Skip to main content
Scientific Reports logoLink to Scientific Reports
. 2015 Jul 24;5:12468. doi: 10.1038/srep12468

Genome-wide high-throughput SNP discovery and genotyping for understanding natural (functional) allelic diversity and domestication patterns in wild chickpea

Deepak Bajaj 1, Shouvik Das 1, Saurabh Badoni 1, Vinod Kumar 2, Mohar Singh 3, Kailash C Bansal 3, Akhilesh K Tyagi 1, Swarup K Parida 1,a
PMCID: PMC4513697  PMID: 26208313

Abstract

We identified 82489 high-quality genome-wide SNPs from 93 wild and cultivated Cicer accessions through integrated reference genome- and de novo-based GBS assays. High intra- and inter-specific polymorphic potential (66–85%) and broader natural allelic diversity (6–64%) detected by genome-wide SNPs among accessions signify their efficacy for monitoring introgression and transferring target trait-regulating genomic (gene) regions/allelic variants from wild to cultivated Cicer gene pools for genetic improvement. The population-specific assignment of wild Cicer accessions pertaining to the primary gene pool are more influenced by geographical origin/phenotypic characteristics than species/gene-pools of origination. The functional significance of allelic variants (non-synonymous and regulatory SNPs) scanned from transcription factors and stress-responsive genes in differentiating wild accessions (with potential known sources of yield-contributing and stress tolerance traits) from cultivated desi and kabuli accessions, fine-mapping/map-based cloning of QTLs and determination of LD patterns across wild and cultivated gene-pools are suitably elucidated. The correlation between phenotypic (agromorphological traits) and molecular diversity-based admixed domestication patterns within six structured populations of wild and cultivated accessions via genome-wide SNPs was apparent. This suggests utility of whole genome SNPs as a potential resource for identifying naturally selected trait-regulating genomic targets/functional allelic variants adaptive to diverse agroclimatic regions for genetic enhancement of cultivated gene-pools.


In chickpea, a plethora of array-based SNP (single nucleotide polymorphism) genotyping approaches [including Illumina GoldenGate/Infinium (Bead Xpress array) and KBioscience Competitive Allele-Specific Polymerase chain reaction (KASPar) assays] are known to greatly expedite the large-scale validation and high-throughput genotyping of previously discovered SNPs in diverse accessions, specifically for genetic diversity studies, phylogenetics and genetic linkage map construction1,2,3,4,5. The implications of RAD-seq (restriction-site-associated DNA sequencing) that relies on barcoded multiplexing and the RE (restriction enzyme)-based NGS (next-generation sequencing) approach for simultaneous mining and genotyping of genome-wide SNPs are well illustrated in chickpea for comprehending the molecular diversity pattern among its diverse desi, kabuli and wild accessions6. Recently, with the emergence of an improved (high barcoded genotype multiplexing capacity), user friendly and resource-saving NGS-based genotyping-by-sequencing (GBS) assay, the potential of simultaneous high-throughput genome-wide SNPs discovery and genotyping have improved manifold, involving diverse crop accessions7,8,9. The fast SNP mining and genotyping capacity of the GBS assay gives it an edge over other available high-throughput genotyping assays that have been utilized in myriad genomics-assisted breeding applications, such as genome-wide high-resolution trait association and genetic mapping in diverse crop plants9,10,11,12,13,14,15,16,17,18,19,20,21. With readily available chickpea draft genome sequences22,6, the large-scale validation and high-throughput genotyping of SNPs at a genome-wide scale employing the GBS approach seems rational and feasible for diverse chickpea accessions. Unfortunately, hitherto, no such efforts have been made in chickpea, specifically in cases of diverse, naturally occurring wild germplasm accessions.

Chickpea [Cicer arietinum (L.)] supposedly originated from the central region of the Fertile Crescent (presently South Eastern Turkey and Syria) along with its close progenitors, the wild annual Cicer species23,24,25,26,27. The genus Cicer consists of 44 annual and perennial species grouped into three gene pools (primary, secondary and tertiary) according to their crossability with cultivated chickpea28,29. The annual wild species (C. reticulatum) representing the primary gene pool with contrasting characteristics of high crossability with cultivated C. arietinum (desi and kabuli) serves as a potential source to broaden the genetic base and enhance the yield-component and stress tolerance traits in the cultivated gene pool30,31,32,33,34,35. To accelerate the process of genetic improvement through inter-specific hybridization, specifically the understanding of molecular diversity and domestication patterns among wild and cultivated Cicer accessions representing each of the three gene pools at a genome-wide scale is essential. Furthermore, this process can be complemented with marker-based introgression of trait-associated novel genes, QTLs (quantitative trait loci) and natural allelic variants scanned from diverse wild gene pools into cultivated accessions.

Significant efforts have been made towards understanding the genetic diversity pattern, population structure and phylogenetic relationship among diverse annual and perennial Cicer species using various random and sequence-based robust SSR (simple sequence repeat) and SNP markers6,29,36,37,38,39,40,41,42,43,44,45,46,47,48. However, the use of a limited number of selected markers in all these studies might provide unrealistic estimates of genetic variability with ambiguous molecular diversity and phylogenetic information among wild and cultivated Cicer accessions. In this context, discovery, large-scale validation and high-throughput genotyping of numerous genome-wide SNP markers using a highly efficient GBS assay could be an attractive approach to more precisely understanding the extent of natural allelic diversity, phylogenetic relationship and domestication pattern among wild and cultivated Cicer accessions. This approach will facilitate mining of novel allelic variants from wild accessions with beneficial traits for introgression breeding and broadening the genetic base of chickpea cultivars through inter-specific hybridization for their genetic enhancement.

The present study was therefore undertaken to discover, validate and genotype SNPs in 93 wild and cultivated Cicer accessions at a genome-wide scale using a GBS assay. The functional significance of identified genome-wide SNPs in trait association and their utility in studying the natural allelic diversity, population genetic structure, phylogeny and LD (linkage disequilibrium) pattern in wild and cultivated Cicer gene pools were determined.

Results and discussion

Genome-wide discovery and high-throughput genotyping of SNPs using a GBS assay

The sequencing of 96-plex ApeKI libraries through a GBS assay generated approximately 279.8 million raw sequence reads from 93 wild and cultivated Cicer accessions (Table S1). This altogether produced 248.3 million (88.7%) high-quality sequence reads (approximately 20-fold sequencing depth of coverage) that varied from 2.02 to 3.97 with a mean of 2.87 million reads per accession (Fig. S1A). On average, approximately 83.5% (ranging from 80.2 to 89.9%) and approximately 86.8% (80.1 to 91.5%) high-quality reads were evenly distributed across 93 Cicer accessions and mapped to unique physical locations on desi and kabuli reference draft genomes, respectively (Fig. S1B). These uniquely mapped non-redundant sequence reads effectively covered approximately 20.7% (153.6 Mb) and 21.5% (159.4 Mb) [ranging from 17.4% (129.3 Mb) to 34.3% (254.1 Mb)] of desi and kabuli chickpea genomes, respectively, with an estimated size of approximately 740 Mb. Notably, 107.6 (14.5%) and 126.7 (17.1%) Mb genomic regions of desi and kabuli chickpea genomes, respectively, represented by >90% and/or all 93 wild and cultivated Cicer accessions, were further utilized for genome-wide discovery and genotyping of SNPs among these accessions. These sequencing data have been submitted to the NCBI SRA (short read achieve) database (http://www.ncbi.nlm.nih.gov/sra) with accession number SRX971856, which will be made freely accessible on the 31st of August 2015.

A total of 82489 high-quality SNPs were identified, including 38511 and 43978 (with read-depth: ≥10, SNP base quality: ≥20, missing data: <10% and heterozygosity: approximately 2% in each accession) from desi and kabuli genomes, respectively. These SNPs were discovered using both reference and de novo-based GBS assays (Table 1, S2, S3). The minor allele frequency (MAF) of GBS-based SNPs detected among 93 Cicer accessions varied from 2 to 21% with an average of 14%. The significant difference in the number of SNPs identified from desi and kabuli genomes could be due to uneven sequence assemblies and lengths of chromosomal pseudomolecules (desi: 124.37 Mb and kabuli: 347.24 Mb) between these two genomes22,6. This difference in sequenced fractions between desi and kabuli genomes encouraged us to employ both these genomes as references in the GBS assay to enrich the large-scale mining and genotyping of SNPs at a genome-wide scale in wild Cicer accessions. A little bit low-quality and uneven whole genome sequence assembly, including a smaller size chromosomal pseudomolecule of desi compared with kabuli has been documented recently by Ruperao et al.49. We therefore, provide details regarding the frequency and characteristics of reference and de novo-based GBS-SNPs mined from the desi genome in the Text S1 and Table S3.

Table 1. Genomic distribution and nucleotide diversity potential of kabuli reference genome- and de novo-based GBS-SNPs among 93 wild and cultivated Cicer accessions.

        Nucleotide diversity
Chromosomes Size (Mb) of chromosomes (pseudomolecules) Number (%) of SNPs mapped Average map density (kb) θπ θω
Ca- kabuli-Chr01 48.36 1934 (16.1) 25.0 0.23 0.34
Ca- kabuli-Chr02 36.63 1118 (9.3) 32.8 0.28 0.34
Ca- kabuli-Chr03 39.99 1295 (10.8) 30.9 0.22 0.33
Ca- kabuli-Chr04 49.19 2418 (20.2) 20.3 0.30 0.35
Ca- kabuli-Chr05 48.17 1232 (10.3) 39.1 0.18 0.34
Ca- kabuli-Chr06 59.46 1900 (15.8) 31.3 0.21 0.33
Ca- kabuli-Chr07 48.96 1551 (12.9) 31.6 0.26 0.32
Ca- kabuli-Chr08 16.48 541 (4.5) 30.5 0.25 0.34
Total 347.24 11989 28.9 0.24 0.34
Ca-kabuli-Scaffold NA 3761 NA 0.21 0.37
Ca-kabuli-de novo NA 28228 NA 0.28 0.35
Total NA 43978 NA 0.24 0.33

Ca-kabuli-Chr: Cicer arietinum kabuli chromosome.

