Next Article in Journal
Efficiency of Wheat Straw Biochar in Combination with Compost and Biogas Slurry for Enhancing Nutritional Status and Productivity of Soil and Plant
Next Article in Special Issue
Responses of Upland Cotton (Gossypium hirsutum L.) Lines to Irrigated and Rainfed Conditions of Texas High Plains
Previous Article in Journal
Effect of Zinc Priming on Salt Response of Wheat Seedlings: Relieving or Worsening?
Previous Article in Special Issue
The Influence of Soil Acidity on the Physiological Responses of Two Bread Wheat Cultivars
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Function and Mechanism of WRKY Transcription Factors in Abiotic Stress Responses of Plants

College of Horticulture and Plant Protection, Yangzhou University, Yangzhou 225009, China
*
Author to whom correspondence should be addressed.
Contributed equally to this work.
Plants 2020, 9(11), 1515; https://doi.org/10.3390/plants9111515
Submission received: 26 September 2020 / Revised: 30 October 2020 / Accepted: 4 November 2020 / Published: 8 November 2020
(This article belongs to the Special Issue Responses of Plants to Environmental Stresses)

Abstract

:
The WRKY gene family is a plant-specific transcription factor (TF) group, playing important roles in many different response pathways of diverse abiotic stresses (drought, saline, alkali, temperature, and ultraviolet radiation, and so forth). In recent years, many studies have explored the role and mechanism of WRKY family members from model plants to agricultural crops and other species. Abiotic stress adversely affects the growth and development of plants. Thus, a review of WRKY with stress responses is important to increase our understanding of abiotic stress responses in plants. Here, we summarize the structural characteristics and regulatory mechanism of WRKY transcription factors and their responses to abiotic stress. We also discuss current issues and future perspectives of WRKY transcription factor research.

Graphical Abstract

1. Introduction

As a fixed-growth organism, plants are exposed to a variety of environmental conditions and may encounter many abiotic stresses, for example, drought, waterlogging, heat, cold, salinity, and Ultraviolet-B (UV-B) radiation. To adapt and counteract the effects of such abiotic stresses, plants have evolved several molecular mechanisms involving signal transduction and gene expression [1,2]. Transcription factors (TFs) are important regulators involved in the process of signal transduction and gene expression regulation under environmental stresses. TFs can be combined with cis-acting elements to regulate the transcriptional efficiency of target genes by inhibiting or enhancing their transcription [3,4]. Accordingly, plants may show corresponding responses to external stresses via TFs regulating target genes. Although some TF families (MYB, bZIP, AP2/EREBP, NAC) are associated with adversity [2,5], WRKY is the most extensively studied TF family in plant stress responses.
The WRKY family is a unique TF superfamily of higher plants and algae, which play important roles in many life processes, particularly in response against biotic and abiotic stress [6,7]. In 1994, the SWEET POTATO FACTOR1 (SPF1) gene of the WRKY family was first found in Impoea batatas [8]. Later, ABF1 and ABF2 were found in wild Avena sativa, and showed regulatory roles in seed germination [9]. A previous study successively cloned WRKY1, WRKY2, and WRKY3 from Petroselinum crispum, named the WRKY TF, and proved for the first time that WRKY protein can regulate plant responses to pathogens [10]. With an increase in available published genomes, many members of the WRKY TF family have been identified in various species, including 104 from Populus [11], 37 from Physcomitrella patens [12], 45 from Hordeum vulgare [13], 55 from Cucumis sativus [14], 74 from Arabidopsis thaliana [15], 83 from Pinus monticola [16], 81 from Solanum lycopersicum [17], and 102 from Oryza sativa [18]. WRKY TFs exist as gene families in plants, and the number of WRKY TFs varies among species. In plants exposed to abiotic stresses (salt, drought, temperature, and so forth), WRKY family members play important roles in diverse stress responses. In addition, these TFs affect the growth and development of plants [19,20]. Therefore, WRKY TFs have attracted broad attention. Although some reviews on WRKY TFs are available, in this review we focus on the structural characteristics and regulatory mechanisms of WRKY TFs and summarize recent progress in understanding the roles of WRKY TFs during exposure to abiotic stresses such as drought, temperature, salt, and UV radiation.

2. Structural Characteristics of WRKY TFs

The WRKY structure consists of two parts: the N-terminal DNA binding domain and the C-terminal zinc-finger structure [21]. The DNA binding domain sequence of WRKY is based on the heptapeptide WRKYGQK (Figure 1), but there are some differences, such as WRKYGQK, WRKYGKK, WRKYGMK, WSKYGQK, WKRYGQK, WVKYGQK, and WKKYGQK [17,22]. Zinc-finger structures mainly include C2H2 type and C2HC type [23], whereas some exist in the form of CX29HXH and CX7CX24HXC [17] (Figure 1). According to the number of WRKY domains and the structure of their zinc-finger motifs, WRKY can be divided into groups I, II, and III [23] (Figure 1). Group I usually contains two WRKY domains and one C2H2 zinc-finger structure. Those in group II and group III contain only one WRKY domain. The difference is that the zinc-finger structure in group II is C2H2 and that in group III is C2HC [19,21,23] (Figure 1). According to the phylogenetic relationship of the amino acid sequence of the primary structure, group II can be further divided into subgroups a–e [7,23,24]. Evolutionary analyses have shown that the WRKY of group II is not generally a single source, mainly including five categories I, IIa + IIb, IIc, IId + IIe, and III [7,24]. In addition, some WRKY proteins contain a glutamate enrichment domain, a proline enrichment domain, and a leucine zipper structure [25].

3. Regulatory Mechanism of WRKY TFs

WRKY family members have diverse regulatory mechanisms. Briefly, WRKY protein can be effectively combined with W-box elements to activate or inhibit the transcription of downstream target genes. Moreover, it can also bind other acting elements to form protein complexes, which enhances the activity of transcription binding [21].
WRKY TFs can effectively activate the expression of downstream genes by binding conserved W-box cis-acting elements in the downstream gene promoter region [21,26]. There are abundant W-box elements in the self-promoter of most WRKY TFs. Therefore, these WRKY TFs can bind with their own promoters to achieve self-regulation or cross-regulation networks by combining with other WRKY TFs [27]. For example, CaWRKY6 of Capsicum frutescens can activate CaWRKY40 and make the plant more tolerant to high temperature and humidity. Glycine max GmWRKY27 not only inhibits the activity of downstream GmNAC29 promoter by independent inhibition, but also cooperatively interacts with GmMYB174 to inhibit the expression of GmNAC29, thereby increasing drought and salt stress resistances [28]. Moreover, chromatin immunoprecipitation (ChIP) studies have shown that when Petroselinum crispum is infected by pathogenic bacteria, PcWRKY1 promoter can effectively bind to itself and the W-box of PcWRKY3 promoter, and transcriptional activation can be achieved through self-negative feedback regulation and cross-regulation with other WRKY proteins [29]. In addition, WRKY TFs can interact with non-W-box elements. For example, Oryza sativa OsWRKY13 can interact with PRE4 (TGCGCTT) elements [30]. Hordeum vulgare HvWRKY46 and Nicotiana tabacum NtWRKY12 can effectively combine with the sucrose response element SURE [31,32]. These results indicate that there are multiple binding modes between WRKY TFs and structural genes. Different binding patterns and preferences of binding sites allow for the regulation of downstream target genes, providing WRKY TFs with versatile functions in the plant transcriptional regulation network.

