Next Article in Journal
Efficient Simultaneous Introduction of Premature Stop Codons in Three Tumor Suppressor Genes in PFFs via a Cytosine Base Editor
Previous Article in Journal
High-Throughput Sequencing Reveals Transcriptome Signature of Early Liver Development in Goat Kids
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genome-Wide Identification and Transcriptional Expression Profiles of PP2C in the Barley (Hordeum vulgare L.) Pan-Genome

1
State Key Laboratory of Crop Stress Biology for Arid Areas, College of Agronomy, Northwest A&F University, Yangling, Xianyang 712100, China
2
College of Agronomy, Jiangxi Agricultural University, Nanchang 330045, China
*
Authors to whom correspondence should be addressed.
Genes 2022, 13(5), 834; https://doi.org/10.3390/genes13050834
Submission received: 7 April 2022 / Revised: 29 April 2022 / Accepted: 3 May 2022 / Published: 7 May 2022
(This article belongs to the Section Plant Genetics and Genomics)

Abstract

:
The gene family protein phosphatase 2C (PP2C) is related to developmental processes and stress responses in plants. Barley (Hordeum vulgare L.) is a popular cereal crop that is primarily utilized for human consumption and nutrition. However, there is little knowledge regarding the PP2C gene family in barley. In this study, a total of 1635 PP2C genes were identified in 20 barley pan-genome accessions. Then, chromosome localization, physical and chemical feature predictions and subcellular localization were systematically analyzed. One wild barley accession (B1K-04-12) and one cultivated barley (Morex) were chosen as representatives to further analyze and compare the differences in HvPP2Cs between wild and cultivated barley. Phylogenetic analysis showed that these HvPP2Cs were divided into 12 subgroups. Additionally, gene structure, conserved domain and motif, gene duplication event detection, interaction networks and gene expression profiles were analyzed in accessions Morex and B1K-04-12. In addition, qRT-PCR experiments in Morex indicated that seven HvMorexPP2C genes were involved in the response to aluminum and low pH stresses. Finally, a series of positively selected homologous genes were identified between wild accession B1K-04-12 and another 14 cultivated materials, indicating that these genes are important during barley domestication. This work provides a global overview of the putative physiological and biological functions of PP2C genes in barley. We provide a broad framework for understanding the domestication- and evolutionary-induced changes in PP2C genes between wild and cultivated barley.

1. Introduction

Phosphorylation and dephosphorylation are a pair of important protein modification processes that can influence multiple metabolic and biochemical reactions [1]. Depending on the primary sequence and catalytic mechanism, protein phosphatases (PPS) are divided into four major classes: phosphoprotein phosphatases (PPP), Mg2+- or Mn2+-dependent protein phosphatase (PPM)/protein phosphatase 2C (PP2C), phosphotyrosine phosphatase (PTP) and aspartate (Asp)-dependent enzyme families [1]. Previous studies showed that PP2C is related to developmental processes and stress responses in plants. In the nucleus, AtPP2C5 acts as an MAPK phosphatase that is involved in seed germination, ABA-inducible gene expression and stomatal closure [2]. In Arabidopsis thaliana, AP2C3 also mediates stomatal development by inducing the ectopic proliferation of epidermal cells [3]. In Komatsu’s study [4], PpABI1B single disruptants (ppabi1b) and PpABI1 double disruptants (ppabi1a/b) and double disruptants (ppabi1a/b) were constructed. The ppabi1a/b of group A PP2C enhanced the tolerance of water-associated stresses by accumulating LEA proteins and soluble sugars at the maximum level [4]. By constructing transgenic AtPP2CA-antisense plants, Sari found that AtPP2CA is a negative regulator of ABA responses during cold acclimation [5].
Barley (H. vulgare) is one of the world’s oldest cultivated crops, and is the fourth most abundant cereal in terms of both area and tonnage harvested (FAO: http://faosta.fao.org, accessed on 1 December 2021). Barley is mainly used for feeding and human consumption, such as in the brewing industry [6]. Compared with wheat, barley shows stronger environmental adaptability. Therefore, it can grow in harsh environments and maintain harvestable yields, allowing growers to avoid food crises. Barley is a diploid plant and has 14 chromosomes. In the Triticeae, barley is an excellent model for plant genetic research due to its abundant genetic diversity and small genome (5.1 G) [7].
It is generally accepted that barley domestication is a single event that happened in the Fertile Crescent [8]. Nevertheless, some other researchers consider Tibet or the Near East as secondary centers of origin [9,10,11]. During the domestication and improvement process, Btr, Ppd, Vrs/int-c, nud, SD and Vrn/Sgh have been regarded as the most critical genes or QTLs controlling barley shattering, photoperiod, row type, naked caryopsis, grain dormancy and vernalization. Among them, shattering has attracted widespread concern as a domestication gene because it is the only trait that directly distinguishes wild and cultivated barley. Non-brittle rachis 1 (btr1) and non-brittle rachis 2 (btr2) are two tightly linked genes located on barley chromosome 3H [12,13]. On the one hand, Btr1 and Btr2 are dominant alleles that generate wild barley with brittle rachises and eventual shattering [14,15]. On the other hand, the genotype of nonbrittle rachis cultivated barley is either btr1Btr2 (btr1-type) or Btr1btr2 (btr2-type), resulting in non-shattering spikes [16]. Another essential gene of domestication is Ppd-H1, which regulates flowering time and is located on the short arm of chromosome 2H [17,18]. According to James’ findings, while mutant alleles at the Ppd-H1 and Ppd-H1 photoperiod loci arose before domestication, mutant vernalization nonresponsive alleles at the VRN-H1 and VRN-H2 loci occurred after domestication [19,20]. However, the time and place of barley domestication are still unclear. Additionally, little is known about the mechanism of domestication. Thus, the study of domestication and evolutionary relationships might reveal information about barley cultivation expansion and subsequent growth, including the crop’s adaptability to new settings and reaction to human selection.
In the past decade, owing to the high degree of genomic variation, research has demonstrated that a single reference genome does not fully represent the diversity of a species [21]. The concept of pan-genomes was first proposed by Tettelin in 2005 [22]. Pan-genomes include multiple genetic resources and show more information about the entire species. Therefore, barley pan-genomes are valuable for research on genetic studies, barley domestication and improvement, adaptive evolution and plant breeding [22]. In recent years, some studies have been conducted to investigate genomic variations and functional genomics, particularly in rice [23] and soybeans [24]. The first barley pan-genomes involving 20 accessions were assembled in 2020, and included one wild barley and nineteen cultivated barley accessions [25]. The cultivated accessions include the reference cultivar Morex, two current or former elite malting varieties, two founder lines of Chinese barley breeding, two genotypes with high transformation efficiency, and a successful German variety [25].
Barley’s domestication process and adaptation to evolution have long been the focus of researchers all over the world. The recent sequencing of the barley pan-genome provides an excellent opportunity to accomplish these missions. In this sense, we identified HvPP2Cs in this barley pan-genome, reported their protein characteristics and chose two accessions, B1K-04-12 and Morex, to represent wild and cultivar materials, respectively. Then, we compared phylogenetics, and multiple structural and functional characterizations of HvPP2Cs between these two accessions. Finally, we analyzed the expression pattern of different tissues and treatments under different stresses. Our first aim in this study was to provide a global overview of the putative physiological and biological functions of PP2C genes in barley. We provide a broad framework for understanding domestication- and evolutionary-induced changes in PP2C genes in wild and cultivated barley. We also intend to provide insights into the genetic improvement of barley breeding strategies.

