Skip to main content

REVIEW article

Front. Plant Sci., 04 July 2022
Sec. Plant Breeding
This article is part of the Research Topic Use of Barley and Wheat Reference Sequences: Downstream Applications in Breeding, Gene Isolation, GWAS and Evolution, Volume II View all 17 articles

Capturing Wheat Phenotypes at the Genome Level

  • 1Department of Biological Sciences, Middle East Technical University, Ankara, Turkey
  • 2Department of Biotechnology, Faculty of Life Sciences, University of Central Punjab, Lahore, Pakistan
  • 3Montana BioAgriculture, Inc., Missoula, MT, United States
  • 4Université Paris-Saclay, INRAE, URGI, Versailles, France
  • 5Department of Environment and Bio-Agriculture, Faculty of Agriculture, Al-Azhar University, Cairo, Egypt
  • 6International Maize and Wheat Improvement Center (CIMMYT), Texcoco, Mexico
  • 7Institute of Plant Breeding and Biotechnology, MNS University of Agriculture, Multan, Pakistan
  • 8Department of Pathology, The National Institute of Agricultural Botany, Cambridge, United Kingdom
  • 9School of Biological Sciences and Institute of Agriculture, University of Western Australia, Perth, WA, Australia
  • 10French Plant Genomic Resource Center, INRAE-CNRGV, Castanet Tolosan, France
  • 11Crop Improvement and Genetics Research, USDA, Agricultural Research Service, Albany, CA, United States
  • 12Council for Agricultural Research and Economics-Research Centre for Genomics and Bioinformatics, Fiorenzuola d’Arda, Italy
  • 13French National Research Institute for Agriculture, Food and the Environment, INRAE, GDEC, Clermont-Ferrand, France
  • 14The John Bingham Laboratory, The National Institute of Agricultural Botany, Cambridge, United Kingdom
  • 15Research & Innovation, Florimond Desprez Group, Cappelle-en-Pévèle, France
  • 16USDA, Agricultural Research Service, Edward T. Schafer Agricultural Research Center, Fargo, ND, United States
  • 17Department of Bioquímica y Biología Molecular, Campus Rabanales C6-1-E17, Campus de Excelencia Internacional Agroalimentario (ceiA3), Universidad de Córdoba, Córdoba, Spain
  • 18University of Western Australia, Perth, WA, Australia
  • 19State Plant Breeding Institute, The University of Hohenheim, Stuttgart, Germany
  • 20International Wheat Genome Sequencing Consortium (IWGSC), Bethesda, MD, United States
  • 21Institute of Evolution and Department of Environmental and Evolutionary Biology, University of Haifa, Haifa, Israel
  • 22Commonwealth Scientific and Industrial Research Organization, Agriculture and Food, Canberra, ACT, Australia
  • 23Department of Languages and Computer Science, ETSI Informática, Campus de Teatinos, Universidad de Málaga, Andalucía Tech, Málaga, Spain
  • 24Department of Crop Science, Washington State University, Pullman, WA, United States
  • 25Atta-ur-Rahman School of Applied Biosciences, National University of Sciences and Technology, Islamabad, Pakistan
  • 26Center of Plant Genome Engineering, Heinrich-Heine-Universität, Düsseldorf, Germany
  • 27Division of Molecular Biology, Centre of Region Haná for Biotechnological and Agriculture Research, Czech Advanced Technology and Research Institute, Palacký University, Olomouc, Czechia
  • 28Institute for Sustainable Agriculture (IAS-CSIC), Consejo Superior de Investigaciones Científicas (CSIC), Córdoba, Spain
  • 29Crop and Soil Science, Texas A&M University, College Station, TX, United States
  • 30Global Crop Diversity Trust, Bonn, Germany
  • 31KWS SAAT SE & Co. KGaA, Einbeck, Germany
  • 32Department of Plant Pathology, Faculty of Agriculture, Assiut University, Assiut, Egypt
  • 33Food and Agriculture Organization of the United Nations, Riyadh, Saudi Arabia
  • 34Molecular Biology, Genetics and Bioengineering, Sabanci University, Istanbul, Turkey
  • 35Institute for Resistance Research and Stress Tolerance, Julius Kühn Institute, Quedlinburg, Germany
  • 36USDA-Agricultural Research Service, Southern Insect Management Research Unit, Stoneville, MS, United States
  • 37Leibniz Institute of Plant Genetics and Crop Plant Research (IPK), Gatersleben, Germany
  • 38Department of Genetics and Bioengineering, Yeditepe University, Istanbul, Turkey
  • 39University of Maryland, Baltimore, MD, United States
  • 40Institute of Biological Sciences (IBS), Gomal University, D. I. Khan, Pakistan
  • 41Ficus Biotechnology, Ostim Teknopark, Ankara, Turkey
  • 42General Directorate of Research, Ministry of Agriculture, Ankara, Turkey
  • 43Murdoch University, Perth, WA, Australia

Recent technological advances in next-generation sequencing (NGS) technologies have dramatically reduced the cost of DNA sequencing, allowing species with large and complex genomes to be sequenced. Although bread wheat (Triticum aestivum L.) is one of the world’s most important food crops, efficient exploitation of molecular marker-assisted breeding approaches has lagged behind that achieved in other crop species, due to its large polyploid genome. However, an international public–private effort spanning 9 years reported over 65% draft genome of bread wheat in 2014, and finally, after more than a decade culminated in the release of a gold-standard, fully annotated reference wheat-genome assembly in 2018. Shortly thereafter, in 2020, the genome of assemblies of additional 15 global wheat accessions was released. As a result, wheat has now entered into the pan-genomic era, where basic resources can be efficiently exploited. Wheat genotyping with a few hundred markers has been replaced by genotyping arrays, capable of characterizing hundreds of wheat lines, using thousands of markers, providing fast, relatively inexpensive, and reliable data for exploitation in wheat breeding. These advances have opened up new opportunities for marker-assisted selection (MAS) and genomic selection (GS) in wheat. Herein, we review the advances and perspectives in wheat genetics and genomics, with a focus on key traits, including grain yield, yield-related traits, end-use quality, and resistance to biotic and abiotic stresses. We also focus on reported candidate genes cloned and linked to traits of interest. Furthermore, we report on the improvement in the aforementioned quantitative traits, through the use of (i) clustered regularly interspaced short-palindromic repeats/CRISPR-associated protein 9 (CRISPR/Cas9)-mediated gene-editing and (ii) positional cloning methods, and of genomic selection. Finally, we examine the utilization of genomics for the next-generation wheat breeding, providing a practical example of using in silico bioinformatics tools that are based on the wheat reference-genome sequence.

The 17 Gbp Wheat Genome: Challenges and Opportunities

Wheat (Triticum aestivum L.) is the key crop for feeding the Earth’ growing population, remaining a staple food in many regions of the world. It is cultivated on more than 220 million hectares worldwide, and global production exceeds 749 million tons annually.1 Bread wheat is a hexaploid species (2n = 6x = 42, genome AABBDD) that evolved via natural hybridization between tetraploid domesticated wheat T. turgidum ssp. dicoccum (contributed the AA and BB sub-genomes) and the wild grass species Aegilops tauschii (DD sub-genome), followed by the domestication of the resulting hexaploid spelt wheat (T. spelta; Petersen et al., 2006). The wheat genome is ~17 Gbp in size and contains a high degree of complexity, particularly in terms of chromosomal duplications and rearrangements, and the very high percentage of repetitive sequences (IWGSC, 2014; Akpinar et al., 2015).

Wheat breeding targets are numerous and varied, given the wide geographic area across which wheat is grown. However, the principal common targets are grain yield (GY), quality determinants, and tolerance to biotic and abiotic stresses. The complexity of the wheat genome makes improving qualitative and quantitative traits through molecular approaches challenging. An example of this is drought tolerance, which is conferred by diverse signaling molecules, including micro-RNA (miRNA), transcriptional factors, quantitative-trait loci (QTL), transcripts, proteomes, ionomes, and other metabolites, resulting in a complex signaling cascade for the control of traits such as abiotic stress (Budak et al., 2015). Furthermore, multiple genes are involved in the production and regulation of these molecules, which leads to a complex signaling cascade, responsible for conferring abiotic/biotic-stress tolerance. Hence, knowledge of the sequence, as well as the precise location, annotation, and casual polymorphisms of the genes involved is vital for utilizing the genomic data in breeding programs, aimed at achieving specific and desired traits or phenotypes.

Due to its large genome size in comparison with other major crops with smaller genomes, efforts to sequence and annotate the wheat genome have been extremely time-consuming, often involving sequencing of individual chromosomes (Berkman et al., 2011, 2012). Thus, the International Wheat Genome Sequencing Consortium (IWGSC) reported a draft sequence of bread wheat (cv. Chinese Spring; CS) in 2014, derived from sequencing flow-sorted chromosomes/chromosome arms. The draft assembly totaled 12.7 Gbp, comprising 124,201 gene loci distributed across A, B, and D sub-genomes (IWGSC, 2014). However, this assembly contained only approximately three-quarters of the whole wheat genome. Furthermore, the genome sequences of the chromosomes/chromosome arms were fragmented with many gaps as well as many incomplete, absent, or incorrectly assigned genes, making it hard for scientists to find and elucidate specific genes (Berkman et al., 2013; IWGSC, 2014; Lai et al., 2015). Despite the incompleteness of this version, it was highly useful for breeders, as it provided valuable information at the chromosomal/chromosome-arm level. A draft whole-genome sequence of wheat was obtained by combining long Pacific Biosciences (PacBio) reads (>10,000 bases long) with short (150 bp) Illumina reads, with 15.34 Gbp and an average contig size of 0.23 Mbp (Zimin et al., 2017). Low-coverage sequence data for 16 varieties were released in 2012 and used as the basis for the first draft wheat pan-genome (Edwards et al., 2012). This study highlighted the fact that, due to gene presence/absence variation, a single reference does not represent the gene content of the species (Bayer et al., 2020; Danilevicz et al., 2020; Golicz et al., 2020).

The high-quality reference sequence of wheat genome was achieved by the International Wheat Genome Sequencing Consortium in several steps. A whole-genome sequence based on Illumina technology and a draft assembly was released in 2016 (IWGSC WGS v0.4), which was comprised of Illumina short sequence reads assembled with NRGene’s DeNovoMagic (Appels et al., 2018).2 This was then combined with physical maps of the chromosome/chromosome arm and other genomic resources that had been developed over 13 years by numerous laboratories around the world, to develop IWGSC RefSeq v1.0. In 2018, the fully annotated reference-genome assembly was released (IWGSC RefSeqv1.1; Appels et al., 2018), with the precise location and annotation of 107,891 high-confidence genes and more than 4 million molecular markers along the 21 chromosomes. The chromosome-scale assembly covered approximately 94% of the bread-wheat genome (cv. Chinese Spring), with a total assembly size of 14.5 Gbp. A key feature of this new genome assembly was the long scaffolds, of which 90% were larger than 4.1 Mbp, the longest super scaffold being 166 Mbp (i.e., larger than the 135 Mbp Arabidopsis thaliana genome and half the size of the rice genome; Appels et al., 2018). Accordingly, RefSeqv1.0, with the highest sequence contiguity, has become a tool for wheat genomics and breeding activities worldwide. In 2020, the release of genome assemblies for 15 additional wheat accessions, with diverse origins across the globe (Walkowiak et al., 2020), has further consolidated the position of wheat in the genomics era, providing additional resources to underpin breeding strategies. The availability of multiple, high-quality genome assemblies for wheat has highlighted the genomic diversity present in the global breeding program. Introgressions from wild relatives, structural rearrangements, and variation in gene content, originating from various breeding efforts aimed for diverse and multiple traits have contributed to wheat genomic diversity (Walkowiak et al., 2020). An advantage of having multiple assemblies is that it enables the discovery of new sequences and genes that were not present in previous versions of the wheat genome, thus creating new opportunities to identify, characterize, and exploit the beneficial alleles/haplotypes present for wheat improvement. IWGSC RefSeq v2.0 available since 25 July 2019, as shown at “IWGSC RefSeq v2.0 now available at URGI.”3

In summary, these genome assemblies represent an essential, highly efficient resource for wheat researchers and breeders, to identify and clone major genes and QTL, to elucidate regulatory regions, including miRNAs and transcription factors, and gene networks involved in yield, as well as biotic- and abiotic-stress tolerance in wheat, thereby facilitating their use in wheat improvement programs (Figure 1). Capturing phenotypes (as QTL) at the genomic level has brought a sharp focus on the repetitive nature of the gene-coding space in the wheat genome, allowing both copy number variation, as well as sequence variation per se, to be correlated with phenotype variation. The distribution of essentially identical genes on different non-syntenic chromosomes provides for further flexibility, as well as complexity, in fine-tuning phenotypes to specific environments.

FIGURE 1
www.frontiersin.org

Figure 1. Overview of the parallel progress in the analysis of the wheat genome and high throughput phenotyping. The top panel provides the timeline for wheat-genome studies, opening up of the next generation omics; the image on the far right is the modeling of wheat granule-bound starch synthase, using Phyre 2 (Kelley et al., 2015). The lower panel emphasizes the progress of both field-based phenomics (image of drone with spectral-recording equipment, kindly provided by S. Kant, Agriculture Victoria, Grains Innovation Park, Horsham, VIC, Australia) and laboratory-based high-throughput analyses (part of Figure 6 of Banerjee et al., 2020), showing false-color composite from hyperspectral data of wheat leaves, kindly provided by S. Kant.

In this paper, we consider the potential of current whole-genome assemblies to improve the accuracy and resolution of genetic mapping, QTL mapping, genome-wide association studies (GWAS), and the use of sequence-based markers for efficient MAS and genomic selection. A key focus for the review is the identification and cloning of major candidate genes (CG), for traditional wheat-breeding programs, and modern ones using Clustered Regularly-Interspaced Short Palindromic-Repeats (CRISPR) associated protein 9 (CRISPR/Cas9)-mediated genome editing, and bioinformatics tools for wheat improvement complement the advances in genotyping and phenotyping. They herald the start of what might be considered a golden era of wheat genomics-assisted breeding, promoting the aim of sustainably intensifying global-food production.

Wheat Reference-Genome and in silico Bioinformatics

The so-called next gene generation sequencing (NGS) technologies, as well as high-resolution optical mapping generate large data sets (Dorado et al., 2019). They have driven improvements in bioinformatics tools used to deal with the challenge of analyzing such large data sets. Advancements have been made possible by a dual-strategy approach, focused on hardware and software. They have included clock-frequency increases for the Central Processing Units (CPU) of computers, node reduction, multicore (a few), many core (higher number) and integration through System on a Chip (SoC) with unified memory between the CPU and Graphics Processing Units (GPU). Machine learning, artificial intelligence, dedicated artificial neural network (ANN) analyses, and massive parallelism are enhanced using multi-core architectures (Gálvez et al., 2016, 2021). All this has contributed to our ability to sequence de novo, assemble, and annotate extremely large and complex genomes.

Abundant transcriptome sequence data have been generated in wheat, being freely available in open-access servers such as Wheat Expression,4 which can be utilized to narrow down the candidate genes identified in GWAS analyses, for further validation projects. The availability of sequenced mutant populations has opened doors to conduct validation studies.5 Predesigned single-nucleotide polymorphism (SNP)-based primers are available in the Ensemble database6 for validating mutations, which can then be combined to develop double- or triple-null mutants, for research and breeding applications. Efforts in this direction are expected if “causal” genes are to be identified. Transcriptomics and other expression analyses have generally been combined with QTL and metaQTL (mQTL), to narrow down and validate the candidate genes (Oyiga et al., 2018; Gálvez et al., 2019; Mérida-García et al., 2020). We expect it to be applied more frequently in marker-assisted and genomic selection. Continued advances in fingerprinting genome regions of interest, through improved designs of SNP arrays and associated statistical analyses of imputation, is establishing haplotype analyses as a more appropriate method to represent QTL, rather than relying on single-candidate genes.

A significant challenge that scientists have faced with the availability of the wheat genome reference (IWGSC RefSeq v 1.0) is its integration with the previously published genetic maps harboring QTL for various traits. Large datasets including physical maps, sequence variations, gene expression, markers, and phenomic data have already been integrated on the IWGSC RefSeq v1.0 (Wheat@URGI portal; Alaux et al., 2018). Furthermore, haplotype-based integration of different marker types and the capacity to align early genetic maps with the reference genome are refined, capitalizing on existing information of trait-linked SNP, DArTseq and/or Genotype By Sequencing (GBS) markers (PRETZEL; Figure 2).7

FIGURE 2
www.frontiersin.org

Figure 2. Aligning genome maps for SNP, DArTseq and/or GBS markers with IWGSC RefSeq 1.0 using PRETZEL https://plantinformatics.io. The right most map provides the location of a QTL for the emergence of additional seminal roots (midpoint = 26.6 cM; Golan et al., 2018) from a Svevo x Zavitan map based on a 90K SNP chip. The second map (from the right) is the durum genome sequence for 1B, available at the URGI with the SNPs annotations included. The second map from the left is the IWGSC RefSeq 1.0 with the HC ver1.1 gene annotation, the 90k SNP annotation, the LC ver1.1 gene annotations and the SSR annotations included. The left most map is the genome sequence for the 1RS.1BL sequence from wheat cv Aikan58 (Ru et al., 2020) with three sources of gene annotations included. The red dots identify the gene models predicted to be located in the QTL identified in the Svevo x Zavitan QTL.

Characterization of Genes and Gene Families Using the Wheat Reference-Genome

As indicated above, the availability of reference genome sequences in many crop species, including wheat, has sparked the publication of many works about genomics and breeding of such species, using bioinformatics tools, with special emphasis on previously unknown areas of the genome. For instance, in silico analyses of the published wheat reference genome, IWGSC RefSeq v1 (Appels et al., 2018), have allowed the identification and characterization of the gene families. They include: (i) Domain of Unknown Function (DUF-966, TaDUF966) gene family, involved in salinity-stress tolerance (Zhou et al., 2020); (ii) MADS-Box gene (TaMADS-box) family members, involved in wheat growth, development, and abiotic stresses (Raza et al., 2021); (iii) Gretchen Hagen3 (TaGH3) gene family, important in various biological processes, including phytohormone responses, growth, development, metabolism, defense, and abiotic-stress tolerance, such as salinity and osmotic ones, with polyploidization contributing to their high number (Jiang et al., 2020); (iv) superoxide dismutase (SOD) gene (TaSOD) family, encoding antioxidant enzymes scavenging reactive oxygen species (ROS), also involved in plant growth, development, and abiotic-stress tolerance, including drought and salinity (Jiang et al., 2019); (v) non-specific lipid transfer proteins (nsLTP/LTP) gene (TansLTP/TaLTP) family, involved in transporting phospholipids across membranes, growth, development, and abiotic stresses, such as drought and salinity, showing high numbers, due to gene duplications (Fang et al., 2020a); (vi) REMorin (REM) gene (TaREM) family, involved in vernalization, plant-microbe interactions, hormonal regulation, development, and tolerance to biotic and abiotic stresses, including cold acclimation (Badawi et al., 2019); (vii) S-phase Kinase-associated Protein 1 (SKP1) gene (TaSKP1) family, encoding core subunits of the Ubiquitin Proteasome 26S (UPS) and have expanded through duplications, being involved in development and stress signaling (El Beji et al., 2019); (viii) subtilase or subtilisin-like protease (SBT) genes (TaSBT), involved in many biological functions, such as defense and tolerance to biotic stresses caused by pathogens, among which are Puccinia striiformis f. sp. tritici, which is the fungus generating the wheat stripe-rust disease (Yang et al., 2020); (ix) highly and structurally conserved “Soluble N-ethylmaleimide Sensitive Factor (NSF; SNF) Attachment Protein (SNAP) REceptor” (SNARE) and Novel Plant SNare (NPSN) gene families (TaSNARE and TaNPSN, respectively), involved in growth and development, regulating vesicle trafficking, fusion, and targeting to vacuoles and exocytosis (Gaggar et al., 2020); (x) “DNA binding with one finger” (Dof) gene (TaDof) family, encoding zinc-finger transcription factors (TaDof), involved in phytohormone response, growth, development, metabolism, defense, and stress responses, including both abiotic (such as salinity and drought) and biotic ones, with their high numbers due to polyploidization, showing many segmental duplications and both miRNA and cis-regulators involvement in modulating their gene expression profiles (Fang et al., 2020b); and (xi) basic leucine ZIPper (bZIP) gene (TabZIP) family, encoding transcription factor, being involved in plant growth, development, metabolism, chlorophyll content, photosynthesis, membrane stability, and tolerance to stresses, including abiotic ones such as drought, salinity, and heat, also involving oxidative stress (Agarwal et al., 2019). These gene families complement many gene models identified by highly conserved domains, using model organisms such as rice and Arabidopsis, and provide a matrix across the wheat genome, to associate with phenotypic variation in order to establish a new matrix of functionally, with annotated gene models predicted to affect wheat phenotype.

Bioinformatics analyses of the published wheat reference genome, IWGSC RefSeq v1.0, have also enabled the comparisons of cDNA, identification, and annotation of genes. These comparisons revealed different metabolic pathways, including starch and sucrose, as well as genes related to abiotic- and biotic-stress tolerance, signaling and transportation (Kaur et al., 2019). Delivering the range of bioinformatics outputs from the wheat genome sequence has required innovation in the production of genotyping microarrays, such as the Axiom Wheat high-density Genotyping Array, Axiom Wheat Breeder’s Genotyping Array, and GeneChip Wheat Genome Array from Affymetrix-Applied Biosystems-Thermo Fisher Scientific.8 The efficacy of the SNP arrays can now be evaluated using a three-way classification system, permitting sorting of SNP into three quality groups (Lange et al., 2020).

The concept of having the wheat genome on a chip is currently under development by the Arbor Biosciences IWGSC exome array group, working with the IWGSC, with a first release in 2019 of the Baits Expert Wheat Exome capture panel, based on the complete high-confidence exon-annotated wheat genome. It included two million probes, targeting more than 200 megabases of high-confidence exons.9 An extra level of fine-tuning for gene regulation of different biological processes, including metabolism, growth, development, transport, cell signaling, structural proteins, and abiotic- and biotic-stress tolerance, is afforded by the characterization of new microRNA (miRNA), including polycistronic miRNA, in cultivated and wild species. The genes targeted by these small RNA can be predicted, and some are monomorphic, whereas others are polymorphic, all represented on the Expert Exome capture chip (Singh et al., 2020).

The small RNAs are spatially and temporally regulated, being involved in post-transcriptional gene regulation, including transcription factors, and thus complement the genome-wide transcriptomics approach established to dissect the dynamics and underlying regulation of key processes, such as wheat-spike development. Genes involved are related to meristem maintenance, initiation and transition, development of flowers, and flowering response to stress (Li et al., 2018b).

Grain Yield and Related Traits

Wheat grain yield is controlled by numerous genetic components, most of which are quantitative in nature. Due to this underlying complexity, QTL mapping is commonly used for the dissection of grain yield and yield components, in order to identify markers for MAS. Prior to the 2014 draft sequence, several QTL studies have reported using redundant SSR-markers for QTL mapping of GY and related traits, as recently reviewed (30, see also Supplementary Table S1), but most of these regions were not incorporated into wheat cultivars in breeding programs using MAS. The IWGSC draft sequence published in 2014 (IWGSC, 2014) enabled the use of genotyping arrays and GBS, with deep coverage to construct high-density linkage maps and identify several candidate genes (Addison et al., 2016; Su et al., 2016; Assanga et al., 2017; Cui et al., 2017; Hussain et al., 2017b; Kuzay et al., 2019; Jin et al., 2020). Major and stable QTL for plant height, anthesis date, flag-leaf length and width, as well as spike length, density and spikelet number per spike were mapped on chromosome 2D and 4B, with individual phenotypic variation (PV) range of 10.10%–30.68%. Other QTLs were mapped on chromosomes 4A and 6D. The markers were associated with candidate genes coding for TGTCTC auxin response elements, F-box protein TIR1, flowering locus T-like protein, MADS-box transcription factor 8 and 12 genes encoding SAUR-like auxin-responsive family proteins (Jin et al., 2020). Three independent studies identified haplotype SNP markers and major stable QTL for seed number per pod (SNPP), thousand-grain weight (TGW), grain length, flag leaf length, width, and area on chromosomes 7A (Su et al., 2016; Hussain et al., 2017b; Kuzay et al., 2019) and 5A (Hussain et al., 2017b). These QTL were associated with candidate genes, such as Wheat ortholog of Aberrant Panicle Organization 1 (WAPO1) and TaGASR7. Among these, a well-studied and reproducible yield QTL on the long arm of chromosome 7A has been located to an 87-kbp region (674,019,191–674,106,327 bp, IWGSC RefSeq v1.0), containing two full and two partial genes. The ortholog of one of these genes (TraesCS7A01G481600) was APO1, which is known to significantly affect panicle attributes (Kuzay et al., 2019). This APO1 ortholog was the best candidate for the spikelets per spike phenotype, being associated with two amino acid changes (C47F and D384N) in the coding region. In the genomic region carrying the chromosome 7A APO1 gene, three major haplotypes were associated with the spikelets per spike phenotype, and two of these show enrichment in modern germplasm (Kuzay et al., 2019; Voss-Fels et al., 2019). More recently, genetic analyses were carried out using a wheat multi-founder population genotyped with a 20K SNP array. They found that allelic variation at the homoeologous location on chromosome 7B was associated with haplotype variation at the WAPO-B1 gene (Corsi et al., 2021). Another recent example of the use of high-density SNP arrays for the genetic mapping of yield components was the use of a 660K SNP array that led to the identification of a stable major QTL for grain number per spike on chromosome 4A that corresponded to 65 putative genes (Cui et al., 2017) and contributed 8.0%–21.2% to PV.

Networks linking quality attributes of the grain to the yield of the grain are evident, as genes initially identified in the quality space have been characterized for their roles in the improvement of yield and related traits. Examples of several such genes include TaGW2 (Su et al., 2011), cell-wall invertase TaCwi-A1 (Ma et al., 2012), TaGASR7-A1 (Dong et al., 2014), TaGS-D1 (Zhang et al., 2014), IAA-glucose hydrolase gene, TaTGW6 (Hu et al., 2016a), and TaTGW-7A (Hu et al., 2016b). Marker-trait associations (MTAs) for yield and related traits have been linked to candidate genes/loci such as Rht-B1, Rht-D1, Vrn-A1, Ppd-D1, TaSus1, TaSus2, TaGS-D1, and TaGW2-6B (Lopes et al., 2015; Zanke et al., 2015). Importantly, a GBS-based GWAS for 768 wheat accessions identified 395 QTL for plant height, DH, SNPP, spike length and number, grain length, grain width, and TGW under seven environments (Pang et al., 2020). These QTL were closely linked with several candidate genes including, but not limited to, yield-related genes: APO1, AUX1, Ehd2, GSN1, GL3, Gn1a, MADS14, MADS15, MADS18, MADS57, Rht-D1, Rht12, TaGA2ox8, TaSus1-7B, Vrn-B1, and Vrn-D1.

Within the matrix of the functional wheat genome, projection of yield-related QTL onto a set of well-defined wheat genes (99,386) identified 32 metaQT including 18 grain-yield mQTL associated with 15,772 genes (28,630 SNP), 37 of which were major candidate genes (Quraishi et al., 2017), including ATPase, GIF1, Ppd-D1, Prog1, Gn1-a, NYC1, emp4, DEP1, GW2, GS2, and Rc3 (Pang et al., 2020).

Drought Tolerance

In candidate gene-based association mapping approaches, resequencing of genes with known or predicted biochemical function is performed, and SNP variation identified within a candidate gene is used to investigate associations with traits (Iehisa et al., 2014; Supplementary Table S2). Using this approach, allelic variation in four drought-related genes has been investigated in wheat. Edae and colleagues (Shokat et al., 2020) reported associations of SNP in three genes, DREB1A, ERA1, and 1-FEH, with multiple agronomic and physiological traits. These are known to be drought stress-induced genes in ABA-dependent (ERA1) and ABA-independent (DREB1A, 1-FEH) pathways. In another CG-based association mapping study, allelic variation in TaSnRK2.8 (a SNF-1 type serine–threonine protein kinase) showed association with plant height, flag leaf width, and water-soluble carbohydrates, under drought conditions (Zhang et al., 2013). Other CG involved in wheat AB/ABA-dependent/ABA-independent signaling pathways have been described, including DREB1, WRKY1, DREB1A, HKT-1, DREB2, DREB3, ERA1-B, ERA1-D, 1-FEH-A, and 1-FEH-B (Budak et al., 2015). Finally, a drought tolerance QTL on chromosome 6D was associated with regulation of expression of late embryogenesis abundant (LEA) genes such as TaABA8, OH1, CYCB2, and CDKA1 (Iehisa et al., 2014), suggesting that these genes play a role in drought tolerance. Several clusters of drought-responsive genes (DREG) have been found on the long arm of the group 5 chromosomes, with orthology to a known QTL of rice (Oryza sativa). In particular, this region contains the genes PSY3, NCED, VRN1, UGDH, and the dehydrin DHN38 that increase their expression under field drought stress conditions (Gálvez et al., 2019). Similarly, transcriptomics analyses of high yielding and drought tolerant United States wheat-cultivars TAM 111 and TAM 112 identified several DREG. For instance, under drought stress, aquaporin, dehydrogenase, kinase, phosphatase synthase, phosphorylase, and sugar transporter were downregulated. On the other hand, dehydrin, ABA-inducible protein kinases, LEA protein, heat-shock protein, caleosin, lyase, amylase, and oxidoreductase were upregulated in such environments (Chu et al., 2021). These studies now clearly engage the very broad aspects of wheat biology, providing a subset of genes for more detailed study in characterizing variation in drought tolerance, facilitating the identification of suitable parents for breeding.

Large-scale haplotypes-based GWAS studies, combined with epistatic interactions, have been undertaken to untangle the genetic architecture of grain yield under multiple stress environments (including mild and severe drought stress), using a large panel of 6,333 advanced lines, from the International Maize and Wheat Improvement Centre (CIMMYT; Sehgal et al., 2020a). This study reported haplotype associations with grain yield, under mild (four datasets) and severe (10 datasets) drought stress environments. Most importantly, the authors identified a significant association of a haplotype block close to the Vrn-B1 flowering time gene on chromosome 5B, with GY in more than 70% of the trials under severe drought stress. Vrn-B1 is significantly correlated with adaptation to low temperature, thus indicating a shared tolerance mechanism for both abiotic stresses.

Heat Tolerance

A candidate gene approach to heat tolerance allowed the examination of variation in Fv/Fm, a parameter for assaying the maximum quantum efficiency potential of Photosystem II (Supplementary Table S3). It was linked to three major QTL mapped on chromosomes 3B and 1D (Sharma et al., 2017), involving two light-reaction genes, chloroplastic NAD(P)H-quinone oxidoreductase subunit 2B (ndhB2) and photosystem I iron–sulfur center (psaC). Two other genes, known to control chloroplastic 3-isopropylmalate dehydrogenase 2 (IMDH2), were also suggested to be involved in metal binding during photosynthesis. Other candidate genes, such as beta-glucosidase 26 (βglu26) and fructokinase 2 (frk2; Sharma et al., 2017), were involved in carbohydrate metabolism. Response to thylakoid membrane damage, plasma-membrane damage and SPAD chlorophyll-content QTL were mapped on chromosomes 1B, 1D, 6A, and 7A. Cell membrane stability was on chromosomes 1D, 2B and 7A (Talukder et al., 2014). These QTL were associated with candidate genes such as srg6, calcium/calmodulin-dependent kinase (CDPK), topoisomerase I (Top1), and aquaporins. A GWAS study identified 20 significant MTA for cell membrane stability on chromosomes 1A, 1B, 2A, 4A, 4B, 6B and 7B, 13 of which were situated solely on 6B (ElBasyoni et al., 2017). All these SNP were then annotated as candidate genes, glycerophosphoric-diester phosphodiesterase (GDE1), TRAF-type zinc-finger protein, SWI3B, and ATPase (Supplementary Table S3). The network nature of grain yield, heat, and drought susceptibility indices, as well as yield stability coefficient, were also reported (Sehgal et al., 2017), as being associated with the flowering time genes Vrn-B1, Ppd-D1, and Vrn-D3.

Salinity Tolerance

Salt tolerance is a complex trait (Hussain et al., 2015). In recent years, SNP-based genotyping platforms, including 9K (Asif et al., 2018), 35K (Hussain et al., 2017a; Chaurasia et al., 2020), 90K (Oyiga et al., 2018) and 660K (Yu et al., 2020; Hu et al., 2021) arrays, have been used to identify novel and major QTL, MTA, and CG that can be used in MAS and genomic selection for salinity tolerance (Supplementary Table S4). Wheat F2 lines (WTSD91 × WN-64) were genotyped using an Axiom 35K SNP array, to develop a high-resolution linkage map, and 49 QTL for sodium ion (NAX) and potassium ion (KC) status under salinity stress were mapped (Hussain et al., 2017a). Two NAX QTL on chromosome 2A coincided with a reported major HKT1 (Nax1/HKT1) QTL (Genc et al., 2010), and two NAX QTL on chromosome 7A which contributed 11.2% and 18.8%, respectively, to phenotypic variation. QTL for KC were located on chromosomes 2A, 3D, 4B, and 6A, whereas a novel Zn QTL on chromosome 7A controlled 11.2% of variation for salt tolerance. The most important Ca2+ (chromosome 6B) and Mg2+ (2A) QTL contributed 11.9% and 8.4%, respectively, to salt tolerance. Furthermore, QTL for Cu, Mn, B, P, S, and Fe were mapped for the first time (Hussain et al., 2017a). Based on SNP and expression analyses, Nax1/HKT1, K+ outward-rectifying channel (SKOR), potassium transporter 12 (KUP12), chloride channel protein (CLC-e), transparent testa 12, glutathione S-transferase U6 (GSTU6), peroxidase 12, auxin transport (BIG), auxin-response factor 5-like, ARF21, NAC78, Mg transporter (NIPA4), and Zn transporter 6 were identified as candidate genes (Hussain et al., 2017a). In another study, QTL for shoot growth, NAX and KC were mapped on chromosome 2B, 5A, 6A, and 7A in a wheat DH population. QTL were associated with several candidate genes. They included salt overly sensitive 1 (SOS1) or sodium/hydrogen exchanger 7 (NHX7), potassium transporter 1 (KUP1), HKT2, pyrophosphate-energized proton pump (H+ pyrophosphatase), KUP12, SKOR and two-pore potassium (TPK) channel and proton pump H+-ATPase 4 (Asif et al., 2018). Importantly, both of these studies (Hussain et al., 2017a; Asif et al., 2018) identified major QTL for KC and NAX on chromosomes 6A and 7A, respectively.

GWAS has been used to identify novel MTA and candidate genes for salinity tolerance (Oyiga et al., 2018; Chaurasia et al., 2020; Yu et al., 2020; Hu et al., 2021). The genotyping of 150 wheat accessions using the 90K SNP array, and a subsequent GWAS, identified 187 SNP and 37 QTL for leaf NAX and KC, including four QTL on chromosomes 2A, 3A and the short arm of chromosome 1B, in addition to novel QTL on the short arm of chromosome 1B and the long arm of 1D (Oyiga et al., 2018). Transcriptomics analyses revealed missense mutations responsible for salt tolerance variations in candidate genes, including ZIP7, SAP8, HAK18, GST1, SWEET17, and KeFC. Chaurasia et al. (2020) reported 42 MTA for shoot fresh/dry weight, chlorophyll content, seedling biomass, K+ and Na+ concentration that contributed 2.4%–42.8% to phenotype variation. These genomic regions were associated with 58 candidate genes, including transparent testa 12, chloroplast iron-superoxide dismutase, serine/threonine protein kinase (Nek6), ethylene responsive transcription factor (ERF3), bHLH30, and GDP-mannose transporter (GONST1).

Furthermore, haplotype analyses have also been coupled with GWAS to identify allelic variation for salt tolerance in wheat (Yu et al., 2020; Hu et al., 2021). A significant number of MTA (102 of 117) for germination and salt tolerance indices were found on chromosomes 1A, 3B, and 6B. They were associated with 53 candidate genes, including abscisic acid-insensitive 5-like (TaABI5-like), DUF674 family protein, SUMO-activating enzyme subunit 1A, glutamate formiminotransferase 1, protein KRI1, and NADH-cytochrome b5 reductase. Haplotype variations and expression of candidate genes under salinity have also been validated (Yu et al., 2020). Another GWAS analysis for yield and related traits under salinity found genomic regions and haplotypes for adult-stage salt tolerance on chromosomes 1B, 3B, 4A, 4D, 5A, 5B, an 7A. The markers were linked to several candidate genes, serine/threonine protein kinases, ricin B-like lectin gene, phytochelatin synthase, MADA-box genes, glycerol-3-phosphate acyltransferase (GPAT), U-box E3 ubiquitin ligase, and lipid-transfer proteins (TaLTPs) [73]. Additionally, 19 QTL for Mg2+ and Ca2+ were mapped at the same location as of Cl-QTL (Genc et al., 2014), with potential candidate genes chloride channel (CLC) and cation chloride co-transporter (CCC).

Frost Tolerance

An important limiting factor for wheat production in North America, North and Eastern Europe, and Russia is low temperature (Babben et al., 2018). As polar regions become more unstable due to climate change, the risk of extreme weather events including freezing temperatures increases (Francis and Skific, 2015). Therefore, resilience to frost is an important crop trait to consider. Frost tolerance is a complex biological process, involving pathways encompassing a large number of genes. The main pathway is frost response and a prolonged period of low temperature (vernalization), which can be regarded as an avoidance mechanism to prevent frost damage to sensitive reproductive organs. Two major frost tolerance loci, Frost Resistance 1 (FR1) and FR2, were identified on the long arm of chromosome 5A (Vágújfalvi et al., 2003). Zhao et al. (2013a) described an additional frost tolerance QTL on chromosome 5B in wheat germplasm from central Europe. During the last decade, several QTL associated with frost tolerance were identified on different wheat chromosomes (i.e., 1A, 1D, 2A, 2B, 3A, 5A, 5B, 6A, 6B, 6D, and 7B; Case et al., 2014; Kruse et al., 2017). The majority of genes assumed to be involved in frost tolerance have been identified on chromosome 5 (Vágújfalvi et al., 2003; Zhao et al., 2013a; Würschum et al., 2017; Babben et al., 2018; see also Supplementary S5). Until recently, only a few studies reported the identification of QTL regions associated with frost tolerance by GWAS (Babben et al., 2018; Kandel et al., 2018). Babben et al. (2018) demonstrated the utilization of the IWGSC RefSeq v1.0 in the specific primer development for highly conserved gene families in wheat. It showed that a candidate-gene association genetics approach is a useful tool for identifying new alleles of genes important for response to flowering time. Sequence analyses concluded that C-repeat binding factors (CBF)-A3, 5, 10, 13, 14, 15, and 18, vernalization response genes (VRN-A1, VRN-B3) and photoperiod response genes (PPD-B1 and PPD-D1) were associated with frost tolerance in wheat. In addition to winter hardiness as described above, an additional and critical frost-resistance phenotype relates to damage caused by transient frosts that generally occur on early spring mornings. While this trait is more complex to study, the above genes described for winter hardiness would be expected to contribute to this tolerance as moderators.

During the last decade, several components encompassing messenger molecules, protein kinases, and phosphatases, as well as transcription factors, which are involved in cold-stress signaling pathways, have been reported by studies using wheat sequence information (Wei and Han, 2017; Babben et al., 2018; Jin et al., 2018; Guo et al., 2019). The CBF (C-repeat binding factor), Inducer of CBF Expression (ICE) and cold-responsive (COR) genes or ICE-CBF-COR are part of the main cold-signaling pathway, playing a major role in controlling frost tolerance for crop species (Jin et al., 2018; Guo et al., 2019). ICE genes belong to the MYC family transcription factor and MYC subfamily of bHLH (Basic Helix–Loop–Helix; Guo et al., 2019). ICE factors are known as positive CBF expression regulators, considered to act upstream of the low-temperature signaling pathway. Two ICE homologs such as TaICE41 and TaICE87 have been identified in wheat (Guo et al., 2019). The TaICE41, TaICE87, and five MYC-like bHLHs were positively regulated upstream of the CBF mediated transcriptional cascade, controlling cold tolerance in wheat (Wei and Han, 2017).

Cloning of Multiple Disease and Insect-Resistance Genes

Examples of the positional or map-based cloning of disease resistance-related genes in wheat are arguably more common than for non-disease resistance traits—presumably due to the gene-for-gene interaction of such major resistance genes with specific avirulent factors in the pathogen, or other major effects (Supplementary S6). These include genes such as Lr21 (Huang et al., 2003), Yr36 (Fu et al., 2009), Yr15 (Klymiuk et al., 2018), YrU1 (Wang et al., 2020b), Fhb1 (Su et al., 2019), Fhb7 (Wang et al., 2020a), SuSr-D1 (Hiebert et al., 2020), Yr7, and Yr5/YrSP (Marchal et al., 2018). They have been cloned in wheat, to improve resistance against leaf and yellow rusts, in addition to fusarium head-blight diseases (Supplementary Table S5). Additionally, various genes were cloned for resistance against stem rust (Sr22, Sr33, Sr35, Sr45, Sr46, Sr60; Periyannan et al., 2013; Saintenac et al., 2013; Steuernagel et al., 2016; Arora et al., 2019; Chen et al., 2020), necrotrophic blotches (Tsn1, Snn1, Stb6, Stb6q; Faris et al., 2010; Shi et al., 2016; Saintenac et al., 2018, 2021), powdery mildew (Pm1a, Pm2, Pm3, Pm3b, Pm5e, Pm21, Pm24, Pm38/Lr34/Yr18/Sr57, Pm41, Pm46/Lr67/Yr46/Sr55, Pm60; Yahiaoui et al., 2004; Krattinger et al., 2009; Hurni et al., 2013; Moore et al., 2015; Sánchez-Martín et al., 2016; Xing et al., 2018; Zou et al., 2018; Li et al., 2020; Lu et al., 2020; Xie et al., 2020), being detailed in Supplementary Table S5. Access to a high-quality genome reference of wheat (IWGSC RefSeq v1.0) has also enabled researchers to explore susceptibility factors that may contribute to the onset of disease. For instance, Henningsen et al. (2021) explored the upregulation of wheat genes with orthology to known susceptibility factors in other plant species, in response to stem-rust fungus. They hypothesized genes that may play a conserved role in susceptibility. Similarly, a study by Corredor-Moreno et al. (2021) combined gene expression analyses and the genome reference of wheat (IWGSC RefSeq v1.0) and provided the basis for showing that the branched-chain amino acid aminotransferase gene (TaBCAT1) contributed to susceptibility to both stripe and stem rust. The authors from both studies suggested that manipulation of susceptibility genes can result in novel strategies to control disease. Fine-mapping and leveraging of available wheat pan-genome datasets, together with TILLING resources, have been utilized to analyze these complex situations. For example, a possible gene driving the complex interactions underlying Sm1-mediated resistance to orange wheat-blossom midge (OWBM, Sitodiplosis mosellana Géhin) wheat insect pest has been identified as a canonical NLR, with kinase and major sperm protein integrated domains (Walkowiak et al., 2020).

GBS-based GWAS identified 27 MTA for powdery mildew (7 MTAs), stem rust (5 MTAs), septoria (3 MTAs) and leaf rust (12 MTAs) resistance on all chromosomes (except for 4B and 5D; Bhatta et al., 2019). MTA were associated with several candidate genes for leaf rust; namely GDSL esterase/lipase, vesicle-associated 1-1-like protein, E3 ubiquitin ligase family protein, phosphatidic acid phosphatase, 12-oxophytodienoate reductase-like protein, septoria (F-box/RNI-like/FBD-like domains-containing protein) and stem rust (zinc transporter, putative). Leaf and stem-rust candidate genes associated with MTA were members of the NLR (nucleotide-binding domain leucine-rich repeat) gene family, nuclear monodehydroascorbate reductase 6 (MDAR6), solanesyl-diphosphate synthase 1 (DSDS1), enhancer of AG-4 protein 2 (AG4), phosphatase 2C (PP2C), and importin-9 (IPT9), being listed in Supplementary Table S5.

Advances in wheat genomics have facilitated the cloning of nine stripe-rust resistance genes (Yr5, Yr7, YrSP, Yr15, Yr18/Lr34, Yr36, Yr46, YrAS2388, and YrU1), out of the >80 genes that have been identified and mapped so far in different genetic backgrounds of wheat (Wang et al., 2020b). Cloning of the broad-spectrum stripe-rust resistance (R)-gene (Yr15), derived from wild emmer wheat, has led to the discovery of a novel protein family, the tandem kinase-pseudokinases (TKP) that emerged as a new class of disease-resistance protein family (TKP), providing plant innate immunity, being present not only in wheat, but also across the whole plant kingdom (Klymiuk et al., 2018). Five plant-disease resistance genes have been identified so far to contain a structure with tandem kinase domains, including three wheat genes, i.e., the wheat-stripe-rust R-gene WTK1 (Yr15; Klymiuk et al., 2018, 2020), wheat-stem-rust R-gene WTK2 (Sr60; Chen et al., 2020), and wheat powdery-mildew [Blumeria graminis f. sp. tritici (Bgt)] R-gene WTK3 (Pm24; Lu et al., 2020). More than 20 WTK copies have been found to be scattered across the three wheat genomes, AA, BB, and DD, including the orthologous group in chr 1 and the paralogous groups on chromosome 6 (Klymiuk et al., 2019). WTK1 orthologs, paralogs, and homologs were found also in the diploid wheat relatives, Triticum urartu (AA), Aegilops speltoides (SS), and A. tauschii (DD), representing the ancestral A, B, and D genomes, respectively, as well as in rye (Secale cereale), barley (Hordeum vulgare), and other cereal species. The protein sequences of TKP were obtained from genome assemblies of wild and cultivated wheat species, being used for phylogenetic analyses (Klymiuk et al., 2018), because it is important for successful deployment of R-genes in wheat breeding programs to identify if a cloned gene differed from other genes localized in the same chromosome region, or may represent different alleles of the same gene. For example, it was found that Yr15-, YrG303-, and YrH52-mediated resistances to yellow rust are encoded by a WTK1 as a single locus (Klymiuk et al., 2020). In future, we expect that many more such cases will be revealed, narrowing down the list of designated R-genes in wheat.

Ten powdery mildew (Pm) genes have been cloned so far, alongside with advances in wheat genomics resources (Supplementary Table S5). Pm3/Pm8, Pm2, Pm21, Pm60, Pm5e, Pm41, and Pm1a encode NLR-immune receptors from different wheat relatives (Yahiaoui et al., 2004; Hurni et al., 2013; Sánchez-Martín et al., 2016; Xing et al., 2018; Zou et al., 2018; Li et al., 2020; Xie et al., 2020), while a tandem kinase protein is encoded by Pm24 (Lu et al., 2020). Furthermore, two non-NLR genes—Pm38 and Pm46, showed broad-spectrum multi-adult plant resistance to powdery mildew and rust diseases. An ABC transporter is encoded by the Lr34/Yr18/Sr57/Pm38 multi-resistance gene (Krattinger et al., 2009), and a hexose transporter is encoded by the Lr67/Yr46/Sr55/Pm46 multi-resistance gene (Moore et al., 2015). The cloning of these Pm genes enables the development of high-throughput diagnostic functional markers that can be used in MAS for fungi-resistance breeding programs (Supplementary Table S5). Some of these Pm genes have been widely used for the protection of wheat cultivars for many years. For example, Triticeae grass Dasypyrum villosum (2n = 2x = 14, VV) harbors Pm21, which confers broad-spectrum resistance, and was transferred in China into wheat cultivars (T6AL.6VS wheat-D. villosum translocation line) in 1995 (Chen et al., 2013). Some of these NLR proteins could be overcome by the fast evolution of virulent Blumeria graminis (Bgt) isolates, especially when the gene is widely deployed in wheat fields. For example, wheat-rye 1BL·1RS translocation carrying Pm8 has lost the resistance function for wheat variety production (Zeng et al., 2014). Different alleles of those cloned Pm genes that might be resistant to different Bgt isolates have been identified and could also be used for MAS. For example, 17 alleles of the Pm3 gene have been identified mediating resistance to distinct race spectra of Bgt. Pm3a has a range of resistance that fully encompasses that of Pm3f, but also extends to additional races (Brunner et al., 2010). Therefore, enriching the Pm gene pools is very important for resistance breeding. Pm24 is a rare natural allele of tandem kinase protein (TKP), with putative kinase-pseudokinase domains, conferring broad-spectrum resistance to wheat powdery-mildew disease. However, there are some other Pm genes that have not been cloned yet, such as Pm30, found in ~80% of Chinese cultivars, as detected by closely linked markers (Cheng et al., 2020). The absence of functional molecular markers is limiting the diagnosis of potential Pm alleles, and their deployment in wheat breeding, via MAS and genome editing.

Two independent GWAS analyses utilizing iSelect 9K and 90K Illumina arrays have reported SNP and genes for Soil-Borne Wheat Mosaic Virus (SBWMV) resistance (Liu et al., 2014, 2020). Liu et al. (2020) completed a GWAS analysis of SBWMV resistance using the 90K Illumina array. Thirty-five SNP in 12 wheat genes and one intergenic SNP in the Sbwm1 region were identified on chromosome 5D, being associated significantly with SBWMV resistance. Resistance to SBWMV was strongly associated with a predicted kinase family protein (Liu et al., 2014). Furthermore, GWAS analyses identified major resistance SNP for Wheat Spindle Streak Mosaic Virus (WSSMV) on chromosome 2D, in addition to regions on 5B and 7D. The 2D genomic region was linked with 18 candidate genes, including 11 NBS-LRR ones (Hourcade et al., 2019), being listed in Supplementary Table S6.

Insect resistance (Supplementary Table S7) has been explored, utilizing the wheat stem sawfly (WSS) transcriptome and its interaction with the regulatory elements, microRNA<--abbreviation indicated above-- and long non-coding RNA (lncRNA). Interestingly, the study found that WSS miRNA may target wheat transcripts and vice versa, thereby potentially modulating plant responses against WSS (Cagirici et al., 2017). The solid-stem trait, associated with WSS resistance, was linked to copy-number variation of a putative Dof Transcription Factor (TdDof) within the 3BL QTL, through the use of high-throughput sequencing in different genetic backgrounds. Transgenic lines over-expressing TdDof firmly established that increased expression of TdDof was responsible for solid stemness, likely through regulation of programmed cell death in pith parenchyma cells (Nilsen et al., 2020). Similarly, genome sequencing in resistant and susceptible cultivars revealed a candidate gene in the Sm1 locus that is known to confer resistance to orange wheat blossom midge (OWBM). This time, knockout mutant lines demonstrated that mutations within this gene resulted in susceptibility against OWBM. The candidate gene contains NB-ARC, and LRR motifs, in addition to a serine/threonine (S/T) kinase that is similar to those found in rust resistance proteins, and a major sperm protein (MSP) domain (Walkowiak et al., 2020).

End-Use Quality Traits

Wheat grain markets and food industries demand not only high yielding and resistant varieties, but also those with specific end-use qualities. End-use quality is, therefore, an important focus in breeding programs. Methods, for testing quality, however, require large amounts of grain and are time-consuming and costly. Significant efforts have been made to identify QTL linked to various end-use quality traits such as grain protein content (GPC), dough rheological properties, and baking quality (Supplementary Table S7). Several comprehensive analyses (Sun et al., 2008; Raman et al., 2009; Carter et al., 2012; Simons et al., 2012) of mapping several quality traits related to protein and starch have been conducted. Sun et al. (2008) analyzed GPC, flour protein content (FPC), grain glutenin macropolymer content, wet gluten content (WGC), dry gluten content (DGC), Zeleny sedimentation volume, flour–water absorption (FWA), dough development time (DDT), and mixing tolerance index and flour paste viscosity (Supplementary Table S7). They identified 30 QTL for starch traits and 15 QTL for protein traits, with QTL clusters for starch traits located on chromosomes 3D, 6B, and 7B, and protein traits on 1D and 3B. Raman et al. (2009) analyzed GPC, milling yield, FPC, flour color, FWA, DDT, dough strength (DS), and dough extensibility (DE). They found several QTL associated with DS, DE, DDT, and FWA, close to glutenin (Glu-B1) locus on chromosome 1B. Simons et al. (2012) analyzed 20 end-use quality traits including six grain, seven milling and flour, four dough mixing strength, and three bread-making traits. They found that the 1DL QTL cluster containing Glu-D1 had a large genetic influence on dough mixing strength and bread-making performance. Furthermore, two QTL clusters located on chromosomes 3B and 4D associated with several milling and baking quality traits (Carter et al., 2012) were reported, being associated with the Wx-B1, Glu-B1 and Glu-D1 genes.

Specific attributes of starch contribute unique properties to certain wheat-breeding lines and the genome level characterization of one such property, udon noodle quality, is detailed in (Appels et al., 2018). The Granule Bound Starch Synthase (GBSS; TraesCS4A01G418200) gene on chromosome 4A is absent from some lines, being the null allele for GBSS-4AL (Wx-B1b) that associates with udon noodle quality. Significant sections of TraesCS4A01G418200 were absent from the exome sequences of 3.9% lines of a set of 644 (hexaploid) wheat varieties and landraces, assessed using 10 SNP identified from snapshot exome sequence data (Appels et al., 2018). The specific deletions within the GBSS-4AL gene mean that the respective lines provide new germplasm sources for wheat breeding. The gene-specific deletions would not be expected to show detrimental effects, due to the deletion of adjoining gene models, and thus be expected to perform successfully at the agronomic level, to satisfy the high-value commercial udon-noodle market. Other attributes of starch, such as a high amylose content for an improved source of fiber in the diet, can now be introduced into commercial wheat lines.

Importantly, a metaQTL analysis (Quraishi et al., 2017) identified stable QTL, by combining 27 quantitative genetic studies with four genetic maps. It located 73 and 82 QTL for baking quality and GPC-related traits, respectively, on a consensus map. They reported 8 metaQTL for baking quality and 6 for GPC. The most precise metaQTL having the smallest confidence intervals were located on chromosomes 3D (3.78 cM) for baking quality and chromosome 2B (5.83 cM) for GPC. The candidate genes identified are listed in Supplementary Table S7.

Recently, high-density SNP arrays and GBS have also been utilized to identify QTL for bread-making quality, using biparental populations (Jin et al., 2016; Boehm et al., 2017; Guo et al., 2020b). These high-resolution genetic maps helped to precisely identify major QTL and candidate genes, thus providing a valuable resource for MAS and genomic selection in wheat. Guo et al. (2020b) used GBS1.0 DArT arrays and 90K iSelect SNP array to map QTL for protein and starch-related traits. The authors reported 26 stable QTL, for GPC, sedimentation volume, DDT, DST, FWA, flour viscosity, break down, and peak time, detected in more than two environments on chromosomes 1A, 1B, 1D, 4B, 5D, 6A, 6B, 6D, 7A, and 7D. These genomic regions were linked to several candidate genes, including embryonic flower 1 (EMF1), trehalose-6-phosphate synthase 6 (TPS), nitrate transporter 1:2 (NRT1; 2 transporters), zinc-finger protein 830 (ZNF830), phospholipid-transporting ATPase, transcription factor (TFIIIB), acylamino-acid-releasing enzyme (APEH), F-box protein (SKIP22), and aldehyde dehydrogenase (ALDH). Similarly, Jin et al. (2016) used the 90K iSelect and the 660K SNP arrays to identify QTL for dough rheology and starch-pasting properties (Jin et al., 2016). Subsequently, 119 additive QTL were mapped on 20 chromosomes (i.e., all except 4D), including 55 and 17 novel QTL for mixolab parameters and 17 for Rapid Visco-Analyzer parameters. SNP markers in these regions were located on eight candidate genes, involved in biosynthesis of fatty acids and amino acids, in addition to starch and sucrose metabolism, i.e., anthranilate phosphoribosyltransferase (AnPRT), 3-ketoacyl-CoA, ornithine aminotransferase, lipoxygenase 2 (LOX2), sucrose-phosphate synthase II (SPS2), lysosomal beta glucosidase, and 5′-methylthioadenosine/S-adenosylhomocysteine nucleosidase (mtnN; Jin et al., 2016). Using the power of GBS, Boehm et al. (2017) identified co-localizing QTL for multiple end-use quality traits (GPC, FWA, and flour yield) on chromosomes 1B, 2D, 7A, and 7B, including allelic variation for the glutenin genes Glu-A1, Glu-B1, Glu-A3, Glu-B3, and Glu-D3.

Current and Potential Methods to Identify and Clone Genes in Wheat

The availability of reference genome in wheat and subsequent construction of several high-density genetic maps developed from the sequence-tagged SNPs (see section Characterization of Genes and Gene Families Using the Wheat Reference-Genome) has opened new opportunities for map-based cloning of the genes. Therefore, here we have discussed the presently used and potential gene cloning methods in bread wheat. Dissecting the genetic and molecular mechanisms regulating grain yield and growth is the key for positional cloning or map-based cloning, as well as wheat breeding and improvement. Traditional forward genetic tools have been widely used to clone genes regulating traits of interest in wheat, e.g., VRN1 (Yan et al., 2003), Gpc-B1 (Uauy et al., 2006), and Lr21 (Huang et al., 2003). Map-based gene cloning, however, usually needs multiple steps such as generating mapping populations, fine mapping to narrow the target region to identify genetic markers co-segregating with the phenotype, screening candidate gene(s) and gene(s) identification by sequencing. This process often requires more time and is labor intensive, especially in wheat. Hence, a limited number of positional cloning studies have been successfully undertaken. Furthermore, BREEDWHEAT program10 summarized 8-year efforts in which novel source of genetic diversity identified and introduced to elite materials to generate superior varieties.

Bulked segregant analysis (BSA) was recommended as a shortcut to identify the linkage of molecular markers with phenotype, being extensively used to map loci that have major effects (Michelmore et al., 1991). In this analysis, DNA of each individual showing extreme phenotype in a segregating population (i.e., F2) are bulked and genotyped, including their parents, with molecular markers (Chhetri et al., 2017). Any marker is considered to be linked with the studied trait if it shows the same allele in the bulk and parent of a similar phenotype. Recently, with the great advances of NGS technologies, several BSA-based modifications have been developed to identify major-effect QTL, regulating quantitative traits. These modifications are based on whole-genome resequencing bulks in a large population, reducing cost of genotyping, time spent, and increasing statistical power of analyses (Zou et al., 2016; Chhetri et al., 2017).

In crops with large genomes such as wheat, complexity reduction is very important to identify and clone target genes more quickly and efficiently. QTL-seq is one such approach that incorporates the potential of BSA. The power of high-throughput whole-genome resequencing to identify genomic regions showed contrasting results of an SNP index in two bulk populations (each with 20–50 individuals), featuring extreme phenotypes (Takagi et al., 2013a). Recently, QTL-seq was used in bread wheat to identify the candidate genomic region tightly linked to the awn inhibitor loci. Diagnostic markers were designed to understand the role of QTL in the awnless trait formation (Wang et al., 2021). Moreover, this approach was applied to identify loci involved in tiller angle in bread wheat, which represents an important factor influencing yield. Also, in this case, functional markers for MAS were developed and validated (Zhao et al., 2020). Multiple QTL-seq (mQTL-seq) used for several mapping populations from crosses with at least one common parent (Das et al., 2015). The use of multiple mapping populations with a broad genetic diversity was critical for the validation of QTL, along with narrowing down the detected QTL. To date, however, this technique has not yet been used in wheat.

Where the cost of whole-genome resequencing becomes prohibitive, bulk segregant RNA sequencing (BSR-seq) can be an alternative strategy for identifying expression QTL (eQTL) regions, generating data of gene expression for genomic loci of interest (Trick et al., 2012). Differential expression of genes in two bulks can be used to identify candidate genes responsible for favorable phenotypes. BSR-seq has been successfully used for mapping of stripe-rust-resistant loci YrMM58 and YrHY1 on chromosome 2AS (Wang et al., 2018b), Yr15 on chromosome 1BS (Ramirez-Gonzalez et al., 2015), and leaf senescence gene (els1) on chromosome 2BS (Li et al., 2018a), in segregating wheat biparental populations. Similarly, BSR-seq enabled fine-mapping of a locus controlling grain-protein content (GPC) in wheat (GPC-B1) to 0.4 cM from the previously reported interval of 30 cM (Trick et al., 2012). This study pinpointed 13–18 candidate genes for GPC in wheat.

NGS platforms have also accelerated the identification and cloning of genes in mutant collections. TEnSeq pipelines are examples of advances that have allowed for rapid gene cloning identification, as recently reviewed (Zhang et al., 2020). Mutagenesis chromosome flow-sorting and short-read sequencing (MutChromSeq) is a recently developed tool (Sánchez-Martín et al., 2016; Hiebert et al., 2020), based on mutagenesis followed by flow sorting of chromosomes and their subsequent sequencing, to identify induced mutations. This rapid approach was successfully described to clone the powdery-mildew resistance locus Pm2 in wheat (Sánchez-Martín et al., 2016). MutChromSeq has the advantage that it does not rely on an assumption that the resistance gene belongs to the NLR class (as for other approaches, see below). Hence, it would be appropriate for identification of non-immune-mediated resistance genes. The most recent application is the cloning of the Med15 protein encoded by SuSr-D1, a suppressor gene of stem-rust resistance (Hiebert et al., 2020) from the wheat cultivar “Canthatch.”

A similar approach to MutChromSeq, which does not require mutagenesis, is the target chromosome-based cloning (TACCA) method. It uses flow-sorted chromosomes, next-generation sequencing, and cultivar-specific de novo assembly. Using this approach, Lr22a, broad-spectrum leaf-rust resistance locus was cloned in wheat. Two SSR markers flanking Lr22a, covering 0.48 cM interval on chromosome 2D, were previously mapped. Sorting chromosome 2D, followed by sequencing and identification of genes, was performed within 4 months (Thind et al., 2017).

Other cloning strategies such as MutMap (mutational mapping) involve mutagenesis, sequencing, and mapping, to identify SNP between wild-type and homozygous mutants, and then zero in on the region containing the gene of interest. Although this approach was initially considered to be applicable only in crops with small genomes, it was successfully utilized to map and clone Ms1 from bread wheat, using F2 plants derived from heterozygous ms1e mutants (Wang et al., 2017). MutMap will be less efficient at identifying the causal mutation, however, if the wild-type reference genome has gaps at the position of the causal mutation. Thus, de novo assembly of the wild-type genome is used in MutMap-Gap (Takagi et al., 2013b), and could be applied to wheat in the future.

Finally, MutRenSeq is a fast gene-cloning tool for the isolation of nucleotide-binding and leucine-rich repeat (NLR) genes (Steuernagel et al., 2016). Chemical mutagenesis, exome capture, and sequencing are required. Most genes associated with disease-resistance encode proteins in the NLR family. Hence, exome capture is necessary to enrich the NLR-specific bait library. Then, resistant wild-type parent and susceptible loss-of-function mutants are sequenced, as a last step. Mutant reads are aligned with the equivalent wild-type pool of the NLR gene family from the parents. This method has been utilized to clone two fungal stem-rust resistance genes (Sr22 and Sr45) and three yellow-rust genes (Yr7, Yr5/YrSP) from bread wheat (Steuernagel et al., 2016; Marchal et al., 2018). This method does not require fine positional mapping and can be applied to isolate NLR-type resistance genes from most crops and their wild relatives. Nevertheless, two major limitations must be taken into account. Firstly, the design of oligonucleotide baits is based on a reference genome sequence. Considering the large-scale presence/absence variations among different accessions, the recent release of pan-genome in wheat is the ideal reference on which to design baits (Walkowiak et al., 2020). Secondly, this approach is limited to isolating only resistance genes, encoding NLR proteins. Therefore, genes that do not belong to the NLR family are missed (Dinh et al., 2020). However, it is possible to add capture baits targeting other classes of genes, thought a priori to be involved in disease resistance, such as wall-associated kinases.

Unlike map-based cloning and MutRenSeq, the association genetics with the R-gene enrichment sequencing (AgRenSeq) method has been developed to align with GWAS platforms (to utilize genome-wide natural variation). Thereby, RenSeq eliminates the need for biparental mapping populations or mutagenesis. This approach was demonstrated successfully to clone the, Sr46 R-gene and to identify the candidate-gene sequence for SrTA1662, using a diverse panel of A. tauschii ssp. strangulata (Arora et al., 2019). It explores a pool of diverse wild relatives, carrying many resistance genes. As a result, it enables cloning of multiple genes at the same time (Dinh et al., 2020). These strategies have many advantages over traditional marker-based mapping. In addition to taking much less time, they find genes or functional nucleotides/haplotypes, responsible for a given agronomic trait. In several cases, genes were rapidly cloned and diagnostic markers were developed. The benefits of cloning genes in wheat, particularly those with a role in disease resistance, have been recently shown (Luo et al., 2021). As demonstrated by Luo et al. (2021), construction of transgene cassettes can simplify breeding bottlenecks, associated with the deployment of multiple genes to be inherited as a single unit. In this case study, a gene cassette containing five previously cloned R-genes (Sr45, Sr50, Sr55, Sr22, and Sr35) provides high levels of resistance to stem rust, suggesting that this could be a viable solution to confer durable multi-pathogen resistance.

Genomic Selection in Wheat to Improve Complex Traits

The relatively recent availability of large numbers of genome-wide molecular markers in wheat genetic resources has led to the application of an alternative marker-assisted approach for wheat genetic improvement, namely genomic selection (Meuwissen et al., 2001; Jannink et al., 2010). For a long time, the lack of high-density markers were a major hindrance to carrying out in-depth genetic and genomic analyses. GS is an advanced form of MAS, wherein genome-wide markers are used to calculate genomic-estimated breeding values (GEBV; Meuwissen et al., 2001). Rather than explicitly identifying and tracking markers associated with genetic loci controlling a given trait, GS aims to use large numbers of genome-wide markers, in conjunction with phenotypic data collected in a collection of lines/varieties (termed the “training set”), to establish parameters that allow forward selection of the progenies derived from the training set over multiple forward generations, in the absence of additional phenotyping (Meuwissen et al., 2001; Jannink et al., 2010). This potentially allows selection to be applied faster and at higher intensities, as more lines can be incorporated for advancing to subsequent generations, without the need for time-consuming phenotyping steps. Additionally, advancement in statistical and bioinformatics methods to deal with high-density marker data for genomic selection has been equally important for plant breeders for the development of GS in wheat. GS is a valuable and attractive plant breeding approach that provides an idea for the conversion of genotypic value to phenotypic value (Nakaya and Isobe, 2012). Although this approach has great potential, plant breeders must carefully consider relationship between training and breeding populations. Additionally, it is important that breeders should investigate all traits of interest during the training phase in order to exclude phenotyping during the breeding cycle.

Since its first use in 2006 (Crossa et al., 2006), GS has been extensively used in wheat, for a wide range of traits with different architectures including grain yield (Crossa et al., 2016; Rutkoski et al., 2016; Saint Pierre et al., 2016; Guo et al., 2020a; Sehgal et al., 2020b), resistances to different diseases as rusts, fusarium head blight, Stagonospora nodorum blotch, Septoria tritici blotch, and tan-spot resistance (Jiang et al., 2017; Juliana et al., 2017a,b), macro- and micro-nutrients (Manickavelu et al., 2017), as well as end-use quality traits (Battenfield et al., 2016; Kristensen et al., 2018). Application of GS for hybrid prediction has also been investigated in wheat (Zhao et al., 2013b; Adhikari et al., 2020), pointing to challenges when predicting hybrids derived from untested parents (Zhao et al., 2015). Using more refined Genotype x Environment (GxE) interaction-based GS models promises to partly reduce this shortcoming (Basnet et al., 2019). The optimization of underlying factors on which genomic selection relies, such as marker density, predictive models, training population size, and the relationship between training and validation population sets, is ongoing (Larkin et al., 2019).

Recent investigations have focused on optimization of GS in genetic resources. In order to harness new diversity from wheat gene banks. Crossa et al. (2016) investigated GS models to predict days to heading and days to maturity, on a large set of wheat landrace accessions (8,416 Mexican landrace accessions and 2,403 Iranian landrace accessions) from CIMMYT gene bank, using two strategies. The first one involved random cross-validation of the data in 20% training (TRN) and 80% testing (TST; TRN20-TST80). In the second strategy, two types of core sets called “diversity” and “prediction,” including 10 and 20%, respectively, of total collections were used. Prediction accuracy of the 20% diversity core set was close to accuracies obtained for 20% training and 80% testing set (0.412–0.654 and 0.182–0.647 for Mexican landraces and Iranian landraces, respectively). For traits controlled by a mix of a few major and many minor genes, it can thus be beneficial to include preexisting knowledge on known candidate genes, to increase accuracy of genome-wide predictions (Bernardo, 2014). The potential of such an approach has been demonstrated when predicting flowering time and plant height for wheat (Zhao et al., 2014). These results suggested a way forward for parental selection in pre-breeding, by predicting the value of all genotyped accessions in a gene bank, followed by pre-breeding programs, based on those genotypes that have the highest predicted value, or harbor promising novel candidate genes or alleles. Once promising parents are identified, efficient pre-breeding programs need to be designed. This is a non-trivial task, mainly depending on diversity of plant genetic resources. Some of the latter have diversity that makes them directly useful as a source of parents for breeding new varieties; however, this tends to be the exception. It is evident that genetic resources must therefore be improved, through appropriate pre-breeding programs, to the point where they can be used as productive parents in breeding programs. Two complementary approaches can be used: (i) pre-breeding of populations created from genetic resources and (ii) pre-breeding of populations created from crosses between plant genetic resources and elite materials. In both cases, genomic prediction enables rapid selection gain, but requires the presence of extensive training populations, related to the base population. Crossing parents selected from agronomic and physiological screening of genetic resources, and advancing through generations, using high throughput phenotyping of physiological parameters, is another approach (Reynolds and Langridge, 2016). Such strategy is complementary to genomic selection of progeny, since many complex physiological traits that may have been used in strategic crossing do not lend themselves to high-throughput progeny screening (Reynolds and Langridge, 2016).

Substantial efforts have shifted to development of high-throughput phenotyping platforms in wheat, which are used to measure different traits, including plant height (Holman et al., 2016), disease resistance (Devadas et al., 2015), growth rate (Holman et al., 2016), and nitrogen deficiency (Devadas et al., 2015). These significant advancements in high-throughput phenotyping have brought a paradigm shift in breeding strategies. Wheat scientists have incorporated high-throughput phenotyping data in GS models, to explore their potential in improving prediction accuracies for complex traits (Rutkoski et al., 2016; Crain et al., 2018). Rutkoski et al. (2016) investigated the role of canopy temperature and green and red normalized difference vegetation index (NDVI), measuring chlorophyll concentration, canopy leaf-area and yield, as secondary traits in GS models, for improving prediction accuracy for grain yield. The authors observed 67% improvement in prediction accuracy, without correcting for days to heading (DTH), and 37% improvement upon correction with DTH. Crain et al. (2018) used over 1.1 million phenotypic data points generated by high-throughput phenotyping on 1,170 advanced CIMMYT lines in drought and heat stress environments, observing an increase in prediction accuracy from 7% to 33%, as compared to the standard univariate model.

Genomic selection has revolutionized animal breeding and will likely be a major source of genetic improvement of crops, including wheat, over the coming decade (Mackay et al., 2021). We envisage that incorporation of additional datatypes and technologies into GS pipelines will open opportunities for further gains to be made. For example, increasing precision of phenotypic characterization in training set via high-throughput phenotyping platforms (Devadas et al., 2015; Holman et al., 2016; Rutkoski et al., 2016; Crain et al., 2018), as well as incorporation of environmental covariates (de los Campos et al., 2020), may lead to improved prediction accuracies. Similarly, incorporating additional molecular or “omics” data may further refine prediction equations. These include: (i) molecular data tagging functionally validated alleles, whether they are natural variants, or novel ones, generated via technologies such as gene editing or TILLING (Krasileva et al., 2017; Hussain et al., 2018) and (ii) integration of transcriptomic and metabolomic data with molecular markers, as has been reported in maize (Schrag et al., 2018). Ultimately, combining these approaches with methods for shortening of wheat generation times will further increase selection intensity. Currently, “speed breeding” methodology, whereby plants are grown under extended photoperiods via the use of supplementary lighting, allows spring wheat generation time to be reduced from four to around 2 months (Watson et al., 2018). Combining speed breeding with “speed vernalization” methods can even more shortened the breeding process. Any further dramatic shortening of cycling times would require development of new approaches, such as generation of recombinant individuals, by in vitro production of gametes and their subsequent fusions (De La Fuente et al., 2013). Combining such in vitro generation-cycling with genomic selection methodologies may well represent an achievable medium-term step-change in genomics-informed breeding.

CRISPR/Cas9-Mediated Genome Editing for Wheat Improvement

Availability of complete genome assemblies of diverse wheat genotypes, originating from different parts of the world, is essential to identify and characterize functions of various wheat genes, for different growth stages and environmental conditions, at the whole-genome level. Furthermore, transcriptomic analyses help to identify genes and gene networks regulating traits in different conditions. Therefore, knockout, knock-in, or activation of such genes through CRISPR/Cas9 gene-editing system provides unique opportunities for wheat genetic improvement (Scheben et al., 2017; Scheben and Edwards, 2017; Hussain et al., 2018). Because of the very large genome size of such species, orthologous gene copies present in the polyploid genome, and the presence of many repetitive sequences, genetic manipulation through CRISPR/Cas9-mediated gene editing system is more challenging in bread and durum wheat, as compared to cereals with smaller genomes. Using wheat-cell suspension cultures which led to InDel mutations, an attempt was made to use the CRISPR/Cas9 method for specific gene modification, in wheat inositol oxygenase (inox) and phytoene desaturase (pds) genes (Upadhyay et al., 2013). The first successful application of CRISPR/Cas9 to generate wheat knockout lines having three homoeoalleles of powdery-mildew resistance locus O gene (TaMLO), by a transient protoplast expression system, was done independently by Shan et al. (2014) and Wang et al. (2014). The latter successfully applied the CRISPR/Cas9 system in bread wheat, for the generation of plants mutated in a single TaMLO-A1 allele, with increased resistance to powdery mildew. A similar strategy has been used to knockout drought-responsive transcription factors in wheat, like dehydration-responsive element-binding protein 2 (TaDREB2) and ethylene-responsive factor 3 (TaERF3), for improved drought signaling (Hussain et al., 2018; Kim et al., 2018).

The knockout of all three homoeoalleles of TaGW2 through the CRISPR/Cas9 system increased the thousand kernel weight (TKW) and seed size (Wang et al., 2018a), implying the utility of the system for crop improvement. Such CRISPR-generated lines can either be exploited as new varieties or used as germplasm. Moreover, recent advances in editing allow simultaneous multiple-gene targeting or genome multiplexing (Xie et al., 2015), opening new horizons for employment of CRISPR/Cas9 in polyploid wheat, carrying many homoeologous and paralogous copies of the same gene, such as α-gliadins. In another study, CRISPR/Cas9 mediated mutations in 35 out of 45 α-gliadin genes, genes controlling gluten content in wheat, generated transgene-free, low-gluten wheat without any off-target mutations (Sánchez-León et al., 2018). Using this genome multiplexing by CRISPR/Cas9, three genes, viz., TaGW2 (grain traits negative regulator), TaMLO (resistance to powdery mildew), and TaLpx-1 (lipoxygenase; offers resistance to Fusarium graminearum), were targeted (Wang et al., 2018a). The first application of zinc-finger nuclease (ZFN)-mediated, non-homologous end joining (NHEJ)-directed loss-of-function gene knockout of acetohydroxyacid synthase (AHAS) in allohexaploid bread wheat through a supplied DNA repair template resulted in resistance to imidazolinone herbicides, due to an amino acid change in the target-gene coding sequence (Ran et al., 2018). Efficient and novel ribonucleoprotein-based (RNP) CRISPR/Cas9 genome editing procedures that required only 7–9 weeks were developed, with no off-target mutations and no transgene integration, implying the efficiency of the system (Liang et al., 2017).

The CRISPR/Cas9 genome editing methodology is also important for pre-breeding, namely, to reduce time to transfer beneficial alleles, increasing success rate. The idea is to directly induce/modify the alleles to beneficial ones in elite wheat germplasm both efficiently and quickly. CRISPR/Cas9-mediated permanent genome integration results in a stable expression of CRISPR/Cas9. However, RNP-based biolistic delivery offers a transient expression of CRISPR/Cas9, and its rapid degradation, which controls off-target mutations (Liang et al., 2017). Thus, RNP-based gene editing has been successfully applied for gene-editing in bread wheat (Liang et al., 2018). Similarly, knockout of three homologs of wheat enhanced disease resistance 1 (TaEDR1), a negative regulator of defense response against powdery mildew, conferred resistance against powdery mildew, without any off-target mutations (Zhang et al., 2017). DNA-virus [e.g., Geminivirus, i.e., wheat dwarf virus (WDV)]-based amplicons were later identified as an efficient construct-delivery method for gene editing, with an enhanced CRISPR/Cas9 expression, as compared to ubiquitin reference gene, proposing that it could be a potential tool for CRISPR-mediated genome editing in wheat (Gil-Humanes et al., 2017). Moreover, CRISPR/Cas9 has been successfully applied by generating heritable, targeted mutations, in wheat male-sterility 1 gene (Ms1), responsible for complete male sterility in commercial wheat cultivars, like Gladius and Fielder (Okada et al., 2019), thus speeding up hybrid-wheat production. These studies demonstrate the utility of the CRISPR/Cas9 system for rapid generation of male sterility in commercial wheat cultivars, for breeding programs. Although CRISPR-mediated genome or gene editing was demonstrated to be successful, its widespread implementation still encounters difficulties, involved in low regeneration efficiency of crops, such as wheat. Recently, growth-regulating factor-grf-interacting factor (GRF-GIF) wheat transformation system has become the game changer by using GRF-GIF chimeric protein construct, which improves regeneration efficiency up to 100% (Debernardi et al., 2020) in the transgenic-wheat lines.

Again, such CRISPR-generated lines can be released either as a variety or can be used as germplasm stocks. Although the utility of this revolutionary technology for crop improvement was demonstrated, regulatory approvals for the use of gene-edited plants still vary among different countries (Jouanin et al., 2018). CRISPR can also be utilized for testing the effect of a mutated allele on the resulting phenotype. If regulations are too strict, like in the European Union, this could be used later on, to search for such alleles in natural populations of wheat progenitors and germplasm stored in gene banks (“natural variation”). Recent initiatives to sequence thousands of gene bank accessions (Milner et al., 2019) can help to facilitate this approach. An additional benefit of using CRISPR/Cas9 is that genome editing for the first time allows direct transfer of favorable alleles into elite breeding material, without typical linkage drag, associated with cross-breeding. Wheat is crossed with maize to induce haploids, and colchicine is applied to get doubled haploid plants that serve as breeding material, or could be introduced as a variety, thus speeding up wheat breeding (Devaux and Cistué, 2016). Taking advantage of the well-established wheat × maize crossing system, maize pollens carrying gRNA for plant height genes (BRI1 and SD1) were crossed to wheat, for inducing site-directed targeted mutagenesis in wheat (Budhagatapalli et al., 2020), without the need of segregating out the transgene. It helped in reducing the genotype-dependent site-directed mutagenesis. This approach can also be used for introducing mutations in multiple genes with one cas9/gRNA-transgenic (pollinator) plant, thus providing an opportunity for multiplex gene-editing in wheat. An important part of genetic analyses is identification of candidate genes and/or diagnostic marker(s) in linkage equilibrium to the trait(s) of selection interest. Many studies for abiotic stresses lack identification of candidate genes, slowing down MAS in wheat.

Concluding Comments

Reviewing the very extensive genome-level analyses undertaken since publication of the CS reference genome sequence has identified the broad importance of considering the network nature of grain yield, heat and drought susceptibility indices and yield stability coefficients (Sehgal et al., 2017). Likewise, their association with flowering time genes (Vrn-B1, Ppd-D1 and Vrn-D3). Observations are consistent with the long history of wheat improvement through breeding at the phenotypic level, and genome-level analyses can now complement this existing knowledge, through refining biological networks and fine-tuning germplasm to micro-environments and defined wavelength environments of LED lighting in speed breeding.

Typically, candidate genes are identified by locating the QTL region on the genome assembly and analyzing the genes residing in the region. If only a few genes are selected to validate their expression under a certain condition, the possibility of human bias (Baxter, 2020) can come into play, slowing down progress of genetic advancement. For example, looking for K+/Na+ transporters for salinity studies could lead to ignoring important genes in tolerance mechanisms. Complementing genomic selection with in silico transcriptomic analyses for all potential genes of interest would cast a wider net to capture a more complete set of genes that are relevant to the phenotype under study or selection.

The role of synthetic wheat in imparting stress tolerance is well known in wheat. Wheat gene banks harbor hundreds of such synthetic wheat lines. In the post reference genome era, extensive genotyping efforts have been undertaken to genotype entire gene bank accessions, generating the so-called “digital gene banks.” For example, CIMMYT has generated GBS data on ~100K accessions stored in its gene bank, in order to bridge the gap between genetic resources and breeding pipelines. Although success has been achieved in quantifying genome contributions of wild germplasm (synthetics and landraces) to the current elite germplasm, it is still unknown how to devise a genome-based strategy to deploy favorable introgressions from synthetic wheat, to enhance breeding value. Development of new genomic selection and machine-learning models and tools will be required to predict the best exotics from gene banks, without having to invest in laborious and costly multi-environmental field testing. These technologies will refine the already-successful breeder pipelines, for establishing new varieties.

Complex metabolic engineering can be exploited to improve cereals like wheat, to produce essential polyunsaturated fatty acids (PUFA), including the healthy omega 3 (ω-3) and omega 6 (ω-6). Genetic engineering and synthetic biology tools can be used to reach such a goal (Kraic et al., 2018). Similarly, these technologies can be used to generate Marker-Free and Transgene Insertion site-Defined (MFTID) transgenic plants. Thus, the lipoxygenase (LOX) gene expression was repressed using an RNA interference (RNAi) cassette reduce lipid peroxidation (improving storability) and increase nutrient quality, such as the amount of healthy fatty acids (e.g., containing linoleic and linolenic residues) of wheat seeds (Cao et al., 2020).

Cereals can also be improved so that their agricultural waste (plant cell wall, being mainly made of cellulose, hemicellulose, and lignin) contains more cellulose and less lignin, thereby allowing its use as feedstock for biofuel or bioproduct production. Currently, high costs and low yields are associated with their use, due to the molecular structure of the natural lignocellulosic biomass. As expected from an evolutionary point of view, it is hard to enzymatically hydrolyze cellulose into glucose, as it is resistant to most microorganism degradation (Rocha-Meneses et al., 2020).

In short, these developments hold interesting potential applications for wheat improvement, within molecular breeding programs, in addition to enhancing yield traits, and biotic- and abiotic-stress tolerance, which are particularly relevant in the present scenario of global warming and climate change.

Author Contributions

HB conceived, designed, and organized the study. All authors listed have made a substantial contribution to the work, drafted, edited, and approved it for publication.

Funding

HB was funded by USDA-NIFA SBIRI and SBIRII. PH and SG were funded by project P18-RT-992 from Junta de Andalucía (Andalusian Regional Government), Spain (Co-funded by FEDER). JC was funded by BBSRC grant BB/P010741/1. MF is supported by the 2Blades Foundation and Grains Research and Development Corporation (project CSP1801-013RTX/9176010). VK is supported by the Ministry of Science and Higher Education of the Russian Federation (grant no. 075-15-2019-1881). GH was supported by funding of the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Germany’s Excellence Strategy—EXC-2048/1—project ID 390686111 and grants 426557363 and 458717903, the European Regional Development Fund (Project ID ZS/2018/06/93171), and the Czech Science Foundation (CZ.02.1.01./0.0/0.0/16_019/0000827, SPP 813103381). BK thanks the Government of Norway (QZA-14/0005) for funding the initiative of “Adapting Agriculture to Climate Change: Collecting, Protecting and Preparing Crop Wild Relatives” (https://www.cwrdiversity.org/project/pre-breeding/).

Conflict of Interest

HB and BA are employed by Montana BioAg Inc., VK is employed by KWS, TU is employed by Ficus Biotechnology, and PD is employed by Florimond Desprez Group.

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

Publisher’s Note

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

Supplementary Material

The Supplementary Material for this article can be found online at: https://www.frontiersin.org/articles/10.3389/fpls.2022.851079/full#supplementary-material

Footnotes

References

Addison, C. K., Mason, R. E., Brown-Guedira, G., Guedira, M., Hao, Y., Miller, R. G., et al. (2016). QTL and major genes influencing grain yield potential in soft red winter wheat adapted to the southern United States. Euphytica 209, 665–677. doi: 10.1007/s10681-016-1650-1

CrossRef Full Text | Google Scholar

Adhikari, A., Basnet, B. R., Crossa, J., Dreisigacker, S., Camarillo, F., Bhati, P. K., et al. (2020). Genome-wide association mapping and genomic prediction of anther extrusion in CIMMYT hybrid wheat breeding program via modeling pedigree, genomic relationship, and interaction with the environment. Front. Genet. 11:586687. doi: 10.3389/fgene.2020.586687

PubMed Abstract | CrossRef Full Text | Google Scholar

Agarwal, P., Baranwal, V. K., and Khurana, P. (2019). Genome-wide analysis of bZIP transcription factors in wheat and functional characterization of a TabZIP under abiotic stress. Sci. Rep. 9, 1–18. doi: 10.1038/s41598-019-40659-7

CrossRef Full Text | Google Scholar

Akpinar, B. A., Lucas, S. J., Vrána, J., Doležel, J., and Budak, H. (2015). Sequencing chromosome 5D of Aegilops tauschii and comparison with its allopolyploid descendant bread wheat (Triticum aestivum). Plant Biotechnol. J. 13, 740–752. doi: 10.1111/pbi.12302

PubMed Abstract | CrossRef Full Text | Google Scholar

Alaux, M., Rogers, J., Letellier, T., Flores, R., Alfama, F., Pommier, C., et al. (2018). Linking the international wheat genome sequencing consortium bread wheat reference genome sequence to wheat genetic and phenomic data. Genome Biol. 19:111. doi: 10.1186/s13059-018-1491-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Appels, R., Eversole, K., Feuillet, C., Keller, B., Rogers, J., Stein, N., et al. (2018). Shifting the limits in wheat research and breeding using a fully annotated reference genome. Science 361:eaar7191. doi: 10.1126/science.aar7191

PubMed Abstract | CrossRef Full Text | Google Scholar

Arora, S., Steuernagel, B., Gaurav, K., Chandramohan, S., Long, Y., Matny, O., et al. (2019). Resistance gene cloning from a wild crop relative by sequence capture and association genetics. Nat. Biotechnol. 37, 139–143. doi: 10.1038/s41587-018-0007-9

PubMed Abstract | CrossRef Full Text | Google Scholar

Asif, M. A., Schilling, R. K., Tilbrook, J., Brien, C., Dowling, K., Rabie, H., et al. (2018). Mapping of novel salt tolerance QTL in an Excalibur × kukri doubled haploid wheat population. Theor. Appl. Genet. 131, 2179–2196. doi: 10.1007/s00122-018-3146-y

PubMed Abstract | CrossRef Full Text | Google Scholar

Assanga, S. O., Fuentealba, M., Zhang, G., Tan, C. T., Dhakal, S., Rudd, J. C., et al. (2017). Mapping of quantitative trait loci for grain yield and its components in a US popular winter wheat TAM 111 using 90K SNPs. PLoS One 12:e0189669. doi: 10.1371/journal.pone.0189669

PubMed Abstract | CrossRef Full Text | Google Scholar

Babben, S., Schliephake, E., Janitza, P., Berner, T., Keilwagen, J., Koch, M., et al. (2018). Association genetics studies on frost tolerance in wheat (Triticum aestivum L.) reveal new highly conserved amino acid substituations in CBF-A3, CBF-A15, VRN3 and PPD1 genes. BMC Genomics 19:409. doi: 10.1186/s12864-018-4795-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Badawi, M. A., Agharbaoui, Z., Zayed, M., Li, Q., Byrns, B., Zou, J., et al. (2019). Genome-wide identification and characterization of the wheat remorin (ta REM) family during cold acclimation. Plant Genome 12:180040. doi: 10.3835/plantgenome2018.06.0040

PubMed Abstract | CrossRef Full Text | Google Scholar

Banerjee, B. P., Joshi, S., Thoday-Kennedy, E., Pasam, R. K., Tibbits, J., Hayden, M., et al. (2020). High-throughput phenotyping using digital and hyperspectral imaging-derived biomarkers for genotypic nitrogen response. J. Exp. Bot. 71, 4604–4615. doi: 10.1093/jxb/eraa143

PubMed Abstract | CrossRef Full Text | Google Scholar

Basnet, B. R., Crossa, J., Dreisigacker, S., Pérez-Rodríguez, P., Manes, Y., Singh, R. P., et al. (2019). Hybrid wheat prediction using genomic, pedigree, and environmental covariables interaction models. Plant Genome 12:180051. doi: 10.3835/plantgenome2018.07.0051

PubMed Abstract | CrossRef Full Text | Google Scholar

Battenfield, S. D., Guzmán, C., Gaynor, R. C., Singh, R. P., Peña, R. J., Dreisigacker, S., et al. (2016). Genomic selection for processing and end-use quality traits in the CIMMYT spring bread wheat breeding program. Plant Genome 9, 1–12. doi: 10.3835/plantgenome2016.01.0005

PubMed Abstract | CrossRef Full Text | Google Scholar

Baxter, I. (2020). We aren’t good at picking candidate genes, and it’s slowing us down. Curr. Opin. Plant Biol. 54, 57–60. doi: 10.1016/j.pbi.2020.01.006

PubMed Abstract | CrossRef Full Text | Google Scholar

Bayer, P. E., Golicz, A. A., Scheben, A., Batley, J., and Edwards, D. (2020). Plant pan-genomes are the new reference. Nat. Plants 6, 914–920. doi: 10.1038/s41477-020-0733-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Berkman, P. J., Skarshewski, A., Lorenc, M. T., Lai, K., Duran, C., Ling, E. Y. S., et al. (2011). Sequencing and assembly of low copy and genic regions of isolated Triticum aestivum chromosome arm 7DS. Plant Biotechnol. J. 9, 768–775. doi: 10.1111/j.1467-7652.2010.00587.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Berkman, P. J., Skarshewski, A., Manoli, S., Lorenc, M. T., Stiller, J., Smits, L., et al. (2012). Sequencing wheat chromosome arm 7BS delimits the 7BS/4AL translocation and reveals homoeologous gene conservation. Theor. Appl. Genet. 124, 423–432. doi: 10.1007/s00122-011-1717-2

PubMed Abstract | CrossRef Full Text | Google Scholar

Berkman, P. J., Visendi, P., Lee, H. C., Stiller, J., Manoli, S., Lorenc, M. T., et al. (2013). Dispersion and domestication shaped the genome of bread wheat. Plant Biotechnol. J. 11, 564–571. doi: 10.1111/pbi.12044

PubMed Abstract | CrossRef Full Text | Google Scholar

Bernardo, R. (2014). Genomewide selection when major genes are known. Crop. Sci. 54, 68–75. doi: 10.2135/cropsci2013.05.0315

PubMed Abstract | CrossRef Full Text | Google Scholar

Bhatta, M., Shamanin, V., Shepelev, S., Stephen Baenziger, P., Pozherukova, V., Pototskaya, I., et al. (2019). Marker-trait associations for enhancing agronomic performance, disease resistance, and grain quality in synthetic and bread wheat accessions in Western Siberia. G3 9, 4209–4222. doi: 10.1534/g3.119.400811

CrossRef Full Text | Google Scholar

Boehm, J. D., Ibba, M. I., Kiszonas, A. M., See, D. R., Skinner, D. Z., and Morris, C. F. (2017). Identification of genotyping-by-sequencing sequence tags associated with milling performance and end-use quality traits in hard red spring wheat (Triticum aestivum L.). J. Cereal Sci. 77, 73–83. doi: 10.1016/j.jcs.2017.07.007

CrossRef Full Text | Google Scholar

Brunner, S., Hurni, S., Streckeisen, P., Mayr, G., Albrecht, M., Yahiaoui, N., et al. (2010). Intragenic allele pyramiding combines different specificities of wheat Pm3 resistance alleles. Plant J. 64, 433–445. doi: 10.1111/j.1365-313X.2010.04342.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Budak, H., Hussain, B., Khan, Z., Ozturk, N. Z., and Ullah, N. (2015). From genetics to functional genomics: improvement in drought signaling and tolerance in wheat. Front. Plant Sci. 6, 1–13. doi: 10.3389/fpls.2015.01012

CrossRef Full Text | Google Scholar

Budhagatapalli, N., Halbach, T., Hiekel, S., Büchner, H., Müller, A. E., and Kumlehn, J. (2020). Site-directed mutagenesis in bread and durum wheat via pollination by cas9/guide RNA-transgenic maize used as haploidy inducer. Plant Biotechnol. J. 18, 2376–2378. doi: 10.1111/pbi.13415

PubMed Abstract | CrossRef Full Text | Google Scholar

Cagirici, H. B., Biyiklioglu, S., and Budak, H. (2017). Assembly and annotation of transcriptome provided evidence of miRNA mobility between wheat and wheat stem sawfly. Front. Plant Sci. 8:1653. doi: 10.3389/fpls.2017.01653

PubMed Abstract | CrossRef Full Text | Google Scholar

Cao, X., Dong, Z., Tian, D., Dong, L., Qian, W., Liu, J., et al. (2020). Development and characterization of marker-free and transgene insertion site-defined transgenic wheat with improved grain storability and fatty acid content. Plant Biotechnol. J. 18, 129–140. doi: 10.1111/pbi.13178

CrossRef Full Text | Google Scholar

Carter, A. H., Garland-Campbell, K., Morris, C. F., and Kidwell, K. K. (2012). Chromosomes 3B and 4D are associated with several milling and baking quality traits in a soft white spring wheat (Triticum aestivum L.) population. Theor. Appl. Genet. 124, 1079–1096. doi: 10.1007/s00122-011-1770-x

PubMed Abstract | CrossRef Full Text | Google Scholar

Case, A. J., Skinner, D. Z., Garland-Campbell, K. A., and Carter, A. H. (2014). Freezing tolerance-associated quantitative trait loci in the brundage × coda wheat recombinant inbred line population. Crop. Sci. 54, 982–992. doi: 10.2135/cropsci2013.08.0526

CrossRef Full Text | Google Scholar

Chaurasia, S., Singh, A. K., Songachan, L. S., Sharma, A. D., Bhardwaj, R., and Singh, K. (2020). Multi-locus genome-wide association studies reveal novel genomic regions associated with vegetative stage salt tolerance in bread wheat (Triticum aestivum L.). Genomics 112, 4608–4621. doi: 10.1016/j.ygeno.2020.08.006

PubMed Abstract | CrossRef Full Text | Google Scholar

Chen, S., Rouse, M. N., Zhang, W., Zhang, X., Guo, Y., Briggs, J., et al. (2020). Wheat gene Sr60 encodes a protein with two putative kinase domains that confers resistance to stem rust. New Phytol. 225, 948–959. doi: 10.1111/nph.16169

PubMed Abstract | CrossRef Full Text | Google Scholar

Chen, P., You, C., Hu, Y., Chen, S., Zhou, B., Cao, A., et al. (2013). Radiation-induced translocations with reduced Haynaldia villosa chromatin at the Pm21 locus for powdery mildew resistance in wheat. Mol. Breed. 31, 477–484. doi: 10.1007/s11032-012-9804-x

CrossRef Full Text | Google Scholar

Cheng, B., Ding, Y. Q., Gao, X., Cao, N., Xin, Z. H., and Zhang, L. Y. (2020). The diversity of powdery mildew resistance gene loci among wheat germplasm in Southwest China. Cereal Res. Commun. 48, 65–70. doi: 10.1007/s42976-020-00015-2

CrossRef Full Text | Google Scholar

Chhetri, M., Bariana, H., Wong, D., Sohail, Y., Hayden, M., and Bansal, U. (2017). Development of robust molecular markers for marker-assisted selection of leaf rust resistance gene Lr23 in common and durum wheat breeding programs. Mol. Breed. 37, 1–8. doi: 10.1007/s11032-017-0628-6

CrossRef Full Text | Google Scholar

Chu, C., Wang, S., Paetzold, L., Wang, Z., Hui, K., Rudd, J. C., et al. (2021). RNA-seq analysis reveals different drought tolerance mechanisms in two broadly adapted wheat cultivars ‘TAM 111’ and ‘TAM 112’. Sci. Rep. 11:4301. doi: 10.1038/s41598-021-83372-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Corredor-Moreno, P., Minter, F., Davey, P. E., Wegel, E., Kular, B., Brett, P., et al. (2021). The branched-chain amino acid aminotransferase TaBCAT1 modulates amino acid metabolism and positively regulates wheat rust susceptibility. Plant Cell 33, 1728–1747. doi: 10.1093/plcell/koab049

PubMed Abstract | CrossRef Full Text | Google Scholar

Corsi, B., Obinu, L., Zanella, C. M., Cutrupi, S., Day, R., Geyer, M., et al. (2021). Identification of eight QTL controlling multiple yield components in a German multi-parental wheat population, including Rht24, WAPO-A1, WAPO-B1 and genetic loci on chromosomes 5A and 6A. Theor. Appl. Genet. 134, 1435–1454. doi: 10.1007/s00122-021-03781-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Crain, J., Mondal, S., Rutkoski, J., Singh, R. P., and Poland, J. (2018). Combining high-throughput phenotyping and genomic information to increase prediction and selection accuracy in wheat breeding. Plant Genome 11:170043. doi: 10.3835/plantgenome2017.05.0043

PubMed Abstract | CrossRef Full Text | Google Scholar

Crossa, J., Burgueño, J., Cornelius, P. L., McLaren, G., Trethowan, R., and Krishnamachari, A. (2006). Modeling genotype × environment interaction using additive genetic covariances of relatives for predicting breeding values of wheat genotypes. Crop. Sci. 46, 1722–1733. doi: 10.2135/cropsci2005.11-0427

CrossRef Full Text | Google Scholar

Crossa, J., Jarquín, D., Franco, J., Pérez-Rodríguez, P., Burgueño, J., Saint-Pierre, C., et al. (2016). Genomic prediction of gene bank wheat landraces. G3 6, 1819–1834. doi: 10.1534/g3.116.029637

CrossRef Full Text | Google Scholar

Cui, F., Zhang, N., Fan, X. L., Zhang, W., Zhao, C. H., Yang, L. J., et al. (2017). Utilization of a Wheat660K SNP array-derived high-density genetic map for high-resolution mapping of a major QTL for kernel number. Sci. Rep. 7, 1–12. doi: 10.1038/s41598-017-04028-6

CrossRef Full Text | Google Scholar

Danilevicz, M. F., Tay Fernandez, C. G., Marsh, J. I., Bayer, P. E., and Edwards, D. (2020). Plant pangenomics: approaches, applications and advancements. Curr. Opin. Plant Biol. 54, 18–25. doi: 10.1016/j.pbi.2019.12.005

PubMed Abstract | CrossRef Full Text | Google Scholar

Das, S., Singh, M., Srivastava, R., Bajaj, D., Saxena, M. S., Rana, J. C., et al. (2015). MQTL-seq delineates functionally relevant candidate gene harbouring a major QTL regulating pod number in chickpea. DNA Res. 23, 53–65. doi: 10.1093/dnares/dsv036

CrossRef Full Text | Google Scholar

De La Fuente, G. N., Frei, U. K., and Lübberstedt, T. (2013). Accelerating plant breeding. Trends Plant Sci. 18, 667–672. doi: 10.1016/j.tplants.2013.09.001

PubMed Abstract | CrossRef Full Text | Google Scholar

de los Campos, G., Pérez-Rodríguez, P., Bogard, M., Gouache, D., and Crossa, J. (2020). A data-driven simulation platform to predict cultivars’ performances under uncertain weather conditions. Nat. Commun. 11, 1–10. doi: 10.1038/s41467-020-18480-y

CrossRef Full Text | Google Scholar

Debernardi, J. M., Tricoli, D. M., Ercoli, M. F., Hayta, S., Ronald, P., Palatnik, J. F., et al. (2020). A GRF–GIF chimeric protein improves the regeneration efficiency of transgenic plants. Nat. Biotechnol. 38, 1274–1279. doi: 10.1038/s41587-020-0703-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Devadas, R., Lamb, D. W., Backhouse, D., and Simpfendorfer, S. (2015). Sequential application of hyperspectral indices for delineation of stripe rust infection and nitrogen deficiency in wheat. Precis. Agric. 16, 477–491. doi: 10.1007/s11119-015-9390-0

CrossRef Full Text | Google Scholar

Devaux, P., and Cistué, L. (2016). “Wheat doubled haploids : production to sequencing - what makes them so appealing ? What is haplodiploidisation ? Doubled haploid production in wheat,” in The World Wheat Book, A History of Wheat Breeding. eds. A. Bonjean, W. J. Angus, and V. M. Ginkel (Paris, France: Lavoisier Tec & doc), 885–938.

Google Scholar

Dinh, H. X., Singh, D., Periyannan, S., Park, R. F., and Pourkheirandish, M. (2020). Molecular genetics of leaf rust resistance in wheat and barley. Theor. Appl. Genet. 133, 2035–2050. doi: 10.1007/s00122-020-03570-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Dong, L., Wang, F., Liu, T., Dong, Z., Li, A., Jing, R., et al. (2014). Natural variation of TaGASR7-A1 affects grain length in common wheat under multiple cultivation conditions. Mol. Breed. 34, 937–947. doi: 10.1007/s11032-014-0087-2

CrossRef Full Text | Google Scholar

Dorado, G., Gálvez, S., Budak, H., Unver, T., and Hernández, P. (2019). “Nucleic-acid sequencing,” in Encyclopedia of Biomedical Engineering. ed. M. Caplan (Amsterdam: Elsevier), 443–460.

Google Scholar

Edwards, D., Wilcox, S., Barrero, R. A., Fleury, D., Cavanagh, C. R., Forrest, K. L., et al. (2012). Bread matters: a national initiative to profile the genetic diversity of Australian wheat. Plant Biotechnol. J. 10, 703–708. doi: 10.1111/j.1467-7652.2012.00717.x

PubMed Abstract | CrossRef Full Text | Google Scholar

El Beji, I. H. S., Mouzeyar, S., Bouzidi, M. F., and Roche, J. (2019). Expansion and functional diversification of SKP1-like genes in wheat (Triticum aestivum L.). Int. J. Mol. Sci. 20:3295. doi: 10.3390/ijms20133295

PubMed Abstract | CrossRef Full Text | Google Scholar

ElBasyoni, I., Saadalla, M., Baenziger, S., Bockelman, H., and Morsy, S. (2017). Cell membrane stability and association mapping for drought and heat tolerance in a worldwide wheat collection. Sustainability 9:1606. doi: 10.3390/su9091606

CrossRef Full Text | Google Scholar

Fang, Z.-W., He, Y.-Q., Liu, Y.-K., Jiang, W.-Q., Song, J.-H., Wang, S. P., et al. (2020a). Bioinformatic identification and analyses of the non-specific lipid transfer proteins in wheat. J. Integr. Agric. 19, 1170–1185. doi: 10.1016/S2095-3119(19)62776-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Fang, Z., Jiang, W., He, Y., Ma, D., Liu, Y., Wang, S., et al. (2020b). Genome-wide identification, structure characterization, and expression profiling of Dof transcription factor gene family in wheat (Triticum aestivum L.). Agronomy 10:294. doi: 10.3390/agronomy10020294

CrossRef Full Text | Google Scholar

Faris, J. D., Zhang, Z., Lu, H., Lu, S., Reddy, L., Cloutier, S., et al. (2010). A unique wheat disease resistance-like gene governs effector-triggered susceptibility to necrotrophic pathogens. Proc. Natl. Acad. Sci. 107, 13544–13549. doi: 10.1073/pnas.1004090107

PubMed Abstract | CrossRef Full Text | Google Scholar

Francis, J., and Skific, N. (2015). Evidence linking rapid Arctic warming to mid-latitude weather patterns. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 373, 1–12. doi: 10.1098/rsta.2014.0170

PubMed Abstract | CrossRef Full Text | Google Scholar

Fu, D., Uauy, C., Distelfeld, A., Blechl, A., Epstein, L., Chen, X., et al. (2009). A kinase-START gene confers temperature-dependent resistance to wheat stripe rust. Science 323, 1357–1360. doi: 10.1126/science.1166289

PubMed Abstract | CrossRef Full Text | Google Scholar

Gaggar, P., Kumar, M., and Mukhopadhyay, K. (2020). Genome-scale identification, in silico characterization and interaction study between wheat SNARE and NPSN gene families involved in vesicular transport. IEEE/ACM Trans. Comput. Biol. Bioinforma. 18, 2492–2501. doi: 10.1109/tcbb.2020.2981896

CrossRef Full Text | Google Scholar

Gálvez, S., Agostini, F., Caselli, J., Hernández, P., and Dorado, G. (2021). BLVector: fast BLAST-like algorithm for manycore CPU with vectorization. Front. Genet. 12:100. doi: 10.3389/fgene.2021.618659

PubMed Abstract | CrossRef Full Text | Google Scholar

Gálvez, S., Ferusic, A., Esteban, F. J., Hernández, P., Caballero, J. A., and Dorado, G. (2016). Speeding-up bioinformatics algorithms with heterogeneous architectures: highly heterogeneous smith-waterman (HHeterSW). J. Comput. Biol. 23, 801–809. doi: 10.1089/cmb.2015.0237

PubMed Abstract | CrossRef Full Text | Google Scholar

Gálvez, S., Mérida-García, R., Camino, C., Borrill, P., Abrouk, M., Ramírez-González, R. H., et al. (2019). Hotspots in the genomic architecture of field drought responses in wheat as breeding targets. Funct. Integr. Genomics 19, 295–309. doi: 10.1007/s10142-018-0639-3

PubMed Abstract | CrossRef Full Text | Google Scholar

Genc, Y., Oldach, K., Verbyla, A. P., Lott, G., Hassan, M., Tester, M., et al. (2010). Sodium exclusion QTL associated with improved seedling growth in bread wheat under salinity stress. Theor. Appl. Genet. 121, 877–894. doi: 10.1007/s00122-010-1357-y

PubMed Abstract | CrossRef Full Text | Google Scholar

Genc, Y., Taylor, J., Rongala, J., and Oldach, K. (2014). A major locus for chloride accumulation on chromosome 5A in bread wheat. PLoS One 9:e98845. doi: 10.1371/journal.pone.0098845

PubMed Abstract | CrossRef Full Text | Google Scholar

Gil-Humanes, J., Wang, Y., Liang, Z., Shan, Q., Ozuna, C. V., Sánchez-León, S., et al. (2017). High-efficiency gene targeting in hexaploid wheat using DNA replicons and CRISPR/Cas9. Plant J. 89, 1251–1262. doi: 10.1111/tpj.13446

PubMed Abstract | CrossRef Full Text | Google Scholar

Golan, G., Hendel, E., Méndez Espitia, G. E., Schwartz, N., and Peleg, Z. (2018). Activation of seminal root primordia during wheat domestication reveals underlying mechanisms of plant resilience. Plant Cell Environ. 41, 755–766. doi: 10.1111/pce.13138

PubMed Abstract | CrossRef Full Text | Google Scholar

Golicz, A. A., Bayer, P. E., Bhalla, P. L., Batley, J., and Edwards, D. (2020). Pangenomics comes of age: from bacteria to plant and animal applications. Trends Genet. 36, 132–145. doi: 10.1016/j.tig.2019.11.006

PubMed Abstract | CrossRef Full Text | Google Scholar

Guo, J., Khan, J., Pradhan, S., Shahi, D., Khan, N., Avci, M., et al. (2020a). Multi-trait genomic prediction of yield-related traits in US soft wheat under variable water regimes. Genes 11:1270. doi: 10.3390/genes11111270

CrossRef Full Text | Google Scholar

Guo, J., Ren, Y., Tang, Z., Shi, W., and Zhou, M. (2019). Characterization and expression profiling of the ICE-CBF-COR genes in wheat. PeerJ 2019, 1–19. doi: 10.7717/peerj.8190

CrossRef Full Text | Google Scholar

Guo, Y., Zhang, G., Guo, B., Qu, C., Zhang, M., Kong, F., et al. (2020b). QTL mapping for quality traits using a highdensity genetic map of wheat. PLoS One 15:e0230601. doi: 10.1371/journal.pone.0230601

CrossRef Full Text | Google Scholar

Henningsen, E. C., Omidvar, V., Coletta, R. D., Michno, J.-M., Gilbert, E., Li, F., et al. (2021). Identification of candidate susceptibility genes to Puccinia graminis f. sp. tritici in wheat. bioRxiv [Preprint], 2021.01.23.427871. doi:doi: 10.1101/2021.01.23.427871

CrossRef Full Text | Google Scholar

Hiebert, C. W., Moscou, M. J., Hewitt, T., Steuernagel, B., Hernández-Pinzón, I., Green, P., et al. (2020). Stem rust resistance in wheat is suppressed by a subunit of the mediator complex. Nat. Commun. 11, 1–10. doi: 10.1038/s41467-020-14937-2

CrossRef Full Text | Google Scholar

Holman, F. H., Riche, A. B., Michalski, A., Castle, M., Wooster, M. J., and Hawkesford, M. J. (2016). High throughput field phenotyping of wheat plant height and growth rate in field plot trials using UAV based remote sensing. Remote Sens. (Basel) 8:1031. doi: 10.3390/rs8121031

CrossRef Full Text | Google Scholar

Hourcade, D., Bogard, M., Bonnefoy, M., Savignard, F., Mohamadi, F., Lafarge, S., et al. (2019). Genome-wide association analysis of resistance to wheat spindle streak mosaic virus in bread wheat. Plant Pathol. 68, 609–616. doi: 10.1111/ppa.12972

CrossRef Full Text | Google Scholar

Hu, M. J., Zhang, H. P., Cao, J. J., Zhu, X. F., Wang, S. X., Jiang, H., et al. (2016a). Characterization of an IAA-glucose hydrolase gene TaTGW6 associated with grain weight in common wheat (Triticum aestivum L.). Mol. Breed. 36, 1–11. doi: 10.1007/s11032-016-0449-z

CrossRef Full Text | Google Scholar

Hu, M. J., Zhang, H. P., Liu, K., Cao, J. J., Wang, S. X., Jiang, H., et al. (2016b). Cloning and characterization of TaTGW-7A gene associated with grain weight in wheat via SLAF-seq-BSA. Front. Plant Sci. 7:1902. doi: 10.3389/fpls.2016.01902

CrossRef Full Text | Google Scholar

Hu, P., Zheng, Q., Luo, Q., Teng, W., Li, H., Li, B., et al. (2021). Genome-wide association study of yield and related traits in common wheat under salt-stress conditions. BMC Plant Biol. 21:27. doi: 10.1186/s12870-020-02799-1

PubMed Abstract | CrossRef Full Text | Google Scholar

Huang, L., Brooks, S. A., Li, W., Fellers, J. P., Trick, H. N., and Gill, B. S. (2003). Map-based cloning of leaf rust resistance gene Lr21 from the large and polyploid genome of bread wheat. Genetics 164, 655–664. doi: 10.1093/genetics/164.2.655

PubMed Abstract | CrossRef Full Text | Google Scholar

Hurni, S., Brunner, S., Buchmann, G., Herren, G., Jordan, T., Krukowski, P., et al. (2013). Rye Pm8 and wheat Pm3 are orthologous genes and show evolutionary conservation of resistance function against powdery mildew. Plant J. 76, 957–969. doi: 10.1111/tpj.12345

PubMed Abstract | CrossRef Full Text | Google Scholar

Hussain, B., Khan, A. S., and Ali, Z. (2015). Genetic variation in wheat germplasm for salinity tolerance at seedling stage: improved statistical inference. Turkish J. Agric. For. 39, 182–192. doi: 10.3906/tar-1404-114

CrossRef Full Text | Google Scholar

Hussain, B., Lucas, S. J., and Budak, H. (2018). CRISPR/Cas9 in plants: at play in the genome and at work for crop improvement. Brief. Funct. Genomics 17, 319–328. doi: 10.1093/bfgp/ely016

PubMed Abstract | CrossRef Full Text | Google Scholar

Hussain, B., Lucas, S. J., Ozturk, L., and Budak, H. (2017a). Mapping QTLs conferring salt tolerance and micronutrient concentrations at seedling stagein wheat. Sci. Rep. 7:15662. doi: 10.1038/s41598-017-15726-6

CrossRef Full Text | Google Scholar

Hussain, W., Stephen Baenziger, P., Belamkar, V., Guttieri, M. J., Venegas, J. P., Easterly, A., et al. (2017b). Genotyping-by-sequencing derived high-density linkage map and its application to QTL mapping of flag leaf traits in bread wheat. Sci. Rep. 7, 1–15. doi: 10.1038/s41598-017-16006-z

CrossRef Full Text | Google Scholar

Iehisa, J. C. M., Matsuura, T., Mori, I. C., Yokota, H., Kobayashi, F., and Takumi, S. (2014). Identification of quantitative trait loci for abscisic acid responsiveness in the D-genome of hexaploid wheat. J. Plant Physiol. 171, 830–841. doi: 10.1016/j.jplph.2014.02.003

PubMed Abstract | CrossRef Full Text | Google Scholar

IWGSC (2014). A chromosome-based draft sequence of the hexaploid bread wheat (Triticum aestivum) genome. Science 345:1251788. doi: 10.1126/science.1251788

CrossRef Full Text | Google Scholar

Jannink, J. L., Lorenz, A. J., and Iwata, H. (2010). Genomic selection in plant breeding: from theory to practice. Brief. Funct. Genomic. Proteomic. 9, 166–177. doi: 10.1093/bfgp/elq001

CrossRef Full Text | Google Scholar

Jiang, Y., Schulthess, A. W., Rodemann, B., Ling, J., Plieske, J., Kollers, S., et al. (2017). Validating the prediction accuracies of marker-assisted and genomic selection of fusarium head blight resistance in wheat using an independent sample. Theor. Appl. Genet. 130, 471–482. doi: 10.1007/s00122-016-2827-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Jiang, W., Yang, L., He, Y., Zhang, H., Li, W., Chen, H., et al. (2019). Genome-wide identification and transcriptional expression analysis of superoxide dismutase (SOD) family in wheat (Triticum aestivum). PeerJ 7:e8062. doi: 10.7717/peerj.8062

PubMed Abstract | CrossRef Full Text | Google Scholar

Jiang, W., Yin, J., Zhang, H., He, Y., Shuai, S., Chen, S., et al. (2020). Genome-wide identification, characterization analysis and expression profiling of auxin-responsive GH3 family genes in wheat (Triticum aestivum L.). Mol. Biol. Rep. 47, 3885–3907. doi: 10.1007/s11033-020-05477-5

PubMed Abstract | CrossRef Full Text | Google Scholar

Jin, J., Liu, D., Qi, Y., Ma, J., and Zhen, W. (2020). Major QTL for seven yield-related traits in common wheat (Triticum aestivum L.). Front. Genet. 11:1012. doi: 10.3389/fgene.2020.01012

PubMed Abstract | CrossRef Full Text | Google Scholar

Jin, H., Wen, W., Liu, J., Zhai, S., Zhang, Y., Yan, J., et al. (2016). Genome-wide QTL mapping for wheat processing quality parameters in a Gaocheng 8901/Zhoumai 16 recombinant inbred line population. Front. Plant Sci. 7:1032. doi: 10.3389/fpls.2016.01032

PubMed Abstract | CrossRef Full Text | Google Scholar

Jin, Y., Zhai, S., Wang, W., Ding, X., Guo, Z., Bai, L., et al. (2018). Identification of genes from the ICE–CBF–COR pathway under cold stress in Aegilops–Triticum composite group and the evolution analysis with those from Triticeae. Physiol. Mol. Biol. Plants 24, 211–229. doi: 10.1007/s12298-017-0495-y

PubMed Abstract | CrossRef Full Text | Google Scholar

Jouanin, A., Boyd, L., Visser, R. G. F., and Smulders, M. J. M. (2018). Development of wheat with hypoimmunogenic gluten obstructed by the gene editing policy in europe. Front. Plant Sci. 9:1523. doi: 10.3389/fpls.2018.01523

PubMed Abstract | CrossRef Full Text | Google Scholar

Juliana, P., Singh, R. P., Singh, P. K., Crossa, J., Huerta-Espino, J., Lan, C., et al. (2017a). Genomic and pedigree-based prediction for leaf, stem, and stripe rust resistance in wheat. Theor. Appl. Genet. 130, 1415–1430. doi: 10.1007/s00122-017-2897-1

CrossRef Full Text | Google Scholar

Juliana, P., Singh, R. P., Singh, P. K., Crossa, J., Rutkoski, J. E., Poland, J. A., et al. (2017b). Comparison of models and whole-genome profiling approaches for genomic-enabled prediction of Septoria tritici blotch, Stagonospora nodorum blotch, and tan spot resistance in wheat. Plant Genome 10, 1–16. doi: 10.3835/plantgenome2016.08.0082

CrossRef Full Text | Google Scholar

Kandel, J. S., Huang, M., Zhang, Z., Skinner, D. Z., and See, D. R. (2018). Genetic diversity of clinal freezing tolerance variation in winter wheat landraces. Agronomy 8:95. doi: 10.3390/agronomy8060095

CrossRef Full Text | Google Scholar

Kaur, P., Yadav, I. S., Yadav, B., Mahato, A., Gupta, O. P., Dolezel, J., et al. (2019). In silico annotation of 458 genes identified from comparative analysis of full length cDNAs and NextGen sequence of chromosome 2A of hexaploid wheat. J. Plant Biochem. Biotechnol. 28, 25–34. doi: 10.1007/s13562-018-0460-z

CrossRef Full Text | Google Scholar

Kelley, L. A., Mezulis, S., Yates, C. M., Wass, M. N., and Sternberg, M. J. E. (2015). The Phyre2 web portal for protein modeling, prediction and analysis. Nat. Protoc. 10, 845–858. doi: 10.1038/nprot.2015.053

PubMed Abstract | CrossRef Full Text | Google Scholar

Kim, D., Alptekin, B., and Budak, H. (2018). CRISPR/Cas9 genome editing in wheat. Funct. Integr. Genomics 18, 31–41. doi: 10.1007/s10142-017-0572-x

PubMed Abstract | CrossRef Full Text | Google Scholar

Klymiuk, V., Fatiukha, A., and Fahima, T. (2019). Wheat tandem kinases provide insights on disease-resistance gene flow and host–parasite co-evolution. Plant J. 98, 667–679. doi: 10.1111/tpj.14264

CrossRef Full Text | Google Scholar

Klymiuk, V., Fatiukha, A., Raats, D., Bocharova, V., Huang, L., Feng, L., et al. (2020). Three previously characterized resistances to yellow rust are encoded by a single locus Wtk1. J. Exp. Bot. 71, 2561–2572. doi: 10.1093/jxb/eraa020

PubMed Abstract | CrossRef Full Text | Google Scholar

Klymiuk, V., Yaniv, E., Huang, L., Raats, D., Fatiukha, A., Chen, S., et al. (2018). Cloning of the wheat Yr15 resistance gene sheds light on the plant tandem kinase-pseudokinase family. Nat. Commun. 9, 1–12. doi: 10.1038/s41467-018-06138-9

CrossRef Full Text | Google Scholar

Kraic, J., Mihálik, D., Klčová, L., Gubišová, M., Klempová, T., Hudcovicová, M., et al. (2018). Progress in the genetic engineering of cereals to produce essential polyunsaturated fatty acids. J. Biotechnol. 284, 115–122. doi: 10.1016/j.jbiotec.2018.08.009

PubMed Abstract | CrossRef Full Text | Google Scholar

Krasileva, K. V., Vasquez-Gross, H. A., Howell, T., Bailey, P., Paraiso, F., Clissold, L., et al. (2017). Uncovering hidden variation in polyploid wheat. Proc. Natl. Acad. Sci. U. S. A. 114, E913–E921. doi: 10.1073/pnas.1619268114

PubMed Abstract | CrossRef Full Text | Google Scholar

Krattinger, S. G., Lagudah, E. S., Spielmeyer, W., Singh, R. P., Huerta-Espino, J., McFadden, H., et al. (2009). A putative ABC transporter confers durable resistance to multiple fungal pathogens in wheat. Science 323, 1360–1363. doi: 10.1126/science.1166453

PubMed Abstract | CrossRef Full Text | Google Scholar

Kristensen, P. S., Jahoor, A., Andersen, J. R., Cericola, F., Orabi, J., Janss, L. L., et al. (2018). Genome-wide association studies and comparison of models and cross-validation strategies for genomic prediction of quality traits in advanced winter wheat breeding lines. Front. Plant Sci. 9:69. doi: 10.3389/fpls.2018.00069

PubMed Abstract | CrossRef Full Text | Google Scholar

Kruse, E. B., Carle, S. W., Wen, N., Skinner, D. Z., Murray, T. D., Garland-Campbell, K. A., et al. (2017). Genomic regions associated with tolerance to freezing stress and snow mold in winter wheat. G3 7, 775–780. doi: 10.1534/g3.116.037622

CrossRef Full Text | Google Scholar

Kuzay, S., Xu, Y., Zhang, J., Katz, A., Pearce, S., Su, Z., et al. (2019). Identification of a candidate gene for a QTL for spikelet number per spike on wheat chromosome arm 7AL by high-resolution genetic mapping. Theor. Appl. Genet. 132, 2689–2705. doi: 10.1007/s00122-019-03382-5

PubMed Abstract | CrossRef Full Text | Google Scholar

Lai, K., Lorenc, M. T., Lee, H. C., Berkman, P. J., Bayer, P. E., Visendi, P., et al. (2015). Identification and characterization of more than 4 million intervarietal SNPs across the group 7 chromosomes of bread wheat. Plant Biotechnol. J. 13, 97–104. doi: 10.1111/pbi.12240

PubMed Abstract | CrossRef Full Text | Google Scholar

Lange, T. M., Heinrich, F., Enders, M., Wolf, M., and Schmitt, A. O. (2020). In silico quality assessment of SNPs—a case study on the axiom® wheat genotyping arrays. Curr. Plant Biol. 21:100140. doi: 10.1016/j.cpb.2020.100140

CrossRef Full Text | Google Scholar

Larkin, D. L., Lozada, D. N., and Mason, R. E. (2019). Genomic selection—considerations for successful implementation in wheat breeding programs. Agronomy 9:479. doi: 10.3390/agronomy9090479

CrossRef Full Text | Google Scholar

Li, M., Dong, L., Li, B., Wang, Z., Xie, J., Qiu, D., et al. (2020). A CNL protein in wild emmer wheat confers powdery mildew resistance. New Phytol. 228, 1027–1037. doi: 10.1111/nph.16761

PubMed Abstract | CrossRef Full Text | Google Scholar

Li, Y., Fu, X., Zhao, M., Zhang, W., Li, B., An, D., et al. (2018b). A genome-wide view of transcriptome dynamics during early spike development in bread wheat. Sci. Rep. 8, 1–16. doi: 10.1038/s41598-018-33718-y

CrossRef Full Text | Google Scholar

Li, M., Li, B., Guo, G., Chen, Y., Xie, J., Lu, P., et al. (2018a). Mapping a leaf senescence gene els1 by BSR-Seq in common wheat. Crop J. 6, 236–243. doi: 10.1016/j.cj.2018.01.004

CrossRef Full Text | Google Scholar

Liang, Z., Chen, K., Li, T., Zhang, Y., Wang, Y., Zhao, Q., et al. (2017). Efficient DNA-free genome editing of bread wheat using CRISPR/Cas9 ribonucleoprotein complexes. Nat. Commun. 8, 1–5. doi: 10.1038/ncomms14261

CrossRef Full Text | Google Scholar

Liang, Z., Chen, K., Zhang, Y., Liu, J., Yin, K., Qiu, J.-L., et al. (2018). Genome editing of bread wheat using biolistic delivery of CRISPR/Cas9 in vitro transcripts or ribonucleoproteins. Nat. Protoc. 13, 413–430. doi: 10.1038/nprot.2017.145

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu, S., Bai, G., Lin, M., Luo, M., Zhang, D., Jin, F., et al. (2020). Identification of candidate chromosome region of Sbwm1 for soil-borne wheat mosaic virus resistance in wheat. Sci. Rep. 10, 1–11. doi: 10.1038/s41598-020-64993-3

CrossRef Full Text | Google Scholar

Liu, S., Yang, X., Zhang, D., Bai, G., Chao, S., and Bockus, W. (2014). Genome-wide association analysis identified SNPs closely linked to a gene resistant to soil-borne wheat mosaic virus. Theor. Appl. Genet. 127, 1039–1047. doi: 10.1007/s00122-014-2277-z

PubMed Abstract | CrossRef Full Text | Google Scholar

Lopes, M. S., Dreisigacker, S., Peña, R. J., Sukumaran, S., and Reynolds, M. P. (2015). Genetic characterization of the wheat association mapping initiative (WAMI) panel for dissection of complex traits in spring wheat. Theor. Appl. Genet. 128, 453–464. doi: 10.1007/s00122-014-2444-2

PubMed Abstract | CrossRef Full Text | Google Scholar

Lu, P., Guo, L., Wang, Z., Li, B., Li, J., Li, Y., et al. (2020). A rare gain of function mutation in a wheat tandem kinase confers resistance to powdery mildew. Nat. Commun. 11:680. doi: 10.1038/s41467-020-14294-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Luo, M., Xie, L., Chakraborty, S., Wang, A., Matny, O., Jugovich, M., et al. (2021). A five-transgene cassette confers broad-spectrum resistance to a fungal rust pathogen in wheat. Nat. Biotechnol. 39, 561–566. doi: 10.1038/s41587-020-00770-x

PubMed Abstract | CrossRef Full Text | Google Scholar

Ma, D., Yan, J., He, Z., Wu, L., and Xia, X. (2012). Characterization of a cell wall invertase gene TaCwi-A1 on common wheat chromosome 2A and development of functional markers. Mol. Breed. 29, 43–52. doi: 10.1007/s11032-010-9524-z

CrossRef Full Text | Google Scholar

Mackay, I. J., Cockram, J., Howell, P., and Powell, W. (2021). Understanding the classics: the unifying concepts of transgressive segregation, inbreeding depression and heterosis and their central relevance for crop breeding. Plant Biotechnol. J. 19, 26–34. doi: 10.1111/pbi.13481

PubMed Abstract | CrossRef Full Text | Google Scholar

Manickavelu, A., Hattori, T., Yamaoka, S., Yoshimura, K., Kondou, Y., Onogi, A., et al. (2017). Genetic nature of elemental contents in wheat grains and its genomic prediction: toward the effective use of wheat landraces from Afghanistan. PLoS One 12:e0169416. doi: 10.1371/journal.pone.0169416

PubMed Abstract | CrossRef Full Text | Google Scholar

Marchal, C., Zhang, J., Zhang, P., Fenwick, P., Steuernagel, B., Adamski, N. M., et al. (2018). BED-domain-containing immune receptors confer diverse resistance spectra to yellow rust. Nat. Plants 4, 662–668. doi: 10.1038/s41477-018-0236-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Mérida-García, R., Gálvez, S., Paux, E., Dorado, G., Pascual, L., Giraldo, P., et al. (2020). High resolution melting and insertion site-based polymorphism markers for wheat variability analysis and candidate genes selection at drought and heat MQTL loci. Agronomy 10:1294. doi: 10.3390/agronomy10091294

CrossRef Full Text | Google Scholar

Meuwissen, T. H. E. E., Hayes, B. J., and Goddard, M. E. (2001). Prediction of total genetic value using genome-wide dense marker maps. Genetics 157, 1819–1829. doi: 10.1093/genetics/157.4.1819

PubMed Abstract | CrossRef Full Text | Google Scholar

Michelmore, R. W., Paran, I., and Kesseli, R. V. (1991). Identification of markers linked to disease-resistance genes by bulked segregant analysis: a rapid method to detect markers in specific genomic regions by using segregating populations. Proc. Natl. Acad. Sci. U. S. A. 88, 9828–9832. doi: 10.1073/pnas.88.21.9828

PubMed Abstract | CrossRef Full Text | Google Scholar

Milner, S. G., Jost, M., Taketa, S., Mazón, E. R., Himmelbach, A., Oppermann, M., et al. (2019). Genebank genomics highlights the diversity of a global barley collection. Nat. Genet. 51, 319–326. doi: 10.1038/s41588-018-0266-x

PubMed Abstract | CrossRef Full Text | Google Scholar

Moore, J. W., Herrera-Foessel, S., Lan, C., Schnippenkoetter, W., Ayliffe, M., Huerta-Espino, J., et al. (2015). A recently evolved hexose transporter variant confers resistance to multiple pathogens in wheat. Nat. Genet. 47, 1494–1498. doi: 10.1038/ng.3439

PubMed Abstract | CrossRef Full Text | Google Scholar

Nakaya, A., and Isobe, S. N. (2012). Will genomic selection be a practical method for plant breeding? Ann. Bot. 110, 1303–1316. doi: 10.1093/aob/mcs109

PubMed Abstract | CrossRef Full Text | Google Scholar

Nilsen, K. T., Walkowiak, S., Xiang, D., Gao, P., Quilichini, T. D., Willick, I. R., et al. (2020). Copy number variation of TdDof controls solid-stemmed architecture in wheat. Proc. Natl. Acad. Sci. U. S. A. 117, 28708–28718. doi: 10.1073/pnas.2009418117

PubMed Abstract | CrossRef Full Text | Google Scholar

Okada, A., Arndell, T., Borisjuk, N., Sharma, N., Watson-Haigh, N. S., Tucker, E. J., et al. (2019). CRISPR/Cas9-mediated knockout of Ms1 enables the rapid generation of male-sterile hexaploid wheat lines for use in hybrid seed production. Plant Biotechnol. J. 17, 1905–1913. doi: 10.1111/pbi.13106

PubMed Abstract | CrossRef Full Text | Google Scholar

Oyiga, B. C., Sharma, R. C., Baum, M., Ogbonnaya, F. C., Léon, J., and Ballvora, A. (2018). Allelic variations and differential expressions detected at quantitative trait loci for salt stress tolerance in wheat. Plant Cell Environ. 41, 919–935. doi: 10.1111/pce.12898

PubMed Abstract | CrossRef Full Text | Google Scholar

Pang, Y., Liu, C., Wang, D., St. Amand, P., Bernardo, A., Li, W., et al. (2020). High-resolution genome-wide association study identifies genomic regions and candidate genes for important agronomic traits in wheat. Mol. Plant 13, 1311–1327. doi: 10.1016/j.molp.2020.07.008

PubMed Abstract | CrossRef Full Text | Google Scholar

Periyannan, S., Moore, J., Ayliffe, M., Bansal, U., Wang, X., Huang, L., et al. (2013). The gene Sr33, an ortholog of barley Mla genes, encodes resistance to wheat stem rust race Ug99. Science 341, 786–788. doi: 10.1126/science.1239028

CrossRef Full Text | Google Scholar

Petersen, G., Seberg, O., Yde, M., and Berthelsen, K. (2006). Phylogenetic relationships of Triticum and Aegilops and evidence for the origin of the A, B, and D genomes of common wheat (Triticum aestivum). Mol. Phylogenet. Evol. 39, 70–82. doi: 10.1016/j.ympev.2006.01.023

PubMed Abstract | CrossRef Full Text | Google Scholar

Quraishi, U. M., Pont, C., Ain, Q., Flores, R., Burlot, L., Alaux, M., et al. (2017). Combined genomic and genetic data integration of major agronomical traits in bread wheat (Triticum aestivum L.). Front. Plant Sci. 8:1843. doi: 10.3389/fpls.2017.01843

PubMed Abstract | CrossRef Full Text | Google Scholar

Raman, R., Allen, H., Diffey, S., Raman, H., Martin, P., and McKelvie, K. (2009). Localisation of quantitative trait loci for quality attributes in a doubled haploid population of wheat (Triticum aestivum L.). Genome 52, 701–715. doi: 10.1139/G09-045

PubMed Abstract | CrossRef Full Text | Google Scholar

Ramirez-Gonzalez, R. H., Segovia, V., Bird, N., Fenwick, P., Holdgate, S., Berry, S., et al. (2015). RNA-Seq bulked segregant analysis enables the identification of high-resolution genetic markers for breeding in hexaploid wheat. Plant Biotechnol. J. 13, 613–624. doi: 10.1111/pbi.12281

PubMed Abstract | CrossRef Full Text | Google Scholar

Ran, Y., Patron, N., Kay, P., Wong, D., Buchanan, M., Cao, Y.-Y., et al. (2018). Zinc finger nuclease-mediated precision genome editing of an endogenous gene in hexaploid bread wheat (Triticum aestivum) using a DNA repair template. Plant Biotechnol. J. 16, 2088–2101. doi: 10.1111/pbi.12941

PubMed Abstract | CrossRef Full Text | Google Scholar

Raza, Q., Riaz, A., Atif, R. M., Hussain, B., Ali, Z., Budak, H., et al. (2021). Genome-wide diversity of MADS-Box genes in bread wheat is associated with its rapid global adaptability. Front. Genet. 12:818880. doi: 10.3389/fgene.2021.818880

PubMed Abstract | CrossRef Full Text | Google Scholar

Reynolds, M., and Langridge, P. (2016). Physiological breeding. Curr. Opin. Plant Biol. 31, 162–171. doi: 10.1016/j.pbi.2016.04.005

PubMed Abstract | CrossRef Full Text | Google Scholar

Rocha-Meneses, L., Ferreira, J. A., Mushtaq, M., Karimi, S., Orupõld, K., and Kikas, T. (2020). Genetic modification of cereal plants: a strategy to enhance bioethanol yields from agricultural waste. Ind. Crop Prod. 150:112408. doi: 10.1016/j.indcrop.2020.112408

CrossRef Full Text | Google Scholar

Ru, Z., Juhasz, A., Li, D., Deng, P., Zhao, J., Gao, L., et al. (2020). 1RS.1BL molecular resolution provides novel contributions to wheat improvement. bioRxiv [Preprint], 1–35. doi:doi: 10.1101/2020.09.14.295733

CrossRef Full Text | Google Scholar

Rutkoski, J., Poland, J., Mondal, S., Autrique, E., Pérez, L. G., Crossa, J., et al. (2016). Canopy temperature and vegetation indices from high-throughput phenotyping improve accuracy of pedigree and genomic selection for grain yield in wheat. G3 6, 2799–2808. doi: 10.1534/g3.116.032888

CrossRef Full Text | Google Scholar

Saint Pierre, C., Burgueño, J., Crossa, J., Fuentes Dávila, G., Figueroa López, P., Solís Moya, E., et al. (2016). Genomic prediction models for grain yield of spring bread wheat in diverse agro-ecological zones. Sci. Rep. 6, 1–11. doi: 10.1038/srep27312

CrossRef Full Text | Google Scholar

Saintenac, C., Cambon, F., Aouini, L., Verstappen, E., Ghaffary, S. M. T., Poucet, T., et al. (2021). A wheat cysteine-rich receptor-like kinase confers broad-spectrum resistance against Septoria tritici blotch. Nat. Commun. 12:433. doi: 10.1038/s41467-020-20685-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Saintenac, C., Lee, W.-S., Cambon, F., Rudd, J. J., King, R. C., Marande, W., et al. (2018). Wheat receptor-kinase-like protein Stb6 controls gene-for-gene resistance to fungal pathogen Zymoseptoria tritici. Nat. Genet. 50, 368–374. doi: 10.1038/s41588-018-0051-x

PubMed Abstract | CrossRef Full Text | Google Scholar

Saintenac, C., Zhang, W., Salcedo, A., Rouse, M. N., Trick, H. N., Akhunov, E., et al. (2013). Identification of wheat gene Sr35 that confers resistance to Ug99 stem rust race group. Science 341, 783–786. doi: 10.1126/science.1239022

PubMed Abstract | CrossRef Full Text | Google Scholar

Sánchez-León, S., Gil-Humanes, J., Ozuna, C. V., Giménez, M. J., Sousa, C., Voytas, D. F., et al. (2018). Low-gluten, nontransgenic wheat engineered with CRISPR/Cas9. Plant Biotechnol. J. 16, 902–910. doi: 10.1111/pbi.12837

PubMed Abstract | CrossRef Full Text | Google Scholar

Sánchez-Martín, J., Steuernagel, B., Ghosh, S., Herren, G., Hurni, S., Adamski, N., et al. (2016). Rapid gene isolation in barley and wheat by mutant chromosome sequencing. Genome Biol. 17, 221. doi: 10.1186/s13059-016-1082-1

CrossRef Full Text | Google Scholar

Scheben, A., and Edwards, D. (2017). Genome editors take on crops. Science 355, 1122–1123. doi: 10.1126/science.aal4680

PubMed Abstract | CrossRef Full Text | Google Scholar

Scheben, A., Wolter, F., Batley, J., Puchta, H., and Edwards, D. (2017). Towards CRISPR/Cas crops—bringing together genomics and genome editing. New Phytol. 216, 682–698. doi: 10.1111/nph.14702

PubMed Abstract | CrossRef Full Text | Google Scholar

Schrag, T. A., Westhues, M., Schipprack, W., Seifert, F., Thiemann, A., Scholten, S., et al. (2018). Beyond genomic prediction: combining different types of omics data can improve prediction of hybrid performance in maize. Genetics 208, 1373–1385. doi: 10.1534/genetics.117.300374

PubMed Abstract | CrossRef Full Text | Google Scholar

Sehgal, D., Autrique, E., Singh, R., Ellis, M., Singh, S., and Dreisigacker, S. (2017). Identification of genomic regions for grain yield and yield stability and their epistatic interactions. Sci. Rep. 7, 1–12. doi: 10.1038/srep41578

CrossRef Full Text | Google Scholar

Sehgal, D., Mondal, S., Crespo-Herrera, L., Velu, G., Juliana, P., Huerta-Espino, J., et al. (2020a). Haplotype-based, genome-wide association study reveals stable genomic regions for grain yield in CIMMYT spring bread wheat. Front. Genet. 11:1427. doi: 10.3389/fgene.2020.589490

CrossRef Full Text | Google Scholar

Sehgal, D., Rosyara, U., Mondal, S., Singh, R., Poland, J., and Dreisigacker, S. (2020b). Incorporating genome-wide association mapping results into genomic prediction models for grain yield and yield stability in CIMMYT spring bread wheat. Front. Plant Sci. 11:197. doi: 10.3389/fpls.2020.00197

CrossRef Full Text | Google Scholar

Shan, Q., Wang, Y., Li, J., and Gao, C. (2014). Genome editing in rice and wheat using the CRISPR/Cas system. Nat. Protoc. 9, 2395–2410. doi: 10.1038/nprot.2014.157

PubMed Abstract | CrossRef Full Text | Google Scholar

Sharma, D. K., Torp, A. M., Rosenqvist, E., Ottosen, C.-O., and Andersen, S. B. (2017). QTLs and potential candidate genes for heat stress tolerance identified from the mapping populations specifically segregating for Fv/Fm in wheat. Front. Plant Sci. 8:1668. doi: 10.3389/fpls.2017.01668

PubMed Abstract | CrossRef Full Text | Google Scholar

Shi, G., Zhang, Z., Friesen, T. L., Raats, D., Fahima, T., Brueggeman, R. S., et al. (2016). The hijacking of a receptor kinase-driven pathway by a wheat fungal pathogen leads to disease. Sci. Adv. 2:e1600822. doi: 10.1126/sciadv.1600822

PubMed Abstract | CrossRef Full Text | Google Scholar

Shokat, S., Sehgal, D., Vikram, P., Liu, F., and Singh, S. (2020). Molecular markers associated with agro-physiological traits under terminal drought conditions in bread wheat. Int. J. Mol. Sci. 21:3156. doi: 10.3390/ijms21093156

PubMed Abstract | CrossRef Full Text | Google Scholar

Simons, K., Anderson, J. A., Mergoum, M., Faris, J. D., Klindworth, D. L., Xu, S. S., et al. (2012). Genetic mapping analysis of bread-making quality traits in spring wheat. Crop. Sci. 52:2182. doi: 10.2135/cropsci2012.03.0175

PubMed Abstract | CrossRef Full Text | Google Scholar

Singh, A. K., Singh, N., Kumar, S., Kumari, J., Singh, R., Gaba, S., et al. (2020). Identification and evolutionary analysis of polycistronic miRNA clusters in domesticated and wild wheat. Genomics 112, 2334–2348. doi: 10.1016/j.ygeno.2020.01.005

PubMed Abstract | CrossRef Full Text | Google Scholar

Steuernagel, B., Periyannan, S. K., Hernández-Pinzón, I., Witek, K., Rouse, M. N., Yu, G., et al. (2016). Rapid cloning of disease-resistance genes in plants using mutagenesis and sequence capture. Nat. Biotechnol. 34, 652–655. doi: 10.1038/nbt.3543

PubMed Abstract | CrossRef Full Text | Google Scholar

Su, Z., Bernardo, A., Tian, B., Chen, H., Wang, S., Ma, H., et al. (2019). A deletion mutation in TaHRC confers Fhb1 resistance to Fusarium head blight in wheat. Nat. Genet. 51, 1099–1105. doi: 10.1038/s41588-019-0425-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Su, Z., Hao, C., Wang, L., Dong, Y., and Zhang, X. (2011). Identification and development of a functional marker of TaGW2 associated with grain weight in bread wheat (Triticum aestivum L.). Theor. Appl. Genet. 122, 211–223. doi: 10.1007/s00122-010-1437-z

PubMed Abstract | CrossRef Full Text | Google Scholar

Su, Z., Jin, S., Lu, Y., Zhang, G., Chao, S., and Bai, G. (2016). Single nucleotide polymorphism tightly linked to a major QTL on chromosome 7A for both kernel length and kernel weight in wheat. Mol. Breed. 36:15. doi: 10.1007/s11032-016-0436-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Sun, H., Lu, J., Fan, Y., Zhao, Y., Kong, F., Li, R., et al. (2008). Quantitative trait loci (QTLs) for quality traits related to protein and starch in wheat. Prog. Nat. Sci. 18, 825–831. doi: 10.1016/j.pnsc.2007.12.013

PubMed Abstract | CrossRef Full Text | Google Scholar

Takagi, H., Abe, A., Yoshida, K., Kosugi, S., Natsume, S., Mitsuoka, C., et al. (2013a). QTL-seq: rapid mapping of quantitative trait loci in rice by whole genome resequencing of DNA from two bulked populations. Plant J. 74, 174–183. doi: 10.1111/tpj.12105

CrossRef Full Text | Google Scholar

Takagi, H., Uemura, A., Yaegashi, H., Tamiru, M., Abe, A., Mitsuoka, C., et al. (2013b). MutMap-gap: whole-genome resequencing of mutant F2 progeny bulk combined with de novo assembly of gap regions identifies the rice blast resistance gene Pii. New Phytol. 200, 276–283. doi: 10.1111/nph.12369

CrossRef Full Text | Google Scholar

Talukder, S. K., Babar, M. A., Vijayalakshmi, K., Poland, J., Prasad, P. V. V., Bowden, R., et al. (2014). Mapping QTL for the traits associated with heat tolerance in wheat (Triticum aestivum L.). BMC Genet. 15:97. doi: 10.1186/s12863-014-0097-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Thind, A. K., Wicker, T., Šimková, H., Fossati, D., Moullet, O., Brabant, C., et al. (2017). Rapid cloning of genes in hexaploid wheat using cultivar-specific long-range chromosome assembly. Nat. Biotechnol. 35, 793–796. doi: 10.1038/nbt.3877

PubMed Abstract | CrossRef Full Text | Google Scholar

Trick, M., Adamski, N. M., Mugford, S. G., Jiang, C. C., Febrer, M., and Uauy, C. (2012). Combining SNP discovery from next- generation sequencing data with bulked segregant analysis (BSA) to fine-map genes in polyploid wheat. BMC Plant Biol. 12:14. doi: 10.1186/1471-2229-12-14

PubMed Abstract | CrossRef Full Text | Google Scholar

Uauy, C., Distelfeld, A., Fahima, T., Blechl, A., and Dubcovsky, J. (2006). A NAC gene regulating senescence improves grain protein, zinc, and iron content in wheat. Science 314, 1298–1301. doi: 10.1126/science.1133649

PubMed Abstract | CrossRef Full Text | Google Scholar

Upadhyay, S. K., Kumar, J., Alok, A., and Tuli, R. (2013). RNA-guided genome editing for target gene mutations in wheat. G3 3, 2233–2238. doi: 10.1534/g3.113.008847

PubMed Abstract | CrossRef Full Text | Google Scholar

Vágújfalvi, A., Galiba, G., Cattivelli, L., and Dubcovsky, J. (2003). The cold-regulated transcriptional activator Cbf3 is linked to the frost-tolerance locus Fr-A2 on wheat chromosome 5A. Mol. Genet. Genomics 269, 60–67. doi: 10.1007/s00438-003-0806-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Voss-Fels, K. P., Keeble-Gagnère, G., Hickey, L. T., Tibbits, J., Nagornyy, S., Hayden, M. J., et al. (2019). High-resolution mapping of rachis nodes per rachis, a critical determinant of grain yield components in wheat. Theor. Appl. Genet. 132, 2707–2719. doi: 10.1007/s00122-019-03383-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Walkowiak, S., Gao, L., Monat, C., Haberer, G., Kassa, M. T., Brinton, J., et al. (2020). Multiple wheat genomes reveal global variation in modern breeding. Nature 2020, 1–7. doi: 10.1038/s41586-020-2961-x

CrossRef Full Text | Google Scholar

Wang, D., Cao, D., Zong, Y., Li, Y., Wang, J., Li, Z., et al. (2021). Bulked QTL-Seq identified a major QTL for the awnless trait in spring wheat cultivars in Qinghai, China. Biotechnol. Biotechnol. Equip. 35, 124–130. doi: 10.1080/13102818.2020.1857661

CrossRef Full Text | Google Scholar

Wang, Y., Cheng, X., Shan, Q., Zhang, Y., Liu, J., Gao, C., et al. (2014). Simultaneous editing of three homoeoalleles in hexaploid bread wheat confers heritable resistance to powdery mildew. Nat. Biotechnol. 32, 947–951. doi: 10.1038/nbt.2969

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, Z., Li, J., Chen, S., Heng, Y., Chen, Z., Yang, J., et al. (2017). Poaceae-specific MS1 encodes a phospholipid-binding protein for male fertility in bread wheat. Proc. Natl. Acad. Sci. U. S. A. 114, 12614–12619. doi: 10.1073/pnas.1715570114

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, W., Pan, Q., He, F., Akhunova, A., Chao, S., Trick, H., et al. (2018a). Transgenerational CRISPR-Cas9 activity facilitates multiplex gene editing in allopolyploid wheat. Cris. J. 1, 65–74. doi: 10.1089/crispr.2017.0010

CrossRef Full Text | Google Scholar

Wang, H., Sun, S., Ge, W., Zhao, L., Hou, B., Wang, K., et al. (2020a). Horizontal gene transfer of Fhb7 from fungus underlies Fusarium head blight resistance in wheat. Science 368:eaba5435. doi: 10.1126/science.aba5435

CrossRef Full Text | Google Scholar

Wang, Y., Zhang, H., Xie, J., Guo, B., Chen, Y., Zhang, H., et al. (2018b). Mapping stripe rust resistance genes by BSR-Seq: YrMM58 and YrHY1 on chromosome 2AS in Chinese wheat lines Mengmai 58 and Huaiyang 1 are Yr17. Crop J. 6, 91–98. doi: 10.1016/j.cj.2017.03.002

CrossRef Full Text | Google Scholar

Wang, H., Zou, S., Li, Y., Lin, F., and Tang, D. (2020b). An ankyrin-repeat and WRKY-domain-containing immune receptor confers stripe rust resistance in wheat. Nat. Commun. 11, 1–11. doi: 10.1038/s41467-020-15139-6

CrossRef Full Text | Google Scholar

Watson, A., Ghosh, S., Williams, M. J., Cuddy, W. S., Simmonds, J., Rey, M. D., et al. (2018). Speed breeding is a powerful tool to accelerate crop research and breeding. Nat. Plants 4, 23–29. doi: 10.1038/s41477-017-0083-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Wei, K., and Han, P. (2017). Comparative functional genomics of the TPR gene family in Arabidopsis, rice and maize. Mol. Breed. 37, 1–21. doi: 10.1007/s11032-017-0751-4

CrossRef Full Text | Google Scholar

Würschum, T., Longin, C. F. H., Hahn, V., Tucker, M. R., and Leiser, W. L. (2017). Copy number variations of CBF genes at the Fr-A2 locus are essential components of winter hardiness in wheat. Plant J. 89, 764–773. doi: 10.1111/tpj.13424

PubMed Abstract | CrossRef Full Text | Google Scholar

Xie, J., Guo, G., Wang, Y., Hu, T., Wang, L., Li, J., et al. (2020). A rare single nucleotide variant in Pm5e confers powdery mildew resistance in common wheat. New Phytol. 228, 1011–1026. doi: 10.1111/nph.16762

PubMed Abstract | CrossRef Full Text | Google Scholar

Xie, K., Minkenberg, B., and Yang, Y. (2015). Boosting CRISPR/Cas9 multiplex editing capability with the endogenous tRNA- processing system. Proc. Natl. Acad. Sci. 112, 3570–3575. doi: 10.1073/pnas.1420294112

PubMed Abstract | CrossRef Full Text | Google Scholar

Xing, L., Hu, P., Liu, J., Witek, K., Zhou, S., Xu, J., et al. (2018). Pm21 from Haynaldia villosa encodes a CC-NBS-LRR protein conferring powdery mildew resistance in wheat. Mol. Plant 11, 874–878. doi: 10.1016/j.molp.2018.02.013

PubMed Abstract | CrossRef Full Text | Google Scholar

Yahiaoui, N., Srichumpa, P., Dudler, R., and Keller, B. (2004). Genome analysis at different ploidy levels allows cloning of the powdery mildew resistance gene Pm3b from hexaploid wheat. Plant J. 37, 528–538. doi: 10.1046/j.1365-313X.2003.01977.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Yan, L., Loukoianov, A., Tranquilli, G., Helguera, M., Fahima, T., and Dubcovsky, J. (2003). Positional cloning of the wheat vernalization gene VRN1. Proc. Natl. Acad. Sci. U. S. A. 100, 6263–6268. doi: 10.1073/pnas.0937399100

PubMed Abstract | CrossRef Full Text | Google Scholar

Yang, Y., Zhang, F., Zhou, T., Fang, A., Yu, Y., Bi, C., et al. (2020). In Silico identification of the full complement of subtilase-encoding genes and characterization of the role of TaSBT1.7 in resistance against stripe rust in wheat. Phytopathology 111, 398–407. doi: 10.1094/phyto-05-20-0176-r

CrossRef Full Text | Google Scholar

Yu, S., Wu, J., Wang, M., Shi, W., Xia, G., Jia, J., et al. (2020). Haplotype variations in QTL for salt tolerance in Chinese wheat accessions identified by marker-based and pedigree-based kinship analyses. Crop J. 8, 1011–1024. doi: 10.1016/j.cj.2020.03.007

CrossRef Full Text | Google Scholar

Zanke, C. D., Ling, J., Plieske, J., Kollers, S., Ebmeyer, E., Korzun, V., et al. (2015). Analysis of main effect QTL for thousand grain weight in European winter wheat (Triticum aestivum L.) by genome-wide association mapping. Front. Plant Sci. 6:644. doi: 10.3389/fpls.2015.00644

PubMed Abstract | CrossRef Full Text | Google Scholar

Zeng, F., Yang, L., Gong, S., Shi, W., Zhang, X., Wang, H., et al. (2014). Virulence and diversity of Blumeria graminis f. sp. tritici populations in China. J. Integr. Agric. 13, 2424–2437. doi: 10.1016/S2095-3119(13)60669-3

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, Y., Bai, Y., Wu, G., Zou, S., Chen, Y., Gao, C., et al. (2017). Simultaneous modification of three homoeologs of TaEDR1 by genome editing enhances powdery mildew resistance in wheat. Plant J. 91, 714–724. doi: 10.1111/tpj.13599

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, Y., Liu, J., Xia, X., and He, Z. (2014). TaGS-D1, an ortholog of rice OsGS3, is associated with grain weight and grain length in common wheat. Mol. Breed. 34, 1097–1107. doi: 10.1007/s11032-014-0102-7

CrossRef Full Text | Google Scholar

Zhang, H., Mao, X., Zhang, J., Chang, X., and Jing, R. (2013). Single-nucleotide polymorphisms and association analysis of drought-resistance gene TaSnRK2.8 in common wheat. Plant Physiol. Biochem. 70, 174–181. doi: 10.1016/j.plaphy.2013.04.010

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, J., Zhang, P., Dodds, P., and Lagudah, E. (2020). How target-sequence enrichment and sequencing (TEnSeq) pipelines have catalyzed resistance gene cloning in the wheat-rust pathosystem. Front. Plant Sci. 11:678. doi: 10.3389/fpls.2020.00678

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhao, Y., Gowda, M., Würschum, T., Longin, C. F. H., Korzun, V., Kollers, S., et al. (2013a). Dissecting the genetic architecture of frost tolerance in central European winter wheat. J. Exp. Bot. 64, 4453–4460. doi: 10.1093/jxb/ert259

CrossRef Full Text | Google Scholar

Zhao, Y., Li, Z., Liu, G., Jiang, Y., Maurer, H. P., Würschum, T., et al. (2015). Genome-based establishment of a high-yielding heterotic pattern for hybrid wheat breeding. Proc. Natl. Acad. Sci. U. S. A. 112, 15624–15629. doi: 10.1073/pnas.1514547112

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhao, Y., Mette, M. F., Gowda, M., Longin, C. F. H., and Reif, J. C. (2014). Bridging the gap between marker-assisted and genomic selection of heading time and plant height in hybrid wheat. Heredity 112, 638–645. doi: 10.1038/hdy.2014.1

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhao, D., Yang, L., Xu, K., Cao, S., Tian, Y., Yan, J., et al. (2020). Identification and validation of genetic loci for tiller angle in bread wheat. Theor. Appl. Genet. 133, 3037–3047. doi: 10.1007/s00122-020-03653-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhao, Y., Zeng, J., Fernando, R., and Reif, J. C. (2013b). Genomic prediction of hybrid wheat performance. Crop. Sci. 53, 802–810. doi: 10.2135/cropsci2012.08.0463

CrossRef Full Text | Google Scholar

Zhou, X., Zhu, X., Shao, W., Song, J., Jiang, W., He, Y., et al. (2020). Genome-wide mining of wheat DUF966 gene family provides new insights into salt stress responses. Front. Plant Sci. 11:1345. doi: 10.3389/fpls.2020.569838

PubMed Abstract | CrossRef Full Text | Google Scholar

Zimin, A. V., Puiu, D., Hall, R., Kingan, S., Clavijo, B. J., and Salzberg, S. L. (2017). The first near-complete assembly of the hexaploid bread wheat genome, Triticum aestivum. Gigascience 6, 1–7. doi: 10.1093/gigascience/gix097

PubMed Abstract | CrossRef Full Text | Google Scholar

Zou, S., Wang, H., Li, Y., Kong, Z., and Tang, D. (2018). The NB-LRR gene Pm60 confers powdery mildew resistance in wheat. New Phytol. 218, 298–309. doi: 10.1111/nph.14964

PubMed Abstract | CrossRef Full Text | Google Scholar

Zou, C., Wang, P., and Xu, Y. (2016). Bulked sample analysis in genetics, genomics and crop improvement. Plant Biotechnol. J. 14, 1941–1955. doi: 10.1111/pbi.12559

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: Wheat, genome-wide association, quantitative trait locus mapping, abiotic-stress tolerance, genomic selection, QTL cloning, disease resistance, CRISPR/Cas9

Citation: Hussain B, Akpınar BA, Alaux M, Algharib AM, Sehgal D, Ali Z, Aradottir GI, Batley J, Bellec A, Bentley AR, Cagirici HB, Cattivelli L, Choulet F, Cockram J, Desiderio F, Devaux P, Dogramaci M, Dorado G, Dreisigacker S, Edwards D, El-Hassouni K, Eversole K, Fahima T, Figueroa M, Gálvez S, Gill KS, Govta L, Gul A, Hensel G, Hernandez P, Crespo-Herrera LA, Ibrahim A, Kilian B, Korzun V, Krugman T, Li Y, Liu S, Mahmoud AF, Morgounov A, Muslu T, Naseer F, Ordon F, Paux E, Perovic D, Reddy GVP, Reif JC, Reynolds M, Roychowdhury R, Rudd J, Sen TZ, Sukumaran S, Ozdemir BS, Tiwari VK, Ullah N, Unver T, Yazar S, Appels R and Budak H (2022) Capturing Wheat Phenotypes at the Genome Level. Front. Plant Sci. 13:851079. doi: 10.3389/fpls.2022.851079

Received: 09 January 2022; Accepted: 19 May 2022;
Published: 04 July 2022.

Edited by:

Jianjun Chen, University of Florida, Gainesville, United States

Reviewed by:

Ahmad M. Alqudah, Aarhus University, Denmark
Hongliang Zhu, China Agricultural University, China

Copyright © 2022 Hussain, Akpınar, Alaux, Algharib, Sehgal, Ali, Aradottir, Batley, Bellec, Bentley, Cagirici, Cattivelli, Choulet, Cockram, Desiderio, Devaux, Dogramaci, Dorado, Dreisigacker, Edwards, El-Hassouni, Eversole, Fahima, Figueroa, Gálvez, Gill, Govta, Gul, Hensel, Hernandez, Crespo-Herrera, Ibrahim, Kilian, Korzun, Krugman, Li, Liu, Mahmoud, Morgounov, Muslu, Naseer, Ordon, Paux, Perovic, Reddy, Reif, Reynolds, Roychowdhury, Rudd, Sen, Sukumaran, Ozdemir, Tiwari, Ullah, Unver, Yazar, Appels and Budak. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Hikmet Budak, hikmet.budak@icloud.com

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