4. WRKY TF Involved in Abiotic Stress Responses

When plants sense stress, the corresponding signaling is activated and transferred to the cell interior. Reactive oxygen species (ROS) and Ca2+ ions are usually exchanged as the signal transduction in the cell. Protein kinases such as MPKs are subsequently activated to regulate the activities of related TFs. Consequently, the plant presents a stress response [31,32]. In response to abiotic stresses, some WRKY TFs can be rapidly differentially expressed, promoting signal transduction and regulating the expression of related genes [33]. The expression patterns and functional identifications of WRKYs in most studies are generally based on transcriptome analyses, real-time fluorescence quantitative PCR, gene chip analyses, and genetic transformation. Hence, WRKY genes can function effectively in most abiotic stress responses or tolerances in various plants (Table 1, Figure 2).

4.1. WRKY TFs and Drought Stress

Drought has a major impact on plant growth and development, resulting in a significant decrease in grain and other types of crop yield [77]. Under drought stress, drought-tolerant plants can accumulate oligosaccharides through sucrose metabolism to improve drought resistance. For example, when Arabidopsis is subjected to drought stress, the expression of AtWRKY53 combined with the Qua-Quine Starch (QQS) promoter sequence is rapidly induced, hydrogen peroxide content is reduced, and the glucose metabolism pathway is significantly enhanced, thereby regulating stomatal opening and ultimately affecting drought tolerance [37]. In Boea hygrometrica, BhWRKY1 effectively regulates the expression of the BhGolS1 gene, and the overexpression of BhGolS1 and BhWRKY1 induces the accumulation of raffinose family oligosaccharides (RFOs) in transgenic Nicotiana tabacum, thus improving the ability of seedlings to resist drought [60].
WRKY protein can directly regulate the expression of drought-resistant genes. For example, in sorghum, SbWRKY30 regulates the drought stress response gene SbRD19 by binding with W-box elements of the SbRD19 promoter, and protects plant cells from the damage of reactive oxygen species by improving ROS scavenging capability, enhancing drought tolerance [66]. TaWRKY2 of wheat can bind to STZ and downstream drought-resistant gene RD29B promoter, with a positive regulatory effect on the expression of RD29B [58]. DREB2A regulates the expression of dehydration stress-related genes [78], while TaWRKY19 can bind to DREB2A promoter, ultimately activating the expression of RD29A, RD29B, and Cor6.6 in transgenic Arabidopsis plants [58]. Similarly, Arabidopsis AtWRKY57 positively regulates the expression of RD29A and NCED3 genes by binding their W-box elements in the promoter regions [39]. In addition, WRKY protein can act on other TFs to play regulatory roles in drought tolerance. For example, TcWRKY53 of Thlaspi arvense significantly inhibits the expression of NtERF5 and NterEBp-1 of the AP2/ERF TF family, thus improving plant resistance to drought stress [63].
WRKY TFs also regulate plant tolerance through abscisic acid (ABA) and ROS-related signaling pathways. During drought stress, higher ABA levels were accumulated in plants, and leaf stomata were closed to reduce transpiration rate, thus regulating water balance in plants. ABA accumulation in cells, integrated with a variety of stress signals, regulates the expression of downstream genes, consequently sensing and responding to the adverse environment [40]. Arabidopsis AtWRKY63 has a specific effect on ABA-mediated stomatal closure and other signal transduction pathways, thus affecting the drought response [40]. GhWRKY21 regulates ABA-mediated cotton drought tolerance by promoting the expression of GhHAB [43]. Overexpression of BdWRKY36 in tobacco reduces the accumulation of ROS, activated NtLEA5, NtNCED1, and NtDREB3 in the ABA biosynthetic pathway, and significantly enhances the drought resistance of plants [48]. In Solanum lycopersicum, SlWRKY81 increases the drought tolerance of plants by inhibiting the accumulation of H2O2, playing a negative regulation role of stomatal closure [72].

4.2. WRKY TFs and Salt Stress

Salt stress is an important abiotic stress affecting crop productivity, particularly in arid and semiarid regions. WRKY TFs play essential roles in regulating the response to salt stress. To date, a total of 47 WRKY genes have been found to be expressed under salt stress in the wheat genome [79]. STZ is a protein related to ZPT2, which acts as a transcriptional inhibitor to downregulate the deactivation of other transcription factors. GmWRKY54 of Glycine max inhibits STZ expression and responds to salt stress by positively regulating the DREB2A-mediated pathway [55]. FcWRKY70 promotes the upregulation of arginine decarboxylase (ADC) expression, which is heterologously expressed in tobacco, and the content of lemon putrescine is significantly increased, thus enhancing the salt tolerance of plants [49]. The IbWRKY47 gene positively regulates stress resistance-related genes and significantly improves the salt tolerance of Ipomoea batatas [68]. MiR156/SPL modulates salt tolerance by upregulation of Malus domestica salt tolerance gene MdWRKY100 [71]. In Sorghum bicolor, SbWRKY50 could directly bind to the upstream promoter of SOS1 and HKT1 and participate in plant salt response by controlling ion homeostasis [67]. In addition, some WRKY genes function as negative regulation factors involved in salt stress resistance. Arabidopsis RPD3-like histone deacetylase HDA9 inhibits salt stress tolerance by regulating the DNA binding and transcriptional activity of WRKY53 [38]. Chrysanthemum CmWRKY17 overexpressed in Arabidopsis allows the plants to be more sensitive to salt stress. The expression level of stress resistance-related genes in transgenic Arabidopsis is lower than that in wild-type plants, indicating that CmWRKY17 may be involved in negatively regulating the salt stress response in Chrysanthemum [80]. The expression of GhWRKY68 is strongly induced in upland cotton and decreases salt tolerance [45]. In contrast, a high expression level of GhWRKY25 enhances the salt tolerance of upland cotton, while transgenic tobacco shows a relatively weaker tolerance to drought stress [44], indicating that the regulatory effects of different WRKY TFs involved in drought response are different.
Plants can also respond to saline–alkali stress through ABA, H2O2, and other signal pathways. In Glycine max, the negative regulatory factor ABI1 in the ABA pathway may be the downstream target gene of GmWRKY13. Genetic transformation experiments in Arabidopsis have shown that overexpression of GmWRKY13 significantly increases the expression of ABI1, but plants show a low tolerance to salt stress [55]. Overexpression of ZmWRKY17 has an inhibitory effect on the sensitivity of exogenous ABA treatment, resulting in a relatively lower tolerance to high levels of salinity [57]. Under salt-induced H2O2 and cytosolic Ca2+ stimulation, the activity of antioxidant enzymes increases, thus improving the tolerance to high-salinity environments [81]. ABA-induced WRKY gene expression is largely related to salt stress. Exogenous application of ABA and NaCl also induce AtWRKY25 and AtWRKY33 in Arabidopsis [33], OsWRKY72 in rice [51], GbWRKY1 in Verbena bonariensis [73], and VpWRKY1/2 [61] and VpWRKY3 [62] in grape.

