Next Article in Journal
Is the Complement System Dysregulated in Preeclampsia Comorbid with HIV Infection?
Previous Article in Journal
N-Acetylcysteine Attenuates Cisplatin Toxicity in the Cerebrum and Lung of Young Rats with Artificially Induced Protein Deficiency
Previous Article in Special Issue
Unraveling a Small Secreted Peptide SUBPEP3 That Positively Regulates Salt-Stress Tolerance in Pyrus betulifolia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genome-Wide Identification, Characterisation, and Evolution of the Transcription Factor WRKY in Grapevine (Vitis vinifera): New View and Update

by
Ekaterina Vodiasova
1,2,*,
Anastasiya Sinchenko
1,
Pavel Khvatkov
1 and
Sergey Dolgov
1,3
1
Federal State Funded Institution of Science “The Labor Red Banner Order Nikita Botanical Gardens—National Scientific Center of the RAS”, Nikita, 298648 Yalta, Russia
2
A.O. Kovalevsky Institute of Biology of the Southern Seas of RAS, 299011 Sevastopol, Russia
3
Branch of Shemyakin and Ovchinnikov Institute of Bioorganic Chemistry, 142290 Puschino, Russia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(11), 6241; https://doi.org/10.3390/ijms25116241
Submission received: 3 May 2024 / Revised: 29 May 2024 / Accepted: 3 June 2024 / Published: 5 June 2024
(This article belongs to the Special Issue Latest Research on Plant Genomics and Genome Editing)

Abstract

:
WRKYs are a multigenic family of transcription factors that are plant-specific and involved in the regulation of plant development and various stress response processes. However, the evolution of WRKY genes is not fully understood. This family has also been incompletely studied in grapevine, and WRKY genes have been named with different numbers in different studies, leading to great confusion. In this work, 62 Vitis vinifera WRKY genes were identified based on six genomes of different cultivars. All WRKY genes were numbered according to their chromosomal location, and a complete revision of the numbering was performed. Amino acid variability between different cultivars was assessed for the first time and was greater than 5% for some WRKYs. According to the gene structure, all WRKYs could be divided into two groups: more exons/long length and fewer exons/short length. For the first time, some chimeric WRKY genes were found in grapevine, which may play a specific role in the regulation of different processes: VvWRKY17 (an N-terminal signal peptide region followed by a non-cytoplasmic domain) and VvWRKY61 (Frigida-like domain). Five phylogenetic clades A–E were revealed and correlated with the WRKY groups (I, II, III). The evolution of WRKY was studied, and we proposed a WRKY evolution model where there were two dynamic phases of complexity and simplification in the evolution of WRKY.

1. Introduction

Living organisms are constantly exposed to various negative factors. Plants have an attached way of life and have been forced to develop special mechanisms to cope with the harmful effects of the environment. One of the ways to increase resistance is through the evolution of an intricate signalling system that allows them to respond rapidly to external stimuli. The key molecules of signalling systems are transcription factors (TFs) that determine the response of plants to biotic and abiotic stress [1,2,3].
One of the largest multigene families of transcription regulators in higher plants is the WRKY TFs [4]. These proteins are important regulators that are involved in the response to various biotic and abiotic factors [5,6,7,8,9,10,11,12,13]. They are integral components of numerous facets of the plant innate immune system and play pivotal roles in plant growth and development [14,15,16].
The WRKY transcription factors possess a DNA-binding domain (DBD) at the N-terminal and a zinc-finger motif at the C-terminal of the protein [17]. The sixty amino acid-long domain is characterised by the signature motif ‘WRKYGQK’. WRKY proteins are divided into four groups according to the number of WRKY domains and zinc finger pattern [18]. The WRKYs belonging to group I have two WRKY DBDs and a Cys2-His2-type zinc finger motif; group II contains proteins with a single DBD and a Cys2-His2-type zinc finger motif; group III possesses proteins with a single DBD and a Cys2-His/Cys-type zinc finger motif; and group IV contains proteins with an incomplete WRKY domain without a zinc finger [19,20]. Concurrently, some WRKY genes with three domains (Glycyrrhiza uralensis, Lupinus angustifolius, Gossypium raimondii, Linum usitatissimum) and even with four (Aquilegia coerulea, Oryza nivara, Solanum lycopersicum) have been identified [16]. Although the WRKY DNA-binding domain is highly conserved, there is increasing evidence of amino acid substitutions in the ‘WRKYGQK’ motif [14]. Such variability has been described in maize [21], banana [22], soybean [23], and rice [24].
To date, WRKY transcription factors have been identified in more than 150 plants, according to the PlantTFDB database. The number of genes encoding this TF varies significantly among plants. The lowest number (<20) of WRKY genes was found for algae or other poorly studied plants (6 WRKY genes in Helianthus annuus) (yellow column in Figure 1B). The limited number of WRKY genes described for some plant species is likely due to the lack of genomic data. For example, only 46 genes were identified for B. napus in 2009 [25]. However, following the availability of the first genomic data in 2016, the number of described genes increased to 287 [26]. Another study identified 278 genes [27], while the PlantTFDB database described 285 genes.
A few plants have a relatively small number of WRKYs, including Amborella trichopoda (32), Coffea canephora (49), Genlisea aurea (38), Carica papaya (49), and Lactuca sativa (50), as reported in PlantTFDB. The maximum number of WRKY genes was identified in Brassica napus (285), Glycine max (296), and Panicum virgatum (275). The distribution of the number of WRKY genes in different plant species, as reported by PlantTFDB, is represented in Figure 1. The majority of plants have between 50 and 130 genes encoding WRKY. Even closely related species exhibit differences in the number of genes. For example, Brassica napus has 285 genes, Brassica oleracea has 191 genes, and Brassica napa has 180 genes.
The considerable diversity in the number of genes present in different plants serves to illustrate the complex evolution of the WRKY gene family. As previously stated, these TFs are involved in the regulation of a multitude of biological processes, including the response to stressful biotic and abiotic factors. Therefore, the large number of duplication events may be an adaptation mechanism to unfavourable external stimuli.
At present, many studies have been carried out on the effect of various stresses on WRKYs expression in different plants [14,15,16,28,29,30,31,32,33]. The principal stress factors under investigation are presented in Figure 1. As WRKYs play a pivotal role in plant defence in response to abiotic stress, a genetic engineering approach could be an effective strategy for enhancing tolerance to some negative factors [34,35,36]. The grapevine (Vitis vinifera L.) is one of the most widely cultivated plant species for commercial purposes. Global production of this crop reaches 70 million tonnes, occupying more than seven million hectares of land for harvesting [37]. In 2016, the value of grapes in agricultural enterprises reached $68 billion, making it the most valuable fruit crop in the world. Vitis vinifera is mainly used to produce a variety of commodities, including wine, table grapes, sultanas, grape juice concentrate, and spirits for industrial use [38]. Adaptation to climate change is an important step in the future of viticulture, which depends heavily on weather and climatic conditions [39]. The primary challenge identified is the heightened risk of water stress, which may result in reduced yield in terms of both quantity and quality, given that a water deficit is observed in the majority of grape-producing regions [40,41,42,43]. Consequently, it is of paramount importance to study the functioning of this transcription factor family in grapevine and to identify WRKY genes that respond to one or another type of biotic and abiotic factor.
To date, several genome-wide analyses of WRKY genes in grapevine have been performed. The number of genes encoding WRKY found varied and appeared to depend on the version of the genomic assembly analysed. A number of papers have identified 59 genes based on the 12X assembly of the V. vinifera cv. Pinot Noir (PN40024) genome sequences and the NCBI GenBank database [37,44,45]. The PlantTFDB transcription factor database also lists 59 WRKY for V. vinifera, the same number of genes analysed in other studies [7,46]. When analysing the NCBI GenBank database, 80 WRKY genes were predicted, which is probably due to the presence of heterozygosity in loci [47]. With the appearance of a more complete genomic assembly of grape (NCBI accession number GCA_030704535) and an improved annotation of the first genomic assembly (GCA_000003745, v3 annotation), a revision of the WRKY genes was conducted, resulting in the identification of 61 genes [48,49]. Nevertheless, all research on the identification of this transcription factor has been conducted on a single Pinot Noir cultivar (PN40024). Consequently, the chromosomal position of some WRKYs remains unconfirmed, and the total number of these genes, including heterozygous variants, remains unknown.
Also, the functional role of different VvWRKY genes was investigated in several studies where the authors in transgenics of A. taliana or N. tabacum revealed the effect of VvWRKY genes by overexpression in plant tissues [50,51,52,53,54,55,56,57,58,59,60,61,62,63]. The function of WRKY transcription factors has also been studied in transgenic grapes [64,65,66,67,68,69]. In these studies, the WRKY gene numbers were determined by the level of homology to Arabidopsis WRKY or by homology with the already known RefSeq or PlantTFDB databases. This has resulted in confusion and a lack of a unified systematics of grape WRKY, as previously mentioned [70]. Consequently, it is difficult to utilise the results of these studies.
This research is devoted to the identification of WRKY using genome-wide analyses based on several grape cultivar assemblies (including diploid). Phylogenetic analyses, protein structure, and intra-varietal variability were employed to perform a comprehensive revision of grape WRKY proteins and propose a novel classification.

2. Results

2.1. The Phylogeny of Grape WRKY Genes

The domain analysis found WRKY DNA-binding domain proteins in each grapevine assembly. However, the number of proteins identified differed between grape cultivars. A total of 234 proteins were found in the diploid assembly for Cabernet Franc, 181 for Cabernet Sauvignon, and 304 for Pinot Noir clone FPS123. In the haploid assemblies of Pinot Noir clone PN40024, the number of proteins identified was also variable and depended on the annotation. These were 89 for the 12X assembly (GCA_000003745.2), 87 for the RefSeq reference assembly (GCA_030704535), 65 for the GenBank reference assembly (GCA_030704535), and 84 for the assembly from GrapeGenomics database. This discrepancy in the number of proteins can be attributed to the presence of multiple variant isoforms for a single gene, the heterozygosity of loci, and the annotation pipeline. Given that some WRKY proteins identified in the 12X assembly and the RefSeq reference assembly have identical IDs, the two datasets were merged into a single entity, and only unique IDs, of which 97 were found, were retained for further analysis. Thus, the combined set of WRKY amino acid sequences from six grape genomic assemblies contained 965 proteins, including all isoforms.
A phylogenetic analysis was conducted to reveal the phylogenetic relationships among all 965 complete amino acid sequences. The analysis identified 62 clusters with 100 bootstrap supports, which were found to be significantly different from neighbouring clusters (Figure 2).
Each cluster corresponds to a specific VvWRKY TF and contains a variable number of amino acid sequences from different grape cultivars. The phylogenetic tree with expanded clusters is presented in Supplementary Figure S1. For each gene encoding the corresponding WRKY protein, its position on the chromosome was determined according to each grape cultivar genome. All gene IDs and chromosome locations are given in Supplementary Table S1. WRKY TFs were numbered according to the ordinal chromosome location number (in most of the assemblies analysed).
The exon-intron structure and conserved motifs were analysed for 62 WRKY proteins of the reference grape genome for Pinot Noir cl. PN40024 v. 5 from the GrapeGenomics Database (Figure 2). The number of exons ranged from 2 up to 8. Twenty-five motifs were found for 62 protein sequences from V. vinifera cultivar Pinot Noir cl. PN40024 v. 5 (GrapeGenomics database). All motif sequences in grape WRKY proteins are represented in Table 1. All WRKYs except for VvWRKY26 were revealed to have three conserved motifs arranged one after the other: 1-3-2. Motif 1 is characterised by the presence of the WRKY signature. Motifs 4 and 13 also contain the WRKY signature. Based on the analysis of the number of DNA-binding domains and the zinc finger motif (Supplementary Figure S2), all studied proteins were assigned to a specific WRKY type for grape described before (groups I, IIa–IIe, III) [37,44,45,47].
A phylogenetic analysis revealed clustering between the different WRKY classes. Five evolutionary groups (A–E) were identified based on the complete amino acid sequences of the proteins. The WRKY groups (I, IIa–IIe, III) were completely correlated with the evolutionary groups. Within some groups, subclades with 100 bootstrap supports were distinguished.
Group A is associated with WRKY type III. It does not contain subclusters, exhibits a similar exon-intron structure (three exons), and has a similar set of conserved motifs. Motifs 11 and 15 in the N-terminal region of the protein are distinctive characteristics of this group.
Group B (the group of WRKY IIc) is comprised of four subclades, each with 100 bootstrap supports (B1–B4). Subgroup B1 is distinguished by a motif 16 in the C-terminal region, which is unique to it and exhibits a three-exon structure. Subgroup B4 is characterised by a simple structure and a two-exon configuration.
Group C is distinguished by genetic diversity, differences in the composition of conserved motifs in different WRKYs, complex structure (the number of exons varies from 4 to 8), and long length. WRKYs of this phylogenetic group belong to the WRKY type I category, as they contain two conserved domains containing the WRKY signature. The C-terminal WRKY domain contains conserved motif 1, while the N-terminal WRKY domain varies in different proteins and is characterised by motif 4 (in VvWRKY11, VvWRKY37, VvWRKY29, VvWRKY42, VvWRKY27, VvWRKY18, VvWRKY15, VvWRKY3) or motif 13 (in VvWRKY61, VvWRKY33, VvWRKY62, VvWRKY49). No clustering into subgroups is observed in C.
In group D, as in group C, WRKY genes exhibit a distinct composition of conserved domains and a complex exon-intron structure (the number of exons varies from 3 to 6). This evolutionary group corresponds to groups IIa + IIb. Three clusters with 100 supports are distinguished, each lacking characteristic conserved motifs or exon-intron structure (D1–D3). The D3 subgroup corresponds to the WRKY IIa type. Group D is distinguished by the presence of two characteristic motifs, 7 and 8, which are located on the N-terminal and C-terminal, respectively. Subgroups D1 and D2 (IIb) are characterised by the presence of a conserved motif 10, which follows the WRKY signature.
Group E is distinguished by a gene structure comprising three exons and the presence of two subgroups, which corresponds to the division into two types, IId and IIe. Subgroup E1 (IId) exhibits characteristic motifs 14 and 18 on the N-terminal and motif 9, followed by motif 1 with tetra amino acid residues (WRKY). Subgroup E2 (IIe, respectively) has the simplest set of conserved motifs of all WRKYs (only common for all 1-3-2 motifs).
There are genes that differ from the others (Figure 2, black arrows). VvWRKY17 and VvWRKY30 represent distinct evolutionary branches and cannot be assigned to either phylogenetic group. VvWRKY44 does not possess motifs 11 and 15, which are characteristic of the entire group A. VvWRKY26 lacks motif 1, which contains tetra-amino acid residues, but has zinc fingers.

