol Biol. 2006;13:815–822. doi: 10.1038/nsmb1135. [DOI] [PubMed] [Google Scholar]
  • 12.Zhang G, Taneja KL, Singer RH, Green MR. Localization of pre-mRNA splicing in mammalian nuclei. Nature. 1994;372:809–812. doi: 10.1038/372809a0. [DOI] [PubMed] [Google Scholar]
  • 13.Bauren G, Wieslander L. Splicing of Balbiani ring 1 gene pre-mRNA occurs simultaneously with transcription. Cell. 1994;76:183–192. doi: 10.1016/0092-8674(94)90182-1. [DOI] [PubMed] [Google Scholar]
  • 14.Wada Y, Ohta Y, Xu M, Tsutsumi S, Minami T, Inoue K, et al. A wave of nascent transcription on activated human genes. Proc Natl Acad Sci USA. 2009;106:18357–18361. doi: 10.1073/pnas.0902573106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Bittencourt D, Dutertre M, Sanchez G, Barbier J, Gratadou L, Auboeuf D. Cotranscriptional splicing potentiates the mRNA production from a subset of estradiol-stimulated genes. Mol Cell Biol. 2008;28:5811–5824. doi: 10.1128/MCB.02231-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Carrillo Oesterreich F, Preibisch S, Neugebauer KM. Global analysis of nascent RNA reveals transcriptional pausing in terminal exons. Mol Cell. 2010;40:571–581. doi: 10.1016/j.molcel.2010.11.004. [DOI] [PubMed] [Google Scholar]
  • 17.Lacadie SA, Rosbash M. Cotranscriptional spliceosome assembly dynamics and the role of U1 snRNA:5′ss base pairing in yeast. Mol Cell. 2005;19:65–75. doi: 10.1016/j.molcel.2005.05.006. [DOI] [PubMed] [Google Scholar]
  • 18.Gornemann J, Kotovic KM, Hujer K, Neugebauer KM. Cotranscriptional spliceosome assembly occurs in a stepwise fashion and requires the cap binding complex. Mol Cell. 2005;19:53–63. doi: 10.1016/j.molcel.2005.05.007. [DOI] [PubMed] [Google Scholar]
  • 19.Lacadie SA, Tardiff DF, Kadener S, Rosbash M. In vivo commitment to yeast cotranscriptional splicing is sensitive to transcription elongation mutants. Genes Dev. 2006;20:2055–2066. doi: 10.1101/gad.1434706. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Alexander RD, Barrass JD, Dichtl B, Kos M, Obtulowicz T, Robert MC, et al. RiboSys, a high-resolution, quantitative approach to measure the in vivo kinetics of pre-mRNA splicing and 3′-end processing in Saccharomyces cerevisiae. RNA. 2010;16:2570–2580. doi: 10.1261/rna.2162610. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Alexander RD, Innocente SA, Barrass JD, Beggs JD. Splicing-dependent RNA polymerase pausing in yeast. Mol Cell. 2010;40:582–593. doi: 10.1016/j.molcel.2010.11.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Auboeuf D, Honig A, Berget SM, O'Malley BW. Coordinate regulation of transcription and splicing by steroid receptor coregulators. Science. 2002;298:416–419. doi: 10.1126/science.1073734. [DOI] [PubMed] [Google Scholar]
  • 23.Kadener S, Cramer P, Nogues G, Cazalla D, de la Mata M, Fededa JP, et al. Antagonistic effects of T-Ag and VP16 reveal a role for RNA pol II elongation on alternative splicing. EMBO J. 2001;20:5759–5768. doi: 10.1093/emboj/20.20.5759. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Kadener S, Fededa JP, Rosbash M, Kornblihtt AR. Regulation of alternative splicing by a transcriptional enhancer through RNA pol II elongation. Proc Natl Acad Sci USA. 2002;99:8185–8190. doi: 10.1073/pnas.122246099. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Kornblihtt AR. Promoter usage and alternative splicing. Curr Opin Cell Biol. 2005;17:262–268. doi: 10.1016/j.ceb.2005.04.014. [DOI] [PubMed] [Google Scholar]
