Open access peer-reviewed chapter

Breaking Yield Ceiling in Wheat: Progress and Future Prospects

Written By

Neeraj Pal, Dinesh Kumar Saini and Sundip Kumar

Submitted: 24 January 2022 Reviewed: 27 January 2022 Published: 25 March 2022

DOI: 10.5772/intechopen.102919

From the Edited Volume

Wheat - Recent Advances

Edited by Mahmood-ur-Rahman Ansari

Chapter metrics overview

230 Chapter Downloads

View Full Metrics

Abstract

Wheat is one of the most important staple crops that contribute considerably to global food and nutritional security. The future projections of the demand for wheat show significant enhancement owing to the population growth and probable changes in diets. Further, historical yield trends show a reduction in the relative rate of gain for grain yield over time. To maintain future food security, there is a strong need to find ways to further increase the yield potential of wheat. Grain yield is a quantitative trait that is highly influenced by the environment. It is determined by various interlinked yield component traits. Molecular breeding approaches have already proven useful in improving the grain yield of wheat and recent advances in high-throughput genotyping platforms now have remodelled molecular breeding to genomics-assisted breeding. Hence, here in this chapter, we have discussed various advancements in understanding the genetics of grain yield, its major components, and summarised the various powerful strategies, such as gene cloning, mining superior alleles, transgenic technologies, advanced genome editing techniques, genomic selection, genome-wide association studies-assisted genomic selection, haplotype-based breeding (HBB), which may be/being used for grain yield improvement in wheat and as the new breeding strategies they could also be utilised to break the yield ceiling in wheat.

Keywords

  • wheat
  • grain yield
  • genomics resources
  • molecular breeding
  • genomics-assisted breeding

1. Introduction

Wheat (Triticum aestivum L.) is the most extensively grown food crop around the world and ranks second after rice [1]. China is the top wheat-growing country which recently in 2020, produced 134,250 thousand tonnes of wheat accounting for approximately 20.66% of the total wheat production around the globe. The top five wheat-growing countries (China, India, Russian Federation, United States of America, and Canada) together account for 63.46% of the world’s wheat production (6,499,759 thousand tonnes in 2020) [2]. It accounts for more than 20% of the calorific intake of humans and supplies more protein (approximately 23%) than all animal sources [1]. The progress for the genetic improvement of grain yield in wheat ranged from 0.3% to 1.0% per year during the last century [3]. Nevertheless, it has been decreased in recent years, largely due to the narrow genetic base of the germplasm used for the development of new genotypes and the lack of adoption of novel breeding strategies. Noticeably, there is a need to increase wheat yield to feed the world population which may be increased from the current 7.5 billion to more than 9 billion by 2050, and this is with the unusual constraints posed by climate change. Under such kind of pressure, wheat breeding programs need to do more to achieve the targeted genetic gain in grain yield ensuring food security in the near future. Many studies have shown that increases in the harvest index (HI), grain weight (GW), grain number per spike (GNPS), and decreases in plant height (PH) are the major traits associated with genetic gain in wheat [4, 5]. Improvement in HI has permitted better partitioning of photosynthetic assimilates to the developing grains, resulting in greater grain yield (GY). The HI of cultivated wheat varieties generally ranges from 0.4 to 0.5 which is already close to the theoretical maximum value of 0.62 [6, 7]. Furthermore, HI values more than 0.5 are very hard to achieve, particularly in unfavourable environmental conditions [8]. This situation again shows that genetic progress in wheat breeding programs may be difficult. Therefore, understanding the changes (either increment or reduction) in yield and related traits is an essential step towards developing new breeding strategies and a further improvement in the grain yield.

Grain yield is the final result of plant growth and development and hence most, if not all, genes are supposed to contribute towards yield either indirectly or directly. Consequently, achieving increased grain yield is a non-trivial task, and the accumulative knowledge from wheat breeding suggests that we would require concurrent improvements of both the ‘source’ and ‘sink’ tissues. Traditional breeding largely depends on empirical phenotypic selection, which has already resulted in the development and release of a large number of high-yielding varieties. However, time consumption, labour intensity, environmental dependence, and low efficiency are prime barriers that nowadays hinder conventional/traditional wheat breeding. High-yielding wheat varieties can result from the uncovering of novel genetic variation, better selection techniques, or the identification of superior genotypes with novel or improved characteristics caused by favourable combinations of superior alleles at multiple loci. In recent years, an impressive number of advancements in genetics and genomics have been made in wheat. Owing to the tremendous effort of IWGSC, the ‘gold standard’ reference genome has become available for wheat cultivar ‘Chinese Spring’. The most comprehensive assembly of this reference line has been recently released in 2018 which gave access to a total of 107,891 high-confidence genes [9]. The genome sequences may assist the identification of important genes at an unprecedented level which is a key aspect in wheat breeding. Different types of molecular markers, such as RFLP, AFLP, SCAR, STS, SSR, CAPS, and GBS-SNPs, have been identified and mapped on the different chromosomes of wheat and highly dense genetic maps have also been developed (available at https://wheat.pw.usda.gov/GG3/) which are being utilised in various genetic studies in wheat [10, 11]. To date, more than 15 different high-throughput GBS strategies have been developed and utilised in various crops including wheat [12]. Moreover, several SNP arrays/assays have also been developed which are flexible in terms of data point and sample number customization, which contributes to its high-density scanning and robust call rates compared to PCR and NGS-based markers. Several high-density SNP genotyping arrays have been utilised for genetic dissection of different traits and marker-assisted breeding in wheat namely the Illumina Wheat 9 K iSelect SNP array [13], the Illumina Wheat 90 K iSelect SNP genotyping array [14], the Wheat 15 K SNP array [15], the Wheat 55 K SNP array, the Axiom Wheat 660 K SNP array, the Axiom HD Wheat genotyping (820 K) array [16], the Wheat Breeders’ 35 K Axiom array [17], and the Wheat 50 K Triticum TraitBreed array [18]. These advancements in genomics have greatly enhanced our understanding of structural and functional aspects of the wheat genome and contributed to wheat improvement in two ways. First, they provided a better understanding of the various biological mechanisms that have led to new or improved screening methods for identifying and selecting superior genotypes more efficiently. Secondly, this new information improved the decision-making process for more efficient breeding strategies. With these advancements, the focus of wheat breeding has gradually switched from phenotype-based to genotype-based selection. Marker-assisted selection (MAS) has improved wheat breeding efficiency to some extent and predominated in breeding programs for decades. Several MAS strategies have been developed—marker-assisted backcrossing (MABC) or introgression of QTL or major genes, selection of complex quantitative traits using molecular markers, and enrichment of favourable alleles in early generations [19]. Availability of high-throughput genotyping platforms and genomics resources now rapidly remodelling marker-assisted breeding to genomics-assisted breeding.

Here in this chapter, we summarise the recent progress in understanding the genetics of grain yield and other related traits together with the new strategies, such as gene cloning and mining of superior alleles, transgenic technologies, genome editing technologies, genomic selection (GS), genome-wide association studies (GWAS)-assisted GS, and haplotype-based breeding (including haplotype-based GWAS and haplotype-based GS), which altogether make it available for genomics-assisted breeding (GAB) in crop improvement and to break the yield ceiling in wheat.

Advertisement

2. Genetics of grain yield and its related traits

Grain yield is a complex polygenic trait, significantly associated with grain number per spike, grain weight, harvest index, number of productive tillers, plant height, days to heading/flowering, etc. The trait is also influenced by the environment and shows a significant level of genotype × environment interaction with low heritability. Previous studies showed that increased yield potential in the major wheat-growing countries was largely associated with increased grains per square meter, harvest index, and biomass, and reduced plant height [4, 5]. Moreover, it has been revealed that the use of dwarfing genes (Rht1, Rht2, Rht8, and Rht24), the 1BL.1RS translocation lines [20, 21, 22, 23], and positive selection of desirable alleles of major genes including grain size (for instances, TaGS3-A1, TaTGW6, TaSus1, TaGW2, TaGW8, and so on), vernalization requirements (Vrn genes), photoperiod response (Ppd-1), etc. resulted into the enormous improvement in wheat grain yield [24, 25]. It is now believed that further improvement in grain yield can be attained only by exploiting untapped genetic variation and depth understanding of its genetic architecture combined with the use of advanced genomics-assisted breeding techniques. QTL mapping has been one of the innovative approaches for understanding the genetic architecture of grain yield and its component traits in wheat. Advancements in molecular marker systems have revolutionised the field of QTL mapping, as hundreds of QTLs for different yield-related traits have been mapped using different bi-parental and multi-parental mapping populations in several countries [26, 27, 28, 29, 30, 31]. The QTL regions identified by the standard interval mapping procedure frequently extend to several centimorgans (cM) on linkage map (on the physical map, it may be equivalent to the several Mbp) which may encompass a large number of genes [31]. Therefore, it becomes very hard to pinpoint the causative locus/candidate gene responsible for a specific trait. Furthermore, the introgression of such large QTL regions based on linked or flanking markers might carry several unwanted genes due to linkage drag, thereby negatively affecting the performance of generated cultivars encompassing the introgressed genomic segments. Therefore, the genetic resolution of the mapping procedures must be increased to allow QTL placement within the shortest possible genomic region using advanced strategies. Fine mapping is an important strategy that can be used for refining the QTL region. Three major factors, such as phenotyping, population size, and the number of markers, mainly regulate the success of QTL dissection, fine mapping, and further cloning of desired QTLs. Advances in NGS technologies have dramatically reduced per sample genotyping cost and offered increased throughput. Moreover, with the latest SNP genotyping platforms such as SNP chips or arrays in place, it is now quite possible to genotype tens of thousands of samples in a short period [32]. Moreover, QTL fine mapping occasionally reveals surprises, for instance, the presence of distinct genes whose combined effects contribute to the QTL identified using standard mapping procedures, distinct upstream non-coding enhancer/modifier sequences that contribute to phenotypic effects of a QTL, and substantial genetic differences between the alleles in the QTL region. Identification of the genes or sequence variants that underlie QTL may help in investigating the contribution of specific genes or structural variants to the overall genetic architecture of grain yield and related traits [26, 33].

