Next Article in Journal
Interactions between Nanoclay, CTAB and Linear/Star Shaped Polymers
Next Article in Special Issue
The B-Type Cyclin CYCB1-1 Regulates Embryonic Development and Seed Size in Maize
Previous Article in Journal
Botrytis cinerea Loss and Restoration of Virulence during In Vitro Culture Follows Flux in Global DNA Methylation
Previous Article in Special Issue
How Cysteine Protease Gene PtCP5 Affects Seed Germination by Mobilizing Storage Proteins in Populus trichocarpa
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Maize PPR278 Functions in Mitochondrial RNA Splicing and Editing

1
National Engineering Laboratory for Crop Molecular Breeding, Institute of Crop Science, Chinese Academy of Agricultural Sciences, Beijing 100081, China
2
State Key Laboratory of Plant Physiology and Biochemistry, China Agricultural University, Beijing 100193, China
3
National Maize Improvement Center, Department of Plant Genetics and Breeding, China Agricultural University, Beijing 100193, China
4
Development Center of Science and Technology, Ministry of Agriculture and Rural Affairs, Beijing 100176, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2022, 23(6), 3035; https://doi.org/10.3390/ijms23063035
Submission received: 28 January 2022 / Revised: 3 March 2022 / Accepted: 8 March 2022 / Published: 11 March 2022
(This article belongs to the Special Issue Molecular Biology of Plant Seed Development)

Abstract

:
Pentatricopeptide repeat (PPR) proteins are a large protein family in higher plants and play important roles during seed development. Most reported PPR proteins function in mitochondria. However, some PPR proteins localize to more than one organelle; functional characterization of these proteins remains limited in maize (Zea mays L.). Here, we cloned and analyzed the function of a P-subfamily PPR protein, PPR278. Loss-function of PPR278 led to a lower germination rate and other defects at the seedling stage, as well as smaller kernels compared to the wild type. PPR278 was expressed in all investigated tissues. Furthermore, we determined that PPR278 is involved in the splicing of two mitochondrial transcripts (nad2 intron 4 and nad5 introns 1 and 4), as well as RNA editing of C-to-U sites in 10 mitochondrial transcripts. PPR278 localized to the nucleus, implying that it may function as a transcriptional regulator during seed development. Our data indicate that PPR278 is involved in maize seed development via intron splicing and RNA editing in mitochondria and has potential regulatory roles in the nucleus.

1. Introduction

Pentatricopeptide repeat (PPR) proteins are widely distributed in plants, with more than 400 members in maize (Zea mays L.) [1], which were also divided into two subfamilies: The P subfamily containing motif P only, and the PLS subfamily harboring the P, L, and S motifs [2,3]. PLS-subfamily members were further classified into PLS, E, E+, and DYW types according to their C-terminal domain [4,5]. PPR proteins have been isolated from species including maize, rice (Oryza sativa L.), and Arabidopsis thaliana L., and functional analysis revealed that they could bind RNA and might play important roles in RNA metabolism, including RNA splicing and editing, cleavage, and maturation [3].
Ongoing work has shown that PPR proteins splice 22 group II introns present within 8 protein-coding genes in mitochondria in maize, including NADH dehydrogenase subunits1, 2, 4, 5, and 7 (nad1, 2, 4, 5, 7), and subunits of a complex involved in the biogenesis of cytochrome c: subunits F (ccmFC), cytochrome c oxidase subunits2 (cox2), and ribosomal protein s3 (rps3) [6,7,8]. Among these spliced genes, nad1 introns 1, 3, and 4; nad2 intron 2; and nad5 introns 2 and 3 are post-transcriptionally modified via trans-splicing, while other introns of target spliced transcripts are modified through cis-splicing [6]. Loss-function of PPR proteins and/or other spliceosomes results in defects in seed and/or seedling development, such as an empty pericarp in the seed, decreased 100-kernel weight, and seedling death, due to abnormal intron splicing of mitochondrial genes, which leads to abnormal assembly of mitochondrial complex Ⅰ and insufficient production of energy molecules [9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26]. These findings highlight the importance of PPR-based RNA splicing in mitochondria for seed development in maize.
Cytidine-to-uridine (C-to-U) RNA editing in mitochondria is an important function of PPR proteins and is required for correcting mutations at the RNA level by recognizing specific transcripts through tandem repeat motifs [27,28]. The DYW domain at the C terminus of some PPR proteins confers catalytic activity during editing; PPR proteins that lack this domain may recruit co-factors and form a larger complex to perform editing functions [4,29]. Indeed, studies in Arabidopsis, rice, and maize have confirmed that PLS-type PPR proteins function as cytidine editors in transcripts of mitochondrial and plastid genes [30]. For example, mutations in Empty pericarp 5 (EMP5), EMP7, EMP9, EMP18, EMP21, Defective kernel 10 (DEK10), DEK36, DEK39, PPR2263, Small kernel 1 (SMK1), SMK4, and SMK6 significantly altered the C-to-U editing efficiency in mitochondrial RNAs, leading to small or defective kernels [12,13,31,32,33,34,35,36,37,38,39,40,41].
More than 99% of PPR proteins across the Arabidopsis genome are predicted to localize in mitochondria and/or plastids, and very few localize in both nuclei and mitochondria [4,5,42]. SUPPRESSOR OF THE ABSCISIC ACID RECEPTOR OVEREXPRESSOR1 (SOAR1, a native regulator of abscisic acid signaling) and GLUTAMINE-RICH PROTEIN23 (GRP23) localize in both nuclei and mitochondria. GRP23 interacts with RNA polymerase II subunit III and PPR PROTEIN LOCALIZED TO NUCLEI AND MITOCHONDRIA1 (PNM1), as well as with the plant-specific Teosinte Branched 1/Cycloidea/Proliferating Cell Factor (TCP) family transcription factors TCP8 and nucleosome assembly protein (NAP1) in nuclei, indicating that this PPR protein likely functions as a regulator in nuclei [43,44,45]. In addition, rice NUCLEAR-LOCALIZED PPR PROTEIN1 (OsNPPR1) and OsNPPR3 also target to the nucleus, where they affect pre-mRNA alternative splicing of nuclear genes related to mitochondrial function [46,47]. Furthermore, DEK43 targets to mitochondria and nuclei, but its functions in the nuclei remain to be investigated [19]. Thus, further study is needed to uncover the functions of PPR proteins with multiple-organelle localization.
In this study, we isolated PPR278, a P-subfamily PPR protein with dual localization in nuclei and cytoplasm of maize. We discovered that splicing of nad2 and nad5 introns and C-to-U editing of 10 mitochondrial transcripts were affected in the ppr278 mutants. RNA-sequencing (RNA-seq) analysis revealed that the expression of nucleus-encoded genes associated with RNA splicing and editing was significantly altered in the ppr278 mutants. Our findings demonstrate that PPR278 affects seed development via effects on RNA intron splicing and editing in mitochondria and on gene expression in nuclei, expanding our limited knowledge about the function of P-subfamily PPR proteins that are capable of splicing introns.

2. Results

2.1. The ppr278-1 Mutants Cause Aberrant Seed Development in Maize

We isolated a kernel mutant from our mutant libraries established by an ethyl methanesulfonate (EMS)-based method, which produced smaller seeds compared to the wild type (WT) (Figure 1A–D), and named it ppr278-1 based on similar phenotypes with other ppr mutants in maize. Only ~5% of mutant kernels could germinate normally on average and grew to the seedling stage (Figure 1E and Figure S1A,B). However, the vigor of mutant seedlings was notably decreased, and homozygous mutant seedlings did not survive to the reproductive stage (Figure S1B). The 100-kernel weight of ppr278 mutant plants was significantly lower than that of WT, which might be due to developmental defects in mutant kernels or disruption of seed-filling processes (Figure 1F,G). The segregation ratio of seeds in mutant ears was 3:1 (normal kernels: mutant kernels, 212:69), implying that the ppr278-1 phenotype is controlled by a single recessive locus (Figure 1A,B).