2.2. The Characteristics of 62 WRKY Classes in Grapes

Phylogenetic analysis enabled the classification of all WRKYs in different grape cultivars and assemblies (Table 2). For each protein, the coding gene, its corresponding locus on the chromosome, and the number of genes encoding each VvWRKY class (Ng) were identified. The number of isoforms (Ni) varied for a single gene from 1 to 36.
In some cases, a large number of isoforms were determined, even in haploid assemblies. For VvWRKY1, VvWRKY10, VvWRKY15, VvWRKY22, VvWRKY29, VvWRKY33, VvWRKY49, and VvWRKY62, the number of isoforms exceeded 10 in the diploid assembly and exceeded 5 in the haploid one. The analysis of haploid assemblies revealed that, with the exception of VvWRKY42, each transcription factor class had a single locus in the genome. For VvWRKY42, three loci were identified in the genome of Pinot Noir clone 40024 (v.5, Grape Genomics), two of which were located on chromosome 12 and one on chromosome 9. Upon analysis of diploid-phased assemblies of three grape cultivars, it was observed that two genes encoding a particular transcription factor were most often identified. This is consistent with the diploid-phased assembly. However, instances were noted where only one locus in the genome or three (potentially indicative of duplication) were detected.
Analyses of the chromosomal location of WRKY genes revealed that they are absent from the third chromosome. There are regions of chromosomes where there are clusters of several genes, but there are also chromosomes with only one or two WRKY genes. WRKY genes of the same class in different grape cultivars are located on the same chromosome, with the exception of the genes encoding the transcription factors VvWRKY42 and VvWRKY56. The VvWRKY42 gene in the Pinot Noir cl. PN40024 is located on chromosomes 9 and 12, while the VvWRKY56 gene in the Pinot Noir cl. FPS123 is located on chromosomes 2 and 16 (Figure 3).
It was observed that some WRKYs were not present in all grape cultivars. For instance, VvWRKY26 and VvWRKY53 were not found in Cabernet Franc cl. 04, while VvWRKY4 and VvWRKY26 were not found in Cabernet Sauvignon cl. 08. Furthermore, differences were also identified between the various assemblies of the reference cultivar Pinot Noir cl. PN40024, likely due to differences in the assembly and annotation pipelines employed. A comprehensive list of all analysed and classified WRKY proteins, along with their respective IDs and chromosomal locations, is provided in Supplementary Table S1.
For each WRKY, the amino acid variability within the class, mutations in DNA-binding hepta-peptide sequences, and the type of zinc finger motif were analysed. Only the 62 amino acid WRKY sequences from V. vinifera cultivar Pinot Noir cl. PN40024, v. 5 (Grape Genomics database) (see Table 2 for gene IDs) were retained to analyse length, position on the chromosomes, number of exons, functional family determination (FF:number), WRKY domain location, and presence of other domains in the proteins, as this is the only assembly in which all WRKYs were detected. The results are presented in Table 3.
The length of VvWRKY ranged from 151 to 746 aa and the number of exons from 2 to 8. The mean value of genetic distances between varieties ranged from 0 (VvWRKY12) to 0.113 (VvWRKY17). Variability of the conserved DNA-binding heptapeptide WRKYGQK was detected. VvWRKY8, VvWRKY13, VvWRKY14, and VvWRKY24 possess the WRKYGKK amino acid sequence, whereas in VvWRKY17, a mutation Arg to Lys has occurred in the characteristic tetrapeptide WRKY. The WRKY domains were analysed, and functional families were identified according to the CATH databases.
The analysis of complete protein sequences rather than domain regions alone, enabled the identification of chimeric forms among VvWRKY transcription factors that include other domains in addition to DBD WRKY. The following chimeric VvWRKYs were identified in grapes: a Zn-cluster domain (IPR018872) followed by a WRKY domain (1); a Frigida-like domain upstream of two WRKY domains (IPRO12474) (2); the structure motif COILS (3), which is a coil located upstream or downstream of the domain; and the signal peptide in the N-terminal and noncytoplasmic domain (VvWRKY17) (4). The LxLxLx repressor and the LxxLL co-activator motifs were identified in eight and twelve VvWRKY proteins, respectively (Table 3).
A phylogenetic analysis of 62 amino acid sequences of VvWRKY domains from V. vinifera cultivar Pinot Noir clone PN40024, v. 5 (Grape Genomics database) revealed a similar tree topology as for the complete protein sequences of this gene family (Figure 4).
Clade B (WRKY group IIc), clade C (I), and clade D (IIa + IIb) are sister lineages and form a single large cluster. This is also observed in the tree based on the complete protein sequences (Figure 2). At the same time, the two trees exhibit a difference in topologies. Clade E (groups WRKY IIe + IId) and clade A (III) constitute a single clade with high bootstrap support, whereas this was not observed when the complete protein sequences were examined.
An association between the clustering of WRKY groups with the functional family of the protein, WRKY signature sequences, and the presence of other domains in the protein with the formation of chimeric WRKY TFs was revealed. Thus, clade D (IIa + IIb) is characterised by the presence of the structural motif COILS Coil. This motif is also present in VvWRKY16 (IIc), VvWRKY21 (IIc), and VvWRKY54 (IIe). The Zn-cluster domain characterises the subclade group E2 (IId). The heptapeptide sequence WRKYGKK is present only in subclade B2 (group IIc).
An analysis of the WRKY gene structure revealed a bipartite distribution of both protein lengths (approximately 300 and 500 amino acids) and exon numbers (three and five, respectively) (Figure 5A,C). This indicates the presence of VvWRKY genes with varying degrees of complexity in their structure. Analyses of exon number and protein length in relation to the evolutionary clades of VvWRKY genes demonstrated that clades C and D exhibited a more complex structure, while the other VvWRKYs exhibited a more simplified gene structure (Figure 5B,C).

2.3. Unified Systematics of Grape WRKY Genes

Each WRKY gene was matched with IDs from the most common databases: PlantTFDB, NCBI (RefSeq, GenBank), Ensembel, Uniprot, and the previously used names of each WRKY transcription factor (Table 4). A homology analysis of all amino acid sequences revealed a correspondence between the new numbering proposed by us (based on the localisation of genes on chromosomes) and all other protein IDs and WRKY numbers used previously. The confusion in gene names is evident from the table. Most of the genes have more than three different names, and some of them have as many as six different names. For example, the gene VvWRKY31 had the names VvWRKY14, VvWRKY30, VvWRKY40, VvWRKY28, and VvWRKY4 in different databases and studies. It was also found that the RefSeq database (which is also included in the reference genome annotation) has genes that have the same numbers but belong to different VvWRKY classes. For example, VvWRKY5, VvWRKY53, and VvWRKY54 are described in the RefSeq database as VvWRKY22 (XP_010658402, XP_002276925, and XP_010662789).

3. Discussion

The plant-specific transcription factor WRKY is involved in plant development and in the response to various biotic and abiotic stresses. This gene family has been studied in many species, including the model plant Arabidopsis and important crops such as rice, cucumber, coffee, tomato, etc. [70,71,72,73]. Despite the large number of studies on grapes, the WRKY genes have not been fully studied in this crop. To date, there are several important studies that have focused on the search for WRKY genes in grapes [37,44,45,47]. These studies were based on analyses of the genome of only one cultivar, V. vinifera cv. Pinot Noir (PN40024) (GCA_000003745), where the assembly has been continuously improved and re-annotated. This has led to a variation in the number of genes identified (from 59 to 80), their length, the number of exons, and the clarification of the different motifs present in the above studies.
To date, two chromosome-level assemblies are available at NCBI for Pinot Noir, which is accepted as a reference. The assembly GCA_000003745, known as 12X, is the first assembly of the grape genome and already has assembly version 3 [74]. This assembly has already been analysed in previous studies. In August 2023, a new assembly, GCA_030704535, appeared, which became the reference assembly for this species and was not previously analysed for WRKY studies [75]. The Grape Genomics database also contains the V. vinifera cultivar Pinot Noir cl. PN40024, v. 5, with its annotation [75,76]. In our work, 62 WRKY genes were identified based on the analysis of all three assemblies mentioned above. The length range of the found VvWRKYs is from 151 to 746 aa and differs from the previously identified ones from 101 to 612 aa [37]. The number of exons ranges from 2 to 8, while the previously identified ranges were 1–17 [45] and 2–7 [37]. The differences in length and exon-intron structure are explained by the incomplete assembly of GCA_000003745 compared to today.

3.1. Grapevine WRKY Numbering

Since there is no unified justification for naming genes in this family, a great deal of confusion has arisen over the years of studying WRKY in grapes. First, some researchers named the gene according to the number of the homologous WRKY gene in a model plant, mainly Arabidopsis [10,50,54,56,57,58,59,60,61,62,63,64,66,77,78,79,80,81]. It seems unreasonable to use homology to another species in this multigene family for numbering, since the number of genes encoding WRKY TFs varies greatly in different plants. The second group of researchers numbered the genes according to their position on the chromosome [37,44,45,70]. Initially, because not all genes had a clear localisation, unknown chromosomes were left in the assemblies, and much confusion arose. A recent study on WRKY gene classification also failed to conclusively determine the position of WRKY59 (an unknown chromosome) [70].
At the same time, a uniform numbering system is necessary because confusion is already emerging. For example, in work on transgenic grapes, it was shown that WRKY70 (VIT_13s0067g03140) is involved in norisoprenoid and flavonol biosynthesis [69]. At the same time, the name WRKY70 is found in the RefSeq database under two IDs: XP_00227272504 (in our study VvWRKY28) and XP_002275401 (in our study VvWRKY45). Thus, naming VvWWRKY70 could lead to errors in further studies. The situation with the VvWRKY35 gene is different. Its role in tolerance to cold and salt stress in transgenic Arabidopsis was previously studied, where it was described as VvWRKY28 [55]. In another paper investigating the functions of WRKYs, this gene (GSVIVT01021397001) is referred to as VvWRKY22 [7], and in a paper investigating WRKYs in strawberries, it is included in the phylogenetic analysis as VvWRKY71 (XP_002272089) [82]. At the same time, another VvWRKY71 with accession number XP_002283603 is deposited in the RefSeq database and analysed under this name in another paper on grape [83]. A recent paper was published showing that VvWRKY71 could promote the biosynthesis of proanthocyanidins, but it is extremely difficult to understand which RefSeq ID was used to annotate the assembled transcriptome [84]. Such situations show the extreme necessity of using a unified WRKY gene numbering system.
Our genome-wide analysis, based on several complete genome assemblies, has determined the chromosomal position for all 62 genes found. Therefore, we propose to name the genes according to their location on the chromosome in the reference genome of V. vinifera cv. Pinot Noir cl. PN40024, v.5 [75,76] with an annotation dated September 2023 (GenBank annotation). For the convenience of researchers, not only was a complete revision of the numbering carried out, but correspondences were also found between the proposed new numbering and all the most common IDs from different databases and previous WRKY gene numbers (Table 4).

3.2. Grape WRKY Diversity in Cultivars

With the development of NGS technologies and the accumulation of a large amount of data, there are studies on the comparison of several genomes within a species, which allows us to assess the presence of intraspecific nucleotide variability in genes. Thus, on the basis of comparative analyses of GATA transcription factors among 19 Arabidopsis genomes, intraspecific amino acid variability was shown [85]. Sequence diversity of the WRKY transcription factor family has been demonstrated in wild and cultivated barley, where the authors showed that haplotype and nucleotide diversity in the majority of WRKY genes were higher in the wild barley population [86]. The high nucleotide variability could lead to further misestimation of relative expression by RT-PCR if primers hit no conservative region of the gene. This situation is probably observed in studies of WRKY expression in response to pathogenic fungal infection and SA treatment. As the authors identified only 57% of the differentially expressed genes (in contrast to 70% of the DEGs in Arabidopsis), they suggested that the reason for this difference was poor primer annealing in the studied cultivar, due to the primer design using the other cultivar [47].
In our study, WRKY transcription factor genes were analysed for the first time in different grape varieties (Cabernet Franc, Cabernet Sauvignon, and Pinot Noir clones FPS123 and PN40024). For some WRKY genes, inter-varietal amino acid variability was detected (Table 3), which was greater than 5% for the genes VvWRKY3, VvWRKY4, VvWRKY17, VvWRKY24, VvWRKY35, VvWRKY42, and VvWRKY54. It should also be noted that some WRKY genes are not present in all varieties. For some genes, there was a significant difference in the number of isoforms in different cultivars (Table 2). For example, the transcription factor VvWRKY1 is encoded by a single gene, but the number of predicted isoforms differed: 12 isoforms for Cabernet Franc, 10 for Cabernet Sauvignon, 15 for Pinot Noir cl. FPS123, but only 4 for Pinot Noir cl. PN40024. This seems to be explained by the fact that for Pinot Noir cl. PN40024, the assembly is haploid, and for the other cultivars, the assembly is diploid. However, VvWRKY33 has 10 isoforms for Pinot Noir cl. FPS123 and only 2 isoforms each for Cabernet Franc and Cabernet Sauvignon. It is obvious that the variability of some WRKYs is quite high, which may indirectly influence, for example, the resistance of different varieties to stress factors. This issue needs further study.

3.3. WRKY Domains in Grapevine

WRKY transcription factors are characterised by a specific gene structure: they contain a DNA-binding domain with a conserved WRKY motif (WRKYGQK) in the L-terminal and a zinc finger motif in the C-terminal. Trp, Tyr, and two Lys residues in the heptapeptide sequence are known to be essential for DNA binding [87,88,89]. However, there is known diversity in the WRKY signature sequence [16,37,45,90]. Similar to previous studies, we detected grape transcription factors with WRKYGKK (VvWRKY8, VvWRKY13, VvWRKY14, VvWRKY24) [37,45,47]. Variants of WRKY domains with such hepta amino acid residues are consistently found in other plants: rice, OsWRKY7 [91], tobacco, NtWRKY12 [92], pepper, CaWRKY39, CaWRKY50, CaWRKY56, and CaWRKY62 [93], cucumber CsWRKY41, CsWRKY44, and CsWRKY54 [94]. Thus, this sequence is likely to be plant-specific, but the efficiency of binding to WK- and W-boxes is not fully understood [91].
Also in grapevine, the VvWRKY17 gene was found to contain the conservative tetrapeptide sequence WKKY instead of the traditional WRKY, which is confirmed by other studies [37,45]. This variant is frequently found in different species, indicating that this mutation is fixed [28,90]. It is likely that genes with this motif have some functionality, and such mutations could result from altered conditions and be needed for the development of plant resistance [16,95,96].
We first discovered VvWRKY26, which contains a WRKY domain but has a previously unknown WRKY signature sequence, WMKGNPH. Whether this gene is functional is unknown and requires a separate study. The zinc-finger motif in the C-terminal also has some variability. The majority of VvWRKYs have C2H2, except for six VvWRKYs that have a C2HC zinc-finger motif, according to us and previous studies [37,44,45,47]. VvWRKY26 has an altered C2HY motif, which also casts doubt on the functionality of this gene.

3.4. WRKY Groups and Evolution Clades

Based on the number of WRKY domains and the type of zinc finger motif, there is an accepted classification of this family of transcription factors, as discussed in the introduction (I–IV). Seven major groups and subgroups have been identified in flowering plants: I, II (IIa, IIb, IIc, IId, IIe), and III [19]. Our results complement previous studies in grapes and show that 12 genes of WRKY group I, 43 genes of WRKY group II, 6 genes of WRKY group III, and VvWRKY26, which probably belongs to group IV because it lacks the C2H2 or C2HC zinc finger motif, are present in grapes. WRKY group II is the most numerous and, according to previous studies, is divided into several subgroups IIa–IIe. We have also shown the presence of two genes, VvWRKY17 and VvWRKY30, which do not belong to the known subgroups but are clearly assigned to group II.
Subtypes IIa–IIe were distinguished as a result of clustering of phylogenetic trees, which led to confusion between the concepts of “WRKY group based on number of domains and type of zinc finger motif” and “WRKY group based on evolution”. This led to confusion. For example, in a previous study, two VvWRKY genes were characterised as NG (non-group), although these genes had one WRKY domain and a C2H2 zinc-finger motif and should have been assigned to group II [37]. This situation arose because it was previously thought that the division into groups I–III reflected the evolution of this family of transcription factors. With the emergence of new data, it became clear that this was not the case. Recent papers investigating the evolution of WRKYs in plants repeatedly state that group II cannot be considered an evolutionary group. If it were a phylogenetically distinct group, then all subgroups (IIa-IIe) should be part of it, but this is not observed: groups IIa + IIb, IIc, and IIe + IId are phylogenetically distinct groups [16,90]. Furthermore, it has been shown that divergence within major evolutionary lineages occurs differently in dicots and monocots, and that the division into subgroups IIa–IIe is not universal for all plants [90].
Therefore, to avoid confusion, in our study we propose to introduce an evolutionary classification—phylogenetic clades A–E, which are distinguished as a result of analyses of the complete amino acid sequences of WRKY TFs (Figure 4) (for ease of comparison, we indicate which WRKY group corresponds to a particular clade). In our work, we have shown that clades B and D have a more complex evolution: there is a divergence into subclades B1–B4 and D1–D3. These subclades have different conserved motifs that may play a role in protein functionality. However, the topology based on phylogenetic analysis of domain regions alone does not reflect the true evolution (Figure 2 and Figure 4). For example, subclade B2 is present on both trees and differs from other genes by a modified WRKY motif (WRKYGKK instead of WRKYGQK). At the same time, subclade B1 is formed only when the whole protein is analysed, as it is characterised by a conserved motif 16 in the C-terminal region. The topology also differs at the clade level. Thus, in domain analysis, clades C and D form a branch, whereas in complete sequence analysis, clade C clusters with B, as confirmed by studies in 30 plant species, from green algae to Arabidopsis [97]. These findings suggest the need to analyse complete amino acid sequences, not just domain regions.

3.5. Chimeric WRKY

We also discovered some new chimeric WRKY TFs in grapes. These are genes that do not contain only the WRKY domain. Such proteins could be involved in the regulation of several seemingly disparate processes and have previously been found in different plant species [16,17,90,98,99]. There are 370 known domain architectures in the InterPro database for the WRKY domain (IPR003657) (as of 30 April 2024). There are 22,847 proteins with a single WRKY domain architecture and 5670 with two WRKY domains. The most common chimeric WRKY TF variant is the Zn cluster domain (IPR018872), followed by the WRKY domain (2975 proteins). In grapes, the Zn cluster domain chimeric protein is characteristic of subgroup IId (subclade E1), which is in agreement with other studies [16]. However, its functional role is not known. Domains such as disease resistance domain, kinase domain, or leucine-rich repeat, which are commonly found in plants, were not found in grapes. However, a structural coil motif was found that is characteristic of clade D. It is likely that these proteins are WRKYs of the immune signalling pathway.
The new chimeric proteins are VvWRKY17 and VvWRKY61. VvWRKY17 has an N-terminal signal peptide region followed by a non-cytoplasmic domain (according to the Phobius prediction). This WRKY transcription factor is likely to have a specific localisation. The localisation of WRKY in different cell compartments has been shown in Glycyrrhiza glabra [11]. This issue requires further investigation.
VvWRKY61 has a Frigida-like domain (IPR012474), followed by two WRKY domains. This family represents proteins similar to the FRIGIDA protein. This protein is located in the nucleus and is required for the regulation of flowering time [100]. AtWRKY75 in Arabidopsis has previously been shown to be involved in the regulation of flowering [101]. Chimeric WRKYs containing the ZF_SBP domain associated with flower development have also been found, and it has been suggested that such chimeric proteins may play an important role in flowering [90]. Thus, a VvWRKY61 TF with a Frigida-like domain and two WRKY domains may regulate flowering in grapes, which requires further investigation.

3.6. Grapevine WRKY Evolution

Our phylogenetic analysis of grapevine WRKY suggests the most likely evolutionary process for this family of transcription factors. There are many theories for the evolution of WRKY genes from the unicellular early green lineage to multicellular plants [11,18,19,24,91,97]. The ancestor of all WRKY genes is thought to be an N-terminal addition of a WRKY-like motif to the BED finger-like C2HC zinc finger domain [97]. Subsequently, gene duplication occurred in the charophyte green algae, and genes containing two WRKY domains were formed and found in the unicellular green alga Chlamydomonas reinhardtii [18,19,70]. More recently, the presence of a gene with a single domain belonging to clade D (IIb) was shown in the filamentous terrestrial alga Klebsormidium flaccidum [102]. Thus, clades C (group I) and D (IIb) are early lineages. Phylogenetic clades A (III), E (IIe and IId), and B (IIc) evolved from clade C (I) due to the loss of the NTWD [91,97]. However, the origin of clade D (IIb and IIa) remains unknown. There are two theories according to which clade D (IIb + IIa) evolved from a single-domain ancestor (IIa + IIb separate hypothesis) or from clade C (I) (Group I hypothesis) [97]. More recently, it has been shown that clade D is characterised by the presence of a V-type intron, whereas clades A, B, and E possess a conserved R-type intron present in CTWDs [91]. As a result, the authors lean towards the “IIa + IIb separate hypothesis”.
Our studies also confirm that the most likely evolutionary pathway was the origin of clade D from a common ancestor. Gene duplication is known to be one of the major evolutionary mechanisms generating new diversity and often leading to gene paralogy [103,104]. In grapevine, approximately 40% of WRKYs originate from tandem or segmental duplications [37]. Duplication events lead to an increase in the number of members of evolutionary clades and reflect an ongoing evolutionary process. There are other evolutionary processes, such as the gain or loss of introns. An intron has a complex structure (donor and acceptor splice sites, branch point, polypyrimidine tract, and appropriate splicing enhancers) and can be several thousand nucleotides long. Therefore, the emergence of new introns by the gradual accumulation of functional sub-elements is unlikely, and it has been shown that introns can also arise by segmental genomic duplication [105]. At the same time, intron loss has been shown to be more likely than intron gain [106], making intron gain rarer. In our study, we showed that clades D and C differ significantly in intron number and protein length from other evolutionary groups (Figure 5). If clade D had evolved from clade C as a result of the loss of NTWD, there would most likely have been a loss of some introns and a reduction in the length of the transcription factor. This is what we observe in genes from clades A, B, and E that evolved from clade C. And we observe the opposite situation: clade D is equivalent to clade C in protein length and number of introns. The WRKY genes from clade D have a large, conserved motif 7 and a coil in the N-terminal structural motif. Considering that the intron gain process is quite rare, the probability that clade D lost NTWD initially and then gained a new domain is quite low.
Based on the probability of evolutionary events, we suggest that there were two dynamic phases of complexity and simplification in the evolution of WRKY (Figure 6). The two clades C and D evolved from a common ancestor as a result of the complexity of the gene structure, then divergence occurred, and a characteristic motif (NTWD or Coil) emerged in each clade. This theory is supported by the fact that subclade D3 (IIa) evolved from clade D (IIb), which is also characterised by a reduction in length and number of introns (Table 3). The proposed evolutionary model is consistent with the fact that macroevolutionary patterns are characterised by the periodicity of opposing dynamics. There is a general pattern of evolution consisting of two distinct evolutionary phases: a short, explosive innovation phase leading to a dramatic increase in genome complexity, and a longer simplification phase leading to either loss of genetic material or adaptive ordering of the genome [106,107,108,109]. Quantitatively, genome evolution is dominated by reduction and simplification, followed by episodes of increasing complexity, which is also consistent with the proposed theory.

4. Materials and Methods

4.1. Genome Data

Two databases representing genomic assemblies of grapes were analysed: the National Center for Biotechnology Information (NCBI) and GrapeGenomics (https://grapegenomics.com/ accessed on 24 February 2024). The NCBI currently contains 19 genomic assemblies of varying quality for different grape cultivars: three assemblies have a chromosomal level, ten have a scaffold level, and six have a contig level. For this study, we analysed the genome assembly of V. vinifera cultivar Pinot Noir cl. PN40024 (GCA_030704535, August 2023), which is the reference for the grape [75], and the first assembly of V. vinifera cultivar Pinot Noir cl. PN40024 is known as 12X (GCA_000003745.2) [74]. The assembly GCA_030704535 has two annotations (RefSeq and GenBank), which are included in our analyses. No predicted proteins are represented for other chromosomal-level assemblies, and thus they were excluded from the analyses. Furthermore, lower-level assemblies (scaffolds or contigs) were not included in the analysis.
Twenty-three genomic assemblies for different cultivars were available in the second database, Grape Genomics. Three assemblies with diploid-phased chromosomal levels and predicted protein data were selected from this database: V. vinifera cultivar Cabernet Franc cl. 04, V. vinifera cultivar Cabernet Sauvignon cl. 08, and V. vinifera cultivar Pinot Noir cl. FPS123 [110]. Also, the reference assembly of V. vinifera cultivar Pinot Noir cl. PN40024, v. 5 was represented in this database [75,76]. Only those assemblies that were at the chromosome level and had protein predictions were included in the analysis. Thus, six genomic assemblies and seven protein annotations were included in the analyses.

4.2. Genome-Wide Analyses and Identification of WRKY

The genome-wide analyses were performed through six grape genome assemblies. For each genomic assembly, all predicted protein sequences and CDS data were downloaded and further analysed. The identification of WRKY proteins was based on finding the WRKY domain signature using InterProScan software, version 5.63-95.0 [111] and the Pfam protein family database [112]. After that, all proteins that had the WRKY DNA-binding domain (accession number PF03106 according to the Pfam database) were checked for containing other domains, and functional families (FF) were detected using the Pfam, InterPro, and CATH databases [113,114]. Zinc finger, LXXLL, and LXLXLX motifs were identified manually.
The exon-intron structure of each WRKY transcription factor-encoding gene was detected using the CDS description and genome annotation of accordance grapevine cultivars from the NCBI and GrapeGenomics databases. A motif analysis was performed using the MEME online tool in classic mode (https://meme-suite.org, accessed on 15 April 2024). The location information of the grape WRKY genes was obtained from the Genome Data Viewer available in NCBI and GrapeGenomics for each assembly. The amino acid diversity analyses of WRKY genes were conducted using MEGA software (version 11.0.13) [115,116]. The average evolutionary divergence over different grape cultivars for each WRKY gene was estimated as the number of amino acid substitutions per site from averaging over all sequence pairs with a bootstrap procedure (1000 replicates). The analysis was conducted using the Poisson correction model [117].
To search for accordance among WRKY proteins from PlantTFDB, NCBI, Uniprot, and Ensembl databases, we used similarity and homology analyses between protein sequences performed using local BLAST v. 2.9.0 with a threshold E-value of 1 × 10−10 [118,119,120,121,122,123].

4.3. Phylogenetic Analyses

A phylogenetic analysis was conducted using IQ-TREE software, version 2.3.0 [124]. The dataset comprised all identified WRKY protein sequences of V. vinifera from six genome assemblies. The multiple alignment of protein sequences was performed using the MAFFT online service (version 7.0) with the scoring matrix BLOSUM62 [125,126]. The best amino acid substitution model, according to BIC, was estimated with ModelFinder [127]. The consistency of the Maximum Likelihood (ML) tree was validated by an ultrafast bootstrap value of 1000 [128]. The final phylogenetic tree was visualised with FigTree version 1.4.4. The rooting of trees was according to the midpoint.

5. Conclusions

Sixty-two WRKY transcription factor genes have been identified in Vitis vinifera grapes. The structure of each sVvWRKY gene was studied, and its chromosomal location was determined. A new numbering of the genes according to their chromosomal location in the reference genome of V. vinifera cv. Pinot Noir cl. PN40024, v.5 was suggested. Inter-varietal amino acid variability was revealed, reaching 5% for some genes. Chimeric VvWRKY genes were also found, which may have specific regulatory functions. Phylogenetic analysis shows that the evolution of WRKY genes went through a phase of complication (intron gain and formation of genes with two domains) and a phase of simplification (loss of introns and reduction of protein length). The data obtained indicate a high functional flexibility of this family, which is consistent with the wide range of processes in which WRKY transcription factors are involved.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms25116241/s1.

