Skip to main content

ORIGINAL RESEARCH article

Front. Genet., 06 April 2022
Sec. Genomics of Plants and the Phytoecosystem
This article is part of the Research Topic Biotechnological and Genomic Approaches for Enhancing Agronomic Performance of Crops View all 19 articles

Genotyping-by-Sequencing Based Molecular Genetic Diversity of Pakistani Bread Wheat (Triticum aestivum L.) Accessions

Shabbir HussainShabbir Hussain1Madiha HabibMadiha Habib1Zaheer AhmedZaheer Ahmed2Bushra SadiaBushra Sadia1Amy BernardoAmy Bernardo3Paul St. AmandPaul St. Amand3Guihua BaiGuihua Bai3Nida GhoriNida Ghori3Azeem I. KhanAzeem I. Khan2Faisal S. Awan
Faisal S. Awan1*Rizwana MaqboolRizwana Maqbool2
  • 1Center of Agricultural Biochemistry and Biotechnology, University of Agriculture, Faisalabad, Pakistan
  • 2Department of Plant Breeding and Genetics, University of Agriculture, Faisalabad, Pakistan
  • 3USDA, Hard Winter Wheat Genetics Research Unit, Manhattan, KS, United States

Spring wheat (Triticum aestivum L.) is one of the most imperative staple food crops, with an annual production of 765 million tons globally to feed ∼40% world population. Genetic diversity in available germplasm is crucial for sustainable wheat improvement to ensure global food security. A diversity panel of 184 Pakistani wheat accessions was genotyped using 123,596 high-quality single nucleotide polymorphism (SNP) markers generated by genotyping-by-sequencing with 42% of the SNPs mapped on B, 36% on A, and 22% on D sub-genomes of wheat. Chromosome 2B contains the most SNPs (9,126), whereas 4D has the least (2,660) markers. The mean polymorphic information content, genetic diversity, and major allele frequency of the population were 0.157, 0.1844, and 0.87, respectively. Analysis of molecular variance revealed a higher genetic diversity (80%) within the sub-population than among the sub-populations (20%). The genome-wide linkage disequilibrium was 0.34 Mbp for the whole wheat genome. Among the three subgenomes, A has the highest LD decay value (0.29 Mbp), followed by B (0.2 Mbp) and D (0.07 Mbp) genomes, respectively. The results of population structure, principal coordinate analysis, phylogenetic tree, and kinship analysis also divided the whole population into three clusters comprising 31, 33, and 120 accessions in group 1, group 2, and group 3, respectively. All groups were dominated by the local wheat accessions. Estimation of genetic diversity will be a baseline for the selection of breeding parents for mutations and the genome-wide association and marker-assisted selection studies.

Introduction

Wheat (Triticum aestivum L.) is amongst the most prominent cereal crops in the world. Wheat is cultivated on more than 224 million hectares (Mha) and provides 20% (Bhatta et al., 2017) of dietary nutrients for 40% world population (Ray et al., 2013; Arzani and Ashraf, 2017; Food and Agriculture Organization of the United Nations, 2019). Moreover, 68% of global wheat is directly utilized for human food while the remaining 32% for livestock consumption and other miscellaneous uses such as biofuel (Mohanty and Swain, 2019). The world population is increasing at the rate of 1.5% per annum and is expected to reach 9.7 billion by 2050 (Lobell et al., 2011). However, a 0.9% annual increase in bread wheat yield currently is not sufficient to ensure global food security (Ray et al., 2013). Pakistan is the fifth densely populated country in the world. Wheat was cultivated on an area of 8,825 thousand ha of land, which produced 24,946 MT of grain in 2020 (Pakistan Economic Survey 2019-20).

Wheat genetic gains can be improved by optimizing crop husbandry (Sener et al., 2009) and harnessing genetic diversity in native germplasm resources (Nielsen et al., 2014; Govindaraj et al., 2015). Dissection of genetic diversity is a prerequisite for plant breeding experiments, such as domestication, inheritance, conservation, and evaluation of wheat germplasm (Peterson et al., 2014). Narrow genetic bases, limited genetic diversity, and continuous reduction in arable farmland, as well as various climate-associated anomalies in the form of biotic and abiotic stresses, pose a continuous threat to world food security in developing countries (Fischer et al., 2014). Moreover, wheat has a long evolutionary history from the days of its early domestication from Einkorn (one of the primitive wheat ancestors) to modern bread wheat and assimilates plenty of genetic variation during this long period of evolution. Farmers’ selection, uniform varietal seed production, continuous selfing, use of modern breeding techniques, domestication, and stringent selection pressure lead to prompt genetic erosion and squat gene pool, which causes continuous losses of favorable alleles in currently used wheat germplasm (Haudry et al., 2007; Sofalian et al., 2008).

Strategic trait-based wheat breeding is the most viable and sustainable solution for crop improvement. The success of a wheat breeding program can be directly associated with the availability of valuable genetic diversity in the program (Rufo et al., 2019; Voss-Fels et al., 2019). Genetic resources such as wild relatives, gene bank accessions, landraces, advanced breeding lines, and induced and natural mutants are considered indispensable genetic resources for maintaining genetic diversity and crop improvement (Ogbonnaya et al., 2013; Das et al., 2016). Natural variation selection, wide crossing, new gene introduction, genetic hybridization, induced mutagenesis, and horizontal and vertical gene transfer can contribute to enriching the genetic diversity of modern wheat varieties to meet the challenges of climate change and global food security (Arya et al., 2013).

DNA markers are considered indispensable tools for the genetic characterization of plants. Various types of DNA-based markers, including randomly amplified polymorphic DNA (RAPD), amplified fragment length polymorphisms (AFLP), restriction fragment length polymorphisms (RFLP), and simple sequence repeats (SSRs) (Röder et al., 1995), were successfully used for genetic characterization in plants. Nowadays, single nucleotide polymorphism (SNP) markers generated through next-generation sequencing (NGS,) such as targeted amplicon sequencing (TAS), Illumina bead chip array, DArT, Genotyping-by-Sequencing (GBS), and kompetitive allele-specific PCR (KASP), are gradually replacing other old marker systems due to high throughput and low cost per data point. GBS is a reduced representation sequencing method that identifies SNP for genotyping and discovers new SNP compared to other array-based genotyping technologies, namely, DArT/Illumina bead chip. An array-based genotyping chip targets and identifies a pre-labeled specific number of SNPs markers. GBS-based sequencing could be effectively used in the breeding of complex genome crops without any prior sequencing information or even in the absence of reference genome in many orphan crops (Getachew et al., 2019). GBS is a technology that can reduce genome complexity by using two restriction enzymes (PstI/MspI) and simultaneously discover and genotype genome-wide variations in complex genome crops, namely, bread wheat (Elshire et al., 2011; Poland et al., 2012a; Poland et al., 2012b). Currently, GBS has been successfully used in many crops to unblock the GD including wheat, barley, rice, maize, cassava, potato, and soybean (Elshire et al., 2011).

SNPs are the most abundant polymorphic markers in both plant and animal genomes (Rimbert et al., 2018). The bi-allelic nature, high level of polymorphisms, ubiquitous presence, uniform distribution across genomes, automated data acquisition, and analysis make SNPs the most suitable marker for genome-wide marker analysis (Verma et al., 2015). Quick advancement in NGS techniques significantly improved sequencing throughput and reduced sequencing cost, making the genome-wide SNP analysis a time- and cost-effective tool for genomic studies (He et al., 2014). Therefore, the NGS-based markers have been widely used for retrieving genetic diversity (GD), harnessing population structure (PS), studying linkage disequilibrium (LD), mapping quantitative trait loci (QTL), conducting genome-wide association studies (GWAS), and genomic selection (GS) in various crops (You et al., 2018).

Although the cost of whole-genome sequencing significantly reduced during the last decade, still it is not feasible to completely sequence all genotypes for routine screening of breeding materials especially for those species with a huge genome such as wheat in most breeding programs. Mourad et al. (2020) genotyped a panel of 103 spring wheat accessions that were collected from five continents and deposited in the USDA gene bank to evaluate the GD, PS, and LD patterns. The panel was genotyped with 36,720 high-quality SNPs, and the whole population was divided into three subpopulations on the basis of analysis of molecular variance (AMOVA), structure analysis, kinship, and principal component analysis. No high LD was observed on a whole bread wheat genome, but at the sub-genomic level, the D genome showed the highest LD decay value compared to the A and B sub-genomes.

Alipour et al. (2017) utilized 16,506 polymorphic GBS-SNPs to dissect the GD of an Iranian wheat diversity panel of 369 wheat genotypes. The B subgenome has the highest number of mapped SNPs compared to the other A and D subgenomes, respectively. The whole population was divided into three subgroups: one for cultivars and two for landraces. Heslot et al. (2013) analyzed 38,412 GBS-SNPs in 365 soft winter wheat to harness GD in advanced breeding lines. Yang et al. (2020) characterized a population of 180 bread wheat genotypes from Asia and Europe to determine the indigenous PS and GD using 24,767 high-quality polymorphic SNPs using the GBS approach and to determine GD of the subjected population. The polymorphic information content (PIC) value of markers ranges from 0.1 to 0.4. Based on cluster and structure analysis, the whole diversity panel was divided into two groups: group 1 comprises European and partial Asian and group 2 consists of the Middle East and partial Asian accessions.

Most Pakistani wheat germplasm has not been characterized using DNA (SNP) markers. Local germplasm is always a key source of resistance against biotic and abiotic anomalies. The primary objective of this study is to investigate the extent and the pattern of GD, PS, LD, PIC, phylogeny, and kinship of the subjected population that will be useful for the selection of breeding parents for various stress breeding strategies through GWAS and association mapping (AM).

Materials and Methods

Deoxy Nucleic Acid Isolation and Genotyping-by-Sequencing Library Preparation

A set of 184 Pakistani spring wheat varieties and germplasm accessions were collected from the Wheat Research Institute, Ayub Agricultural Research Institute (AARI), Pakistan, Faisalabad (Supplementary Table S1). The panel was sown into two 96-well plastic trays in the greenhouse at the Kansas State University, Manhattan, KS, United States. Three pieces of 15-day-old seedling leaf tissues (2 cm) were collected in 1.1 ml 96-deep-well plates with 3 mm stainless beads in each well and immediately freeze-dried for 2 days. Leaf tissues were ground by shaking the plates at 30 cycles per second for 3 min in a Mixer Mill (Retsch GmbH, Haan, Germany). Genomic DNA was isolated using the standard cetyltrimethylammonium bromide (CTAB) method with a slight modification (Zhao et al., 2020). DNA quality was checked with 1% agarose gel and quantified in a FLUOstar Omega microplate reader (BMG LABTECH, Germany) using Quant-iT™ PicoGreen dsDNA assay kits (Thermo Fisher Scientific, Waltham, MA, United States). The genomic DNA of all samples was normalized at a concentration of 20 ng/μL for preparing the GBS library.

GBS libraries of 184 spring wheat samples were prepared using the standard protocol (Poland et al., 2012a). In brief, normalized DNA (200 ng) was digested using restriction enzyme PstI-HF and MspI from New England BioLabs, Inc. (Ipswich, MA, United States) and ligated with barcoded adapters using a T4 DNA ligase (New England BioLabs, Inc., Ipswich, MA, United States). Ligated DNA fragments were pooled and purified using QIAquick PCR Purification Kit (QIAGEN GmbH, Hilden, Germany). Primers complementary on both adapters were used for PCR amplification. PCR product was cleaned using QIAquick PCR Purification Kit and size selected using an E-gel (Thermo Fisher Scientific, Waltham, MA, United States) to select 200–300 bp fragments and then quantified using Qubit 2.0 Fluorometer and Qubit dsDNA HS Assay Kits (Life Technologies Inc., Carlsbad, CA, United States).

Single Nucleotide Polymorphism Calling and Data Imputation

The final library was sequenced in an Ion Proton next-generation sequencer (Thermo Fisher Scientific, Waltham, MA, United States) in the USDA Central Small Grain Genotyping Lab, Kansas State University, Manhattan, KS, United States. SNPs were called using the GBS discovery pipeline v2.0 in Trait Analysis by Association, Evolution, and Linkage (TASSEL) v5.2.63 (Bradbury et al., 2007) by aligning the sequence reads with the International Wheat Genome Sequencing Consortium (IWGSC) reference genome RefSeq V2.0 (IWGSC, 2018). Initially called SNPs were filtered to remove these SNPs with >20% missing data and <0.01% minor allele frequency (MAF) and then imputed for missing data using BEAGLE v5.1, (Browning and Browning, 2009), in TASSEL v5.2.63.

Genetic Diversity, Polymorphic Information Content, and Analysis of Molecular Variance

TASSEL v5.2.67 was used to calculate the evolutionary relationship among the 184 Pakistani accessions. The dendrogram was constructed using the Neighbor-Joining (NJ) distance method (Saitou and Nei, 1987) in TASSEL. Principal coordinate analysis (PCoA) was implemented using the Euclidean distance method in GenAlEx v6.5 (Peakall and Smouse, 2006). Population GD, PIC, major allele frequency (MF), and percentage heterozygosity (HZ) were determined using POWER MARKER v3.25 (Liu and Muse, 2005). Analysis of molecular variance (AMOVA) (Excoffier et al., 1992) and Shannon’s information index (I) were calculated using GenAlEx v6.5 (Peakall and Smouse, 2006).

Analyses of Population Structure, Linkage Disequilibrium, and Kinship Analysis

Population structure was inferred using an admixture model and Bayesian model-based clustering algorithm in STRUCTURE v2.3.4 (Pritchard et al., 2000). The best fit delta K value for the number of subpopulations was determined using K values from 1 to 10 and burned in 10,000 generations and 100,000 Markov chain Monte Carlo of (MCMC) iterations (Zorić et al., 2012; Chen et al., 2012). The output of STRUCTURE analysis was visualized using a STRUCTURE HARVESTER software (Earl and vonHoldt, 2012). TASSEL v5.2.67 was used to calculate LD as squared allele frequency (r2) and physical distance (D) between each pair of SNPs using a sliding window size of 50 with 1,000 permutations. Pairwise LD (r2) values were plotted against the relative physical distances (D), and a locally weighted polynomial curve regression (LOESS) model was fitted to determine genome-wide LD decay using the R package (https://www.r-project.org/). The LDs were calculated for the whole wheat genome and three sub-genomes (A, B, D) (Hill and Weir 1988; Remington et al., 2001). The critical value of r2 beyond which LD likely starts to decay was set at r2 = 0.1. This critical threshold value of r2 was estimated using the 95th percentile in the distribution of r2 below which the relationship among pairs of marker loci is not caused by physical linkage. The intersection of the fitted curve among correlation r2 and physical distance (D) on a chromosome with this critical threshold was considered the estimated LD range. Population kinship heat matrix among all genotypes was calculated using GAPIT in the R package (Wang et al., 2014).

Results

Genomic Distribution of Single Nucleotide Polymorphisms

The panel of 184 wheat accessions generated 202,147,814 sequence reads and 129,180 SNPs with 80% missing data points. A total of 123,596 polymorphic SNPs were retained after removing low-quality SNPs with MAF <0.01, heterozygote rate >0.2, and imputation using the Chinese spring wheat reference genome RefSeq V2.0 (IWGSC, 2018) for downstream analysis.

A total of 51,975 SNPs were mapped on the B genome, 44,400 on the A genome, and 27,221 on the D genome (Figure 1A). These results indicated that the B sub-genome of bread wheat holds the highest number of mapped markers (SNPs) followed by the A and D sub-genomes. However, the genome sequence sizes of all sub-genome are nearly equal to 5.5 Gb each. Chromosome 2B has the highest number of mapped SNPs (9,126), while chromosome 4D holds the lowest number of mapped SNPs (2,660) (Figure 1B). The chromosomal distribution of SNPs within 1 Mb window size is presented in Figure 1C, which depicts a higher density of SNPs on chromosomal arms rather than its centromeric region.

FIGURE 1
www.frontiersin.org

FIGURE 1. (A) Sub-genomic distribution of 123,596 SNPs on three sub-genomes of spring wheat. (B) Chromosomal SNPs distribution on each chromosome of three sub-genomes (A, B, D), (C) Number of SNPs within 1 Mb window size.

Population Structure

A total of 15,779 highly polymorphic SNPs markers with PIC values ranging from 0.30 to 0.37 were used for downstream data analysis. Population structure analysis was performed on 184 Pakistani wheat accessions, distributed into three sub-groups (group 1, group 2, and group 3) based on optimum K = 3. Delta K value gives the optimal number of subpopulations by plotting a graph with the number of clusters (K) (Figures 2A,B). Results of the STRUCTURE analysis were further confirmed by the PCoA analysis (Figure 2C), which was also supported by neighbor-joining phylogenetic tree analysis (Figure 2D).

FIGURE 2
www.frontiersin.org

FIGURE 2. Population structure analysis, (A) Delta K line graph, (B) structure analysis combine bar chart, (C) PCoA analysis of wheat accessions, (D) Neighbor Joining phylogenetic tree analysis.

The panel of 184 accessions was used in this study that includes 118 Pakistani local and 63 accessions from CIMMYT (Centro Internacional de Mejoramiento de Maíz y Trigo; English: International Maize and Wheat Improvement Center, Mexico), two wild wheat relatives, and one accession with unknown origin (Figure 3A). Phylogenetic analysis divided the 184 accessions into three different clusters. Group 1 contained 31 accessions, those dominated by local lines including 23 Pakistani accessions, and only eight CIMMYT accessions. Group 2 consisted of 33 accessions, including 18 Pakistani and 15 from CIMMYT. Group 3 was the largest group that had 120 accessions, including 75 local accessions from Pakistan, 42 from CIMMYT, two wild wheat relatives, and one accession with unknown origin (Figure 3B). A total of 120 local wheat accessions, 63 from CIMMYT, two wild relatives of wheat, and one accession of unknown origin were presented with red, black, blue, and purple colors, respectively (Figure 3C).

FIGURE 3
www.frontiersin.org

FIGURE 3. (A) Origin of wheat accessions, (B) Classification of accessions in three different groups on the basis of structure analysis and phylogenetic tree analysis (C) Distribution of different accession in different sub-groups, the red color showed local genotypes, black color indicated CIMMYT accessions, blue color for wild type and purple color for unknown origin.

Major Allele Frequency, Genetic Diversity, Heterozygosity, Polymorphic Information Content, and Kinship Heat Map Analysis

Major allele frequency (MF), genetic diversity (GD), heterozygosity (HZ), and the polymorphic information content (PIC) of the wheat panel are listed in Table 1. The whole-genome mean MF was 0.87, with 0.86, 0.87, and 0.89 for A, B, and D sub-genomes, respectively. Mean GD was observed as 0.20, 0.19, and 0.16 for sub-genomes A, B, and D, respectively. The HZ was 0.07, 0.06, and 0.08 for A, B, and D genomes, respectively, and the PIC value was observed as 0.16, 0.16, and 0.14 for A, B, and D sub-genomes, respectively, with 0.15 cumulative PIC for whole wheat genome. Results of the current experiment showed that subgenomes A and B do not have any significant differences among MF, GD, HZ, and PIC values. However, the D sub-genome showed a higher and significant difference among the MF, HZ, relatively lower GD, and PIC values compared to the A and B sub-genomes (Table 1).

TABLE 1
www.frontiersin.org

TABLE 1. Mean major allele frequency, genetic diversity, heterozygosity, and polymorphic information content calculated using 123,596 genome-wide SNPs derived from the panel of 184 wheat genotypes.

AMOVA analysis showed 20 and 80% genetic variation among and within the population, respectively. The fixation index Fst was used to genetically differentiate total genetic variability among the sub-populations. A low haploid Nm value (0.497) indicates limited gene flow between the sub-populations (Table 2). Shannon’s information index (I) also reported very low variation among groups than within different groups (Table 3). The percentage of Shannon’s information index (I) and scaled diversity overlapped among and within groups, as presented in Table 3. Kinship analysis also divided the panel into three distinct clusters, suggesting a considerable genetic difference among accessions in this panel. The phylogenetic tree is shown on the top, and the left side of the heat map also confirms the results of structure analysis. However, the intensity of color in the heat map also indicated the high LD regions in the heat map (Figure 4).

TABLE 2
www.frontiersin.org

TABLE 2. Analysis of molecular variance in the panel of 184 genotypes.

TABLE 3
www.frontiersin.org

TABLE 3. Population Shannon information index (I).

FIGURE 4
www.frontiersin.org

FIGURE 4. Heat map of kinship matrix with the dendrogram shown on the top-left side based on 123,596 markers.

Evolution of Linkage Disequilibrium

The scatter plot of (r2) and pairwise physical distance (D) revealed that LD decay increases with an increase in marker physical distances on chromosomes or a genome. The average cutoff value r2 = 0.1 was used to determine the LD decay level. The LD decay value (0.34 Mbp) was observed for the whole genome (Figure 5A). However, among sub-genomes, the LD decay was the highest on the A subgenome (0.29 Mbp), followed by the B (0.2 Mbp) and D (0.07 Mbp) subgenomes, respectively (Figures 5B–D). Chromosome-wide LD of each sub-genome is shown in Table 4. The LD decay of the A sub-genome ranges from 0.08 to 0.12, and chromosome 3A showed the highest value of LD (0.12). The LD decay of the B sub-genome ranges from 0.05 to 0.13, and chromosome 3B showed a maximum value of LD for the 3B chromosome. However, in the case of the D sub-genome, the LD value ranged from 0.01 to 0.16, and chromosome 2D showed the highest value (0.16) of LD. A low LD value in this population may be due to higher heterozygosity in the sub-genome reflecting the actual hemizygous state of SNPs markers on the D genome.

FIGURE 5
www.frontiersin.org

FIGURE 5. Scatter plot showing the linkage disequilibrium (LD) decay across the genomes for 184 diverse Pakistani bread wheat accessions. The genetic distance in megabase pair (Mbp) plotted against the LD estimate (r2) for pairs of SNPs. The solid red line showed the threshold LD value at r2 = 0.1 and the solid green line showed the average LD decay point at which the LD curve intercepts the critical r2. The LOESS LD decay curve was presented with the solid blue curve (A) Genome-wide average LD decay plot using all genomes (B) LD decay plot of the A genome (C) LD decay plot of the B genome; (D) LD decay plot of the D genome.

TABLE 4
www.frontiersin.org

TABLE 4. Chromosome-wise linkage disequilibrium (LD) of individual chromosome.

Discussion

Genetic diversity among the wheat germplasm is exceptionally important for the genetic improvement of wheat cultivars. Wheat widely grows in South Asia (Pakistan, India, and Bangladesh), producing more than 15% of the world’s wheat production. Pakistan ranks seventh in world wheat production. However, the genetic diversity of most Pakistani wheat germplasm has not been well characterized using genome-wide SNPs markers. Information on the genetic diversity of the germplasm would be crucial for breeding programs that can play a critical role for wheat breeders to make efficient use of available germplasm resources in crop improvement (Hawkes, 1981). In this study, 184 Pakistani spring wheat accessions were genotyped using genome-wide SNPs generated by GBS technology. Most SNPs were mapped on the B genome (42%), followed by the A (36%) and D (22%) sub-genomes (Figure 1A), with the maximum number of SNPs on chromosome 2B (9,126) and the minimum number (2,660) on chromosome 4D (Figure 1B). These findings are consistent with previous studies (Alipour et al., 2017; Bhatta et al., 2017; Eltaher et al., 2018; Rufo et al., 2019; Kumar et al., 2020; Yang et al., 2020).

In the current study, the number of SNPs from the D sub-genome (22%) was about half of these present on the B (42%) sub-genome, which agrees with some of the previous reports that the B genome harbors twice as many of the mapped markers as these present on the D sub-genome (Wang et al., 2013; Iehisa et al., 2014; Edae et al., 2015). However, a much lower ratio of the polymorphic markers was found on the D sub-genome, as low as five times compared to the A or B sub-genomes (Allen et al., 2013; Cavanagh et al., 2013). Low polymorphism in the D genome in most of the germplasm was observed not only for SNPs markers but also for RFLP, AFLP, SSR, and DArT markers (Liu and Tsunewaki, 1991; Röder et al., 1998; Peng et al., 2000; Chao et al., 2007; Akhunov et al., 2010; Würschum et al., 2011; Poland et al., 2012a; Berkman et al., 2013; Marcussen et al., 2014; Nielsen et al., 2014; Shavrukov et al., 2014; Edae et al., 2015; Alipour et al., 2017; Eltaher et al., 2018; Rufo et al., 2019). In wheat, A and B are predecessor sub-genomes that accumulated more genetic recombination, duplication, mutation, and gene flow events than those in the relatively newer D subgenome (Berkman et al., 2013). These results suggest that Pakistani wheat genotypes, like the Iranian germplasm (Alipour et al., 2017), hold more genetic diversity on the D genome than other sources of germplasm. Therefore it is a valuable source for crop improvement against various climate anomalies (Jia et al., 2013). Pakistani wheat accessions as T. aestivum L. ssp. Sphaerococcum indigenous from northern Pakistan and northwestern India an early flowering, yellow rustresistant, semi-dwarf plant with a semispherical grain shape (Percival, 1922). This notion points out the ancestral relationship between global populations of Puccinia striiformis f. sp. tritici (Pst) and the putative origin of Pst in the Himalayan, as well as their neighboring plains and foothill regions. The presence of putative alleles, high level of diversity, ability to have the sex-related structure in (Pst) races, and clear signature of recombination further support this hypothesis (Ali et al., 2014). Archeological remains and evolutionary studies indicated that hexaploid spring wheat was already grown in the region of India and Pakistan somewhat between 4000 and 2000 BC, which indicates that the region (India and Pakistan) has the longest cultivation and interactions history with Pst. This statement also indicates that Pakistani wheat accessions were evolved with comparatively more recombination and ultimately enriched source of genetic diversity (Tengberg, 1999; Habib et al., 2020).

PIC is another important parameter for the selection of markers for breeding programs. Botstein et al. (1980) coined a scale range from 0 to 1.0 to categorize multilocus markers such as SSR according to their PIC value. In the current study, PIC values ranged from 0.004 to 0.37, which is low to moderately high with the mean of 0.158 for the whole genome that was smaller than 0.25 reported by previous literature (Chao et al., 2009; Novoselović et al., 2016; Alipour et al., 2017; El-Esawi et al., 2018; Eltaher et al., 2018; Alemu et al., 2020; Mourad et al., 2020). Among the three sub-genomes, the mean PIC for the D sub-genome (0.14) was slightly lower than that of the A or B sub-genomes (0.16) (Table 1) due to limited hybridization event and gene flow in Aegilops tauschii during evolution (Chao et al., 2009; Lopes et al., 2015; Liu Y. et al., 2017; Eltaher et al., 2018; Tyrka et al., 2021). However, much higher PIC (0.26–0.33) were reported in other studies (Allen et al., 2013; Baloch et al., 2017; Chen et al., 2018; Kumar et al., 2020; Yang et al., 2020).

Gene diversity (expected heterozygosity, He) and PIC are considered as primary measures for the dissection of genetic diversity and shedding light on the mutation rate, as well as the evolutionary pressure on a specific allele in a population over the period of time (Botstein et al., 1980; Shete et al., 2000). Overall GD of a population is mainly explained by the distribution of informative markers on a genome (Nielsen et al., 2014). Genetic diversity provides gene diversity of haploid bi-allelic SNPs and ranges of genetic distance, as well as average heterozygosity among individuals of a population (Nei, 1978). As expected, in the current study, the overall mean GD value was greater than PIC (Table 1). In the absence of more polymorphic alleles and ubiquitously even distribution of allele frequency of markers, PIC is always lower than its GD. The PIC values of SNP markers affect the classification of informative markers: highly informative, moderately informative, and slightly informative. Highly informative markers could be used in studying population genetics and GD in various plant species (Salem et al., 2021).

Understanding GD and population structure is a prerequisite to comprehending the genetic variability of germplasm before it can be used in a biotic or abiotic stress breeding program. In the current study, genome-wide mean GD was 0.18, ranging from 0.005 to 0.97 in the panel evaluated, which is slightly lower than 0.26 by Yang et al. (2020) and 0.29 by Mourad et al. (2020) but slightly higher than 0.14 observed by Alipour et al. (2017). The genome-wide heterozygosity (HZ) was 0.0736, ranging from 0 to 0.89. Mean HZ varied greatly among sub-genomes, with the lowest value (0.064) for B and the highest (0.085) for the D sub-genome. A similar trend of HZ values was also reported in other studies (Eltaher et al., 2018; Rimbert et al., 2018; Liu et al., 2019; Kumar et al., 2020).

AMOVA data suggested significant GD within sub-groups (80%), which might result from continuous selection for specific traits under certain environments by wheat breeders. A low level of genetic variability between populations (20%) (Table 2) may be due to schematic gene flow in the form of germplasm exchange among different countries and regions (Arora et al., 2014). The high genetic diversity (80%) within sub-groups means that diversity was attributed to variation within population sub-groups. The results of clustering and AMOVA suggested that inter-cluster cultivars crossing may be useful for developing promising agronomic and disease resistance associated traits in local germplasm. Higher diversity within sub-population than among sub-populations was in line with the findings from other studies (Alipour et al., 2017; Bhatta et al., 2018; Eltaher et al., 2018; Kumar et al., 2020). The Nm value is another important parameter to estimate the gene flow among subpopulations. Nm haploid value of 1 or >1 indicates the low rate of gene flow between subpopulations. In our experiment, the Nm haploid value of 0.497 indicates minor gene flow among three sub-populations because the current germplasm collection does not represent a diverse geographical area. The current population was mainly collected from the Punjab province of Pakistan.

LD is a non-random co-segregation of single/multiple loci among the same or different chromosomes of a genome. The magnitude estimation of LD decay helps determine the resolution of association mapping and the number of SNPs needed for effective association studies and marker-assisted selection (Davey et al., 2011). LD extent varies with sub-genomes or even with chromosomes of the same sub-genome and depends on various factors such as recombination rate, allele frequency, genetic drift, population structure, natural selection, and chromosomal rearrangements (Sukumaran et al., 2015; Habib et al., 2020). Higher values of LD among pairs of markers indicate slower decay rates on a genome or a chromosome, suggesting that fewer markers are needed for GWAS. The current study was conducted on a diverse set of 184 Pakistani spring wheat accessions, and the LD decay at 0.34 Mbp for the whole genome, and 0.29 Mbp, 0.2 Mbp, and 0.07 Mbp for all subgenomes A, B, and D, respectively, observed (Figure 5). In the current study, the A sub-genome showed the highest value of LD (0.29) which shows relatively fewer markers needed for genome coverage and comparatively slower LD decay due to its evolutionary history of origination. However, at the chromosomal level, chromosome 3A showed the highest value (0.12) (Table 4) of LD and chromosome 1A showed the minimum value of LD decay (0.09).

In the case of the B sub-genome chromosomes, 3B and 2B showed maximum and minimum values of LD decay of 0.13 and 0.01, respectively. In the case of D, the sub-genome chromosomes 2D and 4D showed LD values of 0.16 and 0.01, respectively (Table 4). Chromosomal level LD decay helps determine the level of decay, LD hotspots, and number of markers needed for genome coverage on various regions of chromosomes. Results of this study were consistent with the findings of Chao et al. (2010) and Ladejobi et al. (2018) reported the LD decay of 6.4, 4.5, 4, and 4.98 Mbp for A, B, D, and whole-genome, respectively. Aleksandrov et al. (2021) performed a study on Bulgarian bread wheat germplasm to dissect the population structure and linkage disequilibrium on two populations, including one population of old germplasm and one set of modern semi-dwarf cultivars. Old germplasm showed an LD decay of 3.6, 3.3, 3.1, and 3.3 Mb in subgenomes A, B, and D and the whole-genome, respectively. These results support the results of the current study. Li et al. (2019) reported the lowest LD decay value at the D genome, such as 0.35, 0.75, and 0.25 Mb at the A, B, and D subgenomes, respectively. However, the highest LD value was observed in the B sub-genome of the population. Zhang et al. (2013) reported that the highest LD decay rate for the A sub-genome ranges from 25 to 30 cM and on the B and D sub-genomes, LD decay ranges from 15 to 20 cM). Sehgal et al. (2017) reported the LD decay of values of 10, 4, 8, and 5 cM for A, B, D, and whole-genome of the wheat population, respectively. These results also confirm our results of the highest LD decay at the A sub-genome.

Usually, the highest LD value was detected in the D genome in some previous reports (Edae et al., 2015; Sukumaran et al., 2015; Liu W. et al., 2017; Ayana et al., 2018; Bhatta et al., 2018; Jamil et al., 2019; Sandeep et al., 2020). The higher LD value for the D sub-genome might be due to limited and shallow infusion of Aegilops tauschii with tetraploid wheat in evolutionary history. The slow LD decay rate in the D sub-genome suggests that more markers may be needed for GWAS, association mapping, and marker-assisted selection (MAS) than those that are needed for the A and B sub-genomes. High LD value in the A sub-genome indicates a few markers needed to cover the genome.

Conclusion

Analysis of population structure, PIC, LD, PCoA, and NJ phylogenetic tree was used to dissect the genetic diversity of 184 Pakistani genotypes. The whole wheat population was divided into three distinct subpopulations, with group 1 and group 3 dominated by the local lines and group 2 somewhat intermediate type mixture of local and CIMMYT lines. Most of the local genotypes were distributed in all three sub-groups with CIMMYT lines because local accessions were developed by crossing one or both CIMMYT lines as breeding parents. This shows that a huge amount of exotic blood mixing among local germplasm leads toward the ample amount of genetic diversity in local germplasm. Pakistan continuously faces problems such as sudden heatwaves, unusual rainfall, water shortage, and unexpected weather changes due to prompt climate change. Most Pakistani spring wheat genotypes are prone to yellow stripe rust because pathogens get resistant to hot, humid conditions due to unexpected rainfall at the vegetative stage. The results of population structure, phylogenetic tree based genetic distribution, moderately informative nature, and the number of SNP markers, PIC value, and LD regions of this population could be used for the detection of genetic diversity in the local germplasm, and this high genetic diversity would be used for association mapping studies and the selection of crossing parents in national wheat breeding programs. The results of the current genetic diversity study provided the information necessary to broaden the genetic bases and conservation studies. This information will urge national wheat breeders to speed up wheat breeding by using the genetic diversity of local and exotic germplasm. The results of the current study will help dissect dissecting genetic diversity in local germplasm, as well as for GWAS-related studies, to improve the local wheat germplasm of the country.