2.2. PPR278 Encodes a Constitutively Expressed PPR Protein

We used bulked segregate analysis-RNA sequencing (BSR) to isolate PPR278. Endosperm and embryos from M3 generation kernels at 12 days after pollination (DAP) were used to construct RNA-seq libraries of WT and mutant-type pools. High-quality single nucleotide polymorphism (SNP) markers were called by aligning RNA-seq reads to the B73 reference genome (V4). We identified 32 candidate genes including 14 missense variants, 10 synonymous variants, 5 3′ untranslated region (UTR) variants, 2 5′UTR variants, and 1 stop codon-gained variant (Zm00001d015156) (Table S1). Zm00001d015156 (located on chromosome 5) that had the highest ΔSNP index among missense variants and the stop codon-gained variant [48]; therefore, we determined Zm00001d015156 as the candidate gene corresponding to PPR278. Sequence analysis revealed that Zm00001d015156 had one SNP (C-to-T mutation) in the ppr278-1 mutant. This SNP was located in the 15th exon at +2269 bp from the ATG start codon, causing a premature stop codon in Zm00001d015156 (Figure 2A). To test whether the mutation in Zm00001d015156 causes the ppr278-1 phenotypes, we performed co-segregation analysis by genotyping a larger segregating population (281 kernels) and observed that the SNP in the ppr278-1 mutant co-segregated with the low seed germination phenotype (Figure S2). To further test the candidate gene, we screened another mutant (hereafter referred to as ppr278-2) from the mutant libraries, which showed a similar phenotype as the ppr278-1 mutant (Figures S1A and S3). We identified a SNP (C-to-T mutation) at +2404 bp in the 15th exon of Zm00001d015156 in the ppr278-2 mutant, also resulting in a premature stop codon (Figure 2A). Crossing, self-pollinating, and subsequent genetic analyses confirmed ppr278-1 and ppr278-2 as allelic mutants, revealing that Zm00001d015156 is the causal gene for PPR278 (Figure 2B). Expression analysis showed that PPR278 was expressed in multiple tissues, with the highest expression level in 6-DAP kernels (Figure 2D). Thus, we cloned PPR278, encoding a constitutively expressed PPR protein, and confirmed that loss-of-function alleles of PPR278 caused the abnormal phenotype mentioned above.

2.3. PPR278 Belongs to the p-Subfamily and Dually Localizes in Cytoplasm and Nucleus

PPR protein sequences were downloaded from either NCBI (https://www.ncbi.nlm.nih.gov/) (Accessed on 9 March 2022) or PPR energy (https://ppr.plantenergy.uwa.edu.au/) (Accessed on 9 March 2022) databases, and full-length protein sequences were used for phylogenetic analysis, in which PPR278 was clustered into the P-subfamily (Figure S4A) [49,50]. The prediction of PPR energy showed that PPR278 and its Arabidopsis homolog AT4G21880.1 grouped into the P-subfamily clade [42]. Further analysis indicated that PPR278 contained 11 p motifs (Figure 2C, Figures S4B and S5), and amino acid numbers of these motifs ranged from 32 to 37, which differed from typical P-subfamily PPR proteins (Figure S6).
Most PPR proteins target to mitochondria and/or plastids in plants, whereas only a few are predicted to localize in nuclei [4,5,42]. Based on Csbio (http://www.csbio.sjtu.edu.cn/bioinf/Cell-PLoc-2/) (Accessed on 9 March 2022) and Cello (http://cello.life.nctu.edu.tw/) (Accessed on 9 March 2022), PPR278 was predicted to be in the nucleus and cytoplasm (Figure S2C). Subcellular localization experiments using onion (Allium cepa L.) epidermal cells and maize protoplasts revealed that PPR278 targeted to both nuclei and cytoplasm, suggesting that the functions of PPR278 are more complex than those of other PPR proteins targeted to mitochondria or/and plastids (Figure 3A,B). Moreover, the secondary structure of PPR278 consists of numerous paired anti-parallel alpha helices, as predicted by the AlphaFold Protein Structure Database (http://www.alphafold.ebi.ac.uk) (Accessed on 9 March 2022) (Figure 3C), which form the helix-turn-helix motifs necessary for RNA binding [5,51]. Therefore, even though PPR278 is a P-subfamily member, it contains 11 non-canonical P-type PPR repeats, which is a special feature. Another feature of PPR278 is that it is double-targeted in mitochondria and the nucleus. Like other PPR proteins, a series of paired anti-parallel alpha helices of PPR278 provide the ability to bind specific target RNA.

2.4. PPR278 Is Essential for the Cis-Splicing of nad2 Intron 4 and nad5 Introns 1 and 4

PPR proteins are RNA-binding proteins involved in RNA metabolism, including intron splicing and RNA editing, in mitochondria and plastids [3]. To test whether PPR278 is involved in cis-splicing in mitochondria, we performed semi-quantitative reverse transcription PCR (RT-PCR) analysis between mutant and WT plants for 35 mitochondrial genes using 12-DAP kernels. We observed expression changes in four genes, rps2A, nad5, ccmFN, and apocytochrome b (cob), between ppr278-1 and WT plants (Figure 4A). Additionally, RT-PCR analysis showed that the cis-splicing efficiency of nad2 intron 4 (1383 bp) and nad5 introns 1 (870 bp) and 4 (952 bp) was affected in the ppr278-1 and ppr278-2 mutants (Figure 4B,C and Figure S7A). We detected no differences at the transcript level or in cis-splicing for the other mitochondrial genes examined, including 25 tRNAs and 3 rRNAs between ppr278 and WT plants (Figure S7B). PPR278 is not only involved in mitochondrial gene expression (rps2A, nad5, ccmFN, and cob), but it is also necessary for the cis-splicing of nad2 intron 4 and nad5 introns 1 and 4.

2.5. PPR278 Is Required for C-to-U RNA Editing in 10 Mitochondrial Transcripts

RNA editing is a major function of PPR proteins in mitochondria, and nearly 500 C-to-U RNA editing sites in mitochondrion-encoded transcripts were reported in the maize inbred line B73 [35]. Therefore, we examined the editing profiles of the 35 mitochondrion-encoded transcripts in the ppr278 mutants and identified 94 C-to-U RNA editing events (Table 1, Figures S8–S10). We analyzed the C-to-U RNA editing efficiency by directly comparing editing profiles in WT and ppr278 plants and observed decreased efficiency of 67 C-to-U RNA editing sites in seven mitochondrial transcripts and increased editing efficiency in one editing site in rps2B-550 in the ppr278 mutant (Table 1, Figures S8–S10). However, we could not determine the editing efficiency of the remaining 26 C-to-U editing sites in atp8 and nad3 via the peak of C and/or U (Figure 5A and Figure S9A). To understand how atp8 and nad3 are affected in the ppr278 mutant, we sequenced ~40 independent clones from WT and ppr278 plants and analyzed editing efficiency (T/(T + C)%). The T/(T + C) ratio was reduced among all editing sites in atp8 and nad3 in the ppr278 mutant (Figure 5B and Figure S9B). These results indicated that PPR278 is required for editing 94 C-to-U sites of mitochondrial transcripts related to mitochondrial complexes Ⅰ, Ⅳ, and Ⅴ, cytochrome c biogenesis, and ribosomal proteins (Figure S11A).
To further investigate the effects of aberrant C-to-U RNA editing on protein translation, we investigated the coding sequence and corresponding amino acid sequences of all 35 mitochondrion-encoded transcripts in the ppr278 mutant. Nearly 83.16% of C-to-U RNA editing events caused amino acid substitutions (Figure S11A). Notably, the replacement of leucine accounted for more than 40% of all editing events, ranking highest in silenced and non-silenced sites (Figure S11B) [35]. We also discovered that editing failed at 37 bases distributed among atp1, atp4, and ccmFN in the ppr278-1 and ppr278-2 mutants (Table 1). We analyzed the failed editing at these bases in the ppr278 mutants and observed that editing at all bases except for ccmFN-417 and atp1-30 caused amino acid changes. The other failed editing sites in ccmFN resulted in amino acid substitutions at positions 11, 9, 4, 4, 3, 2, and 1, corresponding to Leu-to-Pro, Leu-to-Ser, Try-to-Arg, Phe-to-Ser, Ser-to-Pro, Cys-to-Arg, and Phe-to-Leu, respectively (Figure S10, Table 1). ccmF in Escherichia coli, the homolog of maize ccmFN, encodes a large protein with 11 transmembrane domains [52]. We used TMHMM (http://services.healthtech.dtu.dk/service.php?TMHMM-2.0) (Accessed on 9 March 2022) to analyze the transmembrane domains of ccmFN and determined that the lack of editing in ccmFN disrupted the 1st, 4th, 5th, and 6th transmembrane domains in the ppr278 mutant (Figure S10C). We also analyzed the amino acid conservation of C-to-U editing sites of atp8, nad3, and ccmFN, which indicated that these sites were relatively conserved in terms of amino acids change (Figure 5, Figures S9C and S10D). In total, PPR278 is essential for C-to-U RNA editing in 10 mitochondrial transcripts, including 37 abolished sites in atp1, atp4, and ccmFN. Loss of function of PPR278 lead to a series of amino acid substitution events due to failed C-to-U RNA editing.

