Next Article in Journal
Bioactive Compounds from BRS Violet Grape Pomace: An Approach of Extraction and Microencapsulation, Stability Protection and Food Application
Previous Article in Journal
Prioritization of Physio-Biochemical Selection Indices and Yield-Attributing Traits toward the Acquisition of Drought Tolerance in Chickpea (Cicer arietinum L.)
Previous Article in Special Issue
The Effect of Dwarfing Interstocks on Vegetative Growth, Fruit Quality and Ionome Nutrition of ‘Fuji’ Apple Cultivar ‘Tianhong 2’—A One-Year Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genome-Wide Identification of Bilberry WRKY Transcription Factors: Go Wild and Duplicate

by
Winder Felipez
1,2,†,
Jennifer Villavicencio
2,3,†,
Valeria Oliveira Nizolli
2,
Camila Pegoraro
2,
Luciano da Maia
2 and
Antonio Costa de Oliveira
2,*
1
Instituto de Agroecología y Seguridad Alimentaria, Facultad de Ciências Agrárias, Universidad San Francisco Xavier de Chuquisaca—USFX, Casilla, Correo Central, Sucre 1046, Bolivia
2
Plant Genomics and Breeding Center, Departamento de Fitotecnia, Faculdade de Agronomia Eliseu Maciel, Universidade Federal de Pelotas—UFPel, Pelotas CEP 96010-900, RS, Brazil
3
Carrera de Ingeniería Agroforestal, Facultad de Ciencias Ambientales, Universidad Cientifica del Sur—UCSUR, Antigua Panamericana Sur km 19 Villa el Salvador, Lima CP 150142, Peru
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Plants 2023, 12(18), 3176; https://doi.org/10.3390/plants12183176
Submission received: 24 April 2023 / Revised: 11 July 2023 / Accepted: 20 July 2023 / Published: 5 September 2023

Abstract

:
WRKY transcription factor genes compose an important family of transcriptional regulators that are present in several plant species. According to previous studies, these genes can also perform important roles in bilberry (Vaccinium myrtillus L.) metabolism, making it essential to deepen our understanding of fruit ripening regulation and anthocyanin biosynthesis. In this context, the detailed characterization of these proteins will provide a comprehensive view of the functional features of VmWRKY genes in different plant organs and in response to different intensities of light. In this study, the investigation of the complete genome of the bilberry identified 76 VmWRKY genes that were evaluated and distributed in all twelve chromosomes. The proteins encoded by these genes were classified into four groups (I, II, III, and IV) based on their conserved domains and zinc finger domain types. Fifteen pairs of VmWRKY genes in segmental duplication and four pairs in tandem duplication were detected. A cis element analysis showed that all promoters of the VmWRKY genes contain at least one potential cis stress-response element. Differential expression analysis of RNA-seq data revealed that VmWRKY genes from bilberry show preferential or specific expression in samples. These findings provide an overview of the functional characterization of these proteins in bilberry.

1. Introduction

The bilberry (Vaccinium myrtillus L.), also known as the European wild blueberry, is a fruit that grows in the wild from the forests of Northern Europe to the Caucasus and towards the northern Asia-Pacific [1,2,3]. V. myrtillus is one of the 400 species belonging to the genus Vaccinium of the Ericaceae family, which also includes cultivated blueberry species and wild fruits such as V. corymbosum, V. angustifolium, V. virgatum, and V. macrocarpon [4]. Blueberries are usually diploid (2n = 2x = 24), but polyploid subspecies can be found in North America [2]. Due to their high levels of beneficial nutrients and bioactive phytochemicals present in the skin, pulp, and leaves, especially anthocyanins, the demand for blueberries has been increasing in the market [3,5]. Anthocyanins are responsible for their colors of red, blue, purple, and black fruits [5,6]. In the last two decades, several studies have linked the polyphenols and high natural anthocyanin levels of blueberries to the prevention of chronic diseases such as cancer, coronary heart disease, and neurological disorders [5].
In addition, there has been an increased interest in the Vaccinium genus, and as a result, genomic and transcriptomic resources for this genus have been developed. However, despite this progress, there is still incomplete genome assembly and annotation data, as is the case of blueberries (V. corymbosum) [7]. On the other hand, efforts to understand the genomic regions that control the anthocyanin composition of the berry have led to the generation of the complete genome, as is the case with bilberry (V. myrtillus) [4]. Anthocyanins and flavonoid compounds present in blueberry fruits have been associated with structural genes and transcription factors (TFs) in many species [2]. Advances in genomic resources and the understanding of Vaccinium spp. have made it possible to describe the MYBA TF loci and identify the main regulatory genes of this family that determine anthocyanin production [4]. Recent studies have shown that anthocyanin profiles in Vaccinium spp. are regulated by prevailing light and temperature conditions [8,9]. Blue light induces anthocyanin accumulation in pear fruit (Pyrus communis x pyrifolia cv Red Zaosu) and activates anthocyanin-production-responsive genes [10]. In blueberry, the light intensity regulation of anthocyanin accumulation represents a valuable data set to guide future functional and crop improvement studies [11]. In bilberry, it was demonstrated that blue and red/blue lights have the ability to promote anthocyanin biosynthesis by inducing the expression of key structural genes and accumulation of metabolites involved in the anthocyanin synthesis pathway, as well as the relationship between photosynthesis under different light qualities in blueberry leaves [12]. Several regulatory genes, including WRKY family TFs, have been proposed to control fruit development and ripening processes, as well as their relationship with responses to various stresses [13,14].
The WRKY TFs are one of the largest families of transcriptional regulators in plants and are present in several plant species [15,16]. These TFs are characterized by the WRKY domain, which contains about 60 amino acids that may be located near the N- or C-terminal regions and is composed of a highly conserved WRKYGQK heptapeptide DNA-binding sequence, followed by a zinc finger motif that binds to specific cis-regulatory elements called W-boxes (TTGACT/C) [17,18,19,20,21]. Evidence suggests that W-boxes are present in the promoter region of genes related to plant innate immunity [20]. Based on the number of WRKY domains and the pattern of zinc finger motifs, WRKY genes can be divided into four groups (I, II, III, and IV) [22,23,24]. Group I WRKYs have two domains containing a C2H2 zinc finger motif. Group II representatives have only one WRKY domain and one C2H2 zinc finger motif and can be divided into five subgroups (IIa, IIb, IIc, IId, and IIe). Group III WRKYs also have a single domain, but their zinc finger motif is C2HC [22,23] Recently, Group IV was identified, composed of incomplete WRKYs that do not fit into the previous groups because they lack the zinc finger motif [23,24].
We can find several examples of WRKY genes that act in the biosynthesis of anthocyanins, plant pigments that are present in leaves, flowers, fruits, and roots. Ultraviolet-B (UV-B) radiation promotes the synthesis of anthocyanins in many plants, and several transcription factors respond to UV-B radiation [25]. In M. domestica, MdWRKY72 TF increases anthocyanin synthesis by direct and indirect mechanisms when induced by UV-B radiation [26]. The overexpression of MdWRKY75 led to an increase in anthocyanin levels by binding to the MYB transcription factor promoter, MdMYB1 [27]. Other examples are the MscWRKY12 and MscWRKY19 TFs from M. sylvestris, which are involved in leaf pigmentation during the development of this organ [24], and the involvement of WRKY TFs in the synthesis of anthocyanins in Raphanus sativus L., the carmine radish [28]. Especially in fruits, such as Pyrus spp., their red skin accumulates anthocyanin, and the genes involved can contribute to the improvement of their appearance, as is the case with the combination of PyWRKY26 and PybHLH3 capable of co-directing the PyMYB114 promoter to generate anthocyanin accumulation in red pears [29]. Eight PcWRKYs also participate in color development in red fruits of P. communis, and color fading in some fruits is due to reduced biosynthesis, increased degradation, and the suppression of anthocyanin transport [30]. In addition, by having a better understanding of the genes involved in this biosynthesis, it is possible to overexpress them. In M. domestica, the overexpression of MdWRKY11 [31] and MdWRKY71-L [32] plays a role in the synthesis of anthocyanins in red fruits and demonstrates that TFs participate in the UV-B signaling pathway to regulate the accumulation of anthocyanin in apple. Just like in the apple, the red coloration of mango (Mangifera indica L.) peels results from anthocyanin accumulation, where it was found that MiWRKY1 and MiWRKY81 were upregulated during the light induction accumulation of anthocyanin in mango, indicating that these genes may regulate anthocyanin biosynthesis [33].
The growing interest in the nutraceutical properties of blueberries makes it essential to deepen our understanding of fruit ripening regulation and anthocyanin biosynthesis. In this context, WRKY transcription factors have been identified as a promising tool for manipulating stress tolerance and the synthesis of nutritionally desirable compounds. Therefore, it is crucial to identify the WRKY proteins, from their structural classification to their phylogenetic relationships, as well as their gene structure, conservation of domains and motifs, cis elements, chromosomal mapping, tandem and segmental duplication, and genetic divergence in the complete bilberry genome (Vaccinium myrtillus L.). The detailed characterization of these proteins will provide a comprehensive view of the evolution and modification of the WRKY protein family in the crop and will help to determine the functional features of VmWRKY genes in different plant organs and in response to different intensities of light.

2. Results

2.1. Identification of the WRKY Protein Family in Bilberry

The statistical alignment Hidden Markov Model (HMMsearch) in the bilberry (V. myrtillus) proteome generated 76 WRKY proteins with complementary domains. Two Q-type WRKY proteins and 69 WRKY genes were identified (Table 1), and seven WRKY proteins (VmWRKY14, VmWRKY28, VmWRKY29, VmWRKY32, VmWRKY44, VmWRKY48, and VmWRKY49) were considered recent duplicates due to having more than 98% similarity (Tables S1–S3). The heptapeptide region of the WRKY domain (WRKYGQK) showed variations (WRKYGEK, WRKYG, WRKYGRK, WRKYGKK, and WRKYLQK) within the primary nomenclature of the WRKY family (PF03106). The peptide length ranged from 111 to 740 amino acids (VmWRKY42 and VmWRKY61). The coding sequences (CDS) varied from 336 to 2223 nucleotides. The predicted protein molecular weight ranged from 13,001.347 to 82,707.564 kDa. The isoelectric point varied from 4.68 (VmWRKY27) to 9.89 (VmWRKY28). The instability index ranged from 32.01 (VmWRKY55) to 76.58 (VmWRKY27), and the average hydrophobicity varied from −0.44 (VmWRKY61) to −1.212 (VmWRKY27).

