Next Article in Journal
Omega-3 Fatty Acids in Arterial Hypertension: Is There Any Good News?
Previous Article in Journal
Proteomic and Metabolomic Changes in Psoriasis Preclinical and Clinical Aspects
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genome-Wide Analysis of WRKY Transcription Factors Involved in Abiotic Stress and ABA Response in Caragana korshinskii

Key Laboratory of Plants Adversity Adaptation and Genetic Improvement in Cold and Arid Regions of Inner Mongolia, Inner Mongolia Agricultural University, Hohhot 010018, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(11), 9519; https://doi.org/10.3390/ijms24119519
Submission received: 27 April 2023 / Revised: 23 May 2023 / Accepted: 25 May 2023 / Published: 30 May 2023
(This article belongs to the Section Molecular Plant Sciences)

Abstract

:
The WRKY transcription factor family plays a vital role in plant development and environmental response. However, the information of WRKY genes at the genome-wide level is rarely reported in Caragana korshinskii. In this study, we identified and renamed 86 CkWRKY genes, which were further classified into three groups through phylogenetic analysis. Most of these WRKY genes were clustered and distributed on eight chromosomes. Multiple sequence alignment revealed that the conserved domain (WRKYGQK) of the CkWRKYs was basically consistent, but there were also six variation types (WRKYGKK, GRKYGQK, WRMYGQK, WRKYGHK, WKKYEEK and RRKYGQK) that appeared. The motif composition of the CkWRKYs was quite conservative in each group. In general, the number of WRKY genes gradually increased from lower to higher plant species in the evolutionary analysis of 28 species, with some exceptions. Transcriptomics data and RT-qPCR analysis showed that the CkWRKYs in different groups were involved in abiotic stresses and ABA response. Our results provided a basis for the functional characterization of the CkWRKYs involved in stress resistance in C. korshinskii.

1. Introduction

Due to their static nature, plants must adapt to complex and constantly changing environmental conditions such as drought, salinity, alkalinity and temperature extremes. These environmental factors have profound effects on plant growth and yield [1,2]. Hence, it is pivotal to investigate the molecular mechanisms through which plants respond to stress. Recent research indicates that transcription factors (TFs) play important regulatory functions in plant stress response, for example, genes encoding TFs such as bHLH, MYB, NAC, WRKY and ERF have been found to be up-regulated in wheat under stress conditions [3].
WRKY is one of the largest TF families in plants and plays an important role in the transcriptional regulation of plants in response to a variety of environmental stresses [4,5,6]. They are named as such because of the highly conserved WRKY domain, containing 60 amino acids and a zinc-finger motif (C-X4-7-C-X22-23-H-X-H), and are classified into three groups according to the number of WRKY domains at the amino terminus and the types of zinc-finger motif at the carboxyl terminus. Group I has two WRKY domains and a C2H2 zinc-finger motif, group II has one WRKY domain and a C2H2 zinc-finger motif and group III has one WRKY domain and a C2HC zinc-finger motif. The group II WRKYs can be further divided into five subgroups, from IIa to IIe, based on other conserved amino acid sequences other than the WRKY domain [7]. WRKYs respond to environmental stress by binding to promoters of plant genes through the cis-element, with DNA sequence C/TTGACC/T, termed as W-box [8].
A large number of previous reports have demonstrated that WRKY family members regulate plant responses to various stresses and phytohormones [4,9]. In 1994, the first WRKY gene, SPF1, was identified from sweet potato [10]. With the completion of plant genome sequencing, more and more WRKY genes have been reported at the whole genome level in different species, such as Arabidopsis thaliana, Lilium longiflorum [11], Daucus carota [12], Gossypium raimondii and G. hirsutum [13] and Musa acuminata [14]. It was confirmed that BHWRKY1 enhanced dehydration tolerance in transgenic Nicotiana tabacum [15]. SbWRKY30 regulated the expression of Sbrd19 through binding to the W-box of its promoter to affect the drought tolerance of sorghum [16]. CmWRKY17 overexpression in A. thaliana decreased its tolerance to salt compared to wild type [17]. OsMADS57 and its interacting OSTB1 synergetically activated the transcriptional regulation of OsWRKY94 at low temperatures [18] and WRKYs also helped plants cope with reactive oxygen species [19], UV [20] and so on. In terms of phytohormones, OsWRKY45 was up-regulated by ABA [21]; AtWRKY12 and AtWRKY13 were involved in the GA signaling regulation of plant flowering time [22]; PqWRKY1 was induced by MeJA [23]; AtWRKY46, AtWRKY54, as well as AtWRKY70, were involved in both BR-regulated plant growth and drought response [24].
Caragana korshinskii, a leguminous shrub that predominantly grows in arid and semi-arid areas of China, is very tolerant to a wide range of environmental factors, such as drought, salt and alkali. To adapt to these adverse environments, it has evolved some stress resistance mechanisms. However, little information is available on WRKY TFs at the genomic level in C. korshinskii. Based on the genome data sequenced by our group and collaborators, we conducted a distribution, re-nomenclature, classification, conserved domain and motifs analysis of all CkWRKYs for the first time and explored the CkWRKY genes evolution in different plant species. This study will shed light on understanding the stress resistance mechanism of C. korshinskii via WRKYs by combining the transcriptome data under different stresses.

2. Results

2.1. Identification and Chromsome Distribution of WRKY Genes from C. korshinskii

To identify the WRKY encoding genes in C. korshinskii, we searched the sequences of WRKY genes from the draft genome data of C. korshinskii and the transcriptome data of C. intermedia (a closely related species of C. korshinskii) from previous studies [25]. After removing the incomplete and redundant sequences, a total of 86 putative WRKY genes were identified from the draft genome of C. korshinskii, which were designated as CkWRKY1 to CkWRKY86 based on their chromosomal distribution (Figure 1, Table 1 and Table S1). Among the 86 genes, 80 genes (from CkWRKY1 to CkWRKY80) were almost evenly distributed on the seven chromosomes and had nothing to do with the length of the chromosomes. There was a cluster distribution of some genes on one chromosome, which was similar to other species [26]. In addition, group II WRKYs existed in each chromatin. Only one gene (CkWRKY81) was located on the eighth chromosome and five genes (CkWRKY82~CkWRKY86) could not be found to have an association with any chromosome due to genome assembly problems.
The physical and chemical properties showed that the number of amino acids of these CkWRKYs varied between 139 and 744 aa, the relative molecular weight (MW) was between 15.44 and 81.33 kDa and the isoelectric point (pI) spanned from 4.82 to 10.14, according to the ExPASy online software. All of these CkWRKYs were predicted to be located in the nucleus by subcellular localization online software (Plant-mPLoc).

2.2. Phylogenetic and Multiple Sequence Alignment of the CkWRKYs and AtWRKYs

To classify the 86 CkWRKY proteins, an unrooted phylogenetic tree using the WRKY domain of the 86 CkWRKYs and 72 AtWRKYs was constructed using the Neighbor Joining (NJ) method (Figure 2 and Figure S2). According to the classification and phylogenetic analysis of the WRKYs from A. thaliana, the 86 CkWRKYs were divided into three groups (Figure 2 and Table 1). Group I contained 16 members, and 15 of them contained two conserved WRKY domains, which were further designated as the N-terminal WRKY domain (IN) and the C-terminal WRKY domain (IC); however, CkWRKY73-IN and AtWRKY10-IC contained only one domain [27]. It had been reported that some group I WRKYs contained three or four WRKY domains [28,29], which indicated that the number of conserved WRKY domains varied in group I in different species. Fifty-six members were clustered in group II, which possessed only one WRKY domain, and could be further divided into IIa-IIe subgroups. In addition, 14 WRKYs with one WRKY domain were identified in group III. The group I and group II CkWRKYs harbored a C2H2 zinc finger motif, while group III had a C2HC zinc finger motif, which was consistent with WRKY family members in other plant species.
Sequence alignment analysis of the 86 CkWRKYs showed that each conserved domain had about 60 aa residues, including the highly conserved WRKYGQK motif and the zinc-finger motif. Previous studies have shown that the mutation of this conserved WRKY domain and changes to the zinc-finger structure had an impact on their binding activity [30,31,32].
Among the amino acid sequences of the 86 CkWRKYs, the highly conservative WRKYGQK motif had been found in 76 sequences, while the remaining ten CkWRKYs had six types of mutations (i.e., WRKYGKK, GRKYGQK, WRMYGQK, WRKYGHK, WKKYEEK and RRKYGQK). Among the 16 WRKYs in group I, the WRKY domain of two members (CkWRKY73 and CkWRKY70-IN) were mutated into WKKYEK and RRKYGQK, respectively. In group II, six CkWRKYs had mutations in the WRKY domain; these were CkWRKY13 in subgroup IIa (with GRKYGQK motif) and four members, including CkWRKY20, CkWRKY30, CkWRKY31 and CkWRKY37, in subgroup Iic (with WRKYGKK motif). In addition, CkWRKY30 and CkWRKY3 also had deletion in their zinc-finger structure. Another member of the Iic subgroup, CkWRKY52-Iic, lost its zinc-finger completely. The WRKY domain of the CkWRKY33 in subgroup Iie was mutated into WRMYGQK. The WRKY domain of two members, CkWRKY63 and CkWRKY80, in group III had been mutated into WRKYGHK and WRKYGKK, respectively. In order to understand the variation of the conserved WRKY domain in the different groups of CkWRKYs, the number of variant proteins was compared with the number of total members in the same groups to obtain the variation rate. The variation rate was 12.5% in group I, 10.71% in group II and 14.29% in group III. In addition, half of the mutated conserved motif was WRKYGKK (Figure 3, Tables S2 and S3).

