Next Article in Journal
Variability in Functional Traits along an Environmental Gradient in the South African Resurrection Plant Myrothamnus flabellifolia
Next Article in Special Issue
Path Coefficient and Principal Component Analyses for Biomass Allocation, Drought Tolerance and Carbon Sequestration Potential in Wheat
Previous Article in Journal
Post-Proline Cleaving Enzymes (PPCEs): Classification, Structure, Molecular Properties, and Applications
Previous Article in Special Issue
Drought-Tolerance QTLs Associated with Grain Yield and Related Traits in Spring Bread Wheat
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Genetic Improvement of Wheat for Drought Tolerance: Progress, Challenges and Opportunities

1
African Centre for Crop Improvement, University of Kwa-Zulu Natal, Private Bag X01, Scottsville, Pietermaritzburg 3209, South Africa
2
Agricultural Research Council—Small Grain, Bethlehem 9700, South Africa
*
Author to whom correspondence should be addressed.
Plants 2022, 11(10), 1331; https://doi.org/10.3390/plants11101331
Submission received: 28 March 2022 / Revised: 27 April 2022 / Accepted: 4 May 2022 / Published: 18 May 2022
(This article belongs to the Special Issue Wheat Breeding, Genomic Selection, and Phenomics)

Abstract

:
Wheat production and productivity are challenged by recurrent droughts associated with climate change globally. Drought and heat stress resilient cultivars can alleviate yield loss in marginal production agro-ecologies. The ability of some crop genotypes to thrive and yield in drought conditions is attributable to the inherent genetic variation and environmental adaptation, presenting opportunities to develop drought-tolerant varieties. Understanding the underlying genetic, physiological, biochemical, and environmental mechanisms and their interactions is key critical opportunity for drought tolerance improvement. Therefore, the objective of this review is to document the progress, challenges, and opportunities in breeding for drought tolerance in wheat. The paper outlines the following key aspects: (1) challenges associated with breeding for adaptation to drought-prone environments, (2) opportunities such as genetic variation in wheat for drought tolerance, selection methods, the interplay between above-ground phenotypic traits and root attributes in drought adaptation and drought-responsive attributes and (3) approaches, technologies and innovations in drought tolerance breeding. In the end, the paper summarises genetic gains and perspectives in drought tolerance breeding in wheat. The review will serve as baseline information for wheat breeders and agronomists to guide the development and deployment of drought-adapted and high-performing new-generation wheat varieties.

1. Introduction

Wheat (Triticum aestivum L., 2n = 6x = 42, AABBDD) is a key commodity crop driving food security and the global economy along the value chains. With an increasing world human population and dwindling agricultural lands, the global demand for wheat products will increase by 60% by 2050 [1]. Therefore, wheat yields will need to increase by 1.6% per annum to meet world demands. Recurrent droughts attributable to climate change affect agriculture-based systems through unpredictable rainfall patterns and changes in crop cycles, diseases, and pest dynamics, leading to a reduction in potential yield gains. Consequently, this compromises food security and economic development, among others [2,3].
Low productivity presents a great challenge to wheat farmers due to market fluctuations and price shocks [4]. Hence, there is a need to develop and adopt drought-adapted modern cultivars that are climate-resilient to mitigate the impacts of current and future environmental changes and deliver market-preferred products. Creating drought-tolerant, high yielding and water-use efficient cultivars is the most economical efficient strategy. The use of irrigation water is unaffordable and unsustainable, especially in arid and semi-arid regions, including sub-Saharan African countries. Elucidation of the genetic, agronomic, and environmental components is paramount to determining the response of wheat to drought. Furthermore, the exploitation of genetic resources such as landraces, synthetics and wild relatives will lead to the discovery of a wealth of essential alleles for wheat improvement programs.
Drought stress hampers plant growth, development and yield by changing the inherent agro-physiological and biochemical processes and pathways [5,6]. Drought stress occurs in different patterns and intensities at different crop growth stages [2,7,8]. The impact of drought stress differs according to the genotype, environment and genotype x environment interaction [9]. In addition, drought tolerance is a complex quantitative trait governed by multiple agronomic traits and polygenes [10]. This has posed significant limitations in developing breeding populations and varieties with improved drought tolerance. The temporal and spatial variation in imposing drought stress across different environments and experiments has contributed to poor selection efficiency, slowing genetic progress in drought tolerance breeding. The essential target agronomic traits for improving drought tolerance include early heading, anthesis and maturity [11,12,13,14], as well as root system architecture [15,16]. Physiological traits include relative water content (RWC), canopy temperature (CT), normalized difference vegetative index (NDVI), stem water-soluble carbohydrates (WSC), among others [5,6,17,18,19]. Further, the main biochemical traits associated with drought tolerance include high soluble sugar content, chlorophyll content, reduced gas exchange, high proline content, increased carbohydrates content and reduced superoxide dismutase concentration [6,20,21]. Ideotype breeding with desirable agro-morphological, physiological and biochemical traits can potentially improve drought tolerance in wheat.
Breeding for drought tolerance is an economical approach to increase wheat production and productivity in arid and semi-arid regions [22]. The pace of development of drought-adapted wheat varieties is hindered by several factors, including a lack of robust, high throughput screening techniques, narrow genetic variation, the large genome size (17 gb) of wheat and environmental variance [10,14,23,24]. The success of pre-breeding and creating adequate genetic variability for drought tolerance breeding is affected by (1) the lack of coordinated efforts in exploring and characterising the genetic resources preserved in global genebanks for drought tolerance and (2) the lack of integrated use of genomic and genetic resources and advanced technologies in breeding programs for fast-tracking phenotypic and genotypic selection for drought tolerance [25]. Developments in the state-of-the-art phenotyping and genotyping platforms provide opportunities to enhance selection responses and improve genetic gains in drought tolerance breeding in wheat. Genomic-assisted drought-tolerance breeding has increasingly shown promise, notably in the advent of genotype-by-sequencing (GBS), genome-wide association studies (GWAS), marker–trait association analysis (MTAs) and quantitative trait loci (QTL) mapping. The QTL analyses have increasingly improved the detailed dissection of genes governing complex traits such as drought tolerance and grain yield in wheat [10,26,27]. Additionally, high-throughput molecular markers such as single nucleotide polymorphism (SNPs) are widely used for effective genotyping and marker–trait association. Understanding the underlying genetic, physiological, biochemical, genetic and environmental mechanisms and their interactions are key opportunities for drought tolerance improvement. Therefore, this review aims to document the progress, challenges, and opportunities in breeding for drought tolerance in wheat.

2. Impact of Drought Stress on Wheat Production

Drought stress has a catastrophic effect on agricultural production. Table 1 summarises the impact of drought stress on key agro-physiological traits and grain yield in wheat. Agronomic traits such as grain yield were reportedly reduced by 25% to 62.75%, grain numbers per spike by 38% to 50%, 1000-kernel weight by 16.4% to 19.42% and plant height by 14.7% to 34.45% under drought stress across different studies [5,6,28,29,30]. This suggests that grain yield and grain numbers per spike are more sensitive to drought stress. Physiological traits such as leaf water content, photosynthetic rate and chlorophyll content reduced by up to 73.8%, 32% and 19%, respectively, suggesting that leaf water content was more sensitive to drought stress [31,32]. Root biomass and above-ground plant biomass were reduced by 23% and 45%, in that order [33,34]. Hence, there is a need to breed for new varieties that can maintain or partition more biomass under drought stress. A meta-analysis from 144 studies published between 1980 and 2015 showed that wheat yields have declined by 20.6% [2], while a related study that included 60 published studies between 1980 and 2017 showed a reduction of 27.5% due to drought stress at different growth stages [29]. Recently, several countries in Africa experienced drought events which reduced wheat production by 45% [35]. Several factors such as increased population growth, unsustainable agricultural production and ecological imbalance (soil erosion, depletion of nutrients and water resources, overutilization of natural vegetation, and environmental disasters) have aggravated the impact of drought stress in these countries [36].
Drought stress affects wheat at all stages of crop growth. However, its effect is more devastating at the seedling, tillering, jointing, heading, anthesis and grain filling stages [4,29,37]. Drought at the seedling stage can inflict up to a 50% reduction in root length [38]. Notably, reproductive and grain filling stages are the most sensitive stages to water stress [39,40,41,42]. Drought at anthesis causes abortion of ovules, consequently reducing the number of grains per spike and grain weight and ultimately grain yield [5,43,44]. During grain filling stage, moisture deficit may disrupt nutrient uptake and photosynthesis, leading to the production of shrivelled kernels [41]. Bennett et al. [45] reported that drought stress reduced yield by 65%. The authors reported a reduction in the heritability values of grain yield from 74% to 58% due to drought. Plants show a reduced stomatal opening and gaseous exchange when stressed, which leads to low photosynthetic efficiency and yield gains. Thus, drought stress will inhibit trait inheritance, photosynthesis and grain yield.

3. Challenges in Breeding for Drought Tolerance

Recurrent drought has become the most prominent cause of reduced yield, grain quality and threat to food security and livelihoods [46]. Challenges in breeding for drought tolerance were reported. There are limited research efforts that identified key root traits in selecting and improving drought-tolerant wheat. Notably, improved root system attributes (e.g., deep and wide-spreading) are desirable for breeding drought-tolerant wheat cultivars [47]. Nevertheless, there is a need to identify key root traits to improve or develop cultivars with improved root attributes and aid in marker-assisted selection. Complementarily, there is a lack of simple and efficient phenotyping methods to improve root attributes as they are labor-intensive and require destructive sampling [48,49]. Developing new high-throughput phenotyping methods that will promote systematic phenotyping of root attributes is of paramount importance. There are limited research efforts that identified key agro-physiological traits in selecting and improving drought-tolerant crops. This is because most drought-adaptive and constitutive traits are controlled by polygenic epistatic and unstable QTL, which are highly influenced by genotype-environment interaction [10,20,50,51,52,53]. This renders low selection efficiency for superior genotypes [54]. Genotype–environment interactions are manifested through crossover ranking and rank inconsistencies when using different indices in identifying drought-tolerant genotypes [55]. Identifying genes associated with drought stress tolerance and their expression and bridging the gap between theoretical research and applied crop breeding is another challenge for brought tolerance breeding [56]. This can be tackled by establishing concerted research groups to reveal the genetic, epigenetic, transcriptomic and metabolomic bases of agro-physiological and root attributes associated with drought tolerance in wheat [57,58,59]. The large genome size (17 Gb) of wheat makes it comparatively more difficult to identify genetic loci controlling key agro-physiological traits conferring drought tolerance in wheat due to its complex genetic background. In this regard, identifying stable QTL or establishing marker–trait associations under contrasting water regimes is crucial for improving drought tolerance using marker-assisted selection (MAS). Extending the genetic analysis research into applied breeding beyond QTL detection has been minimal due to the lack of robust phenotyping and the need for translational genetics. Furthermore, existing mapping populations are routinely used, needing the development of new drought-suited populations. Therefore, new mapping populations sourced from genetically and complementary genotypes will provide avenues for improved drought tolerance [60].

4. Opportunities for Drought Tolerance Breeding

4.1. Exploring Mechanisms of Drought Tolerance

There are different mechanisms of drought response, including drought escape, drought avoidance and drought tolerance [12,39]. Some of the potential response strategies used by plants to acclimate to drought stress are described in Table 2. Drought escape is an adaptive trait that enables the plants to grow and complete their life cycle before the beginning of severe drought [61]. Early heading, flowering and maturity, reduced plant height and short growth cycle are unique attributes to escape dry spells [11,12,13]. In particular, early heading, flowering and maturity are major drought escape mechanisms that allow the completion of the life cycle before the onset of terminal drought stress, which is common in most rainfed agro-systems [14]. Drought avoidance includes the ability for enhanced uptake of available water and nutrients by a longer or deeper root system [13]. This mechanism is associated with a slow growth rate, small or closed stomata, decreased leaf area, reduced photosynthetic activity and low cell metabolism [61]. Root traits such as increased root biomass, root length density and rooting depth are key drivers of drought avoidance [62]. Decreased leaf area is one of the drought avoidance attributes which results in reduced water loss through transpiration [63]. Drought tolerance is the ability of the plant to maintain its growth, development and reproduction under drought stress conditions [12]. Early maturity and reduced leaf area are common attributes in drought-adapted genotypes. Therefore, understanding plant response to drought tolerance at all growth stages is paramount for breeding.

4.2. Exploring Selection Indices for Drought Tolerance

Use of the target selection and production environments and water stress management remain fundamental approaches in drought tolerance improvement [36,64]. Abdolshahi et al. [48] reported three approaches for breeding drought tolerance. These include (1) breeding for higher yield under non-stress conditions, (2) breeding for maximum yield under drought-prone environments and (3) breeding for drought tolerance using selection indices (traits). For enhanced selection efficiency under non-stress conditions aimed at improving performance under the target drought-prone environment, the procedure assumes the trait(s) measured in two different environments not as one but as two traits correlated genetically. This is because the physiological and genetic mechanisms and the genes required for superior performance may be different under these environments [36]. High genetic correlation of traits under complementary selection environments guarantees higher selection responses for yield and yield influencing traits. Traits with high heritability, genetic advance and genetic gains are essential for direct and indirect selection for better grain yield under different environmental conditions [24,26,79]. Indirect selection involves a selection of one trait via another, while direct selection involves the per se selection of the target trait [36]. The use of integrative traits accompanied by the development and application of new and advanced technologies could accelerate the phenotypic selection of drought adaptive traits and consequently improve yield in marginal/low-yielding environments.
Agronomic traits such as early heading, anthesis, maturity, spike morphology and reduced plant height have been widely targeted in drought tolerance breeding programs [21,42,80,81]. These traits have been used in direct or indirect selection for grain yield and drought tolerance in wheat [21,26,30,42,55]. Drought response varies across the source populations. Hence evaluating each population is necessary for simultaneous improvement of yield and drought tolerance.
Some physiological traits have been recognised as reliable, cost-effective and non-invasive methods for automated high-throughput phenotyping in crop breeding programs. According to Monneveux et al. [64], physiological traits can be used to select parental genotypes to be used in cross formation. Hence, physiological traits are useful as direct selection criteria for screening populations to eliminate undesirable segregants across generations. Sallam et al. [82] have extensively reviewed drought tolerance-related physiological traits and advances in breeding and genetics research. Key physiological traits are osmotic potential, stay-green, leaf area, relative water content, canopy temperature, normalised difference vegetative index (NDVI), leaf water status and stem water-soluble carbohydrates (WSC) [5,6,17,18]. Stay-green is the ability of a genotype to remain green and continue undertaking photosynthesis due to higher chlorophyll content compared with other genotypes under drought stress [83]. Such genotypes have improved performance under drought conditions with higher grain yield and biomass production [84,85]. SeriM82, a high yielding cultivar released in 1982 exhibited a stay-green phenotype by maintaining green leaf area longer during the grain filling [86]. NDVI is an indirect selection method for stay-green and yield potential [87], while CT denotes the plant’s interaction with the soil and atmosphere whereby plants can mine water under water-restricted conditions [88]. Osmotic adjustments occur when molecular weight accumulates in lower levels of organic solutes [89]. Leaf water status depends on the cell osmotic conditions and water transportation from plant shoots [90]. Canopy temperature [6] and RWC [42] were major yield determinants. Grain yield was associated with CT at both vegetative and grain filling stages [52] and NDVI under rainfed conditions [19]. Furthermore, stem WSC remobilisation during grain filling contributes to grain yield under drought stress [91].
Relatively better yield under drought stress can be achieved by incorporating drought-adaptive biochemical traits from genetically diverse and unrelated parents [30]. Some of the biochemical traits for drought tolerance include soluble sugar content, chlorophyll content, gas exchange, proline content, carbohydrate content, and superoxide dismutase concentration [6,20,21]. Proline content regulates nitrogen accumulation and contributes to membrane stability [92]. Gas exchange is among the key traits susceptible to drought stress. Gas exchange parameters include photosynthetic rate, stomatal conductance, chlorophyll content and water use efficiency [75]. Drought stress tolerance in wheat was associated with high antioxidant enzyme activity, i.e., catalase, glutathione reductase and peroxidase and elevated S-metabolites, i.e., methionine cysteine and glutathione [93,94]. Synthetic derivatives (SYN-DERs) accumulated more soluble sugars, superoxide dismutase concentration, and proline content under drought stress [6]. Proline content was significantly correlated with grain yield suggesting selection efficiency of this trait under drought stress [20]. A reduction in chlorophyll content denotes decreased photosynthesis efficiency [95].
The root and shoot systems are vital for plant growth and development. Exploring genotype plasticity for roots and shoots is useful to improve drought tolerance [26,34,95]. The shoot systems influence plant adaptive response due to differential environmental changes, including drought stress [47]. Roots are primary organs necessary for growth resource acquisition such as water and minerals; thus, wheat varieties with broad environmental adaption and high water use efficiency are the best candidates for breeding. As the soil dries at the surface, a wider and deeper root system ensures access to soil moisture deeper in the profile during water deficit [96]. Thus, breeding for traits such as root length, density, volume, surface area and diameter is an efficient strategy in environments where water deeper in the profile could be available later in the growing season [97,98]. These traits are indirectly involved in water and nutrient acquisition for plant growth [99]. Root length and surface area denote the ability of the plant to acquire soil resources (water and nutrients) [100]. The root length density is used to estimate the soil volume explored by the plant root architecture and consequently the amount of resources (water and nutrients) available to the plant in the soil [97,101]. Root diameter reflects the ability of the plant to adapt to changes in temperature [102], soil texture and water content [103], and mycorrhizal status [104]. This highlights the importance of identifying root system attributes that provide better exploration of the soil profile for resource acquisition and storage, and plant anchorage.
Compared with shallow root genotypes, deep-rooted types have larger-sized grains, higher grain weight and yield. El-Hassouni et al. [105] reported that thousand-kernel weight was 9% higher while grain size and grain yield were 35% higher in deeper rooted genotypes. Root traits can influence stay-green attributes and adaptation to a wide range of climatic or soil conditions [83,106]. Thus, phenotyping of both stay-green and root traits could enable the selection of superior phenotypes for either broad or specific water-stress adaptation.
Plant response or adaptation to drought partially depends on soil properties and the soil water status. Some genotypes respond to drought by maintaining or increasing root growth while decreasing shoot growth [33,107]. Reduced root growth may also occur due to low water status, low oxygen levels (hypoxia or anoxia) and high soil impedance [108]. Early sowing ensures more profound root growth with the ability to access water from deeper soil profiles in drought-prone environments [109]. In this case, increased root versus shoot growth may improve the plant water status under different drought patterns due to enhanced acquisition of water to produce more root tips and maintain the existing shoots. GmbZIP1 has been linked to increased root and shoot growth under drought [110]. Above-ground biomass was reduced by 45% under drought stress compared to non-stressed conditions in Akron, Ohio, USA [33]. Fletcher and Chenu [111,112] investigated the biomass partitioning of 15 elite Australian cultivars released between 1973 and 2012 and found non-significant changes in the plant biomass and green-leaf biomass at flowering. Only new varieties partitioned more stem and spike biomass at the expense of reduced leaves. Changes in tillering ability have also played a role in the partitioning of stems and spikes. For example, during the 1970s, modern varieties showed increased biomass for every litre of water transpired due to high tillering [112], saving up to 500 g of water per plant up to flowering. According to Kirkegaard et al. [113], a 30 cm increase in root depth into the subsoil could extract an extra 10 mm of water in the deeper soils. VRN1 is a gene modulating flowering behavior and balance between shoot and root architecture in wheat and barley [114,115]. A spring wheat cultivar Dharwar Dry released in 1994 exhibited superior shoot and root attributes when assessed at different growth stages under multiple growth conditions. This variety can improve root traits in dryland or water-stressed conditions [116]. There have been fewer studies on genetic analysis of root and shoot attributes due to difficulties in phenotyping as they are labour-intensive and require destructive sampling [48,49]. Genetic variation for phenotypic plasticity is threatened by directional selection within a narrow gene pool composed mainly of elite lines. Yet it is imperative to consider the different and divergent sources of genetic variation to develop drought-tolerant and high-yielding varieties. Studies on trait–marker associations and biomass allocation will improve the selection efficiency in conventional breeding programs [26,34].

