Next Article in Journal
Effect of Meiotic Polyploidisation on Selected Morphological and Anatomical Traits in Interspecific Hybrids of Brassica oleracea × B. napus
Next Article in Special Issue
Grapevine Gene Systems for Resistance to Gray Mold Botrytis cinerea and Powdery Mildew Erysiphe necator
Previous Article in Journal
Biochemical Characterisation and Genetic Structure Provide Insight into the Diversity of the Mediterranean Tomato Ancient Varieties ‘San Marzano’ and ‘Re Fiascone’: New Resources for Breeding
Previous Article in Special Issue
Modeling of Flowering Time in Vigna radiata with Approximate Bayesian Computation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genome-Wide In Silico Analysis and Expression Profiling of Phosphoenolpyruvate Carboxylase Genes in Loquat, Apple, Peach, Strawberry and Pear

1
School of Food and Bioengineering, Fujian Polytechnic Normal University, Fuqing 350300, China
2
Fujian Universities and Colleges Engineering Research Center of Modern Facility Agriculture, Fuqing 350300, China
3
College of Horticulture, Fujian Agriculture and Forestry University, Fuzhou 350002, China
4
Key Laboratory of Horticultural Plant Biology (Ministry of Education), College of Horticulture & Forestry Sciences, Huazhong Agricultural University, Wuhan 430070, China
5
Department of Horticulture, Faculty of Agricultural Sciences and Technology, Bahauddin Zakariya University, Multan 66000, Pakistan
6
Department of Horticulture, College of Agriculture, University of Al-Azhar (Branch Assiut), Assiut 71524, Egypt
7
Plant Production Department (Horticulture—Medicinal and Aromatic Plants), Faculty of Agriculture (Saba Basha), Alexandria University, Alexandria 21531, Egypt
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Agronomy 2022, 12(1), 25; https://doi.org/10.3390/agronomy12010025
Submission received: 24 November 2021 / Revised: 12 December 2021 / Accepted: 20 December 2021 / Published: 23 December 2021
(This article belongs to the Special Issue Improvement of Crops: Current Status and Future Prospects)

Abstract

:
Phosphoenolpyruvate carboxylase (PEPC) genes have multiple potential roles in plant metabolism such as regulation and accumulation of organic acids in fruits, movement of guard cells and stress tolerance, etc. However, the systematic identification and characterization of PEPC genes in Rosaceae species i.e., loquat, apple, peach, strawberry, and pear are yet to be performed. In present study, 27 putative PEPC genes (loquat 4, apple 6, peach 3, strawberry 9, and pear 5) were identified. To further investigate the role of those PEPC genes, comprehensive bioinformatics and expression analysis were performed. In bioinformatic analysis, the physiochemical properties, conserved domains, gene structure, conserved motif, phylogenetic and syntenic analysis of PEPC genes were performed. The result revealed that the PEPcase superfamily domain was conserved in all examined PEPC proteins. Most of the PEPC proteins were predicted to be localized in cytonuclear. Genomic structural and motif analysis showed that the exon and motif number of each PEPC gene ranged dramatically, from 8 to 20, and 7 to 10, respectively. Syntenic analysis indicated that the segmental or whole-genome duplication played a vital role in extension of PEPC gene family in Rosacea species. The Ka and Ks values of duplicated genes depicted that PEPC genes have undergone a strong purifying selection. Furthermore, the expression analysis of PEPC genes in root, mature leaf, stem, full-bloom flower, and ripened fruit of loquat, apple, peach, strawberry, and pear was performed. Some genes were differentially expressed in aforementioned plant tissues, signifying their role in plant metabolism. This study provides the first genome-wide identification, characterization, and expression profiling of PEPC gene family in Rosaceae species, and provides the foundation for further functional analysis.

1. Introduction

Phosphoenolpyruvate (PEP) carboxylase (PEPC) enzyme can be widely found among bacteria, archaea, green algae, cyanobacteria, protozoa, and vascular plants, while absent in animals and fungi cells [1,2]. PEPC is used in the primary fixation reaction for photosynthetic CO2 assimilation in C4 photosynthesis as well as crassulacean acid metabolism (CAM); which take place in mesophyll cells in the presence of bicarbonate (HCO3) and Mg2+ to catalyze an irreversible β-carboxylation reaction of PEP to produce oxaloacetate (OAA) and inorganic phosphate (Pi) [1]. Besides its central role in atmospheric CO2 fixation during photosynthesis, PEPC is also known to have a wide range of non-photosynthetic activities, such as carbon-nitrogen interaction support and fruit ripening [2], seed germination and formation [2,3], and controlling guard cell metabolism for better stomatal functioning [4]. In addition among non-photosynthetic tissues and in leaves of C3 plants, PEPC is important for anaplerotic to replenishing tricarboxylic acid (TCA) cycle with intermediates which are consumed in different biosynthetic pathways and N-assimilation [1,2].
PEPCs are important, as they transport organic acids across membranes [5]. With the help of such transporter proteins, organic acids play a critical role in plants’ primary metabolism and help plants in adaptation according to changing environments, such as stomatal movement, stress responses, and pH regulation [6,7,8,9]. During the past two decades, several experiments have concluded that organic acids specifically malate were released from plant roots to counter the metal toxicity and maintain fruit pH [10,11]. Since the whole-genome sequences of many plant species have been released, various PEPC proteins have been effectively recognized and examined in plants including Arabidopsis thaliana [12], Hordeum vulgare [13], Lotus japonicus [14], Solanum lycopersicum [15], Solanum tuberosum [16], and Triticum aestivum [17]. For instance, three PEPC genes (PPC1- PPC3) were characterized in A. thaliana [12]. Three dicot C4 PEPC (ppc-A, ppc-B, ppc-C) genes were analyzed in Flaveria (Asteraceae), with the ppc-A gene being identified as the gene recruited for use in the C4 photosynthetic pathway [18]. The metabolic functions of PEPC gene family have rarely been reported in the Rosaceae.
Recently, loquat’s (Eriobotrya japonica Lindl.) genome was sequenced using 3rd generation sequencing technology via Nanopore and Hi-C technology [19]. The genomes of Rosaceae species including apple (Malus domestica Borkh.), peach (Prunus persica), strawberry (Fragaria x ananassa Duch.) and pear (Pyrus communis L.) were also accessible. By utilizing the information available, an opportunity to analyze the PEPC gene family in Rosaceae species was undertaken as publishing the first report on PEPC gene family members in loquat, apple, peach, strawberry and pear. In this study, we identified 27 PEPC genes in the genomes of aforementioned five Rosaceae species; phylogenetic relationships, gene duplication and subcellular localization were investigated, and expression of related genes in roots, stem, leaves, flowers and fruits of aforementioned Rosaceae species was observed. Our findings explore molecular features and evolutionary pattern of PEPC gene family and provide groundwork to functionally characterize PEPC genes in Rosaceae species.