Author Contributions

Conceptualisation, E.V. and P.K.; methodology, E.V.; software, A.S. and E.V.; formal analysis, E.V.; writing—original draft preparation, E.V. and A.S.; writing—review and editing, E.V. and P.K.; visualisation, A.S. and E.V.; supervision, S.D. All authors have read and agreed to the published version of the manuscript.

Funding

The study is supported by the Russian Science Foundation Grant no. 23-76-10013.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Publicly available datasets were analyzed in this study. This data can be found here: https://www.ncbi.nlm.nih.gov (accessed on 22 February 2024). The accession numbers of genome data are represented in the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Joshi, R.; Wani, S.H.; Singh, B.; Bohra, A.; Dar, Z.A.; Lone, A.A.; Pareek, A.; Singla-Pareek, S.L. Transcription Factors and Plants Response to Drought Stress: Current Understanding and Future Directions. Front. Plant Sci. 2016, 7, 1029. [Google Scholar]
  2. Joshi, R.; Anwar, K.; Das, P.; Singla-Pareek, S.; Pareek, A. Overview of Methods for Assessing Salinity and Drought Tolerance of Transgenic Wheat Lines. Methods Mol. Biol. 2017, 1679, 83–95. [Google Scholar] [CrossRef]
  3. Takahashi, F.; Shinozaki, K. Long-Distance Signaling in Plant Stress Response. Curr. Opin. Plant Biol. 2019, 47, 106–111. [Google Scholar] [CrossRef]
  4. Bakshi, M.; Oelmüller, R. WRKY Transcription Factors. Plant Signal. Behav. 2014, 9, e27700. [Google Scholar] [CrossRef]
  5. Cai, R.; Zhao, Y.; Wang, Y.; Lin, Y.; Peng, X.; Li, Q.; Chang, Y.; Jiang, H.; Xiang, Y.; Cheng, B. Overexpression of a Maize WRKY58 Gene Enhances Drought and Salt Tolerance in Transgenic Rice. Plant Cell Tissue Organ Cult. (PCTOC) 2014, 119, 565–577. [Google Scholar] [CrossRef]
  6. 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, 1–16. [Google Scholar] [CrossRef]
  7. Zhu, D.; Ma, Q.; Hao, J.; Liu, X. Function Exploration of Grape WRKY Family Proteins Under Abiotic Stresses. Biotechnol. Bull. 2016, 32, 77–83. [Google Scholar] [CrossRef]
  8. Gao, H.; Wang, Y.; Xu, P.; Zhang, Z. Overexpression of a WRKY Transcription Factor TaWRKY2 Enhances Drought Stress Tolerance in Transgenic Wheat. Front. Plant Sci. 2018, 9, 997. [Google Scholar] [CrossRef]
  9. Wang, C.-T.; Ru, J.-N.; Liu, Y.-W.; Yang, J.-F.; Li, M.; Xu, Z.-S.; Fu, J.-D. The Maize WRKY Transcription Factor ZmWRKY40 Confers Drought Resistance in Transgenic Arabidopsis. Int. J. Mol. Sci. 2018, 19, 2580. [Google Scholar] [CrossRef]
  10. Wang, F.-P.; Zhao, P.-P.; Zhang, L.; Zhai, H.; Du, Y.-P. Functional Characterization of WRKY46 in Grape and Its Putative Role in the Interaction between Grape and Phylloxera (Daktulosphaira vitifoliae). Hortic. Res. 2019, 6, 102. [Google Scholar] [CrossRef]
  11. Goyal, P.; Manzoor, M.M.; Vishwakarma, R.A.; Sharma, D.; Dhar, M.K.; Gupta, S. A Comprehensive Transcriptome-Wide Identification and Screening of WRKY Gene Family Engaged in Abiotic Stress in Glycyrrhiza Glabra. Sci. Rep. 2020, 10, 373. [Google Scholar] [CrossRef]
  12. Huang, Y.; Chen, F.; Chai, M.; Xi, X.; Zhu, W.; Qi, J.; Liu, K.; Ma, S.; Su, H.; Tian, Y. Ectopic Overexpression of Pineapple Transcription Factor AcWRKY31 Reduces Drought and Salt Tolerance in Rice and Arabidopsis. Int. J. Mol. Sci. 2022, 23, 6269. [Google Scholar] [CrossRef]
  13. Wang, Y.; Dong, B.; Wang, N.; Zheng, Z.; Yang, L.; Zhong, S.; Fang, Q.; Xiao, Z.; Zhao, H. A WRKY Transcription Factor PmWRKY57 from Prunus Mume Improves Cold Tolerance in Arabidopsis Thaliana. Mol. Biotechnol. 2023, 65, 1359–1368. [Google Scholar] [CrossRef]
  14. Wani, S.H.; Anand, S.; Singh, B.; Bohra, A.; Joshi, R. WRKY Transcription Factors and Plant Defense Responses: Latest Discoveries and Future Prospects. Plant Cell Rep. 2021, 40, 1071–1085. [Google Scholar] [CrossRef]
  15. Khoso, M.A.; Hussain, A.; Ritonga, F.N.; Ali, Q.; Channa, M.M.; Alshegaihi, R.M.; Meng, Q.; Ali, M.; Zaman, W.; Brohi, R.D. WRKY Transcription Factors (TFs): Molecular Switches to Regulate Drought, Temperature, and Salinity Stresses in Plants. Front. Plant Sci. 2022, 13, 1039329. [Google Scholar] [CrossRef]
  16. Goyal, P.; Devi, R.; Verma, B.; Hussain, S.; Arora, P.; Tabassum, R.; Gupta, S. WRKY Transcription Factors: Evolution, Regulation, and Functional Diversity in Plants. Protoplasma 2023, 260, 331–348. [Google Scholar] [CrossRef]
  17. 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]
  18. Zhang, Y.; Wang, L. The WRKY Transcription Factor Superfamily: Its Origin in Eukaryotes and Expansion in Plants. BMC Evol. Biol. 2005, 5, 1–12. [Google Scholar] [CrossRef]
  19. 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]
  20. Xie, Z.; Zhang, Z.-L.; Zou, X.; Huang, J.; Ruas, P.; Thompson, D.; Shen, Q.J. Annotations and Functional Analyses of the RiceWRKYGene Superfamily Reveal Positive and Negative Regulators of Abscisic Acid Signaling in Aleurone Cells. Plant Physiol. 2005, 137, 176–189. [Google Scholar] [CrossRef]
  21. Zhang, T.; Tan, D.; Zhang, L.; Zhang, X.; Han, Z. Phylogenetic Analysis and Drought-Responsive Expression Profiles of the WRKY Transcription Factor Family in Maize. Agric. Gene 2017, 3, 99–108. [Google Scholar] [CrossRef]
  22. Goel, R.; Pandey, A.; Trivedi, P.K.; Asif, M.H. Genome-Wide Analysis of the Musa WRKY Gene Family: Evolution and Differential Expression during Development and Stress. Front. Plant Sci. 2016, 7, 299. [Google Scholar]
  23. Zhou, Q.; Tian, A.; Zou, H.; Xie, Z.; Lei, G.; Huang, J.; Wang, C.; Wang, H.; Zhang, J.; Chen, S. Soybean WRKY-type Transcription Factor Genes, GmWRKY13, GmWRKY21, AndGmWRKY54, Confer Differential Tolerance to Abiotic Stresses in TransgenicArabidopsisplants. Plant Biotechnol. J. 2008, 6, 486–503. [Google Scholar] [CrossRef]
  24. Xu, H.; Watanabe, K.A.; Zhang, L.; Shen, Q.J. WRKY Transcription Factor Genes in Wild Rice Oryza Nivara. DNA Res. 2016, 23, 311–323. [Google Scholar] [CrossRef]
  25. Yang, B.; Jiang, Y.; Rahman, M.H.; Deyholos, M.K.; Kav, N.N. Identification and Expression Analysis of WRKY Transcription Factor Genes in Canola (Brassica napus L.) in Response to Fungal Pathogens and Hormone Treatments. BMC Plant Biol. 2009, 9, 68. [Google Scholar] [CrossRef]
  26. He, Y.; Mao, S.; Gao, Y.; Zhu, L.; Wu, D.; Cui, Y.; Li, J.; Qian, W. Genome-Wide Identification and Expression Analysis of WRKY Transcription Factors under Multiple Stresses in Brassica Napus. PLoS ONE 2016, 11, e0157558. [Google Scholar] [CrossRef]
  27. Chen, H.; Wang, Y.; Liu, J.; Zhao, T.; Yang, C.; Ding, Q.; Zhang, Y.; Mu, J.; Wang, D. Identification of WRKY Transcription Factors Responding to Abiotic Stresses in Brassica napus L. Planta 2022, 255, 3. [Google Scholar] [CrossRef]
  28. Jiang, J.; Ma, S.; Ye, N.; Jiang, M.; Cao, J.; Zhang, J. WRKY Transcription Factors in Plant Responses to Stresses. J. Integr. Plant Biol. 2017, 59, 86–101. [Google Scholar] [CrossRef]
  29. 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]
  30. 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, 231. [Google Scholar] [CrossRef]
  31. Song, Y.; Cui, H.; Shi, Y.; Xue, J.; Ji, C.; Zhang, C.; Yuan, L.; Li, R. Genome-Wide Identification and Functional Characterization of the Camelina Sativa WRKY Gene Family in Response to Abiotic Stress. BMC Genom. 2020, 21, 786. [Google Scholar] [CrossRef]
  32. Du, Z.; You, S.; Zhao, X.; Xiong, L.; Li, J. Genome-Wide Identification of WRKY Genes and Their Responses to Chilling Stress in Kandelia Obovata. Front. Genet. 2022, 13, 875316. [Google Scholar] [CrossRef]
  33. Rai, G.K.; Mishra, S.; Chouhan, R.; Mushtaq, M.; Chowdhary, A.A.; Rai, P.K.; Kumar, R.R.; Kumar, P.; Perez-Alfocea, F.; Colla, G. Plant Salinity Stress, Sensing, and Its Mitigation through WRKY. Front. Plant Sci. 2023, 14, 1238507. [Google Scholar] [CrossRef]
  34. 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]
  35. Zhang, J.; Peng, Y.; Guo, Z. Constitutive Expression of Pathogen-Inducible OsWRKY31 Enhances Disease Resistance and Affects Root Growth and Auxin Response in Transgenic Rice Plants. Cell Res. 2008, 18, 508–521. [Google Scholar] [CrossRef]
  36. Shahzad, R.; Harlina, P.; Cong-Hua, X.; Ewas, M.; Nishawy, E.; Zhenyuan, P.; Foly, M. Overexpression of Potato Transcription Factor (StWRKY1) Conferred Resistance to Phytophthora Infestans and Improved Tolerance to Water Stress. Plant Omics 2016, 9, 149–158. [Google Scholar] [CrossRef]
  37. 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]
  38. Alston, J.M.; Sambucci, O. Grapes in the World Economy. In Grape Genome; Springer: Cham, Switzerland, 2019; pp. 1–24. [Google Scholar] [CrossRef]
  39. Naulleau, A.; Gary, C.; Prévot, L.; Hossard, L. Evaluating Strategies for Adaptation to Climate Change in Grapevine Production–A Systematic Review. Front. Plant Sci. 2021, 11, 607859. [Google Scholar] [CrossRef]
  40. Jones, G.V.; White, M.A.; Cooper, O.R.; Storchmann, K. Climate Change and Global Wine Quality. Clim. Chang. 2005, 73, 319–343. [Google Scholar] [CrossRef]
  41. Schultz, H.R. Climate Change and Viticulture: Research Needs for Facing the Future. J. Wine Res. 2010, 21, 113–116. [Google Scholar] [CrossRef]
  42. Medrano, H.; Tomás, M.; Martorell, S.; Escalona, J.-M.; Pou, A.; Fuentes, S.; Flexas, J.; Bota, J. Improving Water Use Efficiency of Vineyards in Semi-Arid Regions. A Review. Agron. Sustain. Dev. 2015, 35, 499–517. [Google Scholar] [CrossRef]
  43. Mosedale, J.R.; Abernethy, K.E.; Smart, R.E.; Wilson, R.J.; Maclean, I.M.D. Climate Change Impacts and Adaptive Strategies: Lessons from the Grapevine. Glob. Chang. Biol. 2016, 22, 3814–3828. [Google Scholar] [CrossRef]
  44. Wang, L.; Zhu, W.; Fang, L.; Sun, X.; Su, L.; Liang, Z.; Wang, N.; Londo, J.P.; Li, S.; Xin, H. Genome-Wide Identification of WRKY Family Genes and Their Response to Cold Stress in Vitis vinifera. BMC Plant Biol. 2014, 14, 103. [Google Scholar] [CrossRef] [PubMed]
  45. Guo, C.; Guo, R.; Xu, X.; Gao, M.; Li, X.; Song, J.; Zheng, Y.; Wang, X. Evolution and Expression Analysis of the Grape (Vitis vinifera L.) WRKY Gene Family. J. Exp. Bot. 2014, 65, 1513–1528. [Google Scholar] [CrossRef]
