Open access

Recent Advances in Understanding of Alternative Splicing in Neuronal Pathogenesis

Written By

Hung-Hsi Chen, Jung-Chun Lin and Woan-Yuh Tarn

Submitted: 19 November 2010 Published: 29 August 2011

DOI: 10.5772/22477

From the Edited Volume

RNA Processing

Edited by Paula Grabowski

Chapter metrics overview

2,674 Chapter Downloads

View Full Metrics

1. Introduction

The nervous system is an intricate and highly specialized network of neurons. Neuronal differentiation involves complex reprogramming of gene expression. Alternative splicing of precursor mRNAs increases the complexity of transcriptomes and diversifies protein functions at the post-transcriptional level. Indeed, alternative splicing plays an important role in neuronal differentiation, axon guidance, synaptogenesis, synaptic transmission, and plasticity. Because the delicate structure and function of neurons make them particularly susceptible to dysregulation of splicing, aberrant expression or function of splicing factors may cause neuronal disorders. Therefore, it is important to improve our understanding of the mechanisms and physiological functions of alternative splicing regulation in neurons. Regulation of alternative splicing primarily involves the binding of regulatory factors to specific cis-elements of precursor mRNAs, and interplay between splicing factors may lead to fine tuning of splicing regulation, thereby diversifying the cadre of mature products. In addition, transcription rate and the availability of the basal splicing machinery may also influence alternative splicing. Recently, our understanding of the mechanisms underlying alternative splicing have been advanced from studies of several neuronal splicing factors; these studies have utilized genetic knockout or disease models as well as genome-wide analysis of mRNA isoforms. In this chapter, we review current understanding of alternative splicing in neurons.

Advertisement

2. Introduction to alternative splicing

Recent estimates have indicated that as many as 95% of human genes generate alternatively spliced mRNAs (Pan, et al., 2008; E. T. Wang, et al., 2008). Alternative splicing of precursor mRNAs (pre-mRNAs) may alter the coding sequence and hence change the function or stability of the encoded proteins. Moreover, alternative splicing may create pre-mature termination codons within the coding region due to frame shift, thereby inducing mRNA destruction via the nonsense-mediated decay pathway (Isken & Maquat, 2007; McGlincy & Smith, 2008; Moore & Proudfoot, 2009). Alterative splicing may also occur in the 3' untranslated region of a pre-mRNA and thus create or eliminate cis-regulatory elements that may change the kinetics of mRNA decay or translation (Khabar, 2010; Thiele, et al., 2006). Therefore, alternative splicing is a mechanism that not only increases protein diversity but also may post-transcriptionally modulate the level of gene expression. Moreover, a growing body of evidence suggests that coordinated control of alternative splicing of functionally related transcripts allows for proper orchestration of cellular processes and thus the maintenance of homoeostasis (Allen, et al., 2010; Calarco, et al., 2011; Licatalosi & Darnell, 2010).

Alternative splicing plays a critical role in many fundamental biological processes, such as cell differentiation and specification during development and specificity of function in diverse cell types (Keren, et al., 2010; Nilsen & Graveley, 2010). The nervous system also adopts alternative splicing for cell differentiation, morphogenesis, and even for formation of complex neuronal networks and delicate synapse formation/plasticity (Calarco, et al., 2009; Grabowski, 2011; Li, et al., 2007). An extreme case of alternative splicing is the gene encoding Drosophila Down syndrome cell adhesion molecule (Dscam), which potentially could generate >38000 mRNA isoforms by mutually exclusive selection of cassette exons (Hattori, et al., 2008). Dscam encodes neuronal recognition proteins that act as axon guidance receptors. Homotypic interaction between identical Dscam isoforms on opposing membranes causes repulsion between sister neurites (J. W. Park & Graveley, 2007). Therefore, accurate alternative splicing control is critical for establishment of neural circuits in Drosophila. In the mammalian brain, alternative splicing is also an important regulatory mechanism for creating the remarkable capacity for plasticity and adaptation. For example, the splicing-mediated splicing of exon 21 in the ionotropic glutamate receptor N-methyl D-aspartate (NMDA) receptor subunit 1 mRNA affects the membrane trafficking of the NMDA receptor (Ares, 2007). Interestingly, inclusion of exon 21 can be suppressed, likely by the calmodulin-dependent protein kinase IV pathway, which is activated upon cell depolarization (Ares, 2007). Therefore, splicing control provides an intricate and rapid means for regulating mRNA isoform expression.

Alternative splicing is primarily controlled by splicing regulatory factors that bind to cis-elements within exons and/or introns of pre-mRNAs. Their binding may modulate the loading of the spliceosomal components to the splice sites and thereby influence alternative splice site utilization (Chen & Manley, 2009; Witten & Ule, 2011). Several neuron-specific splicing regulatory factors have been discovered; some of them, such as the neuro-oncological ventral antigen (Nova), have been studied intensively (Licatalosi & Darnell, 2010; C. Zhang, et al., 2010). In particular, identification of mRNA targets and potential binding sites of Nova have benefited greatly from recent genome-wide splicing arrays and sequencing technology. Therefore, the study of Nova has provided a detailed picture of a splicing regulatory network as well as the combinatorial action of multiple splicing factors (Nilsen & Graveley, 2010; Z. Wang & Burge, 2008).

The repertoire of splicing factors is adjusted during neuronal differentiation and perhaps for functional specification of different neuronal cell types. For example, the switch in the expression of the polypyrimidine tract-binding protein (PTB) to its neuronal homolog, nPTB, may tune neuronal transcriptomes during differentiation (Coutinho-Mansfield, et al., 2007; Tang, et al., 2011). In addition to RNA-binding factors, altered abundance of basic splicing machinery components may also modulate splice site selection (Calarco, et al., 2011; Saltzman, et al., 2011). The survival of motor neuron (SMN) protein is a key factor for the assembly of spliceosomal small nuclear ribonucleoproteins (snRNPs). SMN deficiency may reduce snRNP abundance and thus influence splicing. Defective Nova and/or SMN proteins are associated with disease (Robert B. Darnell, 2011; Lorson, et al., 2010; Lukong, et al., 2008; G. H. Park, et al., 2010). Therefore, to understand how defects of Nova and SMN induce pathological effects on neurons via misregulated splicing control is an interesting and important question. Therefore, a complete understanding of such diseases will require knowledge of how misregulated splicing control in Nova or SMN causes neuronal pathology.

We hereby discuss our current knowledge of Nova, PTB/nPTB, and SMN with respect to their splicing regulation mechanisms, physiological roles, and functions.

Advertisement

3. Nova

3.1. Nova proteins and expression

Nova was discovered as a target of the autoantibodies in cancer patients with paraneoplastic opsoclonus myoclonus ataxia, a type of rare paraneoplastic neurologic disease that causes motor disorders (Buckanovich, et al., 1993; R. B. Darnell & Posner, 2003; Yang, et al., 1998). The amino acid sequences of the two mammalian Nova proteins (1 and 2) are highly similar throughout their lengths, and both contain three heterogeneous nuclear RNP (hnRNP) K-homology (KH) RNA-binding domains (Buckanovich, et al., 1996; Yang, et al., 1998). In mice, Nova-1 expression is restricted in subcortical regions of developing and mature neurons, whereas Nova-2 is more broadly expressed in the central nervous system (Buckanovich, et al., 1993; Buckanovich, et al., 1996; Yang, et al., 1998). Moreover, Nova-1 and Nova-2 appear to be reciprocally expressed in the nervous system, i.e., the level of Nova-2 is generally low in regions where Nova-1 is abundant, which may reflect their distinct biological functions (Yang, et al., 1998). Nevertheless, Nova-1 and -2 primarily act as splicing regulatory factors and also play a role in alternative polyadenylation and mRNA trafficking in neuronal dendrites (Licatalosi, et al., 2008; Racca, et al., 2010).

3.2. RNA-binding specificity of Nova

The RNA-binding specificity of the Nova proteins has been examined by various methods. Initially, the results of in vitro systematic selection of ligands and RNP immunoprecipitation indicated that Nova-1 binds to long stem-loop RNAs containing UCAU repeats (Buckanovich & Darnell, 1997). Nova-2-selected RNA ligands also appear to form a stem-loop structure encompassing the UCAU motif (Yang, et al., 1998). Structural studies have shown that both Nova-1 and -2 bind to UCAU-containing sequences via their KH3 domains, of which the residues involved in UCAU interaction are conserved (Lewis, et al., 2000). Nevertheless, the KH1 and KH2 domains of both Nova proteins can also bind to the UCAU motif (Musunuru & Darnell, 2004). Perhaps cooperativity between several KH domains promotes protein binding to RNAs (Chmiel, et al., 2006; Valverde, et al., 2008). The similar RNA-binding properties of Nova-1 and Nova-2 suggest that they may regulate common RNA targets.

