Next Article in Journal
Engineered Extracellular Vesicles: Emerging Therapeutic Strategies for Translational Applications
Next Article in Special Issue
Unregulated GmAGL82 due to Phosphorus Deficiency Positively Regulates Root Nodule Growth in Soybean
Previous Article in Journal
The Effect of Plasma-Treated Water on Microbial Growth and Biosynthesis of Gamma-Decalactones by Yarrowia lipolytica Yeast
Previous Article in Special Issue
Differences in the Occurrence of Cell Wall Components between Distinct Cell Types in Glands of Drosophyllum lusitanicum
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Alternative Splicing Variation: Accessing and Exploiting in Crop Improvement Programs

by
Sangam L. Dwivedi
1,
Luis Felipe Quiroz
2,
Anireddy S. N. Reddy
3,
Charles Spillane
2 and
Rodomiro Ortiz
4,*
1
Independent Researcher, Hyderabad 500016, India
2
Agriculture and Bioeconomy Research Centre, Ryan Institute, University of Galway, University Road, H91 REW4 Galway, Ireland
3
Department of Biology and Program in Cell and Molecular Biology, Colorado State University, Fort Collins, CO 80523, USA
4
Department of Plant Breeding, Swedish University of Agricultural Sciences, 23053 Alnarp, SE, Sweden
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(20), 15205; https://doi.org/10.3390/ijms242015205
Submission received: 2 September 2023 / Revised: 9 October 2023 / Accepted: 10 October 2023 / Published: 15 October 2023
(This article belongs to the Special Issue Modern Plant Cell Biotechnology: From Genes to Structure)

Abstract

:
Alternative splicing (AS) is a gene regulatory mechanism modulating gene expression in multiple ways. AS is prevalent in all eukaryotes including plants. AS generates two or more mRNAs from the precursor mRNA (pre-mRNA) to regulate transcriptome complexity and proteome diversity. Advances in next-generation sequencing, omics technology, bioinformatics tools, and computational methods provide new opportunities to quantify and visualize AS-based quantitative trait variation associated with plant growth, development, reproduction, and stress tolerance. Domestication, polyploidization, and environmental perturbation may evolve novel splicing variants associated with agronomically beneficial traits. To date, pre-mRNAs from many genes are spliced into multiple transcripts that cause phenotypic variation for complex traits, both in model plant Arabidopsis and field crops. Cataloguing and exploiting such variation may provide new paths to enhance climate resilience, resource-use efficiency, productivity, and nutritional quality of staple food crops. This review provides insights into AS variation alongside a gene expression analysis to select for novel phenotypic diversity for use in breeding programs. AS contributes to heterosis, enhances plant symbiosis (mycorrhiza and rhizobium), and provides a mechanistic link between the core clock genes and diverse environmental clues.

1. Alternative Splicing Isoforms as Source of Transcriptome and Proteome Diversity Contribute to Phenotypic Variation

Transcript expression and alternative splicing (AS) are two key pre-translational processes, which can generate phenotypic variation for all organisms [1,2]. While transcript expression levels are largely dependent on the interplay between promoter and enhancer activities regulating transcription rates and the rate of RNA degradation (or decay), AS can alter the transcript structure leading to modifications of the encoded protein structure to generate different protein isoforms or protein variants [3]. Alternative splicing can also alter the two-dimensional and three-dimensional structure of RNA transcripts, with possibilities for altered functionality at the non-coding transcript level.
Regardless of the close spatio-temporality shared between transcript expression and AS, it has been considered that these two processes are independent of each other [4,5]. However, the regulatory relationship connecting these processes remains unclear, and emerging evidence in both plants and animals underscores the significant influence of transcription on splicing regulation [6,7,8,9]. Because AS can produce new protein variants, it has been suggested that AS is a major source of transcriptome and proteome diversity in eucaryotes, with ultimate impacts on phenotypic variation [10]. Supporting this, a positive correlation has been indicated between the percentage of genes subject to AS and organismal complexity, measured in terms of unique number of cell types [11]. Nonetheless, the role of AS in evolutionary processes, such as speciation and adaptation, remains largely unexplored [8,9,10].
Most transcriptome research has tended to focus on the relative expression levels of mRNA transcripts, both spatially and temporally. This emphasis has arisen due to the relative ease of investigating transcript expression levels with sequencing technologies and bioinformatic tools [12,13,14,15]. Within the canon of transcriptome research, only a sub-set has a focus on AS variation. However, from a functional viewpoint, there is a paucity of investigations that ascribe clear functional effects between the genome-wide extent of AS over time and space and major phenotypic effects. Hence, it is unclear whether AS is a major source of standing genetic variation that in turn generates phenotypic variation. This lack of clarity likely arises from the experimental challenges associated with functionally characterizing the effects of alternative splice isoforms on phenotype [16,17]. Understanding the functional effects of AS isoforms remains a complex task [10,11]. An increasing number of investigations are revealing the growing significance of AS in evolutionary processes, employing advanced techniques such as whole-transcriptome mRNA sequencing [18].
In plants, the domestication process has generated multiple examples of rapid adaptation via AS [19,20,21,22]. One example of this is the EARLY MATURITY 8 (EAM8) gene in barley [19], which is an orthologue of the circadian core component EARLY FLOWERING 3 (ELF3) in Arabidopsis thaliana. It has been demonstrated that a mutated version of EAM8 (eam8.I), carrying an A to G transition in position 3257 at intron 3, which leads to an AS event with intron retention and a putative truncated protein, is responsible for the early flowering of a barley landrace from the Tibetan plateau, which is a short-season adaptation to high latitudes [19]. Another example of AS is in the domestication of sunflowers, where the domestication process (approximately 5000 years ago) was associated with a large frequency of alternative transcript isoforms generated by AS. In the AS analysis of sunflower domestication, both new combinations of ancestral spliced genes were found and also novel isoforms [20,21]. These examples suggest that AS can be an important component in evolution and domestication contributing to phenotypic variation within and between natural and domesticated populations.
In addition to phenotypic variation, phenotypic plasticity is also a key force in evolution and adaptation [23,24]. While the role of transcript expression is well understood, little is known regarding the potential of AS to generate phenotypic plasticity [10,25]. In plants, the association of environment-triggered AS with environmental stress responses suggests that AS could act as a “molecular thermometer” [26]. However, the role and underlying mechanisms by which AS can produce plastic phenotypes in novel ecological or environmental contexts are largely unexplored.

2. Bioinformatic Tools, Software, and Computational Methods to Quantify and Visualize Splicing Variants

Transcriptome-wide analyses of AS in tissues and plants subjected to various biotic and abiotic stresses as well as in different cultivars have been performed using high-throughput next-generation sequencing technologies, such as Illumina RNA-seq (RNA-seq) [27,28,29], Pacific Biosciences single-molecule real-time (SMRT) long-read Isoform sequencing (PacBio Iso-seq) [30,31], and Oxford Nanopore direct RNA sequencing (dRNA-seq), also called native RNA sequencing [29,32,33,34,35]. All these technologies involve sequencing of fragments of total cellular RNA (ribodepleted/poly(A)) or chromatin-associated RNA converted into cDNAs (Illumina short reads), full-length cDNAs (PacBio Iso-seq and Oxford dRNA-seq), or full-length RNAs (Oxford dRNA-seq). Among these, RNA-seq using the Illumina platform has been widely used as it is cheaper and yields more reads. Large-scale Illumina RNA-seq research allowed the prediction of AS events [36,37,38,39]. However, there are limitations with Illumina short reads. The transcript assemblies from short reads are often inaccurate and produce large numbers of mis-assembled transcripts and missing real transcripts [40,41]. Also, research has shown that it is difficult to reconstruct splice isoforms and quantify differential expression of isoforms using short reads [42,43], which is necessary to determine the nature of the encoded protein and in assessing a splice variant’s role [32,43]. To overcome these limitations with short reads, PacBio Iso-seq, which provides long reads, has been used for accurate identification of full-length splice variants and other post-transcriptional regulatory events, such as alternative transcription start sites and alternative polyadenylation sites [30]. As compared to Illumina RNA-seq, PacBio Iso-seq provides more comprehensive insights into different splicing events and isoform diversity and tissue-/condition-specific splicing regulations. Since 2016, Iso-seq has been used to analyze the splice isoforms in several plants, and this has provided a more detailed and in-depth view of numerous novel splice isoforms [30,31,44,45,46,47]. The most recent annotated splice isoforms in AtRTD3 (Arabidopsis thaliana Reference Transcript Database3) were assembled with PacBio Iso-seq and Illumina using RNA from many organs/tissues that were subjected to different stresses [31]. The Oxford Nanopore sequencing, which also provides long reads of cDNA or RNAs (dRNA-seq), has been increasingly used in recent years to predict splice isoforms and other post-transcriptional processes including base modifications. Other specialized high-throughput technologies, such as Ribo-seq, are used to assess the translation of splice isoforms [48,49]. Although Iso-seq and dRNA-seq approaches can generate full-length transcript sequences, the major issues are the limited depth in coverage and high error rates, which generate many mis-annotated transcripts [30,31,50,51]. Self- or hybrid-correction methods have been used to overcome the effects of sequencing errors in long reads [30,31,51]. Self-correction uses the raw signal and consensus-based calls to reduce errors, while hybrid correction uses Illumina short reads to correct errors in the long reads. Despite the shortcomings of each of these methods, global research of AS in plants revealed enormous complexity of plant transcriptomes and their regulation at the co-/post-transcriptional levels [29,31,32,52,53,54]. In plants, pre-mRNAs of about 80% of intron-containing genes undergo AS, an essential regulatory mechanism in many developmental and physiological processes that affects thousands of genes [28,31,55,56,57]. For example, research has shown that about 25% of genes that respond to cold stress are regulated by AS [58] and 20 splicing regulators of the SR family produce close to 100 distinct transcripts [59,60,61]. Intron retention is the predominant form of AS in plants, whereas exon skipping is the most prevalent AS event in animals [55,62,63]. However, Braunschweig et al. [64] have shown that IR is highly prevalent in mammals. Research has shown that IR is a regulated process that plays a role in development, stress responses, and disease [64,65,66,67].
Accurate reconstruction of transcript isoforms and quantification of the relative abundance of individual splice isoforms are necessary for a comprehensive analysis of transcriptomes and to decipher the biological functions of individual transcripts. Many computational pipelines have been developed to analyze RNA-seq data to identify AS events, estimate isoform abundance, and differential expression of splice variants across tissues/conditions. Some of the tools/pipelines used for AS analysis are shown in Table 1. These methods use different statistical models, and each has advantages and disadvantages [30,68,69,70]. Depending on the type of reads (short or long reads) and sequencing platform, different computational methods are used. These methods involve the alignment of sequence reads to the reference genome (or reference transcriptome in some cases) and allow detection of specific splicing events (exon skipping, intron retention, alternative 3′ and 5′ splice sites, etc.) and full-length splice isoforms in some cases thereby providing insights into their functional implications. There are also de novo assembly tools, but these methods are highly prone to the assembly of erroneous transcripts [71,72]. More recently, machine learning tools especially deep learning methods are being increasingly used to develop models that can accurately predict splicing/AS patterns of pre-mRNAs and gene expression from genome sequences in humans [73,74,75,76,77,78,79]. These methods are yet to be applied to splicing analysis in plants. The deep learning models determine splicing determinants directly from the nucleotide sequence [73], splice site strength in tissues [74], and impact of genetic variation on RNA splicing [74,75]. These emerging methods offer new ways to predict tissue-/condition-specific AS and the effects of genetic variation in plants on the splicing of protein-coding and protein non-coding RNAs and the biological significance of splicing changes.
Different genome browsers including Integrative Genomics Viewer (IGV—https://software.broadinstitute.org/software/igv/; accessed on 9 October 2023), Integrated Genome Browser (IGB—https://www.bioviz.org/; accessed on 9 October 2023), or UCSC Genome Browser (https://genome.ucsc.edu/; accessed on 9 October 2023) allow loading of aligned files (BAM files) to visualize sequence depth corresponding to each exon and intron and AS events. The Sashimi plot tool that is part of MISO (mixture of isoforms—https://miso.readthedocs.io/en/fastmiso/; accessed on 9 October 2023) software, which is also available on IGV (https://software.broadinstitute.org/software/igv/Sashimi; accessed on 9 October 2023) takes RNA-seq alignment files (BAM files) and gene annotations as input and provides a comprehensive view of AS patterns. The output plot shows gene structure including exons and introns, splice junctions, AS events, read coverage, and relative abundance of splice isoforms across tissues/conditions. Isoform expression levels and individual splice events, such as the percent “Splice In” of an AS event across samples, can also be visualized using heatmaps [80]. Absolute quantification of splice isoforms in tissues or in response to signals can also be performed using “Quant AS” using a combination of quantitative PCR and digital PCR.
Table 1. Some commonly used tools to analyze RNA-seq data for alternative splicing.
Table 1. Some commonly used tools to analyze RNA-seq data for alternative splicing.
Tool/Pipeline *Sequencing PlatformSplicing AnalysisURL AddressReference
ASpliIllumina short readsAnnotated and novel AS eventshttps://bioconductor.org/packages/release/bioc/html/ASpli.html; accessed on 9 October 2023[69]
rMATSIllumina short reads; Requires replicatesDifferential AS eventshttps://rnaseq-mats.sourceforge.net/; accessed on 9 October 2023[81]
DEXSeqIllumina short readsDifferential exon usagehttps://bioconductor.org/packages/release/bioc/html/DEXSeq.html; accessed on 9 October 2023[82]
MAJIQIllumina short readsKnown and novel local splice variationshttps://majiq.biociphers.org/; accessed on 9 October 2023[83]
3D RNA-seqIllumina short readsGUI-based pipeline to analyze differential AS and transcript isoformshttps://3drnaseq.hutton.ac.uk/app_direct/3DRNAseq/; accessed on 9 October 2023[31,70]
TAPISPacBio Iso-seqAnalysis of AS events and transcript isoformshttps://bitbucket.org/comp_bio/tapis/src/master/; accessed on 9 October 2023[30]
SUPPA2Illumina short readsDifferential splicing across multiple conditionshttps://github.com/comprna/SUPPA; accessed on 9 October 2023[84]
TAMAPacBio Iso-seqTranscript isoformshttps://github.com/GenomeRIK/tama; accessed on 9 October 2023[31,51,85]
MISOIllumina short readsDifferentially spliced exonshttps://miso.readthedocs.io/en/fastmiso/; accessed on 9 October 2023[86]
SpliceGrapherIllumina short readsDetects patterns of AShttps://splicegrapher.sourceforge.net/; accessed on 9 October 2023[87]
iDiffIRIllumina short readsDifferential intron retentionhttps://bitbucket.org/comp_bio/idiffir/src/master/; accessed on 9 October 2023[88]
DARTSIllumina short reads;
Uses a deep learning model and incorporates the expression of RBP.
Differential AShttps://github.com/Xinglab/DARTS; accessed on 9 October 2023 [79]
SpliceAIIllumina short reads;
Uses a deep learning model
AS events and splice isoformshttps://github.com/Illumina/SpliceAI; accessed on 9 October 2023[73]
PangolinA deep learning model that predicts RNA splicing from DNA sequencePredicts effects of genetic variants on splicing; tissue-specific splicinghttps://github.com/tkzeng/Pangolin; accessed on 9 October 2023[74]
SpliceVault
Web portal
Uses RNA-seq dataGenetic variant’s effect on splicinghttps://kidsneuro.shinyapps.io/splicevault/; accessed on 9 October 2023[76]
* This list is not comprehensive. Older versions of some of the tools are not listed. Also, some tools for which weblinks are inactive are not included. DARTS, deep-learning augmented RNA-seq analysis of transcript splicing; MAJIQ: Modeling Alternative Junction Inclusion Quantification; MISO, Mixture of Isoforms; RBP, RNA-binding protein; rMATS, Replicate Multivariate Analysis of Transcript Splicing; SUPPA2, Sequencing Unified Pipeline for Proximal Alternative splicing analysis2; TAMA, Transcriptome Annotation by Modular Algorithms.

3. Mining Gene Pools for Splicing Isoforms and Diversifying Gene Functions to Obtain Novel Phenotypic Diversity

Alternative splicing allows a gene to encode for various proteins because its exons are put together differently, thus resulting in related but distinct mRNA transcripts. It has been demonstrated that thale cress (Arabidopsis thaliana) uses AS disproportionally as a stress response [28,89]. There are other plants showing a cell memory to environmental stress, such as heat [90], which leads to a response to an increase in temperature. Moreover, a synthesized Brassica hexaploid had significant AS events [91], thus diversifying its gene expression patterns that could improve its adaptability. Furthermore, Zhang et al. [92] indicated that many genes contributing to quantitative traits are likely to be spliced into multiple transcripts causing their variation.
The availability of both genome and transcript sequences in plants enables a thorough analysis of AS in various species, including crops [93]. Multi-variate analysis of transcript splicing (MATS) and replicate MATS (rMATS) are robust and flexible statistical software that detect differential AS between two RNA-Seq samples [94] or replicate RNA-Seq data [81], respectively. The synthetic programming of AS patterns, however, remains underexploited for improving crops [95]. Hence, Pramanik et al. [96] suggested CRISPR/Cas9-mediated engineering for modifying AS with the aim of (de)regulating plant development.
Genome-wide mapping led to the identification of thousands of AS mRNAs isoforms in thale cress [36]. Most of the AS transcripts related to isoforms with premature termination codons, which could shift under abiotic stress. Li et al. [97] did a search of AS affecting reproductive development of young panicles as well as both unfertilized and fertilized florets in rice with the aid of direct RNA sequencing, small RNA sequencing, and degradome sequencing. They found 35,317 AS events, of which in excess of two thirds were novel, and concluded that AS was significantly related to development stages and to complex gene regulation in rice. An RNA-seq survey was able to define AS patterns and to determine that 59.3% of expressed multi-exon genes underwent AS in seedlings, flowers, and young developing fruits of tomato [98]. The use of a single molecule long-read sequencing (Iso-Seq) led to an integrated transcriptome data analysis that facilitated investigating AS in polyploid cotton [99]. This Iso-Seq data analysis was able to identify 15,102 fiber-specific AS events and notice that about 51.4% of homeologous genes produce divergent isoforms in each cotton sub-genome.