4.3. WRKY TFs and Temperature Stress

Both low- and high-temperature stress can reduce crop yield and quality in plants. WRKY TFs play a role in the stress response through different signal transduction pathways. For example, in Verbena bonariensis, VbWRKY32 as a positive regulator, upregulates the transcriptional level of cold response genes, which increases the antioxidant activity, maintains membrane stability, and enhances osmotic regulation ability, thereby improving the survival ability under cold stress [74]. The BcWRKY46 gene of Brassica campestris is strongly induced by low temperature and ABA, activating related genes in the ABA signaling pathway to improve the low-temperature tolerance of plants [59]. CBF TFs regulate the expression of COR, and the overexpressed transgenic lines of CBF1, CBF2, and CBF3 show stronger cold resistance [82]. AtWRKY34 has a negative regulatory effect on the CBF-mediated cold response pathway; it is specifically expressed in mature pollen grains after exposure to low temperatures, resulting in resistance to low temperatures [35]. In addition, plants respond to temperature changes by coordinating organ development in an adverse environment. At low temperatures, rice MADS-Box TF OsMADS57 and its interacting protein OsTB1 synergistically activate the transcriptional regulation of OsWRKY94, preventing tillering by inhibiting transcription of the organ development gene D14 [83].
Due to global climate change, high-temperature stress has attracted significant attention. There is evidence that, to a certain extent, high temperatures will lead to biochemical changes in plants [84]. Thermal stimulation can activate Ca2+ channels to maintain a higher intracellular Ca2+ concentration, thereby activating calmodulin protein expression and inducing thermal-shock protein transcriptional expression [85]. In Arabidopsis, AtWRKY54 significantly responds to heat shock whereas basic leucine zipper factors (bZIPs) respond to prolonged warming [41]. Overexpression of AtWRKY39 can make plants more heat-sensitive. AtWRKY39 is highly homologous to AtWRKY7, and both of them can effectively bind calmodulin in plants, indicating a similar function [36]. In addition, AtWRKY25, AtWRKY26, and AtWRKY33 can improve tolerance to high-temperature stress in transgenic Arabidopsis by regulating the Hsp101 and Zat10 genes [34]. Plants subjected to heat stress can also activate the oxidative stress response through ethylene [86]. Under high-temperature stress, the expressions of AtWRKY25, AtWRKY26, and AtWRKY33 in Arabidopsis are induced by ethylene, the feedback factor EIN2 is transcriptionally regulated, and the effective activation of ethylene signal transduction contribute to relatively stronger heat resistance. In Oryza sativa, HSP101 promoter can activate the expression of the OsWRKY11 gene. Under heat treatment, the leaves wilted more slowly and the green part of the plant was less damaged, which makes it more heat-resistant [50]. In addition, some noncoding RNAs, such as miR396, play a role in the response to heat stress by regulating its target WRKY6 [87].

4.4. WRKY TFs and Other Abiotic Stresses

WRKY TFs are also involved in oxidative stress, mechanical damage, UV radiation, and other abiotic stresses (Figure 3). FcWRKY40 overexpression can significantly enhance the resistance of transgenic tobacco to oxidative stress [88]. When Arabidopsis is treated with ROS, the expressions of AtWRKY30, AtWRKY40, AtWRKY75, AtWRKY6, AtWRKY26, and AtWRKY45 are significantly upregulated [89]. After mechanical injury, the expression levels of AtWRKY11, AtWRKY15, AtWRKY22, AtWRKY33, AtWRKY40, AtWRKY53 [90] and AtWRKY6 [64] are upregulated. Similarly, NaWRKY3 is strongly expressed in tobacco. By contrast, the sensitivity of transgenic plants is increased when NaWRKY3 is knocked out [64]. In two previous studies, UV-B radiation treatment induced three WRKY genes in Arabidopsis and the OsWRKY89 gene in rice, resulting in a thick waxy substance on the leaf surface and improved tolerance to heat [54,91].
In addition, a single WRKY TF can play multiple roles in different stress responses via various signal pathways and regulatory networks. For example, TaWRKY44 expression in tobacco can improve resistance to drought, salt stress, and osmotic stress [92], while PgWRKY62 and PgWRKY33 in Pennisetum glaucum respond to salt and drought simultaneously [75]. BhWRKY1 protein in Boea hygrometrica binds to the promoter of BhGolS1 and is associated with both low-temperature resistance and drought tolerance [60]. IbWRKY2 can interact with IbVQ4, and drought and salt treatment can induce the expression of IbVQ4, thus improving the tolerance of plants to drought and salt stress [69]. MdWRKY30 overexpression enhances tolerance to salt and osmotic stress in transgenic apple callus through transcriptional regulation of stress-related genes [70]. PagWRKY75 negatively regulates the tolerance of 84 K poplar (Populus alba × P. glandulosa) to salt and osmotic stress by reducing the scavenging capacity of ROS and the accumulation of proline, thus actively regulates the rate of leaf water loss [76].

5. Conclusions and Perspectives

As one of the largest TF families, WRKY plays an important and indispensable role in normal life activities of plants. Over the years, it has been shown that WRKY TFs not only participate in plant growth and development, but also show complex regulatory mechanisms and networks involved in external abiotic stresses. A large number of WRKYs have been functionally characterized in model plants, providing abundant functional references for other plants. Given that crops usually face various stresses and WRKYs play important roles in stress responses, further in-depth studies on WRKY genes in more crops are required. As increasing plant genomes have been sequenced, particularly of economically important crops, the genome-wide identification of WRKY genes will facilitate screening for stress resistance-related functional genes in plants. Moreover, previous studies of WRKY gene functions were largely dependent on transcriptomics and functional predictions, whereas more applications of genetic verification combined with new technologies are accelerating the research progress of WRKY’s novel functions. In addition, characterization of the downstream genes regulated by WRKY TFs or WRKY TF self-regulation will help clarify the regulatory network of abiotic stress responses. Furthermore, noncoding RNAs and epigenetic modifications involved in the regulation of WRKY TFs should be explored in future studies. Ultimately, using WRKY TFs to screen for stress-resistant plant cultivars and improve plant stress resistance will significantly benefit agricultural crop yield and quality in the context of aggravated climate change.