Next, ultraviolet cross-linking in conjunction with ribonucleoprotein immunoprecipitation (CLIP) was developed to identify in vivo targets of RNA-binding proteins. CLIP analysis has been applied to identify Nova-associated RNA fragments in mouse brain; approximately 340 Nova-1/-2 CLIP tags (~70 nucleotides each) contained, on average, four YCAY repeats (Y, pyrimidine) (Ule, et al., 2003). Detailed analysis of these tag sequences revealed an overrepresentation of YCAU tetramers flanked by pyrimidines. It is currently believed that YCAY repeats are the principal elements in the mRNAs to which Nova proteins bind. More recently, an unbiased method using CLIP combined with high-throughput sequencing (HITS) of RNA was used to identify genome-wide functional protein-RNA interactions (R. B. Darnell, 2010; Ule, et al., 2005a). A HITS-CLIP analysis showed that Nova-2 tags in mouse cortex, as expected, also harbor ~3.6 YCAY repeats per tag (Ule, et al., 2005a). The results of all these CLIP analyses of Nova coincide with previous observations from RNA selection experiments (Licatalosi, et al., 2008).

Moreover, the aforementioned CLIP analyses also revealed that the largest set of Nova tags is located within introns and even flanking alternative exons (Licatalosi, et al., 2008 which is consistent with the primary role of Nova proteins in regulating pre-mRNA splicing (see below). In addition, Nova tags are also found in protein-coding regions and 3' untranslated region of mRNAs. The results of HITS-CLIP experiments have confirmed the high frequency of potential Nova-binding elements in 3' untranslated regions, thus disclosing a role for Nova in alternative polyadenylation and mRNA transport regulation in brain (Licatalosi, et al., 2008; Racca, et al., 2010).

3.3. Alternative splicing regulated by Nova

Earlier studies have shown that two neuronal transcripts, encoding the inhibitory glycine receptor α2 (GlyRα2) and Nova-1 itself, contain the Nova-1-binding elements adjacent to the alternatively spliced exons and indeed bound to the Nova-1 protein (Buckanovich & Darnell, 1997). The role of Nova-1 in splicing regulation was first analyzed by genetic knockout of Nova-1 in mice (Jensen, et al., 2000); in those mice, selection of GlyRα2 exon 3A is diminished. Further experiments demonstrated that Nova-1 binds to three consecutive YCAY repeats in the intron upstream of exon 3A and thereby increases exon 3A inclusion. Alternative exon selection in another neuronal ionotropic receptor, GABAA Rγ2, is also impaired in Nova-1 null mice (Dredge & Darnell, 2003; Jensen, et al., 2000). Nova-1 was subsequently shown to promote GABAA Rγ2 exon 9 inclusion via a distal downstream intronic YCAY-rich splicing enhancer. Notably, the Nova-1 gene itself harbors five YCAY repeats in its exon 4 that are indeed essential for Nova-1 autoregulation (Dredge, et al., 2005). Inclusion of exon 4 in Nova-1 transcripts was increased in haploinsufficient Nova+/– mice. Accordingly, exon 4 inclusion in a Nova-1 splicing reporter was suppressed upon Nova-1 overexpression. In this case, Nova-1 acts via binding to the YCAY repeats in the alternative exon. Therefore, Nova can function either as a positive or negative splicing regulator via binding to intronic or exonic YCAY elements. Further study by swapping of Nova-1-binding sites between various splicing substrates of Nova-1 indicated that the action of Nova-1 is determined by the position of the Nova-binding elements (Dredge, et al., 2005). This positional effect has also been found in other splicing regulators (see below).

Through CLIP and exon junction arrays, a large number of the potential targets of Nova, including some previously defined targets, have been identified (Ule, et al., 2003). Alternative splicing of several candidates was indeed altered in Nova knockout mice, confirming that those transcripts undergo Nova-mediated splicing control. This systematic and genome-wide identification also indicted that Nova may coregulate a set of synaptic and axonal transcripts by controlling alternative splicing. Such coordinated control of alternative exon usage may perhaps provide a powerful means to rapidly modulate synaptic function in response to stimuli.

More recently, HITS-CLIP analysis established functional RNA binding maps of Nova (Licatalosi, et al., 2008). That analysis predicted ~600 differentially spliced exons in brain that are potentially targeted by Nova-2. The transcripts harboring some of the identified exons showed a severe splicing defect in the neocortex of Nova-2 null mice, where Nova-2 is exclusively expressed. However, minor effects were seen in the spinal cord, cerebellum, and midbrain perhaps owing to redundancy of Nova-1 and Nova-2 in these tissues (Licatalosi, et al., 2008). Other work has indicated that Nova proteins regulate alternative splicing across all members of the spectrin-ankyrin-protein 4.1-CASK scaffold complex and the Cav2.2 voltage-gated calcium channels and thereby may finely modulate synaptic function in a coordinated way (Licatalosi, et al., 2008; Ule, et al., 2005b). A genome-wide search in combination with gene ontology analysis revealed that Nova-2-regulated transcripts encode a group of synaptic proteins located at the cell membrane, most often at cell-cell junctions, and are implicated in synapse biogenesis and synaptic transmission. Therefore, Nova-1 and -2 may function analogously in regulating alternative splicing of pre-mRNAs encoded by functionally related genes in neurons.

Drosophila pasilla is the homolog of mammalian Nova proteins (Seshaiah, et al., 2001). Pasilla localizes primarily to nuclear puncta in Drosophila cells, indicating its role in splicing. Interestingly, identification of the mRNA targets of pasilla also revealed an enrichment of YCAY repeats near pasilla-regulated cassette exons (Brooks, et al., 2011). Like Nova, pasilla suppresses alternative exon inclusion when it binds predominantly upstream of the exon, whereas it activates splicing when it binds downstream of the exon. Therefore, the RNA-binding specificity and regulatory activity of Nova orthologs apparently have been preserved throughout evolution (Irimia, et al., 2011; Jelen, et al., 2007).

3.4. Mechanisms of Nova-mediated splicing regulation

The interplay between Nova proteins and cis-elements was revealed by CLIP-based analyses. CLIP followed by genome-wide detection has identified a considerable number of Nova-binding sites and Nova-regulated transcripts. Bioinformatics analysis revealed the mechanism of Nova-mediated alternative splicing regulation (Ule, et al., 2006). Consistent with previous reports using minigene splicing assays, the position of Nova-binding sites in pre-mRNAs determines the effect of Nova proteins on splicing (Dredge, et al., 2005). For a cassette exon, Nova binding to its downstream intronic YCAY clusters enhances exon inclusion, whereas exon skipping occurs when Nova binds either immediately upstream of or within the exon. Therefore, Nova binding to its target cis-elements may result in an asymmetric action on splicing regulation. Indeed, such a positional effect has also been observed for some other splicing regulatory factors (see below, (Konig, et al., 2010; Xue, et al., 2009; Yeo, et al., 2009)).

Bioinformatic analysis of the Nova HITS-CLIP data using Bayesian networks has further provided a comprehensive view of alternative splicing coordinately regulated by Nova and cofactors (C. Zhang, et al., 2010). This analysis initially predicted more than 600 Nova-regulated alternative splicing events. Gene ontology classification supported the hypothesis that Nova may regulate a subgroup of functionally coherent genes involved in synaptic plasticity. This analysis also revealed that the avidity with which Nova binds YCAY clusters may modulate how Nova affects splicing. Moreover, Nova binding to multiple regions may result in a different effect on alternative exon selection from binding to a single region (Figure 1). For example, Nova binding to both the regulated exon and its upstream intron increases the probability of exon exclusion. Moreover, it is estimated that ~15% of Nova targets harbor binding sites for the splicing factors Fox-1 and Fox-2, implying combinatory splicing regulation by Nova and Fox (C. Zhang, et al., 2010). Therefore, an intriguing issue is how the relative location of Nova and Fox binding sites dictates splicing regulation (Chen & Manley, 2009; Licatalosi & Darnell, 2010).

Figure 1.

Establishment of a Nova regulatory network. The binding sites and target mRNAs of Nova proteins are identified through in vitro RNA selection and in vivo crosslinking followed by sequencing or array analysis. In addition, comparative genome-wide transcriptome profiling of Nova knockout mutants also facilitates identification of the transcripts regulated by Nova. Bioinformatic and ontology analysis helps to establish binding and functional maps for Nova. Methods used are denoted by bold type. The bottom diagram shows a model of Nova-mediated splicing control. Positive and negative cis-elements that appear with high frequency in CLIP datasets are depicted by red and blue rectangles, respectively. Red and blue cylinders are alternative and constitutive exons, respectively.

Previous studies have also revealed how Nova proteins may modulate the activity of the spliceosome. For example, Nova binding to exons may interfere with U1 snRNP binding to the 5' splice site, thereby inhibiting splicing (Dredge, et al., 2005). When Nova binds to the downstream intron of a regulated exon, it may facilitate spliceosome assembly and promote exon inclusion (Dredge, et al., 2005). As described above, Nova and Fox proteins may coregulate alternative splicing of a considerable number of transcripts in a cooperative or antagonistic manner (C. Zhang, et al., 2010). Notably, neuronal depolarization can induce exon 19 exclusion of Fox-1, producing a Fox-1 isoform with higher splicing activity (Lee, et al., 2009). Therefore, under certain circumstances this Fox-1 isoform may function coordinately with Nova to modulate splicing of Fox-1/Nova-coregulated mRNAs. Moreover, the neuron-enriched nPTB can antagonize Nova action to increase inclusion of GlyRα2 exon 3A (Polydorides, et al., 2000). The physical and functional interactions between Nova and other splicing regulatory factors certainly complicate Nova splicing networks, but the detailed mechanisms remain to be investigated.

3.5. Cellular signaling pathways affect Nova expression level and activity

In contrast to the abundant information about Nova-target interactions, we have only rudimentary knowledge of whether Nova’s function can be modulated by cellular signaling pathways. However, several reports have indicated that Nova expression may be regulated at different gene expression levels. It has been shown that the neuronal protein embryonic lethal abnormal visual (nELAV) can increase the stability of the Nova-1 mRNA via binding to its AU-rich elements in the highly conserved 3'-untranslated region (Ratti, et al., 2008; Rossi, et al., 2009). In addition, protein kinase C–induced phosphorylation of nELAV can promote Nova-1 translation (Ratti, et al., 2008). Therefore, nELAV can increase Nova-1 abundance. Moreover, nELAV can modulate the splicing activity of Nova-1 on its target pre-mRNAs (Ratti, et al., 2008). Glucocorticoids can also regulate Nova-1-mediated alternative splicing by downregulating Nova-1 (E. Park, et al., 2009). Moreover, cholinergic stimulation may decrease Nova-2 transcripts but increase Nova-1 transcripts in striatum (Jelen, et al., 2010). Therefore, the expression switch between these two Nova proteins may modify Nova activity in neurons. Finally, it is noteworthy that Nova-1 can autoregulate its exon 4 inclusion by acting as a splicing repressor (Dredge, et al., 2005). Because exon 4 contains multiple phosphorylation sites for serine/threonine kinases, it would be interesting to know whether Nova-1 may modulate its own activity via autoregulation of alternative splicing in response to activation of specific cellular signaling pathways.

3.6. Physiological function and pathological implications of Nova

Early studies of Nova-1 and -2 showed that these proteins have a reciprocal expression pattern in the neocortex and hippocampus in postnatal mouse brain and may have slightly different RNA-binding specificity and/or affinity (Buckanovich & Darnell, 1997; Yang, et al., 1998). Genetic knockout studies then provided further hints to their different physiological roles. Nova-1 knockout mice died 7-10 days after birth owing to a motor deficit caused by apoptotic death of spinal and brainstem neurons, indicating that Nova-1 is an essential gene in mice (Jensen, et al., 2000; Yang, et al., 1998). Nova-2 null mice died in the second postnatal week, whereas double Nova knockout caused perinatal death (Ruggiu, et al., 2009; Ule, et al., 2006). Microarray analysis revealed distinct splicing defects for Nova-1 vs. Nova-2 knockout mice (Ule, et al., 2005b; C. Zhang, et al., 2010). These results indicate that these two Nova genes have non-redundant physiological functions.

Identified targets of Nova have implicated a role for Nova in synaptic plasticity. As predicted, long-term potentiation induced by GABAB receptor-mediated slow inhibitory postsynaptic current in hippocampal neurons is abolished in Nova-2 knockout mice (Huang, et al., 2005). Moreover, a recent report showed a migration deficiency in cortical and Purkinje neurons in Nova-2 null mice (Yano, et al., 2010). These observations are consistent with the role of Nova-2 in alternative splicing regulation of GABA receptor subunits and of disabled-1, a regulatory factor of the reelin signaling pathway essential for cell positioning during neurogenesis. Therefore, Nova-2 can regulate neuronal migration and synaptic plasticity via its control of alternative splicing.

Certain Nova targets have been implicated in genetic disorders. For example, reelin-disabled-1 signaling may be associated with epilepsy, schizophrenia, and autism (Frotscher, 2010; Pardo & Eberhart, 2007). Interestingly, Fox-1, a functional partner of the Nova proteins, has been implicated in autism (Martin, et al., 2007; Smith & Sadee, 2011). Indeed, the genes coregulated by Nova and Fox appear to be more frequently associated with autism (C. Zhang, et al., 2010). Therefore, aberrant regulation of Nova proteins may contribute to autism.

Figure 2.

PTB-mediated splicing regulation and autoregulation. Top diagram: PTB suppresses exon inclusion when bound to silencing elements located upstream or downstream of an alternative exon (red cylinder) or even within the exon. Blue cylinders represent constitutive exons. PTB may compete with U2AF for binding to the intron 3' end, prevent U1 snRNP recognition of the 5' splice site, or loop-out the regulated exon. However, PTB can promote exon inclusion when it binds downstream of an alternative exon or close to a strong constitutive splice site (not depicted). This model suggests a positional effect of PTB binding on splicing. (B) PTB activates exon 11 skipping in its own transcript via binding to phylogenetically conserved CU-rich sequences (green hatched boxes) surrounding exon 11, which fits well with the model shown in panel A. The resulting mRNA contains a premature termination codon (PTC) and is degraded by nonsense-mediated decay. Therefore, PTB downregulates the level of its own mRNA.

Advertisement

4. PTB and nPTB

4.1. PTB/nPTB proteins

PTB is a ubiquitously expressed RNA-binding protein containing four RNA recognition motifs with high affinity for CU-rich sequences (Xue, et al., 2009). nPTB is predominant in neurons although also present in other tissues and is remarkably similar to PTB in domain structure and RNA-binding specificity (Spellman, et al., 2007). Both PTB and nPTB primarily function as splicing regulators, and PTB can also regulate translation of specific mRNAs and internal ribosome entry site-mediated translation (Mitchell, et al., 2001; Sawicka, et al., 2008). PTB localizes primarily to the nucleus but can shuttle between the nucleus and cytoplasm (Michael, et al., 1995). Protein kinase A-mediated phosphorylation of PTB can cause its accumulation in the cytoplasm, thereby promoting its cytoplasmic functions such as translation control and RNA transport (Ma, et al., 2007; Xie, et al., 2003).

4.2. Mechanisms of PTB/nPTB-controlled alternative spicing

Multiple mechanisms underlie PTB/nPTB-induced splicing regulation. In general, PTB and nPTB function as splicing inhibitors. Because CU-rich sequences frequently appear in the 3' end of most constitutive introns, binding of PTB to this region interferes with recognition of the 3' splice site by the essential splicing factor, U2AF, thus preventing spliceosome assembly (Sharma, et al., 2005). Moreover, CU-rich elements are also located in other discrete intronic regions surrounding alternative exons. PTB can loop out a regulated exon via binding to both its upstream and downstream intronic CU-rich sequences and forming homomultimers, which thus drives exon exclusion (Lamichhane, et al., 2010). Consistently, a recent genome-wide mapping of PTB-binding sites revealed a position effect for PTB-mediated splicing regulation (Sawicka, et al., 2008) (Figure 2). When PTB binds near an alternative exon, it generally induces exon skipping. However, PTB can also promote exon inclusion when it binds close to a strong constitutive splice site. Besides self-interaction, PTB can also interact with other splicing factors to form complexes that often compete with the splicing machinery, thereby interfering with splicing (Sharma, et al., 2011). A recent report showed that PTB can interact with a pyrimidine-rich loop of U1 snRNA and alter its recognition of the 5' splice site (Coutinho-Mansfield, et al., 2007). nPTB may use similar mechanisms as PTB to suppress exon inclusion. However, nPTB appears to be a weaker splicing suppressor compared with PTB, and nPTB may interact with additional transcripts that are implicated in neuronal activity (Spellman, et al., 2007). Thus, PTB and nPTB have distinct properties in regulating alternative splicing.

4.3. The PTB/nPTB switch during neuronal differentiation

Immunocytochemistry has shown that PTB is detected in neuronal precursor cells as well as non-neuronal lineages of the brain whereas nPTB is specifically expressed in post-mitotic neurons (Boutz, et al., 2007). Therefore, a switch in expression from PTB to nPTB likely occurs during neuronal differentiation. To date, two post-transcriptional mechanisms have been implicated in mutually exclusive expression of PTB and nPTB (Boutz, et al., 2007; Makeyev, et al., 2007). One mechanism involves alternative splicing-coupled nonsense-mediated decay, and the other involves microRNA-mediated translation control. Skipping of exon 11 of PTB and exon 10 of nPTB generates transcripts containing a premature termination codon that subsequently undergo nonsense-mediated decay. PTB is responsible for such exon suppression by binding to highly conserved CU-rich elements flanking the alternative PTB/nPTB exons (Figure 2). Through this activity, PTB may negatively autoregulate its own expression, perhaps to maintain appropriate levels of the protein and restrict the expression of nPTB in non-neuronal cells. In addition, the neuron-specific microRNA miR-124 can directly target to the PTB mRNA and suppresses its translation (Makeyev, et al., 2007). Therefore, PTB expression is down-regulated in neurons, which thus relieves nPTB suppression.

4.4. The PTB/nPTB switch reprograms specific splicing events

Although nPTB and PTB have overlapping function, their different spatial and timely expression patterns in neurons suggest that they have diverse physiological functions in splicing regulation. Indeed, a recent report showed that ~25% of neuron-specific alternative splicing events may result from a decrease of PTB and increase of nPTB during neuronal differentiation (Boutz, et al., 2007). The possible distinct target specificity of PTB and nPTB may be important for establishing unique and neuron-specific splicing programs during neuronal differentiation. Therefore, the PTB/nPTB switch may have evolved as a post-transcriptional mechanism to fine tune the existing program and alter the transcriptome to promote cell differentiation.

As described above, Nova-1 and Nova-2, although having highly similar sequences, contribute to neuron-specific splicing in different types of neurons (C. Zhang, et al., 2010). Given that Nova-1 can drive its own exon 4 skipping, the Nova-1/Nova-2 reciprocal expression may in part proceed through a negative feedback control mechanism similar to that used by PTB/nPTB (Coutinho-Mansfield, et al., 2007; Dredge, et al., 2005). Therefore, the switch between two highly similar but still distinct splicing factors may provide a potent and rapid means to adjust cellular function in a specific environment.

Advertisement

5. SMN

5.1. SMN genes and expression

Spinal muscular atrophy (SMA) is an autosomal recessive disorder characterized by degeneration of lower motor neurons in the spinal cord with subsequent muscle atrophy (Pearn, 1978). SMA is caused by deletions or mutations of the survival of motor neuron 1 (SMN1) gene (Lefebvre, et al., 1995). In human, the SMN2 gene is almost identical to SMN1 but contains a C to T transition at position 6 in exon 7. This nucleotide change induces SMN2 exon 7 skipping during splicing and results in an unstable truncated SMN protein (Cho & Dreyfuss, 2010). Therefore, SMN2 fails to produce a sufficient amount of functional SMN protein to compensate for the loss of SMN1 (Lorson, et al., 1999; Monani, et al., 1999). Multiple factors have been proposed to regulate exon 7 inclusion/exclusion of the SMN transcripts. In principle, SMN1 exon 7 harbors a splicing enhancer for the splicing activator SF2/ASF, which promotes exon 7 inclusion in the SMN1 transcript, whereas the C to U change in SMN2 pre-mRNA disrupts the binding of SF2/ASF but creates a recognition site for the suppressor hnRNP A1 that excludes exon 7 (Cartegni, et al., 2006; Cartegni & Krainer, 2002; Kashima & Manley, 2003; Kashima, et al., 2007). Besides, other SMN regulators may function via direct binding to exon 7 or even to intronic elements or through a protein complex to modulate exon 7 selection (Doktor, et al., 2011; Nlend Nlend, et al., 2010; Pedrotti & Sette, 2010). SMN splicing regulation has been reviewed elsewhere; this section thus focuses on SMN function in pre-mRNA splicing and regulation.

5.2. SMN and its cellular localization

SMN expression is not restricted to neurons, and in fact it is expressed in all cell types Unlike Nova and PTB, SMN lacks a typical RNA-binding domain but contains a Tudor domain that mediates its interaction with the Sm proteins of spliceosomal snRNPs. Indeed, SMN participates in snRNP biogenesis, which is an important housekeeping function, and also in splicing, transcription, and neuronal mRNA trafficking (Burghes & Beattie, 2009; Coady & Lorson, 2011) (Figure 3). In the nucleus, SMN is particularly concentrated in discrete nuclear bodies that are very close to Cajal bodies (Carvalho, et al., 1999; Young, et al., 2000). This nuclear localization pattern suggests a role for SMN in nuclear snRNP maturation and regeneration. In neurons, SMN forms cytoplasmic granules in neurites and growth cones and participates in active bidirectional transport of mRNAs (Fallini, et al., 2011; Todd, et al., 2010; H. Zhang, et al., 2006). SMN deficiency disrupts Cajal bodies and impairs mRNA trafficking in motor neuron axons (Girard, et al., 2006; Rossoll, et al., 2003; Shpargel & Matera, 2005 indicating that the diverse subcellular distribution of SMN is functionally important. Moreover, a recent observation that SMN associates and colocalizes with the α subunit of “coatomer”, a protein coat for vesicles that mediate intracellular transport, indicates a role for Golgi-associated COPI vesicles in SMN transport (Peter, et al., 2011).

Figure 3.

Functions of SMN in neurons. In the cytoplasm, SMN forms a protein complex to facilitate snRNP biogenesis. Moreover, SMN along with other RNA-binding proteins participates in mRNA transport; the role of COPI-containing vesicles in SMN transport is unclear. In the nucleus, SMN is highly concentrated in Cajal bodies and possibly plays a role in snRNP maturation and multi-snRNP assembly. SMN deficiency may indirectly induce aberrant splicing.

5.3. SMN in snRNP biogenesis

The role of SMN in snRNP biogenesis has been well characterized. Assembly of the heptameric Sm cores with each spliceosomal snRNA occurs in a highly ordered manner (Burghes & Beattie, 2009; Cauchi, 2010; Coady & Lorson, 2011). Initially, the chaperon factor pICln brings methylated Sm protein subcomplexes to the SMN complex, which is composed of SMN, Gemin2-8, and unr-interacting proteins. The SMN complex facilitates the arrangement of the Sm proteins on a single-stranded region of snRNA to form a ring-shaped structure of the snRNP. An initial report showed that ~80% depletion of SMN by RNA interference in cultured mammalian cells had no significant loss-of-function effect on snRNP assembly (Girard, et al., 2006). Perhaps excess SMN complex exists in cells to maintain normal levels of the basal splicing machinery. However, further reduction of SMN differentially impaired snRNP assembly (Gabanella, et al., 2007; Workman, et al., 2009; Z. Zhang, et al., 2008). In particular, the levels of several U12-type spliceosomal snRNAs decreased more significantly in SMA mice, and formation of the U12-type tri-snRNPs was also impeded in lymphoblasts derived from SMA patients (Gabanella, et al., 2007; Workman, et al., 2009; Z. Zhang, et al., 2008). The latter observation is also in accordance with the assumption that SMN is possibly involved in tri-snRNP assembly in Cajal bodies (Carvalho, et al., 1999; Novotny, et al., 2011; Young, et al., 2000). Although direct evidence for defective snRNP assembly or reduced snRNP levels in SMA pathogenesis is lacking, one hint has been provided by the observation that knockdown of SMN or other snRNP assembly factors in zebrafish causes motor axon defects (Winkler, et al., 2005).

5.4. SMN in pre-mRNA splicing

It is clear that SMN has an essential function for snRNP biogenesis, and possibly that insufficient SMN causes various degrees of snRNP assembly defects. However, whether loss of SMN affects splicing of a wide range or a specific set of transcripts and whether the effect of SMN in splicing, if any, indeed results from impaired snRNP assembly have just begun to be investigated.