2.3. The Conserved Motif Analysis

As the conservative motif was sometimes related to the function of WRKY proteins [33], we used MEME to predict the conserved motifs of the CkWRKYs to better characterize this gene family. In our case, overall, 15 motifs were defined from the 86 CkWRKY proteins (Figure 4). Among them, the distribution of motifs 1 (which had the conserved WRKY domain sequences) and 2 were highly conserved in 85 CkWRKYs, while CkWRKY73, which lacked the C-terminal WRKY domain, only had motif 1. Moreover, the conserved motifs of the same group of CkWRKY proteins were basically the same, but there were also slight differences.
Among the 16 group I CkWRKYs, nine motifs (1–6, 9, 10 and 15) were identified. CkWRKY73 only contained motifs 1 and 4 due to the lack of the C-terminal WRKY domain sequence. CkWRKY54 contained five motifs (1–4, 6). CkWRKY7 also contained five motifs (1–5). CkWRKY77 and CkWRKY85 contained six motifs (1–6). CkWRKY12, 17, 45, 47, 50 and 72 contained seven motifs (1–6 and 10). CkWRKY83 and CkWRKY69 also contained seven motifs (1–6 and 15), while CkWRKY16 and CkWRKY84 contained eight motifs (1–6, 10 and 15) and CkWRKY70 contained all nine motifs. The above results indicated that motifs 1 and 4 were indispensable for the group I CkWRKYs.
In the 56 group II CkWRKYs, subgroup IIa had seven CkWRKYs, with seven kinds of motifs (1, 2, 4, 5, 7, 8 and 9) appearing. With the exception of CkWRKY58, which contained all of the motifs, the remaining six CkWRKYs exhibited motif deficiency and motifs 1, 2 and 4 were present in all of the Iia sequences. Subgroup Iib had 14 CkWRKYs, with eight kinds of motifs (1, 2, 4, 5, 7, 8, 9 and 13) appearing. With the exception of CkWRKY9, which lacked motif 8, CkWRKY46, which lacked motif 5, and CkWRKY40, which lacked motifs 5 and 13, the remaining 11 CkWRKYs all contained these eight kinds of motifs. There were 16 CkWRKYs in Iic, with a total of four motifs (1, 2, 4 and 5) found, nine CkWRKYs contained all four of these motifs, except that CkWRKY51 and CkWRKY52 lacked motifs 4 and 5 or 2, and CkWRKY2, 21, 30, 31 and 36 lacked motif 4. There were seven CkWRKYs in Iid, with four motifs (1, 2, 4 and 11) appearing; with the exception of CkWRKY 26, 32 and 53 lacking motif 4, the rest of the CkWRKYs contained all four motifs. There were 12 CkWRKYs in Iie, with four motifs (1, 2, 4 and 5) found; eight members, including CkWRKY4, 23, 48, 33, 55, 66, 79 and 82 lacked motif 5, and the remaining four members contained all four motifs. In summary, among the group II CkWRKYs, Iib had the most diverse motifs, and the motifs in the other subgroups (Iia, Iic and Iie) could be found in Iib. While Iid had the special motif 11, the motif composition of the Iid and Iie subgroups were basically similar, although the arrangement was slightly different. For example, the seven WRKYs of Iid subgroup all had motifs 1, 2 and 11, and their arrangements were completely consistent. In addition, we found that, with the exception of the absence of motif 4 in Iic and Iid, all of the other subgroups contained motifs 1, 2 and 4, indicating that motifs 1 and 2 were conserved in group II, while motif 4 may have emerged late and may play a role in the functional changes of different subgroups. Taken together, the above results indicated that there were more conserved motif changes in the group II WRKYs, suggesting that there might also be more changes in their function.
In the 14 group III CkWRKYs, seven motifs (1, 2, 4, 5, 9, 12 and 14) appeared. These CkWRKYs all contained motifs 1 and 2, while CkWRKY75 also contained motif 13, CkWRKY49 also contained motif 4, CkWRKY80 also contained 5, CkWRKY3, 18, 34, 61, 62 and 63 all contained motif 12, CkWRKY1 and 19 contained motifs 5 and 12 and CkWRKY42, 43 and 74 also contained motifs 5, 12 and 14.
In summary, the results of the motif analysis for the different groups of CkWRKYs showed that motifs 1 and 2 existed simultaneously and were closely adjacent in all CkWRKYs, indicating that these two motifs might be the most primitive and played an indispensable role in the basic function of CkWRKYs. In addition, a pair of closely adjacent motifs, 3 and 6, were found, which only existed simultaneously in all members of the group I genes. According to the currently existing hypothesis, the group I WRKY proteins emerged first, followed by group II and III proteins [34]; it could be inferred that motifs 3 and 6 disappeared during evolution.

2.4. Evolution Analysis of WRKYs among Different Species

To understand the evolution of CkWRKYs in plant lineage, 27 more representative plant species were selected according to the complexity of the genome in the database (PlantTFDB). In this study, 1896 complete sequences of WRKY conserved domains were obtained from 28 species (including C. korshinskii) of plants spanning from algae, mosses, ferns, gymnosperms to angiosperms) (Figure 5).
Among all of the selected algae, the number of WRKYs was the smallest, while the number of WRKYs in angiosperm was relatively large. However, there were exceptions; for example, the angiosperm Amborella trichopoda had 29 WRKYs, which was less than the gymnosperm Pseudotsuga menziesii with 70 WRKYs and also less than Sphagnum palustre with 37 WRKYs. The 64 WRKYs of angiosperm Epipactis helleborine were also less than those of Pseudotsuga menziesii. In addition, the 15 WRKYs of Fern Selaginella moellendorffii were also less than that of Sphagnum palustre.
In general, the number of WRKYs varied greatly from lower to higher plants, ranging between 1 and 175. In algae, Chlamydomonas reinhardtii and Dunaliella parva contained one and five WRKYs, respectively. Among the mosses, 14 and 37 WRKYs were found in Marchantia polymorpha and Sphagnum palustre, respectively. In fern, Selaginella moellendorffii contained 15 WRKYs. Gymnosperms contained 9–70 WRKYs, whereas angiosperms contained 29–175 WRKYs.
As C. korshinskii belongs to leguminous plants, nine leguminous species were selected for comparison with C. korshinskii here (Figure 5). The results showed that the number of WRKYs in leguminous plants had little change, except that Glycine max had 175 WRKYs due to a WGD event [35]. The number of WRKYs in other legumes was within the range of 77–100 and C. korshinskii contained 86 WRKYs, which was relatively in the middle.
In order to explore the evolutionary relationship of WRKYs in plants, the WRKYs from 28 plant species were divided into seven categories (I, IIa, IIb, IIc, IId, IIe and III) based on their homology with A. thaliana (Figure 5 and Figure 6a). Based on the previous WRKYs origin hypothesis, group I appeared first, and then evolved into groups II and III [34]. We found that the aquatic algae only contained group I WRKYs, but group II and III began to appear in terrestrial plant mosses. The result was basically consistent with the above hypothesis. Although Picea glauca had no group I WRKYs, when we checked the initial download sequence, we did find that it contained one sequence with an incomplete WRKY domain. In addition, we also noticed that subgroup IIa did not begin to appear in the angiosperm Epipactis helleborine, while subgroups IIb, IIe and III had already appeared in bryophytes. There were zero in the gymnosperms Pinus taeda and Picea sitchensis, which might indicate that group II appeared earlier than III. Group II genes appeared in terrestrial plants and had occupied a dominant position in terms of quantity, among which the group II genes of Oryza sativa accounted for at least 48.45% and Picea glauca for at most 92.31% in all of the WRKY genes. Subgroups IIb and IIc again accounted for the majority of group II WRKYs, in which IIb accounted for a relatively high proportion, varying between 26% and 39% at the beginning, and then significantly decreasing to 8% in Picea glauca; meanwhile, the proportion of Iic was relatively stable, with at least 22% in Marchantia polymorpha and at most 58% in Triticum urartu (Figure 5 and Figure 6b).
Additionally, in the process of the sequence analysis, we found that the WRKYs sequence contained not only the WRKY conservative domain, but also some other domains (Figure 6c,d). In addition to algae and Selaginella moellendorffii, other terrestrial plants contained the Plant_zn_cluster domain [29,36]. Further analysis found that the Plant_zn_cluster domain was accompanied by the IId subgroup. In addition, we also found some other types of domains in the WRKYs sequence, such as the bZIP, WD40 and NAM domains. Since the other domains appeared in the WRKYs sequence, it was speculated that these minority domains might play a regulatory role in the coordination with the conserved WRKY domains. Among the 28 species, at least one kind of other domain (such as Plant_zn_clust) appeared, with the exception of Chlamydomonas reinhardtii, Dunaliella parva and Selaginella moellendorffii. In addition, Phyllostachys edulis, Populus euphratica and Trifolium pratense, with the largest number of other domains (as many as seven kinds), were all angiosperms. Furthermore, it was observed that Populus euphratica had the largest number of other domains, while the number of WRKYs in Populus euphratica was not the largest, indicating that the number of other domains was not positively correlated with the number of genes. Moreover, because Populus euphratica is a typical drought-resistant species in the northwest region, it was speculated that these extra domains might be related to its drought resistance.