Data Availability Statement

The datasets presented in this study can be found in online repositories. The names of the repository/repositories and accession number(s) can be found below: NCBI SRA; PRJNA783303.

Author Contributions

SH and MH were involved in manuscript witting and data analysis. SH, FA, BS, AK, RM, and ZA were involved in the planning and execution of the experiment. SH, AB, PA, and NG were involved in sample preparation, DNA extraction, SNP library preparation, and GBS sequencing. SH, GB, FA, and RM were involved in reviewing and editing the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Higher Education Commission (HEC) of Pakistan through the International Research Support Initiative Program (IRSIP) (Grant No: 1-8/HEC/HRD/2019/8749).

Conflict of Interest

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.

Publisher’s Note

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.

Acknowledgments

The authors thank the Wheat Research Institute, Ayub Agricultural Research Institute (AARI), and Faisalabad for providing germplasm for this experiment. They are especially thankful to the USDA Central Small Grain Hard Winter Wheat Genotyping Lab, Kansas State University, Manhattan, KS, United States, for GBS and molecular data analysis facilities.

Supplementary Material

The Supplementary Material for this article can be found online at: https://www.frontiersin.org/articles/10.3389/fgene.2022.772517/full#supplementary-material

References

Akhunov, E. D., Akhunova, A. R., Anderson, O. D., Anderson, J. A., Blake, N., Clegg, M. T., et al. (2010). Nucleotide Diversity Maps Reveal Variation in Diversity Among Wheat Genomes and Chromosomes. BMC Genomics 11, 1–22. doi:10.1186/1471-2164-11-702

PubMed Abstract | CrossRef Full Text | Google Scholar

Aleksandrov, V., Kartseva, T., Alqudah, A. M., Kocheva, K., Tasheva, K., Börner, A., et al. (2021). Genetic Diversity, Linkage Disequilibrium and Population Structure of Bulgarian Bread Wheat Assessed by Genome-wide Distributed SNP Markers: From Old Germplasm to Semi-Dwarf Cultivars. Plants 10, 1116. doi:10.3390/plants10061116

PubMed Abstract | CrossRef Full Text | Google Scholar

Alemu, A., Feyissa, T., and Letta, T. (2020). Abeyo, B. Genetic Diversity and Population Structure Analysis Based on the High Density SNP Markers in Ethiopian Durum Wheat (Triticum Turgidum Ssp. Durum). BMC Genet. 21, 1–2. doi:10.1186/s12863-020-0825-x

PubMed Abstract | CrossRef Full Text | Google Scholar

Ali, S., Gladieux, P., Leconte, M., Gautier, A., Justesen, A. F., Hovmøller, M. S., et al. (2014). Origin, Migration Routes and Worldwide Population Genetic Structure of the Wheat Yellow Rust Pathogen Puccinia Striiformis F. Sp. Tritici. Plos. Pathog. 10, e1003903. doi:10.1371/journal.ppat.1003903

PubMed Abstract | CrossRef Full Text | Google Scholar

Alipour, H., Bihamta, M. R., Mohammadi, V., Peyghambari, S. A., Bai, G., and Zhang, G. (2017). Genotyping-by-sequencing (GBS) Revealed Molecular Genetic Diversity of Iranian Wheat Landraces and Cultivars. Front. Plant Sci. 8, 1293. doi:10.3389/fpls.2017.01293

PubMed Abstract | CrossRef Full Text | Google Scholar

Allen, A. M., Barker, G. L., Wilkinson, P., Burridge, A., Winfield, M., Coghill, J., et al. (2013). Discovery and Development of Exome‐based, Co‐dominant Single Nucleotide Polymorphism Markers in Hexaploid Wheat (Triticum aestivum L.). Plant Biotechnol. J. 11, 279–295. doi:10.1111/pbi.12009

PubMed Abstract | CrossRef Full Text | Google Scholar

Arora, A., Kundu, S., Dilbaghi, N., Sharma, I., and Tiwari, R. (2014). Population Structure and Genetic Diversity Among Indian Wheat Varieties Using Microsatellite (SSR) Markers. Aust. J. Crop Sci. 8, 9. doi:10.1080/15427528.2020.1818342

CrossRef Full Text | Google Scholar

Arya, S., Mishra, D. K., and Bornare, S. S. (2013). Screening Genetic Variability in advance Lines for Drought Tolerance of Bread Wheat (Triticum aestivum). Biosc 8, 1193–1196.

Google Scholar

Arzani, A., and Ashraf, M. (2017). Cultivated Ancient Wheats (Triticum spp.): A Potential Source of Health‐beneficial Food Products. Compr. Revi. Food Sci. Food Saf. 16, 477–488. doi:10.1111/1541-4337.12262

CrossRef Full Text | Google Scholar

Ayana, G. T., Ali, S., Sidhu, J. S., Gonzalez Hernandez, J. L., Turnipseed, B., and Sehgal, S. K. (2018). Genome-wide Association Study for Spot Blotch Resistance in Hard winter Wheat. Front. Plant Sci. 9, 926. doi:10.3389/fpls.2018.00926

PubMed Abstract | CrossRef Full Text | Google Scholar

Baloch, F. S., Alsaleh, A., Shahid, M. Q., Çiftçi, V., Sáenz de Miera, E., Aasim, M., et al. (2017). A Whole Genome DArTseq and SNP Analysis for Genetic Diversity Assessment in Durum Wheat from central fertile-crescent. Plos one 12, e0167821. doi:10.1371/journal.pone.0167821

PubMed Abstract | CrossRef Full Text | Google Scholar

Berkman, P. J., Visendi, P., Lee, H. C., Stiller, J., Manoli, S., Lorenc, M. T., et al. (2013). Dispersion and Domestication Shaped the Genome of Bread Wheat. Plant Biotechnol. J. 11, 564–571. doi:10.1111/pbi.12044

PubMed Abstract | CrossRef Full Text | Google Scholar

Bhatta, M., Regassa, T., Rose, D. J., Baenziger, P. S., Eskridge, K. M., Santra, D. K., et al. (2017). Genotype, Environment, Seeding Rate, and Top‐dressed Nitrogen Effects on End‐use Quality of Modern Nebraska winter Wheat. J. Sci. Food Agr. 97, 5311–5318. doi:10.1002/jsfa.8417

CrossRef Full Text | Google Scholar

Bhatta, M., Morgounov, A., Belamkar, V., Poland, J., and Baenziger, P. S. (2018). Unlocking the Novel Genetic Diversity and Population Structure of Synthetic Hexaploid Wheat. BMC Genomics 19, 1–12. doi:10.1186/s12864-018-4969-2

PubMed Abstract | CrossRef Full Text | Google Scholar

Botstein, D., White, R. L., Skolnick, M., and Davis, R. W. (1980). Construction of a Genetic Linkage Map in Man Using Restriction Fragment Length Polymorphisms. Am. J. Hum. Genet. 32, 314.

PubMed Abstract | 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, 26332635. doi:10.1093/bioinformatics/btm308

PubMed Abstract | CrossRef Full Text | Google Scholar

Browning, B. L., and Browning, S. R. (2009). A Unified Approach to Genotype Imputation and Haplotype-phase Inference for Large Data Sets of Trios and Unrelated Individuals. Ameri. J. Hum. Genet. 84, 210–223. doi:10.1016/j.ajhg.2009.01.005

CrossRef Full Text | Google Scholar

Cavanagh, C. R., Chao, S., Wang, S., Huang, B. E., Stephen, S., and Kiani, S. (2013). Genome-wide Comparative Diversity Uncovers Multiple Targets of Selection for Improvement in Hexaploid Wheat Landraces and Cultivars. P. Nat. A. Sci. 110, 8057–8062. doi:10.1073/pnas.1217133110

PubMed Abstract | CrossRef Full Text | Google Scholar

Chao, S., Zhang, W., Dubcovsky, J., and Sorrells, M. (2007). Evaluation of Genetic Diversity and Genome-wide Linkage Disequilibrium Among U.S. Wheat (Triticum aestivum L.) Germplasm Representing Different Market Classes. Crop Sci. 47, 1018–1030. doi:10.2135/cropsci2006.06.0434

CrossRef Full Text | Google Scholar

Chao, S., Zhang, W., Akhunov, E., Sherman, J., Ma, Y., and Luo, M. C. (2009). Analysis of Gene-Derived SNP Marker Polymorphism in US Wheat (Triticum aestivum L.) Cultivars. Mol. Breed. 23, 23–33. doi:10.1007/s11032-008-9210-6

CrossRef Full Text | Google Scholar

Chao, S., Dubcovsky, J., Dvorak, J., Luo, M. C., Baenziger, S. P., and Matnyazov, R. (2010). Population-and Genome-specific Patterns of Linkage Disequilibrium and SNP Variation in spring and winter Wheat (Triticum aestivum L.). BMC Genomics 11, 1–7. doi:10.1186/1471-2164-11-727

PubMed Abstract | CrossRef Full Text | Google Scholar

Chen, X., Min, D., Yasir, T. A., and Hu, Y. G. (2012). Genetic Diversity, Population Structure and Linkage Disequilibrium in Elite Chinese winter Wheat Investigated with SSR Markers. PLoS One 7, e44510. doi:10.1371/journal.pone.0044510

PubMed Abstract | CrossRef Full Text | Google Scholar

Chen, T., Tantasawat, P. A., Wang, W., Gao, X., and Zhang, L. (2018). Population Structure of Chinese Southwest Wheat Germplasms Resistant to Stripe Rust and Powdery Mildew Using the DArT-Seq Technique. Ciência Rural 16, 48. doi:10.1590/0103-8478cr20160066

CrossRef Full Text | Google Scholar

Das, M. K., Bai, G., Mujeeb-Kazi, A., and Rajaram, S. (2016). Genetic Diversity Among Synthetic Hexaploid Wheat Accessions (Triticum aestivum) with Resistance to Several Fungal Diseases. Genet. Resour. Crop Ev. 63, 1285–1296. doi:10.1007/s10722-015-0312-9

CrossRef Full Text | Google Scholar

Davey, J. W., Hohenlohe, P. A., Etter, P. D., Boone, J. Q., Catchen, J. M., and Blaxter, M. L. (2011). Genome-wide Genetic Marker Discovery and Genotyping Using Next-Generation Sequencing. Nat. Rev. Genet. 12, 499–510. doi:10.1038/nrg3012

PubMed Abstract | CrossRef Full Text | Google Scholar

Earl, D. A., and VonHoldt, B. M. (2012). STRUCTURE HARVESTER: A Website and Program for Visualizing STRUCTURE Output and Implementing the Evanno Methods. Conser. Genet. Resour. 4 (2), 359–361. doi:10.1007/s12686-011-9548-7

CrossRef Full Text | Google Scholar

Edae, E. A., Bowden, R. L., and Poland, J. (2015). Application of Population Sequencing (POPSEQ) for Ordering and Imputing Genotyping-By-Sequencing Markers in Hexaploid Wheat. G3: Gen. Genom Genet. 5, 2547–2553. doi:10.1534/g3.115.020362

PubMed Abstract | CrossRef Full Text | Google Scholar

El-Esawi, M. A., Witczak, J., Abomohra, A. E., Ali, H. M., and Elshikh, M. S. (2018). Analysis of the Genetic Diversity and Population Structure of Austrian and Belgian Wheat Germplasm within a Regional Context Based on DArT Markers. Genes 9, 47. doi:10.3390/genes9010047

PubMed Abstract | CrossRef Full Text | Google Scholar

Elshire, R. J., Glaubitz, J. C., Sun, Q., Poland, J. a., Kawamoto, K., Buckler, E. S., et al. (2011). A Robust, Simple Genotyping-By-Sequencing (GBS) Approach for High Diversity Species. PLoS One 6, e19379. doi:10.1371/journal.pone.0019379

PubMed Abstract | CrossRef Full Text | Google Scholar

Eltaher, S., Sallam, A., Belamkar, V., Emara, H. A., Nower, A. A., Salem, K. F. M., et al. (2018). Genetic Diversity and Population Structure of F3:6 Nebraska Winter Wheat Genotypes Using Genotyping-By-Sequencing. Front. Genet. 9, 1–9. doi:10.3389/fgene.2018.00076

PubMed Abstract | CrossRef Full Text | Google Scholar

Excoffier, L., Smouse, P. E., and Quattro, J. M. (1992). Analysis of Molecular Variance Inferred from Metric Distances Among DNA Haplotypes: Application to Human Mitochondrial DNA Restriction Data. Genetics 131, 479–491. doi:10.1093/genetics/131.2.479

PubMed Abstract | CrossRef Full Text | Google Scholar

Fischer, R. A., Byerlee, D., and Edmeades, G. (2014). Crop Yields and Global Food Security. Canberra, ACT: ACIAR, 8–11.

Google Scholar

Food and Agriculture Organization of the United Nations (2019). World Food Situation. Available at: http://www.fao.org/worldfoodsituation/csdb/en/ (Accessed August 19, 2020).

Google Scholar

Getachew, S. E., Ngalle, H. B., Joseph, M. B., and Wosene, G. (2019). Genotyping by Sequencing for Plant Breeding- A Review. Adv. Biotechnol. Microbiol. 14, 555891. doi:10.19080/AIBM.2019.14.555891

CrossRef Full Text | Google Scholar

Govindaraj, M., Vetriventhan, M., and Srinivasan, M. (2015). Importance of Genetic Diversity Assessment in Crop Plants and its Recent Advances: an Overview of its Analytical Perspectives. Genet. Res. Int. 2015, 431487. doi:10.1155/2015/431487

PubMed Abstract | CrossRef Full Text | Google Scholar

Habib, M., Awan, F. S., Sadia, B., and Zia, M. A. (2020). Genome-Wide Association Mapping for Stripe Rust Resistance in Pakistani Spring Wheat Genotypes. Plants 9, 1056. doi:10.3390/plants9091056

PubMed Abstract | CrossRef Full Text | Google Scholar

Haudry, A., Cenci, A., Ravel, C., Bataillon, T., Brunel, D., Poncet, C., et al. (2007). Grinding up Wheat: a Massive Loss of Nucleotide Diversity since Domestication. Mol. Bio. Evol. 24, 1506–1517. doi:10.1093/molbev/msm077

PubMed Abstract | CrossRef Full Text | Google Scholar

Hawkes, J. G. (1981). “Germplasm Collection, Preservation, and Use,” in 2 Plant Breeding Symposium (Iowa: Iowa State University Press).

Google Scholar

He, J., Zhao, X., Andre, L., Lu, Z. H., HongKui, L., and Ziqin, L. (2014). Genotyping-by-Sequencing (GBS), an Ultimate Marker-Assisted Selection (MAS) Tool to Accelerate Plant Breeding. Front. Plant Sci. 5, 484. doi:10.3389/fpls.2014.00484

PubMed Abstract | CrossRef Full Text | Google Scholar

Heslot, N., Rutkoski, J., Poland, J., Jannink, J. L., and Sorrells, M. E. (2013). Impact of Marker Ascertainment Bias on Genomic Selection Accuracy and Estimates of Genetic Diversity. PloS one 8, e74612. doi:10.1371/journal.pone.0074612

PubMed Abstract | CrossRef Full Text | Google Scholar

Hill, W. G., and Weir, B. S. (1988). Variances and Covariances of Squared Linkage Disequilibria in Finite Populations. Theor. Popul. Biol. 33, 54–78. doi:10.1016/0040-5809(88)90004-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Iehisa, J. C., Shimizu, A., Sato, K., Nishijima, R., Sakaguchi, K., Matsuda, R., et al. (2014). Genome-wide Marker Development for the Wheat D Genome Based on Single Nucleotide Polymorphisms Identified from Transcripts in the Wild Wheat Progenitor Aegilops Tauschii. Theor. Appl. Genet. 127, 261–271. doi:10.1007/s00122-013-2215-5

PubMed Abstract | CrossRef Full Text | Google Scholar

IWGSC (2018). Shifting the Limits in Wheat Research and Breeding Using a Fully Annotated Reference Genome. Science 361, 7191. doi:10.1126/science.aar7191

PubMed Abstract | CrossRef Full Text | Google Scholar

Jamil, M., Ali, A., Gul, A., Ghafoor, A., Napar, A. A., Ibrahim, A. M., et al. (2019). Genome-wide Association Studies of Seven Agronomic Traits under Two Sowing Conditions in Bread Wheat. BMC Plant Biol. 19, 1–8. doi:10.1186/s12870-019-1754-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Jia, J., Zhao, S., Kong, X., Li, Y., Zhao, G., He, W., et al. (2013). International Wheat Genome Sequencing, Aegilops Tauschii Draft Genome Sequence Reveals a Gene Repertoire for Wheat Adaptation. Nature 496, 91–95. doi:10.1038/nature12028

PubMed Abstract | CrossRef Full Text | Google Scholar

Kumar, S., Kumari, J., Bhusal, N., Pradhan, A. K., Budhlakoti, N., Mishra, D. C., et al. (2020). Genome-wide Association Study Reveals Genomic Regions Associated with Ten Agronomical Traits in Wheat under Late-Sown Conditions. Front. Plant Sci. 11, 1420. doi:10.3389/fpls.2020.549743

PubMed Abstract | CrossRef Full Text | Google Scholar

Ladejobi, O., Mackay, I. J., Poland, J., Praud, S., Hibberd, J/M., and Bentley, A. R. (2018). Reference Genome Anchoring of High-Density Markers for Association Mapping and Genomic Prediction in European winter Wheat. Front. Plant Sci. 10, 1278. doi:10.3389/fpls.2019.01278

PubMed Abstract | CrossRef Full Text | Google Scholar

Li, L., Peng, Z., Mao, X., Wang, J., Chang, X., Reynolds, M., et al. (2019). Genome-wide Association Study Reveals Genomic Regions Controlling Root and Shoot Traits at Late Growth Stages in Wheat. Ann. Bot. 124, 993–1006. doi:10.1093/aob/mcz041

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu, K., and Muse, S. V. (2005). PowerMaker: an Integrated Analysis Environment for Genetic Maker Analysis. Bioinformatics 21, 2128–2129. doi:10.1093/bioinformatics/bti282

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu, Y. G., and Tsunewaki, K. (1991). Restriction Fragment Length Polymorphism (RFLP) Analysis in Wheat. II. Linkage Maps of the RFLP Sites in Common Wheat. Jpn. J. Genet. 66, 617–633. doi:10.1266/jjg.66.617

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu, J., Rasheed, A., He, Z., Imtiaz, M., Arif, A., Mahmood, T., et al. (2019). Genome-wide Variation Patterns between Landraces and Cultivars Uncover Divergent Selection during Modern Wheat Breeding. Theor. Appl. Genet. 132, 2509–2523. doi:10.1093/aob/mcz04110.1007/s00122-019-03367-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu, W., Maccaferri, M., Chen, X., Laghetti, G., Pignone, D., Pumphrey, M., et al. (2017). Genome-wide Association Mapping Reveals a Rich Genetic Architecture of Stripe Rust Resistance Loci in Emmer Wheat (Triticum Turgidum Ssp. Dicoccum). Theor. Appl. Genet. 130, 2249–2270. doi:10.1007/s00122-017-2957-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu, Y., Lin, Y., Gao, S., Li, Z., Ma, J., Deng, M., et al. (2017). A Genome‐wide Association Study of 23 Agronomic Traits in Chinese Wheat Landraces. Plant J. 91, 861–873. doi:10.1111/tpj.13614

PubMed Abstract | CrossRef Full Text | Google Scholar

Lobell, D. B., Bänziger, M., Magorokosho, C., and Vivek, B. (2011). Nonlinear Heat Effects on African maize as Evidenced by Historical Yield Trials. Nat. Clim. Change. 1, 42–45. doi:10.1038/nclimate1043

CrossRef Full Text | Google Scholar

Lopes, M. S., Dreisigacker, S., Peña, R. J., Sukumaran, S., and Reynolds, M. P. (2015). Genetic Characterization of the Wheat Association Mapping Initiative (WAMI) Panel for Dissection of Complex Traits in spring Wheat. Theor. Appl. Geneti. 128, 453–464. doi:10.1007/s00122-014-2444-2

CrossRef Full Text | Google Scholar

Marcussen, T., Sandve, S. R., Heier, L., Spannagl, M., Pfeifer, M., Jakobsen, K. S., et al. (2014). Ancient Hybridizations Among the Ancestral Genomes of Bread Wheat. Science 345, 6194. doi:10.1126/science.1250092

PubMed Abstract | CrossRef Full Text | Google Scholar

Mohanty, S. K., and Swain, M. R. (2019). “Bioethanol Production from Corn and Wheat: Food, Fuel, and Future,” in Bioethanol Production from Food Crops (Academic Press), 45–59. doi:10.1016/b978-0-12-813766-6.00003-5

CrossRef Full Text | Google Scholar

Mourad, A. M., Belamkar, V., and Baenziger, P. S. (2020). Molecular Genetic Analysis of spring Wheat Core Collection Using Genetic Diversity, Population Structure, and Linkage Disequilibrium. BMC Genomics 21, 1–2. doi:10.1186/s12864-020-06835-0

CrossRef Full Text | Google Scholar

Nei, M. (1978). Estimation of Average Heterozygosity and Genetic Distance from a Small Number of Individuals. Genetics 89, 583–590. doi:10.1093/genetics/89.3.583

PubMed Abstract | CrossRef Full Text | Google Scholar

Nielsen, N. H., Backes, G., Stougaard, J., Andersen, S. U., and Jahoor, A. (2014). Genetic Diversity and Population Structure Analysis of European Hexaploid Bread Wheat (Triticum aestivum L.) Varieties. PloS ONE 9, e94000. doi:10.1371/journal.pone.0094000

PubMed Abstract | CrossRef Full Text | Google Scholar

Novoselović, D., Bentley, A. R., Šimek, R., Dvojković, K., Sorrells, M/E., Gosman, N., et al. (2016). Characterizing Croatian Wheat Germplasm Diversity and Structure in a European Context by DArT Markers. Front. Plant Sci. 7, 184. doi:10.3389/fpls.2016.00184

PubMed Abstract | CrossRef Full Text | Google Scholar

Ogbonnaya, F. C., Abdalla, O., Mujeeb-Kazi, A., Kazi, A. G., Xu, S. S., Gosman, N., et al. (2013). Synthetic Hexaploids: Harnessing Species of the Primary Gene Pool for Wheat Improvement. Plant Breed. Rev. 37, 35–122. doi:10.1002/9781118497869.ch2

CrossRef Full Text | Google Scholar

Peakall, R. O., and Smouse, P. E. (2006). GENALEX 6: Genetic Analysis in Excel. Population Genetic Software for Teaching and Research. Mol. Ecol. Notes. 6, 288–295. doi:10.1111/j.1471-8286.2005.01155.x

CrossRef Full Text | Google Scholar

Peng, J., Korol, A. B., Fahima, T., Röder, M. S., Ronin, Y. I., Li, Y. C., et al. (2000). Molecular Genetic Maps in Wild Emmer Wheat, Triticum Dicoccoides: Genome-wide Coverage, Massive Negative Interference, and Putative Quasi-Linkage. Genome Res. 10, 1509–1531. doi:10.1101/gr.150300

PubMed Abstract | CrossRef Full Text | Google Scholar

Percival, J. (1922). The Wheat Plant; A Monograph. Duckworth and Company. doi:10.1038/109366a0

CrossRef Full Text | Google Scholar

Peterson, G. W., Dong, Y., Horbach, C., and Fu, Y. B. (2014). Genotyping-by-sequencing for Plant Genetic Diversity Analysis: a Lab Guide for SNP Genotyping. Diversity 6, 665–680. doi:10.3390/d6040665

CrossRef Full Text | Google Scholar

Poland, J. A., Endelman, J., Dawson, J., Rutkoski, J., Wu, S., Manes, Y., et al. (2012a). Genomic Selection in Wheat Breeding Using Genotyping-By-Sequencing. Plant Genome 5, 103–113. doi:10.3835/plantgenome2012.06.0006

CrossRef Full Text | Google Scholar

Poland, J. A., Brown, P. J., Sorrells, M. E., and Jannink, J. L. (2012b). Development of High-Density Genetic Maps for Barley and Wheat Using a Novel Two-Enzyme Genotyping-By-Sequencing Approach. PloS One 7, e32253. doi:10.1371/journal.pone.0032253

PubMed Abstract | CrossRef Full Text | Google Scholar

Pritchard, J. K., Stephens, M., and Donnelly, P. (2000). Inference of Population Structure Using Multilocus Genotype Data. Genetics 155, 945–959. doi:10.1093/genetics/155.2.945

PubMed Abstract | CrossRef Full Text | Google Scholar

Ray, D. K., Mueller, N. D., West, P. C., and Jonathan, A. F. (2013). Yield Trends Are Insufficient to Double Global Crop Production by 2050. PLoS One 8, 1–8. doi:10.1371/journal.pone.0066428

PubMed Abstract | CrossRef Full Text | Google Scholar

Remington, D. L., Thornsberry, J. M., Matsuoka, Y., Wilson, L. M., Whitt, S. R., Doebley, J., et al. (2001). Structure of Linkage Disequilibrium and Phenotypic Associations in the maize Genome. P. Nat. A. Sci. 98, 11479–11484. doi:10.1073/pnas.201394398

PubMed Abstract | CrossRef Full Text | Google Scholar

Rimbert, H., Darrier, B., Navarro, J., Kitt, J., Choulet, F., Leveugle, M., et al. (2018). High Throughput Snp Discovery and Genotyping in Hexaploid Wheat. PLoS One 13, 1–9. doi:10.1371/journal.pone.0186329

CrossRef Full Text | Google Scholar

Röder, M. S., Korzun, V., Wendehake, K., Plaschke, J., Tixier, M. H., Leroy, P., et al. (1998). A Microsatellite Map of Wheat. Genetics 149, 2007–2023. doi:10.1093/genetics/149.4.2007

PubMed Abstract | CrossRef Full Text | Google Scholar

Röder, M. S., Plaschke, J., König, S. U., Börner, A., Sorrells, M. E., Tanksley, S. D., et al. (1995). Abundance, Variability and Chromosomal Location of Microsatellites in Wheat. Mol. Gen. Genet. 246 (3), 327–333. doi:10.1007/BF00288605

PubMed Abstract | CrossRef Full Text | Google Scholar

Rufo, R., Alvaro, F., Royo, C., and Soriano, J. M. (2019). From Landraces to Improved Cultivars: Assessment of Genetic Diversity and Population Structure of Mediterranean Wheat Using SNP Markers. PloS One 14, e0219867. doi:10.1371/journal.pone.0219867

PubMed Abstract | CrossRef Full Text | Google Scholar

Saitou, N., and Nei, M. (1987). The Neighbor-Joining Method: a New Method for Reconstructing Phylogenetic Trees. Mol. Biol. Evol. 4, 406–425. doi:10.1093/oxfordjournals.molbev.a040454

PubMed Abstract | CrossRef Full Text | Google Scholar

Salem, K. F., Elabsawy, E. A., Khidr, Y. A., and Elshamy, H. (2021). Assessment Genetic Diversity Among Some Egyptian Bread Wheat (Triticum aestivum L.) Cultivars Using RAPD Markers. J. Genet. Environ. Reso. Conser. 26, 168–175.

Google Scholar

Sandeep, C., Kumar, A., Rodrigues, V., Viswanath, S., Shukla, A. K., and Sundaresan, V. (2020). Morpho-genetic Divergence and Population Structure in Indian Santalum Album L. Trees 34, 1113–1129. doi:10.1007/s00468-020-01963-2

CrossRef Full Text | Google Scholar

Sehgal, D., Autrique, E., Singh, R., Ellis, M., Singh, S., and Dreisigacker, S. (2017). Identification of Genomic Regions for Grain Yield and Yield Stability and Their Epistatic Interactions. Sci. Rep. 7, 1–2. doi:10.1038/srep41578

PubMed Abstract | CrossRef Full Text | Google Scholar

Sener, O., Arslan, M., Soysal, Y., and Erayman, M. (2009). Estimates of Relative Yield Potential and Genetic Improvement of Wheat Cultivars in the Mediterranean Region. J. Agri. Sci. 147, 323–332. doi:10.1017/S0021859609008454

CrossRef Full Text | Google Scholar

Shavrukov, Y., Suchecki, R., Eliby, S., Abugalieva, A., Kenebayev, S., and Langridge, P. (2014). Application of Next-Generation Sequencing Technology to Study Genetic Diversity and Identify Unique SNP Markers in Bread Wheat from Kazakhstan. BMC Plant Biol. 14, 1–3. doi:10.1186/s12870-014-0258-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Shete, S., Tiwari, H., and Elston, R. C. (2000). On Estimating the Heterozygosity and Polymorphism Information Content Value. Theor. Popul. Biol. 57, 265–271. doi:10.1006/tpbi.2000.1452

PubMed Abstract | CrossRef Full Text | Google Scholar

Sofalian, O., Chaparzadeh, N., Javanmard, A., and Hejazi, M. S. (2008). Study the Genetic Diversity of Wheat Landraces from Northwest of Iran Based on ISSR Molecular Markers. Int. J. Agri. Bio. 10, 466–468.

Google Scholar

Sukumaran, S., Dreisigacker, S., Lopes, M., Chavez, P., and Reynolds, M. P. (2015). Genome-wide Association Study for Grain Yield and Related Traits in an Elite spring Wheat Population Grown in Temperate Irrigated Environments. Theor. Appl. Genet. 128, 353–363. doi:10.1007/s00122-014-2435-3

PubMed Abstract | CrossRef Full Text | Google Scholar

Tengberg, M. (1999). Crop Husbandry at Miri Qalat Makran, SW Pakistan (4000–2000 BC). Veg. Hist. Archaeobot. 8, 3–12. doi:10.1007/bf02042836

CrossRef Full Text | Google Scholar

Tyrka, M., Mokrzycka, M., Bakera, B., Tyrka, D., Szeliga, M., Stojałowski, S., et al. (2021). Evaluation of Genetic Structure in European Wheat Cultivars and Advanced Breeding Lines Using High-Density Genotyping-By-Sequencing Approach. BMC Genomics 22, 1–7. doi:10.1186/s12864-020-07351-x

PubMed Abstract | CrossRef Full Text | Google Scholar

Verma, S., Gupta, S., Bandhiwal, N., Kumar, T., Bharadwaj, C., and Bhatia, S. (2015). High-density Linkage Map Construction and Mapping of Seed Trait QTLs in Chickpea (Cicer Arietinum L.) Using Genotyping-By-Sequencing (GBS). Sci. Rep. 5, 1–4. doi:10.1038/srep17512

CrossRef Full Text | Google Scholar

Voss-Fels, K. P., Stahl, A., Wittkop, B., Lichthardt, C., Nagler, S., Rose, T., et al. (2019). Breeding Improves Wheat Productivity under Contrasting Agrochemical Input Levels. Nat. Plants 5, 706–714. doi:10.1038/s41477-019-0445-5

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, N., Thomson, M., Bodles, W. J., Crawford, R. M., Hunt, H. V., Featherstone, A. W., et al. (2013). Genome Sequence of dwarf Birch (Betula Nana) and Cross‐species RAD Markers. Mol. Ecol. 22, 3098–3111. doi:10.1111/mec.12131

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, Q., Tian, F., Pan, Y., Buckler, E., and Zhang, Z. (2014). A SUPER Powerful Method for Genome Wide Association Study. PloS One 9, e107684. doi:10.1371/journal.pone.0107684

PubMed Abstract | CrossRef Full Text | Google Scholar

Würschum, T., Maurer, H. P., Kraft, T., Janssen, G., Nilsson, C., and Reif, J. C. (2011). Genome-wide Association Mapping of Agronomic Traits in Sugar Beet. Theor. Appl. Genet. 123, 1121–1131. doi:10.1007/s00122-011-1653-1

PubMed Abstract | CrossRef Full Text | Google Scholar

Yang, X., Tan, B., Liu, H., Zhu, W., Xu, L., Wang, Y., et al. (2020). Genetic Diversity and Population Structure of Asian and European Common Wheat Accessions Based on Genotyping-By-Sequencing. Front. Genet. 11, 1157. doi:10.3389/fgene.2020.580782

PubMed Abstract | CrossRef Full Text | Google Scholar

You, Q., Yang, X., Peng, Z., Xu, L., and Wang, J. (2018). Development and Applications of a High Throughput Genotyping Tool for Polyploid Crops: Single Nucleotide Polymorphism (SNP) Array. Front. Plant Sci. 9, 104. doi:10.3389/fpls.2018.00104

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, K., Wang, J., Zhang, L., Rong, C., Zhao, F., Peng, T., et al. (2013). Association Analysis of Genomic Loci Important for Grain Weight Control in Elite Common Wheat Varieties Cultivated with Variable Water and Fertiliser Supply. PloS one 8, e57853. doi:10.1371/journal.pone.0057853

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhao, L., Abdelsalam, N. R., Xu, Y., Chen, M. S., Feng, Y., Kong, L., et al. (2020). Identification of Two Novel Hessian Fly Resistance Genes H35 and H36 in a Hard winter Wheat Line SD06165. Theor. Appl. Genet. 133, 2343–2353. doi:10.1007/s00122-020-03602-3

PubMed Abstract | CrossRef Full Text | Google Scholar

Zorić, M., Dodig, D., Kobiljski, B., Quarrie, S., and Barnes, J. (2012). Population Structure in a Wheat Core Collection and Genomic Loci Associated with Yield under Contrasting Environments. Genetica 140, 259–275. doi:10.1007/s10709-012-9677-2

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: genotyping-by-sequencing, genetic diversity, population structure, Triticum aestivum L., single nucleotide polymorphisms, polymorphic information content, Pakistan

Citation: Hussain S, Habib M, Ahmed Z, Sadia B, Bernardo A, Amand PS, Bai G, Ghori N, Khan AI, Awan FS and Maqbool R (2022) Genotyping-by-Sequencing Based Molecular Genetic Diversity of Pakistani Bread Wheat (Triticum aestivum L.) Accessions. Front. Genet. 13:772517. doi: 10.3389/fgene.2022.772517

Received: 08 September 2021; Accepted: 07 January 2022;
Published: 06 April 2022.

Edited by:

Reyazul Rouf Mir, Sher-e-Kashmir University of Agricultural Sciences and Technology, India

Reviewed by:

Ali Raza, Fujian Agriculture and Forestry University, China
Mohd Anwar Khan, Sher-e-Kashmir University of Agricultural Sciences and Technology, India

Copyright © 2022 Hussain, Habib, Ahmed, Sadia, Bernardo, Amand, Bai, Ghori, Khan, Awan and Maqbool. 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: Faisal S. Awan, faisal.saeed@uaf.edu.pk

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.