2. Materials and Methods

2.1. Identification and Characterization of PEPC Genes

The loquat (Eriobotrya japonica) genome sequence was downloaded from the GigaScience Database (http://gigadb.org/dataset/view/id/100711, accessed on 22 December 2020) [19]. The apple, peach, and strawberry genome sequences [20] were downloaded from Phytozome (http://phytozome.jgi.doe.gov/pz/portal.html, accessed on 16 June 2021), and the genome sequences of pear [21] was downloaded from the Genome Database for Rosaceae (GDR) (http://www.rosaceae.org/, accessed on 16 June 2021). While, peptide sequences of Arabidopsis’s [22] PEPC genes were retrieved from TAIR (https://www.arabidopsis.org/, accessed on 22 December 2020), and Arabidopsis’s sequence was used as a query to perform BLAST against the aforementioned species. Furthermore, Pfam online database and HMMER3 software package were used for PEPC domain (PF00311) search and HMM file construction, respectively [23,24]. Using local protein databases, HMM searches were performed of the aforementioned species with HMMER3. Moreover, we checked the physical locations of all candidate PEPC genes and rejected redundant sequences with the same chromosome location. For verification of the presence of PEPC domains, all obtained PEPC protein sequences were analyzed again in the Pfam database via SMART program (http://smart.embl-heidelberg.de/, accessed on 8 January 2021), while proteins lacking PEPC domain were removed.
The physicochemical properties of PEPC proteins were calculated using ExPASy Proteomics Server (http://web.expasy.org/compute_pi/, accessed on 9 January 2021). The WoLF PSORT web server (https://wolfpsort.hgc.jp/, accessed on 9 January 2021) and CELLO version 2.5, subcellular localization predictor, (http://cello.life.nctu.edu.tw/, accessed on 9 January 2020) were used to predict subcellular localization of PEPC proteins. The 3D-structure models of PEPC proteins were predicted through the online tool i-Tasser (https://zhanggroup.org/I-TASSER/, accessed on 26 June 2021).

2.2. Phylogenetic Analyses

Evolutionary and phylogenetic analyses were performed using Molecular Evolutionary Genetics Analysis X (MEGA-X v10.2.6) [25]. Protein sequence alignment was performed using MUSCLE with default parameters; while PEPCs’ phylogenetic tree was constructed by neighbor-joining (NJ) bootstrap = 1000x method.

2.3. Gene Structure, Conserved Motif and Promoter Region Analysis of PEPC Genes

By aligning coding sequences with the corresponding genomic sequences exon-intron association of PEPC genes was obtained. Conserved motifs of all PEPC genes were identified and analyzed using the online MEME suite server (http://meme-suite.org/, accessed on 25 June 2021). The following parameters were set for analysis: maximum numbers of different motifs, 10; minimum width, 10; and maximum width, 50. The promoter region analysis (cis-regulatory elements), was performed through the online PlantCARE database (http://bioinformatics.psb.ugent.be/webtools/plantcare/html/, accessed on 25 June 2021) and visualized using TBtools software package v0.6655 [26].

2.4. Syntenic Analysis of PEPCs in Five Rosaceae Species

By using TBtools software package v0.6655 [26], the combined gff3-file of the genomes of studied species was used to investigate the distribution and mapping of PEPC genes on the chromosomes. By using the MCScanX tool kit [27] duplicated PEPC genes were identified in five Rosaceae species. Concisely, protein sequences from those five Rosaceae species were used for BLASTP analysis (http://www.ncbi.nlm.nih.gov/blast/blast.cgi, accessed on 13 June 2021), with less than 1 × 10−5 E-value. The BLASTP outputs with gene-location files were used as an input for MCScanX to identify syntenic gene pairs and duplication types with default settings. Circos function in TBtools [26] was used for schematic diagram construction of putative PEPC genes duplication, and putative WGD/segmental-duplicated genes or tandem-duplicated genes were connected by links [28].

2.5. Ka and Ks Calculation

MCScanX downstream analysis tools were used to annotate the Ka and Ks substitution rates of syntenic gene pairs. To determine Ka and Ks, KaKs_Calculator 2.0 was used with Nei–Gojobori (NG) method [29,30].

2.6. RNA Isolation and Quantitative RT-PCR Analysis

Five different tissues i.e., stem, root, mature leaf, full-bloom flower, and ripened fruit of loquat, apple, peach, strawberry, and pear were obtained for quantitative RT-PCR evaluation. Total RNA was extracted using a Total RNA kit (TianGen Biotech, Beijing, China). Quantity and quality of RNA were checked through agarose gel electrophoresis as well as by NanoDrop N-1000 spectrophotometer (NanoDrop Technologies, Wilmington, DE, USA). Prime Script RT Reagent Kit with a gDNA Eraser (TaKaRa, Dalian, China) was used to synthesize first-strand cDNA from 1 µg of high-quality total RNA. Real-time qPCR analysis was carried out using high-performance real-time PCR (LightCycler® 96, Roche Applied Science, Penzberg, Germany). The qRT-PCR began with “preincubation” for 5 min at 95 °C, followed by a “2-step amplification” with 40 cycles of 95 °C for 10 s and 60 °C for 30 s, “melting” at 95 °C for 10 s, 65 °C for 1 min and 97 °C for 1 s, and “cooling” at 37 °C for 30 s. The melting curve was created to identify the amplicon specificity. The relative expression level of each gene was measured according to the cycle threshold (Ct), also known as the 2−ΔΔCT method [31]. The validation of 2−ΔΔCt method was carried out by ΔCt variation analysis at different template concentrations [31,32,33]. All analysis consisted of 3 biological and 3 technical replicates. Following the previous studies actin genes [7,29,34,35,36], was selected as the control gene. Table 1 shows all the used primers for qRT-PCR analysis.

3. Results

3.1. Identification and Characterization of PEPC Gene Family in Five Rosaceae Species

We found total 27 putative PEPCs (loquat 4, apple 6, peach 3, strawberry 9, and pear 5) were identified in the genomes of five Rosaceae species. The details about the location of genes on chromosomes, CDS, and peptide sequence length are shown in Table 2.
We characterized the general information of 27 PEPCs in Table 3, which also shows the biochemical and physiological characteristics of respective proteins. PEPC proteins had a length range of 720 to 1144 amino acids, and the molecular weight was predicted from 81.84 kDa to 129.17 kDa. While, theoretical isoelectric point (pI) for PEPC proteins ranged from 5.57 to 7.80, and the grand average of hydropathicity (GRAVY) was noticed as −0.450 to −0.323. Additionally, PEPC proteins show the instability index from 44.36 to 54.34, and the aliphatic index was also noticed as 87.39 to 91.99. All proteins of PEPC genes had flexible structures due to the presence of coils (Figure S1). All proteins had at least two large α helices, while β sheets were not common.
Genomic structural analysis indicated the number of exons for each PEPC gene ranged from 8 to 20. Most of the PEPC genes contained 10 exons, while four PEPC genes i.e., MD13G1049200, MD16G1050300, Prupe.1G302700, and pycom16g04410, contained 20 exons, whereas Prupe.3G118300 and pycom11g23090 contained 8 and 11 exons, respectively. This proposed that the loss and gain of exons had occurred in the PEPCs of five Rosaceae species.
Subcellular localization exploration validated that 27 PEPC proteins were located in the nucleus, mitochondria, vacuole, chloroplast, endoplasmic reticulum, plasma membrane, and Golgi apparatus (Figure 1A). Additionally, among 27 PEPC proteins, 2 putative functional domains were identified, and the PEPcase superfamily domain was present in all PEPC proteins (Figure 1B).

3.2. Phylogenetic Analysis of PEPC Genes in Five Rosaceae Species

Utilizing multiple sequence alignment tools among the protein sequences of E. japonica and four other plant species (M. domestica, P. persica, F. ananassa, and P. cummunis), a phylogenetic tree was created by following neighbor joining (NJ) method. Results demonstrated that all studied PEPCs among five species were clustered into three discrete subgroups (A–C) (Figure 2). All EjPEPC genes were allocated in subgroups A and B.

3.3. Conserved Motif Analysis of PEPC Genes of Five Rosaceae Species

By utilizing online servers of MEME, the distribution of conserved motifs for PEPC genes was thoroughly assessed; a range from 7 to 10 presumed conserved motifs was acknowledged among PEPC proteins. The majority of PEPCs contained 10 motifs, while one PEPC gene i.e., Prupe.3G118300 contained 7 motifs. Figure 3 shows the distribution of conserved motifs. Thus, it can be assumed that during the evolutionary process PEPCs of five Rosaceae species evidently exhibited extreme conservation.

3.4. Promoter Region Analysis of PEPCs in Five Rosaceae Species

To further investigate the transcriptional mechanism of PEPCs, 1000 bps from the upstream region of PEPCs were subjected to promoter analysis (Figure 4). Several plant growth hormones related cis-elements (i.e., ABRE, AuxRE, CGTCA-motif, GARE-motif, TCA-element, TGACG-motif, and TGA-element) were detected in the promoter regions of PEPCs. These cis-elements were responsible for the regulation of abscisic acid, auxins, methyl jasmonate, gibberellins, and salicylic acid. Besides, stress response cis-elements i.e., LTR and TC-rich repeats were also identified in several genes. The ACE, AE-box, GA-motif, Gap-box, GATA-motif, GATT-motif, I-box, LAMP-element, TCCC-motif, and TCT-motif were found as light-responsive cis-elements. Apart from the aforementioned cis-elements, CAT-box was also identified, meristem expression.

3.5. Syntenic Analysis of PEPC Genes in Five Rosaceae Species

The gene duplication among PEPC genes of five Rosaceae species is presented in Figure 5. Among these genes, in E. japonica, all 4 genes were located on different chromosomes i.e., chromosome 4, 11, 12, and 13. In M. domestica, all 6 genes were located on different chromosomes i.e., 3, 9, 11, 13, 16, and 17. Similarly, in P. persica, 3 PEPC genes were located on chromosome 1, 3, and 4. In F. ananassa, chromosome 9, 10, 11, 12, 21, 22, and 24 had one PEPC gene, each, while two genes were located on chromosome 23. In P. communis, all PEPC genes were found on different chromosomes i.e., 3, 9, 11, 16, and 17. A total of 16 (59.25%) PEPC genes in five Rosaceae species revealed WGD/segmental duplication. By this it can be suggested that WGD/segmental duplication is a significant step for the expansion of PEPC gene family in Rosaceae species, because of the duplication process retention of several duplicated genes take place in the genome. For the estimation of evolution rate and selective pressure, Ka (nonsynonymous)/Ks(synonymous) ratio (ω) was used [37]. Table 4 shows analyzed evolutionary pattern among the genomes of studied species, the ω values of gene duplication pairs were calculated to observe and understand selective pressures upon gene duplication, ω value for all PEPC gene pairs was observed to be less than 1, showing the purifying selection has made a strong influence for PEPC evolution occurrence. Thus, it can be concluded that during the domestication of Rosaceae species evolutionary pattern of PEPCs was conserved.

3.6. Expression Patterns of PEPC genes in Five Rosacea Species

Further, we evaluated the relative expressions of PEPC genes in five different tissues of loquat, apple, peach, strawberry, and pear (Figure 6). Relative expression of PEPC genes showed noteworthy differences among all selected tissues for analysis in all species. Briefly, among 4 loquat PEPC genes, expressions of 3 were noticed as relatively less (i.e., EVM0016068.1, EVM0004511.1, and EVM0022212.1), ranging from 0.46 to 3.2. Briefly, two loquat PEPC genes (EVM0016068.1 and EVM0023388.1) showed high expression levels in root and leaf tissues; while, EVM0004511.1 was highly expressed in flower, leaf, and stem. EVM0022212.1 showed maximum expression in leaf tissues. Among 6 apple PEPCs, MD09G1237900 showed higher expressions in leaves and roots, while maximum expressions of MD11G1261900 and MD13G1049200 were observed in leaves and roots, respectively. In case of peach PEPCs, 2 genes (Prupe.1G302700 and Prupe.4G166400) showed high relative expressions in stem, leaves, flowers and fruits. Among 9 strawberry PEPCs, only 5 were differentially expressed in different plant tissues. The maximum transcript level (4.14) was showed by gene_175.31 in stem tissues followed by gene_206.27 in stem and leaf tissues. The gene_214.28 was only expressed in roots among all examined plant tissues. Among 5 PcPEPCs, 2 genes (pycom09g15640 and pycom16g04410) showed high transcript level in roots, while pycom17g23400 exhibited maximum relative expression in stem. The relative expressions of 6 PEPC genes (i.e., MD17G1230800, Prupe.3G118300, gene_89.20, gene_78.22, gene_170.27, gene_121.5) were not detected.

4. Discussion

PEPCs are functionally so much important for the regulation of several physiological processes, such as guard-cells movement, fruit acidity, metal-toxicity tolerance, and mineral nutrition. However, a comprehensive analysis of PEPC genes in Rosaceae species has not been reported yet. In present study, 4, 6, 3, 9 and 5 (total 27) PEPCs were identified in the genomes of loquat [19], apple [20], peach [38], strawberry [39] and pear [21], respectively. In previous studies, PEPC family members were studied in different species [12,13,14,15,16,17]. The conserved domain analysis performed with PEPC proteins of five Rosaceae species revealed that all genes had similar conserved domains (Figure 2).
The current investigation identified 27 PEPC genes in five Rosaceae species. Through phylogenetic analysis, all genes were classified into three subgroups (Figure 3), while, the presence of EjPEPC and MdPEPC genes in the same group of classification denotes that PEPC genes from both species have a close relationship. Novelty in protein functioning induced owing to the result of evolution is primarily due to the gene duplication; which resulted in the loss of functionality for some genes while enhancing protein function for others [40]. Characterization of PEPC gene family in Rosaceae species was done to evaluate their conserved motifs as well as genomic structure, which showed that the PEPC genes almost retained similar exons and conserved motifs (Figure 3). The value of ω is known to be important for diversity measurement [7,41], as we observed lesser than 1 ω value for all the duplicated PEPC gene pairs (Table 2), exhibiting the evolutionary rate was slow and purifying selection during evolution was mainly focused.
Elaborating the functional importance of PEPC genes, it was reported to have plant abiotic stress response mechanism among different species, such as chilling injury and salt stress, induced PEPC gene expression in sorghum, wheat, and Arabidopsis [42,43,44]; while, a study conducted on rice overexpressing PEPC transgenic lines, were noticed with increased photosynthetic rate under temperature and high light stress conditions [45]. In this study, it was observed in the promoter regions of most PEPC genes to have 12 stress-responsive cis-regulated elements such as ABA, MeJA, cold, heat, light, and meristem-related elements, which indicates PEPCs could also be potentially important for phytohormone and stress responses among Rosaceae species (Figure 5).
For Rosaceae species, identified PEPC genes are diverse which may arise from ancient polyploidy as well as recent WGD events. The number of PEPCs among apple and strawberry were noticed almost double among the other three Rosaceae species, indicates they have gone through a recent lineage-specific WGD duplication [46]. Different modes of duplication are key characters for the evolutionary process of a eukaryotic genome to enhance the function of some genes for the betterment of a species as well as deletion of other non-essential genes, such as WGD or segmental duplication, tandem and dispersed duplication [47,48]. However, segmental duplication, which is more like a small-scale duplication, is difficult to distinguish from WGD [49]. There are some reported examples which likely to expand by duplication, such as HCT (Hydroxycinnamoyl Transferase) and CPA (Cation Proton Antiporters) [50,51]. The term tandem duplicates which are adjusted to one another on chromosomes are known as paralogs which may be derived from illegitimate chromosomal recombination [52]. Two large gene families, AP2/ERF and WRKY, expanded primarily through tandem duplication [53,54]. It was reported by several studies that Rosaceae species like apple, pear, and loquat underwent at least a double duplication process [19,20,55]. In the present study, results of synteny analysis also verified that Rosacea species likely derived as a result of segmental duplication which caused the expansion of the PEPC gene family members. The 16 PEPC genes exhibited WGD/segmental duplication in the genomes of examined Rosaceae species (Figure 6). These results indicated that the duplication of PEPC genes is related to WGD/segmental duplication during the process of speciation and domestication.
PEPC genes have been reported among several tissues, including seeds of barley (H. vulgare) and castor (R. communis) [13], seedlings and cell cultures of A. thaliana, tomato (S. lycopersicum) and potato (S. tuberosum) [15,56], root nodules of soybean (G. max) and lotus (L. japonicas) [14,57]. Besides, PEPC proteins are also integral for the non-photosynthetic process like seed germination, fruit ripening, seed formation, and guard cell movement during stomatal opening [2]. In present investigation, we also noticed that EVM0004511.1 has a higher expression level in stem and leaf, but not in root and flower. Besides, other EjPEPCs (EVM0016068.1 and EVM0023388.1) had lower relative expression in flower tissues (Figure 6), which is consistent with the earlier findings in Arabidopsis BTPC, to have lower expression levels in flower and silique [12]. Almost all PEPC genes were predicted to be localized in cytonuclear region, while the higher genetic expressions of EVM0016068.1, EVM0004511.1, EVM0023388.1, EVM0022212.1, MD09G1237900, MD11G1261900, MD16G1050300, gene_117.42, gene_201.40, gene_206.27, pycom09g15640 and pycom11g23090 were recorded in leaves and stem, indicating their possible role in improving plant growth and development [58]. All these findings indicate that PEPCs have various physiological roles in plant species and proposed that members of PEPC gene family may be considered important factors in regulating variety of functions in Rosaceae species.

5. Conclusions

In present study, 27 PEPC genes were identified in five Rosaceae species i.e., loquat, apple, peach, strawberry, and pear. All PEPC genes were subjected to conserved domains, gene structure, conserved motif, phylogenetic, syntenic, and expression analysis. Based on the subcellular localization analysis, it was predicted that all PEPC proteins were localized in cytonuclear region. Syntenic analysis revealed that WGD/segmental duplication played an important role in the expansion of PEPC gene family in Rosacea species. The Ka and Ks values of duplicated genes indicated that PEPC genes had undergone a strong purifying selection. In addition, as the result of expression analysis, some genes were differentially expressed in different plant tissues i.e., root, mature leaf, stem, full-bloom flower, and ripened fruit. This study laid the basis for studying the roles of PEPC genes in developmental processes of Rosaceae species and provide the foundation for further functional analysis such as overexpression, knockout via CRISPR/Cas9 systems, etc.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/agronomy12010025/s1, Figure S1: The predicted structures (3D) of PEPC proteins in five Rosaceae species.

Author Contributions

Conceptualization, M.M.A. and F.C.; methodology, M.M.A., S.M. and A.F.Y.; validation, S.M.A. and F.C.; data curation, M.M.A.; writing—original draft preparation, C.Z., M.M.A. and S.M.A.; writing—review and editing, S.G., S.A. and M.A.A.A.; supervision, F.C.; project administration, F.C.; funding acquisition, F.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by “Fujian Provincial Development and Reform Commission, grant number 2013-772” and “Key Laboratory of Loquat Germplasm Innovation and Utilization, Fujian Province University (Putian), grant number 102/KLh19010A”.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

Authors would like to thank Ali Raza and Yasir Sharif for their valuable suggestions during revision of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Izui, K.; Matsumura, H.; Furumoto, T.; Kai, Y. PHOSPHO ENOL PYRUVATE CARBOXYLASE: A New Era of Structural Biology. Annu. Rev. Plant Biol. 2004, 55, 69–84. [Google Scholar] [CrossRef]
  2. O’Leary, B.; Park, J.; Plaxton, W.C. The remarkable diversity of plant PEPC (phosphoenolpyruvate carboxylase): Recent insights into the physiological functions and post-translational controls of non-photosynthetic PEPCs. Biochem. J. 2011, 436, 15–34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. O’Leary, B.; Fedosejevs, E.T.; Hill, A.T.; Bettridge, J.; Park, J.; Rao, S.K.; Leach, C.A.; Plaxton, W.C. Tissue-specific expression and post-translational modifications of plant- and bacterial-type phosphoenolpyruvate carboxylase isozymes of the castor oil plant, Ricinus communis L. J. Exp. Bot. 2011, 62, 5485–5495. [Google Scholar] [CrossRef] [Green Version]
  4. Cousins, A.B.; Baroli, I.; Badger, M.R.; Ivakov, A.; Lea, P.J.; Leegood, R.C.; von Caemmerer, S. The Role of Phospho enol pyruvate Carboxylase during C4 Photosynthetic Isotope Exchange and Stomatal Conductance. Plant Physiol. 2007, 145, 1006–1017. [Google Scholar] [CrossRef] [Green Version]
  5. Wang, N.; Zhong, X.; Cong, Y.; Wang, T.; Yang, S.; Li, Y.; Gai, J. Genome-wide Analysis of Phosphoenolpyruvate Carboxylase Gene Family and Their Response to Abiotic Stresses in Soybean. Sci. Rep. 2016, 6, 38448. [Google Scholar] [CrossRef] [PubMed]
  6. Pan, T.; Ali, M.M.; Gong, J.; She, W.; Pan, D.; Guo, Z.; Yu, Y.; Chen, F. Fruit Physiology and Sugar-Acid Profile of 24 Pomelo (Citrus grandis (L.) Osbeck) Cultivars Grown in Subtropical Region of China. Agronomy 2021, 11, 2393. [Google Scholar] [CrossRef]
  7. Ma, B.; Yuan, Y.; Gao, M.; Qi, T.; Li, M.; Ma, F. Genome-Wide Identification, Molecular Evolution, and Expression Divergence of Aluminum-Activated Malate Transporters in Apples. Int. J. Mol. Sci. 2018, 19, 2807. [Google Scholar] [CrossRef] [Green Version]
  8. Zhang, X.; Wei, X.; Ali, M.M.; Rizwan, H.M.; Li, B.; Li, H.; Jia, K.; Yang, X.; Ma, S.; Li, S.; et al. Changes in the Content of Organic Acids and Expression Analysis of Citric Acid Accumulation-Related Genes during Fruit Development of Yellow (Passiflora edulis f. flavicarpa) and Purple (Passiflora edulis f. edulis) Passion Fruits. Int. J. Mol. Sci. 2021, 22, 5765. [Google Scholar] [CrossRef]
  9. Meyer, S.; De Angeli, A.; Fernie, A.R.; Martinoia, E. Intra- and extra-cellular excretion of carboxylates. Trends Plant Sci. 2010, 15, 40–47. [Google Scholar] [CrossRef] [Green Version]
  10. Ryan, P.R.; Skerrett, M.; Findlay, G.P.; Delhaize, E.; Tyerman, S.D. Aluminum activates an anion channel in the apical cells of wheat roots. Proc. Natl. Acad. Sci. USA 1997, 94, 6547–6552. [Google Scholar] [CrossRef] [Green Version]
  11. Ma, B.; Liao, L.; Zheng, H.; Chen, J.; Wu, B.; Ogutu, C.; Li, S.; Korban, S.S.; Han, Y. Genes Encoding Aluminum-Activated Malate Transporter II and their Association with Fruit Acidity in Apple. Plant Genome 2015, 8. [Google Scholar] [CrossRef] [Green Version]
  12. Sánchez, R.; Cejudo, F.J. Identification and Expression Analysis of a Gene Encoding a Bacterial-Type Phospho enol pyruvate Carboxylase from Arabidopsis and Rice. Plant Physiol. 2003, 132, 949–957. [Google Scholar] [CrossRef] [Green Version]
  13. Murmu, J.; Plaxton, W.C. Phosphoenolpyruvate carboxylase protein kinase from developing castor oil seeds: Partial purification, characterization, and reversible control by photosynthate supply. Planta 2007, 226, 1299–1310. [Google Scholar] [CrossRef] [PubMed]
  14. Nakagawa, T.; Izumi, T.; Banba, M.; Umehara, Y.; Kouchi, H.; Izui, K.; Hata, S. Characterization and Expression Analysis of Genes Encoding Phosphoenolpyruvate Carboxylase and Phosphoenolpyruvate Carboxylase Kinase of Lotus japonicus, a Model Legume. Mol. Plant-Microbe Interact. 2003, 16, 281–288. [Google Scholar] [CrossRef]
  15. Rontein, D.; Dieuaide-Noubhani, M.; Dufourc, E.J.; Raymond, P.; Rolin, D. The Metabolic Architecture of Plant Cells. J. Biol. Chem. 2002, 277, 43948–43960. [Google Scholar] [CrossRef] [Green Version]
  16. Sima, B.D.; Desjardins, Y. Sucrose supply enhances phosphoenolpyruvate carboxylase phosphorylation level in in vitro Solanum tuberosum. Plant Cell. Tissue Organ Cult. 2001, 67, 235–242. [Google Scholar] [CrossRef]
  17. Duff, S.; Chollet, R. In Vivo Regulation of Wheat-Leaf Phosphoenolpyruvate Carboxylase by Reversible Phosphorylation. Plant Physiol. 1995, 107, 775–782. [Google Scholar] [CrossRef] [Green Version]
  18. Engelmann, S.; Blesing, O.E.; Gowik, U.; Svensson, P.; Westhoff, P. Molecular evolution of C4 phosphoenolpyruvate carboxylase in the genus Flaveria? A gradual increase from C3 to C4 characteristics. Planta 2003, 217, 717–725. [Google Scholar] [CrossRef]
  19. Jiang, S.; An, H.; Xu, F.; Zhang, X. Chromosome-level genome assembly and annotation of the loquat (Eriobotrya japonica) genome. Gigascience 2020, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Daccord, N.; Celton, J.-M.; Linsmith, G.; Becker, C.; Choisne, N.; Schijlen, E.; van de Geest, H.; Bianco, L.; Micheletti, D.; Velasco, R.; et al. High-quality de novo assembly of the apple genome and methylome dynamics of early fruit development. Nat. Genet. 2017, 49, 1099–1106. [Google Scholar] [CrossRef]
  21. Chagné, D.; Crowhurst, R.N.; Pindo, M.; Thrimawithana, A.; Deng, C.; Ireland, H.; Fiers, M.; Dzierzon, H.; Cestaro, A.; Fontana, P.; et al. The Draft Genome Sequence of European Pear (Pyrus communis L. ‘Bartlett’). PLoS ONE 2014, 9, e92644. [Google Scholar] [CrossRef] [PubMed]
  22. Feria, A.B.; Bosch, N.; Sánchez, A.; Nieto-Ingelmo, A.I.; de la Osa, C.; Echevarría, C.; García-Mauriño, S.; Monreal, J.A. Phosphoenolpyruvate carboxylase (PEPC) and PEPC-kinase (PEPC-k) isoenzymes in Arabidopsis thaliana: Role in control and abiotic stress conditions. Planta 2016, 244, 901–913. [Google Scholar] [CrossRef] [PubMed]
  23. Finn, R.D.; Mistry, J.; Tate, J.; Coggill, P.; Heger, A.; Pollington, J.E.; Gavin, O.L.; Gunasekaran, P.; Ceric, G.; Forslund, K.; et al. The Pfam protein families database. Nucleic Acids Res. 2010, 38, D211–D222. [Google Scholar] [CrossRef] [PubMed]
  24. Eddy, S.R. Accelerated Profile HMM Searches. PLoS Comput. Biol. 2011, 7, e1002195. [Google Scholar] [CrossRef] [Green Version]
  25. Kumar, S.; Stecher, G.; Li, M.; Knyaz, C.; Tamura, K. MEGA X: Molecular Evolutionary Genetics Analysis across Computing Platforms. Mol. Biol. Evol. 2018, 35, 1547–1549. [Google Scholar] [CrossRef]
  26. 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]
  27. Wang, Y.; Tang, H.; DeBarry, J.D.; Tan, X.; Li, J.; Wang, X.; Lee, T.-H.; Jin, H.; Marler, B.; Guo, H.; et al. MCScanX: A toolkit for detection and evolutionary analysis of gene synteny and collinearity. Nucleic Acids Res. 2012, 40, e49. [Google Scholar] [CrossRef] [Green Version]
  28. Krzywinski, M.; Schein, J.; Birol, I.; Connors, J.; Gascoyne, R.; Horsman, D.; Jones, S.J.; Marra, M.A. Circos: An information aesthetic for comparative genomics. Genome Res. 2009, 19, 1639–1645. [Google Scholar] [CrossRef] [Green Version]
  29. Gan, X.; Jing, Y.; Shahid, M.Q.; He, Y.; Baloch, F.S.; Lin, S.; Yang, X. Identification, phylogenetic analysis, and expression patterns of the SAUR gene family in loquat (Eriobotrya japonica). Turkish J. Agric. For. 2020, 44, 15–23. [Google Scholar] [CrossRef]
  30. Wang, D.; Zhang, Y.; Zhang, Z.; Zhu, J.; Yu, J. KaKs_Calculator 2.0: A Toolkit Incorporating Gamma-Series Methods and Sliding Window Strategies. Genom. Proteom. Bioinform. 2010, 8, 77–80. [Google Scholar] [CrossRef] [Green Version]
  31. Livak, K.J.; Schmittgen, T.D. Analysis of Relative Gene Expression Data Using Real-Time Quantitative PCR and the 2−ΔΔCT Method. Methods 2001, 25, 402–408. [Google Scholar] [CrossRef]
  32. De Rossi, S.; Di Marco, G.; Bruno, L.; Gismondi, A.; Canini, A. Investigating the Drought and Salinity Effect on the Redox Components of Sulla coronaria (L.) Medik. Antioxidants 2021, 10, 1048. [Google Scholar] [CrossRef]
  33. Xu, L.; Xu, H.; Cao, Y.; Yang, P.; Feng, Y.; Tang, Y.; Yuan, S.; Ming, J. Validation of Reference Genes for Quantitative Real-Time PCR during Bicolor Tepal Development in Asiatic Hybrid Lilies (Lilium spp.). Front. Plant Sci. 2017, 8, 669. [Google Scholar] [CrossRef] [Green Version]
  34. Xue, C.; Guan, S.-C.; Chen, J.-Q.; Wen, C.-J.; Cai, J.-F.; Chen, X. Genome wide identification and functional characterization of strawberry pectin methylesterases related to fruit softening. BMC Plant Biol. 2020, 20, 13. [Google Scholar] [CrossRef]
  35. Lu, B.; Wang, Y.; Zhang, G.; Feng, Y.; Yan, Z.; Wu, J.; Chen, X. Genome-Wide Identification and Expression Analysis of the Strawberry FvbZIP Gene Family and the Role of Key Gene FabZIP46 in Fruit Resistance to Gray Mold. Plants 2020, 9, 1199. [Google Scholar] [CrossRef]
  36. Zhang, M.-Y.; Xue, C.; Hu, H.; Li, J.; Xue, Y.; Wang, R.; Fan, J.; Zou, C.; Tao, S.; Qin, M.; et al. Genome-wide association studies provide insights into the genetic determination of fruit traits of pear. Nat. Commun. 2021, 12, 1144. [Google Scholar] [CrossRef]
  37. Akhunov, E.D.; Sehgal, S.; Liang, H.; Wang, S.; Akhunova, A.R.; Kaur, G.; Li, W.; Forrest, K.L.; See, D.; Šimková, H.; et al. Comparative Analysis of Syntenic Genes in Grass Genomes Reveals Accelerated Rates of Gene Structure and Coding Sequence Evolution in Polyploid Wheat. Plant Physiol. 2012, 161, 252–265. [Google Scholar] [CrossRef] [Green Version]
  38. Verde, I.; Jenkins, J.; Dondini, L.; Micali, S.; Pagliarani, G.; Vendramin, E.; Paris, R.; Aramini, V.; Gazza, L.; Rossini, L.; et al. The Peach v2.0 release: High-resolution linkage mapping and deep resequencing improve chromosome-scale assembly and contiguity. BMC Genom. 2017, 18, 225. [Google Scholar] [CrossRef] [Green Version]
  39. Cheng, H.; Li, J.; Zhang, H.; Cai, B.; Gao, Z.; Qiao, Y.; Mi, L. The complete chloroplast genome sequence of strawberry (Fragaria × ananassa Duch.) and comparison with related species of Rosaceae. PeerJ 2017, 5, e3919. [Google Scholar] [CrossRef] [Green Version]
  40. Nakano, T.; Suzuki, K.; Fujimura, T.; Shinshi, H. Genome-Wide Analysis of the ERF Gene Family in Arabidopsis and Rice. Plant Physiol. 2006, 140, 411–432. [Google Scholar] [CrossRef] [Green Version]
  41. Wang, W.; Zhou, H.; Ma, B.; Owiti, A.; Korban, S.S.; Han, Y. Divergent Evolutionary Pattern of Sugar Transporter Genes is Associated with the Difference in Sugar Accumulation between Grasses and Eudicots. Sci. Rep. 2016, 6, 29153. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. García-Mauriño, S.; Monreal, J.; Alvarez, R.; Vidal, J.; Echevarría, C. Characterization of salt stress-enhanced phosphoenolpyruvate carboxylase kinase activity in leaves of Sorghum vulgare: Independence from osmotic stress, involvement of ion toxicity and significance of dark phosphorylation. Planta 2003, 216, 648–655. [Google Scholar] [CrossRef]
  43. González, M.-C.; Sánchez, R.; Cejudo, F.J. Abiotic stresses affecting water balance induce phosphoenolpyruvate carboxylase expression in roots of wheat seedlings. Planta 2003, 216, 985–992. [Google Scholar] [CrossRef]
  44. Sánchez, R.; Flores, A.; Cejudo, F.J. Arabidopsis phosphoenolpyruvate carboxylase genes encode immunologically unrelated polypeptides and are differentially expressed in response to drought and salt stress. Planta 2006, 223, 901–909. [Google Scholar] [CrossRef]
  45. Bandyopadhyay, A.; Datta, K.; Zhang, J.; Yang, W.; Raychaudhuri, S.; Miyao, M.; Datta, S.K. Enhanced photosynthesis rate in genetically engineered indica rice expressing pepc gene cloned from maize. Plant Sci. 2007, 172, 1204–1209. [Google Scholar] [CrossRef]
  46. Linlin, X.; Xin, Q.; Mingyue, Z.; Shaoling, Z. Genome-Wide analysis of aluminum-activated malate transporter family genes in six rosaceae species, and expression analysis and functional characterization on malate accumulation in Chinese white pear. Plant Sci. 2018, 274, 451–465. [Google Scholar] [CrossRef]
  47. Friedman, R.; Hughes, A.L. Pattern and Timing of Gene Duplication in Animal Genomes. Genome Res. 2001, 11, 1842–1847. [Google Scholar] [CrossRef]
  48. Moore, R.C.; Purugganan, M.D. The early stages of duplicate gene evolution. Proc. Natl. Acad. Sci. USA 2003, 100, 15682–15687. [Google Scholar] [CrossRef] [Green Version]
  49. Wang, Y.; Wang, X.; Paterson, A.H. Genome and gene duplications and gene expression divergence: A view from plants. Ann. N. Y. Acad. Sci. 2012, 1256, 1–14. [Google Scholar] [CrossRef]
  50. Zhou, H.; Qi, K.; Liu, X.; Yin, H.; Wang, P.; Chen, J.; Wu, J.; Zhang, S. Genome-wide identification and comparative analysis of the cation proton antiporters family in pear and four other Rosaceae species. Mol. Genet. Genom. 2016, 291, 1727–1742. [Google Scholar] [CrossRef]
  51. Ma, C.; Zhang, H.; Li, J.; Tao, S.; Qiao, X.; Korban, S.S.; Zhang, S.; Wu, J. Genome-wide analysis and characterization of molecular evolution of the HCT gene family in pear (Pyrus bretschneideri). Plant Syst. Evol. 2017, 303, 71–90. [Google Scholar] [CrossRef]
  52. Freeling, M. Bias in Plant Gene Content Following Different Sorts of Duplication: Tandem, Whole-Genome, Segmental, or by Transposition. Annu. Rev. Plant Biol. 2009, 60, 433–453. [Google Scholar] [CrossRef]
  53. Du, D.; Hao, R.; Cheng, T.; Pan, H.; Yang, W.; Wang, J.; Zhang, Q. Genome-Wide Analysis of the AP2/ERF Gene Family in Prunus mume. Plant Mol. Biol. Rep. 2013, 31, 741–750. [Google Scholar] [CrossRef]
  54. Guo, C.; Guo, R.; Xu, X.; Gao, M.; Li, X.; Song, J.; Zheng, Y.; Wang, X. Evolution and expression analysis of the grape (Vitis vinifera L.) WRKY gene family. J. Exp. Bot. 2014, 65, 1513–1528. [Google Scholar] [CrossRef] [PubMed]
  55. Wu, J.; Wang, Z.; Shi, Z.; Zhang, S.; Ming, R.; Zhu, S.; Khan, M.A.; Tao, S.; Korban, S.S.; Wang, H.; et al. The genome of the pear (Pyrus bretschneideri Rehd.). Genome Res. 2013, 23, 396–408. [Google Scholar] [CrossRef] [Green Version]
  56. Gregory, A.L.; Hurley, B.A.; Tran, H.T.; Valentine, A.J.; She, Y.-M.; Knowles, V.L.; Plaxton, W.C. In vivo regulatory phosphorylation of the phosphoenolpyruvate carboxylase AtPPC1 in phosphate-starved Arabidopsis thaliana. Biochem. J. 2009, 420, 57–65. [Google Scholar] [CrossRef] [Green Version]
  57. Xu, W.; Zhou, Y.; Chollet, R. Identification and expression of a soybean nodule-enhanced PEP-carboxylase kinase gene (NE-PpcK) that shows striking up-/down-regulation in vivo. Plant J. 2003, 34, 441–452. [Google Scholar] [CrossRef] [PubMed]
  58. Ali, M.M.; Alam, S.M.; Anwar, R.; Ali, S.; Shi, M.; Liang, D.; Lin, Z.; Chen, F. Genome-Wide Identification, Characterization and Expression Profiling of Aluminum-Activated Malate Transporters in Eriobotrya japonica Lindl. Horticulturae 2021, 7, 441. [Google Scholar] [CrossRef]
Figure 1. (A) Prediction of subcellular localization of PEPC proteins of five Rosaceae species. Higher signal levels are shown in red, lower signal levels are denoted in blue, while the white color represents no available data. Abbreviations: Nucl—Nucleus; Cyto—Cytonuclear; Mito—Mitochondria; Vacu—Vacuole; Chlo—Chloroplast; E.R.—Endoplasmic reticulum; Plas—Plasma membrane; Golgi—Golgi apparatus. (B) Conserved domain query of PEPC genes in five Rosaceae species genomes.
Figure 1. (A) Prediction of subcellular localization of PEPC proteins of five Rosaceae species. Higher signal levels are shown in red, lower signal levels are denoted in blue, while the white color represents no available data. Abbreviations: Nucl—Nucleus; Cyto—Cytonuclear; Mito—Mitochondria; Vacu—Vacuole; Chlo—Chloroplast; E.R.—Endoplasmic reticulum; Plas—Plasma membrane; Golgi—Golgi apparatus. (B) Conserved domain query of PEPC genes in five Rosaceae species genomes.
Agronomy 12 00025 g001
Figure 2. Phylogenetic tree analysis of PEPC genes in five Rosaceae species using the neighbour-joining (NJ) method.
Figure 2. Phylogenetic tree analysis of PEPC genes in five Rosaceae species using the neighbour-joining (NJ) method.
Agronomy 12 00025 g002
Figure 3. Conserved motifs identified in PEPCs of five Rosaceae species.
Figure 3. Conserved motifs identified in PEPCs of five Rosaceae species.
Agronomy 12 00025 g003
Figure 4. The cis-regulatory elements detected in the promoter sequences of PEPC genes in five Rosaceae species.
Figure 4. The cis-regulatory elements detected in the promoter sequences of PEPC genes in five Rosaceae species.
Agronomy 12 00025 g004
Figure 5. Chromosomal distribution and gene duplication of the PEPC genes in five Rosaceae species. Gene IDs of loquat, apple, peach, strawberry, and pear are labeled in red, blue, green, purple, and black colors, respectively. Red lines represent the putative WGD/segmental-duplication of genes.
Figure 5. Chromosomal distribution and gene duplication of the PEPC genes in five Rosaceae species. Gene IDs of loquat, apple, peach, strawberry, and pear are labeled in red, blue, green, purple, and black colors, respectively. Red lines represent the putative WGD/segmental-duplication of genes.
Agronomy 12 00025 g005
Figure 6. Relative expressions of PEPCs in different tissues of five Rosaceae species i.e., loquat, apple, peach, strawberry, and pear. Same letters indicate non-significant difference among treatments according to Fisher’s least significant difference (LSD) test, when p ≤ 0.05. Vertical bars indicate mean ± standard error (3 biological and 3 technical replicates). ND—not detected.
Figure 6. Relative expressions of PEPCs in different tissues of five Rosaceae species i.e., loquat, apple, peach, strawberry, and pear. Same letters indicate non-significant difference among treatments according to Fisher’s least significant difference (LSD) test, when p ≤ 0.05. Vertical bars indicate mean ± standard error (3 biological and 3 technical replicates). ND—not detected.
Agronomy 12 00025 g006
Table 1. Details about primer sequences of PEPC genes of five Rosaceae species.
Table 1. Details about primer sequences of PEPC genes of five Rosaceae species.
SpecieGeneForward Primer (5′-3′)TmF (°C)Reverse Primer (5′-3′)TmR (°C)
Eriobotrya japonicaEVM0016068.1GCAGTTCTTGGAACCTCTCG59.8CCAATTTCCAGATGCTTGGT57.8
EVM0004511.1TTGCTGATGGAAGCCTTCTT57.8TGAGACACGAGCCATTCTTG59.9
EVM0023388.1GCAGTTCTTGGAACCTCTCG55.8CCAATTTCCAGATGCTTGGT57.8
EVM0022212.1TGGAGCCTCTCGAACTTTGT58.9CATTCTTGCCTACGCTCCTC58.9
EVM0004523.1 (actin)GGAGCGTGGATATTCCTTCA57.8GCTGCTTCCATTCCAATCAT55.8
Malus domesticaMD03G1242000AGGTCACAAGGGATGTTTGC57.8ACCGAGAATCACACGGTAGG59.9
MD09G1237900AACAGCCCCATCTGATGTTC57.8TGCTGCAGATAAACGACCAG57.8
MD11G1261900GAACCCCGATTTGTCGAGTA57.8TGGAACCTTGTTTGTGTCCA55.8
MD13G1049200AACCGCCTGGTTCTGTAATG57.8CCATGAGGTTACGCCACTTT57.8
MD16G1050300TTTGCAGAAAGATGCACGAC59.8CCGAATATCCAACCATCACC57.8
MD17G1230800AGGTCTTTGCTCCAAAAGCA58.8GGAGGAGTCCTTCGGATTTC59.9
MD04G1127400 (actin)CCGTGTTCCCTAGCATTGTT59.9CAGGAGCAACACGAAGTTCA60.0
Prunus persicaPrupe.1G302700TTTGCAGAAAGATGCACGAC55.8TGCTGCAGTAAATCGTCCAG57.8
Prupe.3G118300GTTGAGCTTTTGCAACGTGA55.8TTCTTGCTTCCCATTGATCC55.8
Prupe.4G166400ACCTCCCACTCCACAAGATG59.9GCAAACATCCCTTGTGACCT57.8
Prupe.6G078800 (actin)GGAGCGTGGTTATTCCTTCA60.1TGCAGATTCCATTCCAATCA60.0
Fragaria ananassagene_214.28TCTTGCAGATTGCTGGACAC57.8TGGTCGAACAGTCACATGGT57.8
gene_117.42TCTTGCAGATTGCTGGACAC55.8ACCAGGGGCATACTCACTTG59.9
gene_89.20CTTTTGCAGATTGCTGGACA57.8GTGGTCGAACAGTCACATGG58.7
gene_201.40CTTTTGCAGATTGCTGGACA55.8TGGTCGAACAGTCACATGGT61.9
gene_175.31GCCTGAGAAACAAAGGCAAG57.8TTTCACATGGCATTCACGTT59.9
gene_78.22CGGCCTTTCTCTTGTGAGAC57.8TCTTCGGTTTTGGGAACATC59.7
gene_170.27GGATCAATGGGAAGCAAGAA57.8GGTCCTCCTCCTCTTCCAAC55.8
gene_121.5AAGGGACGTGTGCTTATTGG57.8GCTTGTCTCTCACGTCACCA57.8
gene_206.27GCCTGAGAAACAAAGGCAAG59.9TTTCACATGGCATTCACGTT55.8
gene_191.47 (actin)TGAACTTCGTGTTGCTCCAG60.0ACACCATCCCCAGAGTCAAG60.0
Pyrus communispycom03g18910TCTTGCAGATTGCTGGACAC57.8GCCGGTTTGTTTGATTCACT55.8
pycom09g15640AGATGGTGTTTGCCAAGGTC57.8AGGAGTCGCTGCCTCAAATA57.8
pycom11g23090GCAGTTCTTGGAACCTCTCG59.9CCAATTTCCAGATGCTTGGT55.8
pycom16g04410AAGTATCGGTCGTGGAGGTG59.9CCATGAGGTTACGCCACTTT57.8
pycom17g23400TGGAGCCTCTCGAACTTTGT57.8TGCCTGTCTGATTCTTGTCG57.8
pycom03g07780 (actin)ACCACAGCTGAGCGAGAAAT60.0ATCATGGATGGCTGGAAGAG59.9
Before qRT-PCR analysis, the annealing efficiency of all primers was checked through normal PCR and 60 °C was found to be the optimum temperature for all primer pairs. TmF—Annealing temperature for forward primer; TmR—Annealing temperature for reverse primer.
Table 2. The basic information about PEPC genes in five Rosaceae species.
Table 2. The basic information about PEPC genes in five Rosaceae species.
SpecieGeneChromosomeStart SiteEnd SiteCDS (bps)Protein Length (A.A.)
Eriobotrya japonicaEVM0016068.14516117851668492898965
EVM0004511.11126908874269150912904967
EVM0023388.11233650012336560702898965
EVM0022212.113590634059122272904967
Malus domesticaMD03G1242000332684130326896942904967
MD09G1237900930178715301851282904967
MD11G12619001137648686376550992907968
MD13G1049200133459038346597431261041
MD16G1050300163556349356324631261041
MD17G12308001727925312279315282904967
Prunus persicaPrupe.1G3027001297079302971547731411046
Prupe.3G118300310180567101847712163720
Prupe.4G1664004966050996675022898965
Fragaria ananassagene_214.28921406659214144332898965
gene_117.421011752557117587912898965
gene_89.2011888839088945192898965
gene_201.401220094738201008722898965
gene_175.312117493253174992712886961
gene_78.2222787036278703622886961
gene_170.272316996113170019132886961
gene_121.52312176565121794542889962
gene_206.272420617837206240202886961
Pyrus communispycom03g18910319727998197328152898965
pycom09g15640915712657157181392901966
pycom11g2309011260924352609831732161071
pycom16g04410162776593278360134351144
pycom17g234001721795080218002202853950
Table 3. Summary information of physiological and biochemical properties, and structural analysis of the PEPC proteins in five Rosaceae species.
Table 3. Summary information of physiological and biochemical properties, and structural analysis of the PEPC proteins in five Rosaceae species.
GeneMW (kDa)pIInstability IndexAliphatic IndexGRAVYIntronsExons
EVM0016068.1110,093.82648.1791.77−0.3771010
EVM0004511.1110,394.496.5445.4889.17−0.404910
EVM0023388.1110,081.826.0446.8390.87−0.3911010
EVM0022212.1110,182.065.9745.7888.98−0.387910
MD03G1242000110,304.06647.9191.69−0.3791010
MD09G1237900110,536.416.0545.2788.87−0.4161010
MD11G1261900110,608.47647.391.19−0.3821010
MD13G1049200117,260.816.5152.7687.87−0.4331920
MD16G1050300117,245.636.4654.3287.39−0.452020
MD17G1230800110,225.085.9745.7888.87−0.3921010
Prupe.1G302700117,827.736.8152.2987.72−0.4132020
Prupe.3G11830081,846.165.5747.3988.74−0.42688
Prupe.4G166400110,062.786.0447.2991.98−0.3871010
gene_214.28110,219.735.746.7489.56−0.3871010
gene_117.42110,245.925.8147.0789.46−0.3841010
gene_89.20110,216.865.8946.9989.56−0.3921010
gene_201.40110,217.85.847.1989.56−0.3911010
gene_175.31109,447.165.8645.3488.92−0.3911010
gene_78.22109,389.185.7644.7289.52−0.3671010
gene_170.27109,374.065.8645.0788.71−0.3891010
gene_121.5109,658.56.4944.8987.61−0.4231010
gene_206.27109,402.085.8645.1188.71−0.3891010
pycom03g18910110,058.826.0446.8491.88−0.376910
pycom09g15640110,386.366.2445.189.27−0.411910
pycom11g23090122,259.066.1549.4391.99−0.3231011
pycom16g04410129,170.717.854.3489.66−0.42020
pycom17g23400108,062.615.7944.3690.46−0.3541010
MW: Molecular weight of the amino acid sequence; pI: Theoretical isoelectric point; GRAVY: Grand average of hydropathicity.
Table 4. The Ka (nonsynonymous), Ks (synonymous), and Ka/Ks ratio of duplicated PEPC genes in five Rosaceae species.
Table 4. The Ka (nonsynonymous), Ks (synonymous), and Ka/Ks ratio of duplicated PEPC genes in five Rosaceae species.
Gene 1Gene 2KaKsKa/Ks (ω)SelectionDuplication Mode
MD09G1237900pycom09g156400.0049760.0361380.137707PurifyingSegmental
MD17G1230800pycom17g234000.0110430.0565520.195266PurifyingSegmental
gene_170.27gene_206.274.54E-040.0300530.015101PurifyingSegmental
gene_214.28gene_117.420.0036190.0239420.151148PurifyingSegmental
gene_89.20gene_201.404.52E-040.0014660.308365PurifyingSegmental
EVM0016068.1pycom03g189100.0036290.0997710.036372PurifyingSegmental
EVM0023388.1pycom11g230900.0038550.0648430.059445PurifyingSegmental
MD16G1050300pycom16g044100.0037840.0518720.07294PurifyingSegmental
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhi, C.; Ali, M.M.; Alam, S.M.; Gull, S.; Ali, S.; Yousef, A.F.; Ahmed, M.A.A.; Ma, S.; Chen, F. Genome-Wide In Silico Analysis and Expression Profiling of Phosphoenolpyruvate Carboxylase Genes in Loquat, Apple, Peach, Strawberry and Pear. Agronomy 2022, 12, 25. https://doi.org/10.3390/agronomy12010025

AMA Style

Zhi C, Ali MM, Alam SM, Gull S, Ali S, Yousef AF, Ahmed MAA, Ma S, Chen F. Genome-Wide In Silico Analysis and Expression Profiling of Phosphoenolpyruvate Carboxylase Genes in Loquat, Apple, Peach, Strawberry and Pear. Agronomy. 2022; 12(1):25. https://doi.org/10.3390/agronomy12010025

Chicago/Turabian Style

Zhi, Cao, Muhammad Moaaz Ali, Shariq Mahmood Alam, Shaista Gull, Sajid Ali, Ahmed F. Yousef, Mohamed A. A. Ahmed, Songfeng Ma, and Faxing Chen. 2022. "Genome-Wide In Silico Analysis and Expression Profiling of Phosphoenolpyruvate Carboxylase Genes in Loquat, Apple, Peach, Strawberry and Pear" Agronomy 12, no. 1: 25. https://doi.org/10.3390/agronomy12010025

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