  46. Huang, T.; Yang, J.; Yu, D.; Han, X.; Wang, X. Bioinformatics Analysis of WRKY Transcription Factors in Grape and Their Potential Roles Prediction in Sugar and Abscisic Acid Signaling Pathway. J. Plant Biochem. Biotechnol. 2021, 30, 67–80. [Google Scholar] [CrossRef]
  47. Zhang, Y.; Feng, J. Identification and Characterization of the Grape WRKY Family. BioMed Res. Int. 2014, 2014, 787680. [Google Scholar] [CrossRef] [PubMed]
  48. Agudelo-Romero, A.-P.; Erban, A.; Rego, C.; Carbonell-Bejerano, P.; Nascimento, T.; Sousa, L.; Martínez-Zapater, J.M.; Kopka, J.; Fortes, A.M. Transcriptome and Metabolome Reprogramming in Vitis vinifera Cv. Trincadeira Berries upon Infection with Botrytis Cinerea. J. Exp. Bot. 2015, 66, 1769–1785. [Google Scholar] [CrossRef] [PubMed]
  49. Wu, W.; Fu, P.; Lu, J. Grapevine WRKY Transcription Factors. Fruit Res. 2022, 2, 1–8. [Google Scholar] [CrossRef]
  50. Hao, J.; Ma, Q.; Hou, L.; Zhao, F.; Xin, L. VvWRKY13 Enhances ABA Biosynthesis in Vitis vinifera. Acta Soc. Bot. Pol. 2017, 86, 3546. [Google Scholar] [CrossRef]
  51. Zhao, J.; Zhang, X.; Guo, R.; Wang, Y.; Guo, C.; Li, Z.; Chen, Z.; Gao, H.; Wang, X. Over-Expression of a Grape WRKY Transcription Factor Gene, VlWRKY48, in Arabidopsis Thaliana Increases Disease Resistance and Drought Stress Tolerance. Plant Cell Tissue Organ Cult. (PCTOC) 2018, 132, 359–370. [Google Scholar] [CrossRef]
  52. Guo, R.; Qiao, H.; Zhao, J.; Wang, X.; Tu, M.; Guo, C.; Wan, R.; Li, Z.; Wang, X. The Grape VlWRKY3 Gene Promotes Abiotic and Biotic Stress Tolerance in Transgenic Arabidopsis thaliana. Front. Plant Sci. 2018, 9, 545. [Google Scholar] [CrossRef]
  53. Zhu, D.; Hou, L.; Xiao, P.; Guo, Y.; Deyholos, M.K.; Liu, X. VvWRKY30, a Grape WRKY Transcription Factor, Plays a Positive Regulatory Role under Salinity Stress. Plant Sci. 2019, 280, 132–142. [Google Scholar] [CrossRef]
  54. Hou, L.; Fan, X.; Hao, J.; Liu, G.; Zhang, Z.; Liu, X. Negative Regulation by Transcription Factor VvWRKY13 in Drought Stress of Vitis vinifera L. Plant Physiol. Biochem. 2020, 148, 114–121. [Google Scholar] [CrossRef]
  55. Liu, W.; Liang, X.; Cai, W.; Wang, H.; Liu, X.; Cheng, L.; Song, P.; Luo, G.; Han, D. Isolation and Functional Analysis of VvWRKY28, a Vitis vinifera WRKY Transcription Factor Gene, with Functions in Tolerance to Cold and Salt Stress in Transgenic Arabidopsis thaliana. Int. J. Mol. Sci. 2022, 23, 13418. [Google Scholar] [CrossRef]
  56. Amato, A.; Cavallini, E.; Zenoni, S.; Finezzo, L.; Begheldo, M.; Ruperti, B.; Tornielli, G.B. A Grapevine TTG2-Like WRKY Transcription Factor Is Involved in Regulating Vacuolar Transport and Flavonoid Biosynthesis. Front. Plant Sci. 2017, 7, 1979. [Google Scholar]
  57. Merz, P.R.; Moser, T.; Höll, J.; Kortekamp, A.; Buchholz, G.; Zyprian, E.; Bogs, J. The Transcription Factor VvWRKY33 Is Involved in the Regulation of Grapevine (Vitis vinifera) Defense against the Oomycete Pathogen Plasmopara Viticola. Physiol. Plant. 2015, 153, 365–380. [Google Scholar] [CrossRef]
  58. Ma, Q.; Zhang, G.; Hou, L.; Wang, W.; Hao, J.; Liu, X. Vitis vinifera VvWRKY13 Is an Ethylene Biosynthesis-Related Transcription Factor. Plant Cell Rep. 2015, 34, 1593–1603. [Google Scholar] [CrossRef]
  59. Liu, H.; Yang, W.; Liu, D.; Han, Y.; Zhang, A.; Li, S. Ectopic Expression of a Grapevine Transcription Factor VvWRKY11 Contributes to Osmotic Stress Tolerance in Arabidopsis. Mol. Biol. Rep. 2011, 38, 417–427. [Google Scholar] [CrossRef]
  60. Guillaumie, S.; Mzid, R.; Méchin, V.; Léon, C.; Hichri, I.; Destrac-Irvine, A.; Trossat-Magnin, C.; Delrot, S.; Lauvergeat, V. The Grapevine Transcription Factor WRKY2 Influences the Lignin Pathway and Xylem Development in Tobacco. Plant Mol. Biol. 2010, 72, 215–234. [Google Scholar] [CrossRef]
  61. Marchive, C.; Mzid, R.; Deluc, L.G.; Barrieu, F.; Pirrello, J.; Gauthier, A.; Corio-Costet, M.-F.; Regad, F.; Cailleteau, B.; Hamdi, S.; et al. Isolation and Characterization of a Vitis vinifera Transcription Factor, VvWRKY1, and Its Effect on Responses to Fungal Pathogens in Transgenic Tobacco Plants. J. Exp. Bot. 2007, 58, 1999–2010. [Google Scholar] [CrossRef]
  62. Yongmei, C.; Zhen, Z.; Dan, Z.; Jie, H.; Lixia, H.; Xin, L. VvWRKY13 from Vitis vinifera Negatively Modulates Salinity Tolerance. Plant Cell Tissue Organ Cult. (PCTOC) 2019, 139, 455–465. [Google Scholar] [CrossRef]
  63. Mzid, R.; Zorrig, W.; Ben Ayed, R.; Ben Hamed, K.; Ayadi, M.; Damak, Y.; Lauvergeat, V.; Hanana, M. The Grapevine VvWRKY2 Gene Enhances Salt and Osmotic Stress Tolerance in Transgenic Nicotiana Tabacum. 3 Biotech 2018, 8, 2777. [Google Scholar] [CrossRef]
  64. Marchive, C.; Léon, C.; Kappel, C.; Coutos-Thévenot, P.; Corio-Costet, M.-F.; Delrot, S.; Lauvergeat, V. Over-Expression of VvWRKY1 in Grapevines Induces Expression of Jasmonic Acid Pathway-Related Genes and Confers Higher Tolerance to the Downy Mildew. PLoS ONE 2013, 8, e54185. [Google Scholar] [CrossRef]
  65. Amato, A.; Cavallini, E.; Walker, A.R.; Pezzotti, M.; Bliek, M.; Quattrocchio, F.; Koes, R.; Ruperti, B.; Bertini, E.; Zenoni, S.; et al. The MYB5-Driven MBW Complex Recruits a WRKY Factor to Enhance the Expression of Targets Involved in Vacuolar Hyper-Acidification and Trafficking in Grapevine. Plant J. 2019, 99, 1220–1241. [Google Scholar] [CrossRef]
  66. Jiang, J.; Xi, H.; Dai, Z.; Lecourieux, F.; Yuan, L.; Liu, X.; Patra, B.; Wei, Y.; Li, S.; Wang, L. VvWRKY8 Represses Stilbene Synthase Genes through Direct Interaction with VvMYB14 to Control Resveratrol Biosynthesis in Grapevine. J. Exp. Bot. 2019, 70, 715–729. [Google Scholar] [CrossRef]
  67. Zhang, Z.; Jiang, C.; Chen, C.; Su, K.; Lin, H.; Zhao, Y.; Guo, Y. VvWRKY5 Enhances White Rot Resistance in Grape by Promoting the Jasmonic Acid Pathway. Hortic. Res. 2023, 10, uhad172. [Google Scholar] [CrossRef]
  68. Wang, F.-P.; Zhao, P.-P.; Zhang, L.; Zhai, H.; Abid, M.; Du, Y.-P. The VvWRKY37 Regulates Bud Break in Grape Vine through ABA-Mediated Signaling Pathways. Front. Plant Sci. 2022, 13, 929892. [Google Scholar] [CrossRef]
  69. Wei, Y.; Meng, N.; Wang, Y.; Cheng, J.; Duan, C.; Pan, Q. Transcription Factor VvWRKY70 Inhibits Both Norisoprenoid and Flavonol Biosynthesis in Grape. Plant Physiol. 2023, 193, 2055–2070. [Google Scholar] [CrossRef]
  70. 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]
  71. Ramiro, D.; Jalloul, A.; Petitot, A.-S.; Grossi De Sá, M.F.; Maluf, M.P.; Fernandez, D. Identification of Coffee WRKY Transcription Factor Genes and Expression Profiling in Resistance Responses to Pathogens. Tree Genet. Genomes 2010, 6, 767–781. [Google Scholar] [CrossRef]
  72. Karkute, S.G.; Gujjar, R.S.; Rai, A.; Akhtar, M.; Singh, M.; Singh, B. Genome Wide Expression Analysis of WRKY Genes in Tomato (Solanum lycopersicum) under Drought Stress. Plant Gene 2018, 13, 8–17. [Google Scholar] [CrossRef]
  73. Chen, C.; Chen, X.; Han, J.; Lu, W.; Ren, Z. Genome-Wide Analysis of the WRKY Gene Family in the Cucumber Genome and Transcriptome-Wide Identification of WRKY Transcription Factors That Respond to Biotic and Abiotic Stresses. BMC Plant Biol. 2020, 20, 443. [Google Scholar] [CrossRef] [PubMed]
  74. The French–Italian Public Consortium for Grapevine Genome Characterization. The Grapevine Genome Sequence Suggests Ancestral Hexaploidization in Major Angiosperm Phyla. Nature 2007, 449, 463–467. [Google Scholar] [CrossRef] [PubMed]
  75. Shi, X.; Cao, S.; Wang, X.; Huang, S.; Wang, Y.; Liu, Z.; Liu, W.; Leng, X.; Peng, Y.; Wang, N. The Complete Reference Genome for Grapevine (Vitis viniferaL.) Genetics and Breeding. Hortic. Res. 2023, 10, uhad061. [Google Scholar] [CrossRef] [PubMed]
  76. Velt, A.; Frommer, B.; Blanc, S.; Holtgräwe, D.; Duchêne, É.; Dumas, V.; Grimplet, J.; Hugueney, P.; Kim, C.; Lahaye, M. An Improved Reference of the Grapevine Genome Reasserts the Origin of the PN40024 Highly Homozygous Genotype. G3: Genes Genomes Genet. 2023, 13, jkad067. [Google Scholar] [PubMed]
  77. Jakoby, M.; Weisshaar, B.; Dröge-Laser, W.; Vicente-Carbajosa, J.; Tiedemann, J.; Kroj, T.; Parcy, F. BZIP Transcription Factors in Arabidopsis. Trends Plant Sci. 2002, 7, 106–111. [Google Scholar] [CrossRef] [PubMed]
  78. Mzid, R.; Marchive, C.; Blancard, D.; Deluc, L.; Barrieu, F.; Corio-Costet, M.; Drira, N.; Hamdi, S.; Lauvergeat, V. Overexpression of VvWRKY2 in Tobacco Enhances Broad Resistance to Necrotrophic Fungal Pathogens. Physiol. Plant. 2007, 131, 434–447. [Google Scholar] [CrossRef] [PubMed]
  79. Vannozzi, A.; Wong, D.C.J.; Höll, J.; Hmmam, I.; Matus, J.T.; Bogs, J.; Ziegler, T.; Dry, I.; Barcaccia, G.; Lucchin, M. Combinatorial Regulation of Stilbene Synthase Genes by WRKY and MYB Transcription Factors in Grapevine (Vitis vinifera L.). Plant Cell Physiol. 2018, 59, 1043–1059. [Google Scholar] [CrossRef] [PubMed]
  80. Wang, X.; Tu, M.; Wang, D.; Liu, J.; Li, Y.; Li, Z.; Wang, Y.; Wang, X. CRISPR/Cas9-mediated Efficient Targeted Mutagenesis in Grape in the First Generation. Plant Biotechnol. J. 2018, 16, 844–855. [Google Scholar] [CrossRef]
  81. Ma, T.; Chen, S.; Liu, J.; Fu, P.; Wu, W.; Song, S.; Gao, Y.; Ye, W.; Lu, J. Plasmopara Viticola Effector PvRXLR111 Stabilizes VvWRKY40 to Promote Virulence. Mol. Plant Pathol. 2021, 22, 231–242. [Google Scholar] [CrossRef]
  82. Yue, M.; Jiang, L.; Zhang, N.; Zhang, L.; Liu, Y.; Wang, Y.; Li, M.; Lin, Y.; Zhang, Y.; Zhang, Y.; et al. Importance of FaWRKY71 in Strawberry (Fragaria × Ananassa) Fruit Ripening. Int. J. Mol. Sci. 2022, 23, 12483. [Google Scholar] [CrossRef] [PubMed]
  83. Hou, L.; Wang, W.-J.; Guo, X.-P.; Peining, F.; Xinye, L. Gene Cloning and Expression Analysis of Three WRKYs in Vitis vinifera L. Plant Physiol. J. 2013, 49, 289–296. [Google Scholar]
  84. Feng, J.; Zhang, W.; Wang, W.; Nieuwenhuizen, N.J.; Atkinson, R.G.; Gao, L.; Hu, H.; Zhao, W.; Ma, R.; Zheng, H.; et al. Integrated Transcriptomic and Proteomic Analysis Identifies Novel Regulatory Genes Associated with Plant Growth Regulator-Induced Astringency in Grape Berries. J. Agric. Food Chem. 2024, 72, 4433–4447. [Google Scholar] [CrossRef] [PubMed]
