Next Article in Journal
Novel Insights on the Corpus Luteum Function: Role of Vaspin on Porcine Luteal Cell Angiogenesis, Proliferation and Apoptosis by Activation of GRP78 Receptor and MAP3/1 Kinase Pathways
Next Article in Special Issue
Temperature Regulation of Primary and Secondary Seed Dormancy in Rosa canina L.: Findings from Proteomic Analysis
Previous Article in Journal
A Highly Conserved Iron-Sulfur Cluster Assembly Machinery between Humans and Amoeba Dictyostelium discoideum: The Characterization of Frataxin
Previous Article in Special Issue
Seed Germination in Oil Palm (Elaeis guineensis Jacq.): A Review of Metabolic Pathways and Control Mechanisms
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Emerging Roles of RNA-Binding Proteins in Seed Development and Performance

Shenzhen Branch, Guangdong Laboratory for Lingnan Modern Agriculture, Genome Analysis Laboratory of the Ministry of Agriculture, Agricultural Genomics Institute at Shenzhen, Chinese Academy of Agricultural Sciences, Shenzhen 518120, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2020, 21(18), 6822; https://doi.org/10.3390/ijms21186822
Submission received: 7 August 2020 / Revised: 10 September 2020 / Accepted: 10 September 2020 / Published: 17 September 2020

Abstract

:
Seed development, dormancy, and germination are key physiological events that are not only important for seed generation, survival, and dispersal, but also contribute to agricultural production. RNA-binding proteins (RBPs) directly interact with target mRNAs and fine-tune mRNA metabolism by governing post-transcriptional regulation, including RNA processing, intron splicing, nuclear export, trafficking, stability/decay, and translational control. Recent studies have functionally characterized increasing numbers of diverse RBPs and shown that they participate in seed development and performance, providing significant insight into the role of RBP–mRNA interactions in seed processes. In this review, we discuss recent research progress on newly defined RBPs that have crucial roles in RNA metabolism and affect seed development, dormancy, and germination.

1. Introduction

RNA-binding proteins (RBPs) centrally regulate all aspects of RNA metabolism from synthesis to decay [1]. The functional characterization of RBPs and the resolution of their protein structures indicate that a variety of RNA-binding domains (RBDs) mediate direct interactions between RBPs and their target mRNAs. RBDs include the RNA recognition motif (RRM, also known as RBD or ribonucleoprotein (RNP) domain), the hnRNP K homology (KH) domain, the zinc finger motif, the pentatricopeptide repeat (PPR) motif, Asp-Glu-Ala-Asp (DEAD) boxes, Pumilio/FBF (PUF) domains, and the double-stranded RNA binding domain (dsRBD) [2,3,4,5]. Based on these RBDs, bioinformatic analyses have identified more than 800, about 250 putative, and 336 non-redundant RBP genes in the Arabidopsis, rice, and maize genomes, respectively. Many appear to be plant specific, highlighting their functional diversity and specificity [4,6,7]. Moreover, recent studies have used mRNA–protein interactome capture technology to discover many novel RBPs with non-canonical RNA-binding structures, providing unprecedented insight into the complexity of RNA metabolism in plants [8,9,10].
Transcriptome data have been used to identify more than 12,000 and 17,000 mRNAs in the mature dry seeds of Arabidopsis and rice, respectively [11,12]. These stored mRNAs have a pivotal role in the early stages of seed germination, as many proteins that function in germination are initially translated using stored mRNAs as templates. Germination would be halted if translation were inhibited. Thus, the fate of seed-stored mRNA is important for seed dormancy and germination [13]. Proteomic analysis of mature dry rice seeds has demonstrated that three different types of RBPs participate in seed germination: the KH domain-containing protein NOVA-like, putative RBP RZ-1A, and GR-RBP 1A. GRP1A and RZ-1A significantly increased during seed maturation and desiccation. Their encoded proteins potentially act as molecular chaperones to stabilize seed stored-mRNA and maintain its original integrity and translatability during seed dormancy and under various natural conditions [14,15]. Recent studies have also shown that several different RBPs, including PUF, dsRBD, CSP GRP, RZ, and PPR, have essential functions in seed dormancy and germination [16,17]. Our review therefore focuses on recent advances in understanding the emerging roles of RBPs in seed biology and emphasizes seed development, dormancy, and germination.

2. Structural Characteristics of RBPs

The RRM was the first conserved domain to be identified in RBPs and appears to be common to all eukaryotes [4]. In addition to the RRM, typical RBPs also harbor auxiliary domains rich in glycine, arginine, and serine. These are arranged in a variety of modes and determine the target specificity of individual RBPs [18]. The glycine-rich (GR) RNA-binding proteins (GRPs) are particularly well characterized and are divided into four subclasses (Figure 1). Class I and IV members contain one or two RRMs and a glycine-rich region at the N or C terminus, respectively. Class II members have one CCHC-type zinc finger inserted into the glycine-rich region of Class I; they are also named RZs because they are structural orthologues of RZ-1, a protein identified with a 60S RNP complex from tobacco nuclei [19]. Class III members are referred to as cold shock proteins (CSPs). They contain a cold shock domain (CSD) instead of an RRM, as well as two or more zinc fingers scattered within the glycine-rich region [20]. The CSD binds with RNA and with single- and double-stranded DNA in eukaryotes [21,22]. Protein structural investigations have revealed that, in addition to binding RNA, RRM can also mediate protein–protein interactions, through both of which heterogeneous RNP complexes are formed [23].
The dsRBD is the second most widespread RNA recognition motif and usually comprises 65 to 70 amino acids that form an αβββα fold [24]. How dsRBDs recognize their dsRNA cellular targets is not yet well understood (see review [25] for more details). PUF RBPs are highly conserved across eukaryotes and characterized by Puf domains (also called PUM-HDs; Pumilio-homology domains), whose name derives from the two founding members, Pumilio from Drosophila melanogaster and FBF (Fem-3 mRNA-binding factor) from Caenorhabditis elegans [26] (Figure 1). The canonical PUF RBP binds to a conserved UGURN1-3AU(A⁄U) motif in the 3′-untranslated region (UTR) of the target RNA, mediated by its eight sequential tandem PUF repeats. Each PUF repeat comprises approximately 36 amino acids and specifically recognizes one RNA nucleotide [27,28,29,30,31].
PPR proteins are composed of 2–30 repeats of approximately 35 amino acids each. They are encoded by the nuclear genome but predominantly localized to chloroplasts and/or mitochondria, where they regulate organelle-related RNA metabolism [32] (Figure 1). Based on variations in PPR unit length, PPR proteins are divided into two major types: the P class (P-motif: 35 amino acids long) and the PLS class (PLS-triplets: P-motif, with longer (L) and shorter (S) variants). With the exception of RNA editing, P-class proteins affect nearly all steps of post-transcriptional gene regulation. PLS-class proteins display C-to-U RNA editing ability, which mainly depends on the class-specific extended C-terminal domain. Also, the E or DWY domain in PLS-class proteins may contribute to their RNA association and to target specificity or editing [33].
The KH domain is another common structure in RBPs; it comprises approximately 70 amino acids and appears in multiple copies. KH domain-containing proteins primarily participate in the regulation of transcription and translation [34], and the KH domain itself has various roles, such as interacting with ssDNA or ssRNA and mediating protein–protein interactions [35]. DEAD box RNA helicases constitute a large family that contains 58 members and mainly functions in RNA metabolism and gene translation [36] (Figure 1).

3. PUF-Type RBPs in Seed Development and Performance

PUF RBPs have been reported to affect RNA decay, translational repression, pre-ribosomal RNA processing, and chromosome stability [31]. Some Arabidopsis PUF RBPs (APUMs) are emerging as regulatory factors in seed development and performance (Table 1 and Figure 2). APUM5 was originally identified as causing reduced susceptibility to cucumber mosaic virus (CMV) in a T-DNA mutant screen [37]. APUM5 was highly expressed in silique ends and was markedly upregulated after treatment with abscisic acid (ABA) and NaCl, suggesting that it may function in seeds in response to ABA and abiotic stress. Germination rate was hypersensitive to salt and osmotic stress in seeds that overexpressed APUM5, whereas APUM5-RNAi seeds showed increased tolerance of these stresses [38]. APUM5 was localized to the cytoplasm and shown to regulate abiotic stress response by binding to the 3′-UTRs of abiotic stress-responsive mRNAs and destabilizing them [39]. APUM23 and APUM24 have been reported to affect seed development. Together with APUM25, they are evolutionarily divergent and fail to group with other APUMs on the basis of their amino acid sequences and Puf domain locations [30]. They are localized to the nucleus and function in pre-ribosomal RNA processing. The apum24 mutant is deficient in seed setting, whereas the apum23 knock-out mutant develops normal embryos [16]. Their contrasting mutant phenotypes may arise from different RNA targets: an APUM24-containing complex interacted with INTERNAL TRANSCRIBED SPACER 2 (ITS2), whereas 18S rRNA was the target of APUM23 [40,41]. Additional pleiotropic defects were observed in the apum23 null mutant and the apum24 knock-down mutant, including significantly delayed germination and abnormal development of leaves and roots [40,42]. Notably, apum23 null mutant seeds germinated more slowly than wild-type seeds, regardless of whether or not they were stratified, suggesting the involvement of APUM23 in seed dormancy release and germination [42]. Moreover, dysfunctions in APUM23 also caused hypersensitivity to salt stress, with delayed germination and seedling establishment, possibly resulting from changes in the expression of genes involved in ABA biosynthesis and signaling [43].
Xiang et al. (2014) reported that members of another PUF subgroup, including APUM9, APUM10 and APUM11, may function together with REDUCED DORMANCY5 (RDO5) in the regulation of seed germination and dormancy, as the expression of all three genes was significantly up-regulated in the rdo5 mutant after imbibition. RNAi transgenic lines with simultaneous down-regulation of three APUMs exhibited enhanced dormancy compared with wild type. By contrast, the overexpression (OE) of APUM9 resulted in reduced seed dormancy. Together, these results suggest that APUM9 has a strong negative effect on seed dormancy [17]. Interestingly, a transcriptome comparison of imbibed seeds identified six transcripts that were strikingly down-regulated in APUM9-OE plants compared with wild type. Five (OZF2, ATAF1, FBS1, ABI2, and COL4) were involved in the ABA signaling pathway. Moreover, APUM9 bound to the 3′-UTR of FBS1 and COL4, subsequently destabilizing their neighboring GFP reporter gene in an N. benthamiana transient expression system. It therefore appears likely that APUM9 directly targets ABA signaling genes for degradation, thereby promoting the transition from dormancy to germination [44]. These APUMs function downstream of RDO5 and are required for the RDO5-mediated seed dormancy signaling pathway, as RNAi knockdown of APUM9, APUM10, and APUM11 in an rdo5-1 background completely rescued the reduced dormancy phenotype of rdo5-1 [17]. In the future, it will be exciting to uncover other APUM9 targets and to decipher the mechanism by which APUM9 functions with RDO5 in dormancy regulation.