4. Molecular Mechanisms Regulating Stress-Dependent Gene Splice Variants

Numerous RNA-seq investigations with plants subjected to various biotic and abiotic stresses have revealed that AS of pre-mRNA is widespread. Furthermore, stresses and developmental cues have a profound impact on the splicing patterns of many genes [28,29,31,44,45,46,47,59,100,101,102,103,104,105,106,107]. Despite the prevalence of AS and its role in stress responses, the regulatory mechanisms of splicing and functions of most splice isoforms are not well understood in plants. Decoding the splicing code in plants would require a comprehensive understanding of the rules that dictate splice site choice and the identification of specific mRNA targets of splicing regulators. A variety of factors including splice site strength and the presence of exonic and intronic splicing enhancers and suppressors affect splice site choice, and RNA structural features [108,109,110] also contribute to AS. Limited research with plants has shown that sequence elements are one of the important determinants of splice site choice [111,112,113,114]. Interestingly, the alternatively spliced genes are over-represented in functional categories related to splicing regulators and stress responses [36,103,115,116]. RNA-binding proteins, such as serine-/arginine-rich (SR) and heterogeneous nuclear ribonucleoproteins (hnRNPs), are some of the key regulators of splicing. Alternative splicing of plant pre-mRNAs encoding SR proteins is dramatically altered in response to various stresses [56,59,117,118,119,120,121,122]. The changes in the levels of these splicing regulators in response to stresses may change the splicing of other pre-mRNAs due to auto- and cross-regulation of splicing [111,123,124,125,126,127]. These investigations suggest that altered ratios of splice variants of splicing regulators in response to stresses may have a role in fine-tuning gene expression at the mRNA and protein levels and the adaptation of plants to stresses [28,128]. Also, many stress-responsive genes are associated with significant splicing quantitative trait loci (sQTLs) in Arabidopsis thaliana ecotypes, suggesting a role of AS in plant stress responses [129].
There are several hundred RNA-binding proteins (RBPs) in any given plant species, and the precise roles of most of these proteins in co-/post-transcriptional processes are unknown [130]. Many approaches to identifying the roles of RBPs in splice site choice are available, and a comprehensive review of these methods was recently published [29,131], which is why they are not covered in any detail here. In animals, in vitro splicing assays have greatly contributed to our understanding of the roles of spliceosomal and other splicing regulatory proteins in splicing and elucidating steps in spliceosome assembly and spliceosome composition. However, the lack of a robust plant-derived in vitro splicing system in plants has been a major limitation [132]. Hence, other biochemical, cell biological, genetic, and genomic approaches are used to understand splicing regulation in plants [28,29,103,109,133]. The application of new methodologies, such as the identification of targets of RNA-binding proteins using TRIBE (targets of RNA-binding proteins identified by editing) [133,134] and targeted isoform degradation with CRISPR/Cas13 variants [135], may provide insights into targets of hundreds of uncharacterized RNA-binding proteins and elucidation isoform factions. With TRIBE, the targets of an RNA-binding protein are edited irreversibly by de-aminating adenosine to inosine, which is then recognized as guanosine in cDNAs [136] or modified inosines can be identified directly with Nanopore native RNA sequencing [137]. RNA from the RBP-ADAR-expressing plants is sequenced to identify the RNA targets of the RBP by edited events. Expressing specific isoforms in the mutant background or degrading specific isoforms using CRISPR/Cas13 variants (e.g., Cas13d and Cas13x) that specifically bind RNA [135] open a novel and efficient way to study the functions of splice isoforms.
Emerging evidence suggests that stresses/external cues converge on splicing regulators via different signaling pathways. For example, two proteins (Highly ABA-Induced 1 (HAI1), a protein phosphatase 2C and its interacting RNA-binding protein, HIN1(HAI interactor 1, HIN1), an RNA binding protein) involved in drought acclimation interact with the SR family of splicing factors and regulate splicing [107]. Phytochromes, key light receptors and regulators of many aspects of plant growth and development, interact directly with several splicing regulatory proteins and modulate AS of many pre-mRNAs [103,138,139,140]. The light- and drought-regulated alternatively spliced transcripts contain GAA repeats [107,138] that are known to bind splicing regulators (e.g., SCL33, SCL30, and SR45), suggesting that stress-signaling pathways could converge on these splicing regulators [111,113,141]. A mutant of SR45, which encodes a splicing factor, showed altered responses to abiotic and biotic stresses [142,143]. Like abiotic stresses, biotic stresses also change the splicing patterns of many genes. Recent research shows that pathogens effectors modulate host pre-mRNA splicing by binding to splicing regulators, such as serine-/lysine-/arginine-rich proteins, U1-70K, SR30, SR45, and GRP7, and suppress plant immunity [80,144,145,146], suggesting that pathogens have evolved effectors that target host splicing components and subvert plant immunity. It has been shown that many splicing regulators and spliceosomal proteins form speckles (also called biological condensates or membraneless organelles) and stresses alter the dynamics of proteins in these structures and also the size/shape of these structures [133,147,148,149,150,151,152,153,154,155], suggesting that external signals through the re-organization of speckles and their constituent proteins affect pre-mRNA splicing. However, the mechanisms of stress-induced re-organization of speckles in plants are yet to be understood. Also, the phosphorylation status of many spliceosomal proteins and regulatory splicing factors is known to play an important role in pre-mRNA splicing [156] and stresses may alter phosphorylation status and function of splicing regulators.
Until recently, the splicing code has been thought to consist primarily of exonic and intronic sequence motifs that recruit RBPs that either enhance or suppress the selection of nearby splice sites [55,157]. However, in recent years, most pre-mRNA splicing was found to occur co-transcriptionally in both plants and animals [53,54,158], suggesting that chromatin state may affect splice site choice and AS. Emerging research provides evidence in support of multiple regulatory mechanisms at the chromatin level (open vs. closed chromatin, epigenetic modifications including histone modifications and DNA methylation) and the speed of transcription as key regulators that determine the outcome of AS in plants [28,159,160]. A rice mutant (OsMet1-2) with impaired DNA methylation altered all types of AS events [159]. Also, a mutant with reduced histone H3 lysine 36-specific methyltransferase in rice showed altered intron retention events [160]. In Arabidopsis and rice, open chromatin was found to be associated with intron retention [161]. Higher speeds of transcription in open chromatin regions provide less time for the spliceosomal machinery to recognize and excise introns co-transcriptionally [162,163]. Alternatively, accessible chromatin regions could be the sites of binding for TFs or other regulatory proteins that recruit splicing factors directly or indirectly through chromatin modifications to affect the outcome of splicing [64,164]. The rate of Pol II elongation during transcription was shown to be involved in light-regulated AS of splicing factors [165,166]. A point mutation in Pol II with increased elongation speed increased splicing, indicating a role for Pol II speed in splicing regulation [166]. Furthermore, an increase in two epigenetic changes (H3K4me3 and H3K9ac) increased the rate of transcription elongation and lowered co-transcriptional splicing efficiency [53]. A double mutant, rz-1b rz-1c, of hnRNP-like proteins showed impaired splicing of nascent RNAs, suggesting that these proteins promote splicing at the chromatin level [53]. The direct association of RZ-1C with nascent RNAs further supports its role in co-transcriptional splicing [53]. It has been shown that a shift in temperature alters H3K36me3 methylation and AS [167] and a low temperature changes RNA Pol II elongation kinetics and reduces co-transcriptional splicing [168]. The involvement of chromatin modifiers and a mediator complex in splicing regulation was also reported, and some of these proteins interact with spliceosomal proteins [169,170]. A phosphoprotein phosphatase required for Pol II occupancy was found to promote intron excision [171]. Collectively, these investigations indicate that the epigenetic state of chromatin and the dynamics of transcription modulate AS in plants.
One of the key adaptive changes in response to stresses in plants is the post-transcriptional reprogramming of gene expression [172,173]. The resulting transcript isoforms fine-tune gene expression in profound ways to cope with stresses [90,128,174,175,176]. The research discussed above indicates that stresses/external cues through some yet-to-be-elucidated signaling pathways converge on splicing regulatory proteins and chromatin architecture to modulate AS. An in-depth understanding of splicing code in plants and the roles of splice variants will have applications in fine-tuning gene regulation and developing stress-resilient crops as stresses and developmental cues dramatically alter the levels of splice variants that encode proteins involved in stress responses and plant growth and development [28,29,103].

5. Global Expression of AS Isoforms in Model Plant Arabidopsis and among Diverse Crops

5.1. Arabidopsis

Arabidopsis thaliana, as the main model in plants, has been the subject of intensive investigations to better understand the landscape and functional effects of alternative splice isoforms. Several investigations have demonstrated the widespread extent of AS in Arabidopsis, with initial estimations placing the occurrence of AS events at 11.6% across its genome [116]. However, in recent years and due to the advances in high-throughput sequencing technologies, the estimated rate of intron-containing genes subject to AS has risen to 61–70% in A. thaliana [39,89]. Around 40% of the AS events detected represent intron retention, as the predominant type of AS in Arabidopsis [39,89]. Comparisons between AS events involving intron retention vs. non-retention, as well as with constitutive introns, has revealed that the size of the retained introns was significatively smaller than the non-retained ones, with no differences in constitutive introns [39]. Interestingly, it was found that from the total number of AS events that affected protein-encoding genes, 30.3% have little or no effect on the coding sequences (only one amino acid difference or AS in the 5′ and 3′ regions), while the remaining 69.7% of AS events significatively affected the encoded proteins [31]. An additional layer of complexity in the Arabidopsis genome has emerged as cryptic introns that are characterized by the presence of splice sites within annotated coding exons. Approximately 1300 cryptic introns (around 14.1% of all retained introns) have been detected, with nearly half of them undergoing in-frame splicing, hence possessing the ability to excise amino acid stretches from the full-length protein, generating novel protein isoforms [39]. Furthermore, it has been suggested that in Arabidopsis, AS may modulate upstream ORF production in response to environmental stresses by extending 5′ UTR sequences [89]. Interestingly, not only has it been proposed that transcript expression and AS are independent mechanisms in Arabidopsis, but also that transcript expression and AS act in an exclusive manner, in which the genetic structure of transcript expression-regulated and AS-regulated genes exhibit differential genomic architecture [89]. This may suggest that transcript expression and AS are non-redundant and non-overlapping, yet they are complementary mechanisms to generate phenotypic effects.
In a comparison of the global landscape of AS between A. thaliana and animals, striking divergences in their regulatory roles have been identified. While animals have harnessed AS as a powerful source of transcriptomic and proteomic diversity, primarily facilitating cellular and tissue specialization, plants have shaped AS into a regulatory mechanism to respond to the ever-changing demands of their sessile lifestyle [89]. For plants, fast and efficient adaptation to shifting environmental conditions and stressors necessitates an AS machinery that can orchestrate in situ responses [89,90]. The divergent evolutionary trajectories of these lineages have led to unique molecular regulatory mechanisms, allowing them to exploit the diverse capabilities of AS to meet their specific developmental and physiological requirements [90].

5.2. Grain and Fiber Crops

AS is involved in plant response to abiotic stresses and in various aspects of plant growth, development, and reproduction. Genome-wide association analysis (GWAS) is a powerful approach to identify genomic regions and genes associated with complex traits. GWAS has also been found useful in providing genome-wide summary statistics of AS variants and in genome-wide association analysis of AS variants associated with complex traits. Genomic regions associated with gene expression are commonly referred to as quantitative trait loci (QTL), whereas those regulated by AS variants are referred to as splicing quantitative trait loci (sQTL). An analysis of population-level transcriptome data and GWAS of splicing QTL in developing maize kernels from 368 maize inbred lines unfolded 19,554 unique sQTL for 6570 genes, with distinct protein functions. Natural variation in AS and overall mRNA levels were independently regulated with different cis-sequences used preferentially. Two hundred and fourteen putative trans-acting splicing regulators, including ZmGRP1, controlled the largest trans-cluster, and the knockout of ZmGRP1 modified splicing of several downstream genes. There were 739 sQTL that colocalized with known trait QTL, indicating the significance of AS in diversifying gene function to regulate phenotypic variation [22]. An earlier study involving teosinte (wild ancestor) and maize transcriptomes reported 13,593 highly conserved genes, including 12,030 multi-exonic genes. The two species were no different in number of AS events. Over 60% of the AS in both species were of intron retention (IR) and alternative acceptor (AA) types. The average number of unique AS events per alternatively spliced gene was higher in maize (4.12) than in teosinte (2.26). Ninety-four genes in maize generated 98 IR types with transposable element (TE) sequences, far more than 9 IR with TEs in teosinte. TE insertion is probably an important mechanism for IR-type AS in maize. The AS levels of 3864 genes were significantly different between maize and teosinte, of which 151 AS level-altered genes involved in transcriptional regulation and in stress responses were located in the regions that were targets of selection during maize genetic improvement [177].
GWAS unfolded 35,317 AS events at the early reproductive stage in rice, of which ~67% were of novel AS isoforms, and the intron retention (IR) sub-type was the most abundant [97]. Over 11,000 novel splice isoforms, alternative polyadenylation (APA) of ~11,000 expressed genes and more than 2100 novel genes were reported in sorghum [30,178], whereas 15,102 fiber-specific AS were reported in cotton [99]. To date, a large number of AS events associated with various development stages and molecular functions have been reported in soybean—294,164 AS events across multiple experiments [177]; 1278 AS events associated with nitrate stress in root hairs [179]; and 154,469 AS events in 23,764 genes across development stages [180]. The intron retention form of AS events was predominant in most of the research reported here followed by alternative acceptor sites, alternative donor sites, and exon skipping.
Polyploidization, an evolutionary force, promotes diversity and evolution of new species. A GWAS analysis of synthesized hexaploidy Brassica (2n = 54) and its parents unfolded 7913, 14,447, and 13,205 AS genes that produced 27,540, 70,179, and 60,804 AS isoforms in Brassica rapa (turnip, 2n = 20), B. carinata (Ethiopian mustard, amphidiploid, 2n = 34), and Brassica hexaploidy, respectively. Hexaploid Brassica has 920 new genes. The number of differentially spliced genes between hexaploidy Brassica and its parents was 56. Hexaploid Brassica and its parents had diverse AS patterns of genes, including the gain and loss of AS isoforms [91].
Maize was domesticated in the tropics but is widely grown in temperate environments. How did variation in gene expression, as measured by changes in transcriptomes, enable maize to adapt in temperate environments? A genome-wide association study involving eGWAS and sGWAS based on 572 unique RNA-seq datasets from the roots of 340 maize lines identified 19,602 eQTL associated with the expression of 11,444 genes and 49,897 sQTL for 7614 genes. Genes containing both cis-eQTL and cis-sQTL in LD disproportionately encoded TFs associated with one or more stresses. Further, gene expression data listed transcriptional regulatory networks associated with gene expression, cell propagation, and phase transition powered tropical maize adaption in temperate environments [181].

5.3. Vegetable Crops

A systematic GWAS analysis of AS events in potato plants revealed 226,769 AS events, of which 49% were classified as basic and 51% as complex events, generated from 24,650 genes. The basic events include 19.2% alternative acceptor sites (AAS), 12.1% exon skipping (ES), 8.2% alternative donor site (ADS), and 9.5% intron retention (IR) types. A comparative analysis detected 2929 AS genes conserved among maize, potato, soybean, and tomato plants [182]. The AS landscape in tomatoes consists of 369,911 AS events, identified from 34,419 genomic loci involving 161,913 transcripts. IR-type AS events were predominant (18.9%) followed by AAS (12.9%), ADS (7.3%), and ES (6.0%) within the basic events. The complex AS accounted for 54.9% of total AS events. Sixty-five percent of 35,768 annotated protein-coding genes had pre-mRNAs generating AS isoform transcripts [183]. Thus, it is a useful genomic resource for functional characterization of genes in potato and tomato biology.