2.6. Mutation of PPR278 Alters the Expression of Mitochondrial Function-Related Genes

Considering the nuclear localization of PPR278 and its implication in nuclear gene regulation, we performed RNA-seq analysis using 18-DAP endosperms and embryos from ppr278-1 and WT plants. The materials were collected from segregating ears with two biological replicates (Figure 6A). Those with a fold change and p-value above the threshold (fold change > 2 and p-value < 0.05) were identified as differentially expressed genes (DEGs). We identified 2466 DEGs between ppr278-1 and WT plants, including 1542 up-regulated DEGs and 924 down-regulated DEGs (Figure 6B, Tables S3 and S4).
Gene Ontology (GO; http://systemsbiology.cau.edu.cn/agriGOv2/) (Accessed on 9 March 2022) analysis using the DEGs indicated that both down- and up-regulated DEGs were enriched in the term ‘post-embryonic development’ (GO: 0009791). The reason why this Go term was enriched might be interpreted as genes transcribed actively to regulate post-embryonic development rather than direct results of loss-function-of PPR278. The other down-regulated DEGs were mostly associated with energy-consuming processes, such as ‘metabolic process’, ‘nutrient reservoir activity’, ‘plant and seed development’, and ‘nucleosome assembly’ [12,13], suggesting that energy production in the ppr278-1 mutant is insufficient to maintain plant growth and seed development. The remaining up-regulated DEGs were closely related to ‘mitochondrial part’ (GO: 0044429), which is predicted to function in mitochondria. For example, ‘electron transport chain’ (GO:0022900), ‘mitochondrial respiratory chain complex Ⅰ’ (GO:0005747), ‘II’ (GO: 005749), and ‘III’ (GO:0005750), ‘ATP synthase activity’ (GO:0000276, GO:0042775, GO:0005753), ‘mitochondrial respiratory chain complex assembly’ (GO:0033108), ‘precatalytic spliceosome’ (GO: 0071011), ‘U1 small nuclear ribonucleoprotein particle (snRNP)’(GO:0005685), ‘U2 snRNP’ (GO:0005686), ‘U4/U6×U5 tri-snRNP complex’ (GO:0046540), and ‘pre-catalytic spliceosome’ (GO:0071011) (Figure 6C, Table S4) were enriched in ppr278. These findings demonstrated that the ppr278 mutation significantly altered the expression of genes putatively involved in mitochondrial processes, leading to impaired mitochondrial function and retarded seed development.

3. Discussion

3.1. PPR278 Functions in RNA Cis-Splicing and RNA Editing

P-subfamily PPR proteins are generally involved in RNA splicing [4]. However, three PPR proteins of the P-subfamily function in C-to-U RNA editing in Arabidopsis. For example, PPR596 influences editing efficiency at partial editing sites (rps3eU64RW, rps3eU603FF, rps3eU887SL, rps3eU1344SS, rps3eU1352PL, rps3eU1470SS, rps3eU1534RC, rps3eU1571AV, rps3eU1580SF, rps3eU1598SL, Pseudo-rps14eU99, and Pseudo-rps14Eu194) in mitochondrial transcript rps3 and rps14, while another gene P-type protein, PPR-MODULATING EDITING (PPME), is responsible for C-to-U RNA editing at both sites nad1-898 and nad1-937 [53,54,55]. Here, we characterized PPR278 as a splicing factor that participates in intron splicing of nad2 intron 4 and nad5 introns 1 and 4. In addition, PPR278 was also recognized as an editing factor that is responsible for 94 sites of C-to-U RNA editing in mitochondrial transcripts.
PPR278 is necessary for intron splicing of nad2 intron 4 and nad5 introns 1 and 4, which further highlights the role of P-subfamily PPR proteins in intron splicing. Multiple PPR proteins and other splicing factors may participate in splicing the same intron. For instance, nad2 intron 4 cis-splicing was also abnormal in emp16 and emp12 mutants in maize [9,17]. ZmSMK9 is involved in the cis-splicing of nad5 introns 1 and 4 [18]. Furthermore, EMP25, PPR101, and PPR231 are required for splicing nad5 intron 1 in maize. Thus, mutations in these PPR proteins affect the cis-splicing of mitochondrial nad5 introns [18,21,22]. Similarly, EMP11, DEK2, and EMP8 are associated with nad1 intron splicing [12,13,14,15,16]. In addition, EMP10 and DEK37 are necessary for cis-splicing of nad2 intron 1 [11,56]. These findings indicate that, in its function as a spicing factor, PPR278 may function in spliceosome complexes in mitochondria, possibly in complexes with these other factors that also affect splicing of the same introns.
In addition to its splicing role–and unlike previously reported P-type PPR proteins–PPR278 functions in C-to-U RNA editing in genes associated with other mitochondrial complexes besides complex III (Figure S11A), particularly in ccmFN with 30 abolished C-to-U RNA editing sites in the ppr278 mutant (Figure S10, Table 1). Failed editing of ccmFN abolished the formation of several transmembrane domains (Figure S10C), especially the formation of the 5th transmembrane domain, which is important for protein conformation [32]. ccmFN-302 and ccmFN-1553 are essential for C-to-U editing and the biogenesis of cytochrome c (Figure S10A) [31,32,36]. Additionally, a conserved WWD domain (G [2X] W [2 or 4X] WG [2X] WXWD) consisting of two W-to-R substitutions in sites ccmFN-1357 and ccmFN-1375 is required for accessibility for heme b attachment to cytochromes c1 and c [26,32]. Another conserved histidine (H) in site ccmFN-1342, called PHis2 involved in heme binding, was also fully unedited in the ppr278 mutant [57,58] (Figure S10B). Moreover, atp1-1292 and atp4-59 sites are important for the assembly of mitochondrial complex Ⅴ [31,32,35,36]. Thus, loss of PPR278 function led to 37 abolished C-to-U RNA editing sites and an abnormal mitochondrial complex in maize (Table 1).
In general, P-subfamily members primarily function in splicing and PLS-subfamily members primarily function in editing. PPR278 is the first P-subfamily member determined to be involved in intron splicing of mitochondrial transcripts and C-to-U RNA editing in maize. However, several other PPR proteins act simultaneously as splicing and editing factors in plants. The E-type PPR protein DEK55 plays a crucial role in RNA editing at multiple sites as well as in the splicing of nad1 and nad4 introns in maize, and the DYW-type PPR protein BLX (Baili Xi) controlling mitochondrial RNA editing and splicing is essential for seed development in Arabidopsis [15,16,19,59]. These discoveries have expanded our knowledge about PPR proteins: The function of a PPR protein is not dependent on its classification category. There is a distinguishable divergence function in P-subfamily and PLS-subfamily members, and PPR278 clustered in the P-family by phylogenetic analysis (Figure S4A). The irregular motif p and classification of PPR278 provide a plausible explanation for why PPR278 has both splicing and editing functions in mitochondria. Notably, no C-to-U editing sites were detected in nad2 and nad5 cDNA sequences in the ppr278 mutant, suggesting that nad2 intron 4 and nad5 introns 1 and 4 have no effect on the editing of the exon sequence. This implied that PPR278 might conduct C-to-U RNA editing and intron splicing independently [9].

3.2. PPR278 Is Involved in Regulation of Nuclear Gene Expression

Several nucleus-localized PPR proteins are involved in gene regulation [4,5,42]. These nucleus-localized PPR proteins (SOAR1, GRP23, PNM1, OsNPPR1, and OsNPPR3) interact with transcription factors and/or RNA polymerases to regulate the transcription and/or mRNA processing of nuclear genes associated with mitochondrial function or localized to mitochondria, and likely function as potential regulators of gene expression during embryogenesis in plants [43,44,45,46,47]. In maize, DEK43 targets to nuclei, but its functions in nuclei remain largely unknown [19]. In our study, PPR278 localized to nuclei (Figure 3A,B), and RNA-seq analysis showed that the expression of many genes was significantly changed in the ppr278-1 mutant (Figure 6B).
Genes related to post-embryonic development were dramatically down- and up-regulated in ppr278 mutants (Figure 6C). Moreover, genes associated with plant and seed development were widely down-regulated in the ppr278-1 mutant. This down-regulation might explain the lower germination rate and seedling-lethal phenotype of the ppr278 mutant. Notably, GO terms GO:0006334 (nucleosome assembly essential for a series of energy-consuming biological processes [60]) and GO:0045735 (nutrient reservoir activity associated with kernel development and filling) were also significantly enriched in the dek2, dek10, dek35, dek37, and dek41 mutants, characterized by defective kernels and abnormal mitochondrial function [10,11,12,13,61]. The down-regulated genes in ppr278-1 plants were enriched for both kinds of these GO terms. In addition, 7 and 23 DEGs were highly associated with the GO terms ‘spliceosome’ and ‘mitochondrial respiratory chain’, respectively, and exhibited extensive up-regulation (Figure 6C, Table S4). This enrichment suggests that the mitochondrial oxidation respiratory chain was severely disrupted in ppr278 mutants. It remains to be tested whether the altered expression of nucleus-encoded genes arises from PPR278 binding their transcripts directly in the nucleus or if it is a consequence of arrested mitochondrial function. Based on our results, we conclude that PPR278 has a precise regulatory role in the expression of mitochondrial and nuclear genes and might function in transcription and processing of nuclear mRNA.