4. DsRBD-Type RBPs in Seed Development and Performance

DsRBD-type RBPs are mainly involved in post-transcriptional regulatory events such as the editing, trafficking, and portioning of substrate mRNAs and the biogenesis of various small RNAs. Based on the phylogenetic analyses of dsRBD sequence, 16 dsRBD-type RBP have been identified [45]. Six have been identified and characterized genetically, including RNA polymerase II C-terminal domain phosphatase-like protein 1 (CPL1), CPL2, HYPONASTIC LEAVES1 (HYL1), HUA ENHANCER1 (HEN1), Dicer-like protein (DCL1), DCL2 and DCL3. Abundant evidence confirms their involvement in seed development, germination, and dormancy regulation (Table 1 and Figure 2). In Arabidopsis, CPL1 and CPL2 are nuclear proteins that function as transcriptional repressors, which harbor the dsRBD at their C terminus and an RNA polymerase II C-terminal domain (CTD) phosphatase domain (CPD) at their N terminus [46]. The cpl1 null and dsRBD defective mutants exhibit strong tolerance to ABA and NaCl stress during seed germination, indicating the requirement for CPL1, and especially the dsRBD, in typical seed germination behavior [47]. CPL1 interacts with the KH-domain protein SHINY1 (SHI1), which is also identified and named as HIGH OSMOTIC STRESS GENE EXPRESSION 5 (HOS5) or REGULATOR OF GENE EXPRESSION 3 (RCF3) through different screen strategies [48,49,50]. CPL1-HOS5(RCF3/SHI1) subsequently forms a regulatory complex by recruiting splicing factors serine/arginine-rich protein 40 (RS40) and RS41. This complex suppresses abiotic stress-responsive C-repeat binding factor (CBF)/dehydration-responsive element binding (DREB) transcription factors by inhibiting mRNA capping, thereby preventing the transition from transcription initiation to elongation [48]. By contrast to the cpl1 mutant, the cpl2 mutant with a defect in the dsRBD is hypersensitive to high salinity and shows a lower seed germination rate, suggesting that CPL1 and CPL2 have distinct functions in seed behavior under stress [51]. Interestingly, the KH domain-containing RBP PEPPER (PEP) was revealed to be another CPL1 interactor and shown to control ovule identity, the precursor of seeds [52]. Moreover, CPL1 has been shown to affect seed germination by positively regulating the nonsense-mediated decay (NMD) pathway. UPF1 is one of three key NMD factors, and the UPF1 null mutant displays elongated seeds, hypersensitivity of germination to ABA and glucose, and tolerance of germination to mannose, underscoring the significance of NMD in the control of seed germination [53]. CPL1 interacts with and dephosphorylates the DEAD box RNA helicase eIF4AIII, a core component of the exon junction complex (EJC) that recruits UPF1-3 to mRNA targets to trigger NMD [54]. These findings indicate that the different interactors of CPL1 and CPL2 may determine their target specificity during seed development and germination. They may act primarily through the recognition and elimination of mRNAs with premature translation-termination codons (PTCs) to prevent the production of truncated and potentially deleterious proteins [55].
Recently, CPL1 and CPL2 were reported to affect seed dormancy. The transcript of DELAY OF GERMINATION1 (DOG1), the major quantitative trait locus that specifically controls seed dormancy in Arabidopsis, is tightly controlled on different levels [56,57]. ETHYLENE RESPONSE FACTOR 12 (ERF12) and TOPLES (TPL) repress the expression of DOG1 by directly binding its promoter region, while indirect regulation mode is adopted by histone methyltransferase KYP/SUVH4 [58,59]. On the other hand, CPL1 influences the alternative polyadenylation of DOG1 mRNA. Alternative polyadenylation selectively occurs on the proximal transcription termination site (pTTS) and the distal TTS (dTTS) of the DOG1 sense transcript, resulting in the production of short DOG1 (shDOG1) and long DOG1 (lgDOG1), respectively [60]. Indeed, seeds of the cpl1-8 and cpl1-9 mutants, defective in CPL1, exhibit priority polyadenylation of pTTS, increased expression of shDOG1, and enhanced seed dormancy [61]. However, the CPL1 cofactor that may determine DOG1 target specificity during seed dormancy control has not been identified. More recently, the eceriferum (cer) mutant identified from a stem surface glossiness screen for wax-deficient mutants was shown to harbor a mutation in CER11, and CER11 was demonstrated to be CPL2. Notably, the seed coat of cer11/cpl2 showed reduced mucilage extrusion and deposition, as well as delayed columella secondary cell wall development [62]. Seed mucilage has been reported to affect germination, especially under conditions of limited water availability [63]. Moreover, the secondary cell wall acts as a barrier to block the exchange of gases and fluids between the mature seed and the environment, keeping the embryo dormant and preventing germination, and secondary cell wall defects reduce seed dormancy [64]. Distinct from its roles in the regulation of gene expression, CPL2 also functions in protein sorting in the cer11 context, demonstrated by the fact that secretion of MUCILAGE-MODIFIED 2 (MUM2), a cell-wall modifying enzyme, is delayed in cer11/cpl2 mutant seeds [62]. Therefore, it is likely that CRE11/CPL2 controls the maintenance of seed dormancy and the occurrence of seed germination through a previously unknown role in the cellular secretory trafficking pathway.
Most dsRBD-type RBPs, such as DCLs, HYL1/HYL1 homologs and HENs, are localized to the nucleus and influence the biogenesis of various small RNAs [65,66]. In addition to the dsRBD, some Arabidopsis DCLs contain DExH RNA helicase, RNase III, and/or Piwi/Argonaute/Zwille (PAZ) domains, which mainly function in small RNA processing. Arabidopsis HYL1/DRB1 and its four homologs DRB2–5 harbor two dsRBDs [45]. HYL1 is unusual, containing a nuclear localization sequence and six repeats of 28 amino acids with unknown function [67]. The HYL1 and DCL1 complex is exclusively involved in plant precise microRNA (miRNA) biogenesis. During this process, the first dsRBD of HYL1 interacts with a primary miRNA (pri-miRNA), and the second dsRBD mediates binding with SERRATE (SE) and DCL1 [68,69]. The DRB4–DCL4 complex primarily produces trans-acting (ta) siRNAs and viral siRNAs. DCL2 and DCL3 generate viral siRNAs independent of any dsRBP [70]. The dcl1-11 mutant exhibits germination sensitivity to salt and osmotic stresses, and the germination of dcl2, dcl3 and dcl4 mutant seeds is sensitive to ABA [67,71]. HEN1 is a small RNA methyltransferase that mediates the methylation of miRNAs or siRNAs to protect them from degradation [72]. The hen1-16 and hyponastic leaves1 (hyl1) mutants were identified from ABA sensitivity germination screens and are sensitive to salt and osmotic stresses during germination. HYL1 may affect the ABI5- or ABI3-mediated ABA signaling pathway to control seed germination and seedling establishment by repressing the activity of two mitogen activated protein kinases (MAPKs), SNP1 and MPK3, although the specific mechanism remains unclear [73]. Interestingly, the germination of hyl1 and dcl1-11 mutants was hyposensitive to glucose, which may be explained by alterations in 38 glucose-related pri-miRNAs and their corresponding miRNAs. Notably, the hyl1 mutant showed lower expression of ABI3, ABI4, and ABI5 after glucose treatment, contradictory to the hyl1 ABA hypersensitive seed germination phenotype, suggesting an interaction between ABA, glucose, and miRNAs in the control of seed germination [74]. HYL1 activity is also regulated by phosphorylation/dephosphorylation: CPL1 and CPL2 promote its dephosphorylation and increase its activity, whereas SnRK2 and MPK3 may promote its phosphorylation [75,76,77]. The seed dormancy regulator DOG1 regulates miR156 and miR172 expression by promoting DCL1 and HYL1 expression in lettuce and Arabidopsis seeds [78]. Mutations in miRNA processing genes also affect embryo formation and development. The mutant of DCL1 short integuments (sin1/dcl1-7), the hypomorphic mutant of HEN1 (hen1-8), and the null mutant of HYL1 (hyl1-2) are defective in asymmetric growth of the integuments, leading to abnormal ovule development and reduced female fertility [79,80]. Embryo patterning from the zygote to the globular stage is also severely affected in the dcl1 and hyl1-2 mutants. Transcriptome analysis has identified many miRNAs specifically expressed in the embryo, but their molecular mechanisms of action remain largely unknown [81].
In crops, rice miR156 and wheat miR9678 have been reported to function in seed dormancy through the regulation of the GA and GA/ABA pathways, respectively. Dysfunctions in one MIR156 subfamily or overexpression of MIR9678 increased seed dormancy and suppressed seed pre-harvest sprouting (PHS) [82,83]. These findings suggest that miRNA biogenesis mediated by dsRBD-type RBPs may have an important role in crop seed dormancy and germination.