2. Material and Methods

2.1. Identification of PP2C Genes in H. vulgare

To identify PP2C genes in H. vulgare, the barley pan-genome sequences and their corresponding annotation information were downloaded from https://webblast.ipk-gatersleben.de/downloads/barley_pangenome/, accessed on 1 December 2021 [25]. Then, the known PP2Cs in Arabidopsis thaliana were obtained from the EKPD 1.1 database (http://ekpd.biocuckoo.org/advance.php, accessed on 1 December 2021) [26]. Afterwards, a HMMER (biosequence analysis using profile hidden Markov models) [27] search was conducted on 20 barley pan-genome protein sequences using pfam [28] Stockholm seeds PF00481 and PF13672 [29]. Meanwhile, a blastp [30] procedure was produced with default parameters using 82 AtPP2Cs as a query against these pan-genome protein sequences. The results of the HMMER search and blastp were integrated using a Perl [31] script. Furthermore, conserved domains were predicted using the NCBI CDD/SPARCLE database [32]. Then, the sequences that lacked complete domains belonging to PP2C or PP2C super families were removed. Additionally, the remaining sequences were used as candidates for further analysis.

2.2. Prediction of Protein Physical and Chemical Parameters and Subcellular Localization

The physical and chemical parameters of identified protein sequences were calculated using the Expasy ProtParam tool [33], including molecular weight (MW), theoretical isoelectric point (pI) and amino acid composition. WoLF PSORT online service [34] predicted protein subcellular localization.

2.3. Alignment and Phylogenetic Analysis

Multiple alignments of peptide sequences were carried out using the MEGA version X [35,36] built-in ClustalW method. Then, to find the best protein model for the maximum likelihood (ML) tree, a model test was run in MEGA-X with default parameters. A maximum likelihood tree was then constructed using the best model found in MEGA-CC 10.0.4 [37] with Morex, B1K-04-12 and Arabidopsis thaliana PP2C protein sequences. The topology of the phylogenetic tree was used to classify HvPP2Cs into subgroups, referring to the previous studies on Arabidopsis thaliana and Oryza sativa [38], paper mulberry [39], woodland strawberry [40] and potato [41]. Then, all candidate sequences were renamed based on their subgroup classifications.

2.4. Analysis of Gene Structure, Chromosomal Locations and Conserved Motifs

Conserved motifs were predicted using MEME Suite Program Version 5.4.1 [42] for all HvPP2C sequences. The gene structure and chromosomal location information was extracted from the genome annotation gff3 files. Graphics were displayed using TBtools [43].

2.5. Analysis of Synteny and Detection of Duplication Events

The synteny analysis between the Arabidopsis genome, Morex genome and B1K-04-12 genome was investigated in TBtools using the plugin MCScanX method. The images of synteny analysis and duplication events were then visualized using TBtools software [43].

2.6. Calculation of Ka/Ks Values

The synonymous (Ks) and nonsynonymous (Ka) substitution rates of homologue genes in Morex and B1K-04-12 were calculated using TBtools software [43].

2.7. Analysis of Cis-Acting Elements in the Promoter Regions

The 2000 bp upstream genomic sequences of identified HvPP2C genes were extracted from each genome of 20 barley pan-genomes. Then, the upstream sequences were delivered to the PlantCARE database (http://bioinformatics.psb.ugent.be/webtools/plantcare/html/, accessed on 1 December 2021) to identify cis-acting elements in the promoter regions.

2.8. Network Interaction Analysis

The interaction network in which the HvPP2Cs are involved was investigated using the ExPASy String tool [44]. Then, GO and KEGG annotation and enrichment analysis were implemented using eggNOG-mapper v2 [45,46] and visualized in TBtools [43].

2.9. Transcriptional Analysis

A total of 524 barley RNA-seq data accessions (Table S1) from 13 independent bioprojects (PRJNA294716, PRJNA382490, PRJNA400519, PRJNA428086, PRJNA431836, PRJNA471777, PRJNA541021, PRJNA578897, PRJNA658250, PRJNA668924, PRJNA681455, PRJNA704034 and PRJNA749617) [47,48,49,50,51,52] were obtained from the NCBI Sequence Read Archive (SRA) database. Then, Hisat2 [53] and Stringtie [54,55] were used to analyze the gene expression level. Additionally, the FPKM value (fragments per kilobase of transcript per million fragments mapped) was calculated for each PP2C gene. Morex was used as the reference genome. To obtain a more distinct heatmap, log10-transformed (FPKM+1) values were used to display the results.

2.10. Plant Materials and Growth Conditions

Barley seeds (H. vulgare L. cv. Morex) were surface-sterilized then soaked in distilled water in a greenhouse. The low-pH and aluminum treatments were performed in a hydroponic environment. Briefly, seeds of barley were surface-sterilized in 5% sodium hypochlorite and incubated in the dark at 4 °C for stratification. Then, the seeds were placed on Petri dishes filled with moist filter paper and placed in a growth chamber at 25 °C in the dark. After 48 h, the germinated seeds were transferred to 21 hydroponic containers containing 2.5 L of Hoagland’s nutrient solution for 7 days, and the medium was changed every 24 h to maintain their composition. After 7 days, seedlings were cultured at pH = 6.0, pH = 4.0 and pH = 4.0 with 10 μM of bioavailable Al3+ ions for 1 or 7 days, respectively. A maximum of 12 seedlings were placed in one container, which was considered as one replicate, and each experimental combination was set up as three replicates.

2.11. RNA Isolation and qRT-PCR Analysis

Total RNA was extracted from the roots and shoots of each treated 14-day-old barley seedling using the TIANGEN RNA simple Total RNA Extraction kit (DP419). The quality of total RNA was detected by 1% agarose gel electrophoresis. The concentration of RNA was detected using a NanoDrop-2000 UV spectrophotometer. Samples that had an A260/A280 between 1.8 and 2.0 were used for reverse transcription, and then reverse transcribed into cDNA using an Evo M-MLV reverse transcription kit (Accurate Biotechnology (Hunan) Co., Ltd., Hunan, China). The reverse transcription assay was based on a 20 µL reaction volume containing 1 µg of total RNA. A SYBR Green Premix Pro Taq HS qPCR Kit (Accurate Biotechnology (Hunan) Co., Ltd., Hunan, China) was used for the qPCR experiment. The qPCR assay was based on a 20 µL reaction volume containing 10 µL of 2X SYBR Green Pro Taq HS Premix (ROX plus), 4 µL of each forward and reverse primer and 2 µL of cDNA. The amplification reactions were performed using the QuantStudio 7 Flex Real-time Fluorescent Quantitative PCR System with an initial denaturation at 95 °C for 30 s, followed by 40 cycles of amplification at 95 °C for 5 s and 58 °C for 34 s. Melting curves were acquired using the following procedure: 95 °C for 15 s, 58 °C for 60 s and 95 °C for 15 s. The used primers are listed in Table S2. One-way ANOVA and Waller–Duncan multiple comparisons were performed between treatments using SPSS22 [56]. The results were visualized in R Studio [57].

3. Results

3.1. Genome-Wide Identification of PP2C Genes in the Barley Pan-Genome

A total of 1635 candidate HvPP2C genes were identified from 20 barley pan-genome accessions. The number of each accession is shown in Table 1. Both an HMMER search and the Blastp program were used for identification. The sequences were removed for further study, which revealed that they lacked complete conserved domains belonging to PP2C super families. As a result, Hockett had the most members with 88 HvPP2Cs, followed by Morex with 85 HvPP2Cs. Among all cultivated barley, Igri had the fewest members with 79 HvPP2Cs, followed by Akashinriki, Barke, Golden_Promise and HOR_3365, which have 80 HvPP2Cs. There was just one wild accession, B1K-04-12, among the 20 pan-genome accessions, and it had the smallest HvPP2C group, with only 79 members. The number of HvPP2C genes in each accession is similar to that of Arabidopsis thaliana, which contains a total of 82 sequences (EKPD: http://ekpd.biocuckoo.org/, accessed on 1 December 2021).