4. Materials and Methods

4.1. Plant Materials

The maize (Zea mays L.) ppr278-1 and ppr278-2 mutants were obtained from our EMS-based mutant library. In brief, the mature pollen of inbred line B73 was treated with 0.015% ethyl methanesulfonate solution (vol/vol) in a 1:10 pollen:solution ratio for 30 min. The resulting M1 plants were self-pollinated, and M2 seeds showing a Mendelian segregation ratio were selected. M2 plants were further self-pollinated to obtain M3 progenies for bulked pools. The 12-DAP embryos and endosperms from 20 WT and 20 mutant-type kernels from segregating M3 ears were used for bulked segregant analysis. The plants were grown in the experimental field at Shangzhuang Experimental Base of China Agricultural University under natural conditions (Beijing, China).

4.2. Bulked Segregant Analysis

The 12-DAP embryos and endosperms samples were ground in liquid nitrogen, and total RNA was extracted with TRIzol Reagent (Thermo Fisher SCIENTIFIC, 15596018, Waltham, MA, USA) according to the manufacturer’s instructions. The cDNA library was prepared with VAHTS Stranded mRNA-seq Library Prep Kit for Illumina V2 (Vazyme, NR612, Nanjing, China) and sequenced on the Illumina X-ten platform. Raw reads were aligned to the B73 reference genome (V4) using HISAT2. The alignment files were converted to BAM format and sorted using Samtools. SNP calling, including C-to-T and G-to-A mutations, was performed by Samtools and Bcftools using unique reads. High-confidence SNPs were selected using the following criteria: (1) the SNP was supported by at least five reads, and (2) the SNP was detected in both the mutant pool and its corresponding WT pool. Functional annotation of SNPs was performed using SnpEff to screen SNPs between the two RNA pools. The SNP index was listed in Table S1.

4.3. Candidate Gene Analysis and Validation

We identified 14 genes with missense mutations, 10 genes with synonymous mutations, 5 genes with 3ʹUTR mutations, 2 genes with 5ʹUTR mutations, and 1 gene with stop codon-gained mutation (Zm00001d015156) in WT and mutant pools from our bulked segregant analysis. Then the ΔSNP index (SNP index difference between the mutant bulk and WT for each SNP) was calculated for each SNP. Zm00001d015156 was considered as candidate gene because of its mutant type and the highest SNP index among these genes. To test the degree of linkage between the candidate gene Zm00001d015156 and the mutant phenotype in the M3 segregating population, we designed primers PPR278-1F/R and PPR278-2F/R closely linked with the mutant sites (Table S5). Next, we used heterozygous ppr278-1 and ppr278-2 plants for reciprocal pollination followed by genetic analysis using the kernels from the pollinated ears (Figure 2B).

4.4. Subcellular Localization of PPR278

The full-length PPR278 coding sequence without the stop codon was inserted into the vectors pSuper1300::eGFP and pSPYNE-35S::eGFP. Approximately 1 µg of the recombinant plasmid PPR278::GFP was transformed into onion (Allium cepa L.) epidermal cells and maize protoplasts. After culture for 20 h at 28 °C, the fluorescence signals were detected by a LSM710 confocal laser scanning microscope (Carl Zeiss, Oberkochen, Germany). AtAHL22-mCherry was used as a nuclear marker [62].

4.5. Phylogenetic Analysis

Sequences of PPR278 and its orthologous proteins were downloaded from either NCBI (Bethesda, MD, USA) or PPR energy databases (Perth, Australia) [49,50]. The full-length protein sequences were used to construct a phylogenetic tree with the neighbor-joining method using Mega-X software.

4.6. RT-PCR and qRT-PCR

Total RNA was treated with RQ1-free DNase I (Promega, M610A, Madison, WI, USA) to completely remove DNA contamination. The cDNA was synthesized using oligo (dT) primers (Promega, C1181, Madison, WI, USA). Quantitative reverse transcription PCR (qRT-PCR) was performed using TB Green Premix Ex Taq (Takara, RR420A, Toyobo, Osaka, Japan) with an ABI 7500 system (Thermo Fisher SCIENTIFIC, Waltham, MA, USA), and relative gene expression was calculated with the 2ΔΔCt method as described previously [63]. The data were normalized using ZmActin as a control. The experiments were replicated two times, and the primers used in the qRT-PCR (PPR278-764F and PPR278-1003R) are listed in Table S5.

4.7. Analysis of RNA Splicing by PPR278

Total RNA was extracted and reverse transcribed with random primers as described above. The expression differences of maize mitochondrial transcripts between WT and ppr278-1 plants were compared by PCR using 18s RNA as the reference gene. The primers used for mitochondrial genes were described previously [31]. The expression levels in WT and ppr278-1 plants were normalized using ZmActin. The PCR reaction buffer and DNA polymerase were from Mei5bio (MF743-10, Beijing, China). To further analyze RNA splicing, we chose the primers covering group II introns as described previously [9].

4.8. Analysis of RNA Editing by PPR278

Mitochondrial RNA was reverse transcribed with random primers as described above. The full-length sequences of 35 mitochondrial transcripts in ppr278-1 and ppr278-2 kernels at 12-DAP were reverse transcribed and directly sequenced as described previously [31]. The C-to-U RNA editing status of mitochondrial transcripts from WT and ppr278-1 plants was compared. The editing sites identified in the ppr278-1 mutant were confirmed in the ppr278-2 mutant. The editing efficiency of atp8 and nad3 in WT, ppr278-1, and ppr278-2 plants was estimated by cloning the atp8 and nad3 coding sequences into vector B-zero (TransGen Biotech pEASY-Blunt Zero Cloning Kit, CB501-01, Beijing, China) and sequencing more than 40 independent colonies for each gene.

4.9. RNA-Sequencing Analysis

Total RNA of 18-DAP endosperm and embryo (4 µg) tissues was extracted as described above, and two ppr278-1 or WT biological samples were collected for construction of the RNA-sequencing library. The cDNA library was performed according to Hieff NGS MaxUp II Dual-mode mRNA Library Prep Kit for Illumina (Yeasen, H9001360, Beijing, China) and sequenced by the Illumina NovaSeq platform for 150-nucleotide paired-end reads. Raw reads were aligned to the B73 reference genome (V4) using HISAT2. Data were normalized as fragments per kilobase of exon per million fragments mapped (FPKM), since the sensitivity of RNA-seq depends on the transcript length. The significant DEGs were filtered with the criterion: fold change > 2 and p-value < 0.05.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms23063035/s1.

Author Contributions

X.L., J.L. (Jinsheng Lai) and W.S. conceived this project. Y.C. screened this mutant ppr278, performed the allelic test and wrote the corresponding paragraphs. X.Z. analyzed the BSR data. Z.Y. amplificated gene Zm00001d015156. J.Y. performed other experiments and wrote the manuscript. J.L. (Jinsheng Lai), W.S., J.L. (Jingang Liang) and X.L. revised the manuscript. No conflict of interest declared. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Natural Science Foundation of China (31971957, 91935303, 91935305).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare that they have no conflict of interest.

References

  1. O’Toole, N.; Hattori, M.; Andres, C.; Iida, K.; Lurin, C.; Schmitz-linneweber, C.; Sugita, M.; Small, I. On the expansion of the pentatricopeptide repeat gene family in plants. Mol. Biol. Evol. 2008, 25, 1120–1128. [Google Scholar] [CrossRef] [PubMed]
  2. Barkan, A.; Rojas, M.; Fujii, S.; Yap, A.; Chong, Y.S.; Bond, C.S.; Small, I. A combinatorial amino acid code for RNA recognition by pentatricopeptide repeat proteins. PLoS Genet. 2012, 8, e1002910. [Google Scholar] [CrossRef] [PubMed]
  3. Barkan, A.; Small, I. Pentatricopeptide repeat proteins in plants. Annu. Rev. Plant Biol. 2014, 65, 415–442. [Google Scholar] [CrossRef] [PubMed]
  4. Lurin, C.; Andres, C.; Aubourg, S.; Bellaoui, M.; Bitton, F.; Bruyere, 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] [PubMed] [Green Version]
  5. Schmitz-linneweber, C.; Small, I. Pentatricopeptide repeat proteins: A socket set for organelle gene expression. Trends Plant Sci. 2008, 13, 663–670. [Google Scholar] [CrossRef]
  6. Clifton, S.W.; Minx, P.; Fauron, C.M.; Gibson, M.; Allen, J.O.; Sun, H.; Thompson, M.; Barbazuk, W.B.; Kanuganti, S.; Tayloe, C.; et al. Sequence and comparative analysis of the maize NB mitochondrial genome. Plant Physiol. 2004, 136, 3486–3503. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Brown, G.G.; Francs-Small, C.C.D.; Ostersetzer-Biran, O. Group II intron splicing factors in plant mitochondria. Front. Plant Sci. 2014, 5, 35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Fan, K.J.; Ren, Z.J.; Zhang, X.F.; Liu, Y.; Fu, J.J.; Qi, C.L.; Tatar, W.; Rasmusson, A.G.; Wang, G.Y.; Liu, Y.J. The pentatricopeptide repeat protein EMP603 is required for the splicing of mitochondrial nad1 intron 2 and seed development in maize. J. Exp. Bot. 2020, 72, 6933–6948. [Google Scholar] [CrossRef]
  9. Xiu, Z.; Sun, F.; Shen, Y.; Zhang, X.; Jiang, R.; Bonnard, G.; Zhang, J.; Tan, B.C. Empty pericarp 16 is required for mitochondrial nad2 intron 4 cis-splicing.; complex I assembly and seed development in maize. Plant J. 2016, 85, 507–519. [Google Scholar] [CrossRef]
  10. Chen, X.Z.; Feng, F.; Qi, W.W.; Xu, L.M.; Yao, D.S.; Wang, Q.; Song, R.T. Dek35 encodes a PPR protein that affects cis-splicing of mitochondrial nad4 intron 1 and seed development in maize. Mol. Plant 2017, 10, 427–441. [Google Scholar] [CrossRef] [Green Version]
  11. Dai, D.W.; Luan, S.C.; Chen, X.Z.; Wang, Q.; Feng, Y.; Zhu, C.G.; Qi, W.W.; Song, R.T. Maize Dek37 encodes a P-type PPR protein that affects cis-splicing of mitochondrial nad2 intron 1 and seed development. Genetics 2018, 208, 1069–1082. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Qi, W.; Tian, Z.; Lu, L.; Chen, X.; Chen, X.; Zhang, W.; Song, R.T. Editing of mitochondrial transcripts nad3 and cox2 by Dek10 is essential for mitochondrial function and maize plant development. Genetics 2017, 205, 1489–1501. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Qi, W.; Yang, Y.; Feng, X.; Zhang, M.; Song, R.T. Mitochondrial function and maize kernel development requires Dek2.; a pentatricopeptide repeat protein involved in nad1 mRNA splicing. Genetics 2017, 205, 239–249. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Ren, X.M.; Pan, Z.Y.; Zhao, H.L.; Zhao, J.L.; Cai, M.J.; Li, J.; Zhang, Z.X.; Qiu, F.Z. Empty pericarp11 serves as a factor for splicing of mitochondrial nad1 intron and is required to ensure proper seed development in maize. J. Exp. Bot. 2017, 68, 4571–4581. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Sun, F.; Zhang, X.Y.; Shen, Y.; Wang, H.C.; Liu, R.; Wang, X.M.; Gao, D.H.; Yang, Y.Z.; Liu, Y.W.; Tan, B.C. The pentatricopeptide repeat protein empty pericarp8 is required for the splicing of three mitochondrial introns and seed development in maize. Plant J. 2018, 95, 919–932. [Google Scholar] [CrossRef] [PubMed]
  16. Sun, Y.; Huang, J.Y.; Zhong, S.; Gu, H.Y.; He, S.; Qu, L.J. Novel DYW-type pentatricopeptide repeat (PPR) protein BLX controls mitochondrial RNA editing and splicing essential for early seed development of Arabidopsis. J. Genet. Genom. 2018, 45, 155–168. [Google Scholar] [CrossRef] [PubMed]
  17. Sun, F.; Xiu, Z.H.; Jiang, R.C.; Liu, Y.W.; Zhang, X.Y.; Yang, Y.Z.; Li, X.J.; Zhang, X.; Wang, Y.; Tan, B.C. The mitochondrial pentatricopeptide repeat protein EMP12 is involved in the splicing of three nad2 introns and seed development in maize. J. Exp. Bot. 2019, 70, 963–972. [Google Scholar] [CrossRef] [Green Version]
  18. Pan, Z.Y.; Liu, M.; Xiao, Z.Y.; Ren, X.M.; Zhao, H.L.; Gong, D.M.; Liang, K.; Tan, Z.D.; Shao, Y.Q.; Qiu, F.Z. ZmSMK9, a pentatricopeptide repeat protein, is involved in the cis-splicing of nad5, kernel development and plant architecture in maize. Plant. Sci. 2019, 288, 110205. [Google Scholar] [CrossRef]
  19. Ren, R.C.; Wang, L.L.; Zhang, L.; Zhao, Y.J.; Wu, J.W.; Wei, Y.M.; Zhang, X.S.; Zhao, X.Y. DEK43 is a P-type pentatricopeptide repeat (PPR) protein responsible for the cis-splicing of nad4 in maize mitochondria. J. Integr. Plant Biol. 2020, 62, 299–313. [Google Scholar] [CrossRef] [PubMed]
  20. Zuo, Y.; Feng, F.; Qi, W.W.; Song, R.T. Dek42 encodes an RNA binding protein that affects alternative pre-mRNA splicing and maize kernel development. J. Integr. Plant Biol. 2020, 61, 728–748. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Xiu, Z.H.; Peng, L.; Wang, Y.; Yang, H.H.; Sun, F.; Wang, X.M.; Cao, S.K.; Jiang, R.C.; Wang, L.; Chen, B.Y.; et al. Empty pericarp 24 and empty pericarp 25 are required for the splicing of mitochondrial introns complex Ⅰ assembly and seed development in maize. Front. Plant Sci. 2020, 11, 608550. [Google Scholar] [CrossRef] [PubMed]
  22. Yang, H.H.; Xiu, Z.H.; Wang, L.; Cao, S.K.; Li, X.L.; Sun, F.; Tan, B.C. Two pentatricopeptide repeat proteins are required for the splicing of nad5 introns in maize. Front. Plant Sci. 2020, 11, 732. [Google Scholar] [CrossRef] [PubMed]
  23. Yang, Y.Z.; Ding, S.; Liu, X.Y.; Tang, J.J.; Wang, Y.; Sun, F.; Xu, C.H.; Tan, B.C. EMP32 is required for the cis-splicing of nad7 intron 2 and seed development in maize. RNA Biol. 2020, 18, 499–509. [Google Scholar] [CrossRef]
  24. Yang, Y.Z.; Ding, S.; Wang, Y.; Wang, H.C.; Liu, X.Y.; Sun, F.; Xu, C.; Liu, B.; Tan, B.C. PPR20 is required for the cis-splicing of mitochondrial nad2 intron 3 and seed development in maize. Plant Cell Physiol. 2020, 61, 370–380. [Google Scholar] [CrossRef] [PubMed]
  25. Wang, C.H.; Chen, Z.L.; Yang, Y.Z.; Sun, F.; Ding, S.; Li, X.L.; Xu, C.H.; Tan, B.C. PPR14 interacts with PPR-SMR1 and CRM protein Zm-Mcsf1 to facilitate mitochondrial intron splicing in maize. Front. Plant Sci. 2020, 11, 814. [Google Scholar] [CrossRef]
  26. Liu, R.; Cao, S.K.; Sayyed, A.; Xu, C.H.; Sun, F.; Wang, X.M.; Tan, B.C. The mitochondrial pentatricopeptide repeat protein PPR18 is required for the cis-splicing of nad4 intron 1 and essential to seed development in maize. Int. J. Mol. Sci. 2020, 21, 4047. [Google Scholar] [CrossRef] [PubMed]
  27. Yan, J.J.; Zhang, Q.X.; Yin, P. RNA editing machinery in plant organelles. Sci. China Life Sci. 2018, 61, 162–169. [Google Scholar] [CrossRef] [PubMed]
  28. Gualberto, J.M.; Lamattina, L.; Bonnard, G.; Weil, J.H.; Grienenberger, J.M. RNA editing in wheat mitochondria results in the conservation of protein sequences. Nature 1989, 341, 660–662. [Google Scholar] [CrossRef] [PubMed]
  29. Boussardon, C.; Salone, V.; Avon, A.; Berthome, R.; Hammani, K.; Okuda, K.; Shikanai, T.; Small, I.; Lurin, C. Two interacting proteins are necessary for the editing of the NdhD-1 site in Arabidopsis plastids. Plant Cell 2012, 24, 3684–3694. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Ren, Z.J.; Fan, K.J.; Fang, T.; Zhang, J.J.; Yang, L.; Wang, J.H.; Wang, G.Y.; Liu, Y.J. Maize empty pericarp602 encodes a P-Type PPR protein that is essential for seed development. Plant Cell Physiol. 2019, 60, 1734–1746. [Google Scholar] [CrossRef] [PubMed]
  31. 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]
  32. Sun, F.; Wang, X.; Bonnard, G.; Shen, Y.; Xiu, Z.; Li, X.; Gao, D.; Zhang, Z.; Tan, B.C. Empty pericarp7 encodes a mitochondrial E-subgroup pentatricopeptide repeat protein that is required for ccmFN editing.; mitochondrial function and seed development in maize. Plant J. 2015, 84, 283–295. [Google Scholar] [CrossRef] [PubMed]
  33. Yang, Y.Z.; Ding, S.; Wang, H.C.; Sun, F.; Huang, W.L.; Song, S.; Xu, C.; Tan, B.C. The pentatricopeptide repeat protein EMP9 is required for mitochondrial ccmB and rps4 transcript editing; mitochondrial complex biogenesis and seed development in maize. New Phytol. 2017, 214, 782–795. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Li, X.L.; Huang, W.L.; Yang, H.H.; Jiang, R.C.; Sun, F.; Wang, H.C.; Zhao, J.; Xu, C.H.; Tan, B.C. EMP18 functions in mitochondrial atp6 and cox2 transcript editing and is essential to seed development in maize. New Phytol. 2019, 221, 896–907. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Wang, Y.; Liu, X.Y.; Yang, Y.Z.; Huang, J.; Sun, F.; Lin, J.S.; Gu, Z.Q.; Sayyed, A.; Xu, C.H.; Tan, B.C. Empty pericarp21 encodes a novel PPR-DYW protein that is required for mitochondrial RNA editing at multiple sites, complexes I and V biogenesis, and seed development in maize. PLoS Genet. 2019, 15, e1008305. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Wang, G.; Zhong, M.Y.; Shuai, B.L.; Song, J.D.; Zhang, J.; Han, L.; Ling, H.L.; Tang, Y.P.; Wang, G.F.; Song, R.T. E+ subgroup PPR protein defective kernel 36 is required for multiple mitochondrial transcripts editing and seed development in maize and Arabidopsis. New Phytol. 2017, 214, 1563–1578. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Li, X.J.; Gu, W.; Sun, S.L.; Chen, Z.L.; Chen, J.; Song, W.B.; Zhao, H.M.; Lai, J.S. Defective Kernel 39 encodes a PPR protein required for seed development in maize. J. Integr. Plant Biol. 2018, 60, 45–64. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Sosso, D.; Mbelo, S.; Vernoud, V.; Gendrot, G.; Dedieu, A.; Chambrier, P.; Dauzat, M.; Heurtevin, L.; Guyon, V.; Takenaka, M. PPR2263, a DYW-subgroup pentatricopeptide repeat protein, is required for mitochondrial nad5 and cob transcript editing, mitochondrion biogenesis, and maize growth. Plant Cell 2012, 24, 676–691. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Li, X.J.; Zhang, Y.F.; Hou, M.; Sun, F.; Shen, Y.; Xiu, Z.H.; Wang, X.; Chen, Z.L.; Sun, S.S.; Small, I. Small kernel 1 encodes a pentatricopeptide repeat protein required for mitochondrial nad7 transcript editing and seed development in maize (Zea mays) and rice (Oryza sativa). Plant J. 2014, 79, 797–809. [Google Scholar] [CrossRef] [PubMed]
  40. Wang, H.C.; Sayyed, A.; Liu, X.Y.; Yang, Y.Z.; Sun, F.; Wang, Y.; Wang, M.; Tan, B.C. Small kernel4 is required for mitochondrial cox1 transcript editing and seed development in maize. J. Integr. Plant Biol. 2019, 62, 777–792. [Google Scholar] [CrossRef]
  41. Ding, S.; Liu, X.Y.; Wang, H.C.; Wang, Y.; Tang, J.J.; Yang, Y.Z.; Tan, B.C. SMK6 mediates the C-to-U editing at multiple sites in maize mitochondria. Plant Physiol. 2019, 240, 152992. [Google Scholar] [CrossRef] [PubMed]
  42. Colcombet, J.; Lopez-obando, M.; Heurtevin, L.; Bernard, C.; Martin, K.; Berthomé, R.; Lurin, C. Systematic study of subcellular localization of Arabidopsis PPR proteins confirms a massive targeting to organelles. RNA Biol. 2014, 10, 1557–1575. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Hammani, K.; Gobert, A.; Hleibieh, K.; Choulier, L.; Small, I.; Giege, P. An Arabidopsis dual-localized pentatricopeptide repeat protein interacts with nuclear proteins involved in gene expression regulation. Plant Cell 2011, 23, 730–740. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Mei, C.; Jiang, S.C.; Lu, Y.F.; Wu, F.Q.; Yu, Y.T.; Liang, S.; Feng, X.J.; Comeras, S.P.; Lu, K.; Wu, Z.; et al. Arabidopsis pentatricopeptide repeat protein SOAR1 plays a critical role in abscisic acid signalling. J. Exp. Bot. 2014, 65, 5317–5330. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. 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] [PubMed] [Green Version]
  46. Xue, M.Y.; Liu, L.L.; Yu, Y.F.; Zhu, J.P.; Gao, H.; Wang, Y.H.; Wan, J.M. Lose-of-function of a rice nucleolus-localized pentatricopeptide repeat protein is responsible for the floury endosperm14 mutant phenotypes. Rice 2019, 12, 100–115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Hao, Y.Y.; Wang, Y.L.; Wu, M.M.; Zhu, X.P.; Teng, X.; Sun, Y.L.; Zhu, J.P.; Zhang, Y.Y.; Jing, R.N.; Lei, J.; et al. The nuclear-localized PPR protein OsNPPR1 is important for mitochondrial function and endosperm development in rice. J. Exp. Bot. 2019, 70, 4705–4720. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Fekih, R.; Takagi, H.; Tamiru, M.; Abe, A.; Natsume, S.; Yaegashi, H.; Sharma, S.; Sharma, S.; Kanzaki, H.; Matsumura, H.; et al. MutMap+: Genetic mapping and mutant identification without crossing in rice. PLoS ONE 2013, 8, e68529. [Google Scholar] [CrossRef] [PubMed]
  49. Cheng, S.F.; Gutmann, B.; Zhong, X.; Ye, Y.T.; Fisher, M.F.; Bai, F.Q.; Castleden, I.; Song, Y.; Song, B.; Huang, J.Y.; et al. Redefining the structural motifs that determine RNA binding and RNA editing by pentatricopeptide repeat proteins in land plants. Plant J. 2016, 85, 532–547. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Bernard, G.; Royan, S.; Schallenberg-Rüdinger, M.; Lenz, H.; Castleden, I.R.; McDowell, R.; Vacher, M.A.; Tonti-Filippini, J.; Bond, C.S.; Knoop, V.; et al. The Expansion and Diversification of Pentatricopeptide Repeat RNA-Editing Factors in Plants. Mol. Plant 2020, 13, 215–230. [Google Scholar] [CrossRef]
  51. Yin, P.; Li, Q.; Yan, C.; Liu, Y.; Yu, F.; Wang, Z.; Long, J.; He, J.; Wang, H.W.; Wang, J.; et al. Structural basis for the modular recognition of single-stranded RNA by PPR proteins. Nature 2013, 504, 168–171. [Google Scholar] [CrossRef]
  52. Giege, P.; Grienenberger, J.M.; Bonnard, G. Cytochrome c biogenesis in mitochondria. Mitochondrion 2008, 8, 61–73. [Google Scholar] [CrossRef] [PubMed]
  53. Doniwa, Y.; Ueda, M.; Ueta, M.; Wada, A.; Kadowaki, K.I.; Tsutsumi, N. The involvement of a PPR protein of the P subfamily in partial RNA editing of an Arabidopsis mitochondrial transcript. Gene 2010, 454, 39–46. [Google Scholar] [CrossRef] [PubMed]
  54. Leu, K.C.; Hsieh, M.H.; Wang, H.J.; Hsieh, H.L.; Jauh, G.Y. Distinct role of Arabidopsis mitochondrial P-type pentatricopeptide repeat protein-modulating editing protein, PPME, in nad1 RNA editing. RNA Biol. 2016, 13, 593–604. [Google Scholar] [CrossRef] [Green Version]
  55. Guillaumot, D.; Lopez-obando, M.; Baudry, K.; Avon, A.; Rigaill, G.; Longevialle, A.F.D.; Broche, B.; Takenaka, M.; Berthomé, R.; Jaeger, G.D.; et al. Two interacting PPR proteins are major Arabidopsis editing factors in plastid and mitochondria. Proc. Natl Acad. Sci. USA 2017, 114, 8877–8882. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Cai, M.J.; Li, S.Z.; Sun, F.; Sun, Q.; Zhao, H.L.; Ren, X.M.; Zhao, Y.X.; Tan, B.C.; Zhang, Z.X.; Qiu, F.Z. Emp10 encodes a mitochondrial PPR protein that affects the cis-splicing of nad2 intron 1 and seed development in maize. Plant J. 2017, 91, 132–144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. San Francisco, B.; Bretsnyder, E.C.; Rodgers, K.R.; Kranz, R.G. Heme ligand identification and redox properties of the cytochrome c synthetase, CcmF. Biochemistry 2011, 50, 10974–10985. [Google Scholar] [CrossRef] [Green Version]
  58. San Francisco, B.; Sutherland, M.C.; Kranz, R.G. The CcmFH complex is the system I holocytochrome c synthetase: Engineering cytochrome c maturation independent of CcmABCDE. Mol. Microbiol. 2014, 91, 996–1008. [Google Scholar] [CrossRef] [Green Version]
  59. Ren, R.E.; Yan, X.W.; Zhao, Y.J.; Wei, Y.M.; Lu, X.D.; Zang, J.; Wu, J.W.; Zheng, G.M.; Ding, X.H.; Zhang, X.S.; et al. The novel E-subfamily pentatricopeptide repeat protein DEK55 is responsible for RNA editing at multiple sites and for the splicing of nad1 and nad4 in maize. BMC Plant Biol. 2020, 20, 553. [Google Scholar] [CrossRef] [PubMed]
  60. Gal, C.; Moore, K.M.; Paszkiewicz, K.; Kent, N.A.; Whitehall, S.M. The impact of the HIRA histone chaperone upon global nucleosome architecture. Cell Cycle 2015, 14, 123–134. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Zhu, C.G.; Jin, G.P.; Fang, P.; Zhang, Y.; Feng, X.Z.; Tang, Y.P.; Qi, W.W.; Song, R.T. Maize pentatricopeptide repeat protein DEK41 affects cis-splicing of mitochondrial nad4 intron 3 and is required for normal seed development. J. Exp. Bot. 2019, 70, 3795–3808. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Xiao, C.W.; Chen, F.L.; Yu, X.H.; Lin, C.T.; Fu, Y.F. Over-expression of an AT-hook gene, AHL22, delays flowering and inhibits the elongation of the hypocotyl in Arabidopsis thaliana. Plant Mol. Biol. 2009, 71, 39–50. [Google Scholar] [CrossRef] [PubMed]
  63. Yang, Y.Z.; Tan, B.C. A distal ABA responsive element in AtNECD3 promoter is required for positive feedback regulation of ABA biosynthesis in Arabidopsis. PLoS ONE 2014, 9, e87283. [Google Scholar]
Figure 1. ppr278-1 mutants exhibited defective kernel development and lethal seedlings. (A,B) Ears of self-pollinated heterozygous ppr278-1 plants at 16 days after pollination (16 DAP) (A) and at maturity (B). Red arrows indicate the mutant kernels. Bar = 1 cm. (C,D). Comparison of the length (C) and width (D) of mature seeds of the WT and the ppr278-1 mutant. Bar = 1 cm. (E) Germination of WT and ppr278-1 mutant seeds after dark culture for two days in an incubator. Bar = 1 cm. (F) Comparison of the 100-kernel weight of fully-mature WT and ppr278-1 mutant kernels. Values and error bars represent the mean and standard deviation of three biological replicates. *** p < 0.001 (t-test), representing a significant difference. (G) Longitudinal section of WT (left) and ppr278-1 mutant kernels (right) at 12-DAP. en: endosperm, em: embryo. Bar = 0.5 mm.
Figure 1. ppr278-1 mutants exhibited defective kernel development and lethal seedlings. (A,B) Ears of self-pollinated heterozygous ppr278-1 plants at 16 days after pollination (16 DAP) (A) and at maturity (B). Red arrows indicate the mutant kernels. Bar = 1 cm. (C,D). Comparison of the length (C) and width (D) of mature seeds of the WT and the ppr278-1 mutant. Bar = 1 cm. (E) Germination of WT and ppr278-1 mutant seeds after dark culture for two days in an incubator. Bar = 1 cm. (F) Comparison of the 100-kernel weight of fully-mature WT and ppr278-1 mutant kernels. Values and error bars represent the mean and standard deviation of three biological replicates. *** p < 0.001 (t-test), representing a significant difference. (G) Longitudinal section of WT (left) and ppr278-1 mutant kernels (right) at 12-DAP. en: endosperm, em: embryo. Bar = 0.5 mm.
Ijms 23 03035 g001
Figure 2. Identification of the ppr278 mutation. (A) Schematic of the Zm00001d015156 gene and the location of mutated sites in the ppr278-1 and ppr278-2 mutants. The candidate gene was located on chromosome 5 and identified as Zm00001d015156, which contains 16 exons (blue rectangles) and 15 introns (black lines). A C-to-T mutation in the 15th exon leads to protein termination at positions 2269 and 2404 bp from the ATG start codon in the ppr278-1 and ppr278-2 mutants, respectively. (B) The results of the allelic test of 12-DAP ears of the self-pollinated plants of PPR278/ppr278-1, PPR278/ppr278-2, PPR278/ppr278-1 ×PPR278/ppr278-2, and PPR278/ppr278-2×PPR278/ppr278-1. Red arrows indicate mutant ppr278 seeds. Bar = 1 cm. (C) Schematic of PPR278 protein. Blue rectangles represent the p motifs. (D) qRT-PCR analysis of relative PPR278 expression in different tissues and kernels at different developmental stages.
Figure 2. Identification of the ppr278 mutation. (A) Schematic of the Zm00001d015156 gene and the location of mutated sites in the ppr278-1 and ppr278-2 mutants. The candidate gene was located on chromosome 5 and identified as Zm00001d015156, which contains 16 exons (blue rectangles) and 15 introns (black lines). A C-to-T mutation in the 15th exon leads to protein termination at positions 2269 and 2404 bp from the ATG start codon in the ppr278-1 and ppr278-2 mutants, respectively. (B) The results of the allelic test of 12-DAP ears of the self-pollinated plants of PPR278/ppr278-1, PPR278/ppr278-2, PPR278/ppr278-1 ×PPR278/ppr278-2, and PPR278/ppr278-2×PPR278/ppr278-1. Red arrows indicate mutant ppr278 seeds. Bar = 1 cm. (C) Schematic of PPR278 protein. Blue rectangles represent the p motifs. (D) qRT-PCR analysis of relative PPR278 expression in different tissues and kernels at different developmental stages.
Ijms 23 03035 g002
Figure 3. PPR278 was localized to the nucleus and mitochondria and predicted to form a series of alpha helices. (A,B) Transient expression of PPR278 in onion epidermal cells (A) and maize protoplasts (B). GFP was used as a control (top panels of (A,B)) and was co-expressed with the nuclear marker (bottom panels of (A,B)). Bar = 50 μm (A) and 5 μm (B). (C) Predicted protein structure of PPR278. The purple and red helices indicate the N terminus (N) and C terminus (C) of PPR278, respectively.
Figure 3. PPR278 was localized to the nucleus and mitochondria and predicted to form a series of alpha helices. (A,B) Transient expression of PPR278 in onion epidermal cells (A) and maize protoplasts (B). GFP was used as a control (top panels of (A,B)) and was co-expressed with the nuclear marker (bottom panels of (A,B)). Bar = 50 μm (A) and 5 μm (B). (C) Predicted protein structure of PPR278. The purple and red helices indicate the N terminus (N) and C terminus (C) of PPR278, respectively.
Ijms 23 03035 g003