  • 26.Moldon A, Malapeira J, Gabrielli N, Gogol M, Gomez-Escoda B, Ivanova T, et al. Promoter-driven splicing regulation in fission yeast. Nature. 2008;455:997–1000. doi: 10.1038/nature07325. [DOI] [PubMed] [Google Scholar]
  • 27.Cramer P, Caceres JF, Cazalla D, Kadener S, Muro AF, Baralle FE, et al. Coupling of transcription with alternative splicing: RNA pol II promoters modulate SF2/ASF and 9G8 effects on an exonic splicing enhancer. Mol Cell. 1999;4:251–258. doi: 10.1016/s1097-2765(00)80372-x. [DOI] [PubMed] [Google Scholar]
  • 28.Cramer P, Pesce CG, Baralle FE, Kornblihtt AR. Functional association between promoter structure and transcript alternative splicing. Proc Natl Acad Sci USA. 1997;94:11456–11460. doi: 10.1073/pnas.94.21.11456. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Pagani F, Stuani C, Zuccato E, Kornblihtt AR, Baralle FE. Promoter architecture modulates CFTR exon 9 skipping. J Biol Chem. 2003;278:1511–1517. doi: 10.1074/jbc.M209676200. [DOI] [PubMed] [Google Scholar]
  • 30.Auboeuf D, Dowhan DH, Li X, Larkin K, Ko L, Berget SM, et al. CoAA, a nuclear receptor coactivator protein at the interface of transcriptional coactivation and RNA splicing. Mol Cell Biol. 2004;24:442–453. doi: 10.1128/MCB.24.1.442-453.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Auboeuf D, Dowhan DH, Kang YK, Larkin K, Lee JW, Berget SM, et al. Differential recruitment of nuclear receptor coactivators may determine alternative RNA splice site choice in target genes. Proc Natl Acad Sci USA. 2004;101:2270–2274. doi: 10.1073/pnas.0308133100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Rosonina E, Bakowski MA, McCracken S, Blencowe BJ. Transcriptional activators control splicing and 3′-end cleavage levels. J Biol Chem. 2003;278:43034–43040. doi: 10.1074/jbc.M307289200. [DOI] [PubMed] [Google Scholar]
  • 33.Monsalve M, Wu Z, Adelmant G, Puigserver P, Fan M, Spiegelman BM. Direct coupling of transcription and mRNA processing through the thermogenic coactivator PGC-1. Mol Cell. 2000;6:307–316. doi: 10.1016/s1097-2765(00)00031-9. [DOI] [PubMed] [Google Scholar]
  • 34.Munoz MJ, Perez Santangelo MS, Paronetto MP, de la Mata M, Pelisch F, Boireau S, et al. DNA damage regulates alternative splicing through inhibition of RNA polymerase II elongation. Cell. 2009;137:708–720. doi: 10.1016/j.cell.2009.03.010. [DOI] [PubMed] [Google Scholar]
  • 35.de la Mata M, Alonso CR, Kadener S, Fededa JP, Blaustein M, Pelisch F, et al. A slow RNA polymerase II affects alternative splicing in vivo. Mol Cell. 2003;12:525–532. doi: 10.1016/j.molcel.2003.08.001. [DOI] [PubMed] [Google Scholar]
  • 36.Nogues G, Kadener S, Cramer P, Bentley D, Kornblihtt AR. Transcriptional activators differ in their abilities to control alternative splicing. J Biol Chem. 2002;277:43110–43114. doi: 10.1074/jbc.M208418200. [DOI] [PubMed] [Google Scholar]
  • 37.Ip JY, Schmidt D, Pan Q, Ramani AK, Fraser AG, Odom DT, et al. Global impact of RNA polymerase II elongation inhibition on alternative splicing regulation. Genome Res. 2011;21:390–401. doi: 10.1101/gr.111070.110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Hnilicova J, Hozeifi S, Duskova E, Icha J, Tomankova T, Stanek D. Histone deacetylase activity modulates alternative splicing. PLoS One. 2011;6:16727. doi: 10.1371/journal.pone.0016727. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Batsche E, Yaniv M, Muchardt C. The human SWI/SNF subunit Brm is a regulator of alternative splicing. Nat Struct Mol Biol. 2006;13:22–29. doi: 10.1038/nsmb1030. [DOI] [PubMed] [Google Scholar]
