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 response of roots to salt stress in a salinity-tolerant bread wheat cultivar

  • Nazanin Amirbakhtiar,

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

    Affiliation Department of Agronomy and Plant Breeding, Faculty of Agriculture, Lorestan University, Khorramabad, Iran

  • Ahmad Ismaili ,

    Roles Methodology, Supervision, Writing – review & editing

    shobbar@abrii.ac.ir (ZSS); ismaili.a@lu.ac.ir (AI)

    Affiliation Department of Agronomy and Plant Breeding, Faculty of Agriculture, Lorestan University, Khorramabad, Iran

  • Mohammad Reza Ghaffari,

    Roles Formal analysis, Methodology

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

  • Farhad Nazarian Firouzabadi,

    Roles Writing – review & editing

    Affiliation Department of Agronomy and Plant Breeding, Faculty of Agriculture, Lorestan University, Khorramabad, Iran

  • Zahra-Sadat Shobbar

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

    shobbar@abrii.ac.ir (ZSS); ismaili.a@lu.ac.ir (AI)

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

Abstract

Salt stress is one of the major adverse environmental factors limiting crop productivity. Considering Iran as one of the bread wheat origins, we sequenced root transcriptome of an Iranian salt tolerant cultivar, Arg, under salt stress to extend our knowledge of the molecular basis of salinity tolerance in Triticum aestivum. RNA sequencing resulted in more than 113 million reads and about 104013 genes were obtained, among which 26171 novel transcripts were identified. A comparison of abundances showed that 5128 genes were differentially expressed due to salt stress. The differentially expressed genes (DEGs) were annotated with Gene Ontology terms, and the key pathways were identified using Kyoto Encyclopedia of Gene and Genomes (KEGG) pathway mapping. The DEGs could be classified into 227 KEGG pathways among which transporters, phenylpropanoid biosynthesis, transcription factors, glycosyltransferases, glutathione metabolism and plant hormone signal transduction represented the most significant pathways. Furthermore, the expression pattern of nine genes involved in salt stress response was compared between the salt tolerant (Arg) and susceptible (Moghan3) cultivars. A panel of novel genes and transcripts is found in this research to be differentially expressed under salinity in Arg cultivar and a model is proposed for salt stress response in this salt tolerant cultivar of wheat employing the DEGs. The achieved results can be beneficial for better understanding and improvement of salt tolerance in wheat.

Introduction

Soil salinity is a major environmental factor which limits the growth and development of plants, resulting in decrease in crop productivity and quality[1, 2]. It is estimated that salt stress affects approximately 20% of the irrigated land worldwide and will lead to the loss of 50% of cultivable land by the middle of the twenty-first century[3].

High soil salt concentrations reduce the capability of a plant to absorb water. Moreover, when Na+ and Cl are absorbed in large quantities by roots, both Na+ and Cl adversely influence growth by ruining metabolic processes and reducing photosynthetic efficiency[4, 5]. Therefore, salt stress limits the growth of plants through early-occurring osmotic stress and slowly-occurring ion cytotoxicity[6]. Plants use mechanisms to relieve osmotic stress by decreasing water loss and maximizing water uptake. In addition, plants minimize the adverse consequences of ionic Na+ stress by excretion of Na+ from leaf tissues and by compartmentalization of Na+ largely into vacuoles[7, 8]. Notwithstanding these tolerance mechanisms, salt stress declines crop yields and results in continuous loss of arable land. Therefore, identifying the main genes and mechanisms involved in salinity tolerance in order to engineer crops to improve salt-tolerance mechanisms is necessary to address these challenges[6].

Clarifying the key components in the plant salt tolerance network is essential to engineer more salt tolerant plants. Three kinds of genes are involved in salt stress response in plants including the genes involved in sensing and signaling of the stress, transcriptional regulators and salt-stress related genes. Under salt stress, Na+ enters the cell through non-selective cation channels and other membrane transporters (that most of them are unknown) and inside the cell, Na+ is sensed by an unknown sensory mechanism. At the next step, Ca2+, reactive oxygen species (ROS) and hormones act as the secondary messengers. In the Ca2+ signaling pathway, for example, kinases like Calcineurin B-like proteins (CBLs), CBL-interacting protein kinases (CIPKs) and Calcium-dependent protein kinases (CDPKs) are present which can change the transcriptional profile of the plant. Transcription factors families such as WRKY, MYB, bHLH, bZIP, AP2/ERF and NAC had been shown to be engaged in salt stress response. Finally, these early signaling pathways lead to expression and activation of cellular detoxification mechanisms including Na+ transport mechanisms and osmotic protection strategies[6].

RNA-seq, having high accuracy and sensitivity is one of the most suitable techniques to study the whole transcriptome[9, 10]. It has advantages such as the ability to identify novel genes/transcripts, detect low abundance transcripts, identify genetic variants and detect more differentially expressed genes with higher fold-change in comparison with microarray[10, 11]. Recently, a few RNA-seq studies was used to explore the transcriptome of bread wheat under salt stress: Goyal et al. (2016) found genes involved in providing energy to form proton gradient to drive exceeding cytoplasmic Na+ into vacuoles (like V-ATPase gene), ROS scavengers, genes engaged in energy metabolism adjustment (like ATP citrate synthase) and signaling genes (like Cbl-interacting protein kinase) as the most important genes involved in salt tolerance in root transcriptome of Kharchia local variety[12]. Zhang et al. (2016) introduced some genes such as a NAC transcription factor (homologous to Arabidopsis AtNAC025), a histone-lysine N-methyltransferase (homologous to Arabidopsis AtSDG16), a MYB transcription factor (homologous to Arabidopsis AtMYB333) and TaRSL4 gene (positively associated with root hair development) as necessary genes for salt stress tolerance in bread wheat root[13]. Xiong et al. (2017) compared salt responsive transcriptome of shoot between a salt tolerant bread wheat mutant and the salt sensitive wild type and showed that homeostasis of oxidation-reduction process is important for salt tolerance. They also found “Butanoate metabolism” as a new pathway for salinity response. Moreover, they found key genes for salinity tolerance such as arginine decarboxylase, polyamine oxidase and hormones-related genes to be more induced in salt-tolerant genotype[14]. In spite of the precious insight into the molecular and cellular mechanisms by which bread wheat responds to and tolerate salinity found by these recent research studies, the regulatory mechanisms engaged in harmonizing salt stress tolerance and plant growth are not completely perceived. Thus, a better understanding of salt-tolerance mechanisms would be helpful for breeding salt-tolerant wheat cultivars, in order to stabilize wheat production.

Wheat is the most important crop in Iran and the staple food for most of the people. Given that Iran is known as one of the origins of bread wheat and its wild progenitors [1518], sequencing an Iranian salt tolerant wheat cultivar (such as Arg) can provide new and valuable information. Therefore, in this study, we used the Illumina sequencing to compare the transcriptome of Arg under normal and salt-stressed conditions.

In order to analyze the bread wheat RNASeq data, TGACv1 reference genome of T. aestivum was used. To date, this is the most complete and accurate sequence assembly and annotation of the bread wheat reference accession, Chinese Spring[19]. To the best of our knowledge, it is the first time that this reference genome is employed for RNASeq data analysis under salt stress. The differential gene expression patterns and alternative splicing events were analyzed. In addition, novel transcripts in response to salt stress were identified. Functional categorization of the DEGs was also carried out to determine different metabolic pathways engaged in salt stress response. Overall, this research provides a comprehensive overview of transcriptional regulation in bread wheat under salt stress.

Materials and methods

Plant growth and salt stress treatment

Seeds of bread wheat cultivars of Arg (salt tolerant) and Moghan3 (salt sensitive) were collected from Seed and Plant Improvement Institute, Karaj, Iran. Arg has been produced by the hybridization between cymmit cultivar of Inia and the salt tolerant local line of 1-66-22 in the Seed and Plant Improvement Institute (SPII) of Iran [20].

The seeds were surface sterilized in 1% Sodium hypochlorite (NaClO) for 10 min, then washed in distilled water several times, and finally laid on moistened filter paper. After 2–3 days, the uniformly germinated seeds were grown hydroponically in half-strength Hoagland solution in the green house. The 3-week old plants were salt treated using 150 mM NaCl solution for 12 h. The root samples were taken from both control and salt-treated plants with four biological replicates, each containing three plants. The root samples were immediately frozen in liquid nitrogen and stored at -80°C until analysis.

RNA extraction and Illumina deep sequencing

Total RNA was isolated from the four biological replicates of normal and stressed roots (after 12 h exposure to salt stress) using RNeasy Plant Mini Kit (Qiagen) according to the manufacturer’s instructions (Qiagen). Equal amounts of total RNA of each two biological replicates were pooled for the RNA sequencing. The purity and integrity of RNA was checked by nanodrop, agarose gel electrophoresis and Agilent Bioanalyzer 2100 system (Agilent Technologies Co. Ltd., Beijing, China). The samples with RIN value higher than 9.1 were used for sequencing.

After quality control (QC) of the RNA samples, poly(A) enrichment, RNA fragmentation, random hexamer-primed cDNA synthesis, linker ligation, size selection and PCR amplification were done to prepare cDNA libraries for each sample. Finally, the qualified libraries were fed into HiSeq sequencers after pooling according to its effective concentration and expected data volume. The libraries were sequenced at the Novogene Bioinformatic Institute (Beijing, China) on an Illumina Hiseq 2500 platform and 150bp paired end reads were generated. After sequencing, reads containing adapters, reads with N > 10% (N indicates that the base cannot be determined) and reads having low quality (Qscore< = 5) base, which was over 50% of the total base, were removed.

Quality control and reference-based assembly

The quality of raw fastq data was checked using FastQC toolkit. The high quality reads were submitted for mapping analysis against the release-34 version of wheat reference genome (ftp://ftp.ensemblgenomes.org/pub/release-34/plants/fasta/triticum_aestivum/dna/) using Tophat with default parameters. Assembly was done through Cufflinks using the TopHat mapping files with default parameters. Cuffmerge with default options was used to merge the individual assemblies and produce the final assembly. Furthermore, the novel transcripts were identified by Cuffmerge[21]. The assembled sequences were aligned against the NCBI non-redundant protein database through BlastX with an e-value cut-off of 1e-3 using Blast2GO program.

Identification of differentially expressed genes (DEGs)

The FPKM method was used to calculate the gene/transcript expression in this research. Differential gene expression was defined using Cuffdiff available in Cufflinks package utilizing options,–upper-quartile-norm,–total-hits-norm and–frag-bias-correct. The genes with log2 fold change ≥ 1 (up-regulated genes) and ≤(− 1) (down-regulated genes) with Q-value cut off of ≤ 0.01 were considered as significant differentially expressed transcripts.

Functional annotation and pathway analysis of DEGs

Classification of DEGs to GO terms was done using Blast2GO program at p-values < 0.05 [22]. Online KEGG Automatic Annotation Server (KAAS), http://www.genome.jp/kegg/kaas [23], using single-directional best hit (SBH) method was used to assign KEGG pathways to the DEGs. Furthermore, Mapman (version 3.5.1; http://mapman.gabipd.org/web/guest) [24] was used for pathway analysis of DEGs with P-value cut-off of ≤ 0.05. The DEGs were mapped on Arabidopsis pathway genes to characterize the genes involved in specific pathways.

Alternative splicing analysis

Determination of alternative splicing events was done using AStalavista web tool (version 3; http://genome.crg.es/astalavista/) with default parameters. The outputs prepared for all the AS events were further analyzed manually.

Quantitative Real Time PCR (qRT-PCR) validation

Three replicates were used for quantitative Real-Time PCR. cDNA was synthesized using qScript cDNA Synthesis Kit (Quantabio, USA) according to the manufacturer’s instruction. qRT-PCR was performed for three biological replicates using a LightCycler 96 Real-Time PCR System (Roche Life Science, Germany) and SYBR Premix EX TaqII (Takara Bio Inb,Japan) according to the manufacturer’s instructions. Normalization was done using Actin as an internal control gene as reported in previous studies[12, 24].

The gene specific primers are listed in S1 Table. The relative expression levels of the selected genes were calculated from cycle threshold values using the 2−ΔΔCt procedure [25].

Results and discussion

Sequencing statistics

To obtain a better understanding of the mechanism underlying salt tolerance of Arg cultivar at global transcriptional level, the transcriptome of Arg plants grown in control and NaCl supplemented media were surveyed by RNA-Seq. In total, more than 113.34 million reads were produced. After trimming adapters and filtering out low quality reads, more than 111.73 million clean reads remained for further analysis. Among all the reads, more than 91.3% had Phred-like quality scores at the Q30 level (an error probability of 0.1%) (S2 Table). These results indicated that the quality of sequencing data is high enough for subsequent analysis. The transcriptome raw data of the present research have been submitted at SRA (Sequence Read Achieve) of NCBI with the accession numbers of SRR7755529, SRR7755530, SRR7755531 and SRR7755532.

Reference-based transcriptome assembly was accomplished via TopHat-Cufflinks pipeline[21] utilizing the wheat genome sequence as reference. The high quality reads were mapped to the release-34 version of wheat reference genome (ftp://ftp.ensemblgenomes.org/pub/release-34/plants/fasta/triticum_aestivum/dna/) and the alignment results showed that 85.4–86.4% of the total reads mapped to the wheat reference genome including 74.73–81.03% are uniquely matched (Table 1). The assembly of mapped reads led to the identification of a total of 203080 transcript isoforms and 104013 genes.

thumbnail
Table 1. Summary of Illumina transcriptome reads mapped to the reference genes (The numbers listed in the table are the sum of left and right reads).

https://doi.org/10.1371/journal.pone.0213305.t001

Identification of novel transcripts

One of the major advantages of RNA-seq analysis is to identify novel genes/transcript isoforms [9, 10, 26]. In this study, 26171 novel transcript isoforms and 15060 novel genes were identified. It was found that the average length of novel transcripts was lesser (1690 bp) than that of known (annotated) transcripts (2285bp) similar to what was reported in other plants such as rice and maize[27, 28]. Around 74% of the total assembled transcripts and more than 49.4% of the novel transcripts were assigned with a putative function (S3 Table).

GO analysis of novel transcripts revealed that in biological process category, the novel genes were involved in cellular process, biological regulation, localization, single organism process, metabolic process, cellular component organization or biogenesis, response to stimulus and regulation of biological process (S4 Table). In the molecular function category, most of the novel transcripts were involved in heterocyclic compound binding, small molecule binding, hydrolase activity, ion binding, transferase activity, carbohydrate derivative binding, organic cyclic compound binding, nucleotide binding, hydrolase activity, transferase activity and kinase activity (S5 Table). Furthermore, novel transcripts belonged to cytoplasm, cytoplasmic part, integral component of membrane, intracellular membrane-bounded organelle, plastid, nucleus and mitochondrion cellular component GOSlim categories (S6 Table).

Identification of DEGs

A total of 5128 genes were differentially expressed between salt treated and control bread wheat (Arg cultivar) root samples (S1 Fig). Among these DEGs, 1995 genes were up-regulated and 3133 genes were down-regulated under salt stress. In addition, 109 and 210 genes were unique in salt treated and control plants, respectively (Fig 1A). Among the genes exclusively expressed under salt stress, some important genes were observed which are known to be engaged in abiotic stress response such as transcription factors (e.g. AP2/ERF and MYB), LEA proteins, dehydrins, BURP domain-containing proteins and the genes involved in cell redox homeostasis (S7 Table). Assessing fold change distribution of the DEGs showed that most of the genes had a fold change between 2 and 3 and the least number had a fold change of 6–7 (Fig 1B).

thumbnail
Fig 1. Survey of differentially expressed genes (DEGs) between control and salt stress in bread wheat.