Figure 4. Intron splicing deficiency of nad2 and nad5 in ppr278 mutants. (A) RT-PCR analysis of 35 mitochondrion-encoded transcripts in WT (left) and ppr278-1 (right) plants. RNA was isolated from WT and ppr278-1 endosperm from the same segregating ear at 12-DAP. 18S rRNA served as an internal control. (B) Structures of the maize mitochondrial genes nad2 and nad5. (C) RT-PCR analysis of nad2 and nad5 intron splicing efficiency in WT (left), ppr278-1 (middle), and ppr278-2 (right) plants. A DNA marker is shown in the farther left lane. Red arrows indicate abnormal splicing in the ppr278-1 and ppr278-2 mutants.
Figure 4. Intron splicing deficiency of nad2 and nad5 in ppr278 mutants. (A) RT-PCR analysis of 35 mitochondrion-encoded transcripts in WT (left) and ppr278-1 (right) plants. RNA was isolated from WT and ppr278-1 endosperm from the same segregating ear at 12-DAP. 18S rRNA served as an internal control. (B) Structures of the maize mitochondrial genes nad2 and nad5. (C) RT-PCR analysis of nad2 and nad5 intron splicing efficiency in WT (left), ppr278-1 (middle), and ppr278-2 (right) plants. A DNA marker is shown in the farther left lane. Red arrows indicate abnormal splicing in the ppr278-1 and ppr278-2 mutants.
Ijms 23 03035 g004
Figure 5. C-to-U RNA editing sites and efficiency in atp8 transcripts. (A) Editing profile of five C-to-U RNA editing sites in the atp8 transcripts in the ppr278 mutant. (B) C-to-U editing efficiency in atp8 transcripts; “n” stands for the peak of heterozygous genotype (C/U) in the same editing site, “*” stands for the C-to-U site in atp8. (C). Comparison of amino acid sequences of atp8 in different species.
Figure 5. C-to-U RNA editing sites and efficiency in atp8 transcripts. (A) Editing profile of five C-to-U RNA editing sites in the atp8 transcripts in the ppr278 mutant. (B) C-to-U editing efficiency in atp8 transcripts; “n” stands for the peak of heterozygous genotype (C/U) in the same editing site, “*” stands for the C-to-U site in atp8. (C). Comparison of amino acid sequences of atp8 in different species.
Ijms 23 03035 g005
Figure 6. Transcriptome analysis in the ppr278-1 mutant. (A) Correlation between the replicates of RNA-seq data in WT and ppr278-1 plants. The x- and y-axes represent replicate 1 and replicate 2, respectively. (B) Volcano plot displaying differentially expressed genes (DEGs) between WT and ppr278-1 plants. The y-axis corresponds to the mean expression value of log10 (p-value), and the x-axis displays the log2 (fold change) value. Red dots represent the up-regulated transcripts between WT and ppr278-1 plants. Blue dots represent the down-regulated transcripts. (C) GO terms of the top 33 and 27 up- and down-regulated DEGs, respectively, in WT and ppr278-1 plants. The size of the dot represents the number of genes. The color shows the value of −Log10 (FDR, false discovery rate).
Figure 6. Transcriptome analysis in the ppr278-1 mutant. (A) Correlation between the replicates of RNA-seq data in WT and ppr278-1 plants. The x- and y-axes represent replicate 1 and replicate 2, respectively. (B) Volcano plot displaying differentially expressed genes (DEGs) between WT and ppr278-1 plants. The y-axis corresponds to the mean expression value of log10 (p-value), and the x-axis displays the log2 (fold change) value. Red dots represent the up-regulated transcripts between WT and ppr278-1 plants. Blue dots represent the down-regulated transcripts. (C) GO terms of the top 33 and 27 up- and down-regulated DEGs, respectively, in WT and ppr278-1 plants. The size of the dot represents the number of genes. The color shows the value of −Log10 (FDR, false discovery rate).
Ijms 23 03035 g006
Table 1. Comparison of editing profiles that differ between WT and ppr278.
Table 1. Comparison of editing profiles that differ between WT and ppr278.
GenePosition from ATGC-to-U Editing
Efficiency
WTppr
278-1
ppr
278-2
GenePosition from ATGC-to-U Editing
Efficiency
WTppr
278-1
ppr
278-2
atp130ALeuLeuLeunad3275DSer/PheSer/PheSer/Phe
atp11178ALeu/SerSerSernad3317DSer/PheSer/PheSer/Phe
atp11292ALeu/ProProPronad3344DSer/LeuSer/LeuSer/Leu
atp11490ALeu/ProProPronad3349DArg/TrpArg/TrpArg/Trp
atp11499APhe/SerSerSerrpl16228DIleIleIle
atp456ALeuProProrps2A200DSer/PheSer/PheSer/Phe
atp459APheSerSerrps2A449DAla/ValAla/ValAla/Val
atp471DLeuSerLeu/Serrps2A514DSerSer/ProSer/Pro
atp476DSerProProrps2A541DCysCys/ArgCys/Arg
atp489DLeuSerLeu/Serrps2A548DLeuSer/LeuSer
atp4118DCysArgCys/Argrps2B550IArg/CysCysCys
atp4360DCysCysCysrps1271DSer/LeuSer/LeuSer/Leu
atp4407DLeuSerLeu/Serrps12196DHis/TyrHis/TyrHis/Tyr
atp4428DIleThrIle/Thrrps12221DSer/LeuSer/LeuSer/Leu
atp858DLeu/PheLeu/PheLeu/Pherps12269DSer/LeuSer/LeuSer/Leu
atp8123DLeuLeuLeurps12284DSer/LeuSer/LeuSer/Leu
atp8200DSer/LeuSer/LeuSer/Leurps12289DArg/CysArg/CysArg/Cys
atp8436DPro/Ser
/Leu
Pro/Ser
/Leu
Pro/Ser
/Leu
ccmFN76ASerProPro
atp8437ccmFN77
cox369DLeuLeuLeuccmFN137ALeuProPro
cox3245DLeu/ProLeu/ProLeu/ProccmFN176ALeuSerSer
cox3257DPhe/SerPhe/SerPhe/SerccmFN181ACysArgArg
cox3289DPhe/SerPhe/SerPhe/SerccmFN190ASerProPro
cox3311DSer/PheSer/PheSer/PheccmFN287ALeuSerSer
cox3314DPheSerSer/PheccmFN302ALeuProPro
cox3413DLeuProLeu/ProccmFN401APheSerSer
cox3422DLeuProLeu/ProccmFN410ALeuSerSer
cox3527DPheSer/PheSer/PheccmFN417APhePhePhe
cox3566DPhe/SerSerPhe/SerccmFN743ALeuProPro
cox3754DTrpArgTrp/ArgccmFN752ALeuSerSer
nad35DSer/LeuSer/LeuSer/LeuccmFN790ACysArgArg
nad339DIleIleIleccmFN812ALeuSerSer
nad344DPro/LeuPro/LeuPro/LeuccmFN824ALeuProPro
nad361DPro/Ser
/Leu
Pro/Ser
/Leu
Pro/Ser
/Leu
ccmFN839ALeuSerSer
nad362ccmFN1325ALeuProPro
nad380DPro/LeuPro/LeuPro/LeuccmFN1342ATyrHisHis
nad3137DSer/PheSer/PheSer/PheccmFN1357ATrpArgArg
nad3138DccmFN1375ATrpArgArg
nad3146DSer/PheSer/PheSer/PheccmFN1408ATrpArgArg
nad3185DPro/LeuPro/LeuPro/LeuccmFN1469ALeuSerSer
nad3190DPro/SerPro/SerPro/SerccmFN1489APheLeuLeu
nad3208DPro/Ser
/Leu
Pro/Ser
/Leu
Pro/Ser
/Leu
ccmFN1493ALeuProPro
nad3209ccmFN1505ALeuSerSer
nad3215DPro/LeuPro/LeuPro/LeuccmFN1540ASerProPro
nad3230DSer/PheSer/PheSer/PheccmFN1553APheSerSer
nad3247DPro/SerPro/SerPro/SerccmFN1588ATrpArgArg
nad3251DPro/LeuPro/LeuPro/LeuccmFN1709ALeuProPro
The C-to-U editing efficiency of mitochondrial transcripts in ppr278-1 and ppr278-2 mutants was abolished (A), decreased (D), or increased (I) when compared with the WT. Leu: Leucine, Ser: Serine, Pro: Proline, Phe: Phenylalanine, Cys: Cysteine, Arg: Arginine, Ile: Isoleucine, Thr: Threonine, Try: Tryptophan, Ala: Alanine, Val: Valine, His: Histidine, Tyr: Tyrosine.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yang, J.; Cui, Y.; Zhang, X.; Yang, Z.; Lai, J.; Song, W.; Liang, J.; Li, X. Maize PPR278 Functions in Mitochondrial RNA Splicing and Editing. Int. J. Mol. Sci. 2022, 23, 3035. https://doi.org/10.3390/ijms23063035

AMA Style

Yang J, Cui Y, Zhang X, Yang Z, Lai J, Song W, Liang J, Li X. Maize PPR278 Functions in Mitochondrial RNA Splicing and Editing. International Journal of Molecular Sciences. 2022; 23(6):3035. https://doi.org/10.3390/ijms23063035

Chicago/Turabian Style

Yang, Jing, Yang Cui, Xiangbo Zhang, Zhijia Yang, Jinsheng Lai, Weibin Song, Jingang Liang, and Xinhai Li. 2022. "Maize PPR278 Functions in Mitochondrial RNA Splicing and Editing" International Journal of Molecular Sciences 23, no. 6: 3035. https://doi.org/10.3390/ijms23063035

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