  • 40.Allo M, Buggiano V, Fededa JP, Petrillo E, Schor I, de la Mata M, et al. Control of alternative splicing through siRNA-mediated transcriptional gene silencing. Nat Struct Mol Biol. 2009;16:717–724. doi: 10.1038/nsmb.1620. [DOI] [PubMed] [Google Scholar]
  • 41.Howe KJ, Kane CM, Ares M., Jr Perturbation of transcription elongation influences the fidelity of internal exon inclusion in Saccharomyces cerevisiae. RNA. 2003;9:993–1006. doi: 10.1261/rna.5390803. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Roberts GC, Gooding C, Mak HY, Proudfoot NJ, Smith CW. Co-transcriptional commitment to alternative splice site selection. Nucleic Acids Res. 1998;26:5568–5572. doi: 10.1093/nar/26.24.5568. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Nogues G, Munoz MJ, Kornblihtt AR. Influence of polymerase II processivity on alternative splicing depends on splice site strength. J Biol Chem. 2003;278:52166–52171. doi: 10.1074/jbc.M309156200. [DOI] [PubMed] [Google Scholar]
  • 44.de la Mata M, Kornblihtt AR. RNA polymerase II C-terminal domain mediates regulation of alternative splicing by SRp20. Nat Struct Mol Biol. 2006;13:973–980. doi: 10.1038/nsmb1155. [DOI] [PubMed] [Google Scholar]
  • 45.Barboric M, Lenasi T, Chen H, Johansen EB, Guo S, Peterlin BM. 7SK snRNP/P-TEFb couples transcription elongation with alternative splicing and is essential for vertebrate development. Proc Natl Acad Sci USA. 2009;106:7798–7803. doi: 10.1073/pnas.0903188106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Darzacq X, Shav-Tal Y, de Turris V, Brody Y, Shenoy SM, Phair RD, et al. In vivo dynamics of RNA polymerase II transcription. Nat Struct Mol Biol. 2007;14:796–806. doi: 10.1038/nsmb1280. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Boireau S, Maiuri P, Basyuk E, de la Mata M, Knezevich A, Pradet-Balade B, et al. The transcriptional cycle of HIV-1 in real-time and live cells. J Cell Biol. 2007;179:291–304. doi: 10.1083/jcb.200706018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Mason PB, Struhl K. Distinction and relationship between elongation rate and processivity of RNA polymerase II in vivo. Mol Cell. 2005;17:831–840. doi: 10.1016/j.molcel.2005.02.017. [DOI] [PubMed] [Google Scholar]
  • 49.Ben-Ari Y, Brody Y, Kinor N, Mor A, Tsukamoto T, Spector DL, et al. The life of an mRNA in space and time. J Cell Sci. 2010;123:1761–1774. doi: 10.1242/jcs.062638. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Huranova M, Ivani I, Benda A, Poser I, Brody Y, Hof M, et al. The differential interaction of snRNPs with pre-mRNA reveals splicing kinetics in living cells. J Cell Biol. 2010;191:75–86. doi: 10.1083/jcb.201004030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Wetterberg I, Zhao J, Masich S, Wieslander L, Skoglund U. In situ transcription and splicing in the Balbiani ring 3 gene. EMBO J. 2001;20:2564–2574. doi: 10.1093/emboj/20.10.2564. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Kouzarides T. Chromatin modifications and their function. Cell. 2007;128:693–705. doi: 10.1016/j.cell.2007.02.005. [DOI] [PubMed] [Google Scholar]