4.3. Genetic Variation as a Source of Drought Tolerance

Genetic variation is the pillar for improving quantitative traits such as yield components and drought tolerance [25]. Table 3 contains some of the sources of drought-tolerant genes reported around the world. Genetic variation can be enhanced via the introduction of existing varieties, developing segregating materials through local or international nurseries, hybridisation and mutation breeding [117]. The extent of natural variability changes with time and space due to evolution, natural selection, artificial selection, mutations, gene flow and genetic drift [16]. The development of cultivars with improved adaptation to biotic and abiotic stresses, including drought stress, hinges on identifying suitable genetic resources with adequate and functional genetic variation for target traits. The use of parental lines of divergent genetic backgrounds, including unrelated and complementary genetic resources possessing suitable drought-adaptive and yield-enhancing traits, ensures the development of superior breeding populations [21,30,67].
Important sources of genes for economic traits in wheat include landraces, elite breeding lines, synthetics and wild relatives [6,118,119]. Among these genetic sources, landraces and drought-adapted varieties are ideal for use in pre-breeding and breeding programs. These genetic resources have high cross-compatibility, wide-adaptation and are rich in farmer- and consumer-preferred traits [25]. Furthermore, improved or breeding lines are essential for the creation of genetic variation with less linkage drag associated with undesirable genes or rare alleles. Some wheat genetic resources such as Dharwar Dry (originated from India), Drysdale, Excalibur and Gladius (from Australia) are widely used in developing genetic populations and drought-tolerant lines [120,121,122,123,124]. The reported genetic resources are divergent in transpiration efficiency, drought tolerance, stay-green and high water use efficiency. The major sources of genetic variation in wheat breeding programs are described below.

4.3.1. Landraces

Landraces are genetically heterogenous breeding stocks adapted to their ecologies and farming systems. They are excellent sources of genes for drought tolerance [65,125]. Aragon 03 was one of the landraces selected for its drought tolerance in the 1940s and is widely cultivated in Spain. This cultivar has exhibited durable drought tolerance and improved traits such as higher pre-anthesis biomass production in different conditions [126]. A Japanese variety, ‘Norin10’ had the Rht dwarfing genes (Rht1 and Rht2) [127] while the Aka Komugi landrace harboured the dwarfing Rht8c allele [128] that contribute to drought tolerance. Landraces have not been widely used in breeding programs due to a lack of information on their utility and pedigree, limited availability of descriptors, loss of essential alleles due to evolution and domestication processes and the presence of undesirable alleles that may lead to linkage drag. The exploitation of landraces should be prioritized to deliver important alleles/traits in breeding programs, including drought tolerance.

4.3.2. Synthetics

Synthetic hexaploid wheats (SHWs) developed by artificial hybridization between tetraploid wheat (T. turgidum) and goatgrass (A. tauschii) are valuable sources of drought tolerance genes [129,130]. The major limiting factor in wheat is the narrow genetic variation in the D-genome, thus the SHWs were developed to increase diversity in D-genome for drought tolerance. Useful genes are identified and introgressed via synthetic derivatives or advanced synthetic backcross lines (SBLs). To date, thousands of SHWs have been developed at the International Maize and Wheat Improvement Centre (CIMMYT) using diverse D genome donor species (Aegilops tauschii). The D genome of A. tauschii has close homology to the D genome of the hexaploid wheat, which increases the chances of transferring polygenic traits. Rosyara et al. [131] conducted a study using CIMMYT’s SHWs to investigate the genomic contribution of chromosome D derived from T. tauschii. Their results showed an improved genetic gain for grain yield (25.3%) and maintained higher genetic diversity. CIMMYT’s efforts to widen genetic diversity have produced 1577 SHWs representing 21% of germplasm between 2000 and 2018. Eighty-six varieties were released in 20 countries, with China (34%) having the high adoption rate of these lines, followed by India (7%) [119]. During the past few decades, 30% of yield gains under drought stress were attributed to the use of SHW [132]. Song et al. [130] identified six genotypes that were more drought tolerant than their parents. The genotypes exhibited high antioxidant activities, including superoxide dismutase (SOD), peroxidase (POX) and catalase (CAT) which could have improved drought tolerance under rainfed conditions. These findings demonstrate the importance of SHWs in improving drought tolerance in wheat.

4.3.3. Wild Relatives and Their Progenitors

Wild relatives of wheat and their progenitors, e.g., wild emmer wheat (T. dicoccoides) and Aegilops (A. tauschii) are major sources of drought tolerance genes [133,134,135,136]. Wild-type alleles at the Rht-B1 and Rht-D1, and the presence of the rye translocation (1B.1R) favoured grain yield under drought stress conditions [6]. Wild relative species incurred a lower decline in physiological traits and chlorophyll inflorescence parameters [136]. Like landraces, the major challenge in using wild relatives is the epistatic and pleiotropic effects of some genes associated with rare alleles leading to linkage drag. This can be averted by crossing elite lines to reduce the transfer of rare alleles, simultaneously delivering drought-tolerant genes. This approach will provide avenues for QTL mapping and engineering for drought tolerance in the future.
Table 3. Some of the drought-tolerant wheat genetic sources reported globally.
Table 3. Some of the drought-tolerant wheat genetic sources reported globally.
Variety NamePedigreeCountry/OrganisationYear of ReleaseReferences
KatyaFortunato/No301//Bezostaya 1Bulgaria1983[137]
MufitbeyWariquam//Kloka/Pitic2/3/Warimek/Halberd/4/3 ag3 AroonaTZARI2006[137]
BerkutIrene/Babax//PastorCIMMYT2002[121,138]
Weebil84-CIMMYT-[31]
BabaxBOWjNAC//VEEmBJY/COCCIMMYT1992[51,139]
SeriM82-CIMMYT1982[139]
Pavon F76VICAM-71//CIANO-67/SIETE-CERROS-66/3/KALYANSONA/BLUEBIRDCIMMYT1976[139]
Opata M85-CIMMYT1985[139]
Roelfs F2007-CIMMYT2007[139]
Borlaug100ROELFS-F-2007/4/BOBWHITE/NEELKANT//CATBIRD/3/CATBIRD/5/FRET-2/TUKURU//FRET-2CIMMYT2014[139]
Sitta -CIMMYT -[120]
Dharwar DryDWR39/C306//HD2189India-[120]
Aragon 03-Spain1940[126]
KrichauffWariquam//Kloka/Pitic2/3/Warimek/Halberd/4/3 ag3 AroonaAustralia1996[121]
ExcaliburRAC-l77(Sr26)⁄UNICULM-492⁄⁄RAC-311-SAustralia/University of Adelaide1991[122]
GladiusRAC-875/Kriachauff//Excalibur/Kukri/3/RAC875/Krichauff/4/RAC-875//Excalibur/KukrAustralia/AGT2007[124]
CIMMYT: International Maize and Wheat Improvement Centre. TZARI: Transitional Zone Agricultural Research Institute, Eskisehir. AGT: Australian Grain Technologies.

4.4. International Research Collaborations

Breeding for drought and heat stress tolerance was initiated by the International Wheat Improvement Network (IWIN) led by the International Maize and Wheat Improvement Centre (CIMMYT) established in 1966 in Mexico and the International Center for Agricultural Research in the Dry Areas (ICARDA)/Lebanon established in 1977 and partnered with Egypt in 1979. CIMMYT and ICARDA used diverse genetic sources, including landraces, synthetic hexaploid, and wild relatives of wheat such as goatgrass (A. tauschii) and durum wheat (T. turgidum) in their breeding programs. Both centres adopted complementary breeding technologies involving conventional and molecular breeding approaches to develop high-yielding germplasm with tolerance to heat and drought stress, diseases and insect pests with acceptable end-use qualities [140,141].
Approximately 70% of spring wheat growing areas in developing countries adopted CIMMYT germplasm as either direct release or breeding parents in their new varieties [142]. Moreover, the centre assisted in capacity building through the training of various research experts in global breeding programs. The global coordination of wheat research, human and infrastructure capacity development related to heat and drought stress breeding include the Heat and Drought Wheat Improvement Consortium (HeDWIC), International Spring Wheat Yield Nursery (ISWYN), Semi-arid Wheat Yield Trials (SAWYTs), Elite Spring Wheat Yield Trials (ESWYTs), Wheat Yield Consortium (WYC), and International Wheat Improvement Network (IWIN). These entities are mandated to collectively bring global research expertise and resources through evaluation of wheat under biotic and abiotic stresses in multiple environments, and also to develop new gene combinations with superior varieties, among others. The research consortium has served as a source of new germplasm globally, allowing the use of beneficial gene pools across multiple environments, defined by similar biotic and abiotic stresses and agro-ecologies [143]. For instance, research and development in wheat in South Africa have continued to make use of CIMMYT lines to enhance drought-tolerance improvement and genetic gains through phenotype selection and molecular breeding [10,20,30,144]. The South African Agricultural Research Council-Small Grain (ARC-SG) germplasm bank holds more than 20,000 small grain accessions (mostly imported from global genebanks), including wheat, oats, barley, rye, and tritocosecale, and wheat accounts for 87% of the ARC-SG collections.

4.5. Wheat Variety Registration, Deployment and Impacts

Agriculture has been the pillar of human livelihood since antiquity, with the first cultivation of wheat being about 10,000 years ago. Wheat cultivation began using landraces selected based on better yield and quality characteristics. Continuous production, human and natural selection have led to improved varieties with high yields and farmer-preferred traits with high threshability and non-shattering types [145,146].
CIMMYT is actively working with research and development partners of various national governments to improve maize and wheat for enhanced productivity. This includes developing varieties that can adapt to different mega-environments such as high rainfall, irrigated and dryland environments, and high temperature or heat stress-stricken areas [141]. Following this success and initiation of other global breeding programs, hundreds of varieties are continuously being released and distributed worldwide. These varieties are imported and bred with locally adapted varieties to deliver diverse, desirable traits, including biotic and abiotic stress tolerance.
In the 1940s, the breeding goals of most programs were aimed at yield increase. However, in the 1960s, due to the green revolution, there was a shift in the development of drought-tolerant varieties through breeding local varieties. Following the green revolution initiative, global wheat productivity increased by 3.3% year−1 between 1949 and 1978. The yield increase was due to the expansion of production area rather than improved varieties. However, in the 1960s, wheat yield increase was significantly high due to the adoption of improved varieties, use of irrigation, crop protection chemicals and fertilizers. The largest yield gains were mainly recorded in Africa and Asia (productivity gains of 2.6 and 3.5% year−1, respectively). Between 1982 and 1991, wheat productivity slowed down to 1.5% year−1 except in China, which increased its productivity due to market-oriented reforms in the rural sectors [147].
Table 3 presents some of the drought-tolerant wheat genetic sources reported globally. Aragon 03 was one of the landraces selected from the indigenous landrace population, Catalan de Monte, for its drought tolerance in the 1940s and is widely cultivated in Spain. This cultivar later showed higher biomass under different growth conditions [126]. ‘Norin10’ (released in the 1940s) is a Japanese green revolution variety with dwarfing genes denoted as Rht1 and Rht2, which have semi-dominant genes Rht-Blb and Rht-Dld [127]. These genes conditioned reduced plant height of 60–110 cm compared to other cultivars with plant height taller than 150 cm [148]. Between the 1960s and 1990s, more than 1500 elite wheat varieties distributed through the international nurseries of wheat breeding were released in different countries [141]. Furthermore, there was excess wheat supply and low food prices during this period due to the impact of the first green revolution in South America and Asian countries [149]. Additionally, there was an increase in income from international export markets, as several countries became net importers of wheat [150].
Genotypes SeriM82 (high yielding) and Babax (drought tolerant) developed by CIMMYT were released in 1982 and 1992, respectively, with relatively high yield performance under drought conditions [151]. Other breeding populations derived from these lines were used in assessing the genetic control of yield and drought adaptive traits under a wide range of environments [51,52,152,153,154]. SeriM82 exhibited a narrow root architecture thus capturing water deeper in the soil [155]. Between 1994 and 2014, public and private breeding programs globally released around 63% and 37% of wheat varieties, in that order. Three CIMMYT-bred varieties such as Roelfs F2007, ONAVAS F2009 and Borlaug 100 were released in 2007, 2009 and 2014, showing significantly high yields of 5.95, 5.83 and 6.58 t/ha, respectively [156]. The release of new varieties helped local and international private seed companies to increasing the adoption of improved varieties by more than 50%. For instance, in Ethiopia, the income gained by wheat farmer’s post-adoption of modern wheat varieties in 2016/2017 was estimated at USD 48 million. This has saved the country a monetary value of USD 65 million that could have been spent on importing wheat [157]. This highlights the potential for wheat variety registration/release and deployment for economic development.

5. Breeding Methodologies and Technologies

5.1. Breeding Wheat for Drought-Tolerance—Conventional Approaches

Conventional breeding involves the creation of genetic variation through sexual recombination of genes from contrasting parents and selecting superior progenies for developing improved varieties. Different selection methods are used in the conventional breeding of self-pollinated crops, including wheat. These include bulk selection, pedigree breeding, pure line selection, and single seed descent (SSD) selection methods, among others. The most widely used selection methods in wheat improvement are the bulk selection method, pedigree method and SSD [52,153,158,159], which are briefly outlined below.
Bulk selection involves hybridisation of parents to produce the F1 generation. The F1 is selfed to produce the F2 generation, and the subsequent generations are harvested in bulk to raise the next generations up to the F5 [160]. This method is simple, convenient, cost-effective, and relatively easy and does not require pedigree recording. Line selection is performed at the F6 and the evaluations of lines are performed until F10 to F11 when the population has become homozygous. During this selection method, natural selection is expected to increase the frequency of superior genotypes and improve genotype adaptability to different environmental conditions such as salinity and drought stress [161,162]. The disadvantages of this method include (1) the longer time required to develop a new variety, (2) unavailability of information on trait(s) inheritance since a progeny test is not mandatory, and (3) loss of superior genotypes due to natural selection [160,163].
Single seed descent involves advancing the breeding generations through the use of a single seed after the initial crosses are performed. With this method, F2 to F4 generations are advanced without selection and irrespective of individual plant vigour. Selection is only performed later in the F5 or F6 when the population is presumably homozygous. This method requires little space, labour and effort and allows rapid advancement of the next generation by retaining sufficiently large and random samples from F2 generation. The major advantages of SSD are (1) the homozygosity can be obtained very easily and rapidly (two to three generations per year), (2) it is not affected by natural selection and (3) it is amenable to various selection methods, including speed breeding. The demerits of this method are (1) the selection of plants is based on individual phenotype and not the progeny performance and (2) loss of desirable traits due to selection from a single seed per plant [158].
Pedigree breeding begins with the hybridisation of selected pairs of parents, e.g., “a commercial variety and a genotype chosen based on a particular superior trait”. A single plant is selected from the segregating F2 population. Progeny performance is evaluated with the repeated selections from F3 to F6 until the recombined genes are homozygous and the population is homogeneous [158]. This method requires record keeping to track parent–progeny relationships [164]. During pedigree breeding, phenotypic selection is performed in the early generations (F3–4), and yield tests are conducted later (F5 to F10) when the population has reached adequate homozygosity [159]. According to Allard [164], a new variety may be released for commercial production after testing for five years at least at five representative locations.
The success of any phenotypic selection programmes depends on drought-adaptive and constitutive traits that are highly heritable in the breeding generations. Landraces, breeding lines, synthetics, double haploids (DHs) and recombinant inbred lines (RILs) are largely utilised during drought tolerance breeding [25,52,165]. Integration of drought tolerance genes into the high genetic background is often carried out in conventional breeding. For example, a mapping population was developed from a cross of SeriM82 × Babax, whereby SeriM82 was a high yielding genotype while Babax was a drought-tolerant genotype [52]. Conventional breeding requires several generations of screening to identify contrasting breeding parents and develop stable performing varieties through continuous selection across multiple environmental conditions [163]. Contrasting and target production environments are used to identify superior genotypes with specific or broader adaption [6,10,41,51].
Due to the various limitations associated with phenotypic selection, there is the need for complementary breeding approaches such as marker-assisted selection, genomic selection and genome editing to identify and select superior genotypes. This will enable early generation selection and independent to environmental conditions. The advent of molecular markers and genome editing technologies provides opportunities for phenotyping complex traits, thus reducing labour, time and costs for cultivar development. This will improve selection efficiency in conventional breeding, consequently overcoming the shortcomings of conventional breeding.

5.2. Breeding Wheat for Drought-Tolerance—Genomic Resources

Genomic resources are routinely used to complement phenotypic selection. DNA markers are used for genetic analysis and identify and select superior genotypes heterotic groups, introduce and track genes in the breeding processes. Marker analysis enable to integrate essential traits and genes conferring the adaptation and performance of wheat genotypes, including drought and heat stress conditions. Marker-assisted selection (MAS) will allow efficient selection irrespective of the stage of the plant growth and without the influence of the environment and thus shortening the breeding cycle [166].
Molecular markers such as single nucleotide polymorphisms (SNPs) and diversity array technology (DArT) have revolutionized the application of MAS in wheat breeding programs [167]. Several markers associated with drought-responsive agronomic traits have been identified [6,10,15,143]. However, the genetic control of drought tolerance is complex due to the large number of genes, unstable QTL, epistatic interaction of QTL and the large genome size and the complicated genetic background of wheat [50]. Furthermore, most marker technologies indicate the presence or absence of a gene without detailed information on its expression and effects on a trait [168].
The advent of genomic resources to avail high-density genome-wide genotype-by sequencing (GBS) have allowed genomic prediction and selection of superior genotypes with multi-genetic traits at the early stages of the breeding cycle [169,170]. In wheat, GBS or next-generation genotyping are valuable tools that are widely used to discover SNPs, identify genetic variations, reduce genome complexity and predict genetic gains [170]. Genomic selection (GS) involves the following steps: (1) phenotyping of diverse sets of populations for different quantitative traits and genotyping across the entire genome to predict the performance of the distinct population, (2) genotyping of breeding population to estimate the genomic estimated breeding values (GEBVs), (3) validation of both sets of populations by phenotyping and genotyping and (4) investigation of genetic gains over time [141,170,171]. GS is carried out to accommodate all minor-effect QTL, to identify individuals with the highest GEBVs for target traits and to reduce the number of generations required to select a superior phenotype. The combination of GS with phenotypic selection (PS) showed an improvement of yield by 23%, indicating that genetic gains could be improved by complementing conventional PS with GS in breeding programs [170]. However, GS is yet to be explored for drought tolerance breeding since drought patterns may vary over space and time across wheat varieties and locations.

