Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

The IQD Gene Family in Soybean: Structure, Phylogeny, Evolution and Expression

  • Lin Feng ,

    Contributed equally to this work with: Lin Feng, Zhu Chen

    Affiliation Laboratory of Modern Biotechnology, School of Forestry and Landscape Architecture, Anhui Agricultural University, Hefei, China

  • Zhu Chen ,

    Contributed equally to this work with: Lin Feng, Zhu Chen

    Affiliation Laboratory of Modern Biotechnology, School of Forestry and Landscape Architecture, Anhui Agricultural University, Hefei, China

  • Hui Ma,

    Affiliation Laboratory of Modern Biotechnology, School of Forestry and Landscape Architecture, Anhui Agricultural University, Hefei, China

  • Xue Chen,

    Affiliation Laboratory of Modern Biotechnology, School of Forestry and Landscape Architecture, Anhui Agricultural University, Hefei, China

  • Yuan Li,

    Affiliation Laboratory of Modern Biotechnology, School of Forestry and Landscape Architecture, Anhui Agricultural University, Hefei, China

  • Yiyi Wang,

    Affiliation Laboratory of Modern Biotechnology, School of Forestry and Landscape Architecture, Anhui Agricultural University, Hefei, China

  • Yan Xiang

    xiangyanahau@sina.com

    Affiliations Laboratory of Modern Biotechnology, School of Forestry and Landscape Architecture, Anhui Agricultural University, Hefei, China, Key Laboratory of Crop Biology of Anhui Agriculture University, Hefei, China

Correction

23 Mar 2015: The PLOS ONE Staff (2015) Correction: The IQD Gene Family in Soybean: Structure, Phylogeny, Evolution and Expression. PLOS ONE 10(3): e0119318. https://doi.org/10.1371/journal.pone.0119318 View correction

Abstract

Members of the plant-specific IQ67-domain (IQD) protein family are involved in plant development and the basal defense response. Although systematic characterization of this family has been carried out in Arabidopsis, tomato (Solanum lycopersicum), Brachypodium distachyon and rice (Oryza sativa), systematic analysis and expression profiling of this gene family in soybean (Glycine max) have not previously been reported. In this study, we identified and structurally characterized IQD genes in the soybean genome. A complete set of 67 soybean IQD genes (GmIQD167) was identified using Blast search tools, and the genes were clustered into four subfamilies (IQD I–IV) based on phylogeny. These soybean IQD genes are distributed unevenly across all 20 chromosomes, with 30 segmental duplication events, suggesting that segmental duplication has played a major role in the expansion of the soybean IQD gene family. Analysis of the Ka/Ks ratios showed that the duplicated genes of the GmIQD family primarily underwent purifying selection. Microsynteny was detected in most pairs: genes in clade 1–3 might be present in genome regions that were inverted, expanded or contracted after the divergence; most gene pairs in clade 4 showed high conservation with little rearrangement among these gene-residing regions. Of the soybean IQD genes examined, six were most highly expressed in young leaves, six in flowers, one in roots and two in nodules. Our qRT-PCR analysis of 24 soybean IQD III genes confirmed that these genes are regulated by MeJA stress. Our findings present a comprehensive overview of the soybean IQD gene family and provide insights into the evolution of this family. In addition, this work lays a solid foundation for further experiments aimed at determining the biological functions of soybean IQD genes in growth and development.

Introduction

Ca2 + is a pivotal cytosolic second messenger involved in many physiological processes such as plant growth [1], plant-pathogen interactions [2], photosynthetic electron transport and photophosphorylation [3], regulation of stomatal aperture [4], hormonal regulation [5] and so on. Plants produce calcium signals by adjusting cytoplasm Ca2+ levels at specific times, places and concentrations [6], responding to numerous extracellular stimuli including physical signals (light, temperature, gravity, etc.) and chemical signals (plant hormones, pathogenic bacteria inducing factors, etc.) [7].

The transmission of these intracellular calcium signals relies on the oscillation signal generated by voltage- and ligand-gated Ca2+-permeable channels (influx) and by Ca2+-ATPases and antiporters (efflux) to return to resting Ca2+ levels [8], [9]. The conduction of calcium signals is also dependent on downstream Ca2+ sensors. These Ca2+ sensors detect changes in Ca2+ levels by binding to Ca2+ via domains such as EF hands, which undergo conformational changes [10]. Consequently, calcium signature information is decoded and relayed by these Ca2+ sensors [6], [11][13].

To date, approximately four major classes of Ca2+ sensors have been identified in plants. Most of these sensors contain the Ca2+-binding EF-hand motif, a conserved helix-loop-helix structure that can bind to a single Ca2+ ion [7]. The four major classes of Ca2+ sensors are as follows: class A: calmodulin (CaM), containing four EF-hand motifs; class B: calcineurin B-like (CBL) proteins, possessing three EF-hand motifs; class C: Ca2+-dependent protein kinases (CDPK), containing four EF-hand motifs and a Ca2+-dependent Ser/Thr protein kinase domain and class D: lacking EF-hand motifs [7], [14][19].

Calmodulin (CaM) and calcineurin B-like (CBL) proteins, which lack catalytic activity, are sometimes referred to as “Ca2+ sensor relays” [15], [19], [20]. In contrast, CDPK proteins, which function as catalytic effectors, are referred to as “Ca2+ sensor responders” [18]. Among these Ca2+-binding proteins, calmodulin is the most extensively studied Ca2+ sensor. Calmodulin is small, acid resistant, heat resistant and highly stable. This multifunctional protein is widespread in eukaryotic cells, highly conserved and has at least 30 multiple target proteins or enzymes [21][23].

To mediate intracellular calcium signaling pathways, Ca2+ sensor relays expose their negative hydrophobic surfaces after they undergo conformational changes induced by Ca2+ binding. As a result, the affinity between Ca2+ sensor relays and their effectors are enhanced, and the biochemical activities of target proteins are modulated by Ca2+ sensor relays [6], [12], [14], [19].

In the final phase of the calcium signal transduction process, the target effectors respond to specific extracellular signals by regulating various cellular activities. Calmodulin interacts with numerous target proteins termed calmodulin-binding proteins (CaMBPs), mainly by recognizing and targeting calmodulin-binding domains (CaMBD; basic amphiphilic helices usually composed of 16–35 amino acid residues) in the CaMBPs via its negative hydrophobic pockets [12], [22], [24].

CaMBD amino acid sequences contain three CaMBD motifs that are grouped into two categories, including a Ca2+-independent motif termed the IQ motif and two Ca2+-dependent motifs referred to as the l-5-10 motif and the l-8-14 motif. The number and positions of these motifs in different CaMBPs are variable [25][27]. Due to the diversity of the motif arrangement, there are a variety of diverse CaMBPs with disparate functions, which are implicated in plant development, metabolic regulation, stress reactions, defense reactions, transcriptional regulation and so on [28], [29].

Plant-specific IQ67 domain (IQD) protein families were first identified in Arabidopsis and rice by Abel et al. (2005) [30]. These proteins have two common features in their IQ67 domains (67 conserved amino acid residues) [31]. One feature is 1–3 copies IQ motifs are separated by 11 and 15 residues and overlapped certain regions with 1–4 copies 1-5-10 motif as well as 1-8-14 motif. The other hallmark is a highly conserved exon-intron boundary that interrupts codons 16 and 17 with a 0 phase intron [31][33]. To date, IQD gene families have been identified in four genomes (Arabidopsis, rice, tomato and Brachypodium distachyon), including approximately 30 IQD genes per species (33 in Arabidopsis, 29 in rice, 34 in tomato and 23 in Brachypodium distachyon), and the functions of several members of the IQD family have been reported [30], [34], [35]. Overexpression of Arabidopsis IQD1 can mediate the accumulation of glucosinolate in response to insect herbivory [31]. Arabidopsis IQD22 contributes to the negative feedback regulation of gibberellin (GA) [36]. The tomato SUN gene plays a role in elongating tomato fruit by increasing the vertical division of cells and reducing horizontal cell divisions [37][39].

Soybean serves as a major source of vegetable proteins and edible oil and own the ability to fix atmospheric nitrogen via its intimate symbiosis with microorganisms. This crucial leguminous seed crop has high economic and nutritional value [40], [41], serving as a main food crop for humans and animals in many parts of the world. Nevertheless, soybean production is limited by many biotic stresses. For example, Asian soybean rust (ASR, caused by the fungus Phakopsora pachyrhizi) results in soybean yield losses ranging from 10 to 80% in various countries [42], [43].

In this study, we identified and characterized 67 soybean IQD genes. Among these, we subjected 24 IQD III genes to qRT-PCR analysis to investigate their response to MeJA stress. We determined that all 24 soybean IQD genes are stress-responsive. Our findings lay the foundation of further investigations of the functions of these calmodulin target proteins in soybean.

Results

Identification and annotation of IQD genes in soybean

As described in previous studies, IQD proteins, which are plant-specific calcium-dependent calmodulin-binding proteins, contain 67 amino acid residues in their central regions referred to as the IQ67 domain, including three CaM recruitment motifs exhibiting unique repetitive patterns. The Ca2+-independent IQ motif (IQxxxRGxxxR or its more relaxed version [ILV]QxxxRxxxx [R, K]) is present in 1–3 copies and overlaps with 1–4 copies of the Ca2+-dependent 1-5-10 motif ([FILVW]x3 [FILV]x4 [FILVW]) and the 1-8-14 motif ([FILVW]x6 [FAILVW]x5 [FILVW]) by several conserved basic and hydrophobic amino acid residues flanking these motifs [30], [34]. These features allow the IQ67 domain to fold into a basic amphiphilic helix structure, which enables these proteins to perform specific roles.

To conducted genome-wide identification of IQD gene families in soybean, we performed Glycine max genome BLASTP analysis. Through removing redundant sequences and pattern identificating, a total of 67 IQD genes were identified in the soybean genome, which is twice that of Arabidopsis (Table 1 and 2). We named these 67 IQD genes GmIQD1 to GmIQD67 according to their physical locations (from top to bottom) on chromosomes 1–20 (Table 1).

thumbnail
Table 1. List of 67 IQD genes identified in soybean, their sequence characteristics and subcellular localization.

https://doi.org/10.1371/journal.pone.0110896.t001

thumbnail
Table 2. Number of IQD genes in the soybean, rice, Arabidopsis, tomato and Brachypodium distachyon genomes.

https://doi.org/10.1371/journal.pone.0110896.t002

The physicochemical parameters of each gene were calculated using ExPASy. Although all of the GmIQD genes encode the conserved IQ67 domains (Figure S1), their sequences are highly diverse with respect to size (141–904 aa) and molecular mass (16.3–99.2 kDa; Table 1). Almost all soybean IQD proteins (97%) have relatively high isoelectric points (pI>7.0 with an average of 10.1), except for GmIQD9 (pI 5.4) and GmIQD17 (pI 5.7; Table 1). All soybean IQD proteins were submitted to TargetP and Wolf PSORT to predict their subcellular localizations. Wolf PSORT revealed that fifty-seven soybean IQD proteins are localized to the nucleus, nine to the chloroplast and one to the endoplasmic reticulum. TargetP analysis revealed that fifteen soybean IQD proteins are located in the mitochondria, five in the chloroplast, one in the secretory pathway and forty-six in other compartments (Table 1). The detailed parameters are provided in Table 1.

Phylogenetic and structural analyses of the soybean IQD genes

To infer the similarity and evolutionary ancestry of soybean IQD proteins, we constructed an unrooted phylogenetic tree of the 67 soybean protein sequences. The soybean IQD gene family was categorized into four major subfamilies (subfamily I, II, III and IV; Figure 1a) according to phylogenetic analysis of IQD genes in Arabidopsis, rice, tomato and Brachypodium distachyon [30], [34], [35]. Subfamily I was further divided into four subclasses (clade Ia, Ib, Ic and Id), and subfamily II and III were divided into two subclasses (clade IIa and IIb; clade IIIa and IIIb) based on bootstrap values, the existence and positions of introns and the presence of protein motifs flanking the IQ67 domain (Figure 1 and 2). Subfamily I (containing 27 members) is the largest group, followed by subfamily III (24) and subfamily IV (10). Subfamily II has the fewest IQD gene members (6). This distribution pattern is similar to that observed for IQD genes in Arabidopsis and rice (Table 2). The phylogenetic tree reveals that 62 of the 67 soybean IQD genes form 31 gene pairs with strong bootstrap values (Figure 1a).

thumbnail
Figure 1. Phylogenetic relationships and exon/intron organization of soybean IQD genes.

a: Unrooted tree generated with Clustal X2.0 using the full-length amino acid sequences of the 67 soybean IQD proteins by the Neighbor-Joining (NJ) method with 1,000 bootstrap replicates. Subfamilies and subclasses of IQD genes (I–IV) are highlighted with different colored backgrounds and vertical bars next to the gene names of the tree. b: Exon/intron organization of soybean IQD genes. Green boxes represent exons and black lines represent introns. Untranslated regions (UTRs) are indicated by blue boxes. Numbers 0 and 1 represent the splicing phases. The sizes of exons and introns can be estimated using the scale at the bottom. The exon encoding amino acids 17–67 of the IQ67 domain in each soybean gene is indicated with an asterisk.

https://doi.org/10.1371/journal.pone.0110896.g001

thumbnail
Figure 2. Motif patterns in 67 IQD proteins of soybean.

The schematic soybean IQD proteins were aligned relative to the IQ67 domain (motif 1, light blue box). The lengths of the proteins and motifs can be estimated using the scale. Motifs shared by at least four soybean IQD proteins are depicted at the reference bar on top of each alignment. The positions of putative calmodulin-binding sites predicted by the Calmodulin Target Database are indicated by vertical tick marks below each protein model. Subfamilies and subclasses (I–IV) of IQD proteins are highlighted with colored backgrounds and colored vertical bars as in Figure 1a to the right of the gene names.

https://doi.org/10.1371/journal.pone.0110896.g002

To further examine the structural diversity of the IQD genes in soybean, we deduced the exon/intron organization of individual GmIQD genes (Figure 1b). A comparison of the 67 genomic loci with corresponding cDNA sequences revealed that most of the gene models predicted by GSDS are correct, except for one pair of genes (GmIQD9/-17). Both GmIQD9 and GmIQD17 encode six exons, but GSDS predicted that these genes contain only five exons. This unconformity is caused by the missing annotation of the fifth intron by GSDS. The schematic structures reveal that the coding sequence of each IQD gene contains 2–6 translated exons (Figure 1b), which is similar to that reported in Arabidopsis, rice and Brachypodium distachyon [30], [34]. More than three–fifths of the soybean IQD gene family (41 members) contain five or six protein-coding exons, and one gene (GmIQD56, encoding the smallest protein, comprising 141 aa) contains two exons (Figure 1b). Most closely related soybean IQD members in the same subfamilies share similar intron numbers and exon lengths. Soybean IQD genes in subfamily II and IV possess five and six exons, respectively. Most members in subfamily III contain five exons, except for GmIQD44 (four exons) and GmIQD32, GmIQD41, GmIQD46 and GmIQD47 (six exons). Subfamily I genes harbor 2–5 exons. All introns of most IQD genes are in phase-0 (interrupting two triplet codons exactly); a phase-1 intron (separating the first and second nucleotides of a codon) was found in 15 remaining IQD genes, and no phase-2 intron (splitting the second and third nucleotides of a codon) was found (Figure 1b).

The exon/intron organization of 31 paralogous pairs that clustered together at the terminal branch of the phylogenetic tree was further examined to obtain traceable intron gain/loss information. Although twenty-seven paralogous pairs exhibited conserved exon/intron structures, four pairs (GmIQD16/-27, GmIQD38/-65, GmIQD25/-44 and GmIQD42/-46) showed certain variations (Figure 1b). These differences may have been derived from single intron loss or gain events during the long evolutionary period. Based on analysis of the exon/intron organization of IQD genes from soybean, Arabidopsis [30], rice [30], and Brachypodium distachyon [34], we infered that both GmIQD16 and GmIQD38 gained the third intron; GmIQD46 gained the first intron while GmIQD44 lost the first intron. The second or third exons in the central regions of most members encode amino acids 17–67 of the IQ67 domain, except for GmIQD46 (the fourth exon) and GmIQD56 (the C-terminal exon), with a conserved phase-0 intron separating codons 16 and 17 (Figure 1b and Figure S1).

A total of 67 IQD genes from soybean were subjected to analysis with MEME to reveal conserved domains or motifs shared among related proteins. We identified 10 conserved motifs (Figure 2 and Table S1). Each of the putative motifs was annotated by searching Pfam and SMART. Motif 1 was found to encode the IQ domain. Motif 2 and motif 7 were found to encode proteins of unknown function (DUF4005) and (DUF3982). While the other subfamily-specific motifs have not functional annotation. As expected, most of the closely related members had common motif compositions, suggesting functional similarities among IQD proteins within the same subfamily (Figure 2). The most common motif is motif 1, found in all sixty-seven soybean IQD genes (Figure 2). Motif 8 is mainly present in subfamily I besides one of GmIQD14 exists in subfamily IV. Subfamily III members contain motif 1, motif 10, motif 4 and motif 3 in order, except for GmIQD28 lacking motif 10. Motif 7 is peculiar to subfamily IV. To some extent, these subfamily-specific motifs may contribute to the functional divergence of IQD genes in soybean. The detailed information is shown in table S1. To predict calmodulin-binding sites, we searched the Calmodulin Target Database, which provides various structural and biophysical parameters for the 67 soybean IQD protein sequences. This analysis predicted that all soybean IQD proteins contain multiple IQ motifs and 1–3 strings of high-scoring amino acid residues (Table 3). These IQ motifs and amino acid residues indicate the locations of putative calmodulin interaction sites. Among the 67 IQD protein sequences, the predicted calmodulin binding sites of 50 sequences overlap with the IQ67 domain (Figure 2).

thumbnail
Table 3. Predicted calmodulin binding sites in soybean IQD proteins.

https://doi.org/10.1371/journal.pone.0110896.t003

Chromosomal location and gene duplication

The 67 soybean IQD genes were mapped to all 20 soybean chromosomes. The distribution of soybean IQD genes varies depending on the chromosome and appears to be unequal. Both chromosomes 11 and 18 contain only one soybean IQD gene, while chromosomes 13, which possesses seven IQD genes, has the highest number of IQD genes per chromosome. Although high densities of IQD genes were found on some chromosomal regions, for instance, the bottom of chromosome 13, these is no substantial clustering of soybean IQD genes on the map (Figure 3).

thumbnail
Figure 3. Chromosomal distribution and segmental duplication events for soybean IQD genes.

The 67 IQD genes were mapped to the 20 soybean chromosomes. The duplicated paralogous pairs of IQD genes in the segmental duplicated blocks are indicated with small boxes of the same color and connected by dashed lines of the same color. Red triangles represent soybean IQD genes located on duplicated segments with the corresponding members lost. Black circles indicate soybean IQD genes not located in duplicated regions. Scale represents the length of the chromosome.

https://doi.org/10.1371/journal.pone.0110896.g003

We investigated gene duplication events to further understand the expansion mechanism of the soybean IQD family. Except for three genes (GmIQD11, -39 and -51) located outside of a duplicated block, 64 genes were mapped onto 48 related duplicated blocks (Figure 3 and Table S2). Among these, twenty-two block pairs retained thirty GmIQD gene pairs, whereas the remaining four duplication blocks harbor GmIQD3, -32, -56 and -60 respectively, but lack IQD sisters in their corresponding synteny blocks (Figure 3 and Table S2). Analysis of GmIQD paralogous pairs showed that one pair (GmIQD11/-39) appear to be closely related paralogs, sharing 91.2% identity (Table S3) as well as similar exon–intron organization. However, both of them exist outside of any duplicated blocks. Except for GmIQD11/-39, 30 out of 31 gene pairs have remained in conserved positions on segmental duplicated blocks, indicating that these genes were generated by segmental duplication. Furthermore, we analyzed the adjacent genes to determined whether tandem duplication has taken place. A pair of genes separated by three or fewer genes within a 100-kb region on a chromosome may have resulted from tandem duplication. According to this criterion, no pair was found to be generated by tandem duplication. Therefore, segmental duplication appears to have played a crucial role in the expansion of the IQD gene family in soybean (Figure 3 and Table S2).

To explore the selective constraints on duplicated soybean IQD genes, we calculated the ratio of nonsynonymous versus synonymous substitutions (Ka/Ks) for each pair of duplicated IQD genes. In general, a ratio of 1 indicates that both genes are drifting neutrally; a Ka/Ks ratio >1 indicates accelerated evolution with positive selection, while a ratio <1 indicates functional constraint, with negative or purifying selection of the genes[44]. The Ka/Ks ratios from 31 soybean IQD paralogous pairs (Table 4) were less than 0.6. This result suggests that the soybean IQD gene family has evolved mainly under the influence of strong purifying selection pressure, with limited functional divergence occurring after segmental duplication. Duplication of these 31 paralogous pairs was estimated to have occurred between 6.39 to 17.94 Mya (Table 4), according to the divergence rate of 6.1×10−9 synonymous mutations per synonymous site per year, as previously proposed for soybean [45], [46].

thumbnail
Table 4. Divergence between paralogous IQD gene pairs in soybean.

https://doi.org/10.1371/journal.pone.0110896.t004

Comparative analysis of the IQD genes in soybean, Arabidopsis, rice, tomato and Brachypodium distachyon

The development of comparative genomics has enabled the analysis of the same protein families among different species. We constructed an NJ phylogenetic tree using 184 full-length protein sequence to reveal the evolutionary relationships among soybean, Arabidopsis, rice, tomato and Brachypodium distachyon IQD proteins [34]. In Arabidopsis, the IQD gene family is divided into four subfamilies, with AtIQD33 (containing a C-terminally truncated IQ67 domain) as the outgroup. Therefore, based on their phylogenetic relationships, the combined phylogenetic tree can be divided into five distinct subfamilies (I to V; Figure 4) [30]. In general, IQD I genes comprise the largest subfamily in these plant species, except for Brachypodium distachyon, where both IQD I and III comprise the largest subfamilies. By contrast, IQD V genes comprise the smallest IQD subfamily (Figure 4, Table 2).

thumbnail
Figure 4. Phylogenetic tree of full-length IQD proteins from soybean, Arabidopsis, rice, tomato and Brachypodium distachyon.

The tree was generated with Clustal X2.0 using the NJ method. Dicotyledons (soybean, tomato and Arabidopsis) IQD proteins are marked with colored dots. Monocotyledons (rice and Brachypodium distachyon) are marked with colored triangles. A moss IQD protein (Pp1s38230v6), used as the outgroup, is marked with a black box. Each IQD subfamily is indicated by a specific color.

https://doi.org/10.1371/journal.pone.0110896.g004

To illustrate the paralogous and orthologous relationships among IQD family members, the subfamilies were further divided into subgroups using previously defined clades from studies of Arabidopsis, rice and tomato IQDs, as shown in Figure 4. IQD subfamily I was divided into four subclasses, i.e., a, b, c and d, and clade b was further divided into two clades, b1 and b2. Because one of the IQD Ib clades only contains four IQD genes (BdIQD11, BdIQD19, OsIQD19 and OsIQD20) from monocots, we assigned these four genes to the rice- and Brachypodium distachyon-specific Ib2 clade. The clade containing the genes encoding C-terminal IQ67 domains was defined as Id. Notably, no members of Brachypodium distachyon were detected in this clade, suggesting that Brachypodium distachyon IQD family lost its members of this subgroup during the long period of evolution. Both IQD II and IQD III subfamilies were divided into two subclasses, a and b, which were designated as described by Zejun et al.(2013) and Abel et al. (2005) [35]. The C-terminally truncated IQ67 domain-containing genes (At IQD33, OsIQD28 and BdIQD14) comprise IQD V subfamily (Figure 4) [30], [34].

The combined phylogenetic tree reveals that most genes in the IQD family, especially the duplicated genes, are contained in paralogous pairs in each species, which supports the occurrence of species-specific IQD gene duplication events. By contrast, we identified 20 pairs of orthologous genes from monocotyledons (rice and Brachypodium distachyon) distributed among all of the subfamilies. In addition, two pairs of orthologous genes from dicotyledons (soybean and tomato) stemming from subfamily I (GmIQD56 and SlSUN9) and subfamily III (GmIQD60 and SlSUN13) were found. And AtIQD20 and OsIQD26, members of subfamily I, formed a pair of orthologous genes.

Conserved microsynteny of IQD III genes from soybean, Arabidopsis and tomato

The analysis of microsynteny provides valuable information for identifying gene expansion patterns and inferring gene orthology or paralogy. We combined genetic and phylogenetic analyses to perform microsynteny analysis of three dicotyledons, i.e., soybean, tomato and Arabidopsis.

To provide a basic framework for the identification of IQD III orthologous or paralogous genes, 44 IQD III genes, including 24 predicted soybean IQDs, 10 Arabidopsis IQDs and 10 tomato IQDs, were classified into four distinct clades, clade 1 (thirteen genes), clade 2 (five genes), clade 3 (eleven genes) and clade 4 (fifteen genes), based on phylogenetic analysis (Figure S2). Clade 1, 2 and 3 correspond to IQD IIIa and clade 4 corresponds to IQD IIIb (Figure 4 and S2). Each clade contains at least one gene from soybean, tomato and Arabidopsis, indicating that members from different species may be derived from a common ancestor.

Subsequently, to produce a comparative genetic map, 44 IQD III genes from the three dicot genomes were used as anchor genes. Conserved microsynteny was identified through reciprocal pairwise comparisons of the chromosomal regions containing IQD III genes. Microsynteny relationships among AtIQD3, AtIQD4, AtIQD5, GmIQD32, GmIQD56 or SlSUN13 with other IQD III members in these three dicot genomes were not observed. The map reveals that the 38 conserved syntenic segments diverged into four groups (Figure 5), which were anastomosed with the classification revealed by phylogenetic tree analysis.

thumbnail
Figure 5. Microsynteny map of IQD III genes in soybean, Arabidopsis and tomato.

a, b, c and d: four groups of syntenic segments. The relative positions of all flanking protein-coding genes were defined by the anchored IQD III genes (red). Colored horizontal lines represent chromosome segments of soybean (green), Arabidopsis (yellow) and tomato (purple). Triangles of the same color represent individual genes and their transcriptional orientations. The total number of genes on each segment is indicated to the right below the segment. Colored lines connect the conserved gene pairs among the segments (anchor gene pairs, red; others, blue).

https://doi.org/10.1371/journal.pone.0110896.g005

In clade 1 (Figure 5a), Map a shows a higher level of microsynteny, with both the same and opposite directions. SlSUN11/GmIQD21 and GmIQD53/GmIQD48 exhibit remarkable opposite-direction microsynteny, while GmIQD10/GmIQD18, GmIQD21/GmIQD28, AtIQD7/AtIQD8 and AtIQD8/SlSUN11 are aligned with flanking gene pairs in the same order but discordant transcriptional orientation. In addition, genes in map a were divided into two groups (Figure 5a), i.e., one group with higher levels of microsynteny (GmIQD21, GmIQD28, GmIQD53, GmIQD48, AtIQD7, AtIQD8 and AtIQD8 and SlSUN11) and the other group with lower levels of microsynteny (GmIQD61, GmIQD50, GmIQD10, GmIQD18, AtIQD6 and SlSUN22). These two groups were also detected in the phylogenetic tree of IQD III genes (Figure S2). In clade 2 (Figure 5b), two pairs from soybean and tomato, GmIQD41/SlSUN31 and GmIQD47/SlSUN31, exhibited microsynteny. However, the predicted duplicated pair GmIQD41/GmIQD47 had no detectable linkage with each other. High level of microsynteny exists in Clade 3, with most pairs in reverse order (Figure 5c), especially GmIQD42/GmIQD46, GmIQD45/GmIQD54 and GmIQD25/GmIQD44. GmIQD45/GmIQD44 and GmIQD44/SlSUN33 were identified as having same-direction microsynteny. In clade 4 (Figure 5d), we also observed a higher level of microsynteny. Except for AtIQD2/SlSUN29, SlSUN30/SlSUN29 and GmIQD67/GmIQD36, which are aligned in the opposite direction, most gene pairs in this clade have successive collinearity in order and the same orientation.

Two regions are considered to have originated from a large-scale duplication event when five or more protein-coding gene pairs flanking the anchor point are ligatured with the best non-self match (E-value <1e−10) [47], [48]. Applying this standard, except for the pair GmIQD41/GmIQD47, all soybean IQD III paralogous gene pairs were generated from a large-scale duplication event, which further supports the results of soybean gene duplication analysis (Figure 5, Table S4).

To estimate the degree of conserved gene content and order, the synteny quality was calculated [47]. The average synteny quality of IQD III genes from the three dicotyledons genomes was 18.41% (Table S5d). Due to the large number of syntenic genes between tomato and soybean, the synteny quality between these genomes is 26.39%; this value is higher than that observed in the Sl/At synteny blocks (16.68%). The lowest synteny quality (12.15%) was found between soybean and Arabidopsis (Table S5). Details of this comparative analysis are shown in Table S5.

Expression patterns of soybean IQD genes in various tissues

To gain insight into the expression patterns of soybean IQD genes in various tissues, we searched the RNA-Seq Atlas of Glycine max; this atlas provides high-resolution gene expression data from 14 diverse tissues, including aerial tissues (young leaf, flower, one-cm pod, pod-shell 10-DAF and pod-shell 14-DAF), underground tissues (root and nodule) and seed tissue at various stages of development (seed 10-DAF, seed 14-DAF, seed 21-DAF, seed 25-DAF, seed 28-DAF, seed 35-DAF and seed 42-DAF). Because the expression profiles of eight IQD genes (GmIQD17, -20, -25, -36, -42, -49, -56, -57) weren't obtained in the database, we only examined the expression patterns of fifty-nine IQD genes (Figure 6 and Table S6).

thumbnail
Figure 6. Hierarchical clustering of the expression profiles of soybean IQD genes in 14 tissues.

RNA-seq relative expression data from 14 tissues were used to reconstruct the expression patterns of soybean genes. The raw data was normalized and retrieved from the online database http://soybase.org/soyseq/. The normal relative expression levels of 67 IQD genes are shown in Table S6. YL, young leaf; F, flower; P.1cm, one cm pod; PS.10d, pod shell 10 DAF; PS.14d, pod shell 14 DAF; S.10d, seed 10 DAF; S.14d, seed 14 DAF; S.21d, seed 21 DAF; S.25d, seed 25 DAF; S.28d, seed 28 DAF; S.35d, seed 35 DAF; S.42d, seed 42 DAF; R, root; N, nodule. Gene names in different subfamilies are highlighted with various colors. Genes clustered into four groups (A–D) are indicated by the black vertical bars.

https://doi.org/10.1371/journal.pone.0110896.g006

Most soybean IQD genes exhibit broad expression patterns (Figure 6). Forty-four soybean IQDs are expressed in all of the seven tissues (young leaves, flowers, one-cm pod, pod-shell, roots, nodules and seed). The heat map also revealed that the majority of GmIQDs showed preferential expression. Based on a hierarchical clustering analysis, fifty-nine IQD genes were mainly clustered into four groups (A–D) (Figure 6). Group A showed partial expression in young leaves, group B in roots, group C in nodules and group D in flowers. Eight GmIQDs (GmIQD8, -13, -19, -29, -36, -52, -64 and -67) showed marked high transcript abundance profiles in only a single tissue. Among the fifty-nine soybean IQD genes examined, six showed the highest transcript accumulation in young leaves (GmIQD13, -18, -19, -29, -50 and -61), six in flowers (GmIQD11, -15, -38, -52 -65 and -67), one in roots (GmIQD26) and two in nodules (GmIQD8 and -64; Figure 6). Genes in different subfamilies have their primary abundant transcripts, for instance, GmIQD I in leaves, flowers and nodules, GmIQD II in flowers and roots, GmIQD III in young leaves, flowers and roots and GmIQD IV in roots and young leaves (Figure 6). These subfamily-specific tissue expression patterns may be closely related to gene functions. The expression patterns of the paralogous pairs were also revealed by heat maps; paralogous pairs with high sequence similarity have similar expression patterns. The best examples of this include GmIQD8/-64 and GmIQD6/-26, which are strongly expressed in nodules and root respectively, with little or no expression in other tissues. Expression divergence was also found in paralogous pairs. For example, GmIQD15 is highly expressed in flowers, while its paralog, GmIQD59, is highly expressed in nodules.

Examination of soybean IQD gene expression by qRT-PCR

Since soybean production is limited by stress, it is important to identify the master regulators of stress responses in soybean, as well as their regulatory pathways. According to microsynteny analysis, the high level of microsynteny indicates that IQD III genes existed before the divergence of the three dicotyledon genomes examined (soybean, tomato and Arabidopsis). In addition, IQD III genes in the same clade may share common ancestors and play similar roles in these species. AtIQD1, which belongs to the IQD III subfamily, plays a major role in the response to biotic stress, as it mediates the accumulation of glucosinolate in response to phytophagous insect attack. Jasmonic acid methyl ester (MeJA), the plant hormones and the signal molecules, widely exists in plants, which triggers expression of plants defense genes by exogenous applications and has similar effects with mechanical damage and insect herbivory [49], [50]. Based on these, we subjected 24 members of the soybean IQD III subfamily to real-time quantitative PCR (qRT-PCR) analysis to examine their regulation by MeJA.

The qRT-PCR results show that all 24 genes are MeJA-responsive, but some differences were observed among these genes (Figure 7). Although 23 genes were upregulated by MeJA treatment, GmIQD21 was obviously downregulated (<0.5 folds) at all time points. Eleven of the twenty-three upregulated GmIQD III genes exhibited minor changes in expression (relative expression scale from 0 to 5 and lower), including GmIQD10, -18, -21, -25, -42, -47, -53, -61, -7, -36 and -63. By contrast, 12 genes (GmIQD2, -32, -41, -44, -45, -46, -48, -50, -53, -54, -35, -43, -60 and -67) exhibited major changes in expression (relative expression scales from 0 to 5 up 0 to 80). The expression of six genes (GmIQD35, -36, -47, -54, -63 and -67) peaked relatively early (at 1 h of treatment); GmIQD54 and -67 were strongly upregulated (more than 26-fold and 34-fold, respectively). Eight genes (GmIQD10, -18, -28, -41, -42, -43, -53 and -60) were highly expressed at 4 h; GmIQD28 and -60 had the highest expression level more than 12-fold and GmIQD41 had the highest expression level more than 28-fold. While seven genes (GmIQD7, -25, -32, -44, -46, -48 and -61) exhibited the highest expression levels at 8 h; GmIQD48 were strongly induced by more than 35-fold. Only one gene (GmIQD45) had the highest expression level at 12 h, with a relative expression level approaching 70-fold. We also compared the expression profiles of paralogous pairs. Most paralogs in a pair had different expression profiles. For example, the expression of GmIQD28 peaked at 4 h while its sister gene, GmIQD21, was downregulated at all time points, suggesting that these genes may play diverse roles in the response to MeJA stress.

thumbnail
Figure 7. Expression patterns of 24 selected IQD III genes under MeJA stress using qRT-PCR.

Relative expression levels of 24 IQD genes were examined by qRT-PCR and normalized to the expression of CYP2. Bars represent standard deviations (SD) of three biological replicates. Y-axes indicate the scale of the relative expression levels. X-axes show time courses of MeJA stress treatments for each gene. a and b: data for genes from IIIa and IIIb, respectively.

https://doi.org/10.1371/journal.pone.0110896.g007

Discussion

Structural characteristics of IQD proteins

The plant-specific IQD gene family has previously been comprehensively analyzed in Arabidopsis, rice, tomato and Brachypodium distachyon, this gene family has not been previously identified and annotated in soybean. We identified and characterized 67 IQ67 domain-encoding genes in soybean using genome-wide analysis. The IQD gene family in soybean is by far the largest one compared to that in other plant species (33 in Arabidopsis, 29 in rice, 34 in tomato and 23 in Brachypodium distachyon). At ∼1,150 Mbp, with ∼46,400 predicted coding genes, soybean possesses 9.2-fold larger genome size and 1.75-fold higher gene count than Arabidopsis, which has a genome of 125 Mbp and ∼26,500 coding genes [51]. Given the obvious difference in genome size and estimated gene count between soybean and Arabidopsis, the IQD genes in soybean seems to be highly expanded. The presence of twice as many of these genes in soybean versus Arabidopsis may be mainly due to the recent polyploidy event and segmental duplication events in soybean's evolutionary history. It can be speculated that the presence of more IQD genes in soybean genome may reflect the great needs for these genes coding for calcium signal regulatory components with functions in plant development, defense response or others.

The common feature of IQ67 domain proteins is the arrangement of three IQ motifs separated by 11 and 15 intervening amino acid residues (Figure S1). To date, at least five protein families containing IQ motifs, which play a role in the calcium signaling pathway, have been identified in Arabidopsis. These protein families include the cyclic nucleotide gated channels family (CNGC), the IQ-Motif family (IQM), the CaM-binding transcriptional activator family (CAMTA), the myosin family and the IQD family, which contain one, one, two, five and up to three IQ motifs, respectively [52][54]. The unique spacing of IQ motifs and exon/intron organization of each family suggest that these IQD protein families represent separate classes of putative calmodulin targets. The calmodulin-interacting peptides in AtIQD20 and CNGC proteins, which were experimentally verified, were previously predicted using the algorithm provided by the Calmodulin Target Database successfully [30]. In the current study, using the Calmodulin Target Database, we detected calmodulin-binding sites in all soybean IQD proteins, which strongly suggests that all IQD proteins have the potential to interact with calmodulin (Figure 2 and Table 3). Three aspects of IQD proteins appear to underlie the mechanism of interaction between IQD proteins and calmodulin: the number and specific composition of the IQ, 1-5-10 and 1-8-14 motifs, the predicted calmodulin binding site and the overall tertiary structure of the IQD protein.

Of the 31 soybean IQD paralogs examined, 27 exhibit highly conserved exon-intron structures, which is consistent with the high degree of position and phase conservation broadly found across angiosperms [41]. In addition, the sizes of related introns between paralogs are also highly conserved, indicating that few insertions and deletions have accumulated within introns over the past 13 million years [41]. Most introns in GmIQD genes are in phase-0. This strong bias for phase-0 introns in soybean IQD genes is also found in IQD genes of Arabidopsis, rice and Brachypodium distachyon. The strong bias for one intron phase class, along with the variation in the number of exons (two to six) and the sizes of encoded proteins, suggests that exon shuffling has played a prominent role during the evolution and diversification of IQD genes [30].

65 of 67 soybean IQD proteins have relatively high isoelectric points with an average of 10.1. It is very similar to Arabidopsis (10.3), rice (10.4) and Brachypodium distachyon (10.3) [30], [34]. The extensive presence of the basic isoelectric point and high frequency of serine residues (Arabidopsis: ∼11%, rice: ∼11%, Brachypodium distachyon: ∼11.5% and soybean: ∼12%; Table S7) in IQDs suggest that the basic nature of IQDs is crucial to their biochemical functions [30], [34]. The high isoelectric points are evocative of RNA-binding proteins although IQD proteins don't comprise currently known RNA-binding motifs. Fifty-seven soybean IQD proteins are localized to the nucleus, because of their high content of basic residues revealed by Wolf PSORT. TargetP analysis revealed that fifteen soybean IQD proteins are located in the mitochondria by identifying the presence of mitochondrial targeting peptide (mTP). The contradicting subcellular localization predictions is due to the different algorithm used by Wolf PSORT and TargetP. Most soybean IQD protein members are likely to function in the nucleus, as nucleus specific Ca2+-signatures are reported to generate in plant cells [55][57] and calmodulin and related Ca2+sensor proteins may play a regulatory role in nuclear processes such as transcription [58], [59]. Observably, Arabidopsis IQD1 was reveraled to target to microtubules as well as the cell nucleus and nucleolus [32]. In vitro binding to single-stranded nucleic acids suggests AtIQD1 and other IQD family members may control and fine-tune gene expression and protein sorting by facilitating cellular RNA localization [32].

Phylogenetic analysis and evolution of IQD family genes

IQD proteins are an ancient family of CaM/CML binding proteins that originated during the early evolution of land plants, as IQD genes are present in Physcomitrella patens. ESTs corresponding to IQD proteins for angiosperm species (Arabidopsis, rice, etc.) and at least nine homologous sequences in the gymnosperm pine (Pinus ssp.) corresponding to IQD proteins were identified suggesting that the IQD gene family originated not later than the split of gymnosperms and angiosperms about 300 Myr ago [30]. We performed a genome-wide comparison of plant IQD members from monocots (rice and Brachypodium distachyon) and eudicots (soybean, Arabidopsis and tomato) to explore how the IQD gene family has evolved. The plant IQD members from monocots (rice and Brachypodium distachyon) and eudicots (soybean, Arabidopsis and tomato) appear to be more closely related to each other than to IQD genes of the same species in a different subfamily. This alternating distribution of monocots and eudicots in all subfamilies suggests that an ancestral set of IQD genes have existed before the dicot–monocot split (Figure 4, Table 4). The presence of five distinct subfamilies of IQD genes and the presence of both monocots and eudicots containing members in all five subfamilies indicate IQD genes have diversified before the monocot–eudicot split (Figure 4). These subfamilies include 23 pairs of orthologous genes, suggesting that orthologous genes may have originated from a common ancestor (Figure 4). About half of the orthologous genes (10 pairs; BdIQD1/OsIQD8:N, BdIQD5/OsIQD3:N, BdIQD8/OsIQD10:C, BdIQD9/OsIQD23:C, BdIQD11/OsIQD19:N, BdIQD14/OsIQD28:N, BdIQD17/OsIQD24:N BdIQD18/OsIQD7:N, BdIQD20/OsIQD2:N, SlSUN13/GmIQD60:N) have the same predicted subcellular localization suggesting that the encoded proteins may play similar roles in both species [30], [34]. A total of 87% (20 pairs) of orthologous gene pairs from rice and Brachypodium distachyon are distributed in all subfamilies. However, only two pairs of orthologous genes from dicotyledons (soybean and tomato) are from subfamily I and III. This difference may be due to the fact that both rice and Brachypodium distachyon are in the grass family and are therefore more closely related than Arabidopsis, soybean and tomato, which belong to Cruciferae, Solanaceae and Leguminosae, respectively. The number of soybean genes in each subfamily is greater than that of the other four species examined suggesting that IQD counterparts in soybean may have undergone gene expansion.

The duplication of individual genes, chromosomal segments or entire genomes has been a major force in the evolution of plant genome structure and content during the process of genome evolution [60], [61]. The soybean genome has undergone at least two round of duplication, resulting in the presence of significant features of remnants of a glycine-specific genome duplication that occurred ∼13 Mya and fainter remnants of older polyploidies prior to the divergence of the papilionoids (58–60 Mya) that occurred ∼58 Mya [41], [51]. Thus, 75% of soybean genes are present in multiple copies [41]. Among 67 soybean IQD genes, GmIQD3, -32, -56 and -60 were found as single copies on duplication blocks. These results suggest that segmental duplication has occurred as a continuous process and dynamic changes may have occurred in a chromosomal segment that contained two ancestral IQD genes, leading to corresponding sister gene loss [62]. One paralogs (GmIQD11/-39) shares 91.2% identity and similar exon/intron organization, but exists outside of any duplicated blocks. this pair might have been produced by retrotransposition. A high proportion (approximately 96%) of soybean IQD genes reside preferentially in duplicated segments, suggesting that segmental duplications have played a prominent role in the expansion of the soybean IQD gene family. The duplicated IQD genes in soybean have been preferentially retained at the high rate of 92.5% (62/67), which is distinctly higher than the retention rate (67.3%) of duplicated paralogs in the 1.1-gigabase sequence of the soybean (cv. Williams 82) genome, in which 31,264 genes exist as 15,632 paralog pairs (out of the 46,430 predicted high-confidence genes that were duplicated and retained after the 13-Mya tetraploidy event) [63]. The higher retention rate corroborates previous findings that genes involved in signal transduction are preferentially retained following duplications[64]. Our calculation of the duplication dates of the 31 paralogous pairs revealed that all of the segmental duplication events in the soybean IQD family occurred during the recent whole genome duplication event.

During evolution, eukaryotic genomes have retained genes on corresponding chromosomes (synteny) and in corresponding orders (collinearity) to various degrees. Synteny broadly refers to parallels in gene arrangement in dissimilar genomes. Collinearity, a specific form of synteny, requires genes to occur in largely corresponding orders along the chromosomes of respective genomes. According to the microsynteny analysis, microsynteny relationships among AtIQD3, AtIQD4 or SlSUN13 with other IQD III members in these three dicot genomes were not observed indicating that either these genes are ancient genes without detectable linkage to other IQD genes or that they were formed through complete transposition and loss of their primogenitors. In addition, three different duplicated chromosomal segments (harboring AtIQD5, GmIQD32 and GmIQD6) that lost their sister IQD genes lack detectable microsynteny relationships to all other IQD III genes in the soybean and Arabidopsis genome, respectively. In the four IQD III gene clades, genes from soybean, tomato and Arabidopsis exhibit high levels of microsynteny, which indicates the IQD III genes existed before the divergence of the three dicotyledons genomes (soybean, tomato and Arabidopsis). Microsynteny was detected in most pairs, and alignment in clade 1–3 was discordant, suggesting that these genes may all be present in genome regions that were inverted, expanded or contracted after the divergence. Notably, most gene pairs in clade 4 have successive collinearity in order and the same orientation, which indicates high conservation among these IQD III gene-residing regions, with little rearrangement. The low (18.41%) synteny quality of IQD III genes from the three dicotyledon genomes (soybean, tomato and Arabidopsis) may have been due to the fact that these plants are not closely related; moreover, the gene density differs between Arabidopsis and the two other species. Significantly, the number of synteny blocks (31) within the soybean genome is much more than the number (3 or 4) of synteny blocks between tomato or Arabidopsis genomes, which suggests that soybean IQD III genes may have undergone large-scale duplication events and less rearrangement was followed (Figure 5 and Table S5b). The gene expansion pattern analysis of soybean paralogs indicates that most pairs were generated from a large-scale duplication, which supports the results of soybean gene duplication analysis, with the exception of GmIQD41/-47.

Organ- or tissue-specific expression of IQD genes and expression of GmIQD III genes under MeJA stress treatment

Organ- or tissue-specific expression patterns have been observed for quite a few members of the IQD family. However, the functions of soybean IQD genes remain unclear. We therefore performed a thorough analysis of the RNA-Seq Atlas to investigate organ- or tissue-specific expression of IQD genes and qRT-PCR to examine the expression of GmIQD III members under MeJA stress treatment.

The tissue expression data deficiency of eight soybean IQD genes potentially indicated that these are pseudogenes or express only at specific developmental stages or under special conditions. 65.7% soybean IQD genes constitutively express in all of the seven tissues suggesting that GmIQDs may play roles at multiple developmental stages. Eight GmIQD proteins peak in only one tissue indicating that these tissue-specific calmodulin target proteins may be limited to discrete cells or organs to regulate various cellular activities.

Except for group C comprised of genes from GmIQD I, group A, B and D comprise genes from four subfamilies indicating these soybean IQD genes exhibit similar transcript abundance profiles but are relatively phylogenetically distinct. The analysis indicated that only some members within the same phylogenetic subgroup share a similar expression profile in soybean organs/tissues during development, excluding GmIQD6 and GmIQD26 belong to GmIQD IIb. For instance, GmIQD4, -34, -38 and -65 belong to GmIQD Ib clustered in group with high expression in flowers suggesting their potential roles in flower formation. While the other two GmIQD Ib members (GmIQD8 and -64) were detected in nodules indicating they may involve in fixing atmospheric nitrogen.

Members possessing similar sequences are clustered in the same subfamilies, which may have similar expression patterns or functions. In IQD subgroup Ia, Arabidopsis IQD22 is involved in the negative feedback regulation of GA-responsive DELLA genes [36]. Subgroup Ia members of Arabidopsis IQD26 has higher expression level in parts with divided vigorous growth and microtubule organization of leaves, root and flowers [65]. Eight of the of the thirteen soybean IQD Ia members (GmIQD3, -16, -22,-27, -29, -31 -37 and -66) have high expression in young fleaves. And six of them (GmIQD15, -16, -27, -31 -37 and -66) have high expression in one-cm pod or pod-shell 10-DAF. GmIQD22 have obviously higher expression at seed 21-DAF, the period of seed cell division (3, 4 weeks after flowering) [66]. These founding suggested that soybean IQD Ia members may function in transport of signaling molecules, nutrient transport and cell division.

Mapping and positional cloning of the SUN locus revealed that this member of the IQD II subfamily was generated by duplication of a 24.7-kb region carrying the tomato IQD12 gene, a major gene involved in the control of fruit shape, particularly length, in tomato [35]. SUN is expressed the highest in hypocotyl and shoot apex. Overexpression of SUN causes root reduction when applied auxin and prostrate growth and twisted stems indicating that SUN can affect auxin transport or response [38]. SlSUN1 shows slightly higher expression in the hypocotyl, flower at anthesis and fruit at 10 and 20 DPA [35]. SlSUN12 and SlSUN21 highly are expressed in the hypocotyls and root respectively [35]. SlSUN17 evenly expressed in almost all tissues [35]. Soybean IQD II members showed the similar expression profile (GmIQD11: highest in flowers; GmIQD33, -39 and -40: slight high in flowers and GmIQD6 and -26: highest in roots) indicating that GmIQD II members may play similar role in plant development.

Arabidopsis IQD1 from the IQD III subfamily modulates the expression of several glucosinolate (GS) pathway genes, resulting in the alteration of glucosinolate content and composition to promote resistance to herbivory. Arabidopsis IQD1, the first functionally characterized IQD gene, is expressed in vascular tissues of hypocotyls, leaves, stems, flowers and roots, as revealed by histochemical analysis. Expression pattern analysis of soybean IQD genes revealed that genes from the soybean IQD III subfamily were mainly expressed in young leaves, flowers and roots (Figure 6). Jasmonic acid(JA) treatment leads to elevating levels of specific glucosinolate in Arabidopsis [67], [68]. And overexpression of AtIQD1 causes the accumulation of glucosinolates. However, AtIQD1 expression is independent of JA, as steady-state AtIQD1 mRNA expression levels are not appreciably altered when externally applied JA and are also unaffected in mutants defective in JA synthesis or signaling (JA -jar1 and fad3-2 fad7-2 fad8) [31]. Indeed, AtIQD1 increases resistance against herbivory by augmenting and fine-tuning glucosinolate accumulation [31]. Glucosinolates with important roles in plant defense and human nutrition are a small but diverse class of defense related secondary metabolites in cruciferous species such as Brassica crops and the Arabidopsis thaliana [69], [70]. Obviously, soybean doesn't synthesize glucosinolates. Base on the microsynteny analysis of IQDIII members, we auspiciously found there exsited highly conserved microsynteny relationship between AtIQD1 and soybean IQDIII members. And combined with phylogenetic analysis of IQDIII members, we speculated soybean IQDIII members might have the similar biological function with AtIQD1 in defenses to insect herbivory.

Therefore, we performed qRT-PCR of 24 soybean IQD III with MeJA treatment to detect whether soybean IQD III genes defense to insect herbivory. Compared to AtIQD1, soybean IQD III genes exhibited different responses with the MeJA treatment. The qRT-PCR results showed that 23 of the 24 soybean IQD III subfamily genes were upregulated by MeJA treatment, except GmIQD21, seven genes (GmIQD28, -41, -45, -48, -54, -60 and -67) were strongly induced by MeJA with relative expression more than 10-fold. GmIQD45 even accumulated the highest transcripts approaching 70-fold at 12 h (Figure 7). Based on these, we speculate that IQD III genes in soybean may involve in defense to insect herbivory by JA pathway.

Orthologs may have equivalent functions, as they originated from a single ancestral gene in the last common ancestor of the species. Two pairs of orthologous genes (SlSUN22/AtIQD6 and SlSUN31/AtIQD5) were found between tomato and Arabidopsis (Figure S2). Similar expression patterns of these two pairs in tomato and Arabidopsis have been reported; SlSUN31 and AtIQD5 are almost ubiquitously expressed, whereas SlSUN22 and AtIQD6 are highly expressed in young flower buds [30], [34], [35].

Duplicated genes may face three different fates: nonfunctionalization (one copy becomes silenced); neofunctionalization (one copy acquires a novel, beneficial function while the other copy retains the original function) or subfunctionalization (both copies become partially compromised by the accumulation of mutations) [45], [71]. Paralogs originating from duplication within one organism may have more divergent functions. In the current study, several pairs of paralogs showed similar expression patterns, which suggests that they may share a common or similar function. For example, GmIQD10/GmIQD18 were highly expressed in young leaves, and their expression peaked at 4 h in response to MeJA (Figure 7). Several pairs of paralogs have different expression patterns, suggesting that they play diverse roles in soybean development. For example, GmIQD21/GmIQD28 are mainly expressed in young leaves. Upon MeJA treatment, GmIQD28 was most highly expressed at 4 h while its sister gene GmIQD21 was downregulated at all time points.

In conclusion, IQD proteins play fundamental roles in various plant developmental processes. Therefore, the systematic analysis of the soybean IQD gene family performed in the current study provides an important reference for further characterization of the biological functions of these proteins.

Materials and Methods

Identification of IQD family genes in soybean

To identify IQD proteins in soybean, the Glycine max genome database (release 1.0, http://www.phytozome.net/soybean.php) was searched using Basic Local Alignment Search Tool algorithms (BLASTP), with the published Arabidopsis IQD protein sequences and their IQ67 domain used as initial query sequences. Redundant sequences were then removed manually, and the Hidden Markov Model of the Pfam (http://pfam.sanger.ac.uk/search) [72] and SMART (http://smart.embl-heidelberg.de/) [73] databases were used to confirm each candidate sequence as a member of the IQD family [74]. A total of 33 Arabidopsis, one moss (Physcomitrella patens) and 23 Brachypodium IQD protein sequences were downloaded from Phytozome v9.0 (http://www.phytozome.net/), and 34 tomato IQD protein sequences were retrieved from the tomato WGS chromosomes (2.40; SL2.40) (SGNhttp://solgenomics.net). Finally, 27 rice IQD protein sequences were obtained from the TAIR database (http://rice.plantbiology.msu.edu). Accession numbers of published IQD proteins for Arabidopsis, rice, tomato, Brachypodium and moss were listed in Table S8.

Soybean IQD gene information, including the number of amino acids, ORF lengths and chromosome locations, was obtained from the Phytozome database. Physicochemical parameters including the molecular weight (kDa) and isoelectric point (pI) of each gene product were calculated using compute pI/Mw tool from ExPASy (http://www.expasy.org/tools/) and parameter (resolution) was set to average [75]. Subcellular localization was predicted using the TargetP 1.1 (http://www.cbs.dtu.dk/services/TargetP/) server and WoLF PSORT (http://wolfpsort.org/).

Multiple alignment and phylogenetic analysis of IQD family genes

Multiple sequence alignment of all predicted soybean IQD protein sequences was performed with Clustal X2.0 software using default parameters. Then, based on this alignment, phylogenetic trees were constructed using Clustal X2.0 with the Neighbor-Joining (NJ) method, and bootstrap analysis was conducted using 1,000 replicates [76]. An unrooted NJ tree of 184 the full-length IQD protein sequences from soybean, rice, Arabidopsis, tomato, Brachypodium was constructed using Clustal X2.0 with one moss IQD protein (Pp1s38230v6.1) as the outgroup. The GmIQD genes were classified according to their phylogenetic relationships with the corresponding Arabidopsis and rice IQD genes. For microsynteny analysis of IQD III genes across soybean, tomato and Arabidopsis, a phylogenetic tree was constructed using MEGA 5.0 with the NJ method and bootstrap analysis was conducted using 1,000 replicates.

Genomic structure

Exon and intron structures of individual soybean IQD genes were deduced using GSDS (Gene structure display server; http://gsds.cbi.pku.edu.cn/) via alignment of the cDNAs with their corresponding genomic DNA sequences [77].

Identification of conserved motifs and putative calmodulin-binding sites

Online MEME (Multiple Expectation Maximization for Motif Elicitation) (http://meme.sdsc.edu/meme4_4_0/intro.html) was performed to identify the conserved motif structures encoded by GmIQD genes. The parameters were as followings: number of repetitions - any, maximum number of motifs -10, and the optimum motif width was constrained to between 6 and 200 residues. In addition, each structural motif was annotated using the Pfam (http://pfam.sanger.ac.uk/search) and SMART (http://smart.embl-heidelberg.de/) tools. All IQD protein sequences were examined against the Calmodulin Target Database (http://calcium.uhnres.utoronto.ca/ctdb/ctdb/home.html) to predict putative calmodulin-binding sites.

Chromosomal location and gene duplication

The chromosomal location image of GmIQD genes was generated by MapInspect (http://www.plantbreeding.wur.nl/uk/software_mapinspect.html) according to chromosomal position information provided in the Phytozome database. To identify tandem and segmental duplications, two genes in the same species located in the same clade of the phylogenetic tree were defined as coparalogs. The SoyBase browser (http://soybase.org/gb2/gbrowse/gmax1.01/) [78] was queried to detect the segmental duplication coordinates of the target genes. Coparalogs were deemed to result from segmental duplication if they were located on duplicated chromosomal blocks [79]. Paralogs were deemed to be tandem duplicated genes if two genes were separated by five or fewer genes in a 100-kb region [80]. The local alignment of two protein sequences was calculated using the Smith-Waterman algorithm (http://www.ebi.ac.uk/Tools/psa/).

Calculation of Ka/Ks Values

Amino acid sequences of each paralog pair were first aligned using Clustal X2.0. Then, the multiple sequence alignments of proteins and the corresponding cDNA sequences were converted to codon alignments using PAL2NAL (http://www.bork.embl.de/pal2nal/) [81]. Finally, the resulting codon alignment was subjected to calculation of synonymous (Ks) and non-synonymous (Ka) substitution rates using the CODEML program of PAML [82].

Based on a rate of 6.1×10−9 substitutions per site per year, divergence time (T) was calculated using the Ks value with the formula: T = Ks/(2×6.1×10−9)×10−6 Mya [45].

Microsynteny analysis

For microsynteny analysis, IQD III genes from soybean, tomato and Arabidopsis as the anchors were localized to specific target genomic regions. Then, all protein-coding sequences of 100 kb flanking each anchor point were compared by pairwise BLASTP analysis. The syntenic blocks used to construct synteny analysis maps of the IQD III genes were obtained from the Plant Genome Duplication Database (http://chibba.agtec.uga.edu/duplication/), a web service providing synteny information in terms of collinearity between chromosomes [83][85]. A synteny block is defined as a region where three or more conserved homologs are located within a 100-kb DNA stretch in both genomes. Two regions were considered to have originated from a large-scale duplication event when five or more protein-coding gene pairs flanking the anchor point were ligatured with the best non-self match (E-value<1e−10) [48]. The relative syntenic quality in a region was calculated from the sum of the total number of genes in both conserved gene regions (excluding retroelements and transposons and collapsing tandem duplications) [47].

RNA-Seq atlas analysis

To acquire the tissue-specific transcript data, a list of 67 GmIQD gene names was entered to the RNA-Seq Atlas of Glycine max (http://soybase.org/soyseq/) [86]. The raw digital gene expression counts of the uniquely mappable reads were normalized using a slight variation of the reads/Kb/Million (RPKM) method and the normalized data was download from this database [86]. Hierarchical clustering analysis was conducted using clustering distance “correlation” (Pearson correlation) and the clustering method used “complete” (complete linkage method) in R [87]. A heat map was generated in R using the pheatmap function [87].

Plant growth and treatments

Soybean (Glycine max L.) Williams 82 was used in this study. Seedlings were grown in a growth chamber under the following conditions: temperature, 30°C; photoperiod, 12 h/12 h; photon flux density, 80 µmolm−2 s−1 and relative humidity, 50% [88]. For expression pattern analysis of soybean IQD genes under stress, four-week-old seedlings were treated with 100 µm MeJA in the growth chamber [89]. Jasmonic acid methyl ester (MeJA) (Sigma, 95%) was diluted 1∶10 with 95% ethanol, followed by a further dilution with MilliQ water containing 0.1% Triton X-100, resulting in a final concentration of 100 µmol/L MeJA. Untreated seedlings were used as a control. Leaves of MeJA-treated plants were collected at 0, 1, 2, 4, 8 and 12 h. After collection, the samples were immediately frozen in liquid N2 and stored at −80°C for RNA extraction. Three biological replicates were conducted per sample.

RNA extraction and qRT-PCR analysis

An RNAprep Pure Plant Kit (Tiangen) was used to isolate total RNA from each frozen sample. Possible contaminating genomic DNA was removed using DNaseI supplied in the kit. First-strand cDNA was synthesized from the RNA using a PrimeScript™ RT Master Mix Kit (TaKaRa) according to the manufacturer's instructions. Gene-specific primers for the 24 GmIQD genes were designed using Primer5.0 (Table S9). Primer specificity was first checked using the primer-BLAST tool available on the NCBI website. Subsequently, by performing analysis of melting curves and analysis of visualization of amplicon fragments, we found primers were gene-specific only when corresponding melting curves generated a single sharp peak and the primers demonstrated an electrophoresis pattern of a single amplicon with the correct predicted length. A housekeeping gene constitutively expressed in soybean, CYP2 (cyclophilin) [46], [90][93], was used as a reference for normalization. The qRT-PCR analysis was conducted on an ABI 7500 Real-Time PCR system (Applied Biosystems). The reactions were performed in a 20 µl volume containing 10 µl 2×SYBR® Premix Ex Taq™ II (TaKaRa), 6.0 µl ddH2O, 0. 4 µl ROX Reference Dye II, 2.0 µl diluted cDNA and 0.8 µl of each gene-specific primer. The PCR conditions were as follows: Stage 1: 95°C for 30 s; stage 2: 40 cycles of 5 s at 95°C and 34 s at 60°C; stage 3: 95°C for 15 s, 60°C for 1 min, 95°C for 15 s. At stage 3, a melting curve was generated to estimate the specificity of the reactions. Three biological replicates were used per sample.

The relative expression levels were calculated as 2−ΔΔCT [△CT  =  CT, Target - CT, CYP2. △△CT = △CT, treatment - △CT, CK (0 h)]. The relative expression levels (2− ΔΔCT, CK (0 h)) in the untreated control plants were normalized to 1 as described previously [46], [94], [95]. If an efficiency of amplification was less than 2, the result was proofread. Statistical analyses were conducted using GraphPad Prism 5.01 software [96].

Supporting Information

Figure S1.

Amino acid sequence alignments of IQ67 domains in soybean IQD protein sequences. The multiple alignment results indicate the highly conserved IQD domains among the 67 identified soybean IQD protein sequences. The positions of the conserved IQ calmodulin binding motifs are shown. Identical residues of proteins are marked with an asterisk. The consensus sequence at the bottom was constructed with greater than 50% conservation among the 67 soybean IQD proteins. Red arrow indicates the position of the conserved phase-0 intron, which divides codons 16 and 17 of the IQ67 domain.

https://doi.org/10.1371/journal.pone.0110896.s001

(TIF)

Figure S2.

Phylogenetic tree of full-length IQD III proteins from soybean, Arabidopsis and tomato. The tree was generated with MEGA 5.0 using the NJ method with 1,000 bootstrap replicates. Dicotyledon (soybean, tomato and Arabidopsis) IQD proteins are marked with colored dots. IQD III proteins from soybean, Arabidopsis and tomato were divided into four clades (1–4) presented by different color.

https://doi.org/10.1371/journal.pone.0110896.s002

(TIF)

Table S1.

Detailed information about the 10 motifs in soybean IQD proteins.

https://doi.org/10.1371/journal.pone.0110896.s003

(XLS)

Table S2.

Recent synteny blocks of soybean and soybean (13 Mya) genomes containing IQD genes.

https://doi.org/10.1371/journal.pone.0110896.s004

(XLS)

Table S3.

Pairwise identities between paralogous pairs of IQD genes from soybean.

https://doi.org/10.1371/journal.pone.0110896.s005

(XLS)

Table S4.

The synteny blocks used to construct microsynteny map.

https://doi.org/10.1371/journal.pone.0110896.s006

(XLS)

Table S5.

Number of conserved gene pairs and synteny blocks and relative syntenic quality.

https://doi.org/10.1371/journal.pone.0110896.s007

(XLS)

Table S6.

Transcription of soybean IQD genes, as determined by RNA-seq analysis.

https://doi.org/10.1371/journal.pone.0110896.s008

(XLSX)

Table S7.

Animo acid content of 67 soybean IQD proteins.

https://doi.org/10.1371/journal.pone.0110896.s009

(XLS)

Table S8.

Accession numbers of IQDs for Arabidopsis thaliana, rice, tomato, Brachypodium distachyon and moss.

https://doi.org/10.1371/journal.pone.0110896.s010

(XLS)

Table S9.

List of primer sequences used for qRT-PCR analysis of the 24 soybean IQD III genes.

https://doi.org/10.1371/journal.pone.0110896.s011

(XLS)

Author Contributions

Conceived and designed the experiments: LF YX. Performed the experiments: LF XC. Analyzed the data: LF ZC. Contributed reagents/materials/analysis tools: LF XC YW. Wrote the paper: LF ZC YL.

References

  1. 1. Hepler PK, Vidali L, Cheung AY (2001) Polarized cell growth in higher plants. Annual review of cell and developmental biology 17: 159–187.
  2. 2. Du L, Ali GS, Simons KA, Hou J, Yang T, et al. (2009) Ca2+/calmodulin regulates salicylic-acid-mediated plant immunity. Nature 457: 1154–1158.
  3. 3. Harada A, Shimazaki K-i (2009) Measurement of changes in cytosolic Ca2+ in Arabidopsis guard cells and mesophyll cells in response to blue light. Plant and cell physiology 50: 360–373.
  4. 4. Ng C, Mcainsh MR, Gray JE, Hunt L, Leckie CP, et al. (2001) Calcium-based signalling systems in guard cells. New Phytologist 151: 109–120.
  5. 5. Reddy AS (2001) Calcium: silver bullet in signaling. Plant Science 160: 381–404.
  6. 6. Dodd AN, Kudla J, Sanders D (2010) The language of calcium signaling. Annu Rev Plant Biol 61: 593–620.
  7. 7. Day IS, Reddy VS, Shad Ali G, Reddy AS (2002) Analysis of EF-hand-containing proteins in Arabidopsis. Genome Biol 3: RESEARCH0056.
  8. 8. Evans NH, McAinsh MR, Hetherington AM (2001) Calcium oscillations in higher plants. Current opinion in plant biology 4: 415–420.
  9. 9. Harper JF (2001) Dissecting calcium oscillators in plant cells. Trends in plant science 6: 395–397.
  10. 10. Ali GS, Reddy VS, Lindgren PB, Jakobek JL, Reddy A (2003) Differential expression of genes encoding calmodulin-binding proteins in response to bacterial pathogens and inducers of defense responses. Plant Molecular Biology 51: 803–815.
  11. 11. Snedden WA, Fromm H (2001) Calmodulin as a versatile calcium signal transducer in plants. New Phytologist 151: 35–66.
  12. 12. Bouché N, Yellin A, Snedden WA, Fromm H (2005) Plant-specific calmodulin-binding proteins. Annu Rev Plant Biol 56: 435–466.
  13. 13. Luan S, Kudla J, Rodriguez-Concepcion M, Yalovsky S, Gruissem W (2002) Calmodulins and calcineurin B–like proteins calcium sensors for specific signal response coupling in plants. The Plant Cell Online 14: S389–S400.
  14. 14. Ranty B, Aldon D, Galaud J-P (2006) Plant calmodulins and calmodulin-related proteins: multifaceted relays to decode calcium signals. Plant Signal Behav 1: 96.
  15. 15. Zhang H, Yin W, Xia X (2008) Calcineurin B-Like family in Populus: comparative genome analysis and expression pattern under cold, drought and salt stress treatment. Plant Growth Regulation 56: 129–140.
  16. 16. Zhang C, Bian M, Yu H, Liu Q, Yang Z (2011) Identification of alkaline stress-responsive genes of CBL family in sweet sorghum (Sorghum bicolor L.). Plant Physiol Biochem 49: 1306–1312.
  17. 17. Yu Y, Xia X, Yin W, Zhang H (2007) Comparative genomic analysis of CIPK gene family in Arabidopsis and Populus. Plant Growth Regulation 52: 101–110.
  18. 18. Zuo R, Hu R, Chai G, Xu M, Qi G, et al. (2013) Genome-wide identification, classification, and expression analysis of CDPK and its closely related gene families in poplar (Populus trichocarpa). Mol Biol Rep 40: 2645–2662.
  19. 19. Reddy AS, Ali GS, Celesnik H, Day IS (2011) Coping with stresses: roles of calcium- and calcium/calmodulin-regulated gene expression. Plant Cell 23: 2010–2032.
  20. 20. Batistic O, Kudla J (2004) Integration and channeling of calcium signaling through the CBL calcium sensor/CIPK protein kinase network. Planta 219: 915–924.
  21. 21. Boonburapong B, Buaboocha T (2007) Genome-wide identification and analyses of the rice calmodulin and related potential calcium sensor proteins. BMC Plant Biol 7: 4.
  22. 22. Perochon A, Aldon D, Galaud J-P, Ranty B (2011) Calmodulin and calmodulin-like proteins in plant calcium signaling. Biochimie 93: 2048–2053.
  23. 23. DeFalco T, Bender K, Snedden W (2010) Breaking the code: Ca2+ sensors in plant signalling. Biochem J 425: 27–40.
  24. 24. Chang'en T, Yuping Z (2014) Research Progress in Plant IQ Motif-containing Calmodulin-binding Proteins. Chinese Bulletin of Botany 48: 447–460.
  25. 25. Fischer C, Kugler A, Hoth S, Dietrich P (2013) An IQ domain mediates the interaction with calmodulin in a plant cyclic nucleotide-gated channel. Plant and cell physiology 54: 573–584.
  26. 26. Bähler M, Rhoads A (2002) Calmodulin signaling via the IQ motif. FEBS letters 513: 107–113.
  27. 27. Hoeflich KP, Ikura M (2002) Calmodulin in action: diversity in target recognition and activation mechanisms. Cell 108: 739–742.
  28. 28. Bhattacharya S, Bunick CG, Chazin WJ (2004) Target selectivity in EF-hand calcium binding proteins. Biochimica et Biophysica Acta (BBA)-Molecular Cell Research 1742: 69–79.
  29. 29. Clapperton JA, Martin SR, Smerdon SJ, Gamblin SJ, Bayley PM (2002) Structure of the complex of calmodulin with the target sequence of calmodulin-dependent protein kinase I: studies of the kinase activation mechanism. Biochemistry 41: 14669–14679.
  30. 30. Abel S, Savchenko T, Levy M (2005) Genome-wide comparative analysis of the IQD gene families in Arabidopsis thaliana and Oryza sativa. BMC Evol Biol 5: 72.
  31. 31. Levy M, Wang Q, Kaspi R, Parrella MP, Abel S (2005) Arabidopsis IQD1, a novel calmodulin-binding nuclear protein, stimulates glucosinolate accumulation and plant defense. Plant J 43: 79–96.
  32. 32. Burstenbinder K, Savchenko T, Muller J, Adamson AW, Stamm G, et al. (2012) Arabidopsis Calmodulin-binding Protein IQ67-Domain 1 Localizes to Microtubules and Interacts with Kinesin Light Chain-related Protein-1. Journal of Biological Chemistry 288: 1871–1882.
  33. 33. Abel S, Burstenbinder K, Muller J (2013) The emerging function of IQD proteins as scaffolds in cellular signaling and trafficking. Plant Signal Behav 8: e24369.
  34. 34. Filiz E, Tombuloglu H, Ozyigit II (2013) Genome-wide analysis of IQ67 domain (IQD) gene families in Brachypodium distachyon. Plant Omics 6: 425–432.
  35. 35. Huang Z, Van Houten J, Gonzalez G, Xiao H, van der Knaap E (2013) Genome-wide identification, phylogeny and expression analysis of SUN, OFP and YABBY gene family in tomato. Mol Genet Genomics 288: 111–129.
  36. 36. Zentella R, Zhang ZL, Park M, Thomas SG, Endo A, et al. (2007) Global analysis of della direct targets in early gibberellin signaling in Arabidopsis. Plant Cell 19: 3037–3057.
  37. 37. Xiao H, Jiang N, Schaffner E, Stockinger EJ, van der Knaap E (2008) A retrotransposon-mediated gene duplication underlies morphological variation of tomato fruit. Science 319: 1527–1530.
  38. 38. Wu S, Xiao H, Cabrera A, Meulia T, van der Knaap E (2011) SUN regulates vegetative and reproductive organ shape by changing cell division patterns. Plant Physiol 157: 1175–1186.
  39. 39. Kamenetzky L, Asis R, Bassi S, de Godoy F, Bermudez L, et al. (2010) Genomic analysis of wild tomato introgressions determining metabolism- and yield-associated traits. Plant Physiol 152: 1772–1786.
  40. 40. Soto-Valdez H, Colin-Chavez C, Peralta E. Fabrication and Properties of Antioxidant Polyethylene-based Films Containing Marigold (Tagetes erecta) Extract and Application on Soybean Oil Stability; 2012. DEStech Publications, Inc. pp. 206.
  41. 41. Schmutz J, Cannon SB, Schlueter J, Ma J, Mitros T, et al. (2010) Genome sequence of the palaeopolyploid soybean. Nature 463: 178–183.
  42. 42. Silva DC, Yamanaka N, Brogin RL, Arias CA, Nepomuceno AL, et al. (2008) Molecular mapping of two loci that confer resistance to Asian rust in soybean. Theor Appl Genet 117: 57–63.
  43. 43. Helfer S (2014) Rust fungi and global change. New Phytologist 201: 770–780.
  44. 44. Nekrutenko A, Makova KD, Li W-H (2002) The KA/KS ratio test for assessing the protein-coding potential of genomic regions: an empirical and simulation study. Genome research 12: 198–202.
  45. 45. Lynch M, Conery JS (2000) The evolutionary fate and consequences of duplicate genes. Science 290: 1151–1155.
  46. 46. Chen X, Chen Z, Zhao H, Zhao Y, Cheng B, et al. (2014) Genome-Wide Analysis of Soybean HD-Zip Gene Family and Expression Profiling under Salinity and Drought Treatments. PloS one 9: e87156.
  47. 47. Cannon SB, McCombie WR, Sato S, Tabata S, Denny R, et al. (2003) Evolution and microsynteny of the apyrase gene family in three legume genomes. Molecular Genetics and Genomics 270: 347–361.
  48. 48. Zhang X, Feng Y, Cheng H, Tian D, Yang S, et al. (2011) Relative evolutionary rates of NBS-encoding genes revealed by soybean segmental duplication. Molecular Genetics and Genomics 285: 79–90.
  49. 49. Creelman RA, Tierney ML, Mullet JE (1992) Jasmonic acid/methyl jasmonate accumulate in wounded soybean hypocotyls and modulate wound gene expression. Proceedings of the National Academy of Sciences 89: 4938–4941.
  50. 50. HU W-z, JIANG A-l, YANG H, LIU C-h, HE Y-b (2012) Effect of Jasmonic acid methyl ester treatment on the physiological and biochemical reactions of fresh-cut apple. Science and Technology of Food Industry 16: 086.
  51. 51. Cannon SB, Shoemaker RC (2012) Evolutionary and comparative analyses of the soybean genome. Breeding science 61.
  52. 52. Reddy A, Day IS (2001) Analysis of the myosins encoded in the recently completed Arabidopsis thaliana genome sequence. Genome Biol 2: 1–17.
  53. 53. Gao F, Han X, Wu J, Zheng S, Shang Z, et al. (2012) A heat-activated calcium-permeable channel–Arabidopsis cyclic nucleotide-gated ion channel 6–is involved in heat shock responses. The Plant Journal 70: 1056–1069.
  54. 54. Bürstenbinder K, Savchenko T, Müller J, Adamson AW, Stamm G, et al. (2013) Arabidopsis Calmodulin-binding Protein IQ67-Domain 1 Localizes to Microtubules and Interacts with Kinesin Light Chain-related Protein-1. Journal of Biological Chemistry 288: 1871–1882.
  55. 55. Pauly N, Knight MR, Thuleau P, van der Luit AH, Moreau M, et al. (2000) Cell signalling: control of free calcium in plant cell nuclei. Nature 405: 754–755.
  56. 56. Xiong TC, Jauneau A, Ranjeva R, Mazars C (2004) Isolated plant nuclei as mechanical and thermal sensors involved in calcium signalling. The Plant Journal 40: 12–21.
  57. 57. Lecourieux D, Ranjeva R, Pugin A (2006) Calcium in plant defence-signalling pathways. New Phytologist 171: 249–269.
  58. 58. Anandalakshmi R, Marathe R, Ge X, Herr J, Mau C, et al. (2000) A calmodulin-related protein that suppresses posttranscriptional gene silencing in plants. Science 290: 142–144.
  59. 59. Yoo JH, Park CY, Kim JC, Do Heo W, Cheong MS, et al. (2005) Direct interaction of a divergent CaM isoform and the transcription factor, MYB2, enhances salt tolerance in Arabidopsis. Journal of Biological Chemistry 280: 3697–3706.
  60. 60. Du J, Tian Z, Sui Y, Zhao M, Song Q, et al. (2012) Pericentromeric effects shape the patterns of divergence, retention, and expression of duplicated genes in the paleopolyploid soybean. The Plant Cell Online 24: 21–32.
  61. 61. Ohno S (1970) Evolution by gene duplication: London: George Alien & Unwin Ltd. Berlin, Heidelberg and New York: Springer-Verlag.
  62. 62. Schlueter JA, Dixon P, Granger C, Grant D, Clark L, et al. (2004) Mining EST databases to resolve evolutionary events in major crop species. Genome 47: 868–876.
  63. 63. Shoemaker RC, Schlueter J, Doyle JJ (2006) Paleopolyploidy and gene duplication in soybean and other legumes. Current opinion in plant biology 9: 104–109.
  64. 64. Dreze M, Carvunis A-R, Charloteaux B, Galli M, Pevzner SJ, et al. (2011) Evidence for network evolution in an Arabidopsis interactome map. Science 333: 601–607.
  65. 65. Su-Juan WH-YGZ-QWZ-JLZ-WWZ-JC Isolation and Characterization of Calmodulin-binding Protein AtIQD26 in Arabidopsis thaliana. Progress in Biochemistry and Biophysics 35: 703–711.
  66. 66. Hajduch M, Ganapathy A, Stein JW, Thelen JJ (2005) A systematic proteomic study of seed filling in soybean. Establishment of high-resolution two-dimensional reference maps, expression profiles, and an interactive proteome database. Plant Physiol 137: 1397–1419.
  67. 67. Brader G, Tas É, Palva ET (2001) Jasmonate-dependent induction of indole glucosinolates in arabidopsis by culture filtrates of the nonspecific pathogenerwinia carotovora. Plant Physiology 126: 849–860.
  68. 68. Cipollini D, Enright S, Traw M, Bergelson J (2004) Salicylic acid inhibits jasmonic acid-induced resistance of Arabidopsis thaliana to Spodoptera exigua. Molecular Ecology 13: 1643–1653.
  69. 69. Fahey JW, Zalcmann AT, Talalay P (2001) The chemical diversity and distribution of glucosinolates and isothiocyanates among plants. Phytochemistry 56: 5–51.
  70. 70. Wittstock U, Halkier BA (2002) Glucosinolate research in the Arabidopsis era. Trends in plant science 7: 263–270.
  71. 71. Dittmar K, Liberles D (2011) Evolution after gene duplication: John Wiley & Sons.
  72. 72. Finn RD, Mistry J, Schuster-Böckler B, Griffiths-Jones S, Hollich V, et al. (2006) Pfam: clans, web tools and services. Nucleic acids research 34: D247–D251.
  73. 73. Letunic I, Copley RR, Schmidt S, Ciccarelli FD, Doerks T, et al. (2004) SMART 4.0: towards genomic data integration. Nucleic Acids Research 32: D142–D144.
  74. 74. Bateman A, Coin L, Durbin R, Finn RD, Hollich V, et al. (2004) The Pfam protein families database. Nucleic acids research 32: D138–D141.
  75. 75. Gasteiger E, Gattiker A, Hoogland C, Ivanyi I, Appel RD, et al. (2003) ExPASy: the proteomics server for in-depth protein knowledge and analysis. Nucleic acids research 31: 3784–3788.
  76. 76. Hu R, Qi G, Kong Y, Kong D, Gao Q, et al. (2010) Comprehensive analysis of NAC domain transcription factor gene family in Populus trichocarpa. BMC plant biology 10: 145.
  77. 77. Guo A-Y, Zhu Q-H, Chen X, Luo J-C (2007) GSDS: a gene structure display server]. Yi chuan =  Hereditas/Zhongguo yi chuan xue hui bian ji 29: 1023.
  78. 78. Grant D, Nelson RT, Cannon SB, Shoemaker RC (2010) SoyBase, the USDA-ARS soybean genetics and genomics database. Nucleic acids research 38: D843–D846.
  79. 79. Wei F, Coe E, Nelson W, Bharti AK, Engler F, et al. (2007) Physical and genetic structure of the maize genome reflects its complex evolutionary history. PLoS Genetics 3: e123.
  80. 80. Wang L, Guo K, Li Y, Tu Y, Hu H, et al. (2010) Expression profiling and integrative analysis of the CESA/CSL superfamily in rice. BMC plant biology 10: 282.
  81. 81. Suyama M, Torrents D, Bork P (2006) PAL2NAL: robust conversion of protein sequence alignments into the corresponding codon alignments. Nucleic acids research 34: W609–W612.
  82. 82. Yang Z (2007) PAML 4: phylogenetic analysis by maximum likelihood. Molecular biology and evolution 24: 1586–1591.
  83. 83. Lee T-H, Tang H, Wang X, Paterson AH (2013) PGDD: a database of gene and genome duplication in plants. Nucleic acids research 41: D1152–D1158.
  84. 84. Tang H, Bowers JE, Wang X, Ming R, Alam M, et al. (2008) Synteny and collinearity in plant genomes. Science 320: 486–488.
  85. 85. Tang H, Wang X, Bowers JE, Ming R, Alam M, et al. (2008) Unraveling ancient hexaploidy through multiply-aligned angiosperm gene maps. Genome research 18: 1944–1954.
  86. 86. Severin AJ, Woody JL, Bolon Y-T, Joseph B, Diers BW, et al. (2010) RNA-Seq Atlas of Glycine max: a guide to the soybean transcriptome. BMC plant biology 10: 160.
  87. 87. Team RC (2012) R: A language and environment for statistical computing.
  88. 88. Hyun TK, Eom SH, Jeun YC, Han SH, Kim J-S (2013) Identification of glutamate decarboxylases as a γ-aminobutyric acid (GABA) biosynthetic enzyme in soybean. Industrial Crops and Products 49: 864–870.
  89. 89. Cheng Q, Zhang B, Zhuge Q, Zeng Y, Wang M, et al. (2006) Expression profiles of two novel lipoxygenase genes in Populus deltoides. Plant Science 170: 1027–1035.
  90. 90. Gutierrez N GMJ, Palomino C, et al.. (2011) Assessment of candidate reference genes for expression studies in Vicia faba L. by real-time quantitative PCR. Molecular Breeding: 13–24.
  91. 91. Vívian de Jesus Miranda RRC, Antônio Américo Barbosa Viana , Osmundo Brilhante de Oliveira Neto , Regina Maria Dechechi Gomes Carneiro , Thales Lima Rocha , Maria Fatima Grossi de Sa , Rodrigo Rocha Fragoso (2013) Validation of reference genes aiming accurate normalization of qPCR data in soybean upon nematode parasitism and insect attack. BMC Research Notes 6..
  92. 92. Le DT, Nishiyama R, Watanabe Y, Mochida K, Yamaguchi-Shinozaki K, et al. (2011) Genome-wide expression profiling of soybean two-component system genes in soybean root and shoot tissues under dehydration stress. DNA Res 18: 17–29.
  93. 93. Jian B, Liu B, Bi Y, Hou W, Wu C, et al. (2008) Validation of internal control for gene expression study in soybean by quantitative real-time PCR. BMC Mol Biol 9: 59.
  94. 94. Livak KJ, Schmittgen TD (2001) Analysis of Relative Gene Expression Data Using Real-Time Quantitative PCR and the 2− ΔΔCT Method. methods 25: 402–408.
  95. 95. Peng X, Zhao Y, Cao J, Zhang W, Jiang H, et al. (2012) CCCH-type zinc finger family in maize: genome-wide identification, classification and expression profiling under abscisic acid and drought treatments. PLoS One 7: e40120.
  96. 96. Bryfczynski SP, Pargas RP (2009) GraphPad: a graph creation tool for CS2/CS7. ACM SIGCSE Bulletin 41: 389–389.