  • 53.Radman-Livaja M, Rando OJ. Nucleosome positioning: how is it established and why does it matter? Dev Biol. 2009;339:258–266. doi: 10.1016/j.ydbio.2009.06.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Kolasinska-Zwierz P, Down T, Latorre I, Liu T, Liu XS, Ahringer J. Differential chromatin marking of introns and expressed exons by H3K36me3. Nat Genet. 2009;41:376–381. doi: 10.1038/ng.322. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Schwartz S, Meshorer E, Ast G. Chromatin organization marks exon-intron structure. Nat Struct Mol Biol. 2009;16:990–995. doi: 10.1038/nsmb.1659. [DOI] [PubMed] [Google Scholar]
  • 56.Tilgner H, Nikolaou C, Althammer S, Sammeth M, Beato M, Valcarcel J, et al. Nucleosome positioning as a determinant of exon recognition. Nat Struct Mol Biol. 2009;16:996–1001. doi: 10.1038/nsmb.1658. [DOI] [PubMed] [Google Scholar]
  • 57.Andersson R, Enroth S, Rada-Iglesias A, Wadelius C, Komorowski J. Nucleosomes are well positioned in exons and carry characteristic histone modifications. Genome Res. 2009;19:1732–1741. doi: 10.1101/gr.092353.109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Dhami P, Saffrey P, Bruce AW, Dillon SC, Chiang K, Bonhoure N, et al. Complex exon-intron marking by histone modifications is not determined solely by nucleosome distribution. PLoS One. 2009;5:12339. doi: 10.1371/journal.pone.0012339. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Spies N, Nielsen CB, Padgett RA, Burge CB. Biased chromatin signatures around polyadenylation sites and exons. Mol Cell. 2009;36:245–254. doi: 10.1016/j.molcel.2009.10.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Huff JT, Plocik AM, Guthrie C, Yamamoto KR. Reciprocal intronic and exonic histone modification regions in humans. Nat Struct Mol Biol. 2010;17:1495–1499. doi: 10.1038/nsmb.1924. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Hon G, Wang W, Ren B. Discovery and annotation of functional chromatin signatures in the human genome. PLoS Comput Biol. 2009;5:1000566. doi: 10.1371/journal.pcbi.1000566. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Nahkuri S, Taft RJ, Mattick JS. Nucleosomes are preferentially positioned at exons in somatic and sperm cells. Cell Cycle. 2009;8:3420–3424. doi: 10.4161/cc.8.20.9916. [DOI] [PubMed] [Google Scholar]
  • 63.Schwartz S, Ast G. Chromatin density and splicing destiny: on the cross-talk between chromatin structure and splicing. EMBO J. 2010;29:1629–1636. doi: 10.1038/emboj.2010.71. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Tilgner H, Guigo R. From chromatin to splicing: RNA-processing as a total artwork. Epigenetics. 2010:5. doi: 10.4161/epi.5.3.11319. [DOI] [PubMed] [Google Scholar]
  • 65.Loomis RJ, Naoe Y, Parker JB, Savic V, Bozovsky MR, Macfarlan T, et al. Chromatin binding of SRp20 and ASF/SF2 and dissociation from mitotic chromosomes is modulated by histone H3 serine 10 phosphorylation. Mol Cell. 2009;33:450–461. doi: 10.1016/j.molcel.2009.02.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Barth TK, Imhof A. Fast signals and slow marks: the dynamics of histone modifications. Trends Biochem Sci. 2010;35:618–626. doi: 10.1016/j.tibs.2010.05.006. [DOI] [PubMed] [Google Scholar]