5.3. Gene editing Technologies and Breeding for Drought Tolerance

The wheat genome is a complex build-up of genes from three genomes (AABBDD, 2n = 6x = 42) [172]. This renders complex regulatory pathways to interact, constitute and maintain genetic homeostasis. Genome editing alters particular genomic regions through insertion, deletion and substitution of genes [172,173,174,175,176]. Thus, any new gene(s) inserted or substituted in the genome should be stable and in a desirable direction [177].
In the past 10 years, genome editing has enabled scientists to generate targeted modifications in organisms of interest [178,179]. The genetic engineering approach efficiency in organisms through genome editing involves nucleases: zinc-finger nucleases (ZFNs), transcription activator-like-effector nucleases (TALENs) and the clustered regularly interspaced short palindromic repeat (CRISPR/Cas systems) [179,180] such as CRISPR/Cas9, CRISPR/sgRNA and CRISPR/Cpf1 [181,182,183,184]. Notably, CRISPR/Cas9 could aid in the rapid improvement of drought and locally adapted varieties by integrating genes from wild relatives showing drought tolerance. This will deliver new commercial varieties that still retain stress resistance traits of their wild relatives [185].
The use of genomic resources and technologies could accelerate molecular breeding and improvement of crops for adaptability to abiotic and biotic stresses. Intra-or inter-genus and species transfer of alien genes for drought and heat tolerance can be deployed successfully to improve wheat adaptability. For example, the transfer of foreign genes modulating stress-adaptive traits, including hormones, dehydration-responsive element-binding proteins (DREB), enzymes and deeper rooting genes (DRO1) have proved that wheat adaptation and performance can be improved [186]. An example of intragenus transfer of a gene(s) includes DRO1 and DREB genes. DRO1 controls root growth angle in rice [82,187] and influences the orientation of the root system in wheat [186]. DREB are involved in tolerance to numerous abiotic stresses, including drought, salinity, low temperature and abscisic acid (ABA) in wheat [188], and drought tolerance and growth retardation in rice [189].
The application of these technologies is not limited to drought tolerance studies. For example, CRISPR/Cas9 technologies have confirmed their simplicity, proficiency, flexibility and wide adaptability and applicability in several plant-based applications [190]. For example, they have been used for targeted mutagenesis in chickpeas to unravel responsive genes under drought stress [191]. Loss-of-function mutations are the most genomic modification that occurred during the domestication whereby they were stacked in key genes controlling traits such as flowering, seed shattering, colour and size through the application of CRISPR/Cas9 [192]. This has enabled breeders to retrace thousands of years of crop improvement in the process of de novo domestication.
One of the main limiting factors to genome editing is the plant transformation efficiency, which hinders the transfer of edited material into the target cells [185]. The role of genome editing is to change an organism’s DNA (add, remove or alter) at a particular location in the genome. Therefore, repairing the pathways of DNA double-strand break (DSB) are homology-dependent repair (HDR) in which a donor sequence matching the target is copied, and nonhomologous end-joining (NHEJ) in which rejoining the broken ends can lead to mutations at the break site [175]. The advent and high potential of genome editing in crops are continuing to drive the development of more effective plant transformation approaches.

5.4. Biotechnological Approaches

Molecular Markers, Genetic Engineering and Associated Technologies

Biotechnological approaches such as omics and genetic engineering (GE) technologies have been incorporated into breeding programs for accelerated breeding and increase genetic gains in drought tolerance [193]. Table 4 presents the application of different biotechnology approaches to various crops globally. Transgenic and cisgenic are two GE approaches that have advanced the integration of high value traits [194,195]. Transgenic involves transfer of desirable genes derived from target organisms through vector and tissue culture systems or particle bombardment [196]. Cisgenic is termed as the use of recombinant DNA technology to transfer genes from complementary parents (crossable and sexually compatible species) without altering much of the genetic background of the recipient parent [194]. Cisgenesis is almost similar to conventional breeding as it combines the phenomenon with modern biotechnology approaches. Cisgenesis has more impact when used (1) on traits that are dominantly inherited and (2) when applied in translocation or introgression breeding [197]. The two GE approaches have significantly advanced the improvement of targeted traits in wheat. For example, the expression or overexpression of genes GmDREB and TaPEPKR2 improves drought tolerance [198,199]. Overexpression of TaNAC69 has significantly increased root biomass and longer and deeper roots in bread wheat [200], while the transfer of 1Dy10 has improved end-user quality traits such as bread baking quality in durum wheat [201].
The release of genetically modified (GM) crops is regulated to prevent potential negative and harmful effects to human health and the environment. These regulations are based on transgenic organisms (non-crossable species) and cisgenic organisms (crossable species). However, cisgenesis as an advancement of traditional breeding can be effective if excused from GM regulation [202]. This is because, in GM application, the gene source is more prioritised than the technology of gene transfer [197]. In 2007, GM crops were grown on 114 million hectares and increased to more than 179 million hectares in 2015 (more than 10% of world’s arable land) globally. GM soybean and maize for drought tolerance have been developed and commercialized for production; however, progress in the development of GM wheat is still an infant. For example, in Africa, GM are grown in three countries: South Africa (cotton, maize and soybean (2.3 million ha, 19.13% of Africa’s share)), Burkino Faso (cotton (0.4 mha, 7.02%)) and Sudan (cotton, 0.1 mha, 0.41%), with the common crop being GM cotton [203]. Currently there is no GM wheat commercialised anywhere globally. Khan et al. [204] comprehensively reviewed the progress in developing drought-tolerant transgenic wheat and reported that there is no drought-tolerant transgenic wheat approved for commercialisation. This is because most evaluated transgenic wheat lines fail to perform under stressed environments [205,206].
Marker technologies such as marker-assisted back cross (MABC) breeding enable the transfer of targeted genes/QTL at target loci from complementary parents with two or three backcrosses. MABC is presumably one of the convenient and cost-effective forms of marker-assisted selection (MAS) involving the transfer of targeted traits from a donor parent to a recurrent elite parent without considerable change in the genetic background of the recurrent/recipient parent [207]. This approach is efficient on crops with traits of low heritability values that is difficult to phenotype and select using key quantitative traits, especially those expressed later in the growth stage [194]. However, transferring one gene at a time using MABC can be a challenge, given that quantitative traits are affected by environmental conditions and epistatic and polygenes [25]. Though the MABC approach has enabled accelerated improvement of drought tolerance in wheat [208,209], successful MABC studies for drought tolerance are still marginal.
Table 4. The application of biotechnology approaches on various crops globally.
Table 4. The application of biotechnology approaches on various crops globally.
CropGene Expression GeneTrait DescriptionsReferences
Intragenesis/Transgenesis
WheatOverexpression TaNF-YB4Produces more spikes and increases grain yield.[124]
WheatExpressionGmDREBConfers drought tolerance, produces more leaves, roots and high soluble sugar contents.[198]
WheatExpressionHVA1Improves biomass and water use efficiency.[210]
WheatOverexpressionAtHDG11High grain yield and induce changes on physio-morphological traits such as higher proline content and photosynthesis, lower stomatal density, lower rate of water loss, and increased activities of catalase and superoxide dismutase.[211]
WheatOverexpressionFerritin gene, TaFER-5BImproves leaf iron content and ROS, confers tolerance to drought and temperature.[212]
WheatOverexpressionTaNAC69Increases root biomass and longer and deeper roots.[200]
WheatExpressionPEPCHigher proline, soluble sugar and water use efficiency, more extensive root system as well as increased photosynthetic capacity.[213]
WheatExpressionHaHB4Increases yield and water use efficiency.[214]
WheatOverexpressionCspA and CspBLower rate of water loss and MDA content, higher chlorophyll, proline and grain yield.[215]
WheatOverexpressionTaPEPKR2Enhances drought tolerance, higher root length.[199]
WheatExpressionTaSnRK2.8Enhances tolerance to drought, salt and cold stress. Other traits include longer primary roots and various physiological traits, including higher relative water content, strengthened cell membrane stability, significantly lower osmotic potential, more chlorophyll content, [216]
WheatOverexpressionTabZIP2Reduces spikes and seeds, increases single seed weight.[217]
PotatoSilencingGBSSHigh amylopectin.[218]
PotatoSilencingStAs1, StAs2Limits acrylamide in French fries.[219,220]
PotatoSilencingPpo, R1, PhLPrevents black spot bruise, limit cold-induced degradation of starch and limits acrylamide in French fries.[220,221,222]
AppleExpression HcrVf1, HcrVf2Resistance to scab.[223]
StrawberryOverexpressionPGIPResistance to grey mould.[224]
AlfalfaSilencing ComtReduced lignin levels.[225]
Cisgenesis
Durum wheatExpression1Dy10Improves end-user quality traits such as bread baking quality.[201]
BarleyOverexpressionHvPAPhy_aImproves grain phytase activity.[226]
GrapevineExpressionVVTL-1, NtpIIResistance to fungal disease.[227]
PotatoExpressionR-genesResistance to late blight.[228]
AppleExpressionHcrVf2Resistance to scab.[229]
AppleExpressionRvi6Resistance to scab.[230]
PoplarOverexpressionGrowth genes, PATDifferent growth types (rate of regeneration of transgenic shoots, growth rate, plant size and architecture).[231]

5.5. Genes Associated with Drought Tolerance

Several studies have reported genes linked to important drought-responsive traits (Table 5). A gene is considered a candidate when it is associated with a known or proposed function determining the QTL of a trait of interest. As drought brings about changes in gene expression, identification of the expression of candidate genes under drought stress is important. Wheat genotypes use multiple mechanisms to respond to drought stress and numerous genes condition these mechanisms. This includes genes involved in the coding proteins involved in osmotic adjustment, repairs, transcriptions and regulations [134,232,233,234,235]. Several bio-chemicals assist in regulating dehydration membrane stabilization and osmotic adjustments, among other functions [50,236], thus enabling wheat response to drought at different growth stages [57].
Genes controlling root architecture play a significant role in resource acquisition such as water and nutrients and have been widely targeted in drought tolerance breeding. DEEPER ROOTING1 (DRO1) gene, a rice QTL controlling root growth angle reportedly played a significant role in altering root system architecture, thus improving drought avoidance [186,253]. Rice DRO1 orthologs and wheat DRO1 orthologs share 76% identity, suggesting the possibility of functional similarity and potential contribution in manipulating root surface area (RSA) for drought avoidance in wheat [235]. Deep-rooted plants contribute to drought avoidance by extracting moisture from deeper soil layers [13]. A field study assessed the effect of drought on changes in wheat transcriptome during the early reproductive stage and discovered 309 differentially expressed genes (DEGs) involved in various critical processes such as floral development, photosynthetic activity and stomatal movement [57]. Candidate gene TaELF3 for earliness per se (Eps) locus has proved to play a significant role in regulating flowering time [254]. Green revolution genes such as Rht-Blb and Rht-D1b significantly reduce plant height in wheat [255]. GPC-B1 is an important gene regulating gluten protein content in wheat, thus significantly affecting grain yield [165]. Genes such as TaSNAC8-6A, TaMYB3R1 and TaNAC69 have been reported to contribute to drought response in wheat [200,246,252]. This represents valuable wheat genetic resources for the improvement of drought tolerance. However, the impact of these genes on grain yield remains to be elucidated. Gene pyramiding and stacking by crossing complementary drought tolerance genotypes/traits from different growth stages could boost drought adaptation and grain yield.
Gene cloning in wheat has been achieved using comparative genomics approaches between wheat and rice, resulting in yield-related genes such as TaTGW6 [256], TaGW2-6A [257] and TaGS-D1 [258] and TaSus2 [259] among others. TaMYBsm3 and TaCRT1 are among other genes cloned for wheat adaptation to abiotic stresses such as drought [239,250]. The discovery of these genes has provided useful information in understanding the genetics of wheat adaptation to target environments, yield stability, and its contributing traits’ performance under such conditions. Such breakthroughs can be utilised in MAS and genomic selection to accelerate breeding, variety development and deployment.

5.6. QTLs Associated with Root and Shoot Attributes under Drought Conditions

Genetic improvement of drought-responsive root attributes and their contribution to higher and stable grain yield through the integration of advanced genomic approaches is of immense importance. Initially, QTL mapping involved the use of bi-parental crosses in different genetic backgrounds. This approach enabled estimation of the number of genomic regions controlling the specific traits in defined populations, characterization of the genomic regions with regard to map position, gene function, phenotypic and pleiotropic effects and epistatic interactions with other QTL. However, this was limited in allelic diversity, genomic resolution and the longer time required to develop mapping populations.
Genome-wide association study (GWAS) and/or marker–trait association (MTA) approach overcomes several QTL mapping limitations by producing higher resolution, based on linkage disequilibrium across the genome, exploiting/employing data from diverse genetic backgrounds, making these approaches more efficient [81,143,260,261]. With GWAS/MTAs, an extensive collection of wheat germplasm is genotyped with SNPs or DArT markers throughout the genome to identify associations with the phenotypic trait(s) of interest [49,143,262,263,264]. However, the large and complex wheat genome, and incompatible genome sequence make GWAS and MTAs studies challenging for identifying genomic regions underlying the observed phenotypes. The availability of wheat sequence reference genome has allowed annotation of functional genes [265], thus enhancing understanding of genome architecture, gene expression, the relationship between drought tolerance genes/QTLs and their conditioning factors [262,263,264,266]. In the past 68 years (1947–2015), multi-trait MTAs or genomic regions have significantly contributed to yield gains of 2.63 to 25.7 million tons [262]. Nearly 800 MTAs/GWAS were reported for drought-responsive traits, i.e., agronomic, physiological, roots and its related traits [234]. Significant MTAs were reported for 36 agro-morphological traits [81].
Mwadzingeni et al. [267] have comprehensively documented genomic regions associated with agro-morphological traits. A summary of recent efforts in QTL and association mapping for important root and shoot traits associated with drought tolerance is presented in Table 6. However, genomic studies on the relationship between root and shoot as well as yield traits remain to be elucidated. This is because root and shoot traits are complex, controlled by polygenes, QTL and environmental effects; therefore, it is challenging to quantify under field conditions. Despite the likely importance of roots and shoots in wheat performance and drought tolerance, few genomic studies on these traits have been undertaken [26,49,266]. Wheat genome B and D [268] have shown to be the main genomes influencing root traits response to drought stress suggesting that these regions are pleitropic and have multiple genes influencing root development. Wheat genome D contained the most loci for root traits under drought-stressed conditions [15,49]. Kabir et al. [269] found chromosomes 2A, 3A, 4D and 5A in a DH population and chromosomes 3B, 4A, 4D and 5B in RIL population as the main loci influencing root parameters. Chromosome 4D harbours pleiotropic QTL for root traits in DH and RIL populations, while chromosome 3A had the pleiotropic markers in the DH population. In addition to locus 3A, chromosomes 2B and 2D have shown pleitropic QTL for root parameters [270]. QTL on 1B, 3D, 4D, 5A and 5B for coleoptile length were found while chromosome 3D, 4D and 5A showed pleiotropism for plant height [271]. QTL for coleoptile length were found on chromosomes 3B and 4B [272]. These results show that QTLs for root parameters are genetically complex and highly influenced by the growth medium and the plant genotype. Some favourable alleles are not well recorded from different environments and genetic backgrounds, thus presenting opportunities in deploying specific alleles with the use of molecular markers. Therefore, it is important to develop accurate, reliable and well-defined phenotyping assays and techniques to elucidate the mechanisms underlying tolerance to drought stress and high yield under drought-stressed conditions. Identifying genomic regions associated with important breeding traits under rainfed and drought stress conditions and characterization of their genetic make-up is of paramount importance. This will improve the development of breeding pools with positive and beneficial alleles and introgression through MAS. Current and future breeding programs can devise strategies to accumulate these alleles to increase genetic gains.

6. Progress in Breeding for Drought Tolerance in Wheat

Breeding Progress and Genetic Gains

Breeding progress under drought-free environments has shown an ever-increasing yield trend, and selection based on this may not give adequate results for drought-tolerant varieties [36,277]. This is because it is practically impossible to combine traits/genes responsible for superior performance in all environments into a single variety. Furthermore, drought is a moving target that cannot be addressed through breeding for neither specific nor wide adaptation. On the contrary, progress in breeding and genetic gains under water-limited environments have slowed down over the years due to variable climatic conditions from year to year, lack of breakthrough germplasms, inefficient breeding strategies and challenges in identifying key breeding traits as well as large genotype-by-environment interaction [36,278]. Though, selecting under “intermediate” environments can be a good alternative over either selection at high or low yielding environments, there is no straightforward criterion to determine the intermediateness of a distinct environment. Selection of varieties under both high and low yielding environments could be one of the sustainable approaches to develop varieties suited to both conditions.
To sustain the current and the future wheat demands, it is imperative to assess the genetic gains and their impact on breeding programs. Determining breeding progress and the rate of genetic gains is important for improving the selection efficiency. This considers the effect of the genotype, environment and their interaction. Phenotypic data from representative germplasm samples evaluated across years (historical data) or altogether in an experiment (era trials) is used to realise genetic gains. CIMMYT has played a significant role in breeding and global distribution of wheat varieties adapted to marginal environments for evaluation under multiple environments and collaboration with national breeding programmes (Table 7) [279,280]. Majority of the documented yield progress under marginal environments has been due to testing under several biotic and abiotic stresses including drought stress. Furthermore, much of the reported yield increase was on the basis of selection under optimum conditions. Nevertheless, the rates of genetic gain are still marginal to meet the projected wheat demand by 2050 [22] needing dedicated wheat breeding programs for heat and drought stress tolerance.
Though CIMMYT data represent international yield trends, national yield trends are a prerequisite to serving as a guide for searching for new innovations and germplasm sources. For example, more than 25,000 yield observations from 26 wheat varieties released by the Agricultural Research Council in South Africa to the world from 1992 to 2012 led to countrywide genetic gains of 0.75%, 0.30% and 0.093% for winter, facultative and irrigated spring wheat [3]. Strategies such as high-throughput phenotyping (HTP), use of improved cultivars, genomic selection and integrating traits through crossing at existing loci guided by MAS could greatly increase genetic gains in wheat [286]. Crop management practices such as irrigation expansion, use of insecticides and fertilisers could also contribute greatly to yield increase.
The major limiting factor for breeding progress and genetic gains is the reduced genetic diversity. Over the last few decades, there have been concerns about increased crop uniformity and reduced genetic diversity [287]. Some assessments revealed inconsistency contributed by different plant breeding methods perceived to reduce genetic gains. Van de Wouw [288] asserted that genetic erosion occurred through replacement of landraces with modern varieties and modern breeding approaches or practices. The use of genetic resources such as landraces, obsolete lines and modern pure or breeding lines have potential to contribute to increased genetic gains.

7. Outlook and Conclusions

Wheat breeders and agronomists strive to increase rates of genetic gains for grain yield to support a growing population. Though yield improvement is a key target for most breeding programs globally, current breeding pipelines are not optimised for selecting drought adaptive and constitutive traits and yield. Selection accuracy for quantitative traits may be accelerated by evaluating more selection candidates in multi-environments using well-defined phenotyping assays and phenotyping tools to deliver drought-tolerant varieties. Various selection methods for grain yield and its contributing agro-morphological traits under drought conditions, molecular markers, and genomic regions have been adopted to develop drought-suited varieties. However, simultaneous improvement of drought tolerance and yield through its breeding traits has been perceived as challenging due to inadequate phenotyping techniques, challenges in identifying key breeding traits and large genotype by environment interaction. The large and complex genome size of wheat influencing the expression of multiple genes and QTLs under a given set of conditions makes breeding for quantitative traits complicated. Labour costs, time and space constraints are among major limiting factors.
Strategies such as the cultivation of high-yielding cultivars, high-throughput phenotyping, increased irrigation and water use efficiency are of immense importance. Understanding of environmental variables and agronomic factors determining wheat response to drought and yield is paramount. The exploitation of genetic resources such as landraces, synthetics and wild relatives will lead to the discovery of the wealth of important alleles that can be used in breeding and improvement programs. Advances in molecular markers and marker technologies such as QTL analysis and detection could accelerate genetic selections for significant breeding traits thus reducing the breeding cycle. The availability of wheat sequence reference genome has allowed annotation of functional genes, thus enhancing discovery and understanding of genome architecture, gene expression, the relationship between drought tolerance genes/QTLs and their conditioning factors. Improvement in data analysis techniques will give more power to identify genes and trait associations as well as guide breeders on the yearly status or progress of breeding in their programs. Therefore, elucidation of genetic loci underlying significant breeding traits facilitating wheat adaptation and tolerance to drought and their expression patterns to drought stress will provide a strong foundation for knowledge-based breeding approaches and strategies for improved germplasm for multi-environment and environment-specific niches.