5. GRPs in Seed Development and Performance

GRPs are involved in alternative splicing or transcriptional regulation. Many GRPs function in various stress-related seed germination processes (Table 1). The Arabidopsis genome encodes eight GRPs, and its class III includes four nuclear-localized members (CSP1–4) [84,85,86,87]. All CSPs are predominantly expressed in shoot apices and siliques at both the RNA and protein levels, but they have different roles in plant growth, seed development, and performance [88]. The overexpression of CSP4 leads to reduced silique length and embryo lethality, whereas CSP2 knock-down plants show changes in vegetative development and flowering control [84,87]. Overexpression of CSP1 confers sensitivity to dehydration or salt stress with a delayed germination phenotype. By contrast, CSP2-OE lines exhibit accelerated seed germination under salt stress conditions [89]. CSP2 has also been designated as GRP2 [90]. Intriguingly, GRP2 only affects seed germination through an ABA-independent pathway. It also positively affects seed germination and seedling growth under cold stress [91]. Thus, although the spatial expression profiles of the CSPs are highly similar, their specific biological roles differ, and they deserve further study.
Another two of the Arabidopsis eight GRPs, GRP4 and GRP7, also affect plant stress response. GRP4 is induced by cold stress, markedly repressed by salinity, and slightly repressed by dehydration. Intriguingly, GRP4-OE seeds exhibit delayed germination under salt and dehydration stress, but no noticeable differences were observed under cold or freezing stress [92]. In contrast to GRP4, GRP7 is repressed by ABA, high salt, and mannitol treatment. Its knock-out mutant is hypersensitive to these stresses, displaying delayed seed germination and root development. Moreover, the grp7 mutant has significantly higher expression of RD29A and RAB18 than the wild type in response to stresses and ABA. It has therefore been proposed that GRP7 positively modulates gene expression in the ABA and stress signaling pathway [93]. However, this is inconsistent with the increased seed germination rate observed in another GRP7 knock-out mutant under NaCl or mannitol treatment, which suggested that GRP7 had a negative effect on abiotic stress response. This discrepancy may have arisen from differences between the two mutants or from differences in the stress treatments or plant growth conditions; more accurate experiments are needed to assess the effects of GRP7 on stress tolerance [94]. In addition, GRP7 has been shown to promote seed germination and seedling growth under low temperatures [94].
Arabidopsis RZ-1A, RZ-1B, and RZ-1C are characterized by a zinc finger motif between the RRM domain and the C terminus. RZ-1B and RZ-1C fall into one phylogenetic clade, whereas RZ-1A belongs to a separate group, suggesting that they may have different biological functions [95,96]. RZ-1A-OE seeds showed delayed germination, whereas faster germination and earlier seedling onset were observed in the loss-of-function mutant after salt, dehydration, and freezing treatments. Furthermore, seed germination was affected by ABA or glucose, implying that RZ-1A negatively regulated seed germination through an ABA-dependent pathway [97,98]. By contrast, RZ-1B and RZ-1C redundantly function in the modulation of seed germination and other plant developmental processes under normal growth conditions. Their double mutant exhibited a significantly delayed germination rate and root and leaf defects compared to the wild type. This is consistent with the high levels of RZ-1C expression observed in embryos, endosperm of germinated seeds, and newly initiated leaves and root tips [96].
The rice genome encodes at least six GRPs, designated OsGRP1–OsGRP6. The overexpression of OsGRP1, OsGRP4, and OsGRP6 in the Arabidopsis grp7 knockout mutant rescued the cold-sensitive germination phenotype of the grp7 mutant [99]. Both rice and wheat (Triticum aestivum) harbor three RZ proteins, referred to as OsRZ1–OsRZ3 and TaRZ1–TaRZ3, that are homologous to Arabidopsis RZs. The OsRZs showed noticeable up-regulation only in response to cold stress, whereas the expression of the three TaRZs was affected differently by high salt, drought, and cold stress. Only OsRZ2 affected seed germination under cold stress; the three TaRZ proteins had no significant effect under cold stress but had different effects on germination under high salt and dehydration [100,101]. These studies indicate that some GRPs and RZs of crop plants have important and diverse functions in abiotic stress response.

6. PPR-Type RBPs in Seed Development and Performance

PPR proteins are encoded by the nuclear genome but predominantly targeted to the chloroplasts and/or mitochondria, where they have a significant role in organelle RNA metabolism, biogenesis, and function [102]. Many PPR proteins are involved in embryo and/or kernel development, such as Arabidopsis GLUTAMINE-RICH PROTEIN23 (GRP23) and a series of embryo-defective (EMB) proteins; maize empty pericarp4 (EMP4), EMP5, and PPR8522; and rice opaque and growth retardation 1 (OGR1) [103,104,105,106,107,108]. Barkan and Small (2014) have presented a comprehensive review of the significance and possible biological mechanisms by which PPR proteins function in seed development [109]. Accumulating evidence indicates that PPRs are also involved in the regulation of seed germination (Table 1 and Figure 2). GENOMES UNCOUPLED 1 (GUN1) is a chloroplast-localized P-type PPR protein and an important regulator of plastid-to-nucleus retrograde signaling [110]. During chloroplast retrograde signaling, GUN1 directly interacts with the RNA editing factor MULTIPLE ORGANELLAR RNA EDITING FACTOR 2 (MORF2) to regulate plastid RNA editing and represses nuclear-encoded genes through ABI4 [111]. Seed germination of the gun1 mutant is more sensitive to ABA treatment than abi4, suggesting that the germination regulation of gun1 may be independent of ABI4. Furthermore, a later study showed that both ABA and sucrose delay the seed germination rate of the gun1 mutant [112]. These results suggest a complex interplay between sucrose, ABA, and plastid signaling in seed germination. Many mitochondrially localized PPRs affect energy production and display an altered germination phenotype, especially in response to ABA. For instance, ABA hypersensitive germination 11 (ahg11), slow growth 1 (slg1), and slow growth 2 (slo2) show enhanced ABA hypersensitivity and delayed germination, and appear to act by editing mitochondrially encoded NAD genes of complex I (nad1, nad3, nad4, and/or nad7) [113,114,115,116]. Alternative splicing is another means by which some RRPs alter the expression level of mitochondrial electron transport chain (mETC) subunits. ABA overly sensitive 5 (ABO5) and PPR19 are responsible for the cis-splicing of nad2 intron3 and stabilization of nad1 exon2–exon3 transcripts, respectively [117,118]. The mitochondrial-localized DEAD box RNA helicase ABO6 also impairs the splicing of genes that encode complex I subunits (nad1, nad2, nad4, and nad5) [119]. Likewise, abo5, ppr19, and abo6 mutant seeds also show delayed germination. Disruptions in PENTATRICOPEPTIDE REPEAT PROTEIN FOR GERMINATION ON NaCl (PNG) lead to an ABA hypersensitive phenotype with delayed germination, but it is unclear how PNG changes the transcript abundance of nad1, which encodes a complex I subunit [120]. Some PPRs regulate mETC activity by an unknown mechanism. For example, the ppr40 mutant exhibits hypersensitivity of germination to ABA. No obvious alterations in RNA processes or in the abundance of mETC complex subunits were identified, but the ppr40 mutant had a strongly reduced capacity for complex III electron transport and accumulated more reactive oxygen species (ROS) [121]. In contrast to the above-mentioned increased ABA sensitivity, PPR96 was demonstrated to negatively regulate the ABA pathway because the ppr96 loss-of-function mutant exhibited an ABA insensitive phenotype with a faster germination rate. Future identification of the mitochondrial targets and molecular mechanisms of PPR96 in ABA tolerance will deepen our understanding of mitochondrial function, ABA, and seed germination [122].
Recently, the PPR protein MITOCHONDRIAL EDITING FACTOR11 (MEF11), also known as LOVASTATIN INSENSITIVE1 (LOI1), was isolated from a suppressor screen of the ABA biosynthesis-deficient mutant aba3-1. Dysfunction in MEF11 led to increased seed dormancy. Remarkably, after one month of ripening and stratification, mef11 mutant seeds still germinated slightly more slowly than the wild type. Moreover, hypersensitivity to the germination inhibitor paclobutrazol (a gibberellin biosynthesis inhibitor) and abiotic stresses (NaCl and mannitol) was also observed in mef11 mutant seeds. These alterations in seed behavior may result from a loss of editing of the mitochondrial cytochrome c maturation FN2 (ccmFN2) gene, which encodes a cytochrome c-heme lyase subunit required for cytochrome c generation. Increased respiration and altered production of hydrogen peroxide (H2O2) and nitrogen monoxide (NO) were observed in mef11 mutant seeds, as well as in the mitochondrial function-deficient ppr mutants described above [123,124].
Both ROS and NO function as positive signals that contribute to the transition from seed dormancy to germination, and they probably act through ABA signaling [125,126]. Additional findings suggest that PPR may affect ABA signaling through ROS and/or NO in mitochondrial retrograde signaling [127]. Recently, Nonogaki (2019) outlined a coherent molecular model that integrated mitochondrial function, ROS, NO, DOG1 and seed physiology [128]. However, more experiments are required to confirm and further develop this model. For example, with the exception of ABI4′s role as a retrograde regulator, little is known about the mechanisms of signal perception and transduction from the mitochondria to the nucleus in mitochondrial retrograde signaling.
In rice, the white stripe leaf (wsl) mutant exhibits chlorotic striations on leaves. The WSL gene encodes a chloroplast-localized PPR, and its mutant shows enhanced sensitivity to ABA, high salt, and sugar, with a delayed seed germination rate. Both mis-regulation of splicing in the chloroplast ribosomal protein subunit L2 (RPL2) transcript and decreased accumulation of plastid rRNAs and translation products were observed in the wsl mutant and may have caused its sensitivity to abiotic stresses [129]. OGR1 is targeted to the mitochondrion and acts by editing five mitochondrial transcripts (ccmC, COX2, COX2, NAD2, and NAD4). The ogr1 mutant has pleotropic phenotypes, including delayed seed germination under normal conditions, but its mechanisms of action have not been described in detail [107]. In maize, chloroplast-localized PPR8522 and mitochondrion-targeted EMP4 have been reported to affect seed development and germination. Mutations in both genes are embryo lethal. Embryo rescue experiments showed that ppr8522 and emp4 embryos successfully germinated with sucrose supplementation, suggesting that a metabolic defect affects germination in the two mutant embryos [104,106].
Table 1. List of plant known RBPs involved in seed development and performance.
Table 1. List of plant known RBPs involved in seed development and performance.
TypeNameSpeciesSubcellular
Localization a
Seed DevelopmentSeed Germination
under Different Conditions
Seed DormancyReference
NormalABASalt OsmoticCold
PUFAPUM5A. thalianaCytoplasm [38,39]
APUM23A. thalianaNucleus [42,43]
APUM24A. thalianaNucleus [16,40]
APUM9-11A. thalianaCytoplasm/Nuclear/periphery [17]
dsRBDCPL1A. thalianaNucleus [46,47,52,61]
CPL2A. thalianaCytoplasm/Nucleus [46,51,62]
HYL1A. thalianaNucleus [67,71,73,80,81]
DCL1A. thalianaNucleus [65,71,79,81]
HEN1A. thalianaNucleus [66,71,80]
DCL2-4A. thalianaNucleus [66,71]
GRPCSP1A. thalianaNucleus/Cytoplasm [85,89]
CSP2A. thalianaNucleus/Cytoplasm [84,89,91]
CSP4A. thalianaNucleus/Cytoplasm [87]
RZ-1AA. thalianaNucleus/Cytoplasm [95,97,98]
AtRZ-1BA. thalianaNucleus [95,96]
AtRZ-1CA. thalianaNucleus [95,96]
GRP4A. thalianaUnknown [92]
GRP7A. thalianaNucleus/Cytoplasm [93,94]
OsGRP1O. Sativa L.Unknown [99]
OsGRP4O. sativa L.Unknown [99]
OsGRP6O. sativa L.Unknown [99]
OsRZ2O. sativa L.Unknown [100]
TaRZ1T. aestivum L.Nucleus [101]
TaRZ2T. aestivum L.Nucleus/cytoplasm/ER [101]
TaRZ3T. aestivum L.Nucleus [101]
PPREMBsA. thalianaChloroplast [103]
GUN1A. thalianaChloroplast [110,112]
AHG11A. thalianaMitochondria [113]
SLG1A. thalianaMitochondria [114]
SLO2A. thalianaMitochondria [116]
ABO5A. thalianaMitochondria [118]
ABO6A. thalianaMitochondria [119]
PPR19A. thalianaMitochondria [117]
PNGA. thalianaMitochondria [120]
PPR40A. thalianaMitochondria [121]
PPR96A. thalianaMitochondria [122]
GRP23A. thalianaNucleus [108]
MEF11A. thalianaMitochondria [123,124]
WLSO. sativa L.Chloroplast [129]
OGR1O. sativa L.Mitochondria [107]
EMP4Z maysMitochondria [104]
PPR8522Z. maysChloroplast [106]
a Indicate the RBP subcellular localization verified with experiments.