  • 67.Schor IE, Rascovan N, Pelisch F, Allo M, Kornblihtt AR. Neuronal cell depolarization induces intragenic chromatin modifications affecting NCAM alternative splicing. Proc Natl Acad Sci USA. 2009;106:4325–4330. doi: 10.1073/pnas.0810666106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Gunderson FQ, Johnson TL. Acetylation by the transcriptional coactivator Gcn5 plays a novel role in co-transcriptional spliceosome assembly. PLoS Genet. 2009;5:1000682. doi: 10.1371/journal.pgen.1000682. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Gunderson FQ, Merkhofer EC, Johnson TL. Dynamic histone acetylation is critical for cotranscriptional spliceosome assembly and spliceosomal rearrangements. Proc Natl Acad Sci USA. 2011;108:2004–2009. doi: 10.1073/pnas.1011982108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Kotovic KM, Lockshon D, Boric L, Neugebauer KM. Cotranscriptional recruitment of the U1 snRNP to intron-containing genes in yeast. Mol Cell Biol. 2003;23:5768–5779. doi: 10.1128/MCB.23.16.5768-5779.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Konarska MM, Sharp PA. Electrophoretic separation of complexes involved in the splicing of precursors to mRNAs. Cell. 1986;46:845–855. doi: 10.1016/0092-8674(86)90066-8. [DOI] [PubMed] [Google Scholar]
  • 72.Bindereif A, Green MR. An ordered pathway of snRNP binding during mammalian pre-mRNA splicing complex assembly. EMBO J. 1987;6:2415–2424. doi: 10.1002/j.1460-2075.1987.tb02520.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Jurica MS, Licklider LJ, Gygi SR, Grigorieff N, Moore MJ. Purification and characterization of native spliceosomes suitable for three-dimensional structural analysis. RNA. 2002;8:426–439. doi: 10.1017/s1355838202021088. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Tardiff DF, Rosbash M. Arrested yeast splicing complexes indicate stepwise snRNP recruitment during in vivo spliceosome assembly. RNA. 2006;12:968–979. doi: 10.1261/rna.50506. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Tardiff DF, Lacadie SA, Rosbash M. A genome-wide analysis indicates that yeast pre-mRNA splicing is predominantly posttranscriptional. Mol Cell. 2006;24:917–929. doi: 10.1016/j.molcel.2006.12.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Wahl MC, Will CL, Luhrmann R. The spliceosome: design principles of a dynamic RNP machine. Cell. 2009;136:701–718. doi: 10.1016/j.cell.2009.02.009. [DOI] [PubMed] [Google Scholar]
  • 77.Jobert L, Pinzon N, Van Herreweghe E, Jady BE, Guialis A, Kiss T, et al. Human U1 snRNA forms a new chromatin-associated snRNP with TAF15. EMBO Rep. 2009;10:494–500. doi: 10.1038/embor.2009.24. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Damgaard CK, Kahns S, Lykke-Andersen S, Nielsen AL, Jensen TH, Kjems J. A 5′ splice site enhances the recruitment of basal transcription initiation factors in vivo. Mol Cell. 2008;29:271–278. doi: 10.1016/j.molcel.2007.11.035. [DOI] [PubMed] [Google Scholar]
  • 79.Kwek KY, Murphy S, Furger A, Thomas B, O'Gorman W, Kimura H, et al. U1 snRNA associates with TFIIH and regulates transcriptional initiation. Nat Struct Biol. 2002;9:800–805. doi: 10.1038/nsb862. [DOI] [PubMed] [Google Scholar]
  • 80.O'Gorman W, Thomas B, Kwek KY, Furger A, Akoulitchev A. Analysis of U1 small nuclear RNA interaction with cyclin H. J Biol Chem. 2005;280:36920–36925. doi: 10.1074/jbc.M505791200. [DOI] [PubMed] [Google Scholar]
  • 81.Ares M, Jr, Grate L, Pauling MH. A handful of intron-containing genes produces the lion's share of yeast mRNA. RNA. 1999;5:1138–1139. doi: 10.1017/s1355838299991379. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Furger A, O'Sullivan JM, Binnie A, Lee BA, Proudfoot NJ. Promoter proximal splice sites enhance transcription. Genes Dev. 2002;16:2792–2799. doi: 10.1101/gad.983602. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Eberle AB, Hessle V, Helbig R, Dantoft W, Gimber N, Visa N. Splice-site mutations cause Rrp6-mediated nuclear retention of the unspliced RNAs and transcriptional downregulation of the splicing-defective genes. PLoS One. 2010;5:11540. doi: 10.1371/journal.pone.0011540. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Fong YW, Zhou Q. Stimulatory effect of splicing factors on transcriptional elongation. Nature. 2001;414:929–933. doi: 10.1038/414929a. [DOI] [PubMed] [Google Scholar]