Exon array analysis has shown that splicing of numerous transcripts is affected in various tissues of late-symptomatic SMA mice (Z. Zhang, et al., 2008). It is unclear whether such a widespread splicing defect is caused by reduced levels of SMN. It is suspected that the decrease in SMN level may affect splicing of specific transcripts in motor neurons that are most vulnerable to degeneration in SMA (Briese, et al., 2005; Liu, et al., 2010; Monani, 2005). A recent report showed that the tri-snRNP of the U12-type splicing machinery is most affected in SMA patients and, consistently, the splicing of a subgroup of U12-type introns is affected (Boulisfane, et al., 2011). Because U12-type introns are present in a number of genes involved in cytoskeletal organization, defects in their excision may impair motor neuron function. Finally, the error rate of exon inclusion/skipping is higher in fibroblasts of SMA patients, perhaps owing to poor recognition of the splice sites by a low abundance of functional snRNPs (Fox-Walsh & Hertel, 2009).

5.5. How does SMN deficiency cause SMA pathogenesis

To date, two plausible possibilities have been raised to explain how SMN deficiency causes specific neurological defects of motor neurons. First, as discussed above, inefficient snRNP assembly may affect the splicing of a specific set of transcripts that are critical to motor neuron functions. However, issues such as which transcripts are most sensitive to SMN deficiency and whether their splicing defects lead to SMA pathogenesis remain to be investigated. In addition, it has been shown that SMN forms RNA granules with other RNA-binding proteins such as hnRNP R to deliver β-actin mRNA in motor axons (Glinka, et al., 2010; Rossoll, et al., 2003). A recent report showed that clustering of Cav2.2 calcium channels is impaired in axonal growth cones of SMA animals, and such a defect can be restored by rescue of SMN expression (Jablonka, et al., 2007). It is possible that SMN plays a role in actin filament formation via β-actin mRNA trafficking, and such an activity of SMN is critical for motor neuron functions.

SMA patients and animal models show a variable extent of defects in neuromuscular junction functions, axonal arborization, synaptic transport, and neurodevelopment. Such complexity may result from different residual levels of SMN in different systems as well as multiple cellular functions and interacting partners of SMN (Boyer, et al., 2010; Burghes & Beattie, 2009; Cauchi, 2010; Wu, et al., 2011). Nevertheless, restoration of SMN levels or function is certainly a primary therapeutic strategy for SMA treatment (Kolb & Kissel, 2011). For example, activation of SMN2 transcription and restoration of SMN2 splicing ameliorate symptoms of SMA mice and thus provide promise for future SMA treatment.

Advertisement

6. Conclusions

Our understanding of splicing regulation mechanisms and splicing regulatory networks has been advanced substantially by recent studies using gene inactivation techniques and genome-wide experimental and computational examination of alternative splicing events. In the past decade, CLIP in conjunction with various types of mRNA identification systems has been used extensively for in vitro study of splicing factors and their regulation mechanisms. Ablation of splicing factors in cultured cells by RNA interference has also been widely used for mechanistic studies of alternative splicing of endogenous or reporter minigene transcripts. Nevertheless, we are still at the beginning of our understanding of the mechanistic and, in particular, physiological aspects of alternative splicing regulation.

Our understanding of the physiological consequences of alternative splicing still largely relies on genetic approaches. For example, knockout of splicing factors in animals in combination of mRNA isoform comparison can facilitate the identification of their in vivo targets and biological functions. Study of disease-related splicing factors can in particular provide insights into pathogenesis of aberrant splicing. Moreover, knock-in or knockout of specific mRNA isoforms can help to unveil their functional consequence(s), which is poorly understood, and may even allow delineation of causal effects (Moroy & Heyd, 2007). However, progress has been relatively slow owing to limitations of genetic techniques in mammalian systems. At present, efficient recombination technologies are being developed to facilitate high-throughput gene knockout in embryonic stem cells (Valenzuela, et al., 2003 which may allow large-scale analysis of biological functions of splicing factors as well as mRNA isoforms. Besides more efficient/convenient genetic tools, high-throughput whole-transcriptome sequencing and extensive bioinformatics tools have proved their advantage. With these techniques, we will begin to establish a more accurate paradigm for mRNA splicing regulatory networks with physiological significance.

Advertisement

Acknowledgments

We acknowledge support from intramural funds of the Institute of Biomedical Sciences, Academia Sinica, and the Academia Sinica Investigator Award to W.-Y. Tarn.

References

  1. 1. AllenS. E.DarnellR. B.LipscombeD.2010The neuronal splicing factor Nova controls alternative splicing in N-type and P-type CaV2 calcium channels. Channels (Austin), 46Nov-Dec, 201 483489
  2. 2. AresM.Jr 2007Sing the genome electric: excited cells adjust their splicing. PLoS Biol, 52Feb, 200 e55
  3. 3. BoulisfaneN.CholezaM.RageF.NeelH.SoretJ.BordonneR.2011Impaired minor tri-snRNP assembly generates differential splicing defects of U12-type introns in lymphoblasts derived from a type I SMA patient. Hum Mol Genet, 204Feb 15, 201 641648
  4. 4. BoutzP. L.StoilovP.LiQ.LinC. H.ChawlaG.OstrowK.ShiueL.AresM.Jr BlackD. L.2007A post-transcriptional regulatory switch in polypyrimidine tract-binding proteins reprograms alternative splicing in developing neurons. Genes Dev, 2113Jul 1, 200 16361652
  5. 5. BoyerJ. G.BowermanM.KotharyR.2010The many faces of SMN: deciphering the function critical to spinal muscular atrophy pathogesis. Future Neurol, 56873890
  6. 6. BrieseM.EsmaeiliB.SattelleD. B.2005Is spinal muscular atrophy the result of defects in motor neuron processes? Bioessays, 279Sep, 200 946957
  7. 7. BrooksA. N.YangL.DuffM. O.HansenK. D.ParkJ. W.DudoitS.BrennerS. E.GraveleyB. R.2011Conservation of an RNA regulatory map between Drosophila and mammals. Genome Res, 212Feb, 201 193202
  8. 8. BuckanovichR. J.DarnellR. B.1997The neuronal RNA binding protein Nova-1 recognizes specific RNA targets in vitro and in vivo. Mol Cell Biol, 176Jun, 199 31943201
  9. 9. BuckanovichR. J.PosnerJ. B.DarnellR. B.1993Nova, the paraneoplastic Ri antigen, is homologous to an RNA-binding protein and is specifically expressed in the developing motor system. Neuron, 114Oct, 199 657672
  10. 10. BuckanovichR. J.YangY. Y.DarnellR. B.1996The onconeural antigen Nova-1 is a neuron-specific RNA-binding protein, the activity of which is inhibited by paraneoplastic antibodies. J Neurosci, 163Feb 1, 199 11141122
  11. 11. BurghesA. H.BeattieC. E.2009Spinal muscular atrophy: why do low levels of survival motor neuron protein make motor neurons sick? Nat Rev Neurosci, 108Aug, 200 597609
  12. 12. CalarcoJ. A.SuperinaS.O’HanlonD.GabutM.RajB.PanQ.SkalskaU.ClarkeL.GelinasD.van der KooyD.ZhenM.CirunaB.BlencoweB. J.2009Regulation of vertebrate nervous system alternative splicing and development by an SR-related protein. Cell, 1385Sep 4, 200 898910
  13. 13. CalarcoJ. A.ZhenM.BlencoweB. J.2011Networking in a global world: Establishing functional connections between neural splicing regulators and their target transcripts. RNA, 175May, 201 775791
  14. 14. CartegniL.HastingsM. L.CalarcoJ. A.de StanchinaE.KrainerA. R.2006Determinants of exon 7 splicing in the spinal muscular atrophy genes, SMN1 and SMN2. Am J Hum Genet, 781Jan, 200 6377
  15. 15. CartegniL.KrainerA. R.2002Disruption of an SF2/ASF-dependent exonic splicing enhancer in SMN2 causes spinal muscular atrophy in the absence of SMN1. Nat Genet, 304Apr, 200 377384
  16. 16. CarvalhoT.AlmeidaF.CalapezA.LafargaM.BercianoM. T.Carmo-FonsecaM.1999The spinal muscular atrophy disease gene product, SMN: A link between snRNP biogenesis and the Cajal (coiled) body. J Cell Biol, 1474Nov 15, 199 715728
  17. 17. CauchiR. J.2010SMN and Gemins: ‘we are family’... or are we?: insights into the partnership between Gemins and the spinal muscular atrophy disease protein SMN. Bioessays, 3212Dec, 201 10771089
  18. 18. ChenM.ManleyJ. L.2009Mechanisms of alternative splicing regulation: insights from molecular and genomics approaches. Nat Rev Mol Cell Biol, 1011Nov, 200 741754
  19. 19. ChmielN. H.RioD. C.DoudnaJ. A.2006Distinct contributions of KH domains to substrate binding affinity of Drosophila P-element somatic inhibitor protein. RNA, 122Feb, 200 283291
  20. 20. ChoS.DreyfussG.2010A degron created by SMN2 exon 7 skipping is a principal contributor to spinal muscular atrophy severity. Genes Dev, 245Mar 1, 201 438442
  21. 21. CoadyT. H.LorsonC. L.2011SMN in spinal muscular atrophy and snRNP biogenesis. Wiley Interdisciplinary Reviews: RNA, 201 pp. n/a-n/a
  22. 22. Coutinho-MansfieldG. C.XueY.ZhangY.FuX. D.2007PTB/nPTB switch: a post-transcriptional mechanism for programming neuronal differentiation. Genes Dev, 2113Jul 1, 200 15731577
  23. 23. DarnellR. B.2010HITS-CLIP: panoramic views of protein-RNA regulation in living cells. WIREs RNA, 1266286
  24. 24. DarnellR. B.2011RNA regulation in Neurodegeneration and Cancer, In: Two Faces of Evil: Cancer and Neurodegeneration, Curran, T. & Christen, Y., 103111Springer Berlin Heidelberg, 978-3-642-16602-0,
  25. 25. DarnellR. B.PosnerJ. B.2003Paraneoplastic syndromes involving the nervous system. N Engl J Med, 34916Oct 16, 200 15431554
  26. 26. DoktorT. K.SchroederL. D.VestedA.PalmfeldtJ.AndersenH. S.GregersenN.AndresenB. S.2011SMN2 exon 7 splicing is inhibited by binding of hnRNP A1 to a common ESS motif that spans the 3’ splice site. Hum Mutat, 322Feb, 201 220230
  27. 27. DredgeB. K.DarnellR. B.2003Nova regulates GABA(A) receptor gamma2 alternative splicing via a distal downstream UCAU-rich intronic splicing enhancer. Mol Cell Biol, 2313Jul, 200 46874700
  28. 28. DredgeB. K.StefaniG.EngelhardC. C.DarnellR. B.2005Nova autoregulation reveals dual functions in neuronal splicing. EMBO J, 248Apr 20, 200 16081620
  29. 29. FalliniC.ZhangH.SuY.SilaniV.SingerR. H.RossollW.BassellG. J.2011The Survival of Motor Neuron (SMN) Protein Interacts with the mRNA-Binding Protein HuD and Regulates Localization of Poly(A) mRNA in Primary Motor Neuron Axons. J Neurosci, 3110Mar 9, 201 39143925
  30. 30. Fox-WalshK. L.HertelK. J.2009Splice-site pairing is an intrinsically high fidelity process. Proc Natl Acad Sci U S A, 1066Feb 10, 200 17661771
  31. 31. FrotscherM.2010Role for Reelin in stabilizing cortical architecture. Trends Neurosci, 339Sep, 201 407414
  32. 32. GabanellaF.ButchbachM. E.SaievaL.CarissimiC.BurghesA. H.PellizzoniL.2007Ribonucleoprotein assembly defects correlate with spinal muscular atrophy severity and preferentially affect a subset of spliceosomal snRNPs. PLoS One, 29e921
  33. 33. GirardC.NeelH.BertrandE.BordonneR.2006Depletion of SMN by RNA interference in HeLa cells induces defects in Cajal body formation. Nucleic Acids Res, 341029252932
  34. 34. GlinkaM.HerrmannT.FunkN.HavlicekS.RossollW.WinklerC.SendtnerM.2010The heterogeneous nuclear ribonucleoprotein-R is necessary for axonal beta-actin mRNA translocation in spinal motor neurons. Hum Mol Genet, 1910May 15, 201 19511966
  35. 35. GrabowskiP.2011Alternative splicing takes shape during neuronal development. Curr Opin Genet Dev, (Apr 19, 201 pp.
  36. 36. HattoriD.MillardS. S.WojtowiczW. M.ZipurskyS. L.2008Dscam-mediated cell recognition regulates neural circuit formation. Annu Rev Cell Dev Biol, 24597620
  37. 37. HuangC. S.ShiS. H.UleJ.RuggiuM.BarkerL. A.DarnellR. B.JanY. N.JanL. Y.2005Common molecular pathways mediate long-term potentiation of synaptic excitation and slow synaptic inhibition. Cell, 1231Oct 7, 200 105118
  38. 38. IrimiaM.DenucA.BurgueraD.SomorjaiI.Martin-DuranJ. M.GenikhovichG.Jimenez-DelgadoS.TechnauU.RoyS. W.MarfanyG.Garcia-FernandezJ.2011Stepwise assembly of the Nova-regulated alternative splicing network in the vertebrate brain. Proc Natl Acad Sci U S A, 10813Mar 29, 201 53195324
  39. 39. IskenO.MaquatL. E.2007Quality control of eukaryotic mRNA: safeguarding cells from abnormal mRNA function. Genes Dev, 2115Aug 1, 200 18331856
  40. 40. JablonkaS.BeckM.LechnerB. D.MayerC.SendtnerM.2007Defective Ca2+ channel clustering in axon terminals disturbs excitability in motoneurons in spinal muscular atrophy. J Cell Biol, 1791Oct 8, 200 139149
  41. 41. JelenN.UleJ.ZivinM.2010Cholinergic regulation of striatal Nova mRNAs. Neuroscience, 1692Aug 25, 201 619627
  42. 42. JelenN.UleJ.ZivinM.DarnellR. B.2007Evolution of Nova-dependent splicing regulation in the brain. PLoS Genet, 310Oct, 200 18381847
  43. 43. JensenK. B.DredgeB. K.StefaniG.ZhongR.BuckanovichR. J.OkanoH. J.YangY. Y.DarnellR. B.2000Nova-1 regulates neuron-specific alternative splicing and is essential for neuronal viability. Neuron, 252Feb, 200 359371
  44. 44. KashimaT.ManleyJ. L.2003A negative element in SMN2 exon 7 inhibits splicing in spinal muscular atrophy. Nat Genet, 344Aug, 200 460463
  45. 45. KashimaT.RaoN.DavidC. J.ManleyJ. L.2007hnRNP A1 functions with specificity in repression of SMN2 exon 7 splicing. Hum Mol Genet, 1624Dec 15, 200 31493159
  46. 46. KerenH.Lev-MaorG.AstG.2010Alternative splicing and evolution: diversification, exon definition and function. Nat Rev Genet, 115May, 201 345355
  47. 47. KhabarK. S.2010Post-transcriptional control during chronic inflammation and cancer: a focus on AU-rich elements. Cell Mol Life Sci, 6717Sep, 201 29372955
  48. 48. KolbS. J.KisselJ. T.2011Spinal Muscular Atrophy: A Timely Review. Arch Neurol, (Apr 11, 201 pp.
  49. 49. KonigJ.ZarnackK.RotG.CurkT.KayikciM.ZupanB.TurnerD. J.LuscombeN. M.UleJ.2010iCLIP reveals the function of hnRNP particles in splicing at individual nucleotide resolution. Nat Struct Mol Biol, 177Jul, 201 909915
  50. 50. LamichhaneR.DaubnerG. M.Thomas-CrusellsJ.AuweterS. D.ManatschalC.AustinK. S.ValniukO.AllainF. H.RuedaD.2010RNA looping by PTB: Evidence using FRET and NMR spectroscopy for a role in splicing repression. Proc Natl Acad Sci U S A, 1079Mar 2, 201 41054110
  51. 51. LeeJ. A.TangZ. Z.BlackD. L.2009An inducible change in Fox-1/A2BP1 splicing modulates the alternative splicing of downstream neuronal target exons. Genes Dev, 2319Oct 1, 200 22842293
  52. 52. LefebvreS.BurglenL.ReboulletS.ClermontO.BurletP.ViolletL.BenichouB.CruaudC.MillasseauP.ZevianiM.et al.1995Identification and characterization of a spinal muscular atrophy-determining gene. Cell, 801Jan 13, 199 155165
  53. 53. LewisH. A.MusunuruK.JensenK. B.EdoC.ChenH.DarnellR. B.BurleyS. K.2000Sequence-specific RNA binding by a Nova KH domain: implications for paraneoplastic disease and the fragile X syndrome. Cell, 1003Feb 4, 200 323332
  54. 54. LiQ.LeeJ. A.BlackD. L.2007Neuronal regulation of alternative pre-mRNA splicing. Nat Rev Neurosci, 811Nov, 200 819831
  55. 55. LicatalosiD. D.DarnellR. B.2010RNA processing and its regulation: global insights into biological networks. Nat Rev Genet, 111Jan, 201 7587
  56. 56. LicatalosiD. D.MeleA.FakJ. J.UleJ.KayikciM.ChiS. W.ClarkT. A.SchweitzerA. C.BlumeJ. E.WangX.DarnellJ. C.DarnellR. B.2008HITS-CLIP yields genome-wide insights into brain alternative RNA processing. Nature, 4567221Nov 27, 200 464469
  57. 57. LiuH.ShafeyD.MooresJ. N.KotharyR.2010Neurodevelopmental consequences of Smn depletion in a mouse model of spinal muscular atrophy. J Neurosci Res, 881Jan, 201 111122
  58. 58. LorsonC. L.HahnenE.AndrophyE. J.WirthB.1999A single nucleotide in the SMN gene regulates splicing and is responsible for spinal muscular atrophy. Proc Natl Acad Sci U S A, 9611May 25, 199 63076311
  59. 59. LorsonC. L.RindtH.ShababiM.2010Spinal muscular atrophy: mechanisms and therapeutic strategies. Hum Mol Genet, 19R1Apr 15, 201 R111R118
  60. 60. LukongK. E.ChangK. W.KhandjianE. W.RichardS.2008RNA-binding proteins in human genetic disease. Trends Genet, 248Aug, 200 416425
  61. 61. MaS.LiuG.SunY.XieJ.2007Relocalization of the polypyrimidine tract-binding protein during PKA-induced neurite growth. Biochim Biophys Acta, 17736Jun, 200 912923
  62. 62. MakeyevE. V.ZhangJ.CarrascoM. A.ManiatisT.2007The MicroRNA miR-124 promotes neuronal differentiation by triggering brain-specific alternative pre-mRNA splicing. Mol Cell, 273Aug 3, 200 435448
  63. 63. MartinC. L.DuvallJ. A.IlkinY.SimonJ. S.ArreazaM. G.WilkesK.Alvarez-RetuertoA.WhichelloA.PowellC. M.RaoK.CookE.GeschwindD. H.2007Cytogenetic and molecular characterization of A2BP1/FOX1 as a candidate gene for autism. Am J Med Genet B Neuropsychiatr Genet, 144B7Oct 5, 200 869876
  64. 64. Mc GlincyN. J.SmithC. W.2008Alternative splicing resulting in nonsense-mediated mRNA decay: what is the meaning of nonsense? Trends Biochem Sci, 338Aug, 200 385393
  65. 65. MichaelW. M.SiomiH.ChoiM.Pinol-RomaS.NakielnyS.LiuQ.DreyfussG.1995Signal sequences that target nuclear import and nuclear export of pre-mRNA-binding proteins. Cold Spring Harb Symp Quant Biol, 60663668
  66. 66. MitchellS. A.BrownE. C.ColdwellM. J.JacksonR. J.WillisA. E.2001Protein factor requirements of the Apaf-1 internal ribosome entry segment: roles of polypyrimidine tract binding protein and upstream of N-ras. Mol Cell Biol, 2110May, 200 33643374
  67. 67. MonaniU. R.2005Spinal muscular atrophy: a deficiency in a ubiquitous protein; a motor neuron-specific disease. Neuron, 486Dec 22, 200 885896
  68. 68. MonaniU. R.LorsonC. L.ParsonsD. W.PriorT. W.AndrophyE. J.BurghesA. H.Mc PhersonJ. D.1999A single nucleotide difference that alters splicing patterns distinguishes the SMA gene SMN1 from the copy gene SMN2. Hum Mol Genet, 87Jul, 199 11771183
  69. 69. MooreM. J.ProudfootN. J.2009Pre-mRNA processing reaches back to transcription and ahead to translation. Cell, 1364Feb 20, 200 688700
  70. 70. MoroyT.HeydF.2007The impact of alternative splicing in vivo: mouse models show the way. RNA, 138Aug, 200 11551171
  71. 71. MusunuruK.DarnellR. B.2004Determination and augmentation of RNA sequence specificity of the Nova K-homology domains. Nucleic Acids Res, 321648524861
  72. 72. NilsenT. W.GraveleyB. R.2010Expansion of the eukaryotic proteome by alternative splicing. Nature, 4637280Jan 28, 201 457463
  73. 73. NlendNlend. R.MeyerK.SchumperliD.2010Repair of pre-mRNA splicing: prospects for a therapy for spinal muscular atrophy. RNA Biol, 74Nov 19, 201 430440
  74. 74. NovotnyI.BlazikovaM.StanekD.HermanP.MalinskyJ.2011In vivo kinetics of U4/U6.U5 tri-snRNP formation in Cajal bodies. Mol Biol Cell, 224Feb 15, 201 513523
  75. 75. PanQ.ShaiO.LeeL. J.FreyB. J.BlencoweB. J.2008Deep surveying of alternative splicing complexity in the human transcriptome by high-throughput sequencing. Nat Genet, 4012Dec, 200 14131415
  76. 76. PardoC. A.EberhartC. G.2007The neurobiology of autism. Brain Pathol, 174Oct, 200 434447
  77. 77. ParkE.LeeM. S.BaikS. M.ChoE. B.SonG. H.SeongJ. Y.LeeK. H.KimK.2009Nova-1 mediates glucocorticoid-induced inhibition of pre-mRNA splicing of gonadotropin-releasing hormone transcripts. J Biol Chem, 28419May 8, 200 1279212800
  78. 78. ParkG. H.KariyaS.MonaniU. R.2010Spinal muscular atrophy: new and emerging insights from model mice. Curr Neurol Neurosci Rep, 102Mar, 201 108117
  79. 79. ParkJ. W.GraveleyB. R.2007Complex alternative splicing. Adv Exp Med Biol, 6235063
  80. 80. PearnJ.1978Incidence, prevalence, and gene frequency studies of chronic childhood spinal muscular atrophy. J Med Genet, 156Dec, 197 409413
  81. 81. PedrottiS.SetteC.2010Spinal muscular atrophy: a new player joins the battle for SMN2 exon 7 splicing. Cell Cycle, 919Oct 1, 201 38743879
  82. 82. PeterC. J.EvansM.ThayanithyV.Taniguchi-IshigakiN.BachI.KolpakA.BassellG. J.RossollW.LorsonC. L.BaoZ. Z.AndrophyE. J.2011The COPI vesicle complex binds and moves with survival motor neuron within axons. Hum Mol Genet, 209May 1, 201 17011711
  83. 83. PolydoridesA. D.OkanoH. J.YangY. Y.StefaniG.DarnellR. B.2000A brain-enriched polypyrimidine tract-binding protein antagonizes the ability of Nova to regulate neuron-specific alternative splicing. Proc Natl Acad Sci U S A, 9712Jun 6, 200 63506355
  84. 84. RaccaC.GardiolA.EomT.UleJ.TrillerA.DarnellR. B.2010The Neuronal Splicing Factor Nova Co-Localizes with Target RNAs in the Dendrite. Front Neural Circuits, 45
  85. 85. RattiA.FalliniC.ColombritaC.PascaleA.LaforenzaU.QuattroneA.SilaniV.2008Post-transcriptional regulation of neuro-oncological ventral antigen 1 by the neuronal RNA-binding proteins ELAV. J Biol Chem, 28312Mar 21, 200 75317541
  86. 86. RossiD.AmadioM.CarnevaleBaraglia. A.AzzolinaO.RattiA.GovoniS.PascaleA.CollinaS.2009Discovery of small peptides derived from embryonic lethal abnormal vision proteins structure showing RNA-stabilizing properties. J Med Chem, 5216Aug 27, 200 50175019
  87. 87. RossollW.JablonkaS.AndreassiC.KroningA. K.KarleK.MonaniU. R.SendtnerM.2003Smn, the spinal muscular atrophy-determining gene product, modulates axon growth and localization of beta-actin mRNA in growth cones of motoneurons. J Cell Biol, 1634Nov 24, 200 801812
  88. 88. RuggiuM.HerbstR.KimN.JevsekM.FakJ. J.MannM. A.FischbachG.BurdenS. J.DarnellR. B.2009Rescuing Z+ agrin splicing in Nova null mice restores synapse formation and unmasks a physiologic defect in motor neuron firing. Proc Natl Acad Sci U S A, 1069Mar 3, 200 35133518
  89. 89. SaltzmanA. L.PanQ.BlencoweB. J.2011Regulation of alternative splicing by the core spliceosomal machinery. Genes Dev, 254Feb 15, 201 373384
  90. 90. SawickaK.BushellM.SpriggsK. A.WillisA. E.2008Polypyrimidine-tract-binding protein: a multifunctional RNA-binding protein. Biochem Soc Trans, 36No.Pt 4, (Aug, 200 641647
  91. 91. SeshaiahP.MillerB.MyatM. M.AndrewD. J.2001pasilla, the Drosophila homologue of the human Nova-1 and Nova-2 proteins, is required for normal secretion in the salivary gland. Dev Biol, 2392Nov 15, 200 309322
  92. 92. SharmaS.FalickA. M.BlackD. L.2005Polypyrimidine tract binding protein blocks the 5’ splice site-dependent assembly of U2AF and the prespliceosomal E complex. Mol Cell, 194Aug 19, 200 485496
  93. 93. SharmaS.MarisC.AllainF. H.BlackD. L.2011U1 snRNA directly interacts with polypyrimidine tract-binding protein during splicing repression. Mol Cell, 415Mar 4, 201 579588
  94. 94. ShpargelK. B.MateraA. G.2005Gemin proteins are required for efficient assembly of Sm-class ribonucleoproteins. Proc Natl Acad Sci U S A, 10248Nov 29, 200 1737217377
  95. 95. SmithR. M.SadeeW.2011Synaptic signaling and aberrant RNA splicing in autism spectrum disorders. Front Synaptic Neurosci, 31
  96. 96. SpellmanR.LlorianM.SmithC. W.2007Crossregulation and functional redundancy between the splicing regulator PTB and its paralogs nPTB and ROD1. Mol Cell, 273Aug 3, 200 420434
  97. 97. TangZ. Z.SharmaS.ZhengS.ChawlaG.NikolicJ.BlackD. L.2011Regulation of the Mutually Exclusive Exons 8a and 8 in the CaV1.2 Calcium Channel Transcript by Polypyrimidine Tract-binding Protein. J Biol Chem, 28612Mar 25, 201 1000710016
  98. 98. ThieleA.NagamineY.HauschildtS.CleversH.2006AU-rich elements and alternative splicing in the beta-catenin 3’UTR can influence the human beta-catenin mRNA stability. Exp Cell Res, 31212Jul 15, 200 23672378
  99. 99. ToddA. G.MorseR.ShawD. J.Mc GinleyS.StebbingsH.YoungP. J.2010SMN, Gemin2 and Gemin3 associate with beta-actin mRNA in the cytoplasm of neuronal cells in vitro. J Mol Biol, 4015Sep 3, 201 681689
  100. 100. UleJ.JensenK.MeleA.DarnellR. B.2005aCLIP: a method for identifying protein-RNA interaction sites in living cells. Methods, 374Dec, 2005a), 376386
  101. 101. UleJ.JensenK. B.RuggiuM.MeleA.UleA.DarnellR. B.2003CLIP identifies Nova-regulated RNA networks in the brain. Science, 3025648Nov 14, 200 12121215
  102. 102. UleJ.StefaniG.MeleA.RuggiuM.WangX.TaneriB.GaasterlandT.BlencoweB. J.DarnellR. B.2006An RNA map predicting Nova-dependent splicing regulation. Nature, 4447119Nov 30, 200 580586
  103. 103. UleJ.UleA.SpencerJ.WilliamsA.HuJ. S.ClineM.WangH.ClarkT.FraserC.RuggiuM.ZeebergB. R.KaneD.WeinsteinJ. N.BlumeJ.DarnellR. B.2005bNova regulates brain-specific splicing to shape the synapse. Nat Genet, 378Aug, 2005b), 844852
  104. 104. ValenzuelaD. M.MurphyA. J.FrendeweyD.GaleN. W.EconomidesA. N.AuerbachW.PoueymirouW. T.AdamsN. C.RojasJ.YasenchakJ.ChernomorskyR.BoucherM.ElsasserA. L.EsauL.ZhengJ.GriffithsJ. A.WangX.SuH.XueY.DominguezM. G.NogueraI.TorresR.MacdonaldL. E.StewartA. F.De ChiaraT. M.YancopoulosG. D.2003High-throughput engineering of the mouse genome coupled with high-resolution expression analysis. Nat Biotechnol, 216Jun, 200 652659
  105. 105. ValverdeR.EdwardsL.ReganL.2008Structure and function of KH domains. FEBS J, 27511Jun, 200 27122726
  106. 106. WangE. T.SandbergR.LuoS.KhrebtukovaI.ZhangL.MayrC.KingsmoreS. F.SchrothG. P.BurgeC. B.2008Alternative isoform regulation in human tissue transcriptomes. Nature, 4567221Nov 27, 200 470476
  107. 107. WangZ.BurgeC. B.2008Splicing regulation: from a parts list of regulatory elements to an integrated splicing code. RNA, 145May, 200 802813
  108. 108. WinklerC.EggertC.GradlD.MeisterG.GiegerichM.WedlichD.LaggerbauerB.FischerU.2005Reduced U snRNP assembly causes motor axon degeneration in an animal model for spinal muscular atrophy. Genes Dev, 1919Oct 1, 200 23202330
  109. 109. WittenJ. T.UleJ.2011Understanding splicing regulation through RNA splicing maps. Trends Genet, 273Mar, 201 8997
  110. 110. WorkmanE.SaievaL.CarrelT. L.CrawfordT. O.LiuD.LutzC.BeattieC. E.PellizzoniL.BurghesA. H.2009A SMN missense mutation complements SMN2 restoring snRNPs and rescuing SMA mice. Hum Mol Genet, 1812Jun 15, 200 22152229
  111. 111. WuC. Y.WhyeD.GlazewskiL.ChoeL.KerrD.LeeK. H.MasonR. W.WangW.2011Proteomic assessment of a cell model of spinal muscular atrophy. BMC Neurosci, 1225
  112. 112. XieJ.LeeJ. A.KressT. L.MowryK. L.BlackD. L.2003Protein kinase A phosphorylation modulates transport of the polypyrimidine tract-binding protein. Proc Natl Acad Sci U S A, 10015Jul 22, 200 87768781
  113. 113. XueY.ZhouY.WuT.ZhuT.JiX.KwonY. S.ZhangC.YeoG.BlackD. L.SunH.FuX. D.ZhangY.2009Genome-wide analysis of PTB-RNA interactions reveals a strategy used by the general splicing repressor to modulate exon inclusion or skipping. Mol Cell, 366Dec 25, 200 9961006
  114. 114. YangY. Y.YinG. L.DarnellR. B.1998The neuronal RNA-binding protein Nova-2 is implicated as the autoantigen targeted in POMA patients with dementia. Proc Natl Acad Sci U S A, 9522Oct 27, 199 1325413259
  115. 115. YanoM.Hayakawa-YanoY.MeleA.DarnellR. B.2010Nova2 regulates neuronal migration through an RNA switch in disabled-1 signaling. Neuron, 666Jun 24, 201 848858
  116. 116. YeoG. W.CoufalN. G.LiangT. Y.PengG. E.FuX. D.GageF. H.2009An RNA code for the FOX2 splicing regulator revealed by mapping RNA-protein interactions in stem cells. Nat Struct Mol Biol, 162Feb, 200 130137
  117. 117. YoungP. J.Le T. T.thiMan. N.BurghesA. H.MorrisG. E.2000The relationship between SMN, the spinal muscular atrophy protein, and nuclear coiled bodies in differentiated tissues and cultured cells. Exp Cell Res, 2562May 1, 200 365374
  118. 118. ZhangC.FriasM. A.MeleA.RuggiuM.EomT.MarneyC. B.WangH.LicatalosiD. D.FakJ. J.DarnellR. B.2010Integrative modeling defines the Nova splicing-regulatory network and its combinatorial controls. Science, 3295990Jul 23, 201 439443
  119. 119. ZhangH.XingL.RossollW.WichterleH.SingerR. H.BassellG. J.2006Multiprotein complexes of the survival of motor neuron protein SMN with Gemins traffic to neuronal processes and growth cones of motor neurons. J Neurosci, 2633Aug 16, 200 86228632
  120. 120. ZhangZ.LottiF.DittmarK.YounisI.WanL.KasimM.DreyfussG.2008SMN deficiency causes tissue-specific perturbations in the repertoire of snRNAs and widespread defects in splicing. Cell, 1334May 16, 200 585600

Written By

Hung-Hsi Chen, Jung-Chun Lin and Woan-Yuh Tarn

Submitted: 19 November 2010 Published: 29 August 2011