θπ: Average pair-wise nucleotide diversity.

θω: Watterson’s estimator of segregating sites.

The GBS-based SNPs identified from kabuli chickpea consisted of 15,750 and 28,228 reference genome- and de novo-based SNPs, respectively. The reference genome-based SNPs included 11989 SNPs that were physically mapped on eight kabuli chromosomes with an average map density of 28.9 kb (Table 1, Fig. 1A). The remaining 3761 SNPs were physically mapped on kabuli genome scaffolds. The average SNP map density was highest on kabuli chromosome 4 (20.3 kb) and lowest on chromosome 5 (39.1 kb). A higher proportion of SNPs were physically mapped on kabuli chromosome 4 (20.2%, 2,418) (Table 1). The number of SNPs physically mapped on eight kabuli and desi (Text S1, Table S3, Fig. S2A) chromosomes revealed a direct correlation with their pseudomolecule size (bp). The constructed SNP-based physical maps of desi and kabuli chromosomes could serve as references for faster selection of genome-wide SNPs for manifold high-throughput marker-aided genetic analysis, including targeted mapping of genomes and trait-regulatory genes/QTLs in wild Cicer as well as comparative genome mapping involving chickpea and other legumes. The transitions were more frequent than transversions, which made up slightly more than half (54.7%; 45099 SNPs) of the 82489 identified SNPs (Table S4). A higher frequency of A/G transitions (51.2%, 23108 SNPs) compared to C/G (30.8%, 11,510) and G/T (30.2%, 11290) transversions in both desi and kabuli genomes was apparent. In total, 15750 reference kabuli genome-based SNPs have been submitted to NCBI dbSNP (http://www.ncbi.nlm.nih.gov/SNP/snp_viewTable.cgi?handle=NIPGR) with SNP submission (SS) accession numbers 1399931543 to 1399947292. Collectively, the integrated reference (desi and kabuli)- and de novo-based GBS strategies employed in our study for large-scale mining and high-throughput genotyping of SNPs in wild and cultivated Cicer accessions at a genome-wide scale could have multidimensional applicability in genomics-assisted breeding of chickpea.

Figure 1. Genomic distribution and relative frequency of 11989 GBS-based SNPs structurally annotated on chickpea genome/genes.

Figure 1

(A). The distribution of 11989 GBS-based genome-wide SNPs physically mapped on eight kabuli chickpea chromosomes is illustrated in the Circos circular ideogram. The outermost and innermost circles represent the different colours-coded chromosomes and distribution of SNPs, including non-synonymous SNPs (marked with red dots), respectively. (B) Frequency distribution of 15750 GBS-based SNPs mined in the intergenic regions and diverse coding and non-coding sequence components of 3371 genes annotated from kabuli genome. Number and proportion of SNPs, including synonymous and non-synonymous SNPs annotated in the coding as well as non-coding intronic and regulatory sequences of kabuli chickpea genes and intergenic regions are depicted. The URR (upstream regulatory region) and DRR (downstream regulatory region) of genes were defined as per the gene annotation information of kabuli genome (Varshney et al.6).

Structural and functional annotation of GBS-based SNPs

The structural annotation of 15750 kabuli reference genome-based SNPs detected 7,859 (49.9%) SNPs in the intergenic regions and 7,891 (50.1%) SNPs in 3,371 genes (Fig. 1B). The number of SNPs identified in the kabuli genes ranged from 1 to 7, with a mean of 2.3 SNPs/gene. The highest proportion (47.4%, 3,737 SNPs) of SNPs was annotated in the exons (CDS), followed by introns (33.8%, 2671) and DRRs (12.2%, 963 SNPs) and the lowest was found (6.6%, 520) in URRs (Fig. 1B). A total of 1,736 (46.5%) and 2,001 (53.5%) coding SNPs in 1,090 and 1,171 genes revealed synonymous and non-synonymous substitutions, respectively. The non-synonymous SNPs included 1,951 (97.5%) missense and 50 (2.5%) nonsense SNPs in 1,152 and 49 genes, respectively. The details regarding structural and functional annotation of 12,112 desi reference genome-based GBS-SNPs are noted in the Text S2 and illustrated in the Fig. S2B, S3A. The KOG-based functional annotation (excluding unknown and general functions) of SNP-carrying kabuli genes exhibited their primary roles in the signal transduction mechanism (13.7%), followed by transcription (11.2%) (Fig. S3B). The structurally and functionally annotated SNPs can be deployed to select functionally relevant gene-derived non-synonymous and regulatory SNPs for rapidly establishing efficient marker-trait linkages and identifying genes/QTLs controlling traits of agricultural importance in wild chickpea.

Large-scale validation and polymorphic potential of SNPs

The comparison of 12112 desi (ICC 4958) reference genome-based SNPs mined from 93 wild and cultivated Cicer accessions with the available SNP database (31,019 SNPs) of four chickpea accessions (ICC 4958, ICC 4951, ICC 12968 and ICC 17160) revealed a correlation of 1716 SNPs (7.3%) between past genome resequencing and our present GBS data. These results are based on congruent physical positions and the type/nature of SNPs (Table S5). Among these, a maximum number of SNPs (831 SNPs, 10.7%) were found to be common and polymorphic between ICC 4958 and ICC 17160. The PCR amplicon sequencing of 96 selected SNPs identified by the reference genome- and de novo-based GBS assays efficiently validated 87 (90.1%) SNPs in a representative set of wild and cultivated Cicer accessions (Fig. S4). Therefore, a high degree of reproducibility (100%) as well as in silico and experimental validation success rate (90%) of SNPs discovered from 93 wild and cultivated Cicer accessions by use of both reference (desi and kabuli)- and de novo-based GBS approaches was evident. This suggests the robustness of the GBS assay in rapid discovery and high-throughput genotyping of high-quality and non-erroneous SNPs covering the whole genome in wild chickpea.

A total of 27862 SNPs (MAF ≥0.05), including 12112 and 15750 SNPs identified from desi and kabuli genomes, respectively, exhibited polymorphism among 93 diverse wild and cultivated Cicer accessions with a higher PIC (0.01 to 0.45, mean 0.35) and nucleotide diversity (θπ: 0.26 and θω: 0.25) (Table S6). Inter-specific polymorphism among species (23,794, 85.4% polymorphism and mean PIC: 0.41) was higher than that of intra-specific polymorphism within wild (15,790, 66.4% and 0.35) and cultivated (10,602, 38.1% and 0.27) species. The intra-specific polymorphic potential detected by SNPs was highest in C. judaicum (47.7% polymorphism, mean PIC: 0.39, θπ: 0.30 and θω: 0.29), followed by C. reticulatum (45.2%, 0.37, θπ: 0.27 and θω: 0.28) and lowest in C. arietinum (38.1%, 0.27, 0.16 and 0.18) (Table S6). The SNPs mapped on desi chromosomes (mean θπ: 0.28 and θω: 0.33) showed an almost comparable nucleotide diversity level with those mapped on the kabuli chromosomes (0.24 and 0.34) (Table 1, S3). The nucleotide diversity was highest on chromosomes 8 (mean θπ: 0.34 and θω: 0.33) and 4 (0.30 and 0.35) of desi and kabuli, respectively. The polymorphic potential detected by SNPs among 56 accessions representing the diverse species of the secondary gene pool (73.8% polymorphism, mean PIC: 0.39, θπ: 0.36 and θω: 0.33) was higher in contrast to that among the 36 accessions of the primary gene pool (65.7%, 0.31, 0.27 and 0.30) (Table S6). According to the geographical origin of 93 wild and cultivated Cicer accessions, 76 accessions belonging to different wild species were categorized under the Fertile Crescent (Turkey, Syrian Arab Republic, Jordan, Lebanon and Israel) type, while 17 accessions were grouped under Central Asia (India). Species/accessions originating from the Fertile Crescent had a higher SNP polymorphic potential (90.4% polymorphism, mean PIC: 0.37, θπ: 0.37 and θω: 0.35) than those from Central Asia (44.1%, 0.26, 0.24 and 0.28) (Table S6). The inter- and intra-specific polymorphic potential (66–85%) and PIC (0.35–0.41) detected by genome-wide GBS-based SNPs among 93 wild and cultivated Cicer accessions are higher/comparable to those estimated earlier using large-scale SNP, intron-spanning and microsatellite markers29,39,41,43,47. Therefore, an appreciable sum of polymorphic potential assessed by GBS-based SNPs covering the whole genome by their simultaneous high-throughput discovery and genotyping in wild and cultivated Cicer accessions can be used to assess the natural allelic diversity and domestication pattern in wild and cultivated chickpea.