  • 85.Alexander MR, Wheatley AK, Center RJ, Purcell DF. Efficient transcription through an intron requires the binding of an Sm-type U1 snRNP with intact stem loop II to the splice donor. Nucleic Acids Res. 2010;38:3041–3053. doi: 10.1093/nar/gkp1224. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Sims RJ, 3rd, Millhouse S, Chen CF, Lewis BA, Erdjument-Bromage H, Tempst P, et al. Recognition of trimethylated histone H3 lysine 4 facilitates the recruitment of transcription postinitiation factors and pre-mRNA splicing. Mol Cell. 2007;28:665–676. doi: 10.1016/j.molcel.2007.11.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Vermeulen M, Eberl HC, Matarese F, Marks H, Denissov S, Butter F, et al. Quantitative interaction proteomics and genome-wide profiling of epigenetic histone marks and their readers. Cell. 2010;142:967–980. doi: 10.1016/j.cell.2010.08.020. [DOI] [PubMed] [Google Scholar]
  • 88.Luco RF, Pan Q, Tominaga K, Blencowe BJ, Pereira-Smith OM, Misteli T. Regulation of alternative splicing by histone modifications. Science. 2010;327:996–1000. doi: 10.1126/science.1184208. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Shilatifard A. Molecular implementation and physiological roles for histone H3 lysine 4 (H3K4) methylation. Curr Opin Cell Biol. 2008;20:341–348. doi: 10.1016/j.ceb.2008.03.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Dou Y, Milne TA, Ruthenburg AJ, Lee S, Lee JW, Verdine GL, et al. Regulation of MLL1 H3K4 methyltransferase activity by its core components. Nat Struct Mol Biol. 2006;13:713–719. doi: 10.1038/nsmb1128. [DOI] [PubMed] [Google Scholar]
  • 91.Steward MM, Lee JS, O'Donovan A, Wyatt M, Bernstein BE, Shilatifard A. Molecular regulation of H3K4 trimethylation by ASH2L, a shared subunit of MLL complexes. Nat Struct Mol Biol. 2006;13:852–854. doi: 10.1038/nsmb1131. [DOI] [PubMed] [Google Scholar]
  • 92.Bernstein BE, Kamal M, Lindblad-Toh K, Bekiranov S, Bailey DK, Huebert DJ, et al. Genomic maps and comparative analysis of histone modifications in human and mouse. Cell. 2005;120:169–181. doi: 10.1016/j.cell.2005.01.001. [DOI] [PubMed] [Google Scholar]
  • 93.Barski A, Cuddapah S, Cui K, Roh TY, Schones DE, Wang Z, et al. High-resolution profiling of histone methylations in the human genome. Cell. 2007;129:823–837. doi: 10.1016/j.cell.2007.05.009. [DOI] [PubMed] [Google Scholar]
  • 94.Heintzman ND, Stuart RK, Hon G, Fu Y, Ching CW, Hawkins RD, et al. Distinct and predictive chromatin signatures of transcriptional promoters and enhancers in the human genome. Nat Genet. 2007;39:311–318. doi: 10.1038/ng1966. [DOI] [PubMed] [Google Scholar]
  • 95.Guenther MG, Levine SS, Boyer LA, Jaenisch R, Young RA. A chromatin landmark and transcription initiation at most promoters in human cells. Cell. 2007;130:77–88. doi: 10.1016/j.cell.2007.05.042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Mikkelsen TS, Ku M, Jaffe DB, Issac B, Lieberman E, Giannoukos G, et al. Genome-wide maps of chromatin state in pluripotent and lineage-committed cells. Nature. 2007;448:553–560. doi: 10.1038/nature06008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Schneider R, Bannister AJ, Myers FA, Thorne AW, Crane-Robinson C, Kouzarides T. Histone H3 lysine 4 methylation patterns in higher eukaryotic genes. Nat Cell Biol. 2004;6:73–77. doi: 10.1038/ncb1076. [DOI] [PubMed] [Google Scholar]
  • 98.Baker SP, Grant PA. The SAGA continues: expanding the cellular role of a transcriptional co-activator complex. Oncogene. 2007;26:5329–5340. doi: 10.1038/sj.onc.1210603. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Hassan AH, Prochasson P, Neely KE, Galasinski SC, Chandy M, Carrozza MJ, et al. Function and selectivity of bromodomains in anchoring chromatin-modifying complexes to promoter nucleosomes. Cell. 2002;111:369–379. doi: 10.1016/s0092-8674(02)01005-x. [DOI] [PubMed] [Google Scholar]
  • 100.Martinez E, Palhan VB, Tjernberg A, Lymar ES, Gamper AM, Kundu TK, et al. 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. 2001;21:6782–6795. doi: 10.1128/MCB.21.20.6782-6795.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Zhang P, Du J, Sun B, Dong X, Xu G, Zhou J, et al. Structure of human MRG15 chromo domain and its binding to Lys36-methylated histone H3. Nucleic Acids Res. 2006;34:6621–6628. doi: 10.1093/nar/gkl989. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Cai Y, Jin J, Tomomori-Sato C, Sato S, Sorokina I, Parmely TJ, et al. Identification of new subunits of the multiprotein mammalian TRRAP/TIP60-containing histone acetyltransferase complex. J Biol Chem. 2003;278:42733–42736. doi: 10.1074/jbc.C300389200. [DOI] [PubMed] [Google Scholar]
  • 103.Doyon Y, Selleck W, Lane WS, Tan S, Cote J. Structural and functional conservation of the NuA4 histone acetyltransferase complex from yeast to humans. Mol Cell Biol. 2004;24:1884–1896. doi: 10.1128/MCB.24.5.1884-1896.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Pardo PS, Leung JK, Lucchesi JC, Pereira-Smith OM. MRG15, a novel chromodomain protein, is present in two distinct multiprotein complexes involved in transcriptional activation. J Biol Chem. 2002;277:50860–50866. doi: 10.1074/jbc.M203839200. [DOI] [PubMed] [Google Scholar]
  • 105.Yochum GS, Ayer DE. Role for the mortality factors MORF4, MRGX and MRG15 in transcriptional repression via associations with Pf1, mSin3A and Transducin-Like Enhancer of Split. Mol Cell Biol. 2002;22:7868–7876. doi: 10.1128/MCB.22.22.7868-7876.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Jelinic P, Pellegrino J, David G. A novel mammalian complex containing Sin3B mitigates histone acetylation and RNAPII progression within transcribed loci. Mol Cell Biol. 2011 doi: 10.1128/MCB.00840-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Hayakawa T, Ohtani Y, Hayakawa N, Shinmyozu K, Saito M, Ishikawa F, et al. RBP2 is an MRG15 complex component and downregulates intragenic histone H3 lysine 4 methylation. Genes Cells. 2007;12:811–826. doi: 10.1111/j.1365-2443.2007.01089.x. [DOI] [PubMed] [Google Scholar]
  • 108.Lee N, Erdjument-Bromage H, Tempst P, Jones RS, Zhang Y. The H3K4 demethylase lid associates with and inhibits histone deacetylase Rpd3. Mol Cell Biol. 2009;29:1401–1410. doi: 10.1128/MCB.01643-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Govind CK, Qiu H, Ginsburg DS, Ruan C, Hofmeyer K, Hu C, et al. Phosphorylated Pol II CTD recruits multiple HDACs, including Rpd3C(S), for methylation-dependent deacetylation of ORF nucleosomes. Mol Cell. 2010;39:234–246. doi: 10.1016/j.molcel.2010.07.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Joshi AA, Struhl K. Eaf3 chromodomain interaction with methylated H3-K36 links histone deacetylation to Pol II elongation. Mol Cell. 2005;20:971–978. doi: 10.1016/j.molcel.2005.11.021. [DOI] [PubMed] [Google Scholar]
  • 111.Carrozza MJ, Li B, Florens L, Suganuma T, Swanson SK, Lee KK, et al. Histone H3 methylation by Set2 directs deacetylation of coding regions by Rpd3S to suppress spurious intragenic transcription. Cel. 2005;123:581–592. doi: 10.1016/j.cell.2005.10.023. [DOI] [PubMed] [Google Scholar]
  • 112.Vakoc CR, Mandat SA, Olenchock BA, Blobel GA. Histone H3 lysine 9 methylation and HP1gamma are associated with transcription elongation through mammalian chromatin. Mol Cell. 2005;19:381–391. doi: 10.1016/j.molcel.2005.06.011. [DOI] [PubMed] [Google Scholar]
  • 113.Saint-André V, Batsché E, Rachez C, Muchardt C. Histone H3 lysine 9 trimethylation and HP1γ favor inclusion of alternative exons. Nat Struct Mol Biol. 2011;18:337–344. doi: 10.1038/nsmb.1995. [DOI] [PubMed] [Google Scholar]
  • 114.Piacentini L, Fanti L, Negri R, Del Vescovo V, Fatica A, Altieri F, et al. Heterochromatin protein 1 (HP1a) positively regulates euchromatic gene expression through RNA transcript association and interaction with hnRNPs in Drosophila. PLoS Genet. 2009;5:1000670. doi: 10.1371/journal.pgen.1000670. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Huarte M, Guttman M, Feldser D, Garber M, Koziol MJ, Kenzelmann-Broz D, et al. A large intergenic noncoding RNA induced by p53 mediates global gene repression in the p53 response. Cell. 2010;142:409–419. doi: 10.1016/j.cell.2010.06.040. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Xue Y, Zhou Y, Wu T, Zhu T, Ji X, Kwon YS, et al. Genome-wide analysis of PTB-RNA interactions reveals a strategy used by the general splicing repressor to modulate exon inclusion or skipping. Mol Cell. 2009;36:996–1006. doi: 10.1016/j.molcel.2009.12.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Ferguson M, Henry PA, Currie RA. Histone deacetylase inhibition is associated with transcriptional repression of the Hmga2 gene. Nucleic Acids Res. 2003;31:3123–3133. doi: 10.1093/nar/gkg403. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Rustighi A, Tessari MA, Vascotto F, Sgarra R, Giancotti V, Manfioletti G. A polypyrimidine/polypurine tract within the Hmga2 minimal promoter: a common feature of many growth-related genes. Biochemistry. 2002;41:1229–1240. doi: 10.1021/bi011666o. [DOI] [PubMed] [Google Scholar]
  • 119.Tai HH, Geisterfer M, Bell JC, Moniwa M, Davie JR, Boucher L, et al. CHD1 associates with NCoR and histone deacetylase as well as with RNA splicing proteins. Biochem Biophys Res Commun. 2003;308:170–176. doi: 10.1016/s0006-291x(03)01354-8. [DOI] [PubMed] [Google Scholar]
  • 120.Pryde F, Khalili S, Robertson K, Selfridge J, Ritchie AM, Melton DW, et al. 53BP1 exchanges slowly at the sites of DNA damage and appears to require RNA for its association with chromatin. J Cell Sci. 2005;118:2043–2055. doi: 10.1242/jcs.02336. [DOI] [PubMed] [Google Scholar]
  • 121.Young JI, Hong EP, Castle JC, Crespo-Barreto J, Bowman AB, Rose MF, et al. Regulation of RNA splicing by the methylation-dependent transcriptional repressor methyl-CpG binding protein 2. Proc Natl Acad Sci USA. 2005;102:17551–17558. doi: 10.1073/pnas.0507856102. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Nucleus are provided here courtesy of Taylor & Francis

RESOURCES