(Cut off pvalue: 0.01). (a) Out of 5128 DEGs, 109 genes were uniquely expressed in control and 210 genes were uniquely expressed under salt stress (b) Fold change distribution of 4809 DEGs available in both samples.

https://doi.org/10.1371/journal.pone.0213305.g001

As mentioned in the introduction, three kinds of genes are involved in salt stress response. The first group contains salt responsive genes, which are involved in sensing and signaling of the (S8 Table) were discovered among the up-regulated DEGs which may act as candidate osmosensors in T. aestivum under salt stress[29]. The molecular nature of Na+ sensors is yet unclear, the plasma membrane Na+/H+ antiporter Salt Overly Sensitive1 (SOS1) can be a possible candidate because its cytoplasmic end is assumed to be involved in Na+ sensing. The gene encoding SOS1 (S8 Table) was found among the up-regulated DEGs which may function as Na+sensor[30].

Fluctuation in the cytosolic calcium concentration is one of the primary responses to different stimuli, and elements involved in Ca2+ transport actively take part in retaining this flux and homeostasis[31]. Among the DEGs, there were 3 genes coding for calcium-transporter ATPases. One of them is the up-regulated novel gene which locates in TGACv1_scaffold_641741_U:17189–17824. Orthologue of the mentioned gene in rice is Os.ACA7 (Os10g0418100), which locates in Golgi apparatus and is activated by Calmodulin[31]. Up-regulation of the Ca2+-ATPases in different plant species such as tomato, tobacco, soybean [3235]and Arabidopsis has been reported under salt stress and they may help in dropping the cytosolic calcium level, that was elevated by NaCl stress, and the maintenance of Ca2+ homeostasis[31]. Also, overexpression of N-terminal modified ACA4 in Arabidopsis seedlings resulted in increased salt tolerance in comparison with wild-type plants[36]. Annexins constitute another group of Ca2+ transporters[31]. Annexin D4 is one of the Ca+2 transporters found in this study. Annexins operate downstream of the plasma membrane NADPH oxidases that produce extracellular hydroxyl radicals, which are able to activate Ca+2 influx through annexins [37]. Another Ca2+ transporter discovered in this study was a Na+/Ca2+ exchanger coded by a novel gene, located in GACv1_scaffold_571144_7AS:15506–18943 (Ta.NCL2). It has been reported that AtNCL localizes in the cell membrane, binds Ca2+ and takes part in Ca2+ homeostasis under abiotic stresses in Arabidopsis[38]. Ta.GLR, which encodes a Glutamate receptor, is another gene involved in Ca2+ transport in this study. Glutamate receptors are non-selective cation channels [31]and are responsive to abiotic stresses[39, 40].

Following increase in Ca2+concentration under salt stress, kinases like Calcium-dependent protein kinases (CDPKs)[41], calcineurin B-like proteins (CBLs) and CBL-interacting protein kinases (CIPKs)[42] may become activated, which are able to transduce the signal to downstream protein activity and gene transcription[6]. 19 DEGs encoding CIPKs were discovered (S8 Table), among which 6 genes had been demonstrated to be engaged in salt stress response based on the information available about their orthologues in Arabidopsis [4345] (S9 Table). Six genes encoding Calmodulin were also found among the DEGs (S8 Table). One of the Ca2+-sensing proteins is Calmodulin (CaM) and it has been shown that CaM is involved in transduction of Ca2+signals. After interacting with Ca2+, CaM is subjected to conformational changes and affects the activities of CaM-binding proteins. A number of CaM-binding proteins were supposed to be involved in stress responses in plants, indicating the central role played by CaM in adaptation to detrimental environmental conditions[46].

Among the DEGs, many transcription factors (TFs) were discovered, proposing that TFs play important roles in salt stress response via regulating transcription of the downstream genes responsible for plant tolerance to salt stress. Transcription factors such as AP2/ERF, bZIP, Zn-finger, NAC, MYB and WRKY had been observed to be engaged in the regulation of abiotic stress tolerance in plants and a subset of these transcription factors are discussed in this study[47].

NAC genes are a class of plant-specific transcription factors containing a highly conserved N-terminal domain known as the NAC domain. In this study, 53 NAC domain containing genes were discovered among the DEGs, of which 29 NAC genes were up-regulated and 24 of them were down-regulated under salt stress (S8 Table). Some of these NAC genes have been defined to be involved in salt stress response based on the previous studies in wheat or on the basis of the information regarding their orthologues in Arabidopsis [4851](S9 Table).

Another class of transcription factors are zinc finger proteins (ZFPs), among which four ZFP families of C2H2, CCCH, C3HC4 and C4 play many important regulatory roles in development, growth, stress response and phytohormone response in plants[52]. In this study, about 30 differentially expressed ZF transcription factors were identified which, based on their orthologues in the Arabidopsis, seven salt responsive members were predicted (S9 Table). Among these genes, AtSZF2 (orthologue of TRIAE_CS42_3AL_TGACv1_196305_AA0659290 in Arabidopsis) negatively regulate the expression of salt-responsive genes and play key roles in modulating salt stress tolerance in Arabidopsis plants.[26]

The MYB family, found in all eukaryotes, includes a large and functionally diverse classes of proteins. Most of the MYB proteins function as transcription factors and have been proved to be engaged in regulating different cellular processes, including biotic and abiotic stress response[53]. Based on the achieved results, 48 MYB transcription factors are induced by salt stress.

The third group of salt stress responsive genes are those involved in stress adaptation. Among the DEGs, there were genes coding for Aquaporins (controlling transport of water and ions)[54], LEA proteins, dehydrins and organic osmolytes such as proline in response to osmotic stress. Late-embryogenesis-abundant (LEA) proteins are induced by osmotic stresses in vegetative tissues and cause dehydration tolerance in vegetative tissues of plants[55]. Twenty seven genes coding for LEA proteins were discovered in the DEGs identified in this study. Production of LEA proteins alongside accumulation of organic osmolytes plays key roles in sustaining the low intracellular osmotic potential of plants and thereby attenuates the detrimental effects of salinity stress[56, 57]. Furthermore, peroxidases, catalases, glutaredoxins and Gluthatione-S- transferases were differentially expressed in response to oxidative stress caused by salinity.

For dealing with the ionic stress arising from salinity, the genes coding for plasma membrane Na+/H+ antiporter SOS1, K+ transporters and ABC transporters were available among the up-regulated DEGs. The ABC transporter, Ta.ABAC15, found among the up-regulated DEGs is involved in K+ uptake, and K+ / Na+ homostasis, based on the information about its Arabidopsis orthologue (At1g04120)[58].The genes coding for HAK potassium transporters were also discovered among the DEGs in this study. Horie et al. showed that overexpression of rice Na+-impermeable K+ transporter (OsHAK5) led to salinity tolerance in tobacco bright yellow 2 (BY2) cells[59]. Ion homeostasis during salinity stress needs the preservation of stable K+ attainment and distribution[60] given that K+ accumulation in plant cells equilibrates the poisonous effects of Na+ accumulation.