Author Contributions

T.B.: Conceptualization, original draft preparation, and editing. H.S.: supervision, project administration, content editing. T.J.T.: supervision, project administration and content editing. I.M.: editing initial draft. All authors have read and agreed to the published version of the manuscript.

Funding

The Winter Cereal Trust and AgriSETA/South Africa are thanked for bursary support to the first author. The Agricultural Research Council—Small Grain and University of KwaZulu-Natal/South Africa are acknowledged for the overall research support.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors have no conflict of interest to disclose.

References

  1. Tilman, D.; Balzer, C.; Hill, J.; Befort, B.L. Global food demand and the sustainable intensification of agriculture. Proc. Natl. Acad. Sci. USA 2011, 108, 20260–20264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Daryanto, S.; Wang, L.; Jacinthe, P.-A. Global synthesis of drought effects on maize and wheat production. PLoS ONE 2016, 11, e0156362. [Google Scholar] [CrossRef] [PubMed]
  3. Nalley, L.; Dixon, B.; Chaminuka, P.; Naledzani, Z.; Coale, M.J. The role of public wheat breeding in reducing food insecurity in South Africa. PLoS ONE 2018, 13, e0209598. [Google Scholar] [CrossRef] [PubMed]
  4. Khadka, K.; Raizada, M.N.; Navabi, A. Recent progress in germplasm evaluation and gene mapping to enable breeding of drought-tolerant wheat. Front. Plant Sci. 2020, 11, 1149. [Google Scholar] [CrossRef]
  5. Bayoumi, T.Y.; Eid, M.H.; Metwali, E.M. Application of physiological and biochemical indices as a screening technique for drought tolerance in wheat genotypes. Afr. J. Biotechnol. 2008, 7, 2341–2352. [Google Scholar]
  6. Afzal, F.; Reddy, B.; Gul, A.; Khalid, M.; Subhani, A.; Shazadi, K.; Quraishi, U.M.; Ibrahim, A.M.; Rasheed, A. Physiological, biochemical and agronomic traits associated with drought tolerance in a synthetic-derived wheat diversity panel. Crop Pasture Sci. 2017, 68, 213–224. [Google Scholar] [CrossRef]
  7. Chenu, K.; Deihimfard, R.; Chapman, S.C. Large-scale characterization of drought pattern: A continent-wide modelling approach applied to the Australian wheatbelt-spatial and temporal trends. New Phatol. 2013, 198, 801–820. [Google Scholar] [CrossRef]
  8. Sarto, M.V.M.; Sarto, J.R.W.; Rampim, L.; Bassegio, D.; da Costa, P.F.; Inagaki, A.M. Wheat phenology and yield under drought: A review. Aust. J. Crop. Sci. 2017, 11, 941–946. [Google Scholar] [CrossRef]
  9. Hoffman, M.T.; Carrick, P.J.; Gillson, L.; West, A.G. Drought, climate change and vegetation response in the succulent karoo, South Africa. S. Afr. J. Sci. 2009, 105, 54–60. [Google Scholar]
  10. Mwadzingeni, L.; Shimelis, H.; Rees, D.J.G.; Tsilo, T.J. Genome-wide association analysis of agronomic traits in wheat under drought-stressed and non-stressed conditions. PLoS ONE 2017, 12, e0171692. [Google Scholar] [CrossRef] [Green Version]
  11. Dolferus, R. To grow or not to grow: A stressful decision for plants. Plant Sci. 2014, 229, 247–261. [Google Scholar] [CrossRef] [PubMed]
  12. Aslam, M.; Maqbool, M.A.; Cengiz, R. Mechanisms of drought resistance. In Drought Stress in Maize (Zea mays L.); Springer: Cham, Switzerland, 2015; pp. 19–36. [Google Scholar]
  13. Naeem, M.K.; Ahmad, M.; Kamran, M.; Shah, M.K.N.; Iqbal, M.S. Physiological responses of wheat (Triticum aestivum L.) to drought stress. Int. J. Plant Soil Sci. 2015, 6, 1–9. [Google Scholar]
  14. Shamuyarira, K.W.; Shimelis, H.; Tapera, T.; Tsilo, T.J. Genetic advancement of newly developed wheat populations under drought-stressed and non-stressed conditions. J. Crop Sci. Biotechnol. 2019, 22, 169–176. [Google Scholar] [CrossRef]
  15. Ibrahim, S.E.; Schubert, A.; Pillen, K.; Lèon, J. QTL analysis of drought tolerance for seedling root morphological traits in an advanced backcross of spring wheat. Int. J. Agric. Sci. 2012, 2, 619–626. [Google Scholar]
  16. Ren, Y.; He, X.; Liu, D.; Li, J.; Zhao, X.; Li, B.; Tong, Y.; Zhang, A.; Li, Z. Major quantitative trait loci for seminal root morphology of wheat seedlings. Mol. Breed. 2012, 30, 139–148. [Google Scholar] [CrossRef]
  17. Bowne, J.B.; Erwin, T.A.; Juttner, J.; Schnurbusch, T.; Langridge, P.; Bacic, A.; Roessner, U. Drought responses of leaf tissues from wheat cultivars of differing drought tolerance at the metabolite level. Mol. Plant 2012, 5, 418–429. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Dong, Y.; Liu, J.; Geng, H.; Rasheed, A.; Xiao, Y.; Cao, S.; Fu, L.; Yan, J.; Wen, W.; Zhang, Y.; et al. Genome-wide association of stem water soluble carbohydrates in bread wheat. PLoS ONE 2016, 11, e0164293. [Google Scholar] [CrossRef]
  19. Shi, S.; Azam, F.I.; Li, H.; Chang, X.; Li, B.; Jing, R. Mapping QTL for stay-green and agronomic traits in wheat under diverse water regimes. Euphytica 2017, 213, 246. [Google Scholar] [CrossRef] [Green Version]
  20. Mwadzingeni, L.; Shimelis, H.; Tesfay, S.; Tsilo, T.J. Screening of bread wheat genotypes for drought tolerance using phenotypic and proline analyses. Front. Plant Sci. 2016, 7, 1276. [Google Scholar]
  21. Mkhabela, S.S.; Shimelis, H.; Odindo, A.O.; Mashilo, J. Response of selected drought tolerant wheat (Triticum aestivum L.) genotypes for agronomic traits and biochemical markers under drought-stressed and non-stressed conditions. Acta Agric. Scand. Sect. B-Soil Plant Sci. 2019, 69, 674–689. [Google Scholar] [CrossRef]
  22. Ray, D.K.; Mueller, N.D.; West, P.C.; Foley, J.A. Yield trends are insufficient to double global crop production by 2050. PLoS ONE 2013, 8, e66428. [Google Scholar] [CrossRef] [Green Version]
  23. Brenchley, R.; Spannagl, M.; Pfeifer, M.; Barker, G.L.; D’Amore, R.; Allen, A.M.; McKenzie, N.; Kramer, M.; Kerhornou, A.; Bolser, D.; et al. Analysis of the bread wheat genome using whole-genome shotgun sequencing. Nature 2012, 491, 705–710. [Google Scholar] [CrossRef] [Green Version]
  24. Farzamipour, M.R.; Moghaddam, M.; Aharizad, S.; Rashidi, V. Genetic variation for agronomic characters and drought tolerance among the recombinant inbred lines of wheat from the Norstar × Zagross cross. Int. J. Biosci. 2013, 3, 76–86. [Google Scholar]
  25. Mwadzingeni, L.; Figlan, S.; Shimelis, H.; Mondal, S.; Tsilo, T. Genetic resources and breeding methodologies for improving drought tolerance in wheat. J. Crop Improv. 2017, 31, 648–672. [Google Scholar] [CrossRef]
  26. Mathew, I.; Shimelis, H.; Shayanowako, A.I.T.; Laing, M.; Chaplot, V. Genome-wide association study of drought tolerance and biomass allocation in wheat. PLoS ONE 2019, 14, e0225383. [Google Scholar] [CrossRef] [Green Version]
  27. Ballesta, P.; Mora, F.; Del Pozo, A. Association mapping of drought tolerance indices in wheat: QTL-rich regions on chromosome 4A. Sci. Agric. 2020, 77, e20180153. [Google Scholar] [CrossRef] [Green Version]
  28. Mwadzingeni, L.; Shimelis, H.; Tsilo, T.J. Combining ability and gene action controlling yield and yield components in bread wheat (Triticum aestivum L.) under drought-stressed and nonstressed conditions. Plant Breed. 2018, 137, 502–513. [Google Scholar] [CrossRef]
  29. Zhang, J.; Zhang, S.; Cheng, M.; Jiang, H.; Zhang, X.; Peng, C.; Lu, X.; Zhang, M.; Jin, J. Effect of drought on agronomic traits of rice and wheat: A meta-analysis. Int. J. Environ. Res. Public Health 2018, 15, 839. [Google Scholar] [CrossRef] [Green Version]
  30. Mkhabela, S.S.; Shimelis, H.; Mashilo, J. Genetic differentiation of selected drought and heat tolerant wheat genotypes using simple sequence repeat markers and agronomic traits. S. Afr. J. Plant Soil 2020, 37, 211–219. [Google Scholar] [CrossRef]
  31. Praba, M.L.; Cairns, J.E.; Babu, R.C.; Lafitte, H.R. Identification of physiological traits underlying cultivar differences in drought tolerance in rice and wheat. J. Agron. Crop Sci. 2009, 195, 30–46. [Google Scholar] [CrossRef]
  32. Prasad, G.P.; Pisipati, S.R.; Momčilovié, I.; Ristic, Z. Independent and combined effects of high temperature and drought stress during grain filling on plant yield and chloroplast EF-Tu expression in spring wheat. J. Agron. Crop Sci. 2011, 197, 430–441. [Google Scholar] [CrossRef]
  33. Becker, S.R.; Byrne, P.F.; Reid, S.D.; Bauerle, W.L.; McKay, J.K.; Haley, S.D. Root traits contributing to drought tolerance of synthetic hexaploid wheat in a greenhouse study. Euphytica 2016, 207, 213–224. [Google Scholar] [CrossRef]
  34. Mathew, I.; Shimelis, H.; Mutema, M.; Clulow, A.; Zengeni, R.; Mbava, N.; Chaplot, V. Selection of wheat genotypes for biomass allocation to improve drought tolerance and carbon sequestration into soils. J. Agron. Crop Sci. 2019, 205, 385–400. [Google Scholar] [CrossRef]
  35. Rojas, O.; Vrieling, A.; Rembold, F. Assessing drought probability for agricultural areas in Africa with coarse resolution remote sensing imagery. Remote Sens. Environ. 2011, 115, 343–352. [Google Scholar] [CrossRef]
  36. Keneni, G.; Bekele, E.; Imtiaz, M.; Dagne, K.; Alemaw, G. Challenges associated with crop breeding for adaptation to drought-prone environments. Ethiop. J. Agric. Sci. 2016, 27, 1–24. [Google Scholar]
  37. Baloch, M.J.; Dunwell, J.; Dennet, M.; Rajpar, I.; Jatoi, W.A.; Veesar, N.F. Evaluating spring wheat cultivars for drought tolerance through yield and physiological parameters at booting and anthesis. Afr. J. Biotechnol. 2012, 11, 11559–11565. [Google Scholar] [CrossRef]
  38. Ayalew, H.; Liu, H.; Börner, A.; Kobiljski, B.; Liu, C.; Yan, G. Genome-wide association mapping of major root length QTLs under PEG induced water stress in wheat. Front. Plant Sci. 2018, 9, 1–9. [Google Scholar] [CrossRef]
  39. Farooq, M.; Hussain, M.; Siddique, K.H. Drought stress in wheat during flowering and grain-filling periods. Crit. Rev. Plant Sci. 2014, 33, 331–349. [Google Scholar] [CrossRef]
  40. Allahverdiyev, T.I.; Talai, J.M.; Huseynova, I.M.; Aliyev, J.A. Effect of drought stress on some physiological parameters, yield, yield components of durum (Triticum durum desf.) and bread (Triticum aestivum L.) wheat genotypes. Ekin J. Crop Breed. Genet. 2015, 1, 50–62. [Google Scholar]
  41. Iqbal, J. Morphological, physiological and molecular markers for the adaptation of wheat in drought condition. Asian J. Biotechnol. Genet. Eng. 2019, 2, 1–13. [Google Scholar]
  42. Sobhaninan, N.; Heidari, B.; Tahmasebi, S.; Dadkhodale, A.; McIntyre, C.L. Response of quantitative and physiological traits to drought stress in the SeriM82/Babax wheat population. Euphytica 2019, 215, 1–15. [Google Scholar] [CrossRef]
  43. Pradhan, G.P.; Prasad, P.V.; Fritz, A.K.; Kirkham, M.B.; Gill, B.S. Effects of drought and high temperature stress on synthetic hexaploid wheat. Funct. Plant Biol. 2012, 39, 190–198. [Google Scholar] [CrossRef] [PubMed]
  44. Weldearegay, D.F.; Yan, F.; Jiang, D.; Liu, F. Independent and combined effects of soil warming and drought stress during anthesis on seed set and grain yield in two spring wheat varieties. J. Agron. Crop Sci. 2012, 198, 245–253. [Google Scholar] [CrossRef]
  45. Bennett, D.; Reynolds, M.; Mullan, D.; Izanloo, A.; Kuchel, H.; Langridge, P.; Schnurbusch, T. Detection of two major grain yield QTL in bread wheat (Triticum aestivum L.) under heat, drought and high yield potential environments. Theor. Appl. Genet. 2012, 125, 1473–1485. [Google Scholar] [CrossRef]
  46. Havrlentová, M.; Kraic, J.; Gregusová, V.; Kovácsová, B. Drought Stress in Cereals—A Review. Agriculture (Pol’nohospodárstvo) 2021, 67, 47–60. [Google Scholar] [CrossRef]
  47. Siddiqui, M.N.; Leon, J.; Naz, A.A.; Balivora, A. Genetics and genomics of root system variation in adaptation to drought stress in cereal crops. J. Exp. Bot. 2021, 72, 1007–1019. [Google Scholar] [CrossRef] [PubMed]
  48. Abdolshahi, R.; Nazari, M.; Safarian, A.; Sadathossini, T.S.; Salarpour, M.; Amiri, H. Integrated selection criteria for drought tolerance in wheat (Triticum aestivum L.) breeding programs using discriminant analysis. Field Crops Res. 2015, 174, 20–29. [Google Scholar] [CrossRef]
  49. Khodaee, S.M.M.; Hashemi, M.; Mirlohi, A.; Majidi, M.M.; Sukumaran, S.; Moghaddam, M.E.; Abdollahi, M. Root characteristics of an elite spring wheat panel under contrasting water treatments and their genome-wide association study. Rhizosphere 2021, 19, 100413. [Google Scholar] [CrossRef]
  50. Ashraf, M. Inducing drought tolerance in plants: Recent advances. Biotechnol. Adv. 2010, 28, 169–183. [Google Scholar] [CrossRef]
  51. Pinto, R.S.; Reynolds, M.P.; Mathews, K.L.; McIntyre, C.L.; Olivares-Villegas, J.J.; Chapman, S.C. Heat and drought adaptive QTL in a wheat population designed to minimize confounding agronomic effects. Theor. Appl. Genet. 2010, 121, 1001–1021. [Google Scholar] [CrossRef] [Green Version]
  52. Tahmasebi, S.; Heidari, B.; Pakniyat, H.; McIntyre, L. Mapping QTLs associated with agronomic and physiological traits under terminal drought and heat stress conditions in wheat (Triticum aestivum L.). Genome 2017, 60, 26–45. [Google Scholar] [CrossRef] [PubMed]
  53. Li, L.; Mao, X.; Wang, J.; Chang, X.; Reynolds, M.; Jing, R. Genetic dissection of drought and heat-responsive agronomic traits in wheat. Plant Cell Environ. 2019, 42, 2540–2553. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Langridge, P.; Reynolds, M.P. Genomic tools to assist breeding for drought tolerance. Curr. Opin. Biotechnol. 2015, 32, 130–135. [Google Scholar] [CrossRef]
  55. Semahegn, Y.; Shimelis, H.; Laing, M.; Mathew, I. Evaluation of bread wheat (Triticum aestivum L.) genotypes for yield and related traits under drought stress conditions. Acta Agric. Scand. Sect. B-Soil Plant Sci. 2020, 70, 474–484. [Google Scholar] [CrossRef]
  56. Luo, L.; Xia, H.; Lu, B.-R. Editorial: Crop breeding for drought resistance. Front. Plant Sci. 2019, 10, 314. [Google Scholar] [CrossRef]
  57. Ma, J.; Li, R.; Wang, H.; Li, D.; Wang, X.; Zhang, Y.; Zhen, W.; Duan, H.; Yan, G.; Li, Y. Transcriptomics analyses reveal wheat responses to drought stress during reproductive stages under field conditions. Front. Plant Sci. 2017, 8, 592. [Google Scholar] [CrossRef] [Green Version]
  58. Guo, X.; Xin, Z.; Yang, T.; Ma, X.; Zhang, Y.; Wang, Z.; Ren, Y.; Lin, T. Metabolomics response for drought stress tolerance in chinese wheat genotypes (Triticum aestivum). Plants 2020, 9, 520. [Google Scholar] [CrossRef] [Green Version]
  59. Sun, C.; Ali, K.; Yan, K.; Fiaz, S.; Dormatey, R.; Bi, Z.; Bai, J. Exploration of epigenetics for improvement of drought and other stress resistance in crops: A review. Plants 2021, 10, 1226. [Google Scholar] [CrossRef]
  60. Nezhad, K.Z.; Weber, W.E.; Röder, M.S.; Sharma, S.; Lohwasser, U.; Meyer, R.C.; Saal, B.; Börner, A. QTL analysis for thousand-kernel weight under terminal drought stress in bread wheat (Triticum aestivum L.). Euphytica 2012, 186, 127–138. [Google Scholar] [CrossRef]
  61. Shavrukov, Y.; Kurishbayev, A.; Jatayev, S.; Shvidchenko, V.; Zotova, L.; Koekemoer, F.; De Groot, S.; Soole, K.; Langridge, P. Early flowering as a drought escape mechanism in plants: How can it aid wheat production? Front. Plant Sci. 2017, 8, 1–8. [Google Scholar] [CrossRef]
  62. Kashiwagi, J.; Krishnamurthy, L.; Upadhyaya, H.D.; Krishna, H.; Chandra, S.; Vadez, V.; Serraj, R. Genetic variability of drought-avoidance root traits in the mini-core germplasm collection of chickpea (Cicer arietinum L.). Euphytica 2005, 146, 213–222. [Google Scholar] [CrossRef] [Green Version]
  63. Kapoor, D.; Bhardwaj, S.; Landi, M.; Sharma, A.; Ramakrishnan, M.; Sharma, A. The impact of drought in plant metabolism: How to exploit tolerance mechanisms to increase crop production. Appl. Sci. 2020, 10, 5692. [Google Scholar] [CrossRef]
  64. Monneveux, P.; Jing, R.; Misra, S. Phenotyping for drought adaptation in wheat using physiological traits. Front. Physiol. 2012, 3, 429. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Wasaya, A.; Zhang, X.; Fang, Q.; Yan, Z. Root phenotyping for drought tolerance: A review. Agronomy 2018, 8, 241. [Google Scholar] [CrossRef] [Green Version]