2.2. Analysis of Cis Elements in VmWRKY Gene Promoters

The analyses of promoter sequences of 69 VmWRKY genes generated 108 types of putative cis-regulatory elements, which were categorized into four known and one unknown cis-regulatory action elements (Table S4). The largest category among the known elements comprises light-responsive elements (25%), predominantly represented by cis-action G-Box, Box 4, and GT1-motif. Following that, hormone-responsive elements (9%) with cis-action ABRE, CGTCA-motif, and TGACG-motif were identified. Elements related to development (9%) were found with cis-action CAT-box, O2-site, and GCN4_motif. Furthermore, environment-responsive elements (8%) exhibited cis-action AREM, LTR, and MBS. The elements associated with promoters and binding sites (7%) were predominantly represented by cis-action CAAT-box, TATA-box, and W-box. Notably, the WRKY transcription factors are bound to the W-box to initiate transcription. Additionally, the analysis revealed several other unknown functional elements (40%) with cis-action MYB, MYC, and STRE.

2.3. Phylogeny, Gene Structure, and Motif Analysis of WRKY Protein in Bilberry

The unrooted phylogenetic tree of 69 VmWRKY protein sequences (Figure 1A) displays their relationships, forming eight subgroups (I, IIa, IIb, IIc, IId, IIe, III, and IV). The largest subgroups are group I (14 members) and group IIc (14), followed by group III (12), group IIe (10), group IIb (9), group IId (4), group IIa (4), and group IV (2). In group IV, the protein sequences contain the conserved WRKY domain but lack the zinc finger motif. The primary structure of the 69 VmWRKY gene sequences (Figure 1B) shows variation in the number of introns, ranging from 0 to 9. By conducting the analysis of conserved functional motifs in the protein sequences, a total of 20 functional motifs were identified, distributed across each VmWRKY subgroup (Figure 1C). Each subgroup has specific motif patterns (Figures S1 and S2).

2.4. Chromosomal Localization, VmWRKY Gene Duplication, and Divergence

A total of 64 VmWRKY genes were mapped onto twelve chromosomes of the bilberry genome (Figure 2), and the genes VmWRKY65, VmWRKY66, VmWRKY67, VmWRKY68, and VmWRKY69 are sequenced at the scaffold level, and their location on the chromosome is unknown. The highest number of VmWRKY genes was found on chromosome 12 (11 genes), followed by chromosome 8 (9 genes), chromosome 4 (8 genes), chromosomes 1 and 6 (7 genes), chromosome 3 (5 genes), chromosomes 11 and 9 (4 genes), chromosome 2 (3 genes), and chromosomes 10, 5, and 7 (2 genes). Tandem and segmental duplications were identified (Table 2). Four pairs of VmWRKY genes were identified as tandem duplications (<100 Kb and >70% similarity) (Figure S3), and 12 pairs of VmWRKY genes were identified as segmental duplications (>70% similarity) (Figure S4). This information provides valuable insights into the genetic evolution of the gene family and its domains regarding other model species such as Arabidopsis thaliana (Figure S5). The estimation of divergence time for paralogous VmWRKY gene pairs, based on their synonymous substitution rates, ranged from 0.13 to 83.40 million years for the VmWRKY40/VmWRKY41 and VmWRKY61/VmWRKY63 gene pairs. Other duplicated gene pairs did not yield divergence time estimates with the methods applied using DNA sequences, coding sequences (CDS), and the transcriptome annotation of the genome. Another analysis was applied, and the Nei and Gojobori (NG) method was utilized to determine the divergence between proteins (Figure S6).

2.5. Number of Transcripts and Expression Patterns of VmWRKY Genes in Various Tissues

The expression patterns of each transcriptomic sample under light and organ conditions were identified in fragments per kilobase million (FPKM) out of a total of 36,405 (100%) annotated genes in the bilberry reference genome (Table S5). Gene expression in the light samples was predominantly observed in the red light condition (24,491 = 67%), followed by the control samples (23,741 = 65%) and blue light samples (23,825 = 65%). The multi-dimensional scaling (MDS) analysis of the samples revealed dissimilarity in expression patterns (Figure 3(A1)). The expression of VmWRKY genes was identified in nearly equal proportions in the control sample (61 = 0.17%), red light sample (61 = 0.17%), and blue light sample (59 = 0.16%). However, 56 VmWRKY genes were found to be common across all samples (Figure 3(A2)), while 10 VmWRKY genes (VmWRKY11, VmWRKY42, VmWRKY3, VmWRKY43, VmWRKY2, VmWRKY19, VmWRKY40, VmWRKY33, VmWRKY26, and VmWRKY67) exhibited differential expression in each sample (Figure 3(A3)). Additionally, five genes (VmWRKY11, VmWRKY5, VmWRKY19, VmWRKY12, and VmWRKY34) showed high expression levels across all three samples (Figure 3(A4)).
The gene expression in plant organ samples showed a higher proportion in whole berries (24,268 = 67%) and berry peels (24,449 = 67%), followed by berry pulp (24,190 = 66%), roots (23,370 = 64%), and leaves (22,791 = 63%) (Table S6). The multi-dimensional scaling (MDS) analysis of the samples revealed similarity patterns among three samples and dissimilarity among two samples (Figure 3(B1)). The expressed VmWRKY genes were predominantly identified in roots (62 = 0.17%) and berry peels (60 = 0.16%), followed by leaves (57 = 0.16%), whole berries (57 = 0.16%), and berry pulp (56 = 0.15%). However, 48 VmWRKY genes were found to be common across all samples (Figure 3(B2)). Among the root sample, 41 VmWRKY genes exhibited high differential expression compared to the other samples (Figure 3(B3)). Furthermore, 15 VmWRKY genes were highly expressed in all samples (Figure 3(B4)). Additionally, complementary analyses of raw and processed transcriptomic sample-specific data are presented (Figures S7–S10).

3. Discussion

3.1. Identification of the WRKY Protein Family in Bilberry

Transcription factors exist in the form of a superfamily of genes and play a critical regulatory role in plants’ growth, development, and response to the environment [34]. In recent years, there has been a significant increase in interest in investigating the WRKY protein family in different plant species, including the bilberry crop. With the publication of the complete genome of several species, the identification and analysis of TF families at the whole genome level has become one of the focuses of genomic research. The WRKY TF family plays an important role in plant growth and development and defense mechanisms [35].
In the bilberry genome, it was possible to identify 76 VmWRKY genes. This number is higher than those of the model species Arabidopsis thaliana (72) [22], Vitis vinifera (59) [36], and Rubus occidentalis (60) [37]. In contrast, in other species such as Oryza sativa (109) [19], Triticum aestivum (160) [18], Malus domestica cv Gala (112) [24], and Pyrus Bretschneideri (103) [38], a larger number of genes were identified [19]. The higher number of WRKY genes in some species may be due to polyploidy and genome size [18].

3.2. Analysis of Cis Elements in VmWRKY Gene Promoters

A particular feature of the WRKY TF family is the ability to specifically bind to W-box ((C/T)TGAC(T/C)). Nevertheless, they bind also to other cis-acting elements located in the promoter region of their target genes [19,36,39]. In this study, it was possible to observe that the promoter region of VmWRKY genes contains a total of 108 conserved cis-regulatory elements with diverse functions. Strikingly, 40 VmWRKYs presented one or more W-boxes, indicating the putative autoregulation of these TFs, a number that is higher than the 26 FaWRKYs observed in strawberry [23,40,41,42]. Common promoter and enhancer regions, such as A-Box and CAAT-box, were also identified, as well as TATA-box, which are regions located around −30 of transcription initiation. In addition, the presence of HD-Zip 1 and HD-Zip 3, the elements involved in mesophyll cell differentiation and the site where protein binding occurs, were detected.
The presence of a large number of phytohormone-responsive elements, such as TGA (responsive to auxin), TATC-box (response to gibberellin), SARE (salicylic acid-responsive element), TCA (response to salicylic acid), ABRE (response to abscisic acid), AuxRR-core (responsive to auxin), CGTCA motif (response to MeJA), TGACG motif (response to MeJA), GARE (responsive to gibberellin), and P-box (gibberellin-responsive element), was detected. Additionally, 27 promoter regions associated with the response to light were identified, with G-Box, Box 4, GT1-motif, and TCT-motif being the most abundant within the species. This coincides with what was reported in raspberry (Rubus occidentalis), where the most abundant promoter region was Gbox, ATC, and TCT-motif [37].
The VmWRKY gene promoter regions also showed conserved cis-regulatory elements, which are involved in a variety of functions, such as responses to biotic and abiotic stresses. Among these elements are TC-rich repeats, which act in defense and response to stress; LTRs, which respond to low temperatures; AREs, which act in anaerobic induction; GC motifs, which act in specific anoxic induction; MBSs, which act in response to water stress; the WUN motif, which responds to tissue damage; and AT-rich sequences and AT-rich elements. The presence of diverse cis-acting elements, which mediate responses to environmental stresses and plant hormones, suggests that these WRKY TFs are involved in various biological processes.

3.3. Phylogeny, Gene Structure, and Motif Analysis of WRKY Proteins in Bilberry

Plants have adaptation mechanisms to adverse environments that involve signal transduction and molecular regulation in response to stress. TFs play a key role in this process by activating or inhibiting gene transcription by their specific binding to gene promoter regions. This results in the induction of functional gene expression and contributes to signal transduction in response to stress [43,44].
In the phylogenetic analysis of TFs, the 69 VmWRKY proteins were distributed into seven clusters. Cluster 1 presents nine WRKY proteins, cluster 2 presents four, cluster 3 presents ten, cluster 4 presents four, cluster 5 presents fourteen, cluster 6 presents sixteen, and cluster 7 presents twelve. The proteins belonging to each group (I, II, III, and IV), with group II subdivided into five distinct subgroups (IIa–e), formed clusters with members of only one group, and others were made up of members of two or more groups. During the intragroup evolutionary analysis of VmWRKY genes, it was observed that genes from group IIc were more divergent from the other members within group II (Iia, Iib, Iid, and IIe). Within this branch, IIa and IIb were grouped separately in a branch within group II, similarly to what was found in plum [45] and black raspberry [37]. Also, IId and IIe formed another subgroup. Group I presented almost all members of this group clustered together, with two additional members from group III and group IV. The last group formed was composed of almost all members from group III but included one genotype from group IV (Figure 1). All WRKY TFs from group I in bilberry contain two WRKY domains. Among them, 15 VmWRKYs (21.7%) were assigned to group I, and 47 VmWRKY (68.12%) genes were distributed in group II, which was further classified into five subgroups, IIa, IIb, IIc, IId, and IIe, which contained 7, 9, 14, 5, and 12 VmWRKYs, respectively. The remaining 12 VmWRKYs (17.4%) belonged to group III, and 2 VmWRKYs (2.9%) belonged to group IV. These results are similar to those found in grape (Vitis vinifera) and blackberry (Rubus occidentalis), in which the number of WRKYs genes found in group II were the most abundant with 39 VvWRKYs (66.1%) [36] and 25 RoWRKYs (41.6%), respectively [37].
Group III of WRKY gene members presents a zinc finger motif different from groups I and II and can be considered the most dynamic in terms of evolution [34,46]. In this study, 12 VmWRKYs were identified as group III, which is similar to the 13 AtWRKYs found in Arabidopsis [47] and the 10 found in wild strawberry [48]. However, this count is higher than the 6 VvWRKYs found in grapes [36] and lower than the 28 OsWRKYs in rice [34]. Some WRKYs genes from group III are part of the signaling pathway of the plant’s defense system, having a significant impact on resistance to diseases and drought [34].
Finally, we have VmWRKY68 and VmWRKY69, which were grouped in group IV for not fitting into any of the other groups because the WRKY domain has a partial or lacks a complete zinc finger motif structure. WRKY genes belonging to group IV were also reported in species such as Arabidopsis thaliana with two AtWRKY genes [22] and Pennisetum glaucum with twelve PgWRKY genes [41]. Likewise, M. domestica (10% = 13 MdWRKYs) and Vitis vinifera (2% = 2 VvWRKYs) [49] presented WRKY proteins without a complete domain. This occurrence may mean the loss of the WRKY domain [21,50,51] and modifying the functional properties of the proteins [42]. In studies performed in Hylocereus undatus, the absence of the WRKY domain, the zinc finger or coiled-coil sequence, did not allow the binding between the WRKY protein and the promoter region of the gene [52], suggesting that genes belonging to group IV can originate non-functional proteins.

3.4. Chromosomal Localization, VmWRKY Gene Duplication, and Divergence

Overall, 69 VmWRKY genes were identified in the bilberry genome; however, only 64 VmWRKY genes (VmWRKY1VmWRKY 64) had a known location on the chromosomes, and they were distributed into twelve chromosomes. The WRKY genes (VmWRKY65VmWRKY69) are sequenced at the scaffold level and have no known chromosome locations. And most of the VmWRKYs were abundant on Chr 12. The number of VmWRKY genes is similar to that of the pitaya (Hylocereus undatus), with 70 HuWRKY genes distributed on eleven chromosomes [52]. The raspberry also presents a similar number of WRKY genes (60 RoWRKY); however, it has a lower number of chromosomes (7), and Chr 6 had the largest number of RoWRKY genes, representing 23.33% of the WRKY genes [37]. In the case of grapes (Vitis vinifera), the number of WRKY genes present in this species is also similar (59 VvWRKYs); however, they are mapped to nineteen chromosomes [38], a greater number of chromosomes than that present in bilberry. The differential distribution of the WRKY genes present in a species may imply chromosomal rearrangements and duplication events that took place during its evolution [41].
It has been observed that gene duplication, including tandem duplication, fragment duplication, and genome duplication, is a key factor in gene family amplification in plant genomes [53]. In this particular study, 15 segmental duplication events and four tandem duplication events (VmWRKY37/VmWRKY38, VmWRKY40/VmWRKY41, VmWRKY46/VmWRKY47, and VmWRKY48/VmWRKY49) were identified (Table 2). A high frequency of segmental duplication was also observed in Rubus occidentalis, with five genes containing homologous segments [37]. In Fragaria vesca, segmental duplications were higher than tandem duplications, representing 84.2% and 15.8% of the total duplications, respectively. Compared to what has been observed in V. vinifera [54] and Oryza rufipogon [55], tandem duplication events can significantly contribute to the expansion of VmWRKY genes [52].

3.5. Expression Patterns of WRKY Genes by Induction of Light and Plant Organs

The processing of transcriptomic samples from bilberry plant organs induced by light showed variation in the total gene expression and specific expression of VmWRKY genes in each sample (Table S5). In the case of samples treated with red light (24,491 = 67%), gene expression was higher than in the control and blue light samples. This pattern was also observed in the specific expression of VmWRKY genes. However, 17 VmWRKY genes showed differential expression under red light compared to the control and blue light (Figure 3(A3)). On the other hand, in pear (P. communis x pyrifolia cv ‘Red Zaosu’), exposure to blue light led to an increase in anthocyanin accumulation and the activation of genes responsible for anthocyanin production [10]. In red mango fruit (M. indica L.), UV-B light induction positively regulates anthocyanin accumulation, and the genes MiWRKY1 and MiWRKY81 are involved in this regulation [33]. In general, anthocyanin accumulation is influenced by light availability, and the specific impact of different light qualities varies among plant species [56,57,58]. These results suggest that VmWRKYs are concentrated in the skin and berry of the fruit, which may indicate an association with higher levels of anthocyanins, the main source of organoleptic and antioxidant properties in blueberries, as reported in a gene expression study during bilberry development [59]. However, the expression level in bilberry leaves was high compared to that observed in Tetrastigma (Tetrastigma hemsleyanum), a grape family plant with eight WRKYs, which showed high expression levels in leaves [60].
Gene expression in plant organ samples showed similar levels in the entire berry (24,268 = 67%) and the berry skin (24,449 = 67%), which were comparable to those observed in the berry pulp (24,190 = 66%) and higher than those numbers found in leaves (22,791 = 63%) and roots (23,370 = 64%) (Table S6). However, a greater number of VmWRKY genes were expressed in the root sample (62 = 0.17%), and the differential expression of 41 VmWRKY genes was high compared to the rest of the samples (Figure 3(B3)). The higher number of genes and their differential expression in the root sample could be explained by the role played by WRKY transcription factors, which are mainly involved in development and stress responses, such as salt and water stress tolerance [61]. Since bilberries are sensitive to these conditions, it is important to pay close attention to the presence of VmWRKYs in directly affected organs such as roots, where VmWRKY54 and VmWRKY11 showed higher expression levels (Figure 3(B4)). This is supported by examples of the effectiveness of these transcription factors in Arabidopsis, where the overexpression of AtWRKY46, GmWRKY13, or VvWRKY11 can positively regulate salt and water stress tolerance [62,63,64,65], and in Nicotiana benthamiana, where the overexpression of GhWRKY41 conferred tolerance to water and salt stress [66].
The complete genome of bilberry (Vaccinium myrtillus L.) allowed the characterization of 69 members of the VmWRKY gene family. Segmental and tandem duplications were detected and could enhance biotic/abiotic resistance in the bilberry genome. The average ages of duplications were identified as 8.27 mya (range 013–16.41) for the tandem and 26.43 mya (range 0.48–83.40) for the segmental duplications, suggesting more recent events for tandem duplications than segmental duplications.

4. Materials and Methods

4.1. Identification of WRKY Proteins in the Bilberry

The complete genome sequence of the bilberry (Vaccinium myrtillus isolate NK2018 v1.0 genome sequence) was downloaded from the Genome Database for Vaccinium (GDV) https://www.vaccinium.org/crop/bilberry (accessed on 17 February 2023) [4]. To identify possible candidate amino acid sequences of VmWRKY bilberry. The WRKY domain HMM model (PF03106) was downloaded from Pfam (http://pfam.xfam.org/family/PF03106/hmm) (accessed on 17 February 2023) [67]. The Hmmer software [68] was also used for similarity search in annotated proteins in bilberry using 1 × 10−3 as the upper limit of the e-value. All obtained protein sequences were examined for the presence of the WRKY domain using the Web CD Search Tool (https://www.ncbi.nlm.nih.gov/Structure/cdd/wrpsb.cgi) (accessed on 17 February 2023) [69]. The ExPASY ProtParam (https://web.expasy.org/protparam/) (accessed on 25 February 2023) [70] was used to predict the isoelectric point (pI), molecular weight (MW), and overall average hydrophobicity (GRAVY) of each VmWRKY.

4.2. Analysis of Cis Elements in VmWRKY Gene Promoters

For each VmWRKY gene, a 2000 bp sequence upstream of the start codon was retrieved from the bilberry (Vaccinium myrtillus isolate NK2018 v1.0 genome sequence) by applying integrative genomics viewer—IGV [71]. This sequence was submitted to the PlantCARE website to investigate cis-acting regulatory elements (http://bioinformatics.psb.ugent.be/webtools/plantcare/html/) (accessed on 17 February 2023) [72].

4.3. Multiple Alignment and Phylogenetic Analyses

A phylogenetic tree was constructed to compare bilberry WRKY proteins. Multiple alignment of WRKY protein sequences was performed with ClustalW software using standard parameters [73]. The phylogenetic tree was constructed using BEAST v.2.5 software [74], with the UPGMA clustering method [75]. A bootstrap analysis was conducted using 10,000,000 replicates, and evolutionary distances were calculated using the JTT matrix-based method [76].

4.4. Analysis of Gene Structure and Identification of Conservation of Motifs

To investigate the diversity and structure of VmWRKY family members, genomic sequences for their exon/intron were used and plotted on TBtools [77], based on bilberry genome annotation information (Vaccinium myrtillus isolate NK2018 v1.0 genome sequence). VmWRKY protein sequences were used to identify conserved motifs using the Expectation Maximization Tool for Motive Elicitation MEME (https://meme-suite.org/meme/tools/meme) (accessed on 20 February 2023) [78]. The parameters were as follows: number of repetitions: any; maximum number of motifs: 20; optimal motif widths: 8 to 50 amino acid residues.

4.5. Chromosomal Localization, Gene Duplication, Ka/Ks Calculation, and Divergence Time Estimation

The chromosomal location image of the VmWRKY genes was generated by the TBtools software [77], according to chromosomal position information provided in the genomic database for Vaccinium-GDV (https://www.vaccinium.org/) (accessed on 30 February 2023). To identify specific tandem duplications of VmWRKYs, the following criteria were used: genes within a 100 kbp region on an individual chromosome with a sequence similarity of ≥70% [79]. The pairwise local alignment calculation of two protein sequences was performed by the Smith–Waterman algorithm of EMBOSS Water (http://www.ebi.ac.uk/Tools/psa/) (accessed on 17 February 2023) [80].
For the calculation of non-synonymous substitutions per site (Ka) and the number of synonymous substitutions per synonymous site (Ks), in addition to comparing the selection pressure, a Ka/Ks ratio greater than 1, less than 1, and equal to 1 represent positive, negative, and neutral selection, respectively. For each pair of genes, the value of Ks was used to estimate the time of divergence in millions of years based on a rate of 6.1 × 10−9 replacements per site per year, and the time of divergence (T) was calculated as T = Ks/(2 × 6.1 × 10−9) × 10−6 million years ago (Mya) [81]. The bioinformatics tool used for genetic divergence was the simple Ka/Ks Calculator (NG) from TBtools-II [77].

4.6. Transcriptomic Analysis of VmWRKY Genes in Bilberry

The expression patterns of VmWRKYs genes were analyzed based on published RNA-seq data on NCBI BioProject ID PRJNA739815 [4]. For the analysis of differential expression, five samples of RNA-seq data were collected that were sequenced by the Illumina HiSeq4000: leaf (SRR14876435), root (SRR14876436), whole berry (SRR14876437), berry flesh (SRR14876438), and berry skin (SRR14876439). Also analyzed were RNA-seq samples from Bioproject ID PRJNA747684 [82]. The experimental design of red and blue light was as follows: control (SRR15179770, SRR15179771, and SRR15179772); 6 days of continuous irradiation (red light) (SRR15179767, SRR15179768, and SRR15179769); 6 days of continuous irradiation (blue light) (SRR15179773, SRR15179774, and SRR15179775).
Data processing was performed in the following three steps. (a) Quality control and adjustment of samples: SRA toolkit [83] was used to download the data samples, FastQC [84] was employed to analyze and visualize the quality of readings, and Trimmomatic ver. 0.39 [85] was applied to remove the low quality and library adaptors. (b) The reads were mapped against the bilberry (V. myrtillus) reference genome [4] using the software HISAT2 [86]. (c) For the counting of total reads aligned by gene in the different libraries, FeatureCounts [87] was used. The quantification analysis (d) was performed using the packages limma, edgeR, and DESeq2 in R software [88]. In this protocol, a FPKM normalization method (fragments per kilo base per million mapped reads) was used, which are counts scaled by the total number of reads and the expression of VmWRKY genes per library. The analysis of gene expression proportion was based on the number of genes annotated in the bilberry genome (36,405 genes = 100%), and the expression of each transcriptome sample and VmWRKY genes was proportional to the total annotated genome genes. A heatmap was produced using TBTools with normalized data (log2 counts per FPKM) [77]. Then, a multi-dimensional scaling (MDS) plot was generated to check the repeatability of the sample and the overall difference between samples. Additionally, a MeanVar plot and Biological Coefficient of Variation (BCV) were calculated.

5. Conclusions

The complete genome of bilberry (Vaccinium myrtillus L.) contains 76 identified and 69 characterized members of the VmWRKY gene family, which are located on twelve chromosomes. Additionally, 12 segmental and four tandem duplications were identified in the bilberry genome. The cis-regulatory elements found in the promoters of VmWRKY genes are putatively involved in various functions related to biotic and abiotic stress responses. The differential expression of VmWRKY genes was induced by light, suggesting its involvement in anthocyanin biosynthesis. The characterization of VmWRKY genes in the bilberry genome, their phylogenetic relationships, and differential expression are important for understanding their role in regulating anthocyanin biosynthesis and their adaptation to different environmental conditions.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/plants12183176/s1. Figure S1: Schematic representation of conserved motifs identified in bilberry VmWRKY proteins. Figure S2: Sequence Logo visualization of the identified conserved motives in the WRKY proteins sequence using Expectation Maximization Tool for Motif Elicitation MEME. Figure S3: Pairwise identities between paralogous pairs of VmWRKY genes—tandem duplicates. Figure S4: Pairwise identities between paralogous pairs of VmWRKY genes—segmental duplicates. Figure S5: The unrooted phylogenetic tree, using WRKY proteins from Arabidopsis thaliana genome and WRKY proteins from Vaccinium myrtillus genome. Figure S6: Ks, Ka, and Ks/Ks calculation and divergence time of duplicated bilberry VmWRKY gene pairs—applying the Nei and Gojobori (NG) method. Figure S7: Expression level distribution for all genes in simulated experimental conditions control, red light and blue light in Vaccinium myrtillus—CummeRbund scatter plots. Figure S8: Analysis of gene expression in samples of Vaccinium myrtillus (leaf, root, whole berry, berry flesh, berry skin)—Volcano plots. Figure S9: Analysis of gene expression in samples of Vaccinium myrtillus (leaf, root, whole berry, berry flesh, berry skin)—CummeRbund scatter plots. Figure S10: Analysis of gene expression in simulated experimental conditions control, red light and blue light in Vaccinium myrtillus—Volcano plots, CummeRbund scatter plots. Table S1: Characterization of WRKYs proteins in the Vaccinium myrtillus genome. Table S2: Pairwise identities between transcription proteins of the VmWRKY14, VmWRKY28, VmWRKY29, VmWRKY32, VmWRKY44, VmWRKY48 and VmWRKY49 genes. Table S3: Identification of the conserved WRKY domains of the transcripts (t1–t2) of the VmWRKY14, VmWRKY28, VmWRKY29, VmWRKY32, VmWRKY44, VmWRKY48 and VmWRKY49 genes. Table S4: Cis-Acting elements analysis in the 2 kb promoters of VmWRKY genes. Table S5: Number of readings aligned by the normative method Fragments Per Kilo Base Million of mapped readings -FPKM in control, red light and blue light conditions in Vaccinium myrtillus. Table S6: Number of readings aligned by the normative method Fragments Per Kilo Base Million of mapped readings -FPKM in leaves, roots, whole berry, berry flesh and berry skin samples in Vaccinium myrtillus.

Author Contributions

W.F. and J.V. conceived the study and prepared figures. W.F., J.V. and V.O.N. conducted the analysis and contributed to the main text of the manuscript. A.C.d.O. led the research. All the authors reviewed the manuscript. C.P. and L.d.M. critically reviewed and added information. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Brazilian Ministry of Science and Technology, National Counsel of Technological and Scientific Development (CNPq); the Coordination for the Improvement of Higher Education Personnel (CAPES); the Rio Grande do Sul State Foundation for Research Support (FAPERGS); the INCT Plant Stress (MCTI) and FAO/IAEA Joint Division.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare that there are no conflict of interest.

References

  1. Zoratti, L.; Klemettila, H.; Jaakola, L. Bilberry (Vaccinium myrtillus L.) Ecotypes. In Nutritional Composition of Fruit Cultivars; Simmonds, M.S.J., Ed.; Elsevier Inc.: Philadelphia, PA, USA, 2016; Volume 13, pp. 535–571. ISBN 9780124081178. [Google Scholar]
  2. Nguyen, N.; Suokas, M.; Karppinen, K.; Vuosku, J.; Jaakola, L.; Häggman, H. Recognition of candidate transcription factors related to bilberry fruit ripening by de novo transcriptome and qRT-PCR analyses. Sci. Rep. 2018, 8, 9943. [Google Scholar] [CrossRef]
  3. Samkumar, A.; Karppinen, K.; Dhakal, B.; Martinussen, I.; Jaakola, L. Insights into sugar metabolism during bilberry (Vaccinium myrtillus L.) fruit development. Physiol. Plant. 2022, 174, e13657. [Google Scholar] [CrossRef] [PubMed]
  4. Wu, C.; Deng, C.; Hilario, E.; Albert, N.W.; Lafferty, D.; Grierson, E.R.P.; Plunkett, B.J.; Elborough, C.; Saei, A.; Günther, C.S.; et al. A chromosome-scale assembly of the bilberry genome identifies a complex locus controlling berry anthocyanin composition. Mol. Ecol. Resour. 2022, 22, 345–360. [Google Scholar] [CrossRef] [PubMed]
  5. Colak, N.; Torun, H.; Gruz, J.; Strnad, M.; Subrtova, M.; Inceer, H.; Ayaz, F. Comparison of Phenolics and Phenolic Acid Profiles in Conjunction with Oxygen Radical Absorbing Capacity (ORAC) in Berries of Vaccinium arctostaphylos L. and V. myrtillus L. Pol. J. Food Nutr. Sci. 2016, 66, 85–91. [Google Scholar] [CrossRef]
  6. Su, Z. Anthocyanins and Flavonoids of Vaccinium L. Pharm. Crop. 2012, 3, 7–37. [Google Scholar] [CrossRef]
  7. Gupta, V.; Estrada, A.D.; Blakley, I.; Reid, R.; Patel, K.; Meyer, M.D.; Andersen, S.U.; Brown, A.F.; Lila, M.A.; E Loraine, A. RNA-Seq analysis and annotation of a draft blueberry genome assembly identifies candidate genes involved in fruit ripening, biosynthesis of bioactive compounds, and stage-specific alternative splicing. Gigascience 2015, 4, 5. [Google Scholar] [CrossRef]
  8. Karppinen, K.; Zoratti, L.; Nguyenquynh, N.; Häggman, H.; Jaakola, L. On the Developmental and Environmental Regulation of Secondary Metabolism in Vaccinium spp. Berries. Front. Plant Sci. 2016, 7, 655. [Google Scholar] [CrossRef]
  9. Zoratti, L.; Jaakola, L.; Häggman, H.; Giongo, L. Modification of Sunlight Radiation through Colored Photo-Selective Nets Affects Anthocyanin Profile in Vaccinium spp. Berries. PLoS ONE 2015, 10, e0135935. [Google Scholar] [CrossRef]
  10. Tao, R.; Bai, S.; Ni, J.; Yang, Q.; Zhao, Y.; Teng, Y. The blue light signal transduction pathway is involved in anthocyanin accumulation in ‘Red Zaosu’ pear. Planta 2018, 248, 37–48. [Google Scholar] [CrossRef]
  11. Guo, X.; Shakeel, M.; Wang, D.; Qu, C.; Yang, S.; Ahmad, S.; Song, Z. Metabolome and transcriptome profiling unveil the mechanisms of light-induced anthocyanin synthesis in rabbiteye blueberry (Vaccinium ashei: Reade). BMC Plant Biol. 2022, 22, 223. [Google Scholar] [CrossRef]
  12. Zhang, J.; Li, S.; An, H.; Zhang, X.; Zhou, B. Integrated transcriptome and metabolome analysis reveals the anthocyanin biosynthesis mechanisms in blueberry (Vaccinium corymbosum L.) leaves under different light qualities. Front. Plant Sci. 2022, 13, 1073332. [Google Scholar] [CrossRef] [PubMed]
  13. Zhu, M.; Chen, G.; Zhou, S.; Tu, Y.; Wang, Y.; Dong, T.; Hu, Z. A New Tomato NAC (NAM/ATAF1/2/CUC2) Transcription Factor, SlNAC4, Functions as a Positive Regulator of Fruit Ripening and Carotenoid Accumulation. Plant Cell Physiol. 2014, 55, 119–135. [Google Scholar] [CrossRef] [PubMed]
  14. Huang, G.; Li, T.; Li, X.; Tan, D.; Jiang, Z.; Wei, Y.; Li, J.; Wang, A. Comparative Transcriptome Analysis of Climacteric Fruit of Chinese Pear (Pyrus ussuriensis) Reveals New Insights into Fruit Ripening. PLoS ONE 2014, 9, e107562. [Google Scholar] [CrossRef]
  15. Rushton, P.J.; Somssich, I.E.; Ringler, P.; Shen, Q.J. WRKY transcription factors. Trends Plant Sci. 2010, 15, 247–258. [Google Scholar] [CrossRef]
  16. Ishihama, N.; Yoshioka, H. Post-translational regulation of WRKY transcription factors in plant immunity. Curr. Opin. Plant Biol. 2012, 15, 431–437. [Google Scholar] [CrossRef]
  17. Ülker, B.; Somssich, I.E. WRKY transcription factors: From DNA binding towards biological function. Curr. Opin. Plant Biol. 2004, 7, 491–498. [Google Scholar] [CrossRef]
  18. Okay, S.; Derelli, E.; Unver, T. Transcriptome-wide identification of bread wheat WRKY transcription factors in response to drought stress. Mol. Genet. Genom. 2014, 289, 765–781. [Google Scholar] [CrossRef]
  19. Viana, V.E.; Busanello, C.; da Maia, L.C.; Pegoraro, C.; de Oliveira, A.C. Activation of rice WRKY transcription factors: An army of stress fighting soldiers? Curr. Opin. Plant Biol. 2018, 45, 268–275. [Google Scholar] [CrossRef]
  20. Rajput, L.S.; Aggarwal, S.K.; Mehta, S.; Kumar, S.; Nataraj, V.; Shivakumar, M.; Maheshwari, H.S.; Yadav, S.; Goswami, D. Role of WRKY Transcription Factor Superfamily in Plant Disease Management. In Plant Stress Biol; Giri, B., Sharma, M.P., Eds.; Springer: Singapore, 2020. [Google Scholar] [CrossRef]
  21. Brand, L.H.; Fischer, N.M.; Harter, K.; Kohlbacher, O.; Wanke, D. Elucidating the evolutionary conserved DNA-binding specificities of WRKY transcription factors by molecular dynamics and in vitro binding assays. Nucleic Acids Res. 2013, 41, 9764–9778. [Google Scholar] [CrossRef] [PubMed]
  22. 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]
  23. Chen, P.; Liu, Q.-Z. Genome-wide characterization of the WRKY gene family in cultivated strawberry (Fragaria × ananassa Duch.) and the importance of several group III members in continuous cropping. Sci. Rep. 2019, 9, 8423. [Google Scholar] [CrossRef]
  24. Felipez, W.; de Freitas, K.E.J.; dos Santos, R.S.; Yamamoto, R.R.; de Oliveira, A.C. The roles of WRKY transcription factors in Malus spp. and Pyrus spp. Funct. Integr. Genom. 2022, 22, 713–729. [Google Scholar] [CrossRef] [PubMed]
  25. Yang, T.; Li, K.; Hao, S.; Zhang, J.; Song, T.; Tian, J.; Yao, Y. The Use of RNA Sequencing and Correlation Network Analysis to Study Potential Regulators of Crabapple Leaf Color Transformation. Plant Cell Physiol. 2018, 59, 1027–1042. [Google Scholar] [CrossRef] [PubMed]
  26. Hu, J.; Fang, H.; Wang, J.; Yue, X.; Su, M.; Mao, Z.; Zou, Q.; Jiang, H.; Guo, Z.; Yu, L.; et al. Ultraviolet B-induced MdWRKY72 expression promotes anthocyanin synthesis in apple. Plant Sci. 2020, 292, 110377. [Google Scholar] [CrossRef] [PubMed]
  27. Su, M.; Zuo, W.; Wang, Y.; Liu, W.; Zhang, Z.; Wang, N.; Chen, X. The WKRY transcription factor MdWRKY75 regulates anthocyanins accumulation in apples (Malus domestica). Funct. Plant Biol. 2022, 49, 799–809. [Google Scholar] [CrossRef]
  28. Gao, J.; Peng, H.; Chen, F.; Luo, M.; Li, W. Genome-wide analysis of transcription factors related to anthocyanin biosynthesis in carmine radish (Raphanus sativus L.) fleshy roots. PeerJ 2019, 7, e8041. [Google Scholar] [CrossRef]
  29. Li, C.; Wu, J.; Hu, K.-D.; Wei, S.-W.; Sun, H.-Y.; Hu, L.-Y.; Han, Z.; Yao, G.-F.; Zhang, H. PyWRKY26 and PybHLH3 cotargeted the PyMYB114 promoter to regulate anthocyanin biosynthesis and transport in red-skinned pears. Hortic. Res. 2020, 7, 37. [Google Scholar] [CrossRef] [PubMed]
  30. Wang, Z.; Du, H.; Zhai, R.; Song, L.; Ma, F.; Xu, L. Transcriptome Analysis Reveals Candidate Genes Related to Color Fading of ‘Red Bartlett’ (Pyrus communis L.). Front. Plant Sci. 2017, 8, 455. [Google Scholar] [CrossRef]
  31. Wang, N.; Liu, W.; Zhang, T.; Jiang, S.; Xu, H.; Wang, Y.; Zhang, Z.; Wang, C.; Chen, X. Transcriptomic Analysis of Red-Fleshed Apples Reveals the Novel Role of MdWRKY11 in Flavonoid and Anthocyanin Biosynthesis. J. Agric. Food Chem. 2018, 66, 7076–7086. [Google Scholar] [CrossRef]
  32. Su, M.; Wang, S.; Li, C.; Zhang, Z.; Wang, N.; Li, B.; Chen, X. Ultraviolet-B-induced MdWRKY71-L expression regulates anthocyanin synthesis in apple. Environ. Exp. Bot. 2022, 201, 105000. [Google Scholar] [CrossRef]
  33. Shi, B.; Wu, H.; Zhu, W.; Zheng, B.; Wang, S.; Zhou, K.; Qian, M. Genome-Wide Identification and Expression Analysis of WRKY Genes during Anthocyanin Biosynthesis in the Mango (Mangifera indica L.). Agriculture 2022, 12, 821. [Google Scholar] [CrossRef]
  34. Wang, Y.; Feng, L.; Zhu, Y.; Li, Y.; Yan, H.; Xiang, Y. Comparative genomic analysis of the WRKY III gene family in populus, grape, arabidopsis and rice. Biol. Direct 2015, 10, 48. [Google Scholar] [CrossRef]
  35. Agarwal, P.; Reddy, M.P.; Chikara, J. WRKY: Its structure, evolutionary relationship, DNA-binding selectivity, role in stress tolerance and development of plants. Mol. Biol. Rep. 2011, 38, 3883–3896. [Google Scholar] [CrossRef]
  36. 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]
  37. Wu, Y.; Zhang, S.; Huang, X.; Lyu, L.; Li, W.; Wu, W. Genome-wide identification of WRKY gene family members in black raspberry and their response to abiotic stresses. Sci. Hortic. 2022, 304, 111338. [Google Scholar] [CrossRef]
  38. Huang, X.; Li, K.; Xu, X.; Yao, Z.; Jin, C.; Zhang, S. Genome-wide analysis of WRKY transcription factors in white pear (Pyrus bretschneideri) reveals evolution and patterns under drought stress. BMC Genom. 2015, 16, 1104. [Google Scholar] [CrossRef]
  39. Sun, C.; Palmqvist, S.; Olsson, H.; Borén, M.; Ahlandsberg, S.; Jansson, C. A Novel WRKY Transcription Factor, SUSIBA2, Participates in Sugar Signaling in Barley by Binding to the Sugar-Responsive Elements of the iso1 Promoter. Plant Cell 2003, 15, 2076–2092. [Google Scholar] [CrossRef] [PubMed]
  40. Xu, X.; Chen, C.; Fan, B.; Chen, Z. Physical and Functional Interactions between and WRKY60 Transcription Factors. Plant Cell 2006, 18, 1310–1326. [Google Scholar] [CrossRef] [PubMed]
  41. 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]
  42. Tao, Z.; Kou, Y.; Liu, H.; Li, X.; Xiao, J.; Wang, S. OsWRKY45 alleles play different roles in abscisic acid signalling and salt stress tolerance but similar roles in drought and cold tolerance in rice. J. Exp. Bot. 2011, 62, 4863–4874. [Google Scholar] [CrossRef] [PubMed]
  43. Schwechheimer, C.; Zourelidou, M.; Bevan, M.W. Plant Transcription Factor Studies. Annu. Rev. Plant Biol. 1998, 49, 127–150. [Google Scholar] [CrossRef]
  44. Chen, X.; Li, C.; Wang, H.; Guo, Z. WRKY transcription factors: Evolution, binding, and action. Phytopathol. Res. 2019, 1, 13. [Google Scholar] [CrossRef]
  45. Bao, F.; Ding, A.; Cheng, T.; Wang, J.; Zhang, Q. Genome-Wide Analysis of Members of the WRKY Gene Family and Their Cold Stress Response in Prunus mume. Genes 2019, 10, 911. [Google Scholar] [CrossRef] [PubMed]
  46. Xie, X.; He, Z.; Chen, N.; Tang, Z.; Wang, Q.; Cai, Y. The Roles of Environmental Factors in Regulation of Oxidative Stress in Plant. BioMed Res. Int. 2019, 2019, 21–27. [Google Scholar] [CrossRef] [PubMed]
  47. Kalde, M.; Barth, M.; Somssich, I.E.; Lippok, B.; Shen, L.; Yang, S.; Yang, F.; Guan, D.; He, S.; De Cuyper, C.; et al. Members of the Arabidopsis WRKY Group III Transcription Factors Are Part of Different Plant Defense Signaling Pathways. Mol. Plant-Microbe Interact. 2007, 16, 295–305. [Google Scholar] [CrossRef]
  48. Zhou, H.; Li, Y.; Zhang, Q.; Ren, S.; Shen, Y.; Qin, L.; Xing, Y. Genome-Wide Analysis of the Expression of WRKY Family Genes in Different Developmental Stages of Wild Strawberry (Fragaria vesca) Fruit. PLoS ONE 2016, 11, e0154312. [Google Scholar] [CrossRef]
  49. Meng, D.; Li, Y.; Bai, Y.; Li, M.; Cheng, L. Genome-wide identification and characterization of WRKY transcriptional factor family in apple and analysis of their responses to waterlogging and drought stress. Plant Physiol. Biochem. 2016, 103, 71–83. [Google Scholar] [CrossRef]
  50. Ross, C.A.; Liu, Y.; Shen, Q.J. The WRKY Gene Family in Rice (Oryza sativa). J. Integr. Plant Biol. 2007, 49, 827–842. [Google Scholar] [CrossRef]
  51. Tripathi, P.; Rabara, R.C.; Rushton, P.J. A systems biology perspective on the role of WRKY transcription factors in drought responses in plants. Planta 2014, 239, 255–266. [Google Scholar] [CrossRef]
  52. Chen, C.; Xie, F.; Shah, K.; Hua, Q.; Chen, J.; Zhang, Z.; Zhao, J.; Hu, G.; Qin, Y. Genome-Wide Identification of WRKY Gene Family in Pitaya Reveals the Involvement of HmoWRKY42 in Betalain Biosynthesis. Int. J. Mol. Sci. 2022, 23, 10568. [Google Scholar] [CrossRef]
  53. De Grassi, A.; Lanave, C.; Saccone, C. Genome duplication and gene-family evolution: The case of three OXPHOS gene families. Gene 2008, 421, 1–6. [Google Scholar] [CrossRef] [PubMed]
  54. 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]
  55. Nan, H.; Li, W.; Lin, Y.-L.; Gao, L.-Z. Genome-Wide Analysis of WRKY Genes and Their Response to Salt Stress in the Wild Progenitor of Asian Cultivated Rice, Oryza rufipogon. Front. Genet. 2020, 11, 359. [Google Scholar] [CrossRef]
  56. Bulgakov, V.P.; Avramenko, T.V.; Tsitsiashvili, G.S. Critical analysis of protein signaling networks involved in the regulation of plant secondary metabolism: Focus on anthocyanins. Crit. Rev. Biotechnol. 2017, 37, 685–700. [Google Scholar] [CrossRef]
  57. Gangappa, S.N.; Botto, J.F. The Multifaceted Roles of HY5 in Plant Growth and Development. Mol. Plant 2016, 9, 1353–1365. [Google Scholar] [CrossRef]
  58. Jaakola, L. New insights into the regulation of anthocyanin biosynthesis in fruits. Trends Plant Sci. 2013, 18, 477–483. [Google Scholar] [CrossRef] [PubMed]
  59. Lin, Y.; Wang, Y.; Li, B.; Tan, H.; Li, D.; Li, L.; Liu, X.; Han, J.; Meng, X. Comparative transcriptome analysis of genes involved in anthocyanin synthesis in blueberry. Plant Physiol. Biochem. 2018, 127, 561–572. [Google Scholar] [CrossRef] [PubMed]
  60. Yan, J.; Qian, L.; Zhu, W.; Qiu, J.; Lu, Q.; Wang, X.; Wu, Q.; Ruan, S.; Huang, Y. Integrated analysis of the transcriptome and metabolome of purple and green leaves of Tetrastigma hemsleyanum reveals gene expression patterns involved in anthocyanin biosynthesis. PLoS ONE 2020, 15, e0230154. [Google Scholar] [CrossRef]
  61. 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]
  62. Zhou, Q.-Y.; Tian, A.-G.; Zou, H.-F.; Xie, Z.-M.; Lei, G.; Huang, J.; Wang, C.-M.; Wang, H.-W.; Zhang, J.-S.; Chen, S.-Y. Soybean WRKY-type transcription factor genes, GmWRKY13, GmWRKY21, and GmWRKY54, confer differential tolerance to abiotic stresses in transgenic Arabidopsis plants. Plant Biotechnol. J. 2008, 6, 486–503. [Google Scholar] [CrossRef]
  63. 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] [PubMed]
  64. Zhu, J.; Wen, D.; Yu, Y.; Meudt, H.M.; Nakhleh, L. Bayesian Inference of Phylogenetic Networks from Bi-Allelic Genetic Markers; Bayesian inference of phylogenetic networks from bi-allelic genetic markers. PLoS Comput. Biol. 2018, 14, e1005932. [Google Scholar] [CrossRef] [PubMed]
  65. Ding, Z.J.; Yan, J.Y.; Li, C.X.; Li, G.X.; Wu, Y.R.; Zheng, S.J. Transcription factor WRKY46 modulates the development of Arabidopsis lateral roots in osmotic/salt stress conditions via regulation of ABA signaling and auxin homeostasis. Plant J. 2015, 84, 56–69. [Google Scholar] [CrossRef] [PubMed]
  66. Chu, X.; Wang, C.; Chen, X.; Lu, W.; Li, H.; Wang, X.; Hao, L.; Guo, X. The Cotton WRKY Gene GhWRKY41 Positively Regulates Salt and Drought Stress Tolerance in Transgenic Nicotiana benthamiana. PLoS ONE 2015, 10, e0143022. [Google Scholar] [CrossRef]
  67. Mistry, J.; Chuguransky, S.; Williams, L.; Qureshi, M.; Salazar, G.A.; Sonnhammer, E.L.L.; Tosatto, S.C.; Paladin, L.; Raj, S.; Richardson, L.J.; et al. Pfam: The protein families database in 2021. Nucleic Acids Res. 2021, 49, D412–D419. [Google Scholar] [CrossRef]
  68. Finn, R.D.; Clements, J.; Eddy, S.R. HMMER web server: Interactive sequence similarity searching. Nucleic Acids Res. 2011, 39, 29–37. [Google Scholar] [CrossRef]
  69. Lu, S.; Wang, J.; Chitsaz, F.; Derbyshire, M.K.; Geer, R.C.; Gonzales, N.R.; Gwadz, M.; I Hurwitz, D.; Marchler, G.H.; Song, J.S.; et al. CDD/SPARCLE: The conserved domain database in 2020. Nucleic Acids Res. 2020, 48, D265–D268. [Google Scholar] [CrossRef]
  70. Gasteiger, E.; Gattiker, A.; Hoogland, C.; Ivanyi, I.; Appel, R.D.; Bairoch, A. ExPASy: The proteomics server for in-depth protein knowledge and analysis. Nucleic Acids Res. 2003, 31, 3784–3788. [Google Scholar] [CrossRef]
  71. Robinson, J.T.; Thorvaldsdóttir, H.; Winckler, W.; Guttman, M.; Lander, E.S.; Getz, G.; Mesirov, J.P. Integrative genomics viewer. Nat. Biotechnol. 2011, 29, 24–26. [Google Scholar] [CrossRef]
  72. Lescot, M.; Déhais, P.; Thijs, G.; Marchal, K.; Moreau, Y.; Van de Peer, Y.; Rouzé, P.; Rombauts, S. PlantCARE, a database of plant cis-acting regulatory elements and a portal to tools for in silico analysis of promoter sequences. Nucleic Acids Res. 2002, 30, 325–327. [Google Scholar] [CrossRef]
  73. Larkin, M.A.; Blackshields, G.; Brown, N.P.; Chenna, R.; McGettigan, P.A.; McWilliam, H.; Valentin, F.; Wallace, I.M.; Wilm, A.; Lopez, R.; et al. Clustal W and Clustal X version 2.0. Bioinformatics 2007, 23, 2947–2948. [Google Scholar] [CrossRef] [PubMed]
  74. Bouckaert, R.; Vaughan, T.G.; Barido-Sottani, J.; Duchêne, S.; Fourment, M.; Gavryushkina, A.; Heled, J.; Jones, G.; Kühnert, D.; De Maio, N.; et al. BEAST 2.5: An advanced software platform for Bayesian evolutionary analysis. PLoS Comput. Biol. 2019, 15, e1006650. [Google Scholar] [CrossRef] [PubMed]
  75. Sneath, P.H.A.; Sokal, R.R. Numerical Taxonomy: The Principles and Practice of Numerical Classification; Freeman: San Francicso, CA, USA, 1973; p. 18. [Google Scholar]
  76. Jones, D.T.; Taylor, W.R.; Thornton, J.M. The rapid generation of mutation data matrices from protein sequences. Comput. Appl. Biosci. 1992, 8, 275–282. [Google Scholar] [CrossRef] [PubMed]
  77. Chen, C.J.; Chen, H.; Zhang, Y.; Thomas, H.R.; Frank, M.H.; He, Y.H.; Xia, R. TBtools: An Integrative Toolkit Developed for Interactive Analyses of Big Biological Data. Mol. Plant 2020, 13, 1194–1202. [Google Scholar] [CrossRef]
  78. Bailey, T.L.; Williams, N.; Misleh, C.; Li, W.W. MEME: Discovering and analyzing DNA and protein sequence motifs. Nucleic Acids Res. 2006, 34, 369–373. [Google Scholar] [CrossRef]
  79. Hu, L.; Liu, S. Genome-wide identification and phylogenetic analysis of the ERF gene family in cucumbers. Genet. Mol. Biol. 2011, 34, 624–634. [Google Scholar] [CrossRef]
  80. Madeira, F.; Pearce, M.; Tivey, A.R.N.; Basutkar, P.; Lee, J.; Edbali, O.; Madhusoodanan, N.; Kolesnikov, A.; Lopez, R. Search and sequence analysis tools services from EMBL-EBI in 2022. Nucleic Acids Res. 2022, 50, W276–W279. [Google Scholar] [CrossRef]
  81. Conery, J.S.; Lynch, M. Nucleotide Substitutions and the Evolution of Duplicate Genes. Biocomputing 2001, 167–178. [Google Scholar] [CrossRef]
  82. Samkumar, A.; Jones, D.; Karppinen, K.; Dare, A.P.; Sipari, N.; Espley, R.V.; Martinussen, I.; Jaakola, L. Red and blue light treatments of ripening bilberry fruits reveal differences in signalling through abscisic acid-regulated anthocyanin biosynthesis. Plant Cell Environ. 2021, 44, 3227–3245. [Google Scholar] [CrossRef]
  83. Leinonen, R.; Sugawara, H.; Shumway, M. The Sequence Read Archive. Nucleic Acids Res. 2011, 39, 2010–2012. [Google Scholar] [CrossRef]
  84. Wingett, S.W.; Andrews, S. FastQ Screen: A tool for multi-genome mapping and quality control. F1000Research 2018, 7, 1338. [Google Scholar] [CrossRef] [PubMed]
  85. Bolger, A.M.; Lohse, M.; Usadel, B. Trimmomatic: A flexible trimmer for Illumina sequence data. Bioinformatics 2014, 30, 2114–2120. [Google Scholar] [CrossRef] [PubMed]
  86. Pertea, M.; Kim, D.; Pertea, G.M.; Leek, J.T.; Salzberg, S.L. Transcript-level expression analysis of RNA-seq experiments with HISAT, StringTie and Ballgown. Nat. Protoc. 2016, 11, 1650–1667. [Google Scholar] [CrossRef] [PubMed]
  87. Liao, Y.; Smyth, G.K.; Shi, W. Feature Counts: An efficient general purpose program for assigning sequence reads to genomic features. Bioinformatics 2014, 30, 923–930. [Google Scholar] [CrossRef]
  88. Law, C.W.; Alhamdoosh, M.; Su, S.; Smyth, G.K.; Ritchie, M.E. RNA-seq analysis is easy as 1-2-3 with limma, Glimma and edgeR. F1000Research 2018, 5, 1408. [Google Scholar] [CrossRef]
Figure 1. Phylogenetic relationships and structure of genes encoding the bilberry VmWRKY proteins: (A) The unrooted tree was generated with the BEAST program using the full-length amino acid sequences of the 69 bilberry VmWRKY proteins by the UPGMA method, with 1,000,000 bootstrap replications. VmWRKY protein subfamilies (I, IIa, IIb, IIc, IId, IIe, III, and IV). (B) Exon/intron organization of bilberry VmWRKY genes. Yellow boxes represent exons, and black lines represent introns. Untranslated regions (UTRs) are indicated by blue boxes. The sizes of exons and introns can be estimated using the scale at the bottom. (C) Schematic representation of conserved motifs in bilberry VmWRKY proteins, elucidated from publicly available data (NCBI CDD Domain–Pfam–18,271 PSSMs). Each colored rectangular box represents a motif with the given name and motif consensus.
Figure 1. Phylogenetic relationships and structure of genes encoding the bilberry VmWRKY proteins: (A) The unrooted tree was generated with the BEAST program using the full-length amino acid sequences of the 69 bilberry VmWRKY proteins by the UPGMA method, with 1,000,000 bootstrap replications. VmWRKY protein subfamilies (I, IIa, IIb, IIc, IId, IIe, III, and IV). (B) Exon/intron organization of bilberry VmWRKY genes. Yellow boxes represent exons, and black lines represent introns. Untranslated regions (UTRs) are indicated by blue boxes. The sizes of exons and introns can be estimated using the scale at the bottom. (C) Schematic representation of conserved motifs in bilberry VmWRKY proteins, elucidated from publicly available data (NCBI CDD Domain–Pfam–18,271 PSSMs). Each colored rectangular box represents a motif with the given name and motif consensus.
Plants 12 03176 g001
Figure 2. Chromosomal map and coordinates of VmWRKY gene duplication events: The identity of each linkage group is indicated in the central part of each bar. The putative segmental duplicated genes are connected by red color lines, and the duplicated gene pair in tandem is represented by the blue color on the chromosome.
Figure 2. Chromosomal map and coordinates of VmWRKY gene duplication events: The identity of each linkage group is indicated in the central part of each bar. The putative segmental duplicated genes are connected by red color lines, and the duplicated gene pair in tandem is represented by the blue color on the chromosome.
Plants 12 03176 g002
Figure 3. (A) Gene expression of bilberry transcriptomic light samples (under control, red light, and blue light conditions, measured in FPKM): (1) multi-dimensional scaling (MDS) plot of the light samples; (2) VmWRKY gene expression common to each sample shown in a Venn diagram; (3) differential expression of VmWRKY genes in each sample (values presented in log2, ranging from 0.40 to 2.30); (4) highly expressed VmWRKY genes in all samples (values presented in log2, ranging from 0.35 to 5.66). (B) Gene expression of bilberry transcriptomic organ samples (in leaves, roots, whole berries, berry flesh, and berry skin, measured in FPKM): (1) multi-dimensional scaling (MDS) plot of the organ samples (roots, leaves, berry skin, whole berry, and berry flesh); (2) VmWRKY gene expression common to each sample shown in a Venn diagram; (3) differential expression of VmWRKY genes in each sample (values presented in log2, ranging from −2.00 to 2.00); (4) highly expressed VmWRKY genes in all samples (values presented in log2, ranging from −1.50 to 3.00).
Figure 3. (A) Gene expression of bilberry transcriptomic light samples (under control, red light, and blue light conditions, measured in FPKM): (1) multi-dimensional scaling (MDS) plot of the light samples; (2) VmWRKY gene expression common to each sample shown in a Venn diagram; (3) differential expression of VmWRKY genes in each sample (values presented in log2, ranging from 0.40 to 2.30); (4) highly expressed VmWRKY genes in all samples (values presented in log2, ranging from 0.35 to 5.66). (B) Gene expression of bilberry transcriptomic organ samples (in leaves, roots, whole berries, berry flesh, and berry skin, measured in FPKM): (1) multi-dimensional scaling (MDS) plot of the organ samples (roots, leaves, berry skin, whole berry, and berry flesh); (2) VmWRKY gene expression common to each sample shown in a Venn diagram; (3) differential expression of VmWRKY genes in each sample (values presented in log2, ranging from −2.00 to 2.00); (4) highly expressed VmWRKY genes in all samples (values presented in log2, ranging from −1.50 to 3.00).
Plants 12 03176 g003
Table 1. Information on VmWRKY genes in the bilberry genome (Vaccinium myrtillus L.).
Table 1. Information on VmWRKY genes in the bilberry genome (Vaccinium myrtillus L.).
Gene NameGene IdentifierGroupChr.Start EndStrandCDS(aa)MWpII. IndexGRAVY
Vmy01g15360VmWRKY1IChr1562,585565,972+202867573,184.11015.861457.63496−0.78104
Vmy01g15494VmWRKY2IChr11,805,0331,810,979170456761,989.00226.792652.19947−0.75467
Vmy01g15496VmWRKY3IChr11,815,4001,821,385135645148,425.0534.708362.64989−0.84169
Vmy01g16077VmWRKY4IIbChr18,788,1098,792,213+161753858,725.49976.595845.7948−0.70985
Vmy01g17102VmWRKY5IChr126,099,66126,103,183+142547451,364.86638.940745.66751−0.80696
Vmy01g17169VmWRKY6IicChr127,189,67027,191,92387629132,429.51965.095558.33574−0.73814
Vmy01g18042VmWRKY7IicChr139,683,97639,685,598+101733837,888.74825.965261.04822−0.88817
Vmy02g31145VmWRKY8IicChr225,519,14725,521,976+87629132,444.18196.452870.26117−1.01993
Vmy02g31720VmWRKY9IChr232,825,77832,832,842+161153658,052.0076.611658.72351−0.81287
Vmy02g32314VmWRKY10IibChr239,960,78239,963,444+157852557,472.70525.264748.67124−0.84019
Vmy03g6652VmWRKY11IiaChr33,347,5993,349,848+100833537,183.48697.570754.8606−0.66806
Vmy03g7279VmWRKY12IicChr310,749,22910,756,21080726829,564.14465.462262.2306−0.84179
Vmy03g7725VmWRKY13IicChr317,418,68117,421,41453717820,532.66359.627533.60225−1.04831
Vmy03g7865VmWRKY14IieChr320,460,92720,465,14786728732,027.19935.176964.9216−0.81986
Vmy03g9094VmWRKY15IicChr337,596,33437,598,020+90630133,881.23845.661465.33953−0.84153
Vmy04g26811VmWRKY16IChr42,341,7632,347,090178259364,536.40456.407451.08398−0.73693
Vmy04g26879VmWRKY17IibChr42,979,6262,984,279+146748853,622.54989.045459.18258−0.72992
Vmy04g27403VmWRKY18IChr48,533,6198,536,143143747852,408.70186.744155.2864−1.03536
Vmy04g27528VmWRKY19IibChr410,145,59910,147,984+157552457,082.07276.326745.18626−0.71164
Vmy04g27619VmWRKY20IieChr411,366,86311,367,48662420723,294.81016.599460.09614−0.81063
Vmy04g28682VmWRKY21IicChr430,876,94230,884,97284328031,697.02496.305153.21964−0.76571
Vmy04g28697VmWRKY22IicChr431,197,90731,202,208+56418721,115.91085.789658.50588−0.97701
Vmy04g29204VmWRKY23IChr438,844,92438,855,142+38451356,570.00325.542853.90975−0.86667
Vmy05g13713VmWRKY24IibChr534,491,95634,498,576173457763,130.92696.663950.274−0.78943
Vmy05g14091VmWRKY25IicChr538,520,96638,524,76239613115,231.11719.721137.99313−1.08397
Vmy06g33434VmWRKY26IicChr6535,034539,54279826529,556.11689.188662.50906−0.74113
Vmy06g33447VmWRKY27IieChr6641,618643,25897832537,054.01264.683876.58988−1.212
Vmy06g34052VmWRKY28IidChr66,251,5286262,173151250357,035.15039.830346.33002−0.57475
Vmy06g34072VmWRKY29IChr66,434,3776,437,276+171657162,463.30817.710365.48984−0.9387
Vmy06g34731VmWRKY30IieChr614,371,00914,374,552145248352,681.34655.688453.77598−0.87433
Vmy06g35042VmWRKY31IibChr618,292,32318,295,336+176158663,759.13846.352649.9041−0.69369
Vmy06g35510VmWRKY32IieChr625,953,64925,954,811+81627130,225.8755.213258.61476−0.86236
Vmy07g24811VmWRKY33IieChr710,670,37510,672,822101733837,204.42435.268372.40296−0.86982
Vmy07g25848VmWRKY34IidChr728,979,07328980568+103834537,872.46479.701344.33942−0.63072
Vmy08g19321VmWRKY35IChr88,189,3048,193,703156652157,275.12677.659453.98618−0.78061
Vmy08g19634VmWRKY36IidChr812,539,06712,540,851+97532435,083.26289.582949.67099−0.5071
Vmy08g20363VmWRKY37IibChr826,539,53326,543,458125741846,326.0556.775639.74285−0.76555
Vmy08g20364VmWRKY38IibChr826,567,08726,572,765131443748,350.21478.017844.91215−0.78558
Vmy08g20862VmWRKY39IIIChr833,668,48733,685,457+201367075,108.38686.803552.55225−0.57209
Vmy08g20951VmWRKY40IIIChr835,084,05035,087,270+72624127,218.48979.024045.36892−0.64398
Vmy08g20953VmWRKY41IIIChr835,124,87335,128,367+72624127,276.52588.881145.29091−0.66598
Vmy08g20957VmWRKY42IIIChr835,152,03435,152,55033611113,001.34696.884533.74234−0.95405
Vmy08g20958VmWRKY43IIIChr835,155,90735,158,28635111613,315.64926.395448.45517−0.75517
Vmy09g21848VmWRKY44IiaChr95,768,0525,770,593+98732836,404.08567.602248.30549−0.76738
Vmy09g23773VmWRKY45IChr935,699,79535,704,437165955261,427.45317.172759.53261−0.97409
Vmy09g23833VmWRKY46IIIChr936,446,62736,448,661100833537,488.91475.479657.01851−0.78657
Vmy09g23834VmWRKY47IIIChr936,456,42636,458,32298132636,383.85995.944256.7635−0.7227
Vmy10g9485VmWRKY48IiaChr104,985,3364,988,12787028932,320.86296.325339.20208−0.76332
Vmy10g9486VmWRKY49IiaChr105,039,4285,040,970+54017919,928.30048.656141.61508−0.54078
Vmy11g4858VmWRKY50IIIChr1122,694,87822,709,418+188762868,221.5885.782060.61369−0.63392
Vmy11g4864VmWRKY51IIIChr1122,834,00622,838,06395131635,353.0696.069457.9981−0.70475
Vmy11g5266VmWRKY52IieChr1127,599,93027,601,824+105034937,869.3195.280260.43413−0.75387
Vmy11g5322VmWRKY53IIIChr1128,097,85828,099,773115238342,063.97355.960854.54885−0.72298
Vmy12g1272VmWRKY54IIIChr1211,342,51911,344,263+107135640,375.96775.541446.38343−0.88736
Vmy12g1322VmWRKY55IicChr1211,966,50211,968,411+49816518,670.94359.800032.01515−0.85818
Vmy12g1324VmWRKY56IieChr1211,987,94911,989,52895731834,707.04165.425754.09343−0.53899
Vmy12g2389VmWRKY57IChr1228,589,92328,593,312164454760,328.89288.426156.26289−0.9947
Vmy12g2722VmWRKY58IChr1232,281,67032,282,541+69323025,178.26666.443866.10478−0.94739
Vmy12g2723VmWRKY59IidChr1232,286,55732,288,334+102334038,242.11559.785158.70412−0.78235
Vmy12g371VmWRKY60IIIChr12776,559778,12296332036,043.81894.894038.90625−0.60844
Vmy12g391VmWRKY61IieChr12961,923971,203222374082,707.56434.959253.45962−0.44878
Vmy12g427VmWRKY62IicChr121,332,1361,333,78859719822,659.34699.157349.19899−0.82424
Vmy12g435VmWRKY63IieChr121,390,9281,392,56396932235,739.12325.416961.10155−0.85621
Vmy12g744VmWRKY64IicChr124,792,4184,798,54770823526,134.36287.128850.72255−0.97064
VmyS4054g6302VmWRKY65Iib405411,36114,074+116438741,562.63568.956547.0261−0.54419
VmyS5938g36257VmWRKY66I5938264,671269,317+112837541,777.25066.907744.0112−0.72293
VmyS6208g149VmWRKY67Iic620830,68348,873181260366,813.48359.310643.03997−0.62919
VmyS7930g26553VmWRKY68IV793047,31057,551+148249356,297.42899.015954.64016−0.52312
VmyS8810g36160VmWRKY69IV881023,41630,743117038943,454.7445.650754.54267−0.70103
aa = Amino acid; pI = isoelectric point; MW = molecular weight; GRAVY = large average hydrophobicity; I. index = instability index.
Table 2. Ks, Ka, and Ka/Ks calculation and divergence time of duplicated bilberry VmWRKY gene pairs.
Table 2. Ks, Ka, and Ka/Ks calculation and divergence time of duplicated bilberry VmWRKY gene pairs.
Duplicate Gene Pair KaKsKa_KsDuplicated TypeTime (Mya) *
VmWRKY37/VmWRKY38NaNNaNNaNTandem
VmWRKY40/VmWRKY410.00182290.01160.1571441Tandem0.138938872
VmWRKY46/VmWRKY47NaNNaNNaNTandem
VmWRKY48/VmWRKY490.21536290.43471750.4954089Tandem16.41485794
VmWRKY4/VmWRKY190.18633881.20265910.154939Segmental14.20264977
VmWRKY65/VmWRKY240.01177730.11732450.1003819Segmental0.897657317
VmWRKY11/VmWRKY440.17376590.73098320.2377154Segmental13.24435556
VmWRKY14/VmWRKY323.34641 **4.13312 **0.80965 **Segmental3.5265 **
VmWRKY61/VmWRKY631.09429031.70980920.6400073Segmental83.40627188
VmWRKY59/VmWRKY28NaNNaNNaNSegmental
VmWRKY67/VmWRKY220.9085855NaNNaNSegmental69.25193994
VmWRKY13/VmWRKY25NaNNaNNaNSegmental
VmWRKY62/VmWRKY55NaNNaNNaNSegmental
VmWRKY2/VmWRKY16NaNNaNNaNSegmental
VmWRKY1/VmWRKY30.0063830.04271620.1494286Segmental0.486510442
VmWRKY58/VmWRKY290NaNNaNSegmental0
* Mya, Million years ago. ** The values were computed using the KaKs_Calculator2.0 program, applying the Nei and Gojobori (NG) method. NaN = Not a number (refers to an undefined value or a result that cannot be calculated).
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

Felipez, W.; Villavicencio, J.; Nizolli, V.O.; Pegoraro, C.; da Maia, L.; Costa de Oliveira, A. Genome-Wide Identification of Bilberry WRKY Transcription Factors: Go Wild and Duplicate. Plants 2023, 12, 3176. https://doi.org/10.3390/plants12183176

AMA Style

Felipez W, Villavicencio J, Nizolli VO, Pegoraro C, da Maia L, Costa de Oliveira A. Genome-Wide Identification of Bilberry WRKY Transcription Factors: Go Wild and Duplicate. Plants. 2023; 12(18):3176. https://doi.org/10.3390/plants12183176

Chicago/Turabian Style

Felipez, Winder, Jennifer Villavicencio, Valeria Oliveira Nizolli, Camila Pegoraro, Luciano da Maia, and Antonio Costa de Oliveira. 2023. "Genome-Wide Identification of Bilberry WRKY Transcription Factors: Go Wild and Duplicate" Plants 12, no. 18: 3176. https://doi.org/10.3390/plants12183176

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