As discussed above, several studies have reported hundreds of QTLs in different mapping populations evaluated under different environments. An innovative approach i.e., meta-QTL analysis has emerged which helps in refining the QTL positions by combining the QTL results from independent studies and identifying the most stable and consensus QTLs [34]. The power of this approach lies in detecting regions of the genome that are most often involved in trait variation and reducing the QTL confidence intervals, thus facilitating the identification and characterisation of underlying candidate genes. For the first time in 2010, Zhang and his colleagues [35] conducted a meta-QTL analysis of major QTLs for grain yield and yield-related traits and identified 12 significant MQTLs on chromosomes 1A, 1B, 2A, 2D, 3B, 4A, 4B, 4D, and 5A, few of which also included important known genes, such as Vrn and Rht [35]. Another study reported 16 MQTLs on chromosomes 1B, 2A, 2D, 3B, 4A, 6A, and 6B, related to grain weight [36]. Most recently in 2021, Saini and his colleagues [37] have identified a total of 141 MQTLs responsible for grain yield and related traits, which included 13 breeder’s MQTLs and 24 ortho-MQTLs. This study also identified 1202 high-confidence candidate genes within the physical positions of the MQTL flanking markers [37]. Beside these, recently, various other MQTL studies have been also conducted in wheat [38, 39, 40, 41]. DNA markers tightly linked to these meta-QTLs (MQTLs) may be used as molecular tools for MAS in wheat breeding. Association mapping or GWAS offers an alternative route for identifying genomic regions that have effects across a wider range of germplasm if false associations that are caused by population structure and relatedness can be minimised. With the advancements in high-throughput genotyping technologies, haplotypes and SNP-sets (instead of single SNPs) are being utilised for GWAS, thereby reducing the detection of false positives via overcoming the limitations of multiple testing and enhancing the identification of underlying candidate genes which in turn facilitate gene-based association mapping. Several GWAS studies have been conducted in wheat for grain yield and related traits, which have also resulted in the identification of hundreds of high-confidence candidate genes governing yield-related traits [42, 43, 44, 45, 46]. Combined linkage analysis and joint linkage association mapping (JLAM) have also been used in wheat for genetic dissection of grain yield-related traits. Unlike meta-QTL analysis, meta-GWAS studies have been rare in wheat for yield and related traits. For the first time in 2018, Battenfield and his colleagues [47] described this meta-GWAS approach, which combined GWAS analysis from multi-year unbalanced breeding nurseries and identified the consensus and stable marker-trait associations (MTAs) and underlying candidate genes [47]. The markers, as well as candidate genes identified for grain yield and its component traits, provide important genomic resources for wheat breeding. These genomics resources can be immediately implemented to genomics-assisted breeding in wheat for genetic improvement of grain yield.

Advertisement

3. Gene cloning and allele mining: to be used for MAS

MAS allows a more effective selection of target genotypes which further enable certain traits to be ‘fast-tracked’, resulting in faster line development and variety release. MAS is a more cost-effective approach that can replace phenotyping and thereby allows selection in off-season nurseries as well. Another advantage of using MAS is that the total number of genotypes that need to be tested can be reduced significantly in early generations which allow more efficient use of field or glasshouse space which is generally limited [48]. MAS remains a valid option for major gene or QTL, whereas QTL cloning or gene cloning may become a more routine activity assisted by increased utilisation of high-throughput phenotyping techniques [49], sequencing [50], and identification of high-confidence candidate genes through ‘omics’ profiling [51]. Cloned QTL/gene may provide new opportunities for a more targeted search for novel alleles in wild wheat germplasm and mutants (Table 1).

Genes/QTLsChromosomeProducts/enzymesAssociated yield-related traitsReferences
TaSus22A, 2B, 2DSucrose synthaseEndosperm development[52]
TaCwi-A12ACell wall invertaseKernel weight[53]
TaCWI-5D5DCell wall invertaseKernel weight[54]
TaSAP1-A17AZinc-finger proteinThousand grain weight, number of grains per spike, spike length, peduncle length and spikelet’s per spike[55]
TaGS1a6DGlutamine synthetaseMineral nutrient and grain size[56]
TaTGW-7A7AIndole-3-glycerol-phosphate synthaseThousand grain weight[57]
TaGASR7-A17ASnakin/GASA proteinGrain length[58]
TaGS-D17DGlutamine synthetaseThousand grain weight, grain length[59]
TaCKX6a023DCytokinin oxidase/dehydrogenaseGrain size, grain filling rate, grain weight[60]
Tackx3ACytokinin oxidaseGrain weight and leaf chlorophyll content[61]
TaTPP-6AL16ATrehalose 6-phosphate phosphataseGrain weight[62]
TaFlo2-A12AFLO2 proteinThousand grain weight, grain size[63]
TaSnRK2.31A, 1B, 1DPlant-specific protein kinasePlant height, length of peduncle, penultimate node, thousand grain weight[64]
TaSnRK2.104A, 4B,4DSucrose non-fermenting 1-related protein kinasesThousand grain weight, spike length[65]
6-SFT-A24AFructan 6-fructosyltransferaseThousand grain weight[66]
TaGW2-6A6AE3 ubiquitin ligaseGrain weight, grain size[25]
TaCKX6-D13DCytokinin oxidase/dehydrogenaseThousand grain weight[67]
TaGL3-5A5APutative protein phosphataseGrain length[68]
TaAPO-A17AF-box protein of 429 amino acidsTotal spikelet number per spike[69]
TaTGW6-A13AIndole-3- acetic acid-glucose hydrolaseThousand grain weight[24]
TaGW8-B1a7BE3 ubiquitin ligaseKernel size[70]
TaTAR2.1-3A3ATryptophan amino transferasePlant height, spike number[71]
TaNAC2-5A5ANAC transcription factorSpike number, grain number per spike, and thousand grain weight[72]
TaGS5-3A3ASerine carboxypeptidasesGrain size, grain weight[73]
TaTEF-7A7ATranscript elongation factorGrain number[74]
TaPPH-A7APheophytin pheophorbide hydrolaseThousand grain weight, grain filling[75]
TaNF-YB43BHistone-like transcription factorNumber of spikes per plant[76]
TaNFYA-B16BHistone-like transcription factorNumber of spikes per plant[77]
TaCYP78A37A, 7B, 7DCytochrome P450 CYP78A3Seed size[78]

Table 1.

Cloned genes/QTLs regulating various yield-related traits in wheat.

At present, tremendous sequence information is available in public databases as a result of the sequencing of diverse wheat crop genomes, including reference lines and wild progenitors. This information can be used for mining the novel and superior alleles of agronomically important genes from gene pools to appropriately deploy for the development of high-yielding cultivars. Allele mining also provides insights into the molecular basis of trait variations and identifies the sequence variants associated with superior alleles. Moreover, it helps in the development of allele-specific molecular markers, assisting the introgression of novel alleles via MAS.

Advertisement

4. Transgenic technologies to boost grain yield

Considerable progress has been made in the past for manipulation of genes from diverse sources, including wild relatives and progenitors, and transferring them into wheat to confer increased grain yield, transgenesis can be employed as a powerful alternative for increasing the grain yield through exploiting the genes/traits which does not occur naturally in the wheat species. Transgenic plants refer to plants that contain a gene(s) that has been artificially inserted from an unrelated plant or a completely different species. The increase in grain yield potential through transgenesis involves an ideotypic detail of potential targets for transformation. In 2017, Nadolska-Orczyk and his colleagues [79] reported potential targets for transgenesis which can result in the increased grain yield in wheat. These include ‘transcription factors, regulating spike development, which mainly affect grain number; genes involved in metabolism or signalling of growth regulators—cytokinins, gibberellins, and brassinosteroids—which control plant architecture and consequently stem hardiness and grain yield; genes determining cell division and proliferation mainly impacting grain size; floral regulators influencing inflorescence architecture and consequently seed number, and genes involved in carbohydrate metabolism having an impact on plant architecture and grain yield’. Furthermore, modulated expression of flowering genes, which control vernalization and photoperiod-dependent floral induction, may be good for winter or spring wheat varieties [79, 80]. Besides, augmenting photosynthetic rates of laminar and non-laminar organs and the capability to access and utilise a greater amount of resources, such as nutrients or water, may also be potential targets for transgenesis in wheat for grain yield improvement [81, 82]. Besides, information about specific genotypes as well as climatic and agronomic conditions and consideration of the fact that the majority of the genes are members of multigene families is required for successful implementation of selected potential genes in breeding programs [79].

Transgenic wheat has the capacity to transform agriculture, but progress has been very limited as no transgenic wheat cultivar could be commercially approved so far because of consumers’ concerns. Few promising reports are available where newly developed transgenic wheat showed a significant grain yield advantage [72, 83]. Over-expression of a nitrate-inducible transcription factor (NAC TF) in wheat enhanced root growth and the ability to uptake nitrogen, therefore, increased nitrogen accumulation and grain yield by 10% (on a single plant basis) [72]. In another study, Gonzalez and his colleagues [83] reported that transgenic wheat lines carrying a mutated version of the sunflower TF (HaHB4) can significantly increase grain yield and water use efficiency across a range of environments [83]. Most recently in 2020, Argentina has become the first country to approve a genetically modified wheat variety (HB4). This is a drought-tolerant high-yielding wheat variety jointly developed by Argentine crop inputs manufacturer ‘Bioceres’ and ‘Trigall Genetics’ yielding 20% more than other standard wheat varieties in 10 years trials under drought conditions. The commercial approval of this GMO variety solely depends on approval by Brazil, which imports more than 85% of Argentine wheat [84]. Experts have also raised concerns about the growth and marketing of this GMO wheat variety, citing challenges related to food safety, consumer preferences, environmental effects, and socioeconomic issues. More research is required to determine the true safety of this GMO wheat and to decide, whether they are safe for both the consumers and the environment. At least, most would agree that the possible advantage of producing transgenic wheat, which furnishes the human population with cheaper and more food, makes transgenesis a useful invention.