  66. Farhad, M.; Hakim, M.A.; Alam, M.A.; Barma, N.C.D. Screening wheat genotypes for coleoptile length: A trait for drought tolerance. Am. J. Agric. For. 2014, 2, 237–245. [Google Scholar] [CrossRef] [Green Version]
  67. Liu, E.K.; Mei, X.R.; Yan, C.R.; Gong, D.Z.; Zhang, Y.Q. Effects of water stress on photosynthetic characteristics, dry matter translocation and WUE in two winter wheat genotypes. Agric. Water Manag. 2016, 167, 75–85. [Google Scholar] [CrossRef]
  68. Ding, J.; Huang, Z.; Zhu, M.; Li, C.; Zhu, X.; Guo, W. Does cyclic water stress damage wheat yield more than a single stress? PLoS ONE 2018, 13, e0195535. [Google Scholar] [CrossRef] [Green Version]
  69. Ihsan, M.Z.; El-nakhlawy, F.S.; Ismail, S.M.; Fahad, S.; Daur, I. Wheat phenological development and growth studies as affected by drought and late season high temperature stress under arid environment. Front. Plant Sci. 2016, 7, 795. [Google Scholar] [CrossRef] [Green Version]
  70. Su, Y.; Wu, F.; Ao, Z.; Jin, S.; Qin, F.; Liu, B.; Pang, S.; Liu, L.; Guo, Q. Evaluating maize phenotype dynamics under drought stress using terrestrial lidar. Plant Methods 2019, 15, 11. [Google Scholar] [CrossRef] [Green Version]
  71. Cattivelli, L.; Rizza, F.; Badeck, F.W.; Mazzucotelli, E.; Mastrangelo, A.M.; Francia, E.; Mare, C.; Tondelli, A.; Stanca, A.M. Drought tolerance improvement in crop plants: An integrated view from breeding to ge-nomics. Field Crops Res. 2008, 105, 1–14. [Google Scholar] [CrossRef]
  72. Rebetzke, G.J. Bonnett D.G.Reynolds MP.Awns reduce grain number to increase grain size and harvestable yield in irrigated and rainfed spring wheat. J. Exp. Bot. 2016, 67, 2573–2586. [Google Scholar] [CrossRef] [PubMed]
  73. Ali, M.A.; Zulkiffal, M.; Anwar, J.; Hussain, M.; Farooq, J.; Khan, S.H. Mor-pho-physiological diversity in advanced lines of bread wheat under drought conditions at post-anthesis stage. J. Anim. Plant Sci. 2015, 25, 431–441. [Google Scholar]
  74. Bita, C.; Gerats, T. Plant tolerance to high temperature in a changing environment: Scientific fundamentals and production of heat stress-tolerant crops. Front. Plant Sci. 2013, 4, 273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Borrill, P.; Fahy, B.; Smith, A.M.; Uauy, C. Wheat grain filling is limited by grain filling capacity rather than the duration of flag leaf photosynthesis: A case study us-ing NAM RNAi plants. PLoS ONE 2015, 10, e0134947. [Google Scholar]
  76. Lopes, M.S.; Reynolds, M.P.; Jalal-Kamali, M.R.; Moussa, M.; Feltaous, Y.; Tahir, I.S.A.; Barma, N.; VargasM, M.; Mannes, Y.; Baum, M. The yield correlations of selectable physiological traits in a population of advanced spring wheat lines grown in warm and drought environments. Field Crops Res. 2012, 128, 129–136. [Google Scholar] [CrossRef]
  77. Fang, Y.; Xiong, L. General mechanisms of drought response and their application in drought resistance improvement in plants. Cell. Mol. Life Sci. 2015, 72, 673–689. [Google Scholar] [CrossRef]
  78. Kira, O.; Linker, R.; Gitelson, A. Non-destructive estimation of foliar chlorophyll and carotenoid contents: Focus on informative spectral bands. Int. J. Appl. Earth Obs. Geoinf. 2015, 38, 251–260. [Google Scholar] [CrossRef]
  79. Farshadfar, E.; Jamshidi, B.; Aghaee, M. Biplot analysis of drought tolerance indicators in bread wheat landraces of Iran. Int. J. Agric. Crop Sci. 2012, 4, 226–233. [Google Scholar]
  80. Bilal, M.; Rashid, R.M.; Rehman, S.U.; Iqbal, F.; Ahmed, J.; Abid, M.A.; Ahmed, Z.; Hayat, A. Evaluation of wheat genotypes for drought tolerance. J. Green Physiol. Genet. Genom. 2015, 1, 11–21. [Google Scholar]
  81. Sheoran, S.; Jaiswal, S.; Kumar, D.; Raghav, N.; Sharma, R.; Pawar, S.; Paul, S.; Iquebal, M.A.; Jaiswar, A.; Sharma, P.; et al. Uncovering genomic regions associated with 36 agro-morphological traits in Indian spring wheat using GWAS. Front. Plant Sci. 2019, 10, 1–20. [Google Scholar] [CrossRef]
  82. Sallam, A.; Alqudah, A.M.; Dawood, M.F.; Baenziger, P.S.; Börner, A. Drought stress tolerance in wheat and barley: Advances in physiology, breeding and genetics research. Int. J. Mol. Sci. 2019, 20, 3137. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Christopher, J.T.; Christopher, M.J.; Borrell, A.K.; Fletcher, S.; Chenu, K. Stay-green traits to improve wheat adaptation in well-watered and water-limited environments. J. Exp. Bot. 2016, 67, 5159–5172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Calderini, D.F.; Reynolds, M.P. Changes in grain weight as a consequence of de-graining treatments at pre-and post-anthesis in synthetic hexaploid lines of wheat (Triticum durum x T. tauschii). Funct. Plant Biol. 2000, 27, 183–191. [Google Scholar] [CrossRef]
  85. Chen, J.; Liang, Y.; Hu, X.; Wang, X.; Tan, F.; Zhang, H.; Ren, Z.; Luo, P. Physiological characterization of ‘stay green’wheat cultivars during the grain filling stage under field growing conditions. Acta Physiol. Plant. 2010, 32, 875–882. [Google Scholar] [CrossRef]
  86. Christopher, J.T.; Manschadi, A.M.; Hammer, G.L.; Borrell, A.K. Developmental and physiological traits associated with high yield and stay-green phenotype in wheat. Aust. J. Agric. Res. 2008, 59, 354–364. [Google Scholar] [CrossRef]
  87. Babar, M.A.; Reynolds, M.P.; Van Ginkel, M.; Klatt, A.R.; Raun, W.R.; Stone, M.L. Spectral reflectance to estimate genetic variation for in-season biomass, leaf chlorophyll, and canopy temperature in wheat. Crop Sci. 2006, 46, 1046–1057. [Google Scholar] [CrossRef]
  88. Lopes, M.S.; Reynolds, M.P. Partitioning of assimilates to deeper roots is associated with cooler canopies and increased yield under drought in wheat. Funct. Plant Biol. 2010, 37, 147–156. [Google Scholar] [CrossRef]
  89. Marcińska, I.; Czyczyło-Mysza, I.; Skrzypek, E.; Filek, M.; Grzesiak, S.; Grzesiak, M.T.; Janowiak, F.; Hura, T.; Dziurka, M.; Dziurka, K.; et al. Impact of osmotic stress on physiological and biochemical characteristics in drought-susceptible and drought-resistant wheat genotypes. Acta Physiol. Plant. 2013, 35, 451–461. [Google Scholar] [CrossRef] [Green Version]
  90. Lawlor, D.W.; Cornic, G. Photosynthetic carbon assimilation and associated metabolism in relation to water deficits in higher plants. Plant Cell Environ. 2002, 25, 275–294. [Google Scholar] [CrossRef] [Green Version]
  91. Li, M.; Liu, Y.; Ma, J.; Zhang, P.; Wang, C.; Su, J.; Yang, D. Genetic dissection of stem WSC accumulation and remobilization in wheat (Triticum aestivum L.) under terminal drought stress. BMC Genet. 2020, 21, 50. [Google Scholar] [CrossRef]
  92. Javadi, T.; Arzani, K.; Ebrahimzadeh, H. Study of proline, soluble sugar, and chlorophyll a and b changes in nine Asian and one European pear cultivar under drought stress. In Proceedings of the XXVII International Horticultural Congress-IHC2006: International Symposium on Asian Plants with Unique Horticultural, Seoul, Korea, 30 June 2008; Volume 769, pp. 241–246. [Google Scholar]
  93. Hassan, N.M.; El-Bastawisy, Z.M.; El-Sayed, A.K.; Ebeed, H.T.; Alla, M.M.N. Roles of dehydrin genes in wheat tolerance to drought stress. J. Adv. Res. 2015, 6, 179–188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Islam, M.; Begum, M.C.; Kabir, A.H.; Alam, M.F. Molecular and biochemical mechanisms associated with differential responses to drought tolerance in wheat (Triticum aestivum L.). J. Plant Interact. 2015, 10, 195–201. [Google Scholar] [CrossRef]
  95. Mathew, I.; Shimelis, H.; Mwadzingeni, L.; Zengeni, R.; Mutema, M.; Chaplot, V. Variance components and heritability of traits related to root: Shoot biomass allocation and drought tolerance in wheat. Euphytica 2018, 214, 225. [Google Scholar] [CrossRef]
  96. Ober, E.S.; Alahmad, S.; Cockram, J.; Forestan, C.; Hickey, L.T.; Kant, J.; Maccaferri, M.; Marr, E.; Milner, M.; Pinto, F.; et al. Wheat root systems as a breeding target for climate resilience. Theor. Appl. Genet. 2021, 134, 1645–1662. [Google Scholar] [CrossRef] [PubMed]
  97. Wasson, A.P.; Richards, R.A.; Chatrath, R.; Misra, S.C.; Prasad, S.S.; Rebetzke, G.J.; Kirkegaard, J.A.; Christopher, J.; Watt, M. Traits and selection strategies to improve root systems and water uptake in water-limited wheat crops. J. Exp. Bot. 2012, 63, 3485–3498. [Google Scholar] [CrossRef] [Green Version]
  98. Peng, B.; Liu, X.; Dong, X.; Xue, Q.; Neely, C.B.; Marek, T.; Ibrahim, A.M.; Zhang, G.; Leskovar, D.I.; Rudd, J.C. Root morphological traits of winter wheat under contrasting environments. J. Agron. Crop Sci. 2019, 205, 571–585. [Google Scholar] [CrossRef]
  99. Meister, R.; Rajani, M.S.; Ruzicka, D.; Schachtman, D.P. Challenges of modifying root traits in crops for agriculture. Trends Plant Sci. 2014, 19, 779–788. [Google Scholar] [CrossRef]
  100. Ryser, P.; Eek, L. Consequences of phenotypic plasticity vs. interspecific differences in leaf and root traits for acquisition of aboveground and belowground resources. Am. J. Bot. 2000, 87, 402–411. [Google Scholar] [CrossRef]
  101. Faye, A.; Sine, B.; Chopart, J.L.; Grondin, A.; Lucas, M.; Diedhiou, A.G.; Gantet, P.; Cournac, L.; Min, D.; Audebert, A.; et al. Development of a model estimating root length density from root impacts on a soil profile in pearl millet (Pennisetum glaucum (L.) R. Br). Application to measure root system response to water stress in field conditions. PLoS ONE 2019, 14, e0214182. [Google Scholar] [CrossRef] [Green Version]
  102. Freschet, G.T.; Valverde-Barrantes, O.J.; Tucker, C.M.; Craine, J.M.; McCormack, M.L.; Violle, C.; Fort, F.; Blackwood, C.B.; Urban-Mead, K.R.; Iversen, C.M.; et al. Climate, soil and plant functional types as drivers of global fine-root trait variation. J. Ecol. 2017, 105, 1182–1196. [Google Scholar] [CrossRef] [Green Version]
  103. Eissenstat, D.M.; Kucharski, J.M.; Zadworny, M.; Adams, T.S.; Koide, R.T. Linking root traits to nutrient foraging in arbuscular mycorrhizal trees in a temperate forest. New Phytol. 2015, 208, 114–124. [Google Scholar] [CrossRef] [PubMed]
  104. Valverde-Barrantes, O.J.; Smemo, K.A.; Blackwood, C.B. Fine root morphology is phylogenetically structured, but nitrogen is related to the plant economics spectrum in temperate trees. Funct. Ecol. 2015, 29, 796–807. [Google Scholar] [CrossRef]
  105. El-Hassouni, K.; Alahmad, S.; Belkadi, B.; Filali-Maltouf, A.; Hickey, L.T.; Bassi, F.M. Root system architecture and its association with yield under different water regimes in durum wheat. Crop Sci. 2018, 58, 2331–2346. [Google Scholar] [CrossRef] [Green Version]
  106. Kamal, N.M.; Gorafi, Y.S.A.; Abdelrahman, M.; Abdellatef, E.; Tsujimoto, H. Stay-green trait: A prospective approach for yield potential, and drought and heat stress adaptation in globally important cereals. Int. J. Mol. Sci. 2019, 20, 5837. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Schachtman, D.P.; Goodger, J.Q. Chemical root to shoot signaling under drought. Trends Plant Sci. 2008, 13, 281–287. [Google Scholar] [CrossRef]
  108. Bengough, A.G.; McKenzie, B.M.; Hallett, P.D.; Valentine, T.A. Root elongation, water stress, and mechanical impedance: A review of limiting stresses and beneficial root tip traits. J. Exp. Bot. 2011, 62, 59–68. [Google Scholar] [CrossRef] [Green Version]
  109. Hunt, J.R.; Lilley, J.M.; Trevaskis, B.; Flohr, B.M.; Peake, A.; Fletcher, A.; Zwart, A.B.; Gobbett, D.; Kirkegaard, J.A. Early sowing systems can boost Australian wheat yields despite recent climate change. Nat. Clim. Chang. 2019, 9, 244–247. [Google Scholar] [CrossRef]
  110. Gao, S.Q.; Chen, M.; Xu, Z.S.; Zhao, C.P.; Li, L.; Xu, H.J.; Tang, Y.M.; Zhao, X.; Ma, Y.Z. The soybean GmbZIP1 transcription factor enhances multiple abiotic stress tolerances in transgenic plants. Plant Mol. Biol. 2011, 75, 537–553. [Google Scholar] [CrossRef]
  111. Fletcher, A.; Chenu, K. Change in biomass partitioning and transpiration efficiency in Australian wheat varieties over the last decades. In Proceedings of the 17th Australian Agronomy Conference, Hobart, Australia, 21–24 September 2015; Volume 4. [Google Scholar]
  112. Fletcher, A.; Chenu, K. Recent Improvements in Biomass Partitioning and Transpiration Efficiency of Modern Australian Wheat Varieties–Any Opportunity for the Future? 2016. Available online: https://grdc.com.au/resources-and-publications/grdc-update-papers/tab-content/grdc-update-papers/2016/03/recent-improvements-in-biomass-partitioning-and-transpiration-efficiency-of-wheat (accessed on 13 February 2022).
  113. Kirkegaard, J.A.; Lilley, J.M.; Howe, G.N.; Graham, J.M. Impact of subsoil water use on wheat yield. Aust. J. Agric. Res. 2007, 58, 303–315. [Google Scholar] [CrossRef]
  114. Deng, W.; Casao, M.C.; Wang, P.; Sato, K.; Hayes, P.M.; Finnegan, E.J.; Trevaskis, B. Direct links between the vernalization response and other key traits of cereal crops. Nat. Commun. 2015, 6, 5882. [Google Scholar] [CrossRef]
  115. Voss-Fels, K.P.; Robinson, H.; Mudge, S.R.; Richard, C.; Newman, S.; Wittkop, B.; Stahl, A.; Friedt, W.; Frisch, M.; Gabur, I.; et al. VERNALIZATION1 modulates root system architecture in wheat and barley. Mol. Plant 2018, 11, 226–229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Ghimire, B. In Situ Imaging of Root System Architecture to Improve Drought Tolerance and Yield in Spring Wheat (Triticum aestivum L.). Doctoral dissertation, Washington State University: Vancouver, WA, USA, 2017; 177p. [Google Scholar]
  117. Salimath, P.M.; Toker, C.; Sandhu, J.S.; Kumar, J.; Suma, B.; Yadav, S.S.; Bahl, P.N. Conventional breeding methods. Chickpea Breed. Manag. CAB Int. Wallingford 2007, 369, 369–390. [Google Scholar]
  118. Bhatta, M.; Morgounov, A.; Belamkar, V.; Baenziger, P.S. Genome-wide association study reveals novel genomic regions for grain yield and yield-related traits in drought-stressed synthetic hexaploid wheat. Int. J. Mol. Sci. 2018, 19, 3011. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Aberkane, H.; Payne, T.; Kishi, M.; Smale, M.; Amri, A.; Jamora, N. Transferring diversity of goat grass to farmers’ fields through the development of synthetic hexaploid wheat. Food Secur. 2020, 12, 1017–1033. [Google Scholar] [CrossRef]
  120. Kirigwi, F.M.; Van Ginkel, M.A.R.T.I.N.; Brown-Guedira, G.; Gill, B.S.; Paulsen, G.M.; Fritz, A.K. Markers associated with a QTL for grain yield in wheat under drought. Mol. Breed. 2007, 20, 401–413. [Google Scholar] [CrossRef]
  121. Balouchi, H.R. Screening wheat parents of mapping population for heat and drought tolerance, detection of wheat genetic variation. Int. J. Nucl. Quantum Eng. 2010, 4, 63–73. [Google Scholar]
  122. Edwards, D.; Wilcox, S.; Barrero, R.A.; Fleury, D.; Cavanagh, C.R.; Forrest, K.L.; Hayden, M.J.; Moolhuijzen, P.; Keeble-Gagnère, G.; Bellgard, M.I.; et al. Bread matters: A national initiative to profile the genetic diversity of Australian wheat. Plant Biotechnol. J. 2012, 10, 703–708. [Google Scholar] [CrossRef] [Green Version]
  123. Tausz-Posch, S.; Seneweera, S.; Norton, R.M.; Fitzgerald, G.J.; Tausz, M. Can a wheat cultivar with high transpiration efficiency maintain its yield advantage over a near-isogenic cultivar under elevated CO2? Field Crops Res. 2012, 133, 160–166. [Google Scholar] [CrossRef]
  124. Yadav, D.; Shavrukov, Y.; Bazanova, N.; Chirkova, L.; Borisjuk, N.; Kovalchuk, N.; Ismagul, A.; Parent, B.; Langridge, P.; Hrmova, M.; et al. Constitutive overexpression of the TaNF-YB4 gene in transgenic wheat significantly improves grain yield. J. Exp. Bot. 2015, 66, 6635–6650. [Google Scholar] [CrossRef] [Green Version]
  125. Blum, A.; Shpiler, L.; Golan, G.; Mayer, J. Yield stability and canopy temperature of wheat genotypes under drought-stress. Field Crops Res. 1989, 22, 289–296. [Google Scholar] [CrossRef]