The gene coding for salt overly sensitive 1 (SOS1) was also discovered among the up-regulated DEGs in this research. Plasma membrane-localized SOS1 Na+/H+ antiporter[2], which exports Na+ out of the cell, besides tonoplast-localized NHX1 Na+/H+ antiporter[61] are the two main factors that sustain low cytoplasmic Na+ concentrations in plant cells[6]. As mentioned above, SOS1 besides functioning as antiporter to export Na+ out of the cell, may act as Na+ sensors, too. At the present study, SOS2 was not observed among the DEGs but the CIPK gene which its Arabidopsis orthologue is SOS2-like protein kinase PKS12 was available among the up-regulated DEGs which is engaged in salt stress response in Arabidopsis[44]. It is likely that this gene controls the activity of the Na+/H+ antiporter SOS1. Although the gene coding for Na+/H+ antiporter NHX1 was not available among the DEGs, but hexokinase1 which is able to phosphorylate NHX1 and increase its stability was discovered among the up-regulated DEGs. In fact, this gene increases salt tolerance through increasing compartmentalization of Na+ into vacuole[62].

Gene ontology enrichment analysis for DEGs

In order to study the functions of DEGs, GO terms were extracted utilizing Blast2GO tool and then were exposed to GO enrichment analysis[22]. Annotation of DEGs revealed that a total of 4056(out of 5028) genes were assigned GO terms.

The dominant terms were ‘cell’, ‘cell part’ and ‘membrane’ in cellular component category while ‘catalytic activity’ and ‘binding’ were the most dominant terms in molecular function category. In biological process category, most of the DE genes were classified in metabolic process and cellular process followed by single organism process, biological regulation and response to stimulus (Fig 2). The transcriptional changes of genes categorized in metabolic and cellular processes were previously reported under different environmental conditions[63, 64] proposing that extensive metabolic activities occur in the stress treated plants.

thumbnail
Fig 2. GO classification of DEGs based on sequence homology to 3 main categories of cellular component, molecular function and biological process.

https://doi.org/10.1371/journal.pone.0213305.g002

Not surprisingly, the most enriched biological process terms for over-presented DEGs were electron transport, response to oxidative stress, response to chemical stimulus, carbohydrate metabolic process, plant-type cell-wall organization, transport and establishment of localization which acted as indicators of significant biological processes underlying the specific salinity-stress responses of plants and are in agreement with those reported in salt stress response in previous studies[12, 28, 65].The most enriched molecular function terms for over-presented DEGs were catalytic activity, iron ion binding, tetrapyrrole binding, cation binding, oxidoreductase activity and antioxidant activity. Meanwhile, with regard to the over-representation of cellular component terms among the DEGs, we found extracellular region, membrane, intrinsic to membrane, integral to membrane and intrinsic to plasma membrane to be the most enriched (S10 Table).

Functional annotation and classification of novel DEGs

Blast2GO was used to compare the functional annotation of novel DEGs against the NCBI non redundant (nr) protein database with a cut-off E-value of 1.0 E−3. Of the 544 novel DEGs, 405 genes (74.4%) aligned to nr protein database whereas the remaining 139 genes (25.6%) did not show homology to any sequence in the database. Sequence homology based on GO categorization utilizing Blast2GO tool indicated that out of all the novel DEGs, 259 genes (47.6%) were assigned GO terms and 229 genes (42%) were classified in significant GO terms (S2A Fig). In biological process category, the majority of genes were engaged in metabolic process and cellular process followed by single-organism process, localization, response to stimulus and biological regulation. With respect to the cellular component, cell part, cell and membrane were the dominant groups, followed by membrane part, organelle, organelle part and extracellular region. With regard to molecular function, the top three categories were catalytic activity, binding and transporter activity (S2B Fig).

It is expected that some of the novel DEGs play potential roles in salt stress tolerance such as the genes coding for peroxidase[66], Glutathione S-trasferase[67], Phosphatase 2C[68], Na+/Ca2+ Exchanger-like Protein[38], pathogenesis-related protein[69], salt stress-induced hydrophobic peptide ESI3(Early Salt stress Induced 3)[70], ATP synthase subunit beta[71], Late Embryogenesis Abundant protein[72], WRKY[73], MADS-box[74] and bHLH transcription factors[75] and cytochrome P450 monooxygenase[76] (S11 Table).

KEGG pathway classification of DEGs

For a better understanding of the active biological pathways in the DEGs under salt stress, a single-directional BLAST search against KEGG protein database was done using KAAS server[23, 77]. This is a method to classify gene functions with emphasis on the biochemical pathways. The results indicated that 1744 of the 5128 DEGs could be classified into 227 KEGG pathways, covering the five main KEGG categories of metabolism, environmental information processing, genetic information processing, organismal systems and cellular processes (Fig 3A). These genes belonged mainly to the following KEGG pathways: Transporters, phenylpropanoid biosynthesis, transcription factors, glycosyltransferases, glutathione metabolism, plant hormone signal transduction, cytochrome P450, MAPK signaling pathway–plant, plant-pathogen interaction and exosome (Fig 3B and S12 Table) that are significant pathways in abiotic and biotic stress response in plants and have been reported in previous studies[12, 65, 78]. Phenylpropanoid pathway with the second highest number of genes is a metabolic pathway liable for the synthesis of different plant secondary metabolites having roles in developmental and stress–related processes[79]. In this study, genes such as peroxidases, shikimate O-hydroxycinnamoyl transferases, caffeic acid 3-O-methyltransferases and beta-glucosidases were up-regulated in this pathway. Beta-Glucosidases are known to play a role in abiotic stresses via accumulation of reactive oxygen species (ROS) scavenging flavonols[80]. Plants use accumulation of lignin or alteration of the monomeric composition of lignin in the cell wall to overcome salt stress[81]. Up-regulation of shikimate hydroxycinnamoyl transferase and caffeic acid 3-O-methyltransferase, both engaged in lignification, have been reported under salt stress in previous studies [82, 83].

thumbnail
Fig 3. KEGG classification of the DEGs.

(a) KEGG distribution of the annotated genes into 5 main categories.(b) The top 10 pathways with the highest number of genes.

https://doi.org/10.1371/journal.pone.0213305.g003

Alternative splicing analysis

Alternative splicing (AS) is a fundamental molecular mechanism increasing transcriptome and proteome complexity and diversity in higher eukaryotes. It is reported that alternative splicing is involved in a range of functions in plants, such as growth, development, signal transduction, and responses to biotic and abiotic stress[8489].

In this study, 37% of the multi exonic genes were alternatively spliced at the whole transcriptome level. Alternative splicing has been reported in 61% of Arabidopsis thaliana genes and 21.2 to 33% of rice (Oryza sativa) genes[90, 91]. Our results revealed that 120668 alternative splicing events occurred in our assembly, of which the highest number belonged to intron retention (IR) with 43719 (36%) events, followed by alternate 3 acceptor (AA), alternate 5 donor (AD) and exon skipping (ES) represented by 20578 (17%), 13119 (11%) and 8124 (7%) events, respectively. Intron retention was the most dominant alternative splicing event in our assembly which was in line with the studies in other plants such as sorghum, rice, brachypodium and Arabidopsis.

Functional categorization of the transcripts created via IR AS event indicated their involvement in biological processes such as metabolic process, regulation of biological process, response to stimulus, localization, cellular component organization, developmental process and signaling (S3 Fig).

In addition, alternative splicing analysis was performed on control and salt-treated samples separately. Results showed that 89456 and 87606 alternative splicing events were obtained in salt-treated samples while 86882 and 86721 alternative splicing events were observed in control samples. Therefore, there is an increase in alternate splicing frequency under salt stress conditions, in accordance with the relevant report in Arabidopsis (S4 Fig). [92]. This indicates on the possible roles of AS in plant response to salt stress.

Alternative splicing in DEGs

In this study, we identified 3884 alternative splicing events in the DEGs under salt stress. In total, 30% of DEGs (1482 genes out of 5128 genes) were alternatively spliced and produced 4041 different isoforms. Among the different alternative splicing types, intron retention events predominated (1706), followed by AA (552), AD (372) and ES (197) events. In addition, 1057 events were classified as complex alternative splicing events. The ratio of different AS events in DEGs was similar to the ratio of different AS events at the whole transcriptome assembly (Fig 4).

thumbnail
Fig 4. Analysis of alternative splicing.