  85. Kim, M.; Xi, H.; Park, J. Genome-Wide Comparative Analyses of GATA Transcription Factors among 19 Arabidopsis Ecotype Genomes: Intraspecific Characteristics of GATA Transcription Factors. PLoS ONE 2021, 16, e0252181. [Google Scholar] [CrossRef] [PubMed]
  86. Kan, J.; Gao, G.; He, Q.; Gao, Q.; Jiang, C.; Ahmar, S.; Liu, J.; Zhang, J.; Yang, P. Genome-Wide Characterization of WRKY Transcription Factors Revealed Gene Duplication and Diversification in Populations of Wild to Domesticated Barley. Int. J. Mol. Sci. 2021, 22, 5354. [Google Scholar] [CrossRef] [PubMed]
  87. Maeo, K.; Hayashi, S.; Kojima-Suzuki, H.; Morikami, A.; Nakamura, K. Role of Conserved Residues of the WRKY Domain in the DNA-Binding of Tobacco WRKY Family Proteins. Biosci. Biotechnol. Biochem. 2001, 65, 2428–2436. [Google Scholar] [CrossRef]
  88. Duan, M.-R.; Nan, J.; Liang, Y.-H.; Mao, P.; Lu, L.; Li, L.; Wei, C.; Lai, L.; Li, Y.; Su, X.-D. DNA Binding Mechanism Revealed by High Resolution Crystal Structure of Arabidopsis Thaliana WRKY1 Protein. Nucleic Acids Res. 2007, 35, 1145–1154. [Google Scholar] [CrossRef] [PubMed]
  89. 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]
  90. 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]
  91. Chen, X.; Li, C.; Wang, H.; Guo, Z. WRKY Transcription Factors: Evolution, Binding, and Action. Phytopathol. Res. 2019, 1, 13. [Google Scholar] [CrossRef]
  92. van Verk, M.C.; Pappaioannou, D.; Neeleman, L.; Bol, J.F.; Linthorst, H.J.M. A Novel WRKY Transcription Factor Is Required for Induction of PR-1a Gene Expression by Salicylic Acid and Bacterial Elicitors. Plant Physiol. 2008, 146, 1983–1995. [Google Scholar] [CrossRef] [PubMed]
  93. Zheng, J.; Liu, F.; Zhu, C.; Li, X.; Dai, X.; Yang, B.; Zou, X.; Ma, Y. Identification, Expression, Alternative Splicing and Functional Analysis of Pepper WRKY Gene Family in Response to Biotic and Abiotic Stresses. PLoS ONE 2019, 14, e0219775. [Google Scholar] [CrossRef] [PubMed]
  94. 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, 471. [Google Scholar] [CrossRef] [PubMed]
  95. Jimmy, J.L.; Babu, S. Variations in the Structure and Evolution of Rice WRKY Genes in Indica and Japonica Genotypes and Their Co-Expression Network in Mediating Disease Resistance. Evol. Bioinform Online 2019, 15, 117693431985772. [Google Scholar] [CrossRef] [PubMed]
  96. Khuman, A.; Arora, S.; Makkar, H.; Patel, A.; Chaudhary, B. Extensive Intragenic Divergences amongst Ancient WRKY Transcription Factor Gene Family Is Largely Associated with Their Functional Diversity in Plants. Plant Gene 2020, 22, 100222. [Google Scholar] [CrossRef]
  97. 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]
  98. 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]
  99. Shende, R.; Shinde, R.; Damse, D.; Dande, P. Exploring Biotic and Abiotic Responses in Plants: A Systems Biology Perspective on the Role of WRKY Transcription Factors. Pharma Innov. J. 2023, 12, 2102–2112. [Google Scholar]
  100. Johanson, U.; West, J.; Lister, C.; Michaels, S.; Amasino, R.; Dean, C. Molecular Analysis of FRIGIDA, a Major Determinant of Natural Variation in Arabidopsis Flowering Time. Science 2000, 290, 344–347. [Google Scholar] [CrossRef]
  101. Zhang, L.; Chen, L.; Yu, D. Transcription Factor WRKY75 Interacts with DELLA Proteins to Affect Flowering. Plant Physiol. 2018, 176, 790–803. [Google Scholar] [CrossRef]
  102. Hori, K.; Maruyama, F.; Fujisawa, T.; Togashi, T.; Yamamoto, N.; Seo, M.; Sato, S.; Yamada, T.; Mori, H.; Tajima, N. Klebsormidium Flaccidum Genome Reveals Primary Factors for Plant Terrestrial Adaptation. Nat. Commun. 2014, 5, 3978. [Google Scholar] [CrossRef] [PubMed]
  103. Cannon, S.; Mitra, A.; Baumgarten, A.; Young, N.; May, G. The Roles of Segmental and Tandem Gene Duplication in the Evolution of Large Gene Families in Arabidopsis Thaliana. BMC Plant Biol. 2004, 4, 10. [Google Scholar] [CrossRef] [PubMed]
  104. Wang, Y.; Wang, X.; Paterson, A.H. Genome and Gene Duplications and Gene Expression Divergence: A View from Plants. Ann. N. Y. Acad. Sci. 2012, 1256, 1–14. [Google Scholar] [CrossRef] [PubMed]
  105. Hellsten, U.; Aspden, J.L.; Rio, D.C.; Rokhsar, D.S. A Segmental Genomic Duplication Generates a Functional Intron. Nat. Commun. 2011, 2, 454. [Google Scholar] [CrossRef] [PubMed]
  106. Lin, H.; Zhu, W.; Silva, J.C.; Gu, X.; Buell, C.R. Intron Gain and Loss in Segmentally Duplicated Genes in Rice. Genome Biol. 2006, 7, 1–11. [Google Scholar] [CrossRef] [PubMed]
  107. Wolf, Y.I.; Koonin, E.V. Genome Reduction as the Dominant Mode of Evolution. BioEssays 2013, 35, 829–837. [Google Scholar] [CrossRef] [PubMed]
  108. O’Malley, M.A.; Wideman, J.G.; Ruiz-Trillo, I. Losing Complexity: The Role of Simplification in Macroevolution. Trends Ecol. Evol. 2016, 31, 608–621. [Google Scholar] [CrossRef] [PubMed]
  109. Domazet-Lošo, M.; Široki, T.; Šimičević, K.; Domazet-Lošo, T. Macroevolutionary Dynamics of Gene Family Gain and Loss along Multicellular Eukaryotic Lineages. Nat. Commun. 2024, 15, 2663. [Google Scholar] [CrossRef]
  110. Minio, A.; Cochetel, N.; Vondras, A.M.; Massonnet, M.; Cantu, D. Assembly of Complete Diploid-Phased Chromosomes from Draft Genome Sequences. G3 Gene Genomes Genet. 2022, 12, jkac143. [Google Scholar]
  111. Jones, P.; Binns, D.; Chang, H.-Y.; Fraser, M.; Li, W.; McAnulla, C.; McWilliam, H.; Maslen, J.; Mitchell, A.; Nuka, G. InterProScan 5: Genome-Scale Protein Function Classification. Bioinformatics 2014, 30, 1236–1240. [Google Scholar] [CrossRef]
  112. Mistry, J.; Chuguransky, S.; Williams, L.; Qureshi, M.; Salazar, G.A.; Sonnhammer, E.L.L.; Tosatto, S.C.E.; Paladin, L.; Raj, S.; Richardson, L.J. Pfam: The Protein Families Database in 2021. Nucleic Acids Res. 2021, 49, D412–D419. [Google Scholar] [CrossRef] [PubMed]
  113. Pearl, F.M.G.; Bennett, C.F.; Bray, J.E.; Harrison, A.P.; Martin, N.; Shepherd, A.; Sillitoe, I.; Thornton, J.; Orengo, C.A. The CATH Database: An Extended Protein Family Resource for Structural and Functional Genomics. Nucleic Acids Res. 2003, 31, 452–455. [Google Scholar] [CrossRef] [PubMed]
  114. Hunter, S.; Apweiler, R.; Attwood, T.K.; Bairoch, A.; Bateman, A.; Binns, D.; Bork, P.; Das, U.; Daugherty, L.; Duquenne, L. InterPro: The Integrative Protein Signature Database. Nucleic Acids Res. 2009, 37, D211–D215. [Google Scholar] [CrossRef] [PubMed]
  115. Stecher, G.; Tamura, K.; Kumar, S. Molecular Evolutionary Genetics Analysis (MEGA) for MacOS. Mol. Biol. Evol. 2020, 37, 1237–1239. [Google Scholar] [CrossRef] [PubMed]
  116. Tamura, K.; Stecher, G.; Kumar, S. MEGA11: Molecular Evolutionary Genetics Analysis Version 11. Mol. Biol. Evol. 2021, 38, 3022–3027. [Google Scholar] [CrossRef] [PubMed]
  117. Zuckerkandl, E.; Pauling, L. Evolutionary Divergence and Convergence in Proteins. In Evolving Genes and Proteins; Bryson, V., Vogel, H.J., Eds.; Academic Press: Cambridge, MA, USA, 1965; pp. 97–166. [Google Scholar] [CrossRef]
  118. Ye, J.; McGinnis, S.; Madden, T.L. BLAST: Improvements for Better Sequence Analysis. Nucleic Acids Res. 2006, 34, W6–W9. [Google Scholar] [CrossRef] [PubMed]
  119. Pruitt, K.D.; Tatusova, T.; Brown, G.R.; Maglott, D.R. NCBI Reference Sequences (RefSeq): Current Status, New Features and Genome Annotation Policy. Nucleic Acids Res. 2012, 40, D130–D135. [Google Scholar] [CrossRef] [PubMed]
  120. Jin, J.; Tian, F.; Yang, D.-C.; Meng, Y.-Q.; Kong, L.; Luo, J.; Gao, G. PlantTFDB 4.0: Toward a Central Hub for Transcription Factors and Regulatory Interactions in Plants. Nucleic Acids Res. 2017, 45, D1040–D1045. [Google Scholar] [CrossRef] [PubMed]
  121. Zerbino, D.R.; Achuthan, P.; Akanni, W.; Amode, M.R.; Barrell, D.; Bhai, J.; Billis, K.; Cummins, C.; Gall, A.; Girón, C.G. Ensembl 2018. Nucleic Acids Res. 2018, 46, D754–D761. [Google Scholar] [CrossRef]
  122. UniProt Consortium. UniProt: The Universal Protein Knowledgebase. Nucleic Acids Res. 2017, 45, D158–D169. [Google Scholar] [CrossRef]
  123. Sayers, E.W.; Cavanaugh, M.; Clark, K.; Pruitt, K.D.; Schoch, C.L.; Sherry, S.T.; Karsch-Mizrachi, I. GenBank. Nucleic Acids Res. 2022, 50, D161–D164. [Google Scholar] [CrossRef] [PubMed]
  124. Minh, B.Q.; Schmidt, H.A.; Chernomor, O.; Schrempf, D.; Woodhams, M.D.; von Haeseler, A.; Lanfear, R. IQ-TREE 2: New Models and Efficient Methods for Phylogenetic Inference in the Genomic Era. Mol. Biol. Evol. 2020, 37, 1530–1534. [Google Scholar] [CrossRef] [PubMed]
  125. Katoh, K.; Standley, D.M. MAFFT Multiple Sequence Alignment Software Version 7: Improvements in Performance and Usability. Mol. Biol. Evol. 2013, 30, 772–780. [Google Scholar] [CrossRef] [PubMed]
  126. Katoh, K.; Rozewicki, J.; Yamada, K.D. MAFFT Online Service: Multiple Sequence Alignment, Interactive Sequence Choice and Visualization. Brief. Bioinform. 2019, 20, 1160–1166. [Google Scholar] [CrossRef] [PubMed]
  127. Kalyaanamoorthy, S.; Minh, B.Q.; Wong, T.K.F.; von Haeseler, A.; Jermiin, L.S. ModelFinder: Fast Model Selection for Accurate Phylogenetic Estimates. Nat. Methods 2017, 14, 587–589. [Google Scholar] [CrossRef]
  128. Hoang, D.T.; Chernomor, O.; von Haeseler, A.; Minh, B.Q.; Vinh, L.S. UFBoot2: Improving the Ultrafast Bootstrap Approximation. Mol. Biol. Evol. 2018, 35, 518–522. [Google Scholar] [CrossRef]
Figure 1. The characterisation of the WRKY family of plant transcription factors. The investigation of the influence of abiotic and biotic factors on WRKY TF expression (A). The distribution of WRKY gene number in different plant species (B) and WRKY protein length distribution (C) based on PlantTFDB database data.
Figure 1. The characterisation of the WRKY family of plant transcription factors. The investigation of the influence of abiotic and biotic factors on WRKY TF expression (A). The distribution of WRKY gene number in different plant species (B) and WRKY protein length distribution (C) based on PlantTFDB database data.
Ijms 25 06241 g001
