Next Article in Journal
OsBSK3 Positively Regulates Grain Length and Weight by Inhibiting the Phosphatase Activity of OsPPKL1
Next Article in Special Issue
The S-Gene YUC6 Pleiotropically Determines Male Mating Type and Pollen Size in Heterostylous Turnera (Passifloraceae): A Novel Neofunctionalization of the YUCCA Gene Family
Previous Article in Journal
A Small Gtp-Binding Protein GhROP3 Interacts with GhGGB Protein and Negatively Regulates Drought Tolerance in Cotton (Gossypium hirsutum L.)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Comprehensive Study of the WRKY Transcription Factor Family in Strawberry

by
José Garrido-Gala
1,†,
José-Javier Higuera
2,†,
Antonio Rodríguez-Franco
2,
Juan Muñoz-Blanco
2,
Francisco Amil-Ruiz
3 and
José L. Caballero
2,*
1
Phytoplant Research SLU, 14071 Córdoba, Spain
2
Departamento de Bioquímica y Biología Molecular, Campus Universitario de Rabanales y Campus de Excelencia Internacional Agroalimentario ceiA3, Edificio Severo Ochoa-C6, Universidad de Córdoba, 14071 Córdoba, Spain
3
Unidad de Bioinformática, Servicio Central de Apoyo a la Investigación (SCAI), Universidad de Córdoba, 14071 Córdoba, Spain
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Plants 2022, 11(12), 1585; https://doi.org/10.3390/plants11121585
Submission received: 6 May 2022 / Revised: 10 June 2022 / Accepted: 11 June 2022 / Published: 15 June 2022

Abstract

:
WRKY transcription factors play critical roles in plant growth and development or stress responses. Using up-to-date genomic data, a total of 64 and 257 WRKY genes have been identified in the diploid woodland strawberry, Fragaria vesca, and the more complex allo-octoploid commercial strawberry, Fragaria × ananassa cv. Camarosa, respectively. The completeness of the new genomes and annotations has enabled us to perform a more detailed evolutionary and functional study of the strawberry WRKY family members, particularly in the case of the cultivated hybrid, in which homoeologous and paralogous FaWRKY genes have been characterized. Analysis of the available expression profiles has revealed that many strawberry WRKY genes show preferential or tissue-specific expression. Furthermore, significant differential expression of several FaWRKY genes has been clearly detected in fruit receptacles and achenes during the ripening process and pathogen challenged, supporting a precise functional role of these strawberry genes in such processes. Further, an extensive analysis of predicted development, stress and hormone-responsive cis-acting elements in the strawberry WRKY family is shown. Our results provide a deeper and more comprehensive knowledge of the WRKY gene family in strawberry.

1. Introduction

The WRKY superfamily of transcription factors (WRKY TFs) [1] has an early origin in primitive eukaryotes, with later expansion and evolution in the green lineage driven by extensive tandem and segmental duplication events [2,3,4], thus becoming one of the largest TF families in higher plants [5]. The WRKY TFs are involved in the regulation of various physiological and developmental processes, such as senescence [6,7,8], stem elongation and seed development [9] and flowering [10]. They are also key players in responses to abiotic stresses [11] as well as to biotic stresses, wherein they seem to play major roles in plant immunity [5,12,13].
WRKY TFs regulate their target’s expression by binding to a specific cis-element known as a W-box, with the minimal consensus sequence TTGACY (where Y = C/T), although adjacent sequences are also involved in the binding site preferences [14]. DNA binding by WRKY proteins is mediated by a highly conserved DNA binding domain named the WRKY domain (WD). It is about 60 amino acids long and contains an N-terminal core motif, formed by the almost invariant heptapeptide WRKYGQK, and a distinctive C-terminal zinc-finger (Znf), both required for the DNA binding activity [15]. The amino acids constituting the WRKY core motif are essential for the DNA-binding activity and recognition specificity. Thus amino acid substitutions in this motif can affect the binding to the W-box [14,16,17]. Variations in this conserved sequence have been found in WRKY proteins from different plant species [3], with functional and binding activity characterization in some cases. WRKYs harboring divergent core motifs can bind to novel sequences that deviate from the consensus W-box. For example, NtWRKY12 contains a WRKYGKK core motif, interacting with the WK-box (TTTTCCAC) but not to the consensus W-box [18]. On the other hand, modified motifs can also be unable to bind to the W-box sequence, as GmWRKY167, containing the WRKYEDK core motif [19]. However, it is unclear if the binding specificity depends exclusively on the WRKY core motif. OsWRKY7, which contains the WRKYGKK sequence, binds to the W-box but not with the WK-box [20]. In addition, AtWRKY70, which harbors the conserved WRKYGQK core motif, can bind to the consensus W-box, as well as the novel WT-box (YGACTTTT) [21].
The classical WRKY classification in the three groups was based on both the number of WDs and the pattern of the Znf domain [1]. Group I WRKY proteins are the only ones that have two WDs (I-NT, I-CT), with a C2H2 Znf pattern (CX4–5CX22–23HXH) shared with group II WRKYs. Group II was further divided into subgroups IIa, IIb, IIc, IId and IIe based on the differences in their amino acid sequences from their WD. Group III WRKY proteins contain a WD with a C2HC Znf pattern (CX7CX23HXC). Later studies have proposed a new phylogenetic classification into four major groups (Group I + IIc, Group IIa + IIb, Group IId + IIe and Group III), as well as new hypotheses for the evolution of WRKY genes [22]. The WRKY proteins can also include additional domains. Notoriously, the R protein-WRKY family contains several typical domains of R-proteins, such as Toll-interleukin 1 receptor (TIR), leucine-rich repeat (LRR) and NB-ARC, which are associated with WDs from Group I, Group II and Group III members. Such chimeric genes have been found in multiple plant genomes; however, they are not widespread in plants. Instead, R protein-WRKY genes appear to have evolved on multiple independent occasions as a result of particular genomic rearrangements within specific plant lineages [22].
The genus Fragaria (Rosaceae) comprises about 24 species worldwide, with different geographical distributions and ploidy levels, from diploid to decaploid [23]. The modern cultivated octoploid strawberry (Fragaria × ananassa Duch.) is presumably the most economically important soft berry, with a world production exceeding 8.86 Mt and 384,668 ha harvested (FAOSTAT, 2020). In recent years, many efforts have been made to unravel the genetic background of this species to be used as a molecular breeding tool to identify traits and associated genes of interest for the genetic improvement of this valuable crop. The genome of the diploid (2n = 14) strawberry Fragaria vesca (Fv) was sequenced and published for the first time in 2011 [24] and proposed as a gateway to functional studies of genes within the Rosaceae, particularly for the cultivated strawberry. The first assembly versions were improved and reannotated [25,26,27]. Recently, it has been sequenced de novo using third-generation PacBio Single Molecule, Real-Time (SMRT) sequencing technology and an overall improvement over previous versions has been achieved [28]. Moreover, gene models and genome annotation have been recently updated [29]. Fragaria × ananassa (Fa) is an allo-octoploid hybrid (2n = 8x = 56), which originated around 300 years ago from interspecific crosses of the also octoploids F. virginiana and F. chiloensis. Recently, the genome of Fragaria × ananassa cv. Camarosa has been completely sequenced and annotated, revealing its diploid progenitor species: F. vesca (subsp. bracheata), F. iinumae, F. nipponica and F. viridis [30], although the contribution of the last two species has been debated [31,32,33]. Remarkably, this study has found that the subgenome contributed by Fv is dominant and has replaced large portions of the submissive ones through homoeologous exchanges.
The emergence of all these new data represents a valuable opportunity to conduct new and more comprehensive evolutionary analyses on strawberry, as well as greatly facilitate functional studies to further unravel the roles of members of the WRKY family in regulating the physiology of the strawberry, particularly in key aspects such as fruit ripening and biotic stress. This is especially relevant for the cultivated strawberry, in which many previous studies have been carried out with the limited availability of genetic data on the structure and nucleotide sequence of the octoploid genome, but taking advantage of the synteny and high sequence identity with Fv to use its reference genome as an “anchor” between both species [34,35,36,37]. For F. vesca, we describe extensive new findings not previously covered by Wei et al. (2016) [38] and Zhou et al. (2016) [39]. Moreover, a complete description of the F. ananassa WRKY members, defining the homoeologs and paralogs genes, is presented for the first time. We have further concentrated the expression analysis on fruit ripening, and biotic responses and a comprehensive study regarding the number and density of predicted development, stress and hormone responsive cis-acting elements in the WRKY TFs is shown together with their exon-intron distribution. In consequence, this study updates and expands our knowledge of the members of diploid and octoploid strawberry WRKY TF family, their evolutionary history and their potential roles in specific strawberry tissues and important biological processes such as fruit ripening and defense responses against pathogens.

2. Results and Discussion

2.1. WRKY Members of F. vesca and F. ananassa

The FvWRKY family has been described before using an earlier annotation [38,39]. Here, we have used the latest Fv genome and annotation versions available, which combine a higher quality reference genome [28] and high-fidelity gene models with RNA-seq support using expression data from different F. vesca accessions, tissues and fruit developmental stages [29]. A comprehensive list of the FvWRKY members across different Fv genome annotations is also provided (Table S1). A total of 64 FvWRKY coding genes and their splicing forms (SFs) were confirmed, then named according to their chromosomal locations (Table S2). Some relevant properties of the FvWRKY proteins and their predicted subcellular location are also listed. GRAVY (grand average of hydropathy) values are below 0, indicating that FvWRKYs are hydrophilic and more likely globular-shaped, while LOCALIZER predicted the protein location inside the nucleus for most FvWRKYs.
FvWRKY proteins were classified into groups I, II and III according to their WDs [1]. Core WRKY motif modifications were detected in FvWRKY3 and -8 (WRKYGKK), FvWRKY21 (WKKYGQK), FvWRKY35 (WTKYDQR) and FvWRKY55 (WREYDQR). A more drastic modification is found in FvWRKY37, in which the core motif is found truncated and reduced to WRK. Moreover, differences in some FvWRKYs splicing forms (SF) were observed, affecting the nature of the encoded proteins. Thus, one SF from FvWRKY20, -26 and -43 (Group I) encode for WRKY proteins that have lost their I-NT WDs, while some SF from several FvWRKYs encode proteins containing incomplete WDs. Additionally, one SF from FvWRKY54 and -62 have lost their WDs completely. The regulation of all these alternative transcripts and whether they are efficiently translated into functional proteins remains to be studied.
Additional motifs were also found in some of the FvWRKY proteins (Figure 1). These include a plant zinc cluster domain in FvWRKY2, -7, -9, -10 and -22. Further, TIR, NB-ARC and leucine-rich repeat motifs (LRR) harbored by the FvRWRKY subfamily [22], consisting of FvWRKY35, -55, -61 and -62. Moreover, FvWRKY35 and -55 exhibit an additional WRKY-like domain, lacking the core motif but retaining the Znf portion. Additionally, the loss of the WD in the predicted FvWRKY62.t6 splicing form would turn the translated product into a TIR-NBS-LRR protein.
The recent Fragaria × ananassa Camarosa (Fa) Genome Assembly v1.0 and annotation v1.0.a1 were used to identify the FaWRKY candidates distributed among the four diploid subgenomes, which compose this allo-octoploid [30]. A total of 255 FaWRKY were found by HMMER, and the presence of WD was confirmed in CDD. Homology and shared synteny with the FvWRKY genes were investigated using SynMap2. An additional gene (snap_masked-Fvb3-3-processed-gene-269.16) showed shared synteny with FvWRKY20. This Fa gene was not detected by the HMMER search due to an incomplete protein sequence in the source dataset. FGENESH (www.softberry.com, accessed on 14 April 2022) [40] was used to predict a revised protein sequence, using the source mRNA as input and specific gene-finding parameters for Fv. The corrected protein sequence showed two WDs, and this gene was named FaWRKY20A (Figure S1). Likewise, shared synteny was found between another gene (maker-Fvb3-1-snap-gene-3.50) and FvWRKY21. A genomic DNA track containing this gene (Fvb3-1:336748..339128) was loaded in FGENESH to predict newly revised mRNA and protein sequences and the new FaWRKY member was named FaWRKY21C (Figure S2). Therefore, a total of 257 FaWRKY genes were found in our analysis. Two more FaWRKYs, maker-Fvb3-2-snap-gene-310.32 (FaWRKY21B) and maker-Fvb6-2-snap-gene-308.69 (FaWRKY51A.2), were detected to have incorrect protein predicted sequences in the source dataset. Genomic DNA fragments containing each gene (Fvb3-2:31015005..31018524 and Fvb6-2:30857826..30862505, respectively) were loaded into FGENESH to recover the revised sequences of both transcript and protein (Figures S3 and S4). All the new sequences generated were included in the successive analyses performed below.
The FaWRKY genes were named following three criteria: foremost, due to their homology and shared synteny with the FvWRKYs (FaWRKY1 to FaWRKY64), derived from the SynMap2 analysis; then by a letter indicating the subgenome donor (A, F. nipponica; B, F. iinumae; C, F. viridis; D, F.vesca); and finally, numbering the gene duplications, if any (.1, .2, etc.). The final list of the 257 FaWRKYs is provided in Table S3. Biochemical properties and the subcellular location of the FaWRKY proteins were also calculated and as for their FvWRKY orthologs, GRAVY values indicate that FaWRKYs are hydrophilic and more likely globular-shaped, and located in the nucleus, with few exceptions. Overall, the FaWRKY family shares a high gene structural conservation (Figure S5) and nucleotide sequence identity with their FvWRKY orthologs, with an average of 97.43%. Moreover, FaWRKY proteins were grouped according to their WDs, which match with their respective FvWRKY orthologs. Furthermore, this similarity includes the additional motifs found in many cases. Many WRKY core motif variations, some of them not present in the FvWRKY proteins, as well as truncated Znf motifs, domain loss and presence of additional duplicate motifs also found in the diploid, were detected. The number of WD modifications detected in the FaWRKY proteins is higher than the one found in the FvWRKYs, affecting FaWRKYs belonging to diverse groups, including the R protein-WRKY homologs. In FaWRKY51A.2, only one incomplete WD was found, lacking the Znf portion, and thus was not classified within any WRKY group. Interestingly, FaWRKY26B.2 is a chimeric protein harboring an additional Myb-like DNA-binding domain as well as two unusual, modified WDs (I-N and I-C).