Author Contributions

Conceptualization, W.L. and B.J.; methodology, B.J.; software, S.P.; formal analysis, S.P. and Z.L.; investigation, S.P. and Z.L.; resources, S.P.; writing—original draft preparation, S.P. and W.L.; writing—review and editing, B.J., S.P., Z.L. and W.L.; visualization, B.J. and S.P.; supervision, B.J. and W.L.; project administration, W.L.; funding acquisition, W.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was founded by the Natural Science Foundation of China (grant numbers 31971408, 31670695 and 31971686), and Forestry Technology Innovation and Extension of Jiangsu Province (grant number LYKJ [2019]48).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yoon, Y.; Seo, D.H.; Shin, H.; Kim, H.J.; Kim, C.M.; Jang, G. The role of stress-responsive transcription factors in modulating abiotic stress tolerance in plants. Agronomy 2020, 10, 788. [Google Scholar] [CrossRef]
  2. Ma, Q.; Xia, Z.; Cai, Z.; Li, L.; Cheng, Y.; Liu, J.; Nian, H. GmWRKY16 enhances drought and salt tolerance through an ABA-mediated pathway in Arabidopsis thaliana. Front. Plant Sci. 2019, 9, 9. [Google Scholar] [CrossRef] [Green Version]
  3. Liu, Y.; Yang, T.; Lin, Z.; Guo, B.; Xing, C.; Zhao, L.; Dong, H.; Gao, J.; Xie, Z.; Zhang, S.-L.; et al. A WRKY transcription factor PbrWRKY53 from Pyrus betulaefolia is involved in drought tolerance and AsA accumulation. Plant Biotechnol. J. 2019, 17, 1770–1787. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Shrestha, A.; Khan, A.; Dey, N. cis-trans Engineering: Advances and perspectives on customized transcriptional regulation in plants. Mol. Plant 2018, 11, 886–898. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Yamasaki, K.; Kigawa, T.; Inoue, M.; Watanabe, S.; Tateno, M.; Seki, M.; Shinozaki, K.; Yokoyama, S. Structures and evolutionary origins of plant-specific transcription factor DNA-binding domains. Plant Physiol. Biochem. 2008, 46, 394–401. [Google Scholar] [CrossRef] [PubMed]
  6. Zhang, Y.; Wang, L. The WRKY transcription factor superfamily: Its origin in eukaryotes and expansion in plants. BMC Evol. Biol. 2005, 5, 1. [Google Scholar] [CrossRef] [Green Version]
  7. Pandey, S.P.; Somssich, I.E. The role of WRKY transcription factors in plant immunity. Plant Physiol. 2009, 150, 1648–1655. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. 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 beta-amylase from sweet potato. Mol Gen Genet 1994, 244, 563–571. [Google Scholar] [CrossRef]
  9. Rushton, P.J.; Macdonald, H.; Huttly, A.K.; Lazarus, C.M.; Hooley, R. Members of a new family of DNA-binding proteins bind to a conserved cis-element in the promoters of Amy2 genes. Plant Mol. Biol. 1995, 29, 691–702. [Google Scholar] [CrossRef]
  10. Rushton, P.J.; Torres, J.T.; Parniske, M.; Wernert, P.; Hahlbrock, K.; Somssich, I.E. Interaction of elicitor-induced DNA-binding proteins with elicitor response elements in the promoters of parsley PR1 genes. EMBO J. 1996, 15, 5690–5700. [Google Scholar] [CrossRef] [PubMed]
  11. He, H.; Dong, Q.; Shao, Y.; Jiang, H.; Zhu, S.; Cheng, B.; Xiang, Y. Genome-wide survey and characterization of the WRKY gene family in Populus trichocarpa. Plant Cell Rep. 2012, 31, 1199–1217. [Google Scholar] [CrossRef] [PubMed]
  12. Rensing, S.A.; Lang, D.; Zimmer, A.D.; Terry, A.; Salamov, A.A.; Shapiro, H.; Nishiyama, T.; Perroud, P.-F.; Lindquist, E.; Kamisugi, Y.; et al. The Physcomitrella genome reveals evolutionary insights into the conquest of land by plants. Science 2007, 319, 64–69. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Mangelsen, E.; Kilian, J.; Berendzen, K.W.; Kolukisaoglu, Ü.; Harter, K.; Jansson, C.; Wanke, D. Phylogenetic and comparative gene expression analysis of barley (Hordeum vulgare) WRKY transcription factor family reveals putatively retained functions between monocots and dicots. BMC Genom. 2008, 9, 194. [Google Scholar] [CrossRef] [Green Version]
  14. Ling, J.; Jiang, W.; Zhang, Y.; Yu, H.; Mao, Z.; Gu, X.; Huang, S.; Xie, B. Genome-wide analysis of WRKY gene family in Cucumis sativus. BMC Genom. 2011, 12, 1–20. [Google Scholar] [CrossRef] [Green Version]
  15. Berri, S.; Abbruscato, P.; Faivre-Rampant, O.; Brasileiro, A.C.M.; Fumasoni, I.; Satoh, K.; Kikuchi, S.; Mizzi, L.; Morandini, P.; Pè, M.E.; et al. Characterization of WRKY co-regulatory networks in rice and Arabidopsis. BMC Plant Biol. 2009, 9, 1–22. [Google Scholar] [CrossRef] [Green Version]
  16. Liu, J.-J.; Ekramoddoullah, A.K. Identification and characterization of the WRKY transcription factor family in Pinus monticola. Genome 2009, 52, 77–88. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Huang, S.; Gao, Y.; Liu, J.; Peng, X.; Niu, X.; Fei, Z.; Cao, S.; Liu, Y. Genome-wide analysis of WRKY transcription factors in Solanum lycopersicum. Mol. Genet. Genom. 2012, 287, 495–513. [Google Scholar] [CrossRef]
  18. Ross, C.A.; Liu, Y.; Shen, Q.J. The WRKY gene family in rice (Oryza sativa). J. Integr. Plant Biol. 2007, 49, 827–842. [Google Scholar] [CrossRef]
  19. Rushton, P.J.; Somssich, I.E.; Ringler, P.; Shen, Q.J. WRKY transcription factors. Trends Plant Sci. 2010, 15, 247–258. [Google Scholar] [CrossRef]
  20. Ülker, B.; Somssich, E.I. WRKY transcription factors: From DNA binding towards biological function. Curr. Opin. Plant Biol. 2004, 7, 491–498. [Google Scholar] [CrossRef] [Green Version]
  21. Phukan, U.J.; Jeena, G.S.; Shukla, R.K. WRKY transcription factors: Molecular regulation and stress responses in plants. Front. Plant Sci. 2016, 7, 760. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Xie, Z.; Zhang, Z.-L.; Zou, X.; Huang, J.; Ruas, P.; Thompson, D.; Shen, Q.J. Annotations and functional analyses of the rice WRKY gene superfamily reveal positive and negative regulators of abscisic acid signaling in aleurone cells. Plant Physiol. 2004, 137, 176–189. [Google Scholar] [CrossRef] [Green Version]
  23. Eulgem, T.; Rushton, P.J.; Robatzek, S.; Somssich, I.E. The WRKY superfamily of plant transcription factors. Trends Plant Sci. 2000, 5, 199–206. [Google Scholar] [CrossRef]
  24. Rushton, P.J.; Bokowiec, M.T.; Han, S.; Zhang, H.; Brannock, J.F.; Chen, X.; Laudeman, T.W.; Timko, M.P. Tobacco transcription factors: Novel insights into transcriptional regulation in the Solanaceae. Plant Physiol. 2008, 147, 280–295. [Google Scholar] [CrossRef] [Green Version]
  25. 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 BBA Bioenerg. 2012, 1819, 120–128. [Google Scholar] [CrossRef]
  26. Ciolkowski, I.; Wanke, D.; Birkenbihl, R.P.; Somssich, I.E. Studies on DNA-binding selectivity of WRKY transcription factors lend structural clues into WRKY-domain function. Plant Mol. Biol. 2008, 68, 81–92. [Google Scholar] [CrossRef] [Green Version]
  27. Ezentgraf, U.; Laun, T.; Miao, Y. The complex regulation of WRKY53 during leaf senescence of Arabidopsis thaliana. Eur. J. Cell Biol. 2010, 89, 133–137. [Google Scholar] [CrossRef]
  28. Cai, H.; Yang, S.; Yan, Y.; Xiao, Z.; Cheng, J.; Wu, J.; Qiu, A.; Lai, Y.; Mou, S.; Guan, D.; et al. CaWRKY6 transcriptionally activates CaWRKY40, regulates Ralstonia solanacearum resistance, and confers high-temperature and high-humidity tolerance in pepper. J. Exp. Bot. 2015, 66, 3163–3174. [Google Scholar] [CrossRef] [Green Version]
  29. Turck, F.; Zhou, A.; Somssich, I.E. Stimulus-dependent, promoter-specific binding of transcription factor WRKY1 to Its native promoter and the defense-related gene PcPR1-1 in Parsley. Plant Cell 2004, 16, 2573–2585. [Google Scholar] [CrossRef] [Green Version]
  30. Cai, T.; Flanagan, L.B.; Jassal, R.S.; Black, T.A. Modelling environmental controls on ecosystem photosynthesis and the carbon isotope composition of ecosystem-respired CO2 in a coastal Douglas-fir forest. Plant Cell Environ. 2008, 31, 435–453. [Google Scholar] [CrossRef]
  31. Grierson, C.; Du, J.-S.; Zabala, M.D.T.; Beggs, K.; Smith, C.; Holdsworth, M.; Bevan, M.W. Separate cis sequences and trans factors direct metabolic and developmental regulation of a potato tuber storage protein gene. Plant J. 1994, 5, 815–826. [Google Scholar] [CrossRef] [PubMed]
  32. Sun, C.; Palmqvist, S.; Olsson, H.; Borén, M.; Ahlandsberg, S.; Jansson, C. A novel WRKY transcription factor, SUSIBA2, participates in sugar signaling in barley by binding to the sugar-responsive elements of the iso1 promoter. Plant Cell 2003, 15, 2076–2092. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Jiang, Y.; Deyholos, M.K. Functional characterization of Arabidopsis NaCl-inducible WRKY25 and WRKY33 transcription factors in abiotic stresses. Plant Mol. Biol. 2008, 69, 91–105. [Google Scholar] [CrossRef]
  34. Li, S.; Fu, Q.; Chen, L.; Huang, W.-D.; Yu, D. Arabidopsis thaliana WRKY25, WRKY26, and WRKY33 coordinate induction of plant thermotolerance. Planta 2011, 233, 1237–1252. [Google Scholar] [CrossRef]
  35. Zou, C.; Jiang, W.; Yu, D. Male gametophyte-specific WRKY34 transcription factor mediates cold sensitivity of mature pollen in Arabidopsis. J. Exp. Bot. 2010, 61, 3901–3914. [Google Scholar] [CrossRef] [Green Version]
  36. 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]
  37. Sun, Y.; Yu, D. Activated expression of AtWRKY53 negatively regulates drought tolerance by mediating stomatal movement. Plant Cell Rep. 2015, 34, 1295–1306. [Google Scholar] [CrossRef] [PubMed]
  38. Zheng, Y.; Ge, J.; Bao, C.; Chang, W.; Liu, J.; Shao, J.; Liu, X.; Su, L.; Pan, L.; Zhou, D.-X. Histone deacetylase HDA9 and WRKY53 transcription factor are mutual antagonists in regulation of plant stress response. Mol. Plant 2020, 13, 598–611. [Google Scholar] [CrossRef]
  39. Jiang, Y.; Liang, G.; Yu, D. Activated expression of WRKY57 confers drought tolerance in Arabidopsis. Mol. Plant 2012, 5, 1375–1388. [Google Scholar] [CrossRef] [Green Version]
  40. Ren, X.; Chen, Z.; Liu, Y.; Zhang, H.; Zhang, M.; Liu, Q.; Hong, X.; Zhu, J.-K.; Gong, Z. ABO3, a WRKY transcription factor, mediates plant responses to abscisic acid and drought tolerance in Arabidopsis. Plant J. 2010, 63, 417–429. [Google Scholar] [CrossRef] [Green Version]
  41. Wang, L.; Ma, K.-B.; Lu, Z.-G.; Ren, S.-X.; Jiang, H.-R.; Cui, J.-W.; Chen, G.; Teng, N.-J.; Lam, H.-M.; Jin, B. Differential physiological, transcriptomic and metabolomic responses of Arabidopsis leaves under prolonged warming and heat shock. BMC Plant Biol. 2020, 20, 1–15. [Google Scholar] [CrossRef]
  42. Ren, S.; Ma, K.; Lu, Z.; Chen, G.; Cui, J.; Tong, P.; Wang, L.; Teng, N.; Jin, B. Transcriptomic and metabolomic analysis of the heat-stress response of Populus tomentosa. Carr. For. 2019, 10, 383. [Google Scholar] [CrossRef] [Green Version]
  43. Wang, J.; Wang, L.; Yan, Y.; Zhang, S.; Li, H.; Gao, Z.; Wang, C.; Guo, X. GhWRKY21 regulates ABA-mediated drought tolerance by fine-tuning the expression of GhHAB in cotton. Plant Cell Rep. 2020, 39. [Google Scholar] [CrossRef]
  44. 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 2015, 253, 1265–1281. [Google Scholar] [CrossRef]
  45. Jia, H.; Wang, C.; Wang, F.; Liu, S.; Li, G.; Guo, X. GhWRKY68 reduces resistance to salt and drought in transgenic Nicotiana benthamiana. PLoS ONE 2015, 10, e0120646. [Google Scholar] [CrossRef] [Green Version]
  46. Wang, M.; Vannozzi, A.; Wang, G.; Liang, Y.-H.; Tornielli, G.B.; Zenoni, S.; Cavallini, E.; Pezzotti, M.; Cheng, Z.-M. Genome and transcriptome analysis of the grapevine (Vitis vinifera L.) WRKY gene family. Hortic. Res. 2014, 1, 14016. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Dang, F.F.; Wang, Y.N.; Yu, L.; Eulgem, T.; Lai, Y.; Liu, Z.Q.; Wang, X.; Qiu, A.L.; Zhang, T.X.; Lin, J.; et al. CaWRKY40, a WRKY protein of pepper, plays an important role in the regulation of tolerance to heat stress and resistance to Ralstonia solanacearum infection. Plant Cell Environ. 2013, 36, 757–774. [Google Scholar] [CrossRef]
  48. Sun, J.; Hu, W.; Zhou, R.; Wang, L.; Wang, X.; Wang, Q.; Feng, Z.-J.; Yu, H.; Qiu, D.; He, G.; et al. The Brachypodium distachyon BdWRKY36 gene confers tolerance to drought stress in transgenic tobacco plants. Plant Cell Rep. 2014, 34, 23–35. [Google Scholar] [CrossRef]
  49. Gong, X.; Zhang, J.; Hu, J.; Wang, W.; Wu, H.; Zhang, Q.; Liu, J.-H. FcWRKY70, a WRKY protein of Fortunella crassifolia, functions in drought tolerance and modulates putrescine synthesis by regulating arginine decarboxylase gene. Plant Cell Environ. 2015, 38, 2248–2262. [Google Scholar] [CrossRef]
  50. Wu, X.; Shiroto, Y.; Kishitani, S.; Ito, Y.; Toriyama, K. Enhanced heat and drought tolerance in transgenic rice seedlings overexpressing OsWRKY11 under the control of HSP101 promoter. Plant Cell Rep. 2009, 28, 21–30. [Google Scholar] [CrossRef] [PubMed]
  51. Song, Y.; Chen, L.; Zhang, L.; Yu, D. Overexpression of OsWRKY72 gene interferes in the abscisic acid signal and auxin transport pathway of Arabidopsis. J. Biosci. 2010, 35, 459–471. [Google Scholar] [CrossRef]
  52. Dai, X.; Wang, Y.; Zhang, W.-H. OsWRKY74, a WRKY transcription factor, modulates tolerance to phosphate starvation in rice. J. Exp. Bot. 2016, 67, 947–960. [Google Scholar] [CrossRef] [Green Version]
  53. Yokotani, N.; Sato, Y.; Tanabe, S.; Chujo, T.; Shimizu, T.; Okada, K.; Yamane, H.; Shimono, M.; Sugano, S.; Takatsuji, H.; et al. WRKY76 is a rice transcriptional repressor playing opposite roles in blast disease resistance and cold stress tolerance. J. Exp. Bot. 2013, 64, 5085–5097. [Google Scholar] [CrossRef]
  54. Wang, H.; Hao, J.; Chen, X.; Hao, Z.; Wang, X.; Lou, Y.; Peng, Y.; Guo, Z. Overexpression of rice WRKY89 enhances ultraviolet B tolerance and disease resistance in rice plants. Plant Mol. Biol. 2007, 65, 799–815. [Google Scholar] [CrossRef]
  55. Zhou, Q.-Y.; Tian, A.-G.; Zou, H.-F.; Xie, Z.-M.; Lei, G.; Huang, J.; Wang, C.-M.; Wang, H.-W.; Zhang, J.-S.; Chen, S.-Y. Soybean WRKY-type transcription factor genes, GmWRKY13, GmWRKY21, and GmWRKY54, confer differential tolerance to abiotic stresses in transgenic Arabidopsis plants. Plant Biotechnol. J. 2008, 6, 486–503. [Google Scholar] [CrossRef]
  56. Yan, H.; Jia, H.; Chen, X.; Hao, L.; An, H.; Guo, X. The cotton WRKY transcription factor GhWRKY17 functions in drought and salt stress in transgenic Nicotiana benthamiana through ABA signaling and the modulation of reactive oxygen species production. Plant Cell Physiol. 2014, 55, 2060–2076. [Google Scholar] [CrossRef] [Green Version]
  57. Cai, R.; Dai, W.; Zhang, C.; Wang, Y.; Wu, M.; Zhao, Y.; Ma, Q.; Xiang, Y.; Cheng, B. The maize WRKY transcription factor ZmWRKY17 negatively regulates salt stress tolerance in transgenic Arabidopsis plants. Planta 2017, 246, 1215–1231. [Google Scholar] [CrossRef]
  58. Niu, C.-F.; Wei, W.; Zhou, Q.-Y.; Tian, A.-G.; Hao, Y.-J.; Zhang, W.; Ma, B.; Lin, Q.; Zhang, Z.-B.; Zhang, J.-S.; et al. Wheat WRKY genes TaWRKY2 and TaWRKY19 regulate abiotic stress tolerance in transgenic Arabidopsis plants. Plant Cell Environ. 2012, 35, 1156–1170. [Google Scholar] [CrossRef]
  59. Wang, F.; Hou, X.; Tang, J.; Wang, Z.; Wang, S.; Jiang, F.; Li, Y. A novel cold-inducible gene from Pak-choi (Brassica campestris ssp. chinensis), BcWRKY46, enhances the cold, salt and dehydration stress tolerance in transgenic tobacco. Mol. Biol. Rep. 2011, 39, 4553–4564. [Google Scholar] [CrossRef]
  60. 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]
  61. Li, H.; Xu, Y.; Xiao, Y.; Zhu, Z.; Xie, X.; Zhao, H.; Wang, Y. Expression and functional analysis of two genes encoding transcription factors, VpWRKY1 and VpWRKY2, isolated from Chinese wild Vitis pseudoreticulata. Planta 2010, 232, 1325–1337. [Google Scholar] [CrossRef]
  62. Zhu, Z.; Shi, J.; Cao, J.; He, M.; Wang, Y. VpWRKY3, a biotic and abiotic stress-related transcription factor from the Chinese wild Vitis pseudoreticulata. Plant Cell Rep. 2012, 31, 2109–2120. [Google Scholar] [CrossRef] [PubMed]
  63. Wei, W.; Zhang, Y.; Han, L.; Guan, Z.; Chai, T. A novel WRKY transcriptional factor from Thlaspi caerulescens negatively regulates the osmotic stress tolerance of transgenic tobacco. Plant Cell Rep. 2008, 27, 795–803. [Google Scholar] [CrossRef]
  64. Skibbe, M.; Qu, N.; Galis, I.; Baldwin, I.T. Induced plant defenses in the natural environment: Nicotiana attenuata WRKY3 and WRKY6 coordinate responses to herbivory. Plant Cell 2008, 20, 1984–2000. [Google Scholar] [CrossRef] [Green Version]
  65. Yang, G.; Zhang, W.; Liu, Z.; Yi-Maer, A.Y.; Zhai, M.; Xu, Z. Both JrWRKY2 and JrWRKY7 of Juglans regia mediate responses to abiotic stresses and abscisic acid through formation of homodimers and interaction. Plant Biol. 2017, 19, 268–278. [Google Scholar] [CrossRef]
  66. 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]
  67. Song, Y.; Li, J.; Sui, Y.; Han, G.; Zhang, Y.; Guo, S.; Sui, N. The sweet sorghum SbWRKY50 is negatively involved in salt response by regulating ion homeostasis. Plant Mol. Biol. 2020, 102, 603–614. [Google Scholar] [CrossRef]
  68. Qin, Z.; Hou, F.; Li, A.; Dong, S.; Wang, Q.; Zhang, L. Transcriptome-wide identification of WRKY transcription factor and their expression profiles under salt stress in sweetpotato (Ipomoea batatas L.). Plant Biotechnol. Rep. 2020, 14, 599–611. [Google Scholar] [CrossRef]
  69. Zhu, H.; Zhou, Y.; Zhai, H.; He, S.; Zhao, N.; Liu, Q. A novel sweetpotato WRKY transcription factor, IbWRKY2, positively regulates drought and salt tolerance in transgenic. Arabidopsis 2020, 10, 506. [Google Scholar] [CrossRef] [Green Version]
  70. Dong, Q.; Zheng, W.; Duan, D.; Huang, D.; Wang, Q.; Liu, C.; Li, C.; Gong, X.; Li, C.; Mao, K.; et al. MdWRKY30, a group IIa WRKY gene from apple, confers tolerance to salinity and osmotic stresses in transgenic apple callus and Arabidopsis seedlings. Plant Sci. 2020, 299, 110611. [Google Scholar] [CrossRef] [PubMed]
  71. Ma, Y.; Xue, H.; Zhang, F.; Jiang, Q.; Yang, S.; Yue, P.; Wang, F.; Zhang, Y.; Li, L.; He, P.; et al. The miR156/SPL module regulates apple salt stress tolerance by activating MdWRKY100 expression. Plant Biotechnol. J. 2020, 18. [Google Scholar] [CrossRef]
  72. Ahammed, G.J.; Li, X.; Yang, Y.; Liu, C.; Zhou, G.; Wan, H.; Cheng, Y. Tomato WRKY81 acts as a negative regulator for drought tolerance by modulating guard cell H2O2–mediated stomatal closure. Environ. Exp. Bot. 2020, 171, 103960. [Google Scholar] [CrossRef]
  73. Luo, X.; Li, C.; He, X.; Zhang, X.; Zhu, L.-F. ABA signaling is negatively regulated by GbWRKY1 through JAZ1 and ABI1 to affect salt and drought tolerance. Plant Cell Rep. 2019, 39, 181–194. [Google Scholar] [CrossRef]
  74. Wang, M.-Q.; Huang, Q.-X.; Lin, P.; Zeng, Q.-H.; Li, Y.; Liu, Q.-L.; Zhang, L.; Pan, Y.-Z.; Jiang, B.-B.; Zhang, F. The overexpression of a transcription factor gene VbWRKY32 enhances the cold tolerance in Verbena bonariensis. Front. Plant Sci. 2020, 10, 1746. [Google Scholar] [CrossRef]
  75. Chanwala, J.; Satpati, S.; Dixit, A.; Parida, A.; Giri, M.K.; Dey, N. Genome-wide identification and expression analysis of WRKY transcription factors in pearl millet (Pennisetum glaucum) under dehydration and salinity stress. BMC Genom. 2020, 21, 1–16. [Google Scholar] [CrossRef] [Green Version]
  76. Zhao, K.; Zhang, D.; Lv, K.; Zhang, X.; Cheng, Z.; Li, R.; Zhou, B.; Jiang, T. Functional characterization of poplar WRKY75 in salt and osmotic tolerance. Plant Sci. 2019, 289, 110259. [Google Scholar] [CrossRef]
  77. Anjum, S.A.; Xie, X.-Y.; Wang, L.-C.; Saleem, M.F.; Man, C.; Lei, W. Morphological, physiological and biochemical responses of plants to drought stress. Afr. J. Agr. Res. 2011, 6, 2026–2032. [Google Scholar] [CrossRef]
  78. Sakuma, Y.; Maruyama, K.; Osakabe, Y.; Qin, F.; Seki, M.; Shinozaki, K.; Yamaguchi-Shinozaki, K. Functional analysis of an Arabidopsis transcription factor, DREB2A, involved in drought-responsive gene expression. Plant Cell 2006, 18, 1292–1309. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Hassan, S.; Lethin, J.; Blomberg, R.; Mousavi, H.; Aronsson, H. In silico based screening of WRKY genes for identifying functional genes regulated by WRKY under salt stress. Comput. Biol. Chem. 2019, 83, 107131. [Google Scholar] [CrossRef]
  80. Linxiao, W.; 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]
  81. Shen, Z.; Yao, J.; Sun, J.; Chang, L.; Wang, S.; Ding, M.; Qian, Z.; Zhang, H.; Zhao, N.; Sa, G.; et al. Populus euphratica HSF binds the promoter of WRKY1 to enhance salt tolerance. Plant Sci. 2015, 235, 89–100. [Google Scholar] [CrossRef] [PubMed]
  82. Jaglo-Ottosen, K.R. Arabidopsis CBF1 overexpression induces COR genes and enhances freezing tolerance. Science 1998, 280, 104–106. [Google Scholar] [CrossRef] [Green Version]
  83. 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] [Green Version]
  84. Jin, B.; Wang, L.; Wang, J.; Jiang, K.-Z.; Wang, Y.; Jiang, X.X.; Ni, C.-Y.; Wang, Y.; Teng, N.-J. The effect of experimental warming on leaf functional traits, leaf structure and leaf biochemistry in Arabidopsis thaliana. BMC Plant Biol. 2011, 11, 35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Liu, H.-T.; Li, G.-L.; Chang, H.; Sun, D.-Y.; Zhou, R.-G.; Li, B. Calmodulin-binding protein phosphatase PP7 is involved in thermotolerance in Arabidopsis. Plant Cell Environ. 2007, 30, 156–164. [Google Scholar] [CrossRef] [PubMed]
  86. Larkindale, J.; Huang, B. Effects of abscisic acid, salicylic acid, ethylene and hydrogen peroxide in thermotolerance and recovery for creeping bentgrass. Plant Growth Regul. 2005, 47, 17–28. [Google Scholar] [CrossRef]
  87. Zhao, J.; He, Q.; Chen, G.; Wang, L.; Jin, B. Regulation of non-coding RNAs in heat stress responses of plants. Front. Plant Sci. 2016, 7, 1213. [Google Scholar] [CrossRef] [Green Version]
  88. Gong, X.-Q.; Hu, J.-B.; Liu, J.-H. Cloning and characterization of FcWRKY40, A WRKY transcription factor from Fortunella crassifolia linked to oxidative stress tolerance. Plant Cell Tissue Organ Cult. PCTOC 2014, 119, 197–210. [Google Scholar] [CrossRef]
  89. 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] [Green Version]
  90. Robatzek, S.; Somssich, I.E. A new member of the Arabidopsis WRKY transcription factor family, AtWRKY6, is associated with both senescence- and defence-related processes. Plant J. 2001, 28, 123–133. [Google Scholar] [CrossRef]
  91. 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. 2007, 50, 347–363. [Google Scholar] [CrossRef]
  92. Han, Y.; Zhang, X.; Wang, Y.; Ming, F. The suppression of WRKY44 by GIGANTEA-miR172 pathway is involved in drought response of Arabidopsis thaliana. PLoS ONE 2013, 8, e73541. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The domain of WRKY genes in Arabidopsis thaliana. The WRKY gene family is classified into the I (I N and I C), IIa, IIb, IIc, IId, IIe, and III subfamilies. The aligned conserved domains (DNA binding and zinc-finger structures) are highlighted (left panel) and simplified (right panel).
Figure 1. The domain of WRKY genes in Arabidopsis thaliana. The WRKY gene family is classified into the I (I N and I C), IIa, IIb, IIc, IId, IIe, and III subfamilies. The aligned conserved domains (DNA binding and zinc-finger structures) are highlighted (left panel) and simplified (right panel).
Plants 09 01515 g001
Figure 2. Some WRKY genes involved in the response pathways of major abiotic stresses (drought, salt, cold, heat, oxidative stress, mechanical injury, UV-B).
Figure 2. Some WRKY genes involved in the response pathways of major abiotic stresses (drought, salt, cold, heat, oxidative stress, mechanical injury, UV-B).
Plants 09 01515 g002
Figure 3. WRKY transcription factors in response to abiotic stresses.
Figure 3. WRKY transcription factors in response to abiotic stresses.
Plants 09 01515 g003
Table 1. WRKY transcription factors (TFs) involved in abiotic stress responses in plants.
Table 1. WRKY transcription factors (TFs) involved in abiotic stress responses in plants.
No.GeneSpeciesInduced by FactorsFunctionReferences
1AtWRKY25/26ArabidopsisHeatTolerance to heat[34]
2AtWRKY33ArabidopsisNaCl, mannitol, H2O2Tolerance to heat and NaCl, negative regulator in oxidative stress and abscisic acid (ABA)[33]
3AtWRKY34ArabidopsisColdNegative regulator in cold stress[35]
4AtWRKY39ArabidopsisHeatTolerance to heat[36]
5AtWRKY53ArabidopsisDrought, saltReduced drought resistance and H2O2, sensitive to salt [37,38]
6AtWRKY57ArabidopsisDroughtTolerance to drought[39]
7AtWRKY63ArabidopsisABATolerance to drought, regulated ABA signaling[40]
8AtWRKY54ArabidopsisHeatResponse to heat stress[41]
9POWRKY13Populus tomentosaHeatResponse to heat stress[42]
10GhWRKY21Gossypium hirsutumDroughtTolerance to drought[43]
11GhWRKY25Gossypium hirsutumDroughtTolerance to salt, reduced drought resistance[44]
12GhWRKY68Gossypium hirsutumSalt, droughtReduced salt tolerance and drought resistance, positive regulator in ABA signaling[45]
13VvWRKY24Vitis viniferaColdUpregulated expression at all stages of hypothermia[46]
14CaWRKY40Capsicum annuumHeatTolerance to heat[47]
15BdWRKY36Brachypodium distachyonDroughtTolerance to drought[48]
16FcWRKY70Fortunella crassifoliaSaltTolerance to salt[49]
17OsWRKY11Oryza sativaHeat, droughtTolerance to drought and heat[50]
18OsWRKY72Oryza sativaDrought, NaCl, ABASensitive to salt, drought, sucrose, and ABA[51]
19OsWRKY74Oryza sativaPi deprivation, coldTolerance to cold and Pi deprivation[52]
20OsWRKY76Oryza sativaColdTolerance to cold[53]
21OsWRKY89Oryza sativaABA, UV-BTolerance to UV[54]
22GmWRKY13SoybeanSalt, droughtSensitive to salt and mannitol, negative regulator in ABA signaling[55]
23GmWRKY17SoybeanSaltReduced salt tolerance[56]
24GmWRKY54SoybeanSalt, droughtTolerance to salt and drought[55]
25GmWRKY21Glycine maxNaCl, drought, coldTolerance to cold[55]
26ZmWRKY17Zea maysABA, saltReduced salt tolerance[57]
27TaWRKY2/19Triticum aestivumNaCl, drought, ABATolerance to salt and drought[58]
28BcWRKY46Brassica campestrisNaCl, drought, coldTolerance to salt and drought[59]
29BhWRKY1Boea hygrometricaDehydration, ABATolerance to drought[60]
30VpWRKY1Vitis pseudoreticulataNaCl, ABATolerance to salt[61]
31VpWRKY2Vitis pseudoreticulataCold, NaCl, ABATolerance to salt and cold[61]
32VpWRKY3Vitis pseudoreticulataDrought, ABA, salicylic acid (SA)Tolerance to salt[62]
33TcWRKY53Thlaspi caerulescensCold, PEG, NaClNegative regulator in osmotic stress[63]
34NaWRKY3Nicotiana attenuateMechanical damageSensitive to mechanical damage[64]
35JrWRKY2/7Juglans regiaDrought, coldTolerance to drought and cold[65]
36SbWRKY30Sorghum bicolorSalt, droughtTolerance to salt and drought[66]
37SbWRKY50Sorghum bicolorSaltTolerance to salt[67]
38IbWRKY47Ipomoea batatasSaltTolerance to salt[68]
39IbWRKY2Ipomoea batatasSalt, droughtTolerance to salt and drought[69]
40MdWRKY30Malus domesticaSalt, osmotic stressTolerance to salt and osmotic stress[70]
41MdWRKY100Malus domesticaSaltSensitive to salt[71]
42SlWRKY81Solanum lycopersicumDroughtReduced drought tolerance[72]
43GbWRKY1Gossypium barbadenseSaltTolerance to salt[73]
44VbWRKY32Verbena bonariensisColdTolerance to cold[74]
45PgWRKY33/62Pennisetum glaucumSalt, droughtTolerance to salt and drought[75]
46PagWRKY75Populus albaDroughtNegative regulator in salt and osmotic stress[76]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

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. https://doi.org/10.3390/plants9111515

AMA Style

Li W, Pang S, Lu Z, Jin B. Function and Mechanism of WRKY Transcription Factors in Abiotic Stress Responses of Plants. Plants. 2020; 9(11):1515. https://doi.org/10.3390/plants9111515

Chicago/Turabian Style

Li, Weixing, Siyu Pang, Zhaogeng Lu, and Biao Jin. 2020. "Function and Mechanism of WRKY Transcription Factors in Abiotic Stress Responses of Plants" Plants 9, no. 11: 1515. https://doi.org/10.3390/plants9111515

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