Skip to main content

ORIGINAL RESEARCH article

Front. Plant Sci., 21 June 2019
Sec. Plant Breeding

Identification of Loci and Candidate Genes Responsible for Pod Dehiscence in Soybean via Genome-Wide Association Analysis Across Multiple Environments

\r\nDezhou Hu&#x;Dezhou Hu1†Guizhen Kan&#x;Guizhen Kan1†Wei HuWei Hu1Yali LiYali Li1Derong HaoDerong Hao2Xiao LiXiao Li1Hui Yang,Hui Yang1,3Zhongyi YangZhongyi Yang1Xiaohong HeXiaohong He1Fang Huang*Fang Huang1*Deyue Yu,*Deyue Yu1,3*
  • 1National Center for Soybean Improvement, National Key Laboratory of Crop Genetics and Germplasm Enhancement, Jiangsu Collaborative Innovation Center for Modern Crop Production, Nanjing Agricultural University, Nanjing, China
  • 2Jiangsu Yanjiang Institute of Agricultural Sciences, Nantong, China
  • 3School of Life Sciences, Guangzhou University, Guangzhou, China

Pod dehiscence (shattering) is the main cause of serious yield loss during the soybean mechanical harvesting process. A better understanding of the genetic architecture and molecular mechanisms of pod dehiscence is of great significance for soybean breeding. In this study, genome-wide association analysis (GWAS) with NJAU 355K SoySNP array was performed to detect single nucleotide polymorphisms (SNPs) associated with pod dehiscence in an association panel containing 211 accessions across five environments. A total of 163 SNPs were identified as significantly associated with pod dehiscence. Among these markers, 136 SNPs identified on chromosome 16 were located in the known QTL qPDH1. One, one, three, eleven, three, one, three, three and one SNPs were distributed on chromosome 1, 4, 6, 8, 9, 11, 17, 18, and 20, respectively. Favorable SNPs and six haplotypes were identified based on ten functional SNPs; among those Hap2 and Hap3 were considered as optimal haplotypes. In addition, based on GWAS results, the candidate gene Glyma09g06290 was identified. Quantitative real-time PCR (qRT-PCR) results and polymorphism analysis suggested that Glyma09g06290 might be involved in pod dehiscence. Furthermore, a derived cleaved amplified polymorphic sequences (dCAPS) marker for Glyma09g06290 was developed. Overall, the loci and genes identified in this study will be helpful in breeding soybean accessions resistant to pod dehiscence.

Introduction

Soybean [Glycine max (L.) Merr.] is one of the major oil crops and provides edible oil and abundant protein for human beings. Soybean pod dehiscence (shattering) is when mature pods open along their dorsal or ventral sutures to release seeds (Kang et al., 2005). As pod dehiscence leads to significant yield loss in soybean and may cause more damage to soybean production as climatic conditions become harsher, breeding for soybean varieties with pod-shattering resistance is always one of the important goals for breeders.

In the past two decades, a few QTLs or genes controlling pod dehiscence have been identified. Previously, QTL analysis in cultivated soybean has revealed a possible major QTL and several minor QTLs that regulate pod dehiscence. Bailey et al. (1997) first identified the major QTL by using 140 restriction fragment length polymorphism (RFLP) markers in a population of 120 F4-derived lines (Bailey et al., 1997). Funatsuki et al. (2006) detected this major QTL and named it qPDH1; it was located between the markers Sat_093 and Sat_366 on chromosome 16 and accounted for 50% of the total variance (Funatsuki et al., 2006). Furthermore, this QTL was also reported by several other studies that used different populations and genetic backgrounds (Liu et al., 2007; Kang et al., 2009; Yamada et al., 2009; Suzuki et al., 2010; Ju et al., 2017), suggesting that qPDH1 may be the major QTL associated with pod dehiscence. In addition, other QTLs that were deemed to be minor were also mapped (Bailey et al., 1997; Liu et al., 2007; Kang et al., 2009; Yamada et al., 2009).

Pod dehiscence/seed shattering has been deeply investigated in many crops (Vittori et al., 2019), several genes has been identified in rice (Konishi et al., 2006; Zhou et al., 2012), common bean (Koinange et al., 1996; Rau et al., 2019), Medicago (Fourquin et al., 2013), and cowpea (Suanum et al., 2016; Lo et al., 2018). In soybean, studies predicted that two to four genes were responsible for pod dehiscence (Caviness, 1969; Tsuchiya, 1986, 1987; Kang et al., 2009). To date, two genes located on chromosome 16 have been cloned and characterized. A gene, Pdh1 (Glyma16g25580), detected by fine mapping (located at the qPDH1 QTL) and encoding a dirigent-like protein, promotes pod dehiscence by increasing the torsion of dried pod walls (Funatsuki et al., 2014). In addition, Dong et al. (2014) identified a NAC gene, SHAT1-5 (Glyma16g02200) by analyzing the changes due to domestication between cultivated soybean and wild soybean. This gene activates secondary-cell-wall biosynthesis and promotes the thickening of fiber-cap cells in pod sutures (Dong et al., 2014).

Although linkage mapping has been proved a powerful method to explore regions of the genome that co-segregate with a given trait, it is ordinarily difficult to isolate candidate genes based on a single QTL mapping experiment due to the limited markers (Zhang et al., 2014). Compared to linkage mapping, GWAS can be applied in germplasm collections or naturally occurring populations and provides higher resolution in terms of defining the genomic positions of genes or QTLs (Shirasawa et al., 2013). In recent years, with the advance of next-generation sequencing (NGS), this approach has been successfully applied to important traits in many plant species, including Arabidopsis thaliana (Kobayashi et al., 2016; Kalladan et al., 2017), rice (Huang et al., 2010; Yano et al., 2016) and maize (Kump et al., 2011; Wang X. et al., 2016). In soybean, GWAS has been used to explore markers associated with many traits. For example, Hao et al. (2012) used GWAS to identify 19 SNPs and five haplotypes for soybean yield and yield components (Hao et al., 2012); Chu et al. (2017) mapped the GmMYB29 (Glyma20g35180) gene related to isoflavone biosynthesis in soybean via GWAS (Chu et al., 2017); and Zhang et al. (2018) identified 87 chromosomal regions for seed composition based on 31,850 SNPs by GWAS (Zhang et al., 2018).

Although previous studies have identified many QTLs controlling pod dehiscence in soybean, these QTLs were detected using classical linkage analysis with a limited number of markers, and the allelic variation was also limited by using RILs or other linkage mapping populations. Furthermore, to date, only two pod-dehiscence-related genes have been cloned. In addition, we know little about the other QTLs related to pod dehiscence.

In our previous work, we performed primary mapping of pod-dehiscence-related QTLs due to the low density of the genetic map (Hu, 2017). In this paper, to obtain more accurate information regarding pod-dehiscence-related SNPs, GWAS with the large genome-wide NJAU 355K SoySNP array was performed across multiple environments. The aims of our study were as follows: (i) identify loci and candidate genes involved in pod dehiscence; (ii) identify favorable haplotypes relevant to pod dehiscence. These results will be helpful for MAS in soybean breeding programs.

Materials and Methods

Materials and Field Trials

In this study, the association panel containing 211 soybean accessions was used (including 183 landraces, 25 improved accessions and three accessions with unknown evolution type). This population was collected from different geographic origins that exhibit variant phenotypes for pod dehiscence, and provided by the National Center for Soybean Improvement of China. Experiments to assess association panel were sowed in two places from 2013 to 2017: Jiangpu Experimental Station (32°12 N 118°37 48 E) in 2013 and 2014 (designated as environment Env1 and Env2, respectively); and the Experimental Farm of Jiangsu Yanjiang Institute of Agricultural Sciences (31°58 48 N 120°53 24 E), Nantong, China, in 2015–2017 (designated as Env3, Env4, and Env5, respectively). A complete randomized block design with three replications was used for all trials. Each line of the populations was planted in three rows, and each row was 200 cm long with 50 cm row spacing. At the seedling stage (about 2 weeks after germination), the number of plants in per row was limited to 15–20. Nutrition, water supply, weeding and insect control were maintained throughout the experiment.

Evaluation of Pod Dehiscence

Three plants per line were harvested randomly, and twenty pods per plant were examined. The maturity pods were stored in a kraft paper bag after harvested and then put in the heating ovens as soon as possible. The degree of pod dehiscence was identified by monitoring the percentage of dehiscent pods after heat treatment at 60°C for 3 h (Funatsuki et al., 2006; Yamada et al., 2009). As shown in Supplementary Figure S1, after heat treatment, the completely opened pods were used to calculate PPD. PPD = (number of pod dehiscence/total pods) × 100% (Yamada et al., 2009). The germplasm number, name, origin and phenotypic data of the association panel are listed in Supplementary Table S1.

Phenotypic Data Analysis

The phenotypic data of the association panel was determined using the mean values for each line (Env1 to Env5) with the R software (R Development Core Team, 2015), including descriptive statistics, ANOVA, and h2. Variable components (genotype, year, line, and location) were evaluated in R with the “lme4” package using the BLUP model. The BLUP model was based on the following formula (Merk et al., 2012):

Yik=μ+Gi+Yk+GYik+εik

In this formula, Yik is the trait studied, μ is the overall mean, Gi is the ith genotypic effect, Yk is the effect of the kth year, GYik is the interaction of genotype × year, and εik is the residual error. The BLUP was also included as an environment in the GWAS analysis. The h2 of pod dehiscence was calculated using the following formula:

h2=σg2/(σg2+σge2/n+σe2/nr)

In which h2, broad-sense heritability; σg2, genotype variance; σge2, interaction variance of genotype × environment; σe2, variance of error components; n, the number of environments; and r, the number of replications.

Genome-Wide Association Analysis

The 211 soybean accessions were genotyped with the NJAU 355K SoySNP array containing 282,469 SNPs (Wang J. et al., 2016). After filtering out SNPs with MAFs ≤ 0.05, a total of 201,915 SNPs remained. GWAS was conducted based on the MLM model (Q + K) using the TASSEL software V5.0 (Bradbury et al., 2007). The kinship matrix (K) and population structure (Q) were calculated by TASSEL software V5.0 and STRUCTURE software version 2.3.4, respectively (Pritchard et al., 2000; Falush et al., 2003). A FDR of 0.05 was used as a threshold for significant association (Benjamini and Hochberg, 1995). Manhattan plots and Q–Q plots were generated using the R software package “qqman” (Turner, 2014).

Digital Expression Data Analysis

Two databases were used to identify the tissue expression patterns of candidate genes. RNA-Seq data from 14 tissues, including three vegetative tissues (leaves, root, and nodules) and the whole seeds from 11 stages of reproductive tissue development (flower, pod, and seed), were downloaded from SoyBase1, and soybean microarray expression data was downloaded from Plant Expression Database2 (GEO Accession: GSE26443). The heat maps were generated by using R software packages “heatmap.”

Candidate Gene Selection and Gene Expression Analysis

Based on the LD decay distance of significant SNPs in the GWAS results, we predicted the putative genes related to pod dehiscence according to the soybean genome annotation (Wm82.a1.v1.13). Meanwhile, BLASTP analysis was also performed against Arabidopsis proteins using the amino acid sequences of these putative genes.

The soybean accession Williams 82 was selected to analyze the tissue expression patterns of the putative genes. First, seeds were grown in growth chambers under the conditions of 16/8 h (day/night), 28/25°C (day/night), and 60% relative humidity. Then, roots, stems, leaves, and flowers were sampled during the full-bloom period; pods were sampled on the 7th, 15th, 25th, and 45th days after flowering (DAF), and mature seeds were also collected. The expression levels of the putative genes were detected in 15-day pods in a subset of 18 soybean accessions, representing varieties with high and low PPD. Total RNA was isolated using a Plant RNA Extract Kit (TianGen, Beijing, China), and cDNA was synthesized using a TaKaRa Prime ScriptTM RT reagent kit according to the manufacturer’s instructions. Gene expression in different tissues was performed by qRT-PCR assays using an ABI 7500 system (Applied Biosystems, Foster City, CA, United States) with SYBR Green Real-time Master Mix (Toyobo). Three biological and three technical replicates were used. The constitutively expressed soybean tubulin gene (GenBank accession number: AY907703) was used as a control. The normalized expression, reported as fold change, was calculated for each sample as ΔΔCT=(CTTarget-CTTubulin)genotype-(CTTarget-CTTubulin)calibrator (Livak and Schmittgen, 2001). The primers used are listed in Supplementary Table S2.

Gene Structure Analysis

GSDS4 was used to analyze the gene structure. The protein sequences of homologous genes from different plants were obtained from NCBI database5 (accession numbers are listed in Supplementary Table S3). Sequence alignment was analyzed using ClustalX software version 1.83 (Jeanmougin et al., 1998). A NJ phylogenetic tree was constructed based on protein sequences with MEGA 6.0 software using bootstrap method with 1,000 replications.

Haplotype Analysis

A pair of gene-specific primers was designed (Primer 5.0) to amplify Glyma09g06290 from 42 soybean accessions (forward: 5′-CCAAACAGAGTGAGTGACTTG-3′, reverse: 5′-TGCTACTTTCTTCTTCTAGC-3′). These sequences were aligned using the ClustalX software version 1.8, and the SNPs with MAFs ≥ 0.05 were identified among these accessions. The LD level was calculated by the HapView 4.0 software program (Barrett et al., 2005). The correlation between the expression of Glyma09g06290 and pod dehiscence was calculated by SPSS 20.0 software (Pearson correlation). In addition, the haplotype block was defined by the “Solid Spine of LD” algorithm. LD analysis of 112 significant SNPs located on chromosome 16 was performed using the R software package “LDheatmap” (Shin et al., 2006).

dCAPS Marker Development

The dCAPS marker of Glyma09g06290 was developed based on the S_-500 (C/G). The gene-specific primers were designed by dCAPS Finder 2.06 (forward: 5′-CTCATATCTTT GTGTTGTCCATG-3′, reverse: 5′-ATTGGAGTGATAAGGAAA CTGG-3′). The PCR products were digested with restriction endonuclease NcoI (37°C, 1 h). Digestion products were assayed by 8% polyacrylamide gel electrophoresis to distinguish the polymorphic fragments.

Results

Phenotypic Variation in Association Panel

Descriptive statistics, ANOVA and h2 of the PPD were identified for the association panel in five environments (Table 1). The PPD for the individual accessions in the association panel ranged from 0 to 0.99, and the average value was between 0.15 and 0.27. h2 was estimated as 78.5% in association panel, which was slightly lower than that found in RILs (Kang et al., 2005). The frequency distribution showed a skewed distribution (Figure 1), not a normal distribution, which is similar to the previously reported results (Kang et al., 2009). In addition, the ANOVA results indicated that pod dehiscence was significantly influenced by the genotype effect and the interaction effects of genotype and environment (P < 0.001) in association panel.

TABLE 1
www.frontiersin.org

Table 1. Descriptive statistics, ANOVA and broad-sense heritability of the PPD in five environments of association panel (Env1, Env2, Env3, Env4, and Env5).

FIGURE 1
www.frontiersin.org

Figure 1. Frequency distribution of the PPD in association panel for five environments (Env1, Env2, Env3, Env4, and Env5).

Genome-Wide Association Analysis for Pod Dehiscence

QTL mapping suffers from fundamental limitations; one of these is the limited allelic diversity between recombinant inbred population and parents (Korte and Farlow, 2013). In contrast, by taking advantages of the abundant natural variations of a population, GWAS is helpful for favored allele detecting and functional SNP discovery. In this study, to mine the SNPs significantly associated with pod dehiscence, GWAS was conducted using 201,915 SNPs (NJAU 355K SoySNP array) with the MLM (Q + K) model in an association panel including 211 accessions.

A total of 163 SNPs were identified as significantly associated with pod dehiscence in five environments and BLUP (Figure 2, Table 2, and Supplementary Table S4). Among these SNPs, 136 SNPs were located on chromosome 16 in multiple environments. Among these 136 SNPs, one SNPs was detected in five environments and BLUP. In addition, 51 SNPs were identified in four environments and BLUP; 18 SNPs were mapped in three environments and BLUP; and 19 SNPs were mapped in two environments and BLUP. These 136 SNPs were clustered on chromosome 16 from 29135922 to 29865027 (∼729 kb), which completely overlapped with two QTLs (Pod dehiscence 2-1 and Pod dehiscence 3-4) and partially overlapped with Pod dehiscence 1-7 and Pod dehiscence 4-2 (Bailey et al., 1997; Funatsuki et al., 2006; Kang et al., 2009; Yamada et al., 2009). The gene pdh1, controlling pod dehiscence in soybean, was also located in this cluster. In addition, a few significant SNPs were distributed on other chromosomes, such as chromosome 1, 4, 6, 8, 9, 11, 17, 18, and 20 (Figure 2, Table 2, and Supplementary Table S4). On chromosome 9, two SNPs were identified in both Env3 and Env5 as significant SNPs.

FIGURE 2
www.frontiersin.org

Figure 2. Genome-wide association study of pod dehiscence in association panel using the MLM (Q + K) model across multiple environments. Manhattan plots and Q–Q plots for pod dehiscence in the following environments: (A) Env1; (B) Env2; (C) Env3; (D) Env4; (E) Env5; and (F) BLUP. The horizontal line depicts the significance threshold when FDR = 0.05.

TABLE 2
www.frontiersin.org

Table 2. SNP loci that significantly associated with pod dehiscence based on GWAS in the association panel.

Identification of Favorable SNPs and Haplotypes Relevant to Pod Dehiscence

One SNP was mapped in both five environments and BLUP on chromosome 16; we also found that nine SNPs were mapped in four environments and BLUP on chromosome 16 (the P-value of nine SNPs in Env3 was as low as 9.91 × 10–06); the R2 ranged from 11.9 to 38.5% (Table 2 and Supplementary Table S4). These significant SNPs have more allelic variations and can represent the quality of pod dehiscence in our association panel. Therefore, these SNPs were used to analyze the haplotypes related to pod dehiscence in soybean. We selected 189 of 211 accessions from different geographic origins for further haplotype analysis. Ten significant loci showed very strong LD level and formed one block. The number 1–10 represents these ten SNPs AX-93853844, AX-94151086, AX-93853870, AX-93853873, AX-93853874, AX-94151101, AX-93853876, AX-93853895, AX-93853896, and AX-94151124, respectively (Figure 3A and Supplementary Table S5). Of which, the SNP 8 and 3, 4, 7, 9, 10 showed complete LD. When 8 is T base, 3, 4, 7, 9, 10 is C, T, G, A, T base, respectively; when 8 is A base, 3, 4, 7, 9, 10 is A, C, A, T, C base, respectively (Figure 3B). Therefore, these ten SNPs were subdivided into five SNPs (1, 2, 5, 6, and 8). Based on the five SNPs within the LD block, six haplotype classes were observed in the 189 accessions (Hap1-Hap6) (Figure 3B). Hap2 (TGCAA) was the largest group (n = 83); and Hap1 (CCTTT) was the second largest group (n = 72). The other haplotypes, Hap3 (CCCAA), Hap4 (CGTTT), Hap5 (CCT0T), and Hap6 (CC0TT), comprised seven, fourteen, eight and five accessions, respectively.

FIGURE 3
www.frontiersin.org

Figure 3. Haplotype analysis of ten SNPs significantly associated with pod dehiscence in five environments and the PPD phenotype of five significant SNPs. (A) The physical locations of 10 SNPs and the LD plot based on pairwise R2-values between these SNPs. The color bar was used to indicate the R2-values. (B) Haplotypes observed in the association panel using the 10 SNPs; 0 indicates a base deletion. (C–G) The PPD phenotype of five significant SNPs. (H) Comparison of PPD among the haplotypes Hap1, Hap2, Hap3, Hap5, and Hap6. ∗∗∗Significant at P < 0.001. Two tail t-test was used for statistical analysis. 1–10 represent the SNPs AX-93853844, AX-94151086, AX-93853870, AX-93853873, AX-93853874, AX-94151101, AX-93853876, AX-93853895, AX-93853896, and AX-94151124; SE, standard error.

The phenotypic data of five significant SNPs and six haplotypes were further analyzed. In total, 117 accessions with 1-C exhibited significantly higher values than 1-T; 104 accessions with 2-C exhibited significantly higher values than 2-G; 101 accessions with 5-T and 12 accessions with 5-0 exhibited significantly higher values than 5-C; 98 accessions with 6-T and 14 accessions with 6-0 exhibited significantly higher values than 6-A, 109 accessions with 8-T exhibited significantly higher values than 8-A (Figures 3C–G). In addition, Hap1 and Hap5 had significantly higher PPD than Hap2 and Hap3 (Figure 3H). Considering that soybean varieties with pod-shattering resistance are one of the major goals for breeders, 1-T, 2-G, 5-C, 6-A, and 8-A were the favorable markers; Hap2 and Hap3 should be the favorable haplotypes.

Hap1 and Hap2 were not only the unfavorable and favorable haplotypes but also the largest and second largest groups, respectively. The number of Hap3 and Hap5 was limited (seven and eight accessions, respectively). Therefore, Hap1 and Hap2 were used to analyze the geographical distribution of soybean landraces with different pod dehiscence traits. After filtering out accessions with indeterminate ecoregions, 66 and 76 accessions remained in Hap1 and Hap2, respectively. Hap1 was observed in the three main growing ecoregions in China (Tian et al., 2010) and was mainly (75.8%) distributed in the southern ecoregion (SR); Hap2-containing accessions were mostly observed in the Huang-Huai ecoregion (HR) (34.2%) and SR (34.2%) (Table 3). These results revealed that the soybean landraces of Chinese origin possessing the elite haplotype were mostly distributed in HR and SR.

TABLE 3
www.frontiersin.org

Table 3. Geographic distribution of Hap1 and Hap2 in three soybean-growing ecoregions (NR, HR, and SR) in China.

Identification of Putative Genes

In GWAS results, under the threshold of FDR < 0.05, we detected three significant SNPs on chromosome 9. Moreover, two of them were close to each other and repeatedly identified in different environments. Interestingly, in other regions, there was no significant SNP that was repeatedly detected except on chromosome 16. Notably, one of these three SNPs, AX-93762848, was also detected by using MLMM model (Supplementary Figure S2). It demonstrates that these three SNPs on chromosome 9 were essential to dehiscence trait in soybean. Therefore, the candidate genes were selected based on the SNPs on chromosome 9.

There were 26 genes detected in the 130 kb flanking regions (LD decay distance of cultivated soybeans) of the two SNPs (Table 4; Wang J. et al., 2016). We then performed a BLASTP analysis using the Arabidopsis genome against these 26 genes. The results showed that two genes were without any explicit biological function annotations. Among the other 24 genes with functional annotations, Glyma09g06320 and Glyma09g06390 were predicted to be involved in cellular differentiation and cell wall formation. Additionally, the gene Glyma09g06290 is homologous to At2g18969 which belongs to bHLH (basic helix-loop-helix) gene family. In Arabidopsis, two bHLH transcriptional factors (IND and ALC) were reported to be related to silique dehiscence (Rajani and Sundaresan, 2001; Liljegren et al., 2004). The other genes were forecasted to participate in abiotic and biotic stress, transport, metabolic process, and other functions. We also listed the detail information about the orthologs of these genes in Medicago truncatula, Oryza sativa, and Zea mays, respectively (Supplementary Table S6). Moreover, we identified the expression patterns of these 24 genes in the RNA-Seq and microarray data, respectively (Supplementary Figure S3). The RNA-Seq data showed that Glyma09g06290 was predominately expressed in pod shell and Glyma09g06390 was expressed in several tissues. However, Glyma09g06320 was not expressed according to the RNA-seq data, and this result was confirmed by our qRT-PCR result. In the microarray data, we found that Glyma09g06290 was highly expressed in pod elongation stage and sharply decreased in late seed growth period. By combining the above results, these two genes (Glyma09g06390 and Glyma09g06290) may be related to pod dehiscence.

TABLE 4
www.frontiersin.org

Table 4. Genes detected in the 130 kb flanking region of the significant SNPs on chromosome 9.

Expression Patterns of Putative Genes

According to the Soybase7, we listed the candidate genes’ coding product and metabolic processes (Supplementary Table S7). Moreover, to elucidate the potential functions of the two genes, qRT-PCR was performed to study the expression patterns of these two genes. First, the soybean accession Williams 82 was selected to analyze the tissue expression pattern of the putative genes. Glyma09g06290 was highly expressed in the pod, and the expression of Glyma09g06290 was significantly increased in the late pod growth period, however, the expression of Glyma09g06390 was very low in pod growth period (Figure 4A). These results were similar to the public RNA-Seq data (Supplementary Figure S3). Second, 18 soybean accessions representing varieties with high and low PPD were used to analyze the expression patterns in extreme varieties. The results showed that the expression of Glyma09g06290 rather than that of Glyma09g06390 in the high-PPD varieties was significantly higher than that in the low-PPD varieties in 15-day pods (Figure 4B). Therefore, Glyma09g06290 may be involved in pod dehiscence in soybean.

FIGURE 4
www.frontiersin.org

Figure 4. Expression patterns of Glyma09g06290 and Glyma09g06390 by quantitative real-time polymerase chain reaction. (A) The expression of the putative genes in different tissues; the expression in 7-day pods was used as a control (expression value = 1). (B) The expression in 15-day pods in varieties with high and low PPD. Error bar indicates the standard deviation. ∗∗Significant at 0.001 < P < 0.01, ∗∗∗significant at P < 0.001. Two tail t-test was used for statistical analysis.

Sequence Analysis of Glyma09g06290

The full genomic length of Glyma09g06290 is 1405 bp and only contains one exon with 624 bp. Glyma09g06290 protein is composed of 207 amino acids with molecular mass of 23.23 kDa and isoelectric point of 5.39. Further, phylogenetic trees showed that Glyma09g06290 is homologous to Phaseolus vulgaris Phvul.009g238800 (identity of 78.7%), Medicago truncatula Medtr2g036670 (identity of 61.8%), Populus trichocarpa Potri.018g090200 (identity of 61.8%), M. truncatula Medtr4g127700 (identity of 50.7%) and A. thaliana columbia At2g18969 (identity of 44.9%; Figure 5). By comparing the gene structure of Glyma09g06290 and other plant homologous genes, we identified that they have only one exon and without intron except Medtr4g127700 (Figure 5).

FIGURE 5
www.frontiersin.org

Figure 5. Phylogenetic and gene structure analysis of Glyma09g06290 and other plant gene. A NJ phylogenetic tree was constructed using full protein sequences. Numbers below branches indicate bootstrap value for 1,000 replicates; CDS, coding sequence; UTR, untranslated transcribed region.

Polymorphisms in the Glyma09g06290 Gene Are Associated With Pod Dehiscence

In order to analyze the association between the allelic variation of Glyma09g06290 and pod dehiscence, we sequenced the Glyma09g06290 gene in a subset of 20 accessions with high PPD and 22 accessions with low PPD. An approximately 1.5-kb genomic region, spanning the 5′- to 3′-UTR of Glyma09g06290, was analyzed. A total of three SNPs and two indels (insertions and deletions) were identified, including Site_-500 (located 500 bp upstream from the translation start codon, S_-500), Indel_-230, S_-128 in the 5′-UTR, Indel_766 and S_767 in the 3′-UTR (Figure 6). The association study showed that 5 probable causative sites were significantly associated with variations in pod dehiscence (Figure 6A). Furthermore, based on the five significant variants with strong LD, the 42 soybean genotypes were classified into four haplotype classes (Hap1-Hap4). Hap 1 (n = 24) is the largest group, and Hap2 (n = 13) is the second largest group. Statistically, Hap2 had significantly higher PPD than Hap1 (Figures 6B,C). We then measured the expression of Glyma09g06290 in pods from 13 of 42 soybean accessions. The expression of this gene was positively correlated with pod dehiscence (r2 = 0.58, P < 0.05) (Supplementary Table S8). Therefore, these results suggested that the expression of Glyma09g06290 could partially explain the phenotypic variation in pod dehiscence.

FIGURE 6
www.frontiersin.org

Figure 6. Polymorphisms in Glyma09g06290 are significantly associated with pod dehiscence. (A) Glyma09g06290-based association mapping and pairwise LD analysis. Blue dots represent significant variants (Tassel 5.0, GLM model, P < 0.01). (B) Haplotypes of Glyma09g06290 among 42 soybean accessions; 0 indicates a base deletion. (C) Comparison of PPD between haplotypes Hap1 and Hap2. (D) Products of digestion by gel electrophoresis; M: marker, 50 bp DNA ladder (Tiangen, Beijing, China); 1–10: high-PPD varieties NJAU_C008, NJAU_C014, NJAU_C088, NJAU_C101, NJAU_C121, NJAU_C137, NJAU_C160, NJAU_C180 NJAU_C181, and NJAU_C190. Digested products with size of 166 and 143 bp, 23 bp was not observed because of its small molecular weight; 11–20: low-PPD varieties, NJAU_C054, NJAU_C076, NJAU_C080, NJAU_C082, NJAU_C085, NJAU_C098, NJAU_C165, NJAU_C172, NJAU_C201, and NJAU_C216. Undigested product with size of 166 bp. ∗∗∗Significant at P < 0.001. Two tail t-test was used for statistical analysis. SE, standard error.

Development of a Functional Marker for Glyma09g06290

In this study, five significant SNPs of Glyma09g06290 were identified and associated with pod dehiscence (Figures 6A–C and Supplementary Figure S4). The S_-500-C exhibited significantly lower values than S_-500-G; Indel_-230-0 exhibited significantly lower values than Indel_-230-C; S_-128-C exhibited significantly lower values than S_-128-A; Indel_766-T exhibited significantly lower values than Indel_766-0; S_767-T exhibited significantly lower values than S_767-G (Figure 6B). Therefore, we developed the dCAPS marker based on one of these SNPs (S_-500). As shown in Figure 6D, we selected ten accessions with high PPD and ten accessions with low PPD. Ten accessions with high PPD produced a 166 and 143 bp amplicons by restriction enzyme digestion. The remaining ten accessions with low PPD produced 166 bp amplicons by restriction enzyme digestion. These results confirmed that dCAPS based on the S_-500 is a functional marker and could be useful for plant breeders.

Discussion

Novel Loci Related to Pod Dehiscence Were Identified

The dehiscence of pods (shattering) prior to harvest is a major cause of yield loss in soybean production. Thus, it is extremely important to identify QTLs or genes related to pod dehiscence and apply them to MAS. In this study, by using the association panel which possessed abundant natural variations, GWAS was performed to genotype 211 cultivated soybean accessions with high-quality SNP markers. We identified 163 SNPs. Of which, 136 significant SNPs on chromosome 16 ranged from 29135922 to 29865027 (∼729 kb). These results were consistent with those of a previous study, which showed that the major QTL was located on chromosome 16 (Yamada et al., 2009). Similarly, a very large GWAS also confirmed that shattering in soybean is mainly due to genes located on chromosome 16 (Zhou et al., 2015). Additionally, LD analysis exhibited tight linkage among the 136 associated SNPs (Supplementary Figure S5). One of 136 SNP (AX-93853895) with the lowest P-value and maximum R2 was in the intron of pdh1. Consequently, the formation of this SNP cluster on chromosome 16 may be due to the strong effect of pdh1. Moreover, the domestication gene related to pod dehiscence between cultivated and wild soybean, SHAT1-5, was not mapped in our GWAS results. Thus, the gene effect of SHAT1-5 is likely to be weak in cultivated soybean accessions.

In addition, we also detected 25 significant SNPs except on chromosome 16 (Table 2). For example, on chromosome 8, one region contains 11 significant SNPs that were significantly associated with pod dehiscence, and this region is close to Pod dehiscence 4-1; On chromosome 9, 2 SNPs were co-identified in Env3 and Env5; there are one, one, three, one, three, three and one significant SNPs identified on chromosome 1, 4, 6, 11, 17, 18, and 20, respectively. These SNPs or regions may be the novel loci related to pod dehiscence. Overall, the SNPs identified in this study are helpful to further understanding the genetic basis of soybean pod dehiscence.

Identification of Favorable Haplotypes Related to Pod Dehiscence in Chinese Soybean Germplasms

Haplotype analysis of a candidate gene or peak SNPs have been reported for soybean (Zhang et al., 2014), rice (Famoso et al., 2011), and cotton (Su et al., 2018). For example, in cotton, four peak SNPs located on chromosome D03 were simultaneously associated with four plant architecture component traits. The four peak SNPs revealed five haplotypes, and Hap2 was the most favorable haplotype. In our study, ten significant SNPs were associated with pod dehiscence in Env1-Env5 and BLUP. Haplotype analysis showed that the 10 SNPs could be found as six haplotypes (Hap1-Hap6). Hap2 and Hap3 had lower PPD than the other haplotypes, demonstrating that Hap2 and Hap3 might be significant to breeding soybeans with lower PPD (Figure 3). According to the study, more than half of the analyzed Chinese landraces possesses the pdh1 allele (Funatsuki et al., 2014). In this study, the 189 soybean accessions were selected to represent all three ecological regions of soybean cultivation in China and soybeans with different pod dehiscence qualities. However, as more than 20,000 soybean accessions have been preserved (Wang et al., 2006), Hap2 or Hap3 could be used to discover other excellent soybean germplasms with resistance to pod dehiscence. In addition, two typical haplotypes (Hap1 and Hap2) were used to analyze the geographical distribution of Chinese soybean landraces. We found that the favorable haplotype (Hap2) was mostly observed in the HR and SR. Taken together, the above results could facilitate the development of molecular markers for the breeding of soybean accessions with lower PPD.

Identification of Candidate Genes for Pod Dehiscence in Soybean

Transcription factors play an important role in controlling pod dehiscence. The genetic network directing the morphogenesis of the dehiscence zone in Arabidopsis fruit has been identified (Ballester and Ferrandiz, 2017). For example, two members of the MADS-box transcription factor family, SHATTERPROOF1 (SHP1) and SHATTERPROOF2 (SHP2), act redundantly to control silique dehiscence (Liljegren et al., 2000). FRUITFULL (FUL), which also belongs to the MADS-box family, negatively regulates SHP1/2 expression (Ferrandiz et al., 2000). Moreover, two bHLH transcription factors, INDEHISCENT (IND) and ALCATRAZ (ALC), act downstream of SHP1/2 (Dong and Wang, 2015). Briefly, the four genes SHP1, SHP2, IND, and ALC are expressed at the valve margin to direct dehiscence zone formation. In soybean, the pod-dehiscence-related gene SHAT1-5 also encodes a NAC transcription factor. Interestingly, previously identified genes that related to pod dehiscence or fruit shedding are homologous to those involved in seed shattering in Arabidopsis thaliana. However, Rau et al. (2019) identified a QTL, qPD5.1-Pv, represents a novel locus for the shattering trait in common bean (Rau et al., 2019).

In this study, on chromosome 9, two SNPs (AX-94059226 and AX-93762822) were observed under the threshold of FDR < 0.05 in two environments (Env3 and Env5) (Table 2 and Supplementary Table S4). Moreover, we noticed that the SNP AX-93762848 was close to AX-94059226 and AX-93762822, which were observed in one environment with the FDR value < 0.05 (Supplementary Table S4). Based on these SNPs, we identified a candidate gene, Glyma09g06290, which is located in the 130kb flanking region of two repeatedly observed SNPs, and the expression of Glyma09g06290 was significantly increased in later pod growth stages. Furthermore, different expression patterns of Glyma09g06290 were found in soybean varieties with high and low PPD (Figure 4). Then we sequenced the Glyma09g06290 gene in a subset of 20 accessions with high PPD and 22 accessions with low PPD. Five SNPs were identified and used to analyze haplotypes, and the haplotype analysis showed that various haplotypes of the putative gene display variation in pod dehiscence (Figure 6). However, the biological functions of this gene should be further investigated. In summary, Glyma09g06290 might be a putative gene involved in pod dehiscence in soybean.

In addition, the GWAS results revealed that 25 significant SNPs were identified in only one environment except for the SNPs on chromosome 16 (Table 2). However, eleven of them were consecutively distributed on chromosome 8, showed significant marker-trait association, with FDR as low as 3.23 × 10–03 (P-values = 7.83 × 10–07), and were close to Pod Dehiscence 4-1 (Yamada et al., 2009) (Supplementary Table S4). Interestingly, among these eleven SNPs, seven SNPs were detected within the region of Glyma08g42110 (one of them was detected within the exon). In addition, the Glyma08g42110 is homologous to ATGSL07, which encodes callose synthase 7 (CalS7), a phloem-specific callose synthase responsible for callose deposition in developing sieve elements during phloem formation. ATGSL07 mutant plants exhibited moderate reduction in seedling height and produced aberrant pollen grains and short siliques with aborted embryos in A. thaliana (Xie et al., 2011). Our further work will also focus on the biological functions of this gene.

Conclusion

In this study, 163 SNPs were identified using GWAS across multiple environments. In addition, favorable SNPs and six haplotypes relevant to pod dehiscence were identified. Furthermore, based on GWAS results, we identified a candidate gene Glyma09g06290. Expression patterns and allelic variation showed that Glyma09g06290 was associated with the pod dehiscence. Additionally, we also developed a functional marker in Glyma09g06290 for pod dehiscence in soybean, which may be useful for the breeders. Overall, these results would provide insights for molecular-assisted breeding strategies for resistance to pod dehiscence in soybean.

Data Availability

Publicly available datasets were analyzed in this study. This data can be found here: http://www.soybase.org/.

Author Contributions

DY and FH designed the study. DHu, XL, and GK conducted the GWAS. DHu, DHao, WH, ZY, and XH carried out the field experiments. XL, HY, and YL carried out the qRT-PCR analysis. DHu wrote the manuscript. All authors approved the final version of the manuscript.

Funding

This work was supported partly by the Ministry of Science and Technology (2017YFE0111000), the Key Transgenic Breeding Program of China (2016ZX08009003-004), and the National Natural Science Foundation of China (31571688, 31671715, and 31771821).

Conflict of Interest Statement

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Supplementary Material

The Supplementary Material for this article can be found online at: https://www.frontiersin.org/articles/10.3389/fpls.2019.00811/full#supplementary-material

Abbreviations

ANOVA, analysis of variance; BLUP, best linear unbiased prediction; dCAPS, derived cleaved amplified polymorphic sequences; FDR, false discovery rate; GWAS, genome-wide association analysis; h2, broad-sense heritability; HR, Huang-Huai ecoregion; LD, linkage disequilibrium; MAFs, minor allele frequencies; MAS, marker-assisted selection; NJ, neighbor-joining; NR, northern ecoregion; PPD, the percentage of pod dehiscence; Q–Q plots, quantile–quantile plots; qRT-PCR, quantitative real-time PCR; QTL, quantitative trait locus; RIL, recombinant inbred lines; RNA-Seq, RNA sequencing; SNP, single nucleotide polymorphisms; SR, southern ecoregion.

Footnotes

  1. ^ http://www.soybase.org/soyseq/
  2. ^ http://www.plexdb.org
  3. ^ http://phytozome.jgi.doe.gov/pz/portal.html
  4. ^ http://gsds.cbi.pku.edu.cn/
  5. ^ http://www.ncbi.nlm.nih.gov/
  6. ^ http://helix.wustl.edu/dcaps/dcaps.html
  7. ^ http://www.soybase.org/

References

Bailey, M. A., Mian, M. A. R., Carter, T. E., Ashley, D. A., and Boerma, H. R. (1997). Pod dehiscence of soybean: identification of quantitative trait loci. J. Heredity 88, 152–154. doi: 10.3389/fpls.2016.01852

PubMed Abstract | CrossRef Full Text | Google Scholar

Ballester, P., and Ferrandiz, C. (2017). Shattering fruits: variations on a dehiscent theme. Curr. Opin. Plant Biol. 35, 68–75. doi: 10.1016/j.pbi.2016.11.008

PubMed Abstract | CrossRef Full Text | Google Scholar

Barrett, J. C., Fry, B., Maller, J., and Daly, M. J. (2005). Haploview: analysis and visualization of LD and haplotype maps. Bioinformatics 21, 263–265. doi: 10.1093/bioinformatics/bth457

PubMed Abstract | CrossRef Full Text | Google Scholar

Benjamini, Y., and Hochberg, Y. (1995). Controlling the false discovery rate: a practical and powerful approach to multiple testing. J. R. Stat. Soc. Ser. B Methodol. 57, 289–300.

Google Scholar

Bradbury, P. J., Zhang, Z., Kroon, D. E., Casstevens, T. M., Ramdoss, Y., and Buckler, E. S. (2007). TASSEL: software for association mapping of complex traits in diverse samples. Bioinformatics 23, 2633–2635.

PubMed Abstract | Google Scholar

Caviness, C. E. (1969). Heritability of pod dehiscence and its association with some agronomic characters in soybeans 1. Crop Sci. 9, 207–209.

Google Scholar

Chu, S., Wang, J., Zhu, Y., Liu, S., Zhou, X., Zhang, H., et al. (2017). An R2R3-type MYB transcription factor, GmMYB29, regulates isoflavone biosynthesis in soybean. PLoS Genet. 13:e1006770. doi: 10.1371/journal.pgen.1006770

PubMed Abstract | CrossRef Full Text | Google Scholar

Dong, Y., and Wang, Y. Z. (2015). Seed shattering: from models to crops. Front. Plant Sci. 6:476. doi: 10.3389/fpls.2015.00476

PubMed Abstract | CrossRef Full Text | Google Scholar

Dong, Y., Yang, X., Liu, J., Wang, B. H., Liu, B. L., and Wang, Y. Z. (2014). Pod shattering resistance associated with domestication is mediated by a NAC gene in soybean. Nat. Commun. 5:3352. doi: 10.1038/ncomms4352

PubMed Abstract | CrossRef Full Text | Google Scholar

Falush, D., Stephens, M., and Pritchard, J. K. (2003). Inference of population structure using multilocus genotype data: linked loci and correlated allele frequencies. Genetics 164, 1567–1587.

PubMed Abstract | Google Scholar

Famoso, A. N., Zhao, K., Clark, R. T., Tung, C. W., Wright, M. H., Bustamante, C., et al. (2011). Genetic architecture of aluminum tolerance in rice (Oryza sativa) determined through genome-wide association analysis and QTL mapping. PLoS Genet. 7:e1002221. doi: 10.1371/journal.pgen.1002221

PubMed Abstract | CrossRef Full Text | Google Scholar

Ferrandiz, C., Liljegren, S. J., and Yanofsky, M. F. (2000). Negative regulation of the SHATTERPROOF genes by FRUITFULL during Arabidopsis fruit development. Science 289, 436–438.

PubMed Abstract | Google Scholar

Fourquin, C., Cerro, C. D., Victoria, F. D. C., Vialetteguiraud, A., De Oliveira, A. C., and Ferrandiz, C. (2013). A change in SHATTERPROOF protein lies at the origin of a fruit morphological novelty and a new strategy for seed dispersal in Medicago genus. Plant Physiol. 162, 907–917. doi: 10.1104/pp.113.217570

PubMed Abstract | CrossRef Full Text | Google Scholar

Funatsuki, H., Ishimoto, M., Tsuji, H., Kawaguchi, K., Hajika, M., and Fujino, K. (2006). Simple sequence repeat markers linked to a major QTL controlling pod shattering in soybean. Plant Breed. 125, 195–197.

Google Scholar

Funatsuki, H., Suzuki, M., Hirose, A., Inaba, H., Yamada, T., Hajika, M., et al. (2014). Molecular basis of a shattering resistance boosting global dissemination of soybean. Proc. Natl. Acad. Sci. U.S.A. 111:17797. doi: 10.1073/pnas.1417282111

PubMed Abstract | CrossRef Full Text | Google Scholar

Guang-yu, G., Rui, S., Meng, H., Yong-xin, G., Da-wei, X., Hong-wei, J., et al. (2011). Quantitative trait locus (QTL) analysis of pod related traits in different environments in soybean. Afr. J. Biotechnol. 10, 11848–11854.

Google Scholar

Hao, D., Cheng, H., Yin, Z., Cui, S., Zhang, D., Wang, H., et al. (2012). Identification of single nucleotide polymorphisms and haplotypes associated with yield and yield components in soybean (Glycine max) landraces across multiple environments. Theor. Appl. Genet. 124, 447–458. doi: 10.1007/s00122-011-1719-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Hu, W. (2017). Linkage Analysis and Association Analysis of Pod Shattering and Yield Related Factors in Soybean. Dissertation, Nanjing Agricultural university, Nanjing.

Google Scholar

Huang, X., Wei, X., Sang, T., Zhao, Q., Feng, Q., Zhao, Y., et al. (2010). Genome-wide association studies of 14 agronomic traits in rice landraces. Nat. Genet. 42:961. doi: 10.1038/ng.695

PubMed Abstract | CrossRef Full Text | Google Scholar

Jeanmougin, F., Thompson, J. D., Gouy, M., Higgins, D. G., and Gibson, T. J. (1998). Multiple sequence alignment with Clustal X. Trends Biochem. Sci. 23, 403–405.

Google Scholar

Ju, S. L., Kim, K. R., Ha, B. K., and Kang, S. (2017). Identification of SNPs tightly linked to the QTL for pod shattering in soybean. Mol. Breed. 37:54.

Google Scholar

Kalladan, R., Lasky, J. R., Chang, T. Z., Sharma, S., Juenger, T. E., and Verslues, P. E. (2017). Natural variation identifies genes affecting drought-induced abscisic acid accumulation in Arabidopsis thaliana. Proc. Natl. Acad. Sci. U.S.A. 114, 11536–11541. doi: 10.1073/pnas.1705884114

PubMed Abstract | CrossRef Full Text | Google Scholar

Kang, S. T., Hyeunkyeung, K., Inyoul, B., Mounggun, C., Han, W. Y., Doochull, S., et al. (2005). Genetic analysis of pod dehiscence in soybean. Kor. J. Crop. 50, 281–285.

Google Scholar

Kang, S. T., Kwak, M., Kim, H. K., Choung, M. G., Han, W. Y., Baek, I. Y., et al. (2009). Population-specific QTLs and their different epistatic interactions for pod dehiscence in soybean [Glycine max (L.) Merr.]. Euphytica 166, 15–24.

Google Scholar

Kobayashi, Y., Sadhukhan, A., Tazib, T., Nakano, Y., Kusunoki, K., Kamara, M., et al. (2016). Joint genetic and network analyses identify loci associated with root growth under NaCl stress in Arabidopsis thaliana. Plant Cell Environ. 39, 918–934. doi: 10.1111/pce.12691

PubMed Abstract | CrossRef Full Text | Google Scholar

Koinange, E. M. K., Singh, S. P., and Gepts, P. (1996). Genetic control of the domestication syndrome in common bean. Crop Sci. 36, 1037–1045.

Google Scholar

Konishi, S., Izawa, T., Lin, S., Ebana, K., Fukuta, Y., Sasaki, T., et al. (2006). An SNP caused loss of seed shattering during rice domestication. Science 312, 1392–1396.

PubMed Abstract | Google Scholar

Korte, A., and Farlow, A. (2013). The advantages and limitations of trait analysis with GWAS: a review. Plant Methods 9:29. doi: 10.1186/1746-4811-9-29

PubMed Abstract | CrossRef Full Text | Google Scholar

Kump, K. L., Bradbury, P. J., Wisser, R. J., Buckler, E. S., Belcher, A. R., Oropezarosas, M. A., et al. (2011). Genome-wide association study of quantitative resistance to southern leaf blight in the maize nested association mapping population. Nat. Genet. 43:163. doi: 10.1038/ng.747

PubMed Abstract | CrossRef Full Text | Google Scholar

Liljegren, S. J., Ditta, G. S., Eshed, Y., Savidge, B., Bowman, J. L., and Yanofsky, M. F. (2000). Shatterproof MADS-box genes control seed dispersal in Arabidopsis. Nature 404, 766–770.

PubMed Abstract | Google Scholar

Liljegren, S. J., Roeder, A. H. K., Kempin, S. A., Gremski, K., Østergaard, L., Guimil, S., et al. (2004). Control of fruit patterning in Arabidopsis by INDEHISCENT. Cell 116, 843–853.

PubMed Abstract | Google Scholar

Liu, B., Fujita, T., Yan, Z. H., Sakamoto, S., Xu, D., and Abe, J. (2007). QTL mapping of domestication-related traits in soybean (Glycine max). Ann. Bot. 100, 1027–1038.

PubMed Abstract | Google Scholar

Livak, K. J., and Schmittgen, T. D. (2001). Analysis of relative gene expression data using real-time quantitative PCR and the 2(-Delta Delta C(T)) Method. Methods 25, 402–408.

PubMed Abstract | Google Scholar

Lo, S., Muñoz-Amatriaín, M., Boukar, O., Herniter, I., Cisse, N., Guo, Y.-N., et al. (2018). Identification of QTL controlling domestication-related traits in cowpea (Vigna unguiculata L. Walp). Sci. Rep. 8:6261. doi: 10.1038/s41598-018-24349-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Merk, H. L., Yarnes, S. C., Deynze, V., Tong, N., Menda, N., Mueller, L. A., et al. (2012). Trait diversity and potential for selection indices based on variation among regionally adapted processing tomato germplasm. Autism Res. 137, 427–437.

Google Scholar

Pritchard, J. K., Stephens, M., and Donnelly, P. (2000). Inference of population structure using multilocus genotype data. Genetics 155, 945–959.

PubMed Abstract | Google Scholar

Rajani, S., and Sundaresan, V. (2001). The Arabidopsis myc/bHLH gene ALCATRAZ enables cell separation in fruit dehiscence. Curr. Biol. 11, 1914–1922.

PubMed Abstract | Google Scholar

Rau, D., Murgia, M. L., Rodriguez, M., Bitocchi, E., Bellucci, E., Fois, D., et al. (2019). Genomic dissection of pod shattering in common bean: mutations at non-orthologous loci at the basis of convergent phenotypic evolution under domestication of leguminous species. Plant J. 97, 693–714. doi: 10.1111/tpj.14155

PubMed Abstract | CrossRef Full Text | Google Scholar

Shin, J.-H., Blay, S., McNeney, B., and Graham, J. (2006). LDheatmap: an R function for graphical display of pairwise linkage disequilibria between single nucleotide polymorphisms. J. Stat. Softw. 16, 1–10. doi: 10.18637/jss.v016.c03

CrossRef Full Text | Google Scholar

Shirasawa, K., Fukuoka, H., Matsunaga, H., Kobayashi, Y., Kobayashi, I., Hirakawa, H., et al. (2013). Genome-wide association studies using single nucleotide polymorphism markers developed by re-sequencing of the genomes of cultivated tomato. DNA Res. 20, 593–603. doi: 10.1093/dnares/dst033

PubMed Abstract | CrossRef Full Text | Google Scholar

Su, J., Li, L., Zhang, C., Wang, C., Gu, L., Wang, H., et al. (2018). Genome-wide association study identified genetic variations and candidate genes for plant architecture component traits in Chinese upland cotton. Theor. Appl. Genet. 131, 1299–1314. doi: 10.1007/s00122-018-3079-5

PubMed Abstract | CrossRef Full Text | Google Scholar

Suanum, W., Somta, P., Kongjaimun, A., Yimram, T., Kaga, A., Tomooka, N., et al. (2016). Co-localization of QTLs for pod fiber content and pod shattering in F2 and backcross populations between yardlong bean and wild cowpea. Mol. Breed. 36:80.

Google Scholar

Suzuki, M., Fujino, K., Nakamoto, Y., Ishimoto, M., and Funatsuki, H. (2010). Fine mapping and development of DNA markers for the qPDH1 locus associated with pod dehiscence in soybean. Mol. Breed. 25, 407–418.

Google Scholar

Tian, Z., Wang, X., Lee, R., Li, Y., Specht, J. E., Nelson, R. L., et al. (2010). Artificial selection for determinate growth habit in soybean. Proc. Natl. Acad. Sci. U.S.A. 107, 8563–8568.

Google Scholar

Tsuchiya, T. (1986). Studies on shattering resistance in soybean breeding. Rep. Hokkaido. Prefect. Agric. Exp. Stn 58, 1–53. doi: 10.1270/jsbbs.61.554

PubMed Abstract | CrossRef Full Text | Google Scholar

Tsuchiya, T. (1987). Physiological and genetic analysis of pod shattering in soybeans. JARQ 21, 166–175.

Google Scholar

Turner, S. D. (2014). qqman: an R package for visualizing GWAS results using Q-Q and manhattan plots. BiorxivGoogle Scholar

Vittori, V. D., Gioia, T., Rodriguez, M., Bellucci, E., Bitocchi, E., Nanni, L., et al. (2019). Convergent evolution of the seed shattering trait. Genes 10:68.

Google Scholar

Wang, J., Chu, S., Zhang, H., Zhu, Y., Cheng, H., and Yu, D. (2016). Development and application of a novel genome-wide SNP array reveals domestication history in soybean. Sci. Rep. 6:20728. doi: 10.1038/srep20728

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, X., Wang, H., Liu, S., Ferjani, A., Li, J., Yan, J., et al. (2016). Genetic variation in ZmVPP1 contributes to drought tolerance in maize seedlings. Nat. Genet. 48:1233. doi: 10.1038/ng.3636

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, L., Guan, Y., Guan, R., Li, Y., Ma, Y., Dong, Z., et al. (2006). Establishment of Chinese soybean Glycine max core collections with agronomic traits and SSR markers. Euphytica 151, 215–223.

Google Scholar

Xie, B., Wang, X., Zhu, M., Zhang, Z., and Hong, Z. (2011). CalS7 encodes a callose synthase responsible for callose deposition in the phloem. Plant J. Cell Mol. Biol. 65, 1–14. doi: 10.1111/j.1365-313X.2010.04399.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Yamada, T., Funatsuki, H., Hagihara, S., Fujita, S., Tanaka, Y., Tsuji, H., et al. (2009). A major QTL, qPDH1, is commonly involved in shattering resistance of soybean cultivars. Breed. Sci. 59, 435–440.

Google Scholar

Yano, K., Yamamoto, E., Aya, K., Takeuchi, H., Lo, P. C., Hu, L., et al. (2016). Genome-wide association study using whole-genome sequencing rapidly identifies new genes influencing agronomic traits in rice. Nat. Genet. 48:927. doi: 10.1038/ng.3596

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, D., Song, H., Cheng, H., Hao, D., Wang, H., Kan, G., et al. (2014). The acid phosphatase-encoding gene GmACP1 contributes to soybean tolerance to low-phosphorus stress. PLoS Genet. 10:e1004061. doi: 10.1371/journal.pgen.1004061

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, J., Wang, X., Lu, Y., Bhusal, S. J., Song, Q., Cregan, P. B., et al. (2018). Genome-wide scan for seed composition provides insights into soybean quality improvement and the impacts of domestication and breeding. Mol. Plant 11:460. doi: 10.1016/j.molp.2017.12.016

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhou, Y., Lu, D., Li, C., Luo, J., Zhu, B., Zhu, J., et al. (2012). Genetic Control of seed shattering in rice by the APETALA2 transcription factor shattering abortion1. Plant Cell 24, 1034–1048. doi: 10.1105/tpc.111.094383

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhou, Z., Jiang, Y., Wang, Z., Gou, Z., Lyu, J., Li, W., et al. (2015). Resequencing 302 wild and cultivated accessions identifies genes related to domestication and improvement in soybean. Nat. Biotechnol. 34, 441–441.

Google Scholar

Keywords: soybean, pod dehiscence, GWAS, haplotype analysis, Glyma09g06290, polymorphism analysis

Citation: Hu D, Kan G, Hu W, Li Y, Hao D, Li X, Yang H, Yang Z, He X, Huang F and Yu D (2019) Identification of Loci and Candidate Genes Responsible for Pod Dehiscence in Soybean via Genome-Wide Association Analysis Across Multiple Environments. Front. Plant Sci. 10:811. doi: 10.3389/fpls.2019.00811

Received: 04 March 2019; Accepted: 05 June 2019;
Published: 21 June 2019.

Edited by:

Roberto Papa, Marche Polytechnic University, Italy

Reviewed by:

Saleh Alseekh, Max Planck Institute of Molecular Plant Physiology, Germany
Francesca Sparvoli, Istituto di Biologia e Biotecnologia Agraria (IBBA), Italy
Monica Rodriguez, University of Sassari, Italy

Copyright © 2019 Hu, Kan, Hu, Li, Hao, Li, Yang, Yang, He, Huang and Yu. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Fang Huang, fhuang@njau.edu.cn; Deyue Yu, dyyu@njau.edu.cn

These authors have contributed equally to this work

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.