3.2. Physical and Chemical Feature Prediction and Subcellular Localization of HvPP2Cs

The physical and chemical features of all candidate genes were predicted by Expasy online services, and the results are shown in Table S3. The lengths of the candidate genes varied from 116 amino acids to 1744 amino acids in Hockett. Each accession had a different average length, ranging from 403 amino acids in HOR 3081 to 429 amino acids in Hockett. Furthermore, the theoretical isoelectric point (pI) and average molecular weight (MW) ranged from 4.35 to 11.64 kDa and 12,769.69 to 191,804.35, respectively.
A summary of the subcellular localization of the two chosen accessions in this study is shown in Table 2. The results show that the PP2Cs are randomly distributed in the barley cells. Most of the PP2C genes are located in the chloroplast, and some of them are also located in the cytosol and nucleus. Only a few of them are located in other cellular components such as the cytoskeleton, peroxisome, endoplasmic reticulum and vacuolar membrane. Detailed information on the subcellular localization of each gene in the 20 barley pan-genome accessions is shown in Table S3.

3.3. Phylogenetic Analysis

A phylogenetic ML-tree was constructed with all 1635 HvPP2C sequences (Figure S1). In order to obtain a clear view of the evolutionary relationship, two accessions, Morex and B1K-04-12, were chosen to represent cultivar and wild materials, respectively. Then, another ML-tree was constructed between 82 known Arabidopsis thaliana PP2C sequences, 85 identified barley Morex and 78 B1K-04-12 sequences, using MEGA-X software (Figure 1). The HvPP2Cs were divided into 12 subgroups based on the phylogenetic results, which is consistent with previous studies in Arabidopsis thaliana and Oryza sativa [38], paper mulberry (Broussonetia papyrifera L.) [39], woodland strawberry (Fragaria vesca L.) [40] and potato (Solanum tuberosum L.) [41]. As a result, subgroup A possessed the most members (13 in B1K-04-12 and 14 in Morex), followed by subgroup F (11 in B1K-04-12 and 11 in Morex). Additionally, subgroups J and U, which were identified in Arabidopsis thaliana, did not exist for all barley PP2C members in this study. Compared with B1K-04-12, cultivar Morex has five more members in subgroup K than wild accession B1K-04-12. More research is needed to determine whether these variations in subgroup K were driven by domestication or simply individual differentiations. The code names “Morex” and “FT11”, which were used in the barley pan-genome research (25) of accessions Morex and B1K-04-12, remained in the renaming system in this study. Table S4 shows the MEGA-X model test result of this ML-tree. Additionally, Table S5 shows the subgroup classification and renaming information of the Morex and B1K-04-12 PP2C genes.

3.4. Protein Motifs and Conserved Domain Analysis

Two phylogenetic trees were constructed using B1K-01-12 (Figure 2a) and Morex (Figure 3a) PP2C protein sequences, which represent wild and cultivated barley, respectively. The phylogenetic topology results were used to arrange the order for the protein motif, conserved domain and gene structure analysis.
The MEME motif search tool was used to identify common motifs. Ten conserved motifs were identified both in Morex and B1K-04-12. In wild barley B1K-01-12, seven motifs (1, 3, 4, 5, 6, 7 and 9) were present in the majority of the sequences, whereas other motifs (2, 8 and 10) were specific to the remaining sequences (Figure 2b). Meanwhile, five motifs (1, 2, 3, 6 and 7) were distributed in the majority of the sequences in cultivated barley Morex (Figure 3b). The width, sites and E-value of Morex HvPP2C proteins and their conserved motifs are shown in Figure S2.
The conserved domain was predicted using the NCBI CDD/SPARCLE database (Figure 2c and Figure 3c). All HvPP2C sequences have a conserved domain that belongs to the PP2C family. Most of them have a PP2Cc domain, with the exception of one in Morex (Horvu_MOREX_7H01G494000.1) and one in B1K-04-12 (Horvu_FT11_7H01G466900.1). These two genes have a PLN03145 domain belonging to the PP2C super family. Additionally, domains CAP_ED, PKc_like, MSCRAMM_ClfB and PHA03307 also appear in Morex and B1K-04-12 PP2C protein sequences. In particular, there are five unique domains that only appear in cultivated barley: NB-ARC, PLN03200, Arm, LRR and Rx_N. According to the annotation in the NCBI CDD database, the NB-ARC domain participates in signaling transduction, which regulates plant resistance. The PLN03200 domain is related to cellulose synthase-associated protein. Additionally, the Arm domain has 40 Armadillo/β-catenin-like repeats. These resistance-related domains occurred in the cultivar, suggesting that they might be the result of human breeding.

3.5. Gene Structure and Chromosome Location Analysis

The gene structures of B1K-04-12 and Morex PP2C genes were analyzed (Figure 2d and Figure 3d). In B1K-04-12 and Morex, the number of exons ranged from one to fifteen, with an average of five. Five genes in accession B1K-04-12 have just one exon and no intron, whereas ten genes in Morex have only one exon and no intron. As a result, the genes in the same group share a similar number of exons.
The chromosome locations (Figure 4) were extracted from the genome annotation gff file and displayed in the TBtools software. All 79 HvPP2Cs of B1K-04-12 and 85 HvPP2Cs of Morex were distributed on barley chromosome 1H to 7H, except one, HvMorexPP2C14, which is distributed on chromosome chrUn. As for the distribution location of a single chromosome, HvPP2Cs were largely distributed at both ends of chromosome 1H, 2H, 4H, 5H and 7H in both B1K-04-12 and Morex. The distribution of HvPP2C on chromosomes 3H and 6H, however, was relatively uniform.

3.6. Syntenic Relationships and Ka/Ks Analysis

The syntenic relationships of PP2C genes between Arabidopsis thaliana, wild barley B1K-04-12 and cultivated barley are displayed in Figure 5. The results revealed that the ortholog PP2C gene pairs (red and green lines in Figure 5) are mainly located on Arabidopsis thaliana’s chr1 and chr5 and barley’s chr4H, chr5H and chr6H chromosomes. The remaining sequences between HvPP2Cs and AtPP2Cs show little homology. This indicates that these genes might go through relatively large changes after speciation, preventing them from passing the ortholog identification test. As for the syntenic relationship within barley species between Morex and B1K-04-12 (gray lines in Figure 5), a total of 55 gene pairs were identified as syntenic. They are distributed on seven chromosomes of barley. The synonymous (Ks) and nonsynonymous (Ka) substitution rates of these 55 gene pairs were calculated with TBtools. The meaningless value of Ka/Ks was removed; the results are shown in Table 3.
Among these 55 gene pairs, 25 pairs had a meaningful result, including 24 purified selected gene pairs (with the Ka/Ks value << 1) and one positive selected gene pair (with the Ka/Ks value >> 1). This positive selected gene pair is Horvu_FT11_4H01G404600.1 and Horvu_MOREX_4H01G412600.1 and is renamed as HvFT11PP2C70 and HvMorexPP2C74, which belong to subgroup K and are located on barley chromosome 4H. These two genes both contain a PP2Cc domain and a MSCRAMM_ClfB domain.

3.7. Analysis of Cis-Acting Elements

Cis-acting elements were predicted using HvPP2C upstream 2000 bp sequences in Morex (Figure 6a) and B1K-04-12 (Figure 6b). The summary of predicted cis-elements is shown in Table S6. A total of 3987 cis-elements were discovered in B1K-04-12 (1879) and Morex (2106). Among them, light-responsive elements are the most common, with a total of 1443, accounting for 36% of the total, followed by MeJA- (17%), abscisic acid- (12%), anaerobic induction- (6%), drought induction- (4%) and gibberellin- (4%) responsive elements. These six types of cis-elements account for 80% of the total.

3.8. Network Interaction, GO and KEGG Annotation and Enrichment Analysis

Network interaction analysis was performed using B1K-04-12 (Figure 7a) and Morex (Figure 7b) protein sequences, respectively. A total of 14 genes interacted with each other in B1K-04-12 and Morex.
The GO annotation information is shown in Table S7. The enrichment results are shown in Figure 8. The KEGG annotation information is shown in Table S8. The enrichment results are shown in Figure 9. The GO enrichment results suggest that HvPP2Cs are primarily involved in phosphoric molecular activities and interact with numerous chemicals, such as ester hydrolase, protein serine/threonine and phosphoproteins in biological processes. The KEGG enrichment results suggest that HvPP2Cs are involved in metabolism, protein phosphatases and associated proteins, the MAPK signaling pathway, plant hormone signal transduction, ribosome biogenesis signal transduction and environmental information processing.

3.9. RNA-Seq Analysis of HvPP2C Genes

RNA-Seq data were used to study the expression profiles of barley in 13 published bioprojects. The variety Morex was used as the reference genome. The gene expression level of HvPP2C genes was then normalized using the FPKM (fragments per kilobase million) values. In general, 63 HvPP2C genes were expressed in at least one tissue, stage or experimental treatment.
HvPP2Cs are related to developmental processes. A total of 32 HvPP2Cs were expressed at least in one stage during the early stages of barley development (Figure 10, bioproject PRJNA428086). Among them, two genes (HvMorexPP2C41 and HvMorexPP2C67) exhibited specific expression on day 0 of barley growth, eight genes on day 2 and seven genes on day 5. In addition, another RNA-seq result (Figure S3, bioproject PRJNA668924) showed that HvPP2Cs have different expression levels during inflorescence development and may be required for inflorescence indeterminacy and spikelet determinacy.
HvPP2Cs are also involved in the response to multiple biotic and abiotic stresses. The RNA-seq results demonstrate that multiple HvPP22C genes showed different expression patterns from wildtype to NPR1 overexpressed type and NPR1 knockdown type during their NPR1-mediated acquired resistance (AR) triggered by Pseudomonas syringae pv. tomato DC3000 (Figure S4, bioproject PRJNA431836). In Gao’s study (2018) [51], these ARs are potential resources to improve wheat resistance to Puccinia triticina L. HvPP2C genes also show different expression levels in abiotic stresses, such as high ambient temperatures (Figure S5, PRJNA658250), heavy metal stress (Figure S6, PRJNA382490), aluminum, low pH stresses (Figure S7, PRJNA704034) and salt stress (Figure S8, PRJNA578897).

3.10. Expression Pattern of HvPP2Cs in Aluminum and Low pH Stresses

Summarizing the results of all 13 RNA-seq analyses, the most abundant gene expression differences under experimental treatments were in aluminum and low pH, while many HvPP2Cs also showed tissue specific expression. Seven genes were randomly selected for qRT-PCR experiments from different expressed genes in aluminum and low pH experiment treatments or with tissue-specific expression summarized in other experiments.
To verify the effects of HvPP2Cs during barley stress responses, short- (1 day) and long-term (7 days) aluminum and low pH stresses were taken, and qRT-PCR was performed on barley tissue roots and shoots, respectively (Figure 11). In the shoot, the genes HvMorexPP2C50, HvMorexPP2C68 and HvMorexPP2C72 showed e upregulated expression of pH = 4.0, pH = 4.0 with Al3+ ions and pH = 6.0 short-term treatment. Additionally, HvMorexPP2C46, HvMorexPP2C69 and HvMorexPP2C74 showed upregulated expression of pH = 4.0 with Al3+ ions, and pH = 6.0 short-term treatment. In roots, HvMorexPP2C46, HvMorexPP2C69 and HvMorexPP2C72 showed upregulated expression of pH = 4.0, and pH = 4.0 with Al3+ ions, and HvMorexPP2C74 also showed upregulated expression of pH = 4.0, and pH = 6.0 in short-term treatment. As for long-term treatment, HvMorexPP2C46, HvMorexPP2C69 and HvMorexPP2C72 showed upregulated expression at pH = 4.0, and pH = 4.0 with Al3+ ions, and HvMorexPP2C69, HvMorexPP2C72 and HvMorexPP2C74 showed upregulated expression under the pH = 6.0 condition. Among all the treatments, only two genes (HvMorexPP2C59 and HvMorexPP2C69) showed downregulated expression after short-term pH = 6.0 treatment in shoots. Additionally, only one gene (HvMorexPP2C59) showed downregulated expression after a short-term pH = 6.0 treatment and a long-term pH = 4.0 with Al3+ ion treatment. The expression pattern of HvPP2Cs in aluminum and low pH stresses suggested that HvPP2Cs are sensitive to the stimulation of the external environment, thus confirming that these genes are involved in stress responses in barley.

4. Discussion

To better understand the role of HvPP2Cs in barley domestication and adaptive evolution, the differences in 20 barley pan-genome accessions were compared and summarized between wild accession B1K-04-12 and other cultivated varieties, taking Morex as a representative. First, in terms of gene quantity, wild accession B1K-04-12 has the least number of PP2Cs; only 78 being identified, and other cultivated materials have an average of 82. Second, in terms of distribution, most of these extra sequences of cultivated materials are located on the 7H chromosome and distributed in the cytosol, belonging to the K subfamily. Third, in terms of conserved domains and gene functions, NB-ARC, PLN03200, Arm, LRR and Rx N are five distinct domains that are solely found in cultivated barley. According to the NCBI CDD database annotation, the NB-ARC domain is involved in signaling transduction, which regulates plant resistance, and the PLN03200 domain is related to cellulose synthase-associated protein. Fourth, in terms of selective forces acting on the protein, when calculating the Ka/Ks value of homologous HvPP2C gene pairs of wild and cultivated materials, most homologous gene pairs undergo purification selection (Ka/Ks << 1), only a small proportion experience neutral selection, and a few homologous gene pairs experience positive selection. This suggests that PP2C is highly conserved in barley evolution. This phenomenon is extreme in the PP2C homologous gene pairs of Morex and B1K-04-12, in which only one pair of 25 pairs of effective Ka/Ks values is positively selected, and the remaining 24 pairs are negatively selected, with no neutral selection. Among seven genes used for the qRT-PCR, three of them had a meaningful Ka/Ks value. These include two purified selection genes, HvMorexPP2C50 and HvMorexPP2C72, with Ka/Ks values of 0.1718 and 0.3184, respectively, and one positive selection gene, HvMorexPP2C74, with a Ka/K values of 1.3583. To investigate whether this pair of positively selected genes was genuine, the homologues of the gene Horvu_FT11_4H01G404600.1 in the remaining 19 cultivars were identified. For the gene set composed of these 20 genes, pairwise Ka/Ks values were calculated. Out of a total of 190 results, there were 98 valid Ka/Ks results (Table S9). Among them, 84 results were neutral selection or purification selection, and only 14 results were positive selection gene pairs. All of the positive selection gene pairs were the result of comparing the Horvu_FT11_4H01G404600.1 gene in the wild material B1K-04-12 with the homologous gene pairs in other cultivated materials. Therefore, it can be shown that this gene is indeed involved in the domestication process of barley and is subjected to strong selective forces. qRT-PCR confirmed this positive selection gene HvMorexPP2C74, showing the upregulated expression of pH = 6.0 in both shoot and root after short- or long-term treatments. It also showed the upregulated expression of pH = 4.0 with Al3+ ions in shoot after short-term treatment. This indicates that this gene might be involved in stress responses in barley. In addition to the PP2Cc domain, this gene also contains a MSCRAMM_ClfB domain. There is no published literature on the structure and function of this gene. Therefore, future in-depth studies on the differences between wild and cultivated varieties of this gene will be a good entry point for studying barley domestication.

5. Conclusions

In this study, a total of 1635 HvPP2C genes were identified in 20 barley pan-genome accessions. Two accessions were chosen to represent wild (B1K-04-12) and cultivated (Morex) materials, respectively. Subgroup classification, multiple alignment and phylogenetic analysis of known AtPP2Cs and newly identified HvMorexPP2Cs and HvFT11PP2Cs were carried out. Then, the gene and protein features were analyzed. The results indicate that the HvPP2Cs were conserved during evolution. The expression profile shows that seven HvPP2C genes (HvMorexPP2C46, HvMorexPP2C50, HvMorexPP2C59, HvMorexPP2C68, HvMorexPP2C69, HvMorexPP2C72 and HvMorexPP2C74) are involved in the response to aluminum and low pH stresses. Fourteen positively selected homologous gene pairs were identified between wild accession B1K-04-12 and 14 other cultivars, indicating that these genes are important members of barley domestication. Our findings provide new insights into the functional evolution of HvPP2Cs during barley domestication.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/genes13050834/s1. Table S1. RNA-seq data used in this study. Table S2. Primer used for quantitative real-time PCR. Table S3. Information of PP2C gene family in barley pan-genome accessions. Table S4. Maximum likelihood fits of 56 different amino acid substitution models. Table S5. HvPP2C subgroup classification and rename information. Table S6. The summary of predicted cis-elements. Table S7. The GO annotation information. Table S8. The KEGG annotation information. Table S9. Pairwise Ka/Ks of 20 homologue genes. Figure S1. A phylogenetic ML-tree using all 1656 HvPP2C sequences. Figure S2. The width, sites and E-value and sequence logo of conserved motif of Morex HvPP2C proteins. Figure S3. Expression profiles of barley PP2Cs during inflorescence development. Log2-based fold change was used to create the heatmap. Fold changes in gene expression are shown by colors on the scale. Figure S4. Expression profiles of barley PP2Cs during their NPR1-mediated acquired resistance (AR) triggered by Pseudomonas syringae pv. tomato DC3000. Figure S5. Expression profiles of barley PP2Cs in response to high ambient temperatures. Figure S6. Expression profiles of barley PP2Cs in response to heavy metal stress. Figure S7. Expression profiles of barley PP2Cs in response to aluminum and low pH stresses. Figure S8. Expression profiles of barley PP2Cs in response to salt stress.

Author Contributions

W.S. conceived and designed the experiments; X.-T.W. collected and analyzed the data and wrote the manuscript; Z.-P.X. and F.-L.H. checked the results of the data analysis and revised the article; K.-X.C. conducted the experiments; G.-R.Z. collected the seedlings for the experiments. K.-R.F., X.-H.L., X.-R.L. and Z.T. collected the plant materials. M.-X.W. reviewed the article. All authors have read and agreed to the published version of the manuscript.

Funding

The Fundamental Research Funds for the Central Universities of China, (grant/award number: 2452018101); the Open Funding of State Key Laboratory of Crop Stress Biology for Arid Areas, (grant/award number: CSBAA2017003).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fan, K.; Yuan, S.; Chen, J.; Chen, Y.; Lin, W. Molecular evolution and lineage-specific expansion of the PP2C family in Zea mays. Planta 2019, 250, 1521–1538. [Google Scholar] [CrossRef] [PubMed]
  2. Brock, A.K.; Willmann, R.; Kolb, D.; Grefen, L.; Layunen, H.M.; Bethke, G.; Lee, J.; Nürnberger, T.; Gust, A.A. The Arabidopsis Mitogen-Activated Protein Kinase Phosphatase PP2C5 Affects Seed Germination, Stomatal Aperture, and Abscisic Acid-Inducible Gene Expression. Plant Physiol. 2010, 153, 1098–1111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Umbrasaite, J.; Schweighofer, A.; Kazanaviciute, V.; Magyar, Z.; Ayatollahi, Z.; Unterwurzacher, V.; Choopayak, C.; Boniecka, J.; Murray, J.; Bogre, L. MAPK Phosphatase AP2C3 Induces Ectopic Proliferation of Epidermal Cells Leading to Stomata Development in Arabidopsis. PLoS ONE 2010, 5, e15357. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Komatsu, K.; Suzuki, N.; Kuwamura, M.; Nishikawa, Y.; Sakata, Y. Group A PP2Cs evolved in land plants as key regulators of intrinsic desiccation tolerance. Nat. Commun. 2011, 4, 375–381. [Google Scholar] [CrossRef] [Green Version]
  5. Tähtiharju, S.; Palva, T. Antisense inhibition of protein phosphatase 2C accelerates cold acclimation in Arabidopsis thaliana. Plant J. 2002, 26, 461–470. [Google Scholar] [CrossRef]
  6. The International Barley Genome Sequencing Consortium. A physical, genetic and functional sequence assembly of the barley genome. Nature 2012, 491, 711–716. [Google Scholar] [CrossRef]
  7. Bian, J.; Deng, P.; Zhan, H.; Wu, X.; Nishantha, M.; Yan, Z.; Du, X.; Nie, X.; Song, W. Transcriptional Dynamics of Grain Development in Barley (Hordeum vulgare L.). Int. J. Mol. Sci. 2019, 20, 962. [Google Scholar] [CrossRef] [Green Version]
  8. Badr, A.; Rabey, H.E.; Effgen, S.; Ibrahim, H.H.; Pozzi, C.; Rohde, W.; Salamini, F. On the Origin and Domestication History of Barley (Hordeum vulgare). Mol. Biol. Evol. 2000, 17, 499–510. [Google Scholar] [CrossRef] [Green Version]
  9. Saisho, D.; Purugganan, M.D. Molecular Phylogeography of Domesticated Barley Traces Expansion of Agriculture in the Old World. Genetics 2007, 177, 1765. [Google Scholar] [CrossRef] [Green Version]
  10. Dai, F.; Nevo, E. Tibet is one of the centers of domestication of cultivated barley. Proc. Natl. Acad. Sci. USA 2012, 109, 16969–16973. [Google Scholar] [CrossRef] [Green Version]
  11. Dai, F.; Chen, Z.-H.; Wang, X.; Li, Z.; Jin, G. Transcriptome profiling reveals mosaic genomic origins of modern cultivated barley. Proc. Natl. Acad. Sci. USA 2014, 111, 13403–13408. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Schiemann, E. Genetische Studien an Gerste. Z. Für Indukt. Abstamm. Vererb. 1921, 26, 109–143. [Google Scholar]
  13. Takahashi, R.; Hayashi, J. Linkage study of two complementary genes for brittle rachis in barley. Ohara Inst. Für Landwirtsch. Biol. 1964, 12, 99–105. [Google Scholar]
  14. Komatsuda, T.; Maxim, P.; Senthil, N.; Mano, Y. High-density AFLP map of nonbrittle rachis 1 (btr1) and 2 (btr2) genes in barley (Hordeum vulgare L.). Theor. Appl. Genet. 2004, 109, 986–995. [Google Scholar] [CrossRef]
  15. Senthil, N.; Komatsuda, T. Inter-subspecific maps of non-brittle rachis genes btr1/btr2 using occidental, oriental and wild barley lines. Euphytica 2005, 145, 215–220. [Google Scholar] [CrossRef]
  16. Ubisch, G.V. Analyse eines Falles von Bastardatavismus und Faktoren-koppelung bei Gerste. Z. Für Indukt. Abstamm. Vererb. 1915, 14, 226–237. [Google Scholar] [CrossRef] [Green Version]
  17. Laurie, D.; Pratchett, N.; Bezant, J.H.; Snape, J.W. Genetic analysis of a photoperiod response gene on the short arm of chromosome 2(2H) of Hordeum vulgare (barley). Heredity 1994, 72, 619–627. [Google Scholar] [CrossRef]
  18. Laurie, D.A.; Pratchett, N.; Snape, J.W.; Bezant, J.H. RFLP mapping of five major genes and eight quantitative trait loci controlling flowering time in a winter x spring barley (Hordeum vulgare L.) cross. Genome 1995, 38, 575. [Google Scholar] [CrossRef]
  19. Jones, H.; Civáň, P.; Cockram, J.; Leigh, F.J.; Brown, T.A. Evolutionary history of barley cultivation in Europe revealed by genetic analysis of extant landraces. BMC Evol. Biol. 2011, 11, 320. [Google Scholar] [CrossRef] [Green Version]
  20. Cockram, J.; Hones, H.; O’Sullivan, D.M. Genetic variation at flowering time loci in wild and cultivated barley. Plant Genet. Resour. 2011, 9, 264–267. [Google Scholar] [CrossRef] [Green Version]
  21. Plant pan-genomes are the new reference. Nat. Plants 2020, 6, 914–920. [CrossRef] [PubMed]
  22. Tettelin, H.; Masignani, V.; Cieslewicz, M.; Donati, C.; Medini, D.; Ward, N.; Angiuoli, S.; Crabtree, J.; Jones, A.; Durkin, A. Genome analysis of multiple pathogenic isolates of Streptococcus agalactiae: Implications for the microbial “pan-genome”. Proc. Natl. Acad. Sci. USA 2005, 102, 13950–13955. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Qin, P.; Lu, H.; Du, H.; Wang, H.; Li, S. Pan-genome analysis of 33 genetically diverse rice accessions reveals hidden genomic variations. Cell 2021, 184, 3542–3558.e16. [Google Scholar] [CrossRef] [PubMed]
  24. Liu, Y.; Du, H.; Li, P.; Shen, Y.; Tian, Z. Pan-Genome of Wild and Cultivated Soybeans. Cell 2020, 182, 162–176.e13. [Google Scholar] [CrossRef]
  25. Jayakodi, M.; Padmarasu, S.; Haberer, G.; Bonthala, V.S.; Stein, N. The barley pan-genome reveals the hidden legacy of mutation breeding. Nature 2020, 588, 284–289. [Google Scholar] [CrossRef]
  26. Wang, Y.; Liu, Z.; Cheng, H.; Gao, T.; Pan, Z.; Yang, Q.; Gou, A.; Xue, Y. EKPD: A hierarchical database of eukaryotic protein kinases and protein phosphatases. Nucleic Acids Res. 2014, 42, D496–D502. [Google Scholar] [CrossRef] [Green Version]
  27. Eddy, S.R. Accelerated Profile HMM Searches. PLoS Comput. Biol. 2011, 7, e1002195. [Google Scholar] [CrossRef] [Green Version]
  28. Jaina, M.; Sara, C.; Lowri, W.; Matloob, Q.; Gustavoa, S.; Sonnhammer, E.; Tosatto, S.; Lisanna, P.; Shriya, R.; Richardson, L.J. Pfam: The protein families database in 2021. Nucleic Acids Res. 2020, 49, D412–D419. [Google Scholar]
  29. Das, A.K.; Helps, N.R.; Cohen, P.T.; Barford, D. Crystal structure of the protein serine/threonine phosphatase 2C at 2.0 A resolution. EMBO J. 1996, 15, 6798–6809. [Google Scholar] [CrossRef]
  30. Gertz, E.M. BLAST Scoring Parameters. Manuscript 2005, 1, 1–54. [Google Scholar]
  31. Stajich, J.E. E. The Bioperl Toolkit: Perl Modules for the Life Sciences. Genome Res. 2002, 12, 1611–1618. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Lu, S.; Wang, J.; Chitsaz, F.; Derbyshire, M.K.; Geer, R.C.; Gonzales, N.R.; Gwadz, M.; Hurwitz, D.I.; Marchler, G.H.; Song, J.S.; et al. CDD/SPARCLE: The conserved domain database in 2020. Nucleic Acids Res. 2020, 48, D265–D268. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Duvaud, S.; Gabella, C.; Lisacek, F.; Stockinger, H.; Durinx, C. Expasy, the Swiss Bioinformatics Resource Portal, as designed by its users. Nucleic Acids Res. 2021, 49, W216–W227. [Google Scholar] [CrossRef] [PubMed]
  34. Horton, P.; Park, K.J.; Obayashi, T.; Nakai, K. WoLF PSORT: Protein Localization Prediction Software. Nucleic Acids Res. 2006, 35, 585–587. [Google Scholar] [CrossRef] [Green Version]
  35. Sudhir, K.; Glen, S.; Michael, L.; Christina, K.; Koichiro, T. MEGA X: Molecular Evolutionary Genetics Analysis across computing platforms. Mol. Biol. Evol. 2018, 35, 1547. [Google Scholar]
  36. Nei, M.; Kumar, S. Molecular Evolution and Phylogenetics; Oxford University Press: Oxford, UK, 2000. [Google Scholar]
  37. Sudhir, K.; Glen, S.; Daniel, P.; Koichiro, T.; Tamurasup/sup, K. MEGA-CC: Computing core of molecular evolutionary genetics analysis program for automated and iterative data analysis. Bioinformatics 2012, 28, 2685. [Google Scholar]
  38. Xue, T.; Wang, D.; Zhang, S.; Ehlting, J.; Ni, F.; Jakab, S.; Zheng, C.; Yuan, Z. Genome-wide and expression analysis of protein phosphatase 2C in rice and Arabidopsis. BMC Genomics 2008, 9, 550. [Google Scholar] [CrossRef] [Green Version]
  39. Zhang, B.; Chen, N.; Peng, X.; Shen, S. Identification of the PP2C gene family in paper mulberry (Broussonetia papyrifera) and its roles in the regulation mechanism of the response to cold stress. Biotechnol. Lett. 2021, 43, 1089–1102. [Google Scholar] [CrossRef]
  40. Haider, M.S.; Khan, N.; Pervaiz, T.; Zhogjie, L.; Nasim, M.; Jogaiah, S.; Mushtaq, N.; Jiu, S.; Jinggui, F. Genome-wide identification, evolution, and molecular characterization of the PP2C gene family in woodland strawberry. Gene 2019, 702, 27–35. [Google Scholar] [CrossRef]
  41. Wang, Y.F.; Liao, Y.Q.; Wang, Y.P.; Yang, J.W.; Huai-Jun, S.I. Genome-wide identification and expression analysis of StPP2C gene family in response to multiple stresses in potato (Solanum tuberosum L.). J. Integr. Agric. 2020, 19, 1609–1624. [Google Scholar] [CrossRef]
  42. Bailey, T.L.; Johnson, J.; Grant, C.; Noble, W.S. The MEME Suite. Nucleic Acids Res. 2015, 43, W39–W49. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Chen, C.; Chen, H.; Zhang, Y.; Thomas, H.R.; Frank, M.H.; He, Y.; Xia, R. TBtools: An Integrative Toolkit Developed for Interactive Analyses of Big Biological Data. Mol. Plant 2020, 13, 1194–1202. [Google Scholar] [CrossRef] [PubMed]
  44. Damian, S.; Gable, A.L.; Nastou, K.C.; David, L.; Rebecca, K.; Sampo, P.; Doncheva, N.T.; Marc, L.; Fang, T.; Peer, B. The STRING database in 2021: Customizable protein–protein networks, and functional characterization of user-uploaded gene/measurement sets. Nucleic Acids Res. 2020, 49, D605–D612. [Google Scholar]
  45. Huerta-Cepas, J.; Szklarczyk, D.; Heller, D.; Hernández-Plaza, A.; Forslund, S.K.; Cook, H.; Mende, D.R.; Letunic, I.; Rattei, T.; Jensen, L.J.; et al. eggNOG 5.0: A hierarchical, functionally and phylogenetically annotated orthology resource based on 5090 organisms and 2502 viruses. Nucleic Acids Res. 2019, 47, D309–D314. [Google Scholar] [CrossRef] [Green Version]
  46. Cantalapiedra, C.P.; Ana, H.P.; Ivica, L.; Peer, B.; Jaime, H.C. eggNOG-mapper v2: Functional Annotation, Orthology Assignments, and Domain Prediction at the Metagenomic Scale. Mol. Biol. Evol. 2021, 38, 5825–5829. [Google Scholar] [CrossRef]
  47. Zhong, J.; Esse, G.; Bi, X.; Lan, T.; Korff, M.V. INTERMEDIUM-M encodes an HvAP2L-H5 ortholog and is required for inflorescence indeterminacy and spikelet determinacy in barley. Proc. Natl. Acad. Sci. USA 2021, 118, e2011779118. [Google Scholar] [CrossRef]
  48. Li, G.; Kuijer, H.; Yang, X.; Liu, H.; Zhang, D. MADS1 maintains barley spike morphology at high ambient temperatures. Nature Plants 2021, 7, 1093–1107. [Google Scholar] [CrossRef]
  49. Lai, Y.; Zhang, D.; Wang, J.; Wang, J.; Wang, H. Integrative Transcriptomic and Proteomic Analyses of Molecular Mechanism Responding to Salt Stress during Seed Germination in Hulless Barley. Int. J. Mol. Sci. 2020, 21, 359. [Google Scholar] [CrossRef] [Green Version]
  50. Gao, S.; Zheng, Z.; Powell, J.; Habib, A.; Liu, C. Validation and delineation of a locus conferring Fusarium crown rot resistance on 1HL in barley by analysing transcriptomes from multiple pairs of near isogenic lines. BMC Genomics 2019, 20, 650. [Google Scholar] [CrossRef] [Green Version]
  51. Gao, J.; Bi, W.; Li, H.; Wu, J.; Yu, X.; Liu, D.; Wang, X. WRKY Transcription Factors Associated With NPR1-Mediated Acquired Resistance in Barley Are Potential Resources to Improve Wheat Resistance to Puccinia triticina. Front. Plant Sci. 2018, 9, 1486. [Google Scholar] [CrossRef]
  52. Bélanger, S.; Marchand, S.; Jacques, P.-é.; Meyers, B.; Belzile, F. Differential Expression Profiling of Microspores During the Early Stages of Isolated Microspore Culture Using the Responsive Barley Cultivar Gobernadora. G3 Genes Genomes Genetics 2018, 8, 1603–1614. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Kim, D.; Paggi, J.M.; Park, C.; Bennett, C.; Salzberg, S.L. Graph-based genome alignment and genotyping with HISAT2 and HISAT-genotype. Nat. Biotechnol. 2019, 37, 905–917. [Google Scholar] [CrossRef] [PubMed]
  54. Pertea, M.; Pertea, G.M.; Antonescu, C.M.; Chang, T.C.; Mendell, J.T.; Salzberg, S.L. StringTie enables improved reconstruction of a transcriptome from RNA-seq reads. Nat. Biotechnol. 2015, 33, 290–295. [Google Scholar] [CrossRef] [Green Version]
  55. Pertea, M.; Kim, D.; Pertea, G.M.; Leek, J.T.; Salzberg, S.L. Transcript-level expression analysis of RNA-seq experiments with HISAT, StringTie and Ballgown. Nat. Protoc. 2016, 11, 1650–1667. [Google Scholar] [CrossRef] [PubMed]
  56. IBM Corp. IBM SPSS Statistics for Windows, version 24.0; IBM Corp.: Armonk, NY, USA, 2016.
  57. R Core Team. R: A language and environment for statistical computing. Computing 2011, 1, 12–21. [Google Scholar]
Figure 1. An ML-tree was constructed between 82 known Arabidopsis thaliana PP2C sequences (marked with red stars), 85 new identified barley Morex (marked with blue circles) and 78 B1K-04-12 sequences (marked with yellow triangles) using MEGA-X software. The background colors represent HvPP2C subgroups, which are labelled in the out layer.
Figure 1. An ML-tree was constructed between 82 known Arabidopsis thaliana PP2C sequences (marked with red stars), 85 new identified barley Morex (marked with blue circles) and 78 B1K-04-12 sequences (marked with yellow triangles) using MEGA-X software. The background colors represent HvPP2C subgroups, which are labelled in the out layer.
Genes 13 00834 g001
Figure 2. Phylogenetic, motif, conserved domain and gene structure analysis of HvPP2Cs in wild barley B1K-04-12. (a) Phylogenetic tree, (b) conserved motif distribution, (c) conserved domains and (d) gene structure.
Figure 2. Phylogenetic, motif, conserved domain and gene structure analysis of HvPP2Cs in wild barley B1K-04-12. (a) Phylogenetic tree, (b) conserved motif distribution, (c) conserved domains and (d) gene structure.
Genes 13 00834 g002
Figure 3. Phylogenetic, motif, conserved domain and gene structure analysis of HvPP2Cs in cultivated barley Morex. (a) Phylogenetic tree, (b) conserved motif distribution, (c) conserved domains and (d) gene structure.
Figure 3. Phylogenetic, motif, conserved domain and gene structure analysis of HvPP2Cs in cultivated barley Morex. (a) Phylogenetic tree, (b) conserved motif distribution, (c) conserved domains and (d) gene structure.
Genes 13 00834 g003
Figure 4. Chromosome distribution of HvPP2Cs in (a) B1K-04-12 and (b) Morex.
Figure 4. Chromosome distribution of HvPP2Cs in (a) B1K-04-12 and (b) Morex.
Genes 13 00834 g004
Figure 5. Synteny analysis of PP2C genes between 82 known Arabidopsis thaliana PP2C sequences (located on blue chromosomes), 85 newly identified barley Morex (located on red chromosomes) and 78 B1K-04-12 sequences (located on green chromosomes). The orthologous gene pairs between Arabidopsis and Morex are marked with red lines. The orthologous gene pairs between Arabidopsis and B1K-04-12 are marked with green lines. Additionally, the orthologous gene pairs between Morex and B1K-04-12 are marked with gray lines.
Figure 5. Synteny analysis of PP2C genes between 82 known Arabidopsis thaliana PP2C sequences (located on blue chromosomes), 85 newly identified barley Morex (located on red chromosomes) and 78 B1K-04-12 sequences (located on green chromosomes). The orthologous gene pairs between Arabidopsis and Morex are marked with red lines. The orthologous gene pairs between Arabidopsis and B1K-04-12 are marked with green lines. Additionally, the orthologous gene pairs between Morex and B1K-04-12 are marked with gray lines.
Genes 13 00834 g005
Figure 6. Cis-element analysis of HvPP2Cs in (a) Morex and (b) B1K-04-12.
Figure 6. Cis-element analysis of HvPP2Cs in (a) Morex and (b) B1K-04-12.
Genes 13 00834 g006
Figure 7. Network interaction analysis in B1K-04-12 (a) and Morex (b).
Figure 7. Network interaction analysis in B1K-04-12 (a) and Morex (b).
Genes 13 00834 g007
Figure 8. GO enrichment in B1K-04-12 (a) and Morex (b).
Figure 8. GO enrichment in B1K-04-12 (a) and Morex (b).
Genes 13 00834 g008
Figure 9. KEGG enrichment in B1K-04-12 (a) and Morex (b).
Figure 9. KEGG enrichment in B1K-04-12 (a) and Morex (b).
Genes 13 00834 g009
Figure 10. Expression profile of barley PP2C genes during the early stages of barley development.
Figure 10. Expression profile of barley PP2C genes during the early stages of barley development.
Genes 13 00834 g010
Figure 11. Relative expression levels of seven candidate HvMorexPP2C genes after aluminum and low pH stresses in shoot and root, respectively. The values are given as the means ± SDs of three biological replicates. Additionally, the letters show the multiple comparison results of Waller–Duncan method calculated by SPSS.
Figure 11. Relative expression levels of seven candidate HvMorexPP2C genes after aluminum and low pH stresses in shoot and root, respectively. The values are given as the means ± SDs of three biological replicates. Additionally, the letters show the multiple comparison results of Waller–Duncan method calculated by SPSS.
Genes 13 00834 g011
Table 1. The number of HvPP2C genes in each accession of barley pan-genome.
Table 1. The number of HvPP2C genes in each accession of barley pan-genome.
AccessionsNumber of PP2Cs
Hockett88
Morex85
HOR_308184
HOR_904384
HOR_755283
OUN33383
HOR_1035082
HOR_2159982
ZDM0206482
HOR_1382181
HOR_1394281
HOR_814881
RGT_Planet81
ZDM0146781
Akashinriki80
Barke80
Golden_Promise80
HOR_336580
Igri79
B1K-04-1278
Table 2. Summary of subcellular localization of HvPP2Cs in B1K-04-12 and Morex.
Table 2. Summary of subcellular localization of HvPP2Cs in B1K-04-12 and Morex.
Subcellular LocationB1K-04-12Morex
Chloroplast3538
Cytosol1920
Nuclear1618
Cytoskeleton33
Peroxisome22
Chloroplast_Mitochondrion11
Endoplasmic reticulum11
Vacuolar membrane12
Table 3. Ka/Ks of syntenic gene pairs between B1K-04-12 and Morex.
Table 3. Ka/Ks of syntenic gene pairs between B1K-04-12 and Morex.
Seq_1Seq_2KaKsKa_Ks
Horvu_FT11_4H01G404600.1Horvu_MOREX_4H01G412600.10.00340.00251.3583
Horvu_FT11_2H01G014500.1Horvu_MOREX_2H01G002500.10.00320.00600.5319
Horvu_FT11_2H01G454900.1Horvu_MOREX_2H01G471900.10.00300.00610.4947
Horvu_FT11_2H01G088000.1Horvu_MOREX_2H01G089800.10.00230.00620.3771
Horvu_FT11_6H01G125300.1Horvu_MOREX_6H01G131700.10.00120.00330.3505
Horvu_FT11_4H01G099700.1Horvu_MOREX_4H01G099600.10.00330.00950.3460
Horvu_FT11_5H01G121400.1Horvu_MOREX_5H01G126800.10.00120.00360.3204
Horvu_FT11_3H01G115300.1Horvu_MOREX_3H01G111900.10.00550.01720.3184
Horvu_FT11_7H01G128600.1Horvu_MOREX_7H01G158800.10.00120.00400.3146
Horvu_FT11_3H01G283600.1Horvu_MOREX_3H01G284300.10.00070.00210.3105
Horvu_FT11_6H01G259200.1Horvu_MOREX_6H01G264100.10.00040.00130.3073
Horvu_FT11_4H01G287000.1Horvu_MOREX_4H01G292800.10.00110.00360.2993
Horvu_FT11_7H01G042900.1Horvu_MOREX_7H01G041800.10.00710.03240.2184
Horvu_FT11_7H01G478100.1Horvu_MOREX_7H01G506400.10.00140.00800.1773
Horvu_FT11_7H01G430700.1Horvu_MOREX_7H01G456500.10.00130.00760.1718
Horvu_FT11_5H01G113600.1Horvu_MOREX_5H01G118100.10.00120.00710.1639
Horvu_FT11_2H01G621300.1Horvu_MOREX_2H01G644100.10.00110.00700.1634
Horvu_FT11_5H01G380300.1Horvu_MOREX_5H01G396300.10.00100.00630.1626
Horvu_FT11_3H01G711500.1Horvu_MOREX_3H01G723700.10.00110.00710.1616
Horvu_FT11_1H01G521300.1Horvu_MOREX_1H01G534000.10.00090.00780.1154
Horvu_FT11_4H01G533500.1Horvu_MOREX_4H01G549600.10.00450.04030.1110
Horvu_FT11_1H01G107800.1Horvu_MOREX_1H01G108200.10.00120.01070.1090
Horvu_FT11_1H01G039300.1Horvu_MOREX_1H01G036200.10.00310.03870.0800
Horvu_FT11_5H01G653000.1Horvu_MOREX_5H01G674700.10.00070.01030.0683
Horvu_FT11_5H01G628500.1Horvu_MOREX_5H01G651500.10.00190.06270.0302
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wu, X.-T.; Xiong, Z.-P.; Chen, K.-X.; Zhao, G.-R.; Feng, K.-R.; Li, X.-H.; Li, X.-R.; Tian, Z.; Huo, F.-L.; Wang, M.-X.; et al. Genome-Wide Identification and Transcriptional Expression Profiles of PP2C in the Barley (Hordeum vulgare L.) Pan-Genome. Genes 2022, 13, 834. https://doi.org/10.3390/genes13050834

AMA Style

Wu X-T, Xiong Z-P, Chen K-X, Zhao G-R, Feng K-R, Li X-H, Li X-R, Tian Z, Huo F-L, Wang M-X, et al. Genome-Wide Identification and Transcriptional Expression Profiles of PP2C in the Barley (Hordeum vulgare L.) Pan-Genome. Genes. 2022; 13(5):834. https://doi.org/10.3390/genes13050834

Chicago/Turabian Style

Wu, Xiao-Tong, Zhu-Pei Xiong, Kun-Xiang Chen, Guo-Rong Zhao, Ke-Ru Feng, Xiu-Hua Li, Xi-Ran Li, Zhao Tian, Fu-Lin Huo, Meng-Xing Wang, and et al. 2022. "Genome-Wide Identification and Transcriptional Expression Profiles of PP2C in the Barley (Hordeum vulgare L.) Pan-Genome" Genes 13, no. 5: 834. https://doi.org/10.3390/genes13050834

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