6. Genomic Regions Regulating Splicing of Quantitative Trait Loci (sQTLs)

6.1. Novel Splicing Variants Impacting Flowering and Plant Architecture

Evolutionary transition from wild species to crops or polyploidization may evolve novel splicing variants and may contribute to adaptation and population divergence. Hexaploid wheat is an ideal model for studying variation in the AS landscape in response to domestication and polyploidization. Transcriptome sequencing of roots and leaves of wheat species differing in ploidy levels unfolded ~22% of the genes exhibiting AS events. However, AS events decreased after domestication and polyploidization. The decrease in AS events is consistent with the functional sharing model that proposes complementarity between the two (AS-duplicated genes) in regulating transcriptome plasticity in polyploid crops. Sub-genomes exhibited biased AS response to polyploidization, with ~87% of homeologs showing AS partitioning in hexaploid wheat, and substitution of the D-sub-genome modified ~43% of AS patterns of the A- and B-sub-genomes. Thus, AS variation occurs extensively after polyploidization and domestication in wheat [184].
The regulation of AS in polyploid wheat (tetraploid and hexaploid) and its ancestral diploid grass species unfolded diversity in AS events not only between the endosperm and pericarp and embryo over-development, but also between sub-genomes. The triads of homoeologous chromosomes revealed evolutionary divergence between gene- and transcript-level regulation of embryogenesis. The novel transcript isoforms in young genes were at a more rapid rate than ancient genes, providing a greater understanding of the evolution of regulatory features of AS during embryogenesis and grain development in wheat [185].
The substantial splicing divergence and predominance of divergent splicing transcripts for seed traits between wild and cultivated sunflowers suggest that domestication and selection for seed development affected the evolution of splicing variants in sunflowers. While Helianthus annuus (wild species) contributed to most of the differential splicing patterns, other Helianthus species also contributed to domestication-associated splicing patterns in sunflowers [21]. Significantly more AS isoforms were reported among wild accessions than domesticated sorghum accessions [186].
A multi-silique trait (zws-ms) was discovered in the rapeseed. Such a line forms three independent siliques instead of a commonly observed single silique, with temperature being the most critical factor likely to switch on/off the formation of multi-silique [187]. The pattern of transcriptome variation between zws-ms and its NIL (zws-217), which produces normal siliques (i.e., single silique) grown under optimal conditions, unfolded in a colder environment 11 differentially expressed alternative splicing (DAS) genes, of which 4 were up-regulated and 7 were downregulated in a multi-silique line. Five such genes were associated with the multi-silique trait [188], and two thermos-morphogenesis genes switched off genes controlling the multi-silique trait in cold environments [189].
Do AS events affect flowering and plant architecture in wheat? The two splicing variants of TaNAK1 (TaNAK1.1 and TaNAK1.2) show distinct expression patterns during wheat growth and development, while such an effect has not been observed for TaNAK1.3. TaNAK1 is mainly expressed in developing grains, while TaNAK1.2 is expressed in leaf and flag leaf. Transgenic Arabidopsis over-expressing TaNAK1.1 and TaNAK1.2 showed opposite effects (i.e., TaNAK1.2 positively regulates transition from vegetative to reproductive growth, plant height, branching, seed size, and seed yield, while TaNAK1 negatively regulates these traits) on flowering and plant architecture, resulting in varying seed yield [190].

6.2. Seed Yield and Quality

Spikelet architecture, seed size, and weight are the major determinants of yield in cereal crops. Five splicing variants in TaGS3 and TaGS3.1 to TaGS3.5 showed expression divergence during polyploidization and differential functions to regulate seed size and weight in wheat. TaGS3.1 over-expression significantly reduces seed weight and length by 5.89% and 5.04%, respectively. TaGS3.2-3.4 over-expression relative to wild type (WT) had no significant effect on grain size. TaGS3.5 over-expression significantly increases seed weight and length by 5.7% and 4.3%, respectively [191].
Multiple signaling pathways at transcriptional and post-translational levels control GS3, seed size QTL in rice. The dominant AS isoforms of GS3, GS3.1, and GS3.2 account for about 50% and 40% of total transcripts. GS3.1 over-expression decreases seed size, whereas GS3.2 has no significant effect on seed size. GS3.2 interacts with RGB1 to disrupt GS3.1 activity, thereby implying AS of GS3 decreases the amount of GS3.1 and GS3.2 disrupts the GS3.1 signaling to inhibit the negative effects of GS3.1 to fine-tune grain size in rice [192].
Poor seed filling of inferior spikelets is one of the major limitations in raising rice production. Post-anthesis moderate soil drying promotes starch synthesis and seed filling in inferior spikelets. An assessment of AS events at the grain-filling stage in inferior spikelets under control (irrigated, C) and moderate drought (MD) stress unfolded 16,089 AS events, of which 1840 involving 1392 genes occurred differentially between C and MD treatments, many of which function on spliceosome, starch, and sucrose metabolism, providing new insights into the role of AS to promote seed filling in inferior spikelets under MD in rice [193].
Maintaining yield and quality under low nitrogen conditions is a significant production constraint in cereal. OsGS1;1 enhances nitrogen use efficiency (NUE). SNP polymorphisms in the OsGS1;1 region led to the discovery of AS that generated two functional transcripts, OsGS1;1a and OsGS1;1b. Germplasm containing the OsGS1;1b haplotype had improved NUE, positively affected seed development, and reduced amylose content, providing a new avenue to raise yield and nutritional quality of rice under low N conditions [194]. OsLG3b regulates grain length in tropical Japonica rice. OsLG3b expression is higher during the panicle and seed development stages. SNP polymorphism in the OsLG3b region discovered AS that was found to be associated with grain length and was extensively used in breeding to enhance the productivity of tropical japonica rice [195].
Yellow seed coat color in rapeseed is associated with higher oil content and a higher quality of meal. A comparison of yellow- and black-seeded rapeseed lines at five developmental stages revealed highly similar AS events in the different samples, with the intron retention type being the predominant form of AS patterns. The early and middle stages of seed development were most-affected by AS variants. Twenty-three co-expression modules composed of differentially spliced genes were detected, of which the function of two modules was highly associated with seed coat color. Both the modules in-housed differentially alternative splicing (DAS) candidate genes related to the flavonoid pathway (TT8, TT5, TT12, AHA10, CHI, BAN, and DFR), which could be exploited to develop stable, yellow-seeded rapeseed [196]. A splicing error in the phytic acid synthase gene inositol-1,3,4 triphosphate 5/6-kinase 3 (GmITPK3) caused a low seed phytate phenotype in soybean [197]. Low seed phytate crops improve the bioavailability of micronutrients. Food rich in flavonoids promotes human health and minimizes the risk of old age diseases [198]. BnaPAP2.A7 regulates anthocyanin biosynthesis, with AS (three splicing isoforms) as the main mechanism for the modulation of anthocyanin biosynthesis in rapeseed leaves [199].

6.3. Mineral Nutrient Homeostasis

Very limited knowledge exists about the role of AS in maintaining mineral nutrient homeostasis in plants. Using root transcriptome sequencing of rice grown in the presence or absence of minerals (Fe, Zn, Cu, Mn, and P), Dong et al. [106] noted 13,291 alternatively spliced genes, with a small overlap between differentially expressed genes (DEGs) and DAS genes. Nutrient-specific AS genes represented ~53.3% of multi-exon genes in the rice genome. A group of splicing factors known as serine-/arginine-rich (SR) proteins regulate AS mechanisms. The characterization of mutants in gene-encoding Ser/Arg (SR) proteins in rice unfolded several SR proteins as critical regulators of Zn, Mn, and P nutrition, with highly specific AS targets for each nutrient. For example, three SR protein-encoding genes regulate P uptake and re-mobilization between the leaves and shoots of rice [106]. Clearly, this is an under-explored area of research and must be further investigated to unfold the molecular basis of mineral nutrient homeostasis (DEGs, DAS genes, and interaction between DEG and DAS) in plants.

6.4. Abiotic Stress Adaptation

Alternative splicing variants increase proteome diversity. Heat shock transcription factor (Hsf) under stress may form different transcripts by AS. A novel splice variant TaHsfA2-7-AS, induced by high temperature, regulates thermotolerance in wheat, and its over-expression in Arabidopsis enhances tolerance to heat stress [200]. Plant serine-/arginine-rich (SR) proteins contribute to abiotic stress adaptation by regulating AS of key genes. Over-expression of BrSR45a in Arabidopsis regulates the drought stress response via the AS of target genes in a concentration-dependent manner [201]. Over-expression of AS-related protein from cassava, MeSCL30 in Arabidopsis, enhances tolerance to drought via maintaining ROS homeostasis and increasing the expression of drought-responsive genes [202].
The interplay of AS under stress and across development stages (ear, tassel, and leaf) of a public inbred line B73 under well-watered and drought stress conditions unlocked over 48,000 novel AS isoforms, often with stage- or condition-specific expression. Stress induces large developmental splicing changes in leaves and ears but only a few in tassels. Most of the developmental stage-specific splicing changes affected by stress are tissue-dependent, whereas stage-independent changes frequently overlap between leaves and ears. This suggests that AS is strongly associated with tissue type, developmental stage, and stress condition [203]. Low-temperature stress reduces seed germination, which in turn results in low plant populations and reduces grain yield in maize. A genome-wide analysis of AS during cold stress unlocked approximately 2.05–2.09 AS events per gene on each chromosome, of which seven exclusively expressed AS in cold-tolerant maize inbred lines. Functional validation of the AS gene through mutation reveals that ZmWRKY48 is associated with low-temperature resistance in maize [204].
SR-rich splicing factors play a key role in pre-mRNA splicing to regulate plant growth and development under stress conditions. Butt et al. [205] investigated the role of the plant-specific SR protein RS33 in regulating pre-mRNA splicing and abiotic stress responses in rice. Loss-of-function mutant rs33 showed increased sensitivity to salt and low-temperature stresses. They identified multiple splice isoforms of stress-responsive genes whose AS is regulated by RS33, and the expression of RS33-regulated genes was more under cold than salt stress, indicating plant-specific splicing factor RS33 is crucial in response to abiotic stresses [205]. Multiple abiotic stresses also triggered extensive but different expressional regulations on sweet potato SR genes. Heat stress caused substantial disturbances in both gene transcription and pre-mRNA AS. Tissue- and species-specific AS regulations in response to stresses were noted in sweet potato, unlike Arabidopsis and rice [206].
Multiple abiotic stress induces several DAS genes. Three hundred and fifty-seven DAS genes and their splicing isoforms of candidate genes (RBP45C, LHY, MYB59, SCL30A, RS40, MAJ23.10, and DWF4) were induced in rapeseed [207]. In soybean roots, between 385 and 1429 AS events were differentially spliced under varying water deficits and recovery after severe drought stress [208]. In rice, 764 genotype-specific splicing (GSS) events were identified in salt stress conditions, of which six events in five genes were significantly associated with the shoot’s Na+ content. OsNUC1 and OsRAD23 emerged as candidate genes with splice variants exhibiting significant divergence between the variants for shoot growth under salt stress conditions [209]. In response to drought, heat, and combined (drought and heat) stress in wheat, 200, 3576, and 4056 genes exhibited significant AS pattern changes. The combined stress induced specific AS compared to individual stress, while the B sub-genome exhibited more AS events than on the A and D genomes [210]. The splicing isoforms of candidate genes could be a valuable resource for enhancing abiotic stress adaptation in plants.
The productivity and quality of vegetable crops is impaired in sub-optimal environments. Identifying and exploiting variations in AS events is a powerful approach to cope with these stresses by producing multiple mRNA splicing variants with different sub-cellular localizations, translational efficiency, and coding sequences in vegetable crops [211]. A comparative assessment of AS in tomato plants subjected to varying levels of moisture stress unfolded 464, 512, and 506 AS events under optimal (irrigated), mild, and severe drought stress conditions, respectively. The stage-dependent changes in AS genes may participate in plant tolerance to drought stress [212]. An assessment of differentially expressed genes (DEGs) and differentially alternatively splicing (DAS) events in stress-tolerant (water deficit, low nitrate, or a combination of both) (T270) and intolerant (T250) tomato accessions reveals that transcriptome changes caused by combined stress yielded a low number of DEGs in the tolerant compared to intolerant genotypes. DAS events in combined stress greatly affect the splicing landscape in both genotypes; i.e., stress- and growth-related genes as well as transcription and splicing factors are differentially spliced in the roots and leaves of the two genotypes. This clearly shows that transcriptional and post-transcriptional mechanisms regulate tomato adaptation to growth under drought and low N stress [213].

7. Alternatively Spliced Variants Contribute to Hybrid Vigor

Heterosis, or the superior performance of F1 hybrids vis-à-vis their parental lines, has been widely applied for breeding output to raise the productivity of crops, especially in cross-pollinated grain crops (e.g., maize, pearl millet, and pigeonpea) and in a few self-pollinated crops (e.g., rice, tomato and other vegetables). Differential gene and protein expression between hybrids and their parents regulate hybrid vigor. A genome-wide assessment of AS variants between hybrids and their parents may provide additional means to exploit heterosis in plants. Profiling AS landscape data from immature ears of the maize hybrid ZD808 and its parents (NG5 and CL11) unfolded substantial differential AS events in the hybrid vis-à-vis its parents, which were classified into parental-dominant and novel DAS patterns. NG5-dominant events prevalent in the hybrid accounted for 42% of DAS events and were mainly involved in regulating gene expression associated with carbon/nitrogen metabolism and cell division processes. Cis-regulation was the predominant contributor to AS variation and was involved in biological processes associated with immature ear development in maize [214].
In sorghum, the developing embryo and endosperm show significant and multi-faceted differences in gene expression and AS that may potentially correlate with hybrid vigor. An analysis of genome-wide gene expression between developing embryo and endosperm as well as between F1 hybrids and their parental lines in sorghum uncovered substantial differences in both gene expression and AS events between embryo and endosperm, which were consistent with their biological roles in the two tissues. The hybrids relative to their parents showed substantial and multi-faceted differences in gene expression and AS events, which were distinct and tissue-specific, and may provide transcriptome resources to further elucidate seed yield heterosis in sorghum [215].
A recent study on a sunflower hybrid under control (irrigated) and drought stress conditions revealed that the ‘absence’ alleles at presence/absence variants (PAVs) were disproportionately associated with reduced values of heterosis-related traits, but not those of non-heterotic traits. The expression of gene PAVs differentiating the parental lines was complemented in hybrids, thereby supporting the dominance model of hybrid vigor and yield stability across environments. The consistent expression of many of the PAVs in control and drought stress conditions possibly contributed to heterosis under drought stress. A further comparison of DEGs between hybrid and parental lines revealed that parents responded similarly to drought stress by up-regulating stress response and down-regulating metabolic process genes, while these responses were further strengthened in the hybrid. An inverse relationship between AS changes and expression changes in DEGs implies that AS acts to reinforce expression responses [216].

8. Establishing a Platform for Cataloguing, Curating, and Retrieving Alternative Splicing Isoforms and Gene Expression Quantification Database across Tissues, Development, and Stress Conditions

One of the major challenges in researching AS and transcript expression is the fragmented nature of data; AS isoforms and transcript expression data are scattered across many investigations and datasets, often lacking standardized annotations and metadata [217,217]. Such fragmentation hinders efficient data retrieval, comparison, and interpretation. Furthermore, inconsistencies in data formats and quality pose additional obstacles for researchers [217,218]. To address these challenges, the development of unified platforms for cataloguing, curating, and retrieving AS isoforms and GE quantification data is important. In this context, multiple attempts to generate a unified platform for GE and AS have been published to the date [217,218,219,220,221,222]. The aim of such platforms is to provide researchers with a centralized resource for accessing comprehensive and high-quality data across different biological contexts and investigations.
The accessibility of such platforms requires a focus on user-friendly interfaces, powerful search functionalities, and intuitive data visualization tools. This can allow researchers to better explore and analyze complex AS patterns and transcript expression dynamics. Advanced algorithms and computational tools are being implemented to enable a more comprehensive data analysis, allowing researchers to uncover novel insights into AS and transcript expression [217,218,219,220,221,222]. For instance, the PlantExp platform integrates plant transcript expression and AS profiles from 131,423 uniformly processed publicly available RNA-seq samples that belong to 85 plant species across 24 plant orders [219]. This platform not only allows researchers to investigate and navigate across AS and transcript expression profiles, but also allows a differential and specific expression analysis, an analysis of co-expression networks, a cross-species expression conservation analysis, and an easy visualization of data [219].
Such platforms not only facilitate data-driven research, but also promote collaboration between scientists working on AS and transcript expression. By integrating fragmented data, ensuring data quality and accessibility, and providing powerful analysis tools, such platforms empower researchers to explore the intricate relationship between AS and transcript expression. In addition, scientists can flexibly customize sample groups to re-analyze publicly available RNA-seq datasets and obtain new insights [217,218,219,220,221,222].

9. Alternative Spliced Circadian Clock Genes in Response to Abiotic Stress

Circadian clock genes are a key point of regulation for adaptation to new environments and abiotic stress conditions. Alternative splicing plays a crucial role in the regulation of many core clock genes in plants and represents an important mechanistic link between the core of the circadian clock and diverse environmental inputs [223,224,225,226,227]. One example is partially redundant MYB-related transcription factors CIRCADIAN CLOCK ASSOCIATED 1 (CCA1) and LATE ELONGATED HYPOCOTYL (LHY) [228]. Under cold conditions, a non-functional spliced variant of LHY, which has a premature stop codon, is accumulated [223]. In contrast, the spliced variant of CCA1β, which lacks the MYB-like DNA-binding domain by retaining the fourth intron, is inhibited at low temperatures. The result is that CCA1β interferes with the formation of CCA1α (functional full-length CCA1) and LHY hetero- and homo-dimers [229]. The over-expression of CCA1β reduces freezing tolerance in Arabidopsis, while the over-expression of CCA1α increases tolerance to cold conditions [229]. Thus, the opposite regulation of LHY and CCA1 under low temperatures has revealed the role of AS in ensuring the balance of LHY and CCA1 under acclimation to low temperatures [223,229]. Moreover, extensive thermal fluctuations lead to a significant increase in CCA1β isoforms, causing shifts in their daily timing [230]. Likewise, other stresses, such as drought and Pseudomonas syringae infection, induce similar effects in the accumulation of CCA1β, thereby suggesting that stress-induced instabilities in the central oscillator might be partially compensated by out-of-phase CCA1 intron retention transcript oscillations [231].
In addition to CCA1 and LHY, other genes of the clock have been shown to display AS related to cold stress. For example, the spliced version of TIMING OF CAB EXPRESSION1 (TOC1β) is transcriptionally increased at low temperatures, while the ELF3β is suppressed under the same conditions [225]. Other abiotic stresses are also involved in differential alternative slicing in plants. For instance, heat stress triggers AS to increase the levels of CCA1β, PSEUDO-RESPONSE REGULATOR7β (PRR7β), TOC1β, and ELF3β, while saline conditions do not seem to affect AS of the CCA1, PRR7, TOC1, and ZEITLUPE (ZTL) genes, but reduce the ELF3β variant over the ELE3α, revealing a role of AS in the regulation of ELF3 under salt stress [225]. It has been shown that the splice variants of TOC1 and ELF3 undergo degradation via the nonsense-mediated decay (NMD) pathway, while the splice variants of other clock genes exhibit insensitivity to NMD [225].
Multiple spliceosome components are involved in AS of core plant clock genes. For example, the conserved methyltransferase PROTEIN ARGININE METHYLTRANSFEREASE 5 (PRMT5), involved in histone methylation, regulates AS of PRR9 [230,232]. Another key spliceosome component involved in AS of circadian genes is SNW/SKI-INTERACTING PROTEIN (SKIP). It is proposed that SKIP regulates AS of CCA1, LHY, PRR7, PRR9, and TOC1 by modulating the recognition of the 5′ and 3′ splice donor and acceptor sites. Conversely, loss of SKIP causes a long-period phenotype [224]. Likewise, mutants of SPLICEOSOMAL TIMEKEEPER LOCUS 1 (STIPL1), a homolog of a human spliceosome protein, also cause a long-period phenotype. Additionally, in stipl1 mutants, transcript levels of the spliced variants of CCA1, LHY, PRR9, and TOC1 are altered [233]. It has also been proposed that CCA1 mRNA intron retention modulation, via functional and nonsense/IR transcript ratios, potentially involves the splicing factor SR45 [231].
Core components of the spliceosomal U6 small nuclear ribonucleoprotein complex, SM-like (LSM) genes, also regulate circadian rhythms in plants. Mutations in LSM5 or LSM4 in Arabidopsis extend the circadian period by affecting AS more than constitutive splicing [234]. Another spliceosomal small nuclear ribonucleoprotein assembly factor, GEMIN2, has been suggested to attenuate the effects of temperature on the circadian period by regulation of AS of clock genes, such as CCA1, TOC1, and PRR9 [235]. Despite these discoveries, the complete details of the molecular mechanisms involved in AS effects on circadian components remain unknown.

10. Alternative Splicing Shapes Plant Symbiosis with Mycorrhiza and Rhizobia

Legumes establish a symbiotic relationship with N-fixing soil bacteria, whereas mycorrhiza establish a symbiotic relationship with both monocots and dicots. Rhizobium captures atmospheric N to support plant growth and development, while the bacteria use nutrients from the plants to support their own growth [236]. Mycorrhiza in optimal and stressed environments provide nutrients to host plants to improve biomass yield and quality of edible products under optimal and stressed environments [237]. Recent research as discussed herein states that AS variants contribute to the functioning of symbiosis in plants.

10.1. Mycorrhiza Symbiosis

Numerous genes regulate the formation of symbiotic structures and bidirectional nutrient exchange between host plant and mycorrhiza fungi. Tomatoes have emerged as a model plant for arbuscular mycorrhizal symbiosis (AMS). AMS in tomatoes up-regulated 3174 protein-coding genes, 42% of which were AS isoforms. Symbiosis consistently induced 24 genes from the ortho groups in eight phylogenetically distant angiosperms. Seven additional ortho groups were specifically induced by AMS in all surveyed dicot AMS-host plants, whereas these orthos were absent or not induced in monocots and/or non-AMS hosts, indicating a continuously evolving AMS-responsive network in addition to a conserved core regulatory module. A tomato symbiotic transcriptome database (https://efg.nju.edu.cn/TSTD accessed on 10 July 2023) may serve as a resource for deep deciphering of the AMS regulatory network [238].
AS regulates transcriptome and proteome diversity and therefore may influence symbiosis. Transcriptome profiling of pea roots in symbiosis with arbuscular mycorrhiza and control (nonsymbiotic) showed highly similar AS profiles. The intron retention type accounted for 67% of the AS types, as noted among plant species in general. Eight genes with AS events specific for mycorrhizal roots were identified, four of which were annotated as encoding an apoptosis inhibitor protein, a serine/threonine protein kinase, a dehydrodolichyl diphosphate synthase, and a pre-mRNA-splicing factor ATP-dependent RNA helicase DEAH1. The isoforms of these genes were up-regulated in mycorrhizal roots. Two such genes with mycorrhiza-specific AS were related to splicing and were part of the feedback loops involved in fine-tuning gene expression during mycorrhization [239].

10.2. Rhizobium Symbiosis

The Iso-Seq of soybean root tissues inoculated and uninoculated with Rhizobium unfolded 200,681 transcripts and covered 26,183 gene loci. Most of the multi-exon loci produced more than one splicing variant. Seven thousand and seventy-four DAS events had highly diverse splicing patterns (i.e., defense- and transport-related processes) during nodule development. The profiling of genes with differential isoform uses unlocked 2008 multi-isoform loci that underwent stage-specific or simultaneous major isoform switches after inoculation. In addition, 157 of 1563 high-confidence long non-coding RNAs (lncRNAs) were also differentially expressed during nodule development [240]. A study involving soybean transcriptome data unfolded key transcription patterns of nodule development, which included 9669 core genes and 7302 stage-specific genes and uncovered 2323 genes that undergo AS events during the nodule developmental stage in nodules compared to roots. Stage-specific changes during nodulation were also noted in DNA methylation that impacted the expression of 1864 genes. Thus, there exists an association among gene expression, AS, and DNA methylation in shaping transcriptome complexity and proteome specificity in developing nodules [5]. The assessment of AS events in the pea nodules and root tips unraveled AS isoforms of four genes, PsSIP1, PsIGN, PsWRKY40, and PsPR-10, with pathogens stress response isoforms more highly enriched in nodules than in root tips [241].

11. Applied Aspects of Splice Isoforms in Controlling Agricultural Traits

Alternative splicing produces more than one mRNA from a single pre-RNA molecule in plants, thus increasing transcriptome plasticity and proteome complexity [242] and affecting plant metabolism at different development stages [243]. AS provides therefore means for plants to adapt to changing surrounding environments by regulating their fitness, particularly when they grow under stress [244], e.g., in the response of barley’s clock genes to low temperature [226] or during infection of blast fungus in rice [245]. The recent advances in next-generation sequencing coupled with extensive transcriptomic resources have facilitated the understanding of the role of AS in regulating developmental processes in plants for adapting to stress-prone environments [246].
Splice variants affect agronomic characteristics in crops, e.g., floral development in cereals [247], seed shattering and weight in rice [248], grain size and weight in wheat [191,249], plant architecture in soybean [250], and nutritional quality in rice [210,251], soybean [252,253], tomato [254], and wheat [255]. Genome-wide association genetic analysis (GWAS) can further reveal how AS variants diversify gene function and regulate variation in crops, as shown by Chen et al. [22] in maize. They found ca. 20,000 unique splicing quantitative trait loci for 6570 genes affecting protein functions in 366 inbred lines.

12. Conclusions

AS of pre-mRNA is widespread and the major source of transcriptome and proteome diversity, which in turn generates phenotypic variation. A variety of computational pipelines including deep learning machine tools methods are now available to analyze RNA-seq data to identify AS events, estimate isoform abundance, and differentiate expression of splice variants across tissues/conditions and development stages.
Domestication and polyploidization (Brassica species and wheat) in addition to environmental perturbation cause varying expressions of AS isoforms in plants. Arabidopsis uses AS isoforms as a stress response mechanism to enhance its adaptation to a range of geographically diverse agro-ecologies. To date, many AS quantitative trait loci (sQTL) for a large number of genes with distinct protein functions impacting phenology, plant architecture, biomass yield, or quality, including nutrient homeostasis and stress responses, have been reported in grain (maize, rice, sorghum, and wheat), oil (Brassica species and soybean), and fiber (cotton) crops. Many of these sQTL colocalize with known pQTL impacting phenotypic variation. Evidence also suggests that AS variants contribute to the functioning of symbiosis (mycorrhiza and rhizobium) in plants and heterosis in grain and oil crops and provide a mechanistic link between the core of the circadian clock genes and diverse environmental stimuli.
Though significant advances in the genome-wide expression of AS variants have been made in various crops, applying such advances poses a significant challenge in crop improvement programs, which include but are not limited to (i) a significant bottleneck to establishing cost-effective high-throughput assays to identify AS variants in early breeding generations; (ii) differentiating and quantifying the impact of sQTL from pQTL for genes impacting phenotypic variation; (iii) accurate reconstruction of transcript isoforms and quantification of relative abundance of individual isoforms in deciphering the biological functions of individual transcripts; (iv) identifying common genetic tags (e.g., SNPs, InDels, and structural variation) linked with AS variants and gene expression; and (v) possible adverse effect of combining AS variants with trait gene(s) on phenotypic variation. Until such logistical issues are resolved, the exploitation of AS variants in crop improvement programs will be limited to the discovery and functional characterization of AS variants across tissues or conditions and development stages in plants.

Author Contributions

S.L.D.: Conceptualization, investigation, writing—original draft, and writing—review and editing. L.F.Q.: Investigation, writing—original draft, and editing. A.S.N.R.: Investigation, writing—original draft, and editing. C.S.: Investigation, writing—original draft, and editing. R.O.: Conceptualization, project administration, investigation, writing—original draft, and writing—review and editing. All authors contributed to the article and approved the submitted version. All authors have read and agreed to the published version of the manuscript.

Funding

ASNR acknowledges the support of the National Science Foundation (Grant No. DBI1949036 and Grant No. MCB 2014542).

Conflicts of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Gueroussov, S.; Gonatopoulos-Pournatzis, T.; Irimia, M.; Raj, B.; Lin, Z.-Y.; Gingras, A.-C.; Blencowe, B.J. An alternative splicing event amplifies evolutionary differences between vertebrates. Science 2015, 349, 868–873. [Google Scholar] [CrossRef]
  2. Josephs, E.B. Gene expression links genotype and phenotype during rapid adaptation. Mol. Ecol. 2020, 30, 30–32. [Google Scholar] [CrossRef]
  3. Tellier, M.; Maudlin, I.; Murphy, S. Transcription and splicing: A two-way street. Wiley Interdiscip. Rev. RNA 2020, 11, e1593. [Google Scholar] [CrossRef]
  4. Soergel, D.A.; Lareau, L.F.; Brenner, S.E. Regulation of gene expression by coupling of alternative splicing and NMD. Nonsense-Mediat. Mrna Decay 2006, 623, 175–196. [Google Scholar]
  5. Niyikiza, D.; Piya, S.; Routray, P.; Miao, L.; Kim, W.; Burch-Smith, T.; Gill, T.; Sams, C.; Arelli, P.R.; Pantalone, V.; et al. Interactions of gene expression, alternative splicing, and DNA methylation in determining nodule identity. Plant J. 2020, 103, 1744–1766. [Google Scholar] [CrossRef]
  6. E Grantham, M.; A Brisson, J. Extensive differential splicing underlies phenotypically plastic aphid morphs. Mol. Biol. Evol. 2018, 35, 1934–1946. [Google Scholar] [CrossRef] [PubMed]
  7. Healy, T.M.; Schulte, P.M. Patterns of alternative splicing in response to cold acclimation in fish. J. Exp. Biol. 2019, 222, jeb193516. [Google Scholar] [CrossRef] [PubMed]
  8. Jacobs, A.; Elmer, K.R. Alternative splicing and gene expression play contrasting roles in the parallel phenotypic evolution of a salmonid fish. Mol. Ecol. 2021, 30, 4955–4969. [Google Scholar] [CrossRef]
  9. Singh, P.; Börger, C.; More, H.; Sturmbauer, C. The role of alternative splicing and differential gene expression in cichlid adaptive radiation. Genome Biol. Evol. 2017, 9, 2764–2781. [Google Scholar] [CrossRef] [PubMed]
  10. Singh, P.; Ahi, E.P. The importance of alternative splicing in adaptive evolution. Mol. Ecol. 2022, 31, 1928–1938. [Google Scholar] [CrossRef] [PubMed]
  11. Chen, L.; Bush, S.J.; Tovar-Corona, J.M.; Castillo-Morales, A.; Urrutia, A.O. Correcting for differential transcript coverage reveals a strong relationship between alternative splicing and organism complexity. Mol. Biol. Evol. 2014, 31, 1402–1413. [Google Scholar] [CrossRef] [PubMed]
  12. Brawand, D.; Soumillon, M.; Necsulea, A.; Julien, P.; Csardi, G.; Harrigan, P.; Weier, M.; Liechti, A.; Aximu-Petri, A.; Kircher, M.; et al. The evolution of gene expression levels in mammalian organs. Nature 2011, 478, 343–348. [Google Scholar] [CrossRef]
  13. El Taher, A.; Böhne, A.; Boileau, N.; Ronco, F.; Indermaur, A.; Widmer, L.; Salzburger, W. Gene expression dynamics during rapid organismal diversification in African cichlid fishes. Nat. Ecol. Evol. 2020, 5, 243–250. [Google Scholar] [CrossRef]
  14. Hill, M.S.; Zande, P.V.; Wittkopp, P.J. Molecular and evolutionary processes generating variation in gene expression. Nat. Rev. Genet. 2020, 22, 203–215. [Google Scholar] [CrossRef] [PubMed]
  15. Wray, G.A. The evolutionary significance of cis-regulatory mutations. Nat. Rev. Genet. 2007, 8, 206–216. [Google Scholar] [CrossRef]
  16. Blencowe, B.J. The relationship between alternative splicing and proteomic complexity. Trends Biochem. Sci. 2017, 42, 407–408. [Google Scholar] [CrossRef] [PubMed]
  17. Tress, M.L.; Abascal, F.; Valencia, A. Alternative splicing may not be the key to proteome complexity. Trends Biochem. Sci. 2016, 42, 98–110. [Google Scholar] [CrossRef]
  18. Bedre, R.; Irigoyen, S.; Petrillo, E.; Mandadi, K.K. New era in plant alternative splicing analysis enabled by advances in high-throughput sequencing (HTS) technologies. Front. Plant Sci. 2019, 10, 740. [Google Scholar] [CrossRef]
  19. Xia, T.; Zhang, L.; Xu, J.; Wang, L.; Liu, B.; Hao, M.; Chang, X.; Zhang, T.; Li, S.; Zhang, H.; et al. The alternative splicing of EAM8 contributes to early flowering and short-season adaptation in a landrace barley from the Qinghai-Tibetan Plateau. Theor. Appl. Genet. 2017, 130, 757–766. [Google Scholar] [CrossRef]
  20. Smith, C.C.R.; Rieseberg, L.H.; Hulke, B.S.; Kane, N.C. Aberrant RNA splicing due to genetic incompatibilities in sunflower hybrids. Evolution 2021, 75, 2747–2758. [Google Scholar] [CrossRef] [PubMed]
  21. Smith, C.C.R.; Tittes, S.; Mendieta, J.P.; Collier-Zans, E.; Rowe, H.C.; Rieseberg, L.H.; Kane, N.C. Genetics of alternative splicing evolution during sunflower domestication. Proc. Natl. Acad. Sci. USA 2018, 115, 6768–6773. [Google Scholar] [CrossRef]
  22. Chen, Q.; Han, Y.; Liu, H.; Wang, X.; Sun, J.; Zhao, B.; Li, W.; Tian, J.; Liang, Y.; Yan, J.; et al. Genome-wide association analyses reveal the importance of alternative splicing in diversifying gene function and regulating phenotypic variation in maize. Plant Cell 2018, 30, 1404–1423. [Google Scholar] [CrossRef] [PubMed]
  23. West-Eberhard, M.J. Phenotypic plasticity and the origins of diversity. Annu. Rev. Ecol. Syst. 1989, 20, 249–278. [Google Scholar] [CrossRef]
  24. Ehrenreich, I.M.; Pfennig, D.W. Genetic assimilation: A review of its potential proximate causes and evolutionary consequences. Ann. Bot. 2015, 117, 769–779. [Google Scholar] [CrossRef]
  25. Somero, G.N. RNA thermosensors: How might animals exploit their regulatory potential? J. Exp. Biol. 2018, 221, jeb162842. [Google Scholar] [CrossRef] [PubMed]
  26. Mastrangelo, A.M.; Marone, D.; Laidò, G.; De Leonardis, A.M.; De Vita, P. Alternative splicing: Enhancing ability to cope with stress via transcriptome plasticity. Plant Sci. 2012, 185–186, 40–49. [Google Scholar] [CrossRef]
  27. Sedlazeck, F.J.; Lee, H.; Darby, C.A.; Schatz, M.C. Piercing the dark matter: Bioinformatics of long-range sequencing and mapping. Nat. Rev. Genet. 2018, 19, 329–346. [Google Scholar] [CrossRef] [PubMed]
  28. Jabre, I.; Reddy, A.S.N.; Kalyna, M.; Chaudhary, S.; Khokhar, W.; Byrne, L.J.; Wilson, C.M.; Syed, N.H. Does co-transcriptional regulation of alternative splicing mediate plant stress responses? Nucleic Acids Res. 2019, 47, 2716–2726. [Google Scholar] [CrossRef] [PubMed]
  29. Reddy, A.S.; Huang, J.; Syed, N.H.; Ben-Hur, A.; Dong, S.; Gu, L. Decoding co-/post-transcriptional complexities of plant transcriptomes and epitranscriptome using next-generation sequencing technologies. Biochem. Soc. Trans. 2020, 48, 2399–2414. [Google Scholar] [CrossRef] [PubMed]
  30. Abdel-Ghany, S.E.; Hamilton, M.; Jacobi, J.L.; Ngam, P.; Devitt, N.; Schilkey, F.; Ben-Hur, A.; Reddy, A.S.N. A survey of the sorghum transcriptome using single-molecule long reads. Nat. Commun. 2016, 7, 11706. [Google Scholar] [CrossRef]
  31. Zhang, R.; Kuo, R.; Coulter, M.; Calixto, C.P.G.; Entizne, J.C.; Guo, W.; Marquez, Y.; Milne, L.; Riegler, S.; Matsui, A.; et al. A high-resolution single-molecule sequencing-based Arabidopsis transcriptome using novel methods of Iso-seq analysis. Genome Biol. 2022, 23, 149. [Google Scholar] [CrossRef]
  32. Zhao, L.; Zhang, H.; Kohnen, M.V.; Prasad, K.V.S.K.; Gu, L.; Reddy, A.S.N. Analysis of transcriptome and epitranscriptome in plants using PacBio iso-seq and nanopore-based direct RNA sequencing. Front. Genet. 2019, 10, 253. [Google Scholar] [CrossRef]
  33. Parker, M.T.; Knop, K.; Sherwood, A.V.; Schurch, N.J.; Mackinnon, K.; Gould, P.D.; Hall, A.J.; Barton, G.J.; Simpson, G.G. Nanopore direct RNA sequencing maps the complexity of Arabidopsis mRNA processing and m6A modification. eLife 2020, 9, e49658. [Google Scholar] [CrossRef]
  34. Zhang, S.; Li, R.; Zhang, L.; Chen, S.; Xie, M.; Yang, L.; Xia, Y.; Foyer, C.H.; Zhao, Z.; Lam, H.-M. New insights into Arabidopsis transcriptome complexity revealed by direct sequencing of native RNAs. Nucleic Acids Res. 2020, 48, 7700–7711. [Google Scholar] [CrossRef]
  35. Wang, Y.; Wang, H.; Xi, F.; Wang, H.; Han, X.; Wei, W.; Zhang, H.; Zhang, Q.; Zheng, Y.; Zhu, Q.; et al. Profiling of circular RNA N6-methyladenosine in moso bamboo (Phyllostachys edulis) using nanopore-based direct RNA sequencing. J. Integr. Plant Biol. 2020, 62, 1823–1838. [Google Scholar] [CrossRef]
  36. 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. 2009, 20, 45–58. [Google Scholar] [CrossRef]
  37. Mandadi, K.K.; Scholthof, K.-B.G. Genome-wide analysis of alternative splicing landscapes modulated during plant-virus interactions in Brachypodium distachyon. Plant Cell 2015, 27, 71–85. [Google Scholar] [CrossRef]
  38. Thatcher, S.R.; Zhou, W.; Leonard, A.; Wang, B.-B.; Beatty, M.; Zastrow-Hayes, G.; Zhao, X.; Baumgarten, A.; Li, B. Genome-wide analysis of alternative splicing in Zea mays: Landscape and genetic regulation. Plant Cell 2014, 26, 3472–3487. [Google Scholar] [CrossRef] [PubMed]
  39. Marquez, Y.; Brown, J.W.; Simpson, C.; 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]
  40. Mourão, K.; Schurch, N.J.; Lucoszek, R.; Froussios, K.; MacKinnon, K.; Duc, C.; Simpson, G.; Barton, G.J. Detection and mitigation of spurious antisense expression with RoSA. F1000Research 2019, 8, 819. [Google Scholar] [CrossRef]
  41. Guo, W.; Coulter, M.; Waugh, R.; Zhang, R. The value of genotype-specific reference for transcriptome analyses in barley. Life Sci. Alliance 2022, 5, e202101255. [Google Scholar] [CrossRef] [PubMed]
  42. Kratz, A.; Carninci, P. The devil in the details of RNA-seq. Nat. Biotechnol. 2014, 32, 882–884. [Google Scholar] [CrossRef] [PubMed]
  43. Steijger, T.; The RGASP Consortium; Abril, J. F.; Engström, P.G.; Kokocinski, F.; Hubbard, T.J.; Guigó, R.; Harrow, J.; Bertone, P. Assessment of transcript reconstruction methods for RNA-seq. Nat. Methods 2013, 10, 1177–1184. [Google Scholar] [CrossRef] [PubMed]
  44. Schaarschmidt, S.; Fischer, A.; Lawas, L.M.F.; Alam, R.; Septiningsih, E.M.; Bailey-Serres, J.; Jagadish, S.V.K.; Huettel, B.; Hincha, D.K.; Zuther, E. Utilizing PacBio iso-seq for novel transcript and gene discovery of abiotic stress responses in Oryza sativa L. Int. J. Mol. Sci. 2020, 21, 8148. [Google Scholar] [CrossRef] [PubMed]
  45. Feng, S.; Xu, M.; Liu, F.; Cui, C.; Zhou, B. Reconstruction of the full-length transcriptome atlas using PacBio Iso-Seq provides insight into the alternative splicing in Gossypium australe. BMC Plant Biol. 2019, 19, 1–16. [Google Scholar] [CrossRef]
  46. Minio, A.; Massonnet, M.; Figueroa-Balderas, R.; Vondras, A.M.; Blanco-Ulate, B.; Cantu, D. Iso-seq allows genome-independent transcriptome profiling of grape berry development. G3 Genes|Genomes|Genet. 2019, 9, 755–767. [Google Scholar] [CrossRef]
  47. Wei, J.; Cao, H.; Liu, J.-D.; Zuo, J.-H.; Fang, Y.; Lin, C.-T.; Sun, R.-Z.; Li, W.-L.; Liu, Y.-X. Insights into transcriptional characteristics and homoeolog expression bias of embryo and de-embryonated kernels in developing grain through RNA-Seq and Iso-Seq. Funct. Integr. Genom. 2019, 19, 919–932. [Google Scholar] [CrossRef] [PubMed]
  48. Juntawong, P.; Girke, T.; Bazin, J.; Bailey-Serres, J. Translational dynamics revealed by genome-wide profiling of ribosome footprints in Arabidopsis. Proc. Natl. Acad. Sci. USA 2013, 111, E203–E212. [Google Scholar] [CrossRef]
  49. Reixachs-Solé, M.; Ruiz-Orera, J.; Albà, M.M.; Eyras, E. Ribosome profiling at isoform level reveals evolutionary conserved impacts of differential splicing on the proteome. Nat. Commun. 2020, 11, 1768. [Google Scholar] [CrossRef]
  50. Holmes, I.; Durbin, R. Dynamic programming alignment accuracy. J. Comput. Biol. 1998, 5, 493–504. [Google Scholar] [CrossRef]
  51. Kuo, R.I.; Cheng, Y.; Zhang, R.; Brown, J.W.S.; Smith, J.; Archibald, A.L.; Burt, D.W. Illuminating the dark side of the human transcriptome with long read transcript sequencing. BMC Genom. 2020, 21, 1–22. [Google Scholar] [CrossRef]
  52. Marquardt, S.; Petrillo, E.; A Manavella, P. Cotranscriptional RNA processing and modification in plants. Plant Cell 2022, 35, 1654–1670. [Google Scholar] [CrossRef]
  53. Zhu, D.; Mao, F.; Tian, Y.; Lin, X.; Gu, L.; Gu, H.; Qu, L.-J.; Wu, Y.; Wu, Z. The features and regulation of co-transcriptional splicing in Arabidopsis. Mol. Plant 2020, 13, 278–294. [Google Scholar] [CrossRef]
  54. Li, S.; Wang, Y.; Zhao, Y.; Zhao, X.; Chen, X.; Gong, Z. Global co-transcriptional splicing in Arabidopsis and the correlation with splicing regulation in mature RNAs. Mol. Plant 2019, 13, 266–277. [Google Scholar] [CrossRef] [PubMed]
  55. Reddy, A.S.; Marquez, Y.; Kalyna, M.; Barta, A. Complexity of the alternative splicing landscape in plants. Plant Cell 2013, 25, 3657–3683. [Google Scholar] [CrossRef]
  56. Staiger, D.; Brown, J.W. Alternative splicing at the intersection of biological timing, development, and stress responses. Plant Cell 2013, 25, 3640–3656. [Google Scholar] [CrossRef] [PubMed]
  57. Laloum, T.; Martín, G.; Duque, P. Alternative splicing control of abiotic stress responses. Trends Plant Sci. 2018, 23, 140–150. [Google Scholar] [CrossRef] [PubMed]
  58. Calixto, C.P.; Guo, W.; James, A.B.; Tzioutziou, N.A.; Entizne, J.C.; Panter, P.E.; Knight, H.; Nimmo, H.G.; Zhang, R.; Brown, J.W. rapid and dynamic alternative splicing impacts the Arabidopsis cold response transcriptome. Plant Cell 2018, 30, 1424–1444. [Google Scholar] [CrossRef]
  59. Palusa, S.G.; Ali, G.S.; Reddy, A.S. Alternative splicing of pre-mRNAs of Arabidopsis serine/arginine-rich proteins: Regulation by hormones and stresses. Plant J. 2007, 49, 1091–1107. [Google Scholar] [CrossRef]
  60. Palusa, S.G.; Reddy, A.S.N. Extensive coupling of alternative splicing of pre-mRNAs of serine/arginine (SR) genes with nonsense-mediated decay. New Phytol. 2009, 185, 83–89. [Google Scholar] [CrossRef] [PubMed]
  61. Palusa, S.G.; Reddy, A.S. differential recruitment of splice variants from SR Pre-mRNAs to polysomes during development and in response to stresses. Plant Cell Physiol. 2015, 56, 421–427. [Google Scholar] [CrossRef] [PubMed]
  62. Petrillo, E. Do not panic: An intron-centric guide to alternative splicing. Plant Cell 2023, 35, 1752–1761. [Google Scholar] [CrossRef] [PubMed]
  63. Jia, J.; Long, Y.; Zhang, H.; Li, Z.; Liu, Z.; Zhao, Y.; Lu, D.; Jin, X.; Deng, X.; Xia, R.; et al. Post-transcriptional splicing of nascent RNA contributes to widespread intron retention in plants. Nat. Plants 2020, 6, 780–788. [Google Scholar] [CrossRef]
  64. Braunschweig, U.; Barbosa-Morais, N.L.; Pan, Q.; Nachman, E.N.; Alipanahi, B.; Gonatopoulos-Pournatzis, T.; Frey, B.; Irimia, M.; Blencowe, B.J. Widespread intron retention in mammals functionally tunes transcriptomes. Genome Res. 2014, 24, 1774–1786. [Google Scholar] [CrossRef] [PubMed]
  65. Boothby, T.C.; Zipper, R.S.; van der Weele, C.M.; Wolniak, S.M. Removal of retained introns regulates translation in the rapidly developing gametophyte of Marsilea vestita. Dev. Cell 2013, 24, 517–529. [Google Scholar] [CrossRef]
  66. Yap, K.; Lim, Z.Q.; Khandelia, P.; Friedman, B.; Makeyev, E.V. Coordinated regulation of neuronal mRNA steady-state levels through developmentally controlled intron retention. Genes Dev. 2012, 26, 1209–1223. [Google Scholar] [CrossRef] [PubMed]
  67. Jung, H.; Lee, D.; Lee, J.; Park, D.; Kim, Y.J.; Park, W.-Y.; Hong, D.; Park, P.J.; Lee, E. Intron retention is a widespread mechanism of tumor-suppressor inactivation. Nat. Genet. 2015, 47, 1242–1248. [Google Scholar] [CrossRef] [PubMed]
  68. Mehmood, A.; Laiho, A.; Venäläinen, M.S.; McGlinchey, A.J.; Wang, N.; Elo, L.L. Systematic evaluation of differential splicing tools for RNA-seq studies. Brief. Bioinform. 2019, 21, 2052–2065. [Google Scholar] [CrossRef] [PubMed]
  69. Mancini, E.; Rabinovich, A.; Iserte, J.; Yanovsky, M.; Chernomoretz, A. ASpli: Integrative analysis of splicing landscapes through RNA-Seq assays. Bioinformatics 2021, 37, 2609–2616. [Google Scholar] [CrossRef] [PubMed]
  70. Guo, W.; A Tzioutziou, N.; Stephen, G.; Milne, I.; Calixto, C.P.; Waugh, R.; Brown, J.W.S.; Zhang, R. 3D RNA-seq: A powerful and flexible tool for rapid and accurate differential expression and alternative splicing analysis of RNA-seq data for biologists. RNA Biol. 2020, 18, 1574–1587. [Google Scholar] [CrossRef] [PubMed]
  71. Hsieh, P.-H.; Oyang, Y.-J.; Chen, C.-Y. Effect of de novo transcriptome assembly on transcript quantification. Sci. Rep. 2019, 9, 8304. [Google Scholar] [CrossRef]
  72. Freedman, A.H.; Clamp, M.; Sackton, T.B. Error, noise and bias in de novo transcriptome assemblies. Mol. Ecol. Resour. 2020, 21, 18–29. [Google Scholar] [CrossRef]
  73. Jagannathan, S.; Ramachandran, S.; Rissland, O.S. Slow down to catch up. Cell 2019, 178, 774–776. [Google Scholar] [CrossRef] [PubMed]
  74. Zeng, T.; I Li, Y. Predicting RNA splicing from DNA sequence using Pangolin. Genome Biol. 2022, 23, 103. [Google Scholar] [CrossRef]
  75. Avsec, Ž.; Agarwal, V.; Visentin, D.; Ledsam, J.R.; Grabska-Barwinska, A.; Taylor, K.R.; Assael, Y.; Jumper, J.; Kohli, P.; Kelley, D.R. Effective gene expression prediction from sequence by integrating long-range interactions. Nat. Methods 2021, 18, 1196–1203. [Google Scholar] [CrossRef] [PubMed]
  76. Dawes, R.; Bournazos, A.M.; Bryen, S.J.; Bommireddipalli, S.; Marchant, R.G.; Joshi, H.; Cooper, S.T. SpliceVault predicts the precise nature of variant-associated mis-splicing. Nat. Genet. 2023, 55, 324–332. [Google Scholar] [CrossRef] [PubMed]
  77. Cheng, J.; Nguyen, T.Y.D.; Cygan, K.J.; Çelik, M.H.; Fairbrother, W.G.; Avsec, Ž.; Gagneur, J. MMSplice: Modular modeling improves the predictions of genetic variant effects on splicing. Genome Biol. 2019, 20, 48. [Google Scholar] [CrossRef] [PubMed]
  78. Dawes, R.; Joshi, H.; Cooper, S.T. Empirical prediction of variant-activated cryptic splice donors using population-based RNA-Seq data. Nat. Commun. 2022, 13, 1655. [Google Scholar] [CrossRef] [PubMed]
  79. Zhang, Z.; Pan, Z.; Ying, Y.; Xie, Z.; Adhikari, S.; Phillips, J.; Carstens, R.P.; Black, D.L.; Wu, Y.; Xing, Y. Deep-learning augmented RNA-seq analysis of transcript splicing. Nat. Methods 2019, 16, 307–310. [Google Scholar] [CrossRef]
  80. Huang, J.; Lu, X.; Wu, H.; Xie, Y.; Peng, Q.; Gu, L.; Wu, J.; Wang, Y.; Reddy, A.S.; Dong, S. Phytophthora effectors modulate genome-wide alternative splicing of host mRNAs to reprogram plant immunity. Mol. Plant 2020, 13, 1470–1484. [Google Scholar] [CrossRef]
  81. Shen, S.; Park, J.W.; Lu, Z.-X.; Lin, L.; Henry, M.D.; Wu, Y.N.; Zhou, Q.; Xing, Y. rMATS: Robust and flexible detection of differential alternative splicing from replicate RNA-Seq data. Proc. Natl. Acad. Sci. USA 2014, 111, E5593–E5601. [Google Scholar] [CrossRef]
  82. Reyes, A.; Anders, S.; Weatheritt, R.J.; Gibson, T.J.; Steinmetz, L.M.; Huber, W. Drift and conservation of differential exon usage across tissues in primate species. Proc. Natl. Acad. Sci. USA 2013, 110, 15377–15382. [Google Scholar] [CrossRef]
  83. Vaquero-Garcia, J.; Barrera, A.; Gazzara, M.R.; González-Vallinas, J.; Lahens, N.F.; Hogenesch, J.B.; Lynch, K.W.; Barash, Y. A new view of transcriptome complexity and regulation through the lens of local splicing variations. eLife 2016, 5, e11752. [Google Scholar] [CrossRef]
  84. Trincado, J.L.; Entizne, J.C.; Hysenaj, G.; Singh, B.; Skalic, M.; Elliott, D.J.; Eyras, E. SUPPA2: Fast, accurate, and uncertainty-aware differential splicing analysis across multiple conditions. Genome Biol. 2018, 19, 40. [Google Scholar] [CrossRef] [PubMed]
  85. Coulter, M.; Entizne, J.C.; Guo, W.; Bayer, M.; Wonneberger, R.; Milne, L.; Schreiber, M.; Haaning, A.; Muehlbauer, G.J.; McCallum, N.; et al. BaRTv2: A highly resolved barley reference transcriptome for accurate transcript-specific RNA-seq quantification. Plant J. 2022, 111, 1183–1202. [Google Scholar] [CrossRef]
  86. Katz, Y.; Wang, E.T.; Airoldi, E.M.; Burge, C.B. Analysis and design of RNA sequencing experiments for identifying isoform regulation. Nat. Methods 2010, 7, 1009–1015. [Google Scholar] [CrossRef]
  87. Rogers, M.F.; Thomas, J.; Reddy, A.S.; Ben-Hur, A. SpliceGrapher: Detecting patterns of alternative splicing from RNA-Seq data in the context of gene models and EST data. Genome Biol. 2012, 13, R4. [Google Scholar] [CrossRef] [PubMed]
  88. Filichkin, S.A.; Hamilton, M.; Dharmawardhana, P.D.; Singh, S.K.; Sullivan, C.; Ben-Hur, A.; Reddy, A.S.N.; Jaiswal, P. Abiotic Stresses modulate landscape of poplar transcriptome via alternative splicing, differential intron retention, and isoform ratio switching. Front. Plant Sci. 2018, 9, 5. [Google Scholar] [CrossRef] [PubMed]
  89. Martín, G.; Márquez, Y.; Mantica, F.; Duque, P.; Irimia, M. Alternative splicing landscapes in Arabidopsis thaliana across tissues and stress conditions highlight major functional differences with animals. Genome Biol. 2021, 22, 35. [Google Scholar] [CrossRef] [PubMed]
  90. Chaudhary, S.; Khokhar, W.; Jabre, I.; Reddy, A.S.N.; Byrne, L.J.; Wilson, C.M.; Syed, N.H. Alternative splicing and protein diversity: Plants versus animals. Front. Plant Sci. 2019, 10, 708. [Google Scholar] [CrossRef]
  91. Wang, R.; Liu, H.; Liu, Z.; Zou, J.; Meng, J.; Wang, J. Genome-wide analysis of alternative splicing divergences between Brassica hexaploid and its parents. Planta 2019, 250, 603–628. [Google Scholar] [CrossRef] [PubMed]
  92. Zhang, M.; Liu, Y.-H.; Xu, W.; Smith, C.W.; Murray, S.C.; Zhang, H.-B. Analysis of the genes controlling three quantitative traits in three diverse plant species reveals the molecular basis of quantitative traits. Sci. Rep. 2020, 10, 10074. [Google Scholar] [CrossRef]
  93. Barbazuk, W.B.; Fu, Y.; McGinnis, K.M. Genome-wide analyses of alternative splicing in plants: Opportunities and challenges. Genome Res. 2008, 18, 1381–1392. [Google Scholar] [CrossRef] [PubMed]
  94. Shen, S.; Park, J.W.; Huang, J.; Dittmar, K.A.; Lu, Z.-X.; Zhou, Q.; Carstens, R.P.; Xing, Y. MATS: A Bayesian framework for flexible detection of differential alternative splicing from RNA-Seq data. Nucleic Acids Res. 2012, 40, e61. [Google Scholar] [CrossRef] [PubMed]
  95. Mathur, M.; Kim, C.M.; Munro, S.A.; Rudina, S.S.; Sawyer, E.M.; Smolke, C.D. Programmable mutually exclusive alternative splicing for generating RNA and protein diversity. Nat. Commun. 2019, 10, 2673. [Google Scholar] [CrossRef] [PubMed]
  96. Pramanik, D.; Shelake, R.M.; Kim, M.J.; Kim, J.-Y. CRISPR-mediated engineering across the central dogma in plant biology for basic research and crop improvement. Mol. Plant 2020, 14, 127–150. [Google Scholar] [CrossRef]
  97. Li, H.; Li, A.; Shen, W.; Ye, N.; Wang, G.; Zhang, J. Global survey of alternative splicing in rice by direct rna sequencing during reproductive development: Landscape and genetic regulation. Rice 2021, 14, 75. [Google Scholar] [CrossRef]
  98. Sun, Y.; Xiao, H. Identification of alternative splicing events by RNA sequencing in early growth tomato fruits. BMC Genom. 2015, 16, 948. [Google Scholar] [CrossRef] [PubMed]
  99. Wang, M.; Wang, P.; Liang, F.; Ye, Z.; Li, J.; Shen, C.; Pei, L.; Wang, F.; Hu, J.; Tu, L.; et al. A global survey of alternative splicing in allopolyploid cotton: Landscape, complexity and regulation. New Phytol. 2017, 217, 163–178. [Google Scholar] [CrossRef]
  100. Abdel-Ghany, S.E.; Ullah, F.; Ben-Hur, A.; Reddy, A.S.N. transcriptome analysis of drought-resistant and drought-sensitive sorghum (Sorghum bicolor) genotypes in response to PEG-induced drought stress. Int. J. Mol. Sci. 2020, 21, 772. [Google Scholar] [CrossRef]
  101. Xie, S.-Q.; Han, Y.; Chen, X.-Z.; Cao, T.-Y.; Ji, K.-K.; Zhu, J.; Ling, P.; Xiao, C.-L. ISOdb: A comprehensive database of full-length isoforms generated by iso-seq. Int. J. Genom. 2018, 2018, 9207637. [Google Scholar] [CrossRef]
  102. Ganie, S.A.; Reddy, A.S.N. Stress-induced changes in alternative splicing landscape in rice: Functional significance of splice isoforms in stress tolerance. Biology 2021, 10, 309. [Google Scholar] [CrossRef]
  103. Kathare, P.K.; Xin, R.; Ganesan, A.S.; June, V.M.; Reddy, A.S.N.; Huq, E. SWAP1-SFPS-RRC1 splicing factor complex modulates pre-mRNA splicing to promote photomorphogenesis in Arabidopsis. Proc. Natl. Acad. Sci. USA 2022, 119, e2214565119. [Google Scholar] [CrossRef]
  104. Zhu, G.; Li, W.; Zhang, F.; Guo, W. RNA-seq analysis reveals alternative splicing under salt stress in cotton, Gossypium davidsonii. BMC Genom. 2018, 19, 73. [Google Scholar] [CrossRef] [PubMed]
  105. Li, S.; Yu, X.; Cheng, Z.; Zeng, C.; Li, W.; Zhang, L.; Peng, M. Large-scale analysis of the cassava transcriptome reveals the impact of cold stress on alternative splicing. J. Exp. Bot. 2019, 71, 422–434. [Google Scholar] [CrossRef] [PubMed]
  106. Dong, C.; He, F.; Berkowitz, O.; Liu, J.; Cao, P.; Tang, M.; Shi, S.; Wang, W.; Li, Q.; Whelan, J.; et al. Alternative splicing plays a critical role in maintaining mineral nutrient homeostasis in rice (Oryza sativa). Plant Cell. 2018, 30, 2267–2285. [Google Scholar] [CrossRef] [PubMed]
  107. Chong, G.L.; Foo, M.H.; Lin, W.-D.; Wong, M.M.; Verslues, P.E. Highly ABA-Induced 1 (HAI1)-Interacting protein HIN1 and drought acclimation-enhanced splicing efficiency at intron retention sites. Proc. Natl. Acad. Sci. USA 2019, 116, 22376–22385. [Google Scholar] [CrossRef]
  108. Chen, M.; Manley, J.L. Mechanisms of alternative splicing regulation: Insights from molecular and genomics approaches. Nat. Rev. Mol. Cell Biol. 2009, 10, 741–754. [Google Scholar] [CrossRef] [PubMed]
  109. Assmann, S.M.; Chou, H.-L.; Bevilacqua, P.C. Rock, scissors, paper: How RNA structure informs function. Plant Cell 2023, 35, 1671–1707. [Google Scholar] [CrossRef]
  110. Ding, Y.; Tang, Y.; Kwok, C.K.; Zhang, Y.; Bevilacqua, P.C.; Assmann, S.M. In vivo genome-wide profiling of RNA secondary structure reveals novel regulatory features. Nature 2013, 505, 696–700. [Google Scholar] [CrossRef]
  111. Thomas, J.; Palusa, S.G.; Prasad, K.V.; Ali, G.S.; Surabhi, G.-K.; Ben-Hur, A.; Abdel-Ghany, S.E.; Reddy, A.S. Identification of an intronic splicing regulatory element involved in auto-regulation of alternative splicing of SCL33 pre-mRNA. Plant J. 2012, 72, 935–946. [Google Scholar] [CrossRef] [PubMed]
  112. 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]
  113. Xing, D.; Wang, Y.; Hamilton, M.; Ben-Hur, A.; Reddy, A.S. Transcriptome-wide identification of RNA targets of Arabidopsis SERINE/ARGININE-RICH45 uncovers the unexpected roles of this RNA binding protein in RNA processing. Plant Cell 2015, 27, 3294–3308. [Google Scholar] [CrossRef]
  114. Liu, Z.; Yuan, G.; Liu, S.; Jia, J.; Cheng, L.; Qi, D.; Shen, S.; Peng, X.; Liu, G. Identified of a novel cis-element regulating the alternative splicing of LcDREB2. Sci. Rep. 2017, 7, srep46106. [Google Scholar] [CrossRef] [PubMed]
  115. Wang, B.-B.; Brendel, V. Genomewide comparative analysis of alternative splicing in plants. Proc. Natl. Acad. Sci. USA 2006, 103, 7175–7180. [Google Scholar] [CrossRef] [PubMed]
  116. Iida, K. Genome-wide analysis of alternative pre-mRNA splicing in Arabidopsis thaliana based on full-length cDNA sequences. Nucleic Acids Res. 2004, 32, 5096–5103. [Google Scholar] [CrossRef]
  117. Long, J.C.; Caceres, J.F. The SR protein family of splicing factors: Master regulators of gene expression. Biochem. J. 2009, 417, 15–27. [Google Scholar] [CrossRef] [PubMed]
  118. Lareau, L.F.; Inada, M.; Green, R.E.; Wengrod, J.C.; Brenner, S.E. Unproductive splicing of SR genes associated with highly conserved and ultraconserved DNA elements. Nature 2007, 446, 926–929. [Google Scholar] [CrossRef]
  119. Lazar, G.; Goodman, H.M. The Arabidopsis splicing factor SR1 is regulated by alternative splicing. Plant Mol. Biol. 2000, 42, 571–581. [Google Scholar] [CrossRef]
  120. Rauch, H.B.; Patrick, T.L.; Klusman, K.M.; Battistuzzi, F.U.; Mei, W.; Brendel, V.P.; Lal, S.K. Discovery and expression analysis of alternative splicing events conserved among plant SR proteins. Mol. Biol. Evol. 2013, 31, 605–613. [Google Scholar] [CrossRef]
  121. Reddy, A.S.N.; Shad Ali, G. Plant serine/arginine-rich proteins: Roles in precursor messenger RNA splicing, plant development, and stress responses: Plant SR-rich proteins. Wiley Interdiscip. Rev. RNA 2011, 2, 875–889. [Google Scholar] [CrossRef]
  122. Duque, P. A role for SR proteins in plant stress responses. Plant Signal. Behav. 2011, 6, 49–54. [Google Scholar] [CrossRef] [PubMed]
  123. Reddy, A.S. Alternative splicing of pre-messenger rnas in plants in the genomic era. Annu. Rev. Plant Biol. 2007, 58, 267–294. [Google Scholar] [CrossRef]
  124. Ali, G.S.; Reddy, A.S.N. Regulation of alternative splicing of pre-mRNAs by stresses. Poxviruses 2008, 326, 257–275. [Google Scholar] [CrossRef]
  125. 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]
  126. 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]
  127. 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]
  128. Chaudhary, S.; Jabre, I.; Reddy, A.S.; Staiger, D.; Syed, N.H. Perspective on alternative splicing and proteome complexity in plants. Trends Plant Sci. 2019, 24, 496–506. [Google Scholar] [CrossRef]
  129. Khokhar, W.; Hassan, M.A.; Reddy, A.S.N.; Chaudhary, S.; Jabre, I.; Byrne, L.J.; Syed, N.H. Genome-wide identification of splicing quantitative trait loci (sQTLs) in diverse ecotypes of Arabidopsis thaliana. Front. Plant Sci. 2019, 10, 1160. [Google Scholar] [CrossRef] [PubMed]
  130. Köster, T.; Marondedze, C.; Meyer, K.; Staiger, D. RNA-binding proteins revisited—The emerging Arabidopsis mRNA interactome. Trends Plant Sci. 2017, 22, 512–526. [Google Scholar] [CrossRef]
  131. Burjoski, V.; Reddy, A.S.N. The landscape of RNA-protein interactions in plants: Approaches and current status. Int. J. Mol. Sci. 2021, 22, 2845. [Google Scholar] [CrossRef]
  132. Albaqami, M.; Reddy, A.S.N. Development of an in vitro pre-mRNA splicing assay using plant nuclear extract. Plant Methods 2018, 14, 1. [Google Scholar] [CrossRef]
  133. Zhou, G.; Niu, R.; Zhou, Y.; Luo, M.; Peng, Y.; Wang, H.; Wang, Z.; Xu, G. Proximity editing to identify RNAs in phase-separated RNA binding protein condensates. Cell Discov. 2021, 7, 72. [Google Scholar] [CrossRef]
  134. McMahon, A.C.; Rahman, R.; Jin, H.; Shen, J.L.; Fieldsend, A.; Luo, W.; Rosbash, M. TRIBE: Hijacking an RNA-editing enzyme to identify cell-specific targets of rna-binding proteins. Cell 2016, 165, 742–753. [Google Scholar] [CrossRef]
  135. Tong, H.; Huang, J.; Xiao, Q.; He, B.; Dong, X.; Liu, Y.; Yang, X.; Han, D.; Wang, Z.; Wang, X.; et al. High-fidelity Cas13 variants for targeted RNA degradation with minimal collateral effects. Nat. Biotechnol. 2022, 41, 108–119. [Google Scholar] [CrossRef]
  136. Rahman, R.; Xu, W.; Jin, H.; Rosbash, M. Identification of RNA-binding protein targets with HyperTRIBE. Nat. Protoc. 2018, 13, 1829–1849. [Google Scholar] [CrossRef] [PubMed]
  137. Nguyen, T.A.; Heng, J.W.J.; Kaewsapsak, P.; Kok, E.P.L.; Stanojević, D.; Liu, H.; Cardilla, A.; Praditya, A.; Yi, Z.; Lin, M.; et al. Direct identification of A-to-I editing sites with nanopore native RNA sequencing. Nat. Methods 2022, 19, 833–844. [Google Scholar] [CrossRef] [PubMed]
  138. Lin, B.-Y.; Shih, C.-J.; Hsieh, H.-Y.; Chen, H.-C.; Tu, S.-L. Phytochrome coordinates with a hnRNP to regulate alternative splicing via an exonic splicing silencer. Plant Physiol. 2019, 182, 243–254. [Google Scholar] [CrossRef]
  139. Xin, R.; Zhu, L.; Salomé, P.A.; Mancini, E.; Marshall, C.M.; Harmon, F.G.; Yanovsky, M.J.; Weigel, D.; Huq, E. SPF45-related splicing factor for phytochrome signaling promotes photomorphogenesis by regulating pre-mRNA splicing in Arabidopsis. Proc. Natl. Acad. Sci. USA 2017, 114, E7018–E7027. [Google Scholar] [CrossRef] [PubMed]
  140. Xin, R.; Kathare, P.K.; Huq, E. Coordinated regulation of pre-mRNA Splicing by the SFPS-RRC1 complex to promote photomorphogenesis. Plant Cell 2019, 31, 2052–2069. [Google Scholar] [CrossRef]
  141. Yan, Q.; Xia, X.; Sun, Z.; Fang, Y. Depletion of Arabidopsis SC35 and SC35-like serine/arginine-rich proteins affects the transcription and splicing of a subset of genes. PLOS Genet. 2017, 13, e1006663. [Google Scholar] [CrossRef] [PubMed]
  142. Albaqami, M.; Laluk, K.; Reddy, A.S.N. The Arabidopsis splicing regulator SR45 confers salt tolerance in a splice isoform-dependent manner. Plant Mol. Biol. 2019, 100, 379–390. [Google Scholar] [CrossRef]
  143. Zhang, X.-N.; Shi, Y.; Powers, J.J.; Gowda, N.B.; Zhang, C.; Ibrahim, H.M.M.; Ball, H.B.; Chen, S.L.; Lu, H.; Mount, S.M. Transcriptome analyses reveal SR45 to be a neutral splicing regulator and a suppressor of innate immunity in Arabidopsis thaliana. BMC Genom. 2017, 18, 172. [Google Scholar] [CrossRef]
  144. Huang, J.; Gu, L.; Zhang, Y.; Yan, T.; Kong, G.; Kong, L.; Guo, B.; Qiu, M.; Wang, Y.; Jing, M.; et al. An oomycete plant pathogen reprograms host pre-mRNA splicing to subvert immunity. Nat. Commun. 2017, 8, 2051. [Google Scholar] [CrossRef] [PubMed]
  145. Zhang, Y.; Huang, J.; Ochola, S.O.; Dong, S. Functional analysis of psavr3c effector family from phytophthora provides probes to dissect SKRP mediated plant susceptibility. Front. Plant Sci. 2018, 9, 1105. [Google Scholar] [CrossRef] [PubMed]
  146. Rigo, R.; Bazin, J.; Crespi, M.; Charon, C. Alternative Splicing in the Regulation of Plant–Microbe Interactions. Plant Cell Physiol. 2019, 60, 1906–1916. [Google Scholar] [CrossRef] [PubMed]
  147. Fang, Y.; Hearn, S.; Spector, D.L. Tissue-specific expression and dynamic organization of SR splicing factors in Arabidopsis. Mol. Biol. Cell 2004, 15, 2664–2673. [Google Scholar] [CrossRef] [PubMed]
  148. Ali, G.S.; Golovkin, M.; Reddy, A.S.N. Nuclear localization and in vivo dynamics of a plant-specific serine/arginine-rich protein. Plant J. 2003, 36, 883–893. [Google Scholar] [CrossRef]
  149. Ali, G.S.; Prasad, K.V.S.K.; Hanumappa, M.; Reddy, A.S.N. Analyses of in vivo interaction and mobility of two spliceosomal proteins using FRAP and BiFC. PLoS ONE 2008, 3, e1953. [Google Scholar] [CrossRef]
  150. Ali, G.S.; Reddy, A.S.N. ATP, phosphorylation and transcription regulate the mobility of plant splicing factors. J. Cell Sci. 2006, 119, 3527–3538. [Google Scholar] [CrossRef]
  151. Ali, G.S.; Reddy, A.S.N. Spatiotemporal organization of pre-mRNA splicing proteins in plants. Poxviruses 2008, 326, 103–118. [Google Scholar] [CrossRef]
  152. Bazin, J.; Romero, N.; Rigo, R.; Charon, C.; Blein, T.; Ariel, F.; Crespi, M. Nuclear speckle RNA binding proteins remodel alternative splicing and the non-coding arabidopsis transcriptome to regulate a cross-talk between auxin and immune responses. Front. Plant Sci. 2018, 9, 1209. [Google Scholar] [CrossRef]
  153. Reddy, A.S.; Day, I.S.; Göhring, J.; Barta, A. Localization and dynamics of nuclear speckles in plants. Plant Physiol. 2011, 158, 67–77. [Google Scholar] [CrossRef]
  154. Lorković, Z.J.; Hilscher, J.; Barta, A. Co-localisation studies of Arabidopsis SR splicing factors reveal different types of speckles in plant cell nuclei. Exp. Cell Res. 2008, 314, 3175–3186. [Google Scholar] [CrossRef]
  155. Tillemans, V.; Leponce, I.; Rausin, G.; Dispa, L.; Motte, P. Insights into nuclear organization in plants as revealed by the dynamic distribution of Arabidopsis SR splicing factors. Plant Cell 2006, 18, 3218–3234. [Google Scholar] [CrossRef] [PubMed]
  156. Morton, M.; AlTamimi, N.; Butt, H.; Reddy, A.S.; Mahfouz, M. Serine/arginine-rich protein family of splicing regulators: New approaches to study splice isoform functions. Plant Sci. 2019, 283, 127–134. [Google Scholar] [CrossRef]
  157. Jha, A.; Gazzara, M.R.; Barash, Y. Integrative deep models for alternative splicing. Bioinformatics 2017, 33, i274–i282. [Google Scholar] [CrossRef]
  158. Tilgner, H.; Knowles, D.G.; Johnson, R.; Davis, C.A.; Chakrabortty, S.; Djebali, S.; Curado, J.; Snyder, M.; Gingeras, T.R.; Guigó, R. Deep sequencing of subcellular RNA fractions shows splicing to be predominantly co-transcriptional in the human genome but inefficient for lncRNAs. Genome Res. 2012, 22, 1616–1625. [Google Scholar] [CrossRef] [PubMed]
  159. Wang, X.; Hu, L.; Wang, X.; Li, N.; Xu, C.; Gong, L.; Liu, B. DNA methylation affects gene alternative splicing in plants: An example from rice. Mol. Plant 2015, 9, 305–307. [Google Scholar] [CrossRef]
  160. Wei, G.; Liu, K.; Shen, T.; Shi, J.; Liu, B.; Han, M.; Peng, M.; Fu, H.; Song, Y.; Zhu, J.; et al. Position-specific intron retention is mediated by the histone methyltransferase SDG725. BMC Biol. 2018, 16, 44. [Google Scholar] [CrossRef] [PubMed]
  161. Ullah, F.; Hamilton, M.; Reddy, A.S.; Ben-Hur, A. Exploring the relationship between intron retention and chromatin accessibility in plants. BMC Genom. 2018, 19, 21. [Google Scholar] [CrossRef]
  162. Naftelberg, S.; Schor, I.E.; Ast, G.; Kornblihtt, A.R. Regulation of alternative splicing through coupling with transcription and chromatin structure. Annu. Rev. Biochem. 2015, 84, 165–198. [Google Scholar] [CrossRef] [PubMed]
  163. Saldi, T.; Cortazar, M.A.; Sheridan, R.M.; Bentley, D.L. Coupling of RNA polymerase II transcription elongation with pre-mRNA splicing. J. Mol. Biol. 2016, 428, 2623–2635. [Google Scholar] [CrossRef] [PubMed]
  164. Ullah, F.; Jabeen, S.; Salton, M.; Reddy, A.S.N.; Ben-Hur, A. Evidence for the role of transcription factors in the co-transcriptional regulation of intron retention. Genome Biol. 2023, 24, 53. [Google Scholar] [CrossRef] [PubMed]
  165. Godoy Herz, M.A.G.; Kubaczka, M.G.; Brzyżek, G.; Servi, L.; Krzyszton, M.; Simpson, C.; Brown, J.; Swiezewski, S.; Petrillo, E.; Kornblihtt, A.R. Light regulates plant alternative splicing through the control of transcriptional elongation. Mol. Cell 2019, 73, 1066–1074.e3. [Google Scholar] [CrossRef]
  166. Leng, X.; Ivanov, M.; Kindgren, P.; Malik, I.; Thieffry, A.; Brodersen, P.; Sandelin, A.; Kaplan, C.D.; Marquardt, S. Organismal benefits of transcription speed control at gene boundaries. Embo Rep. 2020, 21, e49315. [Google Scholar] [CrossRef]
  167. Pajoro, A.; Severing, E.; Angenent, G.C.; Immink, R.G.H. Histone H3 lysine 36 methylation affects temperature-induced alternative splicing and flowering in plants. Genome Biol. 2017, 18, 102. [Google Scholar] [CrossRef]
  168. Kindgren, P.; Ivanov, M.; Marquardt, S. Native elongation transcript sequencing reveals temperature dependent dynamics of nascent RNAPII transcription in Arabidopsis. Nucleic Acids Res. 2019, 48, 2332–2347. [Google Scholar] [CrossRef]
  169. Yu, X.; Meng, X.; Liu, Y.; Wang, X.; Wang, T.-J.; Zhang, A.; Li, N.; Qi, X.; Liu, B.; Xu, Z.-Y. The chromatin remodeler ZmCHB101 impacts alternative splicing contexts in response to osmotic stress. Plant Cell Rep. 2018, 38, 131–145. [Google Scholar] [CrossRef]
  170. Wu, F.; Deng, L.; Zhai, Q.; Zhao, J.; Chen, Q.; Li, C. Mediator subunit MED25 couples alternative splicing of JAZ genes with fine-tuning of jasmonate signaling. Plant Cell 2019, 32, 429–448. [Google Scholar] [CrossRef]
  171. Wang, S.; Quan, L.; Li, S.; You, C.; Zhang, Y.; Gao, L.; Zeng, L.; Liu, L.; Qi, Y.; Mo, B.; et al. The PROTEIN PHOSPHATASE4 complex promotes transcription and processing of primary microRNAs in Arabidopsis. Plant Cell 2019, 31, 486–501. [Google Scholar] [CrossRef] [PubMed]
  172. Reddy, A.S.; Ali, G.S.; Celesnik, H.; Day, I.S. Coping with stresses: Roles of calcium- and calcium/calmodulin-regulated gene expression. Plant Cell 2011, 23, 2010–2032. [Google Scholar] [CrossRef]
  173. Peck, S.; Mittler, R. Plant signaling in biotic and abiotic stress. J. Exp. Bot. 2020, 71, 1649–1651. [Google Scholar] [CrossRef]
  174. Dong, J.; Chen, H.; Deng, X.W.; Irish, V.F.; Wei, N. Phytochrome B induces intron retention and translational inhibition of PHYTOCHROME-INTERACTING FACTOR3. Plant Physiol. 2019, 182, 159–166. [Google Scholar] [CrossRef] [PubMed]
  175. Wang, S.; Tian, L.; Liu, H.; Li, X.; Zhang, J.; Chen, X.; Jia, X.; Zheng, X.; Wu, S.; Chen, Y.; et al. Large-scale discovery of non-conventional peptides in maize and arabidopsis through an integrated peptidogenomic pipeline. Mol. Plant 2020, 13, 1078–1093. [Google Scholar] [CrossRef]
  176. Raxwal, V.K.; Simpson, C.G.; Gloggnitzer, J.; Entinze, J.C.; Guo, W.; Zhang, R.; Brown, J.W.; Riha, K. Nonsense-mediated RNA decay factor UPF1 Is critical for posttranscriptional and translational gene regulation in Arabidopsis. Plant Cell 2020, 32, 2725–2741. [Google Scholar] [CrossRef] [PubMed]
  177. Min, X.; Kasamias, T.; Wagner, M.; Ogunbayi, A.; Yu, F. Identification and Analysis of Alternative Splicing in Soybean Plants. In Proceedings of the 14th International Conference on Bioinformatics and Computational Biology, Online, 21–23 March 2022; pp. 1–9. [Google Scholar] [CrossRef]
  178. Panahi, B.; Abbaszadeh, B.; Taghizadeghan, M.; Ebrahimie, E. Genome-wide survey of alternative splicing in Sorghum bicolor. Physiol. Mol. Biol. Plants 2014, 20, 323–329. [Google Scholar] [CrossRef] [PubMed]
  179. Guo, B.; Dai, Y.; Chen, L.; Pan, Z.; Song, L. Genome-wide analysis of the soybean root transcriptome reveals the impact of nitrate on alternative splicing. G3 Genes|Genomes|Genet. 2021, 11, jkab162. [Google Scholar] [CrossRef]
  180. Shen, Y.; Zhou, Z.; Wang, Z.; Li, W.; Fang, C.; Wu, M.; Ma, Y.; Liu, T.; Kong, L.-A.; Peng, D.-L.; et al. Global dissection of alternative splicing in paleopolyploid soybean. Plant Cell 2014, 26, 996–1008. [Google Scholar] [CrossRef]
  181. Sun, G.; Yu, H.; Wang, P.; Lopez-Guerrero, M.; Mural, R.V.; Mizero, O.N.; Grzybowski, M.; Song, B.; van Dijk, K.; Schachtman, D.P.; et al. A role for heritable transcriptomic variation in maize adaptation to temperate environments. Genome Biol. 2023, 24, 55. [Google Scholar] [CrossRef]
  182. Ogungbayi, A.; Lee, J.; Vaghela, V.; Yu, F.; Min, X. Systematic collection and analysis of alternative splicing events in potato plants. J. Plant Sci. 2023, 11, 98–106. [Google Scholar]
  183. Clark, S.; Yu, F.; Gu, L.; Min, X.J. Expanding alternative splicing identification by integrating multiple sources of transcription data in tomato. Front. Plant Sci. 2019, 10, 689. [Google Scholar] [CrossRef] [PubMed]
  184. Yu, K.; Feng, M.; Yang, G.; Sun, L.; Qin, Z.; Cao, J.; Wen, J.; Li, H.; Zhou, Y.; Chen, X.; et al. Changes in alternative splicing in response to domestication and polyploidization in wheat. Plant Physiol. 2020, 184, 1955–1968. [Google Scholar] [CrossRef]
  185. Gao, P.; Quilichini, T.D.; Zhai, C.; Qin, L.; Nilsen, K.T.; Li, Q.; Sharpe, A.G.; Kochian, L.V.; Zou, J.; Reddy, A.S.; et al. Alternative splicing dynamics and evolutionary divergence during embryogenesis in wheat species. Plant Biotechnol. J. 2021, 19, 1624–1643. [Google Scholar] [CrossRef] [PubMed]
  186. Ranwez, V.; Serra, A.; Pot, D.; Chantret, N. Domestication reduces alternative splicing expression variations in sorghum. PLoS ONE 2017, 12, e0183454. [Google Scholar] [CrossRef] [PubMed]
  187. Chai, L.; Zhang, J.; Lu, K.; Li, H.; Wu, L.; Wan, H.; Zheng, B.; Cui, C.; Jiang, J.; Jiang, L. Identification of genomic regions associated with multi-silique trait in Brassica napus. BMC Genom. 2019, 20, 304. [Google Scholar] [CrossRef]
  188. Chai, L.; Zhang, J.; Li, H.; Zheng, B.; Jiang, J.; Cui, C.; Jiang, L. Investigation for a multi-silique trait in Brassica napus by alternative splicing analysis. PeerJ 2020, 8, e10135. [Google Scholar] [CrossRef]
  189. Chai, L.; Zhang, J.; Li, H.; Cui, C.; Jiang, J.; Zheng, B.; Wu, L.; Jiang, L. Investigation of thermomorphogenesis-related genes for a multi-silique trait in Brassica napus by comparative transcriptome analysis. Front. Genet. 2021, 12, 678804. [Google Scholar] [CrossRef] [PubMed]
  190. Wu, B.; Zhang, X.; Hu, K.; Zheng, H.; Zhang, S.; Liu, X.; Ma, M.; Zhao, H. Two alternative splicing variants of a wheat gene TaNAK1, TaNAK1.1 and TaNAK1.2, differentially regulate flowering time and plant architecture leading to differences in seed yield of transgenic Arabidopsis. Front. Plant Sci. 2022, 13, 1014176. [Google Scholar] [CrossRef]
  191. Ren, X.; Zhi, L.; Liu, L.; Meng, D.; Su, Q.; Batool, A.; Ji, J.; Song, L.; Zhang, N.; Guo, L.; et al. Alternative splicing of TaGS3 differentially regulates grain weight and size in bread wheat. Int. J. Mol. Sci. 2021, 22, 11692. [Google Scholar] [CrossRef]
  192. Liu, L.; Zhou, Y.; Mao, F.; Gu, Y.; Tang, Z.; Xin, Y.; Liu, F.; Tang, T.; Gao, H.; Zhao, X. Fine-tuning of the grain size by alternative splicing of GS3 in rice. Rice 2022, 15, 1. [Google Scholar] [CrossRef]
  193. Teng, Z.; Zheng, Q.; Liu, B.; Meng, S.; Zhang, J.; Ye, N. Moderate soil drying-induced alternative splicing provides a potential novel approach for the regulation of grain filling in rice inferior spikelets. Int. J. Mol. Sci. 2022, 23, 7770. [Google Scholar] [CrossRef] [PubMed]
  194. Liu, X.; Tian, Y.; Chi, W.; Zhang, H.; Yu, J.; Chen, G.; Wu, W.; Jiang, X.; Wang, S.; Lin, Z.; et al. Alternative splicing of OsGS1;1 affects nitrogen-use efficiency, grain development, and amylose content in rice. Plant J. 2022, 110, 1751–1762. [Google Scholar] [CrossRef] [PubMed]
  195. Yu, J.; Miao, J.; Zhang, Z.; Xiong, H.; Zhu, X.; Sun, X.; Pan, Y.; Liang, Y.; Zhang, Q.; Rehman, R.M.A.; et al. Alternative splicing of OsLG3b controls grain length and yield in japonica rice. Plant Biotechnol. J. 2018, 16, 1667–1678. [Google Scholar] [CrossRef]
  196. Lin, A.; Ma, J.; Xu, F.; Xu, W.; Jiang, H.; Zhang, H.; Qu, C.; Wei, L.; Li, J. Differences in alternative splicing between yellow and black-seeded rapeseed. Plants 2020, 9, 977. [Google Scholar] [CrossRef] [PubMed]
  197. Qin, D.; Nishida, S.; Tominaga, R.; Ueda, A.; Raboy, V.; Saneoka, H. Aberrant RNA splicing of the phytic acid synthesis gene inositol-1,3,4 trisphosphate 5/6-kinase in a low phytic acid soybean line. Soil Sci. Plant Nutr. 2022, 68, 553–562. [Google Scholar] [CrossRef]
  198. Dwivedi, S.L.; Mattoo, A.K.; Garg, M.; Dutt, S.; Singh, B.; Ortiz, R. Developing germplasm and promoting consumption of anthocyanin-rich grains for health benefits. Front. Sustain. Food Syst. 2022, 6, 867897. [Google Scholar] [CrossRef]
  199. Chen, D.; Liu, Y.; Yin, S.; Qiu, J.; Jin, Q.; King, G.J.; Wang, J.; Ge, X.; Li, Z. Alternatively spliced BnaPAP2.A7 isoforms play opposing roles in anthocyanin biosynthesis of Brassica napus L. Front. Plant Sci. 2020, 11, 983. [Google Scholar] [CrossRef]
  200. Ma, Z.; Li, M.; Zhang, H.; Zhao, B.; Liu, Z.; Duan, S.; Meng, X.; Li, G.; Guo, X. Alternative splicing of TaHsfA2-7 is involved in the improvement of thermotolerance in wheat. Int. J. Mol. Sci. 2023, 24, 1014. [Google Scholar] [CrossRef]
  201. Muthusamy, M.; Yoon, E.K.; Kim, J.; Jeong, M.-J.; Lee, S.I. Brassica rapa SR45a regulates drought tolerance via the alternative splicing of target genes. Genes 2020, 11, 182. [Google Scholar] [CrossRef]
  202. Weng, X.; Zhou, X.; Xie, S.; Gu, J.; Wang, Z.-Y. Identification of cassava alternative splicing-related genes and functional characterization of MeSCL30 involvement in drought stress. Plant Physiol. Biochem. 2021, 160, 130–142. [Google Scholar] [CrossRef] [PubMed]
  203. Thatcher, S.R.; Danilevskaya, O.N.; Meng, X.; Beatty, M.; Zastrow-Hayes, G.; Harris, C.; Van Allen, B.; Habben, J.; Li, B. Genome-wide analysis of alternative splicing during development and drought stress in maize. Plant Physiol. 2015, 170, 586–599. [Google Scholar] [CrossRef]
  204. Zhang, J.; Liang, Y.; Zhang, S.; Xu, Q.; Di, H.; Zhang, L.; Dong, L.; Hu, X.; Zeng, X.; Liu, X.; et al. Global landscape of alternative splicing in maize response to low temperature. J. Agric. Food Chem. 2022, 70, 15715–15725. [Google Scholar] [CrossRef] [PubMed]
  205. Butt, H.; Bazin, J.; Prasad, K.V.S.K.; Awad, N.; Crespi, M.; Reddy, A.S.N.; Mahfouz, M.M. The rice serine/arginine splicing factor RS33 regulates pre-mRNA splicing during abiotic stress responses. Cells 2022, 11, 1796. [Google Scholar] [CrossRef] [PubMed]
  206. Chen, S.; Mo, Y.; Zhang, Y.; Zhu, H.; Ling, Y. Insights into sweet potato SR proteins: From evolution to species-specific expression and alternative splicing. Planta 2022, 256, 72. [Google Scholar] [CrossRef]
  207. Yang, L.; Yang, L.; Zhao, C.; Liu, J.; Tong, C.; Zhang, Y.; Cheng, X.; Jiang, H.; Shen, J.; Xie, M.; et al. Differential alternative splicing genes and isoform co-expression networks of Brassica napus under multiple abiotic stresses. Front. Plant Sci. 2022, 13, 1009998. [Google Scholar] [CrossRef]
  208. Song, L.; Pan, Z.; Chen, L.; Dai, Y.; Wan, J.; Ye, H.; Nguyen, H.T.; Zhang, G.; Chen, H. Analysis of whole transcriptome RNA-seq data reveals many alternative splicing events in soybean roots under drought stress conditions. Genes 2020, 11, 1520. [Google Scholar] [CrossRef]
  209. Yu, H.; Du, Q.; Campbell, M.; Yu, B.; Walia, H.; Zhang, C. Genome-wide discovery of natural variation in pre-mRNA splicing and prioritising causal alternative splicing to salt stress response in rice. New Phytol. 2021, 230, 1273–1287. [Google Scholar] [CrossRef]
  210. Liu, Z.; Qin, J.; Tian, X.; Xu, S.; Wang, Y.; Li, H.; Wang, X.; Peng, H.; Yao, Y.; Hu, Z.; et al. Global profiling of alternative splicing landscape responsive to drought, heat and their combination in wheat (Triticum aestivum L.). Plant Biotechnol. J. 2017, 16, 714–726. [Google Scholar] [CrossRef]
  211. Cao, H.; Wu, T.; Shi, L.; Yang, L.; Zhang, C. Alternative splicing control of light and temperature stress responses and its prospects in vegetable crops. Veg. Res. 2023, 3, 17. [Google Scholar] [CrossRef]
  212. Lee, H.J.; Eom, S.H.; Lee, J.H.; Wi, S.H.; Kim, S.K.; Hyun, T.K. Genome-wide analysis of alternative splicing events during response to drought stress in tomato (Solanum lycopersicum L.). J. Hortic. Sci. Biotechnol. 2019, 95, 286–293. [Google Scholar] [CrossRef]
  213. Ruggiero, A.; Punzo, P.; Van Oosten, M.J.; Cirillo, V.; Esposito, S.; Costa, A.; Maggio, A.; Grillo, S.; Batelli, G. Transcriptomic and splicing changes underlying tomato responses to combined water and nutrient stress. Front. Plant Sci. 2022, 13, 974048. [Google Scholar] [CrossRef] [PubMed]
  214. Hu, X.; Wang, H.; Li, K.; Liu, X.; Liu, Z.; Wu, Y.; Li, S.; Huang, C. Genome-wide alternative splicing variation and its potential contribution to maize immature-ear heterosis. Crop. J. 2020, 9, 476–486. [Google Scholar] [CrossRef]
  215. Zhang, M.; Li, N.; Yang, W.; Liu, B. Genome-wide differences in gene expression and alternative splicing in developing embryo and endosperm, and between F1 hybrids and their parental pure lines in sorghum. Plant Mol. Biol. 2021, 108, 1–14. [Google Scholar] [CrossRef]
  216. Lee, J.S.; Jahani, M.; Huang, K.; Mandel, J.R.; Marek, L.F.; Burke, J.M.; Langlade, N.B.; Owens, G.L.; Rieseberg, L.H. Expression complementation of gene presence/absence polymorphisms in hybrids contributes importantly to heterosis in sunflower. J. Adv. Res. 2022, 42, 83–98. [Google Scholar] [CrossRef]
  217. Chen, M.; Mei, L.; Wang, F.; Dewayalage, I.K.W.B.; Yang, J.; Dai, L.; Yang, G.; Gao, B.; Cheng, C.; Liu, Y.; et al. PlantSPEAD: A web resource towards comparatively analysing stress-responsive expression of splicing-related proteins in plant. Plant Biotechnol. J. 2020, 19, 227–229. [Google Scholar] [CrossRef] [PubMed]
  218. Liu, J.; Lang, K.; Tan, S.; Jie, W.; Zhu, Y.; Huang, S.; Huang, W. A web-based database server using 43,710 public RNA-seq samples for the analysis of gene expression and alternative splicing in livestock animals. BMC Genom. 2022, 23, 702. [Google Scholar] [CrossRef]
  219. Liu, J.; Zhang, Y.; Zheng, Y.; Zhu, Y.; Shi, Y.; Guan, Z.; Lang, K.; Shen, D.; Huang, W.; Dou, D. PlantExp: A platform for exploration of gene expression and alternative splicing based on public plant RNA-seq samples. Nucleic Acids Res. 2022, 51, D1483–D1491. [Google Scholar] [CrossRef] [PubMed]
  220. Liu, J.; Zhang, Y.; Shi, Y.; Zheng, Y.; Zhu, Y.; Guan, Z.; Shen, D.; Dou, D. FungiExp: A user-friendly database and analysis platform for exploring fungal gene expression and alternative splicing. Bioinformatics 2023, 39, btad042. [Google Scholar] [CrossRef] [PubMed]
  221. Tan, S.; Wang, W.; Jie, W.; Liu, J. FishExp: A comprehensive database and analysis platform for gene expression and alternative splicing of fish species. Comput. Struct. Biotechnol. J. 2022, 20, 3676–3684. [Google Scholar] [CrossRef]
  222. Liu, J.; Yin, F.; Lang, K.; Jie, W.; Tan, S.; Duan, R.; Huang, S.; Huang, W. MetazExp: A database for gene expression and alternative splicing profiles and their analyses based on 53 615 public RNA-seq samples in 72 metazoan species. Nucleic Acids Res. 2021, 50, D1046–D1054. [Google Scholar] [CrossRef]
  223. James, A.B.; Syed, N.H.; Bordage, S.; Marshall, J.; Nimmo, G.A.; Jenkins, G.I.; Herzyk, P.; Brown, J.W.; Nimmo, H.G. Alternative splicing mediates responses of the Arabidopsis circadian clock to temperature changes. Plant Cell 2012, 24, 961–981. [Google Scholar] [CrossRef]
  224. 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]
  225. Kwon, Y.-J.; Park, M.-J.; Kim, S.-G.; Baldwin, I.T.; Park, C.-M. Alternative splicing and nonsense-mediated decay of circadian clock genes under environmental stress conditions in Arabidopsis. BMC Plant Biol. 2014, 14, 136. [Google Scholar] [CrossRef] [PubMed]
  226. Calixto, C.P.G.; Simpson, C.G.; Waugh, R.; Brown, J.W.S. Alternative splicing of barley clock genes in response to low temperature. PLoS ONE 2016, 11, e0168028. [Google Scholar] [CrossRef] [PubMed]
  227. Dantas, L.L.B.; Calixto, C.P.G.; Dourado, M.M.; Carneiro, M.S.; Brown, J.W.S.; Hotta, C.T. Alternative splicing of circadian clock genes correlates with temperature in field-grown sugarcane. Front. Plant Sci. 2019, 10, 1614. [Google Scholar] [CrossRef] [PubMed]
  228. Lu, S.X.; Knowles, S.M.; Andronis, C.; Ong, M.S.; Tobin, E.M. CIRCADIAN CLOCK ASSOCIATED1 and LATE ELONGATED HYPOCOTYL function synergistically in the circadian clock of Arabidopsis. Plant Physiol. 2009, 150, 834–843. [Google Scholar] [CrossRef]
  229. Seo, P.J.; Park, M.-J.; Lim, M.-H.; Kim, S.-G.; Lee, M.; Baldwin, I.T.; Park, C.-M. A self-regulatory circuit of CIRCADIAN CLOCK-ASSOCIATED1 underlies the circadian clock regulation of temperature responses in Arabidopsis. Plant Cell 2012, 24, 2427–2442. [Google Scholar] [CrossRef]
  230. 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]
  231. Filichkin, S.A.; Cumbie, J.S.; Dharmawardhana, P.; Jaiswal, P.; Chang, J.H.; Palusa, S.G.; Reddy, A.; 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]
  232. 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]
  233. Jones, M.A.; Williams, B.A.; McNicol, J.; Simpson, C.G.; Brown, J.W.; Harmer, S.L. Mutation of Arabidopsis SPLICEOSOMAL TIMEKEEPER LOCUS1 causes circadian clock defects. Plant Cell 2012, 24, 4066–4082. [Google Scholar] [CrossRef] [PubMed]
  234. 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]
  235. Schlaen, R.G.; Mancini, E.; Sanchez, S.E.; Perez-Santángelo, S.; Rugnone, M.L.; Simpson, C.G.; Brown, J.W.S.; Zhang, X.; Chernomoretz, A.; Yanovsky, M.J. The spliceosome assembly factor GEMIN2 attenuates the effects of temperature on alternative splicing and circadian rhythms. Proc. Natl. Acad. Sci. USA 2015, 112, 9382–9387. [Google Scholar] [CrossRef]
  236. Backer, R.; Rokem, J.S.; Ilangumaran, G.; Lamont, J.; Praslickova, D.; Ricci, E.; Subramanian, S.; Smith, D.L. Plant growth-promoting rhizobacteria: Context, mechanisms of action, and roadmap to commercialization of biostimulants for sustainable agriculture. Front. Plant Sci. 2018, 9, 1473. [Google Scholar] [CrossRef]
  237. Begum, N.; Qin, C.; Ahanger, M.A.; Raza, S.; Khan, M.I.; Ashraf, M.; Ahmed, N.; Zhang, L. Role of arbuscular mycorrhizal fungi in plant growth regulation: Implications in abiotic stress tolerance. Front. Plant Sci. 2019, 10, 1068. [Google Scholar] [CrossRef] [PubMed]
  238. Zeng, Z.; Liu, Y.; Feng, X.-Y.; Li, S.-X.; Jiang, X.-M.; Chen, J.-Q.; Shao, Z.-Q. The RNAome landscape of tomato during arbuscular mycorrhizal symbiosis reveals an evolving RNA layer symbiotic regulatory network. Plant Commun. 2023, 4, 100429. [Google Scholar] [CrossRef]
  239. Zorin, E.A.; Afonin, A.M.; Kulaeva, O.A.; Gribchenko, E.S.; Shtark, O.Y.; Zhukov, V.A. transcriptome analysis of alternative splicing events induced by arbuscular mycorrhizal fungi (Rhizophagus irregularis) in pea (Pisum sativum L.) roots. Plants 2020, 9, 1700. [Google Scholar] [CrossRef]
  240. Liu, J.; Chen, S.; Liu, M.; Chen, Y.; Fan, W.; Lee, S.; Xiao, H.; Kudrna, D.; Li, Z.; Chen, X.; et al. Full-length transcriptome sequencing reveals alternative splicing and lncrna regulation during nodule development in Glycine max. Int. J. Mol. Sci. 2022, 23, 7371. [Google Scholar] [CrossRef] [PubMed]
  241. Zorin, E.A.; Kulaeva, O.A.; Afonin, A.M.; Zhukov, V.A.; Tikhonovich, I.A. Analysis of alternative splicing events in the root tips and nodules of Pisum sativum L. Ecol. Genet. 2019, 17, 53–63. [Google Scholar] [CrossRef]
  242. Muhammad, S.; Xu, X.; Zhou, W.; Wu, L. Alternative splicing: An efficient regulatory approach towards plant developmental plasticity. Wiley Interdiscip. Rev. RNA 2022, 14, e1758. [Google Scholar] [CrossRef] [PubMed]
  243. Lam, P.Y.; Wang, L.; Lo, C.; Zhu, F.-Y. Alternative splicing and its roles in plant metabolism. Int. J. Mol. Sci. 2022, 23, 7355. [Google Scholar] [CrossRef] [PubMed]
  244. Shang, X.; Cao, Y.; Ma, L. Alternative splicing in plant genes: A means of regulating the environmental fitness of plants. Int. J. Mol. Sci. 2017, 18, 432. [Google Scholar] [CrossRef]
  245. Jeon, J.; Kim, K.-T.; Choi, J.; Cheong, K.; Ko, J.; Choi, G.; Lee, H.; Lee, G.-W.; Park, S.-Y.; Kim, S.; et al. Alternative splicing diversifies the transcriptome and proteome of the rice blast fungus during host infection. RNA Biol. 2022, 19, 373–386. [Google Scholar] [CrossRef] [PubMed]
  246. Kim, S.; Kim, T.-H. Alternative splicing for improving abiotic stress tolerance and agronomic traits in crop plants. J. Plant Biol. 2020, 63, 409–420. [Google Scholar] [CrossRef]
  247. Hirsz, D.; Dixon, L.E. The roles of temperature-related post-transcriptional regulation in cereal floral development. Plants 2021, 10, 2230. [Google Scholar] [CrossRef]
  248. Jiang, L.; Ma, X.; Zhao, S.; Tang, Y.; Liu, F.; Gu, P.; Fu, Y.; Zhu, Z.; Cai, H.; Sun, C.; et al. The APETALA2-like transcription factor SUPERNUMERARY BRACT controls rice seed shattering and seed size. Plant Cell 2019, 31, 17–36. [Google Scholar] [CrossRef] [PubMed]
  249. Su, Z.; Hao, C.; Wang, L.; Dong, Y.; Zhang, X. Identification and development of a functional marker of TaGW2 associated with grain weight in bread wheat (Triticum aestivum L.). Theor. Appl. Genet. 2010, 122, 211–223. [Google Scholar] [CrossRef] [PubMed]
  250. Roy, N.S.; Basnet, P.; Ramekar, R.V.; Um, T.; Yu, J.-K.; Park, K.-C.; Choi, I.-Y. Alternative splicing (as) dynamics in dwarf soybean derived from cross of Glycine max and Glycine soja. Agronomy 2022, 12, 1685. [Google Scholar] [CrossRef]
  251. Liu, J.; Wu, X.; Yao, X.; Yu, R.; Larkin, P.J.; Liu, C.-M. Mutations in the DNA demethylase OsROS1 result in a thickened aleurone and improved nutritional value in rice grains. Proc. Natl. Acad. Sci. USA 2018, 115, 11327–11332. [Google Scholar] [CrossRef]
  252. Román, Á.; Andreu, V.; Hernández, M.L.; Lagunas, B.; Picorel, R.; Martínez-Rivas, J.M.; Alfonso, M. Contribution of the different omega-3 fatty acid desaturase genes to the cold response in soybean. J. Exp. Bot. 2012, 63, 4973–4982. [Google Scholar] [CrossRef]
  253. Yuan, F.-J.; Zhu, D.-H.; Tan, Y.-Y.; Dong, D.-K.; Fu, X.-J.; Zhu, S.-L.; Li, B.-Q.; Shu, Q.-Y. Identification and characterization of the soybean IPK1 ortholog of a low phytic acid mutant reveals an exon-excluding splice-site mutation. Theor. Appl. Genet. 2012, 125, 1413–1423. [Google Scholar] [CrossRef] [PubMed]
  254. Chen, L.; Li, W.; Li, Y.; Feng, X.; Du, K.; Wang, G.; Zhao, L. Identified trans-splicing of YELLOW-FRUITED TOMATO 2 encoding the PHYTOENE SYNTHASE 1 protein alters fruit color by map-based cloning, functional complementation and RACE. Plant Mol. Biol. 2019, 100, 647–658. [Google Scholar] [CrossRef] [PubMed]
  255. Luo, M.; Ding, J.; Li, Y.; Tang, H.; Qi, P.; Ma, J.; Wang, J.; Chen, G.; Pu, Z.; Li, W.; et al. A single-base change at a splice site in Wx-A1 caused incorrect RNA splicing and gene inactivation in a wheat EMS mutant line. Theor. Appl. Genet. 2019, 132, 2097–2109. [Google Scholar] [CrossRef] [PubMed]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Dwivedi, S.L.; Quiroz, L.F.; Reddy, A.S.N.; Spillane, C.; Ortiz, R. Alternative Splicing Variation: Accessing and Exploiting in Crop Improvement Programs. Int. J. Mol. Sci. 2023, 24, 15205. https://doi.org/10.3390/ijms242015205

AMA Style

Dwivedi SL, Quiroz LF, Reddy ASN, Spillane C, Ortiz R. Alternative Splicing Variation: Accessing and Exploiting in Crop Improvement Programs. International Journal of Molecular Sciences. 2023; 24(20):15205. https://doi.org/10.3390/ijms242015205

Chicago/Turabian Style

Dwivedi, Sangam L., Luis Felipe Quiroz, Anireddy S. N. Reddy, Charles Spillane, and Rodomiro Ortiz. 2023. "Alternative Splicing Variation: Accessing and Exploiting in Crop Improvement Programs" International Journal of Molecular Sciences 24, no. 20: 15205. https://doi.org/10.3390/ijms242015205

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop