Next Article in Journal
Lessons from the Trials for the Desirable Effects of Sodium Glucose Co-Transporter 2 Inhibitors on Diabetic Cardiovascular Events and Renal Dysfunction
Next Article in Special Issue
Characterizing the Role of TaWRKY13 in Salt Tolerance
Previous Article in Journal
Bovine Lactoferrin Pre-Treatment Induces Intracellular Killing of AIEC LF82 and Reduces Bacteria-Induced DNA Damage in Differentiated Human Enterocytes
Previous Article in Special Issue
Cloning and Expression Analysis of the BocMBF1c Gene Involved in Heat Tolerance in Chinese Kale
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genome-Wide Analysis of the DYW Subgroup PPR Gene Family and Identification of GmPPR4 Responses to Drought Stress

1
Institute of Crop Science, Chinese Academy of Agricultural Sciences (CAAS)/National Key Facility for Crop Gene Resources and Genetic Improvement, Key Laboratory of Biology and Genetic Improvement of Triticeae Crops, Ministry of Agriculture, Beijing 100081, China
2
State Key Laboratory of Crop Stress Biology for Arid Areas, College of Agronomy, Northwest A&F University, Yangling 712100, China
3
Agro-Biotechnology Research Institute, Jilin Academy of Agriculture Sciences, Changchun 130033, China
4
College of Agronomy, Jilin Agricultural University, Changchun 130118, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to the present work.
Int. J. Mol. Sci. 2019, 20(22), 5667; https://doi.org/10.3390/ijms20225667
Submission received: 25 September 2019 / Revised: 25 October 2019 / Accepted: 6 November 2019 / Published: 12 November 2019
(This article belongs to the Special Issue Mapping Abiotic Stress-Tolerance Genes in Plants)

Abstract

:
Pentatricopeptide-repeat (PPR) proteins were identified as a type of nucleus coding protein that is composed of multiple tandem repeats. It has been reported that PPR genes play an important role in RNA editing, plant growth and development, and abiotic stresses in plants. However, the functions of PPR proteins remain largely unknown in soybean. In this study, 179 DYW subgroup PPR genes were identified in soybean genome (Glycine max Wm82.a2.v1). Chromosomal location analysis indicated that DYW subgroup PPR genes were mapped to all 20 chromosomes. Phylogenetic relationship analysis revealed that DYW subgroup PPR genes were categorized into three distinct Clusters (I to III). Gene structure analysis showed that most PPR genes were featured by a lack of intron. Gene duplication analysis demonstrated 30 PPR genes (15 pairs; ~35.7%) were segmentally duplicated among Cluster I PPR genes. Furthermore, we validated the mRNA expression of three genes that were highly up-regulated in soybean drought- and salt-induced transcriptome database and found that the expression levels of GmPPR4 were induced under salt and drought stresses. Under drought stress condition, GmPPR4-overexpressing (GmPPR4-OE) plants showed delayed leaf rolling; higher content of proline (Pro); and lower contents of H2O2, O2 and malondialdehyde (MDA) compared with the empty vector (EV)-control plants. GmPPR4-OE plants exhibited increased transcripts of several drought-inducible genes compared with EV-control plants. Our results provided a comprehensive analysis of the DYW subgroup PPR genes and an insight for improving the drought tolerance in soybean.

Graphical Abstract

1. Introduction

The pentatricopeptide repeat (PPR) proteins containing tandem 30–40 amino acid sequence motifs, constitute a large gene family in plants [1]. Typically, PPR proteins are classed into two major subfamilies (P and PLS). The P subfamily proteins contain only the canonical P motif (35 amino acid residues); the PLS subfamily proteins are confined to contain a series of PLS triplets (L, long variants of P; S, small variants of P) [2]. L motif and S motif contain 31 and 35 or 36 amino acid residues, respectively, which are considered to be variations of the P motif [2]. Most PPR proteins are highly conserved at the C-terminus and usually have three conserved domains at the C-terminus, E, E+ and DYW domains, respectively [3,4]. Based on these C-terminal domains, the PLS subfamily is divided into four subgroups, PLS subgroup, E subgroup, E+ subgroup, and DYW subgroup [2]. The PPR proteins also contain a distinctive feature that is essentially free of intron. Although members of the PPR family are widely distributed in plants and are large in number, they perform special functions [2].
So far, a considerable amount of PPR proteins have been experimentally identified from various plant species. [5]. For example, 441 and 626 PPR members have been identified in the A. thaliana and Populus trichoccapa genome, respectively [5,6,7]. A large number of studies showed that PPR proteins played roles in the growth and development of plants and the formation of organelles [8]. It can participate in the fertility restoration of cytoplasmic male sterility, post-transcriptional processing of RNA and adversity defense [2,7,9,10,11,12]. In maize, EMP12 is targeted to the mitochondria and is essential for the splicing of three nad2 introns and seed development, and the mutant in EMP12 caused defects complex I [13]. It was reported that the nuclear OsNPPR1 encoded a PPR protein, which was involved in the Regulation of mitochondrial development and endosperm development. The mutant named fgr1 exhibited a lower content of starch and delayed seedling growth [14]. Rice PPS1 encodes a DYW motif-containing PPR protein, which is important for C-to-U RNA editing of nad3 transcripts. Knock-down and Knock-out of PPS1 revealed a significant decrease in the editing efficiency of C-to-U at five editing sites of nad3 transcripts, which resulted in some pleiotropic phenotypes, such as dwarfing, developmental retardation and stunted development in vegetative stages; partial pollen was sterile at the reproductive stage [15]. A chloroplast-localized P-class PPR protein, PpPPR_21, was identified as an essential protein for the accumulation of a stable psbl-ycf12 mRNA. The knockout ppr21 mutants displayed decreasing protonemata growth and lower photosynthetic activity [16].
Of particular note, several studies have recently reported roles for PPR genes during abiotic stress [17]. For example, PPR96 from A. thaliana was involved in the response to salt, ABA and oxidative stress, and altered translational levels of abscisic stress responsive genes, implying that PPR96 may take part in the regulation of plant tolerance to abiotics [18]. PGN, an A. thaliana mitochondria-localized PPR protein, was reported to play a role in regulating mitochondrial reactive oxygen species balance in abiotic and biotic stress responses. Inactivation of PGN displayed hypersensitivity to ABA, glucose, and salinity. Loss of PGN resulted in enhanced accumulation of reactive oxygen species (ROS) in seedlings, and the pgn mutant notably elevated levels of ABI4 and ALTERNATIVE OXIDASE1a [19]. In rice, cisc(t) encoded a PPR protein, which was necessary for chloroplast development in early stages under cold stress [20]. The A. thaliana ABA overly sensitive 5 (abo5) mutant showed growth retardation and accumulated increased levers of H2O2 in root tips [21]. AtPPR40 was reported to be involved in salt, ABA and oxidative stress [22].
Previous studies have emphasized the role of PPR genes in the growth and development of plants, and several PPR genes performed potential roles against abiotic stresses. To date, little information about PPR genes relating to abiotic stress response mechanisms in soybean has been reported. Here, a comprehensive genome-wide analysis of the DYW subgroup PPR gene family has been accomplished in soybean. We explored the characteristics of 179 DYW subgroup PPR genes, including intron-exon organization, chromosomal location and phylogenetic relationships. The DYW subgroup PPR genes were divided into three discrete groups (Cluster I to III). An analysis of a complete set of the Cluster I PPR genes/proteins was performed, including classification, chromosomal location, orthologous relationships, duplication analysis and tissue-specific expression patterns. GmPPR4 was further investigated owing to significant up-regulation by drought and salt stresses. Our results showed that overexpression of GmPPR4 improved tolerance to drought stress in soybean. Our study provided an insight into the foundation of the GmPPR4 gene in abiotic stress responses.

2. Result

2.1. Identification of DYW Subgroup PPR Genes in the Soybean Genome

In our study, 205 proteins were identified in the soybean genome (Glycine max Wm82.a2.v1). The candidate members were examined for the required DWY domain, which was unique to DYW subgroup PPR proteins. SMART and Pfam databases were used to verify the presence of the conserved domain in all 205 DYW subgroup PPR proteins, and the results showed that a total of 197 members were identified. Except for the presence of conserved PPR domains, there are great differences in the size and physicochemical properties of the DYW subgroup PPR genes (Table S1). The statistical results showed that the amino acid sequence length of the 179 DYW PPR proteins varied from the lowest of 118 to the highest of 1431, the isoelectric point (pI) varied from 5.28 to 9.36, and the molecular weight (MW) ranged from 13.7 kDa to 160.3 kDa. The detailed information of the 179 DYW subgroup PPR gene sequences is provided in Table S1. For convenience, the 179 DYW subgroup PPR genes were named from GmPPR1 to GmPPR179 according to the order of their chromosomal locations [23,24].
A considerable amount of research showed that the vast majority of PPR genes were found to be lacking or containing several introns in various plant species [5,6,7]. The intron numbers present within the ORF of each of DYW subgroup PPR genes in soybean were decided for analyzing their exon-intron organization. As showed in Figure 1, 59% (105/179) of DYW subgroup PPR genes were predicted to lack introns, 17% (30/179) with one intron, 22% (40/179) with two to five introns, and 2% (4/179) with six or more introns.

2.2. Chromosomal Distribution, Phylogenetic Analysis and Multiple Sequence Alignment

Our study showed 179 DYW subgroup PPR genes were distributed widely and unevenly in all the 20 soybean chromosomes (Figure 2). Chromosome 8 of soybean contained the highest number of DYW subgroup PPR genes, while fewer numbers of DYW subgroup PPR genes were located on chromosome 14 and 19. The detailed position of each DYW subgroup of PPR genes on the soybean chromosomes was obtained from Phytozome (Table S1).
To analyze the evolutionary relationship among DWY subgroup PPR genes, the full-length amino acid sequences of the 179 DYW subgroup PPR proteins were aligned and then the phylogenetic tree was conducted using the Maximum likelihood method by MEGA7.0 software (Figure 3). The phylogenetic analysis categorized the DYW subgroup PPR genes into three distinct clusters (I to III) comprising of 84, 65 and 30 genes, respectively. In order to isolate drought-inducible DYW subgroup PPR genes in soybean, we analyzed the expression patterns of the DYW subgroup PPR genes in soybean transcriptome database. We found that three DWY subgroup PPR genes were highly up-regulated by drought and salt stresses, including GmPPR4 (Glyma.01G011700), GmPPR18 (Glyma.02G174500), and GmPPR111 (Glyma.11G008800). Interestingly, they all clustered into the Cluster I, implying that Cluster I PPR genes may be involved in the abiotic stress of soybean, therefore, we further analyzed the genetic structure and motif of Cluster I PPR genes.
To identify conservation of PPR proteins in dicot plant species, six reported DYW proteins of A. thaliana and six randomly selected proteins of soybean were used to conduct the multiple sequence alignment, and the C-terminal region is displayed in Figure 4. Results showed that PPR proteins were relatively conserved across A. thaliana and soybean, however, only the significant short arrays (no more than four amino acid residues) were absolutely conservative, and 11 members harbored the conservative DYW sequence in C-terminal region that have previously been described as characteristic features of DYW subgroup PPR proteins. In addition, a phylogenic tree was conducted to reveal the relations of 12 selected proteins (Figure S1).

2.3. Gene Structure and Motif Composition of Soybean Cluster I PPR Genes

The exon-intron organization of the 84 Cluster I PPR genes was conducted to provide valuable information for the structure diversity evolution of the PPR family in soybean. As showed in Figure 5 and Table S1, they almost contained very few introns, which was similar to the previous results (Figure 1). However, the result showed that several genes contained more than 10 introns. Fourteen and 17 introns were found in Glyma.13G220400 and Glyma.15G092000, respectively. Many paralogs exhibited the same numbers and sizes of introns such as gene clusters of Glyma.07G057000, Glyma.16G026000, Glyma.02G217200, and Glyma.14G184600; However, only a few of them showed intron gain/loss phenomenon such as Glyma.03G205100 and Glyma.19G202500. On the whole, the intron distribution and the intron phase are in accordance with the alignment cluster of the Cluster I PPR genes.
Also, the Multiple EM for Motif Elicitation (MEME) program was used to search the conserved motifs which were shared with the Cluster I proteins (Figure 6). A total of 10 distinct conserved motifs named 1 to 10, were found. The motifs were defined based on sequence conservation without knowing its structure or function. Motif 1, the most typical DYW domain, comprised of 46 amino acids; the sequence (KNLRVCGDCHSAIKLISKIYNREIIVRDRNRFHHFKBGSCSCGDYW) contains a highly conserved C-terminal region (DYW) (Figure 7). However, this conservative motif is not included in the five proteins, including GmPPR28 (Glyma.03G238000), GmPPR72 (Glyma.08G117400), GmPPR91 (Glyma.09G154200), PPR103 (Glyma.10G093800), and GmPPR165 (Glyma.18G225600), which may be due to the fact that some PPR proteins are truncated at the C-terminus, lacking part, most of or the entire DYW domain. Proteins with higher homology often have the same conserved domain composition, which corresponds to the results of the phylogenetic tree.

2.4. Duplication and Divergence Rate of Soybean Cluster I PPR Genes

Whole genome tandem and segmental duplications play a vital role in multiple copies of genes in a gene family and their subsequent evolution. Among SiPPR genes, 190 (95pairs; ~39.01%) were segmentally duplicated [25]. Our research reached a similar conclusion that this proportion is far higher than of genes generally. Among Cluster I PPR genes, 30 (15 pairs; ~35.7%) were segmentally duplicated (connected by lines in Figure 8). Among 15 paralog pairs, 13 segmental duplication gene pairs were involved in two chromosomes; only two segmental duplications were intra-chromosomal. In particular, three genes have appeared twice, including GmPPR3 (Glyma.01G011300), GmPPR4 (Glyma.01G011700) and GmPPR94 (Glyma.09G209700). The results indicated that gene segmental duplications might mainly contribute to the expansion of the PPR gene family in soybean.
The ratio of non-synonymous (Ka) versus synonymous (Ks) substitution rates (Ka/Ks) was estimated for 15 segmental duplicated gene pairs to evaluate the role of Darwinian positive selection in duplication and divergence of the Cluster I PPR genes (Table S2). The result showed that the Ka/Ks for segmental duplications gene-pairs ranged from 0,14 to 1.03 with an average of 0.61. On the whole, it suggested that the duplication Cluster I PPR genes were under purifying selection pressures since their Ka/Ks rations were estimated as <1 except one gene pair with a rate (Ka/Ks) of 1.03.

2.5. Tissue-Specific Expression Pattern of DYW Subgroup PPR Genes

Different members may exhibit significant diversity in expression abundance among different tissues to adapt to different physiological process. To gain insight into the gene expression patterns within the organism in soybean growth and development, we investigated the relative transcript abundance of the 84 DYW subgroup PPR genes in six soybean tissues using publicly available RNA-seq data from Phytozome database, including flower, leaves, root, root hairs, seed, and stem, revealing that most DYW subgroup PPR genes had a similar pattern of expression in the same tissue. Out of 84 DYW subgroup PPR genes, most of them expressed in almost any tissues, while a total of four genes had no expression abundances to be discovered in any tissues, including Glyma.08G318700, Glyma.18G126700, Glyma.18G128100, and Glyma.18G225600. We found that most PPR genes showed preferential accumulation in leaves compared to other tissues, which were likely to be responsible for their subcellular localization, and most reported PPR proteins were located in mitochondria or chloroplast. A relatively high expression level was revealed in seeds, suggesting that they have potential function in seed development, which has been reported to be involved in endosperm development [14]. As shown in Figure 9, Glyma.11G008800, Glyma.13G220400 and Glyma.15G192000 showed extremely high expression in roots and root hairs, indicating that the DYW subgroup PPR genes may play specific roles in responding to drought stress.

2.6. Cis-Elements Analysis

By examining the expression levels of DYW subgroup PPR genes according to soybean drought- and salt-induced transcriptome database, we found that three genes were highly up-regulated to drought and salt stresses. Cis-elements presented in the upstream region play major roles in regulating the gene expression at the transcriptional level. To further understand the mechanism of the three genes, a set of 12 important cis-elements were identified in their 1.5 Kb 5′ flanking region upstream from the start codons. Various cis-elements including ABA-responsive element (ABRE) and MYB-binding site (MBS) were shown in Figure 10, which suggested that the three candidate genes might be involved in responses abiotic stresses.

2.7. Several Candidates Are Involved in Abiotic Stresses

To comprehensively understand the physiological functions of DYW subgroup PPR genes, we initially examined the expression patterns of three genes in response to salt and drought stresses by qRT-PCR (Figure 11). Under drought treatment, these three genes exhibited similar expression patterns, reaching a peak at 2 h (~ 4.5-fold, 3-fold and 2.5-fold, respectively), indicating that these genes have a similar biological function in the same environment. Similarly, the expression levels of the three genes were enhanced by salt; the expression peaks of GmPPR4/18/111 occurred at 4, 8 and 4 h, respectively, which are equivalent to 3-fold, 2.8-fold and 3.2-fold increases, respectively.

2.8. GmPPR4 Improved Drought Tolerance in Transgenic Soybean Hairy Roots

Among the three genes, GmPPR4 (Glyma.01G011300) clearly responded to drought and salt stresses. For this reason, GmPPR4 was selected for further investigation. To examine the function of GmPPR4 in vivo, transgenic soybean plants which overexpressed GmPPR4 (GmPPR4-OE) were generated into soybean hairy roots [26]. qRT-PCR analysis showed that GmPPR4 accumulated in the GmPPR4-OE plants. (Figure S2), and about 80% of the roots of transgenic soybean plants were positive. To examine whether GmPPR4 plays a role in the drought stress tolerance, we compared drought tolerance of GmPPR4-OE and WT plants at the vegetable stage; the hairy roots of seedlings which grew in soil were withheld from water for 2 weeks. No significant differences were observed for transgenic plants under normal growth conditions, compared to empty vector (EV)-transformed control hairy roots; drought treatment caused obvious differences in the growth of the EV-control and GmPPR4-OE plants; compared with the GmPPR4-OE, the EV-control plants showed wilted leaf under drought for 3 days, and seriously dehydrated leaves were observed after 7 days, whereas the GmPPR4-OE seedlings showed delayed and less leaf rolling during the drought stress process. (Figure 12A).
To explore the potential physiological mechanism responsible for the improved drought tolerance of the GmPPR4-OE seedlings, we compared some stress-related physiological changes in the EV-control and GmPPR4-OE plants under both normal growth and drought conditions. We found that proline accumulation in the transgenic plants was much more evident than in EV-control plants (Figure 12C), while the MDA content was decreased due to drought stress (Figure 12D). Furthermore, we also measured the level of H2O2 and O2- in roots; we found that the drought-treated EV-control roots accumulated much more H2O2 and O2- than the GmPPR4-OE roots (Figure 12E,F).
In addition, we used Trypan blue solution to detect cell activity in EV-control and GmPPR4-OE leaves. As shown in Figure 12B, no plant leaves differed under drought stress conditions, however, the color depth of the GmPPR4-OE leaves was lower than EV-control leaves under drought treatment, which indicated that the cell membrane integrity and stability in the leaves of the EV-control plants was better than that in the leaves of the GmPPR4-OE plants.
NaCl treatment was carried out. However, no obvious differences were observed in EV-control and GmPPR4-OE plants under NaCl treatment.

2.9. GmPPR4-OE Plants Exhibited Increased Transcripts of Some Drought-Inducible Genes

A previous study indicated that several genes play an important role in drought stress [27,28], including DREB2 [29], DREB3 [30], MYB84 [31], bZIP1 [32], bZIP44 [33], NAC11 [34], WRKY13 [35], and WRKY21 [35,36]. We compared the transcripts of several drought-inducible maker genes between GmPPR4-OE and EV-control plants under normal and drought conditions, and we found that drought induced more transcripts of all genes. However, expression differences between GmPPR4-OE and EV-control plants under drought treatment occurred in six genes, including DREB2, DREB3, bZIP1, NAC11, WRKY13, and WRKY21. There was no clear difference in bZIP44 and MYB84 expression with drought treatment between GmPPR4-OE and EV-control plants (Figure 13).

3. Discussion

Soybean is one of the most widely cultivated crops, with a total production of more than 260 million tons in 2010 (FAO data) [37]. Greenhouse and field studies showed that drought stress seriously affects plant development and led to significant reduction in crop yield [38]. Thus, when identifying ideal candidate genes, increasing drought stress resistance is essential for improving soybean yield.
PPR genes comprise a large family, which is ubiquitous to all terrestrial plants. In particular, previous reports indicated that a total of 4000 PPR family genes were identified in the spikemoss (S. moellendorffii) genome [3]. It is estimated that there were more than 1000 PPR genes in soybean. The duplication event was likely to be responsible for the large PPR family size, which has been experimentally verified in many species. Genome-wide identification of PPR family genes has been widely carried out in many species that have been sequenced [5,6,7]. In the present study, we identified 179 DYW subgroup PPR genes in the soybean genome (Glycine max Wm82.a2.v1). Consistent with previous studies, gene structure analysis revealed that most members of 179 DWY subgroup PPR genes were intron-less. Approximately 80% and 65% of PPR genes were free of intron in A. thaliana and rice, respectively [5,6]. The intron-less nature of the great majority of PPR genes may be the result of selection pressures during evolution. However, there are several PPR genes that tend to evolve diverse exon-intron organizations with more than 10 introns; we propose that they have a higher probability of evolving new specialized functions and adjusting to their living environment. The intron-less nature may demonstrate a duplication event of PPR genes [39]; we presumed that the nature also allows alternative splicing and alteration of splicing pattern in plants. In particular, many coding sequences of PPR proteins with extremely high similarity on the genome were divided only by frameshifts, for example, Glyma.13G332400 and Glyma.15G041800. In our study, GmPPR3 (Glyma.01G011300), GmPPR4 (Glyma.01G011700) and GmPPR94 (Glyma.09G209700) shared more than 95% identity in their coding sequence. In particular, many PPR genes have extremely close proximity on the genome, which contributed to the extensive gene-level synteny shared between them [24].
Previous studies have shown that PPR proteins were involved in RNA editing, plant growth and development and abiotic stress. Several PPR proteins have been implicated with trans-splicing of nad intron in A. thaliana, these include: OTP43 [40], SLO4 [41], PPR19 [42], BIR6 [43], MTL1 [44], ABO5 [21], OTP439 [45], and SG3 [46]. Lots of PPR proteins have a similar function in the development of seed or endosperm, such as EMP5 [47], PPR2263 [48], PPR8522 [49], OGR1 [50], AHG11 [17], and OTP43 [40]. In addition, in rice or A. thaliana, six PPR proteins: cisc(t) [20], ABO5 [21], ECB2 [51], PPR40 [22], SVR7 [52], and YS1 [53], were reported to be involved in abiotic stress responses, including ABA, drought, light and cold. In this study, a higher level of H2O2 and O2- were found in GmPPPR4-OE plants compared with EV-control plants under drought stress. In a previous study, the abo5 mutant accumulated higher H2O2 content in roots than wild type [21], and the ahg11 mutants exhibited higher transcript levels of oxidative stress-responsive genes [17]. The above results suggested that PPR genes may play roles in drought stress by regulating the level of H2O2.

4. Methods

4.1. Identification of DYW Subgroup PPR Genes in Soybean

DYW subgroup PPR proteins sequences of soybean were downloaded from Phytozome [54]. Predicted proteins from the soybean genome (Glycine max Wm82.a2.v1) were scanned using HMMER v3 [55] using the Hidden Markov Model (HMM) corresponding to the Pfam of PPR family (PF14432) [56]. The proteins obtained using the raw DYW HMM, a high-quality protein set (E-value < 1 × 10−20), were aligned and used to construct a soybean-specific DYW HMM profile using hmmbuild from the HMMER v3 suite. This new soybean-specific HMM was used to search for all members in all soybean proteins; in addition, all obtained proteins with an E-value lower than 0.01 were selected. The highly matched sequences were reorganized and merged to remove the redundancy. Then, all protein sequences of putative DYW subgroup PPR genes were submitted to the SMART database [57] and Pfam database to confirm the existence of the DYW conserved domain. The sequences lacking a DYW domain were refused in this study. Furthermore, molecular weights and isoelectric points identified that the DYW subgroup PPR proteins were obtained by using tools from EXPASY website [58].

4.2. Chromosomal Location and Phylogenetic Analysis

Positional information of DYW subgroup PPR genes on chromosomes of soybean was obtained from the Phytozome database. All DYW subgroup PPR genes were mapped to 20 chromosomes in soybean.
Multiple sequence alignment of proteins from soybean was performed via ClustalW with a default parameter. A phylogenetic tree was constructed by using the maximum likelihood with MEGA7 software, with the following parameters: Poisson model, pairwise deletion and 1000 bootstrap replications.

4.3. Gene Structure Analysis and the Sequence of PPR Motif Analysis

The gene structures of PPR were illustrated using the online program Gene Structure Display Server [59] by comparing predicted coding sequences with their corresponding genomic DNA sequences.
For sequence analysis of PPR motifs, the MEME online program [60] was used to identify conserved motifs. The scatter diagram was used to display the distribution of PPR motifs at the opposite positions in the PPR proteins by using the TBtools software v0.665 (Guangzhou, China) [61].

4.4. Gene Duplication

All Cluster I PPR proteins were searched for using the BLASTp search (E-value >1e−10) with more than 75% sequence similarity being considered to be a pair of tandem repeat genes. Then, the resulting file and GFF3 files of soybean genome (Glycine max Wm82.a2.v1) were used to analyze the gene duplication by software MCScanX and visualization using CIRCOS [62].

4.5. Tissue-Specific Expression Patterns of DWY subgroup PPR genes

Transcription databases were obtained from Phytozome database to investigate the tissue expression patterns of the soybean DYW subgroup PPR family; TBtools software was used to conduct a visual hierarchical clustering of the 84 DYW subgroup PPR genes under normal conditions. The transcript data are available in Table S3.

4.6. Promoter Sequence Analysis for Potential cis-Elements

For cis-elements analysis, 1.5 kb 5′ upstream region sequences were extracted from the Phytozome database. Then, the potential cis-elements of promotors for each gene were analyzed using PlantCARE database [63].

4.7. Plant Materials and Treatments

Soybean cultivar Zhonghuang 39 was used to analyze the gene expression pattern. The leaves of 15-day soybean seedlings were collected for RNA extraction and used to further qRT-PCR analysis. For drought stress, the soybean seedlings were placed on filter paper, for NaCl stress, the soybean seedlings were subjected to a 200 mM NaCl solution; samples were collected at 0, 1, 2, 4, 8, 12, 24, and 48 h after treatments. All treated leaf samples were frozen immediately in liquid and then stored at −80 °C for subsequent analysis.

4.8. RNA Extraction and qRT-PCR

Total RNA was extracted from plant samples using the Trizol method as the manufacturer’s protocol (TIANGEN, Beijing, China), which was reverse transcribed into cDNA using a PrimeScriptTM RT Reagent Kit (TaKaRa, Shiga, Jappa). All primers used in the study are shown in Table S4 [64].

4.9. Agrobacterium Rhizogenes-Mediated Transformation of Soybean Hairy Roots

Soybean Williams 82, a typical variety, was used to generate GmPPR4-OE soybean hairy roots. The coding sequences of GmPPR4 was ligated into plant transformation vector pCAMBIA3301 with the CaMV 35S promoter. The constructs were transferred into A. rhizogenes strain K599, and Agrobacteriumrhizo strain K599 harboring EV-control and GmPPR4-OE were injected at the cotyledonary node, as previously described [65]. The injected plants were transferred to the greenhouse with high humidity until hairy roots were generated at the infection site and had grown to about 5 cm long. The original main roots were cut off from 0.5 cm below the infection site. Seedlings were transplanted into nutritious soil and cultured normally in the greenhouse for a week (25 °C 16 h light/8 h dark photoperiod).

4.10. Drought and Salt Stress Assays

For drought treatment, one-week-old seedlings were subjected to dehydration for 15 days. After drought treatment, we re-watered soybean plants for 3 days. At the same time, we carried out salt treatment with 200 mM NaCl solution for 3 days [66].

4.11. Measurement of Proline Content, MDA Content, H2O2 Content and O2- Content

The content of proline, MDA, H2O2 and O2- were measured with the corresponding assay kit (Cominbio, Suzhou, China) based on the manufacturer’s protocols; all measurements were from four biological replicates.

4.12. Trypan Blue Staining

The leaves separated from the EV-control and GmPPR4-OE plants were placed on filter paper for the induction of rapid drought for 3 h, which were used to trypan stain, as previously described [23].

5. Conclusions

This study is the first time identifying the presence of 179 DYW subgroup PPR genes in the soybean genome (Glycine max Wm82.a2.v1) sequences, which were designated to GmPPR1 through GmPPR179 on the basis of their chromosomal location. We conducted a comprehensive and systematic analysis of the DYW subgroup PPR family. Based on the gene expression patterns under drought and salt stresses, we found that three PPR genes were highly up-regulated under salt and drought treatment, and GmPPR4 was selected for validating its role in drought stress tolerance. Compared with the EV-control plants, GmPPR4-OE exhibited drought tolerant phenotypes. Our results showed that GmPPR4 could improve tolerance to drought in soybean.

Supplementary Materials

Supplementary materials can be found at https://www.mdpi.com/1422-0067/20/22/5667/s1. Table S1. DYW-subgroup PPR genes in soybean. Detailed genomic information including Phytozome locus, class present, number of introns within ORF, protein, genomic locus (chromosomal location), +/− stand, isoelectric point, and molecular weight (kDa) of the PPR proteins for each PPR gene. Table S2. The Ka/Ks ratios for segmentally duplicated the Cluster I PPR proteins. Table S3. Transcript data of 84 DYW subgroup PPR genes for heat map. Table S4. The sequences of primers used in the study. Figure S1. The phylogenic tree of 12 selected proteins. Figure S2. qRT-PCR analysis of GmPPR4 expression in GmPPR4-OE and EV-control transgenic hairy roots.

Author Contributions

Z.-S.X. coordinated the project, conceived and designed experiments, and edited the manuscript; H.-G.S. performed experiments and wrote the first draft; B.L., X.-Y.S., and J.M. conducted the bioinformatic work and performed experiments; J.C., Y.-B.Z. and M.C. provided analytical tools and managed reagents; D.-H.M., Z.-S.X., and Y.-Z.M. contributed with valuable discussions. All authors have read and approved the final manuscript.

Funding

This research was financially supported the National Transgenic Key Project of the Chinese Ministry of Agriculture (2018ZX0800909B and 2016ZX08002-002), which were used to design the study and write the manuscript, the National Natural Science Foundation of China (31871624) was used to publish the results.

Acknowledgments

We are grateful to Lijuan Qiu and Shi Sun of the Institute of Crop Science, Chinese Academy of Agricultural Sciences for kindly providing soybean seeds.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

NCBINational Center for Biotechnology Information
PPRPentatricopeptiderepeat
qRT-PCRQuantitative real-time PCR
ABAabscisic acid
MDAmalondialdehyde
Proproline

References

  1. Hammani, K.; Bonnard, G.; Bouchoucha, A.; Gobert, A.; Pinker, F.; Salinas, T.; Giegé, P. Helical repeats modular proteins are major players for organelle gene expression. Biochimie 2014, 100, 141–150. [Google Scholar] [CrossRef] [PubMed]
  2. Small, I.D.; Peeters, N. The PPR motif—A TPR-Related motif prevalent in plant organellar proteins. Trends Biochem. Sci. 2000, 25, 45–47. [Google Scholar] [CrossRef]
  3. Cheng, S.; Gutmann, B.; Zhong, X.; Ye, Y.; Fisher, M.F.; Bai, F.; Liu, X. 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]
  4. Boussardon, C.; Avon, A.; Kindgren, P.; Bond, C.S.; Challenor, M.; Lurin, C.; Small, I. The cytidine deaminase signature HxE(x)nCxxC of DYW1 binds zinc and is necessary for RNA editing of ndhD-1. New Phytol. 2014, 203, 1090–1095. [Google Scholar] [CrossRef] [PubMed]
  5. Lurin, C.; Andrés, C.; Aubourg, S.; Bellaoui, M.; Bitton, F.; Bruyère, C.; Lecharny, A. 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]
  6. Chen, G.; Zou, Y.; Hu, J.; Ding, Y. Genome-wide analysis of the rice PPR gene family and their expression profiles under different stress treatments. BMC Genom. 2018, 19, 720. [Google Scholar] [CrossRef] [PubMed]
  7. Xing, H.; Fu, X.; Yang, C.; Tang, X.; Guo, L.; Li, C.; Luo, K. Genome-wide investigation of pentatricopeptide repeat gene family in poplar and their expression analysis in response to biotic and abiotic stresses. Sci. Rep. 2018, 8, 2817. [Google Scholar] [CrossRef] [PubMed]
  8. Klein, R.R.; Klein, P.E.; Mullet, J.E.; Minx, P.; Rooney, W.L.; Schertz, K.F. Fertility restorer locus Rf1 of sorghum (Sorghum bicolor L.) encodes a pentatricopeptide repeat protein not present in the colinear region of rice chromosome 12. Theor. Appl. Genet. 2005, 111, 994–1012. [Google Scholar] [CrossRef] [PubMed]
  9. Cui, X.; Wise, R.P.; Schnable, P.S. The rf2 nuclear restorer gene of male-sterile T-cytoplasm maize. Science 1996, 272, 1334–1336. [Google Scholar] [CrossRef] [PubMed]
  10. Liu, X.; Yu, F.; Rodermel, S. An Arabidopsis pentatricopeptide repeat protein, SUPPRESSOR OF VARIEGATION7, is required for FtsH-mediated chloroplast biogenesis. Plant Physiol. 2010, 154, 1588–1601. [Google Scholar] [CrossRef] [PubMed]
  11. 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]
  12. Banks, J.A.; Nishiyama, T.; Hasebe, M.; Bowman, J.L.; Gribskov, M.; DePamphilis, C.; Ashton, N.W. The Selaginella genome identifies genetic changes associated with the evolution of vascular plants. Science 2011, 332, 960–963. [Google Scholar] [CrossRef] [PubMed]
  13. Sun, F.; Xiu, Z.; Jiang, R.; Liu, Y.; Zhang, X.; Yang, Y.Z.; 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. 2018, 70, 963–972. [Google Scholar] [CrossRef] [PubMed]
  14. Hao, Y.; Wang, Y.; Wu, M.; Zhu, X.; Teng, Y.; Sun, Y.; Li, J. 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]
  15. Xiao, H.; Zhang, Q.; Qin, X.; Xu, Y.; Ni, C.; Huang, J.; Zhu, Y. Rice PPS1 encodes a DYW motif-containing pentatricopeptide repeat protein required for five consecutive RNA-editing sites of nad3 in mitochondria. New Phytol. 2018, 220, 878–892. [Google Scholar] [CrossRef] [PubMed]
  16. Ebihara, T.; Matsuda, T.; Sugita, C.; Ichinose, M.; Yamamoto, H.; Shikanai, T.; Sugita, M. The P-class pentatricopeptide repeat protein PpPPR_21 is needed for accumulation of the psbI-ycf12 dicistronic mRNA in Physcomitrella chloroplasts. Plant J. 2019, 97, 1120–1131. [Google Scholar] [CrossRef] [PubMed]
  17. Murayama, M.; Hayashi, S.; Nishimura, N.; Ishide, M.; Kobayashi, K.; Yagi, Y.; Hirayama, T. Isolation of Arabidopsisahg11, a weak ABA hypersensitive mutant defective in nad4 RNA editing. J. Exp. Bot. 2012, 63, 5301–5310. [Google Scholar] [CrossRef] [PubMed]
  18. 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 Arabidopsisthaliana and confirmation of the responsiveness PPR96 to abiotic stresses. Front. Plant Sci. 2016, 7, 1825. [Google Scholar] [CrossRef] [PubMed]
  19. 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] [PubMed]
  20. Lan, T.; Wang, B.; Ling, Q.; Xu, C.; Tong, Z.; Liang, K.; Wu, W. Fine mapping of cisc(t), a gene for cold-induced seedling chlorosis, and identification of its candidate in rice. Chin. Sci. Bull. 2010, 55, 3149–3153. [Google Scholar] [CrossRef]
  21. 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] [PubMed]
  22. Zsigmond, L.; Rigó, G.; Szarka, A.; Székely, G.; Ötvös, K.; Darula, Z.; Szabados, L. Arabidopsis PPR40 connects abiotic stress responses to mitochondrial electron transport. Plant Physiol. 2008, 146, 1721–1737. [Google Scholar] [CrossRef] [PubMed]
  23. Du, Y.T.; Zhao, M.J.; Wang, C.T.; Gao, Y.; Wang, Y.X.; Liu, Y.W.; Ma, Y.Z. Identification and characterization of GmMYB118 responses to drought and salt stress. BMC Plant Biol. 2018, 18, 320. [Google Scholar] [CrossRef] [PubMed]
  24. Mishra, A.K.; Muthamilarasan, M.; Khan, Y.; Parida, S.K.; Prasad, M. Genome-wide investigation and expression analyses of WD40 protein family in the model plant foxtail millet (Setaria italica L.). PLoS ONE 2014, 91, e86852. [Google Scholar] [CrossRef] [PubMed]
  25. Liu, J.M.; Xu, Z.S.; Lu, P.P.; Li, W.W.; Chen, M.; Guo, C.H.; Ma, Y.Z. Genome-wide investigation and expression analyses of the pentatricopeptide repeat protein gene family in foxtail millet. BMC Genom. 2016, 171, 840. [Google Scholar] [CrossRef] [PubMed]
  26. Gao, Y.; Ma, J.; Zheng, J.C.; Chen, J.; Chen, M.; Zhou, Y.B.; Ma, Y.Z. The Elongation Factor GmEF4 Is Involved in the Response to Drought and Salt Tolerance in Soybean. Int. J. Mol. Sci. 2019, 20, 3001. [Google Scholar] [CrossRef] [PubMed]
  27. Xu, Z.S.; Chen, M.; Li, L.C.; Ma, Y.Z. Functions of the ERF transcription factor family in plants. Botany 2008, 86, 969–977. [Google Scholar] [CrossRef]
  28. Xu, Z.S.; Chen, M.; Li, L.C.; Ma, Y.Z. Functions and Application of the AP2/ERF Transcription Factor Family in Crop Improvement, F. J. Integr. Plant Biol. 2011, 53, 570–585. [Google Scholar] [CrossRef] [PubMed]
  29. Chen, M.; Wang, Q.Y.; Cheng, X.G.; Xu, Z.S.; Li, L.C.; Ye, X.G.; Ma, Y.Z. GmDREB2, a soybean DRE-binding transcription factor, conferred drought and high-salt tolerance in transgenic plants. Biochem. Biophys. Res. Commun. 2007, 353, 299–305. [Google Scholar] [CrossRef] [PubMed]
  30. Nasreen, S.; Amudha, J.; Pandey, S.S. Isolation and characterization of Soybean DREB 3 transcriptional activator. J. Appl. Biol. Biotechnol. 2013, 1, 9–12. [Google Scholar]
  31. Wang, N.; Zhang, W.; Qin, M.; Li, S.; Qiao, M.; Liu, Z.; Xiang, F. Drought tolerance conferred in soybean (Glycine max. L.) by GmMYB84, a novel R2R3-MYB transcription factor. Plant Cell Physiol. 2017, 58, 1764–1776. [Google Scholar] [CrossRef] [PubMed]
  32. Gao, S.Q.; Chen, M.; Xu, Z.S.; Zhao, C.P.; Li, L.; Xu, H.J.; Ma, Y.Z. The soybean GmbZIP1 transcription factor enhances multiple abiotic stress tolerances in transgenic plants. Plant Mol. Biol. 2011, 75, 537–553. [Google Scholar] [CrossRef] [PubMed]
  33. Liao, Y.; Zou, H.F.; Wei, W.; Hao, Y.J.; Tian, A.G.; Huang, J.; Chen, S.Y. Soybean GmbZIP44, GmbZIP62 and GmbZIP78 genes function as negative regulator of ABA signaling and confer salt and freezing tolerance in transgenic Arabidopsis. Planta 2008, 228, 225–240. [Google Scholar] [CrossRef] [PubMed]
  34. Hao, Y.J.; Wei, W.; Song, Q.X.; Chen, H.W.; Zhang, Y.Q.; Wang, F.; Ma, B. Soybean NAC transcription factors promote abiotic stress tolerance and lateral root formation in transgenic plants. Plant J. 2011, 68, 302–313. [Google Scholar] [CrossRef] [PubMed]
  35. Zhou, Q.Y.; Tian, A.G.; Zou, H.F.; Xie, Z.M.; Lei, G.; Huang, J.; Chen, S.Y. Soybean WRKY-type transcription factor genes, GmWRKY13, GmWRKY21, and GmWRKY54, confer differential tolerance to abiotic stresses in transgenic Arabidopsis plants. Plant Biotechnol. J. 2008, 6, 486–503. [Google Scholar] [CrossRef] [PubMed]
  36. Vidal, R.O.; Nascimento, L.C.D.; Maurício Costa Mondego, J.; Amarante Guimarães Pereira, G.; Falsarella Carazzolle, M. Identification of SNPs in RNA-seq data of two cultivars of Glycine max (soybean) differing in drought resistance. Genet. Mol. Biol. 2012, 35, 331–334. [Google Scholar] [CrossRef] [PubMed]
  37. Ohyama, T.; Minagawa, R.; Ishikawa, S.; Yamamoto, M.; Hung, N.V.P.; Ohtake, N.; Sueyoshi, K.; Sato, T.; Nagumo, Y.; Takahashi, Y. Soybean seed production and nitrogen nutrition. In A Comprehensive Survey of International Soybean Research-Genetics, Physiology, Agronomy and Nitrogen Relationships; Board, J.E., Ed.; InTech: Rijeka, Croatia, 2013; pp. 115–157. [Google Scholar]
  38. Frederick, J.R.; Camp, C.R.; Bauer, P.J. Drought-Stress effects on branch and mainstem seed yield and yield components of determinate soybean. Crop Sci. 2001, 41, 759–763. [Google Scholar] [CrossRef]
  39. O’Toole, N.; Hattori, M.; Andres, C.; Iida, K.; Lurin, C.; Schmitz-Linneweber, C.; Small, I. On the expansion of the pentatricopeptide repeat gene family in plants. Mol. Biol. Evol. 2008, 25, 1120–1128. [Google Scholar] [CrossRef] [PubMed]
  40. De Longevialle, A.F.; Meyer, E.H.; Andrés, C.; Taylor, N.L.; Lurin, C.; Millar, A.H.; Small, I.D. The pentatricopeptide repeat gene OTP43 is required for trans-splicing of the mitochondrial nad1 intron 1 in Arabidopsis thaliana. Plant Cell 2007, 19, 3256–3265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Weißenberger, S.; Soll, J.; Carrie, C. The PPR protein SLOW GROWTH 4 is involved in editing of nad4 and affects the splicing of nad2 intron 1. Plant Mol. Biol. 2017, 93, 355–368. [Google Scholar] [CrossRef] [PubMed]
  42. Lee, K.; Han, J.H.; Park, Y.I.; Colas des Francs-Small, C.; Small, I.; Kang, H. The mitochondrial pentatricopeptide repeat protein PPR 19 is involved in the stabilization of NADH dehydrogenase 1 transcripts and is crucial for mitochondrial function and Arabidopsisthaliana development. New Phytol. 2017, 215, 202–216. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Koprivova, A.; des Francs-Small, C.C.; Calder, G.; Mugford, S.T.; Tanz, S.; Lee, B.R.; Kopriva, S. Identification of a pentatricopeptide repeat protein implicated in splicing of intron 1 of mitochondrial nad7 transcripts. J. Biol. Chem. 2010, 285, 32192–32199. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Haïli, N.; Planchard, N.; Arnal, N.; Quadrado, M.; Vrielynck, N.; Dahan, J.; Mireau, H. The MTL1 pentatricopeptide repeat protein is required for both translation and splicing of the mitochondrial NADH DEHYDROGENASE SUBUNIT7 mRNA in Arabidopsis. Plant Physiol. 2016, 170, 354–366. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Des Francs-Small, C.C.; de Longevialle, A.F.; Li, Y.; Lowe, E.; Tanz, S.; Smith, C.; Small, I. The PPR proteins TANG2 and OTP439 are involved in the splicing of the multipartite nad5 transcript encoding a subunit of mitochondrial complex I. Plant Physiol. 2014, 114. [Google Scholar] [CrossRef] [Green Version]
  46. Hsieh, W.Y.; Liao, J.C.; Chang, C.Y.; Harrison, T.; Boucher, C.; Hsieh, M.H. The SLOW GROWTH3 pentatricopeptide repeat protein is required for the splicing of mitochondrial NADH dehydrogenase subunit7 intron 2 in Arabidopsis. Plant Physiol. 2015, 168, 490–501. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Zheng, P.; He, Q.; Wang, X.; Tu, J.; Zhang, J.; Liu, Y.J. Functional analysis for domains of maize PPR protein EMP5 in RNA editing and plant development in Arabidopsis. Plant Growth Regul. 2019, 87, 19–27. [Google Scholar] [CrossRef]
  48. Sosso, D.; Mbelo, S.; Vernoud, V.; Gendrot, G.; Dedieu, A.; Chambrier, P.; Rogowsky, P.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]
  49. 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] [PubMed] [Green Version]
  50. Kim, S.R.; Yang, J.I.; Moon, S.; Ryu, C.H.; An, K.; Kim, K.M.; 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] [PubMed]
  51. Huang, C.; Yu, Q.B.; Li, Z.R.; Ye, L.S.; Xu, L.; Yang, Z.N. Porphobilinogen deaminase HEMC interacts with the PPR-protein AtECB2 for chloroplast RNA editing. Plant J. 2017, 92, 546–556. [Google Scholar] [CrossRef] [PubMed]
  52. Zoschke, R.; Qu, Y.; Zubo, Y.O.; Börner, T.; Schmitz-Linneweber, C. Mutation of the pentatricopeptide repeat-SMR protein SVR7 impairs accumulation and translation of chloroplast ATP synthase subunits in Arabidopsis thaliana. J. Plant Res. 2013, 126, 403–414. [Google Scholar] [CrossRef] [PubMed]
  53. Zhou, W.; Cheng, Y.; Yap, A.; Chateigner-Boutin, A.L.; Delannoy, E.; Hammani, K.; Huang, J. The Arabidopsis gene YS1 encoding a DYW protein is required for editing of rpoB transcripts and the rapid development of chloroplasts during early growth. Plant J. 2009, 58, 82–96. [Google Scholar] [CrossRef] [PubMed]
  54. Finn, R.D.; Clements, J.; Eddy, S.R. HMMER web server: Interactive sequence similarity searching. Nucleic Acids Res. 2011, 39, W29–W37. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Goodstein, D.M.; Shu, S.; Howson, R.; Neupane, R.; Hayes, R.D.; Fazo, J.; Mitros, T.; Dirks, W.; Hellsten, U.; Putnam, N.; et al. Phytozome: A comparative platform for green plant genomics. Nucleic Acids Res. 2012, 40, D1178–D1186. [Google Scholar] [CrossRef] [PubMed]
  56. Finn, R.D.; Bateman, A.; Clements, J.; Coggill, P.; Eberhardt, R.Y.; Eddy, S.R.; Sonnhammer, E.L. Pfam: The protein families database. Nucleic Acids Res. 2013, 42, D222–D230. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Letunic, I.; Doerks, T.; Bork, P. SMART 7: Recent updates to the protein domain annotation resource. Nucleic Acids Res. 2012, 40, D302–D305. [Google Scholar] [CrossRef] [PubMed]
  58. Artimo, P.; Jonnalagedda, M.; Arnold, K.; Baratin, D.; Csardi, G.; de Castro, E.; Duvaud, S.; Flegel, V.; Fortier, A.; Gasteiger, E.; et al. ExPASy: SIB bioinformatics resource portal. Nucleic Acids Res. 2012, 40, W597–W603. [Google Scholar] [CrossRef] [PubMed]
  59. Guo, A.Y.; Zhu, Q.H.; Chen, X.; Luo, J.C. GSDS: A gene structure display server. Yi Chuan 2007, 29, 1023–1026. [Google Scholar] [CrossRef] [PubMed]
  60. Bailey, T.L.; Boden, M.; Buske, F.A.; Frith, M. MEME SUITE: Tools for motif discovery and searching. Nucleic Acids Res. 2009, 37, W202–W208. [Google Scholar] [CrossRef] [PubMed]
  61. Chen, C.; Xia, R.; Chen, H.; He, Y. TBtools, a Toolkit for Biologists integrating various biological data handling tools with a user-friendly interface. bioRxiv 2018, 289660. [Google Scholar] [CrossRef]
  62. Krzywinski, M.; Schein, J.; Birol, I.; Connors, J. Circos: An information aesthetic for comparative genomics. Genome Res. 2009, 19, 1639–1645. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Lescot, M.; Déhais, P.; Thijs, G.; Marchal, K.; Moreau, Y. PlantCARE, a database of plant cis-acting regulatory elements and a portal to tools for in silico analysis of promoter sequences. Nucleic Acids Res. 2002, 30, 325–327. [Google Scholar] [CrossRef] [PubMed]
  64. Cui, X.Y.; Gao, Y.; Guo, J.; Yu, T.F.; Zheng, W.J.; Liu, Y.W.; Ma, Y.Z. BES/BZR Transcription Factor TaBZR2 Positively Regulates Drought Responses by Activation of TaGST1. Plant Physiol. 2019, 180, 605–620. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Wang, F.; Chen, H.W.; Li, Q.T.; Wei, W.; Li, W.; Zhang, W.K.; Man, W.Q. GmWRKY27 interacts with GmMYB174 to reduce expression of GmNAC 29 for stress tolerance in soybean plants. Plant J. 2015, 83, 224–236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Xu, Z.S.; Ni, Z.Y.; Liu, L.; Nie, L.N.; Li, L.C.; Chen, M.; Ma, Y.Z. Characterization of the TaAIDFa gene encoding a CRT/DRE-binding factor responsive to drought, high-salt, and cold stress in wheat. Mol. Genet. Genom. 2008, 280, 497–508. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Relative proportions of intron-containing PPR genes in soybean. (A) Number of introns in the 179 DYW subgroup PPR genes. (B) Number of introns in the 84 Cluster I PPR genes.
Figure 1. Relative proportions of intron-containing PPR genes in soybean. (A) Number of introns in the 179 DYW subgroup PPR genes. (B) Number of introns in the 84 Cluster I PPR genes.
Ijms 20 05667 g001
Figure 2. Chromosomal distribution of the 179 DYW subgroup PPR genes in soybean. (A) The physical location of each member. (B) PPR genes distribution on 20 soybean chromosomes.
Figure 2. Chromosomal distribution of the 179 DYW subgroup PPR genes in soybean. (A) The physical location of each member. (B) PPR genes distribution on 20 soybean chromosomes.
Ijms 20 05667 g002
Figure 3. Phylogenetic tree of DYW subgroup PPR genes. The complete amino acid sequences of the 179 DYW subgroup PPR proteins were aligned by ClustalW and Maximum-likelihood with MEGA7. Three discrete groups (Cluster I to III) were highlighted in different colors.
Figure 3. Phylogenetic tree of DYW subgroup PPR genes. The complete amino acid sequences of the 179 DYW subgroup PPR proteins were aligned by ClustalW and Maximum-likelihood with MEGA7. Three discrete groups (Cluster I to III) were highlighted in different colors.
Ijms 20 05667 g003
Figure 4. Multiple sequence alignment of 12 DYW subgroup PPR proteins from A. thaliana and soybean by DNAMAN. Black or blue shadings represent the 100% or >75% similarity of amino acids. Red rectangle marks highly core short signatures.
Figure 4. Multiple sequence alignment of 12 DYW subgroup PPR proteins from A. thaliana and soybean by DNAMAN. Black or blue shadings represent the 100% or >75% similarity of amino acids. Red rectangle marks highly core short signatures.
Ijms 20 05667 g004
Figure 5. Phylogenetic relationship and structural analysis of the 84 Cluster I PPR genes. The phylogenetic tree was constructed via MEGA7.0 software; the different classes of PPR genes make up separate clades. The schematic diagram was carried out to represent the gene structure. Introns and exons were indicated by black lines and purple boxes, respectively. The lengths of introns and exons of each gene were displayed proportionally.
Figure 5. Phylogenetic relationship and structural analysis of the 84 Cluster I PPR genes. The phylogenetic tree was constructed via MEGA7.0 software; the different classes of PPR genes make up separate clades. The schematic diagram was carried out to represent the gene structure. Introns and exons were indicated by black lines and purple boxes, respectively. The lengths of introns and exons of each gene were displayed proportionally.
Ijms 20 05667 g005
Figure 6. Putative motifs of each Cluster I PPR gene by MEME program and TBtools software. Ten putative motifs were indicated in colored boxes. The length of protein can be estimated using the scale at the bottom.
Figure 6. Putative motifs of each Cluster I PPR gene by MEME program and TBtools software. Ten putative motifs were indicated in colored boxes. The length of protein can be estimated using the scale at the bottom.
Ijms 20 05667 g006
Figure 7. Sequence logo of global multiple alignment of 179 DYW subgroup PPR proteins by MEME program.
Figure 7. Sequence logo of global multiple alignment of 179 DYW subgroup PPR proteins by MEME program.
Ijms 20 05667 g007
Figure 8. Distribution of segmentally duplicated Cluster I PPR genes on soybean chromosomes. Green lines indicate duplicated PPR gene pairs, and bold Photozome Locus indicates the gene appears twice.
Figure 8. Distribution of segmentally duplicated Cluster I PPR genes on soybean chromosomes. Green lines indicate duplicated PPR gene pairs, and bold Photozome Locus indicates the gene appears twice.
Ijms 20 05667 g008
Figure 9. Heat map of expression profiles (in log10-based FPKM) of all DYW subgroup PPR genes from six soybean tissues (flower, leaves, root, root hairs, seed, and stem). The expression abundance of each transcript is represented by the color bar: Red, higher expression; and green, lower expression.
Figure 9. Heat map of expression profiles (in log10-based FPKM) of all DYW subgroup PPR genes from six soybean tissues (flower, leaves, root, root hairs, seed, and stem). The expression abundance of each transcript is represented by the color bar: Red, higher expression; and green, lower expression.
Ijms 20 05667 g009
Figure 10. Putative cis-element in a 1.5 kb 5′ flanking region upstream from the start codon. Different cis-elements were indicated by colored symbols and placed in relative positions on the promoter. The ABA-responsive element (ABRE), light-responsive element (ACE), anaerobic induction element (ARE), auxin responsive element (AuxRR-core), light-responsive element (Box4), MeJA-responsive (CGTCA-motif), light-responsive element (GT1-motif), low temperature responsive element (LTR), MYB-binding site (MBS), light-responsive element (Sp1), salicylic acid responsive element (TCA), and defense and stress responsive element (TC-rich repeat) were analyzed.
Figure 10. Putative cis-element in a 1.5 kb 5′ flanking region upstream from the start codon. Different cis-elements were indicated by colored symbols and placed in relative positions on the promoter. The ABA-responsive element (ABRE), light-responsive element (ACE), anaerobic induction element (ARE), auxin responsive element (AuxRR-core), light-responsive element (Box4), MeJA-responsive (CGTCA-motif), light-responsive element (GT1-motif), low temperature responsive element (LTR), MYB-binding site (MBS), light-responsive element (Sp1), salicylic acid responsive element (TCA), and defense and stress responsive element (TC-rich repeat) were analyzed.
Ijms 20 05667 g010
Figure 11. Expression patterns of three selected PPR genes under salt and drought treatments by qRT-PCR. GmPPR4 (Glyma.01G011700), GmPPR18 (Glyma.02G174500), GmPPR111 (Glyma.11G008800), respectively. The actin gene was used as an internal control. The data are shown as the means ± SD obtained from three biological replicates. ANOVA test demonstrated that there were significant differences (* p < 0.05, ** p < 0.01).
Figure 11. Expression patterns of three selected PPR genes under salt and drought treatments by qRT-PCR. GmPPR4 (Glyma.01G011700), GmPPR18 (Glyma.02G174500), GmPPR111 (Glyma.11G008800), respectively. The actin gene was used as an internal control. The data are shown as the means ± SD obtained from three biological replicates. ANOVA test demonstrated that there were significant differences (* p < 0.05, ** p < 0.01).
Ijms 20 05667 g011
Figure 12. Drought stress analysis of EV-control transgenic plants and GmPPR4-OE transgenic plants. (A) Phenotypes of GmPPR4-OE and EV-control transgenic soybean plants subjected to drought stress without water for 3 days and 7 days. (B) Trypan blue staining of soybean plant leaves placed on filter paper for the induction of rapid drought for 3 h, the dead cells can be strained, but living cells cannot. (C) The proline content, (D) malondialdehyde (MDA) content, (E) H2O2 content, and (F) O2 content of GmPPR4-OE and EV-control transgenic soybean plants under normal and drought conditions. The data are shown as the means ± SD obtained from three biological replicates. ANOVA test demonstrated that there were significant differences (* p < 0.05, ** p < 0.01).
Figure 12. Drought stress analysis of EV-control transgenic plants and GmPPR4-OE transgenic plants. (A) Phenotypes of GmPPR4-OE and EV-control transgenic soybean plants subjected to drought stress without water for 3 days and 7 days. (B) Trypan blue staining of soybean plant leaves placed on filter paper for the induction of rapid drought for 3 h, the dead cells can be strained, but living cells cannot. (C) The proline content, (D) malondialdehyde (MDA) content, (E) H2O2 content, and (F) O2 content of GmPPR4-OE and EV-control transgenic soybean plants under normal and drought conditions. The data are shown as the means ± SD obtained from three biological replicates. ANOVA test demonstrated that there were significant differences (* p < 0.05, ** p < 0.01).
Ijms 20 05667 g012
Figure 13. GmPPR4-OE plants exhibited increased transcripts of some drought-inducible genes. The two-week-old soybean seedlings were placed on filter paper for 3 h. qRT-PCR analysis for drought-inducible genes (AH). The actin gene was used as an internal control. The data were shown as the means ± SD obtained from three biological replicates. ANOVA test demonstrated that there were significant differences (* p < 0.05, ** p < 0.01).
Figure 13. GmPPR4-OE plants exhibited increased transcripts of some drought-inducible genes. The two-week-old soybean seedlings were placed on filter paper for 3 h. qRT-PCR analysis for drought-inducible genes (AH). The actin gene was used as an internal control. The data were shown as the means ± SD obtained from three biological replicates. ANOVA test demonstrated that there were significant differences (* p < 0.05, ** p < 0.01).
Ijms 20 05667 g013

Share and Cite

MDPI and ACS Style

Su, H.-G.; Li, B.; Song, X.-Y.; Ma, J.; Chen, J.; Zhou, Y.-B.; Chen, M.; Min, D.-H.; Xu, Z.-S.; Ma, Y.-Z. Genome-Wide Analysis of the DYW Subgroup PPR Gene Family and Identification of GmPPR4 Responses to Drought Stress. Int. J. Mol. Sci. 2019, 20, 5667. https://doi.org/10.3390/ijms20225667

AMA Style

Su H-G, Li B, Song X-Y, Ma J, Chen J, Zhou Y-B, Chen M, Min D-H, Xu Z-S, Ma Y-Z. Genome-Wide Analysis of the DYW Subgroup PPR Gene Family and Identification of GmPPR4 Responses to Drought Stress. International Journal of Molecular Sciences. 2019; 20(22):5667. https://doi.org/10.3390/ijms20225667

Chicago/Turabian Style

Su, Hong-Gang, Bo Li, Xin-Yuan Song, Jian Ma, Jun Chen, Yong-Bin Zhou, Ming Chen, Dong-Hong Min, Zhao-Shi Xu, and You-Zhi Ma. 2019. "Genome-Wide Analysis of the DYW Subgroup PPR Gene Family and Identification of GmPPR4 Responses to Drought Stress" International Journal of Molecular Sciences 20, no. 22: 5667. https://doi.org/10.3390/ijms20225667

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