  126. Acreche, M.M.; Briceño-Félix, G.; Sánchez, J.A.M.; Slafer, G.A. Radiation interception and use efficiency as affected by breeding in Mediterranean wheat. Field Crops Res. 2009, 110, 91–97. [Google Scholar] [CrossRef]
  127. Swaminathan, M.S. Norman Borlaug and a hunger-free world. Resonance 2014, 19, 109–115. [Google Scholar] [CrossRef]
  128. Lopes, M.S.; El-Basyoni, I.; Baenziger, P.S.; Singh, S.; Royo, C.; Ozbek, K.; Aktas, H.; Ozer, E.; Ozdemir, F.; Manickavelu, A.; et al. Exploiting genetic diversity from landraces in wheat breeding for adaptation to climate change. J. Exp. Bot. 2015, 66, 3477–3486. [Google Scholar] [CrossRef] [PubMed]
  129. Liu, C.; Sukumaran, S.; Claverie, E.; Sansaloni, C.; Dreisigacker, S.; Reynolds, M. Genetic dissection of heat and drought stress QTLs in phenology-controlled synthetic-derived recombinant inbred lines in spring wheat. Mol. Breed. 2019, 39, 34. [Google Scholar] [CrossRef]
  130. Song, Q.; Liu, C.; Bachir, D.G.; Chen, L.; Hu, Y.G. Drought resistance of new synthetic hexaploid wheat accessions evaluated by multiple traits and antioxidant enzyme activity. Field Crops Res. 2017, 210, 91–103. [Google Scholar] [CrossRef]
  131. Rosyara, U.; Kishii, M.; Payne, T.; Sansaloni, C.P.; Singh, R.P.; Braun, H.J.; Dreisigacker, S. Genetic contribution of synthetic hexaploid wheat to CIMMYT’s spring bread wheat breeding germplasm. Sci. Rep. 2019, 9, 12355. [Google Scholar] [CrossRef] [Green Version]
  132. Tang, Y.; Rosewarne, G.M.; Li, C.; Wu, X.; Yang, W.; Wu, C. Physiological factors underpinning grain yield improvements of synthetic-derived wheat in southwestern China. Crop Sci. 2015, 55, 98–112. [Google Scholar] [CrossRef]
  133. Nevo, E.; Chen, G. Drought and salt tolerances in wild relatives for wheat and barley improvement. Plant Cell Environ. 2010, 33, 670–685. [Google Scholar] [CrossRef]
  134. Budak, H.; Hussain, B.; Khan, Z.; Ozturk, N.Z.; Ullah, N. From genetics to functional genomics: Improvement in drought signaling and tolerance in wheat. Front. Plant Sci. 2015, 6, 1012. [Google Scholar] [CrossRef] [Green Version]
  135. Pour-Aboughadareh, A.; Ahmadi, J.; Mehrabi, A.A.; Etminan, A.; Moghaddam, M.; Siddique, K.H. Physiological responses to drought stress in wild relatives of wheat: Implications for wheat improvement. Acta Physiol. Plant. 2017, 39, 106. [Google Scholar] [CrossRef]
  136. Pour-Aboughadareh, A.; Omidi, M.; Naghavi, M.R.; Etminan, A.; Mehrabi, A.A.; Poczai, P. Wild relatives of wheat respond well to water deficit stress: A comparative study of antioxidant enzyme activities and their encoding gene expression. Agriculture 2020, 10, 415. [Google Scholar] [CrossRef]
  137. Alain, B.; William, A.; Maarten, V.G. The World Wheat Book: A History of Wheat Breeding; International Maize and Wheat Improvement Center (CIMMYT): Lavoisier, France, 2011; Volume 2. [Google Scholar]
  138. Sadras, V.O.; 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 Pasture Sci. 2011, 62, 533–549. [Google Scholar] [CrossRef]
  139. Mondal, S.; Dutta, S.; Crespo-Herrera, L.; Huerta-Espino, J.; Braun, H.J.; Singh, R.P. Fifty years of semi-dwarf spring wheat breeding at CIMMYT: Grain yield progress in optimum, drought and heat stress environments. Field Crops Res. 2020, 250, 107757. [Google Scholar] [CrossRef]
  140. Braun, H.J.; Rajaram, S.; Ginkel, M. CIMMYT’s approach to breeding for wide adaptation. In Adaptation in Plant Breeding; Springer: Dordrecht, The Netherlands, 1997; pp. 197–205. [Google Scholar]
  141. Tadesse, W.; Sanchez-Garcia, M.; Assefa, S.G.; Amri, A.; Bishaw, Z.; Ogbonnaya, F.C.; Baum, M. Genetic gains in wheat breeding and its role in feeding the world. Crop Breed. Genet. Genome 2019, 1, 1–28. [Google Scholar]
  142. Lantican, M.; Braun, H.J.; Payne, T.S.; Singh, R.G.; Sonder, K.; Baum, M.; Ginkel, M.V.; Erenstein, O. Impacts of International Wheat Improvement Research: 1994–2014; International Maize and Wheat Improvement Center (CIMMYT): Mexico City, Mexico, 2016; pp. 1–73. [Google Scholar]
  143. Edae, E.A.; Byrne, P.F.; Haley, S.D.; Lopes, M.S.; Reynolds, M.P. Genome-wide association mapping of yield and yield components of spring wheat under contrasting moisture regimes. Theor. Appl. Genet. 2014, 127, 791–807. [Google Scholar] [CrossRef]
  144. Belete, Y.; Shimelis, H.; Laing, M.; Mathew, I. Genetic diversity and population structure of bread wheat genotypes determined via phenotypic and SSR marker analyses under drought-stress conditions. J. Crop Improv. 2021, 35, 303–325. [Google Scholar] [CrossRef]
  145. Shewry, P.R. Wheat. J. Exp. Bot. 2009, 60, 1537–1553. [Google Scholar] [CrossRef]
  146. Shewry, P.R.; Hey, S.J. The contribution of wheat to human diet and health. Food Energy Secur. 2015, 4, 178–202. [Google Scholar] [CrossRef]
  147. Lin, J.Y. Rural reforms and agricultural growth in China. Am. Econ. Rev. 1992, 82, 34–51. [Google Scholar]
  148. Lumpkin, T.A. How a Gene from Japan Revolutionized the World of Wheat: CIMMYT’s Quest for Combining Genes to Mitigate Threats to Global Food Security; Ogihara, Y., Takumi, S., Handa, H., Eds.; Advances in wheat genetics: From genome to field; Springer: Tokyo, Japan, 2015; pp. 13–20. [Google Scholar]
  149. Pingali, P.L. Green revolution: Impacts, limits, and the path ahead. Proc. Natl. Acad. Sci. USA 2012, 109, 12302–12308. [Google Scholar] [CrossRef] [Green Version]
  150. Flachs, A. Green Revolution. Encyclopedia of Food and Agricultural Ethics. 2016. Available online: https://www.researchgate.net/profile/Andrew_Flachs/publication/308013146_Green_Revolution/links/5bfee50c45851523d153196e/Green-Revolution.pdf (accessed on 30 January 2022).
  151. Olivares-Villegas, J.J.; Reynolds, M.P.; McDonald, G.K. Drought-adaptive attributes in the Seri/Babax hexaploid wheat population. Funct. Plant Biol. 2007, 34, 189–203. [Google Scholar] [CrossRef] [PubMed]
  152. Tahmasebi, S.; Heidari, B.; Pakniyat, H.; Jalal Kamali, M.R. Independent and combined effects of heat and drought stress in the Seri M82× Babax bread wheat population. Plant Breed. 2014, 133, 702–711. [Google Scholar] [CrossRef]
  153. Tahmasebi, S.; Heidari, B.; Pakniyat, H.; Dadkhodaie, A. Consequences of 1BL/1RS translocation on agronomic and physiological traits in wheat. Cereal Res. Commun. 2015, 43, 554–566. [Google Scholar] [CrossRef] [Green Version]
  154. Hassan, F.S.C.; Solouki, M.; Fakheri, B.A.; Nezhad, N.M.; Masoudi, B. Mapping QTLs for physiological and biochemical traits related to grain yield under control and terminal heat stress conditions in bread wheat (Triticum aestivum L.). Physiol. Mol. Biol. Plants 2018, 24, 1231–1243. [Google Scholar] [CrossRef] [PubMed]
  155. Manschadi, A.M.; Christopher, J.T.; Hammer, G.L.; Devoil, P. Experimental and modelling studies of drought-adaptive root architectural traits in wheat (Triticum aestivum L.). Plant Biosyst. 2010, 144, 458–462. [Google Scholar] [CrossRef]
  156. Mondal, S. CIMMYT Wheat Breeding: Continuing Genetic Gains through the Development of High Yielding and Nutritious Varieties; CIMMYT: El Batán, Mexico, 2018; Volume 144, pp. 458–462. [Google Scholar]
  157. Klopff, M. Out with the Old, in with the New. International Maize and Wheat Improvement Center. 2021. Available online: https://www.cimmyt.org/blogs/out-with-the-old-in-with-the-new/ (accessed on 29 February 2022).
  158. Breseghello, F.; Coelho, A.S.G. Traditional and modern plant breeding methods with examples in rice (Oryza sativa L.). J. Agric. Food Chem. 2013, 61, 8277–8286. [Google Scholar] [CrossRef]
  159. Xynias, I.N.; Mylonas, I.; Korpetis, E.G.; Ninou, E.; Tsaballa, A.; Avdikos, I.D.; Mavromatis, A.G. Durum wheat breeding in the Mediterranean region: Current status and future prospects. Agronomy 2020, 10, 432. [Google Scholar] [CrossRef] [Green Version]
  160. Acquaah, G. Principles of Plant Genetics and Breeding, 2nd ed.; Wiley-Blackwell: Oxford, UK, 2012; 740p. [Google Scholar]
  161. Al-Azab, K.F. Improving Wheat for Drought Tolerance by Using Hybridization and Mutation Breeding Procedures. Doctoral Dissertation, Faculty of Agriculture Cairo University, Giza, Egypt, 2013; 266p. [Google Scholar]
  162. Farag, H.I.A. Efficiency of three methods of selection in wheat breeding under saline stress conditions. J. Plant Breed. 2013, 17, 85–95. [Google Scholar] [CrossRef]
  163. Acquaah, G. Principles of Plant Genetics and Breeding, 1st ed.; John Wiley & Sons: Hoboken, NJ, USA, 2009; 584p. [Google Scholar]
  164. Allard, R.W. “Plant Breeding”. Encyclopedia Britannica. 2019. Available online: https://www.britannica.com/science/plant-breeding (accessed on 3 March 2022).
  165. Distelfeld, A.; Uauy, C.; Fahima, T.; Dubcovsky, J. Physical map of the wheat high-grain protein content gene Gpc-B1 and development of a high-throughput molecular marker. New Phytol. 2006, 169, 753–763. [Google Scholar] [CrossRef] [Green Version]
  166. Khan, M.A.; Iqbal, M.; Jameel, M.; Nazeer, W.; Shakir, S.; Aslam, M.T.; Iqbal, B. Potentials of molecular based breeding to enhance drought tolerance in wheat (Triticum aestivum L.). Afr. J. Biotechnol. 2011, 10, 11340–11344. [Google Scholar]
  167. Collard, B.C.; Mackill, D.J. Marker-assisted selection: An approach for precision plant breeding in the twenty-first century. Philos. Trans. R. Soc. B Biol. Sci. 2008, 363, 557–572. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Huseynova, I.M.; Rustamova, S.M. Screening for drought stress tolerance in wheat genotypes using molecular markers. Proc. ANAS 2010, 65, 132–139. [Google Scholar]
  169. Tessema, B.B.; Liu, H.; Sørensen, A.C.; Andersen, J.R.; Jensen, J. Strategies using genomic selection to increase genetic gain in breeding programs for wheat. Front. Genet. 2020, 11, 1–12. [Google Scholar] [CrossRef] [PubMed]
  170. Lozada, D.N.; Ward, B.P.; Carter, A.H. Gains through selection for grain yield in a winter wheat breeding program. PLoS ONE 2020, 15, e0221603. [Google Scholar] [CrossRef]
  171. Tsai, H.Y.; Janss, L.L.; Andersen, J.R.; Orabi, J.; Jensen, J.D.; Jahoor, A.; Jensen, J. Genomic prediction and GWAS of yield, quality and disease-related traits in spring barley and winter wheat. Sci. Rep. 2020, 10, 3347. [Google Scholar] [CrossRef]
  172. Gill, B.S.; Appels, R.; Botha-Oberholster, A.M.; Buell, C.R.; Bennetzen, J.L.; Chalhoub, B.; Chumley, F.; Dvorák, J.; Iwanaga, M.; Keller, B.; et al. A workshop report on wheat genome sequencing: International Genome Research on Wheat Consortium. Genetics 2004, 168, 1087–1096. [Google Scholar] [CrossRef] [Green Version]
  173. Sticklen, M. Transgenic, cisgenic, intragenic and subgenic crops. Adv. Crop Sci. Technol. 2015, 3, e123. [Google Scholar]
  174. Puchta, H. Using CRISPR/Cas in three dimensions: Towards synthetic plant genomes, transcriptomes and epigenomes. Plant J. 2016, 87, 5–15. [Google Scholar] [CrossRef] [Green Version]
  175. Carroll, D. Focus: Genome editing: Genome editing: Past, present and future. Yale J. Biol. Med. 2017, 90, 653–659. [Google Scholar]
  176. Khan, M.Z.; Zaidi, S.S.E.A.; Amin, I.; Mansoor, S. A CRISPR way for fast-forward crop domestication. Trends Plant Sci. 2019, 24, 293–296. [Google Scholar] [CrossRef] [PubMed]
  177. Hansen, M.K. Genetic Engineering Is Not an Extension of Conventional Plant Breeding; How Genetic Engineering Differs from Conventional Breeding, Hybridization, Wide Crosses and Horizontal Gene Transfer; Consumer Policy Institute/Consumers Union: New York, NY, USA, 2000; pp. 1–15. [Google Scholar]
  178. Carroll, D. Genome engineering with zinc-finger nucleases. Genetics 2011, 188, 773–782. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Gaj, T.; Gersbach, C.A.; Barba, I.I.I.; Carlos, F. ZFN, TALEN and CRISPR/Cas-based methods for genome engineering. Trends Biotechnol. 2013, 31, 397–405. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Weinthal, D.M.; Gürel, F. Plant genome editing and its applications in cereals. In Genetic Engineering: An Insight into the Strategies and Applications; IntechOpen: London, UK, 2016; pp. 63–73. [Google Scholar]
  181. Belhaj, K.; Chaparro-Garcia, A.; Kamoun, S.; Nekrasov, V. Plant genome editing made easy: Targeted mutagenesis in model and crop plants using the CRISPR/Cas system. Plant Methods 2013, 9, 39. [Google Scholar] [CrossRef] [Green Version]
  182. Mali, P.; Aach, J.; Stranges, P.B.; Esvelt, K.M.; Moosburner, M.; Kosuri, S.; Yang, L.; Church, G.M. CAS9 transcriptional activators for target specificity screening and paired nickases for cooperative genome engineering. Nat. Biotechnol. 2013, 31, 833–838. [Google Scholar] [CrossRef]
  183. Tsai, S.Q.; Wyvekens, N.; Khayter, C.; Foden, J.A.; Thapar, V.; Reyon, D.; Goodwin, M.J.; Aryee, M.J.; Joung, J.K. Dimeric CRISPR RNA-guided FokI nucleases for highly specific genome editing. Nat. Biotechnol. 2014, 32, 569–576. [Google Scholar] [CrossRef] [Green Version]
  184. Zetsche, B.; Gootenberg, J.S.; Abudayyeh, O.O.; Slaymaker, I.M.; Makarova, K.S.; Essletzbichler, P.; Volz, S.E.; Joung, J.; Van Der Oost, J.; Re`gev, A.; et al. Cpf1 is a single RNA-guided endonuclease of a class 2 CRISPR-Cas system. Cell 2015, 163, 759–771. [Google Scholar] [CrossRef] [Green Version]
  185. Martignago, D.; Rico-Medina, A.; Blasco-Escámez, D.; Fontanet-Manzaneque, J.B.; Caño-Delgado, A.I. Drought resistance by engineering plant tissue-specific responses. Front. Plant Sci. 2020, 10, 1–19. [Google Scholar] [CrossRef]
  186. Ashraf, A.; Rehman, O.U.; Muzammil, S.; Léon, J.; Naz, A.A.; Rasool, F.; Ali, G.M.; Zafar, Y.; Khan, M.R. Evolution of Deeper Rooting 1-like homoeologs in wheat entails the C-terminus mutations as well as gain and loss of auxin response elements. PLoS ONE 2019, 14, e0214145. [Google Scholar]
  187. Guseman, J.M.; Webb, K.; Srinivasan, C.; Dardick, C. DRO 1 influences root system architecture in Arabidopsis and Prunus species. Plant J. 2017, 89, 1093–1105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Wei, B.; Jing, R.; Wang, C.; Chen, J.; Mao, X.; Chang, X.; Jia, J. Dreb1 genes in wheat (Triticum aestivum L.): Development of functional markers and gene mapping based on SNPs. Mol. Breed. 2009, 23, 13–22. [Google Scholar] [CrossRef]
  189. Moon, S.J.; Min, M.K.; Kim, J.; Kim, D.Y.; Yoon, I.S.; Kwon, T.R.; Byun, M.O.; Kim, B.G. Ectopic expression of OsDREB1G, a member of the OsDREB1 subfamily, confers cold stress tolerance in rice. Front. Plant Sci. 2019, 10, 1–11. [Google Scholar] [CrossRef] [PubMed]
  190. Joshi, R.K.; Bharat, S.S.; Mishra, R. Engineering drought tolerance in plants through CRISPR/Cas genome editing. 3 Biotech 2020, 10, 400. [Google Scholar] [CrossRef]
  191. Badhan, S.; Ball, A.S.; Mantri, N. First report of CRISPR/Cas9 mediated DNA-free editing of 4CL and RVE7 genes in chickpea protoplasts. Int. J. Mol. Sci. 2021, 22, 396. [Google Scholar] [CrossRef]
  192. Meyer, R.S.; Purugganan, M.D. Evolution of crop species: Genetics of domestication and diversification. Nat. Rev. Genet. 2013, 14, 840–852. [Google Scholar] [CrossRef]
  193. Khan, M.A.; Iqbal, M. Breeding for drought tolerance in wheat (Triticum aestivum L.): Constraints and future prospects. Front. Agric. China 2011, 5, 31–34. [Google Scholar] [CrossRef]
  194. Pandey, K.; Dangi, R.; Prajapati, U.; Kumar, S.; Maurya, N.K.; Singh, A.V.; Pandey, A.K.; Singh, J.; Rajan, R. Advance breeding and biotechnological approaches for crop improvement: A review. Int. J. Chem. Stud. 2019, 7, 837–841. [Google Scholar]
  195. Holme, I.B.; Wendt, T.; Holm, P.B. Intragenesis and cisgenesis as alternatives to transgenic crop development. Plant Biotechnol. J. 2013, 11, 395–407. [Google Scholar] [CrossRef]
  196. Ahmad, P.; Ashraf, M.; Younis, M.; Hu, X.; Kumar, A.; Akram, N.A.; Al-Qurainy, F. Role of transgenic plants in agriculture and biopharming. Biotechnol. Adv. 2012, 30, 524–540. [Google Scholar] [CrossRef]
  197. Jacobsen, E.; Schouten, H.J. Cisgenesis: An important sub-invention for traditional plant breeding companies. Euphytica 2009, 170, 235–247. [Google Scholar] [CrossRef] [Green Version]
