the sr protein family of splicing factors master regulators

13
Biochem. J. (2009) 417, 15–27 (Printed in Great Britain) doi:10.1042/BJ20081501 15 REVIEW ARTICLE The SR protein family of splicing factors: master regulators of gene expression Jennifer C. LONG and Javier F. CACERES 1 Medical Research Council Human Genetics Unit, Institute of Genetics and Molecular Medicine, Western General Hospital, Edinburgh EH4 2XU, U.K. The SR protein family comprises a number of phylogenetically conserved and structurally related proteins with a characteristic domain rich in arginine and serine residues, known as the RS do- main. They play significant roles in constitutive pre-mRNA spli- cing and are also important regulators of alternative splicing. In addition they participate in post-splicing activities, such as mRNA nuclear export, nonsense-mediated mRNA decay and mRNA trans- lation. These wide-ranging roles of SR proteins highlight their importance as pivotal regulators of mRNA metabolism, and if these functions are disrupted, developmental defects or disease may result. Furthermore, animal models have shown a highly specific, non-redundant role for individual SR proteins in the regu- lation of developmental processes. Here, we will review the current literature to demonstrate how SR proteins are emerging as one of the master regulators of gene expression. Key words: alternative splicing, pre-mRNA splicing, RS domain, SR proteins, SR-related proteins, translation regulation. INTRODUCTION Pre-mRNA splicing was discovered in the late 1970s when it was demonstrated that eukaryotic genes contained intervening sequences, or introns, that were not present in the mature mRNA [1,2]. Subsequent studies showed that introns are removed by a macromolecular complex, termed the spliceosome, which consists of five snRNPs [small nuclear RNPs (ribonucleoprotein particles)], U1, U2, U4, U5 and U6, and a large number of protein components (reviewed in [3,4]). Spliceosomal assembly is initiated by the recognition of the 5 and 3 ss (splice sites) by the U1 snRNP and the heterodimeric U2AF (U2 snRNP auxiliary factor) respectively, forming the E complex. Recruitment of the U2 snRNP to the BP (branch-point), in an ATP-dependent manner, results in the formation of the A complex. Subsequent recruitment of the U4/U6 · U5 tri-snRNP forms the B complex, which is followed by a series of structural rearrangements leading to the formation of the catalytically active spliceosomal C complex (reviewed in [5]). The spliceosome is a dynamic structure and more than 300 proteins have been identified in active splicing complexes [6–8] (reviewed in [9]). The aim of this article is to review the contribution of SR protein family members to pre-mRNA splicing, as well as reviewing more recent studies expanding their role in post-splicing activities. Finally, we will discuss how misregulation of SR protein functions can lead to human disease. THE SR PROTEIN FAMILY The SR proteins were first discovered as splicing factors in the early 1990s (reviewed in [10–12]). A protein domain rich in arginine and serine dipeptides, termed the RS domain, was originally observed in three Drosophila splicing regulators, SWAP (suppressor-of-white-apricot) [13], Tra (transformer) [14] and Tra-2 (transformer-2) [15]. Subsequent identification of SF2/ASF (splicing factor 2/alternative splicing factor) [16,17] and SC35 (spliceosomal component 35) [18] revealed that these proteins contained an RS domain, which is also present in the U1 snRNP- associated protein, U1-70K [19,20]. SF2/ASF was the first SR protein to be identified as an activity required to complement an otherwise splicing-deficient HeLa (human cervical carcinoma cell) S100 extract [21] and was also purified from HEK (human embryonic kidney)-293 cells as a factor which could alter 5 ss selection of an SV40 (simian virus 40) early pre-mRNA [22]. The term ‘SR protein’ was coined following identification of additional RS domain-containing pro- teins that were recognized by a monoclonal antibody, mAb 104, which binds to active sites of RNA polymerase II transcription [23]. These novel proteins, which were active in splicing comple- mentation, included the SR proteins SRp20, SRp40, SRp55 and SRp75, named after their apparent molecular mass on an SDS/ PAGE gel, and are conserved across vertebrates and invertebrates [24]. They have a modular structure containing one or two copies of an RRM (RNA recognition motif) at the N-terminus that provides RNA-binding specificity and a C-terminal RS domain that acts to promote protein–protein interactions that facilitate recruitment of the spliceosome [25,26]. The RS domain can also contact the pre-mRNA directly via the BP and the 5 ss, suggesting an alternative way to promote spliceosome assembly [27,28] (reviewed in [29]). Furthermore, the RS domain acts as an NLS (nuclear localization signal), affecting the subcellular localization of SR proteins by mediating the interaction with the SR protein nuclear import receptor, transportin-SR [30–32]. The prototypical SR protein, SF2/ASF, functions in constitutive splicing and also modulates alternative splicing [22,33]. Further studies demonstrated that other SR protein family members Abbreviations used: BP, branch-point; CLIP, cross-linking and immunoprecipitation; CTD, C-terminal domain; Dscam, Down’s syndrome cell adhesion molecule; E1αPDH, E1α pyruvate dehydrogenase; eIF, eukaryotic initiation factor; ESE, exonic splicing enhancer; ESS, exonic splicing silencer; hnRNP, heterogeneous nuclear RNP; ISS, intronic splicing silencer; mTOR, mammalian target of rapamycin; NMD, non-sense-mediated decay; PESE, putative exonic splicing enhancer; RESCUE, relative enhancer and silencer classification by unanimous enrichment; RNAi, RNA interference; RNAP II, RNA polymerase II; RNP, ribonucleoprotein particle; RRM, RNA recognition motif; S6K1, S6 kinase; SC35, spliceosomal component 35; SELEX, selected evolution of ligands through exponential enrichment; SF2/ASF, splicing factor 2/alternative splicing factor; SMA, spinal muscular atrophy; SMN, survival of motor neuron; snRNP, small nuclear RNP; 3 /5 ss, 3 /5 splice site; Tra, transformer ; U2AF, U2 snRNP auxiliary factor. 1 To whom correspondence should be addressed (email [email protected]). c The Authors Journal compilation c 2009 Biochemical Society www.biochemj.org Biochemical Journal

Upload: saeedmoto

Post on 28-Nov-2014

95 views

Category:

Documents


2 download

TRANSCRIPT

Page 1: The SR Protein Family of Splicing Factors Master Regulators

Biochem. J. (2009) 417, 15–27 (Printed in Great Britain) doi:10.1042/BJ20081501 15

REVIEW ARTICLEThe SR protein family of splicing factors: master regulatorsof gene expressionJennifer C. LONG and Javier F. CACERES1

Medical Research Council Human Genetics Unit, Institute of Genetics and Molecular Medicine, Western General Hospital, Edinburgh EH4 2XU, U.K.

The SR protein family comprises a number of phylogeneticallyconserved and structurally related proteins with a characteristicdomain rich in arginine and serine residues, known as the RS do-main. They play significant roles in constitutive pre-mRNA spli-cing and are also important regulators of alternative splicing. Inaddition they participate in post-splicing activities, such as mRNAnuclearexport,nonsense-mediatedmRNAdecayandmRNAtrans-lation. These wide-ranging roles of SR proteins highlight theirimportance as pivotal regulators of mRNA metabolism, and if

these functions are disrupted, developmental defects or diseasemay result. Furthermore, animal models have shown a highlyspecific, non-redundant role for individual SR proteins in the regu-lation of developmental processes. Here, we will review thecurrent literature to demonstrate how SR proteins are emergingas one of the master regulators of gene expression.

Key words: alternative splicing, pre-mRNA splicing, RS domain,SR proteins, SR-related proteins, translation regulation.

INTRODUCTION

Pre-mRNA splicing was discovered in the late 1970s when itwas demonstrated that eukaryotic genes contained interveningsequences, or introns, that were not present in the maturemRNA [1,2]. Subsequent studies showed that introns are removedby a macromolecular complex, termed the spliceosome, whichconsists of five snRNPs [small nuclear RNPs (ribonucleoproteinparticles)], U1, U2, U4, U5 and U6, and a large number ofprotein components (reviewed in [3,4]). Spliceosomal assemblyis initiated by the recognition of the 5′ and 3′ ss (splice sites) bythe U1 snRNP and the heterodimeric U2AF (U2 snRNP auxiliaryfactor) respectively, forming the E complex. Recruitment of theU2 snRNP to the BP (branch-point), in an ATP-dependent manner,results in the formation of the A complex. Subsequent recruitmentof the U4/U6 · U5 tri-snRNP forms the B complex, which isfollowed by a series of structural rearrangements leading to theformation of the catalytically active spliceosomal C complex(reviewed in [5]). The spliceosome is a dynamic structure andmore than 300 proteins have been identified in active splicingcomplexes [6–8] (reviewed in [9]).

The aim of this article is to review the contribution of SR proteinfamily members to pre-mRNA splicing, as well as reviewingmore recent studies expanding their role in post-splicing activities.Finally, we will discuss how misregulation of SR protein functionscan lead to human disease.

THE SR PROTEIN FAMILY

The SR proteins were first discovered as splicing factors in theearly 1990s (reviewed in [10–12]). A protein domain rich inarginine and serine dipeptides, termed the RS domain, wasoriginally observed in three Drosophila splicing regulators, SWAP

(suppressor-of-white-apricot) [13], Tra (transformer) [14] andTra-2 (transformer-2) [15]. Subsequent identification of SF2/ASF(splicing factor 2/alternative splicing factor) [16,17] and SC35(spliceosomal component 35) [18] revealed that these proteinscontained an RS domain, which is also present in the U1 snRNP-associated protein, U1-70K [19,20].

SF2/ASF was the first SR protein to be identified as an activityrequired to complement an otherwise splicing-deficient HeLa(human cervical carcinoma cell) S100 extract [21] and was alsopurified from HEK (human embryonic kidney)-293 cells as afactor which could alter 5′ ss selection of an SV40 (simian virus40) early pre-mRNA [22]. The term ‘SR protein’ was coinedfollowing identification of additional RS domain-containing pro-teins that were recognized by a monoclonal antibody, mAb 104,which binds to active sites of RNA polymerase II transcription[23]. These novel proteins, which were active in splicing comple-mentation, included the SR proteins SRp20, SRp40, SRp55 andSRp75, named after their apparent molecular mass on an SDS/PAGE gel, and are conserved across vertebrates and invertebrates[24]. They have a modular structure containing one or two copiesof an RRM (RNA recognition motif) at the N-terminus thatprovides RNA-binding specificity and a C-terminal RS domainthat acts to promote protein–protein interactions that facilitaterecruitment of the spliceosome [25,26]. The RS domain canalso contact the pre-mRNA directly via the BP and the 5′ ss,suggesting an alternative way to promote spliceosome assembly[27,28] (reviewed in [29]). Furthermore, the RS domain acts asan NLS (nuclear localization signal), affecting the subcellularlocalization of SR proteins by mediating the interaction with theSR protein nuclear import receptor, transportin-SR [30–32].

The prototypical SR protein, SF2/ASF, functions in constitutivesplicing and also modulates alternative splicing [22,33]. Furtherstudies demonstrated that other SR protein family members

Abbreviations used: BP, branch-point; CLIP, cross-linking and immunoprecipitation; CTD, C-terminal domain; Dscam, Down’s syndrome cell adhesionmolecule; E1αPDH, E1α pyruvate dehydrogenase; eIF, eukaryotic initiation factor; ESE, exonic splicing enhancer; ESS, exonic splicing silencer; hnRNP,heterogeneous nuclear RNP; ISS, intronic splicing silencer; mTOR, mammalian target of rapamycin; NMD, non-sense-mediated decay; PESE, putativeexonic splicing enhancer; RESCUE, relative enhancer and silencer classification by unanimous enrichment; RNAi, RNA interference; RNAP II, RNApolymerase II; RNP, ribonucleoprotein particle; RRM, RNA recognition motif; S6K1, S6 kinase; SC35, spliceosomal component 35; SELEX, selectedevolution of ligands through exponential enrichment; SF2/ASF, splicing factor 2/alternative splicing factor; SMA, spinal muscular atrophy; SMN, survival ofmotor neuron; snRNP, small nuclear RNP; 3′/5′ ss, 3′/5′ splice site; Tra, transformer; U2AF, U2 snRNP auxiliary factor.

