Pre-mRNA splicing: Role of epigenetics and implications in disease

Pre-mRNA splicing: Role of epigenetics and implications in disease

Advances in Biological Regulation 52 (2012) 377–388 Contents lists available at SciVerse ScienceDirect Advances in Biological Regulation journal hom...

374KB Sizes 0 Downloads 47 Views

Advances in Biological Regulation 52 (2012) 377–388

Contents lists available at SciVerse ScienceDirect

Advances in Biological Regulation journal homepage: www.elsevier.com/locate/jbior

Pre-mRNA splicing: Role of epigenetics and implications in disease Dilshad H. Khan, Sanzida Jahan, James R. Davie* Manitoba Institute of Child Health, University of Manitoba, Winnipeg, Manitoba, R3E 3P4 Canada

a b s t r a c t Epigenetics refer to a variety of processes that have long-term effects on gene expression programs without changes in DNA sequence. Key players in epigenetic control are histone modifications and DNA methylation which, in concert with chromatin remodeling complexes, nuclear architecture and microRNAs, define the chromatin structure of a gene and its transcriptional activity. There is a growing awareness that histone modifications and chromatin organization influence pre-mRNA splicing. Further there is emerging evidence that pre-mRNA splicing itself influences chromatin organization. In the mammalian genome around 95% of multi-exon genes generate alternatively spliced transcripts, the products of which create proteins with different functions. It is now established that several human diseases are a direct consequence of aberrant splicing events. In this review we present the interplay between epigenetic mechanisms and splicing regulation, as well as discuss recent studies on the role of histone deacetylases in splicing activities. Ó 2012 Elsevier Ltd. All rights reserved.

Introduction RNA splicing is a complex regulatory process that plays a major role in gene expression regime and proteome diversity. Splicing occurs co-transcriptionally, while the pre-mRNA remains bound to chromatin by RNA polymerase II (Brody and Shav-Tal, 2011). Over the last few years, much has been learned about the regulatory network of splicing regulation involving transcription mechanisms and

* Corresponding author. Tel.: þ1 204 7872391; fax: þ1 204 7872190. E-mail address: [email protected] (J.R. Davie). 2212-4926/$ – see front matter Ó 2012 Elsevier Ltd. All rights reserved. doi:10.1016/j.jbior.2012.04.003

378

D.H. Khan et al. / Advances in Biological Regulation 52 (2012) 377–388

chromatin modifications. Splicing reactions can be constitutive, where introns are removed and all the exons are retained in the pre-mRNA, or alternative, where exons and introns can be excised or may be present in variable ways in the mature pre-mRNA. Aberrant regulations of splicing mechanisms, whether constitutive or alternative, have been implicated in a number of diseases and cancers. Targeting splicing as a therapeutic intervention is a rapidly moving field of biomedical research. In this review, we will provide a brief overview of splicing and will discuss the involvement of splicing factors, serine/arginine-rich proteins, in diseases and cancers. We will highlight the interplay between epigenetic mechanisms and splicing regulation, as well as review of recent studies highlighting the role of histone deacetylases in splicing activities. A snapshot: regulation of splicing In human cells, approximately 95% of the multi-exon genes yield alternatively spliced transcripts (Hnilicova and Stanek, 2011; Ip et al., 2011). Pre-mRNA splicing is a highly regulated step-wise process catalyzed by a multicomponent machine (the spliceosome), consisting of RNA and protein (Kalsotra and Cooper, 2011; Sperling et al., 2008; Wahl et al., 2009). The spliceosome consists of uridine rich U1, U2, U4, U5, and U6 small nuclear ribonucleoproteins (snRNPs) and a large number of non-snRNP proteins (Rappsilber et al., 2002; Ritchie et al., 2009; Wahl et al., 2009). The spliceosome recognizes the consensus sequences (e.g. 50 and 30 splice sites) at exon–intron boundaries, assembled onto them, and catalyze the splicing reaction (Black, 2003; Wahl et al., 2009). In addition to these, branch point sequence, polypyrimidine tract, cis- (present within pre-mRNA) and trans-regulatory elements (cellular factors, proteins and RNA) are also necessary for proper splicing regulation (Black, 2003; Wahl et al., 2009). Cis-acting elements, also known as splicing enhancers and silencers, play a major role in spliceosome assembly and splice site selection. They are involved in constitutive as well as alternative splicing mechanisms. Trans-acting factors, such as serine/arginine-rich (SR) proteins or hnRNPs (heterogeneous nuclear ribonucleoproteins), can interact with the spliceosome complex, cis-regulatory elements or proteins bound to cis-elements, thereby controlling the splicing outcomes (Black, 2003; Shepard and Hertel, 2009). Apart from these, several studies have reported alternative splicing can be influenced by transcription apparatus and epigenetic mechanisms. Two general models have been proposed to explain how transcription affects splicing: the kinetic model and recruitment model (Kornblihtt et al., 2004; Munoz et al., 2010). According to the kinetic model, variation in the RNA polymerase II elongation rate modulates the splice site recognition efficiency of spliceosome complex, thus determines the inclusion or exclusion of alternative exons (de la Mata et al., 2003). The recruitment model proposes that splicing and transcription are coupled by physical and functional interaction of RNA polymerase II carboxyl-terminal domain (CTD) with the pre-mRNA processing factors, where CTD functions to recruit the factors (Egloff and Murphy, 2008; Kornblihtt et al., 2004). Presumably, both of the models are involved in alteration of splicing outcomes; however the detailed molecular and mechanical basis of the models still needs further investigation. Role of splicing proteins in human disease and cancer Till the early 1980’s it was unclear how the limited number of human genes could give rise to a large variety of proteins. However, this question was resolved with the realization that alternative splicing could generate several different transcripts and different proteins from a single gene. Alternative splicing was first documented with the Adenovirus 2 major late messages (Leff et al., 1986). About twothirds of the human genome contain one or more alternatively spliced exons as observed by microarray profiling and expressed sequence tags-cDNA sequencing (Johnson et al., 2003). Splice variants created through alternative splicing can have opposing functions and thereby can alter the biological outcome (Bae et al., 2000; Bingle et al., 2000). This is quite evident with human BCL-X, MCL1, CASP-9, CED-4, CASP-2/ICH-1 and HTID-1 genes, which encode both pro- and anti-apoptotic variants (Wu et al., 2003). Though alternative splicing is regulated in response to signal transduction pathways, disease state, cell cycle progression and developmental process, it is the splice site sequences and the proteins that bind to these sequences that direct alternative splicing (Barash et al., 2010; Grabowski and Black, 2001; Irimia and Blencowe, 2012; Kalsotra and Cooper, 2011; Martinez et al., 2012; Palusa et al., 2007; Pick