  198. Gao, S.; Xu, H.; Cheng, X.; Chen, M.; Xu, Z.; Li, L.; Ye, X.; Du, L.; Hao, X.; Ma, Y. Improvement of wheat drought and salt tolerance by expression of a stress-inducible transcription factor GmDREB of soybean (Glycine max). Chin. Sci. Bull. 2005, 50, 2714–2723. [Google Scholar] [CrossRef]
  199. Zang, X.; Geng, X.; He, K.; Wang, F.; Tian, X.; Xin, M.; Yao, Y.; Hu, Z.; Ni, Z.; Sun, Q.; et al. Overexpression of the wheat (Triticum aestivum L.) TaPEPKR2 gene enhances heat and dehydration tolerance in both wheat and Arabidopsis. Front. Plant Sci. 2018, 12, 1710. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  200. Xue, G.G.; Way, H.H.; Richardson, T.; Drenth, J.; Joyce, P.A.; McIntyre, C.L. Overexpression of TaNAC69 leads to enhanced transcript levels of stress up-regulated genes and dehydration tolerance in bread wheat. Mol. Plant 2011, 4, 697–712. [Google Scholar] [CrossRef]
  201. Gadaleta, A.; Giancaspro, A.; Blechl, A.E.; Blanco, A. A transgenic durum wheat line that is free of marker genes and expresses 1DY10. J. Cereal Sci. 2008, 48, 439–445. [Google Scholar] [CrossRef]
  202. Jacobsen, E.; Schouten, H.J. Cisgenesis, a new tool for traditional plant breeding, should be exempted from the regulation on genetically modified organisms in a step by step approach. Potato Res. 2008, 51, 75–88. [Google Scholar] [CrossRef]
  203. The Royal Society. What GM Crops Are Currently Being Grown and Where? 2022. Available online: https://royalsociety.org/topics-policy/projects/gm-plants/what-gm-crops-are-currently-being-grown-and-where (accessed on 19 April 2022).
  204. Khan, S.; Sumera, A.; Shaobo, Y.; Min, S.; Zhenping, Y.; Gao, Z. Development of drought-tolerant transgenic wheat: Achievements and limitations. Int. J. Mol. Sci. 2019, 20, 3350. [Google Scholar] [CrossRef] [Green Version]
  205. Passioura, J.B. Phenotyping for drought tolerance in grain crops: When is it useful to breeders? Funct. Plant Biol. 2012, 39, 851–859. [Google Scholar] [CrossRef]
  206. Araus, J.L.; Serret, M.D.; Lopes, M.S. Transgenic solutions to increase yield and stability in wheat: Shining hope or flash in the pan? J. Exp. Bot. 2019, 70, 1419–1424. [Google Scholar] [CrossRef] [Green Version]
  207. Choudhary, M.; Wani, S.H.; Kumar, P.; Bagaria, P.K.; Rakshit, S.; Roorkiwal, M.; Varshney, R.K. QTLian Breeding for Climate Resilience in Cereals: Progress and prospects. Funct. Integr. Genom. 2019, 19, 685–701. [Google Scholar] [CrossRef]
  208. Rai, N.; Bellundagi, A.; Kumar, P.K.; Kalasapura Thimmappa, R.; Rani, S.; Sinha, N.; Krishna, H.; Jain, N.; Singh, G.P.; Singh, P.K.; et al. Marker-assisted backcross breeding for improvement of drought tolerance in bread wheat (Triticum aestivum L. em Thell). Plant Breed. 2018, 137, 514–526. [Google Scholar] [CrossRef]
  209. Todkar, L.; Singh, G.P.; Jain, N.; Singh, P.K.; Prabhu, K.V. Introgression of drought tolerance QTLs through marker assisted backcross breeding in wheat (Triticum aestivum L.). Indian J. Genet. Plant Breed. 2020, 80, 209–212. [Google Scholar] [CrossRef]
  210. Sivamani, E.; Bahieldin, A.; Wraith, J.M.; Al-Niemi, T.; Dyer, W.E.; Ho, T.H.D.; Qu, R. Improved biomass productivity and water use efficiency under water deficit conditions in transgenic wheat constitutively expressing the barley HVA1 gene. Plant Sci. 2000, 155, 1–9. [Google Scholar] [CrossRef]
  211. Li, L.; Zheng, M.; Deng, G.; Liang, J.; Zhang, H.; Pan, Z.; Long, H.; Yu, M. Overexpression of AtHDG11 enhanced drought tolerance in wheat (Triticum aestivum L.). Mol. Breed. 2016, 36, 23. [Google Scholar] [CrossRef]
  212. Zang, X.; Geng, X.; Wang, F.; Liu, Z.; Zhang, L.; Zhao, Y.; Tian, X.; Ni, Z.; Yao, Y.; Xin, M.; et al. Overexpression of wheat ferritin gene TaFER-5B enhances tolerance to heat stress and other abiotic stresses associated with the ROS scavenging. BMC Plant Biol. 2017, 17, 14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Qin, N.; Xu, W.; Hu, L.; Li, Y.; Wang, H.; Qi, X.; Fang, Y.; Hua, X. Drought tolerance and proteomics studies of transgenic wheat containing the maize C 4 phosphoenolpyruvate carboxylase (PEPC) gene. Protoplasma 2015, 253, 1503–1512. [Google Scholar] [CrossRef]
  214. González, F.F.; Capella, M.; Ribichich, K.K.; Curín, F.; Giacomelli, J.J.; Ayala, F.; Watson, G.; Otegui, M.E.; Chan, R.L. Field-grown transgenic wheat expressing the sunflower gene HaHB4 significantly outyields the wild type. J. Exp. Bot. 2019, 70, 1669–1681. [Google Scholar] [CrossRef] [Green Version]
  215. Yu, T.T.; Xu, Z.Z.; Guo, J.J.; Wang, Y.Y.; Abernathy, B.; Fu, J.J.; Chen, X.; Zhou, Y.Y.; Chen, M.; Ye, X.X.; et al. Improved drought tolerance in wheat plants overexpressing a synthetic bacterial cold shock protein gene SeCspA. Sci. Rep. 2017, 7, srep44050. [Google Scholar] [CrossRef]
  216. Zhang, H.; Mao, X.; Wang, C.; Jing, R. Overexpression of a common wheat gene TaSnRK2. 8 enhances tolerance to drought, salt and low temperature in Arabidopsis. PloS ONE 2010, 5, e16041. [Google Scholar] [CrossRef] [Green Version]
  217. Luang, S.; Sornaraj, P.; Bazanova, N.; Jia, W.; Eini, O.; Hussain, S.S.; Kovalchuk, N.; Agarwal, P.; Hrmova, M.; Lopato, S. The wheat TabZIP2 transcription factor is activated by the nutrient starvation-responsive SnRK3/CIPK protein kinase. Plant Mol. Biol. 2018, 96, 543–561. [Google Scholar] [CrossRef]
  218. de Vetten, N.; Wolters, A.; Raemakers, K.; van der Meer, I.; ter Stege, R.; Heeres, E.; Heeres, P.; Visser, R. A transformation method forobtaining marker-free plants of a cross-pollinating and vegetatively propagated crop. Nat. Biotechnol. 2003, 21, 439–442. [Google Scholar] [CrossRef] [PubMed]
  219. Chawla, R.; Shakya, R.; Rommens, C.M. Tuber-specific silencing of asparagine synthetase-1reduces the acrylamide-forming potential of potatoes grown in the field without affecting tuber shape and yield. Plant Biotechnol. J. 2012, 10, 913–924. [Google Scholar] [CrossRef] [PubMed]
  220. Rommens, C.M.; Yan, H.; Swords, K.; Richael, C.; Ye, J. Low-acrylamide French fries and potato chips. Plant Biotechnol. J. 2008, 6, 843–853. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  221. Rommens, C.M.; Humara, J.M.; Ye, J.; Yan, H.; Richael, C.; Zhang, L.; Perry, R.; Swords, K. Crop improvement through modification of the plant’sown genome. Plant Physiol. 2004, 135, 421–431. [Google Scholar] [CrossRef] [Green Version]
  222. Rommens, C.M.; Ye, J.; Richael, C.; Swords, K. Improving potato storage and processing characteristics through all-native DNA transformation. J. Agric. Food Chem. 2006, 54, 9882–9887. [Google Scholar] [CrossRef]
  223. Joshi, S.G.; Schaart, J.G.; Groenwold, R.; Jacobsen, E.; Schouten, H.J.; Krens, F.A. Functional analysis and expression profiling of HcrVf1 and HcrVf2 for development of scab resistant cisgenic and intragenic apples. Plant Mol. Biol. 2011, 75, 579–591. [Google Scholar] [CrossRef] [Green Version]
  224. Schaart, J.G.; Krens, F.A.; Pelgrom, K.T.B.; Mendes, O.; Rouwendal, G.J.A. Effective production of marker-free transgenic strawberry plants using inducible site-specific recombination and a bifunctional selectable marker-gene. Plant Biotechnol. J. 2004, 2, 233–240. [Google Scholar] [CrossRef]
  225. Weeks, J.T.; Ye, J.; Rommens, C.M. Development of an in planta method for transformation of alfalfa (Medicago sativa). Transgen. Res. 2008, 17, 587–597. [Google Scholar] [CrossRef]
  226. Holme, I.B.; Dionisio, G.; Brinch-Pedersen, H.; Wendt, T.; Madsen, C.K.; Vincze, E.; Holm, P.B. Cisgenic barley with improved phytase activity. Plant Biotechnol. J. 2012, 10, 237–247. [Google Scholar] [CrossRef]
  227. Dhekney, S.A.; Li, Z.T.; Gray, D.J. Grapevines engineered to express cisgenic Vitis vinifera thaumatin-like protein exhibit fungal disease resistance. Vitr. Cell. Dev. Biol.-Plant 2011, 47, 458–466. [Google Scholar] [CrossRef]
  228. Haverkort, A.J.; Struik, P.C.; Visser, R.G.F.; Jacobsen, E.J.P.R. Applied biotechnology to combat late blight in potato caused by Phytophthora infestans. Potato Res. 2009, 52, 249–264. [Google Scholar] [CrossRef]
  229. Vanblaere, T.; Szankowski, I.; Schaart, J.; Schouten, H.; Flachowsky, H.; Broggini, G.A.L.; Gessler, C. The development of a cisgenic apple plant. J. Biotechnol. 2011, 154, 304–311. [Google Scholar] [CrossRef] [PubMed]
  230. Krens, F.A.; Schaart, J.G.; Van der Burgh, A.M.; Tinnenbroek-Capel, I.E.; Groenwold, R.; Kodde, L.P.; Broggini, G.A.; Gessler, C.; Schouten, H.J. Cisgenic apple trees; development, characterization, and performance. Front. Plant Sci. 2015, 6, 1–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  231. Han, K.M.; Dharmawardhana, P.; Arias, R.S.; Ma, C.; Busov, V.; Strauss, S.H. Gibberellin-associated cisgenes modify growth, stature and wood properties in Popu-lus. Plant Biotechnol. J. 2011, 9, 162–178. [Google Scholar] [CrossRef] [PubMed]
  232. Roy, S.J.; Tucker, E.J.; Tester, M. Genetic analysis of abiotic stress tolerance in crops. Curr. Opin. Plant Biol. 2011, 14, 232–239. [Google Scholar] [CrossRef]
  233. Goel, S.; Signh, K.; Grewal, S.; Nath, M. Impact of ¨omics¨ in improving drought tolerance in wheat. Crit. Rev. Plant Sci. 2020, 39, 222–235. [Google Scholar] [CrossRef]
  234. Gupta, P.K.; Balyan, H.S.; Gahlaut, V. QTL analysis for drought-tolerance in wheat: Present status and future possibilities. Agronomy 2017, 7, 5. [Google Scholar] [CrossRef]
  235. Kulkarni, M.; Soolanayakanahally, R.; Ogawa, S.; Uga, Y.; Selvaraj, M.G.; Kagale, S. Drought response in wheat: Key genes and regulatory mechanisms controlling root system architecture and transpiration efficiency. Front. Chem. 2017, 5, 106. [Google Scholar] [CrossRef] [Green Version]
  236. Yang, S.; Vanderbeld, B.; Wan, J.; Huang, Y. Narrowing down the targets: Towards successful genetic engineering of drought-tolerance crops. Mol. Plant 2010, 3, 469–490. [Google Scholar] [CrossRef]
  237. Huseynova, I.M.; Rustamova, S.M.; Mammadov, A.C. Identification of Dreb 1 genes involved in drought tolerance in wheat [Triticum L.]. In Photosynthesis Research for Food; Springer: Berlin/Heidelberg, Germany, 2013; pp. 552–555. [Google Scholar]
  238. Rong, W.; Qi, L.; Wang, A.; Ye, X.; Du, L.; Liang, H.; Xin, Z.; Zhang, Z. The ERF transcription factor TaERF3 promotes tolerance to salt and drought stresses in wheat. Plant Biotechnol. J. 2014, 12, 468–479. [Google Scholar] [CrossRef]
  239. An, Y.Q.; Lin, R.M.; Wang, F.T.; Feng, J.; Xu, Y.F.; Xu, S.C. Molecular cloning of a new wheat calreticulin gene TaCRT1 and expression analysis in plant defense responses and abiotic stress resistance. Genet. Mol. Res. 2011, 10, 3576–3585. [Google Scholar] [CrossRef] [PubMed]
  240. Chang, H.; Chen, D.; Kam, J.; Richardson, T.; Drenth, J.; Guo, X.; Mclntyre, C.L.; Chai, S.; Rae, A.L.; Xue, G.P. Abiotic stress upregulated TaZFP34 represses the expression of type-B response regulator and SHY2 genes and enhances root to shoot ratio in wheat. Plant Sci. 2016, 252, 88–102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  241. Zheng, J.; Yang, Z.; Madgwick, P.J.; Carmo-Silva, E.; Parry, M.A.; Hu, Y.-G. TaER expression is associated with transpiration efficiency traits and yield in bread wheat. PLoS ONE 2015, 10, e0128415. [Google Scholar] [CrossRef] [PubMed]
  242. Ding, W.; Fang, W.; Shi, S.; Zhao, Y.; Li, X.; Xiao, K. Wheat WRKY type transcription factor gene TaWRKY1 is essential in mediating drought tolerance associated with an ABA-dependent pathway. Plant Mol. Biol. Rep. 2016, 34, 1111–1126. [Google Scholar] [CrossRef]
  243. Wang, X.; Zeng, J.; Li, Y.; Rong, X.; Sun, J.; Sun, T.; Li, M.; Wang, L.; Feng, Y.; Chai, R.; et al. Expression of TaWRKY44, a wheat WRKY gene, in transgenic tobacco confers multiple abiotic stress tolerances. Front. Plant Sci. 2015, 6, 615. [Google Scholar] [CrossRef] [Green Version]
  244. Miao, L.L.; LI, Y.Y.; Zhang, H.J.; Zhang, H.J.; Liu, X.L.; Wang, J.Y.; Chang, X.P.; Mao, X.G.; Jing, R.L. TaSnRK2. 4 is a vital regulator in control of thousand-kernel weight and response to abiotic stress in wheat. J. Integr. Agric. 2021, 20, 46–54. [Google Scholar] [CrossRef]
  245. Wang, X.; Ren, Y.; Li, J.; Wang, Z.; Xin, Z.; Lin, T. Knock-down the expression of TaH2B-7D using virus-induced gene silencing reduces wheat drought tolerance. Biol. Res. 2019, 52, 14. [Google Scholar] [CrossRef]
  246. Cai, H.; Tian, S.; Liu, C.; Dong, H. Identification of a MYB3R gene involved in drought, salt and cold stress in wheat (Triticum aestivum L.). Gene 2011, 485, 14–152. [Google Scholar] [CrossRef]
  247. Zhou, S.; Hu, W.; Deng, X.; Ma, Z.; Chen, L.; Huang, C.; Wang, C.; Wang, J.; He, Y.; Yang, G.; et al. Overexpression of the wheat aquaporin gene, TaAQP7, enhances drought tolerance in transgenic tobacco. PLoS ONE 2012, 7, e52439. [Google Scholar] [CrossRef] [Green Version]
  248. Saad, A.S.; Li, X.; Li, H.P.; Huang, T.; Gao, C.S.; Guo, M.W.; Cheng, W.; Zhao, G.Y.; Liao, Y.C. A rice stress-responsive NAC gene enhances tolerance of transgenic wheat to drought and salt stresses. Plant Sci. 2013, 203, 33–40. [Google Scholar] [CrossRef]
  249. Khalid, M.; Afzal, F.; Gul, A.; Amir, R.; Subhani, A.; Ahmed, Z.; Mahmood, Z.; Xia, X.; Rasheed, A.; He, Z. Molecular characterization of 87 functional genes in wheat diversity panel and their association with phenotypes under well-watered and water limited conditions. Front. Plant Sci. 2019, 10, 1–15. [Google Scholar] [CrossRef] [PubMed]
  250. Li, Y.; Zhang, S.; Zhang, N.; Zhang, W.; Li, M.; Liu, B.; Shi, Z. MYB-CC transcription factor, TaMYBsm3, cloned from wheat is involved in drought tolerance. BMC Plant Biol. 2019, 19, 143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  251. Yang, J.; Zhang, G.; An, J.; Li, Q.; Chen, Y.; Zhao, X.; Wu, J.; Wang, Y.; Hao, Q.; Wang, W.; et al. Expansin gene TaEXPA2 positively regulates drought tolerance in transgenic wheat (Triticum aestivum L.). Plant Sci. 2020, 298, 110596. [Google Scholar] [CrossRef] [PubMed]
  252. Mao, H.; Li, S.; Wang, Z.; Cheng, X.; Li, F.; Mei, F.; Chen, N.; Kang, Z. Regulatory changes in TaSNAC8-6A are associated with drought tolerance in wheat seedlings. Plant Biotechnol. J. 2020, 18, 1078–1092. [Google Scholar] [CrossRef] [Green Version]
  253. Uga, Y.; Okuno, K.; Yano, M. Dro1, a major QTL involved in deep rooting of rice under upland field conditions. J. Exp. Bot. 2011, 62, 2485–2494. [Google Scholar] [CrossRef] [Green Version]
  254. Wang, J.; Wen, W.; Hanif, M.; Xia, X.; Wang, H.; Liu, S.; Liu, J.; Yang, L.; Cao, S.; He, Z. TaELF3-1DL, a homolog of ELF3, is associated with heading date in bread wheat. Mol. Breed. 2016, 36, 161. [Google Scholar] [CrossRef]
  255. Zhang, J.; Dell, B.; Biddulph, B.; Drake-Brockman, F.; Walker, E.; Khan, N.; Wong, D.; Hayden, M.; Appels, R. Wild-type alleles of Rht-B1 and Rht-D1 as independent determinants of thousand-grain weight and kernel number per spike in wheat. Mol. Breed. 2013, 32, 771–783. [Google Scholar] [CrossRef]
  256. Hu, M.J.; Zhang, H.P.; Cao, J.J.; Zhu, X.F.; Wang, S.X.; Jiang, H.; Wu, Z.Y.; Lu, J.; Chang, C.; Sun, G.L.; et al. Characterization of an IAA-glucose hydrolase gene TaTGW6 associated with grain weight in common wheat (Triticum aestivum L.). Mol. Breed. 2016, 36, 25. [Google Scholar] [CrossRef]
  257. Qin, L.; Zhao, J.; Li, T.; Hou, J.; Zhang, X.; Hao, C. TaGW2, a good reflection of wheat polyploidization and evolution. Front. Plant Sci. 2017, 8, 1–13. [Google Scholar] [CrossRef] [Green Version]
  258. 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. Mol. Breed. 2014, 34, 1097–1107. [Google Scholar] [CrossRef]
  259. Jiang, Q.; Hou, J.; Hao, C.; Wang, L.; Ge, H.; Dong, Y.; Zhang, X. The wheat (T. aestivum) sucrose synthase 2 gene (TaSus2) active in endosperm development is associated with yield traits. Funct. Integr. Genom. 2011, 11, 49–61. [Google Scholar] [CrossRef] [PubMed]