2.2. Genome-Wide Distribution and Gene Duplications of the Strawberry WRKY Family

Duplicated genes are abundant in plant genomes. These duplicates are retained after whole-genome duplication (WGD) events and have contributed to the evolution of novel functions in plants [41]. Ancient or recent WGD and polyploidy are common among the angiosperms [42,43]. The homologous duplicated genes within the same genome are named paralogs, while the term homoeologs refers to the homologous genes resulting from allopolyploidy [43]. We have used DAGchainer [44], implemented in the SynMap2 web-based tool [45], in order to identify duplicated (paralogous) FvWRKY genes, as well as FaWRKY paralogous and homoeologous genes.
The FvWRKY genes are unevenly distributed among the seven chromosomes, with almost half of them (31 out of 64) located on chromosomes 6 and 7 (Figure 2). This is partly due to gene expansion by tandem and segmental duplications. Remarkably, 14 out of 15 group III WRKY coding genes are located on chromosomes 6 and 7. Sixteen segmental chromosome duplications and three groups of tandem repeats containing FvWRKY paralogous genes were found (Table 1). All the FvWRKY paralogs exhibit low ω values, indicating that they are under strong negative (purifying) selection pressure as observed for other species [46]. Duplicate gene pairs FvWRKY13-50 and FvWRKY17-50; FvWRKY24-53 and FvWRKY24-30; and FvWRKY29-58 and FvWRKY58-64 share a common gene, suggesting that they have evolved as a result of a two-step duplication event [47]. Interestingly, the two tandemly duplicated groups consisting of FvWRKY38-39-40-41-42 and FvWRKY59-60-61 seem to be related to each other. The FvWRKY61 gene, belonging to the FvRWRKY subclass, was identified as a segmental duplicate of FvWRKY39 within a tandem repeat along with FvWRKY59 and -60, pinpointing the origin of this chimeric gene as a likely result of a genetic rearrangement of a group III WRKY gene and an unknown R gene leading to the formation of a novel R protein-WRKY gene. Actually, another non-WRKY gene (FvH4_7g26100) was identified as part of the FvWRKY59-60-61 tandem, sharing partial sequence homology with FvWRKY61, as well as with a TIR-NB-LRR gene (FvH4_7g17700), suggesting a link with the origin of the R protein-like domain found in FvWRKY61 (Figure 3). Tandem or segmental gene duplications involving the other three members of the FvRWRKY group (FvWRKY35, -55 and -62) were not detected, probably because duplication-inherent mechanisms, such as inversions or post-duplication events, have broken the collinear relationships among FvWRKY and the “donor” R genes [42]. However, a candidate process already proposed for the formation and posterior expansion of the R-WRKY protein class, besides WGD events, is gene transposition [22]. TE elements provide the capability to generate new genes by duplicating and recombining gene fragments [48]. In a recent study about the diversification of plant immune receptors, the authors concluded that the integration of exogenous domains into nucleotide-binding leucine-rich repeat (NLR) proteins combine both gene duplication and interchromosomal translocation pointing to TE elements or ectopic recombination as the most likely mechanisms [49]. Such events may also have led to the emergence of the R-WRKY protein families, independently, in several species.
The octoploid Fa genome is composed of four subgenomes (A, B, C and D) highly collinear with F. vesca, and it contains sets of homoeologous WRKY genes derived from the diploid ancestors (Figure 4). Genomic reorganizations and fractionation resulting from polyploidization [26,30] have caused some FaWRKY genes loss, as well as some segmental transpositions and inversions, creating collinearity breaks and hampering the identification of similar paralogous pairs as those found in the diploid. Instead, several new segmental gene duplicates were detected, as well as several non-syntenic gene duplications and triplications (Table 2). Ks values calculated for the octoploid paralogs are quite low in most cases, evidencing that gene duplications are very recent. Several FaWRKY paralogs show ω values greater than 1, which means that these genes may be undergoing positive selection, and sub-functionalization or neo-functionalization processes might be taking place [50]. Taken together, we could hypothesize that they have appeared as a result of either homoeologous exchanges, intrinsic to polyploidization [30], or post duplication events affecting single genomic features as gene transposition [42]. Indeed, TEs can be activated by polyploidy and hybridization events [51] and play important roles in producing segmental duplications in plants and in generating changes in the genome organization and size of the hybrids, which often produce synteny breaks [52,53,54].

2.3. Phylogenomic Analysis of the Strawberry WRKY Family

We have used a phylogenomic approach to study the evolution of the strawberry WRKY family and explore its potential biological implications. Shared synteny among Fragaria vesca (Fv), Arabidopsis thaliana (At) and Vitis vinifera (Vv) WRKY families was investigated using the CoGe web. SynMap3D was used to find collinear WRKY genes shared among the three species (Table S4). In addition, SynMap2 was used to find the collinear WRKY genes shared between Fv-At and Fv-Vv and not found by the SynMap3D analysis (Table S5). Full-length WRKY proteins from the three species were aligned by MUSCLE, and an unrooted phylogenetic tree was constructed and annotated using iTol to integrate both phylogenetic and collinearity relationships (Figure 5). The resulting tree distribution obtained meets the phylogenetic classification of the WRKY proteins into Groups I + IIc, IIa + IIb, IId + IIe and III [22].
Most of the FvWRKY genes share synteny with AtWRKY and VvWRKY members of the corresponding groups. Exceptions include the R protein-WRKY members, absent in Vv, and because At and Fv members harboring similar R domains also carry different WDs as previously described by [22]. The analysis of the ω values obtained for the syntenic At, Vv and FvWRKY genes indicates that purifying selection is likely the main evolutionary driving force for this family in Rosaceae, acting against changes within the protein sequences, which modify or disrupt their functionality [55]. A higher number of collinear relationships are found with VvWRKY than AtWRKY genes, showing lower Ks values overall. This is probably because, unlike At, Vv has a relatively slow evolutionary rate [56,57]. Moreover, like Vv, the latest large gene expansion in the Fragaria lineage, likely took place in the gamma (γ) whole-genome triplication, shared by the core eudicots, since there is no evidence of more recent WGD events [28,56,58].
Synteny and collinearity conservation between FvWRKY and FaWRKY orthologs was also investigated with SynMap2, which found that most of the strawberry WRKYs were located within conserved blocks. Despite the fact that genomic reorganizations are widespread in the Fa genome relative to the diploid, 189 of 257 FaWRKY genes are syntenic and collinear with their FvWRKY orthologs, regardless of their subgenome origin (Table S6). A phylogenetic tree of the diploid and octoploid strawberry WRKY proteins was constructed and annotated with the shared synteny information (Figure S5). The tree distribution obtained for the strawberry full WRKY proteins is also according to the phylogenetic classification of the WRKYs into Groups I + IIc, IIa + IIb, IId + IIe and III.
The ω values show that strawberry WRKY orthologs are predominantly under purifying selection, indicating high conservation of the WRKY gene family within the two Fragaria species. However, some cases of positive and possible neutral selection (due to very low Ks values and thus considered as no evolution) are also uncovered, indicating that some genes may be undergoing neo or subfunctionalization processes. Interestingly, most of the synteny and collinearity breaks between the diploid and octoploid species are observed in those WRKY genes located in chromosomes 6 and 7 (52 out of 68), which appear to have undergone substantial genomic rearrangements in Fa, such as fractionation (loss of genomic features, such as genes) and ectopic recombination with non-homologous chromosomes (Figure S6). Thus, extensive homoeologous gene losses were detected for FaWRKY39, -40, -46 and -61 genes, which retain only one of the homoeologs. Several members of FaWRKY group III are also affected by homoeologous gene losses, perhaps because most of this group is located within these two chromosomes.

2.4. Orthology Relationships and Annotation of the Strawberry WRKYs

Orthologs are pairs of genes found in different species and originated from a speciation event [43]. There is a widespread belief that orthologs are more prone to conserve similar functions than paralogs [59]. This concept, known as the “ortholog conjecture”, is the basis for predicting conserved gene functions. Moreover, syntenic conservation between orthologous genes is less likely to undergo positive selection and, therefore, more likely to retain a conserved function [43,60]. However, inferring orthologs can be challenging, depending on the evolutionary distances and the occurrence of WGD events that produces gene duplication/loss.
Orthologs of Fv and Fa WRKY proteins were investigated using the new eggNOG 5.0 Database and hierarchically classified into orthologous groups (OGs) within several taxonomic scopes (Table S7). Consistent with the high sequence conservation detected by the previous analyses, the putative FvWRKY and FaWRKY orthologs were classified into the same OGs. Moreover, GO functional annotations are essentially shared by both species (Figure 6). The only exceptions were FaWRKY37D and the chimeric protein FaWRKY26B.2, which were assigned to different OGs than their respective orthologs and homoeologs. Moreover, FaWRKY37D was classified into OGs containing non WRKY proteins of unknown function, such as AT2G38570.1. Candidate protein orthologs in several plant species can be identified in the eggNOG 5.0 Database using this hierarchical OG classification. However, we concentrated our attention on the homology with the AtWRKY family, which has been the best functionally characterized so far. The taxonomic scope used was the closest level to fabids (clade including the Rosaceae, to which belongs the genus Fragaria) [61] containing orthologous AtWRKY proteins. Thus, up to 17 one-to-one AtWRKY orthologs were identified, but one-to-many or many-to-many AtWRKY orthologs with the strawberry WRKY family were also found.
For example, several FvWRKY members (FvWRKY37, -38, -39, -40, -41, -42 and -60) were found homologous to the Arabidopsis WRKY54/70 proteins. The phylogenetic reconstruction using all the potential eggNOG homolog genes in several species (Table S8) by the maximum likelihood method shows that the strawberry AtWRKY54/70-like orthologous genes were originated by different rounds of gene duplication (Figure 7). Thus, the strawberry orthologs of the two Vv gamma-paralogs VvWRKY27 and VvWRKY42 experienced successive tandem duplications, originating the FvWRKY37-38-40, FvWRKY39-41-42 and FvWRKY60-61 tandem gene groups. Independently, the AtWRKY54-70 gene duplication likely took place in the more recent At-β or At-α [62], as well as other WGD events in Juglans, Malus and Glycine expanded their WRKY orthologs [63,64,65,66]. Therefore, it is likely that several strawberry WRKY genes, homologous to AtWRKY54-70, share some ancestral functions of the last ones, which still today mainly show redundant roles [6,67,68]. The two Amborella trichopoda orthologs included help to show that divergence between the AtWRKY55-like and AtWRKY54/70-like genes probably predates the gamma event since it is not shared by Amborella [69]. This also would explain the gene FvWRKY59, identified as tandem duplication along with FvWRKY60-61, as an ancient pre-gamma paralog within this tandemly duplicated group.

2.5. Expression Analysis of FvWRKY Genes in Different Tissues, Development Stages and Growth Conditions

Most of the woodland strawberry WRKYs have not been yet functionally characterized nor fully described at the expression level, with few exceptions [70,71,72]. Consequently, to gain insight into their possible biological roles, the expression pattern of the complete FvWRKY family was analyzed in several tissues (Figure S7), considering the transcript expression values collected from previous research [29] (Table S9).
The analysis revealed that four genes, FvWRKY10, -2, -7 and -33, exhibit high transcript accumulation in all samples except pollen, suggesting that they should play important roles in most phases of strawberry physiology. Both transcript abundance and expression patterns of the remaining FvWRKY genes vary largely in different vegetative and reproductive tissues, as well as during receptacle ripening and pathogen challenge. For example, several FvWRKY genes changed their transcripts abundance or expression profiles during ripening in receptacle from two botanical varieties of F. vesca, Ruegen (F7–4) and Yellow Wonder 5AF7, which produce fruits with red or yellow flesh and skin, respectively [73,74]. However, the changes detected in the expression pattern from 15 days to 22 days post-anthesis for FvWRKY2, -23, -29, -36, -47, -50 and -51 were similar in both accessions, pointing out that these genes may play key roles in the strawberry fruit ripening process.
A greater number of FvWRKY genes (49 out of 64) changed their expression pattern in roots infected by the hemibiotrophic oomycete Phytophthora cactorum, with transcript abundance differences that reflect both positive and negative regulation in response to this pathogen [75]. Notorious up-regulation was detected for collinear and orthologous genes of well-known AtWRKY TFs involved in plant defense responses (see Tables S4, S5 and S7), as FvWRKY3 and -8, homologous to AtWRKY50/51, mediators of the SA and JA signaling [76,77]; FvWRKY24, -30 and -53, orthologs of AtWRKY45 and -75; AtWRKY75, and the F. ananassa ortholog FaWRKY24 (previously reported as FaWRKY1) have been described as positive and negative regulators of plant defense in Arabidopsis and F. ananassa, respectively [78,79,80]; FvWRKY29, ortholog of AtWRKY53, which positively modulate SAR in Arabidopsis [81,82]; FvWRKY38, -42 and -60 share homology with AtWRKY54 and -70, its closest homologs, which seem to regulate the balance between SA and JA-dependent defense pathways, acting as a negative regulator of SA biosynthesis [68]; FvWRKY19, -43 and -47 that share orthologous group with AtWRKY25, AtWRKY33 and AtWRKY18/40, respectively, major factors predicted to function as important hubs within the WRKY network of plant defense in Arabidopsis [83,84]. Promoter analysis of these genes show an abundance of predicted cis-acting elements in response to methyl-jasmonate in FvWRKY3, -8, -24, -29, -38, -42, -43, -47, -53 and -60. Moreover, FvWRKY43 and -47 (AtWRKY33 and AtWRKY18/40 orthologs, respectively) contain salicylic-responsive elements. All these cis-acting elements (MeJa and SA response elements) seem to be conserved in their F. ananassa homoeologs (Tables S10 and S11). All in all, these data suggest an outstanding involvement of the WRKY TF family in the woodland strawberry defense response.
A closer look at the transcription pattern (Figure 8) reveals substantial differences in the transcription level of most paralogous pairs, as well as in their expression profiles, which suggest that they are not fully redundant and may have undergone functional divergence, neo or sub-functionalization [85]. Thus, duplicate genes FvWRKY2-9, FvWRKY7-36, FvWRKY13-50, FvWRKY17-50, FvWRKY19-43, FvWRKY24-30, FvWRKY24-53, FvWRKY29-58 and FvWRKY58-64 in receptacle ripening; FvWRKY2-9, FvWRKY11-12 and FvWRKY15-46 in root infection; and members of the tandem duplications FvWRKY38-39-40-41-42 and FvWRKY59-60-61 in both receptacle ripening and root infection, show very different expression patterns.

2.6. Expression Analysis of the FaWRKY Gene Family in Strawberry Tissues, the Fruit Ripening Process and Defense Responses

The recent publication of a high-quality annotated octoploid strawberry genome, and the availability of public RNA-seq data, provide a valuable opportunity to investigate the expression patterns of the FaWRKY gene family in different tissues and growth conditions at the level of homoeologs and paralogs to understand their functions in key strawberry processes, such as fruit ripening and defense responses against pathogens. Thus, we have investigated two publicly available strawberry datasets, featuring both tissue expression profiles and transcriptomic changes along four stages of achene and receptacle ripening [35] and anthracnose defense response of strawberry leaves infected with the hemibiotrophic fungus Colletotrichum fructicola [86]. Both raw and differential gene expression values are provided in Tables S12 and S13, respectively.
The expression pattern for most of the FaWRKY genes varied among the different strawberry tissues investigated, showing many cases of preferential or even tissue-specific expression (Figure S8 and Tables S13–S15). For example, FaWRKY17B/17C.1/17C.2/17D were detected and differentially expressed exclusively in roots, whereas the homoeolog FaWRKY17A was expressed in roots, leaves and green receptacles. On the other hand, FaWRKY24A/24B/24D, among others, showed a higher transcript accumulation in roots than in any other tissues. The strawberry FaWRKY24 homoeologs were identified as orthologs of AtWRKY75, which acts as a negative regulator of root development as well as a positive regulator of Pi stress responses [87].
Fruit ripening is a critical developmental process in strawberry, and fruit size, color and aroma, are important agronomical traits. However, little is known about the involvement of members of the strawberry WRKY gene family in the genetic control of ripening and the major changes undergone in strawberry fruit. Nevertheless, transcriptional reprogramming of FaWRKY genes is evident in both receptacle and achene from immature (green) to mature (red) fruit (Figure S10; Tables S16 and S17). We have found that 74 FaWRKYs were significantly up or down-regulated in receptacles during specific stages of ripening, while 102 were in their respective achenes (Figure 9A). Of these, 24 and 52 FaWRKY genes were differentially expressed only in receptacles or achenes, respectively (Figure 9B). Moreover, this analysis revealed that many paralogous and homoeologous genes were expressed differently in both fruit tissues (Table S18), suggesting that they may have undergone functional divergences in the strawberry fruit ripening.
It has been shown that the hormonal balance between abscisic acid (ABA) and auxins is primarily responsible for the changes leading to ripening in this non-climacteric fruit [34]. Supporting this, the expression in the receptacle of three strawberry WRKYs (Fv gene28720, gene19478 and gene01340) have previously been described to be activated by ABA and repressed by auxins in the receptacle, while another three WRKY genes (Fv gene07210, gene03411 and gene09147) were repressed by auxins and not affected by ABA [34]. Our transcriptomic analysis supports potential roles for five of these genes at some points along the strawberry ripening process. Thus, genes FaWRKY48B/48C (Fv gene28720) and FaWRKY53A/53B (Fv gene01340) were differentially up-regulated in both receptacle and achene, whereas the homoeologous FaWRKY48A was only in achene. On the other hand, FaWRKY24D (Fv gene07210) was differentially up-regulated in both receptacle and achene, but its homoeologous FaWRKY24A was down-regulated in receptacle and up-regulated in achene. Further, genes FaWRKY17A (Fv gene03411) and FaWRKY9D (Fv gene09147) were up-regulated in red receptacle, whereas FaWRKY9B/9C/9D homoeologs were down-regulated in achene. However, none of the FaWRKY13 homoeologs (Fv gene19478) were differentially expressed. Interestingly, ABRE cis-acting elements are abundant in the promoter regions of genes FaWRKY48A/48B/48C, FaWRKY53A/53B, FaWRKY24A/24D and FaWRKY17A, while auxin-responsive elements are present to a lesser extent (Table S10), which seems to be conserved from their FvWRKY orthologs (Table S11).
To obtain further insight into the potential roles of the differentially expressed FaWRKY genes during fruit ripening, we constructed a functional protein association network in STRING using the experimentally known interactions for their Arabidopsis orthologs (Figure S11). The resulting network connects FaWRKY24, FaWRKY57 and FaWRKY47 with clusters of proteins related to the proteasome and developmental processes such as photomorphogenesis and responses to auxin and jasmonic acid. On the other hand, FaWRKY43 and FaWRKY47 (both significantly down-regulated during fruit ripening) interact with proteins involved in the biotic and abiotic stress responses. Several strawberry VQ proteins, which change their expression profiles during fruit ripening [88], appear as interacting nodes with these FaWRKYs and could modulate their transcriptional activity. Finally, FaWRKY11 and FaWRKY47 are associated with members of the bHLH transcription factor family, involved in the regulation of multiple processes such as anthocyanin biosynthesis and cell elongation during strawberry fruit ripening [89,90]. These findings evidence that some WRKY TFs could play important roles in the strawberry fruit ripening and further research is needed to unravel their specific function.
It is known that WRKY TFs have important roles in defense responses against pathogenic fungi [20]. The expression patterns of several FaWRKY genes fluctuated in strawberry leaves challenged with C. fructicola (Figure S12 and Table S17). A total of 53 FaWRKYs were differentially expressed, 41 were up-regulated and 12 were down-regulated (Figure 10), most of which were homologs of well-known defense-related WRKYs. Among the FaWRKY up-regulated genes with higher expression and having known functions in plant defense, we found WRKY8/28-probable orthologs FaWRKY14A, -14B, -14C.2 and -14D; WRKY54/70 orthologs FaWRKY38B, -39A and -60B; WRKY75 orthologs FaWRKY24A and -24D; WRKY50/51 orthologs FaWRKY3C, -8A, -8B, -8C and -8D; WRKY72 ortholog -52A; and WRKY18 ortholog FaWRKY11A. On the other hand, among the down-regulated FaWRKY genes were the WRKY53 orthologs FaWRKY29A, -29B, -29D.1 and -29D.2. Several of these genes are involved in the negative regulation of JA-mediated, but positive regulation of SA-mediated defense responses [77,91,92,93,94,95], while WRKY18 positively regulates the expression of some key JA-signaling genes in Arabidopsis [96,97]. It is worth mentioning that AtWRKY33 high-homology-sharing genes FaWRKY43B.2 and FaWRKY43D are down-regulated along the time course of infection. Arabidopsis WRKY33 is a key positive regulator of the JA-mediated defense response against necrotrophic fungi, also showing an antagonistic effect with the SA-mediated pathway [98], and AtWRKY33 orthologs seem to share similar functions in other plant species, including the woodland strawberry [71]. Although this down-regulation was already noticed by Zhang et al. [86], it contrasts markedly with the results reported in other strawberry plant tissues of either resistant or susceptible cultivars challenged with Colletotrichum gloeosporioides [99] or Colletotrichum acutatum [37,88]. However, it could be explained if subtle differences in the pathogen-specific strategies do exist among Colletotrichum species to manipulate the host defense in different strawberry tissues.
Homoeologs of FaWRKY24 and FaWRKY53 are both AtWRKY75-like factors and, as mentioned before, FaWRKY24 (formerly, FaWRKY1) and AtWRKY75 have been previously reported to play important roles as positive regulators of plant defense in A. thaliana against Pseudomonas syringae [37,79]. Both the Fv paralogs FvWRKY24 and FvWRKY53 and their Fa orthologs FaWRKY24A/24D and FaWRKY53B/53C/53D were also detected to be up-regulated in response to different pathogens. In addition, recent research has shown that WRKY75 is a positive regulator of the JA-mediated defense against the necrotrophic fungi Botrytis cinerea and Alternaria brassicicola [78]. Interestingly, FaWRKY24 (FaWRKY1) homoeologs are also up-regulated during strawberry fruit ripening as well, as they have been reported as negative regulators of the resistance to C. acutatum in strawberry fruit [80]. This illustrates the dual roles that the strawberry AtWRKY75-like factors may play against different pathogens or in different plant tissues.
Among the R protein-WRKY subclass, FaWRKY55D was the only one differentially expressed at early stages of the infection. This is particularly interesting as it could be hypothesized that FaWRKY55D could participate in the activation of the early immune response in strawberry similarly to the RRS1-R/RSP4 pair system (via an integrated WRKY domain in RRS1-R) recently described in Arabidopsis [100,101]. Therefore, it should be included in the repertory of putative candidate strawberry WRKY defense-responsive genes for further study. Several homoeologs of the other FaRWRKY genes (FaWRKY35, FaWRKY61 and FaWRKY62) were expressed in both mock-treated and infected tissues. However, their transcript abundances did not change significantly in the course of infection. Most of them showed relatively high transcript abundances in leaves, roots and fruit tissues also, suggesting that changes in their gene expression, if any, are either not needed to play positive roles against pathogens or these changes are pathogen-specific deployed or involved in other plant processes. Nevertheless, their roles in the strawberry physiology remain unknown, as the vast majority of other R protein-WRKY in several species [100,101].
From the currently available data, a selection of putative candidate FaWRKY genes for further research is shown in Table S19. This set of outstanding genes was selected based on the results shown in the heatmap of FaWRKY genes differentially expressed during strawberry fruit ripening (Figure 9) and in response to C. frutícola infection (Figure 10), and also taking into consideration the abundance of predicted cis-acting elements within the promoter region of each member of the FaWRKY family (Table S10).

3. Conclusions

The present study represents a novel and comprehensive update of the strawberry WRKY family using the ultimate and most accurate genome assemblies and gene annotations available. The use of the most recent data has allowed us to perform more precise analyses and characterization of the WRKY gene family composition, its evolutionary history and gene expression patterns. In the case of the diploid F. vesca, we have made extensive new findings not covered by previous works. As for the cultivated octoploid hybrid Fragaria × ananassa, we have described for the first time the WRKY homoeologs and gene duplications. We have reanalyzed public data to study their expression in several tissues, potential roles in defense and fruit ripening and discussed the results in the light of the precise identification of the strawberry WRKY family, particularly in the case of Fragaria × ananassa.
Thus, a total of 64 WRKY genes are present in the genome of the diploid Fragaria vesca and 257 corresponding WRKY orthologs in the genome of the cultivated allo-octoploid Fragaria × ananassa (cv. Camarosa). Our results show that the strawberry WRKY family preserves a high degree of conservation between both wild-type and cultivated species. Synteny analysis showed that FaWRKY genes are largely syntenic and collinear with their FvWRKY orthologs, despite some gene loses and synteny breaks detected in the octoploid, probably as a result of genomic rearrangements derived from hybridization and polyploidy. The strawberry WRKY family has been expanded by ancient WGD events, originating from several segmental and tandem duplications resulting in several paralogous genes. Moreover, most FaWRKY paralogs are not mirrored in F. vesca and could be inherited from the parental octoploid species F. virginiana and F. chiloensis, in which polyploidization could originate synteny breaks among ancestral paralogs as well as new gene duplications.
Gene expression is related to functionality, and its analysis can shed light on the biological roles of the genes studied. Thus, the expression patterns of most WRKY genes differed in the strawberry tissues here considered, evidencing differences in functional relevance across different tissues and growth conditions. Importantly, many strawberry WRKYs remarkably changed their transcript level or were differentially expressed in defense and fruit ripening stages, indicating their importance in these important biological processes. It is worth noting genes FaWRKY38B, -39A and -60B (AtWRKY54/70 orthologs), FaWRKY24A/D and -43B.2/D (AtWRKY75 and -33 orthologs, respectively), FaWRKY30A/B/D and FaWRKY55D (encoding an R-protein) are related to pathogen defense, and FaWRKY9D, -17A, -48B/C, -53A/B are related with fruit ripening. The genes FaWRKY11 and FaWRKY47, which seem to be related to bHLH transcription factors, and FaWRKY57 are related to members of the VQ protein family, are also relevant. In addition, differences in the expression pattern detected among several FaWRKY paralogs and homoeologs point out either a finely regulated gene expression strategy to achieve putative additive genetic effects or evolutionary functional divergences. These data can help future studies deepen our understanding of the strawberry WRKY TF’s regulatory roles. Future RT-qPCR analyses of expression profiles of candidate WRKY genes under different development and stress conditions (including stress conditions not described in this study, such as salt, heat and osmotic stress) should confirm and expand the data retrieved from publicly available RNA-seq data, as well as help new approaches in breeding programs.

4. Materials and Methods

4.1. Identification of WRKY Family Members

The sequences of Fragaria vesca (Fv; Genome Assembly v4.0.a1 & Annotation v4.0.a2) and Fragaria × ananassa cv. Camarosa (Fa; Genome Assembly v1.0 & Annotation v1.0.a1) were retrieved from The Genome Database for Rosaceae (GDR) website (https://www.rosaceae.org/, accessed on 14 April 2022) [102]. The Hidden Markov Model of the WD (PF03106) was downloaded from the Pfam database [103] and used as a query in HMMER3 search, performed in the freeware tool UGENE v1.21 with default settings [104]. The candidate sequences were further confirmed to include the WD, as well as additional protein domains, using the Conserved Domain Database (CDD) [105]. WRKY sequences from Arabidopsis thaliana (At) [1], Vitis vinifera (Vv) [106] and several species were retrieved from the Plant Transcription Factor Database (PlantTFDB) [107]. Chromosome maps of the strawberry WRKY genes were drawn with MapChart v2.32 [108]. Protein locations were predicted using LOCALIZER (https://localizer.csiro.au/, accessed on 14 April 2022) [109]. General sequence handling and intron-exon structure analysis were performed using TBTools v1.055 WRKY protein properties calculation was carried out in the Freiburg Galaxy server (https://usegalaxy.eu/, accessed on 14 April 2022) [110].

4.2. Phylogenomics

Full WRKY protein sequences were aligned by MUSCLE to generate unrooted phylogenetic trees by the Neighbor-Joining (N-J) method in MEGA 7.0 [111]. The resulting protein trees were annotated with iTOL (https://itol.embl.de/, accessed on 14 April 2022) [112]. The compared synteny among species and strawberry WRKY gene duplications were studied in the CoGE (https://genomevolution.org/coge/, accessed on 14 April 2022) web-platform using LAST to find gene homologies and SynMap2 or SynMap3D to find collinear blocks shared by two or three species, respectively (Table S20) [42,45]. The non-synonymous (Kn) and synonymous (Ks) substitution rates between pairs of syntenic genes were calculated by codeml, implemented in SynMap, or in PAL2NAL (http://www.bork.embl.de/pal2nal/, accessed on 14 April 2022) [113] for duplicated genes lacking syntenic conservation. The non-synonymous (Kn) and synonymous (Ks) substitutions were used to calculate the Kn/Ks ratios (ω) between paralogous strawberry WRKYs and thus estimate the selection pressure. Values of ω > 1 or ω < 1 indicate positive or purifying (negative) selection, respectively, while ω = 1 means neutral (absence of) evolution [114]. Very low substitution values (Ks < 0.01) were considered as the virtual absence of nucleotide mutation and thus not accounted to calculate ω because it may result in inaccurate estimates [115,116]. eggNOG-Mapper v2 (http://eggnog-mapper.embl.de/, accessed on 14 April 2022) and the eggNOG 5.0 Database (http://eggnog5.embl.de/#/app/home, accessed on 14 April 2022) [117,118] were used to classify the WRKY orthologs in strawberry and other species, including Amborella trichopoda, Juglans regia, Malus domestica and Glycine max. GO terms with experimental evidence were acquired from the eggNOG results and plotted using WEGO 2.0 (https://wego.genomics.cn/, accessed on 14 April 2022) [119]. The homology with Arabidopsis proteins was applied in the construction of a protein functional association network in STRING (https://string-db.org/, accessed on 14 April 2022) [120], filtering interactions by high confidence (0.700) and using as active interaction sources “Experiments” and “Co-expression”.

4.3. Identification of Cis-Acting Element Analysis of WRKY Genes

The 2000-bp sequence upstream from the start codon (ATG) of each WRKY gene was obtained from the Fragaria × ananassa cv Camarosa (Annotation v1.0.a1) and Fragaria vesca (Annotation v4.0.a2) genome database for Rosaceae (GDR) website (https://www.rosaceae.org/, accessed on 22 May 2022). These sequences were used to identify cis-acting regulatory elements with the online program PlantCARE (http://bioinformatics.psb.ugent.be/webtools/plantcare/html/, accessed on 22 May 2022, Ghent, Belgium). The analysis of the frequency of the different cis-acting elements in the promoter region of each WRKY gene was obtained by Panda Library (phyton) with Jupyter notebook.

4.4. Expression Analyses of Diploid and Octoploid Strawberry WRKYs

Expression data of FvWRKY genes were taken from a previously published RNA-seq expression analysis in several F. vesca tissues and developmental stages [29]. Transcripts per Million (TPM) values were log10-transformed and depicted using the heatmap function of TBTools. The expression patterns of the FaWRKY genes were obtained through a complete reanalysis of several Fragaria × ananassa RNAseq datasets, including strawberry plant and fruit tissues (3 biological replicates in 54 RNAseq libraries: six independent green receptacle/achene libraries, six white receptacle/achene libraries, six turning receptacle/achene libraries, six red receptacle/achene libraries, three root libraries and three leaf libraries) [35]; and strawberry leaves infected by C. fructicola (three biological replicates) [86]. Detailed experimental procedures can be found within the original studies. Therefore, raw reads from the sequencing platform were processed to retain only high-quality sequences to be subsequently used for the mapping (Cutadapt v1.9, BBDuk v35.43). Sequencing adapters were first clipped from each library, and low-quality bases were trimmed. A Phred quality score of 24 was selected as the threshold, and reads with a length less than 30 nt were filtered out. Reads quality assessment was carried out using FastQC software (v0.11.8) to evaluate the effect of every step of this process. All subsequent analyses were conducted using these high-quality datasets. The remaining ribosomal RNA was detected by SortMeRNA software (v2.1) [121]. Thus, adaptors clipped reads were mapped to SortMeRNA prepackaged databases id98 (from Silva v119 and Rfam) with default parameters. A two-pass mode mapping was carried out by STAR (v2.7.3.a) [122] with parameter “--quantMode GeneCounts” in the second pass to extract raw counts per annotated gene ID according to each particular library’s strands. The obtained expression matrix was then supplied to the R library Deseq2 (v1.28.1) for a differential expression analysis (padj < 0.01 and absolute value of log2 fold change > 1). In addition, hierarchical clustering and heatmap analysis were performed for those genes of interest. Quantification values were z-scored prior to the clustering analysis by Pearson correlation with method “complete” (hclust function in stats package from R). Additional figures were drawn using TBtools.

Supplementary Materials

The following supporting information can be downloaded at: www.mdpi.com/article/10.3390/plants11121585/s1, Figure S1: Original protein product of the snap_masked-Fvb3-3-processed-gene-269.16; Figure S2: Original and revised sequences for FaWRKY21C; Figure S3: Original and revised sequences for FaWRKY21B; Figure S4: Original and revised sequences for FaWRKY51A.2; Figure S5: Gene structure of strawberry WRKY family; Figure S6: Phylogenetic analysis of Fragaria vesca (Fv) and Fragaria × ananassa (Fa) WRKY proteins; Figure S7: Gene retention and fractionation bias in chromosomes 6 and 7 from Fragaria vesca and Fragaria × ananassa cv. Camarosa syntenic genes (1:4 syntenic depth); Figure S8: Expression profiles of FvWRKY family members in different tissues, developmental stages and growth conditions of F. vesca; Figure S9: Expression and hierarchical clustering of FaWRKY genes in strawberry achene and receptacle of green fruits (Green_A and Green_R, respectively), leaves and root tissues; Figure S10: Expression and hierarchical clustering of FaWRKY genes during strawberry fruit ripening in receptacle (A) and achene (B) tissues; Figure S11: Interaction network constructed with the differentially expressed FaWRKYs during ripening, based on experimental and co-expression data with high confidence (0.700) from their Arabidopsis orthologs; Figure S12: Expression and hierarchical clustering of FaWRKY genes in strawberry leaves, inoculated with mock or C. fructicola spores and collected at 24, 72 and 96 h; Table S1: FvWRKY annotations across all the available genome versions; Table S2: Main characteristics of the Fragaria vesca WRKY genes and proteins; Table S3: FaWRKY proteins characteristics; Table S4: Collinear relationships detected by SynMap3D among F. vesca (Fv), A. thaliana (At) and V. vinifera (Vv) WRKY genes; Table S5: Collinear relationships detected by SynMap2 between Fv-At and Fv-Vv WRKY genes; Table S6: Shared synteny between Fv and Fa WRKY orthologs; Table S7: Strawberry (Fv and Fa) WRKY classification into eggNOG hierarchical orthologous groups; Table S8: 23ggnog homologs to the Arabidopsis WRKY54/55/70 proteins; Table S9: FvWRKY expression values (TPM) in 46 strawberry tissues; Table S10: Putative cis-elements in the promoters of 257 FaWRKY genes; Table S11: Putative cis-elements in the promoters of 64 FvWRKY genes; Table S12: Raw counts of FaWRKY genes; Table S13: Expression values (log2 fold change, LFC) of differentially expressed FaWRKY genes in tissues, fruit ripening and defense response against C. fructicola; Table S14: RNA-seq expression z-scores computed for FaWRKY genes in tissues and hieralchical clusters; Table S15: RNA-seq expression z-scores computed for FaWRKY genes in fruit receptacles and hieralchical clusters; Table S16: RNA-seq expression z-scores computed for FaWRKY genes in fruit achenes and hieralchical clusters; Table S17: RNA-seq expression z-scores computed for FaWRKY genes in infected leaves and hieralchical clusters; Table S18: Lists of differentially expressed FaWRKYs in fruit tissues during any stage of ripening; Table S19: A selection of putative candidate FaWRKY genes for further studies; Table S20: Synteny analyses performed in CoGe and persistent links to results.

Author Contributions

J.G.-G., J.-J.H., F.A.-R. and J.L.C. conceived, analyzed and interpreted the data presented in this work. J.G.-G., J.-J.H., F.A.-R., A.R.-F., J.M.-B. and J.L.C., contributed to drafting the manuscript and made critical suggestions. F.A.-R. and J.L.C. contributed equally to the supervision of this work. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by EU PRIMA foundation and Spanish AEI (Med-Berry PCI2019-103396 project) and Junta de Andalucía, Spain (Proyectos de Excelencia P12-AGR2174, and grants to Grupo-BIO278).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All of the related sequence data in this study were downloaded from public databases and detailed in the Material and Methods. The datasets that support the conclusions of this article are included in this article.

Acknowledgments

We thank F.A.-R., and the Central Service for Research Support (SCAI) at the University of Córdoba for its technical support in the bioinformatics data analysis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. 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]
  2. Bakshi, M.; Oelmuller, R. WRKY transcription factors: Jack of many trades in plants. Plant Signal. Behav. 2014, 9, e27700. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. 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] [PubMed] [Green Version]
  4. Zhang, Y.; Wang, L. The WRKY transcription factor superfamily: Its origin in eukaryotes and expansion in plants. BMC Evol. Biol. 2005, 5, 1. [Google Scholar] [CrossRef] [Green Version]
  5. Rushton, P.J.; Somssich, I.E.; Ringler, P.; Shen, Q.J. WRKY transcription factors. Trends Plant Sci. 2010, 15, 247–258. [Google Scholar] [CrossRef]
  6. Besseau, S.; Li, J.; Palva, E.T. WRKY54 and WRKY70 co-operate as negative regulators of leaf senescence in Arabidopsis thaliana. J. Exp. Bot 2012, 63, 2667–2679. [Google Scholar] [CrossRef]
  7. Chen, L.; Xiang, S.; Chen, Y.; Li, D.; Yu, D. Arabidopsis WRKY45 Interacts with the DELLA Protein RGL1 to Positively Regulate Age-Triggered Leaf Senescence. Mol. Plant 2017, 10, 1174–1189. [Google Scholar] [CrossRef] [Green Version]
  8. Doll, J.; Muth, M.; Riester, L.; Nebel, S.; Bresson, J.; Lee, H.C.; Zentgraf, U. Arabidopsis thaliana WRKY25 Transcription Factor Mediates Oxidative Stress Tolerance and Regulates Senescence in a Redox-Dependent Manner. Front. Plant Sci. 2019, 10, 1734. [Google Scholar] [CrossRef] [Green Version]
  9. Zhang, C.Q.; Xu, Y.; Lu, Y.; Yu, H.X.; Gu, M.H.; Liu, Q.Q. The WRKY transcription factor OsWRKY78 regulates stem elongation and seed development in rice. Planta 2011, 234, 541–554. [Google Scholar] [CrossRef]
  10. 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] [Green Version]
  11. Chen, L.; Song, Y.; Li, S.; Zhang, L.; Zou, C.; Yu, D. The role of WRKY transcription factors in plant abiotic stresses. Biochim. Biophys. Acta 2012, 1819, 120–128. [Google Scholar] [CrossRef]
  12. Tsuda, K.; Somssich, I.E. Transcriptional networks in plant immunity. New Phytol. 2015, 206, 932–947. [Google Scholar] [CrossRef]
  13. Alves, M.S.; Dadalto, S.P.; Goncalves, A.B.; de Souza, G.B.; Barros, V.A.; Fietto, L.G. Transcription Factor Functional Protein-Protein Interactions in Plant Defense Responses. Proteomes 2014, 2, 85–106. [Google Scholar] [CrossRef] [Green Version]
  14. Ciolkowski, I.; Wanke, D.; Birkenbihl, R.P.; Somssich, I.E. Studies on DNA-binding selectivity of WRKY transcription factors lend structural clues into WRKY-domain function. Plant Mol. Biol. 2008, 68, 81–92. [Google Scholar] [CrossRef] [Green Version]
  15. 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] [Green Version]
  16. Cheng, X.; Zhao, Y.; Jiang, Q.; Yang, J.; Zhao, W.; Taylor, I.A.; Peng, Y.L.; Wang, D.; Liu, J. Structural basis of dimerization and dual W-box DNA recognition by rice WRKY domain. Nucleic Acids Res. 2019, 47, 4308–4318. [Google Scholar] [CrossRef] [Green Version]
  17. Singh, A.; Sharma, A.K.; Singh, N.K.; Sonah, H.; Deshmukh, R.; Sharma, T.R. Understanding the Effect of Structural Diversity in WRKY Transcription Factors on DNA Binding Efficiency through Molecular Dynamics Simulation. Biology 2019, 8, 83. [Google Scholar] [CrossRef] [Green Version]
  18. Van Verk, M.C.; Pappaioannou, D.; Neeleman, L.; Bol, J.F.; Linthorst, H.J. 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] [Green Version]
  19. Yang, Y.; Zhou, Y.; Chi, Y.; Fan, B.; Chen, Z. Characterization of Soybean WRKY Gene Family and Identification of Soybean WRKY Genes that Promote Resistance to Soybean Cyst Nematode. Sci. Rep. 2017, 7, 17804. [Google Scholar] [CrossRef] [Green Version]
  20. Chen, X.; Li, C.; Wang, H.; Guo, Z. WRKY transcription factors: Evolution, binding, and action. Phytopathol. Res. 2019, 1, 13. [Google Scholar] [CrossRef]
  21. Machens, F.; Becker, M.; Umrath, F.; Hehl, R. Identification of a novel type of WRKY transcription factor binding site in elicitor-responsive cis-sequences from Arabidopsis thaliana. Plant Mol. Biol. 2014, 84, 371–385. [Google Scholar] [CrossRef]
  22. 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] [Green Version]
  23. DiMeglio, L.M.; Staudt, G.; Yu, H.; Davis, T.M. A phylogenetic analysis of the genus Fragaria (strawberry) using intron-containing sequence from the ADH-1 gene. PLoS ONE 2014, 9, e102237. [Google Scholar] [CrossRef] [Green Version]
  24. Shulaev, V.; Sargent, D.J.; Crowhurst, R.N.; Mockler, T.C.; Folkerts, O.; Delcher, A.L.; Jaiswal, P.; Mockaitis, K.; Liston, A.; Mane, S.P.; et al. The genome of woodland strawberry (Fragaria vesca). Nat. Genet. 2011, 43, 109–116. [Google Scholar] [CrossRef]
  25. Darwish, O.; Shahan, R.; Liu, Z.; Slovin, J.P.; Alkharouf, N.W. Re-annotation of the woodland strawberry (Fragaria vesca) genome. BMC Genom. 2015, 16, 29. [Google Scholar] [CrossRef] [Green Version]
  26. Tennessen, J.A.; Govindarajulu, R.; Ashman, T.L.; Liston, A. Evolutionary origins and dynamics of octoploid strawberry subgenomes revealed by dense targeted capture linkage maps. Genome Biol. Evol. 2014, 6, 3295–3313. [Google Scholar] [CrossRef] [Green Version]
  27. Li, Y.; Wei, W.; Feng, J.; Luo, H.; Pi, M.; Liu, Z.; Kang, C. Genome re-annotation of the wild strawberry Fragaria vesca using extensive Illumina- and SMRT-based RNA-seq datasets. DNA Res. 2018, 25, 61–70. [Google Scholar] [CrossRef] [Green Version]
  28. Edger, P.P.; VanBuren, R.; Colle, M.; Poorten, T.J.; Wai, C.M.; Niederhuth, C.E.; Alger, E.I.; Ou, S.; Acharya, C.B.; Wang, J.; et al. Single-molecule sequencing and optical mapping yields an improved genome of woodland strawberry (Fragaria vesca) with chromosome-scale contiguity. Gigascience 2018, 7, gix124. [Google Scholar] [CrossRef] [Green Version]
  29. Li, Y.; Pi, M.; Gao, Q.; Liu, Z.; Kang, C. Updated annotation of the wild strawberry Fragaria vesca V4 genome. Hortic. Res. 2019, 6, 61. [Google Scholar] [CrossRef] [Green Version]
  30. Edger, P.P.; Poorten, T.J.; VanBuren, R.; Hardigan, M.A.; Colle, M.; McKain, M.R.; Smith, R.D.; Teresi, S.J.; Nelson, A.D.L.; Wai, C.M.; et al. Origin and evolution of the octoploid strawberry genome. Nat. Genet. 2019, 51, 541–547. [Google Scholar] [CrossRef] [Green Version]
  31. Edger, P.P.; McKain, M.R.; Yocca, A.E.; Knapp, S.J.; Qiao, Q.; Zhang, T. Reply to: Revisiting the origin of octoploid strawberry. Nat. Genet. 2020, 52, 5–7. [Google Scholar] [CrossRef] [PubMed]
  32. Feng, C.; Wang, J.; Harris, A.J.; Folta, K.M.; Zhao, M.; Kang, M. Tracing the diploid ancestry of the cultivated octoploid strawberry. Mol. Biol. Evol. 2020, 38, 478–485. [Google Scholar] [CrossRef] [PubMed]
  33. Liston, A.; Wei, N.; Tennessen, J.A.; Li, J.; Dong, M.; Ashman, T.L. Revisiting the origin of octoploid strawberry. Nat. Genet. 2020, 52, 2–4. [Google Scholar] [CrossRef] [PubMed]
  34. Medina-Puche, L.; Blanco-Portales, R.; Molina-Hidalgo, F.J.; Cumplido-Laso, G.; Garcia-Caparros, N.; Moyano-Canete, E.; Caballero-Repullo, J.L.; Munoz-Blanco, J.; Rodriguez-Franco, A. Extensive transcriptomic studies on the roles played by abscisic acid and auxins in the development and ripening of strawberry fruits. Funct. Integr. Genom. 2016, 16, 671–692. [Google Scholar] [CrossRef] [PubMed]
  35. Sanchez-Sevilla, J.F.; Vallarino, J.G.; Osorio, S.; Bombarely, A.; Pose, D.; Merchante, C.; Botella, M.A.; Amaya, I.; Valpuesta, V. Gene expression atlas of fruit ripening and transcriptome assembly from RNA-seq data in octoploid strawberry (Fragaria × ananassa). Sci. Rep. 2017, 7, 13737. [Google Scholar] [CrossRef] [PubMed]
  36. Amil-Ruiz, F.; Garrido-Gala, J.; Blanco-Portales, R.; Folta, K.M.; Munoz-Blanco, J.; Caballero, J.L. Identification and validation of reference genes for transcript normalization in strawberry (Fragaria × ananassa) defense responses. PLoS ONE 2013, 8, e70603. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Amil-Ruiz, F.; Garrido-Gala, J.; Gadea, J.; Blanco-Portales, R.; Munoz-Merida, A.; Trelles, O.; de Los Santos, B.; Arroyo, F.T.; Aguado-Puig, A.; Romero, F.; et al. Partial Activation of SA- and JA-Defensive Pathways in Strawberry upon Colletotrichum acutatum Interaction. Front. Plant Sci. 2016, 7, 1036. [Google Scholar] [CrossRef] [Green Version]
  38. Wei, W.; Hu, Y.; Han, Y.T.; Zhang, K.; Zhao, F.L.; Feng, J.Y. The WRKY transcription factors in the diploid woodland strawberry Fragaria vesca: Identification and expression analysis under biotic and abiotic stresses. Plant Physiol. Biochem. 2016, 105, 129–144. [Google Scholar] [CrossRef]
  39. 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]
  40. Solovyev, V.; Kosarev, P.; Seledsov, I.; Vorobyev, D. Automatic annotation of eukaryotic genes, pseudogenes and promoters. Genome Biol. 2006, 7 (Suppl. 1), S10. [Google Scholar] [CrossRef] [Green Version]
  41. Panchy, N.; Lehti-Shiu, M.; Shiu, S.H. Evolution of Gene Duplication in Plants. Plant Physiol. 2016, 171, 2294–2316. [Google Scholar] [CrossRef] [Green Version]
  42. Lyons, E.; Freeling, M. How to usefully compare homologous plant genes and chromosomes as DNA sequences. Plant J. 2008, 53, 661–673. [Google Scholar] [CrossRef]
  43. Glover, N.M.; Redestig, H.; Dessimoz, C. Homoeologs: What Are They and How Do We Infer Them? Trends Plant Sci. 2016, 21, 609–621. [Google Scholar] [CrossRef] [Green Version]
  44. Haas, B.J.; Delcher, A.L.; Wortman, J.R.; Salzberg, S.L. DAGchainer: A tool for mining segmental genome duplications and synteny. Bioinformatics 2004, 20, 3643–3646. [Google Scholar] [CrossRef] [Green Version]
  45. Haug-Baltzell, A.; Stephens, S.A.; Davey, S.; Scheidegger, C.E.; Lyons, E. SynMap2 and SynMap3D: Web-based whole-genome synteny browsers. Bioinformatics 2017, 33, 2197–2198. [Google Scholar] [CrossRef]
  46. Wang, Q.; Wang, M.; Zhang, X.; Hao, B.; Kaushik, S.K.; Pan, Y. WRKY gene family evolution in Arabidopsis thaliana. Genetica 2011, 139, 973–983. [Google Scholar] [CrossRef]
  47. Chanderbali, A.S.; Berger, B.A.; Howarth, D.G.; Soltis, D.E.; Soltis, P.S. Evolution of floral diversity: Genomics, genes and gamma. Philos. Trans. R. Soc. B Biol. Sci. 2017, 372, 20150509. [Google Scholar] [CrossRef] [Green Version]
  48. Sahebi, M.; Hanafi, M.M.; van Wijnen, A.J.; Rice, D.; Rafii, M.Y.; Azizi, P.; Osman, M.; Taheri, S.; Bakar, M.F.A.; Isa, M.N.M.; et al. Contribution of transposable elements in the plant’s genome. Gene 2018, 665, 155–166. [Google Scholar] [CrossRef]
  49. Bailey, P.C.; Schudoma, C.; Jackson, W.; Baggs, E.; Dagdas, G.; Haerty, W.; Moscou, M.; Krasileva, K.V. Dominant integration locus drives continuous diversification of plant immune receptors with exogenous domain fusions. Genome Biol. 2018, 19, 23. [Google Scholar] [CrossRef]
  50. Moore, R.C.; Purugganan, M.D. The evolutionary dynamics of plant duplicate genes. Curr. Opin. Plant Biol. 2005, 8, 122–128. [Google Scholar] [CrossRef]
  51. Vicient, C.M.; Casacuberta, J.M. Impact of transposable elements on polyploid plant genomes. Ann. Bot. 2017, 120, 195–207. [Google Scholar] [CrossRef]
  52. Bashir, T.; Chandra Mishra, R.; Hasan, M.M.; Mohanta, T.K.; Bae, H. Effect of Hybridization on Somatic Mutations and Genomic Rearrangements in Plants. Int. J. Mol. Sci. 2018, 19, 3758. [Google Scholar] [CrossRef] [Green Version]
  53. Hughes, A.L.; Friedman, R.; Ekollu, V.; Rose, J.R. Non-random association of transposable elements with duplicated genomic blocks in Arabidopsis thaliana. Mol. Phylogenet. Evol. 2003, 29, 410–416. [Google Scholar] [CrossRef]
  54. Liu, D.; Hunt, M.; Tsai, I.J. Inferring synteny between genome assemblies: A systematic evaluation. BMC Bioinform. 2018, 19, 26. [Google Scholar] [CrossRef]
  55. Cooper, G.M.; Brown, C.D. Qualifying the relationship between sequence conservation and molecular function. Genome Res. 2008, 18, 201–205. [Google Scholar] [CrossRef] [Green Version]
  56. Jiao, Y.; Leebens-Mack, J.; Ayyampalayam, S.; Bowers, J.E.; McKain, M.R.; McNeal, J.; Rolf, M.; Ruzicka, D.R.; Wafula, E.; Wickett, N.J.; et al. A genome triplication associated with early diversification of the core eudicots. Genome Biol. 2012, 13, R3. [Google Scholar] [CrossRef] [Green Version]
  57. Murat, F.; Armero, A.; Pont, C.; Klopp, C.; Salse, J. Reconstructing the genome of the most recent common ancestor of flowering plants. Nat. Genet. 2017, 49, 490–496. [Google Scholar] [CrossRef]
  58. Xiang, Y.; Huang, C.H.; Hu, Y.; Wen, J.; Li, S.; Yi, T.; Chen, H.; Xiang, J.; Ma, H. Evolution of Rosaceae Fruit Types Based on Nuclear Phylogeny in the Context of Geological Times and Genome Duplication. Mol. Biol. Evol. 2017, 34, 262–281. [Google Scholar] [CrossRef] [Green Version]
  59. Koonin, E.V. Orthologs, paralogs, and evolutionary genomics. Annu. Rev. Genet. 2005, 39, 309–338. [Google Scholar] [CrossRef] [Green Version]
  60. Nehrt, N.L.; Clark, W.T.; Radivojac, P.; Hahn, M.W. Testing the ortholog conjecture with comparative functional genomic data from mammals. PLoS Comput. Biol. 2011, 7, e1002073. [Google Scholar] [CrossRef]
  61. APG (The Angiosperm Phylogeny Group). An update of the Angiosperm Phylogeny Group classification for the orders and families of flowering plants: APG IV. Bot. J. Linn. Soc. 2016, 181, 1–20. [Google Scholar] [CrossRef] [Green Version]
  62. Tiley, G.P.; Ane, C.; Burleigh, J.G. Evaluating and Characterizing Ancient Whole-Genome Duplications in Plants with Gene Count Data. Genome Biol. Evol. 2016, 8, 1023–1037. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Kim, K.D.; El Baidouri, M.; Abernathy, B.; Iwata-Otsubo, A.; Chavarro, C.; Gonzales, M.; Libault, M.; Grimwood, J.; Jackson, S.A. A Comparative Epigenomic Analysis of Polyploidy-Derived Genes in Soybean and Common Bean. Plant Physiol. 2015, 168, 1433–1447. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Schmutz, J.; Cannon, S.B.; Schlueter, J.; Ma, J.; Mitros, T.; Nelson, W.; Hyten, D.L.; Song, Q.; Thelen, J.J.; Cheng, J.; et al. Genome sequence of the palaeopolyploid soybean. Nature 2010, 463, 178–183. [Google Scholar] [CrossRef] [Green Version]
  65. Velasco, R.; Zharkikh, A.; Affourtit, J.; Dhingra, A.; Cestaro, A.; Kalyanaraman, A.; Fontana, P.; Bhatnagar, S.K.; Troggio, M.; Pruss, D.; et al. The genome of the domesticated apple (Malus x domestica Borkh.). Nat. Genet. 2010, 42, 833–839. [Google Scholar] [CrossRef]
  66. Luo, M.C.; You, F.M.; Li, P.; Wang, J.R.; Zhu, T.; Dandekar, A.M.; Leslie, C.A.; Aradhya, M.; McGuire, P.E.; Dvorak, J. Synteny analysis in Rosids with a walnut physical map reveals slow genome evolution in long-lived woody perennials. BMC Genom. 2015, 16, 707. [Google Scholar] [CrossRef] [Green Version]
  67. Chen, J.; Nolan, T.M.; Ye, H.; Zhang, M.; Tong, H.; Xin, P.; Chu, J.; Chu, C.; Li, Z.; Yin, Y. Arabidopsis WRKY46, WRKY54, and WRKY70 Transcription Factors Are Involved in Brassinosteroid-Regulated Plant Growth and Drought Responses. Plant Cell 2017, 29, 1425–1439. [Google Scholar] [CrossRef]
  68. Li, J.; Zhong, R.; Palva, E.T. WRKY70 and its homolog WRKY54 negatively modulate the cell wall-associated defenses to necrotrophic pathogens in Arabidopsis. PLoS ONE 2017, 12, e0183731. [Google Scholar] [CrossRef]
  69. Amborella Genome, P. The Amborella genome and the evolution of flowering plants. Science 2013, 342, 1241089. [Google Scholar] [CrossRef]
  70. Lei, Y.; Sun, Y.; Wang, B.; Yu, S.; Dai, H.; Li, H.; Zhang, Z.; Zhang, J. Woodland strawberry WRKY71 acts as a promoter of flowering via a transcriptional regulatory cascade. Hortic. Res. 2020, 7, 137. [Google Scholar] [CrossRef]
  71. Wei, W.; Cui, M.Y.; Hu, Y.; Gao, K.; Xie, Y.G.; Jiang, Y.; Feng, J.Y. Ectopic expression of FvWRKY42, a WRKY transcription factor from the diploid woodland strawberry (Fragaria vesca), enhances resistance to powdery mildew, improves osmotic stress resistance, and increases abscisic acid sensitivity in Arabidopsis. Plant Sci. An. Int. J. Exp. Plant Biol. 2018, 275, 60–74. [Google Scholar] [CrossRef]
  72. Zhang, W.W.; Zhao, S.Q.; Gu, S.; Cao, X.Y.; Zhang, Y.; Niu, J.F.; Liu, L.; Li, A.R.; Jia, W.S.; Qi, B.X.; et al. FvWRKY48 binds to the pectate lyase FvPLA promoter to control fruit softening in Fragaria vesca. Plant Physiol. 2022, 189, 1037–1049. [Google Scholar] [CrossRef]
  73. Hawkins, C.; Caruana, J.; Li, J.; Zawora, C.; Darwish, O.; Wu, J.; Alkharouf, N.; Liu, Z. An eFP browser for visualizing strawberry fruit and flower transcriptomes. Hortic. Res. 2017, 4, 17029. [Google Scholar] [CrossRef] [Green Version]
  74. Hollender, C.A.; Geretz, A.C.; Slovin, J.P.; Liu, Z. Flower and early fruit development in a diploid strawberry, Fragaria vesca. Planta 2012, 235, 1123–1139. [Google Scholar] [CrossRef]
  75. Toljamo, A.; Blande, D.; Karenlampi, S.; Kokko, H. Reprogramming of Strawberry (Fragaria vesca) Root Transcriptome in Response to Phytophthora cactorum. PLoS ONE 2016, 11, e0161078. [Google Scholar] [CrossRef] [Green Version]
  76. Hussain, R.M.F.; Sheikh, A.H.; Haider, I.; Quareshy, M.; Linthorst, H.J.M. Arabidopsis WRKY50 and TGA Transcription Factors Synergistically Activate Expression of PR1. Front. Plant Sci. 2018, 9, 930. [Google Scholar] [CrossRef]
  77. Gao, Q.M.; Venugopal, S.; Navarre, D.; Kachroo, A. Low oleic acid-derived repression of jasmonic acid-inducible defense responses requires the WRKY50 and WRKY51 proteins. Plant Physiol. 2011, 155, 464–476. [Google Scholar] [CrossRef] [Green Version]
  78. Chen, L.; Zhang, L.; Xiang, S.; Chen, Y.; Zhang, H.; Yu, D. The transcription factor WRKY75 positively regulates jasmonate-mediated plant defense to necrotrophic fungal pathogens. J. Exp. Bot. 2021, 72, 1473–1489. [Google Scholar] [CrossRef]
  79. Encinas-Villarejo, S.; Maldonado, A.M.; Amil-Ruiz, F.; de los Santos, B.; Romero, F.; Pliego-Alfaro, F.; Muñoz-Blanco, J.; Caballero, J.L. Evidence for a positive regulatory role of strawberry (Fragaria × ananassa) Fa WRKY1 and Arabidopsis At WRKY75 proteins in resistance. J. Exp. Bot. 2009, 60, 3043–3065. [Google Scholar] [CrossRef] [Green Version]
  80. Higuera, J.J.; Garrido-Gala, J.; Lekhbou, A.; Arjona-Girona, I.; Amil-Ruiz, F.; Mercado, J.A.; Pliego-Alfaro, F.; Munoz-Blanco, J.; Lopez-Herrera, C.J.; Caballero, J.L. The Strawberry FaWRKY1 Transcription Factor Negatively Regulates Resistance to Colletotrichum acutatum in Fruit Upon Infection. Front. Plant Sci. 2019, 10, 480. [Google Scholar] [CrossRef]
  81. Eulgem, T. Dissecting the WRKY web of plant defense regulators. PLoS Pathog. 2006, 2, e126. [Google Scholar] [CrossRef] [Green Version]
  82. Wang, D.; Amornsiripanitch, N.; Dong, X. A genomic approach to identify regulatory nodes in the transcriptional network of systemic acquired resistance in plants. PLoS Pathog. 2006, 2, e123. [Google Scholar] [CrossRef] [Green Version]
  83. Birkenbihl, R.P.; Kracher, B.; Ross, A.; Kramer, K.; Finkemeier, I.; Somssich, I.E. Principles and characteristics of the Arabidopsis WRKY regulatory network during early MAMP-triggered immunity. Plant J. 2018, 96, 487–502. [Google Scholar] [CrossRef] [Green Version]
  84. Zheng, Z.; Mosher, S.L.; Fan, B.; Klessig, D.F.; Chen, Z. Functional analysis of Arabidopsis WRKY25 transcription factor in plant defense against Pseudomonas syringae. BMC Plant Biol. 2007, 7, 2. [Google Scholar] [CrossRef] [Green Version]
  85. Roulin, A.; Auer, P.L.; Libault, M.; Schlueter, J.; Farmer, A.; May, G.; Stacey, G.; Doerge, R.W.; Jackson, S.A. The fate of duplicated genes in a polyploid plant genome. Plant J. 2013, 73, 143–153. [Google Scholar] [CrossRef]
  86. Zhang, L.; Huang, X.; He, C.; Zhang, Q.Y.; Zou, X.; Duan, K.; Gao, Q. Novel Fungal Pathogenicity and Leaf Defense Strategies Are Revealed by Simultaneous Transcriptome Analysis of Colletotrichum fructicola and Strawberry Infected by This Fungus. Front. Plant Sci. 2018, 9, 434. [Google Scholar] [CrossRef] [Green Version]
  87. Devaiah, B.N.; Karthikeyan, A.S.; Raghothama, K.G. WRKY75 transcription factor is a modulator of phosphate acquisition and root development in Arabidopsis. Plant Physiol. 2007, 143, 1789–1801. [Google Scholar] [CrossRef] [Green Version]
  88. Garrido-Gala, J.; Higuera, J.J.; Munoz-Blanco, J.; Amil-Ruiz, F.; Caballero, J.L. The VQ motif-containing proteins in the diploid and octoploid strawberry. Sci. Rep. 2019, 9, 4942. [Google Scholar] [CrossRef]
  89. Medina-Puche, L.; Martinez-Rivas, F.J.; Molina-Hidalgo, F.J.; Mercado, J.A.; Moyano, E.; Rodriguez-Franco, A.; Caballero, J.L.; Munoz-Blanco, J.; Blanco-Portales, R. An atypical HLH transcriptional regulator plays a novel and important role in strawberry ripened receptacle. BMC Plant Biol. 2019, 19, 586. [Google Scholar] [CrossRef]
  90. Zhao, F.; Li, G.; Hu, P.; Zhao, X.; Li, L.; Wei, W.; Feng, J.; Zhou, H. Identification of basic/helix-loop-helix transcription factors reveals candidate genes involved in anthocyanin biosynthesis from the strawberry white-flesh mutant. Sci. Rep. 2018, 8, 2721. [Google Scholar] [CrossRef] [Green Version]
  91. Bhattarai, K.K.; Atamian, H.S.; Kaloshian, I.; Eulgem, T. WRKY72-type transcription factors contribute to basal immunity in tomato and Arabidopsis as well as gene-for-gene resistance mediated by the tomato R gene Mi-1. Plant J. 2010, 63, 229–240. [Google Scholar] [CrossRef] [PubMed]
  92. Van Verk, M.C.; Bol, J.F.; Linthorst, H.J. WRKY transcription factors involved in activation of SA biosynthesis genes. BMC Plant Biol. 2011, 11, 89. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Hu, Y.; Dong, Q.; Yu, D. Arabidopsis WRKY46 coordinates with WRKY70 and WRKY53 in basal resistance against pathogen Pseudomonas syringae. Plant Sci. Int. J. Exp. Plant Biol. 2012, 185–186, 288–297. [Google Scholar] [CrossRef] [PubMed]
  94. Li, J.; Brader, G.; Kariola, T.; Palva, E.T. WRKY70 modulates the selection of signaling pathways in plant defense. Plant J. 2006, 46, 477–491. [Google Scholar] [CrossRef]
  95. Li, J.; Brader, G.; Palva, E.T. The WRKY70 transcription factor: A node of convergence for jasmonate-mediated and salicylate-mediated signals in plant defense. Plant Cell 2004, 16, 319–331. [Google Scholar] [CrossRef] [Green Version]
  96. Birkenbihl, R.P.; Kracher, B.; Roccaro, M.; Somssich, I.E. Induced Genome-Wide Binding of Three Arabidopsis WRKY Transcription Factors during Early MAMP-Triggered Immunity. Plant Cell 2017, 29, 20–38. [Google Scholar] [CrossRef] [Green Version]
  97. Pandey, S.P.; Roccaro, M.; Schon, M.; Logemann, E.; Somssich, I.E. Transcriptional reprogramming regulated by WRKY18 and WRKY40 facilitates powdery mildew infection of Arabidopsis. Plant J. 2010, 64, 912–923. [Google Scholar] [CrossRef]
  98. Zheng, Z.; Qamar, S.A.; Chen, Z.; Mengiste, T. Arabidopsis WRKY33 transcription factor is required for resistance to necrotrophic fungal pathogens. Plant J. 2006, 48, 592–605. [Google Scholar] [CrossRef]
  99. Wang, F.; Zhang, F.; Chen, M.; Liu, Z.; Zhang, Z.; Fu, J.; Ma, Y. Comparative Transcriptomics Reveals Differential Gene Expression Related to Colletotrichum gloeosporioides Resistance in the Octoploid Strawberry. Front. Plant Sci. 2017, 8, 779. [Google Scholar] [CrossRef] [Green Version]
  100. Mukhi, N.; Brown, H.; Gorenkin, D.; Ding, P.; Bentham, A.R.; Stevenson, C.E.M.; Jones, J.D.G.; Banfield, M.J. Perception of structurally distinct effectors by the integrated WRKY domain of a plant immune receptor. Proc. Natl. Acad. Sci. USA 2021, 118, e2113996118. [Google Scholar] [CrossRef]
  101. Narusaka, M.; Shirasu, K.; Noutoshi, Y.; Kubo, Y.; Shiraishi, T.; Iwabuchi, M.; Narusaka, Y. RRS1 and RPS4 provide a dual Resistance-gene system against fungal and bacterial pathogens. Plant J. 2009, 60, 218–226. [Google Scholar] [CrossRef]
  102. Jung, S.; Lee, T.; Cheng, C.H.; Buble, K.; Zheng, P.; Yu, J.; Humann, J.; Ficklin, S.P.; Gasic, K.; Scott, K.; et al. 15 years of GDR: New data and functionality in the Genome Database for Rosaceae. Nucleic Acids Res. 2019, 47, D1137–D1145. [Google Scholar] [CrossRef] [Green Version]
  103. El-Gebali, S.; Mistry, J.; Bateman, A.; Eddy, S.R.; Luciani, A.; Potter, S.C.; Qureshi, M.; Richardson, L.J.; Salazar, G.A.; Smart, A.; et al. The Pfam protein families database in 2019. Nucleic Acids Res. 2019, 47, D427–D432. [Google Scholar] [CrossRef]
  104. Okonechnikov, K.; Golosova, O.; Fursov, M.; Ugene Team. Unipro UGENE: A unified bioinformatics toolkit. Bioinformatics 2012, 28, 1166–1167. [Google Scholar] [CrossRef] [Green Version]
  105. Marchler-Bauer, A.; Bo, Y.; Han, L.; He, J.; Lanczycki, C.J.; Lu, S.; Chitsaz, F.; Derbyshire, M.K.; Geer, R.C.; Gonzales, N.R.; et al. CDD/SPARCLE: Functional classification of proteins via subfamily domain architectures. Nucleic Acids Res. 2017, 45, D200–D203. [Google Scholar] [CrossRef] [Green Version]
  106. 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]
  107. 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] [Green Version]
  108. Voorrips, R.E. MapChart: Software for the graphical presentation of linkage maps and QTLs. J. Hered. 2002, 93, 77–78. [Google Scholar] [CrossRef] [Green Version]
  109. Sperschneider, J.; Catanzariti, A.M.; DeBoer, K.; Petre, B.; Gardiner, D.M.; Singh, K.B.; Dodds, P.N.; Taylor, J.M. LOCALIZER: Subcellular localization prediction of both plant and effector proteins in the plant cell. Sci. Rep. 2017, 7, 44598. [Google Scholar] [CrossRef] [Green Version]
  110. Afgan, E.; Baker, D.; van den Beek, M.; Blankenberg, D.; Bouvier, D.; Cech, M.; Chilton, J.; Clements, D.; Coraor, N.; Eberhard, C.; et al. The Galaxy platform for accessible, reproducible and collaborative biomedical analyses: 2016 update. Nucleic Acids Res. 2016, 44, W3–W10. [Google Scholar] [CrossRef] [Green Version]
  111. Kumar, S.; Stecher, G.; Tamura, K. MEGA7: Molecular Evolutionary Genetics Analysis Version 7.0 for Bigger Datasets. Mol. Biol. Evol. 2016, 33, 1870–1874. [Google Scholar] [CrossRef] [Green Version]
  112. Letunic, I.; Bork, P. Interactive Tree Of Life (iTOL) v4: Recent updates and new developments. Nucleic Acids Res. 2019, 47, W256–W259. [Google Scholar] [CrossRef] [Green Version]
  113. Suyama, M.; Torrents, D.; Bork, P. PAL2NAL: Robust conversion of protein sequence alignments into the corresponding codon alignments. Nucleic Acids Res. 2006, 34, W609–W612. [Google Scholar] [CrossRef] [Green Version]
  114. Yang, Z.; Nielsen, R. Codon-substitution models for detecting molecular adaptation at individual sites along specific lineages. Mol. Biol. Evol. 2002, 19, 908–917. [Google Scholar] [CrossRef] [Green Version]
  115. Villanueva-Canas, J.L.; Laurie, S.; Alba, M.M. Improving genome-wide scans of positive selection by using protein isoforms of similar length. Genome Biol. Evol. 2013, 5, 457–467. [Google Scholar] [CrossRef] [Green Version]
  116. De La Torre, A.R.; Li, Z.; Van de Peer, Y.; Ingvarsson, P.K. Contrasting Rates of Molecular Evolution and Patterns of Selection among Gymnosperms and Flowering Plants. Mol. Biol. Evol. 2017, 34, 1363–1377. [Google Scholar] [CrossRef] [Green Version]
  117. Huerta-Cepas, J.; Forslund, K.; Coelho, L.P.; Szklarczyk, D.; Jensen, L.J.; von Mering, C.; Bork, P. Fast Genome-Wide Functional Annotation through Orthology Assignment by eggNOG-Mapper. Mol. Biol. Evol. 2017, 34, 2115–2122. [Google Scholar] [CrossRef] [Green Version]
  118. Huerta-Cepas, J.; Szklarczyk, D.; Heller, D.; Hernandez-Plaza, A.; Forslund, S.K.; Cook, H.; Mende, D.R.; Letunic, I.; Rattei, T.; Jensen, L.J.; et al. eggNOG 5.0: A hierarchical, functionally and phylogenetically annotated orthology resource based on 5090 organisms and 2502 viruses. Nucleic Acids Res. 2019, 47, D309–D314. [Google Scholar] [CrossRef] [Green Version]
  119. Ye, J.; Zhang, Y.; Cui, H.; Liu, J.; Wu, Y.; Cheng, Y.; Xu, H.; Huang, X.; Li, S.; Zhou, A.; et al. WEGO 2.0: A web tool for analyzing and plotting GO annotations, 2018 update. Nucleic Acids Res. 2018, 46, W71–W75. [Google Scholar] [CrossRef] [PubMed]
  120. Szklarczyk, D.; Gable, A.L.; Nastou, K.C.; Lyon, D.; Kirsch, R.; Pyysalo, S.; Doncheva, N.T.; Legeay, M.; Fang, T.; Bork, P.; et al. The STRING database in 2021: Customizable protein–protein networks, and functional characterization of user-uploaded gene/measurement sets. Nucleic Acids Res. 2020, 49, D605–D612. [Google Scholar] [CrossRef] [PubMed]
  121. Kopylova, E.; Noe, L.; Touzet, H. SortMeRNA: Fast and accurate filtering of ribosomal RNAs in metatranscriptomic data. Bioinformatics 2012, 28, 3211–3217. [Google Scholar] [CrossRef] [PubMed]
  122. Dobin, A.; Davis, C.A.; Schlesinger, F.; Drenkow, J.; Zaleski, C.; Jha, S.; Batut, P.; Chaisson, M.; Gingeras, T.R. STAR: Ultrafast universal RNA-seq aligner. Bioinformatics 2013, 29, 15–21. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Protein domains (PFAM v32.0) found in FvWRKY proteins. Only the largest splice forms are represented. See Table S2 for more details.
Figure 1. Protein domains (PFAM v32.0) found in FvWRKY proteins. Only the largest splice forms are represented. See Table S2 for more details.
Plants 11 01585 g001
Figure 2. Chromosome mapping of the Fragaria vesca WRKY gene family. Segmental duplicated gene pairs (syntenic paralogs) share the same colors. Tandemly duplicated gene clusters are colored in pink, and underlined genes are collinear with genes outside that syntenic block. Ruler size is in Megabases.
Figure 2. Chromosome mapping of the Fragaria vesca WRKY gene family. Segmental duplicated gene pairs (syntenic paralogs) share the same colors. Tandemly duplicated gene clusters are colored in pink, and underlined genes are collinear with genes outside that syntenic block. Ruler size is in Megabases.
Plants 11 01585 g002
Figure 3. GEvo analyses comparing microsynteny between genomic regions from A. thaliana and F. vesca. (A) Genomic regions containing the genes FvWRKY60, FvWRKY61 and AtWRKY70 with red blocks and connectors show high-scoring sequence pairs between both sequences. The region shown in B is framed. (B) Homology details among FvWRKY60, FvWRKY61 (black arrows) and AtWRKY70. A short, non-WRKY gene (FvH4_7g26100, red arrow) was identified as part of the FvWRKY59-60-61 tandem (see text for details).
Figure 3. GEvo analyses comparing microsynteny between genomic regions from A. thaliana and F. vesca. (A) Genomic regions containing the genes FvWRKY60, FvWRKY61 and AtWRKY70 with red blocks and connectors show high-scoring sequence pairs between both sequences. The region shown in B is framed. (B) Homology details among FvWRKY60, FvWRKY61 (black arrows) and AtWRKY70. A short, non-WRKY gene (FvH4_7g26100, red arrow) was identified as part of the FvWRKY59-60-61 tandem (see text for details).
Plants 11 01585 g003
Figure 4. Chromosome mapping of the Fragaria × ananassa WRKY gene family within each subgenome. Segmental duplicated gene pairs (syntenic paralogs) share the same colors. Non-syntenic gene duplicates are underlined. Tandemly duplicated gene clusters are colored in orange. Ruler size is in Megabases.
Figure 4. Chromosome mapping of the Fragaria × ananassa WRKY gene family within each subgenome. Segmental duplicated gene pairs (syntenic paralogs) share the same colors. Non-syntenic gene duplicates are underlined. Tandemly duplicated gene clusters are colored in orange. Ruler size is in Megabases.
Plants 11 01585 g004
Figure 5. Phylogenetic analysis of Fragaria vesca (Fv), Arabidopsis thaliana (At) and Vitis vinifera (Vv) WRKY proteins. WRKY proteins are clustered into Groups I + IIc, IIa + IIb, IId + IIe and III. R protein-WRKY from Fv and At are clustered within their respective groups. The tree was inferred using the Neighbor-Joining method (1000 bootstrap replicates) and drawn to scale, with branch lengths in the same units as those of the evolutionary distances used to infer the phylogenetic tree. The evolutionary distances were computed using the p-distance method and are in the units of the number of amino acid differences per site. All positions with less than 95% site coverage were eliminated. Connecting lines represent shared synteny between Fv-Vv (red) and Fv-At (blue).
Figure 5. Phylogenetic analysis of Fragaria vesca (Fv), Arabidopsis thaliana (At) and Vitis vinifera (Vv) WRKY proteins. WRKY proteins are clustered into Groups I + IIc, IIa + IIb, IId + IIe and III. R protein-WRKY from Fv and At are clustered within their respective groups. The tree was inferred using the Neighbor-Joining method (1000 bootstrap replicates) and drawn to scale, with branch lengths in the same units as those of the evolutionary distances used to infer the phylogenetic tree. The evolutionary distances were computed using the p-distance method and are in the units of the number of amino acid differences per site. All positions with less than 95% site coverage were eliminated. Connecting lines represent shared synteny between Fv-Vv (red) and Fv-At (blue).
Plants 11 01585 g005
Figure 6. GO functional annotation plot comparing the number and percentage of FvWRKY and FaWRKY proteins sharing the same functions. The y-axis represents the number (right) and the percentage (left) of WRKY genes in Fragaria × ananassa (green) and in Fragaria vesca (blue) for each GO category (x-axis).
Figure 6. GO functional annotation plot comparing the number and percentage of FvWRKY and FaWRKY proteins sharing the same functions. The y-axis represents the number (right) and the percentage (left) of WRKY genes in Fragaria × ananassa (green) and in Fragaria vesca (blue) for each GO category (x-axis).
Plants 11 01585 g006
Figure 7. Phylogenetic tree of the AtWRKY55 and AtWRKY54/70 orthologous genes (red and blue branches, respectively) in strawberry (Fv), soybean (Glyma), walnut (WALNUT), apple (MD), grapevine (Vv) and Amborella (AmTr). Strawberry orthologs to AtWRKY54/70 paralogs underwent tandem duplications after the gamma hexaploidization and independently to the more recent WGD events in other species. The evolutionary history was inferred using the Maximum Likelihood method (100 bootstraps) with optimized parameters (TN + G + I). The tree is drawn to scale, with lengths of branches representing the number of substitutions per site. Analyses were conducted in MEGA7.
Figure 7. Phylogenetic tree of the AtWRKY55 and AtWRKY54/70 orthologous genes (red and blue branches, respectively) in strawberry (Fv), soybean (Glyma), walnut (WALNUT), apple (MD), grapevine (Vv) and Amborella (AmTr). Strawberry orthologs to AtWRKY54/70 paralogs underwent tandem duplications after the gamma hexaploidization and independently to the more recent WGD events in other species. The evolutionary history was inferred using the Maximum Likelihood method (100 bootstraps) with optimized parameters (TN + G + I). The tree is drawn to scale, with lengths of branches representing the number of substitutions per site. Analyses were conducted in MEGA7.
Plants 11 01585 g007
Figure 8. Phylogenetic tree of the FvWRKY family showing expression values during fruit ripening (Ruegen receptacle tissue at 15 and at 22 DPA and Yellow Wonder receptacle tissue at 15 and at 22 DPA: Yw-15D, Yw-22D, Rg-15D and Rg-22D, respectively) and roots (Root, collected from 7-week-old plants grown in aerated hydroponic culture, and Root_P, after 2 days of inoculation with Phytophthora cactorum). FvWRKY paralogs are shown in red (see Table 1. Color scale represents the expression level as log-transformed TPM (Transcripts per Million) values.
Figure 8. Phylogenetic tree of the FvWRKY family showing expression values during fruit ripening (Ruegen receptacle tissue at 15 and at 22 DPA and Yellow Wonder receptacle tissue at 15 and at 22 DPA: Yw-15D, Yw-22D, Rg-15D and Rg-22D, respectively) and roots (Root, collected from 7-week-old plants grown in aerated hydroponic culture, and Root_P, after 2 days of inoculation with Phytophthora cactorum). FvWRKY paralogs are shown in red (see Table 1. Color scale represents the expression level as log-transformed TPM (Transcripts per Million) values.
Plants 11 01585 g008
Figure 9. (A) Heatmap of differentially expressed FaWRKY genes during strawberry fruit ripening stages in receptacle (White_R, Turning_R, and Red_R; left side) and achene (White_A, Turning_A, and Red_A; right side) tissues. Changes in gene expression, with respect to green fruit tissues, are represented as log2 fold change if absolute values were >1 (padj < 0.01), otherwise, they were colored in grey. (B) Venn diagram showing those genes which were differentially expressed in receptacle and achene only, or in both.
Figure 9. (A) Heatmap of differentially expressed FaWRKY genes during strawberry fruit ripening stages in receptacle (White_R, Turning_R, and Red_R; left side) and achene (White_A, Turning_A, and Red_A; right side) tissues. Changes in gene expression, with respect to green fruit tissues, are represented as log2 fold change if absolute values were >1 (padj < 0.01), otherwise, they were colored in grey. (B) Venn diagram showing those genes which were differentially expressed in receptacle and achene only, or in both.
Plants 11 01585 g009
Figure 10. Heatmap of differentially expressed FaWRKY genes in response to C. fructicola infection in leaves. Changes of gene expression in infected strawberry (Inf_24h, Inf_72h, Inf_96h; 24, 72 and 96 h after spore inoculation, respectively) vs. mock-inoculated strawberry are represented as log2 fold change if absolute values were >1 (padj < 0.01), in a color scale from lowest (blue) to highest (red). Otherwise, they were colored grey.
Figure 10. Heatmap of differentially expressed FaWRKY genes in response to C. fructicola infection in leaves. Changes of gene expression in infected strawberry (Inf_24h, Inf_72h, Inf_96h; 24, 72 and 96 h after spore inoculation, respectively) vs. mock-inoculated strawberry are represented as log2 fold change if absolute values were >1 (padj < 0.01), in a color scale from lowest (blue) to highest (red). Otherwise, they were colored grey.
Plants 11 01585 g010
Table 1. FvWRKY gene duplications. Synmap2 did not provide the nucleotide substitution values (Kn, Ks) nor ω for tandemly duplicated genes. Genes marked with asterisks are shared by different duplicate groups.
Table 1. FvWRKY gene duplications. Synmap2 did not provide the nucleotide substitution values (Kn, Ks) nor ω for tandemly duplicated genes. Genes marked with asterisks are shared by different duplicate groups.
Gene 1Gene 2Gene 3Gene 4Gene 5KnKsω
Segmental duplications
FvWRKY2FvWRKY9 0.45836.70820.0683
FvWRKY7FvWRKY36 0.31831.52650.2085
FvWRKY13FvWRKY50 * 0.35502.66630.1331
FvWRKY14FvWRKY48 0.45662.69470.1694
FvWRKY15FvWRKY46 0.41855.74620.0728
FvWRKY17FvWRKY50 * 0.39085.79740.0674
FvWRKY19FvWRKY43 0.36801.81000.2033
FvWRKY24 * FvWRKY53 0.47023.15110.1492
FvWRKY24 * FvWRKY30 0.40205.00960.0802
FvWRKY26FvWRKY51 0.32953.02820.1088
FvWRKY27FvWRKY31 0.72464.11630.1760
FvWRKY27FvWRKY52 0.58802.97600.1976
FvWRKY29FvWRKY58 * 0.47002.24600.2093
FvWRKY30FvWRKY53 0.462877.42160.0060
FvWRKY39 * FvWRKY61 * 0.78492.71590.2890
FvWRKY58 * FvWRKY64 0.586363.43120.0092
Tandem repeats
FvWRKY11FvWRKY12 NANANA
FvWRKY38FvWRKY39 * FvWRKY40FvWRKY41FvWRKY42NANANA
FvWRKY59FvWRKY60FvWRKY61 * NANANA
Table 2. FaWRKY gene duplications. Nucleotide substitution values (Kn, Ks) and ω for tandemly duplicated genes were not provided by Synmap2. PAL2NAL was used to calculate Kn and Ks of the non-syntenic paralogs. ω values were not calculated if Ks < 0.01.
Table 2. FaWRKY gene duplications. Nucleotide substitution values (Kn, Ks) and ω for tandemly duplicated genes were not provided by Synmap2. PAL2NAL was used to calculate Kn and Ks of the non-syntenic paralogs. ω values were not calculated if Ks < 0.01.
Gene 1Gene 2Gene 3KnKsω
Segmental duplications
FaWRKY6C.1FaWRKY6C.2 0.00260.0000NA
FaWRKY43B.1FaWRKY43B.2 0.00170.0055NA
FaWRKY54A.1FaWRKY54A.2 0.00000.0039NA
FaWRKY54D.1FaWRKY54D.2 0.00160.0000NA
FaWRKY56A.1FaWRKY56A.2 0.00830.01090.7615
FaWRKY56D.1FaWRKY56D.2 0.00130.0047NA
FaWRKY57A.1FaWRKY57A.2 0.00000.0087NA
FaWRKY57D.1FaWRKY57D.2 0.00220.0000NA
FaWRKY58A.1FaWRKY58A.2 0.00640.01110.5766
FaWRKY58D.1FaWRKY58D.2 0.00120.0043NA
FaWRKY63A.1FaWRKY63A.2 0.00240.0000NA
FaWRKY64A.1FaWRKY64A.2 0.00260.0063NA
Segmental, non-syntenic
FaWRKY2D.1FaWRKY2D.2 0.01310.01530.8562
FaWRKY2C.1FaWRKY2C.2 0.00000.0000NA
FaWRKY2C.1FaWRKY2C.3 0.01240.01470.8482
FaWRKY2C.2FaWRKY2C.3 0.01240.01470.8482
FaWRKY2B.1FaWRKY2B.2 0.01340.01530.8758
FaWRKY2B.1FaWRKY2B.3 0.01250.01450.8621
FaWRKY2B.2FaWRKY2B.3 0.01330.01031.2913
FaWRKY14C.1FaWRKY14C.2 0.00710.02790.2545
FaWRKY17C.1FaWRKY17C.2 0.00210.0084NA
FaWRKY18C.1FaWRKY18C.2 0.00720.0069NA
FaWRKY26B.1FaWRKY26B.2 0.34250.56310.6082
FaWRKY29D.1FaWRKY29D.2 0.00110.0000NA
FaWRKY35D.1FaWRKY35D.2 0.00240.01010.2376
FaWRKY36D.1FaWRKY36D.2 0.01900.03220.5901
FaWRKY45D.1FaWRKY45D.2 0.00000.0000NA
FaWRKY47C.1FaWRKY47C.2 0.13980.21130.6616
FaWRKY51A.1FaWRKY51A.2 0.04000.04570.8753
FaWRKY51A.1FaWRKY51A.3 0.09650.10910.8845
FaWRKY51A.2FaWRKY51A.3 0.06300.05861.0751
FaWRKY62B.1FaWRKY62B.2 0.02980.02691.1078
FaWRKY63D.1FaWRKY63D.2 0.00220.0053NA
FaWRKY63B.1FaWRKY63B.2 0.01470.01321.1136
FaWRKY64D.1FaWRKY64D.2 0.00250.0032NA
Tandem duplications
FaWRKY11CFaWRKY12C NANANA
FaWRKY11AFaWRKY12A NANANA
FaWRKY55B.1FaWRKY55B.2 NANANA
FaWRKY38AFaWRKY39AFaWRKY41ANANANA
FaWRKY38BFaWRKY42B NANANA
FaWRKY38CFaWRKY42C NANANA
FaWRKY40DFaWRKY41DFaWRKY42DNANANA
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Garrido-Gala, J.; Higuera, J.-J.; Rodríguez-Franco, A.; Muñoz-Blanco, J.; Amil-Ruiz, F.; Caballero, J.L. A Comprehensive Study of the WRKY Transcription Factor Family in Strawberry. Plants 2022, 11, 1585. https://doi.org/10.3390/plants11121585

AMA Style

Garrido-Gala J, Higuera J-J, Rodríguez-Franco A, Muñoz-Blanco J, Amil-Ruiz F, Caballero JL. A Comprehensive Study of the WRKY Transcription Factor Family in Strawberry. Plants. 2022; 11(12):1585. https://doi.org/10.3390/plants11121585

Chicago/Turabian Style

Garrido-Gala, José, José-Javier Higuera, Antonio Rodríguez-Franco, Juan Muñoz-Blanco, Francisco Amil-Ruiz, and José L. Caballero. 2022. "A Comprehensive Study of the WRKY Transcription Factor Family in Strawberry" Plants 11, no. 12: 1585. https://doi.org/10.3390/plants11121585

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