Advertisement

5. Genome-editing technologies

Targeted genome editing has emerged as a powerful tool for studying gene function, correcting defective genes, or introducing novel functionality. Its mechanism involves sequence-specific double-strand breaks (DSBs) in the target DNA, with edits incorporated during the endogenous repair. In the earlier phase of genome editing, to induce the desired double-strand breaks at the target site, the engineering for zinc-finger nucleases (ZFNs) [85] or meganucleases [86] attracted the attention of the researcher community. These genome-editing systems needed specialised competence to produce artificial proteins consisting of customizable DNA-binding domains (sequence-specific), each linked to a non-specific nuclease for target DNA cleavage, and offered researchers with extraordinary tools to perform genetic manipulation. Later, the identification of a novel class of a Flavobacterium okeanokoites catalytic domain (FokI) derived from bacterial proteins termed transcription activator-like effectors (TALEs) further offered new possibilities for precisely targeted genome editing [87]. TALE-based programmable nucleases allowed the cleavage of any DNA sequence of interest with comparatively high frequency. Dimerization of FokI nuclease is needed to make an active nuclease, therefore, every time two modules need to be designed to target closely DNA sequences for generating DSBs at target sites. This dimerization requirement limited the use of these two powerful genomes-editing tools, as designing active nucleases was difficult and very expensive [88].

In 2012, an inexpensive, simple, easy to use, and effective genome-editing system that is clustered regularly interspaced short palindromic repeats (CRISPR)/CRISPR-associated protein 9 (CRISPR/Cas9) was introduced which revolutionised the field of genome editing [89]. The use of this powerful tool allows producing genome-edited plants in a very short period. CRISPR technology can be efficiently utilised for both precisely eliminating the negative regulator genes and augmenting the activity of positive-regulator genes that affect the trait of interest. Nevertheless, there are only a couple of reports available for validation of the CRISPR technique in wheat compared to other crops, such as rice [90]. In these reports, different genes were targeted by CRISPR/Cas9 to address the major biotic, and abiotic stresses along with improving a few agronomic traits in wheat [90]. An exciting advantage of using the CRISPR/Cas9 technology is the possibility of simultaneously editing multiple target genes using a single CRISPR construct. For instance, Wang and his colleagues [91] practiced this multiplexed genome editing in hexaploid wheat for targeting three different genes viz. TaLpx-1, TaGW2, and TaMLO. They placed three sgRNAs (each specific to a different gene) in a tRNA polycistronic cassette under the control of a single promoter to produce knockouts. Multiplex genome-editing tools can be efficiently utilised to address more complex traits (such as grain yield) involving multiple genes in a single attempt [91]. Moreover, this CRISPR/Cas9 mediated multiplex genome editing can also be utilised to mimic the domestication process during evolution in a short time frame, with implications for a convenient and rapid generation of high-yielding wheat varieties. Despite the several advantages of using CRISPR/Cas9, one of the prominently associated challenges is off-target effects, that is undesired mutations at unintended sites induced by genome editing. Various methods have been developed to find off-target mutations both in vitro and in vivo. These include SITE-seq [92], Digenome-seq [93], CIRCLE-seq [94], GUIDE-seq [95], and DISCOVER-seq [96]. In the same way, the engineering of Cas9 proteins has also been performed to enhance the specificity.

5.1 Base editors and prime editing: opening up new avenues for wheat genome engineering

Many crucial agronomic traits are determined by a few base changes or point mutations in a gene [97, 98, 99]. CRISPR/Cas9 mediated gene replacements or gene modifications through homology-directed repair (HDR) has been reported as a practicable approach to correct the point mutations in the target DNA/gene and has the capability for accelerating crop improvement [100, 101]. Yet, the low efficiency of template DNA delivery and the rare occurrence of HDR (endogenous) has always been a difficult task in attaining success in plants. Furthermore, the CRISPR/Cas9 system is amenable for gene knock-in or knock-out, but cannot covert base into another. These challenges highlighted the demand for alternative powerful approaches that can result in precise and stable genome editing in crops. In 2016, a novel approach that is ‘Base editing’ was emerged which allows precise base (nucleotide) substitutions in a programmable manner, without requiring a donor template or disruption of a gene [102]. A base editor is a fusion of catalytically inactive Cas9 domain (Cas9 variants, Cas9 nickase, or dCas9) and an adenosine or cytosine domain that converts one base to another. Nucleotide substitutions or single-base changes may generate elite trait variations in crops which assist in accelerating crop improvement. The base-editing system can revert an SNP or single-base change without gene disruption. In recent years, many adenine and cytosine base editors have emerged as powerful tools for precise genome modifications (A to G or C to T) in eukaryotic genomes [102]. The potential of this approach has been demonstrated in several crops, including wheat [103, 104, 105, 106]. As aforementioned, HDR efficiency is comparatively low in plant cells, so knock-ins of DNA fragments to target sites are challenging. Recently in 2019, Anzalone and co-workers developed a more efficient genome-editing technology that is ‘Prime editing’ which consists of CRISPR-Cas9 nickase–reverse transcriptase fusions programmed with pegRNAs (prime-editing guide RNAs) that enable precise genome editing without inducing DSBs or requiring a donor DNA template (mandatory for genome editing via HDR) in mammalian cells [107]. The prime editors have been adapted for use in wheat via optimization of the codon, promoter, and editing conditions [108]. This optimised suite of prime editors enabled InDels and point mutations in wheat and rice at higher frequencies [108]. Development of new technologies and tools, newly discovered CRISPR/Cas systems, are being continuously reported, inferring that the CRISPR toolbox for wheat genome engineering would expand further in the near future. Researchers have also focused on the development of efficient approaches for eliminating transgenes from genome-edited plants, such as (a) transient expression of DNA and RNA [109], (b) use of CRISPR/Cas9 ribonucleoprotein complexes [110], (c) use of CRISPR-S—an active interference element [111], and (d) programmed self-elimination of the CRISPR/Cas9 constructs [112] to generate transgene-free genome-edited plants. The elimination of transgenes offers the following two advantages—(i) elimination of Cas9 construct from genome-edited plants prevents the induction of genetic changes at undesired loci, (ii) elimination of the transgenes is likely a prerequisite for getting regulatory approval of genome-edited crops for commercial applications. In the future, CRISPR technology may be supposed to accelerate wheat biology research, ultimately facilitating the development of high-yielding wheat varieties.

Advertisement

6. Genomic selection for grain yield improvement

The genetic complexity of grain yield and other yield-related traits limit the power of QTL mapping and association mapping in identifying small effect loci [113]. A powerful breeding strategy that is genomic selection (GS) has been introduced to circumvent this problem which implements whole-genome markers for predictions, and thus can efficiently complement QTL mapping and association analysis in dissecting the complex genetic base of grain yield-related traits in wheat [114, 115]. High-throughput/next-generation genotyping technologies have accelerated the adoption of GS by enabling the development of large sets of DNA marker data at reasonable costs [116]. GS is a potential GAB tool that predicts genomic-estimated breeding values (GEBVs) of individuals (from the breeding population) with genotypic data available via prediction models constructed based on a training population (TP) with available phenotypic and genotypic information [117]. As aforementioned, using the prediction models, the GEBVs of unobserved individuals are predicted, circumventing the omission of the small-effect genomic region (markers) that would fail a threshold (significance) test. Though the effect of each marker is small, a large volume of genotypic information covering the whole genome still has the power to explain all the genetic variance. GS complements conventional breeding approaches and can potentially decrease the requirement of large-scale phenotyping and hasten the rate of genetic gain via shorter breeding cycles [118, 119]. The performance of GS relies mainly on the prediction accuracy, defined as the ‘Pearson’s correlation between the selection criterion and the true breeding value to select individuals with unknown phenotypes’ [120, 121]. Other factors that affect the GS accuracy include gene effects, level of linkage disequilibrium (LD), statistical models, the genetic composition of the TP, relationship between validation population (VP) or selection individuals and TP, and heritability of the target traits [120]. The major objective of GS is to decrease the cost of phenotyping and hasten genetic gains, use of high-throughput phenotyping tools and platforms that enable high-density phenotyping of hundreds to thousands of individuals across time and space using proximal or remote sensing, can increase the intensity and accuracy of selection and, eventually the selection response, as well as reduce phenotyping costs. The main idea of high-throughput phenotyping is to exploit secondary traits, such as canopy temperature, and green normalised difference vegetation index (NDVI) are closely related to grain yield that may be advantageous in early-generation testing of individuals. Data recorded on secondary traits (genetically correlated to grain yield) can be incorporated in multivariate pedigree and GS models, improving indirect selection for GY [122, 123, 124]. Moreover, GS can also be applied to gene bank accessions for germplasm enhancement. Accessions stored in germplasm bank represents an under-exploited rich genetic resource for wheat breeders, superior alleles can be extracted from these accessions which may be exploited for grain yield improvement in wheat [125, 126]. In general, lengthy pre-breeding programs are needed to develop lines that possess favourable alleles/genes from the wild accessions with superior agronomic performance and that may be utilised as parents in breeding programs. Using GS, germplasm enhancement breeding programs can be directly started using wild accessions and landraces. In a recent GS-based study, NGS technologies with multi-environment phenotyping were used to study the contribution of exotic genomes to 984 pre-breeding lines. Significant positive contributions of exotic germplasm to pre-breeding lines derived from crosses of CIMMYT’s best elite lines with exotics were reported [127]. Genomic selection studies conducted in wheat for grain yield and related traits are presented in Table 1. The prediction accuracy of GS for different grain yield-related traits has varied from 0 to 0.98% in wheat (Table 2).

Population type and size*Number of genotyped markersTraitsAccuracy of GEBV usedReferences
Advanced breeding lines from CIMMYT (254)41,371 GBS-SNPsTGW, DTH, and GY0.28–0.45[128]
Two DH populations (165 and 159)1975 and 1483 SNPs (90K SNP)GNPS0.10–0.42[129]
European winter wheat lines (2325)12,642 SNPs (9K SNP)GY0.5–0.65[130]
Winter wheat population (273)40,267 SNPs (90K SNP)GY, TGW, PH and DTH0.33–0.67[131]
Inbred breeding lines (557)12,083 GBS-SNPsDTH and GY0.57[132]
Advanced elite spring wheat lines (287)15,000 SNPs (90 K SNP)GY, TGW and GN0.38–0.63[133]
Lines from multiple families (659)9500 DArT-GBS-SNPsGY0.38–0.41[134]
Winter wheat breeding population from multiple families (861)6600 DArT-GBS-SNPsGY0.39–0.48[135]
Inbred breeding lines (557)12,083 GBS-SNPsGY0.65–0.76[136]
Hybrids obtained by crossing 18 males and 667 females (1888)13,005 SNPs (90 K and 15 K)GY, DTH and PH0.5–0.55[137]
Winter wheat lines (1100)27,000 GBS-SNPsGY0.23–0.55[138]
European winter and spring cultivars (210)GBS-SNPs44 spike morphology traits0.2–0.5[139]
Elite wheat lines (4368)2038 GBS-SNPsDTH, DTM, PH and GY0.35–0.44[140]
Bread wheat lines (10375)18,101 GBS-SNPsGY and TGW0.59–0.98[141]
Double haploid lines (282)7426 GBS-SNPsGY and TGW0.47–0.54[142]
Bread wheat lines (3771)8519 GBS-SNPsDTH, DTM and GY0–0.75[143]
Soft red winter wheat lines (239), Double haploid (100), and Recombinant inbred lines (156)2721 SNPs (9 and 90K)GY, DTH, TGW, GNPS, and PH− 0.14-0.43[144]
F4:6 generation and double haploid winter wheat breeding lines (1114)7300 DArT-GBS-SNPsGY0.45[145]
Winter wheat lines (3282)18,728 GBS-SNPsGY0.25[122]
>6400 breeding lines78,662 GBS-SNPsGY0.41[146]
Advanced breeding lines (456)11,089 GBS-SNPsGY0.33–0.66[147]
Association mapping panel (456), two F5 populations (61 and 501), two DH populations (447 and 759)16,233 GBS-SNPsGY0.21[148]
Advanced bread wheat lines (4302)8443 GBS-SNPsGY0.35–0.43[149]
Winter wheat lines (1325)11,154 SNPs (15 K)GY0.57[150]

Table 2.

Genomic selection studies conducted in wheat for grain yield and related traits.

Figures in parenthesis are the population size.


GY, GNPS, DTH, DTM, PH, and TGW refer to grain yield, grain number per spike, days to heading, days to maturity, plant height, and thousand grain weight, respectively.

6.1 GWAS-assisted GS: making GS more efficient

As discussed above, GWAS estimates marker effects throughout the genome on the target association panel (diverse germplasm) based on prediction models. Based on LD, GWAS may identify new functional variants, including novel MTAs and genes for many agronomically important traits in diverse germplasm. According to a comprehensive simulation study in plants, the use of a few major MTAs/QTLs/genes (each explaining ≥ 10% of the phenotypic variance) as fixed effects in GS models can increase the accuracy of GS for complex quantitative traits [151]. Although, the potential to combine robust and consistent associations identified from GWAS as fixed effects in GS models to increase prediction accuracy for complex traits such as grain yield has not been investigated comprehensively in wheat. The first report of integrating the genetic architecture of GY (revealed through GWAS) into prediction models in wheat has come from the work by Sehgal and co-workers, most recently in 2020 [149]. Firstly, using a haplotype-based genome-wide association study, they identified 58 MTAs for GY. Out of these 58 MTAs, 16 were ‘environment-specific’ with large effects and eight MTAs were consistent across trials and environments. These consistent MTAs were then used as fixed effects in the prediction models which resulted in a 9–10% increase in prediction accuracy for GY [149]. It is suggested that the utility of GS incorporating GWAS results may be noteworthy for GY when GWAS results detect highly robust and significant genomic regions.

Advertisement

7. Haplotype-based breeding (HBB) for grain yield improvement

Due to low heritability and persistent ‘genotype × environment’ interactions, improving grain yield (GY) is a difficult task for the global plant breeding community, especially under stressful environmental conditions [152, 153, 154]. As discussed earlier, GWAS-assisted GS has proven to be an effective method for deciphering the genetic architecture of complex traits, population improvement, and the development of better varieties with a higher yield. However, the problem of ‘missing heritability’, which is widespread in single marker-based GWAS, is not addressed by this approach. The alternative approach to boost the power of GWAS is by constructing haplotypes between neighbouring SNPs on a chromosome. As specific sets of alleles are observed on a single chromosome, haplotypes are inherited jointly with the limited probability of contemporaneous recombination. Haplotypes are implemented in crop improvement in two ways—retrospective and prospective [155]. Plant breeders have to choose the advantageous haplotypes that lead to desirable phenotype(s) for the trait(s) of interest during the long-term selection process. As a result, these advantageous haplotypes in elite crop germplasm can be found utilising the genome resequencing technique to sequence an elite gene pool [156]. Later, molecular markers that characterise these beneficial haplotypes can be produced, and all of these haplotype-defining markers can then be utilised to pick the most ideal combination of haplotypes that govern a certain phenotype. Furthermore, by identifying lines with unique recombination in chromosomal blocks of relevance, these haplotype-related markers can be utilised to distinguish between favourable and unfavourable genetic variation. On the other hand, haplotypes can be employed in a prospective approach, in which a vast collection of ancestral and wild germplasm of specific crop species (not just elite breeding pools) is re-sequenced to find haplotypes with a wider range of genetic variation [153, 155]. The genome-wide haplotypes are employed in this strategy to find novel haplotypes in a wide variety of natural germplasm. For the discovery of QTLs/genes, recent GWA studies based on empirical and simulation data (i.e., better p-values) and allelic effect estimation have demonstrated that haplotype blocks have higher mapping accuracy and power than individual SNPs [153, 155, 156, 157, 158, 159, 160]. Haplotype superiority can be explained by a number of factors. Stephens and his colleagues [161] showed that haplotype blocks are more informative than SNP markers because of their multi-allelic character in nature. The scientists found that haplotype variants were more common than SNPs, implying that recombination and recurrent mutation events occurred within and among haplotype genes (Figure 1). In addition, as compared to individual SNPs, haplotype-based analysis is predicted to reduce the false positives and shows the intricate mechanism of causal haplotypes [162]. Similarly, the haplotype-assisted GS depicts the complex relationships between genotypic information and phenotypes more accurately than individual SNPs. As a result, this method could eventually aid in improving selection gain per unit of time. Because haplotypes can better capture LD and genomic similarities in various lines and may capture local high-order allelic interactions, they may improve the accuracy of genomic prediction [163]. Furthermore, by depicting population structure in the calibration set, prediction accuracy might be enhanced. The superiority of haplotype-based predictions over SNP-based predictions for all studied traits, including yield, test weight, and protein content, was established in a recent GS study that compared the prediction ability computed from haplotypes and SNPs in a set of 383 advanced lines and cultivars of wheat [164]. Based on evidence revealing higher haplotype-assisted genomic prediction efficiency than SNPs, researchers are increasingly embracing haplotype-assisted genomic prediction in crop development programmes.

Figure 1.

Flow diagram indicating how haplotype-based GWAS and haplotype-based GS, when combined with high-throughput genotyping, have the potential to improve gene identification precision and accuracy (modified from Bhat et al. [162]).

Advertisement

8. Conclusions

Significant progress has been made in wheat in developing various genomics resources, including high-throughput molecular markers, dense genetic maps, and next-generation genotyping platforms. The availability of high-quality wheat genome information has also enabled many next-generation sequencing-based approaches for genetic mapping, allele mining, and identification of candidate genes which have enhanced the precision, pace, and efficiency of trait mapping. At present, trait-associated markers, high-throughput genotyping platforms, and expertise are available for deploying genomics-assisted breeding in wheat. We believe that in the coming years, extensive deployment of genome editing, transgenic technology, genomic selection, haplotype-based breeding in combination or alone would be undertaken for crop improvement and breaking the yield ceiling. Various steps involved in generating high-yielding wheat genotypes using genomics-assisted breeding technologies are represented in Figure 2.

Figure 2.

Flowchart demonstrating the steps involved in generating high-yielding wheat genotypes using different genomics-assisted breeding strategies.

Advertisement

Acknowledgments

Thanks are due to the Head, Department of Molecular Biology and Genetic Engineering, G. B. Pant University of Agriculture and Technology, Pantnagar, (India) for providing necessary facilities.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

Funding

This research did not receive any specific grant from funding agencies in the public, commercial, or not-for-profit sectors.

References

  1. 1. FAO. Food and Agricultural Organization [Internet]. 2018. Available from: http://www.fao.org/faostat/en/#data/QC [Accessed: January 21, 2022]
  2. 2. Knoema [Internet] 2020. Available from: https://knoema.com/atlas [Accessed: January 21, 2022]
  3. 3. Graybosch RA, Peterson CJ. Genetic improvement in winter wheat yields in the Great Plains of North America, 1959-2008. Crop Science. 2010;50:1882-1890
  4. 4. Calderini DF, Reynolds MP, Slafer GA. Genetic Gains in Wheat Yield and Main Physiological Changes Associated with them during the 20th Century. New York: Food Product Press; 1999. pp. 352-377
  5. 5. De Vita P, Nicosia OLD, Nigro F, Platani C, Riefolo C, Di Fonzo N, et al. Breeding progress in morpho-physiological, agronomical and qualitative traits of durum wheat cultivars released in Italy during the 20th century. European Journal of Agronomy. 2007;26:39-53
  6. 6. Reynolds MP, Rajaram S, Sayre KD. Physiological and genetic changes of irrigated wheat in the post–green revolution period and approaches for meeting projected global demand. Crop Science. 1999;39:1611-1621
  7. 7. Sadras VO, Lawson C. Genetic gain in yield and associated changes in phenotype, trait plasticity and competitive ability of South Australian wheat varieties released between 1958 and 2007. Crop and Pasture Science. 2011;62:533-549
  8. 8. Fischer RA, Edmeades GO. Breeding and cereal yield progress. Crop Science. 2010;50:S-85
  9. 9. Iwgsc AR, Eversole K, Feuillet C, Keller B, Rogers J, Stein N. Shifting the limits in wheat research and breeding using a fully annotated reference genome. Science. 2018;361:7191
  10. 10. Saini DK, Devi P, Kaushik P. Advances in genomic interventions for wheat biofortification: A review. Agronomy. 2020;10:1-62
  11. 11. Kaur B, Mavi GS, Gill MS, Saini DK. Utilization of KASP technology for wheat improvement. Cereal Research Communications. 2020;48:1-13
  12. 12. Scheben A, Batley J, Edwards D. Genotyping-by-sequencing approaches to characterize crop genomes: Choosing the right tool for the right application. Plant Biotechnology. 2016;15:149-161
  13. 13. Cavanagh CR, Chao S, Wang S, Huang BE, Stephen S, Kiani S, et al. Genome-wide comparative diversity uncovers multiple targets of selection for improvement in hexaploid wheat landraces and cultivars. Proceedings of the National Academy of Sciences. 2013;110:8057-8062
  14. 14. Wang S, Wong D, Forrest K, Allen A, Chao S, Huang BE, et al. Characterization of polyploid wheat genomic diversity using a high-density 90 000 single nucleotide polymorphism array. Plant Biotechnology Journal. 2014;12:787-796
  15. 15. Boeven PH, Longin CFH, Leiser WL, Kollers S, Ebmeyer E, Würschum T. Genetic architecture of male floral traits required for hybrid wheat breeding. Theoretical and Applied Genetics. 2016;129:2343-2357
  16. 16. Winfield MO, Allen AM, Burridge AJ, Barker GL, Benbow HR, Wilkinson PA, et al. High-density SNP genotyping array for hexaploid wheat and its secondary and tertiary gene pool. Plant Biotechnology Journal. 2016;14:1195-1206
  17. 17. Allen AM, Winfield MO, Burridge AJ, Downie RC, Benbow HR, Barker GL, et al. Characterization of a wheat breeders’ Array suitable for high-throughput SNP genotyping of global accessions of hexaploid bread wheat (Triticum aestivum). Plant Biotechnology Journal. 2017;15:390-401
  18. 18. Rasheed A, Xia X. From markers to genome-based breeding in wheat. Theoretical and Applied Genetics. 2019;132:767-784
  19. 19. Savadi S, Prasad P, Kashyap PL, Bhardwaj SC. Molecular breeding technologies and strategies for rust resistance in wheat (Triticum aestivum) for sustained food security. Plant Pathology. 2018;67:771-791
  20. 20. He ZH, Liu L, Xia XC, Liu JJ, Pena RJ. Composition of HMW and LMW glutenin subunits and their effects on dough properties, pan bread, and noodle quality of Chinese bread wheat. Cereal Chemistry. 2005;82:345-350
  21. 21. Tian XL, Zhu ZW, Xie L, Xu DA, Li JH, Fu C, et al. Preliminary exploration of the source, spread and distribution of Rht24 reducing height in bread wheat. Crop Science. 2019;59:19-24
  22. 22. Zhang XK, Yang SJ, Zhou Y, Xia XC, He ZH. Distribution of Rht-B1b, Rht-D1b and Rht8 genes in autumn-sown Chinese wheats detected by molecular markers. Euphytica. 2006;152:109-116
  23. 23. Zhou Y, He ZH, Sui XX, Xia XC, Zhang XK, Zhang GS. Genetic improvement of grain yield and associated traits in the northern China winter wheat region from 1960 to 2000. Crop Science. 2007;47:245-253
  24. 24. Hanif M, Gao F, Liu J, Wen W, Zhang Y, Rasheed A, et al. TaTGW6-A1, an ortholog of rice TGW6, is associated with grain weight and yield in bread wheat. Molecular Breeding. 2016;36:1
  25. 25. Su Z, Hao C, Wang L, Dong Y, Zhang X. Identification and development of a functional marker of TaGW2 associated with grain weight in bread wheat (Triticum aestivum L.). Theoretical and Applied Genetics. 2011;122:211-223
  26. 26. Chen Z, Cheng X, Chai L, Wang Z, Bian R, Li J, et al. Dissection of genetic factors underlying grain size and fine mapping of QTgw.cau-7D in common wheat (Triticum aestivum L.). Theoretical and Applied Genetics. 2020a;133:149-162
  27. 27. Chen Z, Cheng X, Chai L, Wang Z, Du D, Wang Z, Bian R, Zhao A, Xin M, Guo W, Hu Z. Pleiotropic QTL influencing spikelet number and heading date in common wheat (Triticum aestivum L.). Theoretical and Applied Genetics. 2020b;133:1-14
  28. 28. Hu J, Wang X, Zhang G, Jiang P, Chen W, Hao Y, et al. QTL mapping for yield-related traits in wheat based on four RIL populations. Theoretical and Applied Genetics. 2020;133:917-933
  29. 29. Xin F, Zhu T, Wei S, Han Y, Zhao Y, Zhang D, et al. QTL mapping of kernel traits and validation of a major QTL for kernel length-width ratio using SNP and bulked segregant analysis in wheat. Scientific Reports. 2020;10:1-12
  30. 30. Yu M, Liu ZH, Yang B, Chen H, Zhang H, Hou DB. The contribution of photosynthesis traits and plant height components to plant height in wheat at the individual quantitative trait locus level. Scientific Reports. 2020;10:1-10
  31. 31. Liu H, Mullan D, Zhang C, Zhao S, Li X, Zhang A, et al. Major genomic regions responsible for wheat yield and its components as revealed by meta-QTL and genotype–phenotype association analyses. Planta. 2020;252:1-22
  32. 32. Jaganathan D, Bohra A, Thudi M, Varshney RK. Fine mapping and gene cloning in the post-NGS era: Advances and prospects. Theoretical and Applied Genetics. 2020;133:1-20
  33. 33. Holland JB. Genetic architecture of complex traits in plants. Current Opinion in Plant Biology. 2007;10:156-161
  34. 34. Goffinet B, Gerber S. Quantitative trait loci: A meta-analysis. Genetics. 2000;155:463-473
  35. 35. Zhang LY, Liu DC, Guo XL, Yang WL, Sun JZ, Wang DW, et al. Genomic distribution of quantitative trait loci for yield and yield-related traits in common wheat. Journal of Integrative Plant Biology. 2010;52:996-1007
  36. 36. Tyagi S, Mir RR, Balyan HS, Gupta PK. Interval mapping and meta-QTL analysis of grain traits in common wheat (Triticum aestivum L.). Euphytica. 2015;201:367-380
  37. 37. Saini DK, Srivastava P, Pal N, Gupta PK. Meta-QTLs, ortho-meta-QTLs and candidate genes for grain yield and associated traits in wheat (Triticum aestivum L.). Theoretical and Applied Genetics. 2022;5:1-33
  38. 38. Saini DK, Chopra Y, Pal N, Chahal A, Srivastava P, Gupta PK. Meta-QTLs, ortho-MQTLs and candidate genes for nitrogen use efficiency and root system architecture in bread wheat (Triticum aestivum L.). Physiology and Molecular Biology of Plants. 2021a;27:2245-2267
  39. 39. Saini DK, Chahal A, Pal N, Srivastava P, Gupta PK. Meta-analysis reveals consensus genomic regions associated with multiple disease resistance in wheat (Triticum Aestivum L.). 2021;42:11-35. DOI: 10.21203/rs.3.rs-773587/v1
  40. 40. Pal N, Saini DK, Kumar S. Meta-QTLs, ortho-MQTLs and candidate genes for the traits contributing to salinity stress tolerance in common wheat (Triticum aestivum L.). Physiology and Molecular Biology of Plants. 2021;24:1-20
  41. 41. Kumar S, Singh VP, Saini DK, Sharma H, Saripalli G, Kumar S, et al. Meta-QTLs, ortho-MQTLs, and candidate genes for thermotolerance in wheat (Triticum aestivum L.). Molecular Breeding. 2021;41:1-22
  42. 42. Akram S, Arif MAR, Hameed A. A GBS-based GWAS analysis of adaptability and yield traits in bread wheat (Triticum aestivum L.). Journal of Applied Genetics. 2020;62:1-15
