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

Genome-wide identification and characterization of gene family for RWP-RK transcription factors in wheat (Triticum aestivum L.)

  • Anuj Kumar ,

    Contributed equally to this work with: Anuj Kumar, Ritu Batra

    Roles Data curation, Formal analysis, Writing – original draft

    Affiliation Advance Center for Computational & Applied Biotechnology, Uttarakhand Council for Biotechnology (UCB), Dehradun, India

  • Ritu Batra ,

    Contributed equally to this work with: Anuj Kumar, Ritu Batra

    Roles Data curation, Formal analysis, Writing – original draft

    Affiliation Department of Genetics and Plant Breeding, CCS University, Meerut, India

  • Vijay Gahlaut,

    Roles Formal analysis

    Affiliation Department of Plant Molecular Biology, South Campus, University of Delhi, Delhi, India

  • Tinku Gautam,

    Roles Formal analysis

    Affiliation Department of Genetics and Plant Breeding, CCS University, Meerut, India

  • Sanjay Kumar,

    Roles Formal analysis

    Affiliation Bioinformatics Centre, Biotech Park, Lucknow, India

  • Mansi Sharma,

    Roles Formal analysis

    Affiliation ICMR- National Institute of Cancer Prevention and Research, Noida, India

  • Sandhya Tyagi,

    Roles Data curation

    Affiliation Division of Plant Physiology, ICAR-Indian Agricultural Research Institute, New Delhi, India

  • Krishna Pal Singh,

    Roles Formal analysis

    Affiliations Advance Center for Computational & Applied Biotechnology, Uttarakhand Council for Biotechnology (UCB), Dehradun, India, Ch. Charan Singh Haryana Agricultural University, Hisar, India

  • Harindra Singh Balyan,

    Roles Supervision, Writing – review & editing

    Affiliation Department of Genetics and Plant Breeding, CCS University, Meerut, India

  • Renu Pandey,

    Roles Methodology, Supervision, Writing – review & editing

    Affiliation Division of Plant Physiology, ICAR-Indian Agricultural Research Institute, New Delhi, India

  • Pushpendra Kumar Gupta

    Roles Supervision, Writing – review & editing

    pkgupta36@gmail.com

    Affiliation Department of Genetics and Plant Breeding, CCS University, Meerut, India

Abstract

RWP-RKs represent a small family of transcription factors (TFs) that are unique to plants and function particularly under conditions of nitrogen starvation. These RWP-RKs have been classified in two sub-families, NLPs (NIN-like proteins) and RKDs (RWP-RK domain proteins). NLPs regulate tissue-specific expression of genes involved in nitrogen use efficiency (NUE) and RKDs regulate expression of genes involved in gametogenesis/embryogenesis. During the present study, using in silico approach, 37 wheat RWP-RK genes were identified, which included 18 TaNLPs (2865 to 7340 bp with 4/5 exons), distributed on 15 chromosomes from 5 homoeologous groups (with two genes each on 4B,4D and 5A) and 19 TaRKDs (1064 to 5768 bp with 1 to 6 exons) distributed on 12 chromosomes from 4 homoeologous groups (except groups 1, 4 and 5); 2–3 splice variants were also available in 9 of the 37 genes. Sixteen (16) of these genes also carried 24 SSRs (simple sequence repeats), while 11 genes had targets for 13 different miRNAs. At the protein level, MD simulation analysis suggested their interaction with nitrate-ions. Significant differences were observed in the expression of only two (TaNLP1 and TaNLP2) of the nine representative genes that were used for in silico expression analysis under varying levels of N at post-anthesis stage (data for other genes was not available for in silico expression analysis). Differences in expression were also observed during qRT-PCR, when expression of four representative genes (TaNLP2, TaNLP7, TaRKD6 and TaRKD9) was examined in roots and shoots of seedlings (under different conditions of N supply) in two contrasting genotypes which differed in NUE (C306 with low NUE and HUW468 with high NUE). These four genes for qRT-PCR were selected on the basis of previous literature, level of homology and the level of expression (in silico study). In particular, the TaNLP7 gene showed significant up-regulation in the roots and shoots of HUW468 (with higher NUE) during N-starvation; this gene has already been characterized in Arabidopsis and tobacco, and is known to be involved in nitrate-signal transduction pathway.

Introduction

Nitrogen (N) is an essential element for plant growth, productivity and grain quality. It plays a vital role in various metabolic activities within the cell involving synthesis of a variety of macromolecules, such as nucleic acids, proteins, cofactors, chlorophyll and other molecules involved in signaling and storage [1,2]. Only 30% of available soil N is taken up by plant roots in the form of nitrate (NO3) and ammonium (NH4+) ions, amino acids and other organic molecules. The remaining 70% N is lost either through leaching into the soil or in the gaseous form into the atmosphere, the latter also causing environmental pollution. Therefore, efforts are being made to improve N use efficiency (NUE) of crop plants so that high yielding crops can be grown with low N-input without significant yield loss [3].

The NUE includes two major components, N uptake efficiency (NUpE) and N utilization (assimilation) efficiency (NUtE), although N transport and N remobilization (after assimilation) are two other components. Each of these components is controlled by a number of genes [2] including a family of TFs called RWP-RKs, so named due to the presence of a conserved RWP-RK motif [4]. These RWP-RK genes are ubiquitous in plants and have been classified in two sub-families, NLPs (NIN-like proteins) and RKDs (RWP-RK domain proteins). NLPs regulate tissue-specific expression of genes involved in nitrogen use efficiency (NUE) [47] and RKDs regulate expression of genes involved in gametogenesis/embryogenesis [4,8]. The RWP-RK proteins bind to cis-acting elements in the promoter regions of NUE-related genes [including nitrate reductase (NR/NIA1) and nitrite reductase (NiR1)] and the genes responsible for gametogenesis and embryogenesis [9]. Both NLPs and RKDs contain a common but dissimilar DNA binding RWP-RK domain [4,9], but NLPs also carry an additional domain called Phox/Bem1 (PB1), which is an octicosapeptide that allows interactions with additional proteins [4,5]. The N-terminal regions of NLPs respond to nitrate signals and bind specifically to the nitrate responsive elements (NREs) that are found in the promoter regions of nitrate-inducible gene loci [5,9,10].

The function of NLPs under conditions of different NO3 ion concentrations has been investigated in a number of plant species including Arabidopsis (A. thaliana), rice (Oryza sativa) and tobacco (Nicotiana tabacum) [4,6,9]. These studies demonstrated that NLPs control expression of nitrate inducible genes including those associated with NUE [47], and RKDs take part in N signaling [8]. It has been shown that the activity of NLPs is post-translationally modulated by nitrate signaling, so that the suppression of NLPs impairs the nitrate-inducible expression of various genes causing severe growth inhibition [6]. In non-nodulating plants, the most-studied member of RWP-RK family is NLP7, which is involved in nitrate-signal transduction pathway, thus regulating N assimilation [10].

In wheat (Triticum aestivum), the RWP-RK genes have never been subjected to a detailed study. During the present study, we identified and conducted a comprehensive study of 37 wheat RWP-RK genes (19 TaRKDs and 18 TaNLPs) and their encoded proteins. These genes were subjected to a systematic in silico analysis to study the gene structure and promoter sequences; the corresponding proteins were examined for the functional domains, conserved motifs, physicochemical properties, homology modeling, molecular docking, molecular dynamics (MD) simulations and phylogenetic relationships. The study also included in silico expression of these genes in different tissues at different developmental stages under varying levels of N supply. Gene ontology (GO) analysis was also conducted to determine the functions of these genes. Four representative genes (TaRKD6, TaRKD9, TaNLP2, TaNLP7) that were selected on the basis of level of expression (in silico), previous literature and level of homology, were also used for qRT-PCR using two contrasting wheat genotypes (C306 and HUW468), which differ for NUE. The study provides an insight into the family of wheat genes encoding RWP-RK TFs, both at gene and protein levels.

Materials and methods

Analysis of RWP-RK genes

Identification and analysis.

In order to obtain wheat gene sequences that encode RWP-RK TFs, sequences of 18 known RWP-RK genes from the model species Brachypodium distachyon (a closely related species to wheat) were used. Coding DNA sequences (CDSs) of B. distachyon genes were retrieved from Ensembl Plants (http://plants.ensembl.org/) and used in tBLASTx analysis against the recently released wheat genome assembly (IWGSC RefSeq v1.0) available on Ensembl Plants. HMMER tool (available at Ensembl Plants) was also used to retrieve additional genes. Following criteria were used for the identification of wheat orthologs [1113]: (i) high level (>60%) of sequence identity and query coverage along the protein length; (ii) presence of all the domains and motifs available in query sequences. Homoeologous relationships between genes were established on the basis of their chromosome assignment and percentage of protein sequence identity (>90%).

Analysis of the structure of wheat RWP-RK genes was conducted using Gene Structure Display Server (GSDS) v2.0 by comparing their full length CDSs with their corresponding genomic sequences [14]. Intron phases (phase 0, 1, 2) were identified following the criteria used by us earlier [13]. MapInspect (http://www.plantbreeding.wur.nl/UK/software_map-inspect.html) was used to physically map the genes onto individual wheat chromosomes. MCScanX was used to predict the segmental/tandem gene duplications. The values of synonymous (Ks) and non-synonymous (Ka) substitutions were obtained using CDS of wheat genes with respect to CDS of the corresponding Brachypodium genes. One kb genomic region upstream of the translation start site (ATG) (i.e. promoter region) of each gene was analysed for the presence of cis-regulatory response elements using PlantCARE database [15]. Only the response elements on the sense strand showing a matrix value of ≥5 were accepted. Simple sequence repeats (SSRs) were identified within the gene sequences using BatchPrimer3v1.0 (http://probes.pw.usda.gov/batchprimer3/). The miRNAs and their targets in RWP-RK genes were predicted employing web-based psRNATarget server [16] using default parameters.

Synteny and collinearity analyses.

Synteny/collinearity analysis was conducted by comparing a block of 21 genes associated with TaRKD and TaNLP genes (10 genes on either side of the gene of interest) with Brachypodium, rice and sorghum employing URGI database (http://wheat-urgi.versailles.inra,fr/) The results of the above analysis were visualized using the program circos version 0.67 [17].

In silico gene expression profiling and hierarchical clustering

In silico gene expression profiling was based on the assumption that homoeologous genes shared common probe ids. Based on this property, we selected one candidate probe id for each group of homoeologous genes. The selected probe ids were retrieved from the “Affymetrix wheat 61K microarray” platform using the PLEXdb interface [18]. The expression analysis was carried out for the following genes using microarray data in Genevestigator platform [19]: (i) 9 TaRKD (from 3 homoeologous groups) and 18 TaNLP genes (from 6 homoeologous groups) in 15 different tissues, 10 developmental stages and in leaves under varying levels of N availability at post-anthesis, and (ii) wheat orthologs of 4 downstream genes (including 4CL1, CHX17, SIR and AMP dependent synthetase) induced by AtNLP7 in Arabidopsis under varying levels of N. The expression profiles of genes identified from wheat microarray data were used for construction of the heat map using hierarchical clustering tool embedded in Genevestigator platform.

Analysis of RWP-RK protein sequences

Major domains in the predicted protein sequences of wheat genes belonging to RWP-RK family were identified through conserved domain (CD)-search program of conserved domain database (CDD) at NCBI. Manual search was done for RWPXRK domain, which is known to be a characteristic feature of RWP-RK TF family in plants [4]. Physicochemical properties of all the predicted proteins were studied using ProtParam server (https://web.expasy.org/protparam/). The sub-cellular location of wheat genes was predicted using a user-friendly web-server PLANT-mPLoc [20].

3D structure, its evaluation and optimization.

The three dimensional (3D) structures of predicted proteins were deduced using homology modeling [21]. For this purpose, PSI-BLAST was first carried out against protein data bank (PDB) (http://www.rcsb.org/pdb/home/home.do) and swissProt template library (STL) (http://swissmodel.expasy.org/workspace/index.php?func = tools_smtl) to find out the suitable homologous templates (crystal and NMR 3D structures) on the basis of maximum identity, maximum score and minimum e-value. The best template for each gene from PSI-BLAST was selected in SwissModel server to generate the 3D structures of the predicted proteins [22]. The geometric evaluation of the predicted 3D structures of proteins for different genes was performed using PROCHECK and protein structure verification server (PSVS) (htttp://nihserver.mbi.ucla. edu/SAVES/). Ramachandran plots were prepared through calculation of phi (Φ) and psi (ψ) torsion angles.

Alignment of 3D structures of predicted wheat proteins over 3D structures of query proteins.

To check the correct topology of 3D structures of predicted proteins of wheat genes, the generated structures of encoded proteins predicted for one representative gene each belonging to RKD and NLP sub-families of RWP-RK genes were aligned with the 3D structures of proteins encoded by respective query genes belonging to Brachypodium using FATCAT server [23]. The similarity of the generated 3D structures in a globally optimized superimposition environment was measured by comparing RMSD value of the Cα atoms of the generated structures to those of the corresponding 3D structures of the query genes. The function of RWP-RK (as TFs) at the biochemical level was predicted from the 3D structures of proteins using ProFunc server [24].

Molecular docking of 3D structures of the predicted wheat proteins with nitrate ions.

The predicted 3D protein structures of representative wheat proteins TaRKD6-2A and TaNLP7-3A were used for docking studies, since these proteins showed maximum amino acids in the favoured regions in Ramachandran plots (S5 Table). Structure of nitrate ion (CID_94310) was retrieved from PubChem database in SDF format. To find out the interacting residues between 3D protein structures and nitrate ion, docking studies were performed by Surflex Dock program available in SYBYL-X (https://www.certara.com/software/molecular-modeling-and-simulation/sybyl-x-suite/). For docking analysis, hydrogen atoms were added to the predicted 3D protein structures (TaRKD6-2A and TaNLP7-3A) using AMBER7 FF99 charges. The 3D structures, thus obtained, were optimized by applying 100 steps of Powell method and Conjugate gradient algorithm with Tripos force field. Grid generation was done by automated mode of Protomol tool on active regions of both 3D protein structures. A maximum of 20 conformations of nitrate-ion with both the 3D protein structures were generated. On the basis of scoring function, top scoring conformations were selected for protein stability through MD simulations [25,26].

MD simulations.

The MD simulation analysis was conducted using the Desmond v4.8 suite available in Schrodinger-Maestro v11. Energy minimization (100 steps steepest descent followed by 2000 steps conjugate gradient) was undertaken before initiating the MD simulation to remove initial steric clashes. Fourteen (14) Na+ ions were added to each 3D structure and nitrate ion complex with salt atoms that maintain the system at charge neutrality. The system was then solvated using SPC water molecules within Triclinic box in a periodic box condition. Both 3D protein and nitrate complexes had at least 5Å buffer in every direction of the box to permit substantial fluctuations of the conformation during the course of the MD simulation. The complete interaction energy was also calculated [27,28]. The constant pressure during MD simulation was calculated using anisotropic diagonal position scaling. The time-step used was 0.002 ps. The temperature of the system was increased gradually from 100 K to 300 K with 20 ps NPT reassemble. The target pressure was 1 atm. The Berendsen algorithm was used with a scaling factor with time constant of 0.2. The Lennard-Jones cutoff value used was 8Å. SHAKE constraints were applied to all bonds involving hydrogen atoms. Finally, 30 nsec MD simulations were run under the same conditions as the equilibration procedure. The density of the system was maintained near 1 g/cm3. OPLS v2007 force field was used in all calculations.

Phylogenetic analysis

A phylogenetic tree was prepared using MEGA version 6.0 [29] employing the neighbour-joining method. For this purpose, proteins encoded by 37 RWP-RK proteins from wheat, 17 RWP-RK proteins from Brachypodium, 15 RWP-RK proteins from rice and 14 RWP-RK proteins from Arabidopsis were utilized. Initially, all the protein sequences were aligned by multiple sequence alignment (MSA) using ClustalX server 2.1 [30] and then the aligned file was uploaded in MEGA version 6.0 to generate a phylogenetic tree. Bootstrap values for the phylogenetic tree were calculated as percentage of 1000 iterations. The evolutionary distances (expressed as number of aa differences per site) were computed using the method suggested by Nei and Kumar [31].

qRT-PCR analysis

Plant growing conditions.

Seeds of contrasting wheat genotypes, namely C306 (low NUE) and HUW468 (high NUE), were surface sterilized with 0.1% HgCl2 for 2 min followed by 5–6 washings with distilled water. The seedlings were raised in hydroponic solution under controlled conditions at National Phytotron Facility, ICAR-IARI, New Delhi following the method described earlier [32]. Five days after germination, the seedlings were transferred to plastic containers (10 L capacity) in Hoagland solution with low (10 μM) and optimum (7.5 mM) N concentrations. The nutrient solution was changed on every third day in each treatment throughout the experiment. Fresh leaf samples for DNA were collected from 21, 24 and 25 days old seedlings of the control (optimum N) treatment. Samples were also collected from seedlings grown in low N on 21 days. In another set, after 21 days of growth at low N, the plants were completely starved of N (N starvation) and were grown for next three days with no supply of N. Root and shoot samples were collected on third day (24th day) of N starvation. Another batch of N starved seedlings was re-supplied with optimum N; root and shoot samples were then collected after 24 h (25th day). Two replications were used for each treatment in each experiment.

RNA isolation, cDNA synthesis, primer design and qRT-PCR analysis.

Total RNA was isolated from root and shoot tissue using a TRI reagent (Sigma) followed by RNase-free DNase I (Qiagen) treatment for removal of DNA contamination. Reverse transcription reactions were performed using 2.0 μg of total RNA and M-MuLV Reverse Transcriptase (Promega) according to the manufacturer’s instructions.

Primers for the four representative genes including two NLP and two RKD genes (TaNLP2 and TaNLP7, TaRKD6 and TaRKD9) were designed using Primer3 software (S1 Table). qRT-PCR was performed with PikoReal Real-Time PCR Systems (Thermo Scientific) using PowerUp SYBR Green Master Mix (Applied Biosystems) in three technical replicates per biological replicate. The reactions were carried out according to the following conditions: 95°C for 30 sec, 40 cycles of 95°C for 5 sec, and 60°C for 34 sec. Constitutive expression of TaAct2 gene of wheat was used as endogenous control. The transcript abundance for each gene was normalized with the internal control, and 2−ΔΔCt values (fold change) for gene expression under low N (LN), N starvation (NS) or N replete (NR) conditions vs. the control were calculated as follows: 2−ΔΔCt = [(CtLN/NS/NR test − CtLN/NS/NR TaAct) − (Ctcont test − Ctcont TaAct)] [33]. A negative control was also incorporated for each primer pair.

Statistical analysis of qRT-PCR results.

The results of qRT-PCR for expression of each of the four genes separately in roots and shoots were subjected to analysis of variances (ANOVA); for this purpose the means of two replications were used, since replications did not differ. Significance of variances over time (different durations) and space (root and shoot) and also the effect of genetic background (two cultivars) were tested for significance.

Results and discussion

Identification, chromosomal assignment and structure of TaRKD and TaNLP genes

In wheat, we identified 37 RWP-RK genes that were orthologous to only 13 of the 18 Brachypodium genes that were used as queries. The wheat genes corresponding to the remaining five Brachypodium genes must have been lost during evolution or may be available in wheat genotypes, other than Chinese Spring (CS) for which whole genome sequence was used in the present study (Table 1).

thumbnail
Table 1.   A list of RKD and NLP genes of Brachypodium and their putative orthologs in wheat.

https://doi.org/10.1371/journal.pone.0208409.t001

Alternate splicing, which is of common occurrence in plant cells [34] was also examined during the present study involving RWP-RK genes; 2–3 splice variants were available in nine of the 37 genes, the remaining 28 genes each producing a single transcript. Of the nine genes with splice variants, two TaRKD genes and two TaNLP genes each gave three splice variants, while five TaNLP genes produced two splice variants each (Table 1). The occurrence of splice variants was not a surprise, because according to some estimates, >60% of intron-containing genes in plants undergo alternative splicing (AS) producing splice variants. AS has been shown to play an important role in plant growth, development, and responses to external cues [3436].

Triplicate homoeologues in wheat.

The 19 TaRKD genes (with duplicate genes on 2A, 2B and 2D) were orthologs of 7 BdRKD genes, and were distributed on 12 wheat chromosomes belonging to four homoeologous groups [2, 3, 6 and 7; 7A carried four genes and 2A, 2B, 2D (each as tandem repeats) and 7D carried two genes each]. Similarly, the 18 TaNLP genes (orthologs of 6 BdNLP genes) belonged to only five of the seven homoeologous groups, there being no gene on homoeologous groups 1 and 7, and there being two genes each on 4B, 4D and 5A (Table 1 and and Fig 1). The wheat genes and Brachypodium genes were largely present on corresponding homoeologous chromosomes, even though the number of chromosomes in Brachypodium is n = x = 5 (analogous to maize with n = 2x = 10) as against x = 7 in wheat. Three homoeologous groups (2, 3 and 6) carried both RKD and NLP genes of wheat; homoeologous groups 4 and 5 carried only TaNLP genes, while homoeologous group 7 carried only TaRKD genes; the homoeologous group 1 carried no RWP-RK genes.

thumbnail
Fig 1. Distribution of 37 wheat TaRWP-RK genes (19 TaRKD and 18 TaNLP) on 18 chromosomes belonging to six homoeologous groups.

Chromosomes are represented by blue solid vertical bars. TaRKD genes are written in black and TaNLP genes are written in red colour.

https://doi.org/10.1371/journal.pone.0208409.g001

If we assume that we identified all available wheat TaRWP-RK genes during our study, we should have obtained 54 wheat genes against 18 Brachypodium genes; even against 13 Brachypodium genes, which were found to have RWP-RK orthologs, 39 genes were expected but we were able to identify only 37 wheat genes (along with tandem duplicates), because not more than two genes were available for each of the three Brachypodium genes (BdRKD1, BdRKD10 and BdRKD11). The missing genes in wheat included many more RKD genes (corresponding to 7 Brachypodium RKD genes), and relatively fewer NLP genes (corresponding to one Brachypodium gene BdNLP6). These missing genes must have been lost or diverged significantly during evolution of wheat. The data also suggest that the NLP genes are relatively more conserved than RKD genes. As mentioned earlier, NLPs play an important role in the regulation of genes involved in nitrate assimilation and other metabolic/regulatory processes associated with NUE, while RKDs play an important role in gametogenesis and embryogenesis [410,37]. Additional functions and networks involving these RKD and NLP genes may be discovered in future.

It is now well established that Brachypodium chromosomes 1, 2 and 3, each belongs to more than one homoeologous groups of wheat, while chromosomes 4 and 5 correspond to homoeologous groups 5 and 2, respectively. Since wheat genes are largely triplicate in nature, the three genes corresponding to an individual Brachypodium gene should be located on three wheat chromosomes of the same homoeologous group. There were following exceptions to this expectation: (i) TaRKD1-7A, which corresponds to BdRKD1 located on Brachypodium chromosome 2 (homoeologous groups 1 and 3), and (ii) four genes [TaRKD11-7A, which corresponds to BdRKD11 and TaNLP3-4A, TaNLP3-4B and TaNLP3-4D, which correspond to BdNLP3] correspond to Brachypodium chromosome 4 (wheat homoeologous group 5) (see Table 1). A suitable explanation for this discrepancy is not available at present, although this may turn out to be due to possible cryptic translocations involving non-homoeologous wheat chromosomes of groups 1, 3, 4, 5 and 7. Such a situation has been reported while studying the collinearity of the Sh2/A1 orthologous region in rice, sorghum, maize and species of Triticeae [38]. The Sh2, X1, X2 and A1 genes are located on rice chromosome 1 and maize chromosome 3, which are homoeologous to group 3 chromosomes of wheat. Although the genes X2 and A1 have maintained a syntenic position on homoeologous chromosomes in wheat, maize, sorghum and rice, the other two genes (Sh2 and X1) are mapped on chromosome 1A of T. monococcum and Ae. tauschii, 1H of barley and group 1 chromosomes of wheat. The above transfer of Sh2 and X1 genes onto the non-homoeologous group 1 chromosomes in Triticeae species has been attributed to translocation or transposition [38]. In fact, frequent micro-rearrangements between Triticeae and rice have been documented [39] suggesting a need for caution in comparative genomics studies involving model plant systems such as Arabidopsis, rice and Brachypodium.

There are also examples, where three wheat genes corresponding to the same Brachypodium gene belong to more than one homoeologous groups. For instance, the three genes that correspond to gene BdNLP1 belong to homoeologous groups 4 (TaNLP1-4B and TaNLP1-4D) and 5 (TaNLP1-5A). This is not unexpected, because occurrence of a historical translocation between chromosome 4A and 5A has been reported [40], which was also confirmed recently through sequence-based analysis [41]. Known homoeology of Brachypodium chromosome 1 carrying BdNLP1 to four wheat homoeologous groups (2, 4, 5 and 7) also explains this anomaly. We also investigated the occurrence of possible tandem and segmental duplication of wheat RWP-RK genes. A comprehensive gene duplication analysis showed that one pair of TaRKD genes each were arranged in tandem duplication on group two chromosomes i.e. chromosome 2A (TaRKD6a-2A and TaRKD6b-2A), chromosome 2B (TaRKD6a-2B and TaRKD6b-2B) and chromosome 2D (TaRKD6a-2D and TaRKD6b-2D) (Fig 1). Similar tandem duplication events have also been reported in mitogen activated protein kinase kinase kinase (MAPKKK) gene family in wheat. This kind of duplication events are known to contribute towards the expansion of gene families in wheat and other species [42].

Gene lengths.

The length of individual TaRKD genes ranged from 1064 to 5768 bp, while that of TaNLP genes ranged from 2865 to 7340 bp, suggesting that TaNLP genes are little longer, perhaps due to the presence of an additional domain PB1. Accordingly, the coding DNA sequence (CDS) is also longer in TaNLPs (1680 to 2817 bp) than in TaRKDs (615 to 2784 bp). The lengths of wheat genes and the query genes from Brachypodium also varied, such that the length of BdRKD genes

(989 to 6480 bp) and BdNLP genes (3376 to 8010 bp) had a relatively longer range than the corresponding wheat genes. However, the range of the lengths of the CDS of the Brachypodium genes and the putative wheat genes was opposite to the range of the length of the gene sequences of the corresponding genes (Table 2). This may be attributed to the difference in size of the introns and untranslated regions (UTRs) of the genes belonging to Brachypodium and wheat. The average percent similarity of the CDS (73.84% for RKDs and 84.13% for NLPs) was higher than the average percent similarity of the gene sequences (64.18% for RKDs and 73.99% for NLPs) between wheat and Brachypodium. This suggested greater conservation of exonic sequences in the two species (Table 2). The ratio of non-synonymous and synonymous substitutions (Ka/Ks) of CDS of wheat genes with respect to CDS of the corresponding Brachypodium genes was >1 (1.094 to 1.722 for RKDs and 1.107 to 1.793 for NLPs) except for TaRKD9, TaRKD10 and TaNLP2 with Ka/Ks value of 0.120, 0.897 and 0.899, respectively. The >1 value of Ka/Ks indicates that almost all genes evolved under positive selection, which contributed towards speciation whereas TaRKD9, TaRKD10 and TaNLP2 (Ka/Ks ratio <1) genes have been subjected to purifying selection that did not alter the encoded amino acid sequence during 30–40 million years period of speciation [43].

thumbnail
Table 2. Details of CDSs and gene sequences for Brachypodium and wheat along with their percent similarity.

RWP-RK wheat genes with targets for miRNAs are also shown.

https://doi.org/10.1371/journal.pone.0208409.t002

Distribution of exons and introns.

The number of exons in 19 TaRKD genes varied from 1 to 6 (Fig 2), although the number of exons in TaRKD genes belonging to the same homoeologous group did not differ much (3–5 exons) except those belonging to homoeologous groups 2 and 7, which contained 1 to 6 exons. The number of exons in 18 TaNLP genes was generally 4 or 5. As many as 11 genes had 5 exons each and 6 genes had 4 exons each. The only exception was TaNLP7-3B on chromosome 3B with 6 exons (Fig 2).

thumbnail
Fig 2. Structure of TaRKD and TaNLP genes showing distribution of exons (solid red bars), introns (black lines), upstream/downstream regions (solid blue bars) and intron phases marked as 0, 1 and 2.

https://doi.org/10.1371/journal.pone.0208409.g002

The number of exons in corresponding Brachypodium genes differed; 2–6 exons were present in BdRKD genes and 3–5 exons were present in the BdNLP genes. This also suggests higher level of variation in the number of exons in genes belonging to RKD and NLP sub-families in wheat than in Brachypodium. This difference in variation of the number of exons in the genes of the two species could be attributed to deletion, addition, and merging of exons. A comparison of number of exons in the genes belonging to wheat and Brachypodium would suggest that in case of each of the two RKD genes [TaRKD3-7A and TaRKD3-7D], all the exons merged together resulting into solitary exons. Similar results showing merging of adjacent exons in AGPase LS gene of Arabidopsis, chickpea and potato and AGPase SS gene of Arabidopsis were reported by us earlier [13]. Also, an addition of an exon was noticed in TaRKD3-7B, so that the total number of exons in this case is 6 against 5 exons in the corresponding BdRKD3 gene of Brachypodium. Similarly, addition of 1 to 2 exons was noticed in seven TaNLP genes (TaNLP2-5s, TaNLP3-4s and TaNLP7-3B) (Fig 2).

Intron phases were also examined in all the TaRWP-RK genes. In TaRKD genes, the intron phase 1 was more prevalent (37.93% in each) followed by intron phase 0 (36.22%) and intron phase 2 (25.86%). In TaNLP genes, the intron phase 0 (59.70% in each) was most frequent followed by intron phase 1 (40.29%), once again suggesting that TaNLP genes are more conserved than TaRKD genes (Fig 2). Prevalence of intron phase 1 in TaRKD genes suggested that the wheat genes belonging to this sub-family have undergone rapid evolution, as suggested earlier [44].

Simple sequence repeats (SSRs) in TaRKD and TaNLP genes

As many as 24 SSRs were detected in 16 of the 37 genes, including 7 SSRs in 7 TaRKD genes and 17 SSRs in 9 TaNLP genes. Most genes carried each a single SSR except TaNLP1-4B and TaNLP5-6A, each carrying two SSRs, and TaNLP1-4B carrying three SSRs (S2 Table). Tri-nucleotide repeats were most frequent (14 SSRs) followed by hexa-nucleotide repeats (6 SSRs) and tetra -nucleotide repeats (4 SSRs). The abundance of tri-nucleotide repeat SSRs in the present study is in agreement with earlier reports in wheat [45]. SSRs have also been reported in genes encoding other TFs in chickpea [46] and M. truncatula [47]. In future, the polymorphism for SSRs in TaRKD and TaNLP TF genes may be examined in wheat cultivars and the polymorphic SSRs may be utilized for developing markers to be used for MAS in wheat breeding programmes for improvement of NUE in wheat.

Micro RNA (miRNA) and their targets in TaRKD and TaNLP genes

The 37 TaRWP-RK genes were also examined for miRNA targets. Six of the 19 TaRKD genes and five of the 18 TaNLP genes carried targets for 13 different miRNAs; 12 of these 13 miRNAs are being reported in wheat for the first time, the only exception being tae-miR414 (see Table 2 for details). In particular, tae-miR414 targets TaNLP3-4B and TaNLP3-4D, although it is known to target six other wheat genes, which encode proteins that are either involved in metabolic (nucleosome assembly protein I and ATPase subunit 6) and developmental processes (differentially expressed in relation to the extent of cell elongation; nuclear polyadenylated RNA-binding protein NAB3) or represent TFs [TFIIA large subunit (TFIIA-L1) and other DNA-binding protein-like protein] [48]. The miR414 is also known to play important roles in some other species. For instance, it plays a role in N-metabolism in saffron (Crocus sativus L.) [49], drought tolerance in P. patens [50] and plant development in Stevia rebaudiana [51]. Another important tae-miRNA for which targets were found in RWP-RK genes is tae-miR838, which has been reported to have a target site in DCL1 gene of Medicago truncatula, where it expresses in root nodules and thus plays a role in biological nitrogen fixation [52]. These results clearly demonstrate that besides their other functions, miRNAs are also involved in regulation of the expression of genes encoding transcription factors like RWP-RKs [53]. The present study suggests that the expression of tae-miR414 and tae-miR838 in particular may be further examined for their role in regulating the expression of RWP-RK genes in wheat in low-N tolerant and low N-sensitive genotypes or in a genotype responding differently under low and high N-availabilities. This will perhaps help further in manipulation of the regulatory pathways for development of wheat genotypes with improved NUE [3].

Synteny (gene content) and collinearity (gene order)

Synteny and collinearity analysis was undertaken using a block of 21 genes, including 10 genes flanking either side of each of the 19 TaRKD and 18 TaNLP genes. For this purpose, RKD and NLP genes from wheat, Brachypodium, rice and sorghum were utilized, since the data in URGI was available for only these four plant species (Fig 3). Among 37 RWP-RK genes, 9 TaRKD genes showed 10 to 70% synteny and 15 TaNLP genes showed 10 to 100% synteny with corresponding genes in three other genomes. The 9 TaRKD genes showed maximum average percent synteny with Brachypodium genes (53.5%), followed by sorghum genes (50.0%) and rice genes (47.5%). For 15 TaNLP genes, the average synteny did not differ much among three species; 9 TaNLP genes showed mean syntney of 58.1% with Brachypodium genes, 14 TaNLP genes showed mean syntney of 57.1% with rice genes and 12 TaNLP genes showed mean syntney of 56.8% with sorghum genes. The remaining TaRWP-RK genes did not show any synteny with the corresponding genes of the other three species. The collinearity within the synteny blocks for each of the TaRKD and TaNLP genes was disrupted due to insertion, deletion, duplication, and rearrangement of genes (S1 Fig). In the past, the loss of shared synteny and collinearity of some genes across grasses was attributed to gain/loss of genes and other structural changes within chromosome segments [54].

thumbnail
Fig 3. The circos visualization map showing synteny and collinearity among RWP-RK proteins of wheat, Brachypodium, rice and sorghum.

The bands of different colours with variation in thickness depict different levels of synteny between wheat proteins with those of other species.

https://doi.org/10.1371/journal.pone.0208409.g003

Promoter analysis of TaRKD and TaNLP genes(Identification of cis-elements.)

Promoter analysis allowed identification of cis-acting response elements associated with different TaRKD and TaNLP genes. These cis-elements presumably allow spatial and temporal expression of these TF genes, thus indirectly regulating the expression of other genes. Most of these cis-acting response elements were conserved; only in some genes, the basic regulatory elements like TATA-box and CAAT box were absent (TATA box was absent in TaRKD4-6D, TaRKD6a-2B, TaRKD6b-2A, TaRKD6b-2B, TaRKD10-7s, TaRKD11-7A, TaNLP1-5A, TaNLP1-4B, TaNLP1-4D; CAAT box was absent in TaRKD3-7D, TaRKD11-7A, TaNLP3-4D) (Table 3 and S3 Table for details). However, response elements for zein regulatory metabolism and for circadian rhythm were present in TaRKD genes on homoeologous groups 2 (TaRKD6a-2s, TaRKD6b-2A, TaRKD6b-2D), chromosome 7 (TaRKD11-7A) and chromosomes of homoeologous group 3 (TaRKD9-3A and TaRKD9-3D); the response elements for circadian rhythm were also present in TaNLP genes on homoeologous group 6 (TaNLP5-6A and TaNLP5-6D).

thumbnail
Table 3. A summary of cis-acting regulatory elements for abiotic stress in promoters of TaRKD and TaNLP genes of wheat.

https://doi.org/10.1371/journal.pone.0208409.t003

The presence of nitrogen responsive elements (NREs) in the promoter regions of TaRKD3-7D and TaNLP4-2D suggest the role of these genes under N-starvation. One such element is GCN4 motif [ATGA (C/G) TCAT], which is the binding site for the TF GCN4, whose synthesis is regulated at the translational level. In a study in barley, GCN4 motif was shown to function as a negative regulator for hordein genes under low N condition [55].

Another 13 different response elements for abiotic stress were available in the promoters of several TaRKD and TaNLP genes. These motifs (ABRE, GARE and MBS) are known to play a role in response to heat and drought stress [56], but their role in response to N-starvation (an abiotic stress) is yet to be examined. The tissue-specific and light-responsive elements may be responsible for expression of RKD/NLP genes in specific tissues and during plant development [57,58]. This expression of RKD/NLP genes encoding TFs should in turn induce expression of other structural genes encoding proteins that may take part in actual metabolic processes (see next section).

In silico expression of TaRKD and TaNLP genes

Spatial and temporal expression profiles of three TaRKDs and six TaNLPs (each representing three homoeologues; thus making 9 TaRKD and 18 TaNLP genes) were examined in 15 different tissues and at 10 different developmental stages. The expression was also compared under variable doses of N. For the remaining TaRKD genes, no in silico expression analysis could be performed, as no information for these genes was availbale in Genevestigator database.

Expression in different tissues.

In silico expression of 27 wheat RWP-RK genes (three homeologos each of 3 RKD and 6 NLP genes) in 15 different wheat tissues differed (S2 Fig). Similar results were earlier reported in Arabidopsis and rice [4]. TaRKD6, TaRKD9, TaNLP2, and TaNLP3 showed relatively higher but variable expression in all the 15 tissues except in few cases, where the expression was low. The results of several earlier studies are presented with our own results (Table 4). These results confirm the role of TaRKDs in embryogenesis and gametogenesis [59,60]. Similarly, TaNLPs have a role in the development of reproductive organs and roots [4,61].

thumbnail
Table 4. Results of expression analysis obtained during present study and earlier studies.

https://doi.org/10.1371/journal.pone.0208409.t004

In silico expression at different developmental stages.

The expression profiles of different TaRKDs and TaNLPs at 10 different development stages in wheat showed their variable expression (S3 Fig). Following three genes had relatively higher expression at some specific developmental stages: (i) TaNLP2 at booting, tillering and stem elongation stages, (ii) TaRKD9 at booting and germination stages, and (iii) TaRKD6 at the time of booting, germination, ripening, dough and milk developmental stages. Similar results were also reported for ZmNLP2 in maize [61]. Taken together, these results suggested developmental and tissue specific expression of TaRKDs and TaNLPs at different developmental stages.