D.H. Khan et al. / Advances in Biological Regulation 52 (2012) 377–388

379

et al., 2004; Wang and Cooper, 2007). The sequential process for splice site recognition and splice variant formation is modulated by several cis- and trans-acting factors (Cartegni et al., 2002; Wang and Cooper, 2007). In human disease, aberrant pre-mRNA splicing may be a consequence of mutations in cis-acting regulatory sequences and/or in the trans-acting factors. Further altered expression of the trans-acting factors can also impact pre-mRNA splicing. Several human diseases such as spinal muscular atrophy, retinitis pigmentosa, Hutchinson–Gilford progeria syndrome, medium-chain acyl-CoA dehydrogenase deficiency and some cancers are a direct consequence of aberrant splicing events due to mutations in the critical cis-acting sequences guiding splicing events (Dredge et al., 2001; Faustino and Cooper, 2003; Tazi et al., 2009; Zaharieva et al., 2012). Genomic mutation and RNA editing are some of the cis-acting mutations leading to changes in splicing of pre-mRNA (Maas, 2012; Maas et al., 2001). Conversion of adenosine to inosine is the common type of RNA editing in mammalian cells. The RNA splicing machinery recognizes the inosine as a guanine, resulting in the formation or elimination of donor or acceptor splice sites (Dominissini et al., 2011). The serine/arginine-rich (SR) family protein and heterogeneous nuclear ribonucleoprotein (hnRNP) family function as splicing enhancers or silencers by acting as trans-acting factors to influence RNA splicing (Maas et al., 2001; Tacke and Manley, 1999). About 12 SR proteins (SRSF3/SRp20, SRSF2/SC35, SRSF8/SRp46, SRSF11/SRp54, SRSF9/SRp30c, SRSF1/ASF/SF2/SRp30a, SRSF5/SRp40, SRSF6/SRp55, SRSF4/SRp75, SRSF7/9G8, SRSF10/SRp38, SRSF12/SRp35) have been identified in human so far (Manley and Krainer, 2010). These trans-acting factors can bind to cis-acting elements and hence play role in modulating exon inclusion either as positive regulator or negative regulator. Among the splicing regulatory proteins, SR proteins bind to 50 splice site or to exonic enhancer to induce splicing. These proteins also bind to sequences in the introns of pre-mRNA where they interfere with spliceosome assembly in order to repress mRNA splicing. This event was evident from SR proteins involvement in inhibition of adenoviral IIIa pre-mRNA splicing by interfering with U2 snRNP recruitment to the 30 splice site (Kanopka et al., 1996). Recent studies report that SF3B1 is mutated in chronic lymphocytic leukemia (mutation occurring in 15% of patients), with the mutation being associated with the aggressive form of chronic lymphocytic leukemia (poorer prognosis) (Wang et al., 2011). The SF3B1 mutation altered splicing (Quesada et al., 2011; Wang et al., 2011). SF3B1 is involved in the splicing of MCL1 transcripts (Moore et al., 2010). A relatively minute change in SR family protein levels can lead to deregulation of alternative splicing event, resulting in changes in cellular behavior and expression levels of various protein isoforms (Ghigna et al., 1998; Mukherji et al., 2006; Pind and Watson, 2003; Stickeler et al., 1999). Induced expression of certain SR proteins is correlated with preneoplasia and intermediate stage of mouse mammary cancer. For instance, expression of specific SR proteins, such as SRSF5, increases during mammary carcinogenesis (Stickeler et al., 1999). This suggests a role for these splicing regulatory proteins in early tumorigenesis. Several studies have documented changes in SR protein levels during tumorigenesis (Ghigna et al., 1998). The mRNA levels of several SR proteins were lower in tumors than in the normal tissue (e.g. mRNA levels of SRSF1, SRSF5, SRSF6, and SRSF4 were lower in nonfamilial colon adenocarcinomas than in normal) (Ghigna et al., 1998). Alterations in SR proteins expression during tumorigenesis may result in changes in splicing. For example, altered expression of SR proteins in mouse mammary adenocarcinomas results in change in the splicing pattern of CD44, a protein involved cell–cell interactions, producing metastatic CD44 variants (Naor et al., 2002; Stickeler et al., 1999). Effect of SR protein alteration in CD44 is of note as several human cancer cells exhibit CD44 splice variants v5, v6 and v7 by exon inclusion, providing a rationale for using it as biomarker for cancer (Jang et al., 2011; Olsson et al., 2011). In another study, overexpression of polypyrimidine tract-binding protein (PTBP) and SRSF3 was correlated with increased splice variants of the multidrug resistance-associated protein 1 (MRP1/ABCC1) as well as CD44 in ovarian tissue as an early event in the ovarian tumorigenesis (He et al., 2004, 2007). Knockdown of either PTBP or SRSF3 inhibited anchorage independent growth and cell growth of ovarian tumor cells. Analyzing the cell cycle progression after reducing SRSF3 levels using siRNA in ovarian cancer cell lines A2780, IGROV1 and SK-OV-3, apoptosis mediated by down regulation of BCL2 was observed (He et al., 2011). Thereby this result indicates the role of SRSF3 in cell growth, cell survival and cell cycle progression, and is considered as a proto-oncogene (Jia et al., 2010). Interestingly where some SR proteins, such as PTBP, showed increased levels during ovarian tumorigenesis others (SRSF1 and