2.5. Expression Profiles of CkWRKYs under Drought, Salt, Alkali and ABA Treatments

To understand the roles of the CkWRKY gene family under drought, salt, alkali and ABA treatment, we analyzed their expression profiles using the transcriptome data (Log2FC ≥ 1/FC ≥ 2). As shown in Figure 7a,b, the expression of CkWRKY72-I was up-regulated under all four treatments; CkWRKY17-I and CkWRKY50-I were up-regulated under drought, high pH and ABA treatment; CkWRKY35-IIe and CkWRKY69-I were up-regulated under high pH and ABA treatment; CkWRKY10-Iic and CkWRKY59-Iib were up-regulated under drought and salt treatment; CkWRKY47-I was up-regulated under salt and down-regulated under high pH treatment. The expression of the remaining 15 genes—CkWRKY64-Iib, CkWRKY15-Iib, CkWRKY39-Iib, CkWRKY29-Iia, CkWRKY52-Iic, CkWRKY28-Iia, CkWRKY19-III, CkWRKY30-Iic, CkWRKY38-Iic, CkWRKY24-Iic, CkWRKY34-III, CkWRKY76-Iic, CkWRKY67-Iib, CkWRKY41-Iib and CkWRKY77-I—changed only under drought stress. From the diagram (Figure 7b), it was found that some genes responded to multiple treatments. Due to the function of ABA in the stress pathway, the genes that responded to both stress and ABA were counted. For example, five WRKY genes (CkWRKY17, 35, 50, 69, 72) under drought, NaCl and high pH treatment overlapped with the genes under ABA treatment, indicating that these genes might exercise a stress-related function through ABA signal pathway.
In order to understand the response of CkWRKY genes to different stresses, these differentially expressed genes (DEGs) were classified into various groups (Figure 7c); we found that the expression level of the group I genes changed under all four treatments and the expression level of the group II genes changed under two treatments, while the group III genes changed only under one treatment, indicating that the group I genes played a core role in the plant stress response, followed by the group II genes and finally by the group III genes. In addition, the evolutionary analysis (Figure 5a) showed that the group I genes were the most primitive ones of the WRKY gene family and group II and group III had originated from group I, which might show that the more primitive the genes are, the more critical their function.
To analyze the DEGs in different groups under different treatments, the CkWRKYs that changed after four treatments were analyzed. As shown in Figure 7a–d, under drought treatment, a total of 20 DEGs were detected in the transcriptome, including four in group I, fourteen in group II and two in group III; under salt treatment, a total of four genes were changed, including two in group I and two in group II; under high pH treatment, seven genes were affected, including five in group I and two in group II; under ABA treatment, a total of five genes were changed, including four in group I and one in group II. DEGs were mainly clustered around the same branch. For example, the group I differential genes in the transcriptome were clustered together, such as CkWRKY77 and CkWRKY17, as shown in Figure 7a, indicating that similar genes might have similar functions. Among the 16 group I genes in the genome of C. korshinskii, six played a role under four stresses, accounting for 37.5%. For the 56 group II genes, 16 played a role in these stresses, accounting for 28.57%. Among the 14 group III genes, two were responsive to these stresses, accounting for 14.29%. In addition, the Iib genes (42.86%) in group II were the most prevalent in all groups.
The above results show that group I played an Important role In different stresses. The group I genes were clustered together in the phylogenetic analysis, indicating that these genes might possess basic functions because they were the earliest genes of origin. In addition, although the conserved motifs of the group I members were basically the same, their functions were divergent, indicating that other sequences of these group I members might also play an important role.
In order to verify the reliability of the transcriptome results, RT-qPCR experiments were carried out (Figure 8). The results showed that these 24 DEGs had different degrees of expression variation compared with the control under the corresponding treatment, indicating that the transcriptome results had certain reliability. However, the situation was not very consistent, especially under drought treatment.
To investigate whether these 24 DEGs possessed interactions, homologous modeling with A. thaliana was used. The interaction network diagram of the DEGs under stress and ABA treatment is shown in Figure 9. Among the 24 genes, 17 were connected, indicating that these genes might interact. In addition, seven genes might not participate in the interaction because they existed alone in the network. In addition, it was predicted that the maximum number of lines of the interactive network connections was eight with CkWRKY17 and CkWRKY50. In addition, through the statistics of the number of lines for different groups of genes, it was found that there were no obvious rules. Taking six group I genes as an example, three genes (CkWRKY17, 50 and 72) had six lines, two genes (CkWRKY47 and 69) had one line and one gene (CkWRKY77) had zero lines. After observing the relationship between these six genes, it was found that only CkWRKY47 and CkWRKY69 had a predicted interaction. This result showed that the genes within the same group did not have similar and related interaction rules.

3. Discussion

Growing in the arid and semi-arid regions in Northwest China, C. korshinskii possesses drought resistance and adaptability to saline-alkali soil. Studying its adaption mechanism to these environmental factors will lay the foundation for molecular genetic breeding in the future. Previous studies have shown that WRKYs play an important role in stress response in plants, for example, AtWRKY1 regulates stomatal movement in A. thaliana under drought stress [37]; inactivated OsWRKY5 transcription factor enhances the drought tolerance of its mutants through the ABA signaling pathway [38]; GhWRKY1-like mediates drought tolerance in A. thaliana by regulating ABA biosynthesis [39]; the overexpression of CmWRKY17 increases the salt sensitivity of transgenic plants in both in Chrysanthemum morifolium and A. thaliana [17]; the overexpression of PtWRKY39 enhances the salt and alkali tolerance of transgenic tobacco [40]. In our research, through the analysis of the genome and transcriptome results, we evaluated the evolutionary status and structural characteristics of CkWRKYs in different groups among 28 species and found that multiple CkWRKYs genes were responsive to different environmental factors. Finally, through the analysis of two omics data, we found that the group I CkWRKYs possessed the most stress-related genes, suggesting that this group is closely related to stress response, which laid a molecular foundation for the functional research of this group. In brief, our research might provide essential information for understanding the stress tolerance of C. korshinskii at the genetic level.
WRKYs play an important regulatory role in various stresses [4,5], as well as in development [41,42]. With the development of genome sequencing technology, more and more WRKYs in legume species have been found, such as in G. max [43,44], Medicago sativa and M. truncatula [45,46]. In this study, the WRKY genes of C. korshinskii were investigated at the genome level for the first time. Eighty-one genes out of the full-length CkWRKYs were identified and located in eight chromatins; the other five might not be defined in chromatin due to assembly problems, which has been reported in other studies, such as in Andrographis paniculata [27]. In addition, it was observed that most of the CkWRKY genes showed a cluster distribution in chromatins, which had also been reported in other legume studies [26]. In the previous research on CiWRKYs based on the transcriptome, only 28 full-length CiWRKY genes were defined [25], indicating that the genomic research was conducive to the definition and discovery of CkWRKY genes and laid a foundation for the later research of CkWRKY genes.
As the nomenclature of CkWRKYs was mainly based on homology with their orthologs in A. thaliana and other legume species or the order of discovery in this plant, it is quite confusing that one CkWRKY gene might have different names. With the aid of the genome data, we renamed the 86 CkWRKYs according to their distribution on the chromosomes and avoided the dilemma of several CkWRKYs having one AtWRKY ortholog (Table 1). The published CkWRKYs/CiWRKYs and their corresponding new names are summarized in Table S4.
Phylogenetic analysis showed that the WRKY TFs were divided into three groups (group I, II and III), and group II was further divided into subgroups IIa-IIe, of which IIc had the largest number of WRKYs. These results were consistent with those obtained by other researchers [26,46,47,48,49,50]. In previous studies on CkWRKYs, the WRKYGQK conserved motif of one protein sequence was replaced by WKKYEEK, and the WRKYGQK of three protein sequences was replaced by WRKYGKK [25]. In our study, the WRKYGQK conserved domain had six types of mutations (WRKYGKK, GRKYGQK, WRMYGQK, WRKYGHK, WKKYEEK and RRKYGQK) in ten protein sequences, of which WRKYGKK accounted for half of all mutations, which showed that the number and types of conserved domains increased with the identification of more genes in the genome. The greatest conserved domain variation was found in Iic, and these results have also been reported in other species [26], suggesting that the WRKY proteins of IIc might have a variety of biological functions through the conserved domain variation. Motif analysis showed that some motifs were unique to groups I and III, which might lead to the specificity of the gene function of the two groups, and compared with the other groups, IIc only contained up to four conserved motifs. In addition, nine of the sixteen sequences contained these four motifs, which indicated that the motif of the same subgroup was highly conserved; similar conclusions were drawn in the study of Cajanus cajan [51].
Previous studies have shown that WRKY is one of the largest transcription factor families in plants, and the main research focused on higher plant WRKY proteins that have been detected in various organisms, such as spike mosses, single-celled green algae, slime molds and protozoa [7]. In monocots and dicots, such as rice, soybean, wheat and cotton, an especially large number of WRKY proteins with different functions have been confirmed in recent years [21,49,52,53,54]. However, our study was the first to identify and characterize WRKY proteins from the whole genome data of C. korshinskii and to carry out an evolution analysis from lower algae to higher monocotyledons, which had not previously been reported. Our study investigated WRKYs from the perspective of evolution, and the results showed that WRKY genes have been presented since green plants, and gene classification showed that group I genes were first presented in lower algae, and then group II and group III genes were gradually presented, indicating that the other subgroup genes might originate from group I. Some parts of these results were also reported in other studies [34,55,56,57], and our study enriched and validated the previous results and hypotheses.
The present study found that WRKY genes have complex regulatory networks in abiotic stresses and phytohormone responses [4,5,58]. ABA is a phytohormone and plays essential roles in plant responses to abiotic stress [4]. In our study, drought, salt, high pH and ABA were selected to treat C. korshinskii seedlings, and samples were taken for the transcriptome sequencing. Subsequently, the differential WRKYs genes in the transcriptome data before and after treatment were analyzed and summarized and RT-qPCR verification was performed. A total of 24 DEGs were selected, and the expression analysis of these genes revealed different expression patterns for each CkWRKY under different treatments, providing a valuable resource for gene function research. Stress-related WRKYs were also found in the transcriptomes of other species, for example, GaWRKYs were salt-responsive [59], CtWRKYs responded to salt and cold stress [60] and TaWRKYs were responsive to drought [61]. The above results indicated that transcriptome sequencing played an essential role in the gene function research. In addition, the combination of transcriptome and genome analysis was helpful to understand the function of the genes in different subgroups in a more predictable way. In our study, among the DEGs under different treatments, the group I genes responded to a variety of stresses. In addition, the group II genes had the largest number, and subgroups IIb and IIc had the largest number in group II. Our study showed that the DEGs ratio was 16/86 in the group I CkWRKYs, which originated the earliest and responded to a variety of stress signals. The above results showed that although the number of group I genes was small, they played a key role in C. korshinskii, most likely because it appeared first. For example, CkWRKY50-I (formerly CkWRKY33) enhanced the drought tolerance in transgenic A. thaliana [62] and CiWRKY26, a homologous gene of CkWRKY72-I, responded to multiple stresses through expression pattern analysis and was located in the nucleus [25]. We also found that CkWRKY50-I showed a response to drought, and CkWRKY72-I was responsive to various stresses. In other species, studies have also shown that the WRKY genes in group I have multiple functions, for example, GhWRKY3-I could be induced by phytohormones and pathogens [63], NbWRKYs-I [64] were related to defense response and AtWRKY1-I played an important regulatory role in the drought, salt, light and nitrogen signaling pathways [37,65,66].
In conclusion, the definition of WRKY genes based on their genome, the study of gene phylogeny and classification analysis were conducive to the finding of more genes and the classifying of different types of genes. In addition, the evolutionary analysis of 28 different species was helpful to understand the origin relation of the WRKYs. Further, the application of the transcriptome was very helpful for gene function prediction, combined with the gene classification and evolutionary relationship.

4. Materials and Methods

4.1. Genome-Wide Identification and Distribution of CkWRKYs

To comprehensively identify the CkWRKY genes, the genome sequences were obtained from the C. korshinskii genome V1.0, which was completed by our team and Lanzhou University, and the transcriptome data from previous studies [25]. The NCBI Conserved Domains Database (CDD) was used to identify the WRKY domain, then the incomplete WRKY domain and the repetitive sequences were manually removed; finally, the CkWRKYs were identified. After this, ExPASy (http://web.expasy.org/compute_pi accessed on 1 March 2022) and Plant-mPLoc (http://www.csbio.sjtu.edu.cn/bioinf/plant-multi/ accessed on 1 March 2022) were used to predict the molecular weight (MW), isoelectric point (pI) and protein subcellular localization of the defined CkWRKY proteins.
To locate the positions of CkWRKY genes on the chromosomes of C. korshinskii, the MapInspect Software (http://www.softsea.com/download/MapInspect.html accessed on 1 March 2022) was used to align the distribution of the putative CkWRKYs based on the genome annotation.

4.2. Phylogenetic, Conserved Domains and Motif Analysis

The phylogenetic tree of C. korshinskii and A. thaliana was constructed using the truncated amino acid sequences containing a WRKY conserved domain using MEGA7.0 with the Neighbor Joining (NJ) method and the bootstrap test was set for 1000 replications. Multiple sequence alignment files were also obtained using MEGA7.0 to analyze the WRKY conserved domains, and the logo of these sequences was constructed using the WebLogo server (http://weblogo.berkeley.edu/logo.cgi accessed on 20 March 2022). MEME (http://meme.nbcr.net/meme/ accessed on 20 March 2022) analysis was performed to discover the conserved motifs of the full-length protein, with the number of displayed motifs set to 15 and the rest using default settings. Visualization and beautification of the phylogenetic tree, multiple sequence alignment and conserved motif diagram were performed using iTOL (https://itol.embl.de/ accessed on 25 March 2022).

4.3. Evolution Analysis

In order to understand the evolution of CkWRKYs in different species, WRKY family members from 27 other representative species from lower to higher were selected to compare with C. korshinskii for evolutionary analysis. These species were: Chlamydomonas reinhardtii, Dunaliella parva, Marchantia polymorpha, Sphagnum palustre, Selaginella moellendorffii, Pseudotsuga menziesii, Pinus taeda, Picea glauca, Picea sitchensis, Amborella trichopoda, Epipactis helleborine, Zea mays, Oryza sativa, Phyllostachys edulis, Triticum urartu, Aquilegia coerulea, Arabidopsis thaliana, Populus euphratica, Arachis ipaensis, Cajanus cajan, Glycine max, Phaseolus vulgaris, Vigna angularis, Lotus japonicus, Cicer arietinum, Medicago truncatula and Trifolium pratense, respectively. The WRKY sequences of these species were downloaded from PlantTFDB (http://planttfdb.gao-lab.org/family accessed on 1 April 2022), the repetitive sequences were manually removed, then the sequences without conservative WRKY domains were deleted through CDD analysis; the remaining sequences were used for comparative analysis. The evolution of 28 species (covering algae, mosses, ferns, gymnosperms and angiosperms) was analyzed using the TimeTree database (http://timetree.org/ accessed on 1 May 2022).

4.4. Abiotic Stress and ABA Treatments

The seedlings of C. korshinskii were cultured in the growth chamber. The culture conditions were as follows: vegetative soil and vermiculite (V/V = 1:3), photorhythm 16/8 h (light/dark) and culture temperature of about 22 °C. When the seedlings had grown for 3 weeks, they was treated with drought (withholding water), salt (watered with 200 mM NaCl solution), high pH (watered with 200 mM NaHCO3 solution using NaOH to adjust the pH to 10) and ABA (Sprayed with 100 μM ABA solution), then the whole plant of each seedling was sampled at 0, 6 and 12 days, respectively, after drought, salt and ABA treatments, while samples under high pH treatment were collected in at 0, 3 and 6 days due to the premature yellowish color of the seedlings after treatment. The mixture of three seedlings was taken as one sample, and the experiments was repeated for three times. All samples were snap frozen with liquid nitrogen and stored at −80 °C for transcriptome sequencing and real-time quantitative PCR (RT-qPCR).

4.5. Heatmap Analysis and RT-qPCR Verification of the Transcriptome Data

To check the response of CkWRKYs to abiotic stress and ABA, differentially expressed CkWRKYs with over 2-fold changes in the transcriptome data were screened, Heatmaps of these CkWRKYs were constructed using log2-based FPKM values and the results were visualized using TBtools. All of the selected genes were verified through RT-qPCR (using TB Green qPCR Master Mix (TaKaRa)) with a Roche LightCycler 480 system and the primer sequences were listed in Table S5. The results of the RT-qPCR were calculated according to the 2−ΔΔCT method [67] and were normalized using CiEF1α reference gene [GenBank No.: KC679842] [68]. All RT-qPCR assays were carried out with three technical replicates.

5. Conclusions

In this study, we performed a genome-wide identification of WRKY genes in C. korshinskii and identified a total of 86 CkWRKY genes. These genes were classified into three groups—I, II and III—and we further classified the members in group II into five subgroups based on their phylogenetic relationships. The basic features, phylogenetic, conserved domain, motif and species evolution of these genes were analyzed, providing a foundational understanding of the evolutionary relationships within the CkWRKY gene family. The expression profile of the CkWRKYs was studied using the transcriptomic data, and the results revealed that 24 CkWRKYs were influenced by abiotic stresses and ABA. Our results provided a basis for the functional characterization of the CkWRKYs involved in stress resistance in C. korshinskii.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms24119519/s1 [25,69,70,71,72,73,74,75,76,77,78].

Author Contributions

J.L. carried out the experiment and prepared the draft of the manuscript. G.L., Y.W. and G.W. conceived the study and designed the experiments. R.W. prepared the material. J.L. contributed to the data analysis. G.L. revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by The Major Project of Inner Mongolia Natural Science Foundation (2019ZD05); National Natural Science Foundation of China (No. 31860217); Special Project in Major Science and Technology Program of Hohhot (2022-S-M-1-3); Development Fund from the Central Government Guiding Local Science and Technology Program (2022ZY0008); Research Team of Inner Mongolia Agricultural University (BR22-13-05); The University Scientific and Technological Innovation Team Project of Inner Mongolia (NMGIRT2222).

Data Availability Statement

The data used to support the findings of this study are available from the corresponding author upon request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kosova, K.; Vitamvas, P.; Prasil, I.T. Wheat and barley dehydrins under cold, drought, and salinity-what can LEA-II proteins tell us about plant stress response? Front. Plant Sci. 2014, 5, 343. [Google Scholar] [CrossRef] [PubMed]
  2. Boyer, J.S. Plant Productivity and Environment. Science 1982, 218, 443–448. [Google Scholar] [CrossRef] [PubMed]
  3. Li, L.; Han, C.; Yang, J.; Tian, Z.; Jiang, R.; Yang, F.; Jiao, K.; Qi, M.; Liu, L.; Zhang, B.; et al. Comprehensive Transcriptome Analysis of Responses during Cold Stress in Wheat (Triticum aestivum L.). Genes 2023, 14, 844. [Google Scholar] [CrossRef] [PubMed]
  4. Chen, L.; Song, Y.; Li, S.; Zhang, L.; Zou, C.; Yu, D. The role of WRKY transcription factors in plant abiotic stresses. Biochim. Biophys. Acta-Gene Regul. Mech. 2012, 1819, 120–128. [Google Scholar] [CrossRef]
  5. Li, W.; Pang, S.; Lu, Z.; Jin, B. Function and Mechanism of WRKY Transcription Factors in Abiotic Stress Responses of Plants. Plants 2020, 9, 1515. [Google Scholar] [CrossRef]
  6. Eulgem, T.; Somssich, I.E. Networks of WRKY transcription factors in defense signaling. Curr. Opin. Plant Biol. 2007, 10, 366–371. [Google Scholar] [CrossRef]
  7. Rushton, P.J.; Somssich, I.E.; Ringler, P.; Shen, Q.J. WRKY transcription factors. Trends Plant Sci. 2010, 15, 247–258. [Google Scholar] [CrossRef]
  8. Ulker, B.; Somssich, I.E. WRKY transcription factors: From DNA binding towards biological function. Curr. Opin. Plant Biol. 2004, 7, 491–498. [Google Scholar] [CrossRef]
  9. Tripathi, P.; Rabara, R.C.; Rushton, P.J. A systems biology perspective on the role of WRKY transcription factors in drought responses in plants. Planta 2014, 239, 255–266. [Google Scholar] [CrossRef]
  10. Ishiguro, S.; Nakamura, K. Characterization of a cDNA encoding a novel DNA-binding protein, SPF1, that recognizes SP8 sequences in the 5′ upstream regions of genes coding for sporamin and β-amylase from sweet potato. Mol. Gen. Genet. MGG 1994, 244, 563–571. [Google Scholar] [CrossRef]
  11. Kumari, S.; Kanth, B.K.; Ahn, J.Y.; Kim, J.H.; Lee, G.J. Genome-Wide Transcriptomic Identification and Functional Insight of Lily WRKY Genes Responding to Botrytis Fungal Disease. Plants 2021, 10, 776. [Google Scholar] [CrossRef] [PubMed]
  12. Li, M.Y.; Xu, Z.S.; Tian, C.; Huang, Y.; Wang, F.; Xiong, A.S. Genomic identification of WRKY transcription factors in carrot (Daucus carota) and analysis of evolution and homologous groups for plants. Sci. Rep. 2016, 6, 23101. [Google Scholar] [CrossRef] [PubMed]
  13. Dou, L.; Zhang, X.; Pang, C.; Song, M.; Wei, H.; Fan, S.; Yu, S. Genome-wide analysis of the WRKY gene family in cotton. Mol. Genet. Genom. MGG 2014, 289, 1103–1121. [Google Scholar] [CrossRef] [PubMed]
  14. Jia, C.; Wang, Z.; Wang, J.; Miao, H.; Zhang, J.; Xu, B.; Liu, J.; Jin, Z.; Liu, J. Genome-Wide Analysis of the Banana WRKY Transcription Factor Gene Family Closely Related to Fruit Ripening and Stress. Plants 2022, 11, 662. [Google Scholar] [CrossRef]
  15. Wang, Z.; Zhu, Y.; Wang, L.; Liu, X.; Liu, Y.; Phillips, J.; Deng, X. A WRKY transcription factor participates in dehydration tolerance in Boea hygrometrica by binding to the W-box elements of the galactinol synthase (BhGolS1) promoter. Planta 2009, 230, 1155–1166. [Google Scholar] [CrossRef]
  16. Yang, Z.; Chi, X.; Guo, F.; Jin, X.; Luo, H.; Hawar, A.; Chen, Y.; Feng, K.; Wang, B.; Qi, J.; et al. SbWRKY30 enhances the drought tolerance of plants and regulates a drought stress-responsive gene, SbRD19, in sorghum. J. Plant Physiol. 2020, 246, 153142. [Google Scholar] [CrossRef]
  17. Li, P.; Song, A.; Gao, C.; Wang, L.; Wang, Y.; Sun, J.; Jiang, J.; Chen, F.; Chen, S. Chrysanthemum WRKY gene CmWRKY17 negatively regulates salt stress tolerance in transgenic chrysanthemum and Arabidopsis plants. Plant Cell Rep. 2015, 34, 1365–1378. [Google Scholar] [CrossRef]
  18. Chen, L.; Zhao, Y.; Xu, S.; Zhang, Z.; Xu, Y.; Zhang, J.; Chong, K. OsMADS57 together with OsTB1 coordinates transcription of its target OsWRKY94 and D14 to switch its organogenesis to defense for cold adaptation in rice. New Phytol. 2018, 218, 219–231. [Google Scholar] [CrossRef]
  19. Cheong, Y.H.; Chang, H.S.; Gupta, R.; Wang, X.; Zhu, T.; Luan, S. Transcriptional profiling reveals novel interactions between wounding, pathogen, abiotic stress, and hormonal responses in Arabidopsis. Plant Physiol. 2002, 129, 661–677. [Google Scholar] [CrossRef]
  20. Kilian, J.; Whitehead, D.; Horak, J.; Wanke, D.; Weinl, S.; Batistic, O.; D’Angelo, C.; Bornberg-Bauer, E.; Kudla, J.; Harter, K. The AtGenExpress global stress expression data set: Protocols, evaluation and model data analysis of UV-B light, drought and cold stress responses. Plant J. Cell Mol. Biol. 2007, 50, 347–363. [Google Scholar] [CrossRef]
  21. Qiu, Y.; Yu, D. Over-expression of the stress-induced OsWRKY45 enhances disease resistance and drought tolerance in Arabidopsis. Environ. Exp. Bot. 2009, 65, 35–47. [Google Scholar] [CrossRef]
  22. Li, W.; Wang, H.; Yu, D. Arabidopsis WRKY Transcription Factors WRKY12 and WRKY13 Oppositely Regulate Flowering under Short-Day Conditions. Mol. Plant 2016, 9, 1492–1503. [Google Scholar] [CrossRef]
  23. Discovery of WRKY transcription factors through transcriptome analysis and characterization of a novel methyl jasmonate-inducible PqWRKY1 gene from Panax quinquefolius. Plant Cell Tissue Organ Cult. 2013, 114, 269–277. [CrossRef]
  24. Chena, J.; Nolana, T.; Yea, H.; Zhanga, M.; Tongc, H. 3 Arabidopsis WRKY46, WRKY54, and WRKY70 Transcription Factors 4 Are Involved in Brassinosteroid-Regulated Plant Growth and Drought 5 Responses. Plant Cell 2017, 29, 1425–1439. [Google Scholar] [CrossRef] [PubMed]
  25. Wan, Y.; Mao, M.; Wan, D.; Yang, Q.; Yang, F.; Mandlaa; Li, G.; Wang, R. Identification of the WRKY gene family and functional analysis of two genes in Caragana intermedia. BMC Plant Biol. 2018, 18, 31. [Google Scholar] [CrossRef] [PubMed]
  26. Wu, J.; Chen, J.; Wang, L.; Wang, S. Genome-Wide Investigation of WRKY Transcription Factors Involved in Terminal Drought Stress Response in Common Bean. Front. Plant Sci. 2017, 8, 380. [Google Scholar] [CrossRef]
  27. Zhang, R.; Chen, Z.; Zhang, L.; Yao, W.; Xu, Z.; Liao, B.; Mi, Y.; Gao, H.; Jiang, C.; Duan, L.; et al. Genomic Characterization of WRKY Transcription Factors Related to Andrographolide Biosynthesis in Andrographis paniculata. Front. Genet. 2020, 11, 601689. [Google Scholar] [CrossRef]
  28. Song, H.; Sun, W.; Yang, G.; Sun, J. WRKY transcription factors in legumes. BMC Plant Biol. 2018, 18, 243. [Google Scholar] [CrossRef]
  29. Mohanta, T.K.; Park, Y.H.; Bae, H. Novel Genomic and Evolutionary Insight of WRKY Transcription Factors in Plant Lineage. Sci. Rep. 2016, 6, 37309. [Google Scholar] [CrossRef]
  30. Chen, X.; Li, C.; Wang, H.; Guo, Z. WRKY transcription factors: Evolution, binding, and action. Phytopathol. Res. 2019, 1, 13. [Google Scholar] [CrossRef]
  31. Cheng, X.; Zhao, Y.; Jiang, Q.; Yang, J.; Zhao, W.; Taylor, I.A.; Peng, Y.-L.; Wang, D.; Liu, J. Structural basis of dimerization and dual W-box DNA recognition by rice WRKY domain. Nucleic Acids Res. 2019, 47, 4308–4318. [Google Scholar] [CrossRef]
  32. Englbrecht, C.C.; Schoof, H.; Bohm, S. Conservation, diversification and expansion of C2H2 zinc finger proteins in the Arabidopsis thaliana genome. BMC Genom. 2004, 5, 39. [Google Scholar] [CrossRef] [PubMed]
  33. Park, C.Y.; Lee, J.H.; Yoo, J.H.; Moon, B.C.; Choi, M.S.; Kang, Y.H.; Lee, S.M.; Kim, H.S.; Kang, K.Y.; Chung, W.S.; et al. WRKY group IId transcription factors interact with calmodulin. FEBS Lett. 2005, 579, 1545–1550. [Google Scholar] [CrossRef] [PubMed]
  34. Wu, K.L.; Guo, Z.J.; Wang, H.H.; Li, J. The WRKY family of transcription factors in rice and Arabidopsis and their origins. Dna Res. 2005, 12, 9–26. [Google Scholar] [CrossRef] [PubMed]
  35. Song, H.; Wang, P.; Lin, J.Y.; Zhao, C.; Bi, Y.; Wang, X. Genome-Wide Identification and Characterization of WRKY Gene Family in Peanut. Front. Plant Sci. 2016, 7, 534. [Google Scholar] [CrossRef] [PubMed]
  36. Babu, M.M.; Iyer, L.M.; Balaji, S.; Aravind, L. The natural history of the WRKY-GCM1 zinc fingers and the relationship between transcription factors and transposons. Nucleic Acids Res. 2006, 34, 6505–6520. [Google Scholar] [CrossRef]
  37. Qiao, Z.; Li, C.L.; Zhang, W. WRKY1 regulates stomatal movement in drought-stressed Arabidopsis thaliana. Plant Mol. Biol. 2016, 91, 53–65. [Google Scholar] [CrossRef]
  38. Lim, C.; Kang, K.; Shim, Y.; Yoo, S.C.; Paek, N.C. Inactivating transcription factor OsWRKY5 enhances drought tolerance through abscisic acid signaling pathways. Plant Physiol. 2022, 188, 1900–1916. [Google Scholar] [CrossRef]
  39. Hu, Q.; Ao, C.; Wang, X.; Wu, Y.; Du, X. GhWRKY1-like, a WRKY transcription factor, mediates drought tolerance in Arabidopsis via modulating ABA biosynthesis. BMC Plant Biol. 2021, 21, 458. [Google Scholar] [CrossRef]
  40. Niu, Y.; Li, X.; Xu, C.; Ajab, Z.; Liu, Q.; Majeed, Z.; Guan, Q. Analysis of drought and salt-alkali tolerance in tobacco by overexpressing WRKY39 gene from Populus trichocarpa. Plant Signal. Behav. 2021, 16, 1918885. [Google Scholar] [CrossRef]
  41. Liu, L.; Xu, W.; Hu, X.; Liu, H.; Lin, Y. W-box and G-box elements play important roles in early senescence of rice flag leaf. Sci. Rep. 2016, 6, 20881. [Google Scholar] [CrossRef] [PubMed]
  42. Huang, J.; Chen, F.; Wu, S.; Li, J.; Xu, W. Cotton GhMYB7 is predominantly expressed in developing fibers and regulates secondary cell wall biosynthesis in transgenic Arabidopsis. Sci. China Life Sci. 2016, 59, 194–205. [Google Scholar] [CrossRef] [PubMed]
  43. Yang, Y.; Zhou, Y.; Chi, Y.; Fan, B.; Chen, Z. Characterization of Soybean WRKY Gene Family and Identification of Soybean WRKY Genes that Promote Resistance to Soybean Cyst Nematode. Sci. Rep. 2017, 7, 17804. [Google Scholar] [CrossRef] [PubMed]
  44. Mao, P.; Jin, X.; Bao, Q.; Mei, C.; Zhou, Q.; Min, X.; Liu, Z. WRKY Transcription Factors in Medicago sativa L.: Genome-Wide Identification and Expression Analysis Under Abiotic Stress. DNA Cell Biol. 2020, 39, 2212–2225. [Google Scholar] [CrossRef] [PubMed]
  45. Hui, S.; Zhibiao, N. Genome-wide identification and analysis of WRKY transcription factors in Medicago truncatula. Hereditas 2014, 36, 152–168. [Google Scholar] [CrossRef]
  46. 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]
  47. Wei, W.; Hu, Y.; Han, Y.T.; Zhang, K.; Zhao, F.L.; Feng, J.Y. The WRKY transcription factors in the diploid woodland strawberry Fragaria vesca: Identification and expression analysis under biotic and abiotic stresses. Plant Physiol. Biochem. 2016, 105, 129–144. [Google Scholar] [CrossRef]
  48. Song, H.; Wang, P.; Hou, L.; Zhao, S.; Wang, X. Global Analysis of WRKY Genes and Their Response to Dehydration and Salt Stress in Soybean. Front. Plant Sci. 2016, 7, 9. [Google Scholar] [CrossRef]
  49. He, G.H.; Xu, J.Y.; Wang, Y.X.; Liu, J.M.; Li, P.S.; Chen, M.; Ma, Y.Z.; Xu, Z.S. Drought-responsive WRKY transcription factor genes TaWRKY1 and TaWRKY33 from wheat confer drought and/or heat resistance in Arabidopsis. BMC Plant Biol. 2016, 16, 116. [Google Scholar] [CrossRef]
  50. Yin, G.; Xu, H.; Xiao, S.; Qin, Y.; Li, Y.; Yan, Y.; Hu, Y. The large soybean (Glycine max) WRKY TF family expanded by segmental duplication events and subsequent divergent selection among subgroups. BMC Plant Biol. 2013, 13, 148. [Google Scholar] [CrossRef]
  51. Yang, J.; Li, H.; Ma, R.; Chang, Y.; Qin, X.; Xu, J.; Fu, Y. Genome-wide transcriptome analysis and characterization of the cytochrome P450 flavonoid biosynthesis genes in pigeon pea (Cajanus cajan). Planta 2022, 255, 120. [Google Scholar] [CrossRef] [PubMed]
  52. Xiao, L.; Xi, B.; Xiaoli, S.; Dan, Z.; Baohui, L.; Wei, J.; Hua, C.; Lei, C.; Jing, W.; Mengran, H. Expression of wild soybean WRKY20 in Arabidopsis enhances drought tolerance and regulates ABA signalling. J. Exp. Bot. 2013, 64, 2155–2169. [Google Scholar]
  53. Qin, Y.; Tian, Y.; Liu, X. A wheat salinity-induced WRKY transcription factor TaWRKY93 confers multiple abiotic stress tolerance in Arabidopsis thaliana. Biochem. Biophys. Res. Commun. 2015, 464, 428–433. [Google Scholar] [CrossRef] [PubMed]
  54. Liu, X.; Song, Y.; Xing, F.; Wang, N.; Wen, F.; Zhu, C. GhWRKY25, a group I WRKY gene from cotton, confers differential tolerance to abiotic and biotic stresses in transgenic Nicotiana benthamiana. Protoplasma 2016, 253, 1265–1281. [Google Scholar] [CrossRef]
  55. Wu, W.; Zhu, S.; Xu, L.; Zhu, L.; Wang, D.; Liu, Y.; Liu, S.; Hao, Z.; Lu, Y.; Yang, L.; et al. Genome-wide identification of the Liriodendron chinense WRKY gene family and its diverse roles in response to multiple abiotic stress. BMC Plant Biol. 2022, 22, 25. [Google Scholar] [CrossRef]
  56. Rinerson, C.I.; Rabara, R.C.; Tripathi, P.; Shen, Q.J.; Rushton, P.J. The evolution of WRKY transcription factors. BMC Plant Biol. 2015, 15, 66. [Google Scholar] [CrossRef] [PubMed]
  57. Yue, H.; Chang, X.; Zhi, Y.; Wang, L.; Xing, G.; Song, W.; Nie, X. Evolution and Identification of the WRKY Gene Family in Quinoa (Chenopodium quinoa). Genes 2019, 10, 131. [Google Scholar] [CrossRef]
  58. Gang, L.; Shizhuo, Y.; Diqiu, Y.; Yanjuan, J. Arabidopsis WRKY57 functions as a node of convergence for jasmonic acid- and auxin-mediated signaling in jasmonic acid-induced leaf senescence. Plant Cell 2014, 26, 230–245. [Google Scholar]
  59. Fan, X.; Guo, Q.; Xu, P.; Gong, Y.; Shu, H.; Yang, Y.; Ni, W.; Zhang, X.; Shen, X. Transcriptome-wide identification of salt-responsive members of the WRKY gene family in Gossypium aridum. PLoS ONE 2015, 10, e0126148. [Google Scholar] [CrossRef]
  60. Chang, X.; Yang, Z.; Zhang, X.; Zhang, F.; Huang, X.; Han, X. Transcriptome-wide identification of WRKY transcription factors and their expression profiles under different stress in Cynanchum thesioides. PeerJ 2022, 10, e14436. [Google Scholar] [CrossRef]
  61. Okay, S.; Derelli, E.; Unver, T. Transcriptome-wide identification of bread wheat WRKY transcription factors in response to drought stress. Mol. Genet. Genom. 2014, 289, 765–781. [Google Scholar] [CrossRef]
  62. Li, Z.; Liang, F.; Zhang, T.; Fu, N.; Pei, X.; Long, Y. Enhanced tolerance to drought stress resulting from Caragana korshinskii CkWRKY33 in transgenic Arabidopsis thaliana. BMC Genom. Data 2021, 22, 11. [Google Scholar] [CrossRef] [PubMed]
  63. Guo, R.; Yu, F.; Gao, Z.; An, H.; Cao, X.; Guo, X. GhWRKY3, a novel cotton (Gossypium hirsutum L.) WRKY gene, is involved in diverse stress responses. Mol. Biol. Rep. 2011, 38, 49–58. [Google Scholar] [CrossRef] [PubMed]
  64. Sun, S.; Ren, Y.; Wang, D.; Farooq, T.; He, Z.; Zhang, C.; Li, S.; Yang, X.; Zhou, X. A group I WRKY transcription factor regulates mulberry mosaic dwarf-associated virus-triggered cell death in Nicotiana benthamiana. Mol. Plant Pathol. 2022, 23, 237–253. [Google Scholar] [CrossRef]
  65. Wu, X.; Xu, J.; Meng, X.; Fang, X.; Xia, M.; Zhang, J.; Cao, S.; Fan, T. Linker histone variant HIS1-3 and WRKY1 oppositely regulate salt stress tolerance in Arabidopsis. Plant Physiol. 2022, 189, 1833–1847. [Google Scholar] [CrossRef]
  66. Heerah, S.; Katari, M.; Penjor, R.; Coruzzi, G.; Marshall-Colon, A. WRKY1 Mediates Transcriptional Regulation of Light and Nitrogen Signaling Pathways. Plant Physiol. 2019, 181, 1371–1388. [Google Scholar] [CrossRef]
  67. Livak, K.J.; Schmittgen, T. Analysis of relative gene expression data using real-time quantitative PCR and the 2-DDCt method. Methods 2001, 25, 402–408. [Google Scholar] [CrossRef] [PubMed]
  68. Yang, Q.; Yin, J.; Li, G.; Qi, L.; Yang, F.; Wang, R.; Li, G. Reference gene selection for qRT-PCR in Caragana korshinskii Kom. under different stress conditions. Mol. Biol. Rep. 2014, 41, 2325–2334. [Google Scholar] [CrossRef]
  69. Wan, D.; Zhang, J.; Wang, R.; Li, G.; Wan, Y. Cloning and Bioinformatics Analysis of Caragana intermedia WRKY12 gene. Mol. Plant Breed. 2020, 18, 9. (In Chinese) [Google Scholar]
  70. Liu, J.; Wan, Y.; Liu, Y.; Wang, R.; Li, G. Cloning and Bioinformatics Analysis of Caragana intermedia WRKY30. Mol. Plant Breed. 2019, 17, 10. (In Chinese) [Google Scholar]
  71. Liu, J.; Wang, R.; Li, G.; Wan, Y. Cloning and prokaryotic expression of WRKY48 from Caragana intermedia. Open Life Sci. 2022, 17, 131–138. [Google Scholar] [CrossRef] [PubMed]
  72. Mao, M. Functional analysis of Two WRKY Genes in Caragana intermedia. 2017, Inner Mongolia Agricultural University. (In Chinese)
  73. Bai, M.; Wan, Y.; Li, G.; Wang, R. Cloning and Bioinformatics Analysis of CiWRKY17 Gene. Genom. Appl. Biol. 2020, 39, 9. (In Chinese) [Google Scholar]
  74. Yang, Q.; Zhang, T.; Wang, Y.; Li, G.; Ying, J.; Han, X.; Qi, L.; Li, G.; Wang, R. Construction of a Suppression Subtractive Hybridization Library of Caragana korshinskii Under Drought Stress and Cloning of CkWRKY1 Gene. Sci. Silvae Sin. 2013, 49, 62–68. (In Chinese) [Google Scholar]
  75. Wan, Y. Identification of the WRKY Gene Family and Functional Analysis of Two CiWRKYs in Caragana intermedia. 2018, Inner Mongolia Agricultural University. (In Chinese)
  76. Wan, Y.; Mao, M.; Wan, D.; Liu, J.; Wang, G.; Li, G.; Wang, R. Caragana intermedia WRKY75 Altered Arabidopsis thaliana Tolerance to Salt Stress and ABA. Acta Bot. Boreali-Occident. Sin. 2018, 38, 9. (In Chinese) [Google Scholar]
  77. Mao, M.; Wan, Y.; Li, G.; Yang, Q.; Wang, R. Cloning and Expression Analysis of CiWRKY2 Gene in Caragana intermedia. Mol. Plant Breed. 2017, 15, 9. (In Chinese) [Google Scholar]
  78. Zhang, W.; Li, S.; Liu, Y.; Wang, R.; Li, G.; Yang, Q.; Wan, Y. Cloning and Bioinformatics Analysis of Caragana intermedia WRKY15. Chin. J. Grassl. 2019, 9. (In Chinese) [Google Scholar]
Figure 1. Distribution of CiWRKY genes on eight chromosomes of C. korshinskii. (a) Distribution of 81 CkWRKY genes on each chromosome and the five genes marked by (>) in the lower right corner indicated that they had no chromatin to be located. (b) The distribution and quantity statistics of different groups and subgroups of CkWRKYs on each chromosome.
Figure 1. Distribution of CiWRKY genes on eight chromosomes of C. korshinskii. (a) Distribution of 81 CkWRKY genes on each chromosome and the five genes marked by (>) in the lower right corner indicated that they had no chromatin to be located. (b) The distribution and quantity statistics of different groups and subgroups of CkWRKYs on each chromosome.
Ijms 24 09519 g001
Figure 2. Phylogenetic analysis of 158 WRKYs containing the conserved WRKY domains from both C. korshinskii and A. thaliana using the NJ method. These WRKYs were clustered into eight groups or subgroups: IN, IC, IIa, IIb, IIc, IId, IIe and III. CkWRKYs were marked by the solid triangle and AtWRKYs were labeled with the hollow triangle.
Figure 2. Phylogenetic analysis of 158 WRKYs containing the conserved WRKY domains from both C. korshinskii and A. thaliana using the NJ method. These WRKYs were clustered into eight groups or subgroups: IN, IC, IIa, IIb, IIc, IId, IIe and III. CkWRKYs were marked by the solid triangle and AtWRKYs were labeled with the hollow triangle.
Ijms 24 09519 g002
Figure 3. Amino acid sequence alignment and Logo diagram of the typical WRKY domains of C. korshinskii. (a) Alignment of the amino acid sequence. (b) Logo diagram of the conserved WRKY domain; the red box indicated WRKY domain and the blue triangle indicated zinc-finger structure.
Figure 3. Amino acid sequence alignment and Logo diagram of the typical WRKY domains of C. korshinskii. (a) Alignment of the amino acid sequence. (b) Logo diagram of the conserved WRKY domain; the red box indicated WRKY domain and the blue triangle indicated zinc-finger structure.
Ijms 24 09519 g003
Figure 4. Conserved motifs in CkWRKY proteins. (a) Logo diagram of the conserved WRKY motif. (b) Conservative motif distribution in CkWRKYs. (a,b) Fifteen motifs were identified using MEME, as indicated by colored rectangles. The height of the rectangles was proportional to the-log (p-value), truncated at the height of a motif with a p-value of 1 × 10−10.
Figure 4. Conserved motifs in CkWRKY proteins. (a) Logo diagram of the conserved WRKY motif. (b) Conservative motif distribution in CkWRKYs. (a,b) Fifteen motifs were identified using MEME, as indicated by colored rectangles. The height of the rectangles was proportional to the-log (p-value), truncated at the height of a motif with a p-value of 1 × 10−10.
Ijms 24 09519 g004
Figure 5. Evolutionary analysis WRKY proteins of 28 plant species by TimeTree database.
Figure 5. Evolutionary analysis WRKY proteins of 28 plant species by TimeTree database.
Ijms 24 09519 g005
Figure 6. Statistics of the number and other domains of WRKY proteins in 28 species. (a) Statistics of the proportion and number of different groups and subgroups of WRKYs. (b) Statistics of the proportion and number of different subgroups of group II WRKYs. (c) Statistics of protein number in different groups and from different species. (d) Statistics of other domains of WRKY proteins in different groups and from different species.
Figure 6. Statistics of the number and other domains of WRKY proteins in 28 species. (a) Statistics of the proportion and number of different groups and subgroups of WRKYs. (b) Statistics of the proportion and number of different subgroups of group II WRKYs. (c) Statistics of protein number in different groups and from different species. (d) Statistics of other domains of WRKY proteins in different groups and from different species.
Ijms 24 09519 g006
Figure 7. Expression profiles of CkWRKY genes under the treatment of drought, salt, pH and ABA. (a) Heatmaps showing the clustering of these genes were created using log2-based FPKM values, and the scale represented the signal intensities of the FPKM values. (b) The Venn diagram showed the number of overlapped CkWRKY genes under different treatments. (ce) Proportion of differentially expressed CkWRKYs in various groups and subgroups.
Figure 7. Expression profiles of CkWRKY genes under the treatment of drought, salt, pH and ABA. (a) Heatmaps showing the clustering of these genes were created using log2-based FPKM values, and the scale represented the signal intensities of the FPKM values. (b) The Venn diagram showed the number of overlapped CkWRKY genes under different treatments. (ce) Proportion of differentially expressed CkWRKYs in various groups and subgroups.
Ijms 24 09519 g007aIjms 24 09519 g007b
Figure 8. Expression profiles of selected CkWRKYs under treatment of drought, salt, high pH and ABA using RT-qPCR. Note: The results of the RT-qPCR were calculated according to the 2−ΔΔCT method, and were normalized using CiEF1α as the reference gene, * means p < 0.05 and ** means p < 0.01, using Student’s t-test.
Figure 8. Expression profiles of selected CkWRKYs under treatment of drought, salt, high pH and ABA using RT-qPCR. Note: The results of the RT-qPCR were calculated according to the 2−ΔΔCT method, and were normalized using CiEF1α as the reference gene, * means p < 0.05 and ** means p < 0.01, using Student’s t-test.
Ijms 24 09519 g008
Figure 9. Interaction network pattern of 24 differential CkWRKYs under drought, salt, high pH and ABA treatments using STRING (https://cn.string-db.org/cgi/ accessed on 1 January 2020).
Figure 9. Interaction network pattern of 24 differential CkWRKYs under drought, salt, high pH and ABA treatments using STRING (https://cn.string-db.org/cgi/ accessed on 1 January 2020).
Ijms 24 09519 g009
Table 1. The WRKYs identified from C. korshinskii.
Table 1. The WRKYs identified from C. korshinskii.
ChromosomeGene NameORF (aa)pIMW (kDa)Conserved
Heptapeptide
Zinc-Finger TypeDomain
Number
Group/
Subgroup
Ortholog in
A. thaliana
Chr1CkWRKY12925.9433.77WRKYGQKC2HC1IIIAtWRKY70
CkWRKY22327.1126.32WRKYGQKC2H21IIcAtWRKY12
CkWRKY33615.3740.77WRKYGQKC2HC1IIIAtWRKY46
CkWRKY43335.7236.37WRKYGQKC2H21IIeAtWRKY27
CkWRKY53369.6736.59WRKYGQKC2H21IIdAtWRKY15
CkWRKY65946.1264.39WRKYGQKC2H21IIbAtWRKY72
CkWRKY75077.6454.68WRKYGQK/WRKYGQKC2H22IAtWRKY32
CkWRKY83129.8133.88WRKYGQKC2H21IIdAtWRKY11
CkWRKY93918.9842.96WRKYGQKC2H21IIbAtWRKY42
CkWRKY103836.1541.86WRKYGQKC2H21IIcAtWRKY48
CkWRKY115045.5454.71WRKYGQKC2H21IIbAtWRKY42
Chr2CkWRKY122238.1624.85WRKYGQKC2H21IIaAtWRKY40
CkWRKY131445.4516.16GRKYGQKC2H21IIaAtWRKY60
CkWRKY147445.7780.62WRKYGQK/WRKYGQKC2H22IAtWRKY2
CkWRKY154826.0852.90WRKYGQKC2H21IIbAtWRKY6
CkWRKY165876.4863.72WRKYGQK/WRKYGQKC2H22IAtWRKY20
CkWRKY175067.6956.17WRKYGQK/WRKYGQKC2H22IAtWRKY33
CkWRKY183685.3541.30WRKYGQKC2HC1IIIAtWRKY41
CkWRKY193134.8835.23WRKYGQKC2HC1IIIAtWRKY70
CkWRKY202057.7223.28WRKYGKKC2H21IIcAtWRKY51
CkWRKY212508.8427.87WRKYGQKC2H21IIcAtWRKY13
CkWRKY222997.1133.66WRKYGQKC2H21IIaAtWRKY40
Chr3CkWRKY233076.0533.80WRKYGQKC2H21IIeAtWRKY27
CkWRKY243147.1735.06WRKYGQKC2H21IIcAtWRKY23
CkWRKY254298.947.18WRKYGQKC2H21IIbAtWRKY47
CkWRKY263509.6939.23WRKYGQKC2H21IIdAtWRKY21
CkWRKY273399.5136.64WRKYGQKC2H21IIdAtWRKY17
CkWRKY282608.1929.15WRKYGQKC2H21IIaAtWRKY40
CkWRKY292757.0330.97WRKYGQKC2H21IIaAtWRKY40
CkWRKY301755.4920.40WRKYGKK-1IIcAtWRKY50
CkWRKY311555.8118.30WRKYGKK-1IIcAtWRKY50
CkWRKY323369.5737.09WRKYGQKC2H21IIdAtWRKY7
CkWRKY331624.8217.80WRMYGQKC2H21IIeAtWRKY65
Chr4CkWRKY343615.3541.13WRKYGQKC2HC1IIIAtWRKY41
CkWRKY352845.2232.43WRKYGQKC2H21IIeAtWRKY27
CkWRKY362338.9626.64WRKYGQKC2H21IIcAtWRKY13
CkWRKY371866.6521.55WRKYGKKC2H21IIcAtWRKY51
CkWRKY382525.7728.30WRKYGQKC2H21IIeAtWRKY65
CkWRKY393336.1836.99WRKYGQKC2H21IIcAtWRKY71
CkWRKY404825.3953.20WRKYGQKC2H21IIbAtWRKY9
CkWRKY415585.6560.72WRKYGQKC2H21IIbAtWRKY6
CkWRKY423006.2533.56WRKYGQKC2HC1IIIAtWRKY70
CkWRKY433175.9435.47WRKYGQKC2HC1IIIAtWRKY70
CkWRKY443118.5934.92WRKYGQKC2H21IIaAtWRKY40
Chr5CkWRKY455127.0255.86WRKYGQK/WRKYGQKC2H22IAtWRKY4
CkWRKY465647.3262.55WRKYGQKC2H21IIbAtWRKY72
CkWRKY477415.5781.33WRKYGQK/WRKYGQKC2H22IAtWRKY20
CkWRKY483355.8436.67WRKYGQKC2H21IIeAtWRKY27
CkWRKY493076.0234.66WRKYGQKC2HC1IIIAtWRKY70
CkWRKY505636.362.42WRKYGQK/WRKYGQKC2H22IAtWRKY33
CkWRKY513225.235.79WRKYGQKC2H21IIcAtWRKY49
CkWRKY5213910.1415.44WRKYGQK-1IIcAtWRKY75
CkWRKY533189.6235.76WRKYGQKC2H21IIdAtWRKY21
CkWRKY544488.9150.16WRKYGQK/WRKYGQKC2H22IAtWRKY44
CkWRKY552654.8530.29WRKYGQKC2H21IIeAtWRKY35
CkWRKY562968.6233.14WRKYGQKC2H21IIcAtWRKY23
CkWRKY575186.8956.40WRKYGQKC2H21IIbAtWRKY42
Chr6CkWRKY583078.7534.15WRKYGQKC2H21IIaAtWRKY40
CkWRKY594696.2850.69WRKYGQKC2H21IIbAtWRKY6
CkWRKY602665.8728.92WRKYGQKC2H21IIeAtWRKY69
CkWRKY612995.5634.11WRKYGQKC2HC1IIIAtWRKY70
CkWRKY623135.5735.40WRKYGQKC2HC1IIIAtWRKY70
CkWRKY632847.5932.77WRKYGHKC2HC1IIIAtWRKY70
CkWRKY644567.5750.70WRKYGQKC2H21IIbAtWRKY42
CkWRKY653896.442.37WRKYGQKC2H21IIcAtWRKY48
CkWRKY664836.1153.23WRKYGQKC2H21IIeAtWRKY35
CkWRKY676106.3365.55WRKYGQKC2H21IIbAtWRKY6
CkWRKY682246.0224.51WRKYGQKC2H21IIeAtWRKY65
Chr7CkWRKY695506.0160.78WRKYGQK/WRKYGQKC2H22IAtWRKY1
CkWRKY705797.6762.70RRKYGQK/WRKYGQKC2H22IAtWRKY20
CkWRKY712648.1629.90WRKYGQKC2H21IIcAtWRKY71
CkWRKY725747.2163.30WRKYGQK/WRKYGQKC2H22IAtWRKY33
CkWRKY733219.3336.85WKKYEEKC2H21IAtWRKY33
CkWRKY743155.5535.15WRKYGQKC2HC1IIIAtWRKY70
CkWRKY753526.2738.56WRKYGQKC2HC1IIIAtWRKY55
CkWRKY762786.6130.74WRKYGQKC2H21IIcAtWRKY57
CkWRKY773877.7642.48WRKYGQK/WRKYGQKC2H22IAtWRKY3
CkWRKY783129.5334.35WRKYGQKC2H21IIdAtWRKY15
CkWRKY792528.6728.53WRKYGQKC2H21IIeAtWRKY29
CkWRKY802926.7133.07WRKYGKKC2HC1IIIAtWRKY41
Chr8CkWRKY815546.8360.54WRKYGQKC2H21IIbAtWRKY72
ScaffoldCkWRKY821866.0620.79WRKYGQKC2H21IIeAtWRKY27
CkWRKY835936.2665.67WRKYGQK/WRKYGQKC2H22IAtWRKY1
CkWRKY845876.4863.74WRKYGQK/WRKYGQKC2H22IAtWRKY20
CkWRKY855525.6459.82WRKYGQK/WRKYGQKC2H22IAtWRKY3
CkWRKY865176.9156.32WRKYGQKC2H21IIbAtWRKY47
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liu, J.; Li, G.; Wang, R.; Wang, G.; Wan, Y. Genome-Wide Analysis of WRKY Transcription Factors Involved in Abiotic Stress and ABA Response in Caragana korshinskii. Int. J. Mol. Sci. 2023, 24, 9519. https://doi.org/10.3390/ijms24119519

AMA Style

Liu J, Li G, Wang R, Wang G, Wan Y. Genome-Wide Analysis of WRKY Transcription Factors Involved in Abiotic Stress and ABA Response in Caragana korshinskii. International Journal of Molecular Sciences. 2023; 24(11):9519. https://doi.org/10.3390/ijms24119519

Chicago/Turabian Style

Liu, Jinhua, Guojing Li, Ruigang Wang, Guangxia Wang, and Yongqing Wan. 2023. "Genome-Wide Analysis of WRKY Transcription Factors Involved in Abiotic Stress and ABA Response in Caragana korshinskii" International Journal of Molecular Sciences 24, no. 11: 9519. https://doi.org/10.3390/ijms24119519

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