Next Article in Journal
Transcriptomic Analysis Provides Insights into Grafting Union Development in Pecan (Carya illinoinensis)
Next Article in Special Issue
Genome-Wide Identification and Expression Profiling Analysis of the Xyloglucan Endotransglucosylase/Hydrolase Gene Family in Tobacco (Nicotiana tabacum L.)
Previous Article in Journal
Long Non-Coding RNAs in Multiple Myeloma
Previous Article in Special Issue
SlbZIP38, a Tomato bZIP Family Gene Downregulated by Abscisic Acid, Is a Negative Regulator of Drought and Salt Stress Tolerance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genome-Wide Identification and Expression Analysis of the HD-Zip Gene Family in Wheat (Triticum aestivum L.)

1
College of Agronomy, State Key Laboratory of Crop Stress Biology in Arid Areas, Northwest A&F University, Yangling 712100, China
2
College of Life Sciences, Northwest A&F University, Yangling 712100, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Genes 2018, 9(2), 70; https://doi.org/10.3390/genes9020070
Submission received: 11 December 2017 / Revised: 23 January 2018 / Accepted: 26 January 2018 / Published: 1 February 2018
(This article belongs to the Special Issue Genetic Regulation of Abiotic Stress Responses)

Abstract

:
The homeodomain-leucine zipper (HD-Zip) gene family, as plant-specific transcription factors, plays an important role in plant development and growth as well as in the response to diverse stresses. Although HD-Zip genes have been extensively studied in many plants, they had not yet been studied in wheat, especially those involved in response to abiotic stresses. In this study, 46 wheat HD-Zip genes were identified using a genome-wide search method. Phylogenetic analysis classified these genes into four groups, numbered 4, 5, 17 and 20 respectively. In total, only three genes with A, B and D homoeologous copies were identified. Furthermore, the gene interaction networks found that the TaHDZ genes played a critical role in the regulatory pathway of organ development and osmotic stress. Finally, the expression profiles of the wheat HD-Zips in different tissues and under various abiotic stresses were investigated using the available RNA sequencing (RNA-Seq) data and then validated by quantitative real-time polymerase chain reaction (qRT-PCR) to obtain the tissue-specific and stress-responsive candidates. This study systematically identifies the HD-Zip gene family in wheat at the genome-wide level, providing important candidates for further functional analysis and contributing to the better understanding of the molecular basis of development and stress tolerance in wheat.

1. Introduction

The homeodomain-leucine zipper (HD-Zip) gene family is one of the key transcription factors in plants, playing a vital role in various abiotic stresses and signal transduction [1,2,3]. Generally, HD-Zip proteins possess the conserved HD domain, acting as a specific DNA binding site at the C-terminal, together with the adjacent leucine-zipper (LZ) motif that is responsible for protein dimerization [4,5]. Based on the sequence homology, DNA binding specificity and physiological functioning, HD-Zip genes were further divided into four groups, namely HD-Zip I, II, III and IV [6]. Members of groups I and II interacted with similar DNA binding sites of the pseudo-palindromic sequence CAATNATTG [5,7]. However, the members of group II encoded an additional common CPSCE motif consisting of five conserved amino acids—Cys, Pro, Ser, Cys and Glu—which acted as a sensory domain to redox the cell state and were located downstream of the Zip domain [8,9]. Group III and IV contained steroidogenicacute regulatory protein-related lipid transfer (START) and START-associated domain (SAD) domains [10,11,12]. The difference between groups III and IV were mainly dependent on a specific MEKHLA domain located in the C-terminal that responds to oxygen redox and light signaling [13]. In addition, the members of group III recognized sequences GTAATG/CATTAC, whereas members of group IV recognized sequences TAAATGC/TA [12,14,15].
Previous studies have extensively demonstrated that the HD-Zip family plays an important role in regulating diverse developmental and physiological processes in plants [16,17,18]. HD-Zip I proteins were found to be involved in the control of plant growth and development and also regulated the response to abiotic stresses [19]. The TaHDZip I-2 gene regulated flowering and spike development and improved frost tolerance in transgenic barley lines [20]. In Arabidopsis, AtHB1 acted downstream of AtPIF1 to promote hypocotyl elongation, especially in response to short-day photoperiods [21], as well as mediated the leaf cell fate determination [22]. In cotton, the GhHB1 expression level significantly increased in the early developmental roots, then significantly decreased as the roots developed, suggesting it might function in early root development [23]. Additionally, TaHDZipI-3, -4 and -5 genes were found to show the differential expression when wheat subjected to abscisic acid (ABA)treatment, cold and water deficit through binding the specific cis-elements [24].
HD-Zip II proteins regulated auxin signaling, participated in embryonic apical development and responded to light and abiotic stresses [25,26]. In Arabidopsis, some HD-Zip II proteins, including AtHAT1, AtHAT2, AtHAT3, AtHB4 and AtHB2, regulated gene expression to adjust apical embryo development and meristem function [27]. HaHB10 induced specific flowering transition gene expression in the transition stage from the vegetative to the flowering stage and induced accumulation of phytohormones in Arabidopsis under biotic stresses [28]. Additional evidences showed that the EgHOX1 gene in oil palm was up-regulated by exogenous auxin and down-regulated by light [29]. HD-Zip III proteins were reported to impact vascular development, shoot and apical meristem formation, morph-physiological changes in roots and auxin transport [9,30]. CLV3 and HD-Zip III pathways distinctively regulated meristem activity [31]. KANADI interacted with HD-Zip III genes to control lateral root development [32]. PtrHB7 gene in Populus played an important role in balancing secondary growth in xylem cells and phloem tissues [33]. Additionally, some studies have revealed that HD-Zip III genes could be negatively regulated by MicroRNA165/166 [33,34]. HD-Zip IV proteins were the regulators for trichome development, epidermal cell differentiation and root formation [35]. PDF2 played a vital role in the epidermis cells to adjust normal development of the floral organs [36]. OCL4 inhibited trichome development in Arabidopsis and maize and influenced anther cell division and differentiation [37]. In cucumber, CsGL3 and CsGL1 genes regulated trichome initiation and development [38].
Wheat is one of the most important cereal crops in the world, occupying 17% of cultivated lands and serving as the staple food source for 30% of the human population [39]. Abiotic stresses are the main limiting factor for wheat production worldwide, which negatively impact wheat growth and development, resulting in huge yield losses. Between the 2000 and 2008, wheat grain production fell by 5.5% annually due to adverse climates [40]. To improve the stress tolerance of wheat, identification and use of the elite gene resource may help meet the challenges of the changing global climate. Although some studies have studied the function of wheat HD-Zip genes [20,41,42], their genome organization, structure and evolutionary features are not well-understood, especially those involved in the regulatory processes of abiotic stresses. In this study, we systematically characterized the HD-Zip genes in wheat using the latest genome sequences. The genome composition, phylogeny, conserved motifs, chromosome localization, regulatory network and the expression profiles of TaHDZ proteins were systematically analyzed, providing a basis for further investigation of the functions of TaHDZ genes. The results will help reveal the molecular mechanism of development and stresses response in wheat and other cereal crops.

2. Materials and Methods

2.1. Genome-Wide Identification of HD-Zip Gene Family in Wheat

Wheat genome and protein sequences (release 36, accessed in 2017) were obtained from the Ensemble plants database [43] to predict HD-Zip genes. These sequences were first used to construct a local protein database with which to search against known HD-Zip protein sequences collected from A. thaliana (48) and O. sativa (48), through a local protein basic local alignment search (BLASTP) program (https://blast.ncbi.nlm.nih.gov) with an E-value cut-off <10−5 and an identity of 50% as the threshold. The hidden Markov model (HMM) profile of the conserved HD domain of homeobox (PF00046) and the leucine zipper (LZ) domain (PF02183) sequences were download from the PFAM database [44] and used to examine all wheat protein sequences using the HMMER search tool [45]. After manual curating, the obtained protein sequences were checked using theNational Center for Biotechnology Information (NCBI)–Conserved domain database (CDD) search [46] to identify the conserved protein domain with the default parameters. The redundant sequences containing complete HD and LZ domains were further removed by alignment and the remaining ones was considered as putative wheat HD-Zip genes (TaHDZ). The coding sequences of these HD-Zip genes were retrieved from the wheat genome annotation information. Finally, Compute pI/MW tool in ExPASy database [47] was used to calculate the biochemical parameters of TaHDZs. Sub-cellular localization of these genes was predicted by Cell-Ploctool software [48].

2.2. Phylogenetic Analysis and Gene Duplication

The Arabidopsis HD-Zip proteins were downloaded from the Arabidopsis information resource (TAIR) database (http://web.arabidopsis.org) and those of rice were obtained from the rice genome annotation project (http://rice.plantbiology.msu.edu/). To investigate the evolutionary relationships among these TaHDZ genes, the ClutsalX1.83 program [49] was used to align the protein sequences of the wheat, rice and Arabidopsis HD-Zip genes (Supplementary file 1). The phylogenetic tree was constructed using MEGA6.0) [50] with the neighbor-joining (NJ) method and 1000 bootstrap replications.
The chromosome localizations of these genes were analyzed by mapping the gene sequences back to chromosome using the nucleotide basic local alignment search (BLASTN) program (https://blast.ncbi.nlm.nih.gov) with the E-value cutoff <10−5 and the best hits were identified. Gene duplication was investigated following the method as described by Wang et al. [51]. Based on chromosome position and phylogenetic relationship, the homoeologous copies distributed in three sub-genomes A, B or D of wheat were identified. Then, the Circos tool [52] was used to visualize the duplicated regions in the wheat genome and the same colors show the homoeologous chromosomal segments.

2.3. Gene Structure and Protein Conserved Motifs Analysis

The gene structure information was obtained from the Ensemble plants database [43] and displayed in the gene structure display server (GSDS) program [53]. Conserved motifs of these genes were determined using the multiple EM for motif elicitation (MEME) program [54] with the parameters as follow: optimum motif widths of 6–200 residues and a maximum of 20 motifs. The schematic diagram of the amino acid motifs for each TaHDZ gene was drawn accordingly.

2.4. Expression Profile Analysis of TaHDZ Genes

The publicly available wheat RNA-Seq datasets were downloaded from the URGI database [55] and the NCBI sequence read archive (SRA) database (https://www.ncbi.nlm.nih.gov/sra), then used to analyze the expression profiles of the identified TaHDZ genes (Table S2). A total of 5 tissues including root, stem, leaf, spike and grain, as well as four stress treatments including cold (SRR1460552), heat (SRR1542413), drought (SRR1542409) and salt (SRR2306546) were used to identify tissue-specific or stress-responsive ones. Evaluation of the quality of RNA-Seq reads and trimming of the low-quality readings with a Phred quality (Q) score < 20 were performed using FastQCv0.11.5 [56]. Then, TopHat version 2.1.1 [57] was used to map the RNA-Seq reads to the wheat genome (release v36). Cufflinks Version 2.2.1 [58] was used to calculate the value of fragments per kilo base of transcript per million fragments mapped (FPKM) of these genes with the default parameters. The expression level was first normalized and then the log10-transformed values were used for visualizing the heat map using the R software (https://www.r-project.org/).

2.5. Interaction Network of TaHDZ Genes

To predict the regulatory role of TaHDZ genes, the interaction networks of TaHDZs with other wheat genes was constructed based on the orthologous relationship between Arabidopsis and wheat using the AraNetV2 tool [59]. The Arabidopsis gene ontology (GO) biological processes were used as the major reference set. The Cytoscape plugin, BiNGO [60], was used to analyze the orthologous gene and identify the biological pathways of the specific gene sets.

2.6. Plant Materials, Growth Conditions and Abiotic Stress Treatments

Seeds were planted in pots and filled with 1/2 Hoagland’s liquid medium and grown in a greenhouse at 22 °C, with a 16 h photoperiod (12,000 lux) and 8 h dark period. Wheat variety Dekang No. 685 (DK, salt-tolerant) and Chinese spring (CS, salt-susceptible) were used for the salt treatment. Besides, Hanxuan No. 10 (HX, drought-tolerant) and CS were used for the drought treatment. For salt stress, two-week seedlings of DK and CS were treated with 200 mM sodium chloride (NaCl) for 24 h. HX and CS were treated with 19.2% polyethylene glycol (PEG) for 24 h to represent the drought treatment. Then, the plants materials were collected and immediately frozen in liquid nitrogen for RNA extraction. All samples were replicated three times.

2.7. Quantitative Real-Time Polymerase Chain Reaction Analysis

cTotal RNA was isolated from all the collected samples using the Plant RNA isolation kit (OmegaBioTek, Norcross, GA, USA), following the manufacturer’s instructions. The RNA quality was checked using 1.0% (w/v) agarose gel stained with ethidium bromide (EB) and spectrophotometer analysis and then DNase I treatment was conducted to remove the DNA contaminations (Takara, Shiga-ken, Japan). The first strand complementary DNA (cDNA)s were synthesized using the cDNA amplification kit (Vazyme, Nanjing, China). The TaHDZ gene primers and 18S, as the reference gene, were used for quantitative real-time polymerase chain reaction (qRT-PCR), designed using Primer Premier 5.0 software [61] and are listed in Table S1. Then, qRT-PCR was performed using CFX96Touch (Bio-Rad, Hercules, CA, USA). Three biological replicates for each sample were performed and the expression level was evaluated using the 2−∆∆Ct method.

3. Results and Discussion

3.1. Identification of TaHDZ Genes in Wheat

The availability of the complete genome sequence allowed the identification and analysis of gene family at the genome level in wheat. Increasing numbers of gene families such as MAPKKK, Aux/IAA and Annexin have been reported in wheat [34,62,63]. HD-Zip, as one of the plant-specific transcription factor gene families, is vital for regulating plant development and physiological processes as well as in response to abiotic stresses. However, little is known about the HD-Zip gene family in wheat. We identified and characterized the wheat HD-Zip gene family based on a genome-wide search approach. A total of 46 non-redundant genes containing the complete HD and LZ domains were obtained, which were considered as the putative wheat HD-Zip genes. These TaHDZ proteins ranged in length from 100 to 883 amino acids, with molecular weights ranging from 14.62 kDa to 95.73 kDa and the isoelectric points ranged from 4.61 to 10.68.
Subcellular localization analysis indicated that 21 TaHDZ are localized in the nucleus, 20 in the chloroplast, whereas only three and two were found in the peroxisome and mitochondrion, respectively (Table 1). Moreover, the proteins were grouped into 28 clusters based on their phylogenetic relationship. Among them, 15 clusters were assigned to different A, B or D sub-genomes, which were considered as the homoeologous copies of one TaHDZ gene. Finally, there are 28 clusters of HD-Zip gene family in wheat and termed as TaHDZ1-A to TaHDZ28-D according to their chromosomes position (Table 1). The size of HD-Zip family in wheat is similar to that of rice, maize and peal millet, as well as Arabidopsis with 48, 55, 52 and 48, respectively, while are much higher than that of grape (Vitis vinifera L.) with 31 members, chrysanthemum (Chrysanthemum morifolium L.) with 17 members and citrus (Citrus sinensis L.) with 27 members and much lower than that of soybean (Glycine max L.) with 88 members and poplar (Populus trichocarpa L.) with 65 members [9,64,65,66,67]. A previous study reported that the abundance of the HD-Zip gene family was not correlated to genome size, which was mainly dependent on tandem and segmental duplications during the process of genome evolution [65]. Our results show that the HD-Zip gene family did not expand during wheat polyploidization and genome evolution. Besides, fewer duplication events of HD-Zip genes were found in wheat.

3.2. Chromosome Localization Analysis of TaHDZs

Chromosome distribution was further analyzed by BLASTN against wheat genome sequence. Results revealed that the HD-Zip genes were unevenly distributed on wheat chromosomes. In total, 12, 19 and 15 TaHDZ were located on the A, B and D sub-genome, respectively. Chromosome groups two and four had 10 HD-Zip genes each, representing the most abundant regions, followed by group five with seven. At the same time, no TaHDZ genes were found on chromosomes 3A, 5A, 6B, or 7B (Figure 1).
Gene duplication is generally the main factor causing the expansion of the given gene family [68]. As a genetically allohexaploid species, wheat has a complex origin and evolutionary history, derived from three diploid donor species through two naturally interspecific hybridization events. As aresult, each wheat gene generally has three homologous loci arising from polyploidization [69]. Though sequence similarity and chromosome localization analysis, three TaHDZ genes, including TaHDZ9, TaHDZ15 andTaHDZ17, were found to have three copies on each of the A, B and D homoeologous chromosome. Twelve TaHDZ genes (TaHDZ1, -6, -7, -8, -10, -13, -16, -20, -23, -24, -25 and -28) contain two copies in the A, B, or D homoeologous chromosome. Just one copy of the remaining 13 genes was found in wheat chromosomes (Figure 1). Previous studies revealed that inversion and crossover might occur between homologous chromosomes during polyploidization to fractionate the coding region or delete some homologous sequences [70]. Our results indicated considerable homologous genes loss may occur in the wheat HD-Zip gene family, causing the loss of some homologous copies. The retention and dispersion of specific HD-Zip genes in homologous chromosomes could provide some insights into the mechanism of wheat chromosome interaction and genome evolution.

3.3. Phylogenetic Analysis of TaHDZ Genes

To investigate the phylogenetic relationships of the HD-Zip gene family, 46 TaHDZ proteins, together with 48 and 39 publicly available Arabidopsis and rice HDZ proteins, were selected for phylogenetic analysis. Based on the classification criteria in Arabidopsis and rice [6], the wheat HDZ proteins were clustered into four groups, I to IV. Specifically, 20, 17, 4 and 5 TaHDZ proteins were classified into groups I, II, III and IV, respectively. Group I was further divide into eight sub-groups, including group I-а (3), group I-β1 (3), group I-β2 (0), group I-б (7), group I-ε (0), group I-ξ (3), group I-r (4) and group I-φ (0) (Figure 2). Among them, group I was the most abundant in wheat, rice and Arabidopsis, accounting for 43%, 35% and 35%, respectively, similar to that of grape and sorghum having a percentage of 42% and 32%, respectively [9]. In contrast, in maize group II is the largest group, accounting for 32.72% [12,65]. Furthermore, no proteins were categorized into subgroup I-β2 and I-ε in wheat, maize, or rice but more than two members were present in both soybean and Arabidopsis (Table 2). The HD-Zip gene family distribution has been reported to be predominant with a species bias [9,12]. This result indicated that gene loss in these sub-groups might occur during the divergence of dicots and monocots. Additionally, the number of proteins in group I-б in wheat was significantly higher, with seven members, compared to one in maize, two in rice, three in Arabidopsis and two in soybean, suggesting some TaHDZ genes also expanded, although duplication events of HD-Zip genes were unusual in wheat (Table 2 and Figure 3).

3.4. Co-Expression Network between TaHDZ Genes and Other Genes in Wheat

To identify the biological function and interaction relationship between TaHDZs and other wheat genes, their interaction network was constructed using an orthology-based method. A total of nine TaHDZ genes were homologous with Arabidopsis, with 191 gene pairs of network interactions found, suggesting that TaHDZ proteins are widely involved in the wheat metabolic network and regulated diverse biological processes and pathways (Figure 4). Interacting genes were classified into three major gene ontology classes using GO annotations: diverse biological process, cellular component and molecular function (Table S2). Previous studies found that AtHB7 is involved in the regulation of organ development and plays a critical role in the response to osmotic stress [71]. In this study, TaHDZ20-B was found as the orthologous gene of AtHB7 in the wheat HD-Zip family. Further analysis found that they interacted with 71 wheat genes, including the stress-responsive gene LEA, MAPKKK18, MYB and NAC, suggesting they may be mainly involved in the response to abiotic stresses (Figure 4). The identified homology genes and the putative co-expression network analysis of TaHDZs have provided useful information for the further study of the biological function and transduction pathways of HD-Zip genes in wheat.

3.5. Conserved Motifs and Expression Profile Analysis of TaHDZ Genes

Different members of gene families generally exhibit disparities in abundance in different tissues or with different stressors [72]. To gain insight about the putative functions of the TaHDZ genes, the temporal and spatial expression profile of these identified TaHDZ genes were examined using the publicly available RNA-Seq data (Figure 5). Results showed that some tissue-specific genes, such as TaHDZ7-A/B specifically expressed in the wheat spike, were found, although most members of group I had no significant expression differences in the five organs. Furthermore, the stress-responsive group I members were also analyzed. Result found that TaHDZ15-A/B/D were expressed in normal conditions, whereas higher expression was found after 6h of cold and drought stress. Additionally, the TaHDZ8-A/B transcript level increased for up to 6h after exposure to cold, heat, drought and salt stress (Figure 5b), suggesting they play a role in response to stress.
In group II, five genes, TaHDZ10-B/D, TaHDZ25-A/D, TaHDZ28-A/D, TaHDZ3-B and TaHDZ9-A/B/D, had significant expression differentiation in different tissues, whereas TaHDZ25-A/D, TaHDZ26-D, TaHDZ27-D and TaHDZ28-A/D showed significant high expression under stresses. Among them, TaHDZ26-D and TaHDZ27-D were markedly induced by cold and heat stress, respectively. The expression patterns of group I and II genes revealed that wheat HD-Zip I and II genes are relevant for a variety of stresses. Notably, the cold-responsive genes mainly belonged to group I, whereas the salt-responsive genes belonged to group II, consistent with previous results that indicated that group I and II genes play important roles in response to various stresses in plants [73]. For groups III and IV, most of the members showed no significant expression differences in different tissues or with different stresses (Figure 5b). However, TaHDZ13-B/D, belonging to group III, had a relatively high expression in the spike and grain but weak expression in the root and leaf, suggesting involvement in the reproduction process. TaHDZ13-B/D was the one member of group IV that showed significantly up-regulation under heat stress compared to the control, meaning it could be considered heat-responsive gene (Figure 5b).
To obtain insight into the relationship between gene structure and expression of TaHDZs, the conserved motifs in these protein sequences were further predicted using the MEME program. A total of 20 conserved motifs were found (Figure 5c). The identified TaHDZs motifs ranged from 8 to 50 amino acids in length. The details of the sequence of all conserved motifs are shown in Figure S1. The motifs were unevenly distributed in those proteins, with the number of motifs ranging from 3 to 13. It showed that motif one and two, corresponding to the HD domain and motif three, corresponding to the LZ domain, were distributed in all the TaHDZ proteins. Notably, proteins in the same group seemed to share similar motif compositions. For examples, three motifs, including motif six, eight and nine were found to be related to the START domain, which was present in groups III and IV subfamilies, with the exception of TaHDZ5-D. Motif four, encoding the CPSCE domain, was present in each group II member, except for TaHDZ25, -26, -27 and -28. Additionally, the gene structure might also be involved in the control of gene expression patterns in various tissues or abiotic stresses (Figure 5b,c). For examples, motif 11, 17 and 20 only existed in TaHDZ23B/D and TaHDZ1A and these genes had relatively high expression in various tissues or under abiotic stresses. However, group I-б and IV, containing motif five, had relatively low expression in various tissues or under abiotic stresses (Figure 5b,c and Figure S1).

3.6. Expression Profiles of TaHDZ Genes under Abiotic Stress in Wheat by qRT-PCR Analysis

To validate the stress-responsive candidates, 28 TaHDZs showing differential expression underabiotic stresses, based on the RNA-Seq data, were selected to conduct quantitative polymerase chain reaction (qPCR) analysis. Results showed that these genes had differential expression under different stresses and between tolerant and susceptible genotypes. In CS, which is susceptible to salt and drought stress, a total of 21 and 20 TaHDZ genes were induced by salt and drought stress, respectively, whereas 18 were induced by salt stress in the DK genotype (salt-tolerant) and 20 were induced by drought stress in the HX genotype (drought-tolerant). In total, 13 TaHDZ genes were induced by salt stress in both DK and CS. Among them, TaHDZ11 and TaHDZ19 were down-regulated and the others were up-regulated. Sixteen TaHDZ genes were induced by drought in both HX and CS, of which TaHDZ11, TaHDZ17 andTaHDZ18 were rapidly reduced 0.06-fold, 0.39-fold and 0.26-fold under drought stress in CS, whereas these genes were strongly induced to 50-fold, 100-fold and 2.64-fold in HX under drought treatment, suggesting these genes might play a vital role in the response to drought stress in wheat (Figure 6).
A large number of HD-Zip genes have been demonstrated to regulate the abiotic stresses response in model plants [12,25,26,71], providing indications of the biological function of wheat HD-Zip genes using orthology-based predictions. Oshox22 was found to be participate in ABA-mediated signal pathways, regulating drought and salt responses in rice [74]. Through phylogenetic analysis, TaHDZ8 was found as the orthologous gene of Oshox22 in this study. RNA-Seq data showed that TaHDZ8 was only lightly expressed 6h after exposure to abiotic stress (Figure 5c). Using qRT-PCR, the results showed that the expression level of TaHDZ8 in CS under salt or drought stress was 9.50-fold and 18.61-fold higher than those of the control, respectively. The expression level of TaHDZ8 was 33.25-fold higher in DK and 8.13-fold higher in HX than in CS as the control (Figure 6). In addition, AtHB7 and AtHB12 were revealed to be involved in the fine-tuning processes associated with growth and drought stress [71]. Their orthologue counterpart in wheat—TaHDZ20 and TaHDZ28—had significant differential expressions between control and drought stress in susceptible genotypes while no significant difference in tolerant genotype (Figure 6), suggesting these two genes may play the essential role in the regulatory network of drought response in wheat.

4. Conclusions

This study systematically identifies and characterizes the wheat HD-Zip gene family at the whole genome level. A total of 46 wheat HD-Zip genes, belonging to groups I to IV, were identified, and they were unevenly distributed on wheat chromosomes. The gene structure, conserved motif and phylogenetic relationship analysis further supported the classification. Gene duplication analysis found that 13, 12 and 3 TaHDZs were found to have one, two and three homoeologous copies, respectively, suggesting homoeologous gene loss events occurred in this family. Furthermore, the co-expression network between TaHDZs and other wheat genes was constructed. Up to 191 interactions could be found, indicating that TaHDZ proteins are widely involved in the wheat metabolic network and that they regulate diverse biological processes and pathways. Finally, the expression profiles of these TaHDZ genes in different tissues and under various stress conditions were identified by RNA-Seq mapping and further validated by qRT-PCR analysis. In total, thirteen salt-responsive and sixteen drought-responsive genes were obtained, which should be considered as the candidates for future abiotic functional studies. Overall, this study provides the genetic background knowledge for genetic improvement of salt and drought tolerance in wheat, and contributes as well to reveal the molecular mechanism of stress response.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4425/9/2/70/s1, Figure S1: Sequence logos of TaHDZ proteins conserved motifs. A total of 20 motifs were identified using the MEME tool, Table S1: Primer sequences of 28 TaHDZ genes used for qRT-PCR analysis, Table S2: Details of 9 TaHDZ orthologous genes between wheat and Arabidopsis thaliana, Supplementary file 1: The multiple sequence alignment of TaHDZ proteins.

Acknowledgments

This work was supported by the State’s Key Project of Research and Development Program of China (Grant No. 2016YFD0101004 and 2016YFD0100302) and partially supported by the National Natural Science Foundation of China (Grant No. 31401373 and 31561143005).

Author Contributions

X.N. and W.S collected experimental materials. H.Y. and X.N. designed and interpretation of this experiment. H.Y., D.S., M.W. and G.X. carried out experimental work. H.Z. and X.D. analyzed data. H.Y. and D.S. prepared this manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hu, W.; Wang, L.Z.; Tie, W.W.; Yan, Y.; Ding, Z.H.; Liu, J.H.; Li, M.Y.; Ming, P.; Xu, B.Y.; Jin, Z.Q. Genome-wide analyses of the bZip family reveal their involvement in the development, ripening and abiotic stress response in banana. Sci. Rep. 2016, 6, 30203. [Google Scholar] [CrossRef] [PubMed]
  2. Chan, R.L.; Gago, G.M.; Palena, C.M.; Gonzalez, D.H. Homeoboxes in plant development. Biochim. Biophys. Acta 1998, 1442, 1–19. [Google Scholar] [CrossRef]
  3. Mcginnis, W.; Garber, R.L.; Wirz, J.; Kuroiwa, A.; Gehring, W.J. A homologous protein-coding sequence in Drosophila homeotic genes and its conservation in other metazoans. Cell 1984, 37, 403–408. [Google Scholar] [CrossRef]
  4. Ruberti, I.; Giovanna, S.; Lucchetti, S.; Morelli, G. A novel class of plant proteins containing a homeodomain with a closely linked leucine zipper motif. EMBO J. 1991, 10, 1787–1791. [Google Scholar] [PubMed]
  5. Wolfgang, F.; Phillips, J.; Salamini, F.; Bartels, D. Two dehydration-inducible transcripts from the resurrection plant Craterostigma plantagineumencode interacting homeodomain-leucine zipper proteins. Plant J. 1998, 15, 413–421. [Google Scholar]
  6. Ariel, F.D.; Manavella, P.A.; Dezar, C.A.; Chan, R.L. The true story of the HD-Zip family. Trends Plant Sci. 2007, 12, 419–426. [Google Scholar] [CrossRef] [PubMed]
  7. Meijer, A.H.; Scarpella, E.; van Dijk, E.L.; Qin, L.; Taal, A.J.; Rueb, S.; Harrington, S.E.; McCouch, S.R.; Schilperoort, R.A.; Hoge, J.H. Transcriptional repression by Oshox1, a novel homeodomain leucine zipper protein from rice. Plant J. 1997, 11, 263–276. [Google Scholar] [CrossRef] [PubMed]
  8. Tron, A.E.; Bertoncini, C.W.; Chan, R.L.; Gonzalez, D.H. Redox regulation of plant homeodomain transcription factors. J. Biol. Chem. 2002, 277, 34800–34807. [Google Scholar] [CrossRef] [PubMed]
  9. Hu, R.; Chi, X.; Chai, G.; Kong, Y.; He, G.; Wang, X.; Shi, D.; Zhang, D.; Zhou, G. Genome-wide identification, evolutionary expansion and expression profile of homeodomain-leucine zipper gene family in poplar (Populus trichocarpa). PLoS ONE 2012, 7, e31149. [Google Scholar] [CrossRef] [PubMed]
  10. Ponting, C.P.; Aravind, L. START: A lipid-binding domain in StAR, HD-ZIP and signalling proteins. Trends Biochem. Sci. 1999, 24, 130–132. [Google Scholar] [CrossRef]
  11. Schrick, K.; Nguyen, D.; Karlowski, W.M.; Mayer, K.F. START lipid/sterol-binding domains are amplified in plants and are predominantly associated with homeodomain transcription factors. Genome Biol. 2004, 5, R41. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Chen, X.; Zhu, C.; Zhao, H.L.; Zhao, Y.; Cheng, B.J.; Xiang, Y. Genome-wide analysis of soybean HD-ZIP gene family and expression profiling under salinity and drought treatments. PLoS ONE 2014, 9, e87156. [Google Scholar] [CrossRef] [PubMed]
  13. Mukherjee, K.; Bürglin, T.R. MEKHLA, a novel domain with similarity to PAS domains, is fused to plant homeodomain-leucine zipper III proteins. Plant Physiol. 2006, 140, 1142–1150. [Google Scholar] [CrossRef] [PubMed]
  14. Sessa, G.; Steindler, C.; Morelli, G.; Ruberti, I. The Arabidopsis Athb-8,-9 and genes are members of a small gene family coding for highly related HD-ZIP proteins. Plant Mol. Biol. 1998, 38, 609–622. [Google Scholar] [CrossRef] [PubMed]
  15. Abe, M.; Katsumata, H.; Komeda, Y.; Takahashi, T. regulation of shoot epidermal cell differentiation by a pair of homeodomain proteins in Arabidopsis. Development 2003, 130, 635–643. [Google Scholar] [CrossRef] [PubMed]
  16. Di Cristina, M.; Sessa, G.; Dolan, L.; Linstead, P.; Baima, S.; Ruberti, I.; Morelli, G. The Arabidopsis Athb-10 (GLABRA2) is an HD-Zip protein required for regulation of root hair development. Plant J. 1996, 10, 393–402. [Google Scholar] [CrossRef] [PubMed]
  17. Kubo, H.; Peeters, A.J.; Aarts, M.G.; Pereira, A.; Koornneef, M. ANTHOCYANINLESS2, a homeobox gene affecting anthocyanin distribution and root development in Arabidopsis. Plant Cell. 1999, 11, 1217–1226. [Google Scholar] [CrossRef] [PubMed]
  18. Yuan, D.; Tang, Z.; Wang, M.; Gao, W.; Tu, L.; Jin, X.; Chen, L.; He, Y.; Zhang, L.; Zhu, L.; et al. the genome sequence of Sea-Island cotton (Gossypium barbadense) provides insights into the allopolyploidization and development of superior spinnablefibres. Sci. Rep. 2015, 5. [Google Scholar] [CrossRef]
  19. Mao, H.; Yu, L.; Li, Z.; Liu, H.; Han, R. Molecular evolution and gene expression differences within the HD-Zip Transcription Factor Family of Zea mays L. Genetica 2016, 144, 243–257. [Google Scholar] [CrossRef] [PubMed]
  20. Kovalchuk, N.; Chew, W.; Sornaraj, P.; Borisjuk, N.; Yang, N.; Singh, R.; Bazanova, N.; Shavrukov, Y.; Guendel, A.; Munz, E.; et al. The Homeodomain Transcription Factor TaHD-Zipl-2 from wheat regulates frost tolerance, flowering time and spike development in transgenic barley. New Phytol. 2016, 211, 671–687. [Google Scholar] [CrossRef] [PubMed]
  21. Capella, M.; Ribone, P.A.; Arce, A.L.; Chan, R.L. Arabidopsis thaliana HomeoBox 1 (AtHB1), a homedomain-leucine zipper I (HD-Zip I) transcription factor, is regulated by PHYTOCHROME-INTERACTING FACTOR 1 to promote hypocotyl elongation. New Phytol. 2015, 207, 669–682. [Google Scholar] [CrossRef] [PubMed]
  22. Aoyama, T.; Dong, C.H.; Wu, Y.; Carabelli, M.; Sessa, G.; Ruberti, I.; Morelli, G.; Chua, N.H. Ectopic expression of the Arabidopsis transcriptional activator Athb-1 alters leaf cell fate in tobacco. Plant Cell. 1995, 7, 1773–1785. [Google Scholar] [CrossRef] [PubMed]
  23. Ni, Y.; Wang, X.; Li, D.; Wu, Y.; Xu, W.; Li, X. novel cotton homeobox gene and its expression profiling in root development and in response to stresses and phytohormones. Acta Biochim. Biophys. Sin. 2008, 40, 78–84. [Google Scholar] [CrossRef]
  24. Harris, J.C.; Sornaraj, P.; Taylor, M.; Bazanova, N.; Baumann, U.; Lovell, B.; Langridge, P.; Lopato, S.; Hrmova, M. Molecular interactions of the γ-clade homeodomain-leucine zipper class I transcription factors during the wheat response to water deficit. Plant Mol. Biol. 2016, 90, 435–452. [Google Scholar] [CrossRef] [PubMed]
  25. Ge, X.X.; Liu, Z.; Wu, X.M.; Chai, L.J.; Guo, W.W. Genome-wide identification, classification and analysis of HD-Zip gene family in citrus and its potential roles in somatic embryogenesis regulation. Gene 2015, 574, 61–68. [Google Scholar] [CrossRef] [PubMed]
  26. Song, S.; Chen, Y.; Zhao, M.; Zhang, W.H. A novel Medicagotruncatula HD-Zip gene, MtHB2, is involved in abiotic stress responses. Environ. Exp. Bot. 2012, 80, 1–9. [Google Scholar] [CrossRef]
  27. Turchi, L.; Carabelli, M.; Ruzza, V.; Possenti, M.; Sassi, M.; Peñalosa, A.; Sessa, G.; Salvi, S.; Forte, V.; Morelli, G.; et al. Arabidopsis HD-Zip II transcription factors control apical embryo development and meristem function. Development 2013, 140, 2118–2129. [Google Scholar] [CrossRef] [PubMed]
  28. Dezar, C.A.; Giacomelli, J.I.; Manavella, P.A.; Ré, D.A.; Alves-Ferreira, M.; Baldwin, I.T.; Bonaventure, G.; Chan, R.L. HAHB10, a Sunflower HD-ZIP II transcription factor, participates in the induction of flowering and in the control of phytohormone-mediated responses to biotic stress. J. Exp. Bot. 2011, 62, 1061–1076. [Google Scholar] [CrossRef] [PubMed]
  29. Ooi, S.E.; Ramli, Z.; Kulaveerasingam, H.; Ong-Abdullah, M. EgHOX1, a HD-Zip II gene, is highly expressed during early oil palm (Elaeis guineensis Jacq.) somatic embryogenesis. Plant Gene 2016, 8, 16–25. [Google Scholar] [CrossRef]
  30. Franco, D.M.; Silva, E.M.; Saldanha, L.L.; Adachi, S.A.; Schley, T.R.; Rodrigues, T.M.; Dokkedal, A.L.; Nogueira, F.T.; Rolim de Almeida, L.F. Flavonoids modify root growth and modulate expression of SHORT-ROOT and HD-ZIP III. J. Plant Physiol. 2015, 188, 89–95. [Google Scholar] [CrossRef] [PubMed]
  31. Landau, U.; Lior, A.; Leor, E.W. The ERECTA, CLAVATA and Class III HD-ZIP pathways display synergistic interactions in regulating floral meristem activities. PLoS ONE 2015, 10, e0125408. [Google Scholar] [CrossRef] [PubMed]
  32. Hawker, N.P.; Bowman, J.L. Roles for class III HD-Zip and KANADI genes in Arabidopsis root development. Plant Physiol. 2004, 4, 2261–2270. [Google Scholar] [CrossRef] [PubMed]
  33. Zhu, Y.; Song, D.; Sun, J.; Wang, X.; Li, L. PtrHB7, a class III HD-Zip gene, plays a critical role in regulation of vascular cambium differentiation in Populus. Mol. Plant. 2013, 6, 1331–1343. [Google Scholar] [CrossRef] [PubMed]
  34. Carlsbecker, A.; Lee, J.Y.; Roberts, C.J.; Dettmer, J.; Lehesranta, S.; Zhou, J.; Lindgren, O.; Moreno-Risueno, M.A.; Vatén, A.; Thitamadee, S.; et al. Cell signalling by microRNA165/6 directs gene dose-dependent root cell fate. Nature 2010, 465, 316–321. [Google Scholar] [CrossRef] [PubMed]
  35. Li, Z.; Zhang, C.; Guo, Y.; Niu, W.; Wang, Y.; Xu, Y. Genome-wide analysis of HD-Zip genes in grape (Vitis vinifera). Tree Genet. Genomes. 2014, 1, 1–11. [Google Scholar]
  36. Kamata, N.; Okada, H.; Komeda, Y.; Takahashi, T. Mutations in epidermis-specific HD-Zip IV genes affect floral organ identity in Arabidopsis thaliana. Plant J. 2013, 75, 430–440. [Google Scholar] [CrossRef] [PubMed]
  37. Kamata, N.; Okada, H.; Komeda, Y.; Takahashi, T. The HD-Zip IV Transcription Factor OCL4 is necessary for trichome patterning and anther development in maize. Plant J. 2009, 59, 883–894. [Google Scholar]
  38. Pan, Y.; Bo, K.; Cheng, Z.; Weng, Y. The loss-of-function GLABROUS 3 Mutation in Cucumber Is Due to LTR-retrotransposon Insertion in a class IV HD-ZIP transcription factor gene CsGL3 that is epistatic over CsGL1. BMC Plant Biol. 2015, 15, 302. [Google Scholar] [CrossRef] [PubMed]
  39. Gill, B.S.; Appels, R.; Botha-Oberholster, A.M.; Buell, C.R.; Bennetzen, J.L.; Chalhoub, B.; Chumley, F.; Dvorák, J.; Iwanaga, M.; Keller, B.; et al. A workshop report on wheat genome sequencing. Genetics 2004, 168, 1087–1096. [Google Scholar] [CrossRef] [PubMed]
  40. International Wheat Genome Sequencing Consortium. A chromosome-based draft sequence of the hexaploid bread wheat (Triticum aestivum) genome. Science 2014, 345, 1251788. [Google Scholar]
  41. Kovalchuk, N.; Wu, W.; Eini, O.; Bazanova, N.; Pallotta, M.; Shirley, N.; Singh, R.; Ismagul, A.; Eliby, S.; Johnson, A.; et al. The scutellar vascular bundle–specific promoter of the wheat HD-Zip IV transcription factor shows similar spatial and temporal activity in transgenic wheat, barley and rice. Plant Biotechnol. J. 2012, 10, 43–53. [Google Scholar] [CrossRef] [PubMed]
  42. Yang, Y.; Luang, S.; Harris, J.; Riboni, M.; Li, Y.; Bazanova, N.; Hrmova, M.; Haefele, S.; Kovalchuk, N.; Lopato, S. Overexpression of the class I homeodomain transcription factor TaHD-ZipI-5 increases drought and frost tolerance in transgenic wheat. Plant Biotechnol. J. 2017. [Google Scholar] [CrossRef] [PubMed]
  43. Kersey, P.J.; Allen, J.E.; Armean, I.; Boddu, S.; Bolt, B.J.; Carvalho-Silva, D.; Christensen, M.; Davis, P.; Falin, L.J.; Grabmueller, C.; et al. Ensembl Genomes 2016: More genomes, more complexity. Nucleic Acids Res. 2015, 44, D574–D580. [Google Scholar]
  44. Finn, R.D.; Mistry, J.; Schuster-Böckler, B.; Griffiths-Jones, S.; Hollich, V.; Lassmann, T.; Moxon, S.; Marshall, M.; Khanna, A.; Durbin, R.; et al. PFAM: Clans, web tools and services. Nucleic Acids Res. 2006, 34, D247–D251. [Google Scholar] [CrossRef] [PubMed]
  45. Wheeler, T.J.; Eddy, S.R. NHMMER: DNA homology search with profile HMMs. Bioinformatics 2013, 29, 2487–2489. [Google Scholar] [CrossRef] [PubMed]
  46. The Conserved Domain Database. Available online: https://www.ncbi.nlm.nih.gov/cdd (accessed on 15 November 2017).
  47. Compute pI/mw Tool. Available online: https://web.expasy.org/compute_pi/ (accessed on 20 November 2017).
  48. Chou, K.C.; Shen, H.B. Cell-PLoc: A package of web-servers for predicting subcellular localization of proteins in various organisms. Nat. Protoc. 2008, 3, 153–162. [Google Scholar] [CrossRef] [PubMed]
  49. Larkin, M.; Blackshields, G.; Brown, N.; Chenna, R.; McGettigan, P.; McWilliam, H.; Valentin, F.; Wallace, I.M.; Wilm, A.; Lopez, R.; et al. Clustal W and Clustal X version 2.0. Bioinformatics 2007, 23, 2947–2948. [Google Scholar] [CrossRef] [PubMed]
  50. Tamura, K.; Stecher, G.; Peterson, D.; Filipski, A.; Kumar, S. MEGA6: Molecular evolutionary genetics analysis version 6.0. Mol. Boil. Evol. 2013, 30, 2725–2729. [Google Scholar] [CrossRef] [PubMed]
  51. Wang, M.; Yue, H.; Feng, K.; Deng, P.; Song, W.; Nie, X. Genome-wide identification, phylogeny and expressional profiles of mitogen activated protein kinase kinasekinase (MAPKKK) gene family in bread wheat (Triticum aestivum L.). BMC Genom. 2016, 17, 668. [Google Scholar] [CrossRef] [PubMed]
  52. Krzywinski, M.; Schein, J.; Birol1, İ.; Connors, J.; Gascoyne, R.; Horsman, D.; Jones, S.J.; Marra, M.A. Circos: An information aesthetic for comparative genomics. Genome Res. 2009, 19, 1639–1645. [Google Scholar] [CrossRef] [PubMed]
  53. Hu, B.; Jin, J.; Guo, A.Y.; Zhang, H.; Luo, J.; Gao, Ge. GSDS 2.0: An upgraded gene feature visualization server. Bioinformatics 2015, 31, 1296–1297. [Google Scholar] [CrossRef] [PubMed]
  54. Bailey, T.L.; Boden, M.; Buske, F.A.; Frith, M.; Grant, C.E.; Clementi, L.; Ren, J.; Li, W.W.; Noble, W.S. MEME SUITE: Tools for motif discovery and searching. Nucleic Acids Res. 2009, 37, W202–W208. [Google Scholar] [CrossRef] [PubMed]
  55. Seq Repository in URGI Wheat Database. Available online: https://urgi.versailles.inra.fr/files/RNASeqWheat/ (accessed on 30 November 2017).
  56. FastQC. Available online: http://www.bioinformatics.babraham.ac.uk/projects/fastqc/ (accessed on 30 November 2017).
  57. Trapnell, C.; Pachter, L.; Salzberg, S.L. TopHat: Discovering splice junctions with RNA-Seq. Bioinformatics 2009, 25, 1105–1111. [Google Scholar] [CrossRef] [PubMed]
  58. Cufflinks 2.2.1. Available online: http://cole-trapnell-lab.github.io/cufflinks/releases/v2.2.1/ (accessed on 10 November 2017).
  59. Lee, T.; Yang, S.; Kim, E.; Ko, Y.; Hwang, S.; Shin, J.; Shim, J.E.; Shim, H.; Kim, H.; Kim, C.; et al. AraNet v2: An improved database of co-functional gene networks for the study of Arabidopsis thaliana and 27 other nonmodel plant species. Nucleic Acids Res. 2015, 43, D996–D1002. [Google Scholar] [CrossRef] [PubMed]
  60. Maere, S.; Heymans, K.; Kuiper, M. BiNGO: A Cytoscape plugin to assess overrepresentation of gene ontology categories in biological networks. Bioinformatics 2005, 21, 3448–3449. [Google Scholar] [CrossRef] [PubMed]
  61. Primer Premier: A Comprehensive PCR Primer Design Software. Available online: http://www.premierbiosoft.com/primerdesign/ (accessed on 1 December 2017).
  62. Xu, L.; Tang, Y.; Gao, S.; Su, S.; Hong, L.; Wang, W.; Fang, Z.; Li, X.; Ma, J.; Quan, W.; et al. Comprehensive analyses of the annexin gene family in wheat. BMC Genom. 2016, 17, 415. [Google Scholar] [CrossRef] [PubMed]
  63. Qiao, L.; Zhang, X.; Han, X.; Zhang, L.; Li, X.; Zhan, H.; Ma, J.; Luo, P.; Zhang, W.; Cui, L.; et al. A genome-wide analysis of the auxin/indole-3-acetic acid gene family in hexaploid bread wheat (Triticum aestivum L.). Front. Plant Sci. 2015, 6, 770. [Google Scholar] [CrossRef] [PubMed]
  64. Belamkar, V.; Weeks, N.T.; Bharti, A.K.; Farmer, A.D.; Graham, M.A.; Cannon, S.B. Comprehensive characterization and RNA-Seq Profiling of the HD-Zip transcription factor family in soybean (Glycine max) during dehydration and salt stress. BMC Genom. 2014, 15, 950. [Google Scholar] [CrossRef] [PubMed]
  65. Zhao, Y.; Zhou, Y.; Jiang, H.; Li, X.; Gan, D.; Peng, X.; Zhu, S.; Cheng, B. Systematic analysis of sequences and expression patterns of drought-responsive members of the HD-Zip gene family in maize. PLoS ONE 2011, 12, e28488. [Google Scholar] [CrossRef] [PubMed]
  66. Song, A.; Li, P.; Xin, J.; Chen, S.; Zhao, K.; Wu, D.; Fan, Q.; Gao, T.; Chen, F.; Guan, Z. Transcriptome-wide survey and expression profile analysis of putative chrysanthemum HD-Zip I and II genes. Genes 2016, 7, 19. [Google Scholar] [CrossRef] [PubMed]
  67. Jiang, H.Y.; Jing, J.; Huan, L.; Qing, D.; Hanwei, Y.; Defang, G.; Wei, Z.; Suwen, Z. Genome-wide analysis of HD-Zip genes in grape (Vitis vinifera). Tree Genet. Genom. 2015, 11, 827. [Google Scholar] [CrossRef]
  68. Zhang, J.Z. Evolution by gene duplication: An update. Trends Ecol. Evol. 2003, 18, 292–298. [Google Scholar] [CrossRef]
  69. Feldman, M.; Levy, A.A. Allopolyploidy—A shaping force in the evolution of wheat genomes. Cytogenet. Genome Res. 2005, 109, 250–258. [Google Scholar] [CrossRef] [PubMed]
  70. Lynch, M.; Force, A. The probability of duplicate gene preservation by subfunctionalization. Genetics 2000, 154, 459–473. [Google Scholar] [PubMed]
  71. Ré, D.A.; Capella, M.; Bonaventure, G.; Chan, R.L. Arabidopsis AtHB7 and AtHB12 evolved divergently to fine tune processes associated with growth and responses to water stress. BMC Plant Biol. 2014, 14, 150. [Google Scholar] [CrossRef] [PubMed]
  72. Yue, H.; Wang, M.; Liu, S.; Du, X.; Song, W.; Nie, X. Transcriptome-wide identification and expression profiles of the WRKY transcription factor family in Broomcorn millet (Panicum miliaceum L.). BMC Genom. 2016, 17, 343. [Google Scholar] [CrossRef] [PubMed]
  73. Schena, M.; Davis, R.W. structure of homeobox-leucine zipper genes suggests a model for the evolution of gene families. Proc. Natl. Acad. Sci. USA 1994, 91, 8393–8397. [Google Scholar] [CrossRef] [PubMed]
  74. Zhang, S.; Haider, I.; Kohlen, W.; Jiang, L.; Bouwmeester, H.; Meijer, A.H.; Schluepmann, H.; Liu, C.M.; Ouwerkerk, P.B. Function of the HD-Zip I gene Oshox22 in ABA-mediated drought and salt tolerances in rice. Plant Mol. Biol. 2012, 80, 571–585. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Chromosomal localization and gene duplication identified in wheat. Seven chromosomes of wheat A, B and D sub-genomes are displayed in different colors. Duplicated gene pairs are exhibited in linked lines with the corresponding color.
Figure 1. Chromosomal localization and gene duplication identified in wheat. Seven chromosomes of wheat A, B and D sub-genomes are displayed in different colors. Duplicated gene pairs are exhibited in linked lines with the corresponding color.
Genes 09 00070 g001
Figure 2. Phylogenetic analysis of HD-Zip proteins among wheat (46), Arabidopsis (48) and rice (39).
Figure 2. Phylogenetic analysis of HD-Zip proteins among wheat (46), Arabidopsis (48) and rice (39).
Genes 09 00070 g002
Figure 3. The distribution of HD-Zip transcription factors from wheat, maize, Arabidopsis, rice and soybean.
Figure 3. The distribution of HD-Zip transcription factors from wheat, maize, Arabidopsis, rice and soybean.
Genes 09 00070 g003
Figure 4. The co-expression network of TaHDZ genes in wheat according to the orthologues in Arabidopsis.
Figure 4. The co-expression network of TaHDZ genes in wheat according to the orthologues in Arabidopsis.
Genes 09 00070 g004
Figure 5. (a) Phylogenetic relationships; (b) Expression patterns; (c) Conserved motifs compositions of the 46 HD-Zip genes in wheat. (a) The phylogenetic tree was constructed based on the full-length protein sequences using MEGA6.0 (http://web.megasoftware.net/); (b) Hierarchical clustering of the relative expression level of TaHDZ genes. RNA-Seq data of five tissues and four stresses in Chinese spring was used to analysis expression pattern. The heat map was drawn in Log10-transformed expression values. Red or green colors represent decreased or increased expression level in each sample, respectively; (c) Multiple EM for motif elicitation (MEME) program [54] was used to predict conserved motifs. Each motif is represented by a different colored box.
Figure 5. (a) Phylogenetic relationships; (b) Expression patterns; (c) Conserved motifs compositions of the 46 HD-Zip genes in wheat. (a) The phylogenetic tree was constructed based on the full-length protein sequences using MEGA6.0 (http://web.megasoftware.net/); (b) Hierarchical clustering of the relative expression level of TaHDZ genes. RNA-Seq data of five tissues and four stresses in Chinese spring was used to analysis expression pattern. The heat map was drawn in Log10-transformed expression values. Red or green colors represent decreased or increased expression level in each sample, respectively; (c) Multiple EM for motif elicitation (MEME) program [54] was used to predict conserved motifs. Each motif is represented by a different colored box.
Genes 09 00070 g005
Figure 6. The expression profiles of 28 TaHDZ genes that may be involved in the response to salt or drought stress in different wheat varieties using quantitative real-time polymerase chain reaction (qRT-PCR) analysis. For salt stress, two-week-old seedlings of variety Dekang No. 685 (DK) and Chinese Spring (CS) were treated with 200 mM sodium chloride (NaCl) for 24 h. Hanxuan No. 10 (HX) and CS were treated with 19.2% polyethylene glycol (PEG) for 24 h to represent the drought treatment. The control was an untreated seedling. Three biological replicates for each sample were performed and bars represented the standard deviations of the mean. Asterisks on top of the bars indicating statistically significant differences between the stress and counterpart controls (** p< 0.01, Student’s t-test). Gene expression profiles were evaluated using the 2−∆∆C method.
Figure 6. The expression profiles of 28 TaHDZ genes that may be involved in the response to salt or drought stress in different wheat varieties using quantitative real-time polymerase chain reaction (qRT-PCR) analysis. For salt stress, two-week-old seedlings of variety Dekang No. 685 (DK) and Chinese Spring (CS) were treated with 200 mM sodium chloride (NaCl) for 24 h. Hanxuan No. 10 (HX) and CS were treated with 19.2% polyethylene glycol (PEG) for 24 h to represent the drought treatment. The control was an untreated seedling. Three biological replicates for each sample were performed and bars represented the standard deviations of the mean. Asterisks on top of the bars indicating statistically significant differences between the stress and counterpart controls (** p< 0.01, Student’s t-test). Gene expression profiles were evaluated using the 2−∆∆C method.
Genes 09 00070 g006
Table 1. Characteristics of the putative homeodomain-leucine zipper (HD-Zip) proteins in wheat.
Table 1. Characteristics of the putative homeodomain-leucine zipper (HD-Zip) proteins in wheat.
GeneSequence IDLocationAA LengthPIMWSubcellular Location
TaHDZ1-ATraes_1AL_0BE456AC0.11A:80794504-808050378405.6592,041.09Chloroplast
TaHDZ1-BTraes_1BL_43408C9B0.21B:202405337-2024116296205.8068,027.69Chloroplast
TaHDZ2-DTraes_1DL_9FB53E48A.11DL:scaff527273:1-6821289.0014,621.61Nucleus
TaHDZ3-BTraes_1BL_BCA60D8B6.21BL:scaff3858366:4943-62333026.6833,291.33Nucleus
TaHDZ4-ATraes_1AL_1444D461A.11A:192016994-1920175473268.5034,647.92Nucleus
TaHDZ5-DTraes_2DL_036F2A3FC.12DL:scaff9746565:6-7122518.3126,889.50Chloroplast
TaHDZ6-BTraes_2BS_BD0ED621D.22B:203096624-2030980802387.1225,888.92Chloroplast
TaHDZ6-DTraes_2DS_20F748657.22D:47948060-479494992386.7625,755.72Chloroplast
TaHDZ7-ATraes_2AL_CC3E5591E.12A:214032194-2140338142188.8424,308.27Peroxisome
TaHDZ7-BTraes_2BL_419CEED79.12BL:scaff8047670:2819-38021969.5421,615.02Nucleus
TaHDZ8-ATraes_2AL_BFB0C6D4C.12A:92585473-925866971835.1520,209.38Chloroplast
TaHDZ8-BTraes_2BL_B69300543.12BL:scaff8082479:11478-125722595.2029,003.43Nucleus
TaHDZ9-ATraes_2AL_EF9549D16.12AL:scaff6334009:1565-33262278.8425,675.08Nucleus
TaHDZ9-BTraes_2BL_02479C76A.12B:265061413-2650626971599.7818,432.80Nucleus
TaHDZ9-DTraes_2DL_67F1183B2.12D:25397187-253989792308.8425,914.27Nucleus
TaHDZ10-BTRAES3BF043500070CFD_t13B:8112566-81135942288.6425,282.62Nucleus
TaHDZ10-DTraes_3DS_7CCB5ECD2.13D:813820-8145061839.1220,819.69Nucleus
TaHDZ11-BTRAES3BF026400090CFD_t13B:424323223-4243267622229.4224,704.13Nucleus
TaHDZ12-BTRAES3BF023000040CFD_t13B:530484425-5304879597557.8081,462.58Chloroplast
TaHDZ13-BTRAES3BF075200070CFD_t13B:554475326-5544797046748.4873,671.85Chloroplast
TaHDZ13-DTraes_3DL_8AAFB7B06.13D:86091404-860950214468.3349,134.62Nucleus
TaHDZ14-ATraes_4AL_822582A19.14AL:scaff7079911:4919-64382669.2628,237.92Chloroplast
TaHDZ15-ATraes_4AS_F04DD4409.14AS:scaff5975837:1-18022786.7628,363.81Chloroplast
TaHDZ15-BTraes_4BL_BE3E058A6.14BL:scaff7026111:366-30063259.2433,237.48Nucleus
TaHDZ15-DTraes_4DL_88ABAD6C0.14DL:scaff14448085:2765-54093306.1736,411.50Nucleus
TaHDZ16-ATraes_4AL_99A941299.14A:184972869-1849751113184.9434,975.56Mitochondrion
TaHDZ16-BTraes_4BL_ECD20BE67.14B:292827483-2928296703164.9935,099.79Nucleus
TaHDZ17-ATraes_4AS_1EA23DE08.14AS:scaff3077305:961-24712316.2425,530.54Peroxisome
TaHDZ17-BTraes_4BL_BE10705D5.24B:282781983-2827834902336.0425,576.61Peroxisome
TaHDZ17-DTraes_4DL_4798D0BBD.14D:68945470-689469692346.2425,708.82Chloroplast
TaHDZ18-BTraes_4BL_78DD63002.14BL:scaff6966681:8297-103412058.9422,631.64Chloroplast
TaHDZ19-BTraes_5BL_4A3874701.15B:170563162-1705646903556.2737,049.33Nucleus
TaHDZ20-BTraes_5BL_9C32B27E2.15BL:scaff10897212:246-23412495.0227,502.62Nucleus
TaHDZ20-DTraes_5DL_96F9EED93.25DL:scaff4539911:3583-57402495.0227,488.59Nucleus
TaHDZ21-BTraes_5BL_028D02DF6.15B:84457851-844600162994.8632,566.01Chloroplast
TaHDZ22-BTraes_5BL_5DE02D63E.15BL:scaff10833801:1080-27802694.7028,884.88Chloroplast
TaHDZ23-BTraes_5BS_360DD5644.15B:18562939-185690078796.6195,339.96Chloroplast
TaHDZ23-DTraes_5DS_50846FD0C.15D:21777479-217836278836.7395,733.49Chloroplast
TaHDZ24-ATraes_6AL_36AB0312C.16A:185224197-1852263703404.6137,072.67Chloroplast
TaHDZ24-DTraes_6DL_FF4C8C4AB.16D:138591996-1385942813404.6737,147.85Chloroplast
TaHDZ25-ATraes_6AS_3E534A2C1.16AS:scaff4406943:3994-51662268.7224,538.57Chloroplast
TaHDZ25-DTraes_6DS_17B737547.16D:24274468-242757122258.3824,759.74Nucleus
TaHDZ26-DTraes_6DS_D281B7D32.16D:23048583-230496612059.8722,606.56Nucleus
TaHDZ27-DTraes_6DS_F00EB2E01.16D:21598812-215995341929.7920,888.56Chloroplast
TaHDZ28-ATraes_7AL_44206BE21.17A:149156319-14915723716910.6819,423.03Mitochondrion
TaHDZ28-DTraes_7DL_2FE5181AF.17DL:scaff1534355:1875-30771859.5020,757.18Nucleus
ID: identity; AA: amino acids; PI: isoelectric point; MW: molecular weight.
Table 2. The number of HD-Zip genes in wheat, maize, Arabidopsis, rice and soybean.
Table 2. The number of HD-Zip genes in wheat, maize, Arabidopsis, rice and soybean.
SpeciesGroup I-aGroup I-β1Group I-β2Group I-εGroup I-бGroup I-ξGroup I-rGroup I-φGroup IIGroup IIIGroup IVTotal
wheat33007340174546
Maize340015401851555
Rice23002430125839
Arabidopsis423230211051648
Soybean8422282027121986

Share and Cite

MDPI and ACS Style

Yue, H.; Shu, D.; Wang, M.; Xing, G.; Zhan, H.; Du, X.; Song, W.; Nie, X. Genome-Wide Identification and Expression Analysis of the HD-Zip Gene Family in Wheat (Triticum aestivum L.). Genes 2018, 9, 70. https://doi.org/10.3390/genes9020070

AMA Style

Yue H, Shu D, Wang M, Xing G, Zhan H, Du X, Song W, Nie X. Genome-Wide Identification and Expression Analysis of the HD-Zip Gene Family in Wheat (Triticum aestivum L.). Genes. 2018; 9(2):70. https://doi.org/10.3390/genes9020070

Chicago/Turabian Style

Yue, Hong, Duntao Shu, Meng Wang, Guangwei Xing, Haoshuang Zhan, Xianghong Du, Weining Song, and Xiaojun Nie. 2018. "Genome-Wide Identification and Expression Analysis of the HD-Zip Gene Family in Wheat (Triticum aestivum L.)" Genes 9, no. 2: 70. https://doi.org/10.3390/genes9020070

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop