Hostname: page-component-8448b6f56d-tj2md Total loading time: 0 Render date: 2024-04-19T15:36:38.673Z Has data issue: false hasContentIssue false

Genetic diversity and GWAS of agronomic traits using an ICARDA lentil (Lens culinaris Medik.) Reference Plus collection

Published online by Cambridge University Press:  07 July 2021

Karthika Rajendran
Affiliation:
Biodiversity and Crop Improvement Program, International Center for Agricultural Research in the Dry Areas, Rabat, Morocco
Clarice J. Coyne*
Affiliation:
Plant Germplasm Introduction and Testing Research Unit, USDA, ARS, Pullman, WA, USA
Ping Zheng
Affiliation:
Department of Horticulture, Washington State University, Pullman, WA, USA
Gopesh Saha
Affiliation:
Department of Crops and Soils, Washington State University, Pullman, WA, USA
Dorrie Main
Affiliation:
Department of Horticulture, Washington State University, Pullman, WA, USA
Nurul Amin
Affiliation:
Department of Crops and Soils, Washington State University, Pullman, WA, USA
Yu Ma
Affiliation:
Department of Horticulture, Washington State University, Pullman, WA, USA
Ted Kisha
Affiliation:
Plant Germplasm Introduction and Testing Research Unit, USDA, ARS, Pullman, WA, USA
Kirstin E. Bett
Affiliation:
Department of Plant Sciences, University of Saskatchewan, 51 Campus Drive, Saskatoon, SK, S7N 5A8, Canada
Rebecca J. McGee
Affiliation:
Grain Legume Genetics and Physiology Research Unit, USDA, ARS, Pullman, WA, USA
Shiv Kumar
Affiliation:
Biodiversity and Crop Improvement Program, International Center for Agricultural Research in the Dry Areas, Rabat, Morocco
*
*Corresponding author. E-mail: clarice.coyne@usda.gov
Rights & Permissions [Opens in a new window]

Abstract

Genotyping of lentil plant genetic resources holds the promise to increase the identification and utilization of useful genetic diversity for crop improvement. The International Center for Agriculture Research in the Dry Areas (ICARDA) lentil reference set plus collection of 176 accessions was genotyped using genotyping-by-sequencing (GBS) and 22,555 SNPs were identified. The population structure was investigated using Bayesian analysis (STRUCTURE, k = 3) and principal component analysis. The two methods are in concordance. Genome-wide association analysis (GWAS) using the filtered SNP set and ICARDA historical phenotypic data discovered putative markers for several agronomic traits including days to first flower, seeds per pod, seed weight and days to maturity. The genetic and genomic resources developed and utilized in this study are available to the research community interested in exploring the ICARDA reference set plus collection using GWAS.

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
Copyright © The Author(s), 2021. Published by Cambridge University Press on behalf of NIAB

Introduction

Lentil (Lens culinaris Medik.) is an important protein crop. It is a diploid (2n = 2x = 14) and possesses a large genome (~4 Gbp) (Arumuganathan and Earle, Reference Arumuganathan and Earle1991). It ranks fifth, after dry beans, chickpea, dry peas and cowpea, for pulses production in the world (FAOSTAT, 2017). The current global lentil production is estimated at 6.33 million metric tons and falls short of the global demand which is expected to increase soon due to rapid population growth and plant protein market (Reda, Reference Reda2015). In order to bridge the demand-supply gap, efforts are required to accelerate the genetic gain, which is abysmally low mainly due to the narrow genetic base of cultivated lentil. This presents a serious barrier towards developing cultivars for future needs (Lombardi et al., Reference Lombardi, Materne, Cogan, Rodda, Daetwyler, Slater, Forster and Kaur2014; Khazaei et al., Reference Khazaei, Caron, Fedoruk, Diapari, Vandenberg, Coyne, McGee and Bett2016). Integration of genomic tools with conventional breeding approaches would help to alleviate bottlenecks by improving selection efficiency and speed up the breeding process in developing improved cultivars.

Recent efforts on increasing the utilization of plant genetic resources (PGRs) has focused on leveraging genomic tools to unlock the genetic potential of ex situ collections held in national and international gene banks all over the world (McCouch et al., Reference McCouch, McNally, Wang and Hamilton2012, Reference McCouch, Baute, Bradeen, Bramel, Bretting, Buckler, Burke, Charest, Cloutier, Cole, Dempewolf, Dingkuhn, Feuillet, Gepts, Grattapaglia, Guarino, Jackson, Knapp, Langridge, Lawton-Rauh, Lijua, Lusty, Michael, Myles, Naito, Nelson, Pontarollo, Richards, Rieseberg, Ross-Ibarra, Rounsley, Hamilton, Schurr, Stein, Tomooka, van der Knaap, van Tassel, Toll, Valls, Varshney, Ward, Waugh, Wenzl and Zamir2013; Mascher et al., Reference Mascher, Schreiber, Scholz, Graner, Reif and Stein2019). The linkage between genomic characterization and PGR on a global scale can assist with the future challenges to agricultural production such as climate change (Zimmerer and De Haan, Reference Zimmerer and De Haan2017). Even without significant genotypic information for most crops, the USDA germplasm distributions doubled from 2006 to 2012 (Heisey and Rubenstein, Reference Heisey and Rubenstein2015). For plant scientists, and especially plant breeders, access to new positive alleles is paramount for gradual and sustainable genetic gains over the breeding cycles. This requires the utilizing genomic-based tools specifically for genomic-assisted breeding (Varshney et al., Reference Varshney, Singh, Hickey, Xun, Marshall, Wang, Edwards and Ribaut2015), genomic selection (Annicchiarico et al., Reference Annicchiarico, Nazzicari, Pecetti, Romani, Ferrari, Wei and Brummer2017) and breeding-assisted genomics, the recent paradigm switch suggested by Poland (Reference Poland2015).

Recent technological advances are facilitating the expansion of genomic resources for food crops, particularly for pulse crops, in recent years (Varshney, Reference Varshney2016). It is mainly due to the notable reduced costs in sequencing and a surge in bioinformatics tool development (Varshney et al., Reference Varshney, Sinha, Singh, Kumar, Zhang and Bennetzen2020). Many pulse genomes that have been sequenced include pea, lentil, common bean, kabuli chickpea, desi chickpea, cowpea and pigeonpea (Varshney et al., Reference Varshney, Chen, Li, Bharti, Saxena, Schlueter, Donoghue, Azam, Fan, Whaley and Farmer2012, Reference Varshney, Song, Saxena, Azam, Yu, Sharpe, Cannon, Baek, Rosen, Tar'an and Millan2013; Jain et al., Reference Jain, Misra, Patel, Priya, Jhanwar, Khan, Shah, Singh, Garg, Jeena and Yadav2013; Schmutz et al., Reference Schmutz, McClean, Mamidi, Wu, Cannon, Grimwood, Jenkins, Shu, Song, Chavarro, Torres-Torres, Geffroy, Moghaddam, Gao, Abernathy, Barry, Blair, Brick, Chovatia, Gepts, Goodstein, Gonzales, Hellsten, Hyten, Jia, Kelly, Kudrna, Lee, Richard, Miklas, Osorno, Rodrigues, Thareau, Urrea, Wang, Yu, Zhang, Wing, Cregan, Rokhsar and Jackson2014; Ogutcen et al., Reference Ogutcen, Ramsey, von Wettberg and Bett2018; Kreplak et al., Reference Kreplak, Madoui, Cápal, Novák, Labadie, Aubert, Bayer, Kishore, Symes, Main, Klein, Bérard, Fukova, Fournier, d'Agata, Belser, Berrabah, Šimková, Lee, Kougbeadjo, Térézol, Huneau, Turo, Mohellibi, Neumann, Falque, Gallardo-Guerrero, McGee, Tar'an, Bendahmane, Aury, Batley, Le Paslier, Ellis, Warkentin, Coyne, Salse, Edwards, Lichtenzveig, Macas, Doležel, Wincker and Burstin2019; Lonardi et al., Reference Lonardi, Muñoz-Amatriaín, Liang, Shu, Wanamaker, Lo, Tanskanen, Schulman, Zhu, Luo, Alhakami, Ounit, Hasan, Verdier, Roberts, Santos, Ndeve, Doležel, Vrána, Hokin, Farmer, Cannon and Close2019, respectively). Currently, genotyping by sequencing (GBS) is increasingly popular among pulse breeders to screen germplasm quickly and inexpensively (e.g. Guindon et al., Reference Guindon, Martin, Cravero, Gali, Warkentin and Cointry2019; Ma et al., Reference Ma, Marzougui, Coyne, Sankaran, Main, Porter, Mugabe, Smitchger, Zhang, Amin and Rasheed2020). As a high throughput approach, GBS in lentil has facilitated the discovery of genome-wide (SNPs), development of high-density linkage maps and assessment of the genetic diversity in the germplasm collection (Temel et al., Reference Temel, Göl, Akkale, Kahriman and Tanyolac2015; Wong et al., Reference Wong, Verma, Ramsay, Yuan, Caron, Diapari, Vandenberg and Bett2015; Khazaei et al., Reference Khazaei, Fedoruk, Caron, Vandenberg and Bett2017a, Reference Khazaei, Podder, Caron, Kundu, Diapari, Vandenberg and Bettb; Ma et al., Reference Ma, Marzougui, Coyne, Sankaran, Main, Porter, Mugabe, Smitchger, Zhang, Amin and Rasheed2020).