7. Conclusions and Open Questions

Plant RBPs are a huge family with diverse biological roles. Nonetheless, only few of them have been functionally characterized with respect to seed biology, including seed development, dormancy, and germination regulation (Table 1 and Figure 2). Seed germination begins with the metabolism of stored mRNA in dry seeds after imbibition, and these early metabolic events may determine whether or not seeds germinate. RBPs appear to have an important effect on the fate of stored mRNA. Therefore, one pressing task is to systematically identify their substrate mRNAs and better understand their metabolism during early germination. As mentioned above, some RBPs display pleiotropic phenotypes, indicating that they may regulate different targets in a spatiotemporal model. RNA immunoprecipitation (RIP)-seq and/or cross-linking and immunoprecipitation (CLIP)-seq coupled with specific cell type isolation/fraction technology may help to elucidate their targets. In mammals, two newly established technologies, targets of RNA-binding protein identified by editing (TRIBE) and RNA tagging, provide alternative approaches for the genome-wide identification of plant RBP targets in vivo [130,131]. On the other hand, increasing evidence suggests that the regulation of RBPs is important, especially the control of their expression or protein activity during biological processes. Similar to human RNA-binding protein 25 (HsRBM25), Arabidopsis RBM25, a PWI and RRM domain-containing RBP responsible for alternative splicing of the ABA negative core component protein phosphatase 2C HABI, can be activated by ABA signaling not only at the transcriptional level, but also at the protein level through modulation of its phosphorylation status [132,133]. In maize, different types of RBPs, including RNA helicases, PUFs, and pre-mRNA-splicing factors, are preferentially modified by the cereal-specific small ubiquitin-like modifier (SUMO) variant DiSUMO-LIKE (DSUL) during embryogenesis. The DSULylation modification may affect activity and/or stability of the RBPs, resulting in translation/repression of specific mRNAs during embryo development [134]. Therefore, the characterization of RBP regulation at different levels is helpful for uncovering the biological mechanisms in which they are involved. We firmly believe that exciting progress on these questions is forthcoming and will integrate signaling pathways involved in mRNA metabolism during seed-relevant biological processes.

Author Contributions

Writing-Original Draft, L.L.; Review and Editing, Y.X., T.W., and L.D.; Supervision, Y.X. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by grants from the National Natural Science Foundation of China (No. 31670323), and special funds for science technology innovation and industrial development of Shenzhen Dapeng New District (No. KY20160205 and RCTD20180102).