380

D.H. Khan et al. / Advances in Biological Regulation 52 (2012) 377–388

U2AF65) remain unchanged (He et al., 2007). In tissue samples of breast cancer patients increased expression of cell trafficking receptor CD44 variants, CD44v2, CD44v3, CD44v5, and CD44v6 was found in accordance with increased presence of SRSF5 (Huang et al., 2007). Furthermore, in an experimental approach to analyze the expression of different SR proteins in MCF7 and HBL100 breast cell lines, higher level of SR proteins (SRSF4, 5, 6, 9) and altered expression of CD44 variants were found (Pind and Watson, 2003). Though the above mentioned studies link altered levels of certain SR proteins in different cancer cells with altered splicing; however, some SR proteins remained unchanged throughout the tumorigenesis. This raises the question as to the possible mechanisms by which the levels of certain SR proteins change during different stages of tumorigenesis and metastasis. Therefore, it will be important to study the expression profile of all SR proteins and their interacting partners in cancerous and non-cancerous cell in order to decipher their role in disease condition. SR proteins are modified by phosphorylation and acetylation, and these modifications can change the activity of these proteins. Dephosphorylation of SR proteins can modulate the alternative splicing of BCL2L1 (BCL-X) and CASP-9 (Chalfant et al., 2002). Moreover, hyperphosphorylated SRSF10 resulted in the increased G2/M population of SK-OV-3 and BT-474 cell lines (Mukherji et al., 2006). Proteins involved in pre-mRNA splicing are also modified by methylation. The coactivator-associated arginine methyltransferase 1 (CARM1), which methylates H3 at R17 and is a transcriptional coactivator, methylates RNA binding proteins ELAV1/HuR, SNRPB/SmB, SNRPC/U1C and SF3B4/SAP49 (Kuhn et al., 2011). CARM1 has a role in alternative splicing. In the splicing of CD44 pre-mRNA, CARM1 promoted variant exon skipping (Cheng et al., 2007). Understanding the impact of the variety of reversible modifications that the splicing factors undergo on alternative splicing is an area of intense investigation. SRSF1 in disease and cancer SRSF1 is a sequence-specific RNA binding factor that promotes spliceosome formation by binding to exonic splicing enhancers during pre-mRNA splicing (Cho et al., 2011). SRSF1 plays a role in regulating splicing and induction of translation initiation (Michlewski et al., 2008), nonsense mediated mRNA decay (Zhang and Krainer, 2004), mRNA stability and export (Huang et al., 2003; Michlewski et al., 2008; Zhang and Krainer, 2004). Further, SRSF1 has become a crucial factor to be considered as a cancer biomarker as it is found to be associated with tumorigenesis (Ezponda et al., 2010; Karni et al., 2007). SRSF1 is found to be over-expressed in many tumor types, and binds to other oncogenic proteins during tumorigenesis (Pedraza-Farina, 2006). For example, MYC interacts with SRSF1 through its RRM domain (Anczukow et al., 2012). SRSF1 can promote breast cancer and is considered a proto-oncogene (Anczukow et al., 2012; Karni et al., 2007, 2008; Sinclair et al., 2003). Slight changes in the expression of SRSF1 can transform immortal mouse fibroblasts (Karni et al., 2007). Through its function in alternative splicing, SRSF1 plays a role in promoting expression of pro-tumorigenic isoforms of tumor suppressor BIN1 and the kinases Mnk2 and S6K1 in murine immortal fibroblasts (Michlewski et al., 2008). Similarly, increased SRSF1 expression shifts alternative splicing of BCL2L1 (BCL-X), MCL1, and CASP-2 and CASP-9 toward anti-apoptotic forms (Moore et al., 2010). Induction of SRSF1 shifted alternative splicing toward isoforms lacking pro-apoptotic activity, providing evidence that a change in SRSF1 levels can affect several downstream splicing factors and the expression of genes by changing alterative splicing (Anczukow et al., 2012). Thus, overexpression of SRSF1, as observed in many tumor types, results in the altered splicing of several pre-mRNAs, giving rise to proteins that support tumorigenesis and metastasis (Ge et al., 1999; Karni et al., 2007). Epigenetic control of alternative splicing: a two-way communication In recent years, substantial evidences have been accumulated suggesting a major role of epigenetic mechanisms in the alternative splicing regulation. Some of the key players of this regulatory network involve: (1) nucleosome occupancy or positioning and exon–intron architecture, (2) DNA methylation, (3) post-translational modifications (PTM) of histones, (4) chromatin-splicing adaptor complexes, and (5) chromatin remodeling complexes. Nucleosomes and five methyl cytosine (5meC) are nonuniformly distributed along the body of transcribed genes. Nucleosomes and 5meC are enriched on exons relative to introns, and nucleosomes exhibit preferential positioning at exon–intron and intron–

D.H. Khan et al. / Advances in Biological Regulation 52 (2012) 377–388

381

exon boundaries (Chodavarapu et al., 2010; Dhami et al., 2010; Hodges et al., 2009; Schwartz et al., 2009; Shukla and Oberdoerffer, 2012; Spies et al., 2009; Tilgner et al., 2009). The phenomenon of nucleosomal positioning is evolutionary conserved and independent of transcription, GC content or DNA sequences (Andersson et al., 2009; Nahkuri et al., 2009; Schwartz et al., 2009). However, alternatively spliced exons are relatively less enriched in nucleosomes in comparison to the constitutive exons, correlating nucleosome positioning with alternative splicing outcomes (Schwartz et al., 2009). Nucleosomes are suggested to function as ‘speed bumps’ to impose barrier to RNA polymerase II elongation, and thereby affects splicing, which supports the kinetic model of splicing regulation (Carrillo et al., 2011; Schwartz and Ast, 2010). In addition to nucleosomes, genome wide mapping of histone PTMs reveal a non-random distribution pattern with relative enrichment of some specific modifications in exons, attributing their role in splicing (Andersson et al., 2009; Kolasinska-Zwierz et al., 2009; Schwartz et al., 2009; Spies et al., 2009). Trimethylation of H3K36, monomethylation of H3K79, H4K20 and H2BK5 and mono, di, and trimethylation of H3K27 are enriched on exons (Andersson et al., 2009; Dhami et al., 2010; Schwartz et al., 2009; Spies et al., 2009). Some of these modifications have been linked to the splicing regulation via change in chromatin configuration and RNA polymerase II elongation behavior. Histone PTMs can also interact with the splicing factors such as SRSF1 (also called ASF/SF2) and splicing machinery with the help of chromatin-splicing adaptor complexes which act as a scaffold, providing another mechanism of splicing regulation [reviewed in (Luco et al., 2011; Luco and Misteli, 2011)]. Furthermore, chromatin remodeling complexes, such as SWI/SNF (SWItch/Sucrose Non Fermentable) and CHD1 (chromodomain helicase DNA binding protein 1), have been reported to modulate splicing with consequences in splicing efficacy and spliceosome assembly as well as by regulating the rate of RNA polymerase II elongation (Batsche et al., 2006; Sims et al., 2007; Tyagi et al., 2009). These complexes interact with several members of spliceosome complex and splicing proteins and are recruited by specific histone PTMs. For example, trimethylation of histone H3 on Lys4 (H3K4me3), can create a binding site for CHD1 which associates with snRNPs, and facilitates their recruitment (Sims et al., 2007). These studies illustrate the interplay between epigenetic mechanisms and splicing regulation. The MED23 subunit of Mediator interacts with hnRNP L. Mediator is a multiprotein complex that acts as an integrator of various signaling pathways and is a critical component of the RNA polymerase II transcription initiation apparatus (Conaway and Conaway, 2011; Spaeth et al., 2011). MED23 interacts with the splicing machinery and regulates alternative splicing of several hnRNP L targets (Huang et al., 2012). There is emerging evidence that pre-mRNA splicing itself influences chromatin organization (Kim et al., 2011), providing a means of two-way communication between chromatin organization and the splicing regulation. Studies show that splicing mechanisms are required to establish and to maintain epigenetic marks such as trimethylation of H3K36 (de Almeida et al., 2011; Kim et al., 2011). H3K36me3 is considered as a major mark of exons and transcriptional activation (Edmunds et al., 2008; Luco et al., 2010). Genome wide mapping with high throughput sequencing reveals the increased accumulation of H3K36me3 mark toward the 30 end of genes and in intron-containing genes, the enrichment of H3K36me3 mark is higher than intronless genes (Barski et al., 2007; de Almeida et al., 2011; Kolasinska-Zwierz et al., 2009; Schwartz et al., 2009). This observed phenomenon is irrespective of gene sizes, transcription activities and nucleosomal occupancies. Pharmacological inhibition of splicing or siRNA-mediated knockdown of splicing factor (e.g. SAP130) leads to the reduced level of H3K36me3 with consequences of reduced recruitment of histone methyltransferase, HYPB/Setd2 (KMT3A) (de Almeida et al., 2011). However, activation of splicing or change in the splicing patterns (e.g. shift from exon skipping to inclusion), exhibits opposing effects. Furthermore, blockage of splicing by mutating the splice site (e.g. deletion of 30 splice site) or by use of splicing inhibitor such as spliceostatin A cause the relative shift of H3K36me3 mark, which is a generalized observation on thousands of genes containing introns, but not on intronless genes (Kim et al., 2011). Tom Misteli and colleagues reported a connection between H3K36me3 and pre-mRNA splicing (Luco et al., 2010). MRG15, which binds to the polypyrimidine tract-binding protein, binds to H3K36me3. SetD2 (KMT3A), the enzyme catalyzing H3K36me3, binds to the elongating, phosphorylated RNA polymerase II (Sun et al., 2005). Knocking down SetD2 did not impact elongation but

382

D.H. Khan et al. / Advances in Biological Regulation 52 (2012) 377–388

altered pre-mRNA splicing (Edmunds et al., 2008; Luco et al., 2010). MRG15 can form complexes with HDAC2 and KAT5/TIP60, but whether either of these enzymes are co-loaded onto the body of transcribed genes remains to be shown (Doyon et al., 2004). The splicing inhibitor, meayamycin, reduced H3K36me3 levels without altering elongation rates or chromatin-associated RNA (de Almeida et al., 2011). However, in regulation of alternative splicing, the effect of H3K36me3 is not consistent. For neural cell adhesion molecule (NCAM) and fibroblast growth factor receptor 2 (FGFR2) genes, increased accumulation of H3K36me3 has been correlated with exon skipping (Luco et al., 2010; Schor et al., 2009). However, in CD45 and YPEL5 genes, it is reported that H3K36me3 has no role in regulating the splicing outcomes, rather this histone PTM is involved in exon definition (Huff et al., 2010). Reduced levels of H3K36me3 are also observed in alternative exons of these genes than in constitutive ones. This is in contrast with a recent report that alternative exons have higher levels of H3K36me3 relative to constitutive exons (for example, CD44 gene) (de Almeida et al., 2011). However, this may not be a prerequisite or a generalized mechanism for all genes. Depending on cellular contexts or particular splicing events, H3K36me3 may exhibits increased or decreased levels in alternative exons. All of these studies point toward the fact that deposition and maintenance of H3K36me3 is dependent on splicing and localized changes in this epigenetic mark can modulate the splicing activity. However, the detailed mechanisms of regulation still require further validation. HP1g is recruited to the coding region of transcribed genes by the elongating form of RNA polymerase II (Kwon and Workman, 2011a, 2011b). HP1g recruits the histone chaperone complex FACT and has a similar distribution as H3K36me3 (Kwon et al., 2010; Kwon and Workman, 2011a). A recent study demonstrated that H3K9me3 and HP1g, which binds to H3K9me3, was at greater levels on the variant exons of the CD44 gene (Saint-Andre et al., 2011). HP1g played a role in the alternative splicing of CD44 pre-mRNA by slowing down the elongating RNA polymerase II, allowing the inclusion of the variant

A

B

Fig. 1. Histone PTMs alter nucleosome structure facilitating or hindering elongation. A. Nucleosomes with H3K9me3 recruit HP1g. HP1g retards the movement of RNA polymerase II, allowing splicing events for the inclusion of variant exon (E2) to take place. B. HuR inhibits HDAC2 resulting in increased acetylation of nucleosomes. The highly acetylated (Ac) nucleosome will stay in the atypical form established by elongation. The atypical nucleosome will facilitate passage of RNA polymerase II, preventing the formation of spliceosome and the inclusion of alternate exons. In this situation exon 1 (E1) and exon 3 (E3) are spliced together, while alternate exon 2 (E2) is excluded.

D.H. Khan et al. / Advances in Biological Regulation 52 (2012) 377–388

383

exons (Fig. 1). Interestingly HP1g was also bound to the CD44 pre-mRNA in the variant region, thus linking the pre-mRNA to the CD44 chromatin (Saint-Andre et al., 2011). There is a growing list of readers that bridge a gap between pre-mRNA splicing components and histone PTMs along the body of the transcribed gene (Luco et al., 2011). SRSF1 potentially links U1-70K snRNP, involved in early spliceosome assembly, to histone H3 (Hnilicova and Stanek, 2011; Luco et al., 2011). Histone deacetylases: modulator of alternative splicing A role for histone acetylation regulated by lysine acetyltransferases (KATs) and histone deacetylases (HDACs), in RNA splicing and processing is currently gaining momentum, which adds an additional regulatory layer to splicing mechanisms. As such, KATs and HDACs are identified in spliceosome complexes (Rappsilber et al., 2002; Zhou et al., 2002). DEAD-box RNA helicases, for example p68 and p72, that have been shown to be involved in the splicing of pre-mRNAs, are also associated with class I HDACs, HDAC1, 2 and 3 (Mooney et al., 2010; Wilson et al., 2004). Also, an in vitro study revealed that KATs and HDACs are required for spliceosomal rearrangements and assembly (Kuhn et al., 2009). These indicate that HDACs must have a key role in splicing. Multiple studies have demonstrated the association of histone acetylation with alternative splicing regulation (Gunderson et al., 2011; Hnilicova et al., 2011; Kuhn et al., 2009; Schor et al., 2009; Zhou et al., 2011). A recent study by Zhou et al. has suggested that a family of RNA binding protein, Hu proteins, can enhance the localized histone acetylation in regions at the alternative exons of NF1 and FAS genes, involving the functional interaction with the activity of histone deacetylases, in particular, HDAC2 (Zhou et al., 2011). From a mechanistic point of view, Hu proteins are co-transcriptionally recruited to the target genes, directly interact with HDAC2 and inhibit its activity. This localized change in chromatin structure leads to the increase in elongation rate and decreases inclusion of alternative exons in mature mRNA (Fig. 1). During transcription the nucleosome structure is perturbed, forming atypical structures (Czarnota et al., 1997; Locklear et al., 1990). Typically the cysteine residue at position 110 of histone H3 is buried in the interior of the nucleosome. However, the transcribed unfolded nucleosome has its H3 cysteine exposed, offering a tag to isolate and study transcribed nucleosomes (Chen-Cleland et al., 1993; Sun et al., 2002). The atypical nucleosome requires elongation to expose the nucleosome’s cysteinyl-thiol, and histone acetylation will maintain this unfolded structure (Walia et al., 1998). It is conceivable that through histone acetylation maintaining the unfolded nucleosome structure, subsequent rounds of transcription elongation are facilitated (Fig. 1). Changes in the splicing pattern of approximately 700 genes, mostly involved in cell signaling, differentiation and cell cycle regulation, have been identified with HDAC inhibitor treatment by splicing sensitive exon-arrays (Hnilicova et al., 2011). Toward the understanding of the mechanism as to how changes in histone PTMs alter the splicing of several genes, the current literature supports the kinetic model of transcriptional elongation or processivity, where a change in RNA polymerase II elongation rate acts as a sensor to decide the alternative splicing outcomes. For genes with alternative exons that have weak splice sites, a slowly moving RNA polymerase II results in the inclusion of the exons in the RNA. For example, depolarization of neuronal cell membrane induced enhanced acetylation of H3K9 surrounding the exon 18 (E18) of NCAM gene (Schor et al., 2009). However, the increased acetylation status was not detectable in the promoter region of the gene and was localized only in intragenic gene body region. This relaxed chromatin structure leads to an increased rate of RNA polymerase II elongation, preventing the inclusion of this exon in mature mRNA (Fig. 1). Furthermore, removal of depolarization signal can reverse the effect of acetylation and splicing outcomes. Also, the physiological response of membrane depolarization can be functionally mimicked by HDAC inhibitor, TSA treatment, with E18 exon skipping as well as open chromatin structure (Schor et al., 2009). This phenomenon is also observed for other genes such as fibronectin (Hnilicova et al., 2011). HDAC inhibitor mediated alternative splicing of this gene (exon 25 or EDB) has been attributed with increased RNA polymerase II processivity, which is correlated with increased histone H4 acetylation as well as with reduced association of one of the major splicing regulatory protein, SRSF5 (also called SRp40), along the alternative exon of the gene. In addition, siRNAmediated deletion of HDAC1, but not HDAC2, had similar effect as HDAC inhibitor in the splicing

384

D.H. Khan et al. / Advances in Biological Regulation 52 (2012) 377–388

pattern change of this gene, suggesting HDAC1 is primarily involved in the splicing regulation of this gene (Hnilicova et al., 2011). However, it is still not clearly understood how RNA polymerase II slows down at some exons and whether variation in RNA polymerase II kinetics at the alternatively spliced exons occurs at all genes or only in a subset of genes. One possibility is the change in the dynamics of phosphorylation status of RNA polymerase II CTD from elongating Ser-2 phosphorylation to Ser-5 phosphorylation form, facilitates RNA polymerase II to slow down by recruitment of specific phosphatases or kinases (Munoz et al., 2010). However, it is also possible that a change in phosphorylation cycle of RNA polymerase II may not be the major player. Combinatorial effect of chromatin structure, histone PTMs and properties of template DNA sequences, which can act as recruiters for other proteins, may play key roles in determining the elongation rate, which needs to be studied in further detail. Furthermore, it is yet to be known whether change in RNA polymerase kinetics along a gene is a generalized mechanism to determine the splicing outcome in response to HDAC inhibitor treatment. In support of this, recently it has been shown in yeast (Saccharomyces cerevisiae, introncontaining genes, DBP2 and ECM33), the genetic deletion of multiple histone deacetylases Hos3 and Hos2, but not the single deletion of HDAC (either Hos2 or Hos3), caused an enhanced acetylation of H3K9 and H3K14 throughout the body of the genes, with a very slight change in the distribution pattern of RNA polymerase II (Gunderson et al., 2011). However, an aberrant regulation in the cotranscriptional spliceosome complex assembly is detected. Presumably, histone acetylation dynamics influences the dynamics of co-transcriptional spliceosome assembly and can affect splicing without altering the transcriptional elongation rate in yeast, indicating a direct role of chromatin structure in spliceosome complex recruitment. Treatments to manipulate and correct splicing defects Splicing-directed and epigenetic therapies are being actively pursued in animal models and clinical trials to treat diseases that are a consequence of aberrant splicing (Dredge et al., 2001; Osorio et al., 2011; Porensky et al., 2011; Riessland et al., 2010; Sanford et al., 2009; Tazi et al., 2009; Wang et al., 2012; Zaharieva et al., 2012). Morpholino oligonucleotides have been used successfully to alter the splicing of MCL1 (Shieh et al., 2009) and BCL2L1 (Taylor et al., 1999). Morpholino oligonucleotides blocking the 30 acceptor and 50 donor sites of exon 2 of the MCL1 pre-mRNA result increased levels of the MCL-1S transcript, inducing apoptosis in cancer cells (Shieh et al., 2009). An oligonucleotide positioned near the BCL-xL splice donor site, results in a marked increase in the BCL-xS transcript (Taylor et al., 1999). Spinal muscular atrophy is a disease caused by a defect in splicing (Tazi et al., 2009; Ververis and Karagiannis, 2011). Morpholino oligonucleotides have shown promise in treating spinal muscular atrophy (Porensky et al., 2011). There are many HDAC inhibitors in clinical trials (Li and Manley, 2006; Witt et al., 2009). The FDA-approved HDAC inhibitor SAHA (also called Vorinostat and Zolinza) has had success in ameliorating the spinal muscular atrophy phenotype in a mouse model (Riessland et al., 2010). Conclusion Major progress is being made in understanding the dynamic relationship between chromatin organization, transcription elongation and alternative pre-mRNA splicing. In near future, we can look forward to a frontier in the current research field as ‘consanguinity’ of chromatin and pre-mRNA splicing. A comprehensive understanding of the splicing code and interplay between epigenetics and pre-mRNA splicing will provide new insights into the complex network of gene regulations. Further, this research will open doors to develop better therapeutic strategies involving modulation of chromatin and/or splicing to treat human diseases that are a consequence of aberrant splicing. Acknowledgements This work was supported by a grant from the Canadian Institute of Health Research (MOP-9186), a Canada Research Chair to JRD, and a MHRC/CancerCare Manitoba studentship to DK. We acknowledge the strong support of the Manitoba Institute of Child Health and CancerCare Manitoba Foundation.

D.H. Khan et al. / Advances in Biological Regulation 52 (2012) 377–388

385

References de Almeida SF, Grosso AR, Koch F, Fenouil R, Carvalho S, Andrade J, et al. Splicing enhances recruitment of methyltransferase HYPB/Setd2 and methylation of histone H3 Lys36. Nature Structural & Molecular Biology 2011;18:977–83. Anczukow O, Rosenberg AZ, Akerman M, Das S, Zhan L, Karni R, et al. The splicing factor SRSF1 regulates apoptosis and proliferation to promote mammary epithelial cell transformation. Nature Structural & Molecular Biology 2012;19:220–8. Andersson R, Enroth S, Rada-Iglesias A, Wadelius C, Komorowski J. Nucleosomes are well positioned in exons and carry characteristic histone modifications. Genome Research 2009;19:1732–41. Bae J, Leo CP, Hsu SY, Hsueh AJ. MCL-1S, a splicing variant of the antiapoptotic BCL-2 family member MCL-1, encodes a proapoptotic protein possessing only the BH3 domain. Journal of Biological Chemistry 2000;275:25255–61. Barash Y, Calarco JA, Gao W, Pan Q, Wang X, Shai O, et al. Deciphering the splicing code. Nature 2010;465:53–9. 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;129:823–37. Batsche E, Yaniv M, Muchardt C. The human SWI/SNF subunit Brm is a regulator of alternative splicing. Nature Structural & Molecular Biology 2006;13:22–9. Bingle CD, Craig RW, Swales BM, Singleton V, Zhou P, Whyte MK. Exon skipping in Mcl-1 results in a bcl-2 homology domain 3 only gene product that promotes cell death. Journal of Biological Chemistry 2000;275:22136–46. Black DL. Mechanisms of alternative pre-messenger RNA splicing. Annual Review of Biochemistry 2003;72:291–336. Brody Y, Shav-Tal Y. Transcription and splicing: when the twain meet. Transcription 2011;2:216–20. Carrillo OF, Bieberstein N, Neugebauer KM. Pause locally, splice globally. Trends in Cell Biology 2011;21:328–35. Cartegni L, Chew SL, Krainer AR. Listening to silence and understanding nonsense: exonic mutations that affect splicing. Nature Reviews Genetics 2002;3:285–98. Chalfant CE, Rathman K, Pinkerman RL, Wood RE, Obeid LM, Ogretmen B, et al. De novo ceramide regulates the alternative splicing of caspase 9 and Bcl-x in A549 lung adenocarcinoma cells. Dependence on protein phosphatase-1. Journal of Biological Chemistry 2002;277:12587–95. Chen-Cleland TA, Boffa LC, Carpaneto EM, Mariani MR, Valentin E, Mendez E, et al. Recovery of transcriptionally active chromatin restriction fragments by binding to organomercurial-agarose magnetic beads. A rapid and sensitive method for monitoring changes in higher order chromatin structure during gene activation and repression. Journal of Biological Chemistry 1993;268:23409–16. Cheng D, Cote J, Shaaban S, Bedford MT. The arginine methyltransferase CARM1 regulates the coupling of transcription and mRNA processing. Molecular Cell 2007;25:71–83. Cho S, Hoang A, Chakrabarti S, Huynh N, Huang DB, Ghosh G. The SRSF1 linker induces semi-conservative ESE binding by cooperating with the RRMs. Nucleic Acids Research 2011;39:9413–21. Chodavarapu RK, Feng S, Bernatavichute YV, Chen PY, Stroud H, Yu Y, et al. Relationship between nucleosome positioning and DNA methylation. Nature 2010;466:388–92. Conaway RC, Conaway JW. Function and regulation of the mediator complex. Current Opinion in Genetics & Development 2011; 21:225–30. Czarnota GJ, Bazett-Jones DP, Mendez E, Allfrey VG, Ottensmeyer FP. High resolution microanalysis and three-dimensional nucleosome structure associated with transcribing chromatin. Micron 1997;28:419–31. 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 2010;5:e12339. Dominissini D, Moshitch-Moshkovitz S, Amariglio N, Rechavi G. Adenosine-to-inosine RNA editing meets cancer. Carcinogenesis 2011;32:1569–77. Doyon Y, Selleck W, Lane WS, Tan S, Cote J. Structural and functional conservation of the NuA4 histone acetyltransferase complex from yeast to humans. Molecular and Cell Biology 2004;24:1884–96. Dredge BK, Polydorides AD, Darnell RB. The splice of life: alternative splicing and neurological disease. Nature Reviews Neuroscience 2001;2:43–50. Edmunds JW, Mahadevan LC, Clayton AL. Dynamic histone H3 methylation during gene induction: HYPB/Setd2 mediates all H3K36 trimethylation. EMBO Journal 2008;27:406–20. Egloff S, Murphy S. Cracking the RNA polymerase II CTD code. Trends in Genetics 2008;24:280–8. Ezponda T, Pajares MJ, Agorreta J, Echeveste JI, Lopez-Picazo JM, Torre W, et al. The oncoprotein SF2/ASF promotes non-small cell lung cancer survival by enhancing survivin expression. Clinical Cancer Research 2010;16:4113–25. Faustino NA, Cooper TA. Pre-mRNA splicing and human disease. Genes & Development 2003;17:419–37. Ge K, DuHadaway J, Du W, Herlyn M, Rodeck U, Prendergast GC. Mechanism for elimination of a tumor suppressor: aberrant splicing of a brain-specific exon causes loss of function of Bin1 in melanoma. Proceedings of the National Academy of Sciences of the United States of America 1999;96:9689–94. Ghigna C, Moroni M, Porta C, Riva S, Biamonti G. Altered expression of heterogenous nuclear ribonucleoproteins and SR factors in human colon adenocarcinomas. Cancer Research 1998;58:5818–24. Grabowski PJ, Black DL. Alternative RNA splicing in the nervous system. Progress in Neurobiology 2001;65:289–308. Gunderson FQ, Merkhofer EC, Johnson TL. Dynamic histone acetylation is critical for cotranscriptional spliceosome assembly and spliceosomal rearrangements. Proceedings of the National Academy of Sciences of the United States of America 2011; 108:2004–9. He X, Arslan AD, Pool MD, Ho TT, Darcy KM, Coon JS, et al. Knockdown of splicing factor SRp20 causes apoptosis in ovarian cancer cells and its expression is associated with malignancy of epithelial ovarian cancer. Oncogene 2011;30:356–65. He X, Ee PL, Coon JS, Beck WT. Alternative splicing of the multidrug resistance protein 1/ATP binding cassette transporter subfamily gene in ovarian cancer creates functional splice variants and is associated with increased expression of the splicing factors PTB and SRp20. Clinical Cancer Research 2004;10:4652–60. He X, Pool M, Darcy KM, Lim SB, Auersperg N, Coon JS, et al. Knockdown of polypyrimidine tract-binding protein suppresses ovarian tumor cell growth and invasiveness in vitro. Oncogene 2007;26:4961–8.

386

D.H. Khan et al. / Advances in Biological Regulation 52 (2012) 377–388

Hnilicova J, Hozeifi S, Duskova E, Icha J, Tomankova T, Stanek D. Histone deacetylase activity modulates alternative splicing. PLoS ONE 2011;6:e16727. Hnilicova J, Stanek D. Where splicing joins chromatin. Nucleus 2011;2:182–8. Hodges E, Smith AD, Kendall J, Xuan Z, Ravi K, Rooks M, et al. High definition profiling of mammalian DNA methylation by array capture and single molecule bisulfite sequencing. Genome Research 2009;19:1593–605. Huang CS, Shen CY, Wang HW, Wu PE, Cheng CW. Increased expression of SRp40 affecting CD44 splicing is associated with the clinical outcome of lymph node metastasis in human breast cancer. Clinica Chimica Acta 2007;384:69–74. Huang Y, Gattoni R, Stevenin J, Steitz JA. SR splicing factors serve as adapter proteins for TAP-dependent mRNA export. Molecular Cell 2003;11:837–43. Huang Y, Li W, Yao X, Lin QJ, Yin JW, Liang Y, et al. Mediator complex regulates alternative mRNA processing via the MED23 subunit. Molecular Cell 2012;45:459–69. Huff JT, Plocik AM, Guthrie C, Yamamoto KR. Reciprocal intronic and exonic histone modification regions in humans. Nature Structural & Molecular Biology 2010;17:1495–9. 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 Research 2011;21:390–401. Irimia M, Blencowe BJ. Alternative splicing: decoding an expansive regulatory layer. Current Opinion in Cell Biology; 2012. Jang BI, Li Y, Graham DY, Cen P. The role of CD44 in the pathogenesis, diagnosis, and therapy of gastric cancer. Gut and Liver 2011;5:397–405. Jia R, Li C, McCoy JP, Deng CX, Zheng ZM. SRp20 is a proto-oncogene critical for cell proliferation and tumor induction and maintenance. International Journal of Biological Sciences 2010;6:806–26. Johnson JM, Castle J, Garrett-Engele P, Kan Z, Loerch PM, Armour CD, et al. Genome-wide survey of human alternative premRNA splicing with exon junction microarrays. Science 2003;302:2141–4. Kalsotra A, Cooper TA. Functional consequences of developmentally regulated alternative splicing. Nature Reviews Genetics 2011;12:715–29. Kanopka A, Muhlemann O, Akusjarvi G. Inhibition by SR proteins of splicing of a regulated adenovirus pre-mRNA. Nature 1996; 381:535–8. Karni R, de Stanchina E, Lowe SW, Sinha R, Mu D, Krainer AR. The gene encoding the splicing factor SF2/ASF is a proto-oncogene. Nature Structural & Molecular Biology 2007;14:185–93. Karni R, Hippo Y, Lowe SW, Krainer AR. The splicing-factor oncoprotein SF2/ASF activates mTORC1. Proceedings of the National Academy of Sciences of the United States of America 2008;105:15323–7. Kim S, Kim H, Fong N, Erickson B, Bentley DL. Pre-mRNA splicing is a determinant of histone H3K36 methylation. Proceedings of the National Academy of Sciences of the United States of America 2011;108:13564–9. Kolasinska-Zwierz P, Down T, Latorre I, Liu T, Liu XS, Ahringer J. Differential chromatin marking of introns and expressed exons by H3K36me3. Nature Genetics 2009;41:376–81. Kornblihtt AR, de la MM, Fededa JP, Munoz MJ, Nogues G. Multiple links between transcription and splicing. RNA 2004;10: 1489–98. Kuhn AN, van Santen MA, Schwienhorst A, Urlaub H, Luhrmann R. Stalling of spliceosome assembly at distinct stages by smallmolecule inhibitors of protein acetylation and deacetylation. RNA 2009;15:153–75. Kuhn P, Chumanov R, Wang Y, Ge Y, Burgess RR, Xu W. Automethylation of CARM1 allows coupling of transcription and mRNA splicing. Nucleic Acids Research 2011;39:2717–26. Kwon SH, Florens L, Swanson SK, Washburn MP, Abmayr SM, Workman JL. Heterochromatin protein 1 (HP1) connects the FACT histone chaperone complex to the phosphorylated CTD of RNA polymerase II. Genes & Development 2010;24:2133–45. Kwon SH, Workman JL. HP1c casts light on dark matter. Cell Cycle 2011a;10:625–30. Kwon SH, Workman JL. The changing faces of HP1: from heterochromatin formation and gene silencing to euchromatic gene expression: HP1 acts as a positive regulator of transcription. BioEssays 2011b;33:280–9. Leff SE, Rosenfeld MG, Evans RM. Complex transcriptional units: diversity in gene expression by alternative RNA processing. Annual Review of Biochemistry 1986;55:1091–117. Li X, Manley JL. Alternative splicing and control of apoptotic DNA fragmentation. Cell Cycle 2006;5:1286–8. Locklear LJ, Ridsdale JA, Bazett Jones DP, Davie JR. Ultrastructure of transcriptionally competent chromatin. Nucleic Acids Research 1990;18:7015–24. Luco RF, Allo M, Schor IE, Kornblihtt AR, Misteli T. Epigenetics in alternative pre-mRNA splicing. Cell 2011;144:16–26. Luco RF, Misteli T. More than a splicing code: integrating the role of RNA, chromatin and non-coding RNA in alternative splicing regulation. Current Opinion in Genetics & Development 2011;21:366–72. 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. Maas S. Posttranscriptional recoding by RNA editing. Advances in Protein Chemistry and Structural Biology 2012;86: 193–224. Maas S, Patt S, Schrey M, Rich A. Underediting of glutamate receptor GluR-B mRNA in malignant gliomas. Proceedings of the National Academy of Sciences of the United States of America 2001;98:14687–92. Manley JL, Krainer AR. A rational nomenclature for serine/arginine-rich protein splicing factors (SR proteins). Genes & Development 2010;24:1073–4. Martinez NM, Pan Q, Cole BS, Yarosh CA, Babcock GA, Heyd F, et al. Alternative splicing networks regulated by signaling in human T cells. RNA; 2012. 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. Molecular Cell 2003;12:525–32. Michlewski G, Sanford JR, Caceres JF. The splicing factor SF2/ASF regulates translation initiation by enhancing phosphorylation of 4E-BP1. Molecular Cell 2008;30:179–89. Mooney SM, Goel A, D’Assoro AB, Salisbury JL, Janknecht R. Pleiotropic effects of p300-mediated acetylation on p68 and p72 RNA helicase. Journal of Biological Chemistry 2010;285:30443–52.

D.H. Khan et al. / Advances in Biological Regulation 52 (2012) 377–388

387

Moore MJ, Wang Q, Kennedy CJ, Silver PA. An alternative splicing network links cell-cycle control to apoptosis. Cell 2010;142: 625–36. Mukherji M, Brill LM, Ficarro SB, Hampton GM, Schultz PG. A phosphoproteomic analysis of the ErbB2 receptor tyrosine kinase signaling pathways. Biochemistry 2006;45:15529–40. Munoz MJ, de la MM, Kornblihtt AR. The carboxy terminal domain of RNA polymerase II and alternative splicing. Trends in Biochemical Sciences 2010;35:497–504. Nahkuri S, Taft RJ, Mattick JS. Nucleosomes are preferentially positioned at exons in somatic and sperm cells. Cell Cycle 2009;8: 3420–4. Naor D, Nedvetzki S, Golan I, Melnik L, Faitelson Y. CD44 in cancer. Critical Reviews in Clinical Laboratory Sciences 2002;39: 527–79. Olsson E, Honeth G, Bendahl PO, Saal LH, Gruvberger-Saal S, Ringner M, et al. CD44 isoforms are heterogeneously expressed in breast cancer and correlate with tumor subtypes and cancer stem cell markers. BMC Cancer 2011;11:418. Osorio FG, Navarro CL, Cadinanos J, Lopez-Mejia IC, Quiros PM, Bartoli C, et al. Splicing-directed therapy in a new mouse model of human accelerated aging. Science Translational Medicine 2011;3. 106ra107. Palusa SG, Ali GS, Reddy AS. Alternative splicing of pre-mRNAs of Arabidopsis serine/arginine-rich proteins: regulation by hormones and stresses. Plant Journal 2007;49:1091–107. Pedraza-Farina LG. Mechanisms of oncogenic cooperation in cancer initiation and metastasis. Yale Journal of Biology and Medicine 2006;79:95–103. Pick M, Flores-Flores C, Soreq H. From brain to blood: alternative splicing evidence for the cholinergic basis of Mammalian stress responses. Annals of the New York Academy of Sciences 2004;1018:85–98. Pind MT, Watson PH. SR protein expression and CD44 splicing pattern in human breast tumours. Breast Cancer Research and Treatment 2003;79:75–82. Porensky PN, Mitrpant C, McGovern VL, Bevan AK, Foust KD, Kaspar BK, et al. A single administration of morpholino antisense oligomer rescues spinal muscular atrophy in mouse. Human Molecular Genetics; 2011. Quesada V, Conde L, Villamor N, Ordonez GR, Jares P, Bassaganyas L, et al. Exome sequencing identifies recurrent mutations of the splicing factor SF3B1 gene in chronic lymphocytic leukemia. Nature Genetics 2011;44:47–52. Rappsilber J, Ryder U, Lamond AI, Mann M. Large-scale proteomic analysis of the human spliceosome. Genome Research 2002; 12:1231–45. Riessland M, Ackermann B, Forster A, Jakubik M, Hauke J, Garbes L, et al. SAHA ameliorates the SMA phenotype in two mouse models for spinal muscular atrophy. Human Molecular Genetics 2010;19:1492–506. Ritchie DB, Schellenberg MJ, MacMillan AM. Spliceosome structure: piece by piece. Biochimica et Biophysica Acta 2009;1789: 624–33. Saint-Andre V, Batsche E, Rachez C, Muchardt C. Histone H3 lysine 9 trimethylation and HP1gamma favor inclusion of alternative exons. Nature Structural & Molecular Biology 2011;18:337–44. Sanford JR, Wang X, Mort M, Vanduyn N, Cooper DN, Mooney SD, et al. Splicing factor SFRS1 recognizes a functionally diverse landscape of RNA transcripts. Genome Research 2009;19:381–94. Schor IE, Rascovan N, Pelisch F, Allo M, Kornblihtt AR. Neuronal cell depolarization induces intragenic chromatin modifications affecting NCAM alternative splicing. Proceedings of the National Academy of Sciences of the United States of America 2009; 106:4325–30. Schwartz S, Ast G. Chromatin density and splicing destiny: on the cross-talk between chromatin structure and splicing. EMBO Journal 2010;29:1629–36. Schwartz S, Meshorer E, Ast G. Chromatin organization marks exon-intron structure. Nature Structural & Molecular Biology 2009;16:990–5. Shepard PJ, Hertel KJ. The SR protein family. Genome Biology 2009;10:242. Shieh JJ, Liu KT, Huang SW, Chen YJ, Hsieh TY. Modification of alternative splicing of Mcl-1 pre-mRNA using antisense morpholino oligonucleotides induces apoptosis in basal cell carcinoma cells. Journal of Investigative Dermatology 2009;129: 2497–506. Shukla S, Oberdoerffer S. Co-transcriptional regulation of alternative pre-mRNA splicing. Biochimica et Biophysica Acta; 2012. Sims III RJ, 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. Molecular Cell 2007;28: 665–76. Sinclair CS, Rowley M, Naderi A, Couch FJ. The 17q23 amplicon and breast cancer. Breast Cancer Research and Treatment 2003; 78:313–22. Spaeth JM, Kim NH, Boyer TG. Mediator and human disease. Seminars in Cell & Developmental Biology 2011;22:776–87. Sperling J, Azubel M, Sperling R. Structure and function of the Pre-mRNA splicing machine. Structure 2008;16:1605–15. Spies N, Nielsen CB, Padgett RA, Burge CB. Biased chromatin signatures around polyadenylation sites and exons. Molecular Cell 2009;36:245–54. Stickeler E, Kittrell F, Medina D, Berget SM. Stage-specific changes in SR splicing factors and alternative splicing in mammary tumorigenesis. Oncogene 1999;18:3574–82. Sun J-M, Chen HY, Davie JR. Isolation of transcriptionally active chromatin from human breast cancer cells using sulfolink coupling gel chromatography. Journal of Cellular Biochemistry 2002;84:439–46. Sun XJ, Wei J, Wu XY, Hu M, Wang L, Wang HH, et al. Identification and characterization of a novel human histone H3 lysine 36-specific methyltransferase. Journal of Biological Chemistry 2005;280:35261–71. Tacke R, Manley JL. Functions of SR and Tra2 proteins in pre-mRNA splicing regulation. Proceedings of the Society for Experimental Biology and Medicine 1999;220:59–63. Taylor JK, Zhang QQ, Wyatt JR, Dean NM. Induction of endogenous Bcl-xS through the control of Bcl-x pre-mRNA splicing by antisense oligonucleotides. Nature Biotechnology 1999;17:1097–100. Tazi J, Bakkour N, Stamm S. Alternative splicing and disease. Biochimica et Biophysica Acta 2009;1792:14–26.

388

D.H. Khan et al. / Advances in Biological Regulation 52 (2012) 377–388

Tilgner H, Nikolaou C, Althammer S, Sammeth M, Beato M, Valcarcel J, et al. Nucleosome positioning as a determinant of exon recognition. Nature Structural & Molecular Biology 2009;16:996–1001. Tyagi A, Ryme J, Brodin D, Ostlund Farrants AK, Visa N. SWI/SNF associates with nascent pre-mRNPs and regulates alternative pre-mRNA processing. PLoS Genetics 2009;5:e1000470. Ververis K, Karagiannis TC. Potential non-oncological applications of histone deacetylase inhibitors. American Journal of Translational Research 2011;3:454–67. Wahl MC, Will CL, Luhrmann R. The spliceosome: design principles of a dynamic RNP machine. Cell 2009;136:701–18. Walia H, Chen HY, Sun J-M, Holth LT, Davie JR. Histone acetylation is required to maintain the unfolded nucleosome structure associated with transcribing DNA. Journal of Biological Chemistry 1998;273:14516–22. Wang GS, Cooper TA. Splicing in disease: disruption of the splicing code and the decoding machinery. Nature Reviews Genetics 2007;8:749–61. Wang J, Zhang J, Li K, Zhao W, Cui Q. SpliceDisease database: linking RNA splicing and disease. Nucleic Acids Research 2012;40: D1055–9. Wang L, Lawrence MS, Wan Y, Stojanov P, Sougnez C, Stevenson K, et al. SF3B1 and other novel cancer genes in chronic lymphocytic leukemia. The New England Journal of Medicine 2011;365:2497–506. Wilson BJ, Bates GJ, Nicol SM, Gregory DJ, Perkins ND, Fuller-Pace FV. The p68 and p72 DEAD box RNA helicases interact with HDAC1 and repress transcription in a promoter-specific manner. BMC Molecular Biology 2004;5:11. Witt O, Deubzer HE, Milde T, Oehme I. HDAC family: what are the cancer relevant targets? Cancer Letters 2009;277:8–21. Wu JY, Tang H, Havlioglu N. Alternative pre-mRNA splicing and regulation of programmed cell death. Progress in Molecular and Subcellular Biology 2003;31:153–85. Zaharieva E, Chipman K, Soller M. Alternative splicing interference by xenobiotics. Toxicology; 2012. Zhang Z, Krainer AR. Involvement of SR proteins in mRNA surveillance. Molecular Cell 2004;16:597–607. Zhou HL, Hinman MN, Barron VA, Geng C, Zhou G, Luo G, et al. Hu proteins regulate alternative splicing by inducing localized histone hyperacetylation in an RNA-dependent manner. Proceedings of the National Academy of Sciences of the United States of America 2011;108:E627–35. Zhou Z, Licklider LJ, Gygi SP, Reed R. Comprehensive proteomic analysis of the human spliceosome. Nature 2002;419:182–5.