Figure 2. The evolutionary relationship, exon-intron structure, and motif analysis of WRKY transcription factor family members in V. vinifera (a). The midpoint-rooted phylogenetic tree was constructed with the Maximum Likelihood (ML) method and the VT + F + R4 substitution model (on the left) based on the complete amino acid sequences. The different colours and letters (A–E) note the five phylogenetic groups; the letters with numbers indicate the subgroups. The accepted type of WRKY classification is indicated on the right (I, IIa–IIe, III). The yellow boxes represent exons, and lines represent introns (in the middle). All exon and intron lengths are drawn to scale. The 23 different coloured boxes represent diverse conserved motifs identified with MEME (on the right). The numbering of conservative motifs is noted below (b).
Figure 2. The evolutionary relationship, exon-intron structure, and motif analysis of WRKY transcription factor family members in V. vinifera (a). The midpoint-rooted phylogenetic tree was constructed with the Maximum Likelihood (ML) method and the VT + F + R4 substitution model (on the left) based on the complete amino acid sequences. The different colours and letters (A–E) note the five phylogenetic groups; the letters with numbers indicate the subgroups. The accepted type of WRKY classification is indicated on the right (I, IIa–IIe, III). The yellow boxes represent exons, and lines represent introns (in the middle). All exon and intron lengths are drawn to scale. The 23 different coloured boxes represent diverse conserved motifs identified with MEME (on the right). The numbering of conservative motifs is noted below (b).
Ijms 25 06241 g002
Figure 3. Chromosomal location of 62 genes of transcriptional factor WRKY on the nineteen grape chromosomes.
Figure 3. Chromosomal location of 62 genes of transcriptional factor WRKY on the nineteen grape chromosomes.
Ijms 25 06241 g003
Figure 4. The topology of the phylogenetic tree associated with domain characteristics. The midpoint-rooted phylogenetic tree was constructed with the Maximum Likelihood (ML) method and the Q.plant + G4 substitution model based on the domain amino acid sequences. The different colours and letters (A–E) indicate the five phylogenetic groups according to Figure 2. The coloured boxes note the features of different groups of WRKY genes.
Figure 4. The topology of the phylogenetic tree associated with domain characteristics. The midpoint-rooted phylogenetic tree was constructed with the Maximum Likelihood (ML) method and the Q.plant + G4 substitution model based on the domain amino acid sequences. The different colours and letters (A–E) indicate the five phylogenetic groups according to Figure 2. The coloured boxes note the features of different groups of WRKY genes.
Ijms 25 06241 g004
Figure 5. The analysis of length and gene structure. The distribution of VvWRKY protein length (A) and exon number (C) in grapevine. The mean values of protein length (B) and exon number (D) according to belonging to different phylogenetic clades and WRKY groups. The data are presented as mean ± SD.
Figure 5. The analysis of length and gene structure. The distribution of VvWRKY protein length (A) and exon number (C) in grapevine. The mean values of protein length (B) and exon number (D) according to belonging to different phylogenetic clades and WRKY groups. The data are presented as mean ± SD.
Ijms 25 06241 g005
Figure 6. The suggested evolutionary model of the WRKY transcriptional factor family according to the phylogenetic studies in grapevine.
Figure 6. The suggested evolutionary model of the WRKY transcriptional factor family according to the phylogenetic studies in grapevine.
Ijms 25 06241 g006
Table 1. The amino acid consensus sequences of conserved motifs in grapevine WRKYs.
Table 1. The amino acid consensus sequences of conserved motifs in grapevine WRKYs.
MotifMotif Consensus Amino Acid SequencesWidth, aa
1DILDDGYRWRKYGQKPVKGSP21
2GCPVRKQVZRSSEDPSIVITTYEGKHNHP29
3YPRSYYRCTSA11
4DGYNWRKYGQKQVKGSEYPRSYYKCTYPNC30
5KKKAZKTIREPRVAVQTRSEV21
6ERSHDGQITEIIYKGTHNHPKPQPNRRSALG31
7KDELEVLKAELERVREENEKLREMLEQITKBYNALQMHLVEJMQ44
8TVEAATAAITADPNFTAALAAAITSIIG28
9SSGRCHCSKRRKLRVKRSIRVPAISNKIA29
10LPPAATAMASTTSAAASMLLS21
11MASISASAPFPTITLDLTQ19
12PSPLPIARSPYFTIPPGLSPTSLLDSPVLLS33
13EDGYNWRKYGQKQVKGSE18
14DCREIADYAVSKFKKVISJLNRTGHGRFR29
15MEEKLSWEQKTLINELTQGRELAKQLKIHL33
16LLRDHGLLQDIVPSFIRK18
17DEDDEDEPESKRRKKEV18
18QEAIQEAASAGLESVEKLIRLLSHAQDQ28
19QQMKHQADMMYRRSNSGINLKFDGSSCTPTMSSTRSFISSLSMDGSVANL50
20PVKKARVSVRARCDT15
21QGPFGMSHQZVLAQVTAQAAQAQSHMQLQ29
22REDLVVKILRSFEKALSILKCGG23
23NPRSYYKCTNA11
24EKPTDNFEHILNQMQ15
25VPAARNSSHBTAG13
Table 2. The distribution of WRKY protein isoforms (Ni) and encoding genes (Ng) in different grape cultivars and assemblies.
Table 2. The distribution of WRKY protein isoforms (Ni) and encoding genes (Ng) in different grape cultivars and assemblies.
CultivarCabernet FrancCabernet SauvignonPinot Noir
Clone0408FPS123PN40024
AssemblyDiploidHaploid, 12X, and ReferenceHaploid, Assembly v. 5
Grape GenomicsRefSeqGenBankGrape Genomics
Gene NameNiNgNiNgNiNgNiNgNiNgNiNgGene ID
VvWRKY1122102152111141Vitvi01g04492.t001
VvWRKY2421122211131Vitvi01g04490.t001
VvWRKY3213263111111Vitvi01g02157.t001
VvWRKY411--22111121Vitvi01g01680.t002
VvWRKY5426322211111Vitvi02g00039.t001
VvWRKY6222222112121Vitvi02g00114.t001
VvWRKY7322221111111Vitvi02g01847.t001
VvWRKY822222211--11Vitvi04g00133.t001
VvWRKY9422262111131Vitvi04g04524.t001
VvWRKY10142142122211111Vitvi04g00511.t001
VvWRKY11222242111131Vitvi04g04525.t001
VvWRKY12422142111111Vitvi04g00756.t001
VvWRKY13621122211121Vitvi04g00760.t001
VvWRKY14221122111111Vitvi04g01985.t001
VvWRKY153222362512111Vitvi04g01163.t001
VvWRKY16222222111111Vitvi05g00145.t001
VvWRKY17223242211111Vitvi06g01574.t001
VvWRKY18724222111111Vitvi06g00741.t001
VvWRKY19824162311111Vitvi07g00026.t001
VvWRKY20624262311121Vitvi07g00421.t001
VvWRKY21222222111111Vitvi07g00434.t001
VvWRKY2212222122211111Vitvi07g00523.t002
VvWRKY23424242111111Vitvi07g01694.t001
VvWRKY24222242111121Vitvi07g04782.t002
VvWRKY25222222211111Vitvi07g01860.t001
VvWRKY26----22--1111Vitvi08g04106.t001
VvWRKY27623162111111Vitvi08g00793.t001
VvWRKY28221142111111Vitvi08g00868.t001
VvWRKY29925152711121Vitvi08g01134.t001
VvWRKY30221122111111Vitvi08g01221.t001
VvWRKY31119382111111Vitvi09g01122.t001
VvWRKY32624242111111Vitvi10g00063.t001
VvWRKY332222102411111Vitvi10g00270.t001
VvWRKY34214242111111Vitvi10g00618.t001
VvWRKY35422142111111Vitvi10g00732.t001
VvWRKY36222233111111Vitvi10g01078.t001
VvWRKY37624252311111Vitvi11g00694.t002
VvWRKY38424242111111Vitvi11g01188.t001
VvWRKY39222233111111Vitvi12g00048.t001
VvWRKY40222222111111Vitvi12g00148.t001
VvWRKY41221142111111Vitvi12g00388.t001
VvWRKY42112253413333Vitvi12g00664.t003
Vitvi12g04520.t001
Vitvi12g04129.t001
VvWRKY43422233112111Vitvi12g01676.t001
VvWRKY44221122111111Vitvi13g00189.t001
VvWRKY45221122111111Vitvi13g01916.t001
VvWRKY46424242111111Vitvi14g00540.t001
VvWRKY47322222111111Vitvi14g01523.t001
VvWRKY48221122311111Vitvi14g01907.t001
VvWRKY491121102111121Vitvi14g02007.t001
VvWRKY50424262111111Vitvi15g00539.t001
VvWRKY51424242111111Vitvi15g01003.t001
VvWRKY52112222111111Vitvi15g01087.t001
VvWRKY53--2222111111Vitvi15g01090.t001
VvWRKY54332233111111Vitvi16g01132.t001
VvWRKY55112222111111Vitvi16g01133.t001
VvWRKY56222222111111Vitvi16g01213.t001
VvWRKY57332222111111Vitvi17g00102.t001
VvWRKY58326342211111Vitvi1700556.t001
VvWRKY59623152111111Vitvi18g00742.t001
VvWRKY60421162211111Vitvi19g00530.t001
VvWRKY61424211111111Vitvi19g00617.t001
VvWRKY621637222341--51Vitvi19g04652.t002
Total234115181106304129976165628464
Table 3. Characterisation of identified grapevine VvWRKY transcriptional factors.
Table 3. Characterisation of identified grapevine VvWRKY transcriptional factors.
Gene NameORF, aaExonsMean DistanceWRKY Domain LocationDNA-Binding ResiduesZinc FingerFunctional FamilyWRKY Group/CladeOther Features
VvWRKY130530.001162–218WRKYGQKC2H2FF:3IIc/B-
VvWRKY259450.012245–303WRKYGQKC2H2FF:2IIb/DCOILS Coil: 60–101
LxLxLx motif
VvWRKY350240.051247–303
427–484
WRKYGQKC2H2FF:6
FF:1
I/C-
VvWRKY418920.064111–168WRKYGQKC2H2FF:3IIc/B-
VvWRKY533030.030144–200WRKYGQKC2HCFF:7IIe/E-
VvWRKY634230.005133–192WRKYGQKC2HCFF:9III/A-
VvWRKY732330.042254–310WRKYGQKC2H2FF:4IId/EZn-cluster: 204–250
LxxLL motif
VvWRKY816630.005104–161WRKYGKKC2H2FF:3IIc/B-
VvWRKY925840.02197–155WRKYGQKC2H2FF:8IIa/DCOILS Coil: 14–34
LxLxLx motif
VvWRKY1031750.010160–218WRKYGQKC2H2FF:8IIa/DLxxLL motif
LxLxLx motif
VvWRKY1149150.004139–196
354–410
WRKYGQKC2H2-
FF:6
I/C-
VvWRKY1233830.000261–317WRKYGQKC2H2FF:4IId/EZn-cluster: 210–257
VvWRKY1319130.008103–160WRKYGKKC2H2FF:3IIc/B-
VvWRKY1413630.00853–111WRKYGKKC2H2FF:3IIc/B-
VvWRKY1570050.044234–290
452–509
WRKYGQKC2H2FF:6
FF:6
I/CLxLxLx motif
VvWRKY1630930.001155–212WRKYGQKC2H2FF:3IIc/BCOILS Coil: 107–127
LxxLL motif
VvWRKY1718930.11392–148WKKYGQKC2H2-NGSignal peptide: 1–21
Non cytoplasmic domain: 22–189
LxLxLx motif
VvWRKY1860350.004256–312
427–484
WRKYGQKC2H2FF:6
FF:1
I/C-
VvWRKY1934030.001274–331WRKYGQKC2H2FF:4IId/EZn-cluster: 226–270
VvWRKY2024230.03048–105WKKYGQKC2H2FF:4IIe/E-
VvWRKY2130230.003149–206WRKYGQKC2H2FF:3IIc/BCOILS Coil: 102–122
VvWRKY2251260.041269–326WRKYGQKC2H2FF:2IIb/DCOILS Coil: 105–132
VvWRKY2333630.045265–321WRKYGQKC2H2FF:4IId/EZn-cluster: 215–261
LxxLL motif
VvWRKY2419330.068106–163WRKYGKKC2H2FF:3IIc/B-
VvWRKY2522630.026150–206WRKYGQKC2H2FF:3IIc/B-
VvWRKY2623640.00566–109WMKGNPHC2HY-I/C-
VvWRKY2755250.026230–286
393–450
WRKYGQKC2H2FF:6
FF:1
I/C-
VvWRKY2833430.008136–196WRKYGQKC2HC-III/ALxLxLx motif
VvWRKY2947750.029196–250
397–453
WRKYGQKC2H2FF:6
FF:6
I/C-
VvWRKY3029930.006112–168WRKYGQKC2H2FF:6NG-
VvWRKY3131150.020160–218WRKYGQKC2H2FF:8IIa/D-
VvWRKY3253560.004277–334WRKYGQKC2H2FF:2IIb/DCOILS Coil: 99–133
VvWRKY3362650.019270–326
437–493
WRKYGQKC2H2-I/C-
VvWRKY3427830.00378–135WRKYGQKC2H2FF:4IIe/E-
VvWRKY3543830.056181–238WRKYGQKC2H2FF:3IIc/B-
VvWRKY3643830.006220–277WRKYGQKC2H2FF:5IIe/E-
VvWRKY3750040.004190–245
363–419
WRKYGQKC2H2-
FF:6
I/C-
VvWRKY3829730.006226–282WRKYGQKC2H2FF:4IId/EZn-cluster: 175–222
LxxLL motif
LxLxLx motif
VvWRKY3931130.004173–230WRKYGQKC2H2FF:3IIc/B-
VvWRKY4024430.00876–133WRKYGQKC2H2FF:4IIe/E-
VvWRKY4159350.005311–368WRKYGQKC2H2FF:2IIb/DCOILS Coil: 130–171
VvWRKY4240740.069110–166
285–342
WRKYGQKC2H2FF:6
FF:3
I/C-
VvWRKY4348750.005232–290WRKYGQKC2H2FF:2IIb/DCOILS Coil: 81–101
LxLxLx motif
VvWRKY4436430.001175–235WRKYGQKC2HC-III/A-
VvWRKY4531330.003111–170WRKYGQKC2HC-III/ALxxLL motif
VvWRKY4636530.004298–354WRKYGQKC2H2FF:4IId/EZn-cluster: 249–294
VvWRKY4718220.034105–162WRKYGQKC2H2FF:3IIc/B-
VvWRKY4855540.001225–283WRKYGQKC2H2FF:2IIb/DCOILS Coil: 33–81
LxLxLx motif
VvWRKY4952940.046233–289
415–472
WRKYGQKC2H2FF:6
FF:1
I/C-
VvWRKY5022840.033155–211WRKYGQKC2H2FF:3IIc/B-
VvWRKY5134930.004119–178WRKYGQKC2HCFF:9III/A-
VvWRKY5220120.006123–180WRKYGQKC2H2FF:3IIc/BLxxLL motif
VvWRKY5334830.001168–225WRKYGQKC2H2FF:7IIe/E-
VvWRKY5432930.058159–215WRKYGQKC2H2-IIe/ECOILS Coil: 275–295
VvWRKY5518530.01145–101WRKYGQKC2H2-IIe/E-
VvWRKY5636430.006134–193WRKYGQKC2HCFF:9III/A-
VvWRKY5715120.00472–129WRKYGQKC2H2FF:3IIc/B-
VvWRKY5861860.005271–329WRKYGQKC2H2FF:2IIb/DCOILS Coil: 102–129
LxLxLx motif
VvWRKY5934730.004275–331WRKYGQKC2H2FF:4IId/EZn-cluster: 226–271
LxxLL motif
VvWRKY6055160.008300–357WRKYGQKC2H2FF:2IIb/DCOILS Coil: 114–148
LxLxLx motif
VvWRKY6170080.048346–402
516–573
WRKYGQKC2H2FF:6
FF:1
I/CFrigida-like: 81–147 LxLxLx motif
VvWRKY6274650.015318–373
533–590
WRKYGQKC2H2FF:1
FF:6
I/C-
Table 4. WRKY genes identified in the genome of the V. vinifera cultivar Pinot Noir cl. PN40024 in different studies.
Table 4. WRKY genes identified in the genome of the V. vinifera cultivar Pinot Noir cl. PN40024 in different studies.
WRKY NamePlantTFDB
Wang L [44]/Guo [45]
NCBIEnsembelUniprotAccording to
Wu [49]/
Zhang and Feng [47]
RefseqGenBank
VvWRKY1GSVIVT01012196001 VvWRKY11/VvWRKY1XP_002274549.1 VvWRKY57WJZ80117VIT_01s0011g00720F6HF79VvWRKY1/VvWRKY13-1, VvWRKY57-1
VvWRKY2GSVIVT01020060001 VvWRKY19/VvWRKY2XP_010652374.1 VvWRKY72WJZ81033VIT_01s0026g01730D7TND6VvWRKY2/VvWRKY72-3
VvWRKY3GSVIVT01001332001 VvWRKY3/VvWRKY4NP_001268110.1 VvWRKY2WJZ81270VIT_01s0011g00220F6HYH9VvWRKY58/VvWRKY2-3, VvWRKY3-1
VvWRKY4GSVIVT01010525001 VvWRKY8/VvWRKY3XP_002275576.1 VvWRKY75WJZ91720VIT_01s0010g03930D7TB08VvWRKY3/VvWRKY57-2
VvWRKY5GSVIVT01019419001 VvWRKY17/VvWRKY5XP_010658402.1 VvWRKY22WJZ81903VIT_02s0025g00420F6HUN4VvWRKY4/VvWRKY22-3
VvWRKY6GSVIVT01019511001 VvWRKY18/VvWRKY6XP_002272720.1 VvWRKY41WJZ81983VIT_02s0025g01280D7TVE1VvWRKY5/VvWRKY41
VvWRKY7GSVIVT01001286001 VvWRKY2/VvWRKY7XP_059589595.1 VvWRKY21WJZ82421VIT_02s0154g00210D7TN24VvWRKY60/-
VvWRKY8GSVIVT01035426001 VvWRKY53/VvWRKY8XP_002279407.1 VvWRKY50-VIT_04s0008g01470D7STT5VvWRKY6/VvWRKY50
VvWRKY9GSVIVT01035884001 VvWRKY54/VvWRKY9XP_010648680.1 VvWRKY18WJZ85291VIT_04s0008g05750F6H3I5VvWRKY7/VvWRKY18
VvWRKY10GSVIVT01035885001 VvWRKY55/VvWRKY10XP_010648274.1 VvWRKY40WJZ85292VIT_04s0008g05760F6H3I6VvWRKY8/VvWRKY40-2
VvWRKY11GSVIVT01035965001 VvWRKY56/VvWRKY 11XP_010648749.1 VvWRKY32WJZ85365VIT_04s0008g06600F6H336VvWRKY9/VvWRKY32-2
VvWRKY12GSVIVT01033188001 VvWRKY48/VvWRKY12XP_002262775.1 VvWRKY17WJZ85515VIT_04s0069g00920A0A1U8AHW6VvWRKY10/VvWRKY11-1
VvWRKY13GSVIVT01033194001 VvWRKY49/VvWRKY13XP_002263836.1 VvWRKY51WJZ85520VIT_04s0069g00970D7T0E7VvWRKY11/VvWRKY51-3
VvWRKY14GSVIVT01033195001 VvWRKY50/VvWRKY14XP_003631843.1 VvWRKY43WJZ85521VIT_04s0069g00980A0A438DL90
D7T0E8
VvWRKY12/VvWRKY51-1
VvWRKY15GSVIVT01019109001 VvWRKY16/VvWRKY15XP_059592356.1 VvWRKYsusiba2WJZ85872VIT_04s0023g00470F6GX25VvWRKY13/VvWRKY2-1
VvWRKY16GSVIVT01034968001 VvWRKY52/VvWRKY16XP_002279385.1 VvWRKY48WJZ86707VIT_05s0077g00730A0A438GHD0
D7SYJ2
VvWRKY14/VvWRKY48
VvWRKY17GSVIVT01025491001 VvWRKY29/VvWRKY17XP_003632174.3 VvWRKY3WJZ88411VIT_06s0004g00230F6GUN9VvWRKY15/VvWRKY2-4
VvWRKY18GSVIVT01024624001 VvWRKY28/VvWRKY18XP_002272040.1 VvWRKY24WJZ89183VIT_06s0004g07500F6GUH8VvWRKY16/VvWRKY33-2
VvWRKY19GSVIVT01000752001 VvWRKY1/VvWRKY19XP_002282258.1 VvWRKY21WJZ90025VIT_07s0141g00680F6GXM5VvWRKY17/VvWRKY21
VvWRKY20GSVIVT01028129001 VvWRKY34/VvWRKY20XP_002270750.3 VvWRKY65WJZ90472VIT_07s0005g01520D7U2J0VvWRKY18/VvWRKY22-1
VvWRKY21GSVIVT01028147001 VvWRKY35/VvWRKY21XP_002277882.1 VvWRKY23WJZ90489VIT_07s0005g01710D7U2K4VvWRKY19/VvWRKY23
VvWRKY22GSVIVT01028244001 VvWRKY36/VvWRKY22XP_002281194.1 VvWRKY47WJZ90578VIT_07s0005g02570F6HZF7VvWRKY20/VvWRKY47
VvWRKY23GSVIVT01022067001 VvWRKY24/VvWRKY23XP_002283219.1 VvWRKY7WJZ91947VIT_07s0031g00080F6H4G0VvWRKY21/VvWRKY7-2
VvWRKY24GSVIVT01022245001 VvWRKY25/VvWRKY24XP_010652864.1 VvWRKY51WJZ92108VIT_07s0031g01710F6H4B4VvWRKY22/VvWRKY51-2, VvWRKY51-4
VvWRKY25GSVIVT01022259001 VvWRKY26/VvWRKY25XP_002279024.1 VvWRKY13WJZ92120VIT_07s0031g01840A0A438KK47
D7SW85, I0AVQ1
VvWRKY23
VvWRKY26-/--WJZ92812---/-
VvWRKY27GSVIVT01030258001 VvWRKY43/VvWRKY26XP_019077410.1 VvWRKY26WJZ92895VIT_08s0058g00690F6GXS4VvWRKY24/VvWRKY33-1
VvWRKY28GSVIVT01030174001 VvWRKY42/VvWRKY27XP_002272504.1 VvWRKY70WJZ92963VIT_08s0058g01390F6GXW4VvWRKY25/VvWRKY70-2
VvWRKY29GSVIVT01025562001 VvWRKY30/VvWRKY28XP_002275978.1 VvWRKY44WJZ93250VIT_08s0040g03070F6HQV7VvWRKY26/VvWRKY44
VvWRKY30GSVIVT01034148001 VvWRKY51/VvWRKY29XP_002270859.1 VvWRKY49WJZ93336VIT_08s0007g00570A0A438HSE3
D7THM0
VvWRKY27/VvWRKY49
VvWRKY31GSVIVT01015952001 VvWRKY14/VvWRKY30NP_001267919.1 VvWRKY40WJZ95373VIT_09s0018g00240F6HBV8VvWRKY28/VvWRKY4, VvWRKY40-1
VvWRKY32GSVIVT01012682001 VvWRKY12/VvWRKY31XP_002263115.1 VvWRKY31WJZ95843VIT_10s0116g01200F6H7H0VvWRKY29/VvWRKY6-2
VvWRKY33GSVIVT01007006001 VvWRKY4/VvWRKY59XP_010647039.2 VvWRKY20WJZ96075VIT_00s0463g00010A0A438DDY6VvWRKY59/VvWRKY20-1, VvWRKY20-4
VvWRKY34GSVIVT01021252001 VvWRKY21/VvWRKY32XP_002269267.1 VvWRKY65WJZ96565VIT_10s0003g01600D7TJP4VvWRKY30/VvWRKY65-1
VvWRKY35GSVIVT01021397001 VvWRKY22/VvWRKY33XP_002272089.1 VvWRKY 71WJZ96695VIT_10s0003g02810A5BVH3
D7TK05
VvWRKY31/VvWRKY28-1
VvWRKY36GSVIVT01021765001 VvWRKY23/VvWRKY34XP_002269170.1 VvWRKY14WJZ97031VIT_10s0003g05740A0A438BWU0VvWRKY32/VvWRKY14
VvWRKY37GSVIVT01023600001 VvWRKY27/VvWRKY35XP_002276194.1 VvWRKY32WJZ98338VIT_11s0037g00150D7U1A8VvWRKY33/VvWRKY32-1
VvWRKY38GSVIVT01029265001 VvWRKY39/VvWRKY36XP_002266188.1 VvWRKY51WJZ98778VIT_11s0052g00450D0V9L3VvWRKY34/VvWRKY11-2, VvWRKY11-3
VvWRKY39GSVIVT01020864001 VvWRKY20/VvWRKY37XP_002283603.1 VvWRKY71WJZ98954VIT_12s0028g00270E0CU42VvWRKY35/VvWRKY28-2
VvWRKY40-/-XP_002277383.1 VvWRKY65WJZ99058--VvWRKY36/VvWRKY3-2, VvWRKY65-2, VvWRKY65-3
VvWRKY41GSVIVT01030453001 VvWRKY44/VvWRKY38XP_002269696.2 VvWRKY 31WJZ99341VIT_12s0059g00880F6HIC7VvWRKY37/VvWRKY6-1
VvWRKY42GSVIVT01030046001 VvWRKY41/VvWRKY39XP_002272407.1 VvWRKY20WJZ99676VIT_12s0057g00550F6HHL8VvWRKY38/VvWRKY20-3, VvWRKY20-6
VvWRKY43GSVIVT01029688001 VvWRKY40/VvWRKY40XP_010657556.1 VvWRKY9WKA00085VIT_12s0055g00340F6H1R3VvWRKY39/VvWRKY9
VvWRKY44GSVIVT01032662001 VvWRKY46/VvWRKY41XP_002275373.1 VvWRKY55WKA00784VIT_13s0067g03130A0A438E3U8
F6HC34
VvWRKY40/VvWRKY55
VvWRKY45GSVIVT01032661001 VvWRKY45/VvWRKY42XP_002275401.1 VvWRKY70WKA00785VIT_13s0067g03140F6HC33VvWRKY41/VvWRKY70-1
VvWRKY46GSVIVT01036223001 VvWRKY57/VvWRKY43XP_002270614.2 VvWRKY74WKA03253VIT_14s0081g00560F6HVI7VvWRKY42/VvWRKY74
VvWRKY47GSVIVT01033063001 VvWRKY47/VvWRKY44XP_002274387.1 VvWRKY75WKA04134VIT_14s0068g01770D7SVN0
I3RQB5
VvWRKY43/VvWRKY45
VvWRKY48GSVIVT01011356001 VvWRKY9/VvWRKY45XP_002277221.2 VvWRKY72WKA04564VIT_14s0108g00120D7SX70VvWRKY44/VvWRKY72-2
VvWRKY49GSVIVT01011472001 VvWRKY10/VvWRKY46XP_010661104.2 VvWRKY4WKA04672VIT_14s0108g01280-VvWRKY45/-
VvWRKY50GSVIVT01018300001 VvWRKY15/VvWRKY47XP_002270527.1 VvWRKY12WKA05315VIT_15s0021g01310A0A1U6ZIF2VvWRKY46/VvWRKY12-1, VvWRKY12-2
VvWRKY51GSVIVT01027069001 VvWRKY33/VvWRKY48XP_002281031.1 VvWRKY46WKA05852VIT_15s0046g01140F6I6B1VvWRKY47/VvWRKY46
VvWRKY52GSVIVT01026969001 VvWRKY32/VvWRKY49XP_002275528.3 VvWRKY24WKA05934VIT_15s0046g02150D7UCE6
A0A438JLT2
VvWRKY48/VvWRKY24
VvWRKY53GSVIVT01026965001 VvWRKY31/VvWRKY50XP_002276925.1 VvWRKY22WKA05938VIT_15s0046g02190D7UCE2
A0A438JM47
VvWRKY49/VvWRKY22-2
VvWRKY54GSVIVT01028823001 VvWRKY38/VvWRKY51XP_010662789.1 VvWRKY22WKA07335VIT_16s0050g01480E0CUS7VvWRKY50/-
VvWRKY55-/-XP_010662788 VvWRKY27WKA07336---/-
VvWRKY56GSVIVT01028718001 VvWRKY37/VvWRKY52XP_002267793.2 VvWRKY53WKA07458VIT_16s0050g02510E0CUJ8VvWRKY51/VvWRKY53
VvWRKY57GSVIVT01008553001 VvWRKY6/VvWRKY53NP_001268218.1 VvWRKY1WKA07893VIT_17s0000g01280Q5IZC7VvWRKY52/VvWRKY1-2, VvWRKY75
VvWRKY58GSVIVT01008046001 VvWRKY5/VvWRKY54XP_010663394.1 VvWRKY72WKA08379VIT_17s0000g05810D7SIE7VvWRKY53/VvWRKY72-1
VvWRKY59GSVIVT01009441001 VvWRKY7/VvWRKY55XP_002284966.1 VvWRKY7WKA09909VIT_18s0001g10030E0CPR7VvWRKY54/VvWRKY7-1
VvWRKY60GSVIVT01037686001 VvWRKY58/VvWRKY56XP_010644476.1 VvWRKY31WKA12500VIT_19s0090g00840F6HEQ5VvWRKY55/VvWRKY42
VvWRKY61GSVIVT01037775001 VvWRKY59/VvWRKY57XP_010644520.1 VvWRKY20WKA12584VIT_19s0090g01720F6HER4VvWRKY56/VvWRKY20-2, VvWRKY20-5
VvWRKY62GSVIVT01014854001 VvWRKY13/VvWRKY58XP_002265612.1 VvWRKY2-VIT_19s0015g01870F6I4X4VvWRKY57/VvWRKY2-2
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

Vodiasova, E.; Sinchenko, A.; Khvatkov, P.; Dolgov, S. Genome-Wide Identification, Characterisation, and Evolution of the Transcription Factor WRKY in Grapevine (Vitis vinifera): New View and Update. Int. J. Mol. Sci. 2024, 25, 6241. https://doi.org/10.3390/ijms25116241

AMA Style

Vodiasova E, Sinchenko A, Khvatkov P, Dolgov S. Genome-Wide Identification, Characterisation, and Evolution of the Transcription Factor WRKY in Grapevine (Vitis vinifera): New View and Update. International Journal of Molecular Sciences. 2024; 25(11):6241. https://doi.org/10.3390/ijms25116241

Chicago/Turabian Style

Vodiasova, Ekaterina, Anastasiya Sinchenko, Pavel Khvatkov, and Sergey Dolgov. 2024. "Genome-Wide Identification, Characterisation, and Evolution of the Transcription Factor WRKY in Grapevine (Vitis vinifera): New View and Update" International Journal of Molecular Sciences 25, no. 11: 6241. https://doi.org/10.3390/ijms25116241

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