1196

Biochemical Society Transactions (2014) Volume 42, part 4

Regulation of gene expression through production of unstable mRNA isoforms Christopher R. Sibley*1 *Department of Molecular Neuroscience, UCL Institute of Neurology, Queen Square, London WC1N 3BG, U.K.

Biochemical Society Transactions

www.biochemsoctrans.org

Abstract Alternative splicing is universally accredited for expanding the information encoded within the transcriptome. In recent years, several tightly regulated alternative splicing events have been reported which do not lead to generation of protein products, but lead to unstable mRNA isoforms. Instead these transcripts are targets for NMD (nonsense-mediated decay) or retained in the nucleus and degraded. In the present review I discuss the regulation of these events, and how many have been implicated in control of gene expression that is instrumental to a number of developmental paradigms. I further discuss their relevance to disease settings and conclude by highlighting technologies that will aid identification of more candidate events in future.

Principles of splicing Post-transcriptional splicing of RNA is an essential and precisely regulated mechanism that ensures removal of introns and ligation of exons within multi-exon eukaryotic genes in order to produce a translatable mRNA. The spliceosome, a dynamic RNA–protein complex, carries out this splicing of pre-mRNA. Specifically, five snRNAs (small nuclear RNAs) and their associated RBPs (RNAbinding proteins), collectively termed snRNP (small nuclear ribonucleoprotein) complexes, are recruited to cis-elements in the pre-RNA sequence via base pairing to the snRNAs. The cis-elements that are recognized by the snRNPs at exon/intron and intron/exon junctions consist of the 5 splice site (AG/GURAG), 3 splice site (a polypyrimidine tract followed by acceptor AG dinucleotide) and the branch point sequence (YURAY) (Figure 1a). Once an snRNP is recruited to its corresponding cis-element it performs a specific role in the splicing reaction at that site (reviewed comprehensively in [1]). In addition to the snRNPs, over 100 spliceosomeassociated RBPs also recognize the same or additional ciselements in the RNA and influence snRNP recruitment. For example, U2AF65 recognition of the poly-pyrimidine tract is required for recruitment of the U2 snRNP to the branch point sequence at early stages of splicing [2,3] (Figure 1a). Accordingly, the spliceosome is constantly re-modelled as splicing proceeds, allowing additional snRNPs and RBPs to interact with the RNA after previous steps have been completed [1]. Key words: alternative splicing, alternative splicing-coupled nonsense-mediated decay (ASNMD), post-transcriptional control, premature termination codon, regulated intron retention. Abbreviations: ALS, amyotrophic lateral sclerosis; AS-NMD, alternative splicing-coupled NMD; CLIP, cross-linked immunoprecipitation; EJC, exon–junction complex; FTLD, fronto-temporal lobar degeneration; hnRNP, heterogenous ribonucleoprotein; PTBP, polypyrimidine tract-binding protein; PTC, premature termination codon; RBP, RNA-binding protein; Rbfox2, RBP, fox-1 homologue 2; Sam68, Src-associated in mitosis 68 kDa protein; SF2, splicing factor 2; SLC4, solute carrier family 4; snRNA, small nuclear RNA; snRNP, small nuclear ribonucleoprotein; SR, serine/arginine; SURF, SMG1, UPF1, ERF1 and ERF3. 1 email Christopher Sibley: [email protected]

 C The

C 2014 Biochemical Society Authors Journal compilation 

Splicing requires the presence of the 5 splice site, 3 splice site and branch-point sequence being located in appropriate spatial relation to one another [1,4]. However, correct combinations of these cis-acting elements do not always lead to a constitutive splicing event. Instead, alternative splicing is used to vastly expand the repertoire of mRNAs generated from a single transcript. Indeed, 95 % of all human genes are expected to be alternatively spliced through selective use of cassette exons, use of alternative 5 splice sites or 3 splice sites, or through intron retention [5,6] (Figure 1b). Alternative splicing is regulated by additional trans-acting RBPs, often with tissue-specific expression patterns and/or under post-translational control, that recognize certain ciselements surrounding regulated junctions [7–10] (Figure 1a). This includes short SE (splicing enhancer) or SS (splicing silencer) motifs located within exons (ESE and ESS) or introns (ISE and ISS) that promote either exon inclusion or exclusion [11,12]. For example, the SR (serine/arginine) repeat-rich proteins are a conserved family of genes that typically interact with enhancer sites and positively influence snRNP recruitment to the neighbouring splice junctions [13]. SR protein binding to an exon has been demonstrated to enhance recruitment of the U1 snRNP to the 5 splice site [14] and U2 snRNP to upstream branch point region [15, 16], and are required for the assembly of the U4/U6 and U5 tri-snRNP on the pre-mRNA at latter stages of splicing [17]. In opposition, among others, are the hnRNPs (heterogenous ribonucleoproteins), which tend to bind silencer sequences and promote skipping of exons [13]. This is typically achieved by restricting access of other splicesome-associated components to cis-elements in the RNA. For example, hnRNP C has a higher affinity to uridine-rich tracts than U2AF65, but is less tolerable of interspersed cytidines [18]. Composition of the poly-pyrimidine tract will subsequently determine if hnRNP C blocks U2AF65 binding to the site and hence U2 snRNP recruitment. Biochem. Soc. Trans. (2014) 42, 1196–1205; doi:10.1042/BST20140102

RNA UK 2014

Figure 1 Regulated use of cis- and trans-acting factors directs splicing and can lead to various types of alternative-splicing events (a) Cis- and trans-acting factors contributing to the splicing decision during mRNA production. Following transcription, snRNPs and trans-acting RBPs interact with cis-acting elements around splice junctions to dictate which introns are removed and which exons are included during the splicing reaction. (b) Types of alternative splicing events. Red circles highlight the event specified.

 C The

C 2014 Biochemical Society Authors Journal compilation 

1197

1198

Biochemical Society Transactions (2014) Volume 42, part 4

In addition to enhancing the coding information contained within the transcriptome, it has become apparent that production of unstable mRNA isoforms via alternative splicing can be used by the cell as a strategy to quantitatively regulate gene expression. This can be through alternative splicing events that result in targeting of the transcript for NMD (nonsense-mediated decay). This is most commonly referred to as AS-NMD (alternative splicing-coupled NMD) or, alternatively, RUST (regulated unproductive splicing and translation). In addition, this can be achieved by controlling nuclear–cytoplasmic localization of certain transcripts within the cell via regulated intron retention.

AS-NMD NMD is an RNA surveillance pathway that targets PTC (premature termination codon)-containing transcripts for degradation during translation. Recognition of these transcripts involves both nuclear and cytoplasmic events [19] (Figure 2a). First, during pre-mRNA splicing, the EJC (exon–junction complex) is deposited on mature exon–exon junctions. This acts as a positional memory of the splicing event and is additionally required for correct nucleocytoplasmic shuttling of the mRNA [20,21]. Among others, two proteins present in the EJC deposited in the nucleus are the upstream frameshift family members UPF2 and UPF3, which are critical to later NMD activation. After shuttling to the cytoplasm the initial rounds of mRNA translation appear to clear EJC components from the transcript [22–24]. Once a premature termination codon is recognized, the UPF1-containing SURF (SMG1, UPF1, ERF1 and ERF3) complex is deposited on the mRNA at the site of the terminating ribosome [25,26]. Should this be more than ∼50 nucleotides from a proceeding EJC complex that has not been cleared, then UPF1 interacts with UPF2 and UPF3, leading to UPF1 phosphorylation and, as a consequence, translational repression through target degradation [27]. This NMD activation can be through inclusion of a PTC-containing exon [28–30], presence of an upstream ORF [31,32], skipping of a section of the 3 UTR [28] or through retention of a PTC-containing intron in a transcript which remains actively shuttled to the cytosol [28,33–35] (Figure 2b). A further observation is that longer 3 UTRs are more susceptible to NMD, which is likely determined by the distance between the termination codon and poly-A tail [36,37]. Until recently it was expected that the main role of NMD was to clean up errors in splicing, and to eliminate both nonsense and frame-shift mutations that lead to production of damaging truncated proteins. However, this view has changed in lieu of findings demonstrating that up to a third of human transcripts are predicted to be NMD targets [38– 40], and many conserved cassette exons contain PTCs in all three frames to suggest a specific role in triggering NMD [28,39]. Most importantly, many alternative splicing events of PTC-containing exons or within 3 UTRs appear to be both highly conserved and dynamically regulated. Collectively it points towards a regulatory role of NMD, in addition to the  C The

C 2014 Biochemical Society Authors Journal compilation 

surveillance role, in which alternative splicing and translation is coupled to NMD to regulate gene expression [38,41,42]. Furthermore, it is expected that AS-NMD represents a particularly pervasive means by which gene expression can be regulated [38,43]. Although the deliberate production of aberrant mRNA isoforms with low stability at first seems counterproductive, using AS-NMD as means of regulating gene expression has a number of potential benefits to the cell. First, it offers a means by which the expression levels of a specific gene can be fine-tuned without having to interfere with the global activity of transcription factors or translational regulators, many of which bind hundreds, if not thousands, of sites across the genome/transcriptome [10,44–46]. Secondly, continual production of mRNA that is typically just a single event away from generating a translatable mRNA isoform offers a means to generate a rapidly responding molecular switch with which to respond to external cues [44]. Finally, it can permit translation into a protein by directing transcription degradation after just a single round of translation. In doing so, NMD acts as a brake on protein expression as the message, in effect, self-destructs [35,45,46].

What is the meaning of this nonsense? Since its initial suggestion, several AS-NMD candidate events have been identified. Following previous reports demonstrating that one member of the SR protein family regulates its own expression through coupling splicing to NMD [47], Lareau et al. [28] demonstrated that in fact all members of the SR protein family contained ultraconserved elements that were alternatively spliced in order to direct NMD. This included examples of PTC-containing exons, alternative splicing of the 3 UTRs and retention of PTCcontaining introns. It is expected that the presence of these ultraconserved elements represents a negative-feedback loop since stabilization of NMD products through translational inhibition led to decreases in coding variants. Indeed, negative feedback has been confirmed for a tissue-specific hnRNP family member and splicing factor, PTBP1 (polypyrimidine tract-binding protein 1), which regulates its own levels through PTBP1-dependent skipping of exon 11 to create an AS-NMD transcript [48]. PTBP1 is a regulator of alternative splicing, mRNA stability and localization, and this autoregulation is expected to fine-tune PTBP1 protein to appropriate levels. Many other splicing factors similarly contain highly conserved regions consistent with generating NMD products [49–51], some of which demonstrate autoregulation or are implicated in negative-feedback loops [44,52–56]. This includes components of the core splicing machinery [56,57]. Collectively this suggests AS-NMD has a crucial role across evolution in modulating expression levels of many splicing factors, perhaps to regulate tissue-specific splicing patterns via regulation of mRNA stability [34], or to permit rapid response to external cues [44].

RNA UK 2014

Figure 2 NMD targets PTC-containing transcripts for degradation during translation (a) NMD is initiated if a termination codon is detected upstream of an exon–junction complex. The exon–junction complex is deposited at exon–exon junctions following the splicing reaction in the nucleus. Transcripts are exported to the cytosol and

 C The

C 2014 Biochemical Society Authors Journal compilation 

1199

1200

Biochemical Society Transactions (2014) Volume 42, part 4

the initial rounds of translation clear exon–junction complexes from the transcript. Once exon junction complexes are cleared the transcript is recruited to the polysome for active translation. If a premature termination codon is detected during the the initial round of translation then the UPF1-containing SURF complex is recruited to the terminating ribosome. If this is located >50 nucleotides from a downstream exon–junction complex that has yet to be cleared then UPF1 interacts with the exon–junction complex components and triggers a cascade of events leading to transcript degradation. (b) Types of splicing events that can lead to NMD. Red circles highlight the event specified. (c) Intron retention can lead to transcripts that are unable to be exported to the cytosol and are subsequently degraded.

However, AS-NMD is not limited to RBPs. Network analysis following silencing of NMD proteins identifies transcription factors, signalling proteins and metabolic proteins as other protein families with regulation of gene expression via NMD [58,59]. Moreover, many NMD components are also NMD sensitive to suggest autoregulation of the mechanism itself [37,60]. In recent years, several studies have progressed beyond identification of just the regulated sites and successfully distinguished the quality control aspects of NMD from bona fide AS-NMD to demonstrate direct functional consequences of specific events [29,35,44,61]. For example, the transition from epithelial to mesenchymal states is an important step in both development and the progression of epithelial tumour metastasis. Valacca et al. [61] have dissected a role for ASNMD in this transition in which phosphorylation status of the RBP, Sam68 (Src-associated in mitosis 68 kDa protein), regulates splicing of the 3 UTR of the SR protein SF2 (splicing factor 2) in response to extracellular cues. Production of an NMD-sensitive and unstable SF2 transcript, following ERK (extracellular-signal-regulated kinase)-mediated Sam68 phosphorylation, leads to a reduction in the inclusion of exon 11 in the RON proto-oncogene. This generates a constitutively active form of RON that promotes this epithelial to mesenchymal transition. PTBP1 and PTBP2 are two functionally related proteins which regulate alternative splicing of many overlapping and unique exons [44]. The two proteins display opposing expression patterns during neuronal development. With the exception of in neurons, PTBP1 is ubiquitously expressed. In contrast, PTBP2 is restricted to post-mitotic neurons, despite RNA transcripts being found in neural progenitor cells [44]. A switch in usage from PTBP1 to PTBP2 occurs during neuronal differentiation, and this is expected to account for ∼25 % of the changes in alternative splicing that occur during this fate-determining transition. Importantly this switch is regulated by AS-NMD. Both genes contain homologous PTC-containing exons, inclusion of which is promoted by PTBP1. In the case of PTBP1 this is as part of the autoregulatory feedback mechanism to limit PTBP1 levels [48], whereas in PTBP2 this forms the molecular switch implicated in neuronal differentiation [44,62]. Specifically, during neuronal differentiation, elevated miR-124 levels lead to down-regulation of PTBP1 and the subsequent skipping of the PTC-containing PTBP2 exon 10 [44,63]. This results in active translation of PTBP2 and elegantly demonstrates how AS-NMD of a single exon can lead to far-reaching effects on a population of genes. This transition may be even  C The

C 2014 Biochemical Society Authors Journal compilation 

further strengthened by a general reduction of NMD activity which inhibits cell proliferation, inhibits TGFB (transforming growth factor β) signalling and drives expression of NMDtargeting miRNAs to reinforce this decision [30]. Another neuronal-specific regulator of alternative splicing, NOVA, had previously been demonstrated to regulate inclusion of a number of cassette exons and poly-A site choice by binding to YCAY clusters [8,64,65]. By focusing in on nuclear binding events, NOVA has now been shown to bind YCAY clusters deep within long introns while in the nucleus, and regulate multiple AS-NMD events that are ultimately important in determining the steady-state expression of host transcripts through regulation of their stability [29]. This regulation is dynamic since seizure induction, which results in NOVA shuttling to the cytoplasm, led to corresponding changes in inclusion or skipping of these sites and changes to host gene levels as mRNA stability is modified. Indeed, these are present in several transcripts linked to seizure development such as DLG3 (discs large homologue 3) [66], SLC4a10 (solute carrier family 4, sodium bicarbonate transporter, member 10) [67] and SLC4a3 [68]. Importantly, several of these sites are cryptic events not apparent in current gene annotations. This demonstrates the efficiency of NMD in eliminating these transcripts while additionally highlighting that many AS-NMD events may be presently unknown due to limitations in methods used to identify them. Finally, a particularly elegant use of AS-NMD has been demonstrated in the local control of protein synthesis in neurons, which aids axon guidance during development [35]. During embryonic development the commissural neurons of the spinal cord are first attracted to the ventral midline, cross, and then are repelled from the ventral midline in order carry on their intended course of projection [69]. This switch from attraction to repulsion is mediated by cues from the glial-containing floor plate [70], and involves a switch from Robo3.1 to Robo3.2 expression in the growth cone after ventral midline crossing [71]. The two closely related transcripts coding these variants differ only by the presence of one retained intron in the Robo3.2 transcript. This introduces a PTC and makes Robo3.2 an NMD-sensitive transcript. Colak et al. [35] elegantly demonstrate in mouse samples that, unlike Robo3.1, which is translated in the cell body, substantial Robo3.2 mRNA is transported along axons and locally translated in response to floor plate cues after ventral midline crossing. NMD components are additionally associated with the Robo3.2 transcripts and found enriched at the growth cone, suggesting that local regulation of RNA stability by NMD may also be involved. Indeed, this leads to

RNA UK 2014

an excellent exploitation of the NMD machinery since, upon initiation of Robo3.2 translation, it is expected that only a single round of translation occurs before the transcripts are targeted for degradation. This is sufficient to drive adequate Robo3.2 expression to direct repulsion, perhaps one protein per transcript, but not so much that over-repulsion ensues. It is expected that such local control of AS-NMD events will be important for other axonal guidance events given the elevated expression of NMD components in growth cones of other neurons [35,70].

Regulation of AS-NMD AS-NMD must be regulated on a cell-type specific basis in order to ensure appropriate gene expression levels of regulated targets are met. The limitation of AS-NMD to a single round of translation is one manner in which this could be achieved [35]. However, recent work using CLIP (cross-linked immunoprecipitation) from the group of Phillip Sharp has shed additional light on the finetuning mechanism of AS-NMD that permits different steadystate protein concentrations [72]. CLIP analysis of the RNA targets of the RBP, Rbfox2 (RBP, fox-1 homologue 2), revealed an enrichment around cassette exons which had reduced splice site strength, and a preference for binding events located downstream of exons to lead to their inclusion. However, several strongly bound sites failed to reveal corresponding changes in inclusion when looking at RNA-seq data generated following Rbfox2 silencing. Closer inspection revealed that a significant proportion of these bound sites neighboured PTC-containing cassette exons, inclusion of which would lead to instability and degradation of the host transcript, and consequently underestimations of the extent of regulation in RNA-seq data. Many of the genes with these identified events were RBPs, again hinting at the important role of AS-NMD in the regulation of RBP expression levels. Crucially, however, it was found that the protein expression levels of Rbfox2 could modulate the threshold of previously documented autoregulation of ASNMD events in both Tia1 and Ptbp2 leading to corresponding changes in total gene expression [44,52,72]. This importantly sets a precedent for the re-analysis of other CLIP datasets to look for similar genome-wide regulation of AS-NMD events by other RBPs. Further to this, as direct NMD targets, at least seven components of the NMD pathway (UPF1, UPF2, UPF3B, SMG1, SMG5, SMG6 and SMG7) are subject to feedback regulation that is exerted and regulated in both a cell-type and developmental manner [60,73]. Moreover, this feedback could be separated into UPF3B-sensitive and -insensitive branches, corroborating the idea that there are multiple branches of the NMD pathway which have requirements for different combinations of NMD factors [74,75]. The purpose of this feedback is expected to ensure correct buffering of NMD components in response to external stimuli. Supporting this, it was demonstrated that three of these regulated genes (SMG1, SMG5 and SMG6) are rate-limiting to NMD in general, and

blocking the autoregulated up-regulation of one of these following UPF1 depletion severely hampered the NMD response [60]. In contrast, overexpression of one of these ratelimiting factors, in order to induce overactive NMD, resulted in reduced cell proliferation. This supports the hypothesis that maintaining expression of NMD components at celltype specific requirements, and tight feedback regulation of this pathway, are critical for normal cellular homoeostasis [30,60].

Intron retention mediated regulation of gene expression: In addition to AS-NMD, co-ordinated intron retention can also regulate gene expression. Although AS-NMD may ensue if the transcript is efficiently exported to the cytosol and recognized by NMD machinery (Figure 2b) [28,33– 35], in other cases export is inhibited leading to nuclear retention and degradation (Figure 2C) [76–79]. Although the precise mechanism by which transcripts with retained introns are degraded needs further study, it appears that both components of the nuclear exosome (e.g. Dis3, Exosc10 and Exosc9) and nuclear pore-associated proteins (e.g. Tpr) are involved [82]. Intron retention is less well understood than mechanisms underlying AS-NMD of cassette exons, but appears to be dependent on a high GC content of introns, nucleosome density and weak splice sites and linked to reduced availability of splicing components [34,80,81]. Increasing evidence suggests it could be of comparable importance and prevalence to AS-NMD. Indeed, up to 15 % of human protein-coding genes have evidence of such events [76], although many are in 3 UTRs and may yet prove to be processed as AS-NMD targets. The importance of intron retention is highlighted by two key reports. In leucocyte differentiation pathways the differential regulation of 86 intron retention events leads to the NMD of a collection of functionally related genes linked to leucocyte function at specific developmental time points [34]. Meanwhile, in the case of neuronal differentiation, PTBP1 represses a number of 3 UTR intron removal events which leads to their nuclear retention and degradation in an NMD-independent fashion. Following the PTBP1 to PTBP2 switch [44,63], these are removed produce translation competent transcripts that are exported to the cytosol and then participate in the neuronal differentiation program [79]. In fact PTBP1 may have a more general role in retaining premRNA in the nucleus [82], together with the U1 snRNP and U2AF65 [83]. These two studies clearly identify the coupling of alternative splicing to nuclear retention and degradation as a powerful means of regulating gene expression, and it is predicted that other examples will surface in due course. For a more extensive review on intron retention, please see [84].

Unstable mRNA isoforms and disease Splicing events that result in unstable mRNA isoforms have clear relevance to disease settings and may represent novel therapeutic targets. As many as a third of mutations  C The

C 2014 Biochemical Society Authors Journal compilation 

1201

1202

Biochemical Society Transactions (2014) Volume 42, part 4

leading to genetic diseases are due to mutations that create unstable NMD-sensitive transcripts [43,85–87]. Among others, this includes α-thalassaemia [88], β-thalassaemia [89], retinitis pigmentosa [90], ataxia telangiectasia [91], spinal muscular atrophy [92] and numerous cancers [93–96]. Accordingly, drugs which promote PTC read-through are attracting considerable attention as potential therapies [97– 99], although these are likely to be specific to individual cases [100]. In addition, reports have demonstrated disease-relevant disruption of AS-NMD events involved in autoregulation of certain genes. FUS is an ALS (amyotrophic lateral sclerosis)and FTLD (fronto-temporal lobar degeneration)-associated RBP which has been shown to control its own expression through regulated skipping of exon 7, leading to an NMDsensitive transcript [101,102]. However, discordant regulation of this exon was reported with three pathogenic mutations that all lead to cytoplasmic accumulation of FUS. This led to higher levels of inclusion of the exon and the subsequent self-promotion of the cytoplasmic localization. This autoregulation could be re-induced in mutants through use of exon-skipping antisense oligonucleotides, which importantly led to restoration of nuclear FUS localization, suggesting this may be a useful target for ALS/FTLD should appropriate delivery strategies be developed. Similar disruption to the autoregulation of another ALS/FTLDassociated protein, TDP-43 (TAR DNA-binding protein-43), is also expected to be pathogenic [103,104], although this remains to be definitively confirmed. Finally, genetic perturbations to NMD pathway components lead to genome-wide effects on mRNA stability and disruption of gene regulation that contributes to disease. Evident of this, naturally occurring mutations in the UPF3B gene are a direct of cause intellectual disability and other mental disorders [105–107]. This is expected to affect only UPF3B-sensitive NMD targets in one branch of the NMD pathway [74,75]. Indeed only approximately 5 % of human genes are affected, whereas the protein product of its paralogue, UPF3A, is stabilized in order to compensate [108].

genome [109]. It will therefore be interesting to see how many of these cryptic sites are used in additional AS-NMD or intron-retention events as appropriate sequencing approaches to detect them are used. The discovery of unstable alternative mRNA isoforms is made harder by the fact that they are efficiently degraded and difficult to detect with methods such as whole-cell RNA-seq [72]. Re-evaluation of current genome annotations using functional genomics data has recently expanded our knowledge of the transcriptome in the fruitfly [110,111]. This has increased the known number of regulated exons in this species, many of which are candidates for regulation by NMD, and it is hoped that similar analysis will be as fruitful in more complex species. This will allow new coordinates for mapping of next generation sequencing data that will incorporate many new AS-NMD candidate events. Further to this, AS-NMD and intron retention events are carried out in the nucleus and respectively degraded rapidly in the cytoplasm or nucleus. Therefore the use of nuclear RNAseq will help enrich for these characteristic events before their degradation. As an alternative, CLIP-based studies have repeatedly revealed novel cryptic sites that are implicated in the regulation of gene expression via generation of unstable mRNAs [18,29,72]. This is because CLIP can cross-link interactions between nuclear RBPs and the low abundance unstable targets before their degradation. The result is that CLIP studies of candidate regulatory proteins can hugely facilitate cryptic transcript detection. Numerous CLIP studies have now been undertaken [112], and it is expected that re-analysis of existing data will reveal many more regulated splicing events that lead to unstable mRNA isoforms with both biological and disease relevance.

Acknowledgements I thank Jernej Ule and Andrea D’Ambrogio for helpful discussion and a critical reading of the paper during preparation.

Funding Future perspectives The discussed examples make it clear that regulated splicing to produce unstable mRNA isoforms has a fundamental role in regulating gene expression genome-wide and adds an added dimension to regulation of the transcriptome by alternative splicing. It is expected that growing appreciation of the best methods to identify these events will dramatically expand the known cases in the future. Indeed, the precise scale of unstable alternative mRNA isoforms remains unclear. This is partly because much genomic analysis focuses on reference annotations rather than additionally exploring novel splicing events where new candidates may be found. Indeed, recent reports have identified many cryptic splicing events that are tightly repressed and therefore not present in current annotations, but which are clearly under dynamic regulation [18,29,72]. Such cryptic splice sites are widespread across the  C The

C 2014 Biochemical Society Authors Journal compilation 

Funding from the European Research Council supported this work.

References 1 2

3

4

5

Matera, A.G. and Wang, Z. (2014) A day in the life of the spliceosome. Nat. Rev. Mol. Cell Biol. 15, 108–121 CrossRef PubMed Ruskin, B., Zamore, P.D. and Green, M.R. (1988) A factor, U2AF, is required for U2 snRNP binding and splicing complex assembly. Cell: 52, 207–219 CrossRef PubMed Zamore, P.D. and Green, M.R. (1989) Identification, purification, and biochemical characterization of U2 small nuclear ribonucleoprotein auxiliary factor. Proc. Natl. Acad. Sci. U.S.A. 86, 9243–9247 CrossRef PubMed De Conti, L., Baralle, M. and Buratti, E. (2013) Exon and intron definition in pre-mRNA splicing. Wiley Interdiscip. Rev. RNA 4, 49–60 CrossRef PubMed Pan, Q., Shai, O., Lee, L.J., Frey, B.J. and Blencowe, B.J. (2008) Deep surveying of alternative splicing complexity in the human transcriptome by high-throughput sequencing. Nat. Genet. 40, 1413–1415 CrossRef PubMed

RNA UK 2014

6

7

8

9

10

11

12

13

14

15

16

17 18

19 20

21

22

23

24

25

Wang, E.T., Sandberg, R., Luo, S., Khrebtukova, I., Zhang, L., Mayr, C., Kingsmore, S.F., Schroth, G.P. and Burge, C.B. (2008) Alternative isoform regulation in human tissue transcriptomes. Nature 456, 470–476 CrossRef PubMed Tollervey, J.R., Curk, T., Rogelj, B., Briese, M., Cereda, M., Kayikci, M., Konig, J., Hortobagyi, T., Nishimura, A.L., Zupunski, V. et al. (2011) Characterizing the RNA targets and position-dependent splicing regulation by TDP-43. Nat. Neurosci. 14, 452–458 CrossRef PubMed Ule, J., Jensen, K.B., Ruggiu, M., Mele, A., Ule, A. and Darnell, R.B. (2003) CLIP identifies Nova-regulated RNA networks in the brain. Science 302, 1212–1215 CrossRef PubMed Wang, Z., Kayikci, M., Briese, M., Zarnack, K., Luscombe, N.M., Rot, G., Zupan, B., Curk, T. and Ule, J. (2010) iCLIP predicts the dual splicing effects of TIA-RNA interactions. PLoS Biol 8, e1000530 CrossRef PubMed Wang, E.T., Cody, N.A., Jog, S., Biancolella, M., Wang, T.T., Treacy, D.J., Luo, S., Schroth, G.P., Housman, D.E., Reddy, S. et al. (2012) Transcriptome-wide regulation of pre-mRNA splicing and mRNA localization by muscleblind proteins. Cell 150, 710–724 CrossRef PubMed Fairbrother, W.G., Yeh, R.F., Sharp, P.A. and Burge, C.B. (2002) Predictive identification of exonic splicing enhancers in human genes. Science 297, 1007–1013 CrossRef PubMed Ke, S., Shang, S., Kalachikov, S.M., Morozova, I., Yu, L., Russo, J.J., Ju, J. and Chasin, L.A. (2011) Quantitative evaluation of all hexamers as exonic splicing elements. Genome Res. 21, 1360–1374 CrossRef PubMed Busch, A. and Hertel, K.J. (2012) Evolution of SR protein and hnRNP splicing regulatory factors. Wiley Interdiscip. Rev RNA 3, 1–12 CrossRef PubMed Zahler, A.M. and Roth, M.B. (1995) Distinct functions of SR proteins in recruitment of U1 small nuclear ribonucleoprotein to alternative 5 splice sites. Proc. Natl. Acad. Sci. U.S.A. 92, 2642–2646 CrossRef PubMed Lavigueur, A., La Branche, H., Kornblihtt, A.R. and Chabot, B. (1993) A splicing enhancer in the human fibronectin alternate ED1 exon interacts with SR proteins and stimulates U2 snRNP binding. Genes Dev. 7, 2405–2417 CrossRef PubMed Tarn, W.Y. and Steitz, J.A. (1995) Modulation of 5 splice site choice in pre-messenger RNA by two distinct steps. Proc. Natl. Acad. Sci. U.S.A. 92, 2504–2508 CrossRef PubMed Roscigno, R.F. and Garcia-Blanco, M.A. (1995) SR proteins escort the U4/U6.U5 tri-snRNP to the spliceosome. RNA 1, 692–706 PubMed Zarnack, K., Konig, J., Tajnik, M., Martincorena, I., Eustermann, S., Stevant, I., Reyes, A., Anders, S., Luscombe, N.M. and Ule, J. (2013) Direct competition between hnRNP C and U2AF65 protects the transcriptome from the exonization of Alu elements. Cell 152, 453–466 CrossRef PubMed Schoenberg, D.R. and Maquat, L.E. (2012) Regulation of cytoplasmic mRNA decay. Nat. Rev. Genet. 13, 246–259 CrossRef PubMed Sauliere, J., Murigneux, V., Wang, Z., Marquenet, E., Barbosa, I., Le Tonqueze, O., Audic, Y., Paillard, L., Roest Crollius, H. and Le Hir, H. (2012) CLIP-seq of eIF4AIII reveals transcriptome-wide mapping of the human exon junction complex. Nat. Struct. Mol. Biol. 19, 1124–1131 CrossRef PubMed Le Hir, H., Gatfield, D., Izaurralde, E. and Moore, M.J. (2001) The exon-exon junction complex provides a binding platform for factors involved in mRNA export and nonsense-mediated mRNA decay. EMBO J. 20, 4987–4997 CrossRef PubMed Ishigaki, Y., Li, X., Serin, G. and Maquat, L.E. (2001) Evidence for a pioneer round of mRNA translation: mRNAs subject to nonsense-mediated decay in mammalian cells are bound by CBP80 and CBP20. Cell 106, 607–617 CrossRef PubMed Lejeune, F., Ishigaki, Y., Li, X. and Maquat, L.E. (2002) The exon junction complex is detected on CBP80-bound but not eIF4E-bound mRNA in mammalian cells: dynamics of mRNP remodeling. EMBO J. 21, 3536–3545 CrossRef PubMed Dostie, J. and Dreyfuss, G. (2002) Translation is required to remove Y14 from mRNAs in the cytoplasm. Curr. Biol. 12, 1060–1067 CrossRef PubMed Kashima, I., Yamashita, A., Izumi, N., Kataoka, N., Morishita, R., Hoshino, S., Ohno, M., Dreyfuss, G. and Ohno, S. (2006) Binding of a novel SMG-1–Upf1–eRF1–eRF3 complex (SURF) to the exon junction complex triggers Upf1 phosphorylation and nonsense-mediated mRNA decay. Genes Dev. 20, 355–367 CrossRef PubMed

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

Hwang, J., Sato, H., Tang, Y., Matsuda, D. and Maquat, L.E. (2010) UPF1 association with the cap-binding protein, CBP80, promotes nonsense-mediated mRNA decay at two distinct steps. Mol. Cell 39, 396–409 CrossRef PubMed Nagy, E. and Maquat, L.E. (1998) A rule for termination-codon position within intron-containing genes: when nonsense affects RNA abundance. Trends Biochem. Sci. 23, 198–199 CrossRef PubMed 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 ultraconserved DNA elements. Nature 446, 926–929 CrossRef PubMed Eom, T., Zhang, C., Wang, H., Lay, K., Fak, J., Noebels, J.L. and Darnell, R.B. (2013) NOVA-dependent regulation of cryptic NMD exons controls synaptic protein levels after seizure. Elife 2, e00178 CrossRef PubMed Lou, C.H., Shao, A., Shum, E.Y., Espinoza, J.L., Huang, L., Karam, R. and Wilkinson, M.F. (2014) Posttranscriptional control of the stem cell and neurogenic programs by the nonsense-mediated RNA decay pathway. Cell Rep. 6, 748–764 CrossRef PubMed Mendell, J.T., Sharifi, N.A., Meyers, J.L., Martinez-Murillo, F. and Dietz, H.C. (2004) Nonsense surveillance regulates expression of diverse classes of mammalian transcripts and mutes genomic noise. Nat. Genet. 36, 1073–1078 CrossRef PubMed Kim, K.M., Cho, H. and Kim, Y.K. (2012) The upstream open reading frame of cyclin-dependent kinase inhibitor 1A mRNA negatively regulates translation of the downstream main open reading frame. Biochem. Biophys. Res. Commun. 424, 469–475 CrossRef PubMed Bell, T.J., Miyashiro, K.Y., Sul, J.Y., McCullough, R., Buckley, P.T., Jochems, J., Meaney, D.F., Haydon, P., Cantor, C., Parsons, T.D. and Eberwine, J. (2008) Cytoplasmic BK(Ca) channel intron-containing mRNAs contribute to the intrinsic excitability of hippocampal neurons. Proc. Natl. Acad. Sci. U.S.A. 105, 1901–1906 CrossRef PubMed Wong, J.J., Ritchie, W., Ebner, O.A., Selbach, M., Wong, J.W., Huang, Y., Gao, D., Pinello, N., Gonzalez, M., Baidya, K. et al. (2013) Orchestrated intron retention regulates normal granulocyte differentiation. Cell 154, 583–595 CrossRef PubMed Colak, D., Ji, S.J., Porse, B.T. and Jaffrey, S.R. (2013) Regulation of axon guidance by compartmentalized nonsense-mediated mRNA decay. Cell 153, 1252–1265 CrossRef PubMed Buhler, M. and Muhlemann, O. (2005) Alternative splicing induced by nonsense mutations in the immunoglobulin mu VDJ exon is independent of truncation of the open reading frame. RNA 11, 139–146 CrossRef PubMed Yepiskoposyan, H., Aeschimann, F., Nilsson, D., Okoniewski, M. and Muhlemann, O. (2011) Autoregulation of the nonsense-mediated mRNA decay pathway in human cells. RNA 17, 2108–2118 CrossRef PubMed Lewis, B.P., Green, R.E. and Brenner, S.E. (2003) Evidence for the widespread coupling of alternative splicing and nonsense-mediated mRNA decay in humans. Proc. Natl. Acad. Sci. U.S.A. 100, 189–192 CrossRef PubMed Baek, D. and Green, P. (2005) Sequence conservation, relative isoform frequencies, and nonsense-mediated decay in evolutionarily conserved alternative splicing. Proc. Natl. Acad. Sci. U.S.A. 102, 12813–12818 CrossRef PubMed Weischenfeldt, J., Waage, J., Tian, G., Zhao, J., Damgaard, I., Jakobsen, J.S., Kristiansen, K., Krogh, A., Wang, J. and Porse, B.T. (2012) Mammalian tissues defective in nonsense-mediated mRNA decay display highly aberrant splicing patterns. Genome Biol. 13, R35 CrossRef PubMed McGlincy, N.J. and Smith, C.W. (2008) Alternative splicing resulting in nonsense-mediated mRNA decay: what is the meaning of nonsense? Trends Biochem Sci. 33, 385–393 CrossRef PubMed Neu-Yilik, G. and Kulozik, A.E. (2008) NMD: multitasking between mRNA surveillance and modulation of gene expression. Adv. Genet. 62, 185–243 CrossRef PubMed Green, R.E., Lewis, B.P., Hillman, R.T., Blanchette, M., Lareau, L.F., Garnett, A.T., Rio, D.C. and Brenner, S.E. (2003) Widespread predicted nonsense-mediated mRNA decay of alternatively-spliced transcripts of human normal and disease genes. Bioinformatics 19 (Suppl. 1), i118–i121 CrossRef PubMed Boutz, P.L., Stoilov, P., Li, Q., Lin, C.H., Chawla, G., Ostrow, K., Shiue, L., Ares, Jr, M. and Black, D.L. (2007) A post-transcriptional regulatory switch in polypyrimidine tract-binding proteins reprograms alternative splicing in developing neurons. Genes Dev. 21, 1636–1652 CrossRef PubMed  C The

C 2014 Biochemical Society Authors Journal compilation 

1203

1204

Biochemical Society Transactions (2014) Volume 42, part 4

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

 C The

Preitner, N., Quan, J. and Flanagan, J.G. (2013) This message will self-destruct: NMD regulates axon guidance. Cell 153, 1185–1187 CrossRef PubMed Giorgi, C., Yeo, G.W., Stone, M.E., Katz, D.B., Burge, C., Turrigiano, G. and Moore, M.J. (2007) The EJC factor eIF4AIII modulates synaptic strength and neuronal protein expression. Cell 130, 179–191 CrossRef PubMed Sureau, A., Gattoni, R., Dooghe, Y., Stevenin, J. and Soret, J. (2001) SC35 autoregulates its expression by promoting splicing events that destabilize its mRNAs. EMBO J 20, 1785–1796 CrossRef PubMed Wollerton, M.C., Gooding, C., Wagner, E.J., Garcia-Blanco, M.A. and Smith, C.W. (2004) Autoregulation of polypyrimidine tract binding protein by alternative splicing leading to nonsense-mediated decay. Mol. Cell 13, 91–100 CrossRef PubMed Lejeune, F., Cavaloc, Y. and Stevenin, J. (2001) Alternative splicing of intron 3 of the serine/arginine-rich protein 9G8 gene. Identification of flanking exonic splicing enhancers and involvement of 9G8 as a trans-acting factor. J. Biol. Chem. 276, 7850–7858 CrossRef PubMed Kumar, S. and Lopez, A.J. (2005) Negative feedback regulation among SR splicing factors encoded by Rbp1 and Rbp1-like in Drosophila. EMBO J. 24, 2646–2655 CrossRef PubMed 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 with homeostatic control of splicing regulators by alternative splicing and nonsense-mediated decay. Genes Dev. 21, 708–718 CrossRef PubMed Le Guiner, C., Lejeune, F., Galiana, D., Kister, L., Breathnach, R., Stevenin, J. and Del Gatto-Konczak, F. (2001) TIA-1 and TIAR activate splicing of alternative exons with weak 5 splice sites followed by a U-rich stretch on their own pre-mRNAs. J. Biol. Chem. 276, 40638–40646 CrossRef PubMed Rossbach, O., Hung, L.H., Schreiner, S., Grishina, I., Heiner, M., Hui, J. and Bindereif, A. (2009) Auto- and cross-regulation of the hnRNP L proteins by alternative splicing. Mol. Cell Biol. 29, 1442–1451 CrossRef PubMed McGlincy, N.J., Tan, L.Y., Paul, N., Zavolan, M., Lilley, K.S. and Smith, C.W. (2010) Expression proteomics of UPF1 knockdown in HeLa cells reveals autoregulation of hnRNP A2/B1 mediated by alternative splicing resulting in nonsense-mediated mRNA decay. BMC Genomics 11, 565 CrossRef PubMed Sun, S., Zhang, Z., Fregoso, O. and Krainer, A.R. (2012) Mechanisms of activation and repression by the alternative splicing factors RBFOX1/2. RNA 18, 274–283 CrossRef PubMed Rosel-Hillgartner, T.D., Hung, L.H., Khrameeva, E., Le Querrec, P., Gelfand, M.S. and Bindereif, A. (2013) A novel intra-U1 snRNP cross-regulation mechanism: alternative splicing switch links U1C and U1–70K expression. PLoS Genet. 9, e1003856 CrossRef PubMed Saltzman, A.L., Pan, Q. and Blencowe, B.J. (2011) Regulation of alternative splicing by the core spliceosomal machinery. Genes Dev. 25, 373–384 CrossRef PubMed McIlwain, D.R., Pan, Q., Reilly, P.T., Elia, A.J., McCracken, S., Wakeham, A.C., Itie-Youten, A., Blencowe, B.J. and Mak, T.W. (2010) Smg1 is required for embryogenesis and regulates diverse genes via alternative splicing coupled to nonsense-mediated mRNA decay. Proc. Natl. Acad. Sci. U.S.A. 107, 12186–12191 CrossRef PubMed Hurt, J.A., Robertson, A.D. and Burge, C.B. (2013) Global analyses of UPF1 binding and function reveal expanded scope of nonsense-mediated mRNA decay. Genome Res. 23, 1636–1650 CrossRef PubMed Huang, L., Lou, C.H., Chan, W., Shum, E.Y., Shao, A., Stone, E., Karam, R., Song, H.W. and Wilkinson, M.F. (2011) RNA homeostasis governed by cell type-specific and branched feedback loops acting on NMD. Mol. Cell 43, 950–961 CrossRef PubMed Valacca, C., Bonomi, S., Buratti, E., Pedrotti, S., Baralle, F.E., Sette, C., Ghigna, C. and Biamonti, G. (2010) Sam68 regulates EMT through alternative splicing-activated nonsense-mediated mRNA decay of the SF2/ASF proto-oncogene. J. Cell Biol. 191, 87–99 CrossRef PubMed Rahman, L., Bliskovski, V., Kaye, F.J. and Zajac-Kaye, M. (2004) Evolutionary conservation of a 2-kb intronic sequence flanking a tissue-specific alternative exon in the PTBP2 gene. Genomics 83, 76–84 CrossRef PubMed Makeyev, E.V., Zhang, J., Carrasco, M.A. and Maniatis, T. (2007) The microRNA miR-124 promotes neuronal differentiation by triggering brain-specific alternative pre-mRNA splicing. Mol. Cell 27, 435–448 CrossRef PubMed C 2014 Biochemical Society Authors Journal compilation 

64

65

66

67

68

69

70

71

72

73

74

75

76

77

78

79

80 81

82

83

Licatalosi, D.D., Mele, A., Fak, J.J., Ule, J., Kayikci, M., Chi, S.W., Clark, T.A., Schweitzer, A.C., Blume, J.E., Wang, X. et al. (2008) HITS-CLIP yields genome-wide insights into brain alternative RNA processing. Nature 456, 464–469 CrossRef PubMed 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 splicing regulation. Nature 444, 580–586 CrossRef PubMed Qu, M., Aronica, E., Boer, K., Fallmar, D., Kumlien, E., Nister, M., Wester, K., Ponten, F. and Smits, A. (2009) DLG3/SAP102 protein expression in malformations of cortical development: a study of human epileptic cortex by tissue microarray. Epilepsy Res. 84, 33–41 CrossRef PubMed Gurnett, C.A., Veile, R., Zempel, J., Blackburn, L., Lovett, M. and Bowcock, A. (2008) Disruption of sodium bicarbonate transporter SLC4A10 in a patient with complex partial epilepsy and mental retardation. Arch. Neurol. 65, 550–553 CrossRef PubMed Sander, T., Toliat, M.R., Heils, A., Leschik, G., Becker, C., Ruschendorf, F., Rohde, K., Mundlos, S. and Nurnberg, P. (2002) Association of the 867Asp variant of the human anion exchanger 3 gene with common subtypes of idiopathic generalized epilepsy. Epilepsy Res. 51, 249–255 CrossRef PubMed Dickson, B.J. and Zou, Y. (2010) Navigating intermediate targets: the nervous system midline. Cold Spring Harb. Perspect. Biol. 2, a002055 CrossRef PubMed O’Donnell, M., Chance, R.K. and Bashaw, G.J. (2009) Axon growth and guidance: receptor regulation and signal transduction. Annu. Rev. Neurosci. 32, 383–412 CrossRef PubMed Chen, Z., Gore, B.B., Long, H., Ma, L. and Tessier-Lavigne, M. (2008) Alternative splicing of the Robo3 axon guidance receptor governs the midline switch from attraction to repulsion. Neuron 58, 325–332 CrossRef PubMed Jangi, M., Boutz, P.L., Paul, P. and Sharp, P.A. (2014) Rbfox2 controls autoregulation in RNA-binding protein networks. Genes Dev. 28, 637–651 CrossRef PubMed Lestas, I., Vinnicombe, G. and Paulsson, J. (2010) Fundamental limits on the suppression of molecular fluctuations. Nature 467, 174–178 CrossRef PubMed Gehring, N.H., Kunz, J.B., Neu-Yilik, G., Breit, S., Viegas, M.H., Hentze, M.W. and Kulozik, A.E. (2005) Exon-junction complex components specify distinct routes of nonsense-mediated mRNA decay with differential cofactor requirements. Mol. Cell 20, 65–75 CrossRef PubMed Chan, W.K., Huang, L., Gudikote, J.P., Chang, Y.F., Imam, J.S., MacLean, II, J.A. and Wilkinson, M.F. (2007) An alternative branch of the nonsense-mediated decay pathway. EMBO J. 26, 1820–1830 CrossRef PubMed Galante, P.A., Sakabe, N.J., Kirschbaum-Slager, N. and de Souza, S.J. (2004) Detection and evaluation of intron retention events in the human transcriptome. RNA 10, 757–765 CrossRef PubMed Calvanese, V., Mallya, M., Campbell, R.D. and Aguado, B. (2008) Regulation of expression of two LY-6 family genes by intron retention and transcription induced chimerism. BMC Mol. Biol. 9, 81 CrossRef PubMed Buckley, P.T., Lee, M.T., Sul, J.Y., Miyashiro, K.Y., Bell, T.J., Fisher, S.A., Kim, J. and Eberwine, J. (2011) Cytoplasmic intron sequence-retaining transcripts can be dendritically targeted via ID element retrotransposons. Neuron 69, 877–884 CrossRef PubMed Yap, K., Lim, Z.Q., Khandelia, P., Friedman, B. and Makeyev, E.V. (2012) Coordinated regulation of neuronal mRNA steady-state levels through developmentally controlled intron retention. Genes Dev. 26, 1209–1223 CrossRef PubMed Sakabe, N.J. and de Souza, S.J. (2007) Sequence features responsible for intron retention in human. BMC Genomics 8, 59 CrossRef PubMed Amit, M., Donyo, M., Hollander, D., Goren, A., Kim, E., Gelfman, S., Lev-Maor, G., Burstein, D., Schwartz, S., Postolsky, B.G. et al. (2012) Differential GC content between exons and introns establishes distinct strategies of splice-site recognition. Cell Rep. 1, 543–556 CrossRef PubMed Roy, D., Bhanja Chowdhury, J. and Ghosh, S. (2013) Polypyrimidine tract binding protein (PTB) associates with intronic and exonic domains to squelch nuclear export of unspliced RNA. FEBS Lett. 587, 3802–3807 CrossRef PubMed Takemura, R., Takeiwa, T., Taniguchi, I., McCloskey, A. and Ohno, M. (2011) Multiple factors in the early splicing complex are involved in the nuclear retention of pre-mRNAs in mammalian cells. Genes Cells 16, 1035–1049 CrossRef PubMed

RNA UK 2014

84

85 86

87

88

89

90

91

92

93

94

95

96

97

98

99

Ge, Y. and Porse, B.T. (2013) The functional consequences of intron retention: alternative splicing coupled to NMD as a regulator of gene expression. Bioessays 36, 236–243 CrossRef PubMed Linde, L. and Kerem, B. (2011) Nonsense-mediated mRNA decay and cystic fibrosis. Methods Mol. Biol. 741, 137–154 CrossRef PubMed Buratti, E., Chivers, M., Hwang, G. and Vorechovsky, I. (2011) DBASS3 and DBASS5: databases of aberrant 3 - and 5 -splice sites. Nucleic Acids Res. 39, D86–D91 CrossRef PubMed Bidou, L., Allamand, V., Rousset, J.P. and Namy, O. (2012) Sense from nonsense: therapies for premature stop codon diseases. Trends Mol. Med. 18, 679–688 CrossRef PubMed Qadah, T., Finlayson, J., Newbound, C., Pell, N., Pascoe, M., Greenwood, L., Holmes, P., Grey, D., Beilby, J. and Ghassemifar, R. (2012) Molecular and cellular characterization of a new α-thalassemia mutation (HBA2:c.94A>C) generating an alternative splice site and a premature stop codon. Hemoglobin 36, 244–252 CrossRef PubMed Hall, G.W. and Thein, S. (1994) Nonsense codon mutations in the terminal exon of the β-globin gene are not associated with a reduction in β-mRNA accumulation: a mechanism for the phenotype of dominant β-thalassemia. Blood 83, 2031–2037 PubMed Rosenfeld, P.J., Cowley, G.S., McGee, T.L., Sandberg, M.A., Berson, E.L. and Dryja, T.P. (1992) A null mutation in the rhodopsin gene causes rod photoreceptor dysfunction and autosomal recessive retinitis pigmentosa. Nat. Genet. 1, 209–213 CrossRef PubMed Pagani, F., Buratti, E., Stuani, C., Bendix, R., Dork, T. and Baralle, F.E. (2002) A new type of mutation causes a splicing defect in ATM. Nat. Genet. 30, 426–429 CrossRef PubMed Parsons, D.W., McAndrew, P.E., Monani, U.R., Mendell, J.R., Burghes, A.H. and Prior, T.W. (1996) An 11 base pair duplication in exon 6 of the SMN gene produces a type I spinal muscular atrophy (SMA) phenotype: further evidence for SMN as the primary SMA-determining gene. Hum. Mol. Genet. 5, 1727–1732 CrossRef PubMed Perrin-Vidoz, L., Sinilnikova, O.M., Stoppa-Lyonnet, D., Lenoir, G.M. and Mazoyer, S. (2002) The nonsense-mediated mRNA decay pathway triggers degradation of most BRCA1 mRNAs bearing premature termination codons. Hum. Mol. Genet. 11, 2805–2814 CrossRef PubMed Holbrook, J.A., Neu-Yilik, G., Hentze, M.W. and Kulozik, A.E. (2004) Nonsense-mediated decay approaches the clinic. Nat. Genet. 36, 801–808 CrossRef PubMed Wolf, M., Edgren, H., Muggerud, A., Kilpinen, S., Huusko, P., Sorlie, T., Mousses, S. and Kallioniemi, O. (2005) NMD microarray analysis for rapid genome-wide screen of mutated genes in cancer. Cell. Oncol. 27, 169–173 PubMed Karam, R., Carvalho, J., Bruno, I., Graziadio, C., Senz, J., Huntsman, D., Carneiro, F., Seruca, R., Wilkinson, M.F. and Oliveira, C. (2008) The NMD mRNA surveillance pathway downregulates aberrant E-cadherin transcripts in gastric cancer cells and in CDH1 mutation carriers. Oncogene 27, 4255–4260 CrossRef PubMed Peltz, S.W., Welch, E.M., Jacobson, A., Trotta, C.R., Naryshkin, N., Sweeney, H.L. and Bedwell, D.M. (2009) Nonsense suppression activity of PTC124 (ataluren). Proc. Natl. Acad. Sci. U.S.A. 106, E64 CrossRef PubMed Welch, E.M., Barton, E.R., Zhuo, J., Tomizawa, Y., Friesen, W.J., Trifillis, P., Paushkin, S., Patel, M., Trotta, C.R., Hwang, S. et al. (2007) PTC124 targets genetic disorders caused by nonsense mutations. Nature 447, 87–91 CrossRef PubMed Goodier, J.L. and Mayer, J. (2009) PTC124 for cystic fibrosis. Lancet 373, 1426 CrossRef PubMed

100

101

102

103

104

105

106

107

108

109

110

111

112

McElroy, S.P., Nomura, T., Torrie, L.S., Warbrick, E., Gartner, U., Wood, G. and McLean, W.H. (2013) A lack of premature termination codon read-through efficacy of PTC124 (Ataluren) in a diverse array of reporter assays. PLoS Biol 11, e1001593 CrossRef PubMed Rogelj, B., Easton, L.E., Bogu, G.K., Stanton, L.W., Rot, G., Curk, T., Zupan, B., Sugimoto, Y., Modic, M., Haberman, N. et al. (2012) Widespread binding of FUS along nascent RNA regulates alternative splicing in the brain. Sci. Rep. 2, 603 CrossRef PubMed Zhou, Y., Liu, S., Liu, G., Ozturk, A. and Hicks, G.G. (2013) ALS-associated FUS mutations result in compromised FUS alternative splicing and autoregulation. PLoS Genet. 9, e1003895 CrossRef PubMed Ayala, Y.M., De Conti, L., Avendano-Vazquez, S.E., Dhir, A., Romano, M., D’Ambrogio, A., Tollervey, J., Ule, J., Baralle, M., Buratti, E. and Baralle, F.E. (2011) TDP-43 regulates its mRNA levels through a negative feedback loop. EMBO J. 30, 277–288 CrossRef PubMed Buratti, E. and Baralle, F.E. (2011) TDP-43: new aspects of autoregulation mechanisms in RNA binding proteins and their connection with human disease. FEBS J. 278, 3530–3538 CrossRef PubMed Tarpey, P.S., Raymond, F.L., Nguyen, L.S., Rodriguez, J., Hackett, A., Vandeleur, L., Smith, R., Shoubridge, C., Edkins, S., Stevens, C. et al. (2007) Mutations in UPF3B, a member of the nonsense-mediated mRNA decay complex, cause syndromic and nonsyndromic mental retardation. Nat. Genet. 39, 1127–1133 CrossRef PubMed Laumonnier, F., Shoubridge, C., Antar, C., Nguyen, L.S., Van Esch, H., Kleefstra, T., Briault, S., Fryns, J.P., Hamel, B., Chelly, J. et al. (2010) Mutations of the UPF3B gene, which encodes a protein widely expressed in neurons, are associated with nonspecific mental retardation with or without autism. Mol. Psychiatry 15, 767–776 CrossRef PubMed Addington, A.M., Gauthier, J., Piton, A., Hamdan, F.F., Raymond, A., Gogtay, N., Miller, R., Tossell, J., Bakalar, J., Inoff-Germain, G. et al. (2011) A novel frameshift mutation in UPF3B identified in brothers affected with childhood onset schizophrenia and autism spectrum disorders. Mol. Psychiatry 16, 238–239 CrossRef PubMed Nguyen, L.S., Jolly, L., Shoubridge, C., Chan, W.K., Huang, L., Laumonnier, F., Raynaud, M., Hackett, A., Field, M., Rodriguez, J. et al. (2012) Transcriptome profiling of UPF3B/NMD-deficient lymphoblastoid cells from patients with various forms of intellectual disability. Mol. Psychiatry 17, 1103–1115 CrossRef PubMed Kapustin, Y., Chan, E., Sarkar, R., Wong, F., Vorechovsky, I., Winston, R.M., Tatusova, T. and Dibb, N.J. (2011) Cryptic splice sites and split genes. Nucleic Acids Res. 39, 5837–5844 CrossRef PubMed Brown, J.B., Boley, N., Eisman, R., May, G.E., Stoiber, M.H., Duff, M.O., Booth, B.W., Wen, J., Park, S., Suzuki, A.M. et al. (2014) Diversity and dynamics of the Drosophila transcriptome. Nature, doi:10.1038/nature12962 Boley, N., Stoiber, M.H., Booth, B.W., Wan, K.H., Hoskins, R.A., Bickel, P.J., Celniker, S.E. and Brown, J.B. (2014) Genome-guided transcript assembly by integrative analysis of RNA sequence data. Nat. Biotechnol. 32, 341–346 CrossRef PubMed Modic, M., Ule, J. and Sibley, C.R. (2013) CLIPing the brain: studies of protein–RNA interactions important for neurodegenerative disorders. Mol. Cell. Neurosci. 56, 429–435 CrossRef PubMed

Received 14 April 2014 doi:10.1042/BST20140102

 C The

C 2014 Biochemical Society Authors Journal compilation 

1205

Regulation of gene expression through production of unstable mRNA isoforms.

Alternative splicing is universally accredited for expanding the information encoded within the transcriptome. In recent years, several tightly regula...
753KB Sizes 1 Downloads 7 Views