1 To whom correspondence should be addressed (email [email protected]).

c© The Authors Journal compilation c© 2009 Biochemical Society

www.biochemj.org

Bio

chem

ical

Jo

urn

al

Page 2: The SR Protein Family of Splicing Factors Master Regulators

16 J. C. Long and J. F. Caceres

Table 1 ‘Classical’ SR proteins

Protein name Gene name Key domains Splicing role UniProt

SF2/ASF SFRS1 RRM × 2, RS Constitutive and alternative splicing activator Q07955SC35 SFRS2 RRM, RS Constitutive and alternative splicing activator Q01130SRp20 SFRS3 RRM, RS Constitutive and alternative splicing activator P84103SRp75 SFRS4 RRM × 2, RS Constitutive and alternative splicing activator Q08170SRp40 SFRS5 RRM × 2, RS Constitutive and alternative splicing activator Q13243SRp55 SFRS6 RRM × 2, RS Constitutive and alternative splicing activator Q132479G8 SFRS7 RRM, RS, CCHC-type zinc finger Constitutive and alternative splicing activator Q16629

Table 2 Additional SR proteins

Protein name Gene name Key domains Splicing role UniProt

p54 SFRS11 RRM, RS Alternative splicing repressor Q05519SRp30c SFRS9 RRM × 2, RS Constitutive and alternative splicing regulator Q13242SRp38, TASR FUSIP1 RRM, RS General splicing repressor O75494hTra2α TRA2A RRM, RS × 2 Splicing activator Q13595hTra2β SFRS10 RRM, RS × 2 Splicing activator P62995RNPS1 RNPS1 RRM, RS Constitutive and alternative splicing regulator Q15287SRrp35 SRRP35 RRM, RS Negative regulator of alternative splicing Q8WXF0SRrp86, SRrp508 SFRS12 RRM, RS Positive and negative regulator of alternative splicing Q8WXA9U2AF35 U2AF1 RRM, RS, C3H1-type zinc finger × 2 Constitutive splicing factor Q01081U2AF65 U2AF2 RRM × 3, RS Constitutive splicing factor P26368U1-70K SNRP70 RRM, RS Constitutive splicing factor P08621XE7 SFRS17A RRM, RS Alternative splicing regulator Q02040SRp46 SFRS2B RRM, RS Constitutive and alternative splicing regulator Q9BRL6

Table 3 RNA-binding SR-related factors

Protein name Gene name Key domains Splicing role UniProt

Urp ZRSR2 RRM, RS Splicing factor Q15696HCC1/CAPER RBM39 RRM, RS Alternative splicing regulator Q14498hSWAP SFRS16 RS Alternative splicing regulator Q8N2M8Pinin PNN RS Alternative splicing regulator Q9H307SRrp129 SFRS2IP RS Splicing factor Q99590U4/U6 · U5 tri-snRNP-associated 27 kDa protein RY-1 RS Unknown Q8WVK2LUC7B1 LUC7L RS, C2H2-type zinc finger Unknown Q9NQ29Acinus ACIN1 RRM, RS, SAP Unknown Q9UKV3SR-A1 SFRS19/ SCAF1 RS Unknown Q9H7N4ZNF265 ZRANB2 RS, RANBP2-type zinc finger × 2 Alternative splicing regulator O95218SRm160 SRRM1 RS, PWI Constitutive and alternative splicing co-activator Q8IYB3SRm300 SRRM2 RS Constitutive and alternative splicing co-activator Q9UQ35RBM5 RBM5 RRM × 2, RS, RANBP2- and C2H2-type zinc fingers Unknown P52756U2-associated protein SR140 SR140 RRM, RS Unknown O15042RBM23 RBM23 RRM × 2, RS Unknown Q86U06SFRS15 SFRS15 RRM, RS Unknown O95104

could also affect alternative splicing in vitro [34,35]. Thus, thecriteria used to define ‘classical’ SR protein family membersare (i) structural similarity, (ii) dual function in constitutive andalternative splicing, (iii) the presence of a phosphoepitope recog-nised by mAb104; and (iv) their purification using magnesiumchloride (Table 1).

A genome-wide survey in metazoans identified a large numberof RS domain-containing proteins with a role not only in splicingbut also in other cellular processes such as chromatin remodelling,transcription and cell cycle progression [36]. These relatedproteins contain an RS domain but may lack a defined RRM,however a subset can bind RNA through other domains such asthe PWI motif found in the splicing activator SRm160 [37,38](Tables 2–4). These factors are collectively known as SR-likeor SR-related proteins and include both subunits of the U2AF

heterodimer, U1-70K and the splicing coactivators SRm 160/300,among others [39]. It was recently proposed that SR proteinsshould be redefined based on their common structural featuresand their role in pre-mRNA splicing [40]. Based on this, a ‘bonafide’ SR protein has to contain at least one RRM and an RSdomain (irrespective of their positions within the protein) and tofunction in constitutive or alternative splicing, as assayed by eithercomplementation of splicing-deficient S100 HeLa cytoplasmicextracts or in an alternative splicing assay respectively. The humanhomologues of the Drosophila splicing regulators Tra2α [41]and Tra2β [42] contain an RRM flanked by two RS domains,which is not the domain structure found in classical SR proteins(Figure 1). Since both proteins function as sequence-specificsplicing activators [43], they could be classified as SR proteins(Table 2). Other proteins may contain an RS domain but also have

c© The Authors Journal compilation c© 2009 Biochemical Society

Page 3: The SR Protein Family of Splicing Factors Master Regulators

Functional roles of SR proteins in RNA processing 17

Table 4 Other RS-domain containing proteins

Protein name Gene name Key domains Splicing role UniProt

SRrp53 RSRC1 RS, coiled-coil domain Unknown Q96IZ7hPRP5 DDX46 RS, DEAH box Spliceosomal rearrangement Q7L014hPRP16 DHX38 RS, DEAH box Splicing factor Q92620Prp22/HRH1 DHX8 RS, DEAH box Spliceosomal rearrangement Q14562U5-100k/hPRP28 DDX23 RS, DEAD box Spliceosomal rearrangement Q9BUQ8ClkSty-1 CLK1 RS, kinase domain SR protein kinase P49759ClkSty-2 CLK2 RS, kinase domain SR protein kinase P49760ClkSty-3 CLK3 RS, kinase domain SR protein kinase P49761Prp4k PRPF4B RS, kinase domain SR protein kinase Q13523CrkRS CRKRS RS, kinase domain SR protein kinase Q9NYV4CDC2L5 CDC2L5 RS, kinase domain Alternative splicing regulator Q14004Cyclin-L1 CCNL1 RS, cyclin-like domain × 2 Alternative splicing regulator Q9UK58Cyclin-L2 CCNL2 RS, cyclin-like domain × 2 Alternative splicing regulator Q96S94SR-cyp PPIG RS, PPIase cyclophilin-type domain Regulates localisation of SR proteins Q13427CIR CIR RS Alternative splicing regulator Q86X95SRrp130 SFRS18 RS × 2 Unknown Q8TF01

Figure 1 Schematic diagram of SR and SR-related proteins

The domain structures are depicted. DEAH Box, motif characteristic of RNA helicases; RS: arginine/serine-rich domain; PWI: an alternative RNA binding motif; Zn, zinc finger motif. With the exceptionof SRm160 and hPRP5, all proteins are drawn to scale.

other domains required for their enzymatic activities, as is thecase for the RNA helicases HRH1 and hPRP16, that contain aDEAH box domain [44,45] (Table 4).

SR PROTEINS AND TRANSCRIPTION

SR proteins are concentrated in nuclear speckles and are recruitedfrom these sites to nascent sites of RNAP II (RNA polymerase II)transcription [46]. It is well documented that RNA splicing occursco-transcriptionally [47,48]. Interactions between SR-relatedproteins and the CTD (C-terminal domain) of RNAP II havebeen reported [49], and members of the SR protein family wereidentified among the hundreds of proteins present in the RNAP II

complex [50]. It was recently reported that SC35 promotes RNAPII elongation in a subset of genes, confirming the existenceof coupling between transcription and splicing, and perhapssurprisingly, showing that this coupling can be bidirectional [51](reviewed in [52]). In this study [51] it was demonstrated thatSC35 interacts not only with the CTD but also with CDK9(cyclin-dependent kinase 9), which is the kinase component of thetranscriptional elongation factor P-TEFb (positive transcriptionelongation factor b), resulting in phosphorylation of Ser2 in theCTD and leading to transcriptional elongation. This activity ofSR proteins in transcriptional elongation may be functionallyrelated to their reported effect in the maintenance of genomestability. It has been shown that depletion of SF2/ASF, SC35and the SR-related protein RNSP1 results in the formation of

c© The Authors Journal compilation c© 2009 Biochemical Society

Page 4: The SR Protein Family of Splicing Factors Master Regulators

18 J. C. Long and J. F. Caceres

R-loops (RNA:DNA hybrid structures) leading to a hypermutationphenotype [53–55].

The co-transcriptional nature of pre-mRNA splicing underliesa role for the transcriptional machinery in alternative splicingregulation [56]. A kinetic coupling model proposed that changesin the rate of transcriptional elongation affect the timing in whichsplice sites are presented to the splicing machinery, leading todifferential splice site selection [57]. Furthermore, differentialrecruitment of splicing factors to the CTD of RNAP II may alsoinfluence this process [58] (reviewed in [59]).

ROLES OF CLASSICAL SR PROTEINS IN CONSTITUTIVE ANDALTERNATIVE SPLICING

Splice site consensus sequences are generally not sufficient todirect assembly of a functional spliceosome, and auxiliary ele-ments known as ESEs and ISEs (exonic and intronic splicingenhancers respectively) and ESSs and ISSs (exonic and intronicsilencers respectively) are involved in both constitutive and, toperhaps a greater extent, alternative splicing. Binding of SRproteins to ESEs acts as a barrier that prevents exon skipping,thus ensuring the correct 5′ to 3′ linear order of exons in splicedmRNA [60]. Two main models have been proposed to explainthe mechanism by which SR proteins regulate exon inclusion. The‘recruitment model’ focuses on the ability of ESE-bound SRproteins to recruit and stabilize interactions between the U1snRNP at the 5′ ss and U2AF65 at the 3′ ss [61], in a process knownas exon definition [62] (Figure 2A). The hnRNP (heterogeneousnuclear RNP) family comprises a structurally diverse group ofRNA-binding proteins with roles in many aspects of RNA bio-genesis, including pre-mRNA splicing (reviewed in [63]). Inthe ‘inhibitor model’, ESE-bound SR proteins may act byantagonizing the negative activity of hnRNP proteins recognizingESSs [64] (Figure 2B). The SR proteins may also form a networkof protein–protein interactions across introns to juxtapose the 5′

and 3′ ss early in spliceosomal assembly, as shown by the reportedinteractions of SF2/ASF and SC35 with U1-70K at the 5′ ss andwith U2AF35 at the 3′ ss in an RS domain-dependent manner [25](Figure 2C). Additionally, the enhancer-bound RS domain of theSR protein SF2/ASF has been shown to interact directly withRNA at the BP to promote pre-spliceosomal assembly [27,28].SR proteins may also facilitate the recruitment of the U4/U6 · U5tri-snRNP to the pre-spliceosome [65] via RS domain-mediatedinteractions with the SR-related proteins SRrp65 and SRrp110[66]. The function of SF2/ASF in pre-mRNA splicing dependson the context of the pre-mRNA sequence to which it binds,as shown by the fact that SF2/ASF inhibits adenovirus IIIa pre-mRNA splicing when bound to an intronic repressor element [67].The second RRM of SF2/ASF, and in particular a phylogeneticallyconserved heptapeptide, SWQDLKD, which is located in the firstα-helix of this domain [68], is essential for the splice site selectionactivity of SF2/ASF [69,70]. The structure of this domain revealedan atypical RRM fold that binds to RNA in a novel manner [71].

Use of FRAP (fluorescence recovery after photobleaching)approaches revealed a high mobility for SF2/ASF within thenucleus, with kinetics compatible with a diffusion mechanism[72,73]. Advances in imaging have allowed analysis of splicingfactors both in speckles and at other sites in the nucleoplasm byFRET (fluorescence resonance energy transfer) [74,75]. A recentstudy provided a map of SR protein splicing complexes in thenucleus, and showed that they act in exon and intron definitionin vivo [76].

The U12-type class of pre-mRNA introns, also known as AT-ACintrons, are spliced by the less abundant U12-dependent (minor)spliceosome. The 5′ ss and BP sequences are highly conserved in

Figure 2 Roles of SR proteins in splice site selection

(A) SR proteins bound to ESE elements recruit U2AF35 to an upstream 3′ ss and U1-70K tothe downstream 5′ ss. (B) ESSs recruit hnRNP proteins which block 3′ ss selection by U2AF.SR proteins bound to ESEs can antagonize the action of these splicing repressors, therebypromoting splice site selection. (C) SR proteins can facilitate intron bridging interactions bybinding, via the RS domain, to U1-70K and U2AF 35, thereby juxtaposing the 5′ and 3′ ss.

AT-AC introns, unlike the degenerate sequences found in GT-AGintrons (reviewed in [77]). The SR proteins have been shown toparticipate in AT-AC intron splicing where they promote bindingof the U11 and U12 snRNPs to the 5′ ss and BP respectively [78].There is also evidence that SR proteins contact the pre-mRNAof U12-type introns directly via their RS domain, again in ananalogous fashion to that seen in conventional splicing [79].

A delicate interplay of cis-acting sequences and trans-actingfactors modulate the splicing of regulated exons in a combinatorialfashion [80]. SR family proteins antagonize the activity of hnRNPA/B proteins in splice site selection, with an excess of hnRNPA1 favouring distal 5′ ss, whereas SF2/ASF promotes the use ofproximal 5′ ss [81–85]. Thus, the ratio of hnRNP A1 to SR proteinsin the nucleus is of great importance in alternative splicing regul-ation and may have a crucial role in the tissue-specific and devel-opmental control of regulated splicing. Accordingly, the proteinlevels of SF2/ASF and hnRNP A1 have been found to varynaturally over a very wide range in rat tissues and also in immortaland transformed cell lines [35,86]. SF2/ASF and hnRNP A1 havealso been found to have an antagonistic role in the regulation ofthe neuronal-specific N1 exon of the c-src gene [87]. Antagonismbetween hnRNP proteins and SR proteins has also been shownto regulate a highly complex pattern of mutually exclusive exonsin the Dscam (Down’s syndrome cell adhesion molecule) gene inDrosophila [88]. A subset of SR proteins has been shown toactivate alternative splicing of the cTNT (cardiac troponin T)exon 5 by directly interacting with a purine-rich ESE. Thusregulation of the levels of individual SR proteins may contributeto the developmental regulation of alternative splicing in cTNT[89]. Interestingly, individual SR proteins can sometimes haveantagonistic effects on splice site selection, as is the case withSRp20 and SF2/ASF in the regulation of SRp20 pre-mRNAalternative splicing [90] and of SF2/ASF and SC35 in the

c© The Authors Journal compilation c© 2009 Biochemical Society

Page 5: The SR Protein Family of Splicing Factors Master Regulators

Functional roles of SR proteins in RNA processing 19

regulation of β-tropomyosin [91] and human growth hormonepre-mRNA alternative splicing [92]. Other, non-classical, SRproteins, including p54, SRp38 and SRp86, function solely asnegative regulators of alternative splicing, antagonizing classicalSR proteins and promoting exon skipping [93–95].

SR PROTEIN TARGETS

Several approaches have been taken to identify physiologicalRNA targets of SR proteins. One such approach, termed SELEX(selected evolution of ligands through exponential enrichment)involves the selection of high-affinity binding sites from ran-domized pools of RNA sequences [96]. This has resulted in theidentification of high-affinity binding sites for SF2/ASF and SC35[97], SRp40 [98], 9G8 and SRp20 [99] (Table 5). These bindingsites consist of purine-rich sequences that resemble 5′ ss or exonicsequences, known to function as splicing enhancers. SELEX canbe used in conjunction with large-scale bioinformatic screensto identify further potential binding sites. An alternative to theSELEX approach was the development of a functional SELEXstrategy, which involves selection for a sequence that will promotesplicing, rather than binding alone [100]. This can be modifiedfurther by performing the splicing reactions in a S100 extractwith the addition of a single SR protein [101,102]. The motifsidentified for the SR proteins SF2/ASF, SC35, SRp40 and SRp55using functional SELEX were more redundant than those foundby conventional SELEX, suggesting that the specificity of bindingin vivo depends on factors other than just sequence recognition(Table 5). These motifs have been integrated into a web-basedprogram, known as ESE finder, where input sequences can bescanned for potential ESEs responsive to the above proteins [103].

A number of computational approaches have also been usedto define splicing sequence motifs that regulate exon inclusion.RESCUE (relative enhancer and silencer classification byunanimous enrichment)-ESE predicts sequences which couldfunction as ESEs by statistical analysis of exon–intron and splicesite composition [104]. This is based on the observation thatESEs function in a highly position-dependent fashion and arepresent in constitutively spliced exons and absent in introns.This approach identified 238 candidate ESEs that occurred morefrequently in exons with weak splice sites than in exons withstrong splice sites. By sequence similarity, these were condensedto ten RESCUE-ESE motif clusters and were shown to haveenhancer activity in vivo. Another computational study found2000 RNA octamers, identified as PESEs (putative ESEs), thatwere found more frequently in exons than in pseudo-exons orintronless genes [105]. Validation of a subset of these PESEsresulted in 82% exhibiting decreased splicing efficiency whenthe PESE was mutated [106].

ChIP (chromatin immunoprecipitation) can be used to studynascent RNA–protein interactions, but a variation of thistechnique named RIP (RNP immuno-precipitation) provides moreinformation on protein–RNA interactions in vivo [107]. RIP in-volves cross-linking the protein–RNA interactions using formal-dehyde, followed by immuno-precipitation of the protein–RNAcomplexes. After reversing the cross-links, the RNA can be amp-lified by RT–PCR (reverse transcription–PCR). This techniquerelies on random hexamer primers to identify unknown RNAs orcan be used in conjunction with microarray technology. Reasso-ciation of RNA-binding proteins after cell lysis can complicate theanalysis of these results, since observed protein–RNA interactionsmay not necessarily reflect true in vivo interactions [108]. Anadaptation of the SELEX method described above, known asgenomic SELEX, uses real genomic sequences rather than random

Table 5 RNA sequences identified as SR protein binding sites

N: any nucleotide; R: purine; Y: pyrimidine; D: A, G or U; K: U or G; M: A or C; S: G or C; W: Aor U.

SR protein Binding site Method Reference

SF2/ASF RGAAGAAC SELEX [97]AGGACRRAGC SELEX [97]SRSASGA Functional SELEX [102]UGRWG CLIP [114]

SC35 AGSAGAGUA SELEX [97]GUUCGAGUA SELEX [97]GRYYCSYR Functional SELEX [101]UGUUCSAGWU SELEX [99]GWUWCCUGCUA SELEX [99]GGGUAUGCUG SELEX [99]GAGCAGUAGKS SELEX [99]AGGAGAU SELEX [99]UGCNGYY Functional SELEX [211]

SRp20 GGUCCUCUUC Gel shift [214]WCWWC Splicing assay [99]CUCKUCY RNA affinity [211]

SRp75 GAAGGA UV cross-linking [187]

SRp40 GAGCAGUCGGCUC SELEX [98]ACDGS Functional SELEX [102]

SRp55/B52 USCGKM Functional SELEX [102]UCAACCAGGCGAC SELEX [213]

9G8 UCAACA UV cross-linking [215]ACGAGAGAY SELEX [99]GGACGACGAG Functional SELEX [211]

p54 C Rich UV cross-linking [218]

SRp30c GACGAC Functional SELEX [212]AAAGAGCUCGG Functional SELEX [212]CUGGAUU Gel shift [217]

hTra2β (GAA)n SELEX [43]

SRm160 Purine rich (GAA)n Splicing assay [216]

pools, which allows for identification of authentic protein-bindingRNA sequences [109,110].

Previously, a novel technique named CLIP (cross-linking andimmunoprecipitation) was developed in order to identify in vivoRNA targets [111]. CLIP involves an in vivo photo cross-linkingstep to capture the protein–RNA interactions, followed by partialRNase digestion to generate RNA tags of approximately 60nucleotides followed by specific immunoprecipitation of theprotein of interest. An advantage of this method is that by usingan in vivo photo cross-linking step, which induces a covalentprotein–nucleic acid bond, these interactions are preserved inan intact cell. CLIP was used to characterize the in vivo RNAbinding targets of the neuronal-specific splicing factor Nova andhas allowed the generation of an RNA map to predict splicingregulation dependent on this protein [112,113]. It has also beenused for other RNA binding proteins, including SF2/ASF [114](Table 5). The identification of SR protein targets and the studyof how tissue-specific patterns of splicing change, depending onthe complement of SR proteins present, can also be analysed byalternative splicing microarrays [115–117].

The additional importance of structural elements in splice siteselection should also be taken into account. For example, RNAstructure elements associated with alternative splice-site selectionhave been recently identified in the human genome [118]. Inaddition, RNA folding has been shown to affect the recruitment of

c© The Authors Journal compilation c© 2009 Biochemical Society

Page 6: The SR Protein Family of Splicing Factors Master Regulators

20 J. C. Long and J. F. Caceres

SR proteins to mouse and human ESE elements in the fibronectinEDA exon [119,120]. Finally, competing intronic RNA secondarystructures help to define a complex pattern of mutually exclusiveexons in the Dscam gene [121].

SR PROTEINS HAVE FUNCTIONAL SPECIFICITY

Initially, the ability of different individual SR proteins tocomplement splicing-deficient extracts suggested that SR proteinsmay have redundant functions. However, the sequence-specificRNA binding ability of individual SR proteins and differences intheir ability to regulate alternative splicing suggested otherwise[35,122,123].

A growing body of evidence showed that individual SR proteinswere not functionally equivalent in Drosophila, Caenorhabditiselegans and mouse models. For instance, SF2/ASF was shown tobe an essential factor for cell viability in a chicken cell line and itsdepletion could not be rescued by expression of SC35 or SRp40,indicating a non-redundant function of SF2/ASF [124]. Otherstudies have shown that the SR protein B52/SRp55 is essential forDrosophila development [125,126]. B52/SRp55 was shown notto be essential for the splicing of a number of substrates [127],but specific substrates that were mis-spliced in B52-deficient flieswere identified [128,129]. Furthermore, B52/SRp55 regulates theinclusion of alternative exon 2 in eyeless, a master regulatorof eye development in Drosophila, resulting in the production ofa protein isoform that gives rise to a small-eye phenotype. Conver-sely, the canonical eyeless isoform induces eye overgrowth [130].

Use of RNAi (RNA interference) to inhibit SR protein functionduring C. elegans development revealed that depletion of theorthologue of the mammalian SF2/ASF (CeSF2/ASF) resulted inembryonic lethality, which indicates an essential, non-redundant,role for this gene during nematode development. By contrast,RNAi-mediated depletion of other SR genes resulted in noobvious phenotype, which is indicative of functional redundancy[131–133]. The function of SR proteins has also been studied inmouse model systems (recently reviewed in [134]). All SR-nullmice for SRp20 [135], SC35 [136,137] and SF2/ASF [138] showan early embryonic phenotype indicating that SR proteins arenot redundant. However, these essential functions appear to betissue- or developmental stage-specific, as cultured cells fromthe knockout mice are viable. The generation of conditionalknockouts has allowed further characterization of SR proteinfunction in different tissue types or at various developmental timepoints. Deletion of SC35 in mice results in decreased thymus sizeand a major defect in T-cell maturation [136], whereas tissue-specific ablation of SC35 in the heart has been shown to causedilated cardiomyopathy [137]. SF2/ASF has also been shown tohave a role in cardiac function; however its main function is in thedevelopmental process of postnatal heart remodelling [138]. Miceknockouts of other splicing factors, including Nova, U2AF26,hnRNP U and hnRNP C, also result in embryonic lethalityor developmental defects, which highlights the importance ofsplicing for the correct regulation of biological processes such asembryogenesis and tissue maintenance [134].

POST-SPLICING ACTIVITIES OF SR PROTEINS

SR proteins also function in mRNA processing reactions thatoccur after splicing, including mRNA nuclear export, NMD (non-sense-mediated decay) and translation (reviewed in [139]). SRproteins display a nuclear localization pattern and are foundto accumulate in splicing speckles [140]. However, a subset ofSR proteins, which includes SF2/ASF, SRp20 and 9G8, shuttle

continuously between the nucleus and the cytoplasm [141],reminiscent of what was found for a subset of hnRNP proteins[142]. This suggested that the shuttling SR proteins may functionin cytoplasmic processes, or be involved in the transport ofspliced mRNA. Indeed, SRp20, 9G8 and SF2/ASF function inthe nucleocytoplasmic export of mRNA by interacting with themRNA nuclear export receptor TAP/NFX1 [143,144], exhibitinga higher affinity when hypophosphorylated [145].

SR proteins have also been implicated in regulating the NMDpathway, whereby mRNAs containing premature terminationcodons are targeted for degradation. Increased expression of asubset of SR proteins, including SF2/ASF, SC35, SRp40 andSRp55, strongly enhanced NMD [146]. Interestingly, this effectdoes not appear to be dependent on their nucleocytoplasmicshuttling, suggesting a role for SR proteins in enhancing nuclearsteps of NMD. A recent study showed that SF2/ASF has the poten-tial to affect the cellular site of NMD, shifting this process to thenuclear compartment before mRNA release from nuclei [147].

SF2/ASF controls alternative splicing of pre-mRNAs encodingthe kinases MNK2 [MAPK (mitogen-activated protein kinase)-interacting kinase 2] and S6K1 (S6 kinase 1) that are involved intranslational regulation. Increased expression of SF2/ASF resultsin the production of an isoform of MNK2, which promotesMAPK-independent eIF4E (eukaryotic initiation factor 4E) phos-phorylation, and an unusual oncogenic isoform of S6K1, therebyenhancing cap-dependent translation [148]. SR proteins havealso been shown to directly affect translational regulation. SF2/ASF associates with polyribosomes in cytoplasmic extracts andenhances the translation of an ESE-containing luciferase reporterboth in vivo and in vitro [149]. This direct effect of SF2/ASFin the regulation of the translation of SF2/ASF-bound mRNAtargets is mediated by the recruitment of components of themTOR (mammalian target of rapamycin) signalling pathway,resulting in phosphorylation and release of 4E-BP, a competitiveinhibitor of cap-dependent translation [150]. The role of mTORin the activation of S6K1, which phosphorylates eIF4B and S6,promoting translation initiation, may also be enhanced by SF2/ASF [151] (Figure 3). Other SR proteins have also been reportedto function in translation. SRp20 has been shown to function inIRES (internal ribosome entry site)-mediated translation of a viralRNA [152], whereas 9G8 plays a role in translation of unsplicedmRNA containing a CTE (constitutive transport element) [153].

