Next Article in Journal
Synthesis and Crystal Structure of the Europium(II) Hydride Oxide Iodide Eu5H2O2I4 Showing Blue-Green Luminescence
Previous Article in Journal
Antimicrobial Peptidomimetics Prevent the Development of Resistance against Gentamicin and Ciprofloxacin in Staphylococcus and Pseudomonas Bacteria
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Large-Scale Identification and Characterization Analysis of VQ Family Genes in Plants, Especially Gymnosperms

1
Functional and Evolutionary Entomology, Gembloux Agro-Bio Tech, University of Liège, 5030 Gembloux, Belgium
2
Institute of Crop Sciences, Chinese Academy of Agricultural Sciences (CAAS), Beijing 100081, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(19), 14968; https://doi.org/10.3390/ijms241914968
Submission received: 14 August 2023 / Revised: 24 September 2023 / Accepted: 5 October 2023 / Published: 6 October 2023
(This article belongs to the Section Molecular Plant Sciences)

Abstract

:
VQ motif-containing (VQ) proteins are a class of transcription regulatory cofactors widely present in plants, playing crucial roles in growth and development, stress response, and defense. Although there have been some reports on the member identification and functional research of VQ genes in some plants, there is still a lack of large-scale identification and clear graphical presentation of their basic characterization information to help us to better understand this family. Especially in gymnosperms, the VQ family genes and their evolutionary relationships have not yet been reported. In this study, we systematically identified 2469 VQ genes from 56 plant species, including bryophytes, gymnosperms, and angiosperms, and analyzed their molecular and evolutionary features. We found that amino acids are only highly conserved in the VQ domain, while other positions are relatively variable; most VQ genes encode relatively small proteins and do not have introns. The GC content in Poaceae plants is the highest (up to 70%); these VQ proteins can be divided into nine subgroups. In particular, we analyzed the molecular characteristics, chromosome distribution, duplication events, and expression levels of VQ genes in three gymnosperms: Ginkgo biloba, Taxus chinensis, and Pinus tabuliformis. In gymnosperms, VQ genes are classified into 11 groups, with highly similar motifs in each group; most VQ proteins have less than 300 amino acids and are predicted to be located in nucleus. Tandem duplication is an important driving force for the expansion of the VQ gene family, and the evolutionary processes of most VQ genes and duplication events are relatively independent; some candidate VQ genes are preliminarily screened, and they are likely to be involved in plant growth and stress and defense responses. These results provide detailed information and powerful references for further understanding and utilizing the VQ family genes in various plants.

1. Introduction

VQ motif-containing (VQ) protein is known as a transcription regulatory cofactor for interacting with transcription factors (TFs) to regulate gene expression. In 2002, AtVQ23/AtSIB1 was first identified in Arabidopsis [1]. Subsequently, the VQ genes were successively identified via bioinformatics and experimental methods in various plants, such as rice, soybean, maize, grape, and wheat [2,3].
VQ proteins contain the highly conserved VQ motif (FxxxVQxhTG: x represents arbitrary amino acid, and h represents hydrophobic amino acid) [2,4]. The three terminal amino acids in the VQ motif may have different types in different plants. For example, six types (LTG, FTG, VTG, YTG, LTS, and LTD) were identified in Arabidopsis, four types (LTG, FTG, VTG, and ITG) were identified in rice, and five types (LTG, FTG, VTG, ITG, and VMA) were identified in wheat [3]. Most VQ genes in higher plants have no introns and encode relatively small proteins with less than 300 amino acids [2,4]. The lack of introns in these gene sequences makes transcription and translation more efficient, resulting in the production of these small proteins. Most VQ proteins are located in the nucleus, with a few found in the chloroplast and cytoplasm [5,6,7]. Moreover, it has been reported that VQ motif is related to the interaction with WRKY [2]. These basic characteristics provide a certain reference for us to understand and identify VQ family genes.
Jasmonic acid (JA), salicylic acid (SA), and abscisic acid (ABA) are important hormone signaling molecules in plants, which are widely involved in plant growth and development, disease resistance, and stress responses [8,9,10,11]. Research has shown that the expression of many VQ genes in different plants is induced or inhibited by JA, SA, or ABA hormones, as well as treatments such as pathogens, drought, and salt [3,5,6]. This finding indicates that the VQ gene plays an important role in complex signaling pathways in response to JA, SA, and ABA hormones and various stresses. In addition, WRKY, MAPK, and calmodulin (CaM) have been reported to play very important roles in plant life processes and responses to external stimuli [12,13,14,15,16]. By interacting with proteins such as CaM, WRKY, and MAPK, VQ proteins can mediate plant growth and development, as well as defense responses to biotic and abiotic stresses [2,5,17,18,19,20].
Studies have shown that VQ proteins are involved in growth and development. For example, in Arabidopsis, the loss-of-function mutants of AtVQ8 showed yellow–green leaves and delayed growth, while plants overexpressing AtVQ17, AtVQ18, or AtVQ22 showed a stunted phenotype with severely inhibited growth [5]. AtVQ14/IKU1 interacts with WRKY10/MINI3 to regulate the development of seed [21,22]. AtVQ20 interacts with WRKY2 and WRKY34 to regulate the expression of downstream MYB genes, thereby affecting male fertility [23]. AtVQ18 or AtVQ26 can interact with ABI5 and inhibit its transcriptional activation ability to negatively regulate ABA responses during seed germination and seedling establishment [24]. AtVQ29 interacts with PIF1 to enhance its transcriptional activation activity, regulating the elongation of hypocotyls under different spectra [25].
VQ proteins are involved in the responses to abiotic stresses. For example, AtVQ9 interacts with WRKY8 to regulate the expression of AtRD29A, negatively regulating plant salt tolerance [26]. AtVQ15 interacts with CaM, WRKY25, and WRKY51 to regulate plant tolerance to osmotic stress [5,17]. Bamboo PeVQ28 interacts with PeWRKY83 to promote the expression of ABA-related genes and positively regulates salt tolerance in plants [27]. Ectopic overexpression of tomato SlVQ6 in Arabidopsis showed decreased high-temperature tolerance [28]. The hypermorphic mutant of AtVQ10 exhibited enhanced meristem development, increased tolerance to oxidative stress, and reduced sensitivity to NO [29].
VQ proteins are involved in the response to biotic stresses. AtVQ4/MVQ1 is phosphorylated by MPK3/6, which, in turn, affects the interaction between VQ and WRKY to finely regulate the immune response in Arabidopsis [19]. Both AtVQ16/SIB2 and AtVQ23/SIB1 can bind to WRKY33 to regulate plant defense against Botrytis cinerea [5,30]. AtVQ21/MKS1 interacts with MPK4 and WRKY25/33 to regulate the expression of downstream genes. The overexpression of AtVQ21/MKS1 significantly increases resistance to Pseudomonas syringae by participating in the SA pathway [31], while reducing resistance to B. cinerea through the negative regulation of the JA pathway [32,33]. AtVQ22/JAV1 interacts with Ca2+/CaM, JUL1, JAZ8, and WRKY51 to jointly regulate JA synthesis, leading to rapid JA burst and activating plant defense [18,34]. OsVQ25 interacts with OsPUB73 and OsWRKY53 to balance the broad-spectrum disease resistance and growth of rice [35].
Some gymnosperms have high ornamental, medicinal, and economic value. They are also important materials for studying plant evolution. However, the research into the VQ family genes in gymnosperms has not yet been reported. Many gymnosperms have large genomes and numerous repetitive sequences, making it difficult to assemble a complete genome. But, in recent years, with the advancement in technology, some gymnosperms have gradually completed genome sequencing and assembly [36,37,38,39,40]. In view of the prominent role of VQ genes in the growth and development and the responses to environmental stress in angiosperms such as Arabidopsis and rice, VQ genes may also participate in various life processes in gymnosperms. Therefore, studying the morphology and evolution of VQ genes in gymnosperms will be very interesting and meaningful.
Although VQ genes have been structurally and functionally identified in some plants, there is a lack of systematic analysis and intuitive display to clearly show their characteristics in a wider range of plants. Here, in order to identify the detailed information on VQ family genes in the plant kingdom, we characterized a total of 2469 VQ genes from 56 plant species, including bryophytes, gymnosperms, and angiosperms. We carried out a comprehensive bioinformatics analysis, including conserved motifs, basic molecular characterization, and systemic clustering. Importantly, for gymnosperms, we selected Ginkgo biloba, Taxus chinensis, and Pinus tabuliformis as the research objects; identified their VQ gene members; and analyzed their molecular features, gene structures, subcellular locations, chromosome distributions, duplication events, expression levels, synteny blocks, and evolutionary comparisons. Our results provide fundamental information about the characterization and evolution of VQ genes in gymnosperms and angiosperms, which will be valuable for further research into their biological functions and working mechanisms.

2. Results

2.1. Members Identification and Conserved Motif Analysis of VQ Genes in Multiple Plants

The Hidden Markov Model (HMM) of the VQ motif (PF05678) was used to search the coding proteins for putative VQ genes in each species. After manually removing invalid entries and validating the search results, we systematically identified a total of 2469 VQ family genes from 56 plant species, including 3 bryophytes, 3 gymnosperms, and 50 angiosperms. Moreover, their basic information, such as VQ protein sequences (Table S1), chromosome ploidy, genome size, and numbers of total coding genes or VQ genes (Table S2), were summarized and listed for each species. We found that there is not any necessary relationship between genome size and VQ gene numbers (Figure 1a), consistent with previous studies [4,41].
We scanned 2469 VQ proteins and found the most conserved motif is FxxxVQxhTG, where three main terminal amino acids, namely LTG, FTG, and VTG (Figure 1b), are located. The core element ‘VQ’ in FxxxVQxhTG, is slightly changed in some plants, such as FxxxVHxhTG (Table S1), which is in agreement with a previous study [3]. Studies have shown that many VQ proteins contain single- or dual-component nuclear localization signals, and some also contain chloroplast targeting signals [5,6,7]. Moreover, some VQ proteins were reported as containing calmodulin (CaM) binding domains, such as AtVQ15 and AtVQ22 [17,18], or MAPK phosphorylation sites [19]. We predicted conservative motifs using MEME online tool. The top 20 conserved motifs were listed in Figure S1, and these motifs may be related to protein localization and protein interactions. We searched the database for these motifs using the Tomtom program within MEME online tool and did not find any clear functional annotations. Given that most VQ proteins were predicted to be localized in the nucleus, we speculated that there should be motifs associated with nuclear localization. It is known that proteins such as JAV1 and OsVQ25 are located in the nucleus [18,35]. Therefore, we used NLStradamus and PSORT online tools to predict their nuclear localization signals. We found that Motif 4 is directly related to the nuclear localization of these proteins (Figures S1 and S2). Interestingly, we also found a significant overlap between the CaM binding domain [17,18] and Motif 4, suggesting that Motif 4 may both guide nuclear localization and participate in interactions with CaM in these proteins.

2.2. Sequence Similarities, Length, Introns, and GC Content Analysis of VQ Genes in Multiple Plants

After performing multiple sequence alignment using MAFFT (v7.511) software on 2469 VQ proteins, all VQ proteins had an equal alignment length, and the proportion of identical amino acids at each aligned site was calculated. The highest ratio is the maximum sequence similarity of this matching point. This observation can help us to clearly see the overall sequence similarity of VQ family genes in plants. We found that the proportion of alignment sites with maximum sequence similarity ≤ 20% is 99.2%, and except for the VQ motif, the amino acid sequences at other positions are very variable (Figure 1c). Overall, the similarity of VQ proteins is low, but there are also some amino acid sites showing slightly higher similarity. Therefore, for the certain VQ gene, it is still possible to find relatively homologous genes in different species according to their sequence similarity.
We conducted the basic statistics regarding these 2469 VQ genes. Most VQ genes encode relatively small proteins, with 83% and 93% of them having less than 300 and 400 amino acids, respectively (Figure 1d), almost consistent with previous studies [2,4]. In species with multiple VQ genes, moss (Physcomitrella patens) keeps high proportion (more than 70%) of intron-containing VQ genes, while most VQ genes in higher plants do not have introns, no matter their angiosperms or gymnosperms (Figure 2). It has previously been reported that most VQ genes in moss (Physcomitrella patens) have introns [28], but interestingly, we inadvertently found that in different tea varieties (Camellia sinensis), such as ‘Tieguanyin’ (used in this study), ‘Longjing 43’, ’Shuchazao’, ‘Yunkang 10’, and ‘Biyun’, the ratios of intron-free VQ genes are mostly high, except for ‘Longjing 43’, whose proportion is actually as low as 15% (Table S3).
Furthermore, we calculated GC content in the coding region of the VQ gene. Interestingly, we found that the average GC content in the coding sequence of VQ genes among all species is greater than 40%; in commelinids, including Poales and Musa plants, GC content is more than 60%. In Poaceae plants, such as bamboo, sorghum, rice, maize, barley, and wheat, the GC content is up to 70%, which is much higher than the average level of the species (55%), making it a relatively special occurrence in plants (Figure 3). This finding suggests that VQ genes are more stable in Poaceae plants and may play more prominent roles in evolution and function.

2.3. Molecular Features and Gene Structure Analysis of VQ Genes in Gymnosperms

In total, 34, 18, and 64 VQ genes were identified in Ginkgo biloba, Taxus chinensis, and Pinus tabuliformis, respectively, designated as GbVQ1 to GbVQ34, TcVQ1 to TcVQ18, and PtVQ1 to PtVQ64 based on their chromosomal physical location (Table 1 and Table S4). In order to clearly see the distribution of VQ genes on chromosomes, MapChart (v2.3) software was used to produce the chromosome map on chromosomes according to the detailed location information of VQ genes in the GFF file. These VQ genes are not evenly distributed across different chromosomes, and on certain chromosomes, there is no VQ gene distribution (Figure 4). Among them, chromosome 3 and chromosome 10 have the largest number of VQ genes in Ginkgo biloba (n = 9, n = 9), chromosome 1 has the largest number of VQ genes in Taxus chinensis (n = 5), and chromosome 5 and chromosome 8 have the largest number of VQ genes in Pinus tabuliformis (n = 14, n = 18). In addition, we noticed that there are multiple VQ genes stacking within specific chromosomal regions, which may be due to the huge repetitive sequence or genome replication events of gymnosperms.
In gymnosperms, we analyzed the basic molecular features of VQ genes, and the detailed data for each VQ gene, including gene ID, chromosome position, protein length, motif type, isoelectric point, molecular weight, and subcellular localization, were shown in Table S4. The length of their encoded VQ protein ranges from 126 amino acids (GbVQ19) to 519 amino acids (GbVQ32), with an average of 283 amino acids, in Ginkgo biloba; ranges from 110 amino acids (TcVQ12) to 243 amino acids (TcVQ9), with an average of 165 amino acids, in Taxus chinensis; and ranges from 121 amino acids (PtVQ11) to 709 amino acids (PtVQ13), with an average 260 amino acids, in Pinus tabuliformis. The analysis of physiochemical properties further revealed that VQ proteins are widely varied in molecular weight (MW), ranging from 14.43 (GbVQ19) to 54.51 kDa (GbVQ32), with an average of 30.86 kDa, in Ginkgo biloba,; ranging from 12.26 (TcVQ12) to 54.51 kDa (TcVQ9), with an average of 18.37 kDa, in Taxus chinensis; and ranging from 13.08 (PtVQ11) to 76.10 kDa (PtVQ13), with an average of 28.09 kDa, in Pinus tabuliformis. The isoelectric point (pI) of these VQ proteins varies between 4.83 (GbVQ21) and 10.43 (GbVQ3) in Ginkgo biloba (average 7.67), 5.06 (TcVQ5) and 10.27 (TcVQ15) in Taxus chinensis (average 8.55), and 5.07 (PtVQ24) and 11.03 (PtVQ31) in Pinus tabuliformis (average 8.50) (Table 1 and Table S4).
In addition, we identified four VQ motif types (LTG, FTG, VTG, and YTG) in Ginkgo biloba, four VQ motif types (LTG, FTG, VTG, and FTA) in Taxus chinensis, and four VQ motif types (LTG, FTG, MTG, and VTG) in Pinus tabuliformis. Most VQ proteins belong to the LTG type (23/34), with eight FTG type (8/34), two VTG type (2/34), and one YTG type (1/34) found in Ginkgo biloba; many LTG type (13/18), three FTG type (3/18), one VTG type (1/18), and one FTA type (1/18) found in Taxus chinensis; and many LTG type (36/64), twenty-two FTG type (22/64), four MTG type (4/64), and two VTG type (2/64) found in Pinus tabuliformis (Table 1 and Table S4). We also observed slight changes in the core VQ motif from FxxxVQxhTG to FxxxVExhTG (PtVQ37, PtVQ38, PtVQ39) and FxxxVHxhTG (PtVQ28) in Pinus tabuliformis (Table S1). The subcellular localization analysis revealed that most VQ proteins were predicted to be located in the nucleus, while a few were predicted to be located in the cytoplasm/nucleus (Table 1 and Table S4).
Gene structure can provide more information about the evolutionary relationship within a gene family. We conducted a systematic clustering and gene structure analysis of the VQ genes in Ginkgo biloba, Taxus chinensis, and Pinus tabuliformis. We detected 41 conservative motifs using MEME online tool, and we can clearly see the differences between these VQ genes. Motif 1 was identified as the core motif that comprises the VQ domain, which was included in all VQ proteins. It is noteworthy that VQ genes with closer clustering relationships have almost similar conserved motifs, indicating that the phylogenetic classification is relatively reliable and proteins in the same group maybe perform similar functions (Figure 5a,b). We created an exon/intron structure map based on the location information of each VQ gene in the GFF file. The structure of VQ genes was analyzed, and we found 85%, 94%, and 96% of VQ genes to have no introns in Ginkgo biloba, Taxus chinensis, and Pinus tabuliformis, respectively (Figure 5c). These findings indicate that most VQ genes do not contain introns in gymnosperms, which is similar to the results of angiosperms [2].

2.4. Gene Duplication and Collinearity Analysis of VQ Genes in Gymnosperms

Gene duplication is an important event contributing to genome evolution, and it is also an important factor in the expansion of the gene family. It is primarily divided into tandem duplication, segmental duplication, dispersed duplication, and proximal duplication [42]. To better understand the evolutionary mechanisms of VQ genes in gymnosperms, the duplication events of the VQ family gene were evaluated. Following the BLAST and MCScanX results, among the 34 GbVQs, a total of 10 members and 8 gene pairs participated in duplication events in Ginkgo biloba; among the 18 TcVQs, a total of 3 members and 2 gene pairs participated in duplication events in Taxus chinensis, and among the 64 PtVQs, a total of 18 members and 33 gene pairs participated in duplication events in Pinus tabuliformis (Table 2). Ginkgo biloba exhibited five tandem duplication events (5/8), two proximal duplication events (2/8), and one segmental duplication events (1/8); Taxus chinensis exhibited two proximal duplication events (2/2); and Pinus tabuliformis exhibited fourteen dispersed duplication events (14/33), ten tandem duplication events (10/33), and nine proximal duplication events (9/33). In these three gymnosperms, there is an interesting phenomenon: although the distance between two or more VQ gene pairs is a little bigger than 200 kb (here marked as proximal duplication), they are continuously arranged coding genes in one genome block; therefore, this type of gene pair may also be named as tandem duplication (Figure 4 and Table 2). These results suggest that tandem duplication is important to expand the VQ family gene in gymnosperms. While whole-genome duplication (WGD) played a critical role in adaptive evolution in angiosperms [43], few recent WGD events were found in extant gymnosperms [36,37,38,39,40,44], which indicates that these duplicate VQ gene pairs were evolved from independent duplication events or derived from older ancestors.
The selective evolutionary pressure on all VQ gene pairs was investigated by calculating the Ka, Ks, and Ka/Ks ratios of the duplication events. The Ka/Ks values of most duplicated gene pairs (5/8) are less than 1.0, those of two gene pairs (2/8) are slightly greater than 1.0, and that of one gene pair (3/8) (segmental duplication events) is unable to calculate a valid value using KaKs_Calculator 2.0 software in Ginkgo biloba. One duplicated gene pair (1/2) is slightly greater 1.0, and one gene pair (1/2) is unable to calculate a valid value in Taxus chinensis; most duplicated gene pairs (21/33) are less than 1.0, four gene pairs (4/33) are more than 1.0, one gene pair (1/33) is unable to calculate a valid value, and seven duplicated gene pairs (7/33) have Ka = Ks = 0, which means these two genes have no difference in their coding region in Pinus tabuliformis (Table 2). The Ka/Ks values of the majority of VQ gene pairs in gymnosperms are less than 1.0, indicating they have mainly undergone purifying selection pressures during evolution process. The Ka/Ks values of some duplicated gene pairs are greater than 1.0, which shows the presence of positive selection pressure.
To gain insight into the evolution of VQ genes in gymnosperms, the collinear blocks were searched in the chromosomes using MCScanX software. Ginkgo biloba and Pinus tabulaeformis have the most collinear blocks at the whole genome level, followed by Ginkgo biloba and Taxus chinensis and Taxus chinensis and Pinus tabuliformis. This finding indicates that compared to Taxus chinensis, Ginkgo biloba has a higher similarity and closer evolutionary relationship with Pinus tabuliformis. Similarly, the VQ gene pairs in these collinear blocks are the most between Ginkgo biloba and Pinus tabuliformis (n = 12), followed by Ginkgo biloba and Taxus chinensis (n = 5); however, Taxus chinensis and Pinus tabuliformis do not have orthologous VQ gene pairs (n = 0) in fewer collinear blocks (Figure 6). Among all VQ gene pairs mentioned above (n = 17), only one VQ gene (GbVQ29, TcVQ13, and PtVQ38) was shared by these three gymnosperms. In order to better see the conservatism of these VQ genes in different gymnosperms, we introduced the collinearity comparison with another two gymnosperms: Cycas panzhihuaensis [38] and Metasequoia glyptostroboides [40]. We found that this VQ gene and its related collinear block also existed in Metasequoia glyptostroboides and Cycas panzhihuaensis. Moreover, in all collinear blocks, we found five other VQ genes (GbVQ2, GbVQ13, GbVQ17, GbVQ20, and GbVQ34), most of which contain collinear VQ genes in these five gymnosperms (Figure 6). This finding indicates that the genome blocks containing these six VQ genes may have been preserved from ancestors of these gymnosperms. Additionally, through intragenomic collinearity analysis, we found that there are relatively more collinear blocks in Ginkgo biloba, including one VQ gene pair, which belongs to the segmental duplication events. Meanwhile, there are very few collinear blocks in the other two gymnosperms, and no VQ gene pairs were found (Figure S3 and Table 2). All these results suggest that most VQ genes and their duplication events in every gymnosperm mentioned in this study may have evolved independently in the latter stages. Furthermore, almost no collinear blocks and VQ gene pairs were observed between these three gymnosperms and Arabidopsis or rice.

2.5. Expression Patterns of VQ Genes in Different Tissues of Gymnosperms

To characterize the expression patterns of VQ genes, the expression levels were analyzed in different tissues (eight for Ginkgo biloba, four for Taxus chinensis, and four for Pinus tabulaeformis) based on the published data in previous research to gain preliminary insight into their potential functions (Figure 7 and Table S5). Based on the TPM values, we found that 12 VQ genes (GbVQ3, GbVQ10, GbVQ11, GbVQ12, GbVQ13, GbVQ26, GbVQ27, GbVQ29, GbVQ30, GbVQ32, GbVQ33, and GbVQ34), 6 VQ genes (TcVQ2, TcVQ3, TcVQ4, TcVQ9, TcVQ13, and TcVQ14), and 15 VQ genes (PtVQ1, PtVQ9, PtVQ14, PtVQ15, PtVQ29, PtVQ30, PtVQ34, PtVQ37, PtVQ38, PtVQ56, PtVQ57, PtVQ58, PtVQ59, PtVQ60, and PtVQ64) were broadly and prominently expressed in different tissues in Ginkgo biloba, Taxus chinensis and Pinus tabulaeformis, respectively (Figure 7). In contrast, some VQ genes were only expressed in a few tissues, and some even had low or zero expression levels in different tissues (Figure 7 and Tables S6–S8). These VQ genes are denoted as playing different roles in regulating the plant growth and development of gymnosperms. These genes, which are expressed in multiple tissues, often play important roles in plant metabolism, disease resistance, and stress resistance [3,45,46], and they deserve further attention.

2.6. Phylogenetic Analysis of VQ Genes

To detect the evolutionary relationships and classification of the VQ family genes in 56 plant species, circular and unrooted phylogenetic trees were constructed with 2469 VQ proteins. In previous studies, according to structural characteristics, VQ proteins from different species have been clustered into 7 groups [6], 8 groups [3], 9 groups [47], and 10 groups [19]. In this study, we also built the phylogenetic tree for these 2469 VQ proteins from 56 plant species to explore their phylogenetic relationships, in which these proteins could be divided into nine groups (Figure 8 and Figure S4). VQ genes with unknown functions and pathways can often be inferred from species with close genetic relationships. Therefore, we also constructed a species evolution tree based on the taxonomy database of NCBI (Figure S5). The systemic clustering and biological classification of these 56 plant species can help us to better understand the functions and evolution of VQ genes in different plants. Moreover, VQ genes were clustered into 11 groups in three gymnosperms, which is basically consistent with the results of conservative motifs (Figure 5).
To further explore the evolutionary relationship of VQ gene between gymnosperms and angiosperms, as well as between monocotyledons and dicotyledons, nine species, including three gymnosperms (ginkgo, taxus, and Chinese pine) and six angiosperms (monocot: rice, wheat, and maize; dicotyledon: Arabidopsis, soybean, and tomato), were selected for further evolutionary analysis. The phylogenetic tree of VQ proteins from these species was drawn, and it showed that some branches are unique to gymnosperms, some are unique to angiosperms, and some have intersections, indicating that some VQ proteins may appear after the differentiation of angiosperms and gymnosperms (Figure S6). Similarly, monocotyledonous and dicotyledonous plants have their own independent and intersecting branches (Figure S6), which is consistent with previous research [3]. Interestingly, among widely expressed VQ genes, we found that four genes (GbVQ3, GbVQ12, GbVQ33, and GbVQ34) in Ginkgo biloba, one gene (TcVQ2) in Taxus chinensis, and three genes (PtVQ29, PtVQ30, and PtVQ58) in Pinus tabulaeformis were clustered with AtVQ14, AtVQ9, and AtVQ5 (Figure S6), which were reported as influencing seed development, mediating salinity stress responses, and regulating plant defense, respectively [5,22,26]; three genes (GbVQ26, GbVQ27, and GbVQ29) found in Ginkgo biloba, three genes (TcVQ13, TcVQ14, and TcVQ15) in Taxus chinensis, and three genes (PtVQ37, PtVQ38, and PtVQ39) in Pinus tabulaeformis were clustered with AtVQ22, AtVQ27, and AtVQ28 (Figure S6), which are related to plant defense and growth [18,48,49]. Among them, GbVQ29, TcVQ13, and PtVQ38 are the only VQ genes in the collinear block shared by these gymnosperms, which indicates that these genes have a very conservative, important role and status in seed plants. Additionally, GbVQ10, TcVQ3, PtVQ9, PtVQ59, and PtVQ60 were clustered with AtVQ4, which regulates disease resistance [19]; GbVQ30, TcVQ9, PtVQ14, PtVQ15, PtVQ34, and PtVQ64 were clustered with AtVQ15, which negatively regulates osmotic stress tolerance [5,17] (Figure S6). These clusters can serve as references for the potential functions of these VQ genes to some degree.
The different characteristics of VQ genes in different types of plants can provide a molecular perspective for us to better understand plant evolution. All of these results provide important references for the study and utilization of VQ genes in plants, as well as for the evolutionary comparisons between gymnosperms and angiosperms.

3. Discussion

Plants are often affected by various environmental stresses throughout their lifetimes [50]. VQ family genes have been proven to play important roles in growth and development and responses to various abiotic and biotic stresses [2,5]. Therefore, it is of great significance to study the characteristics and functions of VQ genes in a wide range of plants. Up to now, VQ family genes have been identified and analyzed in multiple plant species. However, although there have been some reports on the structure and function of VQ genes, clear and intuitive feature information and knowledge of how to utilize these features are still limited in the plant kingdom, including bryophytes, gymnosperms, and angiosperms. Importantly, the VQ gene family of gymnosperms has not yet been reported. Thus, a comprehensive bioinformatics analysis of VQ genes in multiple plants, especially in gymnosperms, can provide an overall basis for evolutionary and functional studies of VQ genes.

3.1. Molecular Characteristics and Phylogeny of VQ Family Genes

In this study, we systematically identified 2469 VQ genes from 56 plant species. Gene numbers of VQ family are various among different plants. Compared to other plants, plants such as wheat and soybean contain abundant VQ genes (Table S2), which have undergone significant polyploidization and WGD events, respectively [3,51]. In Ginkgo biloba, Taxus chinensis, and Pinus tabulaeformis, 34, 18, and 64 VQ gene members were detected, respectively, but compared to other species, such as Arabidopsis, with 34 members [5], or rice, with 40 members [52], the VQ gene numbers in these three gymnosperms are far lower than expected considering their big genome size. Therefore, it can be concluded that the number of VQ genes has no necessary connection to the genome size, but it is related to species polyploidization and whole-genome replication events.
Except for the conserved VQ domain, VQ protein exhibits significant variability in other regions (Figure 1c). However, through MEME domain scanning, we also found some other motifs in different subgroups (Figure 5 and Figure S1), which may be related to the interaction, modification, and subcellular localization of these VQ proteins, affecting protein function. Although we have enriched multiple motifs in these VQ proteins, their specific functions still need further analysis. Many VQ family genes play important roles in growth and development, as well as biotic and abiotic stress, and interact with various proteins such as MAPK and WRKY. By conducting targeted editing and modification of these enriched candidate loci on the VQ gene, the functions and mechanisms of these motifs can be further revealed. Moreover, given the extensive biological functions of the VQ gene, this also has great potential in plant molecular precision breeding. Our study provides a method and idea for searching and utilizing conserved motifs and loci in gene families. The average length of VQ proteins is fewer than 300 amino acid residues (Figure 1d). This is one of the characteristics of the VQ family genes in plants, which helps us to identify them. Notably, some VQ proteins were found to be longer than 600 amino acids (Figure 1d), and these VQ proteins may contain other domains and participate in more diverse regulatory pathways.
Although the average GC content in the coding region of the VQ gene varies among different plants, the GC content in commelinids plants, including Poales and Musa plants, has significantly increased and is most prominent in Poaceae (GC content > 70%) (Figure 3). The VQ gene, as an ancient transcription regulatory cofactor in plants, plays an important role in the interaction between plants and environment. High GC content means better alkaline and high-temperature tolerance, allowing these VQ genes to remain relatively stable in environmental changes. Poaceae plants are the main food source for humans and many animals, and high GC content of their VQ genes is likely possessed by their ancestors and has been selected and fixed during evolution, which has become one of the characteristics of these plants. As for the deeper reasons why the GC content of VQ genes in Poaceae plants is prominent, further research is needed, and this is a very interesting and meaningful topic.
In the higher eukaryotes, intron-free genes are very common in their genomes [53]. In our study, based on the gene structure analysis, we found that most VQ genes in higher plants are intronless, no matter whether they are in angiosperms or gymnosperms, which is consistent with previous studies, including those of Arabidopsis [5], rice [6], tomato [28], apple [47], and wheat [3]. In contrast, lower plant moss is exactly the opposite [28] (Figure 2). Therefore, many studies speculated that the VQ genes in higher plants lost its intron during evolution [2,3,41]. However, we found that most VQ genes in tea variety ‘Longjing 43’ also contain introns and, at the same time, Marchantia polymorpha, known as a lower bryophyte plant, although there are only seven VQ genes in its genome, six of them (6/7) have no intron. So, the explanation of intron loss from the perspective of evolution history may need further investigation. Moreover, our results revealed that intron-containing VQ genes of gymnosperms are located in different subgroups, suggesting that these introns appear relatively independent, and this is also common in angiosperms [3,41]. Taken together, the identification of gene structure of VQ genes enriched our understanding of the evolution of introns in the plant kingdom.
Phylogenetic trees represent the genetic relationships between gene families from different species and reflect the similarity of protein-coding genes. To further understand the evolutionary relationships between these 2469 VQ genes from 56 plant species, we established a phylogenetic tree based on their protein sequences. They were classified into nine groups from our phylogenetic analysis (Figure 8 and Figure S4) and showed obvious evolutionary characteristics, such as the differentiation between angiosperms and gymnosperms and monocotyledons and dicotyledons. Additionally, in three gymnosperms, the VQ genes were classified into 11 groups, and in each group, they harbored similar type of motifs (Figure 5), suggesting a potential functional similarity. These results highlight the conservatism and diversity among VQ gene families of different plants. Based on the phylogenetic tree results, we can also use VQ genes with known functions to quickly search for homologous genes in specific plants. In theory, gene structure determines its function, and the more similar motifs shared between VQ proteins, the higher probability of their functional similarity. Therefore, systematic clustering combined with motifs analysis will help us to quickly identify homologous genes in different plants and make preliminary judgments on their functions. At the same time, further experiments are needed to verify their detailed functionality.

3.2. Expansion and Duplication Mechanism of VQ Gene Family in Gymnosperms

Genome replication events play an important role in expanding the size of the genome [54] and diversifying gene functions [55]. Chromosome fragment replication is considered to be the main expansion mechanism of gene family [56], and, thus, the evolutionary process can explain the number of specific VQ genes in a species, not the genome size. Previous research indicated that segmental duplication is the major mechanism contributing to the expansion of the VQ gene family in many angiosperms [3,4,41]. In the present study, we found that both tandem duplication and segmental duplication events of VQ genes exist simultaneously in Ginkgo biloba; only proximal duplication events exist in Taxus chinensis. Dispersed duplication, tandem duplication, and proximal duplication events exist in Pinus tabulaeformis; however, segmental duplication events only appear in Ginkgo biloba and account for a small proportion (Table 2). Segmental duplication is an important way of expanding the VQ gene family in angiosperms, while tandem duplication dominates the expansion of the VQ gene family in gymnosperms, which may be caused by different genetic and evolutionary mechanisms between angiosperms and gymnosperms. In addition, tandem duplication can reduce genetic instability caused by single-gene mutations or deletions and enhance plant resistance to environmental stresses, such as stress resistance and disease resistance. Given the importance of VQ genes, this is of great significance for ensuring the normal lives of gymnosperms.
The Ka, Ks, and Ka/Ks ratios of all VQ gene pairs were calculated to investigate whether VQ genes underwent selection pressure. The Ka/Ks values of most VQ gene pairs are less than 1.0, which demonstrates that purifying selection (Ka/Ks < 1) plays an important role in the evolution of VQ gene family in gymnosperms. Moreover, positive selection (Ka/Ks > 1) existed in some VQ genes (Table 2). Most VQ gene pairs are undergoing purification selection, while a few are undergoing positive selection, indicating that the VQ gene family in gymnosperms is in a stable dynamic evolution process.
Furthermore, only a few VQ gene pairs in collinear blocks were detected within and between Ginkgo biloba, Taxus chinensis, and Pinus tabulaeformis. For Taxus chinensis or Pinus tabulaeformis, there were even no VQ gene pairs detected in very few collinear blocks (Figure 6 and Figure S3). The results of synteny analysis indicate that the conservation degree of VQ genes among these three gymnosperms is low and their evolution process is relatively independent, which is similar to the difference between dicotyledons and monocotyledons in angiosperms [41], and this finding may be related to the fact that these three gymnosperms belong to different phytoclasses.

3.3. Expression Patterns of VQ Members in Gymnosperms

Previous studies have demonstrated that VQ genes are involved in regulating plant responses to biotic stresses, abiotic stresses, and growth and development [2,5]. The expression of many VQ genes showed various levels between different tissues and significant changes under pathogen, stress, or hormone treatments [3,5,6]. In this study, we detected the expression level of gymnosperm VQ genes in different tissues by analyzing RNA-seq data. The results showed that some VQ genes are not expressed in any tissues, some are only expressed in certain tissues, and some are widely expressed in different tissues. These widely expressed genes are usually associated with growth and development, hormone response, and plant defense [3,45,46], which requires further experimental verification in gymnosperms.
In Arabidopsis, AtVQ14 is mainly associated with seed development, and its mutation produces small seeds [21,22]. Transgenic plants overexpressing AtVQ5 displayed increased susceptibility to B. cinerea [5]. AtVQ9 is strongly induced via NaCl treatment and negatively regulates the resistance to NaCl stress [26]. Compared to wild type plants, overexpression of AtVQ17, AtVQ18, or AtVQ22 causes highly stunted growth of the transgenic plans [5]. AtVQ22/JAV1 can regulate JA-mediated plant defense and coordinate growth and defense [18]. Among these widely expressed VQ genes in gymnosperms, we found eight VQ genes (GbVQ3, GbVQ12, GbVQ33, GbVQ34, TcVQ2, PtVQ29, PtVQ30, and PtVQ58) as the candidate homologs of AtVQ14/AtVQ9/AtVQ5, nine VQ genes (GbVQ26, GbVQ27, GbVQ29, TcVQ13, TcVQ14, TcVQ15, PtVQ37, PtVQ38, and PtVQ39) as candidate homologs of AtVQ22/AtVQ27/AtVQ28, five VQ genes (GbVQ10, TcVQ3, PtVQ9, PtVQ59, and PtVQ60) as candidate homologs of AtVQ4, and six VQ genes (GbVQ30, TcVQ9, PtVQ14, PtVQ15, PtVQ34, and PtVQ64) as candidate homologs of AtVQ15 (Figure 7 and Figure S6). This finding suggests that these candidate VQ genes may have similar functions and play an important role in growth, development, and response to external environmental stimuli. Among these genes, we found that TcVQ14 is almost identical to TcJAV3, which was reported to be a VQ motif-containing protein and homologous to AtVQ22/JAV1. TcJAV3-TcWRKY26 complex can regulate the expression of the downstream paclitaxel biosynthesis gene DBAT and participate in JA-mediated plant defense [57]. This finding indicates that our research results are trustworthy. In addition, gene duplication can produce gene function redundancy, and most of these repeated or collinear VQ genes showed almost identical expression patterns (Figure 7 and Table 2). Taken together, our study suggested that some VQ genes are involved in growth and development and participate in multiple life processes of gymnosperms.

4. Materials and Methods

4.1. Identification of VQ Genes in Multiple Plants

In the present study, genome sequence and annotation files of the 56 plant species were mainly collected from the Ensembl Plant database (http://plants.ensembl.org/index.html/, accessed on 2 March 2023) [58], NCBI plant genome database (ftp://ftp.ncbi.nlm.nih.gov/genomes/genbank/plant/, accessed on 2 March 2023), phytozome database (https://phytozome-next.jgi.doe.gov/, accessed on 2 March 2023) [59], and National Genomics Data Center (https://ngdc.cncb.ac.cn/, accessed on 2 March 2023). The Hidden Markov Model (HMM) profile of the VQ domain (PF05678) was obtained from the Pfam database (http://pfam.xfam.org/, accessed on 20 March 2023) [60]. The VQ family members were retrieved from plant protein sequences using the HMMSEARCH program of the HMMER (v3.0) software [61]. The online program SMART tool (http://smart.embl-heidelberg.de/, accessed on 25 March 2023) [62], the NCBI Conserved Domains Database (https://www.ncbi.nlm.nih.gov/Structure/cdd/wrpsb.cgi/, accessed on 25 March 2023) [63], and NCBI-BLAST (https://blast.ncbi.nlm.nih.gov/Blast.cgi/, accessed on 25 March 2023) were used to ultimately determine the VQ genes. Default parameters were adopted for all of the software mentioned above.

4.2. Molecular Features and Chromosomal Localization Analysis

The genome information, such as genome size, chromosome ploidy, and the number of total coding genes, were obtained from the Ensembl Plant database (http://plants.ensembl.org/index.html/, accessed on 2 March 2023) and other related published papers. GC content in the coding region, protein length, and sequence similarities of aligned sites of VQ genes were calculated via the Perl program (https://www.perl.org/, accessed on 2 March 2023). The relevant scripts and their input data can be downloaded from https://github.com/ywxkjtsd123/code (accessed on 15 September 2023). After the alignment of 2469 VQ protein sequences, the largest proportion of the same amino acids in each alignment site was calculated. The biophysical properties of the VQ proteins, including the peptide length, isoelectric point (pI), and molecular weight (MW), were estimated using the online program ExPasy (https://web.expasy.org/protparam/, accessed on 7 April 2023) [64]. The physical locations of the VQ genes on the chromosomes were visualized using the MapChart (v2.3) software [65]. The subcellular localization prediction tool DeepLoc-2.0 (https://services.healthtech.dtu.dk/services/DeepLoc-2.0/, accessed on 7 April 2023) [66] was used to predict the likely location of the VQ genes. The online tools NLStradamus (http://www.moseslab.csb.utoronto.ca/NLStradamus/, accessed on 15 September 2023) [67] and PSORT (https://www.genscript.com/psort.html/, accessed on 15 September 2023) [68] were used to predict the nuclear localization signals of proteins.

4.3. Conserved Motifs, Gene Structure, and Phylogenetic Tree Analysis

The conserved motifs of the VQ proteins were detected using the online program MEME (https://meme-suite.org/, accessed on 15 April 2023) with default parameters [69]. Conserved motifs were drawn using the TBtools (v1.123) software [70]. The exon and intron structures were determined via GFF file. After extracting their location information from the annotation file, the gene structures of the VQ genes were visualized using TBtools (v1.123) software. Based on the protein sequences, all multiple sequence alignments were carried out using the MAFFT (v7.511) software [71], using default parameters to study the evolutionary relationships and classification of the VQ genes. Depending on the alignment results, the phylogenetic tree was built using the FastTree 2 software [72] with default parameters. The online tool iTOL (https://itol.embl.de/, accessed on 15 April 2023) [73] was used to draw and adjust the phylogenetic tree. TBtools (v1.123) software was used to integrate phylogenetic trees, conserved motifs results, and gene structure results in three gymnosperms.

4.4. Gene Duplication and Collinearity Analysis

Gene duplication mainly includes segmental duplication, tandem duplication, dispersed duplication, and proximal duplication [42]. Segmental duplicates exist in collinear blocks. Tandem duplicates are defined as closely adjacent to each other on the same chromosome. Proximal duplicates are found on the same chromosome and are close to each other, but they are separated by several other genes. Dispersed duplicates occur on the same or different chromosomes, and they are neither close to each other nor within conserved collinearity blocks [42]. For the definition of tandem duplicates, we added the following concepts based on the literature: two or more VQ genes adjacent to each other within 200 kb can be defined as tandem duplication events [41]. To identify duplication events of the VQ gene family in the gymnosperms, the coding sequences (CDS) of all VQ genes were aligned using BLASTN with an E-value below 1 × 10−15. For the VQ gene with different isoforms, we selected the longest one for analysis. We used the following criteria to search for duplicate VQ gene pairs: both identity and coverage > 75% at the nucleotide level [3,74]. In addition, BLASTP (E-value < 1 × 10−5, top 5 matches) was used for sequence alignment, MCScanX [42] with the default parameters was used to identify collinear blocks within or between species, and the TBtools (v1.123) software and Circos program [75] were used to visualize the collinearity maps and exhibit segmentally duplicated VQ gene pairs. The values of nonsynonymous substitution rate (Ka) and synonymous substitution rate (Ks) of duplicated VQ gene pairs were calculated to evaluate the selection pressure using the KaKs_Calculator 2.0 [76] via the NG method. The relevant scripts and their input data can be obtained from https://github.com/ywxkjtsd123/code (accessed on 15 September 2023).

4.5. Gene Expression Patterns Analysis

The RNA-seq data of VQ genes in different tissues were obtained from previous research [36,37,77,78,79]. The sample name, accession number, and data source were listed in Table S5. Fastqc (https://www.bioinformatics.babraham.ac.uk/projects/fastqc/, accessed on 20 April 2023) was used for quality control, and Trimmomatic (v0.39) [80] was used for data filtering to obtain the clean reads. HISAT2 [81] was used for genome library construction and alignment. Transcript abundance was measured using TPM values. TPM values were calculated via StringTie2 [82] for Ginkgo biloba and Taxus chinensis. Due to the larger genome size and the limitations in computing resources, Bowtie2 [83] was used to build the index library and conduct sequence alignment, and RSEM (v1.3.3) [84] was used to obtain the TPM values for Pinus tabuliformis. A heatmap was generated using log2 (TPM + 1) values via TBtools (v1.123) software. Default parameters were used for all of the software used.

5. Conclusions

In this study, a systematic analysis was performed on the genome-wide identification, molecular characterization, phylogenetic relationship, and expression patterns of VQ genes in multiple plants, especially in gymnosperms. By analyzing the phylogenetic tree, conserved motifs, and duplication events, we have explained the functional divergences and expansion patterns of VQ genes. In conclusion, this study provides the first comprehensive and systematic analysis of the 2469 VQ genes identified in 56 plant species, and we also provided clear data that support the identification and evolution of the VQ gene in gymnosperms. The selection of candidate VQ genes could also provide a reference for future investigations in gymnosperms. Taken together, these findings can help us to better understand the evolution of VQ genes and provide a theoretical basis for further functional research and the practical utilization of VQ gene in various plants.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms241914968/s1.

Author Contributions

F.F. and J.T. designed this study. J.T. and J.Z. performed the investigations, and J.T. analyzed the data. J.T. wrote the manuscript. F.F. and J.T. revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the China Scholarship Council (grant number 202103250061). J.T. was supported by the GSCAAS-ULg Joint PhD Program.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in the study are available in the article and the Supplementary Materials. For further inquiries, you can directly contact the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Morikawa, K.; Shiina, T.; Murakami, S.; Toyoshima, Y. Novel nuclear-encoded proteins interacting with a plastid sigma factor, Sig1, in Arabidopsis thaliana. FEBS Lett. 2002, 514, 300–304. [Google Scholar] [CrossRef]
  2. Yuan, G.; Qian, Y.; Ren, Y.; Guan, Y.; Wu, X.; Ge, C.; Ding, H. The role of plant-specific VQ motif-containing proteins: An ever-thickening plot. Plant Physiol. Biochem. 2021, 159, 12–16. [Google Scholar] [CrossRef]
  3. Zhang, L.; Wang, K.; Han, Y.; Yan, L.; Zheng, Y.; Bi, Z.; Zhang, X.; Zhang, X.; Min, D. Genome-wide analysis of the VQ motif-containing gene family and expression profiles during phytohormones and abiotic stresses in wheat (Triticum aestivum L.). BMC Genom. 2022, 23, 292. [Google Scholar] [CrossRef]
  4. Jiang, S.Y.; Sevugan, M.; Ramachandran, S. Valine-glutamine (VQ) motif coding genes are ancient and non-plant-specific with comprehensive expression regulation by various biotic and abiotic stresses. BMC Genom. 2018, 19, 342. [Google Scholar] [CrossRef]
  5. Cheng, Y.; Zhou, Y.; Yang, Y.; Chi, Y.J.; Zhou, J.; Chen, J.Y.; Wang, F.; Fan, B.; Shi, K.; Zhou, Y.H.; et al. Structural and functional analysis of VQ motif-containing proteins in Arabidopsis as interacting proteins of WRKY transcription factors. Plant Physiol. 2012, 159, 810–825. [Google Scholar] [CrossRef]
  6. Kim, D.Y.; Kwon, S.I.; Choi, C.; Lee, H.; Ahn, I.; Park, S.R.; Bae, S.C.; Lee, S.C.; Hwang, D.J. Expression analysis of rice VQ genes in response to biotic and abiotic stresses. Gene 2013, 529, 208–214. [Google Scholar] [CrossRef]
  7. Garrido-Gala, J.; Higuera, J.J.; Muñoz-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]
  8. Bari, R.; Jones, J.D. Role of plant hormones in plant defence responses. Plant Mol. Biol. 2009, 69, 473–488. [Google Scholar] [CrossRef]
  9. Glazebrook, J.; Chen, W.; Estes, B.; Chang, H.S.; Nawrath, C.; Métraux, J.P.; Zhu, T.; Katagiri, F. Topology of the network integrating salicylate and jasmonate signal transduction derived from global expression phenotyping. Plant J. 2003, 34, 217–228. [Google Scholar] [CrossRef]
  10. Tuteja, N. Abscisic Acid and abiotic stress signaling. Plant Signal. Behav. 2007, 2, 135–138. [Google Scholar] [CrossRef]
  11. Roychoudhury, A.; Paul, S.; Basu, S. Cross-talk between abscisic acid-dependent and abscisic acid-independent pathways during abiotic stress. Plant Cell Rep. 2013, 32, 985–1006. [Google Scholar] [CrossRef] [PubMed]
  12. Khoso, M.A.; Hussain, A.; Ritonga, F.N.; Ali, Q.; Channa, M.M.; Alshegaihi, R.M.; Meng, Q.; Ali, M.; Zaman, W.; Brohi, R.D.; et al. WRKY transcription factors (TFs): Molecular switches to regulate drought, temperature, and salinity stresses in plants. Front. Plant Sci. 2022, 13, 1039329. [Google Scholar] [CrossRef] [PubMed]
  13. Wani, S.H.; Anand, S.; Singh, B.; Bohra, A.; Joshi, R. WRKY transcription factors and plant defense responses: Latest discoveries and future prospects. Plant Cell Rep. 2021, 40, 1071–1085. [Google Scholar] [CrossRef] [PubMed]
  14. Pitzschke, A.; Schikora, A.; Hirt, H. MAPK cascade signalling networks in plant defence. Curr. Opin. Plant Biol. 2009, 12, 421–426. [Google Scholar] [CrossRef] [PubMed]
  15. Taj, G.; Agarwal, P.; Grant, M.; Kumar, A. MAPK machinery in plants: Recognition and response to different stresses through multiple signal transduction pathways. Plant Signal. Behav. 2010, 5, 1370–1378. [Google Scholar] [CrossRef] [PubMed]
  16. Perochon, A.; Aldon, D.; Galaud, J.P.; Ranty, B. Calmodulin and calmodulin-like proteins in plant calcium signaling. Biochimie 2011, 93, 2048–2053. [Google Scholar] [CrossRef]
  17. Perruc, E.; Charpenteau, M.; Ramirez, B.C.; Jauneau, A.; Galaud, J.P.; Ranjeva, R.; Ranty, B. A novel calmodulin-binding protein functions as a negative regulator of osmotic stress tolerance in Arabidopsis thaliana seedlings. Plant J. 2004, 38, 410–420. [Google Scholar] [CrossRef]
  18. Yan, C.; Fan, M.; Yang, M.; Zhao, J.; Zhang, W.; Su, Y.; Xiao, L.; Deng, H.; Xie, D. Injury Activates Ca2+/Calmodulin-Dependent Phosphorylation of JAV1-JAZ8-WRKY51 Complex for Jasmonate Biosynthesis. Mol. Cell 2018, 70, 136–149.e7. [Google Scholar] [CrossRef]
  19. Pecher, P.; Eschen-Lippold, L.; Herklotz, S.; Kuhle, K.; Naumann, K.; Bethke, G.; Uhrig, J.; Weyhe, M.; Scheel, D.; Lee, J. The Arabidopsis thaliana mitogen-activated protein kinases MPK3 and MPK6 target a subclass of ‘VQ-motif’-containing proteins to regulate immune responses. New Phytol. 2014, 203, 592–606. [Google Scholar] [CrossRef]
  20. Weyhe, M.; Eschen-Lippold, L.; Pecher, P.; Scheel, D.; Lee, J. Ménage à trois: The complex relationships between mitogen-activated protein kinases, WRKY transcription factors, and VQ-motif-containing proteins. Plant Signal. Behav. 2014, 9, e29519. [Google Scholar] [CrossRef]
  21. Luo, M.; Dennis, E.S.; Berger, F.; Peacock, W.J.; Chaudhury, A. MINISEED3 (MINI3), a WRKY family gene, and HAIKU2 (IKU2), a leucine-rich repeat (LRR) KINASE gene, are regulators of seed size in Arabidopsis. Proc. Natl. Acad. Sci. USA 2005, 102, 17531–17536. [Google Scholar] [CrossRef] [PubMed]
  22. Wang, A.; Garcia, D.; Zhang, H.; Feng, K.; Chaudhury, A.; Berger, F.; Peacock, W.J.; Dennis, E.S.; Luo, M. The VQ motif protein IKU1 regulates endosperm growth and seed size in Arabidopsis. Plant J. 2010, 63, 670–679. [Google Scholar] [CrossRef] [PubMed]
  23. Lei, R.; Ma, Z.; Yu, D. WRKY2/34-VQ20 Modules in Arabidopsis thaliana Negatively Regulate Expression of a Trio of Related MYB Transcription Factors During Pollen Development. Front. Plant Sci. 2018, 9, 331. [Google Scholar] [CrossRef] [PubMed]
  24. Pan, J.; Wang, H.; Hu, Y.; Yu, D. Arabidopsis VQ18 and VQ26 proteins interact with ABI5 transcription factor to negatively modulate ABA response during seed germination. Plant J. 2018, 95, 529–544. [Google Scholar] [CrossRef] [PubMed]
  25. Li, Y.; Jing, Y.; Li, J.; Xu, G.; Lin, R. Arabidopsis VQ Motif-Containing Protein29 represses seedling deetiolation by interacting with phytochrome-interacting factor1. Plant Physiol. 2014, 164, 2068–2080. [Google Scholar] [CrossRef] [PubMed]
  26. Hu, Y.; Chen, L.; Wang, H.; Zhang, L.; Wang, F.; Yu, D. Arabidopsis transcription factor WRKY8 functions antagonistically with its interacting partner VQ9 to modulate salinity stress tolerance. Plant J. 2013, 74, 730–745. [Google Scholar] [CrossRef] [PubMed]
  27. Cheng, X.; Wang, Y.; Xiong, R.; Gao, Y.; Yan, H.; Xiang, Y. A Moso bamboo gene VQ28 confers salt tolerance to transgenic Arabidopsis plants. Planta 2020, 251, 99. [Google Scholar] [CrossRef] [PubMed]
  28. Ding, H.; Yuan, G.; Mo, S.; Qian, Y.; Wu, Y.; Chen, Q.; Xu, X.; Wu, X.; Ge, C. Genome-wide analysis of the plant-specific VQ motif-containing proteins in tomato (Solanum lycopersicum) and characterization of SlVQ6 in thermotolerance. Plant Physiol. Biochem. 2019, 143, 29–39. [Google Scholar] [CrossRef]
  29. Gayubas, B.; Castillo, M.C.; Ramos, S.; León, J. Enhanced meristem development, tolerance to oxidative stress and hyposensitivity to nitric oxide in the hypermorphic vq10-H mutant in AtVQ10 gene. Plant Cell Environ. 2023, 46, 3445–3463. [Google Scholar] [CrossRef]
  30. Lai, Z.; Li, Y.; Wang, F.; Cheng, Y.; Fan, B.; Yu, J.Q.; Chen, Z. Arabidopsis sigma factor binding proteins are activators of the WRKY33 transcription factor in plant defense. Plant Cell 2011, 23, 3824–3841. [Google Scholar] [CrossRef]
  31. Andreasson, E.; Jenkins, T.; Brodersen, P.; Thorgrimsen, S.; Petersen, N.H.; Zhu, S.; Qiu, J.L.; Micheelsen, P.; Rocher, A.; Petersen, M.; et al. The MAP kinase substrate MKS1 is a regulator of plant defense responses. EMBO J. 2005, 24, 2579–2589. [Google Scholar] [CrossRef] [PubMed]
  32. Petersen, K.; Qiu, J.L.; Lütje, J.; Fiil, B.K.; Hansen, S.; Mundy, J.; Petersen, M. Arabidopsis MKS1 is involved in basal immunity and requires an intact N-terminal domain for proper function. PLoS ONE 2010, 5, e14364. [Google Scholar] [CrossRef] [PubMed]
  33. Fiil, B.K.; Petersen, M. Constitutive expression of MKS1 confers susceptibility to Botrytis cinerea infection independent of PAD3 expression. Plant Signal. Behav. 2011, 6, 1425–1427. [Google Scholar] [CrossRef] [PubMed]
  34. Ali, M.R.M.; Uemura, T.; Ramadan, A.; Adachi, K.; Nemoto, K.; Nozawa, A.; Hoshino, R.; Abe, H.; Sawasaki, T.; Arimura, G.I. The Ring-Type E3 Ubiquitin Ligase JUL1 Targets the VQ-Motif Protein JAV1 to Coordinate Jasmonate Signaling. Plant Physiol. 2019, 179, 1273–1284. [Google Scholar] [CrossRef] [PubMed]
  35. Hao, Z.; Tian, J.; Fang, H.; Fang, L.; Xu, X.; He, F.; Li, S.; Xie, W.; Du, Q.; You, X.; et al. A VQ-motif-containing protein fine-tunes rice immunity and growth by a hierarchical regulatory mechanism. Cell Rep. 2022, 40, 111235. [Google Scholar] [CrossRef]
  36. Liu, H.; Wang, X.; Wang, G.; Cui, P.; Wu, S.; Ai, C.; Hu, N.; Li, A.; He, B.; Shao, X.; et al. The nearly complete genome of Ginkgo biloba illuminates gymnosperm evolution. Nat. Plants 2021, 7, 748–756. [Google Scholar] [CrossRef]
  37. Xiong, X.; Gou, J.; Liao, Q.; Li, Y.; Zhou, Q.; Bi, G.; Li, C.; Du, R.; Wang, X.; Sun, T.; et al. The Taxus genome provides insights into paclitaxel biosynthesis. Nat. Plants 2021, 7, 1026–1036. [Google Scholar] [CrossRef] [PubMed]
  38. Liu, Y.; Wang, S.; Li, L.; Yang, T.; Dong, S.; Wei, T.; Wu, S.; Liu, Y.; Gong, Y.; Feng, X.; et al. The Cycas genome and the early evolution of seed plants. Nat. Plants 2022, 8, 389–401. [Google Scholar] [CrossRef]
  39. Niu, S.; Li, J.; Bo, W.; Yang, W.; Zuccolo, A.; Giacomello, S.; Chen, X.; Han, F.; Yang, J.; Song, Y.; et al. The Chinese pine genome and methylome unveil key features of conifer evolution. Cell 2022, 185, 204–217.e14. [Google Scholar] [CrossRef]
  40. Fu, F.; Song, C.; Wen, C.; Yang, L.; Guo, Y.; Yang, X.; Shu, Z.; Li, X.; Feng, Y.; Liu, B.; et al. The Metasequoia genome and evolutionary relationship among redwoods. Plant Commun. 2023, 4, 100643. [Google Scholar] [CrossRef]
  41. Xu, K.; Wang, P. Genome-wide identification and expression analysis of the VQ gene family in Cucurbita pepo L. PeerJ 2022, 10, e12827. [Google Scholar] [CrossRef] [PubMed]
  42. Wang, Y.; Tang, H.; Debarry, J.D.; Tan, X.; Li, J.; Wang, X.; Lee, T.H.; Jin, H.; Marler, B.; Guo, H.; et al. MCScanX: A toolkit for detection and evolutionary analysis of gene synteny and collinearity. Nucleic Acids Res. 2012, 40, e49. [Google Scholar] [CrossRef]
  43. Jiao, Y.; Wickett, N.J.; Ayyampalayam, S.; Chanderbali, A.S.; Landherr, L.; Ralph, P.E.; Tomsho, L.P.; Hu, Y.; Liang, H.; Soltis, P.S.; et al. Ancestral polyploidy in seed plants and angiosperms. Nature 2011, 473, 97–100. [Google Scholar] [CrossRef] [PubMed]
  44. Li, Z.; Baniaga, A.E.; Sessa, E.B.; Scascitelli, M.; Graham, S.W.; Rieseberg, L.H.; Barker, M.S. Early genome duplications in conifers and other seed plants. Sci. Adv. 2015, 1, e1501084. [Google Scholar] [CrossRef] [PubMed]
  45. Pedrosa, A.M.; Martins, C.d.P.S.; Gonçalves, L.P.; Costa, M.G. Late Embryogenesis Abundant (LEA) Constitutes a Large and Diverse Family of Proteins Involved in Development and Abiotic Stress Responses in Sweet Orange (Citrus sinensis L. Osb.). PLoS ONE 2015, 10, e0145785. [Google Scholar] [CrossRef]
  46. 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] [PubMed]
  47. Dong, Q.; Zhao, S.; Duan, D.; Tian, Y.; Wang, Y.; Mao, K.; Zhou, Z.; Ma, F. Structural and functional analyses of genes encoding VQ proteins in apple. Plant Sci. 2018, 272, 208–219. [Google Scholar] [CrossRef] [PubMed]
  48. Lan, X.; Wang, X.; Tao, Q.; Zhang, H.; Li, J.; Meng, Y.; Shan, W. Activation of the VQ Motif-Containing Protein Gene VQ28 Compromised Nonhost Resistance of Arabidopsis thaliana to Phytophthora Pathogens. Plants 2022, 11, 858. [Google Scholar] [CrossRef]
  49. Kumari, A.; Kaladhar, V.C.; Yadav, N.; Singh, P.; Reddy, K.; Gupta, K.J. Nitric oxide regulates mitochondrial biogenesis in plants. Plant Cell Environ. 2023, 46, 2492–2506. [Google Scholar] [CrossRef]
  50. Suzuki, N.; Rivero, R.M.; Shulaev, V.; Blumwald, E.; Mittler, R. Abiotic and biotic stress combinations. New Phytol. 2014, 203, 32–43. [Google Scholar] [CrossRef]
  51. 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]
  52. Li, N.; Li, X.; Xiao, J.; Wang, S. Comprehensive analysis of VQ motif-containing gene expression in rice defense responses to three pathogens. Plant Cell Rep. 2014, 33, 1493–1505. [Google Scholar] [CrossRef] [PubMed]
  53. Mourier, T.; Jeffares, D.C. Eukaryotic intron loss. Science 2003, 300, 1393. [Google Scholar] [CrossRef] [PubMed]
  54. Adams, K.L.; Wendel, J.F. Polyploidy and genome evolution in plants. Curr. Opin. Plant Biol. 2005, 8, 135–141. [Google Scholar] [CrossRef] [PubMed]
  55. Rensing, S.A. Gene duplication as a driver of plant morphogenetic evolution. Curr. Opin. Plant Biol. 2014, 17, 43–48. [Google Scholar] [CrossRef] [PubMed]
  56. Kaltenegger, E.; Leng, S.; Heyl, A. The effects of repeated whole genome duplication events on the evolution of cytokinin signaling pathway. BMC Evol. Biol. 2018, 18, 76. [Google Scholar] [CrossRef] [PubMed]
  57. Chen, L.; Wu, L.; Yang, L.; Yu, H.; Huang, P.; Wang, Y.; Yao, R.; Zhang, M. TcJAV3-TcWRKY26 Cascade Is a Missing Link in the Jasmonate-Activated Expression of Taxol Biosynthesis Gene DBAT in Taxus chinensis. Int. J. Mol. Sci. 2022, 23, 13194. [Google Scholar] [CrossRef]
  58. Bolser, D.M.; Staines, D.M.; Perry, E.; Kersey, P.J. Ensembl Plants: Integrating Tools for Visualizing, Mining, and Analyzing Plant Genomic Data. Methods Mol. Biol. 2017, 1533, 1–31. [Google Scholar] [CrossRef]
  59. Goodstein, D.M.; Shu, S.; Howson, R.; Neupane, R.; Hayes, R.D.; Fazo, J.; Mitros, T.; Dirks, W.; Hellsten, U.; Putnam, N.; et al. Phytozome: A comparative platform for green plant genomics. Nucleic Acids Res. 2012, 40, D1178–D1186. [Google Scholar] [CrossRef]
  60. Mistry, J.; Chuguransky, S.; Williams, L.; Qureshi, M.; Salazar, G.A.; Sonnhammer, E.L.L.; Tosatto, S.C.E.; Paladin, L.; Raj, S.; Richardson, L.J.; et al. Pfam: The protein families database in 2021. Nucleic Acids Res. 2021, 49, D412–D419. [Google Scholar] [CrossRef]
  61. Potter, S.C.; Luciani, A.; Eddy, S.R.; Park, Y.; Lopez, R.; Finn, R.D. HMMER web server: 2018 update. Nucleic Acids Res. 2018, 46, W200–W204. [Google Scholar] [CrossRef] [PubMed]
  62. Letunic, I.; Khedkar, S.; Bork, P. SMART: Recent updates, new developments and status in 2020. Nucleic Acids Res. 2021, 49, D458–D460. [Google Scholar] [CrossRef] [PubMed]
  63. Marchler-Bauer, A.; Derbyshire, M.K.; Gonzales, N.R.; Lu, S.; Chitsaz, F.; Geer, L.Y.; Geer, R.C.; He, J.; Gwadz, M.; Hurwitz, D.I.; et al. CDD: NCBI’s conserved domain database. Nucleic Acids Res. 2015, 43, D222–D226. [Google Scholar] [CrossRef] [PubMed]
  64. Gasteiger, E.; Gattiker, A.; Hoogland, C.; Ivanyi, I.; Appel, R.D.; Bairoch, A. ExPASy: The proteomics server for in-depth protein knowledge and analysis. Nucleic Acids Res. 2003, 31, 3784–3788. [Google Scholar] [CrossRef] [PubMed]
  65. Voorrips, R.E. MapChart: Software for the graphical presentation of linkage maps and QTLs. J. Hered. 2002, 93, 77–78. [Google Scholar] [CrossRef] [PubMed]
  66. Thumuluri, V.; Almagro Armenteros, J.J.; Johansen, A.R.; Nielsen, H.; Winther, O. DeepLoc 2.0: Multi-label subcellular localization prediction using protein language models. Nucleic Acids Res. 2022, 50, W228–W234. [Google Scholar] [CrossRef] [PubMed]
  67. Nguyen Ba, A.N.; Pogoutse, A.; Provart, N.; Moses, A.M. NLStradamus: A simple Hidden Markov Model for nuclear localization signal prediction. BMC Bioinform. 2009, 10, 202. [Google Scholar] [CrossRef] [PubMed]
  68. Horton, P.; Park, K.J.; Obayashi, T.; Fujita, N.; Harada, H.; Adams-Collier, C.J.; Nakai, K. WoLF PSORT: Protein localization predictor. Nucleic Acids Res. 2007, 35, W585–W587. [Google Scholar] [CrossRef]
  69. Bailey, T.L.; Johnson, J.; Grant, C.E.; Noble, W.S. The MEME Suite. Nucleic Acids Res. 2015, 43, W39–W49. [Google Scholar] [CrossRef]
  70. Chen, C.; Chen, H.; Zhang, Y.; Thomas, H.R.; Frank, M.H.; He, Y.; Xia, R. TBtools: An Integrative Toolkit Developed for Interactive Analyses of Big Biological Data. Mol. Plant 2020, 13, 1194–1202. [Google Scholar] [CrossRef]
  71. Katoh, K.; Standley, D.M. MAFFT multiple sequence alignment software version 7: Improvements in performance and usability. Mol. Biol. Evol. 2013, 30, 772–780. [Google Scholar] [CrossRef] [PubMed]
  72. Price, M.N.; Dehal, P.S.; Arkin, A.P. FastTree 2—Approximately maximum-likelihood trees for large alignments. PLoS ONE 2010, 5, e9490. [Google Scholar] [CrossRef] [PubMed]
  73. Letunic, I.; Bork, P. Interactive Tree Of Life (iTOL) v5: An online tool for phylogenetic tree display and annotation. Nucleic Acids Res. 2021, 49, W293–W296. [Google Scholar] [CrossRef] [PubMed]
  74. Kovach, A.; Wegrzyn, J.L.; Parra, G.; Holt, C.; Bruening, G.E.; Loopstra, C.A.; Hartigan, J.; Yandell, M.; Langley, C.H.; Korf, I.; et al. The Pinus taeda genome is characterized by diverse and highly diverged repetitive sequences. BMC Genom. 2010, 11, 420. [Google Scholar] [CrossRef] [PubMed]
  75. Krzywinski, M.; Schein, J.; Birol, I.; Connors, J.; Gascoyne, R.; Horsman, D.; Jones, S.J.; Marra, M.A. Circos: An information aesthetic for comparative genomics. Genome Res. 2009, 19, 1639–1645. [Google Scholar] [CrossRef] [PubMed]
  76. Wang, D.; Zhang, Y.; Zhang, Z.; Zhu, J.; Yu, J. KaKs_Calculator 2.0: A toolkit incorporating gamma-series methods and sliding window strategies. Genom. Proteom. Bioinform. 2010, 8, 77–80. [Google Scholar] [CrossRef] [PubMed]
  77. Niu, S.H.; Liu, S.W.; Ma, J.J.; Han, F.X.; Li, Y.; Li, W. The transcriptional activity of a temperature-sensitive transcription factor module is associated with pollen shedding time in pine. Tree Physiol. 2019, 39, 1173–1186. [Google Scholar] [CrossRef]
  78. Zhang, X.; Cao, H.; Wang, H.; Zhang, R.; Jia, H.; Huang, J.; Zhao, J.; Yao, J. Effects of graphene on morphology, microstructure and transcriptomic profiling of Pinus tabuliformis Carr. roots. PLoS ONE 2021, 16, e0253812. [Google Scholar] [CrossRef]
  79. Li, J.; Han, F.; Yuan, T.; Li, W.; Li, Y.; Wu, H.X.; Wei, H.; Niu, S. The methylation landscape of giga-genome and the epigenetic timer of age in Chinese pine. Nat. Commun. 2023, 14, 1947. [Google Scholar] [CrossRef]
  80. Bolger, A.M.; Lohse, M.; Usadel, B. Trimmomatic: A flexible trimmer for Illumina sequence data. Bioinformatics 2014, 30, 2114–2120. [Google Scholar] [CrossRef]
  81. Kim, D.; Paggi, J.M.; Park, C.; Bennett, C.; Salzberg, S.L. Graph-based genome alignment and genotyping with HISAT2 and HISAT-genotype. Nat. Biotechnol. 2019, 37, 907–915. [Google Scholar] [CrossRef]
  82. Pertea, M.; Pertea, G.M.; Antonescu, C.M.; Chang, T.C.; Mendell, J.T.; Salzberg, S.L. StringTie enables improved reconstruction of a transcriptome from RNA-seq reads. Nat. Biotechnol. 2015, 33, 290–295. [Google Scholar] [CrossRef]
  83. Langmead, B.; Salzberg, S.L. Fast gapped-read alignment with Bowtie 2. Nat. Methods 2012, 9, 357–359. [Google Scholar] [CrossRef]
  84. Li, B.; Dewey, C.N. RSEM: Accurate transcript quantification from RNA-Seq data with or without a reference genome. BMC Bioinform. 2011, 12, 323. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Members identification and sequence characteristics of VQ genes. (a) The relationship between genome size and VQ gene numbers in 56 plant species. (b) The conserved VQ motif. (c) The maximum sequence similarity in each alignment site. The red frame is the region of the VQ motif, and the X-axis represents alignment site. (d) The length of VQ proteins. The X-axis represents 2469 VQ proteins, and the Y-axis shows amino acid numbers of VQ protein.
Figure 1. Members identification and sequence characteristics of VQ genes. (a) The relationship between genome size and VQ gene numbers in 56 plant species. (b) The conserved VQ motif. (c) The maximum sequence similarity in each alignment site. The red frame is the region of the VQ motif, and the X-axis represents alignment site. (d) The length of VQ proteins. The X-axis represents 2469 VQ proteins, and the Y-axis shows amino acid numbers of VQ protein.
Ijms 24 14968 g001
Figure 2. The ratio of single exon (intron-free) VQ genes. The labels above the column mean ‘numbers of single-exon VQ genes/numbers of all VQ genes’.
Figure 2. The ratio of single exon (intron-free) VQ genes. The labels above the column mean ‘numbers of single-exon VQ genes/numbers of all VQ genes’.
Ijms 24 14968 g002
Figure 3. The average GC content of genes in the coding sequence. The blue column is just for VQ genes, and the red column is for all coding genes in each species.
Figure 3. The average GC content of genes in the coding sequence. The blue column is just for VQ genes, and the red column is for all coding genes in each species.
Ijms 24 14968 g003
Figure 4. Chromosomal distribution of VQ genes in three gymnosperms. Chromosome numbers are listed at the top. The length of the chromosome is displayed in megabase (Mb) scale.
Figure 4. Chromosomal distribution of VQ genes in three gymnosperms. Chromosome numbers are listed at the top. The length of the chromosome is displayed in megabase (Mb) scale.
Ijms 24 14968 g004
Figure 5. Phylogenetic tree, conserved motifs, and gene structure of VQ protein in three gymnosperms. (a) Phylogenetic tree of 116 VQ proteins in three gymnosperms based on the results of sequence alignment. (b) A total of 41 conserved motifs of the VQ protein. Each specific motif is indicated by a different colored box. (c) Gene structure (exon–intron) of the VQ genes. Exons are indicated by green rectangles, and lines connecting two exons represent introns.
Figure 5. Phylogenetic tree, conserved motifs, and gene structure of VQ protein in three gymnosperms. (a) Phylogenetic tree of 116 VQ proteins in three gymnosperms based on the results of sequence alignment. (b) A total of 41 conserved motifs of the VQ protein. Each specific motif is indicated by a different colored box. (c) Gene structure (exon–intron) of the VQ genes. Exons are indicated by green rectangles, and lines connecting two exons represent introns.
Ijms 24 14968 g005
Figure 6. Collinear analysis between different gymnosperms. The grey lines in the background represent the collinear regions between different plant genomes, while the red lines highlight the collinear VQ gene pairs. The numbers represent chromosomes.
Figure 6. Collinear analysis between different gymnosperms. The grey lines in the background represent the collinear regions between different plant genomes, while the red lines highlight the collinear VQ gene pairs. The numbers represent chromosomes.
Ijms 24 14968 g006
Figure 7. Expression level of VQ genes in different tissues of three gymnosperms. Expression levels of these VQ genes were obtained using RNA-seq data and measured using TPM values. Heatmap was generated using log2 (TPM + 1) values.
Figure 7. Expression level of VQ genes in different tissues of three gymnosperms. Expression levels of these VQ genes were obtained using RNA-seq data and measured using TPM values. Heatmap was generated using log2 (TPM + 1) values.
Ijms 24 14968 g007
Figure 8. Unrooted phylogenetic tree of VQ proteins in 56 plant species. Multiple sequences alignment was performed using MAFFT (v7.511) software, and the phylogenetic tree was constructed via FastTree. Different colors indicate different groups, and these 2469 VQ proteins were clustered into nine subgroups.
Figure 8. Unrooted phylogenetic tree of VQ proteins in 56 plant species. Multiple sequences alignment was performed using MAFFT (v7.511) software, and the phylogenetic tree was constructed via FastTree. Different colors indicate different groups, and these 2469 VQ proteins were clustered into nine subgroups.
Ijms 24 14968 g008
Table 1. Summary of molecular characterization and subcellular localization of VQ proteins in three gymnosperms.
Table 1. Summary of molecular characterization and subcellular localization of VQ proteins in three gymnosperms.
SpeciesNumberMotif TypeLength (aa)MW (kDa)pISubcellular Localization
LTGFTGVTGMTGYTGFTANucleusCytoplasm/Nucleus
Ginkgo biloba342382-1-283 ± 9830.86 ± 10.257.67 ± 1.73268
Taxus chinensis181331--1165 ± 4318.37 ± 4.818.55 ± 1.73171
Pinus tabuliformis64362224--260 ± 8828.09 ± 9.108.50 ± 1.604915
Protein length, MW; pI, mean ± SD.
Table 2. Duplication events of VQ genes in three gymnosperms.
Table 2. Duplication events of VQ genes in three gymnosperms.
SpeciesGene 1Gene 2Duplication TypeKaKsKa/Ks
Ginkgo bilobaGbVQ5GbVQ6Tandem0.1030730.1290920.798447
GbVQ5GbVQ7Proximal0.0953080.0947411.005980
GbVQ6GbVQ7Proximal0.0576740.0593490.971771
GbVQ14GbVQ15Tandem0.0442800.0523440.845953
GbVQ14GbVQ16Tandem0.0454200.0334701.357020
GbVQ15GbVQ16Tandem0.0620360.0675610.918212
GbVQ26GbVQ27Tandem0.0403190.0412620.977136
GbVQ2GbVQ20SegmentalNANANA
Taxus chinensisTcVQ10TcVQ12ProximalNANANA
TcVQ11TcVQ12Proximal0.0985320.0944281.043470
Pinus tabuliformisPtVQ3PtVQ4Proximal0.0542390.1795280.302121
PtVQ5PtVQ17Dispersed0.0354090.0349281.013750
PtVQ6PtVQ13DispersedNANANA
PtVQ6PtVQ7Proximal0.0047650.0281180.169469
PtVQ7PtVQ13Dispersed0.1421770.1904260.746623
PtVQ7PtVQ8Tandem00NA
PtVQ8PtVQ6Proximal0.0047650.0281180.169469
PtVQ8PtVQ13Dispersed0.1793570.1939090.924953
PtVQ16PtVQ5Dispersed0.0099380.0086091.154340
PtVQ16PtVQ17Proximal0.0250900.0441060.568853
PtVQ18PtVQ19Proximal0.0321030.0174211.842760
PtVQ18PtVQ20Dispersed0.0357650.0652270.548313
PtVQ19PtVQ20Tandem0.0458740.0537540.853421
PtVQ21PtVQ22Tandem0.0984280.2256480.436201
PtVQ29PtVQ30Proximal0.0013460NA
PtVQ31PtVQ32Dispersed0.0014850.0253030.058707
PtVQ37PtVQ38Tandem00NA
PtVQ44PtVQ61Dispersed0.0102090.0054171.884470
PtVQ44PtVQ62Dispersed0.0283670.0304570.931388
PtVQ44PtVQ45Tandem0.0257400.0681540.377675
PtVQ44PtVQ46Proximal0.0450120.0634650.709246
PtVQ45PtVQ46Tandem0.0598900.1011270.592221
PtVQ48PtVQ49Tandem00NA
PtVQ48PtVQ50Proximal0.0046490.0244700.189988
PtVQ49PtVQ50Proximal0.0049470.0260180.190129
PtVQ54PtVQ55Dispersed00NA
PtVQ56PtVQ57Tandem00NA
PtVQ59PtVQ60Tandem00NA
PtVQ61PtVQ62Tandem0.0318800.0361300.882364
PtVQ61PtVQ45Dispersed0.0266130.0711330.374138
PtVQ61PtVQ46Dispersed0.0487330.0695100.701090
PtVQ62PtVQ45Dispersed0.0213520.0655540.325719
PtVQ62PtVQ46Dispersed0.0589540.0853320.690883
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Tian, J.; Zhang, J.; Francis, F. Large-Scale Identification and Characterization Analysis of VQ Family Genes in Plants, Especially Gymnosperms. Int. J. Mol. Sci. 2023, 24, 14968. https://doi.org/10.3390/ijms241914968

AMA Style

Tian J, Zhang J, Francis F. Large-Scale Identification and Characterization Analysis of VQ Family Genes in Plants, Especially Gymnosperms. International Journal of Molecular Sciences. 2023; 24(19):14968. https://doi.org/10.3390/ijms241914968

Chicago/Turabian Style

Tian, Jinfu, Jiahui Zhang, and Frédéric Francis. 2023. "Large-Scale Identification and Characterization Analysis of VQ Family Genes in Plants, Especially Gymnosperms" International Journal of Molecular Sciences 24, no. 19: 14968. https://doi.org/10.3390/ijms241914968

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