Next Article in Journal
Extracellular Vesicles from Ovarian Carcinoma Cells Display Specific Glycosignatures
Next Article in Special Issue
Histone Deacetylase Inhibitors Activate Tristetraprolin Expression through Induction of Early Growth Response Protein 1 (EGR1) in Colorectal Cancer Cells
Previous Article in Journal
Direct and/or Indirect Roles for SUMO in Modulating Alpha-Synuclein Toxicity
Previous Article in Special Issue
Functional Integration of mRNA Translational Control Programs
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Pre-mRNA Splicing in Plants: In Vivo Functions of RNA-Binding Proteins Implicated in the Splicing Process

Department of Molecular Cell Physiology, Faculty of Biology, Bielefeld University, D-33615 Bielefeld, Germany
*
Author to whom correspondence should be addressed.
Biomolecules 2015, 5(3), 1717-1740; https://doi.org/10.3390/biom5031717
Submission received: 29 May 2015 / Revised: 16 July 2015 / Accepted: 20 July 2015 / Published: 24 July 2015
(This article belongs to the Special Issue RNA-Binding Proteins—Structure, Function, Networks and Disease)

Abstract

:
Alternative pre-messenger RNA splicing in higher plants emerges as an important layer of regulation upon exposure to exogenous and endogenous cues. Accordingly, mutants defective in RNA-binding proteins predicted to function in the splicing process show severe phenotypic alterations. Among those are developmental defects, impaired responses to pathogen threat or abiotic stress factors, and misregulation of the circadian timing system. A suite of splicing factors has been identified in the model plant Arabidopsis thaliana. Here we summarize recent insights on how defects in these splicing factors impair plant performance.

1. Introduction

The realization of the genetic information encoded in eukaryotic genomes critically depends on the removal of intervening sequences (introns), the non-coding sequences, and the joining of the flanking exons through the splicing process [1,2]. The introns are delineated by short consensus sequences at the borders between the introns and exons, the splice sites (Figure 1). Splicing of the primary transcripts, the pre-messenger RNAs (pre-mRNAs), is accomplished by the spliceosome, a large complex consisting of several non-coding RNAs and a plethora of protein factors [3,4,5,6]. The process of pre-mRNA splicing is best understood in mammals and yeast, mainly through biochemical purification of spliceosomal components and functional tests using in vitro splicing assays.
Figure 1. Splicing signals involved in pre-mRNA splicing. Introns are delineated by short consensus sequences, the 5' and 3' splice sites (R = Purine, A/G), the polypyrimidine tract (Y = Pyrimidine, C/T), and the branch point sequence. Intronic splicing silencers (ISS), intronic splicing enhancers (ISE), exonic splicing silencers (ESS), and exonic splicing enhancers (ESE) are additional sequence elements serving as docking sites for protein factors that promote or impair assembly of the spliceosome at particular splice sites and thus control the splicing process. The green boxes denote exons. Introns are represented by lines with the consensus sequences indicated (see text for details).
Figure 1. Splicing signals involved in pre-mRNA splicing. Introns are delineated by short consensus sequences, the 5' and 3' splice sites (R = Purine, A/G), the polypyrimidine tract (Y = Pyrimidine, C/T), and the branch point sequence. Intronic splicing silencers (ISS), intronic splicing enhancers (ISE), exonic splicing silencers (ESS), and exonic splicing enhancers (ESE) are additional sequence elements serving as docking sites for protein factors that promote or impair assembly of the spliceosome at particular splice sites and thus control the splicing process. The green boxes denote exons. Introns are represented by lines with the consensus sequences indicated (see text for details).
Biomolecules 05 01717 g001

1.1. The Splicing Process

The spliceosome assembles at each intron in a precise order, briefly summarized here [4]. The spliceosome comprises small nuclear ribonucleoprotein complexes (snRNPs) which are composed of small nuclear RNAs associated with distinct sets of proteins. The U rich small nuclear RNAs (U snRNAs) are short non-coding, non-polyadenylated transcripts. The Sm class of snRNAs, U1, U2, U4, and U5, associates with Sm proteins, named after the patient Smith with an autoimmune disease who developed antibodies against these proteins. The Like-Sm U6 snRNA associates with the related Sm-like proteins Like Sm 2 (Lsm2) to Lsm8 to form the U6 snRNP [7].
The splicing process is initiated by the interaction of the U1 snRNP with the 5' splice site through base pairing of the U1 snRNA to complementary sequences in the pre-mRNAs (Figure 2). The two subunits of U2 auxillary factor (U2AF) interact with the 3' splice site. The 35 kDa subunit U2AF35 binds to the intron/exon border and the 65 kDa subunit U2AF65 binds to the polypyrimidine tract upstream of the intron/exon border, leading to the so-called complex E. Subsequently, the U2 snRNP binds to the branch point via base pairing of U2 snRNA, defining complex A. An assembly of the U4, U5, and U6 snRNPs (the so-called U4/U6.U5 snRNP) is then docked onto the U2 snRNP, leading to the precatalytic complex B that comprises all five snRNPs. After major rearrangements and the release of the U1 and U4 snRNPs, the activated complex Bact is formed. In the first step of the splicing reaction, the pre-mRNA is cleaved at the 5' splice site and lariat formation of the intron sequence takes place by ligating its 5' end to the branch point adenosine. This leads to the formation of complex C that then catalyzes the second step of the splicing reaction, the cleavage at the 3' splice site, the ligation of the exons, and the release of the spliced mRNA. The lariat is degraded and the snRNPs are recycled. Throughout the catalytic cycle, ATP-dependent RNA helicases are involved in the remodeling of the spliceosome, e.g., through reconfiguring RNA-RNA interactions [8].
Figure 2. The spliceosomal cycle. The splicing process is initiated by the interaction of the U1 snRNP with the 5' splice site through base pairing of the U1 snRNA to complementary sequences. The U2AF subunit U2AF35 binds to the intron/exon border and the subunit U2AF65 binds to the polypyrimidine tract upstream of the intron/exon border, leading to the complex E. Subsequently, the U2 snRNP binds to the branch point, defining complex A. The U4/U6.U5 snRNP is then docked onto the U2 snRNP, leading to the precatalytic complex B. After major rearrangements and the release of the U1 and U4 snRNPs, the activated complex Bact is generated. In the first step of the splicing reaction, the pre-mRNA is cleaved at the 5' splice site and lariat formation of the intron sequence takes place by ligating its 5' end to the branch point adenosine. This leads to the formation of complex C that then catalyzes the second step of the splicing reaction, the cleavage at the 3' splice site, the ligation of the exons, and the release of the spliced mRNA. The lariat is degraded and the snRNPs are recycled. The proteins with a demonstrated role in pre-mRNA splicing in Arabidopsis are shown where in the spliceosomal cycle they are proposed to act, based on the homology to their human counterparts. The boxes denote exons; the thin line denotes the intron. NTC, Nineteen Complex.
Figure 2. The spliceosomal cycle. The splicing process is initiated by the interaction of the U1 snRNP with the 5' splice site through base pairing of the U1 snRNA to complementary sequences. The U2AF subunit U2AF35 binds to the intron/exon border and the subunit U2AF65 binds to the polypyrimidine tract upstream of the intron/exon border, leading to the complex E. Subsequently, the U2 snRNP binds to the branch point, defining complex A. The U4/U6.U5 snRNP is then docked onto the U2 snRNP, leading to the precatalytic complex B. After major rearrangements and the release of the U1 and U4 snRNPs, the activated complex Bact is generated. In the first step of the splicing reaction, the pre-mRNA is cleaved at the 5' splice site and lariat formation of the intron sequence takes place by ligating its 5' end to the branch point adenosine. This leads to the formation of complex C that then catalyzes the second step of the splicing reaction, the cleavage at the 3' splice site, the ligation of the exons, and the release of the spliced mRNA. The lariat is degraded and the snRNPs are recycled. The proteins with a demonstrated role in pre-mRNA splicing in Arabidopsis are shown where in the spliceosomal cycle they are proposed to act, based on the homology to their human counterparts. The boxes denote exons; the thin line denotes the intron. NTC, Nineteen Complex.
Biomolecules 05 01717 g002

1.2. Alternative Splicing

Not every splice site of a pre-mRNA is used in each instance the gene is transcribed. This conditional splicing is known as alternative splicing [9,10]. It enables the cell to obtain different mRNA isoforms from the same gene. By these means, the coding capacity of the genome is greatly enhanced. In humans, for example, alternative splicing expands the proteome five times compared to the number of protein-coding genes [11]. Alternative splicing events include exon skipping (ES), alternative 5' or 3' splice site selection, retention of introns (IR), or combinations of these events. Depending on the exclusion or inclusion of exons, protein variants with different domain composition arise from the mRNA isoforms. The corresponding protein variants may have distinct functions or localize to different subcellular compartments. Furthermore, alternative splicing can alter the open reading frame (ORF) and lead to mRNA variants with premature termination codons (PTCs). Such PTC-containing transcripts can be eliminated via the nonsense-mediated decay (NMD) pathway, a specialized RNA decay pathway that removes PTC-containing mRNAs that upon translation may give rise to aberrant proteins [12]. In effect, alternative splicing-NMD can thus also lead to altered transcript levels.

1.3. Regulation of Splicing

The use of the splice sites, i.e., the assembly of the spliceosome at defined splice sites, is governed by cis regulatory elements that can be located both in the introns and the exons. These elements function by recruiting additional RNA-binding proteins during the assembly and the catalytic cycle of the spliceosome. These factors either promote or inhibit the use of splice sites. Accordingly, the regulatory sequence motifs are termed exonic splicing enhancers, exonic splicing silencers, intronic splicing enhancers, and intronic splicing silencers (Figure 1). The trans-acting proteins belong mainly to two broad categories. The SR (Serine Arginine rich) proteins consist of one or two RNA recognition motifs (RRMs) that contact the RNA targets and a domain enriched in serine and arginine residues involved in protein-protein interaction [13]. The hnRNPs (heterogeneous nuclear ribonucleoparticle proteins) are a class of RNA-binding proteins that associate with heterogeneous nuclear RNA and mRNA in the cytoplasm and participate in multiple steps of pre-mRNA processing in addition to their role in splicing [14]. How the splicing regulatory elements impact the choice of splice sites may vary depending on the sequence context in their surroundings. For example, SR proteins commonly repress splicing when they are bound by intronic sequences and activate splicing when they are bound by exonic sequences [15]. Mammalian hnRNPA1 inhibits splicing when recruited to introns or to exons [15]. The hnRNP-like proteins of the neuro-oncological ventral antigen (NOVA) family recognize YCAY (Y = pyrimidine, C or U) motifs that can act as exonic splicing silencers or serve as intronic splicing silencers when located in the intron preceeding an alternatively spliced exon or as an intronic splicing silencer when located in the intron following an alternatively spliced exon [16].
The importance of correct splicing patterns for the function of the cell is underscored by the observation that in humans, mis-regulation of splicing is a common feature of many diseases [17,18,19]. These disorders can be caused by mutations that disrupt the splicing of specific genes or by a general loss of spliceosomal function, affecting entire networks of downstream genes.