Acknowledgments

We apologize to those whose work was not cited because of space constrains.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Moore, M.J. From birth to death: The complex lives of eukaryotic mRNAs. Science 2005, 309, 1514–1518. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Owttrim, G.W. RNA helicases and abiotic stress. Nucleic Acids Res. 2006, 34, 3220–3230. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Tam, P.P.; Barrette-Ng, I.H.; Simon, D.M.; Tam, M.W.; Ang, A.L.; Muench, D.G. The Puf family of RNA-binding proteins in plants: Phylogeny, structural modeling, activity and subcellular localization. BMC Plant Biol. 2010, 10, 44. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Lorkovic, Z.J.; Barta, A. Genome analysis: RNA recognition motif (RRM) and K homology (KH) domain RNA-binding proteins from the flowering plant Arabidopsis thaliana. Nucleic Acids Res. 2002, 30, 623–635. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Sachetto-Martins, G.; Franco, L.O.; De Oliveira, D.E. Plant glycine-rich proteins: a family or just proteins with a common motif? Biochim. Biophys. Acta 2000, 1492, 1–14. [Google Scholar] [CrossRef]
  6. Walker, N.S.; Stiffler, N.; Barkan, A. POGs/PlantRBP: A resource for comparative genomics in plants. Nucleic Acids Res. 2007, 35, D852–D856. [Google Scholar] [CrossRef]
  7. Zhang, J.; Zhao, Y.; Xiao, H.; Zheng, Y.; Yue, B. Genome-wide identification, evolution, and expression analysis of RNA-binding glycine-rich protein family in maize. J. Integr. Plant Biol. 2014, 56, 1020–1031. [Google Scholar] [CrossRef]
  8. Koster, 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]
  9. Reichel, M.; Liao, Y.; Rettel, M.; Ragan, C.; Evers, M.; Alleaume, A.M.; Horos, R.; Hentze, M.W.; Preiss, T.; Millar, A.A. In planta determination of the mRNA-binding proteome of Arabidopsis etiolated seedlings. Plant Cell 2016, 28, 2435–2452. [Google Scholar] [CrossRef] [Green Version]
  10. Zhang, Z.; Boonen, K.; Ferrari, P.; Schoofs, L.; Janssens, E.; van Noort, V.; Rolland, F.; Geuten, K. UV crosslinked mRNA-binding proteins captured from leaf mesophyll protoplasts. Plant Methods 2016, 12, 42. [Google Scholar] [CrossRef] [Green Version]
  11. Howell, K.A.; Narsai, R.; Carroll, A.; Ivanova, A.; Lohse, M.; Usadel, B.; Millar, A.H.; Whelan, J. Mapping metabolic and transcript temporal switches during germination in rice highlights specific transcription factors and the role of RNA instability in the germination process. Plant Physiol. 2009, 149, 961–980. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Nakabayashi, K.; Okamoto, M.; Koshiba, T.; Kamiya, Y.; Nambara, E. Genome-wide profiling of stored mRNA in Arabidopsis thaliana seed germination: Epigenetic and genetic regulation of transcription in seed. Plant J. 2005, 41, 697–709. [Google Scholar] [CrossRef] [PubMed]
  13. Rajjou, L.; Gallardo, K.; Debeaujon, I.; Vandekerckhove, J.; Job, C.; Job, D. The effect of alpha-amanitin on the Arabidopsis seed proteome highlights the distinct roles of stored and neosynthesized mRNAs during germination. Plant Physiol. 2004, 134, 1598–1613. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Masaki, S.; Yamada, T.; Hirasawa, T.; Todaka, D.; Kanekatsu, M. Proteomic analysis of RNA-binding proteins in dry seeds of rice after fractionation by ssDNA affinity column chromatography. Biotechnol. Lett. 2008, 30, 955–960. [Google Scholar] [CrossRef]
  15. Sano, N.; Masaki, S.; Tanabata, T.; Yamada, T.; Hirasawa, T.; Kashiwagi, M.; Kanekatsu, M. RNA-binding proteins associated with desiccation during seed development in rice. Biotechno. Lett. 2013, 35, 1945–1952. [Google Scholar] [CrossRef]
  16. Shanmugam, T.; Abbasi, N.; Kim, H.S.; Kim, H.B.; Park, N.I.; Park, G.T.; Oh, S.A.; Park, S.K.; Muench, D.G.; Choi, Y.; et al. An Arabidopsis divergent pumilio protein, APUM24, is essential for embryogenesis and required for faithful pre-rRNA processing. Plant J. 2017, 92, 1092–1105. [Google Scholar] [CrossRef] [Green Version]
  17. Xiang, Y.; Nakabayashi, K.; Ding, J.; He, F.; Bentsink, L.; Soppe, W.J.J. Reduced Dormancy5 encodes a protein phosphatase 2C that is required for seed dormancy in Arabidopsis. Plant Cell 2014, 26, 4362–4375. [Google Scholar] [CrossRef] [Green Version]
  18. Lorković, Z.J. Role of plant RNA-binding proteins in development, stress response and genome organization. Trends Plant Sci. 2009, 14, 229–236. [Google Scholar] [CrossRef]
  19. Hanano, S.; Sugita, M.; Sugiura, M. Isolation of a novel RNA-binding protein and its association with a large ribonucleoprotein particle present in the nucleoplasm of tobacco cells. Plant Mol. Biol. 1996, 31, 57–68. [Google Scholar] [CrossRef]
  20. Czolpinska, M.; Rurek, M. Plant glycine-rich proteins in stress response: An emerging, still prospective story. Front. Plant Sci. 2018, 9, 302. [Google Scholar] [CrossRef]
  21. Graumann, P.L.; Marahiel, M.A. A superfamily of proteins that contain the cold-shock domain. Trends Biochem. Sci. 1998, 23, 286–290. [Google Scholar] [CrossRef]
  22. Manival, X.; Ghisolfi-Nieto, L.; Joseph, G.; Bouvet, P.; Erard, M. RNA-binding strategies common to cold-shock domain- and RNA recognition motif-containing proteins. Nucleic Acids Res. 2001, 29, 2223–2233. [Google Scholar] [CrossRef] [PubMed]
  23. Maris, C.; Dominguez, C.; Allain, F.H. The RNA recognition motif, a plastic RNA-binding platform to regulate post-transcriptional gene expression. FEBS J. 2005, 272, 2118–2131. [Google Scholar] [CrossRef]
  24. Tian, B.; Bevilacqua, P.C.; Diegelman-Parente, A.; Mathews, M.B. The double-stranded-RNA-binding motif: Interference and much more. Nat. Rev. Mol. Cell. Biol. 2004, 5, 1013–1023. [Google Scholar] [CrossRef] [PubMed]
  25. Chen, Y.; Varani, G. Engineering RNA-binding proteins for biology. FEBS J. 2013, 280, 3734–3754. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Zhang, B.; Gallegos, M.; Puoti, A.; Durkin, E.; Fields, S.; Kimble, J.; Wickens, M.P. A conserved RNA-binding protein that regulates sexual fates in the C. elegans hermaphrodite germ line. Nature 1997, 390, 477–484. [Google Scholar] [CrossRef]
  27. Edwards, T.A.; Pyle, S.E.; Wharton, R.P.; Aggarwal, A.K. Structure of Pumilio reveals similarity between RNA and peptide binding motifs. Cell 2001, 105, 281–289. [Google Scholar] [CrossRef] [Green Version]
  28. Wang, X.; Zamore, P.D.; Hall, T.M. Crystal structure of a pumilio homology domain. Mol. Cell 2001, 7, 855–865. [Google Scholar] [CrossRef]
  29. Wang, X.; McLachlan, J.; Zamore, P.D.; Hall, T.M. Modular recognition of RNA by a human pumilio-homology domain. Cell 2002, 110, 501–512. [Google Scholar] [CrossRef] [Green Version]
  30. Francischini, C.W.; Quaggio, R.B. Molecular characterization of Arabidopsis thaliana PUF proteins-binding specificity and target candidates. FEBS J. 2009, 276, 5456–5470. [Google Scholar] [CrossRef]
  31. Wang, M.; Oge, L.; Perez-Garcia, M.D.; Hamama, L.; Sakr, S. The PUF Protein Family: Overview on PUF RNA targets, biological functions, and post transcriptional regulation. Int. J. Mol. Sci. 2018, 19, 410. [Google Scholar] [CrossRef] [Green Version]
  32. Takenaka, M.; Zehrmann, A.; Verbitskiy, D.; Hartel, B.; Brennicke, A. RNA editing in plants and its evolution. Annu. Rev. Genet. 2013, 47, 335–352. [Google Scholar] [CrossRef]
  33. Lurin, C.; Andrés, C.; Aubourg, S.; Bellaoui, M.; Bitton, F.; Bruyère, C.; Caboche, M.; Debast, C.; Gualberto, J.; Hoffmann, B.; et al. Genome-wide analysis of Arabidopsis pentatricopeptide repeat proteins reveals their essential role in organelle biogenesis. Plant Cell 2004, 16, 2089–2103. [Google Scholar] [CrossRef] [Green Version]
  34. Grishin, N.V. KH domain: One motif, two folds. Nucleic Acids Res. 2001, 29, 638–643. [Google Scholar] [CrossRef]
  35. Makeyev, A.V.; Liebhaber, S.A. The poly(C)-binding proteins: A multiplicity of functions and a search for mechanisms. RNA 2002, 8, 265–278. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Mingam, A.; Toffano-Nioche, C.; Brunaud, V.; Boudet, N.; Kreis, M.; Lecharny, A. DEAD-box RNA helicases in Arabidopsis thaliana: Establishing a link between quantitative expression, gene structure and evolution of a family of genes. Plant Biotechnol. J. 2004, 2, 401–415. [Google Scholar] [CrossRef]
  37. Huh, S.U.; Kim, M.J.; Paek, K.H. Arabidopsis Pumilio protein APUM5 suppresses cucumber mosaic virus infection via direct binding of viral RNAs. Proc. Natl. Acad. Sci. USA 2013, 110, 779–784. [Google Scholar] [CrossRef] [Green Version]
  38. Huh, S.U.; Paek, K.H. APUM5, encoding a Pumilio RNA binding protein, negatively regulates abiotic stress responsive gene expression. BMC Plant Biol. 2014, 14, 75. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Arae, T.; Morita, K.; Imahori, R.; Suzuki, Y.; Yasuda, S.; Sato, T.; Yamaguchi, J.; Chiba, Y. Identification of Arabidopsis CCR4-NOT complexes with pumilio RNA-binding proteins, APUM5 and APUM2. Plant Cell Physiol. 2019, 60, 2015–2025. [Google Scholar] [CrossRef]