The results described in this section demonstrate that SRprotein function is not restricted to nuclear mRNA splicing, and itseems sensible that proteins already bound to spliced mRNA mayfunction in subsequent processing events as they are already inplace to facilitate future interactions. However, it also highlightsthe requirement for exquisite regulation of SR proteins in order tomaintain their role in cytoplasmic processing of mRNAs withoutdisrupting nuclear processes, which are highly sensitive to therelative concentration of splicing factors.

SR PROTEIN REGULATION

A dynamic cycle of phosphorylation and dephosphorylation isrequired for pre-mRNA splicing [154], this being related, at leastin part, to the phosphorylation status of SR proteins. The RSdomain of SR proteins is extensively phosphorylated on serineresidues and this plays an important role in regulating the sub-cellular localization and activity of SR proteins (reviewed in [40]).For instance, phosphorylation of the RS domain in SF2/ASF actsto enhance protein–protein interactions with other RS domain-containing splicing factors, such as U1-70K [155], whereasdephosphorylation of SR and SR-related proteins is required forsplicing catalysis to proceed [156,157].

c© The Authors Journal compilation c© 2009 Biochemical Society

Page 7: The SR Protein Family of Splicing Factors Master Regulators

Functional roles of SR proteins in RNA processing 21

Figure 3 Role of SF2/ASF in translation

SF2/ASF-bound mRNAs recruit the mTOR kinase resulting in the phosphorylation and release of 4E-BP, leading to in enhanced translation initiation. The mTOR kinase phosphorylates S6K1, whichpromotes translation initiation, and this may also be enhanced by SF2/ASF. The Mnk2b isoform induced by SF2/ASF-dependent alternative splicing also leads to translation activation.

Several protein kinase families have been shown to phos-phorylate the RS domain of SR proteins, including the SRPK(SR protein kinase) family [158,159], the Clk/Sty family of dual-specificity kinases [160] and topoisomerase I [161]. SRPK1 is aserine-specific kinase that binds to a ‘docking motif’ in SF2/ASFthat restricts phosphorylation to the N-terminus of the RS domain[162]. In contrast, Clk/Sty can phosphorylate the whole of the RSdomain, resulting in a hyperphosphorylated state [163]. The phos-phorylation status of the RS domain of SR proteins is alsoimportant in the post-splicing activities of SR proteins. A hypo-phosphorylated RS domain is required for the interaction of nuc-leocytoplasmic shuttling SR proteins with the TAP/NFX1 nuclearexport receptor [145]. SR protein kinases present in the cytoplasmare required to re-phosphorylate the RS domain before the SRprotein can return to the nucleus [164]. RS domain dephosphoryl-ation also plays an important role in sorting SR proteins in thenucleus, where shuttling SR proteins and non-shuttling SR pro-teins are recycled via different pathways [165]. In the cytoplasm,dephosphorylation of the RS domain enhances mRNA binding ofSF2/ASF and contributes to its role in translation [166]. SR proteinphosphorylation is also important in developmental regulation, asdemonstrated in the nematode Ascaris lumbricoides [167].

Importantly, alternative splicing is extensively regulated bysignal transduction pathways, whereby signalling cascades canlink the splicing machinery to the exterior environment [168].For instance, the SR protein SRp38 is dephosphorylated uponheat shock by the phosphatase PP1 and becomes a potent splicingrepressor [169,170]. Two other well described examples are theinsulin-induced promotion of protein kinase C beta II alternativesplicing as a result of SRp40 phosphorylation by Akt [171], andthe growth factor induced alternative splicing of the fibronectinEDA exon, via phosphorylation of SF2/ASF and 9G8 by Akt[172]. Interestingly, growth factors not only modify the alternativesplicing pattern of the fibronectin gene but also affect its trans-lation in an SR protein-dependent fashion, providing an examplewhere modification of SR protein activity in response to extra-cellular stimulation leads to a concerted regulation of splicing andtranslation [173]. Caffeine regulates the alternative splicing of a

subset of cancer-associated genes, including the tumor suppressorKLF6. This response is mediated by the SR protein, SC35, whichis in turn induced by caffeine, and its overexpression is sufficientto recapitulate this regulated event [174]. Another example of thetight regulation of the SR protein family members is exemplifiedby the common existence of unproductive splicing of SR genes.This is associated with ultraconserved elements that overlap alter-natively spliced exons and target the resulting mRNAs fordegradation by NMD [175,176].

SR PROTEINS AND HUMAN DISEASE

Disruption of the many roles of SR family proteins can lead tohuman disease. Approximately 15% of mutations that cause ge-netic disease affect pre-mRNA splicing [177], targeting conservedsplicing signals including the 5′ ss, 3′ ss and BP, as well asenhancer and silencer sequences. Indeed, analysis of a databaseof 50 single-base substitutions associated with exon-skipping inhuman genes revealed that more than 50% of these mutationsdisrupted at least one of the target motifs for the SR proteinsSF2/ASF, SRp40, SRp55 and SC35 [178,179] (reviewed in [180]).

Cancer

There is emerging evidence that establishes a connection betweenthe mis-expression of SR proteins and the development of can-cerous tissues, mainly as a result of change in the alternativesplicing patterns of key transcripts. Increased expression of SRproteins usually correlates with cancer progression, as shown byelevated expression of SF2/ASF, SC35 and SRp20 in malignantovarian tissue [181] and of several classical SR proteins in breastcancer [182]. However, the mRNA levels of SF2/ASF, SRp40,SRp55 and SRp75 are lower in non-familial colon adenocarcino-mas than adjacent non-pathological tissue, suggesting the levelsof SR proteins in cancerous tissues may be tissue-specific [183].

SF2/ASF was found to be upregulated in several human tu-mours, including lung, colon, kidney, liver, pancreas and breasttumours [148]. Accordingly, gene amplification of SFRS1, which

c© The Authors Journal compilation c© 2009 Biochemical Society

Page 8: The SR Protein Family of Splicing Factors Master Regulators

22 J. C. Long and J. F. Caceres

codes for SF2/ASF, is commonly found in breast cancers [184].Furthermore, increased expression of SF2/ASF transformsimmortal rodent fibroblasts and leads to the formation of sarcomasin nude mice, whereas downregulation of SF2/ASF reverses thesephenotypes. Other SR proteins, such as SC35 and SRp55, didnot have transforming activity, indicating a highly specific role ofSF2/ASF in cancer development. Altogether, these results supportthe notion that SFRS1 is a proto-oncogene [148]. Another RNAtarget for SF2/ASF that can explain its transforming activity isthe proto-oncogene Ron (macrophage-stimulating 1 receptor).SF2/ASF regulates the alternative splicing of Ron pre-mRNAby binding to an ESE in exon 12 and promoting skipping of exon11 [185]. This results in production of �Ron, a constitutivelyactive isoform which confers increased motility on expressingcells, a characteristic required for tumour metastasis. Importantly,abnormal accumulation of �Ron occurs in breast and colontumours and the levels of SF2/ASF mirror those of �Ron [185].

HIV

HIV-1 uses a combination of several alternative 5′ and 3′ ssto generate more than 40 different mRNAs from its full-lengthgenomic pre-mRNA [186]. Several SR proteins have been shownto regulate different splicing events affecting the viral transcripts.For instance, SRp75 binds a viral ESE [187], whereas SF2/ASFand SRp40 bind a guanosine-adenosine-rich ESE identified inexon 5 of HIV-1 leading to its inclusion [188]. Furthermore,HIV infection induces changes in the levels of splicing factors,including SR proteins, that regulate viral alternative splicing andtherefore virus replication [189]. Current drugs used to treat HIV-infected patients involve the use of combinations of retroviralsthat specifically target viral proteins such as reverse transcrip-tase, protease and gp120 (reviewed by [190]). HIV viral produc-tion is tightly linked with alternative splicing of the viral HIV-1pre-mRNA. Therefore, an alternative and novel approach to cir-cumvent the problem of resistance of HIV-1 to current inhibitors isto target the role of SR proteins in HIV pre-mRNA splicing [191].A screen for chemical inhibitors of pre-mRNA splicing identifiedindole derivatives that specifically inhibit ESE-dependent splicingby interacting directly and selectively with individual SR proteins[192]. One such small chemical compound was shown to preventthe production of key viral HIV-1 regulatory proteins whosesplicing depends on weak 3′ ss [193].

SMA (spinal muscular atrophy)

SMA is a severe hereditary neurodegenerative disorder that resultsfrom the lack of a functional SMN1 (survival of motor neuron 1)gene product, which is a key component of the snRNP biogenesispathway. An SMN1 paralogue, the centromeric SMN2 gene,differs by a single nucleotide change, a C > T transition in exon7, that causes substantial skipping of this exon and results in theproduction of a non-functional protein. This exon-skipping eventhas been attributed either to the loss of an SF2/ASF-dependentexonic splicing enhancer [194] or to the creation of an hnRNPA/B-dependent exonic splicing silencer [195].

Several therapeutic approaches, which focus on altering thesplicing of SMN2 to induce exon 7 inclusion and would resultin functional SMN protein in affected patients, have made useof antisense technology [196]. The first uses bifunctional ASOs(antisense oligonucleotides) which are comprised of oligonuc-leotides complementary to exon 7, with a non-complementarytail containing exonic-splicing enhancer motifs recognized by SRproteins. This approach has been shown to mediate the bindingof SF2/ASF to SMN2 exon 7 and promote exon inclusion [197].

An alternative strategy has recently been developed based onbifunctional U7 snRNAs that contain both an antisense sequencetargeting exon 7 and a splicing enhancer sequence to improverecognition of the exon. These RNAs are stably introduced intocells and the U7 snRNAs become incorporated into snRNPs,inducing a prolonged restoration of SMN protein in SMA fibro-blasts [198]. Another approach uses ESSENCE (exon-specificsplicing enhancement by small chimaeric effectors) molecules,which also contain an antisense moiety complementary to thetarget exon and a minimal RS domain peptide designed to mimicthe effect of SR proteins. The ESSENCE molecules have alsobeen shown to restore SMN2 levels to that of wild-type SMN1levels by exon inclusion [199]. Interestingly, the antisense moietyalone stimulated exon 7 inclusion, and functional full-lengthSMN protein was produced in primary fibroblasts from a type ISMA patient [200,201]. An in vivo delivery system has beendeveloped for bifunctional RNAs using a viral vector [202].

Other human diseases

It has been demonstrated that SR proteins are autoantigens inpatients with systemic lupus erythematosus [203]. SR familyproteins have also been shown to have regulatory roles in thesplicing of several pre-mRNAs associated with human disease.For example, SF2/ASF and SRp40 bind to an ISS and promoteexclusion of exon 9 of CFTR (cystic fibrosis transmembraneconductance regulator) [204]. Lack of exon 9 correlates with theoccurrence of monosymptomatic and full forms of cystic fibrosisdisease [205]. SC35 has a role in the aberrant splicing of theE1αPDH (E1α pyruvate dehydrogenase) mRNA, resulting in adefect of mitochondrial energy metabolism. An intronic mutationof the E1αPDH gene that activates a cryptic 5′ ss leads to mis-spliced mRNA and defective protein [206]. SC35 has been shownto significantly activate splicing at this cryptic site. Accordingly,RNAi-mediated depletion of SC35 in primary fibroblasts fromthe affected patient could restore the normal E1αPDH splicingpattern [207]. It has also been recently reported that the expressionof SRp20 is elevated in bipolar patients, which may explain theaberrant splicing of glucocorticoid receptor α in these patients[208].

CONCLUSIONS AND PERSPECTIVES

This article has reviewed the many roles of SR proteins in geneexpression. An obvious question is, why are SR proteins involvedin so many cellular functions? These are very abundant nuclearproteins and a subset of them shuttle to the cytoplasm where theyare involved in NMD and translation regulation. Their functionin different biochemical activities may underlie the extensivenetwork of coupling amongst gene expression machines [209]. Itshould be noted that SR proteins are not the only proteins couplingnuclear and cytoplamic RNA processing events, since the EJC(exon junction complex), a multiprotein complex deposited asa consequence of pre-mRNA splicing, links pre-mRNA splicingwith mRNA export, NMD and translation (reviewed in [210]).Individual SR proteins regulate subsets of pre-mRNAs via spli-cing in the nucleus and post-splicing processes in the cytoplasm.The next few years will see considerable efforts to identify physio-logical RNA targets of SR proteins and gain a better understandingof the many cellular functions of these master regulators of RNAprocessing.

ACKNOWLEDGEMENTS

We are grateful to Sonia Guil (Barcelona) for critical reading of this manuscript.

c© The Authors Journal compilation c© 2009 Biochemical Society

Page 9: The SR Protein Family of Splicing Factors Master Regulators

Functional roles of SR proteins in RNA processing 23

FUNDING

This work was supported by the Medical Research Council and by the European AlternativeSplicing Network of Excellence, EURASNET [grant number S18238].

REFERENCES

1 Berget, S. M., Moore, C. and Sharp, P. A. (1977) Spliced segments at the 5′ terminus ofadenovirus 2 late mRNA. Proc. Natl. Acad. Sci. U.S.A. 74, 3171–3175

2 Chow, L. T., Gelinas, R. E., Broker, T. R. and Roberts, R. J. (1977) An amazing sequencearrangement at the 5′ ends of adenovirus 2 messenger RNA. Cell 12, 1–8

3 Kramer, A. (1996) The structure and function of proteins involved in mammalianpre-mRNA splicing. Annu. Rev. Biochem. 65, 367–409

4 Will, C. L. and Luhrmann, R. (2001) Spliceosomal UsnRNP biogenesis, structure andfunction. Curr. Opin. Cell Biol. 13, 290–301

5 Matlin, A. J. and Moore, M. J. (2007) Spliceosome assembly and composition.Adv. Exp. Med. Biol. 623, 14–35

6 Rappsilber, J., Ryder, U., Lamond, A. I. and Mann, M. (2002) Large-scale proteomicanalysis of the human spliceosome. Genome Res. 12, 1231–1245

7 Zhou, Z., Licklider, L. J., Gygi, S. P. and Reed, R. (2002) Comprehensive proteomicanalysis of the human spliceosome. Nature 419, 182–185

8 Bessonov, S., Anokhina, M., Will, C. L., Urlaub, H. and Luhrmann, R. (2008) Isolationof an active step I spliceosome and composition of its RNP core. Nature 452,846–850

9 Jurica, M. S. and Moore, M. J. (2003) Pre-mRNA splicing: awash in a sea of proteins.Mol. Cell 12, 5–14

10 Fu, X. D. (1995) The superfamily of arginine/serine-rich splicing factors. RNA 1,663–680

11 Graveley, B. R. (2000) Sorting out the complexity of SR protein functions. RNA 6,1197–1211

12 Bourgeois, C. F., Lejeune, F. and Stevenin, J. (2004) Broad specificity of SR(serine/arginine) proteins in the regulation of alternative splicing of pre-messenger RNA.Prog. Nucleic Acid Res. Mol. Biol. 78, 37–88

13 Chou, T. B., Zachar, Z. and Bingham, P. M. (1987) Developmental expression of aregulatory gene is programmed at the level of splicing. EMBO J. 6, 4095–4104

14 Boggs, R. T., Gregor, P., Idriss, S., Belote, J. M. and McKeown, M. (1987) Regulation ofsexual differentiation in D. melanogaster via alternative splicing of RNA from thetransformer gene. Cell 50, 739–747

15 Amrein, H., Gorman, M. and Nothiger, R. (1988) The sex-determining gene tra-2 ofDrosophila encodes a putative RNA binding protein. Cell 55, 1025–1035

16 Ge, H., Zuo, P. and Manley, J. L. (1991) Primary structure of the human splicing factorASF reveals similarities with Drosophila regulators. Cell 66, 373–382

17 Krainer, A. R., Mayeda, A., Kozak, D. and Binns, G. (1991) Functional expression ofcloned human splicing factor SF2: homology to RNA-binding proteins, U1 70K, andDrosophila splicing regulators. Cell 66, 383–394

18 Fu, X. D. and Maniatis, T. (1990) Factor required for mammalian spliceosome assemblyis localized to discrete regions in the nucleus. Nature 343, 437–441

19 Spritz, R. A., Strunk, K., Surowy, C. S., Hoch, S. O., Barton, D. E. and Francke, U. (1987)The human U1-70K snRNP protein: cDNA cloning, chromosomal localization,expression, alternative splicing and RNA-binding. Nucleic Acids Res. 15,10373–10391

20 Theissen, H., Etzerodt, M., Reuter, R., Schneider, C., Lottspeich, F., Argos, P.,Luhrmann, R. and Philipson, L. (1986) Cloning of the human cDNA for the U1RNA-associated 70K protein. EMBO J. 5, 3209–3217

21 Krainer, A. R., Conway, G. C. and Kozak, D. (1990) Purification and characterization ofpre-mRNA splicing factor SF2 from HeLa cells. Genes Dev. 4, 1158–1171

22 Ge, H. and Manley, J. L. (1990) A protein factor, ASF, controls cell-specific alternativesplicing of SV40 early pre-mRNA in vitro. Cell 62, 25–34

23 Roth, M. B., Murphy, C. and Gall, J. G. (1990) A monoclonal antibody that recognizes aphosphorylated epitope stains lampbrush chromosome loops and small granules in theamphibian germinal vesicle. J. Cell Biol. 111, 2217–2223

24 Zahler, A. M., Lane, W. S., Stolk, J. A. and Roth, M. B. (1992) SR proteins: a conservedfamily of pre-mRNA splicing factors. Genes Dev. 6, 837–847

25 Wu, J. Y. and Maniatis, T. (1993) Specific interactions between proteins implicated insplice site selection and regulated alternative splicing. Cell 75, 1061–1070

26 Kohtz, J. D., Jamison, S. F., Will, C. L., Zuo, P., Luhrmann, R., Garcia-Blanco, M. A. andManley, J. L. (1994) Protein-protein interactions and 5′-splice-site recognition inmammalian mRNA precursors. Nature 368, 119–124

27 Shen, H., Kan, J. L. and Green, M. R. (2004) Arginine-serine-rich domains bound atsplicing enhancers contact the branchpoint to promote prespliceosome assembly.Mol. Cell 13, 367–376

28 Shen, H. and Green, M. R. (2004) A pathway of sequential arginine-serine-richdomain-splicing signal interactions during mammalian spliceosome assembly.Mol. Cell 16, 363–373

29 Hertel, K. J. and Graveley, B. R. (2005) RS domains contact the pre-mRNA throughoutspliceosome assembly. Trends Biochem. Sci. 30, 115–118

30 Caceres, J. F., Misteli, T., Screaton, G. R., Spector, D. L. and Krainer, A. R. (1997) Role ofthe modular domains of SR proteins in subnuclear localization and alternative splicingspecificity. J. Cell Biol. 138, 225–238

31 Kataoka, N., Bachorik, J. L. and Dreyfuss, G. (1999) Transportin-SR, a nuclear importreceptor for SR proteins. J. Cell Biol. 145, 1145–1152

32 Lai, M. C., Lin, R. I., Huang, S. Y., Tsai, C. W. and Tarn, W. Y. (2000) A humanimportin-β family protein, transportin-SR2, interacts with the phosphorylated RSdomain of SR proteins. J. Biol. Chem. 275, 7950–7957

33 Krainer, A. R., Conway, G. C. and Kozak, D. (1990) The essential pre-mRNA splicingfactor SF2 influences 5′ splice site selection by activating proximal sites. Cell 62, 35–42

34 Fu, X. D., Mayeda, A., Maniatis, T. and Krainer, A. R. (1992) General splicing factors SF2and SC35 have equivalent activities in vitro, and both affect alternative 5′ and 3′ splicesite selection. Proc. Natl. Acad. Sci. U.S.A. 89, 11224–11228

35 Zahler, A. M., Neugebauer, K. M., Lane, W. S. and Roth, M. B. (1993) Distinct functionsof SR proteins in alternative pre-mRNA splicing. Science 260, 219–222

36 Boucher, L., Ouzounis, C. A., Enright, A. J. and Blencowe, B. J. (2001) A genome-widesurvey of RS domain proteins. RNA 7, 1693–1701

37 Blencowe, B. J., Issner, R., Nickerson, J. A. and Sharp, P. A. (1998) A coactivator ofpre-mRNA splicing. Genes Dev. 12, 996–1009

38 Szymczyna, B. R., Bowman, J., McCracken, S., Pineda-Lucena, A., Lu, Y., Cox, B.,Lambermon, M., Graveley, B. R., Arrowsmith, C. H. and Blencowe, B. J. (2003) Structureand function of the PWI motif: a novel nucleic acid-binding domain that facilitatespre-mRNA processing. Genes Dev. 17, 461–475

39 Blencowe, B. J., Bowman, J. A., McCracken, S. and Rosonina, E. (1999) SR-relatedproteins and the processing of messenger RNA precursors. Biochem. Cell Biol. 77,277–291

40 Lin, S. and Fu, X. D. (2007) SR proteins and related factors in alternative splicing.Adv. Exp. Med. Biol. 623, 107–122

41 Dauwalder, B., Maya-Manzanares, F. and Mattox, W. (1996) A human homologue of theDrosophila sex determination factor transformer-2 has conserved splicing regulatoryfunctions. Proc. Natl. Acad. Sci. U.S.A. 93, 9004–9009

42 Beil, B., Screaton, G. and Stamm, S. (1997) Molecular cloning of htra2-β-1 andhtra2-β-2, two human homologs of tra-2 generated by alternative splicing. DNA CellBiol. 16, 679–690

43 Tacke, R., Tohyama, M., Ogawa, S. and Manley, J. L. (1998) Human Tra2 proteins aresequence-specific activators of pre-mRNA splicing. Cell 93, 139–148

44 Ohno, M. and Shimura, Y. (1996) A human RNA helicase-like protein, HRH1, facilitatesnuclear export of spliced mRNA by releasing the RNA from the spliceosome. Genes Dev.10, 997–1007

45 Zhou, Z. and Reed, R. (1998) Human homologs of yeast prp16 and prp17 revealconservation of the mechanism for catalytic step II of pre-mRNA splicing. EMBO J. 17,2095–2106

46 Misteli, T., Caceres, J. F. and Spector, D. L. (1997) The dynamics of a pre-mRNAsplicing factor in living cells. Nature 387, 523–527

47 Bauren, G. and Wieslander, L. (1994) Splicing of Balbiani ring 1 gene pre-mRNA occurssimultaneously with transcription. Cell 76, 183–192

48 Beyer, A. L. and Osheim, Y. N. (1988) Splice site selection, rate of splicing, andalternative splicing on nascent transcripts. Genes Dev. 2, 754–765

49 Yuryev, A., Patturajan, M., Litingtung, Y., Joshi, R. V., Gentile, C., Gebara, M. andCorden, J. L. (1996) The C-terminal domain of the largest subunit of RNA polymerase IIinteracts with a novel set of serine/arginine-rich proteins. Proc. Natl. Acad. Sci. U.S.A.93, 6975–6980

50 Das, R., Yu, J., Zhang, Z., Gygi, M. P., Krainer, A. R., Gygi, S. P. and Reed, R. (2007) SRproteins function in coupling RNAP II transcription to pre-mRNA splicing. Mol. Cell 26,867–881

51 Lin, S., Coutinho-Mansfield, G., Wang, D., Pandit, S. and Fu, X. D. (2008) The splicingfactor SC35 has an active role in transcriptional elongation. Nat. Struct. Mol. Biol. 15,819–826

52 Fededa, J. P. and Kornblihtt, A. R. (2008) A splicing regulator promotes transcriptionalelongation. Nat. Struct. Mol. Biol. 15, 779–781

53 Li, X. and Manley, J. L. (2005) Inactivation of the SR protein splicing factor ASF/SF2results in genomic instability. Cell 122, 365–378

54 Xiao, R., Sun, Y., Ding, J. H., Lin, S., Rose, D. W., Rosenfeld, M. G., Fu, X. D. and Li, X.(2007) Splicing regulator SC35 is essential for genomic stability and cell proliferationduring mammalian organogenesis. Mol. Cell Biol. 27, 5393–5402

55 Li, X., Niu, T. and Manley, J. L. (2007) The RNA binding protein RNPS1 alleviatesASF/SF2 depletion-induced genomic instability. RNA 13, 2108–2115

c© The Authors Journal compilation c© 2009 Biochemical Society

Page 10: The SR Protein Family of Splicing Factors Master Regulators

24 J. C. Long and J. F. Caceres

56 Hicks, M. J., Yang, C. R., Kotlajich, M. V. and Hertel, K. J. (2006) Linking splicing to PolII transcription stabilizes pre-mRNAs and influences splicing patterns. PLoS Biol. 4,e147

57 de la Mata, M., Alonso, C. R., Kadener, S., Fededa, J. P., Blaustein, M., Pelisch, F.,Cramer, P., Bentley, D. and Kornblihtt, A. R. (2003) A slow RNA polymerase II affectsalternative splicing in vivo. Mol. Cell 12, 525–532

58 de la Mata, M. and Kornblihtt, A. R. (2006) RNA polymerase II C-terminal domainmediates regulation of alternative splicing by SRp20. Nat. Struct. Mol. Biol. 13, 973–980

59 Kornblihtt, A. R. (2007) Coupling transcription and alternative splicing. Adv. Exp.Med. Biol. 623, 175–189

60 Ibrahim, E. C., Schaal, T. D., Hertel, K. J., Reed, R. and Maniatis, T. (2005)Serine/arginine-rich protein-dependent suppression of exon skipping by exonic splicingenhancers. Proc. Natl. Acad. Sci. U.S.A. 102, 5002–5007

61 Graveley, B. R., Hertel, K. J. and Maniatis, T. (2001) The role of U2AF35 and U2AF65 inenhancer-dependent splicing. RNA 7, 806–818

62 Robberson, B. L., Cote, G. J. and Berget, S. M. (1990) Exon definition may facilitatesplice site selection in RNAs with multiple exons. Mol. Cell Biol. 10, 84–94

63 Martinez-Contreras, R., Cloutier, P., Shkreta, L., Fisette, J. F., Revil, T. and Chabot, B.(2007) hnRNP proteins and splicing control. Adv. Exp. Med. Biol. 623, 123–147

64 Zhu, J., Mayeda, A. and Krainer, A. R. (2001) Exon identity established throughdifferential antagonism between exonic splicing silencer-bound hnRNP A1 andenhancer-bound SR proteins. Mol. Cell 8, 1351–1361

65 Roscigno, R. F. and Garcia-Blanco, M. A. (1995) SR proteins escort the U4/U6 · U5tri-snRNP to the spliceosome. RNA 1, 692–706

66 Makarova, O. V., Makarov, E. M. and Luhrmann, R. (2001) The 65 and 110 kDaSR-related proteins of the U4/U6 · U5 tri-snRNP are essential for the assembly of maturespliceosomes. EMBO J. 20, 2553–2563

67 Kanopka, A., Muhlemann, O. and Akusjarvi, G. (1996) Inhibition by SR proteins ofsplicing of a regulated adenovirus pre-mRNA. Nature 381, 535–538

68 Birney, E., Kumar, S. and Krainer, A. R. (1993) Analysis of the RNA-recognition motifand RS and RGG domains: conservation in metazoan pre-mRNA splicing factors.Nucleic Acids Res. 21, 5803–5816

69 Dauksaite, V. and Akusjarvi, G. (2004) The second RNA-binding domain of the humansplicing factor ASF/SF2 is the critical domain controlling adenovirus E1A alternative5′-splice site selection. Biochem. J. 381, 343–350

70 Dauksaite, V. and Akusjarvi, G. (2002) Human splicing factor ASF/SF2 encodesfor a repressor domain required for its inhibitory activity on pre-mRNA splicing.J. Biol. Chem. 277, 12579–12586

71 Tintaru, A. M., Hautbergue, G. M., Hounslow, A. M., Hung, M. L., Lian, L. Y., Craven,C. J. and Wilson, S. A. (2007) Structural and functional analysis of RNA and TAPbinding to SF2/ASF. EMBO Rep. 8, 756–762

72 Kruhlak, M. J., Lever, M. A., Fischle, W., Verdin, E., Bazett-Jones, D. P. and Hendzel,M. J. (2000) Reduced mobility of the alternate splicing factor (ASF) through thenucleoplasm and steady state speckle compartments. J. Cell Biol. 150, 41–51

73 Phair, R. D. and Misteli, T. (2000) High mobility of proteins in the mammalian cellnucleus. Nature 404, 604–609

74 Chusainow, J., Ajuh, P. M., Trinkle-Mulcahy, L., Sleeman, J. E., Ellenberg, J. andLamond, A. I. (2005) FRET analyses of the U2AF complex localize the U2AF35/U2AF65interaction in vivo and reveal a novel self-interaction of U2AF35 RNA 11, 1201–1214

75 Rino, J., Desterro, J. M., Pacheco, T. R., Gadella, Jr, T. W. and Carmo-Fonseca, M.(2008) Splicing factors SF1 and U2AF associate in extra-spliceosomal complexes.Mol. Cell Biol. 28, 3045–3057

76 Ellis, J. D., Lleres, D., Denegri, M., Lamond, A. I. and Caceres, J. F. (2008) Spatialmapping of splicing factor complexes involved in exon and intron definition. J. Cell Biol.181, 921–934

77 Will, C. L. and Luhrmann, R. (2005) Splicing of a rare class of introns by theU12-dependent spliceosome. Biol. Chem. 386, 713–724

78 Hastings, M. L. and Krainer, A. R. (2001) Functions of SR proteins in the U12-dependentAT-AC pre-mRNA splicing pathway. RNA 7, 471–482

79 Shen, H. and Green, M. R. (2007) RS domain-splicing signal interactions in splicing ofU12-type and U2-type introns. Nat. Struct. Mol. Biol. 14, 597–603

80 Smith, C. W. and Valcarcel, J. (2000) Alternative pre-mRNA splicing: the logic ofcombinatorial control. Trends Biochem. Sci. 25, 381–388

81 Mayeda, A. and Krainer, A. R. (1992) Regulation of alternative pre-mRNA splicing byhnRNP A1 and splicing factor SF2. Cell 68, 365–375

82 Eperon, I. C., Ireland, D. C., Smith, R. A., Mayeda, A. and Krainer, A. R. (1993) Pathwaysfor selection of 5′ splice sites by U1 snRNPs and SF2/ASF. EMBO J. 12, 3607–3617

83 Caceres, J. F., Stamm, S., Helfman, D. M. and Krainer, A. R. (1994) Regulation ofalternative splicing in vivo by overexpression of antagonistic splicing factors. Science265, 1706–1709

84 Yang, X., Bani, M. R., Lu, S. J., Rowan, S., Ben David, Y. and Chabot, B. (1994) The A1and A1B proteins of heterogeneous nuclear ribonucleoparticles modulate 5′ splice siteselection in vivo. Proc. Natl. Acad. Sci. U.S.A. 91, 6924–6928

85 Eperon, I. C., Makarova, O. V., Mayeda, A., Munroe, S. H., Caceres, J. F., Hayward, D. G.and Krainer, A. R. (2000) Selection of alternative 5′ splice sites: role of U1 snRNP andmodels for the antagonistic effects of SF2/ASF and hnRNP A1 Mol. Cell Biol. 20,8303–8318

86 Hanamura, A., Caceres, J. F., Mayeda, A., Franza, B. R. and Krainer, A. R. (1998)Regulated tissue-specific expression of antagonistic pre-mRNA splicing factors. RNA 4,430–444

87 Rooke, N., Markovtsov, V., Cagavi, E. and Black, D. L. (2003) Roles for SR proteins andhnRNP A1 in the regulation of c-src exon N1. Mol. Cell Biol. 23, 1874–1884

88 Olson, S., Blanchette, M., Park, J., Savva, Y., Yeo, G. W., Yeakley, J. M., Rio, D. C. andGraveley, B. R. (2007) A regulator of Dscam mutually exclusive splicing fidelity.Nat. Struct. Mol. Biol. 14, 1134–1140

89 Ramchatesingh, J., Zahler, A. M., Neugebauer, K. M., Roth, M. B. and Cooper, T. A.(1995) A subset of SR proteins activates splicing of the cardiac troponin T alternativeexon by direct interactions with an exonic enhancer. Mol. Cell Biol. 15, 4898–4907

90 Jumaa, H. and Nielsen, P. J. (1997) The splicing factor SRp20 modifies splicing of itsown mRNA and ASF/SF2 antagonizes this regulation. EMBO J. 16, 5077–5085

91 Gallego, M. E., Gattoni, R., Stevenin, J., Marie, J. and Expert-Bezancon, A. (1997) TheSR splicing factors ASF/SF2 and SC35 have antagonistic effects on intronicenhancer-dependent splicing of the β-tropomyosin alternative exon 6A. EMBO J. 16,1772–1784

92 Solis, A. S., Peng, R., Crawford, J. B., Phillips, III, J. A. and Patton, J. G. (2008) Growthhormone deficiency and splicing fidelity: two serine/arginine-rich proteins, ASF/SF2and SC35, act antagonistically. J. Biol. Chem. 283, 23619–23626

93 Zhang, W. J. and Wu, J. Y. (1996) Functional properties of p54, a novel SR protein activein constitutive and alternative splicing. Mol. Cell Biol. 16, 5400–5408

94 Cowper, A. E., Caceres, J. F., Mayeda, A. and Screaton, G. R. (2001) Serine-arginine(SR) protein-like factors that antagonize authentic SR proteins and regulate alternativesplicing. J. Biol. Chem. 276, 48908–48914

95 Barnard, D. C., Li, J., Peng, R. and Patton, J. G. (2002) Regulation of alternativesplicing by SRrp86 through coactivation and repression of specific SR proteins. RNA 8,526–533

96 Tuerk, C. and Gold, L. (1990) Systematic evolution of ligands by exponentialenrichment: RNA ligands to bacteriophage T4 DNA polymerase. Science 249, 505–510

97 Tacke, R. and Manley, J. L. (1995) The human splicing factors ASF/SF2 and SC35possess distinct, functionally significant RNA binding specificities. EMBO J. 14,3540–3551

98 Tacke, R., Chen, Y. and Manley, J. L. (1997) Sequence-specific RNA binding by an SRprotein requires RS domain phosphorylation: creation of an SRp40-specific splicingenhancer. Proc. Natl. Acad. Sci. U.S.A. 94, 1148–1153

99 Cavaloc, Y., Bourgeois, C. F., Kister, L. and Stevenin, J. (1999) The splicing factors 9G8and SRp20 transactivate splicing through different and specific enhancers. RNA 5,468–483

100 Coulter, L. R., Landree, M. A. and Cooper, T. A. (1997) Identification of a new class ofexonic splicing enhancers by in vivo selection. Mol. Cell Biol. 17, 2143–2150

101 Liu, H. X., Chew, S. L., Cartegni, L., Zhang, M. Q. and Krainer, A. R. (2000) Exonicsplicing enhancer motif recognized by human SC35 under splicing conditions.Mol. Cell Biol. 20, 1063–1071

102 Liu, H. X., Zhang, M. and Krainer, A. R. (1998) Identification of functional exonic splicingenhancer motifs recognized by individual SR proteins. Genes Dev. 12, 1998–2012

103 Cartegni, L., Wang, J., Zhu, Z., Zhang, M. Q. and Krainer, A. R. (2003) ESEfinder:a web resource to identify exonic splicing enhancers. Nucleic Acids Res. 31, 3568–3571

104 Fairbrother, W. G., Yeh, R. F., Sharp, P. A. and Burge, C. B. (2002) Predictiveidentification of exonic splicing enhancers in human genes. Science 297,1007–1013

105 Zhang, X. H. and Chasin, L. A. (2004) Computational definition of sequence motifsgoverning constitutive exon splicing. Genes Dev. 18, 1241–1250

106 Zhang, X. H., Kangsamaksin, T., Chao, M. S., Banerjee, J. K. and Chasin, L. A. (2005)Exon inclusion is dependent on predictable exonic splicing enhancers. Mol. Cell Biol.25, 7323–7332

107 Niranjanakumari, S., Lasda, E., Brazas, R. and Garcia-Blanco, M. A. (2002) Reversiblecross-linking combined with immunoprecipitation to study RNA-protein interactionsin vivo. Methods 26, 182–190

108 Mili, S. and Steitz, J. A. (2004) Evidence for reassociation of RNA-binding proteins aftercell lysis: implications for the interpretation of immunoprecipitation analyses. RNA 10,1692–1694

109 Singer, B. S., Shtatland, T., Brown, D. and Gold, L. (1997) Libraries for genomic SELEX.Nucleic Acids Res. 25, 781–786

c© The Authors Journal compilation c© 2009 Biochemical Society

Page 11: The SR Protein Family of Splicing Factors Master Regulators

Functional roles of SR proteins in RNA processing 25

110 Lorenz, C., von Pelchrzim, F. and Schroeder, R. (2006) Genomic systematic evolution ofligands by exponential enrichment (Genomic SELEX) for the identification ofprotein-binding RNAs independent of their expression levels. Nat. Protoc. 1, 2204–2212

111 Ule, J., Jensen, K. B., Ruggiu, M., Mele, A., Ule, A. and Darnell, R. B. (2003) CLIPidentifies Nova-regulated RNA networks in the brain. Science 302, 1212–1215

112 Ule, J., Ule, A., Spencer, J., Williams, A., Hu, J. S., Cline, M., Wang, H., Clark, T.,Fraser, C., Ruggiu, M., Zeeberg, B. R., Kane, D., Weinstein, J. N., Blume, J. and Darnell,R. B. (2005) Nova regulates brain-specific splicing to shape the synapse.Nat. Genet. 37, 844–852

113 Ule, J., Stefani, G., Mele, A., Ruggiu, M., Wang, X., Taneri, B., Gaasterland, T., Blencowe,B. J. and Darnell, R. B. (2006) An RNA map predicting Nova-dependent splicingregulation. Nature 444, 580–586

114 Sanford, J. R., Coutinho, P., Hackett, J. A., Wang, X., Ranahan, W. and Caceres, J. F.(2008) Identification of nuclear and cytoplasmic mRNA targets for the shuttling proteinSF2/ASF, PLoS ONE 3, e3369

115 Johnson, J. M., Castle, J., Garrett-Engele, P., Kan, Z., Loerch, P. M., Armour, C. D.,Santos, R., Schadt, E. E., Stoughton, R. and Shoemaker, D. D. (2003) Genome-widesurvey of human alternative pre-mRNA splicing with exon junction microarrays. Science302, 2141–2144

116 Pan, Q., Shai, O., Misquitta, C., Zhang, W., Saltzman, A. L., Mohammad, N., Babak, T.,Siu, H., Hughes, T. R., Morris, Q. D., Frey, B. J. and Blencowe, B. J. (2004) Revealingglobal regulatory features of Mammalian alternative splicing using a quantitativemicroarray platform. Mol. Cell 16, 929–941

117 Blanchette, M., Green, R. E., Brenner, S. E. and Rio, D. C. (2005) Global analysis ofpositive and negative pre-mRNA splicing regulators in Drosophila. Genes Dev. 19,1306–1314

118 Shepard, P. J. and Hertel, K. J. (2008) Conserved RNA secondary structures promotealternative splicing. RNA 14, 1463–1469

119 Muro, A. F., Caputi, M., Pariyarath, R., Pagani, F., Buratti, E. and Baralle, F. E. (1999)Regulation of fibronectin EDA exon alternative splicing: possible role of RNA secondarystructure for enhancer display. Mol. Cell Biol. 19, 2657–2671

120 Buratti, E., Muro, A. F., Giombi, M., Gherbassi, D., Iaconcig, A. and Baralle, F. E. (2004)RNA folding affects the recruitment of SR proteins by mouse and human polypurinicenhancer elements in the fibronectin EDA exon. Mol. Cell Biol. 24, 1387–1400

121 Graveley, B. R. (2005) Mutually exclusive splicing of the insect Dscam pre-mRNAdirected by competing intronic RNA secondary structures. Cell 123, 65–73

122 Screaton, G. R., Caceres, J. F., Mayeda, A., Bell, M. V., Plebanski, M., Jackson, D. G.,Bell, J. I. and Krainer, A. R. (1995) Identification and characterization of three membersof the human SR family of pre-mRNA splicing factors. EMBO J. 14, 4336–4349

123 Wang, J. and Manley, J. L. (1995) Overexpression of the SR proteins ASF/SF2 and SC35influences alternative splicing in vivo in diverse ways. RNA 1, 335–346

124 Wang, J., Takagaki, Y. and Manley, J. L. (1996) Targeted disruption of an essentialvertebrate gene: ASF/SF2 is required for cell viability. Genes Dev. 10, 2588–2599

125 Ring, H. Z. and Lis, J. T. (1994) The SR protein B52/SRp55 is essential for Drosophiladevelopment. Mol. Cell Biol. 14, 7499–7506

126 Peng, X. and Mount, S. M. (1995) Genetic enhancement of RNA-processing defects by adominant mutation in B52, the Drosophila gene for an SR protein splicing factor.Mol. Cell Biol. 15, 6273–6282

127 Hoffman, B. E. and Lis, J. T. (2000) Pre-mRNA splicing by the essential Drosophilaprotein B52: tissue and target specificity. Mol. Cell Biol. 20, 181–186

128 Kim, S., Shi, H., Lee, D. K. and Lis, J. T. (2003) Specific SR protein-dependent splicingsubstrates identified through genomic SELEX. Nucleic Acids Res. 31, 1955–1961

129 Rasheva, V. I., Knight, D., Bozko, P., Marsh, K. and Frolov, M. V. (2006) Specific role ofthe SR protein splicing factor B52 in cell cycle control in Drosophila. Mol. Cell Biol. 26,3468–3477

130 Fic, W., Juge, F., Soret, J. and Tazi, J. (2007) eye development under the control ofSRp55/B52-mediated alternative splicing of eyeless. PLoS ONE 2, e253

131 Kawano, T., Fujita, M. and Sakamoto, H. (2000) Unique and redundant functions of SRproteins, a conserved family of splicing factors, in Caenorhabditis elegans development.Mech. Dev. 95, 67–76

132 Longman, D., Johnstone, I. L. and Caceres, J. F. (2000) Functional characterization ofSR and SR-related genes in Caenorhabditis elegans. EMBO J. 19, 1625–1637

133 Longman, D., McGarvey, T., McCracken, S., Johnstone, I. L., Blencowe, B. J. andCaceres, J. F. (2001) Multiple interactions between SRm160 and SR family proteins inenhancer- dependent splicing and development of C. elegans. Curr. Biol. 11, 1923–1933

134 Moroy, T. and Heyd, F. (2007) The impact of alternative splicing in vivo: mouse modelsshow the way. RNA 13, 1155–1171

135 Jumaa, H., Wei, G. and Nielsen, P. J. (1999) Blastocyst formation is blocked in mouseembryos lacking the splicing factor SRp20 Curr. Biol. 9, 899–902

136 Wang, H. Y., Xu, X., Ding, J. H., Bermingham, Jr, J. R. and Fu, X. D. (2001) SC35 plays arole in T cell development and alternative splicing of CD45. Mol. Cell 7, 331–342

137 Ding, J. H., Xu, X., Yang, D., Chu, P. H., Dalton, N. D., Ye, Z., Yeakley, J. M., Cheng, H.,Xiao, R. P., Ross, J. et al. (2004) Dilated cardiomyopathy caused by tissue-specificablation of SC35 in the heart. EMBO J. 23, 885–896

138 Xu, X., Yang, D., Ding, J. H., Wang, W., Chu, P. H., Dalton, N. D., Wang, H. Y.,Bermingham, Jr, J. R., Ye, Z., Liu, F. et al. (2005) ASF/SF2-regulated CaMKIIδalternative splicing temporally reprograms excitation-contraction coupling in cardiacmuscle. Cell 120, 59–72

139 Huang, Y. and Steitz, J. A. (2005) SRprises along a messenger’s journey. Mol. Cell 17,613–615

140 Lamond, A. I. and Spector, D. L. (2003) Nuclear speckles: a model for nuclearorganelles. Nat. Rev. Mol. Cell Biol. 4, 605–612

141 Caceres, J. F., Screaton, G. R. and Krainer, A. R. (1998) A specific subset of SRproteins shuttles continuously between the nucleus and the cytoplasm. Genes Dev. 12,55–66

142 Dreyfuss, G., Matunis, M. J., Pinol-Roma, S. and Burd, C. G. (1993) hnRNP proteinsand the biogenesis of mRNA. Annu. Rev. Biochem. 62, 289–321

143 Huang, Y., Gattoni, R., Stevenin, J. and Steitz, J. A. (2003) SR splicing factors serve asadapter proteins for TAP-dependent mRNA export. Mol. Cell 11, 837–843

144 Huang, Y. and Steitz, J. A. (2001) Splicing factors SRp20 and 9G8 promote thenucleocytoplasmic export of mRNA. Mol. Cell 7, 899–905

145 Huang, Y., Yario, T. A. and Steitz, J. A. (2004) A molecular link between SR proteindephosphorylation and mRNA export. Proc. Natl. Acad. Sci. U.S.A. 101, 9666–9670

146 Zhang, Z. and Krainer, A. R. (2004) Involvement of SR proteins in mRNA surveillance.Mol. Cell 16, 597–607

147 Sato, H., Hosoda, N. and Maquat, L. E. (2008) Efficiency of the pioneer round oftranslation affects the cellular site of nonsense-mediated mRNA decay. Mol. Cell 29,255–262

148 Karni, R., de Stanchina, E., Lowe, S. W., Sinha, R., Mu, D. and Krainer, A. R. (2007) Thegene encoding the splicing factor SF2/ASF is a proto-oncogene. Nat. Struct. Mol. Biol.14, 185–193

149 Sanford, J. R., Gray, N. K., Beckmann, K. and Caceres, J. F. (2004) A novel role forshuttling SR proteins in mRNA translation. Genes Dev. 18, 755–768

150 Michlewski, G., Sanford, J. R. and Caceres, J. F. (2008) The splicing factor SF2/ASFregulates translation initiation by enhancing phosphorylation of 4E-BP1. Mol. Cell 30,179–189

151 Holz, M. K., Ballif, B. A., Gygi, S. P. and Blenis, J. (2005) mTOR and S6K1 mediateassembly of the translation preinitiation complex through dynamic protein interchangeand ordered phosphorylation events. Cell 123, 569–580

152 Bedard, K. M., Daijogo, S. and Semler, B. L. (2007) A nucleo-cytoplasmic SR proteinfunctions in viral IRES-mediated translation initiation. EMBO J. 26, 459–467

153 Swartz, J. E., Bor, Y. C., Misawa, Y., Rekosh, D. and Hammarskjold, M. L. (2007) Theshuttling SR protein 9G8 plays a role in translation of unspliced mRNA containing aconstitutive transport element. J. Biol. Chem. 282, 19844–19853

154 Mermoud, J. E., Cohen, P. T. and Lamond, A. I. (1994) Regulation of mammalianspliceosome assembly by a protein phosphorylation mechanism. EMBO J. 13,5679–5688

155 Xiao, S. H. and Manley, J. L. (1997) Phosphorylation of the ASF/SF2 RS domain affectsboth protein-protein and protein-RNA interactions and is necessary for splicing. GenesDev. 11, 334–344

156 Tazi, J., Kornstadt, U., Rossi, F., Jeanteur, P., Cathala, G., Brunel, C. and Luhrmann, R.(1993) Thiophosphorylation of U1-70K protein inhibits pre-mRNA splicing. Nature 363,283–286

157 Cao, W., Jamison, S. F. and Garcia-Blanco, M. A. (1997) Both phosphorylation anddephosphorylation of ASF/SF2 are required for pre-mRNA splicing in vitro. RNA 3,1456–1467

158 Gui, J. F., Lane, W. S. and Fu, X. D. (1994) A serine kinase regulates intracellularlocalization of splicing factors in the cell cycle. Nature 369, 678–682

159 Wang, H. Y., Lin, W., Dyck, J. A., Yeakley, J. M., Songyang, Z., Cantley, L. C. and Fu,X. D. (1998) SRPK2: a differentially expressed SR protein-specific kinase involved inmediating the interaction and localization of pre-mRNA splicing factors in mammaliancells. J. Cell Biol. 140, 737–750

160 Colwill, K., Pawson, T., Andrews, B., Prasad, J., Manley, J. L., Bell, J. C. and Duncan,P. I. (1996) The Clk/Sty protein kinase phosphorylates SR splicing factors and regulatestheir intranuclear distribution. EMBO J. 15, 265–275

161 Rossi, F., Labourier, E., Forne, T., Divita, G., Derancourt, J., Riou, J. F., Antoine, E.,Cathala, G., Brunel, C. and Tazi, J. (1996) Specific phosphorylation of SR proteins bymammalian DNA topoisomerase I. Nature 381, 80–82

162 Ngo, J. C., Giang, K., Chakrabarti, S., Ma, C. T., Huynh, N., Hagopian, J. C., Dorrestein,P. C., Fu, X. D., Adams, J. A. and Ghosh, G. (2008) A sliding docking interaction isessential for sequential and processive phosphorylation of an SR protein by SRPK1.Mol. Cell 29, 563–576

c© The Authors Journal compilation c© 2009 Biochemical Society

Page 12: The SR Protein Family of Splicing Factors Master Regulators

26 J. C. Long and J. F. Caceres

163 Ngo, J. C., Chakrabarti, S., Ding, J. H., Velazquez-Dones, A., Nolen, B., Aubol, B. E.,Adams, J. A., Fu, X. D. and Ghosh, G. (2005) Interplay between SRPK and Clk/Stykinases in phosphorylation of the splicing factor ASF/SF2 is regulated by a dockingmotif in ASF/SF2. Mol. Cell 20, 77–89

164 Ding, J. H., Zhong, X. Y., Hagopian, J. C., Cruz, M. M., Ghosh, G., Feramisco, J.,Adams, J. A. and Fu, X. D. (2006) Regulated cellular partitioning of SR protein-specifickinases in mammalian cells. Mol. Biol. Cell 17, 876–885

165 Lin, S., Xiao, R., Sun, P., Xu, X. and Fu, X. D. (2005) Dephosphorylation-dependentsorting of SR splicing factors during mRNP maturation. Mol. Cell 20, 413–425

166 Sanford, J. R., Ellis, J. D., Cazalla, D. and Caceres, J. F. (2005) Reversiblephosphorylation differentially affects nuclear and cytoplasmic functions of splicingfactor 2/alternative splicing factor. Proc. Natl. Acad. Sci. U.S.A. 102, 15042–15047

167 Sanford, J. R. and Bruzik, J. P. (1999) Developmental regulation of SR proteinphosphorylation and activity. Genes Dev. 13, 1513–1518

168 Lynch, K. W. (2007) Regulation of alternative splicing by signal transduction pathways.Adv. Exp. Med. Biol. 623, 161–174

169 Shin, C., Feng, Y. and Manley, J. L. (2004) Dephosphorylated SRp38 acts as a splicingrepressor in response to heat shock. Nature 427, 553–558

170 Shi, Y. and Manley, J. L. (2007) A complex signaling pathway regulates SRp38phosphorylation and pre-mRNA splicing in response to heat shock. Mol. Cell 28, 79–90

171 Patel, N. A., Kaneko, S., Apostolatos, H. S., Bae, S. S., Watson, J. E., Davidowitz, K.,Chappell, D. S., Birnbaum, M. J., Cheng, J. Q. and Cooper, D. R. (2005) Molecular andgenetic studies imply Akt-mediated signaling promotes protein kinase Cβ II alternativesplicing via phosphorylation of serine/arginine-rich splicing factor SRp40. J. Biol.Chem. 280, 14302–14309

172 Blaustein, M., Pelisch, F., Coso, O. A., Bissell, M. J., Kornblihtt, A. R. and Srebrow, A.(2004) Mammary epithelial-mesenchymal interaction regulates fibronectin alternativesplicing via phosphatidylinositol 3-kinase. J. Biol. Chem. 279, 21029–21037

173 Blaustein, M., Pelisch, F., Tanos, T., Munoz, M. J., Wengier, D., Quadrana, L., Sanford,J. R., Muschietti, J. P., Kornblihtt, A. R., Caceres, J. F. et al. (2005) Concerted regulationof nuclear and cytoplasmic activities of SR proteins by AKT. Nat. Struct. Mol. Biol. 12,1037–1044

174 Shi, J., Hu, Z., Pabon, K. and Scotto, K. W. (2008) Caffeine regulates alternative splicingin a subset of cancer-associated genes: a role for SC35. Mol. Cell Biol. 28, 883–895

175 Lareau, L. F., Inada, M., Green, R. E., Wengrod, J. C. and Brenner, S. E. (2007)Unproductive splicing of SR genes associated with highly conserved and ultraconservedDNA elements. Nature 446, 926–929

176 Ni, J. Z., Grate, L., Donohue, J. P., Preston, C., Nobida, N., O’Brien, G., Shiue, L., Clark,T. A., Blume, J. E. and Ares, Jr, M. (2007) Ultraconserved elements are associated withhomeostatic control of splicing regulators by alternative splicing and nonsense-mediated decay. Genes Dev. 21, 708–718

177 Krawczak, M., Reiss, J. and Cooper, D. N. (1992) The mutational spectrum of singlebase-pair substitutions in mRNA splice junctions of human genes: causes andconsequences. Hum. Genet. 90, 41–54

178 Valentine, C. R. (1998) The association of nonsense codons with exon skipping. Mutat.Res. 411, 87–117

179 Liu, H. X., Cartegni, L., Zhang, M. Q. and Krainer, A. R. (2001) A mechanism for exonskipping caused by nonsense or missense mutations in BRCA1 and other genes. Nat.Genet. 27, 55–58

180 Cartegni, L., Chew, S. L. and Krainer, A. R. (2002) Listening to silence and understandingnonsense: exonic mutations that affect splicing. Nat. Rev. Genet. 3, 285–298

181 Fischer, D. C., Noack, K., Runnebaum, I. B., Watermann, D. O., Kieback, D. G.,Stamm, S. and Stickeler, E. (2004) Expression of splicing factors in human ovariancancer. Oncol. Rep. 11, 1085–1090

182 Stickeler, E., Kittrell, F., Medina, D. and Berget, S. M. (1999) Stage-specific changes inSR splicing factors and alternative splicing in mammary tumorigenesis. Oncogene 18,3574–3582

183 Ghigna, C., Moroni, M., Porta, C., Riva, S. and Biamonti, G. (1998) Altered expressionof heterogenous nuclear ribonucleoproteins and SR factors in human colonadenocarcinomas. Cancer Res. 58, 5818–5824

184 Sinclair, C. S., Rowley, M., Naderi, A. and Couch, F. J. (2003) The 17q23 amplicon andbreast cancer. Breast Cancer Res. Treat. 78, 313–322

185 Ghigna, C., Giordano, S., Shen, H., Benvenuto, F., Castiglioni, F., Comoglio, P. M.,Green, M. R., Riva, S. and Biamonti, G. (2005) Cell motility is controlled by SF2/ASFthrough alternative splicing of the Ron protooncogene. Mol. Cell 20, 881–890

186 Stoltzfus, C. M. and Madsen, J. M. (2006) Role of viral splicing elements and cellularRNA binding proteins in regulation of HIV-1 alternative RNA splicing. Curr. HIV Res. 4,43–55

187 Exline, C. M., Feng, Z. and Stoltzfus, C. M. (2008) Negative and positive mRNA splicingelements act competitively to regulate human immunodeficiency virus type 1 vif geneexpression. J. Virol. 82, 3921–3931

188 Asang, C., Hauber, I. and Schaal, H. (2008) Insights into the selective activation ofalternatively used splice acceptors by the human immunodeficiency virus type-1bidirectional splicing enhancer. Nucleic Acids Res. 36, 1450–1463

189 Dowling, D., Nasr-Esfahani, S., Tan, C. H., O’Brien, K., Howard, J. L., Jans, D. A.,Purcell, D. F., Stoltzfus, C. M. and Sonza, S. (2008) HIV-1 infection induces changes inexpression of cellular splicing factors that regulate alternative viral splicing and virusproduction in macrophages. Retrovirology 5, 18

190 Barbaro, G., Scozzafava, A., Mastrolorenzo, A. and Supuran, C. T. (2005) Highly activeantiretroviral therapy: current state of the art, new agents and their pharmacologicalinteractions useful for improving therapeutic outcome. Curr. Pharm. Des. 11,1805–1843

191 Soret, J., Gabut, M. and Tazi, J. (2006) SR proteins as potential targets for therapy.Prog. Mol. Subcell. Biol. 44, 65–87

192 Soret, J., Bakkour, N., Maire, S., Durand, S., Zekri, L., Gabut, M., Fic, W., Divita, G.,Rivalle, C., Dauzonne, D. et al. (2005) Selective modification of alternative splicing byindole derivatives that target serine-arginine-rich protein splicing factors. Proc. Natl.Acad. Sci. U.S.A. 102, 8764–8769

193 Bakkour, N., Lin, Y. L., Maire, S., Ayadi, L., Mahuteau-Betzer, F., Nguyen, C. H.,Mettling, C., Portales, P., Grierson, D., Chabot, B., Jeanteur, P., Branlant, C., Corbeau, P.and Tazi, J. (2007) Small-molecule inhibition of HIV pre-mRNA splicing as a novelantiretroviral therapy to overcome drug resistance. PLoS Pathog. 3, 1530–1539

194 Cartegni, L. and Krainer, A. R. (2002) Disruption of an SF2/ASF-dependent exonicsplicing enhancer in SMN2 causes spinal muscular atrophy in the absence of SMN1.Nat. Genet. 30, 377–384

195 Kashima, T. and Manley, J. L. (2003) A negative element in SMN2 exon 7 inhibitssplicing in spinal muscular atrophy. Nat. Genet. 34, 460–463

196 Wirth, B., Brichta, L. and Hahnen, E. (2006) Spinal muscular atrophy: from gene totherapy. Semin. Pediatr. Neurol. 13, 121–131

197 Skordis, L. A., Dunckley, M. G., Yue, B., Eperon, I. C. and Muntoni, F. (2003)Bifunctional antisense oligonucleotides provide a trans-acting splicing enhancer thatstimulates SMN2 gene expression in patient fibroblasts. Proc. Natl. Acad. Sci. U.S.A.100, 4114–4119

198 Marquis, J., Meyer, K., Angehrn, L., Kampfer, S. S., Rothen-Rutishauser, B. andSchumperli, D. (2007) Spinal muscular atrophy: SMN2 pre-mRNA splicing corrected bya U7 snRNA derivative carrying a splicing enhancer sequence. Mol. Ther. 15,1479–1486

199 Cartegni, L. and Krainer, A. R. (2003) Correction of disease-associated exon skipping bysynthetic exon-specific activators. Nat. Struct. Biol. 10, 120–125

200 Hua, Y., Vickers, T. A., Okunola, H. L., Bennett, C. F. and Krainer, A. R. (2008) Antisensemasking of an hnRNP A1/A2 intronic splicing silencer corrects SMN2 splicing intransgenic mice. Am. J. Hum. Genet. 82, 834–848

201 Hua, Y., Vickers, T. A., Baker, B. F., Bennett, C. F. and Krainer, A. R. (2007) Enhancementof SMN2 exon 7 inclusion by antisense oligonucleotides targeting the exon. PLoS. Biol.5, e73

202 Baughan, T., Shababi, M., Coady, T. H., Dickson, A. M., Tullis, G. E. and Lorson, C. L.(2006) Stimulating full-length SMN2 expression by delivering bifunctional RNAs via aviral vector. Mol. Ther. 14, 54–62

203 Neugebauer, K. M., Merrill, J. T., Wener, M. H., Lahita, R. G. and Roth, M. B. (2000) SRproteins are autoantigens in patients with systemic lupus erythematosus. Importance ofphosphoepitopes. Arthritis Rheum. 43, 1768–1778

204 Buratti, E., Stuani, C., De Prato, G. and Baralle, F. E. (2007) SR protein-mediatedinhibition of CFTR exon 9 inclusion: molecular characterization of the intronic splicingsilencer. Nucleic Acids Res. 35, 4359–4368

205 Pagani, F., Buratti, E., Stuani, C., Romano, M., Zuccato, E., Niksic, M., Giglio, L.,Faraguna, D. and Baralle, F. E. (2000) Splicing factors induce cystic fibrosistransmembrane regulator exon 9 skipping through a nonevolutionary conserved intronicelement. J. Biol. Chem. 275, 21041–21047

206 Mine, M., Brivet, M., Touati, G., Grabowski, P., Abitbol, M. and Marsac, C. (2003)Splicing error in E1α pyruvate dehydrogenase mRNA caused by novel intronic mutationresponsible for lactic acidosis and mental retardation. J. Biol. Chem. 278, 11768–11772

207 Gabut, M., Mine, M., Marsac, C., Brivet, M., Tazi, J. and Soret, J. (2005) The SRprotein SC35 is responsible for aberrant splicing of the E1α pyruvate dehydrogenasemRNA in a case of mental retardation with lactic acidosis. Mol. Cell Biol. 25, 3286–3294

208 Watanuki, T., Funato, H., Uchida, S., Matsubara, T., Kobayashi, A., Wakabayashi, Y.,Otsuki, K., Nishida, A. and Watanabe, Y. (2008) Increased expression of splicing factorSRp20 mRNA in bipolar disorder patients. J. Affect. Disord. 110, 62–69

209 Maniatis, T. and Reed, R. (2002) An extensive network of coupling among geneexpression machines. Nature 416, 499–506

210 Tange, T. O., Nott, A. and Moore, M. J. (2004) The ever-increasing complexities of theexon junction complex. Curr. Opin. Cell Biol. 16, 279–284

c© The Authors Journal compilation c© 2009 Biochemical Society

Page 13: The SR Protein Family of Splicing Factors Master Regulators

Functional roles of SR proteins in RNA processing 27

211 Schaal, T. D. and Maniatis, T. (1999) Selection and characterization of pre-mRNAsplicing enhancers: identification of novel SR protein-specific enhancer sequences.Mol. Cell Biol. 19, 1705–1719

212 Tian, H. and Kole, R. (2001) Strong RNA splicing enhancers identified by a modifiedmethod of cycled selection interact with SR protein. J. Biol. Chem. 276,33833–33839

213 Shi, H., Hoffman, B. E. and Lis, J. T. (1997) A specific RNA hairpin loop structure bindsthe RNA recognition motifs of the Drosophila SR protein B52. Mol. Cell Biol. 17,2649–2657

214 Lou, H., Neugebauer, K. M., Gagel, R. F. and Berget, S. M. (1998) Regulation ofalternative polyadenylation by U1 snRNPs and SRp20. Mol. Cell Biol. 18, 4977–4985

215 Lynch, K. W. and Maniatis, T. (1996) Assembly of specific SR protein complexes ondistinct regulatory elements of the Drosophila doublesex splicing enhancer. Genes Dev.10, 2089–2101

216 Eldridge, A. G., Li, Y., Sharp, P. A. and Blencowe, B. J. (1999) The SRm160/300 splicingcoactivator is required for exon-enhancer function. Proc. Natl. Acad. Sci. U.S.A. 96,6125–6130

217 Simard, M. J. and Chabot, B. (2002) SRp30c is a repressor of 3′ splice site utilization.Mol. Cell Biol. 22, 4001–4010

218 Kennedy, C. F., Kramer, A. and Berget, S. M. (1998) A role for SRp54 during intronbridging of small introns with pyrimidine tracts upstream of the branch point.Mol. Cell Biol. 18, 5425–5434

Received 24 July 2008/4 September 2008; accepted 4 September 2008Published on the Internet 12 December 2008, doi:10.1042/BJ20081501

c© The Authors Journal compilation c© 2009 Biochemical Society