Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Transcriptome analysis of bread wheat leaves in response to salt stress

  • Nazanin Amirbakhtiar,

    Roles Data curation, Formal analysis, Investigation, Methodology, Validation, Visualization, Writing – original draft

    Affiliations Plant Production and Genetic Engineering Department, Faculty of Agriculture, Lorestan University, Khorramabad, Iran, National Plant Gene Bank of Iran, Seed and Plant Improvement Institute (SPII), Agricultural Research, Education and Extension Organization (AREEO), Karaj, Iran

  • Ahmad Ismaili,

    Roles Supervision, Writing – review & editing

    Affiliation Plant Production and Genetic Engineering Department, Faculty of Agriculture, Lorestan University, Khorramabad, Iran

  • Mohammad-Reza Ghaffari,

    Roles Formal analysis

    Affiliation Department of Systems Biology, Agricultural Biotechnology Research Institute of Iran (ABRII), Agricultural Research, Education and Extension Organization (AREEO), Karaj, Iran

  • Raheleh Mirdar Mansuri,

    Roles Formal analysis, Visualization

    Affiliation Department of Systems Biology, Agricultural Biotechnology Research Institute of Iran (ABRII), Agricultural Research, Education and Extension Organization (AREEO), Karaj, Iran

  • Sepideh Sanjari,

    Roles Writing – review & editing

    Affiliation Department of Systems Biology, Agricultural Biotechnology Research Institute of Iran (ABRII), Agricultural Research, Education and Extension Organization (AREEO), Karaj, Iran

  • Zahra-Sadat Shobbar

    Roles Conceptualization, Funding acquisition, Methodology, Project administration, Supervision, Validation, Writing – review & editing

    shobbar@abrii.ac.ir

    Affiliation Department of Systems Biology, Agricultural Biotechnology Research Institute of Iran (ABRII), Agricultural Research, Education and Extension Organization (AREEO), Karaj, Iran

Abstract

Salinity is one of the main abiotic stresses limiting crop productivity. In the current study, the transcriptome of wheat leaves in an Iranian salt-tolerant cultivar (Arg) was investigated in response to salinity stress to identify salinity stress-responsive genes and mechanisms. More than 114 million reads were generated from leaf tissues by the Illumina HiSeq 2500 platform. An amount of 81.9% to 85.7% of reads could be mapped to the wheat reference genome for different samples. The data analysis led to the identification of 98819 genes, including 26700 novel transcripts. A total of 4290 differentially expressed genes (DEGs) were recognized, comprising 2346 up-regulated genes and 1944 down-regulated genes. Clustering of the DEGs utilizing Kyoto Encyclopedia of Genes and Genomes (KEGG) indicated that transcripts associated with phenylpropanoid biosynthesis, transporters, transcription factors, hormone signal transduction, glycosyltransferases, exosome, and MAPK signaling might be involved in salt tolerance. The expression patterns of nine DEGs were investigated by quantitative real-time PCR in Arg and Moghan3 as the salt-tolerant and susceptible cultivars, respectively. The obtained results were consistent with changes in transcript abundance found by RNA-sequencing in the tolerant cultivar. The results presented here could be utilized for salt tolerance enhancement in wheat through genetic engineering or molecular breeding.

Introduction

Plant growth and productivity are seriously threatened by abiotic stresses [1]. Among abiotic stresses, salt stress is considered a serious threat to crop yield worldwide [2]. Wheat is the third most important cereal crop in the world [3], and salinity levels of 6–8 dsm-1 cause to decline wheat yield [4]. A practical approach to minimize salinity’s impact on global wheat production is to enhance salt tolerance in wheat cultivars.

Ion toxicity, nutrient limitations, and oxidative and osmotic stresses are the adverse effects of salinity stress on crops [5]. Plant salt tolerance is achieved through integrated responses at physiological, cellular, molecular, and metabolic levels. At the molecular level, genes coding for transcription factors, ion transporters, protein kinases, and osmolytes are involved in salt tolerance [6, 7]. Some signaling pathways, including plant hormones, salt overly sensitive (SOS), calcium, mitogen-activated protein kinase (MAPK), and proline metabolism, play critical roles in salt stress tolerance, as well [812]. Salinity tolerance, as a quantitative trait, is under the control of multiple genes [13]. Thus, it is necessary to discover key components underlying the salt tolerance network to improve it through genetic engineering.

RNA-sequencing provides a much more accurate measurement of transcript levels and isoforms compared to other transcriptomic methods [14]. A few studies applied RNA-sequencing technology to inspect the transcriptome profile of shoots under salt conditions in bread wheat in recent years. Comparing the shoot expression profiling in a salinity tolerant mutant of Triticum aestivum L and its susceptible wild type exposed to salt stress resulted in discovering some salt tolerance involved genes like polyamine oxidase, arginine decarboxylase, and hormones-associated genes, which were further up-regulated in the mutant. They also succeeded in finding "Butanoate metabolism" as a novel salt stress-response pathway and indicated that oxidation-reduction (redox) homeostasis was essential for salt tolerance [15]. In another study, Mahajan et al. (2017) performed RNA-sequencing to prepare transcriptome profiling of flag leaves in the salt-tolerant cultivar of Kharcha in response to salt stress. They indicated that the up-regulated genes under salt stress were related to different biological processes like ion transport, phytohormones signaling, signal transduction, osmoregulation, flavonoid biosynthesis, and ROS homeostasis [16]. Luo et al. (2019) compared young and old leaf transcriptome of a salt-tolerant bread wheat cultivar and a high-yielding cultivar with lower salt tolerance in response to salinity. They found that the polyunsaturated fatty acid (PUFA) metabolism was the most significant term/pathway in the salt-tolerant wheat cultivar according to the enriched GO terms and the Kyoto Encyclopedia of Genes and Genomes (KEGG) pathways analysis. They suggested that PUFAs could promote salt tolerance through the photosynthetic system and JA-related pathways [17].

Zhang et al. (2016) compared root transcriptome response of a salt-tolerant and a salt-sensitive cultivar and identified two NAC transcription factors (TFs), a MYB TF (homologous to AtMYB33), a gene positively associated with root hair development (Ta.RSL4) and a gene coding for histone-lysine N-methyl transferase (homologous to Arabidopsis AtSDG16) as essential genes for salinity tolerance in Triticum aestivum [18]. Amirbakhtiar et al. (2019) evaluated transcriptome profile of a salt tolerant bread wheat cultivar in response to salinity. They identified pathways related to transporters, phenylpropanoid biosynthesis, TFs, glycosyltransferases, glutathione metabolism and plant hormone signal transduction as the most important pathways involved in salt stress response [19]. Mahajan et al. (2020) sequenced root transcriptome of a salt tolerant wheat cultivar at anthesis stage. They showed that genes involved in ROS homeostasis, ion transport, signal transduction, ABA biosynthesis and osmoregulation up-regulated in response to salt stress. They also indicated that genes coding for expansin, dehydrins, xyloglucan endotransglucosylase and peroxidases, engaged in root growth improvement, up-regulated under salt stress [20]. Despite the valuable insight discovered by recent researches about the cellular and molecular mechanisms engaged in salinity stress response and tolerance in bread wheat, many aspects are still uncovered. In the current study, considering Iran as one of the origin lands of Triticum aestivum and its wild lineages [2124], deep transcriptome sequencing was used for an Iranian salt-tolerant wheat cultivar (Arg) under normal and salinity conditions to complement the insights regarding molecular mechanisms involved in bread wheat salt-tolerance. We succeeded in providing a panel of the regulatory mechanisms at transcriptional level in the leaves of the salt-tolerant wheat cultivar (Arg) under salinity stress by identifying all differentially expressed genes, novel salt-responsive genes, and diverse metabolic pathways involved in response to salinity stress.

Materials and methods

Wheat culture conditions and salinity treatment

Seeds of the bread wheat salt-tolerant (Arg) and salt-sensitive (Moghan3) genotypes were kindly supplied by Seed and Plant Improvement Institute (SPII), Karaj, Iran. After surface sterilizing the seeds in 1% sodium hypochlorite, they were grown on moist filter paper for approximately 72 hours. The uniform germinated seeds were then selected and transferred to half-strength Hoagland’s culture solution in the greenhouse. NaCl solution (150 mM) was used to treat the three-week old plants for 12 and 72 hours. The leaves of the control and salt-stressed plants were collected separately. The number of biological replicates was four, and each replicate included three independent plants. The samples were frozen instantly in liquid nitrogen and kept at -80°C.

Measurements of Na+ and K+ concentrations

The leaves of the plants exposed to salt stress for 72 hr were harvested and dried at 70°C for 48 hr. Flame spectrophotometry method was used to measure Na+ and K+ concentrations [25].

RNA isolation and Illumina sequencing

RNA was extracted from wheat leaves with four biological replicates under normal and salinity conditions utilizing RNeasy Plant Mini Kit (Qiagen). Equal quantities of the total RNA of every two biological replicates of Arg cultivar were pooled together to prepare two replicates for the RNA sequencing. Agarose gel electrophoresis, nanodrop, and Agilent Bioanalyzer 2100 system (Agilent Technologies Co. Ltd., Beijing, China) were used to control the quantity, quality, and integrity of RNA. The RIN value of the samples used for sequencing was more than or equal to 6.9. cDNA library preparation and sequencing were performed using an Illumina Hiseq 2500 platform at the Novogene Bioinformatics Institute (Beijing, China). The generated reads were paired-end with 150bp size. After sequencing, adapter-containing reads, poly-N-containing reads (N > 10%), and low quality (Qscore< = 5) base-containing reads were eliminated.

Read mapping and reference-based assembly

The FastQC toolkit was used to assess the quality of raw fastq data. Tophat software with standard parameters was utilized to map the high-quality reads to the wheat reference genome (ftp://ftp.ensemblgenomes.org/pub/release-34/plants/fasta/triticum_aestivum/dna/). Cufflinks with default settings was applied to create assembly based on the Tophat mapping files. The individual assemblies were merged using Cuffmerge with default options, and a final assembly was produced. The novel transcripts were recognized via Cuffmerge [26]. For functional annotation, Blast2GO via BlastX with 1e-3 as an e-value cut-off was used to align all the transcripts against NCBI’s non-redundant protein database. This software was also utilized to obtain the gene ontology (GO) terms of transcripts with a p-value cut-off of 0.05.

Differential gene expression analysis

The gene/transcript expression was calculated using the FPKM method. Significantly differentially expressed genes (DEGs) were identified applying Cuffdiff provided in the Cufflinks package based on |log2 fold change| ≥ 1 and Q-value cut off ≤ 0.01.

Functional annotation and pathway enrichment analysis of DEGs

The Online KEGG Automatic Annotation Server (KAAS), http://www.genome.jp/kegg/kaas [27], was utilized to identify metabolic pathways in which DEGs were engaged. The pathway analysis of the DEGs was performed applying Mapman (version 3.5.1; http://mapman.gabipd.org/web/guest) [28] with a p-value limit of ≤ 0.05.

Confirmation of RNA-sequencing results by Real-Time PCR analysis

The extracted RNAs from three biological replicates were reverse-transcribed with qScript cDNA Synthesis Kit (Quantabio, USA) for first-strand cDNA synthesis based on the manufacturer’s protocol. LightCycler® 96 Real-Time PCR System (Roche Life Science, Germany) and SYBR Premix EX TaqII (Takara Bio Inc, Japan) were utilized to perform Quantitative Real-Time PCR. Actin was used in place of an internal control gene in the RT-PCR experiment to normalize the gene expression value [17, 19, 29, 30]. Genes with log2 fold change ≥ 1 or log2 fold change ≤ -1 were considered as significant DEGs.

Specific primers for the selected genes are listed in S1 Table. Each gene’s transcript level in every genotype under control conditions was utilized as the calibrator for each time point. The 2−ΔΔCt procedure [31] was applied to calculate the relative expressions of the selected genes.

Results

Na+ and K+ content

The salt stress led to a significant increase in Na+ content and a significant decrease in the K+/Na+ ratio in the roots and leaves of both genotypes. A significant increase was observed in the K+ content of the roots in both genotypes under salinity stress, while no significant change was observed in their leaves. As expected, less Na+ content and more K+/Na+ ratio were observed in the leaves of the tolerant cultivar (Arg) compared to those of the susceptible cultivar (Moghan 3) under salinity stress. A higher accumulation and maintenance of Na+ ion in the roots than in the leaves may act as a tolerance mechanism by maintaining the essential osmotic potential for absorbing water into the root and limiting Na+ ion flux into the shoot [32]. The higher accumulation of Na+ ion in the root than in the leaf blade is a salt tolerance mechanism in the grasses, limiting the transfer of sodium ions into photosynthetic cells and active meristem tissues [33]. The higher amount of Na+ in the roots of Arg compared to those of Moghan 3 and the less amount of Na+ in the leaves of Arg compared to those of Moghan 3 under salinity stress indicate that root can be considered as an important barrier to prevent the transfer of Na+ to the leaves in Arg (S1 and S2 Figs). This result is consistent with the results obtained by Davenport et al. (2007) [34].

Sequencing statistics and reference-based analysis

A total of 114.29 million raw reads were obtained by transcriptome sequencing. After removing adapters and low-quality reads, a total of 112.6 million clean reads were produced, while more than 88.1% of clean reads had Phred-like quality scores at the Q30 level (Table 1). Accession numbers of SRR7975953, SRR7968059, SRR7968053, and SRR7920873 at the SRA (Sequence Read Achieve) of NCBI include the raw transcriptome reads generated in the current study.

Mapping the cleaned high-quality reads to the wheat reference genome (ftp://ftp.ensemblgenomes.org/pub/release-34/plants/fasta/triticum_aestivum/dna/) showed that around 81.9%-85.7% of the reads were mapped successfully to the wheat reference sequence, including 72.8%-79.3% uniquely matched (Table 2). The aligned reads were assembled using cufflinks while 187003 and 98819 transcript isoforms and genes were identified, respectively.

thumbnail
Table 2. Results of mapping reads to the reference genes.

https://doi.org/10.1371/journal.pone.0254189.t002

Exploration of novel transcripts via mRNA sequencing

The discovery of novel genes/transcripts is one of the main benefits of RNA- sequencing experiments [14, 35, 36]. The current study identified 27800 and 16339 novel transcript isoforms and genes, respectively. Conforming with other crops, including rice and maize [37, 38], the mean length of the novel transcripts (1609 bp) was less than that of the annotated transcripts (2304 bp). Based on the gene ontology analysis results, a putative function was assigned to more than 53.1% of the novel transcripts.

The GO analysis of the novel transcripts revealed that these novel genes would play a role in biological processes, including stimulus responses, localization, biogenesis, and biological regulation (S2 Table). Molecular function classification showed that the novel transcripts were enriched in some terms such as transferase, oxidoreductase, catalytic, and hydrolase activities; small molecule and ion binding; carbohydrate derivative binding; and organic cyclic and heterocyclic compound binding (S3 Table). The novel transcripts were also enriched in some cellular component categories, such as intracellular membrane-bounded organelle, an integral component of membrane, cytoplasm, mitochondrion, nucleus, and chloroplast (S4 Table).

Identification of DEGs involved in salt stress response

In total, 4290 genes were differentially regulated under salinity stress, of which 2346 and 1944 were up- and down-regulated genes, respectively (S5 Table). Among the DEGs, 110 and 98 genes were exclusively expressed under salt-stressed and normal conditions, respectively (S3A Fig). Some essential genes engaged in response to abiotic stresses, including LEA proteins, dehydrins, bHLH transcription factor, phosphatase 2C, peroxidase, and calcium-transporting ATPase plasma membrane-type (S6 Table), were observed among the genes exclusively expressed under salinity stress. Surveying the fold change distribution of the DEGs indicated that the most and least number of the genes had a fold change of 2–3 and 6–7, respectively (S3B Fig).

GO categorization of DEGs

The gene ontology analysis indicated that GO terms were assigned to 3594 out of 4290 genes. In biological process classification, the majority of the genes were involved in the metabolic process (23.4%) followed by cellular process (20.4%), single-organism process (17.4%), stimulus-response (6.7%), and biological regulation (6.6%). In molecular function categorization, catalytic activity (45.9%), binding (37.7%) and transporter activity (7.5%) were the most frequent terms and in the cellular component category, cell (22%), cell part (21.9%), and membrane (16.1%) were the most dominant terms (Fig 1).

thumbnail
Fig 1. GO classification of the DEGs in Arg cultivar.

Percentage of the transcripts in cellular component, molecular function, and biological process classifications are represented.

https://doi.org/10.1371/journal.pone.0254189.g001

Functional identification of novel DEGs

Comparing the functional annotation of the novel salt responsive genes against NCBI’s non-redundant (nr) protein database utilizing the Blast2GO tool indicated that 320 (70%) out of the 457 novel DEGs were aligned to the NR protein database. In contrast, the rest of the genes (30%) displayed no homology to database sequences. The GO classification showed that 230 novel DEGs (50.3%) were assigned to GO terms, and 205 novel DEGs (44.9%) were grouped in significant GO terms (S4A Fig). In the cellular component category, cell part, cell, and membrane were the prevailing clusters. However, the top three classes were catalytic activity, binding, and transporter activity concerning the molecular function. In biological process categorization, most of the genes were involved in the metabolic and cellular processes followed by regulating the single-organism process, regulating the biological process and responding to a stimulus (S4B Fig).

The genes coding for Phosphatase 2C [39], Delta-1-pyrroline-5-carboxylate synthase [40], MYB transcription factor [41], Sodium/Calcium exchanger [42], Ribulose bisphosphate carboxylase large chain [43], Late Embryogenesis Abundant protein [44], Glutahione S-trasferase [45], and cytochrome P450 monooxygenase [46] (S7 Table) with potential roles in salt stress response were observed among the novel DEGs.

KEGG pathway classification of DEGs

In an attempt to map DEGs to various biological pathways, a single-directional BLAST search showed that 1503 out of the 4290 DEGs were categorized into 227 KEGG pathways, located in the five chief KEGG classes (Fig 2A). Pathways relating to phenylpropanoid biosynthesis, transporters, transcription factors, plant hormone signal transduction, glycosyltransferases, exosome, MAPK signaling pathway, peptidases, cytochrome P450, and sucrose and starch metabolism included the highest number of DEGs (Fig 2B, S8 Table). The involvement of these pathways in environmental stress response was confirmed in previous reports [29, 47, 48].

thumbnail
Fig 2. KEGG categorization of the DEGs.

(A) Categorization of the DEGs into five chief KEGG classes. (B) The top 10 pathways with the highest gene number.

https://doi.org/10.1371/journal.pone.0254189.g002

The phenylpropanoid pathway with the highest gene number is responsible for synthesizing diverse secondary metabolites in plants such as lignin, flavonoids, and coumarins playing roles in developmental and stress–associated processes [49, 50]. In the first step of this pathway, cinnamic acid is synthesized from phenylalanine by the rate-limiting enzyme of phenylalanine ammonia-lyase (PAL) [51]. In this study, 29 up-regulated DEGs coding for PAL were mapped in this pathway.

Plants utilize deposition of lignin or modification of monomeric lignin composition in the cell wall to defeat salinity stress [52]. In the present study, the up-regulated DEGs coding for shikimate hydroxycinnamoyl transferase, cinnamoyl-CoA reductase, and caffeic acid 3-O-methyltransferase, which were all involved in lignification, were mapped in the phenylpropanoid pathway, while their over-expression was also reported under salinity stress in prior researches [5355].

Functional analysis of salt-regulated genes using Mapman

The putative function of the salt-regulated genes was searched utilizing Mapman to visualize salt-induced alterations in diverse metabolic processes. Metabolic pathway overviews based on the results of mapping salt-responsive genes indicated that photosynthesis and cell wall biosynthesis pathways were among the enriched pathways (Fig 3, S9 Table). Most of the genes encoded chlorophyll-binding proteins in the photosynthesis pathway, showing down-regulation under salt stress. The decrease in photosynthesis efficiency under abiotic stresses was reported in previous studies [56, 57].

thumbnail
Fig 3. Metabolic pathway overview of the DEGs in Arg cultivar under salinity stress utilizing Mapman.

The down- and up-regulated genes are shown in red and blue, respectively.

https://doi.org/10.1371/journal.pone.0254189.g003

Mapping the DEGs to the cellular pathways indicated that the misc pathway, including genes regarding abiotic stress-related various enzyme families, was enriched under salt stress (S5 Fig., S9 Table). Most of the misc pathway genes are Germin-like proteins (GLPs), which code for ubiquitous plant glycoproteins and belong to the cupin superfamily. One of the main roles of the proteins mentioned above is triggering the abiotic stress-tolerance in many plant species. Li et al. (2016) revealed that GLP transcripts were plentiful after treatment with high salinity, PEG6000, abscisic acid, and methyl viologen in soybean. Arabidopsis plants overexpressing a GLP from soybean indicated enhanced drought, salt, and oxidative tolerance [58]. Furthermore, Arabidopsis transgenic plants, which overexpressed genes encoding peanut GLPs, showed increased tolerance to salinity. Complementary studies also showed that PR-defense genes and antioxidant coding genes, which can increase salt tolerance, showed up-regulation in transgenic plants [59].

Investigating the secondary metabolite pathways revealed that the genes playing roles in terpenoid, lignin, phenols, isoflavonoid, and wax metabolic pathways were significantly enriched under salt stress (S6 Fig, S9 Table). Furthermore, the stress response pathways showed that the transcription regulators and peroxidases and the genes relating to brassinosteroid signaling pathways were enriched in Arg cultivar under salt stress (S7 Fig., S9 Table).

Confirmation of gene expression patterns by qRT-PCR

The expression pattern of nine candidate salt-regulated genes was examined by qRT-PCR to validate the RNA-sequencing results (Fig 4). The high consistency between qRT-PCR and RNA sequencing results was observed (R2 = 0.98), confirming the identified DEGs in the present research. The candidate genes’ expression profile was assessed in the two salt contrasting genotypes to obtain further insight. Based on the obtained results, Ta.bHLH35, Ta.CIPK23, and Ta.P5CS were up-regulated significantly in the tolerant cultivar after 12 hr of salt stress, while the increase in the expression of these genes was much less in the sensitive cultivar than in the tolerant cultivar and was not significant (Fig 4A, 4B and 4F). Ta.ERF061 showed significant up-regulation after 12 hr of exposure to salt stress in both cultivars. However, at the time point of 72 hr, a more severe decrease in expression was observed in the tolerant cultivar compared to the sensitive cultivar, which can be related to the quicker response of the tolerant cultivar to salt stress. (Fig 4C). For Ta.HSFB1 at the time point of 12 hr, while the tolerant cultivar indicated up-regulation, the sensitive cultivar showed down-regulation (Fig 4D). For another transcription factor, Ta.NAC, significant up-regulation was observed in both cultivars at the two-time points, and there was no significant difference between the cultivars (Fig 4E). For the gene encoding salt response protein, while the tolerant cultivar showed up-regulation at the two-time points, the sensitive cultivar indicated up-regulation after 12 hr of exposure to salt treatment and down-regulation after 72 hr of exposure to salinity (Fig 4G). Furthermore, for the gene coding for RUBISCO small chain involved in photosynthesis, a more severe decrease was observed in the tolerant cultivar compared to the sensitive cultivar (Fig 4H). The decrease in this gene expression may be due to the need to change the energy flow from the biosynthesis of photosynthesis-engaged macromolecules toward respiratory paths to supply the energy needed to overcome stress. A regulatory gene called phosphoglycerate kinase, with a possible negative effect on stress tolerance, showed more severe down-regulation in the tolerant cultivar than in the sensitive cultivar after 12 hr of exposure to salt stress (Fig 4I).

thumbnail
Fig 4.

Validation of the candidate genes by qRT-PCR including bHLH transcription factor 35 (A), calcineurin B-like protein (CBL)-interacting protein kinase 23 (B), ethylene responsive factor 061 (C), heat shock transcription factor B1 (D), NAC transcription factor (E), pyrroline-5-carboxylate synthetase (F), salt response protein (G), Ribuluse biphosphate carboxylase small chain (H), and Phosphoglycerate Kinase (I). Refer to S1 Table to find the gene ensemble IDs.

https://doi.org/10.1371/journal.pone.0254189.g004

Discussion

Next-generation sequencing technologies with the ability to characterize transcriptome profiles of different organisms under different conditions can reveal the molecular basis of salt stress response in plants. In general, genes engaged in salt stress response can be divided into three classes, comprising stress sensing and signaling-related genes, transcriptional regulators, and salinity-stress associated genes [60].

Signal transduction paths play crucial roles in the response of plants to different stresses. Variation in cytosol’s calcium concentration is one of the early responses to various stimuli, and calcium transporting elements actively maintain this flux and homeostasis [61]. In the current research, two genes encoding calcium-transporting ATPases were up-regulated under salt stress. One of them is a novel gene (represented as Ta.ACA7 in Fig 5 and S10 Table), expressed only under salt stress. Orthologous of the forenamed gene in rice, Os.ACA7 (Os10g0418100), is activated by calmodulin (CaM) [61]. Ca2+-ATPases are involved in maintaining Ca2+ homeostasis [61], and the up-regulation of them has been observed in different plant species, including tomato, tobacco, Arabidopsis, and soybean, in response to salinity stress [6265]. Genes coding for glutamate receptors (GLRs), known as non-selective cation channels that can be engaged in Ca+2 transport [61], were also observed among the DEGs in this study (Fig 5, S10 Table). Glutamate receptors are responsive to abiotic stresses based on the previous reports [66, 67]. After an increase in the Ca2+ concentration under salinity, CBL-interacting protein kinases (CIPKs) [68] with the ability to transduce the signal to downstream protein activity and gene transcription may become activated [69]. Among the DEGs, 13 genes coding for CIPKs were discovered. One gene coding for CaM was up-regulated in response to salinity (Fig 5, S10 Table). CaM, known as a Ca2+-sensing protein, is involved in the transduction of Ca2+signals. Conformational changes occur in CaM after interacting with Ca2+, and then, CaM influences the activities of the proteins which bind to it. Several CaM-binding proteins are engaged in plant responses to salinity stress, showing that CaM plays a central role in stress adaptation in plants [70]. A differentially expressed CaM-binding gene in the current study is Ta.MLO (Fig 5, S10 Table), encoding a plant-specific seven-transmembrane domain protein. A study on the MLO family in rice concluded that environmental stresses might provoke alteration in the H2O2 level via interaction between MLO and CaM. The resulting H2O2 might act as a messenger to stimulate the responsive genes’ expression to acclimatize to the stress [71]. CBP60, a CaM-binding transcription factor, was up-regulated under salinity stress in the present study (Fig 5, S10 Table). It has been shown that the overexpression of CBP60 (At5g26920) in Arabidopsis resulted in increased defense response, hypersensitivity to ABA, and drought tolerance, possibly through activating salicylic acid accumulation [72].

thumbnail
Fig 5. The model proposed for a response to salinity stress in Arg cultivar.

Dark blue and purple colors were used to exhibit stress sensing and signaling-related genes and transcription factors, respectively. Light blue, light green, and pink colors were utilized to depict genes involved in the reaction to osmotic, ionic, and oxidative stresses caused by salinity, respectively.

https://doi.org/10.1371/journal.pone.0254189.g005

Previous reports indicated that the overexpression of GDSL esterase lipases (GLIPs) could release fatty acids acting as hormone signal transduction molecules [73]. It has also been reported that excessive GLIPs exhibited improved salinity stress tolerance in Oryza sativa and Arabidopsis thaliana [74, 75]. Five genes coding for Ta.GLIPs were up-regulated under salinity stress in the present study (Fig 5, S10 Table).

Receptor-like kinases (RLKs), as the largest gene family in plants, play crucial roles in signaling networks [76]. Wall-associated kinases (WAKs), as a subfamily of RLKs, function as a signaling linker between the cytoplasm and the extracellular region [77]. It has been reported that WAKs are engaged in regulating plant adaptation to abiotic stresses. Arabidopsis plants overexpressing AtWAK1 showed increased aluminum tolerance [78], and Arabidopsis plants with the impaired expression of AtWAKL4 indicated more hypersensitivity to excessive Na+, K+, Cu+2, and Zn+2 [79]. In the current study, six genes coding for WAKs were up-regulated under salt stress (Fig 5, S10 Table). LecRLKs, another subfamily of RLKs, can be engaged in salinity tolerance, including a plasma membrane-localized LecRLK from Pisum sativum. Tobacco plants overexpressing PsLecRLK showed enhanced salt tolerance by increasing ROS scavenging activity and activating water channels, leading to reduced ROS accumulation and enhanced water uptake [80]. In the present research, three genes coding for LecRLKs were up-regulated in response to salinity stress (Fig 5, S10 Table).

Many TFs were observed among the DEGs, indicating their crucial roles in salt stress response. They regulate the expression of downstream genes liable for salinity stress tolerance in plants. ERFs, bZIPs, Zn-fingers, NACs, MYBs, and WRKYs were found among the differentially expressed TFs, and some of them were discussed here.

MYB TFs are known as one of the largest and most diverse families of TFs in plants [81, 82]. The involvement of MYB TFs in salt tolerance has been reported in previous studies [83, 84]. Twenty-seven genes coding for MYBs were observed among the DEGs in the present research (Fig 5, S10 Table).

Plant basic leucine zipper (bZIP) TFs are involved in regulating abiotic stress signaling pathways mediated by abscisic acid (ABA) in plants [85]. Tomato SlbZIP38 regulates drought and salinity tolerance negatively via regulating ABA signaling [86]. The overexpression of cotton GhABF2, encoding a bZIP TF, significantly increased tolerance to drought and salinity in Arabidopsis and cotton [87]. Two genes coding for bZIPs were differentially expressed in the current study (Fig 5, S10 Table).

Four families of zinc finger proteins (ZFP), including C2H2, CCCH, C3HC4, and C4, have crucial roles in regulating phytohormone and stress response in plants [88]. The engagement of zinc finger TFs in salt tolerance has been reported in previous studies. Transgenic rice overexpressing OsZFP213 indicated improved salt tolerance via enhancing ROS scavenging ability [89]. Tobacco plants overexpressing GhZFP1, a CCCH-type zinc finger protein from cotton, showed increased tolerance to salinity stress and resistance to Rhizoctonia solani [90]. In the present study, around 17 differentially expressed zinc finger TFs were identified (Fig 5, S10 Table).

TIFY proteins are engaged in regulating many plant processes, including response to stresses. JAZ proteins, working as the jasmonic acid signaling pathway’s key regulators, are the best-characterized sub-group of TIFY proteins. Two genes coding for TIFY were found among the DEGs (Fig 5, S10 Table). The involvement of TIFY TFs in wheat salt tolerance was reported in a previous study [91].

In the present study, 31 genes coding for WRKY TFs were differentially expressed under salt stress, among which only one gene showed down-regulation (S10 Table). WRKY TFs are engaged in increasing salinity tolerance in plants via regulating stomatal conductance, ROS levels, and auxin and ABA signaling [92].

In addition, 28 NAC domain-containing genes were differentially regulated under salt stress in the current study, among which only four genes were down-regulated (Fig 5, S10 Table). NAC TFs take part in complicated signaling networks related to stress response in plants [93]. Rice OsNAC022, induced by drought, high salinity, and ABA, enhanced drought and salinity stress tolerance via regulating an ABA-dependent pathway in transgenic plants [94]. TsNAC1 from a halophyte called Thellungiella halophila targeted positive ion transportation regulators and improved salt tolerance in both T. halophila and Arabidopsis [95].

Some ethylene response factors (ERFs) bind to dehydration-responsive elements, function as a central regulatory hub, and incorporate ethylene, abscisic acid, jasmonate, and redox signaling in abiotic stress response in plants [96]. In the present study, 15 genes relating to ERF transcription factors were differentially expressed under salinity stress (S10 Table). Previous studies have shown that the overexpression of ERFs by increasing salt-responsive genes’ expression leads to salt tolerance in plants [97, 98].

We also identified transcripts encoding homeodomain-containing transcription factors (HOX) 7 and 22, which were significantly up-regulated under salt stress (Fig 5, S10 Table). According to the previous reports, the HOX family members as regulators of plant growth and development were remarkably enriched in NaCl-induced transcripts in Oryza sativa [99, 100]. It has also been reported that ABA, GA, SA, and auxin enhance the transcript levels of some HOXs [99].

A high ratio of cytosolic K+/Na+ is necessary to keep ionic homeostasis under stress and increases salinity tolerance in wheat (Oyiga et al., 2016). Plants utilize various methods at different levels to retain this ratio in the cytosol. One selected approach in plants is sending out Na+ from the roots. SOS1, a plasma membrane Na+/H+ antiporter, drives Na+ out from the root. Evaluating the transcriptome response of the root in Arg cultivar under salt stress showed the up-regulation of SOS1 under salinity stress [19]. Sustaining a high ratio of K+/Na+ in the cell cytoplasm can also be performed by sequestrating Na+ into the vacuoles of root cells done by the tonoplast Na+/H+ antiporter (NHX1). Hexokinase1 phosphorylates NHX1 and increases its stability [101]. Transcriptome response analysis of the root in Arg cultivar under salt stress showed a significant increase in the expression of hexokinase1 [19]. Evaluating the transcriptome response of leaves to salt stress showed that 22 genes involved in transporting sodium, potassium, or both significantly responded to salt stress. Among the up-regulated sodium transporters in leaves, a gene indicated severe up-regulation under salt stress (represented as Ta.HKT1 in Fig 5 and S10 Table). An orthologue of the mentioned gene in Arabidopsis, At4g10310, encodes the sodium transporter HKT1. This transporter shows a central role in plant tolerance to salinity. It loads Na+ into the phloem sap in shoots and unloads it in roots, leading to eliminating large quantities of Na+ from the shoot [102]. The other sodium transporter is represented as Ta.HKT4 in Fig 5 and S10 Table. The orthologue of the forenamed gene in rice is Os04g0607500 that encodes for the cation transporter HKT4. OsHKT4 acts as a low-affinity sodium transporter and is possibly engaged in regulating K+/Na+ homeostasis [103]. Seven genes, coding for potassium transporters, were differentially expressed under salinity stress. Among the potassium transporters, we can refer to a gene represented as Ta.KT in Fig 5 and S10 Table. The orthologue in Arabidopsis, At2g30070, encodes AtKT1 and acts as a high-affinity potassium transporter [104].

In order to deal with the osmotic stress caused by salinity, the genes encoding for LEA proteins (Ta.LEA) and dehydrins (Ta.Dhn) as well as the genes involved in biosynthesis of organic osmolytes like proline (pyrroline-5-carboxylate synthase; Ta.P5CS) and glycine betaine (betaine aldehyde dehydrogenase; Ta.BADH) were up-regulated under salinity stress (Fig 5, S10 Table).

Furthermore, peroxiredoxin (Ta.PRX), peroxidases (Ta.POX), Cytochrome P450 (Ta.CYP450), and Glutathione-S- transferases (Ta.GST) were differentially regulated in reaction to oxidative stress arising from salinity (Fig 5, S10 Table). Cytochrome P450 (CYP) includes a superfamily of heme-containing proteins that take part in redox homeostasis and numerous biosynthetic pathways [105]. Eight genes belonging to the CYP71 family, as the largest CYP family in plants, were up-regulated in the present research. This result is in line with those obtained by another study on transcriptome response of wheat leaf to salinity stress [16], indicating the CYP71 family might play a role in salinity stress tolerance.

Supporting information

S1 Fig. Comparison of means for changes in Na+ content (a), K+ content (b) and K+/Na+ ratio (c) in the leaves of Arg and Moghan3 under the normal and salt stressed conditions at the probability level of 5%.

https://doi.org/10.1371/journal.pone.0254189.s001

(DOCX)

S2 Fig. Comparison of means for changes in Na+ content (a), K+ content (b) and K+/Na+ ratio (c) in the roots of Arg and Moghan3 under the normal and salinity treated conditions at the probability level of 5%.

https://doi.org/10.1371/journal.pone.0254189.s002

(DOCX)

S3 Fig. Survey of DEGs observed between the control and salt treated samples (cut off p-value: 0.01).

(a) Out of 4290 DEGs, 110 and 98 genes were exclusively expressed under the salt stress (STL) and control (CL) conditions, respectively (b) Fold change distribution of 4082 DEGs present in both normal and salt treated samples.

https://doi.org/10.1371/journal.pone.0254189.s003

(DOCX)

S4 Fig.

(a) Annotation statistics of the novel DEGs. (b) GO classification of the novel DEGs under salt stress.

https://doi.org/10.1371/journal.pone.0254189.s004

(DOCX)

S5 Fig. Cellular pathway overview of DEGs in T. aestivum under salinity stress using Mapman.

Blue: up-regulated genes and red: down-regulated genes.

https://doi.org/10.1371/journal.pone.0254189.s005

(DOCX)

S6 Fig. Secondary metabolite pathway overview of the DEGs in T. aestivum under salinity stress using Mapman.

Blue: up-regulated genes and red: down-regulated genes.

https://doi.org/10.1371/journal.pone.0254189.s006

(DOCX)

S7 Fig. Stress response pathways overview of the DEGs in T. aestivum under salinity stress using Mapman.

Blue: up-regulated genes and red: down-regulated genes.

https://doi.org/10.1371/journal.pone.0254189.s007

(DOCX)

S1 Table. The primers used for Real Time PCR.

https://doi.org/10.1371/journal.pone.0254189.s008

(XLSX)

S2 Table. Biological process classification of the novel transcripts.

https://doi.org/10.1371/journal.pone.0254189.s009

(XLSX)

S3 Table. Molecular function classification of the novel transcripts.

https://doi.org/10.1371/journal.pone.0254189.s010

(XLSX)

S4 Table. Cellular component classification of the novel transcrips.

https://doi.org/10.1371/journal.pone.0254189.s011

(XLSX)

S5 Table. List of the differentially expressed genes.

https://doi.org/10.1371/journal.pone.0254189.s012

(XLSX)

S6 Table. List of the genes exclusively expressed under salt stress.

https://doi.org/10.1371/journal.pone.0254189.s013

(XLSX)

S7 Table. List of the novel differentially expressed genes.

https://doi.org/10.1371/journal.pone.0254189.s014

(XLSX)

S8 Table. KEGG pathway classification of the DEGs.

https://doi.org/10.1371/journal.pone.0254189.s015

(XLSX)

S9 Table. Results of functional analysis of the salt-regulated genes using Mapman.

https://doi.org/10.1371/journal.pone.0254189.s016

(XLSX)

Acknowledgments

The authors are grateful to Seed and Plant Improvement Institute (SPII) for providing the seeds, Miss. Saeedeh Asari for her technical assistance and Mr. Mohammad Jedari to help in creating the artworks.

References

  1. 1. Maurel C, Boursiac Y, Luu D-T, Santoni V, Shahzad Z, Verdoucq L. Aquaporins in plants. Physiological reviews. 2015;95(4):1321–58. pmid:26336033
  2. 2. Kong W, Zhong H, Gong Z, Fang X, Sun T, Deng X, et al. Meta-analysis of salt stress transcriptome responses in different rice genotypes at the seedling stage. Plants. 2019;8(3):64.
  3. 3. F.A.O. FAO cereal supply and demand brief. http://wwwfaoorg/worldfoodsituation/csdb/en/. 2015.
  4. 4. Royo A, Abió D. Salt tolerance in durum wheat cultivars. Spanish Journal of Agricultural Research. 2003;1(3):27–35.
  5. 5. Shrivastava P, Kumar R. Soil salinity: a serious environmental issue and plant growth promoting bacteria as one of the tools for its alleviation. Saudi journal of biological sciences. 2015;22(2):123–31. pmid:25737642
  6. 6. Kasuga M, Liu Q, Miura S, Yamaguchi-Shinozaki K, Shinozaki K. Improving plant drought, salt, and freezing tolerance by gene transfer of a single stress-inducible transcription factor. Nature biotechnology. 1999;17(3):287–91. pmid:10096298
  7. 7. Tuteja N. Mechanisms of high salinity tolerance in plants. Methods in enzymology. 428: Elsevier; 2007. p. 419–38. pmid:17875432
  8. 8. Danquah A, de Zelicourt A, Colcombet J, Hirt H. The role of ABA and MAPK signaling pathways in plant abiotic stress responses. Biotechnology advances. 2014;32(1):40–52. pmid:24091291
  9. 9. Ji H, Pardo JM, Batelli G, Van Oosten MJ, Bressan RA, Li X. The Salt Overly Sensitive (SOS) pathway: established and emerging roles. Molecular plant. 2013;6(2):275–86. pmid:23355543
  10. 10. Kishor PK, Sangam S, Amrutha R, Laxmi PS, Naidu K, Rao K, et al. Regulation of proline biosynthesis, degradation, uptake and transport in higher plants: its implications in plant growth and abiotic stress tolerance. Curr Sci. 2005;88(3):424–38.
  11. 11. Knight H. Calcium signaling during abiotic stress in plants. International review of cytology. 195: Elsevier; 1999. p. 269–324.
  12. 12. Zhu J-K. Salt and drought stress signal transduction in plants. Annual review of plant biology. 2002;53(1):247–73. pmid:12221975
  13. 13. Chinnusamy V, Jagendorf A, Zhu J-K. Understanding and improving salt tolerance in plants. Crop Science. 2005;45(2):437–48.
  14. 14. Wang Z, Gerstein M, Snyder M. RNA-Seq: a revolutionary tool for transcriptomics. Nature reviews genetics. 2009;10(1):57–63. pmid:19015660
  15. 15. Xiong H, Guo H, Xie Y, Zhao L, Gu J, Zhao S, et al. RNAseq analysis reveals pathways and candidate genes associated with salinity tolerance in a spaceflight-induced wheat mutant. Scientific reports. 2017;7(1):1–13.
  16. 16. Mahajan MM, Goyal E, Singh AK, Gaikwad K, Kanika K. Transcriptome dynamics provide insights into long-term salinity stress tolerance in Triticum aestivum cv. Kharchia Local. Plant Physiology and Biochemistry. 2017;121:128–39. pmid:29102901
  17. 17. Luo Q, Teng W, Fang S, Li H, Li B, Chu J, et al. Transcriptome analysis of salt-stress response in three seedling tissues of common wheat. The Crop Journal. 2019;7(3):378–92.
  18. 18. Zhang Y, Liu Z, Khan AA, Lin Q, Han Y, Mu P, et al. Expression partitioning of homeologs and tandem duplications contribute to salt tolerance in wheat (Triticum aestivum L.). Scientific reports. 2016;6(1):1–10.
  19. 19. Amirbakhtiar N, Ismaili A, Ghaffari MR, Nazarian Firouzabadi F, Shobbar Z-S. Transcriptome response of roots to salt stress in a salinity-tolerant bread wheat cultivar. PloS one. 2019;14(3):e0213305. pmid:30875373
  20. 20. Mahajan MM, Goyal E, Singh AK, Gaikwad K, Kanika K. Shedding light on response of Triticum aestivum cv. Kharchia Local roots to long-term salinity stress through transcriptome profiling. Plant Growth Regulation. 2020;90(2):369–81.
  21. 21. Aghaei MJ, Mozafari J, Taleei AR, Naghavi MR, Omidi M. Distribution and diversity of Aegilops tauschii in Iran. Genetic resources and crop evolution. 2008;55(3):341–9.
  22. 22. Dudnikov AJ, Kawahara T. Aegilops tauschii: genetic variation in Iran. Genetic Resources and Crop Evolution. 2006;53(3):579–86.
  23. 23. NAKAI Y. Isozyme variations in Aegilops and Triticum, IV. The origin of the common wheats revealed from the study on esterase isozymes in synthesized hexaploid wheats. The Japanese Journal of Genetics. 1979;54(3):175–89.
  24. 24. TSUNEWAKI K. Comparative gene analysis of common wheat and its ancestral species. II. Waxiness, growth habit and awnedness. Jpn J Bot. 1966;19:175–229.
  25. 25. Lachica M, Aguilar A, Yañez J. Foliar analysis: Analytical methods used in the Estacion Experimental del Zaidin. iI. An Edafol Agribiol. 1973.
  26. 26. Trapnell C, Roberts A, Goff L, Pertea G, Kim D, Kelley DR, et al. Differential gene and transcript expression analysis of RNA-seq experiments with TopHat and Cufflinks. Nature protocols. 2012;7(3):562–78. pmid:22383036
  27. 27. Moriya Y, Itoh M, Okuda S, Yoshizawa AC, Kanehisa M. KAAS: an automatic genome annotation and pathway reconstruction server. Nucleic acids research. 2007;35(suppl_2):W182–W5. pmid:17526522
  28. 28. Thimm O, Bläsing O, Gibon Y, Nagel A, Meyer S, Krüger P, et al. MAPMAN: a user-driven tool to display genomics data sets onto diagrams of metabolic pathways and other biological processes. The Plant Journal. 2004;37(6):914–39. pmid:14996223
  29. 29. Goyal E, Amit SK, Singh RS, Mahato AK, Chand S, Kanika K. Transcriptome profiling of the salt-stress response in Triticum aestivum cv. Kharchia Local. Scientific reports. 2016;6(1):1–14.
  30. 30. Ma X, Gu P, Liang W, Zhang Y, Jin X, Wang S, et al. Analysis on the transcriptome information of two different wheat mutants and identification of salt-induced differential genes. Biochemical and biophysical research communications. 2016;473(4):1197–204. pmid:27074578
  31. 31. Livak KJ, Schmittgen TD. Analysis of relative gene expression data using real-time quantitative PCR and the 2− ΔΔCT method. methods. 2001;25(4):402–8. pmid:11846609
  32. 32. Renault S, Croser C, Franklin JA, Zwiazek JJ. Effects of NaCl and Na2SO4 on red-osier dogwood (Cornus stolonifera Michx) seedlings. Plant and soil. 2001;233(2):261–8.
  33. 33. Hasegawa PM, Bressan RA, Zhu J-K, Bohnert HJ. Plant cellular and molecular responses to high salinity. Annual review of plant biology. 2000;51(1):463–99. pmid:15012199
  34. 34. Davenport RJ, MUÑOZ-MAYOR A, Jha D, Essah PA, Rus A, Tester M. The Na+ transporter AtHKT1; 1 controls retrieval of Na+ from the xylem in Arabidopsis. Plant, cell & environment. 2007;30(4):497–507.
  35. 35. Jain M. Next-generation sequencing technologies for gene expression profiling in plants. Briefings in functional genomics. 2011;11(1):63–70. pmid:22155524
  36. 36. Sun J, Jiang H, Xu Y, Li H, Wu X, Xie Q, et al. The CCCH-type zinc finger proteins AtSZF1 and AtSZF2 regulate salt stress responses in Arabidopsis. Plant and Cell Physiology. 2007;48(8):1148–58. pmid:17609218
  37. 37. Lu X, Chen D, Shu D, Zhang Z, Wang W, Klukas C, et al. The differential transcription network between embryo and endosperm in the early developing maize seed. Plant physiology. 2013;162(1):440–55. pmid:23478895
  38. 38. Shankar R, Bhattacharjee A, Jain M. Transcriptome analysis in different rice cultivars provides novel insights into desiccation and salinity stress responses. Scientific reports. 2016;6:23719. pmid:27029818
  39. 39. Singh A, Jha SK, Bagri J, Pandey GK. ABA inducible rice protein phosphatase 2C confers ABA insensitivity and abiotic stress tolerance in Arabidopsis. PloS one. 2015;10(4):e0125168. pmid:25886365
  40. 40. Karthikeyan A, Pandian SK, Ramesh M. Transgenic indica rice cv. ADT 43 expressing a Δ 1-pyrroline-5-carboxylate synthetase (P5CS) gene from Vigna aconitifolia demonstrates salt tolerance. Plant Cell, Tissue and Organ Culture (PCTOC). 2011;107(3):383–95.
  41. 41. Yang A, Dai X, Zhang W-H. A R2R3-type MYB gene, OsMYB2, is involved in salt, cold, and dehydration tolerance in rice. Journal of experimental botany. 2012;63(7):2541–56. pmid:22301384
  42. 42. Wang P, Li Z, Wei J, Zhao Z, Sun D, Cui S. A Na+/Ca2+ exchanger-like protein (AtNCL) involved in salt stress in Arabidopsis. Journal of Biological Chemistry. 2012;287(53):44062–70. pmid:23148213
  43. 43. Yang L, Ma C, Wang L, Chen S, Li H. Salt stress induced proteome and transcriptome changes in sugar beet monosomic addition line M14. Journal of plant physiology. 2012;169(9):839–50. pmid:22498239
  44. 44. Jia F, Qi S, Li H, Liu P, Li P, Wu C, et al. Overexpression of Late Embryogenesis Abundant 14 enhances Arabidopsis salt stress tolerance. Biochemical and biophysical research communications. 2014;454(4):505–11. pmid:25450686
  45. 45. Qi Y, Liu W, Qiu L, Zhang S, Ma L, Zhang H. Overexpression of glutathione S-transferase gene increases salt tolerance of Arabidopsis. Russian journal of plant physiology. 2010;57(2):233–40.
  46. 46. Wang C, Yang Y, Wang H, Ran X, Li B, Zhang J, et al. Ectopic expression of a cytochrome P450 monooxygenase gene PtCYP714A3 from Populus trichocarpa reduces shoot growth and improves tolerance to salt stress in transgenic rice. Plant biotechnology journal. 2016;14(9):1838–51. pmid:26970512
  47. 47. Zhang F, Zhou Y, Zhang M, Luo X, Xie J. Effects of drought stress on global gene expression profile in leaf and root samples of Dongxiang wild rice (Oryza rufipogon). Bioscience Reports. 2017;37(3):11.
  48. 48. Zhou Y, Yang P, Cui F, Zhang F, Luo X, Xie J. Transcriptome analysis of salt stress responsiveness in the seedlings of Dongxiang wild rice (Oryza rufipogon Griff.). PloS one. 2016;11(1):e0146242. pmid:26752408
  49. 49. Baxter HL, Stewart CN Jr. Effects of altered lignin biosynthesis on phenylpropanoid metabolism and plant stress. Biofuels. 2013;4(6):635–50.
  50. 50. Vogt T. Phenylpropanoid biosynthesis. Molecular plant. 2010;3(1):2–20. pmid:20035037
  51. 51. Hisano H, Nandakumar R, Wang Z-Y. Genetic modification of lignin biosynthesis for improved biofuel production. In Vitro Cellular & Developmental Biology-Plant. 2009;45(3):306–13.
  52. 52. Neves G, Marchiosi R, Ferrarese M, Siqueira-Soares R, Ferrarese-Filho O. Root growth inhibition and lignification induced by salt stress in soybean. Journal of Agronomy and crop science. 2010;196(6):467–73.
  53. 53. Fernandez-Garcia N, Lopez-Perez L, Hernandez M, Olmos E. Role of phi cells and the endodermis under salt stress in Brassica oleracea. New Phytologist. 2009;181(2):347–60. pmid:19121032
  54. 54. Ghosh R, Choi B, Cho B-K, Lim H-S, Park S-U, Bae H-J, et al. Characterization of developmental-and stress-mediated expression of cinnamoyl-CoA reductase in kenaf (Hibiscus cannabinus L.). The scientific world journal. 2014;2014:10.
  55. 55. Kim J, Choi B, Cho B-K, Lim H-S, Kim JB, Natarajan S, et al. Molecular cloning, characterization and expression of the caffeic acid O-methyltransferase (COMT) ortholog from kenaf (Hibiscus cannabinus). Plant Omics. 2013;6(4):246.
  56. 56. Moradi F, Ismail AM. Responses of photosynthesis, chlorophyll fluorescence and ROS-scavenging systems to salt stress during seedling and reproductive stages in rice. Annals of botany. 2007;99(6):1161–73. pmid:17428832
  57. 57. Munns R. Comparative physiology of salt and water stress. Plant, cell & environment. 2002;25(2):239–50. pmid:11841667
  58. 58. Li Y, Zhang D, Li W, Mallano AI, Zhang Y, Wang T, et al. Expression study of soybean germin-like gene family reveals a role of GLP7 gene in various abiotic stress tolerances. Canadian Journal of Plant Science. 2016;96(2):296–304.
  59. 59. Wang T, Chen X, Zhu F, Li H, Li L, Yang Q, et al. Characterization of peanut germin-like proteins, AhGLPs in plant development and defense. PLos one. 2013;8(4):e61722. pmid:23626720
  60. 60. Zhang X, Zhen J, Li Z, Kang D, Yang Y, Kong J, et al. Expression profile of early responsive genes under salt stress in upland cotton (Gossypium hirsutum L.). Plant Molecular Biology Reporter. 2011;29(3):626–37.
  61. 61. Singh A, Kanwar P, Yadav AK, Mishra M, Jha SK, Baranwal V, et al. Genome-wide expressional and functional analysis of calcium transport elements during abiotic stress and development in rice. The FEBS journal. 2014;281(3):894–915. pmid:24286292
  62. 62. Chung WS, Lee SH, Kim JC, Do Heo W, Kim MC, Park CY, et al. Identification of a calmodulin-regulated soybean Ca2+-ATPase (SCA1) that is located in the plasma membrane. The Plant Cell. 2000;12(8):1393–407. pmid:10948258
  63. 63. Geisler M, Frangne N, Gomes E, Martinoia E, Palmgren MG. The ACA4 gene of Arabidopsis encodes a vacuolar membrane calcium pump that improves salt tolerance in yeast. Plant physiology. 2000;124(4):1814–27. pmid:11115896
  64. 64. Perez-Prat E, Narasimhan ML, Binzel ML, Botella MA, Chen Z, Valpuesta V, et al. Induction of a putative Ca2+-ATPase mRNA in NaCl-adapted cells. Plant physiology. 1992;100(3):1471–8. pmid:16653146
  65. 65. Wimmers LE, Ewing NN, Bennett AB. Higher plant Ca (2+)-ATPase: primary structure and regulation of mRNA abundance by salt. Proceedings of the National Academy of Sciences. 1992;89(19):9205–9. pmid:1384045
  66. 66. Kang J, Mehta S, Turano FJ. The putative glutamate receptor 1.1 (AtGLR1. 1) in Arabidopsis thaliana regulates abscisic acid biosynthesis and signaling to control development and water loss. Plant and Cell Physiology. 2004;45(10):1380–9. pmid:15564521
  67. 67. Meyerhoff O, Müller K, Roelfsema MRG, Latz A, Lacombe B, Hedrich R, et al. AtGLR3. 4, a glutamate receptor channel-like gene is sensitive to touch and cold. Planta. 2005;222(3):418–27. pmid:15864638
  68. 68. Weinl S, Kudla J. The CBL–CIPK Ca2+-decoding signaling network: function and perspectives. New Phytologist. 2009;184(3):517–28.
  69. 69. Deinlein U, Stephan AB, Horie T, Luo W, Xu G, Schroeder JI. Plant salt-tolerance mechanisms. Trends in plant science. 2014;19(6):371–9. pmid:24630845
  70. 70. Virdi AS, Singh S, Singh P. Abiotic stress responses in plants: roles of calmodulin-regulated proteins. Frontiers in plant science. 2015;6:809. pmid:26528296
  71. 71. Nguyen VN, Vo KT, Park H, Jeon J-S, Jung K-H. A systematic view of the MLO family in rice suggests their novel roles in morphological development, diurnal responses, the light-signaling pathway, and various stress responses. Frontiers in plant science. 2016;7:1413. pmid:27729915
  72. 72. Wan D, Li R, Zou B, Zhang X, Cong J, Wang R, et al. Calmodulin-binding protein CBP60g is a positive regulator of both disease resistance and drought tolerance in Arabidopsis. Plant cell reports. 2012;31(7):1269–81. pmid:22466450
  73. 73. Chaires M, Gupta D, Joshee N, Cooper KK, Basu C. RNA-seq analysis of the salt stress-induced transcripts in fast-growing bioenergy tree, Paulownia elongata. Journal of Plant Interactions. 2017;12(1):128–36.
  74. 74. Mansuri RM, Shobbar Z-S, Jelodar NB, Ghaffari MR, Nematzadeh G-A, Asari S. Dissecting molecular mechanisms underlying salt tolerance in rice: a comparative transcriptional profiling of the contrasting genotypes. Rice. 2019;12(1):13. pmid:30830459
  75. 75. Naranjo MA, Forment J, RoldÁN M, Serrano R, Vicente O. Overexpression of Arabidopsis thaliana LTL1, a salt-induced gene encoding a GDSL-motif lipase, increases salt tolerance in yeast and transgenic plants. Plant, cell & environment. 2006;29(10):1890–900. pmid:16930315
  76. 76. Ye Y, Ding Y, Jiang Q, Wang F, Sun J, Zhu C. The role of receptor-like protein kinases (RLKs) in abiotic stress response in plants. Plant cell reports. 2017;36(2):235–42. pmid:27933379
  77. 77. Kanneganti V, Gupta AK. Wall associated kinases from plants—an overview. Physiology and Molecular Biology of Plants. 2008;14(1–2):109–18. pmid:23572878
  78. 78. Sivaguru M, Ezaki B, He Z-H, Tong H, Osawa H, Baluška F, et al. Aluminum-induced gene expression and protein localization of a cell wall-associated receptor kinase in Arabidopsis. Plant physiology. 2003;132(4):2256–66. pmid:12913180
  79. 79. Hou X, Tong H, Selby J, DeWitt J, Peng X, He Z-H. Involvement of a cell wall-associated kinase, WAKL4, in Arabidopsis mineral responses. Plant physiology. 2005;139(4):1704–16. pmid:16286448
  80. 80. Vaid N, Pandey P, Srivastava VK, Tuteja N. Pea lectin receptor-like kinase functions in salinity adaptation without yield penalty, by alleviating osmotic and ionic stresses and upregulating stress-responsive genes. Plant Molecular Biology. 2015;88(1–2):193–206. pmid:25863480
  81. 81. Riechmann JL, Heard J, Martin G, Reuber L, Jiang C-Z, Keddie J, et al. Arabidopsis transcription factors: genome-wide comparative analysis among eukaryotes. Science. 2000;290(5499):2105–10. pmid:11118137
  82. 82. Rosinski JA, Atchley WR. Molecular evolution of the Myb family of transcription factors: evidence for polyphyletic origin. Journal of Molecular Evolution. 1998;46(1):74–83. pmid:9419227
  83. 83. Zhang P, Wang R, Yang X, Ju Q, Li W, Lü S, et al. The R2R3-MYB transcription factor AtMYB49 modulates salt tolerance in Arabidopsis by modulating the cuticle formation and antioxidant defense. Plant, Cell & Environment. 2020;43:1925–43.
  84. 84. Zhang X, Chen L, Shi Q, Ren Z. SlMYB102, an R2R3-type MYB gene, confers salt tolerance in transgenic tomato. Plant Science. 2020;291:110356. pmid:31928668
  85. 85. Wang Z, Yan L, Wan L, Huai D, Kang Y, Shi L, et al. Genome-wide systematic characterization of bZIP transcription factors and their expression profiles during seed development and in response to salt stress in peanut. BMC genomics. 2019;20(1):1–14.
  86. 86. Pan Y, Hu X, Li C, Xu X, Su C, Li J, et al. SlbZIP38, a tomato bZIP family gene downregulated by abscisic acid, is a negative regulator of drought and salt stress tolerance. Genes. 2017;8(12):402. pmid:29261143
  87. 87. Liang C, Meng Z, Meng Z, Malik W, Yan R, Lwin KM, et al. GhABF2, a bZIP transcription factor, confers drought and salinity tolerance in cotton (Gossypium hirsutum L.). Scientific reports. 2016;6(1):1–14.
  88. 88. Li W-T, He M, Wang J, Wang Y-P. Zinc finger protein (ZFP) in plants-A review. Plant Omics. 2013;6(6):474.
  89. 89. Zhang Z, Liu H, Sun C, Ma Q, Bu H, Chong K, et al. A C2H2 zinc-finger protein OsZFP213 interacts with OsMAPK3 to enhance salt tolerance in rice. Journal of plant physiology. 2018;229:100–10. pmid:30055519
  90. 90. Guo YH, Yu YP, Wang D, Wu CA, Yang GD, Huang JG, et al. GhZFP1, a novel CCCH-type zinc finger protein from cotton, enhances salt stress tolerance and fungal disease resistance in transgenic tobacco by interacting with GZIRD21A and GZIPR5. New Phytologist. 2009;183(1):62–75. pmid:19402879
  91. 91. Ebel C, BenFeki A, Hanin M, Solano R, Chini A. Characterization of wheat (Triticum aestivum) TIFY family and role of Triticum Durum TdTIFY11a in salt stress tolerance. PloS one. 2018;13(7):e0200566. pmid:30021005
  92. 92. Phukan UJ, Jeena GS, Shukla RK. WRKY transcription factors: molecular regulation and stress responses in plants. Frontiers in plant science. 2016;7:760. pmid:27375634
  93. 93. Puranik S, Sahu PP, Srivastava PS, Prasad M. NAC proteins: regulation and role in stress tolerance. Trends in plant science. 2012;17(6):369–81. pmid:22445067
  94. 94. Hong Y, Zhang H, Huang L, Li D, Song F. Overexpression of a stress-responsive NAC transcription factor gene ONAC022 improves drought and salt tolerance in rice. Frontiers in Plant Science. 2016;7:4. pmid:26834774
  95. 95. Liu C, Wang B, Li Z, Peng Z, Zhang J. TsNAC1 is a key transcription factor in abiotic stress resistance and growth. Plant physiology. 2018;176(1):742–56. pmid:29122985
  96. 96. Müller M, Munné-Bosch S. Ethylene response factors: a key regulatory hub in hormone and stress signaling. Plant physiology. 2015;169(1):32–41. pmid:26103991
  97. 97. Rong W, Qi L, Wang A, Ye X, Du L, Liang H, et al. The ERF transcription factor Ta ERF 3 promotes tolerance to salt and drought stresses in wheat. Plant biotechnology journal. 2014;12(4):468–79. pmid:24393105
  98. 98. Wang X, Hou C, Zheng K, Li Q, Chen S, Wang S. Overexpression of ERF96, a small ethylene response factor gene, enhances salt tolerance in Arabidopsis. Biologia Plantarum. 2017;61(4):693–701.
  99. 99. Bhattacharjee A, Khurana JP, Jain M. Characterization of rice homeobox genes, OsHOX22 and OsHOX24, and over-expression of OsHOX24 in transgenic Arabidopsis suggest their role in abiotic stress response. Frontiers in plant science. 2016;7:627. pmid:27242831
  100. 100. Mirdar Mansuri R, Shobbar Z-S, Babaeian Jelodar N, Ghaffari MR, Nematzadeh G-A, Asari S. Dissecting molecular mechanisms underlying salt tolerance in rice: a comparative transcriptional profiling of the contrasting genotypes. Rice (19398425). 2019;12(1). pmid:30830459
  101. 101. Sun M-H, Ma Q-J, Hu D-G, Zhu X-P, You C-X, Shu H-R, et al. The glucose sensor MdHXK1 phosphorylates a tonoplast Na+/H+ exchanger to improve salt tolerance. Plant physiology. 2018;176(4):2977–90. pmid:29440593
  102. 102. Berthomieu P, Conéjéro G, Nublat A, Brackenbury WJ, Lambert C, Savio C, et al. Functional analysis of AtHKT1 in Arabidopsis shows that Na+ recirculation by the phloem is crucial for salt tolerance. The EMBO journal. 2003;22(9):2004–14. pmid:12727868
  103. 103. Garciadeblás B, Senn ME, Bañuelos MA, Rodríguez-Navarro A. Sodium transport and HKT transporters: the rice model. The Plant Journal. 2003;34(6):788–801. pmid:12795699
  104. 104. Kim EJ, Kwak JM, Uozumi N, Schroeder JI. AtKUP1: an Arabidopsis gene encoding high-affinity potassium transport activity. The Plant Cell. 1998;10(1):51–62. pmid:9477571
  105. 105. Long W, Zou X, Zhang X. Transcriptome analysis of canola (Brassica napus) under salt stress at the germination stage. PloS one. 2015;10(2):e0116217. pmid:25679513