In silico expression under variable N inputs.

The expression of different TaRKD and TaNLP genes was examined under variable doses of N using data available in Genevestigator. None of the TaRKD genes showed significant differential expression (fold change ≥2 and p value 0.05) under variable doses of N supply. However, in Chlamydomonas, MID gene containing RWP-RK motif (a characteristic of the wheat RKD genes) was shown to play a role in gamete formation under N starvation [8]. Similar studies need to be conducted in wheat and other eukaryotes also [4].

In contrast to TaRKDs, two TaNLPs (TaNLP1, and TaNLP2) exhibited up-regulation at low doses of N (50 kg/ha) relative to that at higher doses of N (200 kg N/ha; Fig 4A and 4B). These results are in agreement with those reported in an earlier study in Arabidopsis, where AtNLP5, AtNLP8 and AtNLP9 showed relatively higher expression during N-starvation [4]. Thus, it appears that the low doses of N induce higher expression, while high doses inhibit expression of NLP genes. However, these in silico results need to be validated in wet-lab studies in wheat.

thumbnail
Fig 4. Results of in silico expression analysis of TaRKD and TaNLP genes of wheat under varying doses of nitrogen in four different experiments.

(a) Experiment (1)-(3) showing expression at 21 daa under 200 kg N/ha vs. 50 kg N/ha (used as control),using genotype Istabraq in 1st, Hereward in 2nd and Soissons in 3rd, and experiment (4) showing expression at 7 daa under192 kg N/ha vs. 48 kg N/ha (used as control), and (b) Experiment (1)-(2) showing expression at 7 daa at 200 kg N/ha vs. 50 kg N/ha (used as control) and experiment (3)-(4) showing expression at 21 daa under 200 kgN/ha vs. 50 kg N/ha (used as control) using genotype Riband in 1st and 3rd and Soissons in 2nd and 4th. Genes showing significant differential expression are indicated with red colour.

https://doi.org/10.1371/journal.pone.0208409.g004

In contrast to our own results, where no other genes except TaNLP1 and TaNLP2 showed differential expression, in several other studies, NLP7 was reported to be the most important NLP gene. For instance, under condition of low dose of nitrate, AtNLP7 induced differential expression of a large number of genes including nitrate transporter (NRT2.1), nitrate reductase (NIA1), nitrite reductase (NiR) and glutamine synthetase 2 (GS) [10,62]. Also, nlp7 mutants (but no other NLP mutant) in Arabidopsis exhibited abnormal phenotype (rosette structure, delayed growth and flowering) under N starvation. We also identified 13 wheat orthologs of a number of downstream Arabidopsis genes that are differentially expressed; only four downstream genes (4CL1, CHX17, SIR and AMP dependent synthetase) showed significant differential expression under variable doses of nitrogen (data not presented). Since, TaNLP7 did not show significant variation in expression in the wheat microarray data set used by us during the present study, the role of TaNLP7 TF in regulation of the expression of a number of downstream wheat genes needs to be further examined.

Characterization of proteins

The lengths of predicted proteins for 19 TaRKD genes were more variable (204 aa for TaRKD4-6B to 927 aa for TaRKD3-7B) than the predicted proteins for the 18 TaNLP genes (559 aa for TaNLP5-6B to 938 aa for each of the three TaNLP4 genes on group 2 chromosomes) (Table 5). The same pattern is available in BdRKD (127aa to 609 aa) and BdNLP proteins (742 aa to 953 aa). Also, relative to TaRKDs (52.99 to 81.73, mean 66.74), TaNLPs had higher similarity (73.30 to 89.78; mean 81.31) with corresponding Brachypodium proteins, once again suggesting higher level of conservation of NLPs (also inferred above at the gene level).

thumbnail
Table 5. Details of Brachypodium and wheat RWP-RK proteins sequences and their conserved domains (with per cent similarity of wheat proteins with Brachypodium proteins).

https://doi.org/10.1371/journal.pone.0208409.t005

Functional domains.

More information is available for RKD domain (46-49aa) in both RKD and NLP proteins relative to NLP’s PB1 domain (71-93aa) that is separated from RKD domain by 127-273aa (Table 5). The RKD domain was present in 17 of the 19 TaRKDs (except TaRKD3-7A and TaRKD3-7D) and occurred towards the C-terminus except in those encoded by genes on homoeologous group 3 chromosomes, where the RKD domain was available towards the N-terminus. The RKD domain contains a consensus sequence RWPXRK that has been described as a characteristic feature of RWP-RK family of TFs in plants [4]. However, N-terminal RKD domain in two TaRKD proteins, namely TaRKD3-7A and TaRKD3-7D, is deficient for more than half of the downstream region with RWPXRK, which is responsible for DNA-binding function. A similar situation was reported in OsNLP6 [4]. One may speculate that these proteins (deficient for more than half of RKD domain) must have lost their DNA binding properties and must have either acquired new function or become non-functional [4]. Regarding PB1 domain, its role in interaction of NLPs with other proteins has been demonstrated in Arabidopsis, but no such information is available for wheat NLPs.

Physicochemical properties of proteins.

The molecular weight of predicted TaRKD proteins varied from 23.75 to 103.69 KD and that of TaNLP proteins ranged from 98.56.26 to 102.31 KD. The isoelectric point (pI) of all TaRKD proteins was within the alkaline range except in case of TaRKD10, TaRKD11, TaRKD3-7B and TaRKD3-7D proteins, where the pI was within the acidic range. However, the situation of the TaNLP proteins differed, where the pI ranged from high (alkaline range) to low (acidic range) [63]. All TaRKD and TaNLP proteins had unstable nature, making it difficult to obtain these proteins in pure crystalline state for a study of their crystalline structure [64]. However, the higher aliphatic index for the TaRKD proteins (range: 71.63–86.10) relative to that of TaNLP proteins (range: 71.02–79.28), suggests relatively higher stability of the TaRKDs at wider range of temperatures [65]. Grand Average of Hydropathy (GRAVY) ranged from -0.153 to -0.629 for TaRKD proteins and from -0.319 to -0.469 for TaNLP proteins suggesting hydrophilic nature of these proteins (S4 Table). The physicochemical properties of the RWP-RK proteins from other plant species are not reported and hence the results of the present study could not be discussed.

3D structure.

In silico 3D structures were determined for eight TaRKD and six TaNLP representative proteins (Fig 5), which shared 50–80% similarity with the corresponding Brachypodium structures used as template. This level of structural similarity was adequate for analysis of 3D protein structures (a minimum of 30% similarity is needed) that were determined using Swiss-Model algorithm (S5 Table). Ramachandran plots indicated a high proportion of amino acids in the favoured region (i.e. 78.3% in TaRKD9-3A to 96.3% in TaRKD6-2A; 75.9% in TaNLP5-6A to 84.7% in TaNLP3-4A) suggesting satisfactory geometry of the predicted 3D structures (S5 Table). The use of Ramachandran plot for evaluation of the accuracy of protein structures is known and has been emphasized in several recent studies [27, 66]. The 3D structures of TaRKDs and TaNLPs have been submitted to Protein Model Database (PMDB) (https://bioinformatics.cineca.it/PMDB/) [67], which can be freely accessed (S5 Table).

thumbnail
Fig 5. 3D structures of TaRKD and TaNLP proteins.

In all figures, spirals are helices, broad strips with arrow-head are β-pleated sheets and thin loops are coils.

https://doi.org/10.1371/journal.pone.0208409.g005

Alignment, subcellular localization and functional annotation of 3D structures.

The 3D structures of three TaRKDs and four TaNLPs had significant per cent similarity with corresponding 3D structures of BdRKDs [range of similarity = 36.97% (TaRKD3-7A) to 88.14% (TaRKD6-2A)] and BdNLPs [range of similarity = 75.77% (TaNLP4-2A) to 95.99 (TaNLP7-3A)], respectively (S4 Fig and S6 Table). The 3D structures of the remaining four proteins from RKD sub-family (TaRKD1-7A, TaRKD4-6A, TaRKD9-3A, TaRKD10-7A) and two proteins from NLP sub-family (TaNLP1-4B and TaNLP5-6A), however, did not have significant similarity (<30%) with 3D structures of the respective Brachypodium proteins. This may be attributed to the possible divergence of the amino acid sequences of the corresponding wheat and Brachypodium proteins.

All TaRKDs and TaNLPs (except TaRKD3-7s and TaRKD6b-2B) were localized in the nucleus providing support for their function as transcription factors (S7 Table) [27]. The functional annotation analysis also suggested their involvement in different activities including DNA binding, metal ion binding, protein binding and ATP binding, thus influencing several biological processes including cellular, metabolic and biosynthetic processes (S5 Fig). In particular, NLPs should respond to nitrate signals, through binding to cis-elements of relevant genes [6].

Molecular docking and molecular dynamics simulations analysis

Since, the protein structures of the two TFs (TaRKD6-2A and TaNLP7-3A) showed maximum amino acids in the favoured regions in Ramachandran plots (S5 Table), we used these two representative protein structures, one from each sub-family, for molecular docking and molecular simulation analysis. The active region of TaRKD6-2A was situated between helix H2 and loop L3. Amino acid residues from THR46 –LYS57 of H2 and GLY62 –ARG65 of loop3 were involved in the formation of this active region. Nitrate ion that functions as a ligand did not show any hydrogen bond formation but showed good contact with LYS51 and LYS55 within 3 Å in active regions (Fig 6A). The active site of TaNLP7-3A covered helix H8 and loops L2 and L6. Nitrate ion formed one hydrogen bond (2.04 Å) with SER94 and four salt bridge interactions involving the following three amino acid residues: ASP397 (5.40Å), ARG97 (2.05Å & 3.03Å) and ASP96 (3.93Å) (Fig 6B). The docking scores of TaRKD6-2A (4.59) and TaNLP7-3A (5.37) showed compact binding affinity with pocket fitting in the active regions suggesting stable conformation of both the protein-nitrate ion complexes.

thumbnail
Fig 6.

Interaction of nitrate ion with specific amino acid residues of (a) TaRKD6-2A and (b) TaNLP3-7A proteins.

https://doi.org/10.1371/journal.pone.0208409.g006

The effect of nitrate ion on the binding position of both the above protein structures (TaRKD6-2A and TaNLP7-3A) showed stability values in the acceptable range of <3.0 Å (0.02Å – 0.07Å for TaRKD6-2A and 0.03Å – 0.08Å for TaNLP7-3A). RMSD values suggested that nitrate ion remained bound in the binding pocket and stabilized both the protein structures during simulation analysis (Fig 7A and 7B). The RMSD values of complex involving TaRKD6-2A protein showed fluctuations during 1 nsec to 13 nsec trajectory between 4.2Å – 8.5Å, after which RMSD values got stabilized and converged at 2.1Å – 2.4Å distance in fixed range. In the complex involving TaNLP7-3A, protein backbone atoms showed higher RMSD values relative to TaRKD6-2A during 1 nsec to 19 nsec at 4.05Å - 6.88Å distance. This may be due to the availability of more loop regions in the 3D structure of TaNLP7-3A. These loop regions may be responsible for higher fluctuations in the 3D structure. After 19 nsec, RMSD values decreased and were stabilized at 1.25Å – 2.82Å distance (Fig 7C and 7D). The RMSD values obtained during 0 nsec to 1 nsec trajectory in the initial stages of simulation analysis were not considered due to large thermal changes.

thumbnail
Fig 7.

RMSD plots of nitrate ion during binding with TaRKD6-2A (a and c) and TaNLP3-7A (b and d) obtained using Prizm software.

https://doi.org/10.1371/journal.pone.0208409.g007

Phylogenetic analysis of proteins

For phylogenetic analysis, we used a total of 83 RWP-RK proteins including 37 from wheat, 17 from Brachypodium, 15 from rice and 14 from Arabidopsis. The 83 proteins made two major clusters, cluster I largely including NLPs and cluster II largely including RKDs (Fig 8). The only outliers were OsRKD1, three TaRKD9s and BdRKD9, together forming one sub-cluster and OsNLP6 forming another sub-cluster within cluster I. The inclusion of five RKD proteins with NLPs in cluster I may be attributed to evolving nature of RKDs relative to NLPs, which are conserved. A separate sub-cluster with OsNLP6 alone may be attributed to the absence of downstream half of the protein [4]. In this respect, our results partly differ from those of an earlier study, where NLPs and RKDs from six species were shown to clearly segregate in two independent clusters with no exceptions [4]. Another noticeable feature of our analysis at cluster level is that wheat proteins in both the clusters clustered with corresponding proteins of Brachypodium along with some rice proteins. As expected, all Arabidopsis proteins clustered together making separate sub-clusters in each of the two clusters. This suggested differentiation among these proteins following the divergence of monocots and dicots from a common ancestor.

thumbnail
Fig 8.

Phylogenetic tree constructed using proteins sequences of RKDs and NLPs belonging to four plant species (A. thaliana, B. distachyon, O. sativa and T. aestivum). Red, magenta, blue and green colours represent proteins sequences of RKDs and NLPs belonging to A. thaliana, O. sativa, B. distachyon and T. aestivum, respectively. The branch length represents the magnitude of genetic change.

https://doi.org/10.1371/journal.pone.0208409.g008

qRT-PCR expression analysis of TaNLP and TaRKD genes

The four representative genes (TaNLP2, TaNLP7, TaRKD6 and TaRKD9) in two wheat cultivars were used for qRT-PCR and were found to differ in expression patterns in roots and shoots under three different N regimes (Fig 9). The results differed not only for four genes, but also in two wheat genotypes, which included C306 (with low NUE) and HUW468 (with high NUE). Following is the summary of the results of qRT-PCR that are presented in Fig 9. (i) In root tissue, no major change was noticed in the expression of two TaNLP genes and also in TaRKD6 gene, but ~10 fold increase in the expression of TaRKD9 was noticed in HUW468 on N restoration; (ii) In shoot tissue, the expression of TaNLP2 in C306 increased ~25 fold under low N, and in HUW468 it increased ~20 fold under N starvation; the expression of TaNLP7 increased ~60 fold in C306 on N restoration, and 100 fold in HUW468 on N starvation; in shoot tissue, the expression of both TaRKD genes declined (-60 fold in C306 under low N in TaRKD6, and -15 fold in HUW468 on N restoration in TaRKD9).

thumbnail
Fig 9. Relative expression level of two TaNLP and two TaRKD genes in root and shoot tissues of wheat seedlings.

(A) C306 root, (B) C306 shoot, (C) HUW 468 root, and (D) HUW 468 shoot. Four treatments are shown by four different colours. For details of treatments, see text. C-control, LN- low N, NS- N starvation and NR- N restoration.

https://doi.org/10.1371/journal.pone.0208409.g009

The above summary of results suggests that the expression of two TaNLP genes is N-dose dependent, tissue specific and differs in two genotypes, which differ for NUE. Firstly, the expression of the two TaNLP genes was higher in the shoot tissues suggesting their involvement in N translocation/mobilization rather than in the uptake of N. Secondly, the expression of same genes differed under different doses of N in the two genotypes; in the genotype HUW468 with higher NUE, their expression was higher under N starvation but in C306 with low NUE, their expression was higher at low or optimum dose of N. These results clearly suggest that TaNLP genes may help in N translocation/mobilization in genotype HUW468 with relatively high NUE even in the absence of N, but in the genotype with low NUE, presence of N is necessary to induce these genes to help in N translocation/mobilization in the shoot.

Although the results of our qRT-PCR in seedlings cannot be compared with those of in silico analysis at post-anthesis stage, similar conclusions can be drawn from these two studies. For instance, in silico analysis of expression of two TaNLP genes (TaNLP1 and TaNLP2) in adult leaves of the genotype Herebard, was relatively low at high dose of N (200Kg N/ha) than at a low dose of N (50 KgN/ha) suggesting their role in NUE. NLP genes have also been implicated in nitrate assimilation and other metabolic/regulatory processes associated with NUE [6] and AtNLP7 has been shown to have a role in the control of the expression of nitrate transporter (NRT1.1), adaptation to limited supply of nitrogen and in signaling F-box (AFB3) genes in Arabidopsis [7]. However, in wheat, the role of TaNLPs in controlling the down-stream genes is yet to be fully understood.

The results of qRT-PCR were also subjected to ANOVA, which suggested significant variation in expression due to genes (four genes), due to tissues (root and shoot) and due to treatments including supply of different levels of N. Interactions were largely non-significant, except for the following: (i) N x V in case of TaNLP7 in shoot, (ii) V x D in case of TaNLP2 and TaRKD6 in root and (iii) N x D in case of TaNLP2 in root. This suggests that the expression of RWP-RK genes is controlled by several interdependent factors which form a network, and does not merely depends upon level of N supply.

RKD genes are known to be primarily involved in egg and sperm development, as shown in Marchantia polymorpha [68], Arabidopsis and wheat [69,70] and in post-zygotic embryogenesis as shown in Arabidopsis [71]. However, in our qRT-PCR analysis, TaRKD9 was found to respond to N restoration after N starvation, suggesting a positive role of this gene in response to nitrogen signaling. The presence of nitrogen response elements in the promoter of TaRKD3-7D also suggests that TaRKD genes may be involved in response to N starvation. Therefore, the response to nitrogen status may be considered to be a novel function for at least some TaRKD genes, which deserves further study. However, in Chlamydomonas reinhardii, the change in expression level of RKD genes in response to N starvation is known [8], but its role in higher plants needs to be examined. In wheat also, role of TaRKD1 and TaRKD2 in egg development has been examined, but there is no report on a role of RKD genes in nitrogen signaling, although N response elements were detected in TaRKD3-7D. Alteration of expression of TaRKD genes in shoot and not in root also suggest a role of these genes in regulating the expression of many downstream genes, which may be involved in shoot development. Thus, there is a need to understand the role of the TaRKD genes in vegetative tissues and their control on the expression of down-stream genes in response to N application.

Supporting information

S1 Fig. Representative figure showing synteny and collinearity of wheat TaNLP1-4D gene with respective genes of Brachypodium, rice and sorghum.

TaNLP1-4D gene (with a green boundary) in wheat is connected with corresponding gene in Brachypodium, rice and sorghum by a thick green line.

https://doi.org/10.1371/journal.pone.0208409.s001

(TIF)

S2 Fig. Results of in silico expression analysis, hierarchical clustering of TaRKDs and TaNLPs in different parts of a plant.

https://doi.org/10.1371/journal.pone.0208409.s002

(TIF)

S3 Fig. Results of in silico expression analysis, hierarchical clustering of TaRKDs and TaNLPs under different development stages.

https://doi.org/10.1371/journal.pone.0208409.s003

(TIF)

S4 Fig. Superimposed 3D structures of TaRKD and TaNLP proteins over 3D structures of BdRKD and BdNLP proteins (shown in grey colour).

https://doi.org/10.1371/journal.pone.0208409.s004

(TIF)

S5 Fig.

Gene ontology analysis of TaRKD and TaNLP proteins: (a) predicted biological process and, (b) predicted biochemical functions.

https://doi.org/10.1371/journal.pone.0208409.s005

(TIF)

S1 Table. List of primers for representative genes used in quantitative real time-PCR (qRT-PCR) expression profiling.

https://doi.org/10.1371/journal.pone.0208409.s006

(DOCX)

S2 Table. Simple sequence repeats (SSRs) identified in TaRKD and TaNLP genes.

https://doi.org/10.1371/journal.pone.0208409.s007

(DOCX)

S3 Table. The starting position of various regulatory elements from transcription start site identified in 1kb upstream promoter region of RWP-RK genes in wheat.

https://doi.org/10.1371/journal.pone.0208409.s008

(DOCX)

S4 Table. Physicochemical properties of TaRKD and TaNLP proteins.

https://doi.org/10.1371/journal.pone.0208409.s009

(DOCX)

S5 Table. Homology modeling and structure validation of TaRKD and TaNLP proteins using Swiss-Model and Protein Structure Validation Suite (PSVS) respectively, along with their PMDB accessions.

https://doi.org/10.1371/journal.pone.0208409.s010

(DOCX)

S6 Table. Predicted values of different parameters after superimposition of 3D structures of TaRKD and TaNLP proteins over 3D structure of BdRKD and BdNLP proteins.

https://doi.org/10.1371/journal.pone.0208409.s011

(DOCX)

S7 Table. Sub-cellular localization of TaRKD and TaNLP proteins.

https://doi.org/10.1371/journal.pone.0208409.s012

(DOCX)

Acknowledgments

AK received help from AKTU, Lucknow and UCB, Dehradun. Thanks are due to Dr. Kalpana Singh in conducting tandem/segmental duplication analysis using MCScanX. The supercomputer facilities provided by the ‘Bioinformatics Resources and Applications Facility’ (BRAF) at C-DAC, Pune and the facilities provided by Bioinformatics Infrastructure Facility (BIF) Laboratory, CCS University, Meerut made this study possible.

References

  1. 1. Wang R, Tischner R, Gutiérrez RA, Hoffman M, Xing X, Chen M, et al. Genomic analysis of the nitrate response using a nitrate reductase-null mutant of Arabidopsis. Plant Physiol. 2004; 136; 2512–2522. pmid:15333754
  2. 2. Gojon A. Nitrogen nutrition in plants: rapid progress and new challenges. J. Exp. Bot. 2017; 68(10); 2457–62. pmid:30053117
  3. 3. Balyan HS, Gahlaut V, Kumar A, Jaiswal V, Dhariwal R, Tyagi S, et al. Nitrogen and phosphorus use efficiencies in wheat: Physiology, phenotyping, genetics and breeding. Plant Breed. Rev. 2015; 40; 67–234.
  4. 4. Chardin C, Girin T, Roudier F, Meyer C, Krapp A. The plant RWP-RK transcription factors: key regulators of nitrogen responses and of gametophyte development. J. Exp. Bot. 2014; 65; 5577–5587. pmid:24987011
  5. 5. Konishi M, Yanagisawa S. Emergence of a new step towards understanding the molecular mechanisms underlying nitrate-regulated gene expression. J. Exp. Bot. 2014; 65; 5589–600. pmid:25005135
  6. 6. Konishi M. and Yanagisawa S. Arabidopsis NIN-like transcription factors have a central role in nitrate signalling. Nat. Commun. 2013a; 4; 1617.
  7. 7. Yu L, Wu J, Tang H, Yuan Y, Wang S, Wang Y, et al. Overexpression of Arabidopsis NLP7 improves plant growth under both nitrogen-limiting and -sufficient conditions by enhancing nitrogen and carbon assimilation. Sci. Rep. 2016; 6:27795. pmid:27293103
  8. 8. Lin H, Goodenough UW. Gametogenesis in the Chlamydomonas reinhardtii minus mating type is controlled by two genes, MID and MTD1. Genetics.2007; 176; 913–925. pmid:17435233
  9. 9. Konishi M, Yanagisaw S. An NLP-binding site in the 3’ flanking region of the nitrate reductase gene confers nitrate-inducible expression in Arabidopsis thaliana (L.) J. Soil Sci. Plant Nutr. 2013b; 59; 612–620.
  10. 10. Castaings L, Camargo A, Pocholle D, Gaudon V, Texier Y, Boutet-Mercey S, et al. The nodule inception-like protein 7 modulates nitrate sensing and metabolism in Arabidopsis. Plant J. 2009; 57; 426–435. pmid:18826430
  11. 11. Dhaliwal AK, Mohan A, Gill KS. Comparative analysis of ABCB1 reveals novel structural and functional conservation between monocots and dicots. Front. Plant Sci. 2014; 5; 657. pmid:25505477
  12. 12. Kumar A, Kumar S, Kumar U, Suravajhala P, Gajula MNVP. Functional and structural insights into novel DREB1A transcription factors in common wheat (Triticum aestivum L.): A molecular modeling approach. Comp. Biol Chem. 2016a; 64; 217–216.
  13. 13. Batra R, Saripalli G, Mohan A, Gupta S, Gill KS, Varadwaj PK, et al. Comparative analysis of AGPase genes and encoded proteins in eight monocots and three dicots with emphasis on wheat. Front. Plant. Sci. 2017; 8; 1–16.
  14. 14. Hu B, Jin J, Guo AY, Zhang H, Luo J, Gao G. GSDS 2.0: an upgraded gene feature visualization server. Bioinformatics. 2015; 31; 1296–1297. pmid:25504850
  15. 15. Lescot M, Déhais P, Thijs G, Marchal K, Moreau Y, de Peer YV, et al. PlantCARE, a database of plant cis-acting regulatory elements and a portal to tools for in silico analysis of promoter sequences. Nucleic Acids Res. 2002; 30; 325–327. pmid:11752327
  16. 16. Dai X, Zhao PX. psRNA Target: a plant small RNA target analysis server. Nucleic Acids Res. 2011; 39; 155–159.
  17. 17. Krzywinski M, Schein J, Birol I, Connors J, Gascoyne R, Horsman D, et al. Circos: an information esthetic for comparative genomics. Genome Res. 2009; 19; 1639–1645. pmid:19541911
  18. 18. Dash S, Van Hemert J, Hong L, Wise RP, Dickerson JA. PLEXdb: gene expression resources for plants and plant pathogens. Nucleic Acids Res. 2012; 40 (D1); D1194–D1201.
  19. 19. Hruz T, Laule O, Szabo G, Wessendorp F, Bleuler S, Oertle L, et al. Genevestigator v3: a reference expression database for the meta-analysis of transcriptomes. Adv. Bioinformatics. 2008; 420747; 1–5.
  20. 20. Chou KC, Shen HB. Plant-mPLoc: a top-down strategy to augment the power for predicting plant protein subcellular localization. PLoS ONE. 2010; 5:e11335. pmid:20596258
  21. 21. Kumar A, Mishra DC, Rai A, Sharma MK, Gajula MNVP. In silico analysis of protein-protein interaction between resistance and virulence protein during leaf rust disease in wheat (Triticum aestivum L). Worl. Res. J. Pept. Prot. 2013; 2; 52–58.
  22. 22. Biasini M, Bienert S, Waterhouse A, Arnold K, Studer G, Schmidt T, et al. SWISS-MODEL:modelling protein tertiary and quaternary structure using evolutionary information. Nucleic Acids Res. 2014; 42; W252–W258. pmid:24782522
  23. 23. Ye Y, Godzik A. Flexible structure alignment by chaining aligned fragment pairs allowing twists. Bioinformatics. 2003; 19(S2); ii246–ii55.
  24. 24. Laskowski RA, Watson JD, Thornton JM. ProFunc: a server for predicting protein function from 3D structure. Nucleic Acids Res. 2005; 33 (W); W89–W93. pmid:15980588
  25. 25. Kumar A, Kumar S, Kumar A, Sharma N, Sharma M, Singh KP, et al. Homology modeling, molecular docking and molecular dynamics based functional insights into rice urease bound to urea. Proc. Nat. Acad. Sci. Bio. India. 2017.
  26. 26. Gajula MNVP, Kumar A, Ijaq J. Protocol for molecular dynamics simulations of proteins. Bio. Protoc. 2016; 6:e2051.
  27. 27. Kumar A, Kumar S, Kumar U, Suravajhala P, Gajula MNVP. Functional and structural insights into novel DREB1A transcription factors in common wheat (Triticum aestivum L.): A molecular modeling approach. Comp. Biol. Chem. 2016a; 64; 217–216.
  28. 28. Gajula MNVP, Steinhoff HJ, Kumar A, Kumar AP, Siddiq EA. Displacement of the tyrosyl radical in RNR enzyme: A sophisticated computational approach to analyze experimental data. In Proceedings of International Conference on Bioinformatics and Computational Biology (BICOB–2015), 2015; 7; 211–219, March 9–11, Honolulu, Hawaii, USA.
  29. 29. Tamura K, Stecher G, Peterson D, Filipski A, Kumar S. MEGA6: Molecular evolutionary genetics analysis version 6.0. Mol. Biol. Evol. 2013; 30; 2725–2729. pmid:24132122
  30. 30. Larkin MA, Blackshields G, Brown NP, Chenna R, McGettigan PA, McWilliam H, et al. Clustal W and Clustal X version 2.0. Bioinformatics. 2007; 23; 2947–2948. pmid:17846036
  31. 31. Nei M, Kumar S. Molecular evolution and phylogenetics. Oxford University Press, New York. 2000; 25(3); 569–570.
  32. 32. Pandey R, Lal MK, Vengavasi K. Differential response of hexaploid and tetraploid wheat to interactive effects of elevated [CO2] and low phosphorus. Plant Cell Rep. 2018; 37 (9); 1231–1244. pmid:29868985
  33. 33. Schmittgen TD, Livak KJ. Analyzing real-time PCR data by the comparative CT method. Nat. Protoc. 2008; 3; 1101–1108. pmid:18546601
  34. 34. Shang X, Cao Y, Ma L. Alternative Splicing in Plant Genes: A Means of Regulating the Environmental Fitness of Plants. Int. J. Mol. Sci. 2017; 18(2); 432.
  35. 35. Reddy ASN, Marquez Y, Kalyna M, Bartab A. Complexity of the Alternative Splicing Landscape in Plants. Plant Cell. 2013; 25; 3657–3683. pmid:24179125
  36. 36. Staiger D, Brown JWS. Alternative splicing at the intersection of biological timing, development, and stress responses. Plant Cell. 2013; 25; 3640–3656. pmid:24179132
  37. 37. Tedeschi F, Rizzo P, Rutten T, Altschmied L, Bäumlein H. RWP-RK domain-containing transcription factors control cell differentiation during female gametophyte development in Arabidopsis. New Phytol. 2017; 213; 1909–1924. pmid:27870062
  38. 38. Gill BS. The colinearity of the Sh2/A1 orthologous region in rice, sorghum and maize is interrupted and accompanied by genome expansion in the triticeae. Genetics. 2002; 160(3); 1153–1162. pmid:11901130
  39. 39. Feuillet C, Keller B. High gene density is conserved at syntenic loci of small and large grass genomes. Proc. Natl. Acad. Sci. USA. 1999; 96(14); 8265–8270. pmid:10393983
  40. 40. Devos KM, Dubcovsky J, Dvořák J, Chinoy CN, Gale MD. Structural evolution of wheat chromosomes 4A, 5A, and 7B and its impact on recombination. Theor. Appl. Genet. 1995; 91; 282–288. pmid:24169776
  41. 41. Ma J, Stiller J, Berkman PJ, Wei Y, Rogers J, et al. (2013) Sequence-Based Analysis of Translocations and Inversions in Bread Wheat (Triticum aestivum L.). PLoS ONE 8(11): e79329. pmid:24260197
  42. 42. Wang M, Yue H, Feng K, Deng P, Song W, Nie X. Genome-wide identification, phylogeny and expressional profiles of mitogen activated protein kinase kinase kinase (MAPKKK) gene family in bread wheat (Triticum aestivum L.). 2016; BMC Genomics; 17:668 pmid:27549916
  43. 43. Huo N, Vogel JP, Lazo GR, You FM, Ma Y, McMahon S, et al. Structural characterization of Brachypodium genome and its syntenic relationship with rice and wheat. Plant Mol. Biol. 2009; 70; 47–61. pmid:19184460
  44. 44. Long M, Deutsch M. Association of intron phases with conservation at splice site sequences and evolution of spliceosomal introns. Mol. Biol. Evol. 1999; 16; 1528–1534. pmid:10555284
  45. 45. Gupta PK, Rustgi S, Sharma S, Singh R, Kumar N, Balyan HS. Transferable EST-SSR markers for the study of polymorphism and genetic diversity in bread wheat. Mol. Genet. Genomics. 2003; 270 (4); 315–323. pmid:14508680
  46. 46. Kujur A, Bajaj D, Saxena MS, Tripathi S, Upadhyaya HD, Gowda CLL et al. Functionally relevant microsatellite markers from chickpea transcription factor genes for efficient genotyping applications and trait association mapping. DNA Res. 2013; 20; 355–374. pmid:23633531
  47. 47. Liu W, Jia X, Liu Z, Zhang Z, Wang Y, Liu Z, et al. Development and characterization of transcription factor gene derived microsatellite (TFGM) markers in Medicago truncatula and their transferability in leguminous and non leguminous species. Molecules. 2015; 20(5); 8759–771. pmid:25988608
  48. 48. Han Y, Luan F, Zhu H, Shao Y, Chen A, Lu C, et al. Computational identification of microRNAs and their targets in wheat (Triticum aestivum L.). Sci. China Ser. C. Life Sci. 2009; 52; 1091–1100.
  49. 49. Zinati Z, Shamloo-dashtpagerdi R, Behpouri A. In silico identification of miRNAs and their target genes and analysis of gene co-expression network in saffron. Mol. Biol. Res. Commu. 2016; 5 (4); 233–246.
  50. 50. Wan P, Wu J, Zhou Y, Xiao J, Feng J, Zhao W, et al. Computational analysis of drought stress-associated miRNAs and miRNA co-regulation network in Physcomitrella patens. Genomics Proteomics Bioinformatics. 2011; 9(1–2); 37–44. pmid:21641561
  51. 51. Guleria P, Yadav SK. Identification of miR414 and expression analysis of conserved miRNAs from Stevia rebaudiana. Genomics Proteomics Bioinformatics. 2011; 9; 211–217. pmid:22289477
  52. 52. Tworak A, Urbanowicz A, Podkowinski J, Kurzynska-Kokorniak A, Koralewska N, Figlerowicz M. Six Medicago truncatula dicer-like protein genes are expressed in plant cells and upregulated in nodules. Plant Cell Rep. 2016; 35 (5); 1043–1052. pmid:26825594
  53. 53. Jones-Rhoades MW, Bartel DP, Bartel B. MicroRNAs and their regulatory roles in plants. Annu. Rev. Plant Biol. 2006; 57; 19–53. pmid:16669754
  54. 54. Bennetzen JL, Chen M. Grass genomic synteny illuminates plant genome function and evolution. Rice. 2008; 1; 109–118.
  55. 55. Muller M, Knudsen S. The nitrogen response of a barley C-hordein promoter is controlled by positive and negative regulation of the GCN4 and endosperm box. The Plant J. 1993; 4(2); 343–355. pmid:8220485
  56. 56. Joo J, Lee YH, Kim YK, Nahm BH, Song SI. Abiotic stress responsive rice ASR1 and ASR3 exhibit different tissue-dependent sugar and hormone-sensitivities. Mol. Cells. 2013; 35; 421–435. pmid:23620302
  57. 57. Fankhauser C, Chory J. Light Control of Plant Development. Annu. Rev. Cell Dev. Biol. 1997; 13; 203–229. pmid:9442873
  58. 58. Morishima A. Identification of preferred binding sites of a light-inducible DNA-binding factor (MNF1) within 5’-upstream sequence of C4-type phosphoenolpyruvate carboxylase gene in maize. Plant Mol. Biol. 1998; 38; 633–646. pmid:9747808
  59. 59. Li X, Han JD, Fang YH, Bai SN, Rao GY. Expression analyses of embryogenesis-associated genes during somatic embryogenesis of Adiantum capillus-veneris L. in vitro: New Insights into the evolution of reproductive organs in land plants. Front. Plant. Sci. 2017; 1–12.
  60. 60. Waki T, Hiki T, Watanabe R, Hashimoto T, Nakajima K. The Arabidopsis RWP-RK protein RKD4 triggers gene expression and pattern information in early embryogenesis. Curr. Biol. 2011; 2; 1277–1281.
  61. 61. Ge M, Liu Y, Jiang L, Wang Y, Lv Y, Zhou L, et al. Genome-wide analysis of maize NLP transcription factor family revealed the roles in nitrogen response. J. Plant Growth Regul. 2017; 84(1); 95–101.
  62. 62. Marchive C, Roudier F, Castaings L, Bréhaut V, Blondet E, Colot V, et al. Nuclear retention of the transcription factor NLP7 orchestrates the early response to nitrate in plants. Nat. Commun. 2013; 4; 1713. pmid:23591880
  63. 63. Guruprasad K, Reddy BV, Pandit MW. Correlation between stability of a protein and its dipeptide composition: A novel approach for predicting in vivo stability of a protein from its primary sequence. Protein Engg. 1990; 4; 155–161.
  64. 64. Dawar C, Jain S, Kumar S. Insight into the 3D structure of ADP-glucose pyrophosphorylase from rice (Oryza sativa L.). J. Mol. Model. 2013; 19; 3351–3367. pmid:23674369
  65. 65. Gupta SK, Rai AK, Kanwar SS, Sharma TR. Comparative analysis of zinc finger proteins involved in plant disease resistance. PLoS One. 2012; 7:e42578. pmid:22916136
  66. 66. Kumar A, Sharma M, Kumar S, Tyagi P, Wani SH, Gajula MNVP, et al. Functional and structural insights into candidate genes associated with nitrogen and phosphorus nutrition in wheat (Triticum aestivum L.). Int J Biol Macromol. 2018; 118(Pt A); 76–91. pmid:29879411
  67. 67. Castrignanò T, De Meo PD, Cozzetto D, Talamo IG, Tramontano A. The PMDB Protein Model Database. Nucleic Acids Res. 2006; 34; D306–D309. pmid:16381873
  68. 68. Koi S, Hisanaga T, Sato K, Shimamura M, Yamato KT, Ishizaki K, et al. An Evolutionarily Conserved Plant RKD Factor Controls Germ Cell Differentiation. Curr Biol. 2016; 26; 1775–1781. pmid:27345165
  69. 69. Wuest SE, Vijverberg K, Schmidt A, Weiss M, Gheyselinck J, Lohr M, et al. Arabidopsis female gametophyte gene expression map reveals similarities between plant and animal gametes. Curr. Biol. 2010; 20, 506–512. pmid:20226671
  70. 70. Koszegi D, Johnston AJ, Rutten T, Czihal A, Altschmied L, Kumlehn J, et al. Members of the RKD transcription factor family induce an egg cell-like gene expression program. Plant J. 2011; 67; 280–291. pmid:21457369
  71. 71. Jeong S, Palmer TM, Lukowitz W. The RWP-RK factor GROUNDED promotes embryonic polarity by facilitating YODA MAP kinase signaling. Curr. Biol. 2011; 21; 1268–1276. pmid:21802295