  43. 43. Basile SML, Ramírez IA, Crescente JM, Conde MB, Demichelis M, Abbate P, et al. Haplotype block analysis of an Argentinean hexaploid wheat collection and GWAS for yield components and adaptation. BMC Plant Biology. 2019;19:553
  44. 44. Gupta PK, Balyan HS, Sharma S, Kumar R. Genetics of yield, abiotic stress tolerance and biofortification in wheat (Triticum aestivum L.). Theoretical and Applied Genetics. 2020;133:1-34
  45. 45. Muhammad A, Hu W, Li Z, Li J, Xie G, Wang J, et al. Appraising the genetic architecture of kernel traits in hexaploid wheat using GWAS. International Journal of Molecular Sciences. 2020;21:5649
  46. 46. Li F, Wen W, Liu J, Zhang Y, Cao S, He Z, et al. Genetic architecture of grain yield in bread wheat based on genome-wide association studies. BMC Plant Biology. 2019;19:168
  47. 47. Battenfield SD, Sheridan JL, Silva LD, Miclaus KJ, Dreisigacker S, Wolfinger RD, et al. Breeding-assisted genomics: Applying meta-GWAS for milling and baking quality in CIMMYT wheat breeding program. PLoS One. 2018;13:0204757
  48. 48. Collard BC, Mackill DJ. Marker-assisted selection: An approach for precision plant breeding in the twenty-first century. Philosophical Transactions of the Royal Society B: Biological Sciences. 2008;363:557-572
  49. 49. Araus JL, Cairns JE. Field high-throughput phenotyping: The new crop breeding frontier. Trends in Plant Science. 2014;19:52-61
  50. 50. Berkman PJ, Lai K, Lorenc MT, Edwards D. Next-generation sequencing applications for wheat crop improvement. American Journal of Botany. 2012;99:365-371
  51. 51. Shah T, Xu J, Zou X, Cheng Y, Nasir M, Zhang X. Omics approaches for engineering wheat production under abiotic stresses. International Journal of Molecular Sciences. 2018;19:2390
  52. 52. Jiang Q, Hou J, Hao C, Wang L, Ge H, Dong Y, et al. The wheat (T. aestivum) sucrose synthase 2 gene (TaSus2) active in endosperm development is associated with yield traits. Functional & Integrative Genomics. 2011;11:49-61
  53. 53. Ma D, Yan J, He Z, Wu L, Xia X. Characterization of a cell wall invertase gene TaCwi-A1 on common wheat chromosome 2A and development of functional markers. Molecular Breeding. 2012;2943-52
  54. 54. Jiang Y, Jiang Q, Hao C, Hou J, Wang L, Zhang H, et al. A yield-associated gene TaCWI, in wheat: Its function, selection and evolution in global breeding revealed by haplotype analysis. Theoretical and Applied Genetics. 2015;128:131-143
  55. 55. Chang J, Zhang J, Mao X, Li A, Jia J, Jing R. Polymorphism of TaSAP1-A1 and its association with agronomic traits in wheat. Planta. 2013;237:1495-1508
  56. 56. Guo Y, Sun J, Zhang G, Wang Y, Kong F, Zhao Y, et al. Haplotype, molecular marker and phenotype effects associated with mineral nutrient and grain size traits of TaGS1a in wheat. Field Crops Research. 2013;154:119-125
  57. 57. Hu MJ, Zhang HP, Liu K, Cao JJ, Wang SX, Jiang H, et al. Cloning and characterization of TaTGW-7A gene associated with grain weight in wheat via SLAF-seq-BSA. Frontiers in Plant Science. 2016;7:1902
  58. 58. Dong L, Wang F, Liu T, Dong Z, Li A, Jing R, et al. Natural variation of TaGASR7-A1 affects grain length in common wheat under multiple cultivation conditions. Molecular Breeding. 2014;34:947
  59. 59. Zhang Y, Liu J, Xia X, He Z. TaGS-D1, an ortholog of rice OsGS3, is associated with grain weight and grain length in common wheat. Molecular Breeding. 2014;34:1097-1107
  60. 60. Lu J, Chang C, Zhang HP, Wang SX, Sun G, Xiao SH, et al. Identification of a novel allele of TaCKX6a02 associated with grain size, filling rate and weight of common wheat. PLoS One. 2015;10:0144765
  61. 61. Chang C, Lu J, Zhang HP, Ma CX, Sun G. Copy number variation of cytokinin oxidase gene Tackx4 associated with grain weight and chlorophyll content of flag leaf in common wheat. PLoS One. 2015;10:0145970
  62. 62. Zhang P, He Z, Tian X, Gao F, Xu D, Liu J, et al. Cloning of TaTPP-6AL1 associated with grain weight in bread wheat and development of functional marker. Molecular Breeding. 2017a;37:1-8
  63. 63. Sajjad M, Ma X, Khan SH, Shoaib M, Song Y, Yang W, et al. TaFlo2-A1, an ortholog of rice Flo2, is associated with thousand grain weight in bread wheat (Triticum aestivum L.). BMC Plant Biology. 2017;17:1-11
  64. 64. Miao L, Mao X, Wang J, Liu Z, Zhang B, Li W, et al. Elite haplotypes of a protein kinase gene TaSnRK2.3 associated with important agronomic traits in common wheat. Frontiers. Plant Science. 2017;8:368
  65. 65. Zhang ZG, Lv GD, Li B, Wang JJ, Zhao Y, Kong FM, et al. Isolation and characterization of the TaSnRK2.10 gene and its association with agronomic traits in wheat (Triticum aestivum L.). PLoS One. 2017b;12:0174425
  66. 66. Yue A, Li A, Mao X, Chang X, Li R, Jing R. Identification and development of a functional marker from 6-SFT-A2 associated with grain weight in wheat. Molecular Breeding. 2015;35:63
  67. 67. Zhang L, Zhao YL, Gao LF, Zhao GY, Zhou R, Zhang BS, et al. TaCKX6-D1, the ortholog of rice OsCKX2, is associated with grain weight in hexaploid wheat. New Phytologist. 2012;195:574-584
  68. 68. Yang J, Zhou Y, Wu Q, Chen Y, Zhang P, Zhang YE, et al. Molecular characterization of a novel TaGL3-5A allele and its association with grain length in wheat (Triticum aestivum L.). Theoretical and Applied Genetics. 2019a;132:1799-1814
  69. 69. Muqaddasi QH, Brassac J, Koppolu R, Plieske J, Ganal MW, Röder MS. TaAPO-A1, an ortholog of rice ABERRANT PANICLE ORGANIZATION 1, is associated with total spikelet number per spike in elite European hexaploid winter wheat (Triticum aestivum L.) varieties. Scientific Reports. 2019;9:1-12
  70. 70. Yan X, Zhao L, Ren Y, Dong Z, Cui D, Chen F. Genome-wide association study revealed that the TaGW8 gene was associated with kernel size in Chinese bread wheat. Scientific Reports. 2019;9:1-10
  71. 71. Shao A, Ma W, Zhao X, Hu M, He X, Teng W, et al. The auxin biosynthetic TRYPTOPHAN AMINOTRANSFERASE RELATED TaTAR2.1-3A increases grain yield of wheat. Plant Physiology. 2017;174:2274-2288
  72. 72. He X, Qu B, Li W, Zhao X, Teng W, Ma W, et al. The nitrate-inducible NAC transcription factor TaNAC2-5A controls nitrate response and increases wheat yield. Plant Physiology. 2015;169:1991-2005
  73. 73. Ma L, Li T, Hao C, Wang Y, Chen X, Zhang X. TaGS 5-3A, a grain size gene selected during wheat improvement for larger kernel and yield. Plant Biotechnology Journal. 2016;14:1269-1280
  74. 74. Zheng J, Liu H, Wang Y, Wang L, Chang X, Jing R, et al. TEF-7A, a transcript elongation factor gene, influences yield-related traits in bread wheat (Triticum aestivum L.). Journal of Experimental Botany. 2014;65:5351-5365
  75. 75. Wang H, Wang S, Chang X, Hao C, Sun D, Jing R. Identification of TaPPH-7A haplotypes and development of a molecular marker associated with important agronomic traits in common wheat. BMC Plant Biology. 2019;19:1-2
  76. 76. Yadav D, Shavrukov Y, Bazanova N, Chirkova L, Borisjuk N, Kovalchuk N, et al. Constitutive overexpression of the TaNF-YB4 gene in transgenic wheat significantly improves grain yield. Journal of Experimental Botany. 2015;66:6635-6650
  77. 77. Qu B, He X, Wang J, Zhao Y, Teng W, Shao A, et al. A wheat CCAAT box-binding transcription factor increases the grain yield of wheat with less fertilizer input. Plant Physiology. 2015;167:411-423
  78. 78. Ma M, Wang Q, Li Z, Cheng H, Li Z, Liu X, et al. Expression of TaCYP78A3, a gene encoding cytochrome P450-CYP78A3 protein in wheat (Triticum aestivum L.), affects seed size. The Plant Journal. 2015;83:312-325
  79. 79. Nadolska-Orczyk A, Rajchel IK, Orczyk W, Gasparis S. Major genes determining yield-related traits in wheat and barley. Theoretical and Applied Genetics. 2017;130:1081-1098
  80. 80. Araus JL, Serret MD, Lopes MS. Transgenic solutions to increase yield and stability in wheat: Shining hope or flash in the pan? Journal of Experimental Botany. 2019;70:1419-1424
  81. 81. Sivamani E, Bahieldin A, Wraith JM, Al-Niemi T, Dyer WE, Ho THD, et al. Improved biomass productivity and water use efficiency under water deficit conditions in transgenic wheat constitutively expressing the barley HVA1 gene. Plant Science. 2000;155:1-9
  82. 82. Kulkarni M, Soolanayakanahally R, Ogawa S, Uga Y, Selvaraj MG, Kagale S. Drought response in wheat: Key genes and regulatory mechanisms controlling root system architecture and transpiration efficiency. Frontiers in Chemistry. 2017;5:106
  83. 83. González FG, Capella M, Ribichich KF, Curín F, Giacomelli JI, Ayala F, et al. Field-grown transgenic wheat expressing the sunflower gene HaHB4 significantly outyields the wild type. Journal of Experimental Botany. 2019;70:1669-1681
  84. 84. Bioceres Crop Solutions [Internet]. 2020. Available from: https://investors.biocerescrops.com/news/news-details/2020/Bioceres-Crop-Solutions-Corp.Announces-Regulatory-Approval-of-Drought-Tolerant-HB4-Wheat-in-Argentina/default.aspx [Accessed: December 19, 2020]
  85. 85. Urnov FD, Rebar EJ, Holmes MC, Zhang HS, Gregory PD. Genome editing with engineered zinc finger nucleases. Nature Reviews Genetics. 2010;11:636-646
  86. 86. Silva G, Poirot L, Galetto R, Smith J, Montoya G, Duchateau P, et al. Meganucleases and other tools for targeted genome engineering: Perspectives and challenges for gene therapy. Current Gene Therapy. 2011;11:11-27
  87. 87. Joung JK, Sander JD. TALENs: A widely applicable technology for targeted genome editing. Nature Reviews Molecular Cell Biology. 2013;14:49-55
  88. 88. Osakabe Y, Osakabe K. Genome editing with engineered nucleases in plants. Plant and Cell Physiology. 2015;56:389-400
  89. 89. Jinek M, Chylinski K, Fonfara I, Hauer M, Doudna JA, Charpentier E. A programmable dual-RNA–guided DNA endonuclease in adaptive bacterial immunity. Science. 2012;337:816-821
  90. 90. Kumar R, Kaur A, Pandey A, Mamrutha HM, Singh GP. CRISPR-based genome editing in wheat: A comprehensive review and future prospects. Molecular Biology Reports. 2019;46:1-13
  91. 91. Wang W, Pan Q, He F, Akhunova A, Chao S, Trick H, et al. Transgenerational CRISPR-Cas9 activity facilitates multiplex gene editing in allopolyploid wheat. The CRISPR Journal. 2018;1:65-74
  92. 92. Cameron P, Fuller CK, Donohoue PD, Jones BN, Thompson MS, Carter MM, et al. Mapping the genomic landscape of CRISPR–Cas9 cleavage. Nature Methods. 2017;14:600-606
  93. 93. Kim D, Bae S, Park J, Kim E, Kim S, Yu HR, et al. Digenome-seq: Genome-wide profiling of CRISPR-Cas9 off-target effects in human cells. Nature Methods. 2015;12:237-243
  94. 94. Tsai SQ, Nguyen NT, Malagon-Lopez J, Topkar VV, Aryee MJ, Joung JK. CIRCLE-seq: A highly sensitive in vitro screen for genome-wide CRISPR–Cas9 nuclease off-targets. Nature Methods. 2017;14:607
  95. 95. Tsai SQ, Zheng Z, Nguyen NT, Liebers M, Topkar VV, Thapar V, et al. GUIDE-seq enables genome-wide profiling of off-target cleavage by CRISPR-Cas nucleases. Nature Biotechnology. 2015;33:187-197
  96. 96. Wienert B, Wyman SK, Richardson CD, Yeh CD, Akcakaya P, Porritt MJ, et al. Unbiased detection of CRISPR off- targets in vivo using DISCOVER-Seq. Science. 2019;364:286-289
  97. 97. Doebley JF, Gaut BS, Smith BD. The molecular genetics of crop domestication. Cell. 2006;127:1309-1321
  98. 98. Li J, Sun Y, Du J, Zhao Y, Xia L. Generation of targeted point mutations in rice by a modified CRISPR/Cas9 system. Molecular Plant. 2017;10:526-529
  99. 99. Sun Y, Zhang X, Wu C, He Y, Ma Y, Hou H, et al. Engineering herbicide-resistant rice plants through CRISPR/Cas9-mediated homologous recombination of acetolactate synthase. Molecular Plant. 2016;9:628-631
  100. 100. Ma X, Zhang Q, Zhu Q, Liu W, Chen Y, Qiu R, et al. A robust CRISPR/Cas9 system for convenient, high-efficiency multiplex genome editing in monocot and dicot plants. Molecular Plant. 2015;8:1274-1284
  101. 101. Wang M, Lu Y, Botella JR, Mao Y, Hua K, Zhu JK. Gene targeting by homology-directed repair in rice using a geminivirus-based CRISPR/Cas9 system. Molecular Plant. 2017;10:1007-1010
  102. 102. Mishra R, Joshi RK, Zhao K. Base editing in crops: Current advances, limitations and future implications. Plant Biotechnology Journal. 2020;18:20-31
  103. 103. Li C, Zong Y, Wang Y, Jin S, Zhang D, Song Q, et al. Expanded base editing in rice and wheat using a Cas9-adenosine deaminase fusion. Genome Biology. 2018;19:59
  104. 104. Lu Y, Zhu JK. Precise editing of a target base in the rice genome using a modified CRISPR/Cas9 system. Molecular Plant. 2017;10:523-525
  105. 105. Tang X, Ren Q, Yang L, Bao Y, Zhong Z, He Y, et al. Single transcript unit CRISPR 2.0 systems for robust Cas9 and Cas12a mediated plant genome editing. Plant Biotechnology Journal. 2019;17:1431-1445
  106. 106. Zong Y, Wang Y, Li C, Zhang R, Chen K, Ran Y, et al. Precise base editing in rice, wheat and maize with a Cas9-cytidine deaminase fusion. Nature Biotechnology. 2017;35:438
  107. 107. Anzalone AV, Randolph PB, Davis JR, Sousa AA, Koblan LW, Levy JM, et al. Search-and-replace genome editing without double-strand breaks or donor DNA. Nature. 2019;576:149-157
  108. 108. Lin Q, Zong Y, Xue C, Wang S, Jin S, Zhu Z, et al. Prime genome editing in rice and wheat. Nature Biotechnology. 2020;38(5):582-585
  109. 109. Zhang Y, Liang Z, Zong Y, Wang Y, Liu J, Chen K, et al. Efficient and transgene-free genome editing in wheat through transient expression of CRISPR/Cas9 DNA or RNA. Nature Communications. 2020;7:12617
  110. 110. Liang Z, Chen K, Li T, Zhang Y, Wang Y, Zhao Q, et al. Efficient DNA-free genome editing of bread wheat using CRISPR/Cas9 ribonucleoprotein complexes. Nature Communications. 2017;8:14261
  111. 111. Lu HP, Liu SM, Xu SL, Chen WY, Zhou X, Tan YY, et al. CRISPR-S: An active interference element for a rapid and inexpensive selection of genome-edited, transgene-free rice plants. Plant Biotechnology. 2017J;15:1371
  112. 112. He Y, Zhu M, Wang L, Wu J, Wang Q, Wang R, et al. Programmed self-elimination of the CRISPR/Cas9 construct greatly accelerates the isolation of edited and transgene-free rice plants. Molecular Plant. 2018;11:1210-1213
  113. 113. Korte A, Farlow A. The advantages and limitations of trait analysis with GWAS: A review. Plant Methods. 2013;9:29
  114. 114. Mirdita V, He S, Zhao Y, Korzun V, Bothe R, Ebmeyer E, et al. Potential and limits of whole genome prediction of resistance to Fusarium head blight and Septoria tritici blotch in a vast central European elite winter wheat population. Theoretical and Applied Genetics. 2015;128:2471-2481
  115. 115. Bentley AR, Scutari M, Gosman N, Faure S, Bedford F, Howell P, et al. Applying association mapping and genomic selection to the dissection of key traits in elite European wheat. Theoretical and Applied Genetics. 2014;127:2619-2633
  116. 116. Patel DA, Zander M, Dalton-Morgan J, Batley J. Advances in plant genotyping:Where the future will take us. In: Batley J, editor. Plant Genotyping: Methods and Protocols. New York: Springer New York; 2015. pp. 1-11
  117. 117. Meuwissen THE, Hayes BJ, Goddard ME. Prediction of total genetic value using genome-wide dense marker maps. Genetics. 2001;157:1819-1829
  118. 118. Heffner EL, Lorenz AJ, Jannink JL, Sorrells ME. Plant breeding with genomic selection: Gain per unit time and cost. Crop Science. 2010;50:1681-1690
  119. 119. Nakaya A, Isobe SN. Will genomic selection be a practical method for plant breeding? Annals of Botany. 2012;110:1303-1316
  120. 120. Desta ZA, Ortiz R. Genomic selection: Genome-wide prediction in plant improvement. Trends Plant Science. 2014;19:592-601
  121. 121. Sandhu KS, Aoun M, Morris C, Carter AH. Genomic selection for end-use quality and processing traits in soft white winter wheat breeding program with machine and deep learning models. Biology. 2021;10:689
  122. 122. Sun J, Poland JA, Mondal S, Crossa J, Juliana P, Singh RP, et al. High-throughput phenotyping platforms enhance genomic selection for wheat grain yield across populations and cycles in early stage. Theoretical and Applied Genetics. 2019;132:1705-1720
  123. 123. Crain J, Mondal S, Rutkoski J, Singh RP, Poland J. Combining high-throughput phenotyping and genomic information to increase prediction and selection accuracy in wheat breeding. The Plant Genome. 2018;11:1-14
  124. 124. Sandhu KS, Mihalyov PD, Lewien MJ, Pumphrey MO, Carter AH. Combining genomic and phenomic information for predicting grain protein content and grain yield in spring wheat. Frontiers in Plant Science. 2021;12:170
  125. 125. McCouch SR, McNally KL, Wang W, Sackville HR. Genomics of gene banks: A case study in rice. American Journal of Botany. 2012;99:407-423
  126. 126. Yu X, Li X, Guo T, Zhu C, Wu Y, Mitchell SE, et al. Genomic prediction contributing to a promising global strategy to turbocharge gene banks. Nature Plants. 2016;2:1-7
  127. 127. Singh S, Vikram P, Sehgal D, Burgueño J, Sharma A, Singh SK, et al. Harnessing genetic potential of wheat germplasm banks through impact-oriented-prebreeding for future food and nutritional security. Scientific Reports. 2018;8:1-11
  128. 128. Poland J, Endelman J, Dawson J, Rutkoski J, Wu S, Manes Y, et al. Genomic selection in wheat breeding using genotyping-by-sequencing. The Plant Genome. 2012;5:103-113
  129. 129. Thavamanikumar S, Dolferus R, Thumma BR. Comparison of genomic selection models to predict flowering time and spike grain number in two hexaploid wheat doubled haploid populations. G3: Genes, Genomes, Genetics. 2015;5:1991-1998
  130. 130. He S, Schulthess AW, Mirdita V, Zhao Y, Korzun V, Bothe R, et al. Genomic selection in a commercial winter wheat population. Theoretical and Applied Genetics. 2016;129:641-651
  131. 131. Huang M, Cabrera A, Hoffstetter A, Griffey C, Van Sanford D, Costa J, et al. Genomic selection for wheat traits and trait stability. Theoretical and Applied Genetics. 2016;129:1697-1710
  132. 132. Rutkoski J, Poland J, Mondal S, Autrique E, Pérez LG, Crossa J, et al. Canopy temperature and vegetation indices from high-throughput phenotyping improve accuracy of pedigree and genomic selection for grain yield in wheat. G3: Genes, Genomes, Genetics. 2016;6:2799-2808
  133. 133. Sukumaran S, Crossa J, Jarquin D, Lopes M, Reynolds MP. Genomic prediction with pedigree and genotype × environment interaction in spring wheat grown in South and West Asia, North Africa, and Mexico. G3: Genes, Genomes, Genetics. 2017;7:481-495
  134. 134. Michel S, Ametz C, Gungor H, Epure D, Grausgruber H, Löschenberger F, et al. Genomic selection across multiple breeding cycles in applied bread wheat breeding. Theoretical and Applied Genetics. 2016;129:1179-1189
  135. 135. Michel S, Ametz C, Gungor H, Akgöl B, Epure D, Grausgruber H, et al. Genomic assisted selection for enhancing line breeding: Merging genomic and phenotypic selection in winter wheat breeding programs with preliminary yield trials. Theoretical and Applied Genetics. 2017;130:363-376
  136. 136. Sun J, Rutkoski JE, Poland JA, Crossa J, Jannink JL, Sorrells ME. Multitrait, random regression, or simple repeatability model in high-throughput phenotyping data improve genomic prediction for wheat grain yield. The Plant Genome. 2017;10:1-15
  137. 137. Basnet BR, Crossa J, Dreisigacker S, Pérez-Rodríguez P, Manes Y, Singh RP, et al. Hybrid wheat prediction using genomic, pedigree, and environmental covariables interaction models. The Plant Genome. 2019;12:1-13
  138. 138. Belamkar V, Guttieri MJ, Hussain W, Jarquín D, El-basyoni I, Poland J, et al. Genomic selection in preliminary yield trials in a winter wheat breeding program. G3: Genes, Genomes, Genetics. 2018;8:2735-2747
  139. 139. Guo Z, Zhao Y, Röder MS, Reif JC, Ganal MW, Chen D, et al. Manipulation and prediction of spike morphology traits for the improvement of grain yield in wheat. Scientific Reports. 2018;8:1-10
  140. 140. Juliana P, Singh RP, Poland J, Mondal S, Crossa J, Montesinos-López OA, et al. Prospects and challenges of applied genomic selection—A new paradigm in breeding for grain yield in bread wheat. The Plant Genome. 2018;11:1-17
  141. 141. Norman A, Taylor J, Edwards J, Kuchel H. Optimising genomic selection in wheat: Effect of marker density, population size and population structure on prediction accuracy. G3: Genes, Genomes, Genetics. 2018;8:2889-2899
  142. 142. Hu X, Carver BF, Powers C, Yan L, Zhu L, Chen C. Effectiveness of genomic selection by response to selection for winter wheat variety improvement. The Plant Genome. 2019;12:1-15
  143. 143. Krause MR, González-Pérez L, Crossa J, Pérez-Rodríguez P, Montesinos-López O, Singh RP, et al. Hyperspectral reflectance-derived relationship matrices for genomic prediction of grain yield in wheat. G3: Genes, Genomes, Genetics. 2019;9:1231-1247
  144. 144. Lozada DN, Mason RE, Sarinelli JM, Brown-Guedira G. Accuracy of genomic selection for grain yield and agronomic traits in soft red winter wheat. BMC Genetics. 2019;20:82
  145. 145. Michel S, Löschenberger F, Ametz C, Pachler B, Sparry E, Bürstmayr H. Simultaneous selection for grain yield and protein content in genomics-assisted wheat breeding. Theoretical and Applied Genetics. 2019;132:1745-1760
  146. 146. Juliana P, Singh RP, Braun HJ, Huerta-Espino J, Crespo-Herrera L, Govindan V, et al. Genomic selection for grain yield in the CIMMYT wheat breeding program—Status and perspectives. Frontiers in Plant Science. 2020;11:1418
  147. 147. Lozada DN, Godoy JV, Ward BP, Carter AH. Genomic prediction and indirect selection for grain yield in US Pacific northwest winter wheat using spectral reflectance indices from high-throughput phenotyping. International Journal of Molecular Sciences. 2020a;21:165
  148. 148. Lozada DN, Ward BP, Carter AH. Gains through selection for grain yield in a winter wheat breeding program. PLoS One. 2020b;15:0221603
  149. 149. Sehgal D, Rosyara U, Mondal S, Singh R, Poland J, Dreisigacker S. Incorporating genome-wide association mapping results into genomic prediction models for grain yield and yield stability in CIMMYT spring bread wheat. Frontiers in Plant Science. 2020;11:197
  150. 150. Tsai HY, Cericola F, Edriss V, Andersen JR, Orabi J, Jensen JD, et al. Use of multiple traits genomic prediction, genotype by environment interactions and spatial effect to improve prediction accuracy in yield data. PLoS One. 2020;15:0232665
  151. 151. Bernardo R. Genome wide selection when major genes are known. Crop Science. 2014;54:68-75. DOI: 10.2135/cropsci2013.05.0315
  152. 152. Sehgal D, Autrique E, Singh R, Ellis M, Singh S, Dreisigacker S. Identification of genomic regions for grain yield and yield stability and their epistatic interactions. Scientific Reports. 2017;7:1-2
  153. 153. Sehgal D, Mondal S, Crespo-Herrera L, Velu G, Juliana P, Huerta-Espino J, et al. Haplotype-based, genome-wide association study reveals stable genomic regions for grain yield in CIMMYT spring bread wheat. Frontiers in Genetics. 2020;11:1427
  154. 154. Quarrie SA, Pekic Quarrie S, Radosevic R, Rancic D, Kaminska A, Barnes JD, et al. Dissecting a wheat QTL for yield present in a range of environments: From the QTL to candidate genes. Journal of Experimental Botany. 2006;57:2627-2637
  155. 155. Bevan MW, Uauy C, Wulff BB, Zhou J, Krasileva K, Clark MD. Genomic innovation for crop improvement. Nature. 2017;543:346-354
  156. 156. Qian L, Hickey LT, Stahl A, Werner CR, Hayes B, Snowdon RJ, et al. Exploring and harnessing haplotype diversity to improve yield stability in crops. Frontiers in Plant Science. 2017;8:1534
  157. 157. Ledesma-Ramírez L, Solís-Moya E, Iturriaga G, Sehgal D, Reyes-Valdes MH, Montero-Tavera V, et al. GWAS to identify genetic loci for resistance to yellow rust in wheat pre-breeding lines derived from diverse exotic crosses. Frontiers in Plant Science. 2019;10:1390
  158. 158. Li F, Wen W, Liu J, Zhang Y, Cao S, He Z, et al. Genetic architecture of grain yield in bread wheat based on genome-wide association studies. BMC Plant Biology. 2019;19:1-9
  159. 159. Shokat S, Sehgal D, Liu F, Singh S. GWAS analysis of wheat pre-breeding germplasm for terminal drought stress using next generation sequencing technology. International Journal of Molecular Sciences. 2020;21:3156
  160. 160. Brinton J, Ramirez-Gonzalez RH, Simmonds J, Wingen L, Orford S, Griffiths S, et al. A haplotype-led approach to increase the precision of wheat breeding. Communications Biology. 2020;3:1-1
  161. 161. Stephens M, Smith NJ, Donnelly P. A new statistical method for haplotype reconstruction from population data. The American Journal of Human Genetics. 2001;68:978-989
  162. 162. Bhat JA, Yu D, Bohra A, Ganie SA, Varshney RK. Features and applications of haplotypes in crop breeding. Communications Biology. 2021;4:1-2
  163. 163. Hamazaki K, Iwata H. RAINBOW: Haplotype-based genome-wide association study using a novel SNP-set method. PLoS Computational Biology. 2020;16:e1007663
  164. 164. Sallam AH, Conley E, Prakapenka D, Da Y, Anderson JA. Improving prediction accuracy using multi-allelic haplotype prediction and training population optimization in wheat. G3: Genes, Genomes, Genetics. 2020;10:2265-2273

Written By

Neeraj Pal, Dinesh Kumar Saini and Sundip Kumar

Submitted: 24 January 2022 Reviewed: 27 January 2022 Published: 25 March 2022