1.4. Pre-mRNA Splicing in Plants

Posttranscriptional control of gene expression by regulated pre-mRNA splicing has recently moved in the focus of plant molecular biology, as alternative splicing has been found to reshape the transcriptome in particular in response to biotic and abiotic stress [20,21,22]. Accordingly, aberrant pre-mRNA splicing leads to severe phenotypes including impaired pathogen defense and abnormal developmental programs [23,24,25,26,27]. Currently, it is estimated that 61% of all multi-exon genes undergo alternative splicing in the model plant Arabidopsis thaliana, whereas in mammals alternative splicing has been found for more than 90% of intron containing genes [28].
Knowledge on pre-mRNA splicing in higher plants is still limited. While the basic principles are conserved there appear to also be differences in detecting introns. A mammalian intron was not efficiently spliced in plants, but some plant introns were correctly spliced in HeLa cell extracts [29,30,31]. One of the differences may lie in the vastly differing lengths of introns from animals that are about 5 kb on average and of plant introns with an average length of 160 bp [32].
In Arabidopsis, genes for counterparts of most spliceosomal proteins have been identified through homology searches. Soon after the Arabidopsis genome was available, 395 genes for factors predicted to function in the splicing process were identified [33]. Later, Koncz and co-workers systematically mined the TAIR10 (The Arabidopsis Information Resource 10) version of the Arabidopsis genome for homologs of components of biochemically purified spliceosomal complexes in humans, Drosophila, or yeast that were identified via mass spectrometry [34]. This search yielded another hundred candidate splicing factors. In many cases, several paralogs of mammalian and yeast spliceosomal components have been found in Arabidopsis, pointing to some redundancy within these families [33,34]. Indeed, numerous mutants have been identified in Arabidopsis with defects in predicted counterparts of mammalian and yeast splicing factors while such mutants would not be viable in the absence of a compensatory gene.
Here, we summarize recent advances in understanding proteins involved in pre-mRNA splicing and alternative splicing in the reference plant Arabidopsis thaliana. We will not provide a complete inventory of splicing factors. Rather, we mainly focus on proteins with a demonstrated in vivo function in pre-mRNA splicing through mutant analysis or transgenic approaches and summarize the phenotypes generated by misexpression of the splicing factors. Arabidopsis thaliana proteins named according to their mammalian counterparts are designated with “At-” for clarity. In keeping with the convention in the plant community, Arabidopsis proteins are given in uppercase letters, the corresponding transcripts are given in uppercase and italics, and mutants are given in lowercase and italics.

2. Proteins Involved in Pre-mRNA Splicing in Arabidopsis thaliana

2.1. SR Proteins

The first proteins shown to regulate pre-mRNA splicing in plants were members of the SR protein family [35]. Originally, SR proteins in plants were identified through their cross-reaction with the mAB104 antibody directed against a phosphorylated serine in the RS domain in mammalian SR proteins [36,37]. Of note is the high number of SR proteins in plants. Arabidopsis thaliana has 18 SR proteins that represent six different gene families [35,38,39,40,41]. Three of them comprise orthologs of metazoan SR proteins. The family of ASF/SF2 orthologs comprises At-SR30, At-SR34, At-SR34a, and At-SR34b. The family of 9G8 orthologs comprises At-RSZ21, At-RSZ22, and At-RSZ22a that feature one zinc knuckle. In addition, there is one ortholog of SC35. The other three SR families are plant-specific. The RS2Z family comprises At-RS2Z32 and At-RS2Z33 with two zinc knuckles between the RRM and the RS domain and a C-terminal serine and proline-rich extension [42]. The RS family comprises At-RS31, At-RS31a, At-RS40, and At-RS41. The SCL family harbors At-SCL28, At-SCL30, At-SCL30a, and At-SCL33.
Ectopic expression of the SR proteins At-SR30 and At-RS2Z33 in transgenic plants produces a range of morphological and physiological phenotypes and impacts alternative splicing of a suite of other transcripts [43,44]. Recently, alternative splicing of At-SR30 and At-RS31 has been shown to be regulated by retrograde signaling from the chloroplast to the nucleus, a process that synchronizes gene expression in the nucleus with the needs of the chloroplast compartment [45]. Mutants deficient in At-SR34b are sensitive to Cadmium (Cd) due an increased uptake of toxic Cd2+ ions into the root [46]. In sr34b, splicing and stability of the mRNA-encoding IRON REGULATED TRANSPORTER 1 (IRT1), a transporter of Cd2+ ions located in the plasma membrane of root epidermal cells, are altered, suggesting that altered posttranscriptional regulation of IRT1 in sr34b is linked to the Cd sensitivity phenotype.
Several SR proteins undergo auto-regulation through alternative splicing to PTC-containing isoforms [44,47,48,49]. About half of the PTC-containing isoforms are degraded via NMD and thus lead to alterations in transcript abundance [50,51]. Several of these alternative splicing events in SR genes are evolutionary conserved, underscoring their physiological relevance [41,52]. Recently, regulatory elements within an intron of At-SCL33 have been identified that mediate auto-regulation of At-SCL33 alternative splicing [53]. At-SCL33 binds to a conserved sequence motif in this intron with four GAAG motifs that are necessary for correct splicing. The identification of binding sites for the SR protein will help to identify additional downstream targets of At-SCL33.
In addition to the canonical SR proteins, Arabidopsis harbors the SR-like protein SR45 with two RS domains, one of each side of the RRM, that shows homology to RNA binding protein S1, serine-rich domain (RNPS1) of the mammalian exon junction complex, a protein complex loaded onto mRNAs upon splicing [54]. The sr45 mutant exhibits developmental abnormalities, including narrow leaves, an altered number of petals and stamens in the flowers, and short roots [55]. The use of alternative 3' splice sites leads to two alternative splice isoforms which differ by eight amino acids. The short splice variant SR45.2 specifically complements the defect of the sr45 mutant in root growth [56]. SR45.1 harbouring the T218SPQRKTG peptide specifically complements the defect of the sr45 mutant in petal development [56]. The mutation of T218 to alanine abolished the complementation whereas the mutation of T218 to aspartic or glutamic acid, which mimics the charge of a phosphate group, still allowed rescue of the mutant [57]. This has been attributed to phosphorylation of T218. Additionally, the sr45 mutant displays hypersensitivity to glucose and the phytohormone abscisic acid (ABA), which is complemented by both splice isoforms [58]. SR45 has recently also been implicated in splicing of the circadian clock gene CCA1 (CIRCADIAN CLOCK ASSOCIATED 1). Through alternative splicing at its fourth intron, the CCA1 transcript can give rise to transcript isoforms retaining intron 4 that would generate truncated protein isoforms due to PTCs. In the sr45 mutant, less of the intron-retained forms are detected compared to wild-type plants. Because recombinant SR45 protein is able to bind to the retained CCA1 intron in vitro, it may contribute to this alternative splicing event [59]. It is not yet known whether the sr45 mutant exhibits a defect in circadian clock function. Furthermore, SR45 regulates alternative splicing of several SR transcripts, indicating extensive hierarchical regulation among those splicing regulators [55,60]. In the At-SR30 pre-mRNA, SR45 binds to the 5' end of intron 10, whose splicing is altered in the sr45 mutant. SR45 interacts with the U1 snRNP protein U1-70K involved in the recognition of 5' splice sites and with U2AF35b involved in the recognition of 3' splice sites, suggesting that SR45 helps to recruit U1snRNP to the 5' splice site and At-U2AF to the 3' splice site [60].
A hallmark of SR function is regulation by posttranslational modification. A large proportion of the SR proteins can undergo phosphorylation [61,62]. In particular, a LAMMER protein kinase, named for a conserved EHLAMMERILG motif in their catalytic subdomain, can phosphorylate SR proteins, and overexpression of the LAMMER kinase leads to aberrant splicing of At-SR30 and At-SR34 as well as other transcripts [63].

2.2. hnRNPs

The second type of splicing regulators described in plants corresponds to the hnRNPs [64].

2.2.1. Polypyrimidine Tract Binding Proteins

In Arabidopsis, three homologs of the polypyrimidine tract binding proteins (PTBs) that interact with the pyrimidine rich motifs at the 3' splice sites have been identified. In mammals, PTBs either activate or repress the usage of specific splice sites in the vicinity of their binding sites. PTBs can also act in a combinatorial manner with other hnRNP proteins or SR proteins [65].
Mammalian PTB negatively auto-regulates by alternative splicing where skipping of the first exon produces an NMD substrate [66,67]. Arabidopsis At-PTB1 and At-PTB2 proteins similarly show negative auto-regulation by alternative splicing of their own pre-mRNAs where inclusion of a PTC-containing cassette exon creates an NMD substrate [68]. Furthermore, At-PTB1 and At-PTB2 reciprocally cross-regulate through the same mechanism. In transgenic Arabidopsis, overexpression or knock-down of At-PTB1 or At-PTB2 alters the splicing pattern of a plethora of transcripts [69,70]. Using a transient expression system it was shown that At-PTB1 promoted skipping of an alternative exon in a splice reporter construct whereas At-U2AF65 promoted its exclusion, indicating that At-PTB1 and At-U2AF65 can act antagonistically, similar to what is observed in mammals [70].
A physiological role for At-PTBs in seed germination was found through the control of PHYTOCHROME INTERACTION FACTOR 6 (PIF6) splicing. A PIF6 alternative splice variant is predicted to encode a short protein variant lacking the DNA binding domain due to exon skipping [71]. Down-regulation of AtPTB1 or AtPTB2 increased the exon skipping, and the shift to the short protein variant accelerated seed germination in the presence of the phytohormone abscisic acid (ABA) that retards germination. Up-regulation of At-PTB1 or At-PTB2 in turn reduced exon skipping, favoring the full length PIF6 protein, and delayed seed germination in the presence of ABA.
The more distantly related At-PTB3 did not exhibit a significant impact on alternative splicing [69].

2.2.2. Glycine-Rich RNA Binding Proteins

Higher plants harbor a gene family encoding small RNA-binding proteins with an N-terminal RRM and a glycine stretch at the C-terminus that are simplified versions of mammalian hnRNPs [72,73]. These glycine-rich RNA-binding proteins are involved in responses to diverse environmental stresses. Two of the family members, AtGRP7 and AtGRP8, have a role in alternative splicing. These proteins are controlled by the circadian clock [74]. A reverse genetic approach has shown that AtGRP7 plays a regulatory role in the circadian timing system [75]. AtGRP7 negatively auto-regulates through binding to its own mRNA, leading to the usage of an alternative cryptic splice site within the intron that generates an alternative transcript isoform subject to NMD [76]. AtGRP8 uses the same negative feedback loop to control its own expression. Additionally, AtGRP7 and AtGRP8 are able to reciprocally cross-regulate each other’s expression [77]. Thus, they are the first example of a posttranscriptional feedback loop in the circadian timing system of any organism.
When transgenic plants with elevated or reduced levels of AtGRP7 were investigated with an alternative splicing panel based on Reverse Transcription–PCR [78], significant changes of alternatively spliced isoforms were found for 20% of the analyzed splicing events [79]. In particular, altered AtGRP7 levels lead to changes in the choice of alternative 5' splice sites. Several of the identified transcripts are bound in vivo by AtGRP7 but not by a mutant version with the conserved arginine 49 in the RRM that is important for RNA binding exchanged for glutamine [79,80]. This indicates that AtGRP7 impacts alternative splicing of these transcripts via direct interaction. Several of these alternative splicing events are also influenced by the paralog AtGRP8, indicating that both proteins act on a shared set of downstream targets. Notably, AtGRP7 also boosts plant defense reactions against pathogenic Pseudomonas syringae bacteria. It binds to several defense-related transcripts in vivo [81,82]. Upon infection, pathogenic Pseudomonas syringae bacteria inject type III effector protein HopU1, an ADP riboslyase, into plant cells as part of their virulence strategy [83]. HopU1 ADP ribosylates the R49 in the RNA recognition motif of AtGRP7 to disable RNA binding and subvert immune responses [84].

2.3. Spliceosomal Proteins

Apart from the involvement of SR proteins and hnRNPs in the regulation of splicing, mutants deficient in spliceosome-associated proteins have been found to exhibit splicing defects in Arabidopsis.

2.3.1. General snRNP Proteins

The snRNPs of the spliceosome contain general proteins in addition to proteins specific for the respective snRNP. SmD3b mutants defective in the general snRNP core protein At-SmD3b display a range of physiological phenotypes including impaired root growth, an altered number of floral organs, and late flowering [85]. These phenotypes correlated with altered splicing patterns of several pre-mRNAs.
Mutants defective in the U6 snRNP core protein At-Lsm4 also show splicing defects [86]. In particular, several genes of the circadian clock are aberrantly spliced. Accordingly, the lsm4 mutant shows a long period of circadian transcript oscillations [87]. The lsm5 mutant shows a global defect in pre-mRNA splicing [88,89]. A reduced level of U6 snRNA in lsm5 suggests that At-Lsm5 contributes to U6 stability [88]. In particular, At-Lsm5 affects the splicing of transcripts associated with salinity stress, a major threat to plants. Accordingly, constitutive overexpression of At-Lsm5 improves salt tolerance in transgenic plants. This correlates with an increased splicing accuracy and efficiency for stress-responsive genes [89].
Arginine methylation is an important posttranslational modification of proteins catalyzed by protein arginine methyltransferases. While many cellular proteins including histones can be modified by these enzymes, arginine methylation is frequently observed in RNA-binding proteins with glycine-arginine-rich motifs. Protein arginine methyltransferase 5 methylates the spliceosomal U snRNP proteins B, B', D1, and D3. This modification is crucial for the assembly of the snRNP complexes [90,91]. In Arabidopsis, prmt5 mutant plants show a range of defects including late flowering. The transition from vegetative to reproductive growth is an important developmental decision in the life of the plant and accordingly is exquisitely controlled. In Arabidopsis, flowering is repressed by the key floral repressor FLOWERING LOCUS C (FLC) until the plant encounters favorable conditions. Late flowering of prmt5 mutants correlates with elevated FLC levels. Furthermore, the prmt5 mutant shows reduced inhibition of hypocotyl elongation in response to light and impaired root growth in the presence of high levels of NaCl, pointing to reduced salt stress tolerance [92,93]. The prmt5 mutant has also a defect in circadian timekeeping, showing a longer period of leaf movement and gene expression rhythms [94,95]. At-PRMT5 affects splicing of the core clock gene PSEUDORESPONSE REGULATOR 9 (PRR9), suggesting that the circadian defect in prmt5 can be partly attributed to changes in PRR9 splice patterns [94,96]. Moreover, At-PRMT5 deficiency causes global splicing defects. About 20% of investigated alternative splicing events were altered in the prmt5 mutant plants. Moreover, constitutive splicing was affected, although to a lesser extent than alternative splicing [93]. In prmt5, methylation of the U snRNP At-SmD1, At-SmD3, and At-LSm4 proteins is disturbed, which may be responsible for the observed splicing defects [96].
Recently, Arabidopsis At-PRMT4, a representative of type I protein arginine methyltransferases that generates monomethylated and asymmetrically dimethylated arginine residues, was analyzed for a role in splicing control. This was prompted by the observation that among substrates for human PRMT4, also known as CARM1, are several splicing factors including the U1C protein, a component of the U1 snRNP, SAP49 and CA150, which are associated with the U2 snRNP, as well as the spliceosomal protein SmB. Moreover, PRMT4/CARM1 regulates splicing by promoting exon skipping [97]. Single mutations in either At-PRMT4a or At-PRMT4b did not cause major developmental defects, pointing to the redundancy of the two proteins [98]. Double mutants deficient in both At-PRMT4a and At-PRMT4b show strikingly similar phenotypes as prmt5 mutants, including late flowering, reduced sensitivity to light, and reduced salt stress tolerance [93]. However, they are not affected in circadian timekeeping. Transcript profiling of the prmt4a prmt4b mutant revealed for the first time a global role for At-PRMT4 in the control of alternative splicing in any organism and also uncovered a function in constitutive splicing. Both At-PRMT4 and At-PRMT5 appear to act on introns with weak 5' splice sites that deviate from the consensus sequence. Thus, At-PRMT4 and At-PRMT5 function may be required to stabilize the interaction of the weak splice sites with U1 snRNA [93]. A genome-wide survey will be required to test whether the mammalian counterparts may have a similar bias towards weak 5' splice sites.

2.3.2. U snRNP Specific Proteins

In addition to the general spliceosomal proteins, the function of several U snRNP-specific proteins has been characterized in Arabidopsis. STABILIZED 1 (STA1) is a homolog of the human U5 snRNP-associated 102-kDa protein, also known as PRPF6 (Precursor RNA Processing factor 6) [99]. STA1 affects splicing of transcripts that are induced by low temperature exposure of plants including the COR15A (COLD REGULATED 15A) transcript. COR15A encodes a polypeptide located in the chloroplast that protects chloroplast membranes against damage at low temperatures [100]. Because the sta1 mutant is cold-sensitive, STA1 appears to be required for correct splicing during the cold stress response.
The spliceosomal protein PRPF31 regulates the formation of the U4/U6.U5 snRNP. While prp31 null mutants in the Arabidopsis homolog of PRPF31 are embryonic lethal, mutants with a reduced level of At-PRP31 also show an impaired tolerance to low temperature [101]. At-PRP31 is required for splicing of cold-induced genes, especially under cold stress. Expression of STA1 and At-BRR2, a homolog of the U5 snRNP-specific RNA helicase that acts at specific steps of the splicing cycle to catalyze RNA-RNA rearrangements and spliceosome remodeling, is up-regulated in the prp31 mutant. This suggests a concerted action of these splicing factors in the low temperature response.
A genetic screen for Arabidopsis mutants with defects in the development of female gametes, based on aberrant expression of an egg cell specific reporter gene, yielded mutants defective in several snRNP proteins [102]. While homozygous mutants are lethal, the heterozygous mutants show severe defects during development of the female gametes. In the female gametophyte, two gametic cells, the egg cell and central cell, as well as several nongametic accessory cells, the two synergid cells neighboring the egg cell and three antipodal cells, develop. In lachesis (lis) mutants, supernumerary egg cells are observed due to misspecification of the synergids, the central cell and the antipodal cells that also express features of the central cell, i.e., they aberrantly activate a gametic cell fate in a non-gametic cell [102]. LIS codes for a homolog of the U4/U6-specific PRPF4. LIS is up-regulated in gametic cells and down-regulated in the accessory cells. Mutants defective in CLOTHO (CLO)/GAMETOPHYTIC FACTOR 1 (GFA1) show diminished female gametophyte development and arrest of early embryo development [103,104]. GFA1 encodes a homolog of the U5–116kD protein and physically interacts with homologs of two other components of the U5 snRNP, the U5-200kD protein At-BRR2, and the U5-220kD protein At-PRP8. In atropos (ato) mutants with a defect in the homolog of SF3A3, a component of the U2 snRNP, the fate of the egg cell and central cell is also compromised [104]. Collectively, the phenotypes of these mutants indicate that correct splicing is essential for cell fate determination in the female gametophyte. Surprisingly, mutations in factors crucial for the splicing process have severe consequences only in a very limited developmental window, pointing to novel mechanisms of spliceosome function yet to be described.
Mutants in several spliceosomal proteins show alterations in the timing of the transition to flowering through misexpression of the floral repressor FLC. Antisense transcripts originating from a promoter downstream of the FLC gene play a crucial role in down-regulation of FLC expression through chromatin modification. These COOLAIR antisense transcripts themselves undergo alternative splicing and alternative polyadenylation either at a promoter proximal poly(A) site or a distal poly(A) site opposite to the FLC transcription start site. The use of the promoter proximal poly(A) site in the COOLAIR antisense transcript correlates with increased histone H3 lysine 4 demethylation and reduced expression of the FLC sense transcript [105].
A crucial role in FLC antisense transcript regulation has been found for the homolog of yeast PRP8, the largest and most highly conserved protein of the spliceosome [106,107]. PRP8 is necessary for several rearrangements occurring throughout the spliceosomal cycle.
While a complete loss-of-function mutant Arabidopsis At-PRP8 is lethal, a hypomorphic mutation reduces splicing of the COOLAIR introns and usage of the proximal poly(A) site [108]. This leads to increased histone methylation in the FLC gene and concomitant up-regulation of FLC transcription. Accordingly, the transition to flowering is delayed in prp8.
Arabidopsis harbors two homologs of a tetratricopeptide repeat protein with similarity to the yeast U1 snRNP-specific pre-mRNA processing protein PRP39. The prp39a mutant flowers later than wild-type plants and shows an elevated FLC level [109]. Whether splicing is affected in the prp39a mutant has not yet been investigated.
A key component promoting the transition to flowering in response to increasing day length in spring and summer is SUPPRESSOR OF CONSTANS 1 (SOC1). SOC1 activates floral meristem identity genes in the shoot apical meristem to initiate flower development. An accumulation of fully spliced and incompletely spliced SOC1 transcript isoforms is observed in the early flowering 9 (elf9) mutant that shows premature transition to flowering [110]. ELF9 codes for a homolog of human Tat stimulatory factor 1 (Tat-SF1) necessary for HIV replication and yeast CUS2, a splicing factor that aids the assembly of the splicing-competent U2 snRNP.
SUA (SUPPRESSOR OF abi3-5) is a splicing factor that affects seed maturation. The phytohormone ABA prevents germination, and during seed germination the ABA signaling component ABSCISIC ACID INSENSITIVE3 (ABI3) is down-regulated. ABI3 undergoes developmentally regulated alternative splicing that is under control of SUA, a homolog of the human splicing regulator RBM5 [111]. RBM5 interacts with U2AF65, and SUA has also been shown to interact with At-U2AF65, suggesting that it may act at an early step of the splicing process.

2.3.3. The Nineteen Complex

A non-snRNP complex consisting solely of proteins is essential for the activation of the spliceosome, designated the nineteen complex (NTC) for its core component PRP19 [112]. Arabidopsis harbors two orthologs of PRP19 that interact with At-CDC5, the ortholog of the NTC components CELL DIVISION CYCLE 5 (CDC5) and PLEIOTROPIC REGULATORY LOCUS 1 (PRL1) [113]. Interestingly, several Arabidopsis orthologs of NTC components have been identified in genetic screens for components involved in pathogen defense. Plants sense the presence of pathogenic bacteria through dedicated receptors in the cell membrane, and the activation of appropriate signaling chains induces a set of defense responses in the plant. Transcriptional activation of defense genes largely relies on a key component, NONEXPRESSOR OF PATHOGENESIS RELATED GENES 1 (NPR1). Accordingly, npr1 mutants are impaired in the induction of pathogenesis-related genes. A second site mutation of the SNC1 (SUPPRESSOR OF NPR1-1, CONSTITUTIVE1) gene encoding a resistance (R) protein restores pathogen resistance in the npr1 background [114]. In turn, a suppressor screen for mutants reverting the pathogen resistance phenotype of the snc1 mutant revealed several mutants with a constitutive pathogen defense response even in the absence of a bacterial pathogen. Accordingly, the corresponding proteins are known as MOS (MODIFIER OF snc1) proteins. In this screen, At-BCAS2 was identified and designated MOS4. Moreover, through a proteomics approach to identified proteins associated with epitope-tagged MOS4 in vivo, the orthologs of PRP19 were recovered and designated MAC3A (MOS4 associated complex 3A) and MAC3B [113,115]. In the mos4 and cdc5 mutants and the mac3a mac3b double mutant the SNC1 splicing pattern is altered, providing evidence that the MOS complex is involved in alternative splicing, similar to the NTC [116].
Recently, aberrant splicing of SUPPRESSOR OF NPR1-1, CONSTITUTIVE4 (SNC4), and CHITIN ELICITOR RECEPTOR KINASE 1 (CERK1), encoding receptor-like kinases involved in pathogen defense, has been found in the sua mutant [117]. Accordingly, sua mutants show enhanced susceptibility to Pseudomonas syringae infection. Thus, the RBM5 counterpart SUA appears to be involved both in germination (see Section 2.3.2) and response to pathogens.
Other proteins associated with the MOS (MODIFIER OF snc1) complex are MAC5A and MAC5B, with similarity to human RBM22 (RNA Binding Motif Protein 22) [118]. The zinc-finger protein RBM22 interacts with the U6 snRNA and the pre-mRNA in the spliceosomal Bact and C complexes [34,119]. This invites the speculation that MAC5 may also play a role in splicing in Arabidopsis.
Furthermore, the homolog of the human NTC component SNW/Ski-interacting protein domain protein (SKIP) and yeast Prp45 has been functionally characterized in Arabidopsis [120]. Arabidopsis At-SKIP interacts with the splicing factor SR45 and has a global effect on splicing. In particular, the core clock genes PRR7 and PRR9 are aberrantly spliced in skip mutants. This may contribute to the long period of the circadian clock in skip mutants. In addition, skip mutants are hypersensitive to both salt and osmotic stress [121]. A genome-wide analysis showed that At-SKIP impacts alternative splicing of numerous genes, particularly upon salt stress, and is required for splicing of genes associated with salt stress tolerance.

2.3.4. The Nineteen Related Complex

The NTR (nineteen related complex) is another non-snRNP complex consisting solely of proteins. It contributes to spliceosome disassembly. Arabidopsis SPLICEOSOMAL TIMEKEEPER LOCUS1 (At-STIPL1) is a homolog of the NTR component TFP11 in humans and Ntr1p in yeast [122]. In the Arabidopsis stipl1 mutant splicing of many introns is reduced. At the physiological level, stipl1 shows a long period of the circadian clock. In accordance with this, the accumulation of circadian transcripts is altered. In particular, retention of intron 3 in the core clock gene PRR9 is increased, which leads to a non-functional protein.

2.4. RNA Export Factors

In yeast and animals, the TREX (TRANSCRIPTION EXPORT) complex links transcription elongation to mRNA export. Human TREX has also been found to associate with the splicing machinery [123]. Recently, orthologs of components of the THO complex, a subunit of the TREX complex, have been connected to splicing in Arabidopsis. The hrp1-5 mutant impaired in HYPER RECOMBINATION1 (HPR1)/THO1 is affected in bulk nuclear mRNA accumulation and thus RNA export [124]. In an allelic mutant, alternative splicing patterns of At-RS31, At-RS40, and At-SR34b are altered [125]. At-HRP1 co-localizes with At-SR33 in nuclear speckles. The exact mechanism of At-HRP1 function in splicing control remains to be determined. Mutants with reduced expression of Arabidopsis At-THO2 also display splicing defects [126].
Mutants defective in the K homology (KH)-domain RNA-binding protein HOS5 (HIGH OSMOTIC STRESS GENE EXPRESSION 5) are impaired in mRNA export, particularly under salt stress conditions [127]. HOS5 interacts with At-RS40 and At-RS41 in nuclear speckles, and intron retention in many stress-related genes is observed under salt stress but not under normal conditions in the hos5 mutant [127].

2.5. Miscellaneous Other RNA Binding Proteins

Arabidopsis RCF1 (REGULATOR OF C-REPEAT BINDING FACTOR GENE EXPRESSION 1) encodes a DEAD-box RNA helicase that is induced by cold [128]. Many cold-responsive genes are misspliced in the rcf1-1 mutant under cold stress. Accordingly, the mutant is hypersensitive to cold stress.
The SR45a protein initially was considered to represent an SR protein, but now has been recognized to be a homolog of Transformer-2, a factor involved in sex-specific splicing of pre-mRNA from the sex determination gene doublesex in Drosophila [38,129]. SR45a interacts with several spliceosomal proteins, including U1-70K that recognizes the 5' splice site, U2AF35b that recognizes the 3' splice site, At-SR45, and At-SCL28, as well as the ortholog of PRP38 that contributes to maturation of the spliceosome before the first step of the splicing reaction. Thus, SR45a may participate in the splicing process in Arabidopsis through bridging the 5' end and 3' end of introns [130].
At-RTF2 encodes an evolutionarily conserved protein containing a Replication termination factor2 (Rtf2) domain [131]. While a homozygous null mutation in At-RTF2 is embryo-lethal, RNA-seq performed on hypomorphic atrtf2 alleles showed that 13%–16% of all introns are retained.
ROOT INITIATION DEFECTIVE1 (RID1) encodes a DEAH-box RNA helicase similar to the splicing factors Prp22 in yeast and DEAH-box polypeptide8 (DHX8) in humans that is involved in remodeling of the spliceosome [132]. In the rid1-1 mutant, several transcripts were aberrantly spliced and the mutant shows defects in leaf and root morphogenesis.
Mutants deficient in the subunits of the cap-binding complex (CBC), At-CBP20, and At-CBP80, which interact with the m7G cap of pre-mRNAs, also show widespread changes in splicing patterns [133,134]. In particular, splicing of the first intron is influenced in mutants deficient in the CBC.

3. Factors with a Dual Role in Splicing of Pre-mRNAs and microRNA Precursors

Several Arabidopsis splicing factors affect splicing of pre-mRNAs and biogenesis of microRNAs (miRNAs). miRNAs, 21–24 nucleotide-long single-stranded RNAs, are generated from RNA polymerase II transcripts with internal stem-loop structures, the pri (primary)-miRNAs that are processed in a step-wise manner by RNase III-like proteins to yield the mature miRNAs. Plant miRNA biogenesis differs from miRNA biogenesis in animals. While in animals miRNAs frequently localize to introns of pre-mRNAs or non-coding RNAs, most plant miRNAs are derived from independent transcription units. In Arabidopsis, both conversion of pri-miRNAs to stem-loop pre (precursor)-miRNAs and further processing to miRNA/miRNA* duplexes take place in the nuclear compartment and are executed by the same enzyme, DICER-LIKE1 (DCL1) [135]. Mature miRNAs are loaded into Argonaute (AGO)-containing RNA-induced silencing complexes and guide AGO proteins to complementary sites in their targets, leading to down-regulation of the transcripts predominantly by AGO cleavage in plants, but also by inhibition of translation [136].
Several RNA-binding proteins contribute to processing of the pri-miRNAs: the double-stranded RNA-binding protein HYPONASTIC LEAVES1 (HYL1), the zinc finger protein SERRATE (SE), a homolog of ARSENITE RESISTANT 2, and the G-patch domain protein TOUGH (TGH) [137]. Together with DCL1, these proteins form the so-called Microprocessor complex that is tethered to the pri-miRNA via interaction with the cap binding proteins At-CBP20 and At-CBP80 [134,138,139]. SE and HYL1 promote accurate pri-miRNA processing by DCL1 [140,141]. TGH enhances DCL1 activity without influencing the processing accuracy [137]. DAWDLE (DDL) is a forkhead-associated (FHA) domain protein that is assumed to recruit DCL1 to pri-miRNAs [142]. DDL is a homolog of SNIP1 (Smad nuclear interacting protein 1) that is recruited to the spliceosome at the transition from complex A to complex B in humans [143]. Up to now, DDL has not yet been associated with the splicing process in Arabidopsis.
The subunits of the CBC, At-CBP20, and At-CBP80 thus have a dual role in alternative splicing and pri-miRNA processing [133,134]. Similarly, SE is required for correct splicing of numerous pre-mRNAs [144]. There was a substantial overlap in the transcripts targeted by either SE or the CBC, and both SE and the nuclear CBC appeared to act in the selection of 5' splice sites and mainly on the first introns. Aberrant alternative splicing patterns were also found in the hyl1-2 and dcl1-7 mutants but these mutants shared little overlap with the se mutant. Thus, SE and CBC have distinct effects on alternative splicing from the other microRNA processing components DCL1 and HYL1. In addition to the general processing components, AtGRP7 also affects both pre-mRNAs and the processing of a subset of pri-miRNAs [145]. Furthermore, the NTC components CDC5 and PRL1 have been found to interact with the DCL-1 complex and enhance pri-miRNA processing [146,147]. Recently, STA1 has also been implicated in miRNA biogenesis, as intron-containing pri-miRNAs are incorrectly spliced in sta1, providing another example of a protein with dual function in pre-mRNA and pri-miRNA splicing [148].

4. Conclusions/Outlook

Relying on Arabidopsis thaliana as a model organism with ample genetic resources has greatly advanced our knowledge of plant splicing factors. In the absence of an in vitro splicing system, the analysis of mutants has proven revealing in elucidating a function of candidate RNA binding proteins in alternative splicing. Additionally, proteomic analysis of oligo(dT)-captured mRNPs has globally detected mRNA binding proteins in Arabidopsis [149]. However, mechanistic details on how individual factors work at the molecular level remain to be resolved. Furthermore, other splicing regulators may employ noncanonical RNA binding domains and thus may have been missed in computational predictions. For example, hundreds of proteins previously unknown to bind RNA have been identified in a proteomic approach in human cells [150]. Moreover, about half of the RRM-containing proteins in Arabidopsis appear to be specific to plants [151]. Thus, plants may have evolved their own specialized splicing regulators that are not recovered based on sequence similarity. Such additional regulators await experimental determination from genome-wide transcript profiling and target gene identification.

Acknowledgments

Katja Meyer is supported by the German National Academic Foundation. Work in our laboratory is supported by the DFG through STA653 and SPP1530. We acknowledge support for the Article Processing Charge by the Deutsche Forschungsgemeinschaft and the Open Access Publication Fund of Bielefeld University.

Author Contributions

Katja Meyer and Tino Koester commented on the text and prepared figures. Dorothee Staiger wrote the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Berget, S.M.; Moore, C.; Sharp, P.A. Spliced segments at the 5' terminus of adenovirus 2 late mRNA. Proc. Natl. Acad. Sci. USA 1977, 74, 3171–3175. [Google Scholar] [CrossRef] [PubMed]
  2. Chow, L.T.; Gelinas, R.E.; Broker, T.R.; Roberts, R.J. An amazing sequence arrangement at the 5' ends of adenovirus 2 messenger RNA. Cell 1977, 12, 1–8. [Google Scholar] [CrossRef]
  3. Wahl, M.C.; Will, C.L.; Lührmann, R. The spliceosome: Design principles of a dynamic RNP machine. Cell 2009, 136, 701–718. [Google Scholar] [CrossRef] [PubMed]
  4. Will, C.L.; Lührmann, R. Spliceosome Structure and Function. Cold Spring Harbor Perspect. Biol. 2010. [Google Scholar] [CrossRef] [PubMed]
  5. Matera, A.G.; Wang, Z. A day in the life of the spliceosome. Nat. Rev. Mol. Cell Biol. 2014, 15, 108–121. [Google Scholar] [CrossRef] [PubMed]
  6. Chen, W.; Moore, M.J. Spliceosomes. Curr. Biol. 2015, 25, R181–R183. [Google Scholar] [CrossRef] [PubMed]
  7. Tharun, S. Roles of eukaryotic LSM proteins in the regulation of mRNA function. Int. Rev. Cell. Mol. Biol. 2009, 272, 149–189. [Google Scholar] [PubMed]
  8. Cordin, O.; Beggs, J.D. RNA helicases in splicing. RNA Biol. 2013, 10, 83–95. [Google Scholar] [CrossRef] [PubMed]
  9. Blencowe, B.J. Alternative splicing: New insights from global analyses. Cell 2006, 126, 37–47. [Google Scholar] [CrossRef] [PubMed]
  10. Smith, C.W.J.; Valcarel, J. Alternative pre-mRNA splicing: The logic of combinatorial control. Trends Biochem. Sci. 2000, 25, 381–388. [Google Scholar] [CrossRef]
  11. Nilsen, T.W.; Graveley, B.R. Expansion of the eukaryotic proteome by alternative splicing. Nature 2010, 463, 457–463. [Google Scholar] [CrossRef] [PubMed]
  12. McGlincy, N.J.; Smith, C.W. Alternative splicing resulting in nonsense-mediated mRNA decay: What is the meaning of nonsense? Trends Biochem. Sci. 2008, 33, 385–393. [Google Scholar] [CrossRef] [PubMed]
  13. Shen, H.; Kan, J.L.; Green, M.R. Arginine-serine-rich domains bound at splicing enhancers contact the branchpoint to promote prespliceosome assembly. Mol. Cell 2004, 13, 367–376. [Google Scholar] [CrossRef]
  14. Han, S.P.; Tang, Y.H.; Smith, R. Functional diversity of the hnRNPs: Past, present and perspectives. Biochem. J. 2010, 430, 379–392. [Google Scholar] [CrossRef] [PubMed]
  15. Wang, Y.; Xiao, X.; Zhang, J.; Choudhury, R.; Robertson, A.; Li, K.; Ma, M.; Burge, C.B.; Wang, Z. A complex network of factors with overlapping affinities repress splicing through intronic elements. Nat. Struct. Mol. Biol. 2013, 20, 36–45. [Google Scholar] [CrossRef] [PubMed]
  16. Ule, J.; Stefani, G.; Mele, A.; Ruggiu, M.; Wang, X.; Taneri, B.; Gaasterland, T.; Blencowe, B.J.; Darnell, R.B. An RNA map predicting NOVA-dependent splicing regulation. Nature 2006, 444, 580–586. [Google Scholar] [CrossRef] [PubMed]
  17. Tazi, J.; Bakkour, N.; Stamm, S. Alternative splicing and disease. Biochim. Biophys. Acta Mol. Basis Dis. 2009, 1792, 14–26. [Google Scholar] [CrossRef] [PubMed]
  18. Liu, M.M.; Zack, D.J. Alternative splicing and retinal degeneration. Clin. Genet. 2013, 84, 142–149. [Google Scholar] [CrossRef] [PubMed]
  19. Ladomery, M. Aberrant alternative splicing is another hallmark of cancer. Int. J. Cell Biol. 2013. [Google Scholar] [CrossRef] [PubMed]
  20. Reddy, A.S. Alternative splicing of pre-messenger RNAs in plants in the genomic Era. Ann. Rev. Plant Biol. 2007, 58, 267–294. [Google Scholar] [CrossRef] [PubMed]
  21. Reddy, A.S.N.; Marquez, Y.; Kalyna, M.; Barta, A. Complexity of the alternative splicing landscape in plants. Plant Cell 2013, 25, 3657–3683. [Google Scholar] [CrossRef] [PubMed]
  22. Filichkin, S.A.; Priest, H.D.; Givan, S.A.; Shen, R.; Bryant, D.W.; Fox, S.E.; Wong, W.K.; Mockler, T.C. Genome-wide mapping of alternative splicing in Arabidopsis thaliana. Genome Res. 2010, 20, 45–58. [Google Scholar] [CrossRef] [PubMed]
  23. Syed, N.H.; Kalyna, M.; Marquez, Y.; Barta, A.; Brown, J.W.S. Alternative splicing in plants-coming of age. Trends Plant Sci. 2012, 17, 616–623. [Google Scholar] [CrossRef] [PubMed]
  24. Staiger, D.; Brown, J.W.S. Alternative splicing at the intersection of biological timing, development, and stress responses. Plant Cell 2013, 25, 3640–3656. [Google Scholar] [CrossRef] [PubMed]
  25. Staiger, D.; Korneli, C.; Lummer, M.; Navarro, L. Emerging role for RNA-based regulation in plant immunity. New Phytol. 2013, 197, 394–404. [Google Scholar] [CrossRef] [PubMed]
  26. Loraine, A.; McCormick, S.; Estrada, A.; Patel, K.; Qin, P. High-throughput sequencing of Arabidopsis thaliana pollen cDNA uncovers novel transcription and alternative splicing. Plant Physiol. 2013, 162, 1092–1109. [Google Scholar] [CrossRef] [PubMed]
  27. Staiger, D. Shaping the Arabidopsis transcriptome through alternative splicing. Adv. Bot. 2015. [Google Scholar] [CrossRef]
  28. Marquez, Y.; Brown, J.W.S.; Simpson, C.G.; Barta, A.; Kalyna, M. Transcriptome survey reveals increased complexity of the alternative splicing landscape in Arabidopsis. Genome Res. 2012, 22, 1184–1195. [Google Scholar] [CrossRef] [PubMed]
  29. Brown, J.W.S.; Feix, G.; Frendewey, D. Accurate in vitro splicing of two pre-mRNA plant introns in a HeLa cell nuclear extract. Embo J. 1986, 5, 2749–2758. [Google Scholar] [PubMed]
  30. Barta, A.; Sommergruber, K.; Thompson, D.; Hartmuth, K.; Matzke, M.; Matzke, A.M. The expression of a nopaline synthase—Human growth hormone chimaeric gene in transformed tobacco and sunflower callus tissue. Plant Mol. Biol. 1986, 6, 347–357. [Google Scholar] [CrossRef] [PubMed]
  31. Wiebauer, K.; Herrero, J.J.; Filipowicz, W. Nuclear pre-mRNA processing in plants: Distinct modes of 3'-splice-site selection in plants and animals. Mol. Cell. Biol. 1988, 8, 2042–2051. [Google Scholar] [PubMed]
  32. Iwata, H.; Gotoh, O. Comparative analysis of information contents relevant to recognition of introns in many species. BMC Genomics 2011. [Google Scholar] [CrossRef] [PubMed]
  33. Wang, B.B.; Brendel, V. The ASRG database: Identification and survey of Arabidopsis thaliana genes involved in pre-mRNA splicing. Genome Biol. 2004. [Google Scholar] [CrossRef] [Green Version]
  34. Koncz, C.; deJong, F.; Villacorta, N.; Szakonyi, D.; Koncz, Z. The spliceosome-activating complex: Molecular mechanisms underlying the function of a pleiotropic regulator. Front. Plant Sci. 2012. [Google Scholar] [CrossRef] [PubMed]
  35. Reddy, A.S. Plant serine/arginine-rich proteins and their role in pre-mRNA splicing. Trends Plant Sci. 2004, 9, 541–547. [Google Scholar] [CrossRef] [PubMed]
  36. Lopato, S.; Mayeda, A.; Krainer, A.R.; Barta, A. Pre-mRNA splicing in plants: Characterization of Ser/Arg splicing factors. Proc. Natl. Acad. Sci. USA 1996, 93, 3074–3079. [Google Scholar] [CrossRef] [PubMed]
  37. Lazar, G.; Schaal, T.; Maniatis, T.; Goodman, H.M. Identification of a plant serine-arginine-rich protein similar to the mammalian splicing factor SF2/ASF. Proc. Natl. Acad. Sci. USA 1995, 92, 7672–7676. [Google Scholar] [CrossRef] [PubMed]
  38. Barta, A.; Kalyna, M.; Reddy, A.S. Implementing a rational and consistent nomenclature for Serine/Arginine-rich protein splicing factors (SR Proteins) in plants. Plant Cell 2010, 22, 2926–2929. [Google Scholar] [CrossRef] [PubMed]
  39. Duque, P. A role for SR proteins in plant stress responses. Plant Signal. Behav. 2011, 6, 49–54. [Google Scholar] [CrossRef] [PubMed]
  40. Richardson, D.N.; Rogers, M.F.; Labadorf, A.; Ben-Hur, A.; Guo, H.; Paterson, A.H.; Reddy, A.S. Comparative analysis of Serine/Arginine-rich proteins across 27 eukaryotes: Insights into sub-family classification and extent of alternative splicing. PLoS ONE 2011, 6, e24542. [Google Scholar] [CrossRef] [PubMed]
  41. Kalyna, M.; Lopato, S.; Voronin, V.; Barta, A. Evolutionary conservation and regulation of particular alternative splicing events in plant SR proteins. Nucleic Acids Res. 2006, 34, 4395–4405. [Google Scholar] [CrossRef] [PubMed]
  42. Lopato, S.; Forstner, C.; Kalyna, M.; Hilscher, J.; Langhammer, U.; Indrapichate, K.; Lorkovic, Z.J.; Barta, A. Network of interactions of a novel plant-specific Arg/Ser-rich protein, atRSZ33, with atSC35-like splicing factors. J. Biol. Chem. 2002, 277, 39989–39998. [Google Scholar] [CrossRef] [PubMed]
  43. Kalyna, M.; Lopato, S.; Barta, A. Ectopic expression of atRSZ33 reveals its function in splicing and causes pleiotropic changes in development. Mol. Biol. Cell 2003, 14, 3565–3577. [Google Scholar] [CrossRef] [PubMed]
  44. Lopato, S.; Kalyna, M.; Dorner, S.; Kobayashi, R.; Krainer, A.R.; Barta, A. atSRp30, one of two SF2/ASF-like proteins from Arabidopsis thaliana, regulates splicing of specific plant genes. Genes Dev. 1999, 13, 987–1001. [Google Scholar] [CrossRef] [PubMed]
  45. Petrillo, E.; Godoy Herz, M.A.; Fuchs, A.; Reifer, D.; Fuller, J.; Yanovsky, M.J.; Simpson, C.; Brown, J.W.S.; Barta, A.; Kalyna, M.; et al. A Chloroplast Retrograde Signal Regulates Nuclear Alternative Splicing. Science 2014, 344, 427–430. [Google Scholar] [CrossRef] [PubMed]
  46. Zhang, W.; Du, B.; Liu, D.; Qi, X. Splicing factor SR34b mutation reduces cadmium tolerance in Arabidopsis by regulating iron-regulated transporter 1 gene. Biochem. Biophys. Res. Commun. 2014, 455, 312–317. [Google Scholar] [CrossRef] [PubMed]
  47. Lazar, G.; Goodman, H. The Arabidopsis splicing factor SR1 is regulated by alternative splicing. Plant Mol. Biol. 2000, 42, 571–581. [Google Scholar] [CrossRef] [PubMed]
  48. Golovkin, M.; Reddy, A.S. The plant U1 small nuclear ribonucleoprotein particle 70K protein interacts with two novel Serine/Arginine-rich proteins. Plant Cell 1998, 10, 1637–1648. [Google Scholar] [PubMed]
  49. Lopato, S.; Waigmann, E.; Barta, A. Characterization of a novel arginine/serine-rich splicing factor in Arabidopsis. Plant Cell 1996, 8, 2255–2264. [Google Scholar] [CrossRef] [PubMed]
  50. Palusa, S.G.; Reddy, A.S. Extensive coupling of alternative splicing of pre-mRNAs of Serine/Arginine (SR) genes with nonsense-mediated decay. New Phytol. 2010, 185, 83–89. [Google Scholar] [CrossRef] [PubMed]
  51. Kalyna, M.; Simpson, C.G.; Syed, N.H.; Lewandowska, D.; Marquez, Y.; Kusenda, B.; Marshall, J.; Fuller, J.; Cardle, L.; McNicol, J.; et al. Alternative splicing and nonsense-mediated decay modulate expression of important regulatory genes in Arabidopsis. Nucleic Acids Res. 2012, 40, 2454–2469. [Google Scholar] [CrossRef] [PubMed]
  52. Iida, K.; Go, M. Survey of conserved alternative splicing events of mRNAs encoding SR Proteins in land plants. Mol. Biol. Evol. 2006, 23, 1085–1094. [Google Scholar] [CrossRef] [PubMed]
  53. Thomas, J.; Palusa, S.G.; Prasad, K.V.S.K.; Ali, G.S.; Surabhi, G.-K.; Ben-Hur, A.; Abdel-Ghany, S.E.; Reddy, A.S.N. Identification of an intronic splicing regulatory element involved in autoregulation of alternative splicing of the SCL33 pre-mRNA. Plant J. 2012, 72, 935–946. [Google Scholar] [PubMed]
  54. Lykke-Andersen, J.; Shu, M.D.; Steitz, J.A. Communication of the position of exon-exon junctions to the mRNA surveillance machinery by the protein RNPS1. Science 2001, 293, 1836–1839. [Google Scholar] [CrossRef] [PubMed]
  55. Ali, G.S.; Palusa, S.G.; Golovkin, M.; Prasad, J.; Manley, J.L.; Reddy, A.S.N. Regulation of plant developmental processes by a novel splicing factor. PLoS ONE 2007, 2, e471. [Google Scholar] [CrossRef] [PubMed]
  56. Zhang, X.N.; Mount, S.M. Two alternatively spliced isoforms of the Arabidopsis thaliana SR45 protein have distinct roles during normal plant development. Plant Physiol. 2009, 150, 1450–1458. [Google Scholar] [CrossRef] [PubMed]
  57. Zhang, X.-N.; Mo, C.; Garrett, W.M.; Cooper, B. Phosphothreonine 218 is required for the function of SR45.1 in regulating flower petal development in Arabidopsis. Plant Signal. Behav. 2014. [Google Scholar] [CrossRef]
  58. Carvalho, R.F.; Carvalho, S.D.; Duque, P. The plant-specific SR45 protein negatively regulates glucose and ABA signaling during early seedling development in Arabidopsis. Plant Physiol. 2010, 154, 772–783. [Google Scholar] [CrossRef] [PubMed]
  59. Filichkin, S.A.; Cumbie, J.S.; Dharmawadhana, J.P.; Jaiswal, P.; Chang, J.H.; Palusa, S.G.; Reddy, A.S.N.; Megraw, M.; Mockler, T.C. Environmental stresses modulate abundance and timing of alternatively spliced circadian transcripts in Arabidopsis. Mol. Plant 2015, 8, 207–227. [Google Scholar] [CrossRef] [PubMed]
  60. Day, I.S.; Golovkin, M.; Palusa, S.G.; Link, A.; Ali, G.S.; Thomas, J.; Richardson, D.N.; Reddy, A.S.N. Interactions of SR45, an SR-like protein, with spliceosomal proteins and an intronic sequence: Insights into regulated splicing. Plant J. 2012, 71, 936–947. [Google Scholar] [CrossRef] [PubMed]
  61. De la Fuente van Bentem, S.; Anrather, D.; Roitinger, E.; Djamei, A.; Hufnagl, T.; Barta, A.; Csaszar, E.; Dohnal, I.; Lecourieux, D.; Hirt, H. Phosphoproteomics reveals extensive in vivo phosphorylation of Arabidopsis proteins involved in RNA metabolism. Nucleic Acids Res. 2006, 34, 3267–3278. [Google Scholar] [CrossRef] [PubMed]
  62. Feilner, T.; Hultschig, C.; Lee, J.; Meyer, S.; Immink, R.G.; Koenig, A.; Possling, A.; Seitz, H.; Beveridge, A.; Scheel, D.; et al. High throughput identification of potential Arabidopsis mitogen-activated protein kinases substrates. Mol. Cell. Proteomics 2005, 4, 1558–1568. [Google Scholar] [CrossRef] [PubMed]
  63. Savaldi-Goldstein, S.; Aviv, D.; Davydov, O.; Fluhr, R. Alternative splicing modulation by a LAMMER Kinase impinges on developmental and transcriptome expression. Plant Cell 2003, 15, 926–938. [Google Scholar] [CrossRef] [PubMed]
  64. Wachter, A.; Rühl, C.; Stauffer, E. The role of polypyrimidine tract-binding proteins and other hnRNP proteins in plant splicing regulation. Front. Plant Sci. 2012. [Google Scholar] [CrossRef] [PubMed]
  65. Izquierdo, J.M.; Majós, N.; Bonnal, S.; Martínez, C.; Castelo, R.; Guigó, R.; Bilbao, D.; Valcárcel, J. Regulation of fas alternative splicing by antagonistic effects of TIA-1 and PTB on exon definition. Mol. Cell 2005, 19, 475–484. [Google Scholar] [CrossRef] [PubMed]
  66. Boutz, P.L.; Stoilov, P.; Li, Q.; Lin, C.H.; Chawla, G.; Ostrow, K.; Shiue, L.; Ares, M., Jr.; Black, D.L. A post-transcriptional regulatory switch in polypyrimidine tract-binding proteins reprograms alternative splicing in developing neurons. Genes Dev. 2007, 21, 1636–1652. [Google Scholar] [CrossRef] [PubMed]
  67. Spellman, R.; Llorian, M.; Smith, C.W. Crossregulation and functional redundancy between the splicing regulator PTB and its paralogs nPTB and ROD1. Mol. Cell 2007, 27, 420–434. [Google Scholar] [CrossRef] [PubMed]
  68. Stauffer, E.; Westermann, A.; Wagner, G.; Wachter, A. Polypyrimidine tract-binding protein homologues from Arabidopsis underlie regulatory circuits based on alternative splicing and downstream control. Plant J. 2010, 64, 243–255. [Google Scholar] [CrossRef] [PubMed]
  69. Rühl, C.; Stauffer, E.; Kahles, A.; Wagner, G.; Drechsel, G.; Rätsch, G.; Wachter, A. Polypyrimidine tract binding protein homologs from arabidopsis are key regulators of alternative splicing with implications in fundamental developmental processes. Plant Cell 2012, 24, 4360–4375. [Google Scholar] [CrossRef] [PubMed]
  70. Simpson, C.G.; Lewandowska, D.; Liney, M.; Davidson, D.; Chapman, S.; Fuller, J.; McNicol, J.; Shaw, P.; Brown, J.W.S. Arabidopsis PTB1 and PTB2 proteins negatively regulate splicing of a mini-exon splicing reporter and affect alternative splicing of endogenous genes differentially. New Phytol. 2014, 203, 424–436. [Google Scholar] [CrossRef] [PubMed]
  71. Penfield, S.; Josse, E.M.; Halliday, K.J. A role for an alternative splice variant of PIF6 in the control of Arabidopsis primary seed dormancy. Plant Mol. Biol. 2010, 73, 89–95. [Google Scholar] [CrossRef] [PubMed]
  72. Mangeon, A.; Junqueira, R.M.; Sachetto-Martins, G. Functional diversity of the plant glycine-rich proteins superfamily. Plant Signal. Behav. 2010, 5, 99–104. [Google Scholar] [CrossRef] [PubMed]
  73. Ciuzan, O.; Hancock, J.; Pamfil, D.; Wilson, I.; Ladomery, M. The evolutionarily conserved multifunctional glycine-rich RNA-binding proteins play key roles in development and stress adaptation. Physiol. Plant. 2015, 153, 1–11. [Google Scholar] [CrossRef] [PubMed]
  74. Carpenter, C.D.; Kreps, J.A.; Simon, A.E. Genes encoding glycine-rich Arabidopsis thaliana proteins with RNA-binding motifs are influenced by cold treatment and an endogenous circadian rhythm. Plant Physiol. 1994, 104, 1015–1025. [Google Scholar] [CrossRef] [PubMed]
  75. Staiger, D.; Heintzen, C. The circadian system of Arabidopsis thaliana: Forward and reverse genetic approaches. Chronobiol. Int. 1999, 16, 1–16. [Google Scholar] [CrossRef] [PubMed]
  76. Staiger, D.; Zecca, L.; Wieczorek Kirk, D.A.; Apel, K.; Eckstein, L. The circadian clock regulated RNA-binding protein AtGRP7 autoregulates its expression by influencing alternative splicing of its own pre-mRNA. Plant J. 2003, 33, 361–371. [Google Scholar] [CrossRef] [PubMed]
  77. Schöning, J.C.; Streitner, C.; Meyer, I.M.; Gao, Y.; Staiger, D. Reciprocal regulation of glycine-rich RNA-binding proteins via an interlocked feedback loop coupling alternative splicing to nonsense-mediated decay in Arabidopsis. Nucleic Acids Res. 2008, 36, 6977–6987. [Google Scholar] [CrossRef] [PubMed]
  78. Simpson, C.G.; Fuller, J.; Maronova, M.; Kalyna, M.; Davidson, D.; McNicol, J.; Barta, A.; Brown, J.W. Monitoring changes in alternative precursor messenger RNA splicing in multiple gene transcripts. Plant J. 2008, 53, 1035–1048. [Google Scholar] [CrossRef] [PubMed]
  79. Streitner, C.; Köster, T.; Simpson, C.G.; Shaw, P.; Danisman, S.; Brown, J.W.S.; Staiger, D. An hnRNP-like RNA-binding protein affects alternative splicing by in vivo interaction with target transcripts in Arabidopsis thaliana. Nucleic Acids Res. 2012, 40, 11240–11255. [Google Scholar] [CrossRef] [PubMed]
  80. Köster, T.; Haas, M.; Staiger, D. The RIPper case: Identification of RNA-binding protein targets by RNA immunoprecipitation. Methods Mol. Biol. 2014, 1158, 107–121. [Google Scholar] [PubMed]
  81. Nicaise, V.; Joe, A.; Jeong, B.; Korneli, C.; Boutrot, F.; Wested, I.; Staiger, D.; Alfano, J.R.; Zipfel, C. Pseudomonas HopU1 affects interaction of plant immune receptor mRNAs to the RNA-binding protein GRP7. EMBO J. 2013, 32, 701–712. [Google Scholar] [CrossRef] [PubMed]
  82. Hackmann, C.; Korneli, C.; Kutyniok, M.; Köster, T.; Wiedenlübbert, M.; Müller, C.; Staiger, D. Salicylic acid-dependent and -independent impact of an RNA-binding protein on plant immunity. Plant Cell Environ. 2014, 37, 696–706. [Google Scholar] [CrossRef] [PubMed]
  83. Fu, Z.Q.; Guo, M.; Jeong, B.R.; Tian, F.; Elthon, T.E.; Cerny, R.L.; Staiger, D.; Alfano, J.R. A type III effector ADP-ribosylates RNA-binding proteins and quells plant immunity. Nature 2007, 447, 284–288. [Google Scholar] [CrossRef] [PubMed]
  84. Jeong, B.-R.; Lin, Y.; Joe, A.; Guo, M.; Korneli, C.; Yang, H.; Wang, P.; Yu, M.; Cerny, R.L.; Staiger, D.; et al. Structure function analysis of an ADP-ribosyltransferase type III effector and its RNA-binding target in plant immunity. J. Biol. Chem. 2011, 286, 43272–43281. [Google Scholar] [CrossRef] [PubMed]
  85. Swaraz, A.M.; Park, Y.D.; Hur, Y. Knock-out mutations of Arabidopsis SmD3-b induce pleotropic phenotypes through altered transcript splicing. Plant Sci. 2011, 180, 661–671. [Google Scholar] [CrossRef] [PubMed]
  86. Zhang, Z.; Zhang, S.; Zhang, Y.; Wang, X.; Li, D.; Li, Q.; Yue, M.; Li, Q.; Zhang, Y.E.; Xu, Y.; et al. Arabidopsis floral initiator SKB1 confers high salt tolerance by regulating transcription and pre-mRNA splicing through altering histone H4R3 and small nuclear ribonucleoprotein LSM4 methylation. Plant Cell 2011, 23, 396–411. [Google Scholar] [CrossRef] [PubMed]
  87. Perez-Santángelo, S.; Mancini, E.; Francey, L.J.; Schlaen, R.G.; Chernomoretz, A.; Hogenesch, J.B.; Yanovsky, M.J. Role for LSM genes in the regulation of circadian rhythms. Proc. Natl. Acad. Sci. USA 2014, 111, 15166–15171. [Google Scholar] [CrossRef] [PubMed]
  88. Golisz, A.; Sikorski, P.J.; Kruszka, K.; Kufel, J. Arabidopsis thaliana LSM proteins function in mRNA splicing and degradation. Nucleic Acids Res. 2013, 41, 6232–6249. [Google Scholar] [CrossRef] [PubMed]
  89. Cui, P.; Zhang, S.; Ding, F.; Ali, S.; Xiong, L. Dynamic regulation of genome-wide pre-mRNA splicing and stress tolerance by the Sm-like protein LSm5 in Arabidopsis. Genome Biol. 2014. [Google Scholar] [CrossRef] [PubMed]
  90. Meister, G.; Eggert, C.; Buhler, D.; Brahms, H.; Kambach, C.; Fischer, U. Methylation of Sm proteins by a complex containing PRMT5 and the putative U snRNP assembly factor pICln. Curr. Biol. 2001, 11, 1990–1994. [Google Scholar] [CrossRef]
  91. Friesen, W.J.; Paushkin, S.; Wyce, A.; Massenet, S.; Pesiridis, G.S.; van Duyne, G.; Rappsilber, J.; Mann, M.; Dreyfuss, G. The methylosome, a 20S complex containing JBP1 and pICln, produces dimethylarginine-modified Sm proteins. Mol. Cell Biol. 2001, 21, 8289–8300. [Google Scholar] [CrossRef] [PubMed]
  92. Pei, Y.; Niu, L.; Lu, F.; Liu, C.; Zhai, J.; Kong, X.; Cao, X. Mutations in the Type II protein arginine methyltransferase AtPRMT5 result in pleiotropic developmental defects in Arabidopsis. Plant Physiol. 2007, 144, 1913–1923. [Google Scholar] [CrossRef] [PubMed]
  93. Hernando, C.E.; Sanchez, S.E.; Mancini, E.; Yanovsky, M.J. Genome wide comparative analysis of the effects of PRMT5 and PRMT4/CARM1 arginine methyltransferases on the Arabidopsis thaliana transcriptome. BMC Genomics 2015. [Google Scholar] [CrossRef] [PubMed]
  94. Sanchez, S.E.; Petrillo, E.; Beckwith, E.J.; Zhang, X.; Rugnone, M.L.; Hernando, C.E.; Cuevas, J.C.; Godoy Herz, M.A.; Depetris-Chauvin, A.; Simpson, C.G.; et al. A methyl transferase links the circadian clock to the regulation of alternative splicing. Nature 2010, 468, 112–116. [Google Scholar] [CrossRef] [PubMed]
  95. Hong, S.; Song, H.R.; Lutz, K.; Kerstetter, R.A.; Michael, T.P.; McClung, C.R. Type II protein arginine methyltransferase 5 (PRMT5) is required for circadian period determination in Arabidopsis thaliana. Proc. Natl. Acad. Sci. USA 2010, 107, 21211–21216. [Google Scholar] [CrossRef] [PubMed]
  96. Deng, X.; Gu, L.; Liu, C.; Lu, T.; Lu, F.; Lu, Z.; Cui, P.; Pei, Y.; Wang, B.; Hu, S.; et al. Arginine methylation mediated by the Arabidopsis homolog of PRMT5 is essential for proper pre-mRNA splicing. Proc. Natl. Acad. Sci. USA 2010, 107, 19114–19119. [Google Scholar] [CrossRef] [PubMed]
  97. Cheng, D.; Côté, J.; Shaaban, S.; Bedford, M.T. The arginine methyltransferase CARM1 regulates the coupling of transcription and mRNA processing. Mol. Cell 2007, 25, 71–83. [Google Scholar] [CrossRef] [PubMed]
  98. Niu, L.; Zhang, Y.; Pei, Y.; Liu, C.; Cao, X. Redundant requirement for a pair of PROTEIN ARGININE METHYLTRANSFERASE4 homologs for the proper regulation of Arabidopsis flowering time. Plant Physiol. 2008, 148, 490–503. [Google Scholar] [CrossRef] [PubMed]
  99. Lee, B.H.; Kapoor, A.; Zhu, J.; Zhu, J.K. STABILIZED1, a stress-upregulated nuclear protein, is required for pre-mRNA splicing, mRNA turnover, and stress tolerance in Arabidopsis. Plant Cell 2006, 18, 1736–1749. [Google Scholar] [CrossRef] [PubMed]
  100. Thalhammer, A.; Bryant, G.; Sulpice, R.; Hincha, D.K. Disordered cold regulated15 proteins protect chloroplast membranes during freezing through binding and folding, but do not stabilize chloroplast enzymes in vivo. Plant Physiol. 2014, 166, 190–201. [Google Scholar] [CrossRef] [PubMed]
  101. Du, J.-L.; Zhang, S.-W.; Huang, H.-W.; Cai, T.; Li, L.; Chen, S.; He, X.-J. The splicing factor PRP31 is involved in transcriptional gene silencing and stress response in Arabidopsis. Mol. Plant 2015, 8, 1053–1068. [Google Scholar] [CrossRef] [PubMed]
  102. Gross-Hardt, R.; Kägi, C.; Baumann, N.; Moore, J.M.; Baskar, R.; Gagliano, W.B.; Jürgens, G.; Grossniklaus, U. LACHESIS restricts gametic cell fate in the female gametophyte of Arabidopsis. PLoS Biol. 2007, 5, e47. [Google Scholar] [CrossRef] [PubMed]
  103. Liu, M.; Yuan, L.; Liu, N.Y.; Shi, D.Q.; Liu, J.; Yang, W.C. GAMETOPHYTIC FACTOR 1, involved in pre-mRNA splicing, is essential for megagametogenesis and embryogenesis in Arabidopsis. J. Integr. Plant Biol. 2009, 51, 261–271. [Google Scholar] [CrossRef] [PubMed]
  104. Moll, C.; von Lyncker, L.; Zimmermann, S.; Kägi, C.; Baumann, N.; Twell, D.; Grossniklaus, U.; Groß-Hardt, R. CLO/GFA1 and ATO are novel regulators of gametic cell fate in plants. Plant J. 2008, 56, 913–921. [Google Scholar] [CrossRef] [PubMed]
  105. Hornyik, C.; Duc, C.; Rataj, K.; Terzi, L.C.; Simpson, G.G. Alternative polyadenylation of antisense RNAs and flowering time control. Biochem. Soc. Trans. 2010, 38, 1077–1081. [Google Scholar] [CrossRef] [PubMed]
  106. Pena, V.; Rozov, A.; Fabrizio, P.; Lührmann, R.; Wahl, M.C. Structure and function of an RNase H domain at the heart of the spliceosome. EMBO J. 2008, 27, 2929–2940. [Google Scholar] [CrossRef] [PubMed]
  107. Grainger, R.J.; Beggs, J.D. Prp8 protein: At the heart of the spliceosome. RNA 2005, 11, 533–557. [Google Scholar] [CrossRef] [PubMed]
  108. Marquardt, S.; Raitskin, O.; Wu, Z.; Liu, F.; Sun, Q.; Dean, C. Functional consequences of splicing of the antisense transcript COOLAIR on FLC transcription. Mol. Cell 2014, 54, 156–165. [Google Scholar] [CrossRef] [PubMed]
  109. Wang, C.; Tian, Q.; Hou, Z.; Mucha, M.; Aukerman, M.; Olsen, O.A. The Arabidopsis thaliana AT PRP39-1 gene, encoding a tetratricopeptide repeat protein with similarity to the yeast pre-mRNA processing protein PRP39, affects flowering time. Plant Cell Rep. 2007, 26, 1357–1366. [Google Scholar] [CrossRef] [PubMed]
  110. Song, H.-R.; Song, J.-D.; Cho, J.-N.; Amasino, R.M.; Noh, B.; Noh, Y.-S. The RNA binding protein ELF9 directly reduces SUPPRESSOR OF OVEREXPRESSION OF CO1 transcript levels in Arabidopsis, possibly via nonsense-mediated mRNA decay. Plant Cell 2009, 21, 1195–1211. [Google Scholar] [CrossRef] [PubMed]
  111. Sugliani, M.; Brambilla, V.; Clerkx, E.J.; Koornneef, M.; Soppe, W.J. The conserved splicing factor SUA controls alternative splicing of the developmental regulator ABI3 in Arabidopsis. Plant Cell 2010, 22, 1936–1946. [Google Scholar] [CrossRef] [PubMed]
  112. Hogg, R.; McGrail, J.C.; O’Keefe, R.T. The function of the NineTeen Complex (NTC) in regulating spliceosome conformations and fidelity during pre-mRNA splicing. Biochem. Soc. Trans. 2010, 38, 1110–1115. [Google Scholar] [CrossRef] [PubMed]
  113. Palma, K.; Zhao, Q.; Cheng, Y.T.; Bi, D.; Monaghan, J.; Cheng, W.; Zhang, Y.; Li, X. Regulation of plant innate immunity by three proteins in a complex conserved across the plant and animal kingdoms. Genes Dev. 2007, 21, 1484–1493. [Google Scholar] [CrossRef] [PubMed]
  114. Zhang, Y.; Goritschnig, S.; Dong, X.; Li, X. A gain-of-function mutation in a plant disease resistance gene leads to constitutive activation of downstream signal transduction pathways in suppressor of Npr1-1, constitutive 1. Plant Cell 2003, 15, 2636–2646. [Google Scholar] [CrossRef] [PubMed]
  115. Monaghan, J.; Xu, F.; Gao, M.; Zhao, Q.; Palma, K.; Long, C.; Chen, S.; Zhang, Y.; Li, X. Two Prp19-like U-box proteins in the MOS4-associated complex play redundant roles in plant innate immunity. PLoS Pathog. 2009, 5, e1000526. [Google Scholar] [CrossRef] [PubMed]
  116. Xu, F.; Xu, S.; Wiermer, M.; Zhang, Y.; Li, X. The cyclin L homolog MOS12 and the MOS4-associated complex are required for proper splicing of plant resistance genes. Plant J. 2012, 70, 916–928. [Google Scholar] [CrossRef] [PubMed]
  117. Zhang, Z.; Liu, Y.; Ding, P.; Li, Y.; Kong, Q.; Zhang, Y. Splicing of receptor-like kinase-encoding SNC4 and CERK1 is regulated by two conserved splicing factors that are required for plant immunity. Mol. Plant 2014, 7, 1766–1775. [Google Scholar] [CrossRef] [PubMed]
  118. Monaghan, J.; Xu, F.; Xu, S.; Zhang, Y.; Li, X. Two putative RNA-binding proteins function with unequal genetic redundancy in the MOS4-associated complex. Plant Physiol. 2010, 154, 1783–1793. [Google Scholar] [CrossRef] [PubMed]
  119. Rasche, N.; Dybkov, O.; Schmitzova, J.; Akyildiz, B.; Fabrizio, P.; Luhrmann, R. Cwc2 and its human homologue RBM22 promote an active conformation of the spliceosome catalytic centre. EMBO J. 2012, 31, 1591–1604. [Google Scholar] [CrossRef] [PubMed]
  120. Wang, X.; Wu, F.; Xie, Q.; Wang, H.; Wang, Y.; Yue, Y.; Gahura, O.; Ma, S.; Liu, L.; Cao, Y.; et al. SKIP Is a Component of the Spliceosome Linking Alternative Splicing and the Circadian Clock in Arabidopsis. Plant Cell 2012, 24, 3278–3295. [Google Scholar] [CrossRef] [PubMed]
  121. Feng, J.; Li, J.; Gao, Z.; Lu, Y.; Yu, J.; Zheng, Q.; Yan, S.; Zhang, W.; He, H.; Ma, L.; et al. SKIP confers osmotic tolerance during salt stress by controlling alternative gene splicing in Arabidopsis. Mol. Plant 2015, 8, 1038–1052. [Google Scholar] [CrossRef] [PubMed]
  122. Jones, M.A.; Williams, B.A.; McNicol, J.; Simpson, C.G.; Brown, J.W.S.; Harmer, S.L. Mutation of Arabidopsis SPLICEOSOMAL TIMEKEEPER LOCUS1 causes circadian clock defects. Plant Cell 2012, 24, 4907–4916. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Dias, A.P.; Dufu, K.; Lei, H.; Reed, R. A role for TREX components in the release of spliced mRNA from nuclear speckle domains. Nat. Commun. 2010. [Google Scholar] [CrossRef] [PubMed]
  124. Xu, C.; Zhou, X.; Wen, C.-K. HYPER RECOMBINATION1 of the THO/TREX complex plays a role in controlling transcription of the REVERSION-TO-ETHYLENE SENSITIVITY1 gene in Arabidopsis. PLoS Genet. 2015, 11, e1004956. [Google Scholar] [CrossRef] [PubMed]
  125. Furumizu, C.; Tsukaya, H.; Komeda, Y. Characterization of EMU, the Arabidopsis homolog of the yeast THO complex member HPR1. RNA 2010, 16, 1809–1817. [Google Scholar] [CrossRef] [PubMed]
  126. Francisco-Mangilet, A.G.; Karlsson, P.; Kim, M.-H.; Eo, H.J.; Oh, S.A.; Kim, J.H.; Kulcheski, F.R.; Park, S.K.; Manavella, P.A. THO2, core member of the THO/TREX complex, is required for micro RNA production in Arabidopsis. Plant J. 2015, 82, 1018–1029. [Google Scholar] [CrossRef] [PubMed]
  127. Chen, T.; Cui, P.; Chen, H.; Ali, S.; Zhang, S.; Xiong, L. A KH-domain RNA-binding protein interacts with FIERY2/CTD phosphatase-like 1 and splicing factors and is important for pre-mRNA splicing in Arabidopsis. PLoS Genet. 2013, 9, e1003875. [Google Scholar] [CrossRef] [PubMed]
  128. Guan, Q.; Wu, J.; Zhang, Y.; Jiang, C.; Liu, R.; Chai, C.; Zhu, J. A DEAD box RNA helicase is critical for pre-mRNA splicing, cold-responsive gene regulation, and cold tolerance in Arabidopsis. Plant Cell 2013, 25, 342–356. [Google Scholar] [CrossRef] [PubMed]
  129. Tanabe, N.; Yoshimura, K.; Kimura, A.; Yabuta, Y.; Shigeoka, S. Differential expression of alternatively spliced mRNAs of Arabidopsis SR protein homologues, atSR30 and atSR45a, in response to environmental stress. Plant Cell Physiol. 2007, 48, 1036–1049. [Google Scholar] [CrossRef] [PubMed]
  130. Tanabe, N.; Kimura, A.; Yoshimura, K.; Shigeoka, S. Plant-specific SR-related protein atSR45a interacts with spliceosomal proteins in plant nucleus. Plant Mol. Biol. 2009, 70, 241–252. [Google Scholar] [CrossRef] [PubMed]
  131. Sasaki, T.; Kanno, T.; Liang, S.-C.; Chen, P.-Y.; Liao, W.-W.; Lin, W.-D.; Matzke, A.J.M.; Matzke, M. An Rtf2 domain-containing protein influences pre-mrna splicing and is essential for embryonic development in Arabidopsis thaliana. Genetics 2015, 200, 523–535. [Google Scholar] [CrossRef] [PubMed]
  132. Ohtani, M.; Demura, T.; Sugiyama, M. Arabidopsis ROOT INITIATION DEFECTIVE1, a DEAH-box RNA helicase involved in pre-mRNA splicing, is essential for plant development. Plant Cell 2013, 25, 2056–2069. [Google Scholar] [CrossRef] [PubMed]
  133. Raczynska, K.D.; Simpson, C.G.; Ciesiolka, A.; Szewc, L.; Lewandowska, D.; McNicol, J.; Szweykowska-Kulinska, Z.; Brown, J.W.; Jarmolowski, A. Involvement of the nuclear cap-binding protein complex in alternative splicing in Arabidopsis thaliana. Nucleic Acids Res. 2010, 38, 265–278. [Google Scholar] [CrossRef] [PubMed]
  134. Laubinger, S.; Sachsenberg, T.; Zeller, G.; Busch, W.; Lohmann, J.U.; Ratsch, G.; Weigel, D. Dual roles of the nuclear cap-binding complex and SERRATE in pre-mRNA splicing and microRNA processing in Arabidopsis thaliana. Proc. Natl. Acad. Sci. USA 2008, 105, 8795–8800. [Google Scholar] [CrossRef] [PubMed]
  135. Rogers, K.; Chen, X. Biogenesis, turnover, and mode of action of plant microRNAs. Plant Cell 2013, 25, 2383–2399. [Google Scholar] [CrossRef] [PubMed]
  136. Brodersen, P.; Sakvarelidze-Achard, L.; Bruun-Rasmussen, M.; Dunoyer, P.; Yamamoto, Y.Y.; Sieburth, L.; Voinnet, O. Widespread translational inhibition by plant miRNAs and siRNAs. Science 2008, 320, 1185–1190. [Google Scholar] [CrossRef] [PubMed]
  137. Ren, G.; Xie, M.; Dou, Y.; Zhang, S.; Zhang, C.; Yu, B. Regulation of miRNA abundance by RNA binding protein TOUGH in Arabidopsis. Proc. Natl. Acad. Sci. USA 2012, 109, 12817–12821. [Google Scholar] [CrossRef] [PubMed]
  138. Kim, S.; Yang, J.-Y.; Xu, J.; Jang, I.-C.; Prigge, M.J.; Chua, N.-H. Two CAP BINDING PROTEINS CBP20 and CBP80 are involved in processing primary microRNAs. Plant Cell Physiol. 2008, 49, 1634–1644. [Google Scholar] [CrossRef] [PubMed]
  139. Gregory, B.D.; O’Malley, R.C.; Lister, R.; Urich, M.A.; Tonti-Filippini, J.; Chen, H.; Millar, A.H.; Ecker, J.R. A link between RNA metabolism and silencing affecting Arabidopsis development. Dev. Cell 2008, 14, 854–866. [Google Scholar] [CrossRef] [PubMed]
  140. Kurihara, Y.; Takashi, Y.; Watanabe, Y. The interaction between DCL1 and HYL1 is important for efficient and precise processing of pri-miRNA in plant microRNA biogenesis. RNA 2006, 12, 206–212. [Google Scholar] [CrossRef] [PubMed]
  141. Manavella, P.A.; Hagmann, J.; Ott, F.; Laubinger, S.; Franz, M.; Macek, B.; Weigel, D. Fast-forward genetics identifies plant CPL phosphatases as regulators of miRNA processing factor HYL1. Cell 2012, 151, 859–870. [Google Scholar] [CrossRef] [PubMed]
  142. Yu, B.; Bi, L.; Zheng, B.; Ji, L.; Chevalier, D.; Agarwal, M.; Ramachandran, V.; Li, W.; Lagrange, T.; Walker, J.C.; et al. The FHA domain proteins DAWDLE in Arabidopsis and SNIP1 in humans act in small RNA biogenesis. Proc. Natl. Acad. Sci. USA 2008, 105, 10073–10078. [Google Scholar] [CrossRef] [PubMed]
  143. Dziembowski, A.; Ventura, A.P.; Rutz, B.; Caspary, F.; Faux, C.; Halgand, F.; Laprévote, O.; Séraphin, B. Proteomic analysis identifies a new complex required for nuclear pre-mRNA retention and splicing. EMBO J. 2004, 23, 4847–4856. [Google Scholar] [CrossRef] [PubMed]
  144. Raczynska, K.D.; Stepien, A.; Kierzkowski, D.; Kalak, M.; Bajczyk, M.; McNicol, J.; Simpson, C.G.; Szweykowska-Kulinska, Z.; Brown, J.W.S.; Jarmolowski, A. The SERRATE protein is involved in alternative splicing in Arabidopsis thaliana. Nucleic Acids Res. 2013, 42, 1224–1244. [Google Scholar] [CrossRef] [PubMed]
  145. Köster, T.; Meyer, K.; Weinholdt, C.; Smith, L.M.; Lummer, M.; Speth, C.; Grosse, I.; Weigel, D.; Staiger, D. Regulation of pri-miRNA processing by the hnRNP-like protein AtGRP7 in Arabidopsis. Nucleic Acids Res. 2014, 42, 9925–9936. [Google Scholar] [CrossRef] [PubMed]
  146. Zhang, S.; Xie, M.; Ren, G.; Yu, B. CDC5, a DNA binding protein, positively regulates posttranscriptional processing and/or transcription of primary microRNA transcripts. Proc. Natl. Acad. Sci. USA 2013, 110, 17588–17593. [Google Scholar] [CrossRef] [PubMed]
  147. Zhang, S.; Liu, Y.; Yu, B. PRL1, an RNA-binding protein, positively regulates the accumulation of miRNAs and siRNAs in Arabidopsis. PLoS Genet. 2014, 10, e1004841. [Google Scholar] [CrossRef] [PubMed]
  148. Ben Chaabane, S.; Liu, R.; Chinnusamy, V.; Kwon, Y.; Park, J.-H.; Kim, S.Y.; Zhu, J.-K.; Yang, S.W.; Lee, B.-H. STA1, an Arabidopsis pre-mRNA processing factor 6 homolog, is a new player involved in miRNA biogenesis. Nucleic Acids Res. 2013, 41, 1984–1997. [Google Scholar] [CrossRef] [PubMed]
  149. Schmidt, F.; Marnef, A.; Cheung, M.-K.; Wilson, I.; Hancock, J.; Staiger, D.; Ladomery, M. A proteomic analysis of oligo(dT)-bound mRNP containing oxidative stress-induced Arabidopsis thaliana RNA-binding proteins ATGRP7 and ATGRP8. Mol. Biol. Rep. 2010, 37, 839–845. [Google Scholar] [CrossRef] [PubMed]
  150. Castello, A.; Fischer, B.; Eichelbaum, K.; Horos, R.; Beckmann, B.M.; Strein, C.; Davey, N.E.; Humphreys, D.T.; Preiss, T.; Steinmetz, L.M.; et al. Insights into RNA biology from an atlas of mammalian mRNA-binding proteins. Cell 2012, 149, 1393–1406. [Google Scholar] [CrossRef] [PubMed]
  151. Lorkovic, Z.J.; Barta, A. Genome analysis: RNA recognition motif (RRM) and K homology (KH) domain RNA-binding proteins from the flowering plant Arabidopsis thaliana. Nucleic Acids Res. 2002, 30, 623–635. [Google Scholar] [CrossRef] [PubMed]

Share and Cite

MDPI and ACS Style

Meyer, K.; Koester, T.; Staiger, D. Pre-mRNA Splicing in Plants: In Vivo Functions of RNA-Binding Proteins Implicated in the Splicing Process. Biomolecules 2015, 5, 1717-1740. https://doi.org/10.3390/biom5031717

AMA Style

Meyer K, Koester T, Staiger D. Pre-mRNA Splicing in Plants: In Vivo Functions of RNA-Binding Proteins Implicated in the Splicing Process. Biomolecules. 2015; 5(3):1717-1740. https://doi.org/10.3390/biom5031717

Chicago/Turabian Style

Meyer, Katja, Tino Koester, and Dorothee Staiger. 2015. "Pre-mRNA Splicing in Plants: In Vivo Functions of RNA-Binding Proteins Implicated in the Splicing Process" Biomolecules 5, no. 3: 1717-1740. https://doi.org/10.3390/biom5031717

Article Metrics

Back to TopTop