The International Centre for Agricultural Research in the Dry Areas (ICARDA) has a global mandate for the genetic improvement of lentil. The ICARDA lentil reference set (Kumar et al., Reference Kumar, Rajendran, Kumar, Hamwieh and Baum2015), representing the major production and geographical (51 countries) regions, was phenotyped for economically important traits, but was genotyped with only microsatellites. The objectives of this project were to (1) construct a public available lentil SNP genotype set for internationally available lentil PGRs, (2) explore the population structure and diversity, and (3) assess the genotyped collection for possible marker identification (allelic contribution/function) for agronomic traits using genome-wide association study (GWAS) by data mining historical data collected by ICARDA.

Materials and methods

Plant material and field data

In this study, the ICARDA Reference Plus collection of 176 lentil lines (130 Generation Challenge Program (GCP)) reference set (Furman, Reference Furman2006; Kumar et al., Reference Kumar, Rajendran, Kumar, Hamwieh and Baum2015), plus 39 abiotic stress-tolerant lines and seven recombinant inbred lines parents were selected based on phenotypic diversity from the world lentil germplasm collection held by ICARDA (online Supplementary Table S1). The field data presented in online Supplementary Table S2 were historic data collected by ICARDA (e.g. Migicovsky et al., Reference Migicovsky, Gardner, Money, Sawler, Bloom, Moffett, Chao, Schwaninger, Fazio, Zhong and Myles2016; González et al., Reference González, Weise, Zhao, Philipp, Arend, Börner, Oppermann, Graner, Reif and Schulthess2018a). Plant materials were grown using an α-lattice design with two replicates at two ICARDA experiment stations: (1) Tel Hadya, Syria and (2) Terbol, Lebanon, from 2007 to 2011. During the crop growing period, all crop management practices typical for the area were followed. Lines were phenotyped for days to first flower (number of days from sowing to the appearance of the first flower); plant height (average height of five plants from the ground to the tip of the foliage at maturity); seeds per pod (average number of seeds in 10 randomly chosen dry pods); days to maturity (number of days from sowing until 90% of the pods were golden brown); biomass yield of each plot (weight of dried mature plants in a plot); seed yield (seed yield of a plot after threshing, expressed as kg/ha); straw yield (calculated as the difference between biomass yield and seed yield); harvest index (calculated as the ratio of seed to biomass yield); and hundred-seed weight (average weight of two samples of 100 randomly chosen seeds in g). Phenotypic values were also combined across years and averaged in cases of replication for a particular accession (online Supplementary Table S2; Supplementary Fig. S1).

Genotyping

DNA was extracted, using a DNeasy Plant Kit (QIAGEN, Valencia, CA), from a single plant per accession grown in the greenhouse at the USDA-ARS Western Regional Plant Introduction Station in Pullman, WA in 2013. DNA was quantified using a spectrophotometer (Nano-Drop Technologies, Wilmington, DE, USA).

The two-enzyme genotyping-by-sequencing procedure of Poland et al. (Reference Poland, Brown, Sorrells and Jannink2012) was followed using the modifications of Wong et al. (Reference Wong, Verma, Ramsay, Yuan, Caron, Diapari, Vandenberg and Bett2015). Briefly, 200 ng of genomic DNA was double-digested with PstI and MspI and ligated to two adapters, of which one contained a barcode sequence. Samples were pooled, PCR amplified and cleaned up using a column (Qiagen QIAquick PCR Purification Kit). Four libraries of 48 bar-coded samples were sequenced in four lanes using an Illumina HiSeq2000 by the Vincent J. Coates Genomics Sequencing Laboratory at the University of California, Berkeley.

Genotypic data analysis

The sequencing data were processed to remove low-quality data and analysed using ‘Stacks’ software (Catchen et al., Reference Catchen, Hohenlohe, Bassham, Amores and Cresko2013). Unfiltered Fastq sequence Illumina data were assigned to individual samples via the barcode sequence using ‘Stacks’ software (Catchen et al., Reference Catchen, Amores, Hohenlohe, Cresko and Postlethwait2011, Reference Catchen, Hohenlohe, Bassham, Amores and Cresko2013). The RAD-Tags algorithm was used to examine raw reads from Illumina sequencing runs by checking that the barcode and the RAD cut-site are intact, and at the same time de-multiplex the data. The default parameters of the ‘Stack’ software were used to call the SNPs. ‘Stacks’ uses short-read sequence data to identify and genotype loci in a set of individuals and in-house scripts generated a VCF format genotype data file.

A second analysis was conducted using FreeBayes software (Garrison and Marth, Reference Garrison and Marth2012) to call the variants using the lentil reference genome, Lc1.2 of ‘CDC Redberry’ (Ramsay et al., Reference Ramsay, Chan, Sharpe, Cook, Penmetsa, Chang, Coyne, McGee, Main, Edwards, Kaur, Vandenberg and Bett2016; Bett, Reference Bett2020). Analyses of pipeline details are presented in Yu et al. (Reference Yu, Zheng, Bhamidimarri, Liu and Main2017). Briefly, the SNPs were identified with the lentil reference genome version Lc1.2 using BamTools (Barnett et al., Reference Barnett, Garrison, Quinlan, Strömberg and Marth2011) and the FreeBayes variant caller (https://github.com/ekg/freebayes) (Garrison, Reference Garrison2012; Garrison and Marth, Reference Garrison and Marth2012). The 50% missing data were used as a minimum to keep the variant in the final filtered VCF file to give future users of the SNP data flexibility on filtering without resorting to reanalysing the raw data (Glaubitz et al., Reference Glaubitz, Casstevens, Lu, Harriman, Elshire, Sun and Buckler2014; O'Leary et al., Reference O'Leary, Puritz, Willis, Hollenbeck and Portnoy2018).

Structure and diversity

The genetic structure of the collection was determined using the software STRUCTURE (Pritchard et al., Reference Pritchard, Stephens and Donnelly2000). PGDSpider software was used to convert the *.VCF file to the STRUCTURE software format (Lischer and Excoffier, Reference Lischer and Excoffier2012). Parameter set included a 10,000 burn-in period, 10,000 Markov Chain Monte Carlo (MCMC) replications, admixture model, run for each subpopulation (K) value ranging from 1 to 7. The best K value was determined by plotting the rate of change in the log probability of data (ΔK) against the successive K values (Evanno et al., Reference Evanno, Regnaut and Goudet2005) implemented in STRUCTURE HARVESTER (Earl and von Holdt, Reference Earl and von Holdt2012). The K value was considered to be optimum while ΔK reaches the maximum. A tree was constructed in NTSys-pc using Prevosti's Distance substituting the probability of assignment to each population at K = 3 for allele frequency (Rohlf Reference Rohlf2009). The distance matrices were used to produce a dendrogram based on clustering using the unweighted pair-group method with arithmetic averages (UMGMA) in the SAHN module of NTSYS-PC program version 2.02 k (Rohlf, Reference Rohlf2009). A tree view was created using the distance matrix and UPGMA (Sneath and Sokal, Reference Sneath and Sokal1973) clustering method modules in STRUCTURE. Genetic diversity between subpopulations as determined by the STRUCTURE software was calculated using Analysis of Molecular Variance (AMOVA) which calculates PhiPT (Excoffier et al., Reference Excoffier, Smouse and Quattro1992). Phi-statistics is a modified version of Wright's F that refers to the relative contributions of between-subpopulation separation to the overall genetic variation in the whole sample. The variance components are used to calculate phi-statistics which are analogous to Wright's F-statistics, ΦST = (σ 2a + σ 2b)/σ 2T (Schneider et al., Reference Schneider, Roessli and Excoffier2000). AMOVA was calculated using the ‘Distance’ and ‘AMOVA’ functions in GenAlEx 6.5 (Peakall and Smouse, Reference Peakall and Smouse2006, Reference Peakall and Smouse2012). Principal component analysis (PCA) was conducted using the ‘PCA’ module in TASSEL using the SNP data and graphed using SigmaPlot Version 13.0 (Systat Software, San Jose, CA, USA) (Bradbury et al., Reference Bradbury, Zhang, Kroon, Casstevens, Ramdoss and Buckler2007).

GWAS

The GWAS was conducted using phenotypic data means collected from 2007 to 2011 (online Supplementary Table S2) held in the lentil germplasm database of ICARDA using the GLM-PCA batch commands in the software TASSEL 5.2.29 (Bradbury et al., Reference Bradbury, Zhang, Kroon, Casstevens, Ramdoss and Buckler2007). SNPs were filtered to a maximum of 15% of the lines missing the SNP call with a minimum allele frequency of 0.05 (Glaubitz et al., Reference Glaubitz, Casstevens, Lu, Harriman, Elshire, Sun and Buckler2014; O'Leary et al., Reference O'Leary, Puritz, Willis, Hollenbeck and Portnoy2018). Marker-trait associations (MTAs) were analysed for hundred-seed weight, days to flower, plant height, days to maturity, seeds per pod, biomass yield, seed yield, and harvest index using a generalized linear model and a population stratification (structure) correction based on principal component (3) analysis (PCA) (Price et al., Reference Price, Patterson, Plenge, Weinblatt, Shadick and Reich2006). The significance of associations between SNPs and traits was based on the threshold P < 1.57 × 10−4, a modified Bonferroni correction calculated by dividing 1 by the total number of SNPs (6373) in the analysis (Li et al., Reference Li, Chen, Xu, Gao, Yan, Qiao, Li, Li, Li, Xiao and Zhang2016).

Results

The non-reference (de novo) pipeline identified 11,225 SNPs in the 176 accessions originating from 51 countries. The SNP dataset from the de novo analysis was further filtered allowing for 15% missing data which left 1021 SNPs, i.e. SNP calls were available in 85% of the accessions. Reanalysing the variants using FreeBayes and the lentil reference genome version Lc1.2 (Ramsay et al., Reference Ramsay, Chan, Sharpe, Cook, Penmetsa, Chang, Coyne, McGee, Main, Edwards, Kaur, Vandenberg and Bett2016; Ogutcen et al., Reference Ogutcen, Ramsey, von Wettberg and Bett2018) increased the SNPs identified to 22,555. These SNPs were filtered allowing for the same 15% missing data and increased the SNPs sixfold to 6373 versus 1021 de novo called variants. Finally, allowing for no missing data, the number of SNPs was 4195 using the reference-based FreeBayes analysis versus zero SNPs for the de novo Stack pipeline.

The collection of 176 accessions was analysed for subpopulation structure using two methods: a Bayesian clustering method (online Supplementary Table S3) and PCA based on the SNP genotypes. In the Bayesian approach, first proposed by Pritchard et al. (Reference Pritchard, Stephens and Donnelly2000), the Evanno method used to determine the number of subpopulations supports k = 3 with far lower support for k = 5 (Evanno et al., Reference Evanno, Regnaut and Goudet2005). Both Bayesian and PCA methods indicated three subpopulations (Figs. 1 and 2). Further, the UPGMA tree is also in agreement with the STRUCTURE software clustering subpopulations (Fig. 3; online Supplementary Figs. S1, S2; Supplementary Table S3). One cluster of the dendrogram contained most of the admixture accessions (Fig. 1). Genetic diversity among and between subpopulations was calculated using AMOVA (Excoffier et al., Reference Excoffier, Smouse and Quattro1992). The partition of total genetic diversity within the three (k = 3) subpopulations was 74% and among the three subpopulations was 26%. ΦPT value was 0.256, P > 0.001.

Fig. 1. The subpopulations of K = 3 as determined by the ad hoc statistic ΔK based on the rate of change in the log probability of data between successive K = 1–7 (Evanno et al., Reference Evanno, Regnaut and Goudet2005).

Fig. 2. Dendrogram based on UPGMA and the subpopulations (K = 3) calculated using the Bayesian clustering method of the software STRUCTURE based on SNP data for 176 lentil accessions (Pritchard et al., Reference Pritchard, Stephens and Donnelly2000).

Fig. 3. Principal component analysis of 176 lentil accessions at K = 3 based on SNP genotyping. Colours correspond to greater than 50% association with a subpopulation and colours correspond to Fig. 3. White accessions are admixtures. Further views available in online Supplementary Material (Fig. S2).

Frequency histograms of the phenotypic data are presented in online Supplementary Fig. S1. The MTAs were identified across five chromosomes for four traits: days to first flower, days to maturity, seeds per pod, and 100-seed weight (Table 1; online Supplementary Fig. S3). The range of variance explained by the MTAs for the four traits ranged from R 2 = 0.10 to 0.17. The associations are moderate (Table 1). For days to flower, two MTAs were found on chromosome 3. The most MTAs were found for days to maturity, eight MTAs on chromosomes 2, 3, 5, 6 and 7. Two MTAs were found for seeds per pod on chromosomes 2 and 7. For 100-seed weight, one MTA was identified on chromosome 2.

Table 1. Significant SNP markers identified using genome-wide associations for four traits based on ICARDA's historical phenotypic data collected from 2007 to 2011

Discussion

Genotyping of PGRs has spurred discoveries of the genetic control of important agronomic traits in several crops. For example, SNP genotyping of the entire USDA + 18 K soybean collection has been used in 124 research studies (Song et al., Reference Song, Hyten, Jia, Quigley, Fickus, Nelson and Cregan2015). Genotyping-by-sequencing has been a useful approach for genotyping and analysis of lentil PGRs. Wong et al. (Reference Wong, Verma, Ramsay, Yuan, Caron, Diapari, Vandenberg and Bett2015) were the first to report the use of the two-enzyme GBS approach of Poland et al. (Reference Poland, Brown, Sorrells and Jannink2012) in lentil and identified 266,356 genome-wide SNPs before filtering. After filtering to 20% missing SNP data, the final dataset was 32,019 SNPs. In comparison, filtering the ICARDA Lentil Reference Plus SNP set to 20% missing data in the present study resulted in fewer SNPs but still a useful set of 11,171. Recently, a single enzyme GBS of lentil resulted in the discovery of 6693 SNPs (Pavan et al., Reference Pavan, Bardaro, Fanelli, Marcotrigiano, Mangini, Taranto, Catalano, Montemurro, De Giovanni, Lotti and Ricciardi2019). While fair genome coverage was gained through these GBS experimental designs, a lentil exome capture resource has now been developed (Ogutcen et al., Reference Ogutcen, Ramsey, von Wettberg and Bett2018). This approach will enhance the coverage of the transcribed genes and perhaps improve the success in identifying causal genes and alleles. However, the cost is significantly higher for the lentil exome capture sequencing approach compared to GBS. Therefore, GBS is still a good option for SNP discovery and diversity studies in lentil.

The ICARDA Lentil Reference Plus collection subpopulation (K = 3) result is similar to a larger lentil genetic diversity study of 352 accessions by Khazaei et al. (Reference Khazaei, Caron, Fedoruk, Diapari, Vandenberg, Coyne, McGee and Bett2016). However, the populations they reported (K = 3) separated into three geographic regions: South Asia, Mediterranean and northern temperate. In contrast, the ICARDA Reference Plus collection's subpopulations did not stratify geographically. One aspect to consider is the ICARDA Reference Plus collection includes 20% breeding lines involving parents from different geography and elite lines thus listed as unknown origin in the ICARDA database. Lombardi et al. (Reference Lombardi, Materne, Cogan, Rodda, Daetwyler, Slater, Forster and Kaur2014) also reported three subpopulations studying 505 cultivars and landraces but found weaker geographic clustering outside of a breeding program cluster. Idrissi et al. (Reference Idrissi, Piergiovanni, Toklu, Houasli, Udupa, De Keyser, Van Damme and De Riek2018) reported subpopulations of two studying 74 Mediterranean lentil landraces stratifying geographically detecting a northern gene pool composed of Turkish, Italian and Greek landraces, and a southern gene pool composed of Moroccan landraces. A larger study of the genetic diversity of the Mediterranean ex situ lentil collection (n = 349) held at the Italian National Research Council also partitioned into two subpopulations with lower support for three and five subpopulations using the Evanno method (Pavan et al., Reference Pavan, Bardaro, Fanelli, Marcotrigiano, Mangini, Taranto, Catalano, Montemurro, De Giovanni, Lotti and Ricciardi2019).

Two other diversity studies have been published on the ICARDA lentil collection using genetic markers. The work of Hamwieh et al. (Reference Hamwieh, Udupa, Sarker, Jung and Baum2009) detected two clusters in the ICARDA core collection using 26 microsatellites and UPGMA and PCA analyses. Their collection of 109 accessions (52% wild Lens spp.) overlaps our set by only five accessions. An earlier study of an ICARDA set (308 accessions) that included the Lens wild relatives (175 accessions) used 22 expressed sequence tags (ESTs) and found eight subpopulations (K = 8) using the same Evanno method to select the best fitting model (Alo et al., Reference Alo, Furman, Akhunov, Dvorak and Gepts2011). Lens culinaris ssp. culinaris accessions clustered into two subpopulations, the other six subpopulations were single taxon wild accessions. The Alo et al. (Reference Alo, Furman, Akhunov, Dvorak and Gepts2011) ICARDA set overlaps our study by nine accessions.

For GWAS, precise phenotypic quantitative trait values are required. The data available for our test were the historical data from ICARDA collected over 5 years, not an experiment per se. Further, quantitative trait data, particularly days to flower, days to maturity and seed weight are known to have a genotype by environment interaction (Abbo et al., Reference Abbo, Ladizinsky and Weeden1992; Singh et al., Reference Singh, Singh, Gil, Kumar and Sarker2009; Kahriman et al., Reference Kahriman, Temel, Aydogan and Tanyolac2015). Also, our GWAS experiment might have been affected by the use of phenotypic means from different years. Nonetheless, interesting single-SNP defined regions were identified for four traits: days to first flower, days to maturity, seeds per pod and seed weight. These results will be useful for future meta-analyses as more lentil GWAS studies are published for agronomic traits. Days to first flower was highly significantly correlated with days to maturity and the chromosome 3 MTAs for these two traits are in the same region. One hundred seed weight was significantly negatively correlated with seeds per pod and had no correlation to days to maturity. Comparison with earlier QTL studies is limited by the lack previously of a lentil consensus linkage map so numbering of linkage groups was not consistent. An early genetic study using isozymes, RAPDs and RFLPs, found four linkage groups with factors controlling seed weight in three wide crosses with contrasting seed sizes (L. culinaris × L. orientalis) (Abbo et al., Reference Abbo, Ladizinsky and Weeden1992). A major QTL for seed weight was identified using SSRs explaining 48.4% of the variance (Verma et al., Reference Verma, Goyal, Chahota, Sharma, Abdin and Bhatia2015). Three QTL for seed weight were reported on two linkage groups with one QTL explaining 34–50% of the variance (Jha et al., Reference Jha, Bohra, Jha, Rana, Chahota, Kumar and Sharma2017). Our GWAS MTA on chromosome 2 explains less of the variance (12.9%) than these other studies. However, another lentil GWAS study recently found one significant MTA lentil seed weight using SSRs which explained a similar amount of the variance (Singh et al., Reference Singh, Dikshit, Mishra, Aski and Kumar2019).

Days to flowering has been an important selection criterion for lentil breeders (Erskine et al., Reference Erskine, Ellis, Summerfield, Roberts and Hussain1990). Sarker et al. (Reference Sarker, Erskine, Sharma and Tyagi1999) originally reported and named a lentil early flowering gene (sn). Weller et al. (Reference Weller, Liew, Hecht, Rajandran, Laurie, Ridge, Wenden, Vander Schoor, Jaminon, Blassiau and Dalmais2012) establish the importance of the Hr locus (orthologue of Early Flowering 3) on photoperiod response of flowering in lentil located on chromosome 3 (Bett, unpublished data). The one significant MTA for days to flowering identified in the ICARDA Reference Plus collection was located also on chromosome 3 in our study. Once the lentil reference genome is published, it can be determined if this MTA maps close to the Hr locus. A bi-parental mapping population reported a major flowering time QTL explained 60% of the variance (Kahriman et al., Reference Kahriman, Temel, Aydogan and Tanyolac2015). Three other QTL for days to flower have been reported on a bi-parental SNP-based map (Fedoruk et al., Reference Fedoruk, Vandenberg and Bett2013). A single major locus for days to flower was recently reported in a wide cross between L. culinaris × L. odemensis (Polanco et al., Reference Polanco, de Miera, González, García, Fratini, Vaquero, Vences and de la Vega2019). A lentil GWAS study on flowering time in 324 L. culinaris lines using 255,714 SNP markers identified three MTAs (two on chromosome 2, one on chromosome 5) using the CDC Redberry Lc1.2 assembly (Ramsay et al., Reference Ramsay, Chan, Sharpe, Cook, Penmetsa, Chang, Coyne, McGee, Main, Edwards, Kaur, Vandenberg and Bett2016; Neupane, Reference Neupane2019). Flowering time MTAs were also reported for days to flower and days to maturity using GWAS with unmapped SSR and EST markers by Kumar et al. (Reference Kumar, Gupta, Biradar, Gupta, Dubey and Singh2018a, Reference Kumar, Gupta, Gupta and Singhb).

The data mining of historical genebank phenotypic data for GWAS is relatively new and mostly untested. Nguyen and Norton (Reference Nguyen and Norton2020) recently reviewed this approach for GWAS and genomic selection. Two examples of this wealth of data on barley and wheat were published by Leibniz Institute of Plant Genetics and Crop Plant Research (IPK) (González et al., Reference González, Weise, Zhao, Philipp, Arend, Börner, Oppermann, Graner, Reif and Schulthess2018a; Philipp et al., Reference Philipp, Weise, Oppermann, Börner, Graner, Keilwagen, Kilian, Zhao, Reif and Schulthess2018). González et al. (Reference González, Philipp, Schulthess, Weise, Zhao, Börner, Oppermann, Graner and Reif2018b) published a strategy to utilize historical phenotypic data collected during seed regeneration to assemble large mapping populations of accessions to discover the genetic effects. Their proposed strategy is not crop-specific and can be used as a guide for the phenotypic evaluation of basically any collection with quality phenotypic data. Utilizing phenotypic historic data from ex situ genebanks was thought to ‘elevate them to bio-digital resource centres’. A successful application of the historical data approach was used for a GWAS study confirming the association between fruit colour and the MYB1 locus in apple (Migicovsky et al., Reference Migicovsky, Gardner, Money, Sawler, Bloom, Moffett, Chao, Schwaninger, Fazio, Zhong and Myles2016).

Incorporating this genotype data of the ICARDA Reference Plus collection into genebank databases will bring the world's plant science research community closer to ‘unlocking’ genetic diversity within these collections (Tanksley and McCouch, Reference Tanksley and McCouch1997). Linking the genotypic data to ex situ PGR accessions has been limited based on current genebank database software schema (Postman et al., Reference Postman, Hummer, Bretting, Kinard, Bohning, Emberland, Sinnott, Mackay, Cyr, Millard, Gardner, Weaver, Ayala-Silva, Franko, Mackay and Guarino2010; van Treuren and van Hintum, Reference van Treuren and van Hintum2014). Finkers et al. (Reference Finkers, Chibon, van Treuren, Visser and Hintum2015) proposed using semantic web technology. A USDA effort (www.breedinginsight.org) was undertaken to link genomic data directly to GRIN Global databases. The most advanced effort to link genotypes to PGR accessions has been developed, the database ‘Germinate’ (Raubach et al., Reference Raubach, Kilian, Dreher, Amri, Bassi, Boukar, Cook, Cruickshank, Fatokun, El Haddad, Humphries, Jordan, Kehel, Kumar, Labarosa, Loi, Mace, McCouch, McNally, Marshall, Mikwa, Milne, Odeny, Plazas, Prohens, Rieseberg, Schafleitner, Sharma, Stephen, Tin, Abou Togola, Emily Warchefsky, Peter Werner and Shaw2021). This version integrates both the phenotypic and genotypic data with the PGR accession. Currently, this lentil accession genotypic data is available for download from the PulseDB database and in the future will be linked to the ICARDA genebank database. Seed of the ICARDA Reference Plus collection is available for requestors directly from ICARDA (https://www.icarda.org/).

Data availability

The lentil SNP data set in vcf format file as well as corresponding FASTA sequences are available for downloading on the Pulse Crops Database (https://www.pulsedb.org/). All raw sequence data are available through the National Center for Biotechnology Information (NCBI) Sequence Read Archive (SRA) with BioProject number: PRJNA639210 (http://www.ncbi.nlm.nih.gov/bioproject/639210).

Supplementary material

The supplementary material for this article can be found at https://doi.org/10.1017/S147926212100006X

Acknowledgements

The authors thank the USAID for a linkage grant (KR, SK, RJM, CJC), CRP-Grain Legumes (KR, SK) and the Northern Pulse Growers Association (DM, RJM, CJC) and for funding and support from USDA ARS Project Nos. 5348-21000-017-00D (CJC), #5348-21000-024-00D (RJM). The authors further thank the ‘Lentil Genome Sequencing (LenGen) Project’ and its Project Leaders (KE Bett and DR Cook), and the researcher(s) responsible for generating the data. This work used the Vincent J. Coates Genomics Sequencing Laboratory at UC Berkeley, supported by NIH S10 Instrumentation Grants S10RR029668 and S10RR027303.

Footnotes

Present address: VIT School of Agricultural Innovations and Advanced Learning (VAIAL), Vellore Institute of Technology, Vellore 632014, Tamil Nadu, India.

Present address: Brotherton Seed Co., 451 S Milwaukee Ave, Moses Lake, WA, USA.

§

Present address: Department of Crop and Soil Sciences, Breeder Seed Production Center, Bangladesh Agricultural Research Institute, Debiganj, Panchagarh, Bangladesh.

References

Abbo, S, Ladizinsky, G and Weeden, NF (1992) Genetic analysis and linkage study of seed weight in lentil. Euphytica 58: 259266.CrossRefGoogle Scholar
Alo, F, Furman, BJ, Akhunov, E, Dvorak, J and Gepts, P (2011) Leveraging genomic resources of model species for the assessment of diversity and phylogeny in wild and domesticated lentil. Journal of Heredity 102: 315329.CrossRefGoogle ScholarPubMed
Annicchiarico, P, Nazzicari, N, Pecetti, L, Romani, M, Ferrari, B, Wei, Y and Brummer, EC (2017) GBS-based genomic selection for pea grain yield under severe terminal drought. The Plant Genome 10: 2.CrossRefGoogle ScholarPubMed
Arumuganathan, K and Earle, ED (1991) Nuclear DNA content of some important plant species. Plant Molecular Biology 9: 208218.CrossRefGoogle Scholar
Barnett, DW, Garrison, EK, Quinlan, AR, Strömberg, MP and Marth, GT (2011) Bamtools: a C++ API and toolkit for analyzing and managing BAM files. Bioinformatics (Oxford, England) 27: 16911692.CrossRefGoogle Scholar
Bett, KE (2020) Lentil Genome Sequencing (LenGen) Project. Available at https://knowpulse.usask.ca/study/2653517 (accessed 17 June 2020).Google Scholar
Bradbury, PJ, Zhang, Z, Kroon, DE, Casstevens, TM, Ramdoss, Y and Buckler, ES (2007) TASSEL: software for association mapping of complex traits in diverse samples. Bioinformatics (Oxford, England) 23: 26332635.CrossRefGoogle ScholarPubMed
Catchen, JM, Amores, A, Hohenlohe, P, Cresko, W and Postlethwait, JH (2011) Stacks: building and genotyping loci de novo from short-read sequences. G3: Genes, Genomes, Genetics 1: 171182.CrossRefGoogle ScholarPubMed
Catchen, J, Hohenlohe, PA, Bassham, S, Amores, A and Cresko, WA (2013) Stacks: an analysis tool set for population genomics. Molecular Ecology 22: 31243140.CrossRefGoogle ScholarPubMed
Earl, DA and von Holdt, BM (2012) STRUCTURE HARVESTER: a website and program for visualizing STRUCTURE output and implementing the Evanno method. Conservation Genetics Resources 4: 359361.CrossRefGoogle Scholar
Erskine, W, Ellis, RH, Summerfield, RJ, Roberts, EH and Hussain, A (1990) Characterization of responses to temperature and photoperiod for time to flowering in a world lentil collection. Theoretical and Applied Genetics 80: 193199.CrossRefGoogle Scholar
Evanno, G, Regnaut, S and Goudet, J (2005) Detecting the number of clusters of individuals using the software STRUCTURE: a simulation study. Molecular Ecology 14: 26112620.CrossRefGoogle ScholarPubMed
Excoffier, L, Smouse, PE and Quattro, JM (1992) Analysis of molecular variance inferred from metric distances among DNA haplotypes: application to human mitochondrial DNA restriction data. Genetics 131: 479491.CrossRefGoogle ScholarPubMed
FAOSTAT (2017) Available at: http://www.fao.org/faostat/en/#data/QC/visualize (accessed 20 September 2017).Google Scholar
Fedoruk, MJ, Vandenberg, A and Bett, KE (2013) Quantitative trait loci analysis of seed quality characteristics in lentil using single nucleotide polymorphism markers. The Plant Genome 6. doi: 10.3835/plantgenome2013.05.0012.CrossRefGoogle Scholar
Finkers, R, Chibon, PY, van Treuren, R, Visser, R and Hintum, TV (2015) Genebanks and genomics: how to interconnect data from both communities? Plant Genetic Resources 13: 9093.CrossRefGoogle Scholar
Furman, BJ (2006) Methodology to establish a composite collection: case study in lentil. Plant Genetic Resources 4: 212.CrossRefGoogle Scholar
Garrison, E (2012) FreeBayes source repository. Available at https://github.com/ekg/freebayes.Google Scholar
Garrison, E and Marth, G (2012) Haplotype-based variant detection from short-read sequencing. arXiv preprint arXiv: 1207.3907.Google Scholar
Glaubitz, JC, Casstevens, TM, Lu, F, Harriman, J, Elshire, RJ, Sun, Q and Buckler, ES (2014) TASSEL-GBS: a high capacity genotyping by sequencing analysis pipeline. PLoS ONE 9: e90346.CrossRefGoogle ScholarPubMed
González, MY, Weise, S, Zhao, Y, Philipp, N, Arend, D, Börner, A, Oppermann, M, Graner, A, Reif, JC and Schulthess, AW (2018a) Unbalanced historical phenotypic data from seed regeneration of a barley ex situ collection. Scientific Data 5: 110.CrossRefGoogle Scholar
González, MY, Philipp, N, Schulthess, AW, Weise, S, Zhao, Y, Börner, A, Oppermann, M, Graner, A and Reif, JC (2018b) Unlocking historical phenotypic data from an ex situ collection to enhance the informed utilization of genetic resources of barley (Hordeum sp.). Theoretical and Applied Genetics 131: 20092019.CrossRefGoogle Scholar
Guindon, MF, Martin, E, Cravero, V, Gali, KK, Warkentin, TD and Cointry, E (2019) Linkage map development by GBS, SSR, and SRAP techniques and yield-related QTLs in pea. Molecular Breeding 39: 54.CrossRefGoogle Scholar
Hamwieh, A, Udupa, SM, Sarker, A, Jung, C and Baum, M (2009) Development of new microsatellite markers and their application in the analysis of genetic diversity in lentils. Breed Science 59: 7786.CrossRefGoogle Scholar
Heisey, P and Rubenstein, DR (2015) Using crop genetic resources to help agriculture adapt to climate change: economics and policy. USDA-ERS Economic Information Bulletin 139: 123.Google Scholar
Idrissi, O, Piergiovanni, A, Toklu, F, Houasli, C, Udupa, S, De Keyser, E, Van Damme, P and De Riek, J (2018) Molecular variance and population structure of lentil (Lens culinaris Medik.) landraces from Mediterranean countries as revealed by simple sequence repeat DNA markers: implications for conservation and use. Plant Genetic Resources 16: 249259.CrossRefGoogle Scholar
Jain, M, Misra, G, Patel, RK, Priya, P, Jhanwar, S, Khan, AW, Shah, N, Singh, VK, Garg, R, Jeena, G and Yadav, M (2013) A draft genome sequence of the pulse crop chickpea (Cicer arietinum L.). The Plant Journal 74: 715729.CrossRefGoogle Scholar
Jha, R, Bohra, A, Jha, UC, Rana, M, Chahota, RK, Kumar, S and Sharma, TR (2017) Analysis of an intraspecific RIL population uncovers genomic segments harbouring multiple QTL for seed relevant traits in lentil (Lens culinaris L.). Physiology and Molecular Biology of Plants 23: 675684.CrossRefGoogle Scholar
Kahriman, A, Temel, HY, Aydogan, A and Tanyolac, MB (2015) Major quantitative trait loci for flowering time in lentil. Turkish Journal of Agriculture and Forestry 39: 588595.CrossRefGoogle Scholar
Khazaei, H, Caron, CT, Fedoruk, M, Diapari, M, Vandenberg, A, Coyne, CJ, McGee, R and Bett, KE (2016) Genetic diversity of cultivated lentil (Lens culinaris Medik.) and its relation to the world's agro-ecological zones. Frontiers in Plant Science 7: 1093.CrossRefGoogle ScholarPubMed
Khazaei, H, Fedoruk, M, Caron, CT, Vandenberg, A and Bett, KE, (2017a) Single nucleotide polymorphism markers associated with seed quality characteristics of cultivated lentil. The Plant Genome 11: 170051.CrossRefGoogle Scholar
Khazaei, H, Podder, R, Caron, CT, Kundu, SS, Diapari, M, Vandenberg, A and Bett, KE (2017b) Marker-trait association analysis of iron and zinc concentration in lentil (Lens culinaris Medik.) seeds. The Plant Genome 10. doi: 10.3835/plantgenome2017.02.0007.CrossRefGoogle Scholar
Kreplak, J, Madoui, M-A, Cápal, P, Novák, P, Labadie, K, Aubert, G, Bayer, P, Kishore, KG, Symes, RA, Main, D, Klein, A, Bérard, A, Fukova, I, Fournier, C, d'Agata, L, Belser, C, Berrabah, W, Šimková, H, Lee, HT, Kougbeadjo, A, Térézol, M, Huneau, C, Turo, CJ, Mohellibi, N, Neumann, P, Falque, M, Gallardo-Guerrero, K, McGee, R, Tar'an, B, Bendahmane, A, Aury, J-M, Batley, J, Le Paslier, MC, Ellis, THN, Warkentin, T, Coyne, CJ, Salse, J, Edwards, D, Lichtenzveig, J, Macas, J, Doležel, J, Wincker, P and Burstin, J (2019) A reference genome for pea provides insight into legume evolution. Nature Genetics 51: 14111422.CrossRefGoogle Scholar
Kumar, S, Rajendran, K, Kumar, J, Hamwieh, A and Baum, M (2015) Current knowledge in lentil genomics and its application for crop improvement. Frontiers in Plant Science 6: 78.CrossRefGoogle ScholarPubMed
Kumar, J, Gupta, S, Biradar, RS, Gupta, P, Dubey, S and Singh, NP (2018a) Association of functional markers with flowering time in lentil. Journal of Applied Genetics 59: 921.CrossRefGoogle Scholar
Kumar, J, Gupta, S, Gupta, DS and Singh, NP (2018b) Identification of QTLs for agronomic traits using association mapping in lentil. Euphytica 214: 75.CrossRefGoogle Scholar
Li, F, Chen, B, Xu, K, Gao, G, Yan, G, Qiao, J, Li, J, Li, H, Li, L, Xiao, X and Zhang, T (2016) A genome-wide association study of plant height and primary branch number in rapeseed (Brassica napus). Plant Science 242: 169177.CrossRefGoogle Scholar
Lischer, HEL and Excoffier, L (2012) PGDSpider: an automated data conversion tool for connecting population genetics and genomics programs. Bioinformatics (Oxford, England) 28: 298299.CrossRefGoogle ScholarPubMed
Lombardi, M, Materne, M, Cogan, NOI, Rodda, M, Daetwyler, HD, Slater, AT, Forster, JW and Kaur, S (2014) Assessment of genetic variation within a global collection of lentil (Lens culinaris Medik.) cultivars and landraces using SNP markers. BMC Genetics 15: 150.CrossRefGoogle ScholarPubMed
Lonardi, S, Muñoz-Amatriaín, M, Liang, Q, Shu, S, Wanamaker, SI, Lo, S, Tanskanen, J, Schulman, AH, Zhu, T, Luo, MC, Alhakami, H, Ounit, R, Hasan, AM, Verdier, J, Roberts, PA, Santos, JPR, Ndeve, A, Doležel, J, Vrána, J, Hokin, SA, Farmer, AD, Cannon, SB and Close, TJ (2019) The genome of cowpea (Vigna unguiculata [L.] Walp.). The Plant Journal 98: 767782.CrossRefGoogle Scholar
Ma, Y, Marzougui, A, Coyne, CJ, Sankaran, S, Main, D, Porter, LD, Mugabe, D, Smitchger, JA, Zhang, C, Amin, M and Rasheed, N (2020) Dissecting the genetic architecture of Aphanomyces root rot resistance in lentil by QTL mapping and Genome-Wide Association Study. International Journal of Molecular Sciences 21: 2129.CrossRefGoogle ScholarPubMed
Mascher, M, Schreiber, M, Scholz, U, Graner, A, Reif, JC and Stein, N (2019) Genebank genomics bridges the gap between the conservation of crop diversity and plant breeding. Nature Genetics 51: 10761081.CrossRefGoogle ScholarPubMed
McCouch, SR, McNally, KL, Wang, W and Hamilton, RS (2012) Genomics of gene banks: a case study in rice. American Journal of Botany 99: 407423.CrossRefGoogle ScholarPubMed
McCouch, S, Baute, GJ, Bradeen, J, Bramel, P, Bretting, PK, Buckler, E, Burke, JM, Charest, D, Cloutier, S, Cole, G, Dempewolf, H, Dingkuhn, M, Feuillet, C, Gepts, P, Grattapaglia, D, Guarino, L, Jackson, S, Knapp, S, Langridge, P, Lawton-Rauh, A, Lijua, Q, Lusty, C, Michael, T, Myles, S, Naito, K, Nelson, RL, Pontarollo, R, Richards, CM, Rieseberg, L, Ross-Ibarra, J, Rounsley, S, Hamilton, RS, Schurr, U, Stein, N, Tomooka, N, van der Knaap, E, van Tassel, D, Toll, J, Valls, J, Varshney, RK, Ward, J, Waugh, R, Wenzl, P and Zamir, D (2013) Agriculture: feeding the future. Nature 499: 2324.CrossRefGoogle ScholarPubMed
Migicovsky, Z, Gardner, KM, Money, D, Sawler, J, Bloom, JS, Moffett, P, Chao, CT, Schwaninger, H, Fazio, G, Zhong, GY and Myles, S (2016) Genome to phenome mapping in apple using historical data. The Plant Genome 9: 115.CrossRefGoogle ScholarPubMed
Neupane, S (2019) Flowering time response of diverse lentil (Lens culinaris Medik.) germplasm grown in multiple environments. MS Thesis, University of Saskatchewan.Google Scholar
Nguyen, GN and Norton, SL (2020) Genebank phenomics: a strategic approach to enhance value and utilization of crop germplasm. Plants 9: 817.CrossRefGoogle ScholarPubMed
Ogutcen, E, Ramsey, L, von Wettberg, EJB and Bett, K (2018) Capturing variation in Lens: development and utility of an exome capture array for lentil. Applications in Plant Science 6: e01165.CrossRefGoogle ScholarPubMed
O'Leary, SJ, Puritz, JB, Willis, SC, Hollenbeck, CM and Portnoy, DS (2018) These aren't the loci you're looking for: principles of effective SNP filtering for molecular ecologists. Molecular Ecology 27: 31933206.CrossRefGoogle Scholar
Pavan, S, Bardaro, N, Fanelli, V, Marcotrigiano, AR, Mangini, G, Taranto, F, Catalano, D, Montemurro, C, De Giovanni, C, Lotti, C and Ricciardi, L (2019) Genotyping by sequencing of cultivated lentil (Lens culinaris Medik.) highlights population structure in the Mediterranean gene pool associated with geographic patterns and phenotypic variables. Frontiers in Genetics 10: 872.CrossRefGoogle ScholarPubMed
Peakall, R and Smouse, PE (2006) GENALEX 6: genetic analysis in Excel. Population genetic software for teaching and research. Molecular Ecology Notes 6: 288295.CrossRefGoogle Scholar
Peakall, R and Smouse, PE (2012) Genalex 6.5: genetic analysis in Excel. Population genetic software for teaching and research – an update. Bioinformatics (Oxford, England) 28: 25372539.CrossRefGoogle ScholarPubMed
Philipp, N, Weise, S, Oppermann, M, Börner, A, Graner, A, Keilwagen, J, Kilian, B, Zhao, Y, Reif, JC and Schulthess, AW (2018) Leveraging the use of historical data gathered during seed regeneration of an ex situ genebank collection of wheat. Frontiers in Plant Science 9: 609.CrossRefGoogle Scholar
Polanco, C, de Miera, LES, González, AI, García, P, Fratini, R, Vaquero, F, Vences, FJ and de la Vega, MP (2019) Construction of a high-density interspecific (Lens culinaris x L. odemensis) genetic map based on functional markers for mapping morphological and agronomical traits, and QTLs affecting resistance to Ascochyta in lentil. PLoS ONE 14: e0214409.CrossRefGoogle Scholar
Poland, JA (2015) Breeding-assisted genomics. Current Opinion in Plant Biology 24: 119124.CrossRefGoogle ScholarPubMed
Poland, JA, Brown, PJ, Sorrells, ME and Jannink, JL (2012) Development of high-density genetic maps for barley and wheat using a novel two-enzyme genotyping-by-sequencing approach. PLoS ONE 7: e32253.CrossRefGoogle ScholarPubMed
Postman, J, Hummer, K, Bretting, P, Kinard, G, Bohning, M, Emberland, G, Sinnott, Q, Mackay, M, Cyr, P, Millard, M, Gardner, C, Weaver, B, Ayala-Silva, T, Franko, T, Mackay, M and Guarino, L (2010) GRIN‐Global: An international project to develop a global plant genebank information management system. Acta Horticulturae 859: 4956.CrossRefGoogle Scholar
Price, AL, Patterson, NJ, Plenge, RM, Weinblatt, ME, Shadick, NA and Reich, D (2006) Principal components analysis corrects for stratification in genome-wide association studies. Nature Genetics 38: 904909.CrossRefGoogle ScholarPubMed
Pritchard, JK, Stephens, M and Donnelly, P (2000) Inference of population structure using multilocus genotype data. Genetics 155: 945959.CrossRefGoogle ScholarPubMed
Ramsay, L, Chan, C, Sharpe, AG, Cook, DR, Penmetsa, RV, Chang, P, Coyne, C, McGee, R, Main, D, Edwards, D, Kaur, S, Vandenberg, A and Bett, KE (2016) Lens culinaris CDC Redberry genome assembly v1.2. Retrieved from https://knowpulse.usask.ca/genome-assembly/Lc1.2.Google Scholar
Raubach, S, Kilian, B, Dreher, K, Amri, A, Bassi, FM, Boukar, O, Cook, D, Cruickshank, A, Fatokun, C, El Haddad, N, Humphries, A, Jordan, D, Kehel, Z, Kumar, S, Labarosa, SJ, Loi, NH, Mace, E, McCouch, S, McNally, K, Marshall, DF, Mikwa, EO, Milne, I, Odeny, DA, Plazas, M, Prohens, J, Rieseberg, LH, Schafleitner, R, Sharma, S, Stephen, G, Tin, HQ, Abou Togola, A, Emily Warchefsky, E, Peter Werner, P and Shaw, PD (2021) From bits to bites: advancement of the germinate platform to support genetic resources collections and pre-breeding informatics for crop wild relatives. Crop Science 62: 129.Google Scholar
Reda, A (2015) Lentil (Lens culinaris Medik.) current status and future prospect of production in Ethiopia. Advances in Plants & Agriculture Research 2: 00040.CrossRefGoogle Scholar
Rohlf, FJ (2009) Numeric Taxonomy and Multivariate Analysis System (NTSYSpc), version 2.21c. Setauket, New York: Exeter Software.Google Scholar
Sarker, A, Erskine, W, Sharma, B and Tyagi, MC (1999) Inheritance and linkage relationship of days to flower and morphological loci in lentil (Lens culinaris Medikus subsp. culinaris). Journal of Heredity 90: 270275.CrossRefGoogle Scholar
Schmutz, J, McClean, PE, Mamidi, S, Wu, GA, Cannon, SB, Grimwood, J, Jenkins, J, Shu, S, Song, Q, Chavarro, C, Torres-Torres, M, Geffroy, V, Moghaddam, SM, Gao, D, Abernathy, B, Barry, K, Blair, M, Brick, MA, Chovatia, M, Gepts, P, Goodstein, DM, Gonzales, M, Hellsten, U, Hyten, DL, Jia, G, Kelly, JD, Kudrna, D, Lee, R, Richard, MMS, Miklas, PN, Osorno, JM, Rodrigues, J, Thareau, V, Urrea, CA, Wang, M, Yu, Y, Zhang, M, Wing, RA, Cregan, PB, Rokhsar, DS and Jackson, SA (2014) A reference genome for common bean and genome-wide analysis of dual domestications. Nature Genetics 46: 707713.CrossRefGoogle ScholarPubMed
Schneider, S, Roessli, D and Excoffier, L (2000) Arlequin Ver 2000: A Software for Population Genetics Data Analysis. Geneva, Switzerland: Genetics and Biometry Laboratory, University of Geneva.Google Scholar
Singh, S, Singh, I, Gil, RK, Kumar, S and Sarker, A (2009) Genetic studies for yield and component characters in large seeded exotic lines of lentil. Journal of Food Legumes 22: 229232.Google Scholar
Singh, A, Dikshit, HK, Mishra, GP, Aski, M and Kumar, S (2019) Association mapping for grain diameter and weight in lentil using SSR markers. Plant Gene 20: 100204.CrossRefGoogle Scholar
Sneath, PHA and Sokal, RR (1973) Numerical Taxonomy. San Francisco: W. H. Freeman and Company.Google Scholar
Song, Q, Hyten, DL, Jia, G, Quigley, CV, Fickus, EW, Nelson, RL and Cregan, PB (2015) Fingerprinting soybean germplasm and its utility in genomic research. G3: Genes, Genomes, Genetics 5: 19992006.CrossRefGoogle ScholarPubMed
Tanksley, SD and McCouch, SR (1997) Seed banks and molecular maps: unlocking genetic potential from the wild. Science (New York, N.Y.) 277: 10631066.CrossRefGoogle Scholar
Temel, HY, Göl, D, Akkale, HBK, Kahriman, A and Tanyolac, MB (2015) Single nucleotide polymorphism discovery through Illumina-based transcriptome sequencing and mapping in lentil. Turkish Journal of Agriculture and Forestry 39: 470488.CrossRefGoogle Scholar
van Treuren, R and van Hintum, TJ (2014) Next-generation genebanking: plant genetic resources management and utilization in the sequencing era. Plant Genetic Resources 12: 298307.CrossRefGoogle Scholar
Varshney, RK (2016) Exciting journey of 10 years from genomes to fields and markets: some success stories of genomics-assisted breeding in chickpea, pigeonpea and groundnut. Plant Science 242: 98107.CrossRefGoogle ScholarPubMed
Varshney, RK, Chen, W, Li, Y, Bharti, AK, Saxena, RK, Schlueter, JA, Donoghue, MT, Azam, S, Fan, G, Whaley, AM and Farmer, AD (2012) Draft genome sequence of pigeonpea (Cajanus cajan), an orphan legume crop of resource-poor farmers. Nature Biotechnology 30: 83.CrossRefGoogle Scholar
Varshney, RK, Song, C, Saxena, RK, Azam, S, Yu, S, Sharpe, AG, Cannon, S, Baek, J, Rosen, BD, Tar'an, B and Millan, T (2013) Draft genome sequence of chickpea (Cicer arietinum) provides a resource for trait improvement. Nature Biotechnology 31: 240246.CrossRefGoogle ScholarPubMed
Varshney, RK, Singh, VK, Hickey, JM, Xun, X, Marshall, DF, Wang, J, Edwards, D and Ribaut, JM (2015) Analytical and decision support tools for genomics-assisted breeding. Trends in Plant Science 21: 354363.CrossRefGoogle ScholarPubMed
Varshney, RK, Sinha, P, Singh, VK, Kumar, A, Zhang, Q and Bennetzen, JL (2020) 5Gs for crop genetic improvement. Current Opinion in Plant Biology 56: 190196.CrossRefGoogle ScholarPubMed
Verma, P, Goyal, R, Chahota, RK, Sharma, TR, Abdin, MZ and Bhatia, S (2015) Construction of a genetic linkage map and identification of QTLs for seed weight and seed size traits in lentil (Lens culinaris Medik.). PLoS ONE 10: e0139666.CrossRefGoogle Scholar
Weller, JL, Liew, LC, Hecht, VF, Rajandran, V, Laurie, RE, Ridge, S, Wenden, B, Vander Schoor, JK, Jaminon, O, Blassiau, C and Dalmais, M (2012) A conserved molecular basis for photoperiod adaptation in two temperate legumes. Proceedings of the National Academy of Sciences 109: 2115821163.CrossRefGoogle ScholarPubMed
Wong, MML, Verma, NG, Ramsay, L, Yuan, HY, Caron, C, Diapari, M, Vandenberg, A and Bett, KE (2015) Classification and characterization of species within the genus Lens using genotyping-by-sequencing (GBS). PLoS ONE 10: e0122025.CrossRefGoogle Scholar
Yu, LX, Zheng, P, Bhamidimarri, S, Liu, XP and Main, D (2017) The impact of genotyping-by-sequencing pipelines on SNP discovery and identification of markers associated with verticillium wilt resistance in autotetraploid alfalfa (Medicago sativa L.). Frontiers in Plant Science 8: 89.CrossRefGoogle Scholar
Zimmerer, KS and De Haan, S (2017) Agrobiodiversity and a sustainable food future. Nature Plants 3: 17047.CrossRefGoogle Scholar
Figure 0

Fig. 1. The subpopulations of K = 3 as determined by the ad hoc statistic ΔK based on the rate of change in the log probability of data between successive K = 1–7 (Evanno et al., 2005).

Figure 1

Fig. 2. Dendrogram based on UPGMA and the subpopulations (K = 3) calculated using the Bayesian clustering method of the software STRUCTURE based on SNP data for 176 lentil accessions (Pritchard et al., 2000).

Figure 2

Fig. 3. Principal component analysis of 176 lentil accessions at K = 3 based on SNP genotyping. Colours correspond to greater than 50% association with a subpopulation and colours correspond to Fig. 3. White accessions are admixtures. Further views available in online Supplementary Material (Fig. S2).

Figure 3

Table 1. Significant SNP markers identified using genome-wide associations for four traits based on ICARDA's historical phenotypic data collected from 2007 to 2011

Supplementary material: File

Rajendran et al. supplementary material

Figures S1-S4 and Tables S1-S5
Download Rajendran et al. supplementary material(File)
File 5 MB