  260. Abou-Elwafa, S.F.; Shehzad, T. Genetic diversity, GWAS and prediction for drought and terminal heat stress tolerance in bread wheat (Triticum aestivum L.). Genet. Resour. Crop Evol. 2021, 68, 711–728. [Google Scholar] [CrossRef]
  261. Gizaw, S.A.; Godoy, J.G.V.; Garland-Campbell, K.; Carter, A.H. Genome-wide association study of yield and component traits in Pacific Northwest winter wheat. Crop Sci. 2018, 58, 2315–2330. [Google Scholar] [CrossRef]
  262. Ain, Q.-U.; Rasheed, A.; Anwar, A.; Mahmood, T.; Imtiaz, M.; Mahmood, T.; Xia, X.; He, Z.; Quraishi, U.M. Genome-wide association for grain yield under rainfed conditions in historical wheat cultivars from Pakistan. Front. Plant Sci. 2015, 6, 743. [Google Scholar] [CrossRef] [Green Version]
  263. Shokat, S.; Sehgal, D.; Vikram, P.; Liu, F.; Singh, S. Molecular markers associated with agro-physiological traits under terminal drought conditions in bread wheat. Int. J. Mol. Sci. 2020, 21, 3156. [Google Scholar] [CrossRef]
  264. Rahimi, Y.; Bihamta, M.R.; Taleei, A.; Alipour, H.; Ingvarsson, P.K. Genome-wide association study of agronomic traits in bread wheat reveals novel putative alleles for future breeding programs. BMC Plant Biol. 2019, 19, 541. [Google Scholar] [CrossRef] [Green Version]
  265. IWGSC Website. 2018. Available online: http://www.wheatgenome.org (accessed on 19 September 2021).
  266. Li, L.; Peng, Z.; Mao, X.; Wang, J.; Chang, X.; Reynolds, M.; Jing, R. Genome-wide association study reveals genomic regions controlling root and shoot traits at late growth stages in wheat. J. Exp. Bot. 2019, 124, 993–1006. [Google Scholar] [CrossRef]
  267. Mwadzingeni, L.; Shimelis, H.; Dube, E.; Laing, M.D.; Tsilo, T.J. Breeding wheat for drought tolerance: Progress and technologies. J. Integr. Agric. 2016, 15, 935–943. [Google Scholar] [CrossRef] [Green Version]
  268. Liu, X.; Li, R.; Chnag, X.; Jing, R. Mapping QTLs for seedling root traits in a double haploid wheat population under different water regimes. Euphytica 2013, 189, 51–66. [Google Scholar] [CrossRef]
  269. Kabir, M.R.; Liu, G.; Guan, P.; Wang, F.; Khan, A.A.; Ni, Z.; Yao, Y.; Hu, Z.; Xin, M.; Peng, H.; et al. Mapping QTLs associated with root traits using two different populations in wheat (Triticum aestivum L.). Euphytica 2015, 206, 175–190. [Google Scholar] [CrossRef]
  270. Liu, R.X.; Wu, F.K.; Xin, Y.I.; Yu, L.N.; Wang, Z.Q.; Liu, S.H.; Mei, D.E.N.G.; Jian, M.A.; Wei, Y.M.; Zheng, Y.L.; et al. Quantitative trait loci analysis for root traits in synthetic hexaploid wheat under drought stress conditions. J. Integr. Agric. 2020, 19, 1947–1960. [Google Scholar] [CrossRef]
  271. Yu, J.B.; Bai, G.H. Mapping quantitative trait loci for long coleoptile in Chinese wheat landrace Wangshuibai. Crop Sci. 2010, 50, 43–50. [Google Scholar] [CrossRef]
  272. Singh, K.; Shukla, S.; Kadam, S.; Semwal, V.K.; Singh, N.K.; Khanna-Chopra, R. Genomic regions and underlying candidate genes associated with coleoptile length under deep sowing conditions in a wheat RIL population. J. Plant Biochem. Biotechnol. 2015, 24, 324–330. [Google Scholar] [CrossRef]
  273. Navakode, S.; Weidner, A.; Lohwasser, U.; Röder, M.S.; Börner, A. Molecular mapping of quantitative trait loci (QTLs) controlling aluminium tolerance in bread wheat. Euphytica 2009, 166, 283–290. [Google Scholar] [CrossRef]
  274. Petrarulo, M.; Marone, D.; Ferragonio, P.; Cattivelli, L.; Rubiales, D.; De Vita, P.; Mastrangelo, A.M. Genetic analysis of root morphological traits in wheat. Mol. Genet. Genom. 2015, 290, 785–806. [Google Scholar] [CrossRef]
  275. Ahmad, I.; Ali, N.; Ahmad, H. Association Mapping of Root Traits for Drought Tolerance in Bread Wheat; IntechOpen: London, UK, 2017. [Google Scholar]
  276. Rufo, R.; Salvi, S.; Royo, C.; Soriano, J.M. Exploring the genetic architecture of root-related traits in mediterranean bread wheat landraces by genome-wide association analysis. Agronomy 2020, 10, 613. [Google Scholar] [CrossRef]
  277. Abdolshahi, R.; Safarian, A.; Nazari, M.; Pourseyedi, S.; Mohamadi-Nejad, G. Screening drought-tolerant genotypes in bread wheat (Triticum aestivum L.) using different multivariate methods. Arch. Agron. Soil Sci. 2013, 59, 685–704. [Google Scholar] [CrossRef]
  278. Abdipour, M.; Ebrahimi, M.; Izadi-Darbandi, A.; Mastrangelo, A.M.; Najafian, G.; Arshad, Y.; Mirniyam, G. Association between grain size and shape and quality traits, and path analysis of thousand grain weight in Iranian bread wheat landraces from different geographic regions. Not. Bot. Horti Agrobot. Cluj-Napoca 2016, 44, 228–236. [Google Scholar] [CrossRef] [Green Version]
  279. Lopes, M.S.; Reynolds, M.P.; Manès, Y.; Singh, R.P.; Crossa, J.; Braun, H.J. Genetic yield gains and changes in associated traits of CIMMYT spring bread wheat in a “historic” set representing 30 years of breeding. Crop Sci. 2012, 52, 1123–1131. [Google Scholar] [CrossRef]
  280. Manès, Y.; Gomez, H.; Puhl, L.; Reynolds, M.; Braun, H.; Trethowan, R. Genetic yield gains of CIMMYT international semi-arid wheat trials from 1994 to 2010. Crop Sci. 2012, 52, 1543–1552. [Google Scholar] [CrossRef]
  281. Gerard, G.S.; Crespo-Herrera, L.A.; Crossa, J.; Mondal, S.; Velu, G.; Juliana, P.; Huerta-Espino, J.; Vargas, M.; Rhandawa, M.S.; Bhavani, S.; et al. Grain yield genetic gains and changes in physiological related traits for CIMMYT’s High Rainfall Wheat Screening Nursery tested across international environments. Field Crops Res. 2020, 249, 107742. [Google Scholar] [CrossRef] [PubMed]
  282. Crespo-Herrera, L.A.; Crossa, J.; Huerta-Espino, J.; Autrique, E.; Mondal, S.; Velu, G.; Vargas, M.; Braun, H.J.; Singh, R.P. Genetic yield gains in CIMMYT’s international elite spring wheat yield trials by modeling the genotype× environment interaction. Crop Sci. 2017, 57, 789–801. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  283. Crespo-Herrera, L.A.; Crossa, J.; Huerta-Espino, J.; Vargas, M.; Mondal, S.; Velu, G.; Payne, T.S.; Braun, H.; Singh, R.P. Genetic gains for grain yield in CIMMYT’s semi-arid wheat yield trials grown in suboptimal environments. Crop Sci. 2018, 58, 1890–1898. [Google Scholar] [CrossRef] [PubMed]
  284. Trethowan, R.M.; van Ginkel, M.; Rajaram, S. Progress in breeding wheat for yield and adaptation in global drought affected environments. Crop Sci. 2002, 42, 1441–1446. [Google Scholar] [CrossRef]
  285. Lantican, M.; Pingali, P.; Rajaram, S. Growth in Wheat Yield Potential in Marginal Environments. In Proceedings of the Kronstad Symposium, Ciudad Obregón, Mexico, 15–17 March 2001; pp. 73–80. [Google Scholar]
  286. Mohammadi, R. Breeding for increased drought tolerance in wheat: A review. Crop Pasture Sci. 2018, 69, 223–241. [Google Scholar] [CrossRef]
  287. Bonnin, I.; Bonneuil, C.; Goffaux, R.; Montalent, P.; Goldringer, I. Explaining the decrease in the genetic diversity of wheat in France over the 20th century. Agric. Ecosyst. Environ. 2014, 195, 183–192. [Google Scholar] [CrossRef]
  288. Van de Wouw, M.; Kik, C.; van Hintum, T.; van Treuren, R.; Visser, B. Genetic erosion in crops: Concept, research results and challenges. Plant Genet. Resour. 2010, 8, 1–15. [Google Scholar] [CrossRef] [Green Version]
Table 1. Impact of drought stress on agro-physiological traits in wheat.
Table 1. Impact of drought stress on agro-physiological traits in wheat.
Agronomic TraitReduction (%)Location/CountryReferences
Plant height
Number of tillers per plant
Grains per spike
Grain yield
1000-grain weight
34.45%
25.43%
38.10%
62.75%
19.42%
Pakistan[6]
Biomass
Grain yield
27.05%
25%
China[29]
Grain numbers per spike
Individual grain weight [mg]
Leaf photosynthetic rate
Chlorophyll content
Spikelet fertility
48%
35%
32%
19%
29%
Kansas State University, USA[32]
Plant height
Days to 50% heading
Number of effective tillers
Spike length
1000-grain weight
Grain yield
Biomass
Harvest index
14.7%
4.78
36.3%
23.7%
16.4%
43.2%
32.9%
12.7%
Egypt[5]
Number of grains per spike50%South Africa[30]
Root bimass23%South Africa[34]
Grain yield40%South Africa[28]
Above-ground biomass45%Colorado State University, USA[33]
Leaf water content (LWC) in cultivars Seri M82 and
Weebil4, respectively
64.9%
73.8%
Philippines[31]
LWC in cultivars Kukri and
Excalibur, respectively
72.6–54.4%
74.5–50.5%
Australia[17]
Table 2. Agro-morphological traits responsive to drought stress in plants.
Table 2. Agro-morphological traits responsive to drought stress in plants.
TraitFunction in PlantsReferences
Early growthReduces moisture evaporation from the soil surface and increases soil water available for transpiration and growth[64]
Root system architecture
(RSA)
Plays a vital role in the growth, development and overall productivity of the plants[65]
Long and thick stem
internodes
Plays a vital role in the storage of carbon products[64]
Long coleoptilesFavoured by deep sowing, functions to avoid extreme hot temperatures from the soil surface, and avoid soil drying. Covers the emerging shoot or first leaf during germination[64,66]
Tiller numbersDetermines the development of reproductive organs
(spike, spikelets, and florets)
[4]
Heading and anthesisImprove the translocation of assimilates [67,68]
Longer grain filling Associated with drought tolerance[69]
Spike photosynthetic capacityContributes to remobilization during grain filling[64]
Reduced plant heightAssociated with resistance to lodging, reduce the moisture demand and prevent moisture loss due to transpiration[70]
Stomatal conductance Increased water intake[71]
Presence of awnsContribute to photosynthesis and efficient water use. Influences spike length, increases grain size and grain yield under drought stress[72,73]
Cell membrane stabilityEnables continuous leaf functioning at high temperature[74]
Delayed leaf senescenceInfluences grain yield[75]
Canopy temperatureEnables plants to extract moisture from deeper soil profiles [76]
Leaf rollingHelps plants acclimate to moisture deficit[77]
Chlorophyll contentSpecifies a plant’s photosynthetic capacity and accelerates plant productivity, and plant physiological and phenological status [78]
Large grain sizeEmergence, early groundcover, initial biomass[64]
Table 5. Putative genes associated with target agro-physiological traits conditioning drought tolerance.
Table 5. Putative genes associated with target agro-physiological traits conditioning drought tolerance.
Gene CropFunction References
Dreb1WheatDrought tolerance[237]
TaERF3WheatDrought and salt tolerance
Drought avoidance
[238]
VRN1Wheat and barleyRegulates flowering behaviour
Balance between shoot and root architecture
Modulate plant morphology
[114,115]
TaCRT1WheatCalreticulin Ca2+ binding protein[239]
Ppd-D1 and
Ppd-D1a
WheatFlowering time determinant gene, involved drought tolerance
Photoperiod insensitive allele
[6]
TaZFP22, TaZFP34, and TaZFP46WheatRoot expressed and drought induced Q-type C2H2 zinc finger transcriptional repressors in wheat[240]
TaER1 and TaER2WheatHigh transpiration efficiency and grain yield[241]
TaNAC69WheatPhotoperiod insensitivity [200]
TaWRKY 1WheatConfer drought tolerance[242]
TaWRKY44WheatConfer tolerance to multiple abiotic stresses such as drought, salt, and osmotic stress [243]
TaSnRK2Wheat Encodes sucrose non-fermenting 1-related protein kinase and adapt to various environmental conditions with significant correlation to spike length and thousand kernel weight[244]
TaH2B-7DWheat Confer drought tolerance, increases relative electrolyte leakage rate and malonaldehyde (MDA) content[245]
TaWRKY8WheatGrain yield and abiotic stress tolerance[53]
TaMYB3R1WheatSalt, vernalisation and drought tolerance in wheat[246]
TaAQP7WheatDrought tolerance in Arabidopsis[247]
SNAC1WheatSalt and water stress tolerance in transgenic wheat[248]
GmbZIP1Wheat Drought tolerance[110]
TaSST-D1 and TaSST-A1 WheatWater-soluble carbohydrates,
Increased thousand-kernel weight (TKW), plant height and drought adaptability
[249]
TaMYBsm3WheatDrought adaptation[250]
TaEXPA2WheatConfer drought tolerance[251]
TaSNAC8-6AWheatDrought adaptation[252]
Table 6. Putative QTL regions for drought-related traits in wheat mapping populations under either individual or drought, heat and non-stressed conditions.
Table 6. Putative QTL regions for drought-related traits in wheat mapping populations under either individual or drought, heat and non-stressed conditions.
ChromosomesAssociated Roots and Related Trait[s]Study ApproachCollection or Population TypeReferences
1D, 2A, 2B, 2D, 3A, 4A, 4B, 5A, 5B, 5D, 6D, 7A, 7DAverage root diameter, number of root crossing, number of root forks, number of root tips, root volume, surface root areasQTL mappingAdvanced backcross population[15]
1A, 2B, 3A, 3D, 4B, 4D, 6A, 6B, 6D, 7BMaximum root length, primary root length, lateral root length, root tip number, total root lengthQTL mappingRIL population[16]
2A, 2B, 3A, 4B, 4D, 5A, 6A, 6D, 7BTotal root length, total root surface area, total root volume, number of root tips, main root lengthQTL mappingDH population[269]
1B, 2B, 3B, 4A, 4D, 5A, 5B, 7A RIL population
3B, 4DRoot re-growth, root tolerance index, aluminium toleranceQTL mappingD-genome substitution lines[273]
5AStay greenQTL mappingDH population[19]
1A, 1B, 2A, 3A, 6A, 6BRoot length, root volume, root surface area, number of tipsQTL mappingRIL population[274]
2B, 2D, 3B, 3D, 4B, 4A, 6D, 7B, 7DMaximum root length, root fresh weight, ratio of root water loss, total root length, total root surface area, total root volume, number of root tips, number of root forksQTL mappingRIL population[269]
1A, 1B, 2B, 2D, 3A, 3B, 5A, 5B, 6A, 7AMaximum root length, seminal root number, total root length, project root length, root surface area, seminal root angle, grain yieldQTL mappingDH population[268]
1B, 2A, 2B, 2D, 3D, 4A, 4B, 5A, 5B, 6A, 6B, 6D, 7ARoot depth at booting and mid-grain fill stage, root dry weight at booting and mid-grain fill stageGWASCore collection[266]
2A, 2B, 2D, 5BTotal root length, root fresh weight, maximum root length, nodal roots, root density, root diameterMTACore collection[275]
1B, 2A, 2B, 3B, 5A, 5B, 6A, 7ATotal root number, root dry weight, seminal root angle, seed weight, seed lengthGWASLandraces[276]
DH: Doubled haploid; MTA: Marker–trait association; QTL: Quantitative trait loci; RIL: Recombinant inbred line.
Table 7. Rates of yield gains realised from international breeding programs using CYMMIT lines under marginal environments.
Table 7. Rates of yield gains realised from international breeding programs using CYMMIT lines under marginal environments.
YearsRate of Yield IncreaseTarget EnvironmentReferences
2007–20160.93% (40 kg ha−1 yr−1)Drought prone[281]
2006–20142.7% (88 kg ha−1 yr−1)Drought prone[282]
2002–20141.8% (0.15–3.5 t ha−1 yr−1)Drought prone[283]
1965–201417.7 to 25.6 kg−1Drought prone[139]
1994–20100.7% (2.07–2.7 t ha−1 yr−1)Drought prone [280]
1991–19970.09% (2.1 kg ha−1 yr−1)Low yielding environments [284]
1979–19993.48% (2.3 to 3.5 t ha−1 yr−1)Drought prone[285]
1979–19980.19% (5.3 kg ha−1 yr−1)Drought prone [284]
1979–19952.75% (70.5 kg ha−1 yr−1)Drought prone[285]
1977–20080.5% (251 to 291 g m−1 yr−1)Drought prone[279]
1964–19781.54% (2.3 to 4.3 t ha−1 yr−1)Drought prone[285]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bapela, T.; Shimelis, H.; Tsilo, T.J.; Mathew, I. Genetic Improvement of Wheat for Drought Tolerance: Progress, Challenges and Opportunities. Plants 2022, 11, 1331. https://doi.org/10.3390/plants11101331

AMA Style

Bapela T, Shimelis H, Tsilo TJ, Mathew I. Genetic Improvement of Wheat for Drought Tolerance: Progress, Challenges and Opportunities. Plants. 2022; 11(10):1331. https://doi.org/10.3390/plants11101331

Chicago/Turabian Style

Bapela, Theresa, Hussein Shimelis, Toi John Tsilo, and Isack Mathew. 2022. "Genetic Improvement of Wheat for Drought Tolerance: Progress, Challenges and Opportunities" Plants 11, no. 10: 1331. https://doi.org/10.3390/plants11101331

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop