Next Article in Journal
Mechanistic Insights into Stereospecific Antifungal Activity of Chiral Fungicide Prothioconazole against Fusarium oxysporum F. sp. cubense
Previous Article in Journal
Thalamic T-Type Calcium Channels as Targets for Hypnotics and General Anesthetics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genome-Wide Identification and Analysis of bZIP Gene Family and Resistance of TaABI5 (TabZIP96) under Freezing Stress in Wheat (Triticum aestivum)

College of Life Science, Northeast Agricultural University, Harbin 150030, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(4), 2351; https://doi.org/10.3390/ijms23042351
Submission received: 11 January 2022 / Revised: 2 February 2022 / Accepted: 15 February 2022 / Published: 21 February 2022
(This article belongs to the Topic Plant Functional Genomics and Crop Genetic Improvement)

Abstract

:
The basic leucine zipper (bZIP) regulates plant growth and responds to stress as a key transcription factor of the Abscisic acid (ABA) signaling pathway. In this study, TabZIP genes were identified in wheat and the gene structure, physicochemical properties, cis-acting elements, and gene collinearity were analyzed. RNA-Seq and qRT-PCR analysis showed that ABA and abiotic stress induced most TabZIP genes expression. The ectopic expression of TaABI5 up-regulated the expression of several cold-responsive genes in Arabidopsis. Physiological indexes of seedlings of different lines under freezing stress showed that TaABI5 enhanced the freezing tolerance of plants. Subcellular localization showed that TaABI5 is localized in the nucleus. Furthermore, TaABI5 physically interacted with cold-resistant transcription factor TaICE1 in yeast two-hybrid system. In conclusion, this study identified and analyzed members of the TabZIP gene family in wheat. It proved for the first time that the gene TaABI5 affected the cold tolerance of transgenic plants and was convenient for us to understand the cold resistance molecular mechanism of TaABI5. These results will provide a new inspiration for further study on improving plant abiotic stress resistance.

1. Introduction

Wheat is one of the most important crops and is an important plant protein resource around the world. Low temperature stresses will seriously affect wheat’s growth state and then lead to yield reduction. Chilling stress and freezing stress can damage cell membranes and increase ROS in the body, thus affecting the growth and physiological state of plants, especially the yield and quality of wheat and other field crops [1]. Under the stress condition, plants can produce stress response to adapt to the adverse factors through a variety of pathways, so as to reduce the negative influence of the environment on themselves [2]. Plant transcription factors (TFs) play crucial roles in the regulatory and biological processes under various environmental stresses [3]. TFs are sequence-specific binding proteins that bind to the promoter regions of specific target genes to regulate their transcription [4,5]. It controls cell processes such as signal transduction, cell morphogenesis and resistance to environmental stress in plants [6,7].
The basic leucine zipper (bZIP) transcription factors family member is one of the most diverse and largest transcription factors in eukaryotes [8,9]. Leucine zipper is a dimerization motif, consisting of a helical structure composed of seven nucleotide repeats of leucine or other large hydrophobic amino acids. These helical structures ensure the stability and specificity of the dimerization reaction [10]. The domain of bZIP is 60 to 80 amino acids in length and contains two regions with diverse functions [3]. And DNA binding base region is highly conserved, with an invariant N-x7-R/K motif. Plant bZIP protein preferentially binds to the DNA sequence of ACGT core, such as A-box (TACGTA), C-box (GACGTC), and G-box (CACGTG). The binding specificity is regulated by flanking nucleotides [11].
To date, the bZIP gene family in many plants has been extensively studied. The bZIP genes are involved in important regulatory processes of plant growth and development, such as regulating root development [12], promoting anthocyanin accumulatio [13], and integrating transient abscisic acid and glucose signals [14]. Similarly, bZIP genes also participate in regulating the physiological responses of plants under stress. bZIP genes can affect salt and drought tolerance of many species, such as tomato [15], rice [16], soybean [17], and panax ginseng [18]. The cold signal of pear is regulated by PbbZIP [19]. mrna30280 and mrna11827, as FvbZIP genes, are involved in the heat stress response pathway [20]. A variety of bZIP proteins are also involved in regulating different hormones. MabZIP74 can interact with MaMAPK11-3 and regulate the expression of ethylene synthesis gene MAaco1/4 in bananas [21] and OsbZIP81 positively regulates JA levels by targeting JA regulatory genes [22]. The bZIP transcription factor BZI-1 of tobacco is involved in auxin-mediated growth [23]. VvbZIP60 contributes importantly to SA signaling against by up-regulating PR1 in grape [24].
The bZIP gene in wheat also has a variety of biological functions. Overexpression of TabZIP28 enhanced heat tolerance of transgenic Arabidopsis thaliana. [25]. As a positive regulator of wheat stripe rust resistance, TabZIP74 is involved in root development through mRNA splicing [26]. TabZIP3 is involved in the salt stress response of wheat [27]. As a bZIP transcription factor, TaABP1 had better drought resistance and was strongly expressed under ABA treatment [28]. TabZIP15 can improve salt tolerance of wheat [29]. TaFDL2 interacts with TabZIP8 protein to jointly regulate ABA signaling in response to drought stress [30]. The osmotic stress tolerance of wheat was closely related to the expression level and trend of TaDREB1-D [31]. TabZIP14-B is involved in salt tolerance and frost resistance in plants [32]. These results suggest that bZIP transcription factors are involved in various hormone regulation in different species, and the TabZIP gene may affect plant growth and stress tolerance through hormonal signaling pathways.
Representatively, Abscisic acid (ABA) has often been reckoned as the universal stress phytohormone due to its significant influence on plant growth regulation during biotic and abiotic stress responses [33]. The bZIP transcription factor in plants is the largest of the known ABA-induced DNA-binding protein family and can effectively trigger downstream genes during abiotic stress [8]. ABA senses stress signals and forms a complex with PYL/PYL/RCAR receptor and PP2C phosphatase, preventing PP2C phosphatase from dephosphorylating SnRK2 kinase, resulting in phosphorylation and activation of bZIP transcription factor by SnRK2 kinase, such as ABA-insensitive 5 (ABI5) and ABRE binding factor/ABRE binding protein (ABFs/AREBs) [34]. Thereinto, ABI5 (ABA-INSENSITIVE 5) was named after the fifth ABA-non-responsive mutant identified by Finkelstein et al. [35]. As a key transcription factor in the ABA signaling pathway, ABI5 is also involved in plant growth [36] and metabolism [37]. Simultaneously, it is also involved in the regulation of various stress responses. For instance, AtABI5 can improve transgenic cotton resistance to drought stress by regulating ROS scavenging [38]. Overexpression of AtABI5 increased the accumulation of DGAT1 and TAG proteins, and regulated the growth of Arabidopsis thaliana seedlings under stress [39]. ABI5 can improve the drought resistance of rice [40] and barley [41]. The salt tolerance of Siberian wildrye [42] and maize [43] was also regulated by ABI5. Studies in several species have shown that the response and regulation of ABI5 are mainly reflected in osmotic stress such as drought and salt, but there are few reports on the cold resistance mechanism of wheat.
In this study, 227 members of the TabZIP gene family were identified and divided into nine groups. TaABI5, as a representative of group A, showed excellent performance under various stresses. Therefore, we cloned the TaABI5 gene, studied its role in cold resistance through ectopic expression in Arabidopsis and further explored the low-temperature resistance mechanism of TaABI5 by interacting with a cold-resistance transcription factor TaICE1 [44] protein. The resistance of transgenic lines and wild-type plants to abiotic stress was evaluated by several physiological parameters. This study will provide a theoretical reference for the study of bZIP genes in plant abiotic resistance.

2. Results

2.1. Identification of bZIP Gene Family in Wheat

After removing erroneous and redundant sequences, 227 TabZIP genes were identified for subsequent study. Nomenclature of these genes was given based on the chromosome position, and all genes were renamed from TabZIP1 to TabZIP227. The basic information of 227 TabZIP genes was analyzed in this study, including the protein sequence lengths, relative molecular weight (MW), and isoelectric point (pI). As shown in Table S1, the amino acid lengths of its coding protein are between 129 (TabZIP57) and 652 (TabZIP22), and the average size of amino acid sequences was 314 aa. Isoelectric points of those proteins ranged from 4.36 (TabZIP208) to 12.14 (TabZIP140). The MW range from 14,098.74 Da (TabZIP57) to 68,897.15 Da (TabZIP22), with an average value of 33,878.42. The subcellular location shows that all TabZIP genes are locate in the nucleus, which is related to the biological function of the gene family as transcription factors. We found that all of the TabZIP protein’s GRAVY is negative, indicating that these proteins are hydrophilic. Also, these are essential physical and chemical properties as transcription factors.

2.2. Phylogeny Analysis of TabZIP

In order to investigate the phylogenetic relationship of TabZIP proteins, the phylogenetic tree was constructed based on multiple sequence alignment of 227 wheat TabZIP members and their homologous proteins in Arabidopsis (Figure 1). Finally, the bZIP genes of wheat and Arabidopsis were divided into 13 groups according to the Arabidopsis classification system. The group J was the branch with the most members, containing 56 TabZIP genes. Intriguingly, it was found that only the bZIP protein of Arabidopsis was found in group D. The B and G subfamilies contain only wheat genes and no Arabidopsis genes.

2.3. Gene Structure and Motif Analysis of TabZIP

In order to better understand the functions and phylogenetic relationship of members of the TabZIP gene family, structural analysis was performed by comparing each bZIP gene in wheat to understand the phenomena of structural diversity. The results showed that 13.6% of TabZIP genes (31/227) had no introns in the structure, and these genes that did not contain introns were closely related to each other. The remaining TabZIP genes contained 1~14 introns (Figure 2).
In this study, a total of 10 high confidence motifs was predicted, and only one gene (TabZIP9) was not identified to contain any motifs. TabZIP genes were divided into different subfamilies (A, C, E, F, G, H, J, L, and M) according to the phylogenetic tree and divergence of motifs. Motif composition and distribution of TabZIP family members were highly regular, with the number of motifs in all TabZIP genes ranging from 1 to 5 (Figure 2). The motif characteristics from the same group were relatively similar, such as S subfamily. However, TabZIP34, TabZIP174, and TabZIP137 contain motif3 absent in the other genes in the same subgroup. Motif1 was distributed in all members of the TabZIP family, consistent with the result that the Motif1 sequence is a bZIP conserved domain sequence. And about 17.6% of TabZIP members (40/227) only contain motif1.

2.4. Chromosomal Localization and Synteny Analysis of TabZIP Genes

All the TabZIP genes were located on 21 chromosomes, except for TabZIP227 (Figure 3a). Each chromosome contains some bZIP genes, but they are unevenly distributed and vary widely. Chromosome 5A contained the largest number of TabZIP genes (17.6%), whereas chromosome 1A contained only nine members (3.9%). Chromosomes 5A, 5B, and 5D contain denser genes than others do in the same genome. Moreover, most of the genes were distributed near the ends of the chromosomes.
Besides, to understand the evolutionary relationship of the bZIP family in wheat, collinearity relationships were displayed by comparing wheat with four other species. These species include two dicotyledons (Arabidopsis and potato) and two monocotyledons (maize and rice). A total of 2, 5, 38, and 22 homologous pairs were identified in wheat and these species, respectively (Figure 3b).

2.5. Expansion and Evolutionary Analysis

Gene duplication plays a crucial role in genome evolution. In this study, 227 duplicated genes in TabZIP genes were identified, including whole genome duplication (WGD), proximal duplication (PD), tandem duplication (TD), and transposed duplication (TRD). There were 213 duplicated genes (223 pairs) in wheat genome (Figure 4). About 95% of the duplicated genes were identified as WGDs (211 gene pairs), only one PD gene pair (TabZIP84 and TabZIP86) was identified. The remaining duplicated genes were TDs (8 gene pairs) and TRDs (3 gene pairs). The non-synonymous mutation rate to synonymous mutation rate (Ka/Ks) is essential for exploring genomic evolution [33]. The results showed that the Ka/Ks ratios of all wheat bZIP genes are from 0 to 1.1387, and only 3 pairs of genes’ ratios were greater than 1 (Table S2). The earliest genetic differentiation appeared 76.7703 million years ago.

2.6. Cis-acting Element Analysis

Promoter cis-element analysis provides convenience for understanding gene function and spatial specific expression. There were many cis-elements in the upstream sequence of wheat TabZIP genes (Figure 5). In addition to cis-elements, such as CAT-box and CCGTCC-box, related to growth and development, there were also many photo-responsive elements, such as SP1, GT1-motif, and G-box and hormone-responsive elements, such as TGACG-motif, CGTCA-motif (methyl jasmonate-induced), ABRE (ABA response element), and as-1 (Auxin response element). However, the most common classes of cis-elements were those that respond to abiotic stresses, such as low-temperature response (LTR), involved in anaerobic inducible enhancer (GC-Motif), antioxidant response (ARE), dehydration stress (DRE), drought (MYB, and MYC), and osmotic stress response. Finally, it was not difficult to find that the high-frequency cis-acting elements in the promoter of TabZIP genes concentrated in stress-related elements, such as drought and osmotic stress.

2.7. Expression Analysis of TabZIP Gene and RNA-Seq under Abiotic Stress

The potential role of wheat bZIP family in different abiotic stresses was analyzed based on RNA-Seq data. The RNA-Seq data were analyzed under the drought and high temperature of about 40 °C. Transcriptional levels of TabZIP genes at 1 h and 6 h after treatment were revealed. These treatments included drought for 1 h (D-1) and 6 h (D-6), high temperature for 1 h (H-1) and 6 h (H-6) at 40 °C, and the combination of the two stresses for 1 h (CS-1) and 6 h (CS-6). The TabZIP genes had a high expression in each type or time for treatment (Figure 6). Among them, red represented high/up-regulated expression of the genes, low/down-regulated expression genes were shown in blue, and yellow represents no significant difference in gene expression. The genes in group Ⅰ were significantly up-regulated under drought treatment. Except for Ⅲ groups of genes, the genes with high expression under heat stress were lower in drought condition. The number of differentially expressed genes was similar to heat treatment. It was noteworthy that TabZIP179 was the most significantly up-regulated gene under drought treatment, and it was also significantly up-regulated in H-6, while the most significantly up-regulated gene in H-1 was TabZIP191. In CS-1, almost all genes had the low expression, and genes with high expression in CS-6 were mainly in group Ⅱ. In group Ⅲ, the highly expressed genes were densely distributed in all treatments except CS-1.
In general, gene expression level and pattern indicate gene function. To clarify the expression characteristic of wheat TabZIP gene family. We examined the relative expression levels of genes in each subgroup of the TabZIP under abiotic stresses (salt, drought, and cold) and hormones (ABA, JA, and SA), as shown in Figure 7. The expression levels of TabZIP96 and TabZIP22 were up-regulated under salt stress, and their expression trends were similar, which increased first and then decreased, and reached the peak at 24 h. Most genes respond to drought. The relative expression of TabZIP19 and TabZIP103 after PEG treatment was significantly increased by 10 and 15 times, respectively. The expression of the Only 3 genes (TabZIP96, TabZIP19, and TabZIP103) were significantly induced by low temperature. Besides, the hormone treatment results showed that ABA-induced all genes, and some responded rapidly to up-regulation in the early stage, and some responded violently in the late stage. Most genes were also involved in the response of JA, such as TabZIP96, TabZIP107, and TabZIP19. While the expression level changes of most genes after SA treatment were not significant.

2.8. Overexpression of TaABI5 in A. thaliana Strengthens Plant Freezing Tolerance

Due to the particular function of genes in subgroup A under abiotic stress, we cloned the entire length of gene TaABI5 (TabZIP96) in group A. In order to investigate the potential effect of TaABI5 on low-temperature tolerance in plants, we constructed overexpression vector 35S: TaABI5 and transformed it into wild-type A. thaliana plants (WT; col-0). Overexpression of the TaABI5 gene in A. thaliana plants resulted in significantly enhanced freezing stress resistance (Figure 8). There was no significant phenotype difference in all lines at 4 °C, but leaf curl and wilting occurred in all lines when the temperature dropped to −10 °C. It was easier to be observed that the mutant abi5 with functional loss showed a cold-sensitive phenotype at −10°C and that the OE lines were slightly stronger than the wild-type. After seven days of recovery growth, survival rates of the three transgenic lines were 43%, 45%, and 44%, respectively, significantly higher than that of WT plants (25%), as shown in Figure 8a. The malondialdehyde (MDA) contents of the three transgenic lines were 33, 31, and 33 nmol g−1 FW after treatment at −10°C (Figure 8b), respectively, which were lower than that of the wild-type lines (42 nmol g−1 FW). Electrolyte leakage rates, another indicator of membrane permeability, were 60%, 62%, and 61% in OE lines, which were lower than WT lines (78%), as shown in Figure 8b.
Transcriptional levels of low temperature response-related genes were detected by qRT-PCR. As shown in the Figure 8, transgenic plants overexpressing TaABI5 may up-regulate the expression of several key stress-responsive genes compared with the wild type under normal conditions. These stress-responsive genes include RD29A, COR47, COR15A, and KIN1 (Figure 8c). As expected, the expression of these genes was down-regulated in abi5 mutant plants. After 4 °C treatment for three days, these stress-responsive genes were significantly induced. Thereinto, AtRD29A gene was highly up-regulation. AtCOR15A was highly expressed under −10 °C stress. Furthermore, the expression of stress-responsive genes in transgenic plants was more than those of wild-type.

2.9. TaABI5 Regulates Low-Temperature Stress Response by Participating in ROS Metabolism

In order to study the effect of TaABI5 on plants under low temperature stress, four weeks old A. thaliana were subjected to low temperature stress. Cell membrane damage caused by excessive ROS is a fundamental cause of osmotic stress, such as cold stress [45]. We examined the levels of superoxide (O2−), hydrogen peroxide (H2O2), and antioxidant enzymes activities of abi5, transgenic plants and wild-type Arabidopsis under cold stress by 3,3′-diaminobenzidine staining (DAB) and Nitroblue tetrazolium staining (NBT). CAT and POD showed high activity in OE lines, which was consistent with DAB and NBT staining results. As shown in Figure 9, the staining of NBT and DAB in transgenic lines was weaker than that of WT and abi5.

2.10. Subcellular Localization of TaABI5 and Y2H Assay

The bZIP genes perform the molecular function as a transcription factor. It can activate the transcription of downstream genes. The transcription activation of TaABI5 has been detected. Full-length TaABI5 fused with GAL4 DNA-binding domain (BD-TaABI5) was transformed into yeast strains. As shown in Figure 10a, no yeast colony growth was observed in the negative control (BD-PGBKT7 empty vector) on the SD/-Trp/-His/-Ade medium, but BD-TaABI5 activated the expression of reporter genes with strong transcription activation activity.
In order to explore the regulatory mechanism of TaABI5 in plant response to cold stress, the full-length TaICE1 into the Gal4 activation domain of the prey vector (AD-TaICE1). BD-TaABI5 and AD-TaICE1 were co-transformed into yeast strain AH109. As shown in Figure 10b, only the yeast co-transferred by TaABI5 and TaICE1 could survive on QDO medium with 40 mM Aureobasidin A (AbA). The results demonstrate that TaABI5 physically associated with TaICE1.

2.11. Analysis of the Subcellular Localization of TaABI5

The subcellular localization of TaABI5 was explored to understand more functions. TaABI5: GFP was transformed into the leaf cells of N. benthamiana. As shown in Figure 11, only the nucleus of tobacco cells emitted green fluorescence in the confocal laser imaging, which indicated that TaABI5 fused with green fluorescence protein was located in the nucleus.

3. Discussion

As wheat is a very important economic crop, it is necessary to study the regulation mechanism of its frost resistance. The bZIP genes is a large transcription factor in plants and can participate in a variety of physiological and biochemical reactions [46]. The family also plays an important role in regulating plant stress tolerance [47]. This study focused on the identification and analysis of bZIP gene family members in wheat, and attempted to explore the effect of bZIP gene on cold tolerance of wheat.
In the previous report by Agarwal et al., 191 TabZIP genes were identified [48]. Nevertheless, 36 new genes have been added to the current version (Table S1). In this study, 227 TabZIP gene family members were identified and analyzed from the wheat genome database. The results were compared to those of previously identified species Arabidopsis thaliana (75 genes) [3], Solanum lycopersicum (69 genes) [49], B.distachyon (96 genes) [50], Oryza sativa (89 genes) [9], Zea mays (125 genes) [51], Glycine max (131 genes) [52], Coniothyrium minitans (34 genes) [53], Vitis vinifera (55 genes) [54], Malus domestica (116 genes) [55], Fragaria vesca (50 genes) [20], Sorghum bicolor (92 genes) [47], and Prunus persica (47 genes) [56]. The number of bZIP gene family members increased, but was relatively small compared with B. napus contain 247 bZIP genes [57]. It is not unusual for gene families to change in size due to gene duplication, deletion, pseudogenization, and functional diversification [58]. The bZIP gene of rice [9], apple [55], and grapevine [54] can be divided into the same 10 subgroups as that of Arabidopsis [46], which is basically consistent with the classification of wheat in our study. According to its evolutionary branch and gene motifs, this study divided wheat bZIP gene into nine groups for subsequent specific analysis (Figure 1). Different branches indicate functional differences or that these subfamilies have specific functions [59]. Group D contained only AtbZIP genes but no TabZIP genes, indicating that the TabZIP genes in these groups might have been lost or differentiated into other groups during the evolution of wheat [60]. Groups B and G contain only the TabZIP gene, suggesting that this clade may be specific to wheat [61].
Gene duplication can provide the evolutionary potential for new functions in species [62]. This study revealed the gene duplication patterns of the bZIP gene family in wheat and visualized their chromosomal locations, as shown in Figure 3. The TabZIP was a large gene family distributed on each chromosome (Figure 3a). The TabZIP genes in wheat is closely related to monocotyledonous crops, such as potato and rice. Most plants have experienced a genome wide duplication event or a segmental duplication event [63], a large-scale chromosome doubling event that results in the retention of a large number of chromosomes doubling segments in the genome [64]. According to predecessors, it is speculated that hexaploid wheat is a cross between tetraploid wheat and Aegilops [65]. Therefore, hexaploid wheat has a large number of the oldest and most conserved fragment replication events [66]. Segmental genome duplications or whole genome duplication events may explain the expansion of the bZIP gene family [57]. This study identified 211 gene pairs with WGD, accounting for about 95% of the total gene pairs. Wheat also contained eight gene pairs with TD (Figure 4). The eight gene pairs (14 genes) formed a gene cluster, in which genes had similar sequences and similar functions [67]. Notably, only one pair of genes had PD events. This means that these two genes (TabZIP84 and TabZIP86) may have greater adaptive ability under stronger selection pressure [68]. Ka/Ks ratio can determine whether there is selection pressure on the protein-coding gene, and the ratio is essential for exploring genomic evolution [69]. Most genes were purifying selection (Ka/Ks < 1) and a few were positively selected (Ka/Ks > 1) [69]. The study results indicate that 98.6% of gene pairs had a Ka/Ks ratio less than 1, indicating the evolution of the wheat TabZIP family was relatively conserved [70]. This result also indicated that there was no significant functional division after fragment duplication and tandem duplication, suggesting that purifying selection played an important role in the evolution of the wheat genome. However, there were five genes (three gene pairs) with Ka/Ks ratios greater than 1, which were supported by Darwinian selection and led to adaptive evolution [71]. And these genes were strongly positively selected, which have recently evolved rapidly and were of great importance to the evolution of species [72]. Overall, the TabZIP gene family members in wheat genomes were relatively evolutionarily conserved. This provided a theoretical basis for further exploring the function and properties of TabZIP genes.
In order to explore the specific expression pattern of TabZIP gene in wheat, transcriptome analysis found that the expression of the entire gene family responds to abiotic stresses such as drought and heat to varying degrees (Figure 6). The expression levels of most TabZIP genes were significantly up-regulated after drought and heat combined stress. At the same time, their promoter cis-acting elements were analyzed and found to contain a variety of components in response to abiotic stress, including low temperature (Figure 5). These results suggest that TabZIP gene has a high probability of responding to environmental stress and may be involved in the regulation of plant response to abiotic stress [73]. In order to confirm the reliability of the above analysis and verify the above speculation, we applied three kinds of abiotic stress and three kinds of hormone treatment to indoor wheat seedlings at the three-leaf stage [74]. The qRT-PCR data and promoter cis-acting elements analysis showed that the TabZIP genes were up-regulated after ABA treatment. We also found that cis-acting elements associated with ABA response were highly enriched in promoters of a large number of genes (Figure 5). This result is consistent with the biological identity of bZIP as a transcription factor [75]. The bZIP genes are involved in ABA signaling and ABA plays an important role in the process of plant resistance to stress [76]. Therefore, we speculate that the bZIP gene also has a prominent contribution in regulating plant stress resistance. Cis-acting element results also found many elements related to abiotic stress response, such as LTR, ARE, DRE, MYC, and MYB. The results showed that there were a large number of photoinduction-related elements in the TabZIP gene promoter, which may be closely related to light response and light sensitivity. The functions of light response need to be further explored.
Subsequently, we examined the expression levels of TabZIP genes (one gene from each subfamily was randomly selected) under different abiotic stress treatments. TabZIP gene family showed different expression patterns in response to salt, drought, and low temperature (Figure 7). The expression of TabZIP107 and TabZIP30 increased significantly after 48 h of stress treatment, suggesting that these two genes may play a role in the late stage of salt stress. In drought treatment, all genes were up-regulated in different amplitude, and the gene (TabZIP19) in group G was up-regulated 17 times rapidly after drought stress. The above findings deserve further confirmation and exploration. Most genes were insensitive to cold stress. However, the gene TabZIP96 with differential expression appeared in group A, and its expression was significantly up-regulated by more than 20 times of low temperature stress. We predict that this gene may play a role in the later stage of plants responding to low temperature stress, providing a good adaptive basis for plants to tolerate freezing stress conditions at lower temperatures [77]. Simultaneously, to further explore the relationship between TabZIP genes and hormone regulation, the expression pattern of TabZIP gene family was analyzed under different hormone treatments (Figure 7). All genes were induced under ABA, and the expression trend was similar to that under JA treatment. These results suggest that TabZIP genes may play a pivotal role in these two hormone pathways, and JA responsive genes are also regulated by bZIP transcription factors [78]. This result is consistent with the existence of a large number of JA related cis-acting elements, such as TGACG-motif and CGTCA-motif in Figure 5. ABI5 encodes a bZIP transcription factor that physically interacts with JAZ1 in the JA pathway [79]. This confirmed that bZIP gene is involved in the regulation of crosstalk of two hormone signals [80,81].
In order to deeply explore the effect of bZIP gene on cold resistance of wheat, we screened a gene involved in low temperature response through real-time fluorescence quantitative results and verified the biological function of the gene (Figure 7). TabZIP96 (TaABI5), a group representative member of the TabZIP gene family, was cloned as a homologous gene of AtABI5 in wheat [82]. The gene response to various stress conditions and showed obvious transcriptional characteristics. Phenotypic analysis and related physiological indicators of cold resistance showed that overexpression of TaABI5 could enhance cold resistance of transgenic Arabidopsis thaliana (Figure 8).
The cold tolerance of plants was negatively correlated with the degree of cell damage. The stronger the plant’s tolerance, the lower the degree of cell damage. The root cause of cell destruction is the accumulation of ROS [83,84]. High levels of O2− and H2O2 can lead to accumulation of ROS [85]. DAB and NBT staining can indicate O2− and H2O2 levels [86]. And the higher the level of cell damage, the deeper DAB and NBT staining [87]. Both CAT and POD activities were stronger than those of wild-type and mutant plants, indicating a strong ROS scavenging ability [88] (Figure 9d). Staining depth of each line of transgenic plants were lower than that of the WT, and the mutants with functional loss were the darkest in Figure 9. Which means that plants have a strong ability to scavenge reactive oxygen species to resist oxidation [89]. The low temperature destroys the structure of the cell membrane [90,91]. Freezing damage to plant cells is more serious, so that intracellular regionalization is highly damaged [92]. Electron leakage rate can reflect the disorder degree of ion transfer reaction of cells, and the side exposure of membrane permeability [87,93]. Otherwise, it was demonstrated that overexpression of TaABI5 enhanced the expression of downstream cold-responsive genes in Arabidopsis thaliana, especially under low temperature and freezing conditions (Figure 8c). These results once again proved that functional TaABI5 could enhance the frost resistance of plants.
By observing the phenotype of each line under freezing stress, the growth state and survival rate of OE line were better than other lines (Figure 9c). To further explore how TaABI5 is involved in the cold resistance pathway, we found that TaABI5 physically interacts with TaICE1 in the yeast two-hybrid system (Figure 10b). Subcellular localization experiments confirmed that TaABI5 was located in the nucleus, which was in line with the biological role of transcription factor (Figure 11). This suggests that these two proteins recognize each other in the nucleus and bind to form protein complexes that are involved in each other’s regulatory pathways, although their respective functions are different. We speculated that TaABI5 and TaICE1 jointly regulate the tolerance of wheat to low-temperature stress. Based on these results, we concluded that the effect of TaABI5 on plant cold stress tolerance may be realized through the interaction with TaICE1. Analogously, research has shown that MdABI4 interacts with MdICE1 protein modules jointly regulating growth of plants [94]. ABI5 in Arabidopsis demonstrates physical interaction with MED25 [95]. MED25 is involved in different regulatory pathways as an intermediary subunit, its functions are complex [96]. Therefore, the functions of ABI5 in other aspects and its interaction network need to be further explored and studied. Studies have shown that ABI5 binds directly to the promoter regions of EM1 and EM6, promoting their transcription [97]. EM1 and EM6 are late embryogenesis abundant proteins (LEA), which belong to a class of cold-responsive protein [93]. Furthermore, ABI5 can regulate the expression of RAB18, COR15A and COR6.6 [98]. RAB18 has a certain sense in the low temperature resistance pathway [99]. COR15A and COR6.6 are well-known cold-response genes [100]. The cold resistance of TaABI5 may also directly affect the expression of these cold tolerance genes. The cold-resistant regulatory network of TaABI5 in wheat needs to be further confirmed and supplemented. In conclusion., it can be indicated that ABI5 had an important position in plant under cold stress, and TaABI5 also plays a certain function in cold response pathway of wheat.

4. Materials and Methods

4.1. Identification of the Members of TabZIP Gene Family in Wheat

The Hidden Markov model of bZIP domain (PF00170) was downloaded by PFAM (protein domain family) database (http://pfam.xfam.org, accessed on 11 October 2021), and the bZIP genome sequence information of wheat genome 2.0 was downloaded by IWGSC (http://www.wheatgenome.org/, accessed on 11 October 2021). To identify TabZIP genes in wheat, Hmmsearch tool was used to retrieve a domain similar to the bZIP domain in wheat, and the value was set to 1×10−5. These protein sequences were submitted to Pfam and CDD (Conserved Domains Database, https://www.ncbi.nlm.nih.gov/Structure/cdd/cdd.shtmlis, accessed on 11 October 2021) to remove redundant sequences for further verification of the inclusion of the bZIP domain.

4.2. Sequence Analysis of TabZIP Genes in Wheat

All CDS sequences of TabZIP genes were submitted to ExPASy (https://www.expasy.org, accessed on 11 October 2021) for gene length, amino acid length, relative molecular weight, isoelectric point, hydrophilicity, stability, and other physicochemical properties analysis. The gff3 file was used to extract the intron-exon and UTR (non-coding region) distribution and other genes structure information, which was finally visualized by TBtools [101]. And the software MEME was used to analyze the conservative motifs of the proteins [102]. The maximum number of motifs to be found was 10, motif width was 6–20, and other parameters were default values. Promoters affect gene function to a certain extent. The 2000 bp sequence upstream of the TabZIP genes transcription start site was submitted to PlantCARE (http://bioinformatics.psb.ugent.be/webtools/plantcare/html, accessed on 11 October 2021) to analyze the promoters enrichment type of the gene family members.

4.3. Phylogenetic Analysis and Chromosomal Location

MUSCLE in MEGA Version 7 software was used to align the full-length sequence of TabZIP gene family members, and 1000 was used as the bootstrapping to build the neighbor joining tree [103]. The locations and chromosome lengths of all TabZIP genes were obtained from the Ensembl Plants database and the data were located using TBtools. In addition, to study the evolutionary relationship of bZIP genes between wheat and Arabidopsis, the bZIP gene sequences of Arabidopsis and wheat were used to map evolutionary trees. The genome and protein sequences of them were downloaded from the Ensembl Plants database using iTOL (https://itol.embl.de, accessed on 12 October 2021) for beautification.

4.4. Duplication, Ka/Ks, and Synteny Analysis

Blastall tool was used to blast the CDS sequence of 227 genes of TabZIP, and the expectation value is 1 × 10−20 [104]. MCScanX was used to analyze gene replication and synteny events. The proteins sequence of Arabidopsis (A. thaliana), potato (S. tuberosum), maize (Z. mays), and rice (O. sativa) were downloaded from Ensembl Plants. The evolutionary relationship between genes is performed by synonymous substitution rate (Ks) and nonsynonymous substitution rate (Ka). Ka/Ks ratio in wheat was calculated by KaKs_calculator [105]. All duplicated gene pairs of TabZIP genes in wheat were classified into the following four categories: segmental, tandem, proximal, and dispersed duplications. Finally, collinear regions and names of TabZIP gene family were visualized by Circos software. The divergence time of gene duplication (MYA) was calculated by the formula T = Ks/2λ with a value of =6.5 × 10−9 synonymous substitutions per site per year [106,107].

4.5. Expression Analysis Based on RNA-Seq

NCBI (National Center for Biotechnology Information) database was used to obtain the transcriptome data (accession number: SRP045409), and transcriptome enrichment analysis was performed on members of the TabZIP family. Control (SRR1542405, SRR1542404), drought treatment for 1 h (SRR1542407, SRR1542406), 6 h (SRR1542409, SRR1542408), heat treatment for 1 h (SRR1542411, SRR1542410), 6 h (SRR1542411, SRR1542410), drought and heat combined treatment for 1 h (SRR1542415, SRR1542414), and 6 h (SRR1542417, SRR1542416). HISAT software was used to map the reads from all samples to the wheat genome [108].

4.6. Plant material and Treatments

Winter wheat (Triticum aestivum) cultivar Dongnongdongmai1 (Dn1) was obtained from the Wheat Breeding Institute of Northeast Agricultural University, Harbin, China. 100 uM ABA (Abscisic-acid), 100 uM MeJA (Methyl Jasmonate), 100 uM SA (Salicylic acid) [109,110], and Cold (4 °C), 200 mM NaCl solution and 20% PEG solution were used to treat wheat seedlings at the tri-leaf stage for 48 h [111]. The seedlings grew in a greenhouse with 22 h of light and 16 h of dark. Total RNA was extracted from the leaves of unstressed and stressed wheat seedlings at 0, 6, 12, 24, and 48 h.
Arabidopsis was selected as the Columbian type. The abi5 (SALK_013163, Col-0 background) mutant was a kind gift from Dr. Gang Wu (State Key Laboratory of Subtropical Silviculture, School of Agriculture and Food Sciences, Zhejiang Agriculture and Forestry University) [112]. The culturing condition was 24 C, 16/8 photoperiod [113]. Four-week-old Arabidopsis thaliana was treated at 4 °C for 3 days, and then the plants were subjected to −10 °C for 2 h. Finally, phenotypic analysis was performed 7 days after resuming culture conditions [44].

4.7. Total RNA Extraction and qRT-PCR

Total RNA was extracted with Trizol [114], reverse transcribed to cDNA with the TransScript One-Step gDNA Removal and cDNA Synthesis SuperMix Kit (Applied Biosystems, Shanghai, China). One gene from each subfamily was selected for qRT-PCR under different treatments to analyze the expression characteristics of TabZIP gene family members. As shown in Table S3, the primers were designed on NCBI. The ChamQ™ Universal SYBR qPCR Master MIX (Vazyme, Nanjing, China) was used. Three biological replicates were set for each sample. It was used to analyze the expression characteristics of TabZIP genes under different stresses. The relative expression level was calculated using the 2−∆∆ct method.

4.8. Cloning of TaABI5 and Plant Transformation

The full-length CDS sequence of TaABI5 was cloned and inserted into a plant expression vector (Pcambia230035Su) by user method so that the expression of TaABI5 was driven by CaMV35S promoter, and the recombinant plasmid of plant overexpression vector was transformed into Agrobacterium tumefaciens strain EHA105. The 35S:TaABI5 constructs was transformed into Arabidopsis wild-type (Col-0) plants. Finally, 1/2MS medium containing kanamycin resistance was used to screen T3 generation positive plants for subsequent experiments.

4.9. Determination of Physiological Indices and ROS Metabolism Assays

Malondialdehyde (MDA) content in Arabidopsis leaves was detected using the method previously reported [115]. Detection of catalase (CAT) activity according to the method of Aebi [116], Peroxidase (POD) activity was measured by the method of Chance [117]. In addition, plant resistance to low temperature stress was related to the accumulation of reactive oxygen species (ROS) [1]. 3,3′-diaminobenzidine (DAB) and nitro blue tetrazolium (NBT) tissue staining can understand ROS accumulation by visualizing H2O2 and O2− content, specific methods and steps have been described by predecessors [118].

4.10. Statistical Analysis

Graphpad Prizm 7.0 statistical software was used for data analysis, ANOVA was used to test significance, * p < 0.05, ** p < 0.01, *** p < 0.001, and **** p < 0.0001. All experiments were repeated three times. The values presented the means ± one standard deviation (SD) of three biological replicates.

4.11. Two-Hybrid Assay and Subcellular Localization

The CDS sequences of TaABI5 and TaICE1 were cloned and full-length cDNA were respectively inserted into bait vector pGBDT7 and prey vectors pGADT7 for yeast two-hybrid experiments. As both of them are transcription factors with transcriptional activation activities, transcriptional activation experiments are indispensable. The recombinant pGBDT7:TaABI5 was transformed into strain AH109 (Takara, Shiga, Japan).
In the yeast two-hybrid experiment, recombinant plasmids pGADT7:TaICE1 and pGBDT7:TaABI5 were co-transformed into yeast strain AH109 and cultured on SD/-Trp-Leu medium for 60–80 h, then transferred to SD/-Trp-His-Leu-Ade medium with AbA (400 ng/mL) (Takara, Shiga, Japan) for 80–90 h.
The full-length TaABI5 sequence was inserted into the green fluorescent protein (GFP) vector. The pTaABI5:GFP vector was constructed using the above methods. Nicotiana benthamiana epidermal cells were infected by Agrobacterium-mediated. The transient expression of TaABI5 was detected using laser confocal microscopy.

5. Conclusions

A total of 227 TabZIP genes were identified in wheat, adding 36 new genes from previous studies. The TabZIP gene family was divided into nine branches according to the gene structure and motif. The promoters of these genes contained many cis-acting elements related to abiotic stress and showed different expression patterns under different abiotic stress conditions. We demonstrated that TaABI5 increased the expression of downstream stress-response genes, and that overexpression of TaABI5 enhanced the resistance to low temperature and freezing stress in transgenic plants. For another, TaABI5 located in the nucleus and physically interacted with TaICE1. These results indicated that TabZIPs had certain significance in wheat under low temperature stress, and laid a good foundation for further exploring the regulation mechanism of wheat stress resistance.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms23042351/s1.

Author Contributions

Y.L. collect samples, complete experiments, data analysis and write papers. J.X., Y.J. and H.C. help to collect samples and assist experiments. Y.B., D.W., D.Z. and J.Y. participate in revising articles. J.C. design ideas and provide project support. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (31971831), National Natural Science Foundation of China (31701348), and Heilongjiang Provincial Natural Science Foundation (LH2021C022).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

We are grateful to the Laboratory of Plant Physiology and Molecular Biology, the College of Life Sciences, and the Northeast Agricultural University for the experimental materials and technical assistance provided. Thanks to Wu Gang’s team from the State Key Laboratory of Subtropical Silviculture, School of Agriculture and Food Sciences, Zhejiang Agriculture and Forestry University, for providing experimental materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ruelland, E.; Vaultier, M.N.; Zachowski, A.; Hurry, V. Cold Signalling and Cold Acclimation in Plants; Chapter 2; Elsevier Science and Technology: Amsterdam, The Netherlands, 2009. [Google Scholar]
  2. Jan, N.; Andrabi, K.I. Cold resistance in plants: A mystery unresolved. Electron. J. Biotechnol. 2009, 12, 14–15. [Google Scholar] [CrossRef]
  3. Jakoby, M.; Weisshaar, B.; Droge-Laser, W.; Vicente-Carbajosa, J.; Tiedemann, J.; Kroj, T.; Parcy, F. bZIP transcription factors in Arabidopsis. Trends Plant Sci. 2002, 7, 106–111. [Google Scholar] [CrossRef]
  4. Riechmann, J.L.; Heard, J.; Martin, G.; Reuber, L.; Jiang, C.; Keddie, J.; Adam, L.; Pineda, O.; Ratcliffe, O.J.; Samaha, R.R.; et al. Arabidopsis transcription factors: Genome-wide comparative analysis among eukaryotes. Science 2000, 290, 2105–2110. [Google Scholar] [CrossRef] [PubMed]
  5. Wray, G.A.; Hahn, M.W.; Abouheif, E.; Balhoff, J.P.; Pizer, M.; Rockman, M.V.; Romano, L.A. The evolution of transcriptional regulation in eukaryotes. Mol. Biol. Evol. 2003, 20, 1377–1419. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Ahmad, M.; Yan, X.; Li, J.; Yang, Q.; Jamil, W.; Teng, Y.; Bai, S. Genome wide identification and predicted functional analyses of NAC transcription factors in Asian pears. BMC Plant Biol. 2018, 18, 214. [Google Scholar] [CrossRef] [Green Version]
  7. Xiang, Y.; Tang, N.; Du, H.; Ye, H.; Xiong, L. Characterization of OsbZIP23 as a key player of the basic leucine zipper transcription factor family for conferring abscisic acid sensitivity and salinity and drought tolerance in rice. Plant Physiol. 2008, 148, 1938–1952. [Google Scholar] [CrossRef] [Green Version]
  8. Landschulz, W.H.; Johnson, P.F.; McKnight, S.L. The leucine zipper: A hypothetical structure common to a new class of DNA binding proteins. Science 1988, 240, 1759–1764. [Google Scholar] [CrossRef] [Green Version]
  9. Nijhawan, A.; Jain, M.; Tyagi, A.K.; Khurana, J.P. Genomic survey and gene expression analysis of the basic leucine zipper transcription factor family in rice. Plant Physiol. 2008, 146, 333–350. [Google Scholar] [CrossRef] [Green Version]
  10. Izawa, T.; Foster, R.; Chua, N.H. Plant bZIP protein DNA binding specificity. J. Mol. Biol. 1993, 230, 1131–1144. [Google Scholar] [CrossRef]
  11. Meshi, T.; Moda, I.; Minami, M.; Okanami, M.; Iwabuchi, M. Conserved Ser residues in the basic region of the bZIP-type transcription factor HBP-1a(17): Importance in DNA binding and possible targets for phosphorylation. Plant Mol. Biol. 1998, 36, 125–136. [Google Scholar] [CrossRef]
  12. Oyama, T.; Shimura, Y.; Okada, K. The Arabidopsis HY5 gene encodes a bZIP protein that regulates stimulus-induced development of root and hypocotyl. Genes Dev. 1997, 11, 2983–2995. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Luo, Q.J.; Mittal, A.; Jia, F.; Rock, C.D. An autoregulatory feedback loop involving PAP1 and TAS4 in response to sugars in Arabidopsis. Plant Mol. Biol. 2012, 80, 117–129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Kim, S.; Kang, J.Y.; Cho, D.I.; Park, J.H.; Kim, S.Y. ABF2, an ABRE-binding bZIP factor, is an essential component of glucose signaling and its overexpression affects multiple stress tolerance. Plant J. 2004, 40, 75–87. [Google Scholar] [CrossRef] [PubMed]
  15. Pan, Y.; Hu, X.; Li, C.; Xu, X.; Su, C.; Li, J.; Song, H.; Zhang, X. SlbZIP38, a Tomato bZIP Family Gene Downregulated by Abscisic Acid, Is a Negative Regulator of Drought and Salt Stress Tolerance. Genes 2017, 8, 402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Yang, S.; Xu, K.; Chen, S.; Li, T.; Xia, H.; Chen, L.; Liu, H.; Luo, L. A stress-responsive bZIP transcription factor OsbZIP62 improves drought and oxidative tolerance in rice. BMC Plant Biol. 2019, 19, 260. [Google Scholar] [CrossRef] [Green Version]
  17. Yang, Y.; Yu, T.F.; Ma, J.; Chen, J.; Zhou, Y.B.; Chen, M.; Ma, Y.Z.; Wei, W.L.; Xu, Z.S. The Soybean bZIP Transcription Factor Gene GmbZIP2 Confers Drought and Salt Resistances in Transgenic Plants. Int. J. Mol. Sci. 2020, 21, 670. [Google Scholar] [CrossRef] [Green Version]
  18. Li, H.; Chen, J.; Zhao, Q.; Han, Y.; Zhang, M. Basic Leucine zipper (bZIP) transcription factor genes and their response to drought stress in ginseng, Panax ginseng C.A. Meyer. BMC Genom. 2020, 22, 316. [Google Scholar] [CrossRef]
  19. Ma, M.; Chen, Q.; Dong, H.; Zhang, S.; Huang, X. Genome-wide identification and expression analysis of the bZIP transcription factors, and functional analysis in response to drought and cold stresses in pear (Pyrus breschneideri). BMC Plant Biol. 2021, 21, 583. [Google Scholar] [CrossRef]
  20. Wang, X.L.; Chen, X.; Yang, T.B.; Cheng, Q.; Cheng, Z.M. Genome-Wide Identification of bZIP Family Genes Involved in Drought and Heat Stresses in Strawberry (Fragaria vesca). Int. J. Genom. 2017, 2017, 3981031. [Google Scholar]
  21. Liang, S.M.; Chen, S.C.; Liu, Z.L.; Shan, W.; Kuang, J.F. MabZIP74 interacts with MaMAPK11-3 to regulate the transcription of MaACO1/4 during banana fruit ripening. Postharvest Biol. Technol. 2020, 169, 111293. [Google Scholar] [CrossRef]
  22. Liu, D.; Shi, S.; Hao, Z.; Xiong, W.; Luo, M. OsbZIP81, A Homologue of Arabidopsis VIP1, May Positively Regulate JA Levels by Directly Targetting the Genes in JA Signaling and Metabolism Pathway in Rice. Int. J. Mol. Sci. 2019, 20, 2360. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Heinekamp, T.; Strathmann, A.; Kuhlmann, M.; Froissard, M.; Muller, A.; Perrot-Rechenmann, C.; Droge-Laser, W. The tobacco bZIP transcription factor BZI-1 binds the GH3 promoter in vivo and modulates auxin-induced transcription. Plant J. 2004, 38, 298–309. [Google Scholar] [CrossRef] [PubMed]
  24. Yu, Y.H.; Jiao, Z.L.; Bian, L.; Wan, Y.T.; Yu, K.K.; Zhang, G.H.; Guo, D.L. Overexpression of Vitis vinifera VvbZIP60 enhances Arabidopsis resistance to powdery mildew via the salicylic acid signaling pathway. Sci. Hortic. 2019, 256, 108640. [Google Scholar] [CrossRef]
  25. Xiaoli, G. Isolation and Function Analysis of Heat Stress Related Transcription Factor Gene TabZIP28 in Wheat (Triticum aestivum). J. Agric. Biotechnol. 2016, 24, 157–167. [Google Scholar]
  26. Wang, F.; Lin, R.; Li, Y.; Wang, P.; Feng, J.; Chen, W.; Xu, S. TabZIP74 Acts as a Positive Regulator in Wheat Stripe Rust Resistance and Involves Root Development by mRNA Splicing. Front. Plant Sci. 2019, 10, 1551. [Google Scholar] [CrossRef]
  27. Guo, G.; Yang, Y.; Cao, L.; Liu, W.; Caili, B.I. RF2 Basic Leucine Zipper Transcription Factor TabZIP3 Involved in Salt Stress Response in Wheat. J. Agric. Sci. Technol. 2019, 21, 20. [Google Scholar]
  28. Cao, X.Y.; Chen, M.; Zhao-Shi, X.U.; Chen, Y.F.; Lian-Cheng, L.I.; Yue-Hua, Y.U.; Liu, Y.N.; You-Zhi, M.A. Isolation and Functional Analysis of the bZIP Transcription Factor Gene TaABP1 from a Chinese Wheat Landrace. J. Integr. Agric. 2012, 11, 12. [Google Scholar] [CrossRef]
  29. Bi, C.; Yu, Y.; Dong, C.; Yang, Y.; Zhai, Y.; Du, F.; Xia, C.; Ni, Z.; Kong, X.; Zhang, L. The bZIP transcription factor TabZIP15 improves salt stress tolerance in wheat. Plant Biotechnol. J. 2021, 19, 209–211. [Google Scholar] [CrossRef]
  30. Wang, B.; Li, L.; Peng, D.; Liu, M.; Wei, A.; Li, X. TaFDL2-1A interacts with TabZIP8-7A protein to cope with drought stress via the abscisic acid signaling pathway. Plant Sci. 2021, 311, 111022. [Google Scholar] [CrossRef]
  31. Liu, M.; Wang, Z.; Xiao, H.M.; Yang, Y. Characterization of TaDREB1 in wheat genotypes with different seed germination under osmotic stress. Hereditas 2018, 155, 26. [Google Scholar] [CrossRef]
  32. Zhang, L.; Zhang, L.; Xia, C.; Gao, L.; Hao, C.; Zhao, G.; Jia, J.; Kong, X. A Novel Wheat C-bZIP Gene, TabZIP14-B, Participates in Salt and Freezing Tolerance in Transgenic Plants. Front. Plant Sci. 2017, 8, 710. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Verslues, P.E.; Zhu, J.K. Before and beyond ABA: Upstream sensing and internal signals that determine ABA accumulation and response under abiotic stress. Biochem. Soc. Trans. 2005, 33, 375–379. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Novikova, G.V.; Stepanchenko, N.S.; Moshkov, A. At the beginning of the route: ABA perception and signal transduction in plants. Russ. J. Plant Physol. 2009, 56, 727–741. [Google Scholar] [CrossRef]
  35. Brocard, I.M.; Lynch, T.J.; Finkelstein, R.R. Regulation and role of the Arabidopsis abscisic acid-insensitive 5 gene in abscisic acid, sugar, and stress response. Plant Physiol. 2002, 129, 1533–1543. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Lin, J.H.; Liu, Z.G.; Zhang, N.; Yan, R.; Zhang, N.; He, X.M.; Wang, W.T.; Liu, D.; Wu, J. Stress response of ABI5 to BR stress and its regulation on hypocotyls growth in Arabidopsis thaliana. Yi Chuan 2021, 43, 901–909. [Google Scholar]
  37. Su, M.; Huang, G.; Zhang, Q.; Wang, X.; Li, C.; Tao, Y.; Zhang, S.; Lai, J.; Yang, C.; Wang, Y. The LEA protein, ABR, is regulated by ABI5 and involved in dark-induced leaf senescence in Arabidopsis thaliana. Plant Sci. 2016, 247, 93–103. [Google Scholar] [CrossRef]
  38. Mittal, A.; Gampala, S.S.; Ritchie, G.L.; Payton, P.; Burke, J.J.; Rock, C.D. Related to ABA-Insensitive3(ABI3)/Viviparous1 and AtABI5 transcription factor coexpression in cotton enhances drought stress adaptation. Plant Biotechnol. J. 2014, 12, 578–589. [Google Scholar] [CrossRef] [Green Version]
  39. Kong, Y.; Chen, S.; Yang, Y.; An, C. ABA-insensitive (ABI) 4 and ABI5 synergistically regulate DGAT1 expression in Arabidopsis seedlings under stress. FEBS Lett. 2013, 587, 3076–3082. [Google Scholar] [CrossRef] [Green Version]
  40. Zou, M.; Guan, Y.; Ren, H.; Zhang, F.; Chen, F. A bZIP transcription factor, OsABI5, is involved in rice fertility and stress tolerance. Plant Mol. Biol. 2008, 66, 675–683. [Google Scholar] [CrossRef]
  41. Collin, A.; Daszkowska-Golec, A.; Kurowska, M.; Szarejko, I. Barley ABI5 (Abscisic Acid INSENSITIVE 5) Is Involved in Abscisic Acid-Dependent Drought Response. Front. Plant Sci. 2020, 11, 1138. [Google Scholar] [CrossRef]
  42. De, Y.; Shi, F.; Gao, F.; Mu, H.; Yan, W. Siberian Wildrye (Elymus sibiricus L.) Abscisic Acid-Insensitive 5 Gene Is Involved in Abscisic Acid-Dependent Salt Response. Plants 2021, 10, 1351. [Google Scholar] [CrossRef] [PubMed]
  43. Yan, F.; Deng, W.; Wang, X.; Yang, C.; Li, Z. Maize (Zea mays L.) homologue of ABA-insensitive (ABI) 5 gene plays a negative regulatory role in abiotic stresses response. Plant Growth Regul. 2012, 68, 383–393. [Google Scholar] [CrossRef]
  44. Peng, K.; Tian, Y.; Sun, X.; Song, C.; Ren, Z.; Bao, Y.; Xing, J.; Li, Y.; Xu, Q.; Yu, J. tae-miR399- UBC24 Module Enhances Freezing Tolerance in Winter Wheat via a CBF Signaling Pathway. J. Agric. Food Chem. 2021, 69, 13398–13415. [Google Scholar] [CrossRef] [PubMed]
  45. Morell, S.; Follmann, H.; De Tullio, M.; Haberlein, I. Dehydroascorbate and dehydroascorbate reductase are phantom indicators of oxidative stress in plants. FEBS Lett. 1997, 414, 567–570. [Google Scholar] [CrossRef] [Green Version]
  46. Droge-Laser, W.; Snoek, B.L.; Snel, B.; Weiste, C. The Arabidopsis bZIP transcription factor family-an update. Curr. Opin. Plant Biol. 2018, 45, 36–49. [Google Scholar] [CrossRef]
  47. Wang, J.; Zhou, J.; Zhang, B.; Vanitha, J.; Ramachandran, S.; Jiang, S.Y. Genome-wide expansion and expression divergence of the basic leucine zipper transcription factors in higher plants with an emphasis on sorghum. J. Integr. Plant Biol. 2011, 53, 212–231. [Google Scholar] [CrossRef]
  48. Agarwal, P.; Baranwal, V.K.; Khurana, P. Genome-wide Analysis of bZIP Transcription Factors in wheat and Functional Characterization of a TabZIP under Abiotic Stress. Sci. Rep. 2019, 9, 4608. [Google Scholar] [CrossRef] [Green Version]
  49. Li, D.; Fu, F.; Zhang, H.; Song, F. Genome-wide systematic characterization of the bZIP transcriptional factor family in tomato (Solanum lycopersicum L.). BMC Genom. 2015, 16, 771. [Google Scholar] [CrossRef] [Green Version]
  50. Liu, X.; Chu, Z. Genome-wide evolutionary characterization and analysis of bZIP transcription factors and their expression profiles in response to multiple abiotic stresses in Brachypodium distachyon. BMC Genom. 2015, 16, 227. [Google Scholar] [CrossRef] [Green Version]
  51. Wei, K.; Chen, J.; Wang, Y.; Chen, Y.; Chen, S.; Lin, Y.; Pan, S.; Zhong, X.; Xie, D. Genome-wide analysis of bZIP-encoding genes in maize. DNA Res. 2012, 19, 463–476. [Google Scholar] [CrossRef] [Green Version]
  52. Liao, Y.; Zou, H.F.; Wei, W.; Hao, Y.J.; Tian, A.G.; Huang, J.; Liu, Y.F.; Zhang, J.S.; Chen, S.Y. Soybean GmbZIP44, GmbZIP62 and GmbZIP78 genes function as negative regulator of ABA signaling and confer salt and freezing tolerance in transgenic Arabidopsis. Planta 2008, 228, 225–240. [Google Scholar] [CrossRef] [PubMed]
  53. Xu, Y.; Wang, Y.; Zhao, H.; Wu, M.; Zhang, J.; Chen, W.; Li, G.; Yang, L. Genome-Wide Identification and Expression Analysis of the bZIP Transcription Factors in the Mycoparasite Coniothyrium minitans. Microorganisms 2020, 8, 1045. [Google Scholar] [CrossRef] [PubMed]
  54. Liu, J.; Chen, N.; Chen, F.; Cai, B.; Dal Santo, S.; Tornielli, G.B.; Pezzotti, M.; Cheng, Z.M. Genome-wide analysis and expression profile of the bZIP transcription factor gene family in grapevine (Vitis vinifera). BMC Genom. 2014, 15, 281. [Google Scholar] [CrossRef] [Green Version]
  55. Dong, M. Genome-wide identification and expression analysis of the bZIP gene family in apple (Malus domestica). Tree Genet. Genomes 2016, 12, 82. [Google Scholar]
  56. Wang, X.L.; Zhong, Y.; Cheng, Z.M.; Xiong, J.S. Divergence of the bZIP Gene Family in Strawberry, Peach, and Apple Suggests Multiple Modes of Gene Evolution after Duplication. Int. J. Genom. 2015, 2015, 536943. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Zhou, Y.; Xu, D.; Jia, L.; Huang, X.; Ma, G.; Wang, S.; Zhu, M.; Zhang, A.; Guan, M.; Lu, K.; et al. Genome-Wide Identification and Structural Analysis of bZIP Transcription Factor Genes in Brassica napus. Genes 2017, 8, 288. [Google Scholar] [CrossRef] [Green Version]
  58. Pan, Z.; Chen, L.; Wang, F.; Song, W.; Cao, A.; Xie, S.; Chen, X.; Jin, X.; Li, H. Genome-Wide Identification and Expression Analysis of the Ascorbate Oxidase Gene Family in Gossypium hirsutum Reveals the Critical Role of GhAO1A in Delaying Dark-Induced Leaf Senescence. Int. J. Mol. Sci. 2019, 20, 6167. [Google Scholar] [CrossRef] [Green Version]
  59. Nei, M. Phylogenetic analysis in molecular evolutionary genetics. Annu. Rev. Genet. 1996, 30, 371–403. [Google Scholar] [CrossRef]
  60. De Souza, S.J.; Long, M.; Gilbert, W. Introns and gene evolution. Genes Cells 1996, 1, 493–505. [Google Scholar] [CrossRef] [Green Version]
  61. Long, M.; VanKuren, N.W.; Chen, S.; Vibranovski, M.D. New gene evolution: Little did we know. Annu. Rev. Genet. 2013, 47, 307–333. [Google Scholar] [CrossRef] [Green Version]
  62. Sampedro, J.; Lee, Y.; Carey, R.E.; DePamphilis, C.; Cosgrove, D.J. Use of genomic history to improve phylogeny and understanding of births and deaths in a gene family. Plant J. 2005, 44, 409–419. [Google Scholar] [CrossRef] [PubMed]
  63. Crow, K.D.; Wagner, G.P. What is the role of genome duplication in the evolution of complexity and diversity? Mol. Biol. Evol. 2006, 23, 887–892. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Hurles, M. Gene duplication: The genomic trade in spare parts. PLoS Biol. 2004, 2, e206. [Google Scholar] [CrossRef] [PubMed]
  65. McFadden, E.S.; Sears, E.R. The origin of Triticum spelta and its free-threshing hexaploid relatives. J. Hered. 1946, 37, 107. [Google Scholar] [CrossRef]
  66. Paux, E.; Roger, D.; Badaeva, E.; Gay, G.; Bernard, M.; Sourdille, P.; Feuillet, C. Characterizing the composition and evolution of homoeologous genomes in hexaploid wheat through BAC-end sequencing on chromosome 3B. Plant J. 2006, 48, 463–474. [Google Scholar] [CrossRef] [PubMed]
  67. Ye, Y.N.; Hua, Z.G.; Huang, J.; Rao, N.; Guo, F.B. CEG: A database of essential gene clusters. BMC Genom. 2013, 14, 769. [Google Scholar] [CrossRef] [Green Version]
  68. Pace, J.N.; Sen, S.K.; Batzer, M.A.; Feschotte, C. Repair-mediated duplication by capture of proximal chromosomal DNA has shaped vertebrate genome evolution. PLoS Genet. 2009, 5, e1000469. [Google Scholar] [CrossRef] [Green Version]
  69. Hurst, L.D. The Ka/Ks ratio: Diagnosing the form of sequence evolution. Trends Genet. 2002, 18, 486. [Google Scholar] [CrossRef]
  70. Li, J.; Zhang, Z.; Vang, S.; Yu, J.; Wong, G.K.; Wang, J. Correlation between Ka/Ks and Ks is related to substitution model and evolutionary lineage. J. Mol. Evol. 2009, 68, 414–423. [Google Scholar] [CrossRef]
  71. Xing, Y.; Lee, C. Can RNA selection pressure distort the measurement of Ka/Ks? Gene 2006, 370, 1–5. [Google Scholar] [CrossRef]
  72. Parmley, J.L.; Hurst, L.D. How common are intragene windows with KA > KS owing to purifying selection on synonymous mutations? J. Mol. Evol. 2007, 64, 646–655. [Google Scholar] [CrossRef] [PubMed]
  73. Lashgari, G.; Kantar, R.S.; Tannous, B.A. Secreted Reporters for Monitoring Multiple Promoter Function. Methods Mol. Biol. 2017, 1651, 33–47. [Google Scholar] [PubMed]
  74. Rascio, A.; Fiorillo, F. African Journal of Plant Science Indoor characterization of three durum wheat genotypes exposed to drought and heat stress during early vegetative growth stages. Afr. J. Plant Sci. 2020, 14, 436–442. [Google Scholar]
  75. Lu, F.; Lionnet, T. Transcription Factor Dynamics. Cold Spring Harb. Perspect. Biol. 2021, 13, a040949. [Google Scholar] [CrossRef] [PubMed]
  76. Umezawa, T.; Nakashima, K.; Miyakawa, T.; Kuromori, T.; Tanokura, M.; Shinozaki, K.; Yamaguchi-Shinozaki, K. Molecular basis of the core regulatory network in ABA responses: Sensing, signaling and transport. Plant Cell Physiol. 2010, 51, 1821–1839. [Google Scholar] [CrossRef] [PubMed]
  77. Oberschelp, G.; Guarnaschelli, A.B.; Teson, N.; Harrand, L.; Podesta, F.E.; Margarit, E. Cold acclimation and freezing tolerance in three Eucalyptus species: A metabolomic and proteomic approach. Plant Physiol. Biochem. 2020, 154, 316–327. [Google Scholar] [CrossRef] [PubMed]
  78. Wang, Y.; Hou, Y.; Qiu, J.; Wang, H.; Wang, S.; Tang, L.; Tong, X.; Zhang, J. Abscisic acid promotes jasmonic acid biosynthesis via a ‘SAPK10-bZIP72-AOC’ pathway to synergistically inhibit seed germination in rice (Oryza sativa). New Phytol. 2020, 228, 1336–1353. [Google Scholar] [CrossRef]
  79. Ju, L.; Jing, Y.; Shi, P.; Liu, J.; Chen, J.; Yan, J.; Chu, J.; Chen, K.M.; Sun, J. JAZ proteins modulate seed germination through interaction with ABI5 in bread wheat and Arabidopsis. New Phytol. 2019, 223, 246–260. [Google Scholar] [CrossRef]
  80. Wang, X.; Li, Q.; Xie, J.; Huang, M.; Jiang, D. Abscisic acid and jasmonic acid are involved in drought priming-induced tolerance to drought in wheat. Crop J. 2020, 9, 120–132. [Google Scholar] [CrossRef]
  81. Pan, J.; Hu, Y.; Wang, H.; Guo, Q.; Chen, Y.; Howe, G.A.; Yu, D. Molecular Mechanism Underlying the Synergetic Effect of Jasmonate on Abscisic Acid Signaling during Seed Germination in Arabidopsis. Plant Cell 2020, 32, 3846–3865. [Google Scholar] [CrossRef]
  82. Utsugi, S.; Ashikawa, I.; Nakamura, S.; Shibasaka, M. TaABI5, a wheat homolog of Arabidopsis thaliana ABA insensitive 5, controls seed germination. J. Plant Res. 2020, 133, 245–256. [Google Scholar] [CrossRef] [PubMed]
  83. Jacobson, M.D. Reactive oxygen species and programmed cell death. Trends Biochem. Sci. 1996, 21, 83–86. [Google Scholar] [CrossRef]
  84. Mehdy, M.C. Active Oxygen Species in Plant Defense against Pathogens. Plant Physiol. 1994, 105, 467–472. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Bolwell, G.P.; Blee, K.A.; Butt, V.S.; Davies, D.R.; Gardner, S.L.; Gerrish, C.; Minibayeva, F.; Rowntree, E.G.; Wojtaszek, P. Recent advances in understanding the origin of the apoplastic oxidative burst in plant cells. Free Radic. Res. 1999, 31 (Suppl. S1), S137–S145. [Google Scholar] [CrossRef]
  86. Wu, J.; Yang, R.; Yang, Z.; Yao, S.; Zhao, S.; Wang, Y.; Li, P.; Song, X.; Jin, L.; Zhou, T.; et al. ROS accumulation and antiviral defence control by microRNA528 in rice. Nat. Plants 2017, 3, 16203. [Google Scholar] [CrossRef] [PubMed]
  87. Bajji, M.; Lutts, S.; Kinet, J. Water deficit effects on solute contribution to osmotic adjustment as a function of leaf ageing in three durum wheat (Triticum durum Desf.) cultivars performing differently in arid conditions. Plant Sci. 2001, 160, 669–681. [Google Scholar] [CrossRef]
  88. Lu, Q.; Guo, F.; Xu, Q.; Cang, J. LncRNA improves cold resistance of winter wheat by interacting with miR398. Funct. Plant Biol. 2020, 47, 544–557. [Google Scholar] [CrossRef]
  89. Zhang, X.Z. The measurement and mechanism of lipid peroxidation and SOD, POD and CAT activities in biological system. Res. Methodol. Crop Physiol. 1992, 7, 208–211. [Google Scholar]
  90. Agurla, S.; Gahir, S.; Munemasa, S.; Murata, Y.; Raghavendra, A.S. Mechanism of Stomatal Closure in Plants Exposed to Drought and Cold Stress. Adv. Exp. Med. Biol. 2018, 1081, 215–232. [Google Scholar]
  91. Chen, M.; Thelen, J.J. ACYL-LIPID DESATURASE2 is required for chilling and freezing tolerance in Arabidopsis. Plant Cell 2013, 25, 1430–1444. [Google Scholar] [CrossRef] [Green Version]
  92. Dowgert, M.F.; Steponkus, P.L. Behavior of the Plasma Membrane of Isolated Protoplasts during a Freeze-Thaw Cycle. Plant Physiol. 1984, 75, 1139–1151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Thomashow, M.F. Plant Cold Acclimation: Freezing Tolerance Genes and Regulatory Mechanisms. Annu. Rev. Plant Physiol. Plant Mol. Biol. 1999, 50, 571–599. [Google Scholar] [CrossRef] [Green Version]
  94. An, J.P.; Xu, R.R.; Liu, X.; Su, L.; Yang, K.; Wang, X.F.; Wang, G.L.; You, C.X. Abscisic acid insensitive 4 interacts with ICE1 and JAZ proteins to regulate ABA signaling-mediated cold tolerance in apple. J. Exp. Bot. 2022, 73, 980–997. [Google Scholar] [CrossRef] [PubMed]
  95. Chen, R.; Jiang, H.; Li, L.; Zhai, Q.; Qi, L.; Zhou, W.; Liu, X.; Li, H.; Zheng, W.; Sun, J.; et al. The Arabidopsis mediator subunit MED25 differentially regulates jasmonate and abscisic acid signaling through interacting with the MYC2 and ABI5 transcription factors. Plant Cell 2012, 24, 2898–2916. [Google Scholar] [CrossRef] [Green Version]
  96. Qu, L.J. A High-Throughput Screening System for Arabidopsis Transcription Factors and Its Application to Med25-Dependent Transcriptional Regulation. Mol. Plant 2011, 4, 546–555. [Google Scholar]
  97. Zinsmeister, J.; Lalanne, D.; Terrasson, E.; Chatelain, E.; Vandecasteele, C.; Vu, B.L.; Dubois-Laurent, C.; Geoffriau, E.; Signor, C.L.; Dalmais, M.; et al. ABI5 Is a Regulator of Seed Maturation and Longevity in Legumes. Plant Cell 2016, 28, 2735–2754. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Skubacz, A.; Daszkowska-Golec, A.; Szarejko, I. The Role and Regulation of ABI5 (ABA-Insensitive 5) in Plant Development, Abiotic Stress Responses and Phytohormone Crosstalk. Front. Plant Sci. 2016, 7, 1884. [Google Scholar] [CrossRef] [Green Version]
  99. Lang, V.; Palva, E.T. The expression of a rab-related gene, rab18, is induced by abscisic acid during the cold acclimation process of Arabidopsis thaliana (L.) Heynh. Plant Mol. Biol. 1992, 20, 951–962. [Google Scholar] [CrossRef]
  100. Chinnusamy, V.; Zhu, J.; Zhu, J.K. Cold stress regulation of gene expression in plants. Trends Plant Sci. 2007, 12, 444–451. [Google Scholar] [CrossRef]
  101. Chen, C.; Chen, H.; Zhang, Y.; Thomas, H.R.; Frank, M.H.; He, Y.; Xia, R. TBtools: An Integrative Toolkit Developed for Interactive Analyses of Big Biological Data. Mol. Plant 2020, 13, 1194–1202. [Google Scholar] [CrossRef]
  102. 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]
  103. Kumar, S.; Stecher, G.; Tamura, K. MEGA7: Molecular Evolutionary Genetics Analysis Version 7.0 for Bigger Datasets. Mol. Biol. Evol. 2016, 33, 1870–1874. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Mount, D.W. Using the Basic Local Alignment Search Tool (BLAST). CSH Protoc. 2007, 2007, pdb.top17. [Google Scholar] [CrossRef] [PubMed]
  105. Wang, D.; Zhang, Y.; Zhang, Z.; Zhu, J.; Yu, J. KaKs_Calculator 2.0: A toolkit incorporating gamma-series methods and sliding window strategies. Genom. Proteom. Bioinform. 2010, 8, 77–80. [Google Scholar] [CrossRef] [Green Version]
  106. Lynch, M.; Conery, J.S. The evolutionary fate and consequences of duplicate genes. Science 2000, 290, 1151–1155. [Google Scholar] [CrossRef] [Green Version]
  107. Yang, Z.; Gu, S.; Wang, X.; Li, W.; Tang, Z.; Xu, C. Molecular evolution of the CPP-like gene family in plants: Insights from comparative genomics of Arabidopsis and rice. J. Mol. Evol. 2008, 67, 266–277. [Google Scholar] [CrossRef]
  108. Kim, D.; Langmead, B.; Salzberg, S.L. HISAT: A fast spliced aligner with low memory requirements. Nat. Methods 2015, 12, 357–360. [Google Scholar] [CrossRef] [Green Version]
  109. Dong, W.; Wang, M.; Xu, F.; Quan, T.; Peng, K.; Xiao, L.; Xia, G. Wheat oxophytodienoate reductase gene TaOPR1 confers salinity tolerance via enhancement of abscisic acid signaling and reactive oxygen species scavenging. Plant Physiol. 2013, 161, 1217–1228. [Google Scholar] [CrossRef] [Green Version]
  110. Wang, Y.; Yuan, G.; Yuan, S.; Duan, W.; Wang, P.; Bai, J.; Zhang, F.; Gao, S.; Zhang, L.; Zhao, C. TaOPR2 encodes a 12-oxo-phytodienoic acid reductase involved in the biosynthesis of jasmonic acid in wheat (Triticum aestivum L.). Biochem. Biophys. Res. Commun. 2016, 470, 233–238. [Google Scholar] [CrossRef]
  111. Xin, M.; Wang, Y.; Yao, Y.; Xie, C.; Peng, H.; Ni, Z.; Sun, Q. Diverse set of microRNAs are responsive to powdery mildew infection and heat stress in wheat (Triticum aestivum L.). BMC Plant Biol. 2010, 10, 123. [Google Scholar] [CrossRef] [Green Version]
  112. Guo, C.; Jiang, Y.; Shi, M.; Wu, X.; Wu, G. ABI5 acts downstream of miR159 to delay vegetative phase change in Arabidopsis. New Phytol. 2021, 231, 339–350. [Google Scholar] [CrossRef] [PubMed]
  113. Wang, J.; Ma, X.M.; Kojima, M.; Sakakibara, H.; Hou, B.K. Glucosyltransferase UGT76C1 finely modulates cytokinin responses via cytokinin N-glucosylation in Arabidopsis thaliana. Plant Physiol. Biochem. 2013, 65, 9–16. [Google Scholar] [CrossRef] [PubMed]
  114. Nakayama, D.G.; Santos, J.C.; Kishi, L.T.; Pedezzi, R.; Santiago, A.C.; Soares-Costa, A.; Henrique-Silva, F. A transcriptomic survey of Migdolus fryanus (sugarcane rhizome borer) larvae. PLoS ONE 2017, 12, e0173059. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Du, Z.; Bramlage, W.J. Modified thiobarbituric acid assay for measuring lipid oxidation in sugar-rich plant tissue extracts. J. Agric. Food. Chem. 1992, 40, 1566–1570. [Google Scholar] [CrossRef]
  116. Aebi, H. Catalase in vitro‖ on antioxidative defenses. Methods Enzymol. 1984, 105, 121–126. [Google Scholar]
  117. Maehly, A.; Chance, B. Assay of catalases and peroxidases. Methods Enzymol. 1955, 2, 764–775. [Google Scholar]
  118. Dutilleul, C.; Garmier, M.; Noctor, G.; Mathieu, C.; Chetrit, P.; Foyer, C.H.; de Paepe, R. Leaf mitochondria modulate whole cell redox homeostasis, set antioxidant capacity, and determine stress resistance through altered signaling and diurnal regulation. Plant Cell 2003, 15, 1212–1226. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The phylogenetic tree of bZIP genes in wheat and A. thaliana. Different colors indicate different subfamilies (A–M).
Figure 1. The phylogenetic tree of bZIP genes in wheat and A. thaliana. Different colors indicate different subfamilies (A–M).
Ijms 23 02351 g001
Figure 2. Distribution of introns-Exons and Conserved motifs of the TabZIP gene family in wheat. (Left) phylogenetic relationship and location of Conserved motifs on each gene and (right) distribution of exons/introns of the gene family in wheat.
Figure 2. Distribution of introns-Exons and Conserved motifs of the TabZIP gene family in wheat. (Left) phylogenetic relationship and location of Conserved motifs on each gene and (right) distribution of exons/introns of the gene family in wheat.
Ijms 23 02351 g002
Figure 3. Collinearity analysis of TabZIP gene family. (a) Chromosome locations of TabZIP genes. The chromosome number is on the left side of each chromosome. The scale on the left represents chromosome length. (b) Synteny analysis of TabZIP genes between Triticum aestivum and four representative plant species. Gray lines indicate the collinear bars within the Triticum aestivum and other plant genomes, whereas the red lines highlight the syntenic TabZIP gene pairs.
Figure 3. Collinearity analysis of TabZIP gene family. (a) Chromosome locations of TabZIP genes. The chromosome number is on the left side of each chromosome. The scale on the left represents chromosome length. (b) Synteny analysis of TabZIP genes between Triticum aestivum and four representative plant species. Gray lines indicate the collinear bars within the Triticum aestivum and other plant genomes, whereas the red lines highlight the syntenic TabZIP gene pairs.
Ijms 23 02351 g003
Figure 4. The syntenic pairs of wheat TabZIP genes from different duplication mode diagrams. Syntenic pairs from different duplication modes were linked by corresponding color lines. Red, blue, purple, and green represent syntenic regions with segmental, tandem, transposed, and proximal duplication. Others syntenic pairs genes were shown with pale yellow lines.
Figure 4. The syntenic pairs of wheat TabZIP genes from different duplication mode diagrams. Syntenic pairs from different duplication modes were linked by corresponding color lines. Red, blue, purple, and green represent syntenic regions with segmental, tandem, transposed, and proximal duplication. Others syntenic pairs genes were shown with pale yellow lines.
Ijms 23 02351 g004
Figure 5. The cis-element of promoter in TabZIP gene family. The depth of the color represents the number of times the cis-element occurs.
Figure 5. The cis-element of promoter in TabZIP gene family. The depth of the color represents the number of times the cis-element occurs.
Ijms 23 02351 g005
Figure 6. Transcriptional profiles of TabZIP gene family members in wheat under heat stress (H-1/H-6), drought stress (D-1/D-6), and combined stress (CS-1/CS-6). Grids with colors indicate low/down expression (blue), high/up expression(red), and non-expression/no change (yellow). According to the similarity of expression characteristics, some genes were classified into Group I, Group II and Group III respectively.
Figure 6. Transcriptional profiles of TabZIP gene family members in wheat under heat stress (H-1/H-6), drought stress (D-1/D-6), and combined stress (CS-1/CS-6). Grids with colors indicate low/down expression (blue), high/up expression(red), and non-expression/no change (yellow). According to the similarity of expression characteristics, some genes were classified into Group I, Group II and Group III respectively.
Ijms 23 02351 g006
Figure 7. Relative expression analysis of 9 genes in the TabZIP gene family under cold stress, drought stress, salt stress, ABA treatment, JA treatment, and SA treatment in wheat. All values in the figure are mean ± SD (n = 3), p values (versus 0 h) were corrected with ANOVA, p < 0.05 indicates significant difference, indicated by *; p < 0.001 indicates high significance, indicated by **.
Figure 7. Relative expression analysis of 9 genes in the TabZIP gene family under cold stress, drought stress, salt stress, ABA treatment, JA treatment, and SA treatment in wheat. All values in the figure are mean ± SD (n = 3), p values (versus 0 h) were corrected with ANOVA, p < 0.05 indicates significant difference, indicated by *; p < 0.001 indicates high significance, indicated by **.
Ijms 23 02351 g007
Figure 8. Analysis of cold resistance and expression of key stress response genes in transgenic plants. (a) Phenotype and survival rate of Arabidopsis lines under freezing stress. (b) MDA content and electrolyte leakage were measured. (c) Heat maps of stress response genes expression at 24 °C, 4 °C and −10 °C in different plant lines, respectively. All values in the figure are mean ± SD (n = 3), p values (versus WT) were corrected with ANOVA, * p < 0.05, ** p < 0.01, and **** p < 0.0001.
Figure 8. Analysis of cold resistance and expression of key stress response genes in transgenic plants. (a) Phenotype and survival rate of Arabidopsis lines under freezing stress. (b) MDA content and electrolyte leakage were measured. (c) Heat maps of stress response genes expression at 24 °C, 4 °C and −10 °C in different plant lines, respectively. All values in the figure are mean ± SD (n = 3), p values (versus WT) were corrected with ANOVA, * p < 0.05, ** p < 0.01, and **** p < 0.0001.
Ijms 23 02351 g008
Figure 9. Overexpression of TaABI5 in A. thaliana enhanced tolerance to freezing stresses and ROS-scavenging ability. (a) Histochemical staining assays with NBT. (b) Histochemical staining assays with DAB. (c) The 2-week-old transgenic, wild type, and abi5 seedlings were exposed to freezing treatment. The plants were acclimated at 4 °C for 3 d, then treated at −10 °C for 2 h, and restored to normal culture for 7 d. (d) CAT and POD of Arabidopsis lines under freezing stress. All values in the figure are mean ± SD (n = 3), p values (versus WT) were corrected with ANOVA, ** p < 0.01, *** p < 0.001, and **** p < 0.0001.
Figure 9. Overexpression of TaABI5 in A. thaliana enhanced tolerance to freezing stresses and ROS-scavenging ability. (a) Histochemical staining assays with NBT. (b) Histochemical staining assays with DAB. (c) The 2-week-old transgenic, wild type, and abi5 seedlings were exposed to freezing treatment. The plants were acclimated at 4 °C for 3 d, then treated at −10 °C for 2 h, and restored to normal culture for 7 d. (d) CAT and POD of Arabidopsis lines under freezing stress. All values in the figure are mean ± SD (n = 3), p values (versus WT) were corrected with ANOVA, ** p < 0.01, *** p < 0.001, and **** p < 0.0001.
Ijms 23 02351 g009
Figure 10. Y2H assay of interaction between TaICE1 and TaABI5. (a) Transcription activation assay of TaABI5 in yeast. (b) TaICE1 and TaABI5 were co-transformed into yeast. Colonies were assayed for SD/-Leu-Trp autotrophy and SD/-Leu-Trp-His-Ade with 40 mM Aureobasidin (AbA) autotrophy.
Figure 10. Y2H assay of interaction between TaICE1 and TaABI5. (a) Transcription activation assay of TaABI5 in yeast. (b) TaICE1 and TaABI5 were co-transformed into yeast. Colonies were assayed for SD/-Leu-Trp autotrophy and SD/-Leu-Trp-His-Ade with 40 mM Aureobasidin (AbA) autotrophy.
Ijms 23 02351 g010
Figure 11. Fluorescence of TaABI5 in tobacco leaf cells. Subcellular localization of TaABI5 (top) and empty vector (bottom) in dark field, bright field, and merged field. Scale bars = 25 µm.
Figure 11. Fluorescence of TaABI5 in tobacco leaf cells. Subcellular localization of TaABI5 (top) and empty vector (bottom) in dark field, bright field, and merged field. Scale bars = 25 µm.
Ijms 23 02351 g011
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liang, Y.; Xia, J.; Jiang, Y.; Bao, Y.; Chen, H.; Wang, D.; Zhang, D.; Yu, J.; Cang, J. Genome-Wide Identification and Analysis of bZIP Gene Family and Resistance of TaABI5 (TabZIP96) under Freezing Stress in Wheat (Triticum aestivum). Int. J. Mol. Sci. 2022, 23, 2351. https://doi.org/10.3390/ijms23042351

AMA Style

Liang Y, Xia J, Jiang Y, Bao Y, Chen H, Wang D, Zhang D, Yu J, Cang J. Genome-Wide Identification and Analysis of bZIP Gene Family and Resistance of TaABI5 (TabZIP96) under Freezing Stress in Wheat (Triticum aestivum). International Journal of Molecular Sciences. 2022; 23(4):2351. https://doi.org/10.3390/ijms23042351

Chicago/Turabian Style

Liang, Yi, Jingqiu Xia, Yunshuang Jiang, Yuzhuo Bao, Huichan Chen, Duojia Wang, Da Zhang, Jing Yu, and Jing Cang. 2022. "Genome-Wide Identification and Analysis of bZIP Gene Family and Resistance of TaABI5 (TabZIP96) under Freezing Stress in Wheat (Triticum aestivum)" International Journal of Molecular Sciences 23, no. 4: 2351. https://doi.org/10.3390/ijms23042351

APA Style

Liang, Y., Xia, J., Jiang, Y., Bao, Y., Chen, H., Wang, D., Zhang, D., Yu, J., & Cang, J. (2022). Genome-Wide Identification and Analysis of bZIP Gene Family and Resistance of TaABI5 (TabZIP96) under Freezing Stress in Wheat (Triticum aestivum). International Journal of Molecular Sciences, 23(4), 2351. https://doi.org/10.3390/ijms23042351

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