Genome-wide SNP-based molecular diversity and population genetic structure in wild chickpea

The use of 27862 genome-wide SNPs (MAF ≥0.05) for studying the molecular diversity among 93 wild and cultivated Cicer accessions belonging to six populations (POP I–POP VI) exhibited a wider range of genetic distances from 0.06 (POP I: ICC 4951 between POP I: ICCV 92311) to 0.64 (POP I: IC 296132 between POP IV: ILWC 32) with a mean of 0.43 (Table S7). This is higher/comparable to the molecular diversity level documented in earlier studies41,44,45,50 but lower than the data obtained using large-scale microsatellite and SNP markers47,29. The genetic distance among the accessions within wild species based on 15790 genome-wide SNPs varied from 0.06 (POP IV: ILWC 285 and POP IV: IG 136796) to 0.64 (POP II: ILWC 288 and POP VI: C. microphyllum) with an average of 0.42 (Table S7). The accessions belonging to POP III (mean genetic distance: 0.31) had the highest molecular diversity, followed by POP II (0.28), POP IV (0.23) and POP I (0.19) (Table 2). The species/accessions included under the secondary gene pool (mean genetic distance: 0.28) exhibited greater genetic diversity than those in the primary gene pool (0.21). The accessions/species originating from the Fertile Crescent (mean genetic distance: 0.27) possessed greater diversity than those from Central Asia (0.20). The SNPs revealing a wider genetic base and broader molecular diversity among wild and cultivated Cicer accessions could have significance in the selection of useful diverse parental wild and cultivated accessions as well as in the precise identification of true inter-specific hybrids in introgression breeding programs. Consequently, it will be helpful to scan the transfer of target genomic (gene) regions governing diverse agronomic traits (yield contributing and abiotic/biotic stress tolerance traits) from wild Cicer gene pools51 into the backgrounds of cultivated (desi and kabuli) species for their genetic improvement. The GBS-based SNPs developed at a genome-wide scale have the potential to distinguish among wild and cultivated Cicer accessions and would thus be of importance for chickpea variety improvement.

Table 2. Polymorphic and molecular diversity potential estimated within six structured populations using GBS-based genome-wide SNPs.

Populations Number of accessions PIC (polymorphism information content) Minor allele frequency (MAF) Nucleotide diversity Genetic distance
θπ θω Range Mean
POP I 12 0.08–0.38 (0.27) 0.10 0.16 0.18 0.06 (ICC4951–ICCV92311) to 0.32 (ICC4958–ICCV95334) 0.19
POP II 24 0.06–0.42 (0.36) 0.17 0.28 0.29 0.01 (ILWC253–ILWC242) to 0.44 (ILWC233–ILWC258) 0.28
POP III 15 0.04–0.45 (0.39) 0.21 0.30 0.29 0.11 (ILWC9–ILWC22) to 0.53 (ILWC51–ILWC9) 0.31
POP IV 19 0.05–0.38 (0.35) 0.18 0.26 0.27 0.07 (IG136796–ILWC285) to 0.39 (ILWC42–ILWC32) 0.23
POP V 22 0.06–0.37 (0.30) 0.16 0.25 0.26 0.12 (ICC182–ILWC20) to 0.40 (ILWC283–ILWC273) 0.21
POP VI 1 NA NA NA NA NA NA
All POP 93 0.010.45 (0.35) 0.15 0.26 0.25 0.06 (ICC4951ICCV92311) to 0.53 (ILWC51ILWC9) 0.30

The determination of the population genetic structure among 93 wild and cultivated Cicer accessions using 27862 genome-wide SNPs at varying levels of possible population numbers (K = 1 to 10) with 20 replications revealed the most apparent inflection of average LnP(D) (log-likelihood) at one of the best replicates of K = 6. The population numbers (K) were further validated using the second order statistics of STRUCTURE by ΔK estimation. Overall, these analyses support the classification of 93 accessions into six distinct populations (POP I–VI) with a high-resolution population structure (Fig. 2A). POP I consisted of 12 cultivated desi and kabuli accessions of C. arietinum. POP II consisted of 24 wild accessions from C. reticulatum (16) and C. echinospermum (8). POP III, IV, V and VI included accessions from C. judaicum (22 accessions), C. bijugum (18), C. pinnatifidum (16) and C. microphyllum (1), respectively. The high-resolution assignment of 93 accessions into six populations was comparable with their clustering patterns and phylogeny obtained by the neighbor-joining unrooted phylogram (Fig. 2B) and PCA (Fig. 2C). The classification of species/accessions representing members of primary (C. arietinum, C. reticulatum and C. echinospermum) and secondary (C. bijugum, C. pinnatifidum and C. judaicum) gene pools into six diverse populations is comparable with previous molecular diversity, population genetic structure and evolutionary studies6,29,36,38,39,40,41,42,44,45,46,47,52,53,54,55. The inclusion of accessions from wild annual Cicer species (C. reticulatum and C. echinospermum) into one population (POP II) infers that geographical origin (Turkey), including phenotypic characteristics (growth habits), rather than species/gene pools of origination had a greater impact on the clustering patterns in a wild population51.

Figure 2. Genome-wide SNP-based molecular diversity, phylogeny and population genetic structure among 93 wild and cultivated Cicer accessions.

Figure 2

(A) Population genetic structure among 93 wild and cultivated accessions using 27862 genome-wide SNPs. These SNPs assigned 93 accessions into six populations (POP I, POP II, POP III, POP IV, POP V and POP VI) that majorly grouped as per their species and gene pools of origination. The accessions represented by vertical bars along the horizontal axis were classified into K colour segments based on their estimated membership fraction in each K cluster. Six diverse colours represent different population groups based on optimal population number K = 6. (B) Unrooted phylogram illustrating the genetic relationships (Nei’s genetic distance) among 93 wild and cultivated accessions belonging to seven Cicer species using 27862 genome-wide SNPs. The phylogenetic tree clearly differentiated 93 accessions into six diverse groups, which correspond to their species and gene pools of origination. (C) Principal component analysis (PCA) differentiating the 93 wild and cultivated accessions belonging to seven Cicer species into six populations (POP I, POP II, POP III, POP IV, POP V and POP VI) as determined by population genetic structure. The PC1, PC2, PC3, PC4, PC5 and PC6 explained 10.2%, 7.8%, 3.7%, 3.4%, 2.7%, 2.4% and 2.3% of the total variance, respectively.

The ability of SNPs to detect polymorphism among the six populations revealed that the highest polymorphic potential occurred in POP III (mean PIC: 0.39, MAF: 0.21, θπ: 0.30 and θω: 0.29), followed by POP II and lowest in POP I (Table 2). A narrow genetic base and lower molecular diversity among cultivated desi and kabuli accessions of C. arietinum (POP I) was observed despite being closely related to the more diverse C. reticulatum and C. echinospermum (POP II). This is possibly due to the combined impacts of sequential evolutionary bottlenecks (such as founder effects, adaption-based selection pressure and modern breeding efforts) during the evolutionary divergence and domestication of cultivated gene pools with wild gene pools in South Eastern Turkey24,25,26,29,56. These observations are well supported by our estimation of higher molecular diversity in the wild gene pools originating from the Fertile Crescent than the gene pools from Central Asia. The molecular genetic variation among and within six populations of 93 Cicer accessions using 27862 genome-wide SNPs revealed a wider level of significant genetic and population differentiation based on pair-wise FST (P < 0.001) that varied from 0.17 (POP I and POP V) to 0.48 (POP III and POP V) with an average of 0.39. Higher divergence between populations (FST: 0.43) compared to that estimated among accessions within populations (0.35) was evident. The SNPs physically mapped on the desi chromosomes (FST varied from 0.02 to 0.42, mean: 0.28) showed a lower potential for population differentiation compared to those on kabuli chromosomes (0.02 to 0.47, mean: 0.37) (Fig. S5). The maximum average FST of 0.26 and 0.33 were detected by SNPs mapped on desi and kabuli chromosomes 8 and 4, respectively.

All 93 wild and cultivated Cicer accessions clearly belonged to a structured population, with approximately 93% inferred ancestry derived from one of the model-based populations and the remaining approximately 7% containing admixed ancestry. The highest admixed ancestry (approximately 9%) was observed in POP IV, followed by POP II (approximately 5%) and lowest in POP I (approximately 1%). The occurrence of admix ancestry among six populations gave clues regarding origination and domestication of wild and cultivated Cicer from a common ancestor along with founder crops at the Fertile Crescent (Eastern Mediterranean region) approximately 10000–12000 years ago24,57. Diverse admixture patterns observed among the six populations could have evolved through complex domestication patterns involving inter-crossing/introgression coupled with various strong adaptive selection pressures among wild and cultivated Cicer accessions after their subsequent evolutionary divergence from progenitors. POP IV (C. bijugum) had the highest admixed ancestry (approximately 20%) with POP V (C. pinnatifidum), which is expected because they both belong to secondary gene pools. POP I showed the greatest admixtures (approximately 10%) with POP II and POP VI. The admixture was not detected in 11.8% (11 accessions) of chickpea accessions, while 15% (14) of the accessions exhibited >10% admixed ancestry. The highest admixtures among the members of the primary gene pool POP I (C. arietinum) and POP II (C. reticulatum and C. echinospermum) reflect their close phylogenetic relationship during their domestication at the archaeological sites of South Eastern Turkey24,29. A higher admixed ancestry of POP VI (Indian originated perennial wild accession C. microphyllum) with POP I (C. arietinum) and POP II (C. reticulatum and C. echinospermum) primary gene pools is in line with an earlier study29.

To infer the precise domestication patterns among 93 wild and cultivated Cicer accessions, the molecular diversity level, phylogeny and population genetic structure-related information revealed by genome-wide GBS-based SNPs were correlated with their qualitative and quantitative agromorphological traits, evaluated at two diverse agro-climatic zones in North Western India. A wider level of phenotypic variation, for instance, in three quantitative traits, BN (5 to 64), PN (1 to 309) and SW (1.0 to 70 g), was observed among 93 wild and cultivated Cicer accessions belonging to six populations (POP I to POP VI) based on two years of multi-location replicated field data (Table S8). All 81 wild accessions belonging to five populations also had higher phenotypic diversity levels for three agromorphological traits (BN: 5–64, PN: 1–309 and SW: 1–18 g) (Table S8). A wider phenotypic diversity among Cicer accessions was further supported by their broader Euclidean genetic distance (1.56 to 7.89, mean: 4.63) and Shannon-Weaver diversity coefficient (0.26 to 0.79, mean: 0.48). The dendrogram and PCA-based clustering of Cicer accessions into six populations (POP I to POP VI) and trends of phenotypic diversity level observed among the six populations using the phenotypic diversity coefficient remained similar to those determined by the molecular diversity distance matrix and population structure using genome-wide GBS-based SNPs. A wider phenotypic diversity among wild Cicer gene pools might be due to the diverse geographical origination and domestication of accessions included under wild species across various agro-climatic regions of the world. The correlation between phenotypic (agromorphological) and molecular diversity-based domestication pattern across wild and cultivated accessions could assist us in deciphering the effect of natural adaptive selection at a genome-wide scale on Cicer gene pools that are adapted to diverse agro-climatic regions. This will eventually lead to the delineation of trait-associated target genomic regions and natural allelic variants functionally relevant for chickpea genetic enhancement.

Recently, the added advantages of selecting genome-wide markers from LD/haplotype blocks of chromosomes and their effective use in detecting a wider spectrum of natural allelic diversity have been well demonstrated in HapMap projects of many crop plants58,59,60. In this context, the genome-wide GBS-based SNP genotyping data utilized in our study for realistic estimation of molecular diversity, genetic structure, phylogeny and domestication pattern among 93 wild and cultivated Cicer accessions seem rational. This is relevant considering the suitability of genome-wide GBS-based SNPs in distinct differentiation of these accessions belonging to wild and cultivated gene pools into six populations and correspondence of phylogenetic relationships with their species/gene pools and geographical origination. This precise diversity-related information generated by us could eventually facilitate diverse genomics-assisted breeding applications, including association and genetic (QTL) mapping to identify functionally relevant genes/QTLs regulating important agronomic traits in wild chickpea.

Genome-wide and population-specific LD patterns in wild chickpea

The LD estimates (average r2) and extent of LD decay using all possible pair-combinations of the 4672 and 11989 genome-wide desi and kabuli SNPs (physically mapped on eight chromosomes) were determined among 93 wild and cultivated Cicer accessions belonging to six populations (POP I-POP VI). In the entire population and across eight desi and kabuli chickpea chromosomes, a proportion of 4.4–21% SNP-pairs exhibited significant LD (P < 0.0001) (Table 3, S9), indicating a moderate LD level in a diversity panel of wild and cultivated chickpea accessions. The significant LD (%) was highest in desi (13.3%) and kabuli (10.2%) chromosome 4. The LD estimates ranged from 0.53 in POP I to 0.85 in POP III with an average of 0.74 (Table 4). The LD estimates varied from 0.36 in desi chromosomes 4 and 7 to 0.67 in chromosome 2 with an average of 0.52, while in kabuli, it ranged from 0.38 in chromosome 7 to 0.67 in chromosome 1 with a mean density of 0.48 (Table 3, S9). The primary gene pool (0.80) had a higher LD estimate than the secondary gene pool (0.76). The LD estimate in the species/accessions originating from the Fertile Crescent (0.73) was higher than that in the species/accessions originating from Central Asia (0.59) (Table 5).

Table 3. Chromosome-wise LD estimates in 93 wild and cultivated Cicer accessions using kabuli reference genome-based GBS-SNP.

Chromosomes Number of marker-pairs used aNumber of significant marker-pairs in LD bSignificant LD (%) Extent of LD (r2)
Ca-kabuli-Chr01 1165945 52455 4.5 0.67
Ca-kabuli-Chr02 624404 37643 6.0 0.40
Ca-kabuli-Chr03 836571 23832 2.8 0.42
Ca-kabuli-Chr04 1516603 155216 10.2 0.39
Ca-kabuli-Chr05 758296 7816 1.0 0.51
Ca-kabuli-Chr06 1148703 39687 3.5 0.58
Ca-kabuli-Chr07 1053625 33542 3.2 0.38
Ca-kabuli-Chr08 1053625 5822 4.0 0.46
Total chromosomes 1019720 44500 4.4 0.48

aPercentage of SNP marker locus-pairs in significant (P < 0.001) LD.

bSignificant threshold (P < 0.001) at which pair-wise LD estimates is significant statistically.

Table 4. LD estimates in a structured population encompassing 93 wild and cultivated Cicer accessions using GBS-based SNPs.

Populations Number of marker-pairs used aNumber of significant marker-pairs in LD bSignificant LD (%) Extent of LD (r2)
POP I 831775 124766 15 0.53
POP II 831775 216260 26 0.82
POP III 831775 191308 23 0.85
POP IV 831775 158037 19 0.69
POP V 831775 141402 17 0.66
Total populations 831775 174672 21 0.74

aPercentage of SNP marker locus-pairs in significant (P < 0.001) LD.

bSignificant threshold (P < 0.001) at which pair-wise LD estimates is significant statistically.

Table 5. Gene pools- and geographical area-wise LD estimates 93 wild and cultivated Cicer accessions using GBS-based SNPs.

Characteristics Number of marker-pairs used aNumber of significant marker-pairs in LD bSignificant LD (%) Extent of LD (r2)
Primary gene pool(36 accessions) 831775 24254 2.9 0.80
Secondary gene pool(76 accessions) 831775 32708 3.9 0.76
Geographical area(Central Asia) 831775 19131 2.3 0.59
Geographical area(Fertile Crescent) 831775 41650 5.0 0.73

aPercentage of SNP marker locus-pairs in significant (P < 0.001) LD.

bSignificant threshold (P < 0.001) at which pair-wise LD estimates is significant statistically.

The LD decay of 4672 desi and 11989 kabuli genome-based SNPs was determined individually across six populations by pooling the r2 estimates across eight chromosomes and plotting their average r2 against the 50 kb uniform physical distance of 0–1000 kb (Fig. 3, S6). A non-linear regression curve depicted a decreasing trend of LD decay with an increase in the physical distance (kb) in the six populations. No significant decay of LD (r2 ≤ 0.1) was observed in any of the six populations (POP I–POP VI) up to a 1000 kb physical distance for both desi and kabuli chromosomes (Fig. 3, S6). However, the r2 decreased to half of its maximum value at an approximately 450–500 kb physical distance in desi chromosomes and 500–550 kb in kabuli chromosomes (Fig. 3, S6). The LD estimates and chromosomal LD decay estimated in the six populations using the genome-wide GBS-based SNPs are much higher compared to those reported earlier in cross-pollinated58,61,62,63,64,65 and self-pollinated66,67 plant species.

Figure 3. LD decay (mean r2) estimated in five structured populations (POP I–POP V) using 11989 kabuli genome (chromosome)-derived GBS-SNPs.

Figure 3

The average r2 estimated among SNPs spaced with uniform 50 kb physical intervals from 0 to 1000 kb are illustrated by the plotted curved lines. The LD decay for POP VI could not be calculated due to inclusion of only one wild accession within this population group. The dotted lines were plotted between pooled r2 (across entire five populations) and physical distance (kb) based on nonlinear regression model considering the r2 = 1 at marker physical distance of 0 kb and the trend of LD decay were estimated in wild and cultivated Cicer accessions.

These findings reflect the substantial reduction of genetic diversity in wild and cultivated Cicer gene pools compared to that of other self- and cross-pollinated plant species, which is possibly due to four sequential evolutionary bottlenecks during the domestication of chickpea24,25,26,27. The diverse populations POP III (C. judaicum) and POP II (C. reticulatum and C. echinospermum) exhibited almost comparable LD decay (400–450 kb in desi and kabuli chromosomes) with other less diverse populations (450–550 kb) (Fig. 3, S6). These observations overall suggest that the SNP genome coverage/density as well as the molecular diversity, genetic structure and domestication pattern had significant contributions towards shaping the LD patterns in wild and cultivated Cicer gene pools68,69. The number as well as characteristics of wild and cultivated Cicer accessions included under a population could also be an important contributing factor for such comparable LD decay between cultivated and wild Cicer gene pools. However, the effective correlation of potential bias introduced by sequence reads mapping procedures and/or the non-randomness of missing GBS-based SNP data scanned from diverse genotyped accessions with their degree of LD decay at a genome-wide scale cannot be overruled. The population and chromosome-specific LD patterns determined in 93 wild and cultivated Cicer accessions by use of genome-wide GBS-based SNPs could assist us in anticipating the marker density required to more precisely estimate natural allelic diversity over various genomic regions and for efficient high-resolution genome-wide association study (GWAS) and genetic/QTL mapping, which will result in the identification of potential trait-regulatory genes/QTLs in chickpea.

Functional significance of genome-wide SNPs

To understand the functional significance of the SNPs identified from 93 wild and cultivated Cicer accessions at a genome-wide scale, the SNPs, including non-synonymous and regulatory SNPs differentiating accessions belonging to each of six wild species (with potential known sources of yield contributing and stress tolerance traits), from accessions of cultivated desi and kabuli were documented systematically (Figs 4I and 5I). For instance, a total of 11194 SNPs (7253 genic SNPs), including 1392 non-synonymous and 2643 regulatory SNPs discriminating the accessions belonging to wild C. reticulatum species from desi cultivated C. arietinum species were identified (Fig. 4I). Likewise, 9574 SNPs (6005 genic SNPs), including 1043 non-synonymous and 2153 regulatory SNPs showing differentiation between wild C. judaicum and sensitive kabuli cultivated C. arietinum species were detected (Fig. 5I). The identified informative SNP-carrying genes (1800 genes between desi vs. wild species and 1925 genes between kabuli vs. wild species) were functionally annotated. Furthermore, differential expression pattern of these genes were determined through in silico digital expression profiling using the global transcript profiling data from the same genes available in diverse vegetative and reproductive desi accession (ICC 4958) tissues.

Figure 4. The functional relevance of non-synonymous and regulatory SNPs-carrying genes differentiating the accessions belonging to six wild Cicer species from accessions of cultivated C. arietinum desi (I) demonstrated by their non-synonymous (Ka) to synonymous (Ks) substitution rates and time of divergence (million years ago/Mya).

Figure 4

Functional annotation of non-synonymous and regulatory SNPs-carrying genes discriminated the wild species/accessions from desi (II) accessions. The differential expression profiles (preferential and tissue-specific expression) of non-synonymous and regulatory SNPs-carrying genes discriminated the wild accessions from desi (III) accessions assayed in diverse vegetative and reproductive tissues of ICC 4958 are represented through Hierarchical cluster display. The average log signal expression values of genes in various tissues are marked with a colour scale at the top. The low, medium and high level of gene expression are indicated with green, black and red colors, respectively. At the top and right side of expression map, the tissues and genes, respectively utilized for expression profiling are mentioned. Ca: C. arietinum, Cr: C. reticulatum, Ce: C. echinospermum, Cj: C. judaicum, Cb: C. bijugum, Cp: C. pinnatifidum and Cm: C. microphyllum.

Figure 5. The functional relevance of non-synonymous and regulatory SNPs-carrying genes differentiating the accessions belonging to six wild Cicer species from accessions of cultivated C. arietinum kabuli (I) demonstrated by their non-synonymous (Ka) to synonymous (Ks) substitution rates and time of divergence (million years ago/Mya).

Figure 5

 Functional annotation of non-synonymous and regulatory SNPs-carrying genes discriminated the wild species/accessions from kabuli (II) accessions. The differential expression profiles (preferential and tissue-specific expression) of non-synonymous and regulatory SNPs-carrying genes discriminated the wild accessions from kabuli (III) accessions assayed in diverse vegetative and reproductive tissues of ICC 4958 are represented through Hierarchical cluster display. The average log signal expression values of genes in various tissues are marked with a colour scale at the top. The low, medium and high level of gene expression are indicated with green, black and red colors, respectively. At the top and right side of expression map, the tissues and genes, respectively utilized for expression profiling are mentioned. Ca: C. arietinum, Cr: C. reticulatum, Ce: C. echinospermum, Cj: C. judaicum, Cb: C. bijugum, Cp: C. pinnatifidum and Cm: C. microphyllum.

A significant proportion of these regulatory and non-synonymous SNP-carrying genes corresponded to transcription factors (16%), abiotic (3%)/biotic (3–4%) stress-responsive proteins and signal transduction enzymes (8%) (Figs 4II and 5II). Notably, 294 and 205 SNPs revealing non-synonymous (missense and nonsense) amino acid substitutions within the functional domains encoding transcription factors and biotic/abiotic stress-responsive genes discriminated the wild Cicer species/accessions from sensitive accessions of cultivated desi and kabuli, respectively (Fig. S4A, S4B). Selected 586 desi and 800 kabuli non-synonymous and regulatory SNP-containing genes, respectively exhibited significant differential expression (≥two-fold) in diverse vegetative and reproductive tissues of ICC 4958 (Figs 4III and 5III). A pronounced differential expression (≥10-fold) of these 119 desi and 168 kabuli SNP-carrying genes was apparent. Approximately 28–35% of these genes with SNPs had preferential and tissue-specific expression (≥4-fold), particularly in the roots, flower buds and young pods of ICC 4958 (Figs 4III and 5III). The significance of non-synonymous and regulatory SNPs in governing differential expression (particularly tissue-specific expression) and the possible transcriptional mechanism of genes associated with multiple agronomic traits have been demonstrated in rice, soybean and chickpea70,71,72,73,74.

To assess the potential of identified genome-wide SNPs for fine-mapping/map-based cloning of genes underlying QTLs, the markers linked to/flanking the 20 known QTLs for yield component and stress tolerance traits mapped previously on diverse inter- and intra-specific genetic linkage maps75,76,77,78,79,80,81,82,83,84,85,86,87 were selected. The genetic positions of markers flanking these known QTL intervals were compared/correlated with those of a high-density inter-specific genetic linkage map (ICC 4958 × PI 489777)88, which was used earlier as a reference for constructing physical maps and chromosome pseudomolecules for the kabuli genome6. The integration of genetic position-related information of the QTLs mapped on the reference inter-specific genetic linkage map (ICC 4958 × PI 489777) with that of physical map of kabuli chromosome pseudomolecules resulted in the ability to define the physical positions of markers flanking the 10 known stress tolerance trait-regulating QTL intervals mapped on five kabuli chromosomes.

The structural and functional annotation as well as in silico digital expression profiling (using the available global transcriptome sequencing data in diverse vegetative and reproductive tissues of ICC 4958) of SNP-carrying genes harboring these 10 known QTLs revealed significant differential expression, including preferential and tissue-specific expression of some of the selected genes with non-synonymous and regulatory SNPs. For instance, one “QTL-hotspot” region [ICCM0249 (54.9 cM)-GA24 (79.2 cM)] governing drought tolerance traits, particularly the root traits (58.2% PVE) mapped on an intra-specific genetic map (ICC 4958 × ICC 1882) of linkage group (LG) 4, was selected to evaluate the potential of our identified genome-wide GBS-based SNPs for fine mapping. A major genomic region (9102618–16180940 bp spanning 7.08 Mb on chromosome 4) harboring this robust QTL was defined by integrating that region with inter-specific genetic linkage (ICC 4958 × PI 489777) and physical map of the kabuli genome. The comprehensive GBS-based SNP analysis at this 7.08 Mb target QTL interval among 93 wild and cultivated Cicer accessions detected 375 SNPs. The structural annotation of these SNPs on the kabuli genome revealed the presence of 161 SNPs in the intergenic regions with 214, including 36 (29 genes) non-synonymous and 58 (31) regulatory, SNPs in the 60 genes. The functional annotation of 60 non-synonymous and regulatory SNP-carrying genes exhibited maximum correspondence to transcription factors (25%) and stress-responsive genes (33%). The differential expression analysis of 60 non-synonymous and regulatory SNP-containing genes based on in silico digital expression profiling showed pronounced differential up-/down-regulation (2 to 18.5-fold) of 51 genes in at least one of the tissues of ICC 4958 compared to others. The preferential and tissue-specific expression (≥4-fold) in the roots compared to the leaves of ICC 4958 was evident. The LD estimates and LD decay using 439 SNPs (with significant LD at MAF ≥0.05) identified and annotated on a 7.08 Mb drought responsive QTL region of kabuli chromosome 4 were determined. The LD estimates in the QTL region varied from 0.09 to 0.52 with a mean of 0.21, which is much lower than the LD estimates measured on all of kabuli chromosome 4 (mean LD: 0.39) using 2418 SNPs. The LD decay (r2 decreased to half of its maximum value) determined at a 200 kb physical distance from the target QTL is much faster compared to all eight kabuli chromosomes, including kabuli chromosome 4 (500–550 kb).

The estimation of the proportionate distribution of Ka to Ks substitution rates across SNP-carrying genes conserved between each of the wild and cultivated Cicer species depicted that a larger fraction (>90%) of such genes had Ka/Ks < 1.0, indicating negative/purifying selection pressure (Figs 4I and 5I). This is consistent with the average fraction of non-synonymous to synonymous SNPs (Ka/Ks < 1.0) estimated previously in chickpea genes/transcripts6,22,89,90. The Ka/Ks was highest between C. arietinum [desi (0.37)/kabuli (0.40)] and C. judaicum, and lowest between C. arietinum [desi (0.14)/kabuli (0.18)] and C. microphyllum. The time of divergence estimation based on synonymous substitution rates (Ks) among conserved SNP-carrying genes between each of the wild and cultivated Cicer species revealed the highest Ks between C. arietinum (desi/kabuli) and C. judaicum, giving their approximate divergence time at 0.61 and 0.65 Mya, respectively (Figs 4I and 5I). This was followed by that between C. arietinum [desi (0.58 Mya)/kabuli (0.62 Mya)] and C. bijugum and by the minimum between C. arietinum [desi (0.35 Mya)/kabuli (0.38 Mya)] and C. microphyllum (Figs 4I and 5I). The Ka/Ks-based estimates of divergence time measured among wild and cultivated Cicer species (specifically between C. arietinum and C. reticulatum) is consistent with a number of previous studies6,22,89,90.

The wild accessions constituting six annual and perennial Cicer species from diverse geographical origins selected for genome-wide discovery and high-throughput SNP genotyping using GBS assay had considerable phenotypic diversity with regard to multiple agromorphological and disease resistance traits based on their multi-location replicated phenotyping at two diverse agro-climatic zones of North Western India by Singh et al.51. For instance, the early flowering C. reticulatum accession ILWC 36 originating from Turkey (52 days), C. judaicum accessions (ILWC 4 and ILWC 273) from Lebanon (41 days) and C. pinnatifidum accessions (ILWC 226 and ILWC 251) also from Turkey (41 days) displayed multiple resistance against Ascochyta blight and root knot nematodes. One accession of C. judaicum (ILWC 256) originally from Jordan containing a higher number of pods/plant (284) vis-a-vis cultivated desi and kabuli Cicer species showed resistance to both Ascochyta blight and Botrytis gray mold. Therefore, large-scale non-synonymous and regulatory SNPs in the stress-responsive known/candidate genes and transcription factors differentiating the wild Cicer accessions (with potential known sources of various seed and pod yield contributing and stress tolerance traits) from sensitive cultivated desi and kabuli accessions scanned by us could potentially be utilized as functional markers for various large-scale genetic analyses in chickpea. The practical significance of these functional alleles (SNPs) mined at a genome-wide scale was apparent for fine-mapping/map-based cloning of genes underlying QTLs as well as for understanding natural allelic diversity and domestication pattern among diverse wild and cultivated Cicer species/accessions across gene pools based on their population genetic structure and LD patterning. These findings could help us to identify potential natural targets for adaptive trait genetic enhancement in cultivated gene pools that are domesticated in diverse agro-climatic regions. This would eventually accelerate genomics-assisted breeding, including hybridization breeding programs by introgressing diverse traits of agronomic importance from the wild Cicer gene pools into cultivated Cicer species for genetic improvement.

Methods

Mining and high-throughput genotyping of SNPs

For genome-wide discovery and genotyping of SNPs, 93 diverse Cicer accessions representing one annual cultivated C. arietinum (12 accessions), five annual wild species, namely C. reticulatum (16), C. echinospermum (8), C. judaicum (22), C. bijugum (19) and C. pinnatifidum (15) and one perennial species C. microphyllum were selected (Table S1). The genomic DNA from the young leaf samples of these accessions were isolated using QIAGEN DNeasy 96 Plant Kit (QIAGEN, USA). Two accessions (one each representing the C. reticulatum and C. judaicum) were used as biological replicates to evaluate the reproducibility of GBS-based SNPs. To constitute 96-plex GBS libraries, the genomic DNA of 95 accessions was digested with ApeKI and ligated to adapters carrying unique barcodes, while one sample (without genomic DNA) was considered as control. These libraries were pooled together and sequenced (100-bp single end) using Illumina HiSeq2000 adopting the methods of Elshire et al.8 and Spindel et al.13. The high-quality FASTQ sequence reads generated from accessions were de-multiplexed according to their unique barcodes and mapped to reference desi (ICC 495822) and kabuli (CDC Frontier6) draft chickpea genome sequences using Bowtie v2.1.0. To identify accurate and high-quality SNPs (with SNP base-quality ≥20 supported by minimum sequence read-depth of 10) from 95 accessions, the sequence alignment map files generated from desi and kabuli genomes were analyzed individually using reference-based GBS pipeline of STACKS. The sequence reads remained unaligned on desi and kabuli reference genomes were further analyzed individually using the de novo genotyping approach of STACKS v1.0 (http://creskolab.uoregon.edu/stacks) to identify the high-quality and non-erroneous SNPs from 93 accessions.

Annotation and validation of SNPs

To deduce the genomic distribution of GBS-based SNPs across chickpea chromosomes, the total SNPs, including non-synonymous SNPs were plotted individually based on their unique physical positions (bp) on eight chromosomes (pseudomolecules) of desi and kabuli genomes and visualised using Circos visualisation tool. For structural and functional annotation of SNPs in diverse coding and non-coding sequence components of genes and genomes (chromosomes/pseudomolecules and scaffolds), the physical positions (bp) of reference-based SNPs were correlated with the GFF of desi22 and kabuli6 chickpea genome annotation. The outputs were further analysed using the customised perl scripts, single-nucleotide polymorphism effect predictor (SnpEff v3.1h; http://snpeff.sourceforge.net), PFAM database v27.0 (http://pfam.sanger.ac.uk) and KOGnitor NCBI database (ftp://ftp.ncbi.nih.gov/pub/COG/KOG).

The SNP genotyping data obtained from 93 wild and cultivated Cicer accessions was correlated with an available in silico SNP database of four chickpea accessions (ICC 4958, ICC 4951, ICC 12968 and ICC 17160) based on their nature/types and congruent physical positions on the desi (ICC 4958) genome22 to validate the GBS-derived SNPs. For experimental validation of SNPs, the primers designed from 200-bp flanking sequences of 96 selected SNPs were PCR amplified using the genomic DNA of a representative set of wild and cultivated Cicer accessions. The PCR amplicons were sequenced and aligned to detect SNPs among accessions as per Kujur et al.80,81.

Evaluation of polymorphic potential, molecular diversity and population genetic structure

The SNP genotyping information (minor allele frequency/MAF ≥5% with <10% missing data) generated from 93 wild and cultivated Cicer accessions was analysed with PowerMarker v3.51, MEGA v5.0 and TASSEL v3.0 (http://www.maizegenetics.net). Based on these outcomes, the PIC (polymorphism information content), average pair-wise nucleotide diversity (θπ) and Watterson’s estimator of segregating sites (θω) were measured and an unrooted neighbour-joining (NJ)-based phylogenetic tree (with 1000 bootstrap replicates) among accessions was constructed. To determine optimal value of population number (K) and population structure among 93 accessions, the SNP genotyping information was analyzed in STRUCTURE v2.3.4 following Kujur et al.80 and Saxena et al.29. The principal component analysis (PCA) among accessions was performed using TASSEL v3.0 and GAPIT.

Assessment of phenotypic diversity

The replicated multilocation (New Delhi and Himachal Pradesh) field phenotyping information of 93 wild and cultivated Cicer accessions for diverse yield contributing qualitative and quantitative traits [plant pigmentation, plant hairiness, seed colour, days to flowering and maturity time, plant height, 100-seed weight (SW), branch number/plant (BN) and pod number/plant (PN)] were obtained from Singh et al.51. The detailed analysis of these agronomic traits based on diverse statistical measures (mean, standard deviation and coefficient of variation) were performed following Singh et al.51. The standardization of phenotypic data, estimation of genetic distance based on Euclidean coefficient (EUCLID), PCA, Shannon-Weaver diversity index and cluster analysis/tree construction among Cicer accessions were executed using NTSYS-pc V2.1 and POPGENE V1.32, and the methods of Zhang et al.91.

Determination of LD patterns

To access the genome-wide and population-specific LD patterns (r2; average correlation coefficient among pairs of alleles across a pair of SNP loci) individually in desi and kabuli chickpea, the genotyping information of SNPs physically mapped on eight chromosomes were analysed using PLINK and the full-matrix approach of TASSEL, following Zhao et al.59 and Saxena et al.29. The LD decay was estimated by plotting average r2 against 50 kb uniform physical intervals across eight desi and kabuli chromosomes using aforementioned methods.

Evaluation of functional significance of SNPs

To determine the functional significance of SNPs, the gene-derived SNPs (including non-synonymous coding and regulatory SNPs) differentiating accessions belonging to each of the wild Cicer species from accessions of cultivated (desi and kabuli) species were functionally annotated. To determine the differential expression pattern of these SNPs-carrying genes, in silico digital expression profiling was employed. For this, the global transcript profiling data (whole genome transcriptome sequencing) available for selected informative non-synonymous and regulatory SNPs-carrying genes in different vegetative (shoot, root and mature leaf) and reproductive (flower bud and young pod) tissues of chickpea (ICC 4958; Chickpea Transcriptome Database; http://www.nipgr.res.in/ctdb.html) were obtained. The differential expression profiling data assayed by SNPs-carrying genes in diverse vegetative and reproductive tissues of ICC 4958 was compared and a heat map was constructed using the TIGR MultiExperiment Viewer (MeV, http://www.tm4.org/mev.html).

The potential of GBS-based SNPs identified in our study for fine mapping/map-based cloning of QTLs was further assessed. For this, the genetic positions of markers linked/flanking the known QTLs regulating various yield component and stress tolerant traits mapped on eight LGs/chromosomes of diverse chickpea intra- and inter-specific genetic linkage maps were correlated and integrated with that of available high-resolution reference genetic linkage map (ICC 4958 × PI 489777)88 and physical map of kabuli genome6. The SNPs localized in the known trait-regulating QTL intervals based on their physical positions were structurally and functionally annotated. To determine the differential expression pattern of informative SNPs-carrying genes harbouring these known QTLs, the in silico digital expression profiling and heat map construction were performed following aforesaid methods.

Estimation of divergence time among wild and cultivated chickpea

For estimation of divergence time, the synonymous (Ks) and non-synonymous (Ka) substitution rates of the aligned amino acid sequences and their corresponding cDNA sequences of SNPs-carrying genes conserved across each of seven wild and cultivated Cicer species were analyzed using the CODEML program in PAML interface tool of PAL2NAL (http://www.bork.embl.de/pal2nal). The time (million years ago, Mya) of divergence among Cicer species was measured using a synonymous mutation rate of λ substitutions per synonymous site per year as T = Ks/2λ (λ = 6.5 × 10−9)92.

Additional Information

How to cite this article: Bajaj, D. et al. Genome-wide high-throughput SNP discovery and genotyping for understanding natural (functional) allelic diversity and domestication patterns in wild chickpea. Sci. Rep. 5, 12468; doi: 10.1038/srep12468 (2015).

Supplementary Material

Supplementary Information
srep12468-s1.pdf (7.9MB, pdf)

Acknowledgments

The authors gratefully acknowledge the financial support for this study provided by a research grant from the Department of Biotechnology (DBT), Government of India (102/IFD/SAN/2161/2013-14). We thank the Vice Chancellor, CSKHPKV, Palampur, Himachal Pradesh, India for providing off-season nursery facility for phenotyping of Wild Cicer accessions at its Regional Station, Sangla. SD acknowledges the DBT for Junior Research Fellowship award.

Footnotes

Author Contributions D.B. conducted all experiments and drafted the manuscript. S.D., S.B. and V.K. involved in SNP genotyping and allelic diversity data analysis. M.S. and K.C.B. introduced and collected wild Cicer accessions from the global gene bank and hot spots, and also helped in constitution of diversity panel and their multilocation phenotyping. S.K.P. and A.K.T. conceived and designed the study, guided data analysis and interpretation, participated in drafting and correcting the manuscript critically and gave the final approval of the version to be published. All authors have read and approved the final manuscript.

References

  1. Gaur R. et al. High-throughput SNP discovery and genotyping for constructing a saturated linkage map of chickpea (Cicer arietinum L.). DNA Res. 19, 357–373 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Hiremath P. J. et al. Large-scale development of cost-effective SNP marker assays for diversity assessment and genetic mapping in chickpea and comparative mapping in legumes. Plant Biotechnol. J. 10, 716–732 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Thudi M., Li Y., Jackson S. A., May G. D. & Varshney R. K. Current state-of-art of sequencing technologies for plant genomics research. Brief Funct. Genomics 11, 3–11 (2012). [DOI] [PubMed] [Google Scholar]
  4. Roorkiwal M. et al. Single nucleotide polymorphism genotyping for breeding genetics and applications in chickpea and pigeonpea using the BeadXpress platform. Plant Genome 6, 1–10 (2013). [Google Scholar]
  5. Stephens A. et al. Genetic marker discovery, intraspecific linkage map construction and quantitative trait locus analysis of Ascochyta blight resistance in chickpea (Cicer arietinum L.). Mol. Breed. 33, 297–313 (2014). [Google Scholar]
  6. Varshney R. K. et al. Draft genome sequence of chickpea (Cicer arietinum) provides a resource for trait improvement. Nat. Biotechnol. 31, 240–246 (2013). [DOI] [PubMed] [Google Scholar]
  7. Schnable P. S. et al. The B73 maize genome: complexity, diversity and dynamics. Science 326, 1112–1115 (2009). [DOI] [PubMed] [Google Scholar]
  8. Elshire R. J. et al. A robust, simple genotyping-by-sequencing (GBS) approach for high diversity species. PLoS One 6, e19379 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Poland J. A., Brown P. J., Sorrells M. E. & Jannink J. L. Development of high-density genetic maps for barley and wheat using a novel two-enzyme genotyping-by-sequencing approach. PLoS One 7, e32253 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Sonah H. et al. An improved genotyping by sequencing (GBS) approach offering increased versatility and efficiency of SNP discovery and genotyping. PLoS One 8, e54603 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Thurber C. S., Ma J. M., Higgins R. H. & Brown P. J. Retrospective genomic analysis of sorghum adaptation to temperate-zone grain production. Genome Biol. 14, R68 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Morris G. P. et al. Population genomic and genome-wide association studies of agroclimatic traits in sorghum. Proc. Natl. Acad. Sci. USA 110, 453–458 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Spindel J. et al. Bridging the genotyping gap: using genotyping by sequencing (GBS) to add high-density SNP markers and new value to traditional bi-parental mapping and breeding populations. Theor. Appl. Genet. 126, 2699–2716 (2013). [DOI] [PubMed] [Google Scholar]
  14. Byrne S. et al. Genome wide allele frequency fingerprints (GWAFFs) of populations via genotyping by sequencing. PLoS One 8, e57438 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Crossa J. et al. Genomic prediction in maize breeding populations with genotyping-by-sequencing. G3 (Bethesda) 3, 1903–1926 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Mascher M. et al. Barley whole exome capture: a tool for genomic research in the genus Hordeum and beyond. Plant J. 76, 494–505 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Uitdewilligen J. G. et al. A next-generation sequencing method for genotyping-by-sequencing of highly heterozygous autotetraploid potato. PLoS One 8, e62355 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Bastien M., Sonah H. & Belzile F. Genome wide association mapping of Sclerotinia sclerotiorum resistance in soybean with a genotyping-by-sequencing approach. Plant Genome 7, 1–13 (2014). [Google Scholar]
  19. Huang B. E., Raghavan C., Mauleon R., Broman K. W. & Leung H. Efficient imputation of missing markers in low-coverage genotyping-by-sequencing data from multi-parental crosses. Genetics 197, 401–404 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Liu H. et al. An evaluation of genotyping by sequencing (GBS) to map the Breviaristatum-e (ari-e) locus in cultivated barley. BMC Genomics 15, 104 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Tardivel A., Sonah H., Belzile F. & O’Donoughue L. S. Rapid identification of alleles at the soybean maturity gene E3 using genotyping by sequencing and a haplotype-based approach. Plant Genome 7, 1–9 (2014). [Google Scholar]
  22. Jain M. et al. A draft genome sequence of the pulse crop chickpea (Cicer arietinum L.). Plant J. 74, 715–729 (2013). [DOI] [PubMed] [Google Scholar]
  23. Ladizinsky G. & Adler A. The origin of chickpea (Cicer arietinum L.). Euphytica 25, 211–217 (1976a). [Google Scholar]
  24. Abbo S., Berger J. & Turner N. C. Evolution of cultivated chickpea: four bottlenecks limit diversity and constrain adaptation. Funct. Plant Biol. 30, 1081–1087 (2003). [DOI] [PubMed] [Google Scholar]
  25. Berger J. D., Abbo S. & Turner N. C. Ecogeography of Annual Wild Cicer Species. Crop Sci. 43, 1076–1090 (2003). [Google Scholar]
  26. Berger J. D., Buck R., Henzell J. M. & Turner N. C. Evolution in the genus Cicer vernalisation response and low temperature pod set in chickpea (C. arietinum L.) and its annual wild relatives. Aust. J. Agric. Res. 56, 1191–1200 (2005). [Google Scholar]
  27. Toker C. A note on the evolution of kabuli chickpeas as shown by induced mutations in Cicer reticulatum Ladizinsky. Genet. Resour. Crop Evol. 56, 7–12 (2009). [Google Scholar]
  28. Singh R., Sharma P., Varshney R. K. & Sharma S. K. Chickpea improvement: role of wild species and genetic markers. Biotechnol. Genet. Eng. 25, 267–313 (2008). [DOI] [PubMed] [Google Scholar]
  29. Saxena M. S. et al. Natural allelic diversity, genetic structure and linkage disequilibrium pattern in wild chickpea. PLoS One 9, e107484 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Singh K. B., Malhotra R. S. & Saxena M. C. Additional sources of tolerance to cold in cultivated and wild Cicer species. Crop Sci. 35, 1491–1497 (1995). [Google Scholar]
  31. Winter P. et al. A linkage map of the chickpea (Cicer arietinum L.) genome based on recombinant inbred lines from a C. arietinum × C. reticulatum cross: localization of resistance genes for fusarium wilt races 4 and 5. Theor. Appl. Genet. 101, 1155–1163 (2000). [Google Scholar]
  32. Collard B. C., Pang E. C., Ades P. K. & Taylor P. W. Preliminary investigation of QTLs associated with seedling resistance to Ascochyta blight from Cicer echinospermum, a wild relative of chickpea. Theor. Appl. Genet. 107, 719–729 (2003). [DOI] [PubMed] [Google Scholar]
  33. Dwivedi S. L. et al. Using genomics to exploit grain legume biodiversity in crop improvement. Plant Breed. Rev. 26, 171–357 (2005). [Google Scholar]
  34. Shah T. M. et al. Evaluation of Cicer species for resistance to Ascochyta blight. Pak. J. Bot. 37, 431–438 (2005). [Google Scholar]
  35. Upadhyaya H. D. et al. Genomic tools and germplasm diversity for chickpea improvement. Plant Genet. Resour. 9, 45–58 (2011). [Google Scholar]
  36. Sudupak M., Akkaya M. & Kence A. Analysis of genetic relationships among perennial and annual Cicer species growing in Turkey using RAPD markers. Theor. Appl. Genet. 105, 1220–1228 (2002). [DOI] [PubMed] [Google Scholar]
  37. Sudupak M. A., Akkaya M. S. & Kence A. Genetic relationships among perennial and annual Cicer species growing in Turkey assessed by AFLP fingerprinting. Theor. Appl. Genet. 108, 937–944 (2004). [DOI] [PubMed] [Google Scholar]
  38. Nguyen T. T., Taylor P. W. J., Redden R. J. & Ford R. Genetic diversity estimates in Cicer using AFLP analysis. Plant Breed. 123, 173–179 (2004). [Google Scholar]
  39. Cingilli H., Altinkut A. & Akcin A. The use of microsatellite markers in the annual and perennial Cicer species growing in Turkey. Biology 60, 93–98 (2005). [Google Scholar]
  40. Shan F., Clarke H. C., Plummer J. A., Yan G. & Siddique K. H. M. Geographical patterns of the genetic variation in the world collections of wild annual Cicer characterized by amplified fragment length polymorphisms. Theor. Appl. Genet. 110, 381–391 (2005). [DOI] [PubMed] [Google Scholar]
  41. Sethy N. K., Choudhary S., Shokeen B. & Bhatia S. Identification of microsatellite markers from Cicer reticulatum: molecular variation and phylogenetic analysis. Theor. Appl. Genet. 112, 347–357 (2006). [DOI] [PubMed] [Google Scholar]
  42. Upadhyaya H. D. et al. Genetic structure, diversity, and allelic richness in composite collection and reference set in chickpea (Cicer arietinum L.). BMC Plant Biol. 8, 106 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Nayak S. N. et al. Integration of novel SSR gene-based SNP marker loci in the chickpea genetic map and establishment of new anchor points with Medicago truncatula genome. Theor. Appl. Genet. 120, 1415–1441 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Bharadwaj C. et al. Molecular diversity and phylogeny in geographical collection of chickpea (Cicer sp.) accessions. J. Genet. 90, e94–100 (2011). [PubMed] [Google Scholar]
  45. Gujaria N. et al. Development and use of genic molecular markers (GMMs) for construction of a transcript map of chickpea (Cicer arietinum L.). Theor. Appl. Genet. 122, 1577–1589 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Saeed A. et al. Genetic diversity of Iranian accessions, improved lines of chickpea (Cicer arietinum L.) and their wild relatives by using simple sequence repeats. Plant Mol. Biol. Rep. 29, 848–858 (2011). [Google Scholar]
  47. Choudhary P. et al. Genetic structure and diversity analysis of the primary gene pool of chickpea using SSR markers. Genet. Mol. Res. 11, 891–905 (2012). [DOI] [PubMed] [Google Scholar]
  48. Roorkiwal M. et al. Allele diversity for abiotic stress responsive candidate genes in chickpea reference set using gene based SNP markers. Front. Plant Sci. 5, 248 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Ruperao P. et al. A chromosomal genomics approach to assess and validate the desi and kabuli draft chickpea genome assemblies. Plant Biotechnol. J. 12, 778–786 (2014). [DOI] [PubMed] [Google Scholar]
  50. Choudhary S., Sethy N. K., Shokeen B. & Bhatia S. Development of chickpea EST-SSR markers and analysis of allelic variation across related species. Theor. Appl. Genet. 118, 591–608 (2009). [DOI] [PubMed] [Google Scholar]
  51. Singh M. et al. Characterization and evaluation of wild annual Cicer species for agro-morphological traits and major biotic stresses under northwestern Indian conditions. Crop Sci. 54, 229–239 (2014). [Google Scholar]
  52. Choumane W., Winter P., Weigand F. & Kahl G. Conservation and variability of sequence-tagged microsatellite sites from chickpea (Cicer arietinum L.) within the genus Cicer. Theor. Appl. Genet. 101, 269–278 (2000). [Google Scholar]
  53. Chowdhury M. A., Vandenberg V. & Warkentin T. Cultivar identification and genetic relationship among selected breeding lines and cultivars in chickpea (Cicer arietinum L.). Euphytica 127, 317–325 (2002). [Google Scholar]
  54. Iruela M. et al. Phylogenetic analysis in the genus Cicer and cultivated chickpea using RAPD and ISSR markers. Theor. Appl. Genet. 104, 643–651 (2002). [DOI] [PubMed] [Google Scholar]
  55. Rajesh N. et al. Identification of an STMS marker for the double-podding gene in chickpea. Theor. Appl. Genet. 105, 604–607 (2002). [DOI] [PubMed] [Google Scholar]
  56. Kumar J. & Abbo S. Genetics of flowering time in chickpea and its bearing on productivity in semiarid environments. In: Advances in Agronomy. Elsevier, pp. 107–138 (2001). [Google Scholar]
  57. Tanno K. & Willcox G. How fast was wild wheat domesticated? Science 311, 1886 (2006). [DOI] [PubMed] [Google Scholar]
  58. Gore M. A. et al. A first-generation haplotype map of maize. Science 326, 1115–1117 (2009). [DOI] [PubMed] [Google Scholar]
  59. Zhao K. et al. Genome-wide association mapping reveals a rich genetic architecture of complex traits in Oryza sativa. Nat. Commun. 2, 467 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Huang X. et al. Genome-wide association study of flowering time and grain yield traits in a worldwide collection of rice germplasm. Nat. Genet. 44, 32–39 (2012). [DOI] [PubMed] [Google Scholar]
  61. Hyten D. L. et al. Highly variable patterns of linkage disequilibrium in multiple soybean populations. Genetics 175, 1937–1944 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Lam H. M. et al. Resequencing of 31 wild and cultivated soybean genomes identifies patterns of genetic diversity and selection. Nat. Genet. 42, 1053–1059 (2010). [DOI] [PubMed] [Google Scholar]
  63. Xiao Y. et al. Genetic structure and linkage disequilibrium pattern of a rapeseed (Brassica napus L.) association mapping panel revealed by microsatellites. Theor. Appl. Genet. 125, 437–447 (2012). [DOI] [PubMed] [Google Scholar]
  64. Riedelsheimer C. et al. Genome-wide association mapping of leaf metabolic profiles for dissecting complex traits in maize. Proc. Natl. Acad. Sci. USA 109, 8872–8877 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Sakiroglu M. et al. Patterns of linkage disequilibrium and association mapping in diploid alfalfa (M. sativa L.). Theor. Appl. Genet. 125, 577–590 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Atwell S. et al. Genome-wide association study of 107 phenotypes in Arabidopsis thaliana inbred lines. Nature 465, 627–631 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Branca A. et al. Whole-genome nucleotide diversity, recombination, and linkage disequilibrium in the model legume Medicago truncatula. Proc. Natl. Acad. Sci. USA 108, E864–E870 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Haseneyer G. et al. DNA polymorphisms and haplotype patterns of transcription factors involved in barley endosperm development are associated with key agronomic traits. BMC Plant Biol. 10, 5 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Varshney R. K. et al. Genome wide association analyses for drought tolerance related traits in barley (Hordeum vulgare L.). Field Crops Res. 126, 171–180 (2012). [Google Scholar]
  70. Bao J. S., Corke H. & Sun M. Nucleotide diversity in starch synthase IIa and validation of single nucleotide polymorphisms in relation to starch gelatinization temperature and other physicochemical properties in rice (Oryza sativa L.). Theor. Appl. Genet. 113, 1171–1183 (2006). [DOI] [PubMed] [Google Scholar]
  71. Konishi S. et al. An SNP caused loss of seed shattering during rice domestication. Science 312, 1392–1396 (2006). [DOI] [PubMed] [Google Scholar]
  72. Mao H. et al. Linking differential domain functions of the GS3 protein to natural variation of grain size in rice. Proc. Natl. Acad. Sci. USA 107, 19579–19584 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Li S. et al. Natural variation in PTB1 regulates rice seed setting rate by controlling pollen tube growth. Nat. Comm. 4, 2793 (2013). [DOI] [PubMed] [Google Scholar]
  74. Zhang D. et al. The acid phosphatase-encoding gene GmACP1 contributes to soybean tolerance to low-phosphorus stress. PloS Genetics 10, e1004061 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Cobos M. J. et al. Population derived from a kabuli x desi cross. Ann. Appl. Biol. 151, 33–42 (2007). [Google Scholar]
  76. Cobos M. J. et al. Genetic analysis of agronomic traits in a wide cross of chickpea. Field Crops Res. 111, 130–136 (2009). [Google Scholar]
  77. Radhika P. et al. Development of an integrated intraspecific map of chickpea (Cicer arietinum L.) using two recombinant inbred line populations. Theor. Appl. Genet. 115, 209–216 (2007). [DOI] [PubMed] [Google Scholar]
  78. Hossain S., Ford R., McNeil D., Pittock C. & Panozzo J. F. Inheritance of seed size in chickpea (Cicer arietinum L.) and identification of QTL based on 100-seed weight and seed size index. Aust. J. Crop Sci. 4, 126–135 (2010). [Google Scholar]
  79. Gowda S. J. M. et al. Mapping of QTLs governing agronomic and yield traits in chickpea. J. Appl. Genet. 52, 9–21 (2011). [DOI] [PubMed] [Google Scholar]
  80. Kujur A. et al. Functionally relevant microsatellite markers from chickpea transcription factor genes for efficient genotyping applications and trait association mapping. DNA Res. 20, 355–374 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Kujur A. et al. An efficient and cost-effective approach for genic microsatellite marker-based large-scale trait association mapping: identification of candidate genes for seed weight in chickpea. Mol. Breed. 34, 241–265 (2014). [Google Scholar]
  82. Sabbavarapu M. M. et al. Molecular mapping of QTLs for resistance to Fusarium wilt (race 1) and Ascochyta blight in chickpea (Cicer arietinum L.). Euphytica 193, 121–133 (2013). [Google Scholar]
  83. Vadez V. et al. Assessment of ICCV 2 x JG 62 chickpea progenies shows sensitivity of reproduction to salt stress and reveals QTLs for seed yield and yield components. Mol. Breed. 30, 9–21 (2012). [Google Scholar]
  84. Ali L. et al. Mapping and identification of a Cicer arietinum NSP2 gene involved in nodulation pathway. Theor. Appl. Genet. 127, 481–488 (2014). [DOI] [PubMed] [Google Scholar]
  85. Thudi M. et al. Genetic dissection of drought and heat tolerance in chickpea through genome-wide and candidate gene-based association mapping approaches. PloS One 9, e96758 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Varshney R. K. et al. Genetic dissection of drought tolerance in chickpea (Cicer arietinum L.). Theor. Appl. Genet. 127, 445–462 (2014a). [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Varshney R. K. et al. Marker-assisted backcrossing to introgress resistance to Fusarium wilt race 1 and Ascochyta blight in C214, an elite cultivar of chickpea. Plant Genome 7, 1–11 (2014b). [Google Scholar]
  88. Thudi M. et al. Novel SSR markers from BAC-end sequences, DArT arrays and a comprehensive genetic map with 1,291 marker loci for chickpea (Cicer arietinum L.). PLoS One 6, e27275 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Agarwal G. et al. Comparative analysis of kabuli chickpea transcriptome with desi and wild chickpea provides a rich resource for development of functional markers. PloS One 7, e52443 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Jhanwar S. et al. Transcriptome sequencing of wild chickpea as a rich resource for marker development. Plant Biotechnol. J. 10, 690–702 (2012). [DOI] [PubMed] [Google Scholar]
  91. Zhang Y. et al. Genetic diversity assessment of sesame core collection in China by phenotype and molecular markers and extraction of a mini-core collection. BMC Genet. 13, 102 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Lynch M. & Conery J. S. The evolutionary fate and consequences of duplicate genes. Science 290, 1151–1155 (2000). [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary Information
srep12468-s1.pdf (7.9MB, pdf)

Articles from Scientific Reports are provided here courtesy of Nature Publishing Group

RESOURCES