  40. Maekawa, S.; Ishida, T.; Yanagisawa, S. Reduced expression of APUM24, encoding a novel rRNA processing factor, induces sugar-dependent nucleolar stress and altered sugar responses in Arabidopsis thaliana. Plant Cell 2018, 30, 209–227. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Zhang, C.; Muench, D.G. A Nucleolar PUF RNA-binding protein with specificity for a unique RNA sequence. J. Biol. Chem. 2015, 290, 30108–30118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Abbasi, N.; Kim, H.B.; Park, N.I.; Kim, H.S.; Kim, Y.K.; Park, Y.I.; Choi, S.B. APUM23, a nucleolar Puf domain protein, is involved in pre-ribosomal RNA processing and normal growth patterning in Arabidopsis. Plant J. 2010, 64, 960–976. [Google Scholar] [CrossRef] [PubMed]
  43. Huang, K.C.; Lin, W.C.; Cheng, W.H. Salt hypersensitive mutant 9, a nucleolar APUM23 protein, is essential for salt sensitivity in association with the ABA signaling pathway in Arabidopsis. BMC Plant Biol. 2018, 18, 40. [Google Scholar] [CrossRef] [Green Version]
  44. Nyikó, T.; Auber, A.; Bucher, E. Functional and molecular characterization of the conserved Arabidopsis PUMILIO protein, APUM9. Plant Mol. Biol. 2019, 100, 199–214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Hiraguri, A.; Itoh, R.; Kondo, N.; Nomura, Y.; Aizawa, D.; Murai, Y.; Koiwa, H.; Seki, M.; Shinozaki, K.; Fukuhara, T. Specific interactions between Dicer-like proteins and HYL1/DRB-family dsRNA-binding proteins in Arabidopsis thaliana. Plant Mol. Biol. 2005, 57, 173–188. [Google Scholar] [CrossRef]
  46. Koiwa, H.; Hausmann, S.; Bang, W.Y.; Ueda, A.; Kondo, N.; Hiraguri, A.; Fukuhara, T.; Bahk, J.D.; Yun, D.J.; Bressan, R.A.; et al. Arabidopsis C-terminal domain phosphatase-like 1 and 2 are essential Ser-5-specific C-terminal domain phosphatases. Proc. Natl. Acad. Sci. USA 2004, 101, 14539–14544. [Google Scholar] [CrossRef] [Green Version]
  47. Xiong, L.; Lee, H.; Ishitani, M.; Tanaka, Y.; Stevenson, B.; Koiwa, H.; Bressan, R.A.; Hasegawa, P.M.; Zhu, J.K. Repression of stress-responsive genes by FIERY2, a novel transcriptional regulator in Arabidopsis. Proc. Natl. Acad. Sci. USA 2002, 99, 10899–10904. [Google Scholar] [CrossRef] [Green Version]
  48. Chen, T.; Cui, P.; Chen, H.; Ali, S.; Zhang, S.; Xiong, L. A KH-domain RNA-binding protein interacts with FIERY2/CTD phosphatase-like 1 and splicing factors and is important for pre-mRNA splicing in Arabidopsis. PLoS Genet. 2013, 9, e1003875. [Google Scholar] [CrossRef] [Green Version]
  49. Jeong, I.S.; Fukudome, A.; Aksoy, E.; Bang, W.Y.; Kim, S.; Guan, Q.; Bahk, J.D.; May, K.A.; Russell, W.K.; Zhu, J.; et al. Regulation of abiotic stress signalling by Arabidopsis C-terminal domain phosphatase-like 1 requires interaction with a k-homology domain-containing protein. PLoS ONE 2013, 8, e80509. [Google Scholar] [CrossRef] [Green Version]
  50. Jiang, J.; Wang, B.; Shen, Y.; Wang, H.; Feng, Q.; Shi, H. The Arabidopsis RNA binding protein with K homology motifs, SHINY1, interacts with the C-terminal domain phosphatase-like 1 (CPL1) to repress stress-inducible gene expression. PLoS Genet. 2013, 9, e1003625. [Google Scholar] [CrossRef] [Green Version]
  51. Ueda, A.; Li, P.; Feng, Y.; Vikram, M.; Kim, S.; Kang, C.H.; Kang, J.S.; Bahk, J.D.; Lee, S.Y.; Fukuhara, T.; et al. The Arabidopsis thaliana carboxyl-terminal domain phosphatase-like 2 regulates plant growth, stress and auxin responses. Plant Mol. Biol. 2008, 67, 683–697. [Google Scholar] [CrossRef] [PubMed]
  52. Rodríguez-Cazorla, E.; Ortuño-Miquel, S.; Candela, H.; Bailey-Steinitz, L.J.; Yanofsky, M.F.; Martínez-Laborda, A.; Ripoll, J.J.; Vera, A. Ovule identity mediated by pre-mRNA processing in Arabidopsis. PLoS Genet. 2018, 14, e1007182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Yoine, M.; Ohto, M.A.; Onai, K.; Mita, S.; Nakamura, K. The lba1 mutation of UPF1 RNA helicase involved in nonsense-mediated mRNA decay causes pleiotropic phenotypic changes and altered sugar signalling in Arabidopsis. Plant J. 2006, 47, 49–62. [Google Scholar] [CrossRef]
  54. Cui, P.; Chen, T.; Qin, T.; Ding, F.; Wang, Z.; Chen, H.; Xiong, L. The RNA Polymerase II C-terminal domain phosphatase-like protein FIERY2/CPL1 interacts with eIF4AIII and is essential for nonsense-mediated mRNA decay in Arabidopsis. Plant Cell 2016, 28, 770–785. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Dai, Y.; Li, W.; An, L. NMD mechanism and the functions of Upf proteins in plant. Plant Cell Rep. 2016, 35, 5–15. [Google Scholar] [CrossRef]
  56. Bentsink, L.; Jowett, J.; Hanhart, C.J.; Koornneef, M. Cloning of DOG1, a quantitative trait locus controlling seed dormancy in Arabidopsis. Proc. Natl. Acad. Sci. USA 2006, 103, 17042–17047. [Google Scholar] [CrossRef] [Green Version]
  57. Soppe, W.J.J.; Bentsink, L. Seed dormancy back on track, its definition and regulation by DOG1. New Phytol. 2020. [Google Scholar] [CrossRef] [Green Version]
  58. Li, X.; Chen, T.; Li, Y.; Wang, Z.; Cao, H.; Chen, F.; Li, Y.; Soppe, W.J.J.; Li, W.; Liu, Y. ETR1/RDO3 regulates seed dormancy by relieving the inhibitory effect of the ERF12-TPL complex on DELAY OF GERMINATION1 expression. Plant Cell 2019, 31, 832–847. [Google Scholar] [CrossRef] [Green Version]
  59. Zheng, J.; Chen, F.; Wang, Z.; Cao, H.; Li, X.; Deng, X.; Soppe, W.J.J.; Li, Y.; Liu, Y. A novel role for histone methyltransferase KYP/SUVH4 in the control of Arabidopsis primary seed dormancy. New Phytol. 2012, 193, 605–616. [Google Scholar] [CrossRef]
  60. Cyrek, M.; Fedak, H.; Ciesielski, A.; Guo, Y.; Sliwa, A.; Brzezniak, L.; Krzyczmonik, K.; Pietras, Z.; Kaczanowski, S.; Liu, F.; et al. Seed dormancy in Arabidopsis is controlled by alternative polyadenylation of DOG1. Plant Physiol. 2016, 170, 947–955. [Google Scholar] [CrossRef] [Green Version]
  61. Kowalczyk, J.; Palusinska, M.; Wroblewska-Swiniarska, A.; Pietras, Z.; Szewc, L.; Dolata, J.; Jarmolowski, A.; Swiezewski, S. Alternative polyadenylation of the sense transcript controls antisense transcription of DELAY OF GERMINATION 1 in Arabidopsis. Mol. Plant 2017, 10, 1349–1352. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Shi, L.; Dean, G.H. ECERIFERUM11/C-TERMINAL DOMAIN PHOSPHATASE-LIKE2 affects secretory trafficking. Plant cell 2019, 181, 901–915. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Penfield, S.; Meissner, R.C.; Shoue, D.A.; Carpita, N.C.; Bevan, M.W. MYB61 is required for mucilage deposition and extrusion in the Arabidopsis seed coat. Plant Cell 2001, 13, 2777–2791. [Google Scholar] [CrossRef] [Green Version]
  64. Debeaujon, I.; Léon-Kloosterziel, K.M.; Koornneef, M. Influence of the testa on seed dormancy, germination, and longevity in Arabidopsis. Plant Physiol. 2000, 122, 403–414. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Papp, I.; Mette, M.F.; Aufsatz, W.; Daxinger, L.; Schauer, S.E.; Ray, A.; van der Winden, J.; Matzke, M.; Matzke, A.J. Evidence for nuclear processing of plant micro RNA and short interfering RNA precursors. Plant Physiol. 2003, 132, 1382–1390. [Google Scholar] [CrossRef] [Green Version]
  66. Pontes, O.; Vitins, A.; Ream, T.S.; Hong, E.; Pikaard, C.S.; Costa-Nunes, P. Intersection of small RNA pathways in Arabidopsis thaliana sub-nuclear domains. PLoS ONE 2013, 8, e65652. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Lu, C.; Fedoroff, N. A mutation in the Arabidopsis HYL1 gene encoding a dsRNA binding protein affects responses to abscisic acid, auxin, and cytokinin. Plant Cell 2000, 12, 2351–2366. [Google Scholar] [CrossRef] [Green Version]
  68. Voinnet, O. Origin, biogenesis, and activity of plant microRNAs. Cell 2009, 136, 669–687. [Google Scholar] [CrossRef] [Green Version]
  69. Yang, S.W.; Chen, H.Y.; Yang, J.; Machida, S.; Chua, N.H.; Yuan, Y.A. Structure of Arabidopsis HYPONASTIC LEAVES1 and its molecular implications for miRNA processing. Structure 2010, 18, 594–605. [Google Scholar] [CrossRef] [Green Version]
  70. Curtin, S.J.; Watson, J.M.; Smith, N.A.; Eamens, A.L.; Blanchard, C.L.; Waterhouse, P.M. The roles of plant dsRNA-binding proteins in RNAi-like pathways. FEBS Lett. 2008, 582, 2753–2760. [Google Scholar] [CrossRef]
  71. Zhang, J.F.; Yuan, L.J.; Shao, Y.; Du, W.; Yan, D.W.; Lu, Y.T. The disturbance of small RNA pathways enhanced abscisic acid response and multiple stress responses in Arabidopsis. Plant Cell Environ. 2008, 31, 562–574. [Google Scholar] [CrossRef]
  72. Li, J.; Yang, Z.; Yu, B.; Liu, J.; Chen, X. Methylation protects miRNAs and siRNAs from a 3′-end uridylation activity in Arabidopsis. Curr. Biol. 2005, 15, 1501–1507. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Lu, C.; Han, M.H.; Guevara-Garcia, A.; Fedoroff, N.V. Mitogen-activated protein kinase signaling in postgermination arrest of development by abscisic acid. Proc. Natl. Acad. Sci. USA 2002, 99, 15812–15817. [Google Scholar] [CrossRef] [Green Version]
  74. Duarte, G.T.; Matiolli, C.C.; Pant, B.D.; Schlereth, A.; Scheible, W.R.; Stitt, M.; Vicentini, R.; Vincentz, M. Involvement of microRNA-related regulatory pathways in the glucose-mediated control of Arabidopsis early seedling development. J. Exp. Bot. 2013, 64, 4301–4312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Manavella, P.A.; Hagmann, J.; Ott, F.; Laubinger, S.; Franz, M.; Macek, B.; Weigel, D. Fast-forward genetics identifies plant CPL phosphatases as regulators of miRNA processing factor HYL1. Cell 2012, 151, 859–870. [Google Scholar] [CrossRef] [Green Version]
  76. Raghuram, B.; Sheikh, A.H.; Rustagi, Y.; Sinha, A.K. MicroRNA biogenesis factor DRB1 is a phosphorylation target of mitogen activated protein kinase MPK3 in both rice and Arabidopsis. FEBS J. 2015, 282, 521–536. [Google Scholar] [CrossRef]
  77. Yan, J.; Wang, P. The SnRK2 kinases modulate miRNA accumulation in Arabidopsis. PLoS Genet. 2017, 13, e1006753. [Google Scholar] [CrossRef] [Green Version]
  78. Huo, H.; Wei, S.; Bradford, K.J. Delay of Germination1 (DOG1) regulates both seed dormancy and flowering time through microRNA pathways. Proc. Natl. Acad. Sci. USA 2016, 113, E2199–E2206. [Google Scholar] [CrossRef] [Green Version]
  79. Robinson-Beers, K.; Pruitt, R.E.; Gasser, C.S. Ovule development in wild-type Arabidopsis and two female-sterile mutants. Plant Cell 1992, 4, 1237–1249. [Google Scholar] [CrossRef]
  80. Wei, S.J.; Chai, S.; Zhu, R.M.; Duan, C.Y.; Zhang, Y.; Li, S. HUA ENHANCER1 mediates ovule development. Front. Plant Sci. 2020, 11, 397. [Google Scholar] [CrossRef]
  81. Armenta-Medina, A.; Lepe-Soltero, D.; Xiang, D.; Datla, R.; Abreu-Goodger, C.; Gillmor, C.S. Arabidopsis thaliana miRNAs promote embryo pattern formation beginning in the zygote. Dev. Biol. 2017, 431, 145–151. [Google Scholar] [CrossRef] [PubMed]
  82. Guo, G.; Liu, X. Wheat miR9678 affects seed germination by generating phased siRNAs and modulating abscisic acid/gibberellin signaling. Plant Cell 2018, 30, 796–814. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Miao, C.; Wang, Z.; Zhang, L.; Yao, J.; Hua, K.; Liu, X.; Shi, H. The grain yield modulator miR156 regulates seed dormancy through the gibberellin pathway in rice. Nat. Commun. 2019, 10, 3822. [Google Scholar] [CrossRef] [Green Version]
  84. Fusaro, A.F.; Bocca, S.N.; Ramos, R.L.B.; Barrôco, R.M.; Magioli, C.; Jorge, V.C.; Coutinho, T.C.; Rangel-Lima, C.M.; De Rycke, R.; Inzé, D.; et al. AtGRP2, a cold-induced nucleo-cytoplasmic RNA-binding protein, has a role in flower and seed development. Planta 2007, 225, 1339–1351. [Google Scholar] [CrossRef] [PubMed]
  85. Juntawong, P.; Sorenson, R.; Bailey-Serres, J. Cold shock protein 1 chaperones mRNAs during translation in Arabidopsis thaliana. Plant J. 2013, 74, 1016–1028. [Google Scholar] [CrossRef]
  86. Karlson, D.; Imai, R. Conservation of the cold shock domain protein family in plants. Plant Physiol. 2003, 131, 12–15. [Google Scholar] [CrossRef] [Green Version]
  87. Yang, Y.; Karlson, D.T. Overexpression of AtCSP4 affects late stages of embryo development in Arabidopsis. J. Exp. Bot. 2011, 62, 2079–2091. [Google Scholar] [CrossRef] [Green Version]
  88. Nakaminami, K.; Hill, K.; Perry, S.E.; Sentoku, N.; Long, J.A.; Karlson, D.T. Arabidopsis cold shock domain proteins: Relationships to floral and silique development. J. Exp. Bot. 2009, 60, 1047–1062. [Google Scholar] [CrossRef] [Green Version]
  89. Park, S.J.; Kwak, K.J.; Oh, T.R.; Kim, Y.O.; Kang, H. Cold shock domain proteins affect seed germination and growth of Arabidopsis thaliana under abiotic stress conditions. Plant Cell Physiol. 2009, 50, 869–878. [Google Scholar] [CrossRef]
  90. Sasaki, K.; Kim, M.H.; Imai, R. Arabidopsis COLD SHOCK DOMAIN PROTEIN2 is a RNA chaperone that is regulated by cold and developmental signals. Biochem. Biophys. Res. Commun. 2007, 364, 633–638. [Google Scholar] [CrossRef]
  91. Kim, J.Y.; Park, S.J.; Jang, B.; Jung, C.H.; Ahn, S.J.; Goh, C.H.; Cho, K.; Han, O.; Kang, H. Functional characterization of a glycine-rich RNA-binding protein 2 in Arabidopsis thaliana under abiotic stress conditions. Plant J. 2007, 50, 439–451. [Google Scholar] [CrossRef] [PubMed]
  92. Kwak, K.J.; Kim, Y.O.; Kang, H. Characterization of transgenic Arabidopsis plants overexpressing GR-RBP4 under high salinity, dehydration, or cold stress. J. Exp. Bot. 2005, 56, 3007–3016. [Google Scholar] [CrossRef] [PubMed]
  93. Cao, S.; Jiang, L.; Song, S.; Jing, R.; Xu, G. AtGRP7 is involved in the regulation of abscisic acid and stress responses in Arabidopsis. Cell. Mol. Biol. Lett. 2006, 11, 526–535. [Google Scholar] [CrossRef] [PubMed]
  94. Kim, J.S.; Jung, H.J.; Lee, H.J.; Kim, K.A.; Goh, C.-H.; Woo, Y.; Oh, S.H.; Han, Y.S.; Kang, H. Glycine-rich RNA-binding protein 7 affects abiotic stress responses by regulating stomata opening and closing in Arabidopsis thaliana. Plant J. 2008, 55, 455–466. [Google Scholar] [CrossRef]
  95. Kim, W.Y.; Kim, J.Y.; Jung, H.J.; Oh, S.H.; Han, Y.S.; Kang, H. Comparative analysis of Arabidopsis zinc finger-containing glycine-rich RNA-binding proteins during cold adaptation. Plant Physiol. Biochem. 2010, 48, 866–872. [Google Scholar] [CrossRef]
  96. Wu, Z.; Zhu, D.; Lin, X.; Miao, J.; Gu, L.; Deng, X.; Yang, Q.; Sun, K.; Zhu, D.; Cao, X.; et al. RNA Binding Proteins RZ-1B and RZ-1C play critical roles in regulating pre-mRNA splicing and gene expression during development in Arabidopsis. Plant Cell 2016, 28, 55–73. [Google Scholar] [CrossRef]
  97. Kim, Y.O.; Kim, J.S.; Kang, H. Cold-inducible zinc finger-containing glycine-rich RNA-binding protein contributes to the enhancement of freezing tolerance in Arabidopsis thaliana. Plant J. 2005, 42, 890–900. [Google Scholar] [CrossRef]
  98. Kim, Y.O.; Pan, S.; Jung, C.H.; Kang, H. A zinc finger-containing glycine-rich RNA-binding protein, atRZ-1a, has a negative impact on seed germination and seedling growth of Arabidopsis thaliana under salt or drought stress conditions. Plant Cell Physiol. 2007, 48, 1170–1181. [Google Scholar] [CrossRef] [Green Version]
  99. Kim, J.Y.; Kim, W.Y.; Kwak, K.J.; Oh, S.H.; Han, Y.S.; Kang, H. Glycine-rich RNA-binding proteins are functionally conserved in Arabidopsis thaliana and Oryza sativa during cold adaptation process. J. Exp. Bot. 2010, 61, 2317–2325. [Google Scholar] [CrossRef] [Green Version]
  100. Kim, J.Y.; Kim, W.Y.; Kwak, K.J.; Oh, S.H.; Han, Y.S.; Kang, H. Zinc finger-containing glycine-rich RNA-binding protein in Oryza sativa has an RNA chaperone activity under cold stress conditions. Plant Cell Environ. 2010, 33, 759–768. [Google Scholar]
  101. Xu, T.; Gu, L.; Choi, M.J.; Kim, R.J.; Suh, M.C.; Kang, H. Comparative functional analysis of wheat (Triticum aestivum) zinc finger-containing glycine-rich RNA-binding proteins in response to abiotic stresses. PLoS ONE 2014, 9, e96877. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Lee, K.; Kang, H. Emerging roles of RNA-Binding proteins in plant growth, development, and stress responses. Mol. Cells 2016, 39, 179–185. [Google Scholar] [PubMed] [Green Version]
  103. Bryant, N.; Lloyd, J.; Sweeney, C.; Myouga, F.; Meinke, D. Identification of nuclear genes encoding chloroplast-localized proteins required for embryo development in Arabidopsis. Plant Physiol. 2011, 155, 1678–1689. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Gabotti, D.; Caporali, E.; Manzotti, P.; Persico, M.; Vigani, G.; Consonni, G. The maize pentatricopeptide repeat gene empty pericarp4 (emp4) is required for proper cellular development in vegetative tissues. Plant Sci. 2014, 223, 25–35. [Google Scholar] [CrossRef]
  105. Liu, Y.J.; Xiu, Z.H.; Meeley, R.; Tan, B.C. Empty pericarp5 encodes a pentatricopeptide repeat protein that is required for mitochondrial RNA editing and seed development in maize. Plant Cell 2013, 25, 868–883. [Google Scholar] [CrossRef] [Green Version]
  106. Sosso, D.; Canut, M.; Gendrot, G.; Dedieu, A.; Chambrier, P.; Barkan, A.; Consonni, G.; Rogowsky, P.M. PPR8522 encodes a chloroplast-targeted pentatricopeptide repeat protein necessary for maize embryogenesis and vegetative development. J. Exp. Bot. 2012, 63, 5843–5857. [Google Scholar] [CrossRef] [Green Version]
  107. Kim, S.R.; Yang, J.I.; Moon, S.; Ryu, C.H.; An, K.; Kim, K.M.; Yim, J.; An, G. Rice OGR1 encodes a pentatricopeptide repeat-DYW protein and is essential for RNA editing in mitochondria. Plant J. 2009, 59, 738–749. [Google Scholar] [CrossRef]
  108. Ding, Y.H.; Liu, N.Y.; Tang, Z.S.; Liu, J.; Yang, W.C. Arabidopsis GLUTAMINE-RICH PROTEIN23 is essential for early embryogenesis and encodes a novel nuclear PPR motif protein that interacts with RNA polymerase II subunit III. Plant Cell 2006, 18, 815–830. [Google Scholar] [CrossRef] [Green Version]
  109. Barkan, A.; Small, I. Pentatricopeptide repeat proteins in plants. Annu. Rev. Plant Biol. 2014, 65, 415–442. [Google Scholar] [CrossRef]
  110. Koussevitzky, S.; Nott, A.; Mockler, T.C.; Hong, F.; Sachetto-Martins, G.; Surpin, M.; Lim, J.; Mittler, R.; Chory, J. Signals from chloroplasts converge to regulate nuclear gene expression. Science 2007, 316, 715–719. [Google Scholar] [CrossRef]
  111. Zhao, X.; Huang, J.; Chory, J. GUN1 interacts with MORF2 to regulate plastid RNA editing during retrograde signaling. Proc. Natl. Acad. Sci. USA 2019, 116, 10162–10167. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Cottage, A.; Mott, E.K.; Kempster, J.A.; Gray, J.C. The Arabidopsis plastid-signalling mutant gun1 (genomes uncoupled1) shows altered sensitivity to sucrose and abscisic acid and alterations in early seedling development. J. Exp. Bot. 2010, 61, 3773–3786. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Murayama, M.; Hayashi, S.; Nishimura, N.; Ishide, M.; Kobayashi, K.; Yagi, Y.; Asami, T.; Nakamura, T.; Shinozaki, K.; Hirayama, T. Isolation of Arabidopsis ahg11, a weak ABA hypersensitive mutant defective in nad4 RNA editing. J. Exp. Bot. 2012, 63, 5301–5310. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Yuan, H.; Liu, D. Functional disruption of the pentatricopeptide protein SLG1 affects mitochondrial RNA editing, plant development, and responses to abiotic stresses in Arabidopsis. Plant J. 2012, 70, 432–444. [Google Scholar] [CrossRef]
  115. Zhu, Q.; Dugardeyn, J.; Zhang, C.; Takenaka, M.; Kuhn, K.; Craddock, C.; Smalle, J.; Karampelias, M.; Denecke, J.; Peters, J.; et al. SLO2, a mitochondrial pentatricopeptide repeat protein affecting several RNA editing sites, is required for energy metabolism. Plant J. 2012, 71, 836–849. [Google Scholar] [CrossRef] [Green Version]
  116. Zhu, Q.; Dugardeyn, J.; Zhang, C.; Mühlenbock, P.; Eastmond, P.J.; Valcke, R.; De Coninck, B.; Öden, S.; Karampelias, M.; Cammue, B.P.A.; et al. The Arabidopsis thaliana RNA editing factor SLO2, which affects the mitochondrial electron transport chain, participates in multiple stress and hormone responses. Mol. Plant 2014, 7, 290–310. [Google Scholar] [CrossRef] [Green Version]
  117. Lee, K.; Han, J.H.; Park, Y.I.; Colas des Francs-Small, C.; Small, I.; Kang, H. The mitochondrial pentatricopeptide repeat protein PPR19 is involved in the stabilization of NADH dehydrogenase 1 transcripts and is crucial for mitochondrial function and Arabidopsis thaliana development. New Phytol. 2017, 215, 202–216. [Google Scholar] [CrossRef] [Green Version]
  118. Liu, Y.; He, J.; Chen, Z.; Ren, X.; Hong, X.; Gong, Z. ABA overly-sensitive 5 (ABO5), encoding a pentatricopeptide repeat protein required for cis-splicing of mitochondrial nad2 intron 3, is involved in the abscisic acid response in Arabidopsis. Plant J. 2010, 63, 749–765. [Google Scholar] [CrossRef]
  119. He, J.; Duan, Y.; Hua, D.; Fan, G.; Wang, L.; Liu, Y.; Chen, Z.; Han, L.; Qu, L.J.; Gong, Z. DEXH box RNA helicase-mediated mitochondrial reactive oxygen species production in Arabidopsis mediates crosstalk between abscisic acid and auxin signaling. Plant Cell 2012, 24, 1815–1833. [Google Scholar] [CrossRef] [Green Version]
  120. Laluk, K.; Abuqamar, S.; Mengiste, T. The Arabidopsis mitochondria-localized pentatricopeptide repeat protein PGN functions in defense against necrotrophic fungi and abiotic stress tolerance. Plant Physiol. 2011, 156, 2053–2068. [Google Scholar] [CrossRef] [Green Version]
  121. Zsigmond, L.; Rigó, G.; Szarka, A.; Székely, G.; Ötvös, K.; Darula, Z.; Medzihradszky, K.F.; Koncz, C.; Koncz, Z.; Szabados, L. Arabidopsis PPR40 connects abiotic stress responses to mitochondrial electron transport. Plant Physiol. 2008, 146, 1721–1737. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Liu, J.M.; Zhao, J.Y.; Lu, P.P.; Chen, M.; Guo, C.H.; Xu, Z.S.; Ma, Y.Z. The E-subgroup pentatricopeptide repeat protein family in Arabidopsis thaliana and confirmation of the responsiveness PPR96 to abiotic stresses. Front. Plant Sci. 2016, 7, 1825. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Kobayashi, K.; Suzuki, M.; Tang, J.; Nagata, N.; Ohyama, K.; Seki, H.; Kiuchi, R.; Kaneko, Y.; Nakazawa, M.; Matsui, M.; et al. Lovastatin insensitive 1, a novel pentatricopeptide repeat protein, is a potential regulatory factor of isoprenoid biosynthesis in Arabidopsis. Plant Cell Physiol. 2007, 48, 322–331. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Sechet, J.; Roux, C.; Plessis, A.; Effroy, D.; Frey, A.; Perreau, F.; Biniek, C.; Krieger-Liszkay, A.; Macherel, D.; North, H.M.; et al. The ABA-deficiency suppressor locus HAS2 encodes the PPR protein LOI1/MEF11 involved in mitochondrial RNA editing. Mol. Plant 2015, 8, 644–656. [Google Scholar] [CrossRef] [Green Version]
  125. Bailly, C.; El-Maarouf-Bouteau, H.; Corbineau, F. From intracellular signaling networks to cell death: The dual role of reactive oxygen species in seed physiology. C. R. Biol. 2008, 331, 806–814. [Google Scholar] [CrossRef]
  126. Nonogaki, H. Seed biology updates-highlights and new discoveries in seed dormancy and germination research. Front. Plant Sci. 2017, 8, 524. [Google Scholar] [CrossRef] [Green Version]
  127. Locato, V.; Cimini, S.; De Gara, L. ROS and redox balance as multifaceted players of cross-tolerance: Epigenetic and retrograde control of gene expression. J. Exp. Bot. 2018, 69, 3373–3391. [Google Scholar] [CrossRef]
  128. Nonogaki, H. The long-standing paradox of seed dormancy unfolded? Trends Plant Sci. 2019, 24, 989–998. [Google Scholar] [CrossRef]
  129. Tan, J.; Tan, Z.; Wu, F.; Sheng, P.; Heng, Y.; Wang, X.; Ren, Y.; Wang, J.; Guo, X.; Zhang, X.; et al. A novel chloroplast-localized pentatricopeptide repeat protein involved in splicing affects chloroplast development and abiotic stress response in rice. Mol. Plant 2014, 7, 1329–1349. [Google Scholar] [CrossRef] [Green Version]
  130. Lapointe, C.P.; Wilinski, D.; Saunders, H.A.J.; Wickens, M. Protein-RNA networks revealed through covalent RNA marks. Nat. Methods 2015, 12, 1163–1170. [Google Scholar] [CrossRef] [Green Version]
  131. 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]
  132. Laloum, T.; Martín, G.; Duque, P. Alternative splicing control of abiotic stress responses. Trends Plant Sci. 2018, 23, 140–150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Zhan, X.; Qian, B.; Cao, F.; Wu, W.; Yang, L.; Guan, Q.; Gu, X.; Wang, P.; Okusolubo, T.A.; Dunn, S.L.; et al. An Arabidopsis PWI and RRM motif-containing protein is critical for pre-mRNA splicing and ABA responses. Nat. Commun. 2015, 6, 8139. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Chen, J.; Kalinowska, K.; Muller, B.; Mergner, J.; Deutzmann, R.; Schwechheimer, C.; Hammes, U.Z.; Dresselhaus, T. DiSUMO-LIKE interacts with RNA-binding proteins and affects cell-cycle progression during maize embryogenesis. Curr. Biol. 2018, 28, 1548–1560. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Structure diagram for different types of RNA-binding protein (RBP) from Arabidopsis thaliana described in this review (the left). The well-known RNA-binding domains (RBDs) and auxiliary motifs in plants (the right).
Figure 1. Structure diagram for different types of RNA-binding protein (RBP) from Arabidopsis thaliana described in this review (the left). The well-known RNA-binding domains (RBDs) and auxiliary motifs in plants (the right).
Ijms 21 06822 g001
Figure 2. The model of various RBPs involved in ABA-related signaling pathway in the Arabidopsis seed germination and/or dormancy. All PRBs are showed in pink box, the targets of RBP are in yellow box, and the blue boxes are the other components in the signaling transduction. All proteins are normal font, and mRNAs are showed with italic font. Arrows with lines or dashed lines indicate the proteins and/or genes are directly or indirectly regulated by upstream factors, respectively. Lines with bars indicate the negative regulation in the signaling. P, phosphate; PP2C, protein phosphatase 2C; SnRK2, SNF1-related protein kinases 2; RD22, responsive to dehydration 22; RAB18, ABA-responsive gene 18; ATAF1, Arabidopsis transcription activation factor 1; ABI2, the phosphatase ABA Insensitive 2; ABI4, ABA-INSENTIVE 4 (a nuclear transcriptional factor); CBF/DREB, C-repeat binding factor/dehydration-responsive element (DRE) binding proteins.
Figure 2. The model of various RBPs involved in ABA-related signaling pathway in the Arabidopsis seed germination and/or dormancy. All PRBs are showed in pink box, the targets of RBP are in yellow box, and the blue boxes are the other components in the signaling transduction. All proteins are normal font, and mRNAs are showed with italic font. Arrows with lines or dashed lines indicate the proteins and/or genes are directly or indirectly regulated by upstream factors, respectively. Lines with bars indicate the negative regulation in the signaling. P, phosphate; PP2C, protein phosphatase 2C; SnRK2, SNF1-related protein kinases 2; RD22, responsive to dehydration 22; RAB18, ABA-responsive gene 18; ATAF1, Arabidopsis transcription activation factor 1; ABI2, the phosphatase ABA Insensitive 2; ABI4, ABA-INSENTIVE 4 (a nuclear transcriptional factor); CBF/DREB, C-repeat binding factor/dehydration-responsive element (DRE) binding proteins.
Ijms 21 06822 g002

Share and Cite

MDPI and ACS Style

Lou, L.; Ding, L.; Wang, T.; Xiang, Y. Emerging Roles of RNA-Binding Proteins in Seed Development and Performance. Int. J. Mol. Sci. 2020, 21, 6822. https://doi.org/10.3390/ijms21186822

AMA Style

Lou L, Ding L, Wang T, Xiang Y. Emerging Roles of RNA-Binding Proteins in Seed Development and Performance. International Journal of Molecular Sciences. 2020; 21(18):6822. https://doi.org/10.3390/ijms21186822

Chicago/Turabian Style

Lou, Lijuan, Ling Ding, Tao Wang, and Yong Xiang. 2020. "Emerging Roles of RNA-Binding Proteins in Seed Development and Performance" International Journal of Molecular Sciences 21, no. 18: 6822. https://doi.org/10.3390/ijms21186822

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