Bar chart showing the percentage of different types of alternative splicing events at the whole trascriptome assembly and differentially expressed genes AD: Alternate-5′Donor; AA: Alternate-3′-Acceptor, IR: Intron Retention and ES: Exon Skipping.

https://doi.org/10.1371/journal.pone.0213305.g004

Functional categorization of the alternatively spliced DEGs revealed their involvement in different pathways related to stress responses. Genes related to the biological processes of transport, nucleic acid metabolic processes, localization, response to stimulus, response to stress, metal ion transport, response to oxidative stress, signaling process, regulation of RNA metabolic process and transcription were enriched in the alternatively spliced DEGs (S5 Fig). We discovered alternative splicing events in many important genes engaged in salt stress response such as CBL-interacting kinases, serine-threonine kinases, different transcription factors (MYBs, NACs, bZIPs, HSTFs and WRKYs), phosphatases 2c, peroxidases, Gluthatione S-transferases, LEA proteins, Calcium transporting ATPases, Calcium binding proteins, antiporters, salt tolerant-related proteins, salt response proteins etc. (S13 Table).

Metabolic pathways involved in salt stress response

As a complementary approach to the GO analysis, we searched the putative functions of the salt responsive genes identified using MapMan[24], allowing the visualization of salt induced changes in different metabolic processes. The results of mapping the DEGs to the metabolic pathway overview revealed that lipid and sucrose metabolism pathways were among the enriched pathways (Fig 5 and S14 Table). In sucrose metabolism pathway, genes encoding for hexokinase, cell wall invertase and sucrose synthase are up-regulated under salt stress. Abiotic stresses usually lead to sugar accumulation [93]. Acumulation of glucose, sucrose and fructose under high salinity plays a key role in carbon storage, osmotic regulation, homeostasis and ROS scavenging[94]. Sun et al.[62] reported that exogenous application of glucose enhanced salt tolerance in apple and hexokinase1, acting as glucose sensor, contributed to Glc-mediated salinity tolerance. Hexokinase1 interacted with Tonoplast-localized Na+/H+ exchanger (NHX1) and phosphorylated it. Phosphorylation improved the stability of Na+/H+ exchanger and enhanced its Na+/H+ transport activity in transgenic apple overexpressing NHX1. Overexpression of the cell wall invertase gene from Chenopodium rubrum improved drought tolerance in tomato. This gene critically acts at the integration point of metabolic, hormonal, and stress signals, preparing a novel strategy to conquer drought-induced limitations to crop yield[95].

thumbnail
Fig 5. Metabolic pathways overview of DEGs in T.aestivum under salt stress using Mapman.

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

https://doi.org/10.1371/journal.pone.0213305.g005

The cellular overview pathway showed that genes coding for abiotic stress related miscellaneous enzyme families (misc) and glutaredoxins were up-regulated in T. aestivum under salinity stress (S6 Fig and S14 Table). Glutaredoxins (GRXs) are small redox proteins, which use glutathione to catalyze the reduction of disulfide bonds of substrate proteins to maintain cellular redox homeostasis. Furthermore, diverse functions such as transcriptional regulation of defense responses, flower development, oxidative stress response, redox signaling, hormonal regulation, iron homeostasis, and environmental adaptation have been reported for various plant GRXs[96].

The secondary metabolite pathway overview revealed that the genes involved in terpenoid, lignin, phenols and isoflavonoid metabolic pathways were significantly enriched in this salinity tolerant bread wheat variety under salinity stress (S7 Fig and S14 Table). Terpenoids are the largest and most diverse group of chemicals produced by plants. Plants employ terpenoids in basic functions such as growth and development but the majority of terpenoids are used for protection against the abiotic and biotic stresses. Increase in the expression of genes encoding terpenoids and their role in coping with salt stress have been reported in Mangrove plants[97].

In addition, the stress response pathways indicated that the genes involved in ethylene signalling pathways and genes coding for transcription regulators and peroxidases were found to be enriched in T. aestivum under salinity stress (S8 Fig and S14 Table).

Validation of differential gene expression using qRT-PCR

To further validate the RNA-Seq expression profiling, nine salt responsive genes were selected for qRT-PCR (Fig 6). The qRT-PCR results were highly consistent with those of RNA sequencing (R2 = 0.98). Therefore, the DEGs identified in this study can be considered to have a high accuracy. To achieve further insight, the expression pattern of these genes was compared between two cultivars. The results of qRT-PCR analysis showed that there was no significant difference in the expression of the selected genes between Arg and Moghan3 cultivars or weaker response was observed in the susceptible (Moghan3) cultivar except for SOS1 which showed higher expression in Moghan3 (Fig 6). It is probable that the sensitive genotype has not been able to reduce the sodium level using other approaches, so up-regulation of this antiporter might diminish the sodium level via sodium excretion.

thumbnail
Fig 6. Validation of selected genes using qRT-PCR.

(a) Calcineurin B-like protein (CBL)-CBL-interacting protein kinase 31 (Ta.CIPK31); (b) NAC transcription factor 15 (Ta. NAC15); (c) Heat Shock transcription factor C1b (Ta. HsfC1b);(d) Salt Overly Sensitive 1 (Ta. SOS1); (e) Dehydration-responsive protein RD22 (Ta. RD22) (f) bHLH transcription factor (Ta. bHLH); (g) late embryogenesis abundant protein 3 (Ta. LEA3); (h) salt tolerant-related protein (Ta. STRP); (i) Proline oxidase (Ta. POX). The gene ensemble Ids are mentioned in the S1 Table.

https://doi.org/10.1371/journal.pone.0213305.g006

In summary, this study prepares a comprehensive overview of the transcriptome changes of an Iranian salt tolerant bread wheat cultivar under salt stress. Using TGACv1 reference genome of the bread wheat for analysis, more than 85% of the total reads were mapped to this reference genome. Moreover, around 26171 novel transcripts were identified which can improve the genome annotation of T. aestivum. Multiple genes and several key pathways were recognized to be involved in salt tolerance. In addition, 3884 alternative splicing events were identified in the salt responsive genes and IR was the most dominant event. Overall, the achieved results could improve the current understanding of salt stress response in root tissue of bread wheat, but further research studies will be required to examine the application of the detected genes as biomarkers for marker-assisted breeding or cadidates for genetic engineering in order to obtain salt tolerant plants.

Conclusion

This study presents a comprehensive overview of the transcriptome changes of an Iranian salt tolerant bread wheat cultivar, Arg, under salt stress, which can help understanding the molecular basis of salinity tolerance in T. aestivum. A model is proposed for salt stress response in Arg cultivar employing the DEGs (Fig 7) (S15 Table and. S9 Fig). Based on the achieved results, salinity-induced osmotic and ionic stress might be sensed by mechanosensitive ion channels (e.g. Ta.Msc) and membrane Na+/H+ antiporter (Ta.SOS1), respectively. After sensing the stress, signaling cascades are triggered[6]. To this end, Ca+2 has been reported to serve as a secondary messenger, so an increase in cytosolic Ca+2 concentrations is expected[98]. In this study, the genes coding for Ca2+ transporters such as Ta.ANN4, Ta.ACA7 and Ta.NCL2 were appeared to be up-regulated, which may adjust the Ca+2 cytosolic concentrations. Ta.GLR, which encodes a non-selective cation channel were also induced in Arg under salt stress, and is supposed to be involved in Ca2+ transport. The genes coding for CaM, CIPK and CPK were also up-regulated, which are involved in Ca+2 signaling pathway [41, 42, 46]. The genes coding for transcription factors such as MYB, NAC, bHLH, WRKY, bZIPs and AP2/ERF were observed among the DEGs. Some of these genes have been proved to be involved in salt stress response based on the information about their orthologues in Arabidopsis (S1 Table). These transcription factors can regulate the expression of the genes engaged in dealing with osmotic, ionic and oxidative stresses arising from salinity[6]. The genes coding for Aquaporins (Ta.TP4-1-like and Ta.NIP1-1-like), LEA proteins (Ta.Wrab18, Ta.LEA1, Ta.LEA3, Ta.LEA-D34-Like and Ta.LEA14-A) and dehydrins (Ta.DHN3, Ta.DHN4, Ta.DHN7 and Ta.DHN9), P5CS (involved in proline synthesis) (Ta.P5CS) with increased expression and proline oxidase (Ta.ProDH) (involved in proline degradation) with decreased expression can alleviate the osmotic stress. In order to deal with the ionic stress, plasma membrane Na+/H+ antiporter SOS1, K+ transporters (such as Ta.HAK25) and ABC transporters (such as Ta.ABAC15) were significantly up-regulated under salt stress. The gene coding for SOS2-like protein kinase PKS12 is likely to control the activity of the Na+/H+ antiporter SOS1. Although the transcript level of Na+/H+ antiporter NHX1 was not increased under salinity in this study, but the protein encoded by up-regulated Ta.HXK1 is able to phosphorylate Ta-NHX1, leading to higher compartmentalization of Na+ into vacuole. In addition, the genes coding for catalases (Ta.CAT), glutaredoxins (Ta.GRXC1) and Gluthatione-S- transferases (Ta.GST) appear to deal with oxidative stress (Fig 7). We hope the attained results could be useful toward achieving salt tolerant cultivars through molecular breeding or genetic engineering.

thumbnail
Fig 7. Model proposed for salt stress response in Arg.

Sensing and signaling components and transcription factors highlighted in dark blue and purple, respectively. Mechanisms engaged in response to osmotic, ionic and oxidative stresses arising from salinity depicted in light blue, light green and pink, respectively.

https://doi.org/10.1371/journal.pone.0213305.g007

Supporting information

S1 Fig. Hierachical clustering for all DEGs using Heatmapper online software (http://www.heatmapper.ca).

Each row reperesents a separate gene expression and each column a separate mRNA sample.

https://doi.org/10.1371/journal.pone.0213305.s001

(DOCX)

S2 Fig.

(a) Annotation statistics of novel DEGs. (b) GO annotation clusters of novel differentially expressed genes between normal and salt stress.

https://doi.org/10.1371/journal.pone.0213305.s002

(DOCX)

S3 Fig. GOSlim terms (in biological process category) assigned to the transcripts produced through IR AS event.

https://doi.org/10.1371/journal.pone.0213305.s003

(DOCX)

S4 Fig. Alternavive splicing events happened in normal and salt-treated samples.

https://doi.org/10.1371/journal.pone.0213305.s004

(DOCX)

S5 Fig. Gene ontology enrichment for biological process category for the differentially expressed genes having alternative splicing events.

https://doi.org/10.1371/journal.pone.0213305.s005

(DOCX)

S6 Fig. Cellular response pathways overview of differentially expressed genes in Triticum aestivum under salt stress.

blue,up-regulated genes and red,down-regulated genes.

https://doi.org/10.1371/journal.pone.0213305.s006

(DOCX)

S7 Fig. Secondary metabolic pathways overview of differentially expressed genes in Triticum aestivum under salinity stress.

blue,up-regulated genes and red,down-regulated genes.

https://doi.org/10.1371/journal.pone.0213305.s007

(DOCX)

S8 Fig. Biotic stress pathways overview of differentially expressed genes in Triticum aestivum under salinity stress.

blue,up-regulated genes and red,down-regulated genes.

https://doi.org/10.1371/journal.pone.0213305.s008

(DOCX)

S9 Fig. Hierachical clustering for DEGs located in the proposed model using Heatmapper online software ((http://www.heatmapper.ca).

https://doi.org/10.1371/journal.pone.0213305.s009

(DOCX)

S3 Table. The percentage of annotated and unannotated known and novel transcripts.

https://doi.org/10.1371/journal.pone.0213305.s012

(XLSX)

S4 Table. Biological process terms assigned to novel transcripts.

https://doi.org/10.1371/journal.pone.0213305.s013

(XLSX)

S5 Table. Molecular function terms assigned to novel transcripts.

https://doi.org/10.1371/journal.pone.0213305.s014

(XLSX)

S6 Table. Cellular component terms assigned to novel transcripts.

https://doi.org/10.1371/journal.pone.0213305.s015

(XLSX)

S7 Table. List of transcripts exclusively expressed under salt stress.

https://doi.org/10.1371/journal.pone.0213305.s016

(XLSX)

S8 Table. List of differentially expressed genes.

https://doi.org/10.1371/journal.pone.0213305.s017

(XLSX)

S9 Table. List of some published salt responsive genes among the DEGs detected by RNA-Seq.

https://doi.org/10.1371/journal.pone.0213305.s018

(XLS)

S10 Table. Gene ontology terms assigned to differentially expresssed genes.

https://doi.org/10.1371/journal.pone.0213305.s019

(XLSX)

S11 Table. List of novel differentially expressed genes.

https://doi.org/10.1371/journal.pone.0213305.s020

(XLSX)

S12 Table. KEGG pathway classification of DEGs.

https://doi.org/10.1371/journal.pone.0213305.s021

(XLSX)

S13 Table. List of differentially expressed genes having alternative splicing events.

https://doi.org/10.1371/journal.pone.0213305.s022

(XLSX)

S14 Table. The results of mapping the DEGs to the metabolic pathway overview.

https://doi.org/10.1371/journal.pone.0213305.s023

(DOCX)

S15 Table. List of some salt responsive genes located in the proposed model.

https://doi.org/10.1371/journal.pone.0213305.s024

(DOCX)

Acknowledgments

The authors are grateful to the Seed and Plant Improvement Institute (SPII), Karaj, for providing plant material used in this research and Miss. Saeedeh Asari for her technical assistance.

References

  1. 1. Hussain TM, Hazara M, Sultan Z, Saleh BK, Gopal GR. Recent advances in salt stress biology a review. Biotechnology and Molecular Biology Reviews. 2008;3(1):8–13.
  2. 2. Qiu Q-S, Guo Y, Dietrich MA, Schumaker KS, Zhu J-K. Regulation of SOS1, a plasma membrane Na+/H+ exchanger in Arabidopsis thaliana, by SOS2 and SOS3. Proceedings of the National Academy of Sciences. 2002;99(12):8436–41.
  3. 3. Mahajan S, Tuteja N. Cold, salinity and drought stresses: an overview. Archives of biochemistry and biophysics. 2005;444(2):139–58. pmid:16309626
  4. 4. Flowers T, Yeo A. Breeding for salinity resistance in crop plants: where next? Functional Plant Biology. 1995;22(6):875–84.
  5. 5. Mäser P, Eckelman B, Vaidyanathan R, Horie T, Fairbairn DJ, Kubo M, et al. Altered shoot/root Na+ distribution and bifurcating salt sensitivity in Arabidopsis by genetic disruption of the Na+ transporter AtHKT1. FEBS letters. 2002;531(2):157–61. pmid:12417304
  6. 6. 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
  7. 7. Blumwald E. Sodium transport and salt tolerance in plants. Current opinion in cell biology. 2000;12(4):431–4. pmid:10873827
  8. 8. Munns R, Tester M. Mechanisms of salinity tolerance. Annu Rev Plant Biol. 2008;59:651–81. pmid:18444910
  9. 9. Jain M. Next-generation sequencing technologies for gene expression profiling in plants. Briefings in functional genomics. 2011;11(1):63–70. pmid:22155524
  10. 10. Wang Z, Gerstein M, Snyder M. RNA-Seq: a revolutionary tool for transcriptomics. Nature reviews genetics. 2009;10(1):57–63. pmid:19015660
  11. 11. Zhao S, Fung-Leung W-P, Bittner A, Ngo K, Liu X. Comparison of RNA-Seq and microarray in transcriptome profiling of activated T cells. PloS one. 2014;9(1):e78644. pmid:24454679
  12. 12. 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;6:27752. pmid:27293111
  13. 13. 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;6:21476. pmid:26892368
  14. 14. 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.
  15. 15. 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.
  16. 16. Dudnikov AJ, Kawahara T. Aegilops tauschii: genetic variation in Iran. Genetic Resources and Crop Evolution. 2006;53(3):579–86.
  17. 17. 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.
  18. 18. 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.
  19. 19. Clavijo BJ, Venturini L, Schudoma C, Accinelli GG, Kaithakottil G, Wright J, et al. An improved assembly and annotation of the allohexaploid wheat genome identifies complete families of agronomic genes and provides genomic evidence for chromosomal translocations. Genome research. 2017;27(5):885–96. pmid:28420692
  20. 20. Sefidab AA, Vahabzadeh M, Heravan EM, Akbari A, Afyoni D, Saberi M, et al. Arg, A New Bread Wheat Cultivar for Moderate Climate Zones of Iran with Salinity of Soil and Water. Seed and Plant Improvment Journal. 2012;28(4):724–6.
  21. 21. 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. pmid:22383036
  22. 22. Conesa A, Götz S, García-Gómez JM, Terol J, Talón M, Robles M. Blast2GO: a universal tool for annotation, visualization and analysis in functional genomics research. Bioinformatics. 2005;21(18):3674–6. pmid:16081474
  23. 23. 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.
  24. 24. 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
  25. 25. 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
  26. 26. 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
  27. 27. 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
  28. 28. 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
  29. 29. Peyronnet R, Tran D, Girault T, Frachisse J-M. Mechanosensitive channels: feeling tension in a world under pressure. Frontiers in plant science. 2014;5:558. pmid:25374575
  30. 30. Zhu J-K. Regulation of ion homeostasis under salt stress. Current opinion in plant biology. 2003;6(5):441–5. pmid:12972044
  31. 31. 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
  32. 32. 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
  33. 33. Geisler M, Frangne N, Gomès 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
  34. 34. 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
  35. 35. 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.
  36. 36. Geisler M, Axelsen KB, Harper JF, Palmgren MG. Molecular aspects of higher plant P-type Ca 2+-ATPases. Biochimica et Biophysica Acta (BBA)-Biomembranes. 2000;1465(1):52–78.
  37. 37. Laohavisit A, Shang Z, Rubio L, Cuin TA, Véry A-A, Wang A, et al. Arabidopsis annexin1 mediates the radical-activated plasma membrane Ca2+-and K+-permeable conductance in root cells. The Plant Cell. 2012;24(4):1522–33. pmid:22523205
  38. 38. 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
  39. 39. 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
  40. 40. 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
  41. 41. Harmon AC, Gribskov M, Harper JF. CDPKs–a kinase for every Ca 2+ signal? Trends in plant science. 2000;5(4):154–9. pmid:10740296
  42. 42. Weinl S, Kudla J. The CBL–CIPK Ca2+‐decoding signaling network: function and perspectives. New Phytologist. 2009;184(3):517–28. pmid:19860013
  43. 43. Albrecht BE, Schulz B, Harter K, Luan S, Bock R, rg Kudla J. Alternative complex formation of the Ca 2-regulated protein kinase CIPK1 controls abscisic acid-dependent and inde-pendent stress responses in Arabidopsis. The Plant Journal. 2006;48:857–72. pmid:17092313
  44. 44. Kim K-N, Cheong YH, Grant JJ, Pandey GK, Luan S. CIPK3, a calcium sensor–associated protein kinase that regulates abscisic acid and cold signal transduction in Arabidopsis. The Plant Cell. 2003;15(2):411–23. pmid:12566581
  45. 45. Pandey GK, Cheong YH, Kim B-G, Grant JJ, Li L, Luan S. CIPK9: a calcium sensor-interacting protein kinase required for low-potassium tolerance in Arabidopsis. Cell research. 2007;17(5):411–21. pmid:17486125
  46. 46. Virdi AS, Singh S, Singh P. Abiotic stress responses in plants: roles of calmodulin-regulated proteins. Frontiers in plant science. 2015;6.
  47. 47. Hirayama T, Shinozaki K. Research on plant abiotic stress responses in the post‐genome era: Past, present and future. The Plant Journal. 2010;61(6):1041–52. pmid:20409277
  48. 48. Balazadeh S, Kwasniewski M, Caldana C, Mehrnia M, Zanor MI, Xue G-P, et al. ORS1, an H 2 O 2-responsive NAC transcription factor, controls senescence in Arabidopsis thaliana. Molecular plant. 2011;4(2):346–60. pmid:21303842
  49. 49. He XJ, Mu RL, Cao WH, Zhang ZG, Zhang JS, Chen SY. AtNAC2, a transcription factor downstream of ethylene and auxin signaling pathways, is involved in salt stress response and lateral root development. The Plant Journal. 2005;44(6):903–16. pmid:16359384
  50. 50. Kawaura K, Mochida K, Ogihara Y. Genome-wide analysis for identification of salt-responsive genes in common wheat. Functional & integrative genomics. 2008;8(3):277–86.
  51. 51. Mahmood K, El-Kereamy A, Kim S-H, Nambara E, Rothstein SJ. ANAC032 positively regulates age-dependent and stress-induced senescence in Arabidopsis thaliana. Plant and Cell Physiology. 2016;57(10):2029–46. pmid:27388337
  52. 52. Li W-T, He M, Wang J, Wang Y-P. Zinc finger protein (ZFP) in plants-A review. Plant Omics. 2013;6(6):474.
  53. 53. Stracke R, Werber M, Weisshaar B. The R2R3-MYB gene family in Arabidopsis thaliana. Current opinion in plant biology. 2001;4(5):447–56. pmid:11597504
  54. 54. Afzal Z, Howton T, Sun Y, Mukhtar MS. The roles of aquaporins in plant stress responses. Journal of developmental biology. 2016;4(1):9.
  55. 55. Chinnusamy V, Jagendorf A, Zhu J-K. Understanding and improving salt tolerance in plants. Crop Science. 2005;45(2):437–48.
  56. 56. Tarczynski MC, Jensen RG, Bohnert HJ. Stress protection of transgenic tobacco by production of the osmolyte mannitol. Science. 1993;259(5094):508–10. pmid:17734171
  57. 57. Verslues PE, Agarwal M, Katiyar‐Agarwal S, Zhu J, Zhu JK. Methods and concepts in quantifying resistance to drought, salt and freezing, abiotic stresses that affect plant water status. The Plant Journal. 2006;45(4):523–39. pmid:16441347
  58. 58. Lee EK, Kwon M, Ko J-H, Yi H, Hwang MG, Chang S, et al. Binding of sulfonylurea by AtMRP5, an Arabidopsis multidrug resistance-related protein that functions in salt tolerance. Plant Physiology. 2004;134(1):528–38. pmid:14684837
  59. 59. Horie T, Sugawara M, Okada T, Taira K, Kaothien-Nakayama P, Katsuhara M, et al. Rice sodium-insensitive potassium transporter, OsHAK5, confers increased salt tolerance in tobacco BY2 cells. Journal of bioscience and bioengineering. 2011;111(3):346–56. pmid:21084222
  60. 60. Schroeder JI, Ward JM, Gassmann W. Perspectives on the Physiology and Structure of Inward-Rectifying K+ Channels in Higher Plants: Biophysical Implications for K Uptake. Annual review of biophysics and biomolecular structure. 1994;23(1):441–71.
  61. 61. Blumwald E, Poole RJ. Na+/H+ antiport in isolated tonoplast vesicles from storage tissue of Beta vulgaris. Plant physiology. 1985;78(1):163–7. pmid:16664191
  62. 62. 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:pp. 01472.2017.
  63. 63. Bowman MJ, Park W, Bauer PJ, Udall JA, Page JT, Raney J, et al. RNA-Seq transcriptome profiling of upland cotton (Gossypium hirsutum L.) root tissue under water-deficit stress. PLoS One. 2013;8(12):e82634. pmid:24324815
  64. 64. Mousavi S, Alisoltani A, Shiran B, Fallahi H, Ebrahimie E, Imani A, et al. De novo transcriptome assembly and comparative analysis of differentially expressed genes in Prunus dulcis Mill. in response to freezing stress. PloS one. 2014;9(8):e104541. pmid:25122458
  65. 65. 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
  66. 66. Kala S. Effect Of Nacl Salt Stress On Antioxidant Enzymes Of Isabgol (Plantago Ovata Forsk.) Genotypes. International Journal of Scientific & Technology Research. 2015;4(2):40–3.
  67. 67. 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.
  68. 68. 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
  69. 69. Wu J, Kim SG, Kang KY, Kim J-G, Park S-R, Gupta R, et al. Overexpression of a pathogenesis-related protein 10 enhances biotic and abiotic stress tolerance in rice. The plant pathology journal. 2016;32(6):552. pmid:27904462
  70. 70. Gulick PJ, Shen W, An H. ESI3, a stress-induced gene from Lophopyrum elongatum. Plant physiology. 1994;104(2):799. pmid:8159795
  71. 71. Zhang X, Takano T, Liu S. Identification of a mitochondrial ATP synthase small subunit gene (RMtATP6) expressed in response to salts and osmotic stresses in rice (Oryza sativa L.). Journal of experimental botany. 2005;57(1):193–200. pmid:16317034
  72. 72. Bhardwaj R, Sharma I, Kanwar M, Sharma R, Handa N, Kaur H, et al. LEA proteins in salt stress tolerance. Salt Stress in Plants: Springer; 2013. p. 79–112.
  73. 73. Yang X, Li H, Yang Y, Wang Y, Mo Y, Zhang R, et al. Identification and expression analyses of WRKY genes reveal their involvement in growth and abiotic stress response in watermelon (Citrullus lanatus). PloS one. 2018;13(1):e0191308. pmid:29338040
  74. 74. Yin W, Hu Z, Hu J, Zhu Z, Yu X, Cui B, et al. Tomato (Solanum lycopersicum) MADS-box transcription factor SlMBP8 regulates drought, salt tolerance and stress-related genes. Plant Growth Regulation. 2017;83(1):55–68.
  75. 75. Kavas M, Baloğlu MC, Atabay ES, Ziplar UT, Daşgan HY, Ünver T. Genome-wide characterization and expression analysis of common bean bHLH transcription factors in response to excess salt concentration. Molecular genetics and genomics. 2016;291(1):129–43. pmid:26193947
  76. 76. 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
  77. 77. Kanehisa M, Goto S. KEGG: kyoto encyclopedia of genes and genomes. Nucleic acids research. 2000;28(1):27–30. pmid:10592173
  78. 78. 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:BSR20160509.
  79. 79. Baxter HL, Stewart Jr CN. Effects of altered lignin biosynthesis on phenylpropanoid metabolism and plant stress. Biofuels. 2013;4(6):635–50.
  80. 80. Baba SA, Vishwakarma RA, Ashraf N. Functional Characterization of CsBGlu12, a β-Glucosidase from Crocus sativus, Provides Insights into Its Role in Abiotic Stress through Accumulation of Antioxidant Flavonols. Journal of Biological Chemistry. 2017;292(11):4700–13. pmid:28154174
  81. 81. 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.
  82. 82. 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
  83. 83. 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.
  84. 84. Balasubramanian S, Sureshkumar S, Lempe J, Weigel D. Potent induction of Arabidopsis thaliana flowering by elevated growth temperature. PLoS genetics. 2006;2(7):e106. pmid:16839183
  85. 85. Dinesh-Kumar S, Baker BJ. Alternatively spliced N resistance gene transcripts: their possible role in tobacco mosaic virus resistance. Proceedings of the National Academy of Sciences. 2000;97(4):1908–13.
  86. 86. Egawa C, Kobayashi F, Ishibashi M, Nakamura T, Nakamura C, Takumi S. Differential regulation of transcript accumulation and alternative splicing of a DREB2 homolog under abiotic stress conditions in common wheat. Genes & genetic systems. 2006;81(2):77–91.
  87. 87. Iida K, Go M. Survey of conserved alternative splicing events of mRNAs encoding SR proteins in land plants. Molecular biology and evolution. 2006;23(5):1085–94. pmid:16520337
  88. 88. Jordan T, Schornack S, Lahaye T. Alternative splicing of transcripts encoding Toll-like plant resistance proteins–what's the functional relevance to innate immunity? Trends in plant science. 2002;7(9):392–8. pmid:12234730
  89. 89. Kazan K. Alternative splicing and proteome diversity in plants: the tip of the iceberg has just emerged. Trends in plant science. 2003;8(10):468–71. pmid:14557042
  90. 90. Marquez Y, Brown JW, Simpson C, Barta A, Kalyna M. Transcriptome survey reveals increased complexity of the alternative splicing landscape in Arabidopsis. Genome research. 2012;22(6):1184–95. pmid:22391557
  91. 91. Zhang G, Guo G, Hu X, Zhang Y, Li Q, Li R, et al. Deep RNA sequencing at single base-pair resolution reveals high complexity of the rice transcriptome. Genome research. 2010;20(5):646–54. pmid:20305017
  92. 92. Ding F, Cui P, Wang Z, Zhang S, Ali S, Xiong L. Genome-wide analysis of alternative splicing of pre-mRNA under salt stress in Arabidopsis. BMC genomics. 2014;15(1):431.
  93. 93. Krasensky J, Jonak C. Drought, salt, and temperature stress-induced metabolic rearrangements and regulatory networks. Journal of experimental botany. 2012;63(4):1593–608. pmid:22291134
  94. 94. Singh M, Kumar J, Singh S, Singh VP, Prasad SM. Roles of osmoprotectants in improving salinity and drought tolerance in plants: a review. Reviews in Environmental Science and Bio/Technology. 2015;14(3):407–26.
  95. 95. Albacete A, Cantero-Navarro E, Großkinsky DK, Arias CL, Balibrea ME, Bru R, et al. Ectopic overexpression of the cell wall invertase gene CIN1 leads to dehydration avoidance in tomato. Journal of experimental botany. 2014;66(3):863–78. pmid:25392479
  96. 96. Wu Q, Yang J, Cheng N, Hirschi KD, White FF, Park S. Glutaredoxins in plant development, abiotic stress response, and iron homeostasis: From model organisms to crops. Environmental and Experimental Botany. 2017.
  97. 97. Basyuni M, Baba S, Inafuku M, Iwasaki H, Kinjo K, Oku H. Expression of terpenoid synthase mRNA and terpenoid content in salt stressed mangrove. Journal of plant physiology. 2009;166(16):1786–800. pmid:19535167
  98. 98. Laohavisit A, Richards S, Shabala L, Chen C, Colaco R, Swarbreck S, et al. Salinity-induced calcium signaling and root adaptation in Arabidopsis thaliana require the calcium regulatory protein annexin1. Plant Physiology. 2013:pp. 113.217810.