Skip to main content

REVIEW article

Front. Genet., 28 October 2022
Sec. Genomics of Plants and the Phytoecosystem
This article is part of the Research Topic Genomics-Based Breeding of Crops for Food and Nutritional Security in 21st Century - Volume 2 View all 12 articles

Improving drought tolerance in maize: Tools and techniques

  • 1Department of Agronomy, Iowa State University, Ames, IA, United States
  • 2School of Agricultural, Earth and Environmental Sciences, University of KwaZulu-Natal, Pietermaritzburg, South Africa

Drought is an important constraint to agricultural productivity worldwide and is expected to worsen with climate change. To assist farmers, especially in sub-Saharan Africa (SSA), to adapt to climate change, continuous generation of stress-tolerant and farmer-preferred crop varieties, and their adoption by farmers, is critical to curb food insecurity. Maize is the most widely grown staple crop in SSA and plays a significant role in food security. The aim of this review is to present an overview of a broad range of tools and techniques used to improve drought tolerance in maize. We also present a summary of progress in breeding for maize drought tolerance, while incorporating research findings from disciplines such as physiology, molecular biology, and systems modeling. The review is expected to complement existing knowledge about breeding maize for climate resilience. Collaborative maize drought tolerance breeding projects in SSA emphasize the value of public-private partnerships in increasing access to genomic techniques and useful transgenes. To sustain the impact of maize drought tolerance projects in SSA, there must be complementary efforts to train the next generation of plant breeders and crop scientists.

Introduction

Maize is the primary staple food for more than 900 million people globally and the third most important source of calories after rice and wheat (Shiferaw et al., 2011; Adebayo and Menkir, 2015). With a decrease in rice production in China and India, and an increase in the demand for dairy and meat, global demand for maize is projected to double by 2050 (Rosegrant et al., 2009). Maize is also the most widely grown staple crop in SSA and serves an important role in food security, but it is also highly susceptible to drought, with 15%–20% of its yield lost to drought each year (FAOSTAT, 2010; Bankole et al., 2017; Lunduka et al., 2017; FAO, 2021). Because of maize yield loss to drought, from 2005 to 2915, developing countries experienced a revenue loss of up to USD 29 billion (FAOSTAT, 2010; Liu and Qin, 2021). Changes in rainfall magnitude, distribution, and timing, as well as an increase in temperature, all interact to destabilize maize production further (IPCC, 2012; Meseka et al., 2018). Data from more than 20,000 trials in Africa revealed a 1% reduction in maize yield for each “degree day” above 30°C (Lobell et al., 2011; Bankole et al., 2017; Lunduka et al., 2017). When other factors are not limiting, the combination of heat and drought stress causes a 1.7% yield reduction for each excess degree day (Lobell et al., 2011).

An assessment of the impact of physical drought on maize revealed an alarming vulnerability in SSA (Kamali et al., 2018). On average, 5 to 10 drought events were experienced between 1970 and 2004 in most parts of SSA (Fisher et al., 2015). This makes drought a key constraint to maize production in the region (Heisey and Edmeades, 1999). Further, climate change is expected to exacerbate the impact of drought in the region and cause a reduction in maize production by almost 22% in 2050 (Schlenker and Lobell, 2010; Challinor et al., 2016; Tesfaye et al., 2018; Barbosa et al., 2021). Therefore, meeting the increased demand while stabilizing production requires a strategy that includes the genetic improvement of maize for drought tolerance (Messina et al., 2021). A recent review by Sheoran et al. (2022) discussed new breeding technologies and approaches to improve maize drought tolerance. This review presents an overview of a broad range of tools and techniques, the integration of plant physiology, molecular biology, and systems modeling, and examples of private-public partnerships in developing drought tolerant maize for Africa. The review also discusses the importance of plant breeding education to address the shortage of plant breeders in SSA.

Breeding for drought tolerance in maize

Drought is a key abiotic stress that causes low-income countries to lose billions of dollars (FAO, 2021). Levitt (1972) provided functional definitions for water deficit and drought stress which can be used to develop breeding targets. Water deficit occurs when plant transpiration cannot fully meet the atmospheric demand due to a lack of water in the environment. This deficit causes damage and induces a stress response proportional to the rate of deficit (Blum, 2014).

The impact of drought stress depends on the interaction of plant, environmental, and management factors: the crop development stage, rate of water deficit development, peak intensity of the deficit, and planting density. Regarding the development stage, maize is most sensitive to drought stress during flowering (Bolanos and Edmeades, 1993). Severe water deficits during the period of a few days before silking to roughly 25 days can eliminate yield entirely (Claassen and Shaw, 1970). Drought stress causes a delay in ear growth and silking, increasing the anthesis-silking interval (ASI) to the point where it inhibits fertilization. The result is a barren ear or one with few kernels (Sah et al., 2020). Even with successful pollination, kernel abortion beginning as early as 2–3 days after pollination can reduce kernel number (Westgate and Boyer, 1986). Drought stress at the start of grain filling can also significantly lower or eliminate yield (Barker et al., 2005). Since the crop reaches full size prior to flowering, water use is at a maximum. Drought stress induces premature leaf senescence and reduces ear growth, with severe stress causing complete desiccation. Consequently, kernel weight is significantly reduced due to lowered photosynthate accumulation.

According to Ribaut et al. (2009), maize responds to and mitigates the impact of water deficit using three primary strategies: drought escape, drought avoidance, and drought tolerance. Drought escape is a strategy to prevent the coincidence of water deficit with key developmental stages and is primarily achieved by early flowering and maturation. Drought avoidance, on the other hand, is the capacity to avoid or reduce plant water deficit but maintaining turgor through an increase in water uptake (using a deeper and/or a larger root system for example) and/or a reduction in water use (for instance, decreased stomatal conductance). Drought tolerance is the ability to maintain plant function during water deficit, which can be achieved by alleviating oxidative stress, for instance.

The goal of a drought tolerance breeding program is two-fold: 1) reduce the gap between yields in optimal and stress conditions without sacrificing yield potential, and 2) improve yield stability for a range of stress conditions. According to Lunduka et al. (2017), a drought tolerant maize variety is one that gives at least a yield of 30% of its potential under water-stress, especially during the flowering and grain-filling stages. In addition, Messina et al. (2021) found that although root systems architecture and yield have changed because of breeding for maize drought tolerance, the uptake of water has not changed.

While drought escape does create an advantage under drought conditions, using earlier maturing maize varieties which generally yield less than full season varieties (White et al., 1922) in non-drought years creates a lower yield potential (Ke and Ma, 2021). To improve both yield potential and stability, the use of hybrids with a maturity suited to the wettest part of the year in the target environment might be an option. Together, the capacity of a plant to avoid or reduce water deficit (dehydration avoidance), sustain function under water deficit (dehydration tolerance), improve crop productivity and provide avenues for plant improvement.

Even without drought tolerance as an explicit breeding goal, selecting for high yield potential under well-watered conditions has consistently led to increased yield in both water deficit and non-deficient environments (Castleberry et al., 1984). Selecting for high yield potential extends tolerance to other abiotic stresses as well, such as heat, cold, and low soil fertility. The improvement in drought tolerance and the rate of genetic gain has been evaluated through experiments testing historic cultivars under drought conditions applied at various growth stages, demonstrating a gain of 124 kg ha−1 yr−1 for flowering stress and 91 kg ha−1 yr−1 for mid grain fill stress (Barker et al., 2005). Nevertheless, the ability of newer hybrids to tolerate drought stress is primarily due to the adaptation of parent lines to higher planting densities (Tollenaar and Lee, 2006).

Although selecting for high yield potential in favorable environments has led to increases in drought tolerance, the correlation of yield for hybrids grown under well-watered and water-stressed conditions is reduced with the degree of stress. In addition, the degree of genotype by environment interaction necessitates screening materials under both well-watered and water stressed conditions. Therefore, because drought is unpredictable in the target environment, sites in rain-free environments, which use irrigation and planting date to control the timing and severity of drought, are needed (Campos et al., 2004).

Choosing testing environments

The rate of achievable genetic gain is highly dependent on the choice of testing environments, especially how closely the selection environments mirror the target environments (Cooper et al., 2006). Since the distribution and amount of rainfall, temperature, soil water holding capacity, and the developmental stage of the crop all interact to create distinct drought scenarios, an important preliminary step in developing a breeding program for drought tolerance is capturing this information to identify the target population of environments (TPE) for which the crop will be adapted (Cooper et al., 1997).

Environmental characterization and TPE delineation are crucial for several reasons. Information on the TPE can be used to explain and predict a considerable portion of genotype by environment (G x E) interactions. This is necessary since the effect of a particular allele can be different depending on environmental conditions (Chenu et al., 2009). In one drought scenario a trait can be advantageous, while in others it can be detrimental. For instance, breeding for water use efficiency can improve yield in very dry environments, while reducing the potential yield in mild drought conditions (Tardieu, 2012).

Characterizing the TPE is also crucial in identifying appropriate managed stress environments for multi-environment trials (METs) since the expected performance gain is dependent on the similarity between environments represented in METs and the TPE (Ribaut et al., 2009). Furthermore, by weighting phenotype data from METs based on how representative individual trials of the TPE are, selection gain over generations can be improved. This improvement in gain was demonstrated by Podlich and Cooper (1998) using a genetic simulation model. Comparing the drought scenarios of the TPE with other drought-prone regions facilitates methodological and adapted germplasm exchange to other parts of the world (Chenu et al., 2011). Increases in temperature and changes in weather patterns brought about by climate change highlight the need to characterize the target environment in advance, since the environmental conditions could be significantly different by the time a variety is developed and disseminated, resulting in lower-than-expected yields (Challinor et al., 2016). These seasonal variability and resource constraints often lead to multi-environment breeding trials offering biased representation of the TPE. One way to avoid this is use of weighted analysis based on representative trials which can help breeders select for germplasm better adapted to the TPE. Managed-environment trials are another way to evaluate performance in representative environments or for stresses, allowing detailed assessment of germplasm, traits, or genes of interest (Chenu, 2015).

As mentioned above, crop simulation models have been used to characterize TPEs, and to evaluate how well MET locations fit the TPE. Studies using these models help to improve breeding efficiency for multiple crops in regions around the world. For instance, the “Cerrado” environments of central Brazil have been characterized for rice and maize production (Heinemann et al., 2008). More recently, maize growing environments in Eastern and Southern Africa were characterized using the APSIM model, leading to the distinction of four environmental types to which breeding objectives can be catered (Seyoum et al., 2017).

In addition to similarity to the target environment, site homogeneity and the ability to manage water inputs determine the success of testing (Blum, 2011b). Differences in soil texture, effective rooting depth, micronutrient concentration, salinity, pH, and the presence of pathogens all increase residual variability, thus minimizing the precision in estimating genotypic means. Regarding the ability to manage water inputs, use of desert and off-season environments, and rainout shelters provide options to prevent the effect of precipitation (Blum, 2011b). One issue with the use of desert environments is the effect of temperature extremes. Estimating drought tolerance is confounded by the occurrence of extreme temperatures during ear growth (Otegui and Andrade, 2000). Recent improvements in modeling techniques have increased their effectiveness. Liu et al. (2021) concluded that DSSAT CERES-Maize can adequately simulate regional maize yields using the CERES-Maize module calibrated to regional soil and daily weather databases. Adnan et al. (2020) successfully use the CERES-Maize model to generate data for GEI and stability studies of maize genotype in the absence of observed field data, and Ramirez-Villegas et al. (2020) pointed out the varying important roles of crop modeling in breeding efforts, including assessing genotypic adaptability and stability, characterizing and identifying target breeding environments, identifying tradeoffs among traits for such environments, and making predictions of the likely breeding value of the genotypes. Recognizing the successes from simulation modeling, Hajjapoor et al. (2022) pointed out the problems that still exist in identifying MET environments that fit TPEs, to deal with Genotype-by-Environment-by-Management (G x E x M) interactions and proposed a simple step-by-step approach to bring the capacity of process-based models to better define target population of environment, within which the clustering of subunits will allow for the reduction of G x E × M interactions.

Incorporating secondary traits

Although maximizing grain yield is the primary objective of breeding for improved drought tolerance, reduced genotypic variance and high G × E interaction contribute to reduced heritability of yield when testing under drought conditions. Incorporating secondary traits into the selection process, that is, traits which more directly reflect the physiological effect of drought stress, can increase selection efficiency, and improve gain (Ribaut et al., 2009). Edmeades et al. (1996a) developed criteria for ideal secondary traits for drought screening. In addition to high heritability, high genetic variability, genetic correlation with yield, and no association with yield loss under non-limiting conditions, ideal traits should be simple, cheap, non-destructive, and fast to assay. Using these criteria, only ASI, ears per plant (EPP), barrenness, kernels per ear, and stay green have been found to be suitable secondary traits. Of those traits, ASI and EPP have been identified as the best performing traits, with ASI being the most widely used. While tassel growth is not as affected by drought stress, silk emergence and thus ASI can be used as an indicator for ear and plant growth rates during flowering. When testing under conditions that reduce yield by more than 50%, the incorporation of these secondary traits into a selection index has led to a selection efficiency on a par with testing under optimal conditions (Chapman et al., 1997). Using yield alone for selection resulted in significantly lower gains but the use of secondary traits for selection has resulted in improved genetic gains in other instances as well (Campos et al., 2004).

Application of genomic mapping tools

While the use of secondary traits provides more heritable selection targets, the advancing field of genomics has been used to further improve selection efficiency by identifying and selecting the genomic regions responsible for improved secondary traits and tolerant phenotypes. By developing a better understanding of the genetic and physiological basis of drought tolerance, and using this information during selection, the value of genomics in improving drought tolerance has been demonstrated in maize (Tuberosa et al., 2007; Tsonev et al., 2009), in addition to several other crops (Tuberosa and Salvi, 2006; Mir et al., 2012). With marker-assisted selection, the initial step is to identify markers and genes associated with drought tolerance. These associations have been identified using a variety of approaches. In addition, the use of parental haplotype sharing can help increase the power, precision, and accuracy in Quantitative Trait Loci (QTL) mapping (Jansen, Jannink, and Beavis, 2003).

Most understanding of the genetic basis of drought tolerance was gained from traditional QTL mapping, using a relatively limited number of markers in bi-parental populations (Cattivelli et al., 2008). While a large number of associations have been found, there are limitations with traditional QTL mapping: 1) the resulting QTL are comparatively large due to limited genetic resolution; 2) extra time is required to develop a mapping population; 3) by using a mapping population, only a small proportion of the total allelic diversity expected in the germplasm pool is sampled; and 4) QTL for the same trait can segregate differently in other mapping populations.

Using linkage mapping, numerous QTL relating to morphological traits such as flowering and tassel size, as well as physiological parameters such as ABA and carbohydrate metabolism have been identified. Several of these QTL studies have been summarized by Ribaut (2006). To identify genome regions and candidate genes consistent across populations, which convey drought tolerance, QTL from various studies have been analyzed and compiled into consensus maps (Tuberosa et al., 2002; Sawkins et al., 2004; Hao et al., 2009; Semagn et al., 2013; Zhao et al., 2018).

Linkage disequilibrium (LD)-based association mapping has been used to overcome some of the constraints of linkage mapping. The advantages of LD association mapping are summarized by Mir et al. (2012) and include: less time and resources are required since a natural germplasm collection can be used, and it provides a higher mapping resolution. LD studies also have the advantage of being able to simultaneously evaluate the varying effects of multiple alleles in multiple backgrounds at one time (Buckler and Thornsberry, 2002). Since diverse germplasm is used for association mapping, identified markers are more likely to convey drought tolerance in multiple backgrounds which is particularly valuable to breeders.

Although linkage and LD-mapping both have unique advantages, the approaches are complementary (Myles et al., 2009) and have been combined to better identify QTL associated with drought tolerance. Using the combined techniques, SNP markers were identified, which better explained the phenotypic variation regarding ASI compared to either technique alone (Lu et al., 2010). However, association mapping is not without limitations. One drawback is the difficulty in detecting associations with traits underpinned by many rare variants with a large effect size, or by many common variants that have a small effect (Korte and Farlow, 2013). Since the effect size of the allelic variants as well as their frequency in the sampling population determine the phenotypic variance, rare variants, and a combination of alleles with small effect sizes are difficult to associate (Korte and Farlow, 2013).

As sequencing technologies have advanced and costs have dropped, genome-wide association studies (GWAS) have become a common approach to uncover the genetic basis of drought tolerance (Yamada and Dwiyanti, 2013). GWAS has been successful in identifying genome regions associated with drought tolerance. Using 60,000 SNPs on a hybrid testcross background allowed dominant alleles to be detected. Farfan et al. (2015) identified 10 quantitative trait variants (QTVs) for flowering time, plant and ear height, and yield. Under both well-watered and stressed conditions, three of these QTVs explained 5–10% of yield variation. Many of the QTVs also co-located with QTL from other studies, which confirmed their association with drought tolerance. Rather than using a high-density coverage of the maize genome, other association studies have taken a more targeted approach and instead have used a comparatively limited number of SNPs (1536) selected from candidate genes associated with drought response (Setter et al., 2011).

While Setter et al. (2011) identified several significant loci and candidate genes, it is possible that many were missed due to a lack of genome coverage. To enlarge the panel of markers for the collection without additional genotyping, Zhang et al. (2016) used imputation based on 556,809 SNPs from 368 diverse inbred lines that were previously genotyped using RNA sequencing. This method resulted in the identification of 26 new loci associated with metabolic and physiological traits in leaf tissue, and only one of the six loci significantly associated with drought-related metabolites from the Setter study (Setter et al., 2011) was still significant. Positional cloning of QTL in conjunction with association mapping helped to identify genes and noncoding sequences associated with flowering time, a trait strongly associated with drought response (Salvi et al., 2007).

Nested Association Mapping (NAM), which relies on a specifically designed population, also combines the advantages of linkage and association mapping while minimizing the disadvantages of each (Yu et al., 2008). By crossing 25 inbred lines to the B73 inbred and selfing the F2 populations to the F6 generation, 200 recombinant inbred lines (RILs) were created for each of the 25 populations. These RILs were genotyped using the same 1106 markers to identify recombination blocks and the parents were sequenced, resulting in 5000 RILs (known as the US-NAM population) that were genotyped at high density and could be compared and analyzed together (Yu et al., 2008). Researchers from the Institute of Crop Science at the Chinese Academy of Agricultural Sciences used both the US-NAM and China NAM populations to identify 52 candidate genes with differential expression under two water treatments, allowing them to make genomic predictions of drought-related traits with a mean accuracy of 0.57 (Li et al., 2016).

Candidate genes have also been identified by incorporating functional genomics technologies, such as transcriptome and metabolome analysis. These approaches help reveal how biochemical, physiological, and regulatory networks change in response to stress, and expose differences in drought response among tissues and genotypes (Mir et al., 2012). By studying gene expression in pre-fertilization ears using both cDNA microarray and genome-survey technology, Zinselmeier et al. (2002) demonstrated differential gene expression among tissues, and identified several genes not previously associated with drought stress response. Microarray technology also facilitated the identification of 22 differentially expressed genes which co-located on the genetic map with QTL previously associated with drought tolerance (Marino et al., 2008).

Like gene expression analysis, examining changes in the plant metabolome can connect the agronomic phenotype with the underlying genetics, and can be useful in identifying genes that are not as affected by environmental factors (Riedelsheimer et al., 2012). A distinct advantage of profiling the metabolome (and proteome) over the use of transcriptomics is that effects from posttranscriptional and posttranslational regulation can be accounted for (Tuberosa et al., 2007). By examining metabolic changes in maize leaves and ears due to water deficit and dissecting the genetic basis of those traits using GWAS, Zhang et al. (2016) identified 23 metabolite-associated loci and validated 10 as responsive to drought stress. Using the same technique on leaves alone, Riedelsheimer et al. (2012) identified 26 SNPs strongly associated with changes in distinct metabolites, which explained up to 32% of the observed genetic variance.

More exciting research results have been achieved in recent years on the genetics of drought tolerance/resistance. The identification of naturally occurring loci or genes associated with drought tolerance can serve as direct targets for both engineering and selecting improved maize for drought regions. Increased expression of the NAC gene (ZmNAC111) with the MITE (significantly associated with natural variation in maize drought tolerance) insertion in its promoter enhanced drought tolerance in maize seedlings (Mao et al., 2015). Wang et al. (2016) reported that transgenic maize with enhanced expression ZmVPP1 (also with natural variation) exhibited improved seedling drought tolerance. In a study of a NAC-encoding gene of ZmNAC080308, a functional marker developed for use in predicting drought stress tolerance in a US maize inbred line panel showed that lines carrying Hap2 produced greater than 10% grain yield than those carrying Hap1 under drought stress condition (Wang et al., 2021). Wang et al. (2022) identified and reported the overexpression of a transcription factor, ZmERF21, is tightly associated with drought tolerance in maize seedlings, expressed by significantly increased chlorophyll content and activities of antioxidant enzymes under drought conditions. Another GWAS study (Khan et al., 2022) identified candidate genes and their key variations that will contribute to an understanding of the genetic basis of drought tolerance, especially for the female inflorescence, and will be important in facilitating drought-tolerant maize breeding.

Bridging the gene-to-phenotype gap

Advances in molecular biology have provided tools and strategies to associate genomic regions with traits that convey improved yield in a drought scenario. With genomic selection, these associations can be used for phenotypic prediction and breeding decisions. However, predicting the effects of genes across scales of biological organization is made difficult by the complex interaction of genes and environmental factors (Hammer et al., 2006). Much of the time the effect of a particular allele on a complex trait such as drought tolerance is confounded by gene interactions such as epistasis and pleiotropy (Hammer et al., 2006). In addition to gene-gene interactions, the drought scenario strongly influences whether an allele will convey a significant advantage. In many cases, the effect of a trait on yield is a trade-off which depends on environmental conditions (Tardieu and Tuberosa, 2010). For instance, reduced transpiration and biomass accumulation protect against drought stress, yet reduce yield potential (Blum, 2009).

Modeling can be incorporated to address the challenges presented by gene-gene and gene-environment interactions in the physiological and genetic dissection of traits which convey drought tolerance, and ultimately in developing improved cultivars (Tardieu and Tuberosa, 2010). In addition to improving efficiency by dissecting complex traits into more measurable targets, which helps in the development of phenotyping strategies, using crop growth and development models to evaluate traits and predict phenotypes in the TPE is useful for assessing breeding strategies and allocating resources (Hammer et al., 2006; Messina et al., 2011). With the addition of environment modeling, the range of drought scenarios that comprise a TPE over time can be accounted for, helping to efficiently identify associations between traits and the set of environmental conditions which maximize yield.

Not only has modeling been used to enhance the physiological and genetic dissection of drought tolerance, it also has been used to augment a maize breeding program. By extending the concept of fitness landscapes to the characterization of yield-trait performance landscapes and extending the E (NK) model of trait genetic architecture to incorporate biophysical, physiological, and statistical components, Messina et al. (2011) developed a graphical representation of the associated yield-trait performance landscape that could be used in selection decisions. Doubled haploid (DH) lines were selected based on their relative position in the performance landscape, their predicted performance, and their potential to contribute to further yield improvement. This approach showed that an understanding of yield-trait performance landscapes can be used to improve genomic selection and phenotyping strategies.

Studies which use modeling have improved the predictive power of the combined effects of major QTL. In one study, a model based on the combined effects of the major QTLs was able to predict 74% of the variability for maize leaf elongation rate under a range of temperature and water deficit conditions (Reymond et al., 2003). Modeling has also been combined with phenotypic analysis to identify QTL for drought traits and to characterize genotypes (Tardieu, 2006). The development of response curves, which more quantitatively define the relationship between the phenotypic trait and environmental conditions, are unique to each genotype and can be compared to select optimal genotypes for a particular range of conditions. A similar approach has been used to identify common QTL for leaf growth and ASI under drought conditions, suggesting that the genetic determination of leaf growth and silk elongation rate is at least partially shared (Welcker et al., 2006). By extending the APSIM crop model to include genotype-specific parameters, Chenu et al. (2009) simulated the effects of QTL on leaf and silk elongation, and ultimately yield. The study demonstrated the high level of QTL × environment interaction, opening the possibility of exploiting these interactions for drought tolerance (Chenu et al., 2009).

Marker assisted selection in breeding

While many drought tolerance candidate genes and QTL have been documented, few have been validated to produce a clear genetic gain when selected for in diverse germplasm under field conditions (Ribaut et al., 2009). This is in part due to many studies being limited to putative association based on colocalization of candidate genes and QTL along genetic maps. Another reason is that QTL or genes affecting drought tolerance are not distinguished based on how they are expressed under contrasting drought treatments. For breeding, QTL which are constitutively expressed and affect the yield consistently should be given priority, since they show limited interaction with the environment (Vargas et al., 2006). However, relatively simple heritable constitutive plant morphological and developmental traits are often ignored when evaluated by functional genomics, even though they have a considerable effect on performance under drought stress (Blum, 2011a). Other factors which limit the practical use of previous genetic investigations are the germplasm used for the mapping population, and the drought treatment used. Non-commercially viable lines are often used for association studies and identified QTL may have a null or negative effect when introgressed into elite materials (Monneveux and Ribaut, 2006). Regarding the drought treatment, many of the expressed genes identified in association studies are due to imposing rapid stress in a laboratory setting, but the slow progress of drought stress in field conditions results in minimal expression of these stress-responsive genes (Barker et al., 2005).

Although most information gathered from the genetic dissection of drought tolerance has not directly led to the development of improved cultivars, marker assisted selection has been used successfully. By introgressing favorable alleles at five genomic regions, MABC-derived hybrids were selected, which yielded 50% more than control hybrids under severe water stress conditions. Under mild stress and well-watered conditions, the hybrids performed as well as the control (Ribaut and Ragot, 2006). While the experiment did improve the recipient line through the introgression of drought QTL, further improvement is limited after the recurrent parent is fixed for the new QTL (Mir et al., 2012). To address this limitation, MARS has been used to take advantage of desirable alleles in multiple lines. The value of MARS in breeding for improved maize grain yield in SSA was demonstrated by Beyene et al. (2016) who estimated the rate of genetic gain under drought and non-drought environments. Since the value of a set of QTL alleles depends on the germplasm used in breeding, another MAS strategy was proposed by Podlich et al. (2004) to attempt to account for the presence of epistasis and G × E interaction. The “Mapping As You Go” (MAYG) strategy works by re-estimating the value of each QTL allele upon creation of a new set of germplasm, and its effectiveness has been estimated by simulation (Podlich et al., 2004).

Rather than selection based on known markers with significant associations to drought traits, the use of genomic selection (GS) or genome-wide selection (GWS) to develop drought tolerance promises improved gains (Bernardo and Yu, 2007). The application of GS in yield trials of tropical maize lines across multiple locations in SSA produced selection candidates at lower cost than phenotypic selection (Beyene et al., 2019). The cost aspect is important because in situations where doubled haploids (DH) lines are used, the capacity to phenotype testcross materials across multiple sites can limit progress (Beyene et al., 2021).

Genomic Selection described as “test-half-predict half” approach uses random markers to genotype a phenotyped training population. The marker and phenotyping data are used to develop breeding values of alternative alleles, which are fitted as random effects in a linear model. Selection in each recurrent generation is based on the sum of those breeding values, known as the genomic estimated breeding value (GEBV) (Meuwissen et al., 2001). This approach has considerable advantages, such as high selection accuracy when selecting on markers alone, and no prior knowledge of QTL positions is needed (Resende et al., 2014). The efficiency of the GS approach has been compared experimentally to selection based on yield alone along with selection incorporating secondary traits (Ziyomo and Bernardo, 2013). Compared to selection based on yield, secondary trait assisted selection was slightly higher or lower depending on the trait, while GS was significantly more efficient.

Sources of genetic diversity

There are many sources of genetic diversity which potentially hold alleles promoting drought resistance (Barbosa et al., 2021). While conventional germplasm bases contain enormous levels of allelic polymorphism (Guo et al., 2004), the development of mutagenic and transgenic lines create a virtually endless supply of genetic diversity. Most sources of natural variation have remained relatively untapped. Globally, only 5% of the available maize germplasm is used commercially (Brown, 1975), and exotic germplasm constitutes only 1% of the US germplasm base (Goodman, 1983). This section aims to present resources that have been used or with potential in breeding for or dissecting drought tolerance.

An ideal germplasm base 1) segregates for the trait of interest, 2) has a high probability of containing desired alleles, and 3) those alleles should be relatively easy to introgress into cultivars (Blum, 2011b). The existing drought tolerance, genetic diversity, and ease of introgression make the agronomic germplasm pool an ideal first choice. Agronomic germplasm has been used to develop the Pioneer AQUAmax® product line of drought-tolerant maize hybrids, improving upon the drought tolerance of commercial hybrids that already possess a high degree of tolerance (Cooper et al., 2014).

Alleles from landraces have been introgressed into elite varieties, leading to improved drought tolerance. Researchers at the International Institute of Tropical Agriculture (IITA) in Ibadan, Nigeria, crossed six landraces pre-screened for drought tolerance with an elite maize variety (AK9443-DMRSR), resulting in some of the BC1F2 populations expressing improved yield potential under random-drought conditions (Meseka et al., 2013). In dry, marginal growing regions of Kenya, farmers use local landraces rather than hybrids because they are believed to produce better under low or no input use (Sammons, 1987), and possess resistance to biotic and abiotic stress.

While wild germplasm and related species, such as Zea maxicana or Tripsacum floridanum, are excellent sources of novel genes for drought tolerance improvement (Singh, 2010), there are some difficulties with their use. First, the technical difficulty in making a wide cross between an agronomic genotype and the donor. Second, the need to eliminate the introgression of negative traits carried by the donor, and third, the value of genes or alleles conveying drought resistance once introgressed (Blum, 2011b). Since adaptation to drought stress in wild type and related species enhances survival rather than optimizing yield, the genes which impart that resistance may not be beneficial when introgressed into breeding germplasm.

A molecular approach particularly suited to addressing this complication and identifying beneficial alleles from wild sources is advanced backcross QTL (AB-QTL) analysis (Tanksley and Nelson, 1996). After developing backcross families and eliminating lines displaying yield-reducing characteristics in the BC1 and BC2 generations, QTL analysis is performed. This approach has the benefit of identifying valuable QTL and developing superior genotypes simultaneously and has been used to analyze more conventional crosses as well (Ho et al., 2002).

Transgenic and gene-editing tools

The application of mutagenesis techniques to improve drought tolerance in maize is limited (Blum, 2011b; Gao et al., 2014; Ruswandi et al., 2014). Therefore, transgenic approaches are useful in functional analysis of genes affecting stress response and adaptation (Yang et al., 2010), and can act as a bridge to move valuable genes into breeding germplasm (Blum, 2011b). Nelson et al. (2007) transformed plants to overexpress transcription factor ZmNFYB2 resulting in increased drought tolerance and yield. Plants engineered to overexpress the ZmAsr1 gene, a putative transcription factor, resulted in lines with improved water use efficiency and dry weight accumulation (Jeanneau et al., 2002). In another study, Nuccio et al. (2015) engineered plants to overexpress a gene encoding rice trehalose-6-phosphate phosphatase (TPP), which increased sucrose concentration in ear spikelets, kernel set, and harvest index, leading to improved yield in mild and severe drought. Guo et al. (2014) generated transgenic maize overexpressing the ARGOS1 (ZAR1) gene resulting in enhanced maize organ growth, grain yield, and drought-stress tolerance.

While the transgenic approach has helped to identify genes and mechanisms that improve drought tolerance, there are obstacles to their use in released cultivars. Regarding transgenic events tested in laboratory conditions, the effect of the transgene might not yield an advantage in the TPE because of environmental interaction (Bänziger and Araus, 2007). When placed in agronomic germplasm, the expression of a single gene may not significantly alter the final phenotype due to dampening and compensation from other processes (Sinclair and Purcell, 2005). The mechanism of tolerance can also be dependent on the developmental stage (Flowers, 2004).

Despite these obstacles, there has been some success with the use of transgenic cultivars to mitigate the impact of drought stress on crops. One released cultivar developed using transgenics is the Droughtgard™ hybrids from Monsanto. By isolating and transferring a cold-shock protein gene (cspB) from the soil bacteria “Bacillus subtilis”, yield improvements from 11%–21% were achieved under drought conditions, with no negative effects under normal conditions. The protein cspB acts as a chaperone for other proteins and is believed to help disentangle RNA which has folded abnormally due to drought. Yield improvement due to the expression of cspB is primarily due to an increase in the number of kernels per plant (Castiglioni et al., 2008).

Novel genome editing tools, such as CRISPR-Cas9, have also been used to create genetic diversity resulting in improved drought response. Using CRIPR-Cas9, the native maize GOS2 promoter was used by Shi et al. (2017) to both replace and supplement the native ARGOS8 promoter. This created variants with altered expression of ARGOS8, a negative regulator of ethylene responses. Some variants achieved a yield gain of five bushels per acre under flowering stress conditions, with no yield penalty under well-watered conditions, demonstrating the viability of the editing tool (Shi et al., 2017). Gene knockout study of ahb2 in maize via CRISPR/Cas9 resulted in quicker closure of stomata in response to water deficit stress, and three independent homozygous lines for the i1, d2 and d35 alleles that are tolerant to drought stress have been obtained (Liu et al., 2020). Gene editing approach will most likely be more attractive since crop varieties derived from this method could be considered non-genetically modified and thus be more acceptable.

Breeding for maize drought tolerance in SSA

Decades of partnerships with public and private sector institutions by the International Maize and Wheat Improvement Center (CIMMYT) have resulted in successful breeding and deployment of elite stress-tolerant maize cultivars across SSA (Edmeades et al., 1996b; Prasanna et al., 2021). CIMMYT leveraged molecular tools such as QTL and MARS to improve maize tolerance to drought, as well as other important stresses such as nitrogen use efficiency, maize streak virus, and maize lethal necrosis (Semagn et al., 2015). Advances in molecular and systems biology provided new opportunities to accelerate the maize improvement progress in SSA (Wossen et al., 2017). As a result, the adoption of improved maize varieties tolerant to drought and other stresses has increased across the region (Chivasa et al., 2022).

Two projects that stand out in SSA are public-private funded collaborations: the Drought Tolerant Maize for Africa (DTMA) and Water Efficient Maize for Africa (WEMA) (Oikeh et al., 2014; Nasser et al., 2020). DTMA is implemented jointly by CIMMYT and the International Institute for Tropical Agriculture, in collaboration with national agricultural research systems in participating nations. WEMA is a partnership involving the Bill & Melinda Gates Foundation, USAID, the Howard G. Buffett Foundation, CIMMYT, Monsanto, the National Agricultural Research Systems (NARS) from the participating nations (Kenya, Uganda, Tanzania, South Africa, Ethiopia and Mozambique), and led by the African Agricultural Technology Foundation (AATF) a nonprofit organization. The projects used conventional breeding, double haploid technology (DHT), transgenic technology, and marker assisted breeding. The objectives of WEMA and DTMA projects were to develop and distribute drought tolerant white maize hybrids that would yield more under drought stress than commercially used varieties (Oikeh et al., 2014; Nasser et al., 2020).

In the DTMA project, researchers used MARS to improve locally adapted germplasm and identify genomic regions associated with drought tolerance. The DTMA project resulted in the registration of 160 drought tolerant maize hybrids for release across 13 SSA countries between 2007 and 2013, with an adoption rate of between 9%–61% in six countries (Fisher et al., 2015). In 2020, 27 new, multiple-stress-tolerant maize hybrids and open-pollinated varieties were released by small and medium sized enterprises and national agricultural research system (NARS) partners for commercialization in SSA.

In the WEMA project, lines used to create the hybrids were also developed using DHT and selection achieved through MARS (Edge et al., 2017). About 106 non-transgenic drought tolerant maize hybrids were released across five participating countries (Edge et al., 2017), and sold under the name, DroughtTEGO, as well as several genetically modified (GM) varieties marketed under the brand name TELA, which are both drought tolerant and insect resistant from the Bt gene (Oikeh et al., 2014; Edge et al., 2017).

Plant breeding education

The shortage of plant breeders in SSA presents a serious challenge because it limits the development of improved varieties of crops for regional food security (Suza et al., 2016). To sustain the success of the maize drought tolerance projects, a pipeline of plant breeders and crop scientists must be in place in SSA. Support for human capacity enhancement can be modeled using successful partnerships such as Improved Master of Science in Cultivar Development for Africa and Plant Breeding E-Learning in Africa (Suza et al., 2016). In addition, centers in Africa, such as the African Centre for Crop Improvement (ACCI) in South Africa and West Africa Centre for Crop Improvement (WACCI) in Ghana, with funding largely from the Alliance for a Green Revolution in Africa (AGRA) have contributed plant breeding capacity in Africa. Several of the breeders from ACCI and WACCI have identified germplasm that contain useful alleles for drought tolerance improvement of maize inbred lines and hybrids under drought stress (Derera et al., 2008; Adebayo and Menkir, 2015; Meseka et al., 2018; Nasser et al., 2020). Evaluation of these germplasm has been achieved mainly through the utilization of conventional breeding via indirect selection using secondary traits drought selection indices. Other studies involving plant breeders from ACCI and WACCI also utilized marker-assisted recurrent selection and reported genetic gain under drought stress ranging from 22.7 kg ha−1 yr−1–118 kg ha−1 yr−1 (Beyene et al., 2016; Bankole et al., 2017; Masuka et al., 2017).

Conclusion

Since the development of hybrid maize, progress in breeding for yield potential has consistently improved drought tolerance. However, new strategies are needed to address the increase in demand and challenges brought about by climate change. While the goal is improved yield under drought conditions and yield stability, selecting on yield alone is inefficient due to the low heritability under stress conditions. Physiological dissection of yield into more heritable and selectable secondary traits can lead to better phenotyping strategies and improved gains. Advances in molecular biology provide new tools to understand the genetic basis of drought tolerance. Such tools have helped to identify associations between QTL and drought tolerance, yet gene interactions such as epistasis and pleiotropy, and gene by environment interactions complicate the use of genomic information for breeding. Integrating molecular genetics with physiology will help untangle the complex network of gene and environmental interactions, promoting the identification of the most promising loci conveying drought tolerance. Ecophysiological modeling can be used to better understand how genetic variability translates into the final phenotype by incorporating genomic information with known principles of crop growth and development. By integrating the vast amount of environmental, genetic, and physiological knowledge and applying it to a well-chosen germplasm base, plant breeders will be able to meet the challenge of drought stress and continue to deliver improved varieties in the future. The DTMA and WEMA projects in SSA emphasize the importance of multidisciplinary approaches that incorporate multiple breeding tools and approaches In addition, financial investment in human and institutional infrastructure capacity strengthening are needed to sustain the application and adoption innovations such as DTMA and WEMA in Africa.

Author contributions

MM and WS designed the manuscript. All authors contributed to writing the manuscript.

Funding

This work was supported by Iowa State University (ISU).

Acknowledgments

Siddique Aboobucker for reviewing the manuscript. We apologize to those colleagues whose important contributions may have been omitted.

Conflict of interest

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

Publisher’s note

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

References

Adebayo, M. A., and Menkir, A. (2015). Combining ability of adapted and exotic drought-tolerant maize inbred lines under full irrigation and rainfed conditions in Nigeria. J. Crop Improv. 29, 117–130. doi:10.1080/15427528.2014.980484

CrossRef Full Text | Google Scholar

Adnan, A. A., Diels, J., Jibrin, J. M., Kamara, A. Y., Shaibu, A. S., Craufurd, P., et al. (2020). CERES-Maize model for simulating genotype-by-environment interaction of maize and its stability in the dry and wet savannas of Nigeria. Field Crops Res. 253, 107826. doi:10.1016/j.fcr.2020.107826

PubMed Abstract | CrossRef Full Text | Google Scholar

Bankole, F., Menkir, A., Olaoye, G., Crossa, J., Hearne, S., Unachukwu, N., et al. (2017). Genetic gains in yield and yield related traits under drought stress and favorable environments in a maize population improved using marker assisted recurrent selection. Front. Plant Sci. 8, 808. doi:10.3389/fpls.2017.00808

PubMed Abstract | CrossRef Full Text | Google Scholar

Bänziger, M., and Araus, J.-L. (2007). “Recent advances in breeding maize for drought and salinity stress tolerance,” in Advances in molecular breeding toward drought and salt tolerant crops. Editors M. A. Jenks, P. M. Hasegawa, and S. M. Jain (Netherlands: Springer), 587–601.

Google Scholar

Barbosa, P. A. M., Fritsche-Neto, R., Andrade, M. C., Petroli, C. D., Burgueño, J., Galli, G., et al. (2021). Introgression of maize diversity for drought tolerance: Subtropical maize landraces as source of new positive variants. Front. Plant Sci. 12, 691211. doi:10.3389/fpls.2021.691211

PubMed Abstract | CrossRef Full Text | Google Scholar

Barker, T., Campos, H., Cooper, M., Dolan, D., Edmeades, G., Habben, J., et al. (2005). in Improving drought tolerance in maizePlant Breeding Reviews. Editor J. Janick (Oxford, UK: John Wiley & Sons), 173–253.

Google Scholar

Bernardo, R., and Yu, J. (2007). Prospects for genomewide selection for quantitative traits in maize. Crop Sci. 47 (3), 1082–1090. doi:10.2135/cropsci2006.11.0690

CrossRef Full Text | Google Scholar

Beyene, Y., Gowda, M., Olsen, M., Robbins, K. R., Pérez-Rodríguez, P., Alvarado, G., et al. (2019). Empirical comparison of tropical maize hybrids selected through genomic and phenotypic selections. Front. Plant Sci. 10, 1502. doi:10.3389/fpls.2019.01502

PubMed Abstract | CrossRef Full Text | Google Scholar

Beyene, Y., Gowda, M., Pérez-Rodríguez, P., Olsen, M., Robbins, K. R., Burgueño, J., et al. (2021). Empirical comparison of tropical maize hybrids selected through genomic and phenotypic selections. Front. Plant Sci. 12, 1502. doi:10.3389/fpls.2019.01502

CrossRef Full Text | Google Scholar

Beyene, Y., Semagn, K., Crossa, J., Mugo, S., Atlin, G. N., Tarekegne, A., et al. (2016). Improving maize grain yield under drought stress and non-stress environments in sub-saharan Africa using marker-assisted recurrent selection. Crop Sci. 56 (1), 344–353. doi:10.2135/cropsci2015.02.0135

CrossRef Full Text | Google Scholar

Blum, A. (2011a). Drought resistance - is it really a complex trait? Funct. Plant Biol. 38 (10), 753–757. doi:10.1071/FP11101

PubMed Abstract | CrossRef Full Text | Google Scholar

Blum, A. (2009). Effective use of water (EUW) and not water-use efficiency (WUE) is the target of crop yield improvement under drought stress. Field Crops Res. 112 (2), 119–123. doi:10.1016/j.fcr.2009.03.009

CrossRef Full Text | Google Scholar

Blum, A. (2014). Genomics for drought resistance – getting down to Earth. Funct. Plant Biol. 41 (11), 1191–1198. doi:10.1071/FP14018

PubMed Abstract | CrossRef Full Text | Google Scholar

Blum, A. (2011b). Plant breeding for water-limited environments. New York: Springer.

Google Scholar

Bolanos, J., and Edmeades, G. O. (1993). Eight cycles of selection for drought tolerance in lowland tropical maize. I. Responses in grain yield, biomass, and radiation utilization. Field Crops Res. 31 (3-4), 233–252. doi:10.1016/0378-4290(93)90064-t

CrossRef Full Text | Google Scholar

Brown, W. L. (1975). "A broader germplasm base in corn and sorghum", in: Proceedings of the 30th Annual Corn and Sorghum Research Conference.

Google Scholar

Buckler, E. S., and Thornsberry, J. M. (2002). Plant molecular diversity and applications to genomics. Curr. Opin. Plant Biol. 5 (2), 107–111. doi:10.1016/S1369-5266(02)00238-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Campos, H., Cooper, M., Habben, J. E., Edmeades, G. O., and Schussler, J. R. (2004). Improving drought tolerance in maize: A view from industry. Field Crops Res. 90 (1), 19–34. doi:10.1016/j.fcr.2004.07.003

CrossRef Full Text | Google Scholar

Castiglioni, P., Warner, D., Navarro, S., Back, S., Fernandes, M., Targolli, J., et al. (2008). Bacterial RNA chaperones confer abiotic stress tolerance in plants and improved grain yield in maize under water-limited conditions. Plant Physiol. 147 (2), 446–455. doi:10.1104/pp.108.118828

PubMed Abstract | CrossRef Full Text | Google Scholar

Castleberry, R. M., Crum, C. W., and Krull, C. F. (1984). Genetic yield improvement of U.S. maize cultivars under varying fertility and climatic environments. Crop Sci. 24 (1), 33–36. doi:10.2135/cropsci1984.0011183X002400010008x

CrossRef Full Text | Google Scholar

Cattivelli, L., Rizza, F., Badeck, F.-W., Mazzucotelli, E., Mastrangelo, A. M., Francia, E., et al. (2008). Drought tolerance improvement in crop plants: An integrated view from breeding to genomics. Field Crops Res. 105 (1), 1–14. doi:10.1016/j.fcr.2007.07.004

CrossRef Full Text | Google Scholar

Challinor, A. J., Koehler, A. K., Ramirez-Villegas, J., Whitfield, S., and Das, B. (2016). Current warming will reduce yields unless maize breeding and seed systems adapt immediately. Nat. Clim. Chang. 6 (10), 954–958. doi:10.1038/NCLIMATE3061

CrossRef Full Text | Google Scholar

Chapman, S. C., Crossa, J., and Edmeades, G. O. (1997). Genotype by environment effects and selection for drought tolerance in tropical maize. I. Two mode pattern analysis of yield. Euphytica 95 (1), 01–09. doi:10.1023/A:1002918008679

CrossRef Full Text | Google Scholar

Chenu, K., Chapman, S. C., Tardieu, F., McLean, G., Welcker, C., and Hammer, G. L. (2009). Simulating the yield impacts of organ-level quantitative trait loci associated with drought response in maize: A "gene-to-phenotype" modeling approach. Genetics 183 (4), 1507–1523. doi:10.1534/genetics.109.105429

PubMed Abstract | CrossRef Full Text | Google Scholar

Chenu, K. (2015). “Chapter 13 - characterizing the crop environment – nature, significance and applications,” in Crop physiology. Editors Victor O. Sadras, and Daniel F. Calderini. Second Edition (Academic Press), 321–348.

CrossRef Full Text | Google Scholar

Chenu, K., Cooper, M., Hammer, G. L., Mathews, K. L., Dreccer, M. F., and Chapman, S. C. (2011). Environment characterization as an aid to wheat improvement: Interpreting genotype–environment interactions by modelling water-deficit patterns in north-eastern Australia. J. Exp. Bot. 62 (6), 1743–1755. doi:10.1093/jxb/erq459

PubMed Abstract | CrossRef Full Text | Google Scholar

Chivasa, W., Worku, M., Teklewold, A., Setimela, P., Gethi, J., Magorokosho, C., et al. (2022). Maize varietal replacement in Eastern and Southern Africa: Bottlenecks, drivers and strategies for improvement. Glob. Food Sec. 32, 100589. doi:10.1016/j.gfs.2021.100589

PubMed Abstract | CrossRef Full Text | Google Scholar

Claassen, M. M., and Shaw, R. H. (1970). Water deficit effects on corn. II. Grain components. Agron. J. 62 (5), 652–655. doi:10.2134/agronj1970.00021962006200050032x

CrossRef Full Text | Google Scholar

Cooper, M., Gho, C., Leafgren, R., Tang, T., and Messina, C. (2014). Breeding drought-tolerant maize hybrids for the US corn-belt: Discovery to product. J. Exp. Bot. 65 (21), 6191–6204. doi:10.1093/jxb/eru064

PubMed Abstract | CrossRef Full Text | Google Scholar

Cooper, M., Stucker, R. E., DeLacy, I. H., and Harch, B. D. (1997). Wheat breeding nurseries, target environments, and indirect selection for grain yield. Crop Sci. 37 (4), 1168–1176. doi:10.2135/cropsci1997.0011183X003700040024x

CrossRef Full Text | Google Scholar

Cooper, M., van Eeuwijk, F. A., Chapman, S., Podlich, D. W., and Löffler, C. (2006). “Genotype-by-environment interactions under water-limited conditions,” in Drought adaptation in cereals. Editor J.-M. Ribaut (New York: The Haworth Press), 51–95.

Google Scholar

Derera, J., Tongoona, P., Vivek, B. S., and Laing, M. D. (2008). Gene action controlling grain yield and secondary traits in southern African maize hybrids under drought and non-drought environments. Euphytica 162, 411–422. doi:10.1007/s10681-007-9582-4

CrossRef Full Text | Google Scholar

Edge, M., Oikeh, S. O., Kyetere, D., Mugo, S., and Mashingaidze, K. (2017). “Water efficient maize for Africa: A public-private partnership in technology transfer to smallholder farmers in sub-saharan Africa,” in From agroscience to agribusiness. Editors N. Kalaitzandonakes, E. Carayannis, E. Grigoroudis, and S. Rozakis (Cham: Springer International Publishing), 391–412.

CrossRef Full Text | Google Scholar

Edmeades, G. O., Bänziger, M., Beck, D. L., Bolaños, J. A., and Ortega, C. A. (1996). “Development and per se performance of CIMMYT maize populations asdrought-tolerant sources,” in Developing drought and low N-tolerant Maize Proceedings of a symposium El Batán, Mexico, March 25–9, 1996. G. O. Edmeades, M. Bänziger, H. R. Mickelson, and C. B. Peña-Valdivia (Mexico: CIMMYT.

Google Scholar

Edmeades, G. O., Bolanos, J., and Chapman, S. C. (1996). “Value of secondary traits in selecting for drought tolerance in tropical maize,” in Developing drought- and low N-tolerant maize. Proceedings of a symposium, March 25-29, 1996, CIMMYT, El Batán, Mexico. Mexico, D.F. Editors G. O. Edmeades, M. Bänziger, H. R. Mickelson, and C. B. Peña-Valdivia (Mexico: CIMMYT).

Google Scholar

FAO (2021). The impact of disasters and crises on agriculture and food security. Rome: Food and agriculture organization of the United Nations.

Google Scholar

Faostat, (2010). Statistical database of the food and agriculture organization of the united nations. Rome: FAO.

Google Scholar

Farfan, I. D. B., De La Fuente, G. N., Murray, S. C., Isakeit, T., Huang, P.-C., Warburton, M., et al. (2015). Genome wide association study for drought, aflatoxin resistance, and important agronomic traits of maize hybrids in the sub-tropics. PLoS One 10 (2), e0117737. doi:10.1371/journal.pone.0117737

PubMed Abstract | CrossRef Full Text | Google Scholar

IPCC (2012). “Changes in climate extremes and their impacts on the natural physical environment,” in Managing the risks of extreme events and disasters to advance climate change adaptation. Editors C. B. B. V. Field, T. F. Stocker, D. Qin, D. J. Dokken, K. L. Ebiet al. (Cambridge, England: Cambridge University Press).

Google Scholar

Fisher, M., Abate, T., Lunduka, R. W., Asnake, W., Alemayehu, Y., and Madulu, R. B. (2015). Drought tolerant maize for farmer adaptation to drought in sub-Saharan Africa: Determinants of adoption in eastern and southern Africa. Clim. Change 133 (2), 283–299. doi:10.1007/s10584-015-1459-2

CrossRef Full Text | Google Scholar

Flowers, T. J. (2004). Improving crop salt tolerance. J. Exp. Bot. 55 (396), 307–319. doi:10.1093/jxb/erh003

PubMed Abstract | CrossRef Full Text | Google Scholar

Gao, Z., Liu, H., Wang, H., Li, N., Wang, D., Song, Y., et al. (2014). Generation of the genetic mutant population for the screening and characterization of the mutants in response to drought in maize. Chin. Sci. Bull. 59 (8), 766–775. doi:10.1007/s11434-013-0031-6

CrossRef Full Text | Google Scholar

Guo, M., Rupe, M. A., Wei, J., Winkler, C., Goncalves-Butruille, M., Weers, B. P., et al. (2014). Maize ARGOS1 (ZAR1) transgenic alleles increase hybrid maize yield. J. Exp. Bot. 65 (1), 249–260. doi:10.1093/jxb/ert370

PubMed Abstract | CrossRef Full Text | Google Scholar

Guo, M., Rupe, M. A., Zinselmeier, C., Habben, J., Bowen, B. A., and Smith, O. S. (2004). Allelic variation of gene expression in maize hybrids. Plant Cell 16 (7), 1707–1716. doi:10.1105/tpc.022087

PubMed Abstract | CrossRef Full Text | Google Scholar

Hajjapoor, A., Nelson, W. C. D., and Vadez, V. (2022). How process-based modeling can help plant breeding deal with G x E x M interactions. Field Crops Res. 283, 108554. doi:10.1016/j.fcr.2022.108554

CrossRef Full Text | Google Scholar

Hammer, G., Cooper, M., Tardieu, F., Welch, S., Walsh, B., van Eeuwijk, F., et al. (2006). Models for navigating biological complexity in breeding improved crop plants. Trends Plant Sci. 11 (12), 587–593. doi:10.1016/j.tplants.2006.10.006

PubMed Abstract | CrossRef Full Text | Google Scholar

Hao, Z., Li, X., Liu, X., Xie, C., Li, M., Zhang, D., et al. (2009). Meta-analysis of constitutive and adaptive QTL for drought tolerance in maize. Euphytica 174 (2), 165–177. doi:10.1007/s10681-009-0091-5

CrossRef Full Text | Google Scholar

Heinemann, A. B., Dingkuhn, M., Luquet, D., Combres, J. C., and Chapman, S. (2008). Characterization of drought stress environments for upland rice and maize in central Brazil. Euphytica 162 (3), 395–410. doi:10.1007/s10681-007-9579-z

CrossRef Full Text | Google Scholar

Heisey, P. W., and Edmeades, G. O. (1999). Maize production in drought-stressed environments: Technical options and research resource allocation. Mexico D.F.: CIMMYT.Maize production in drought-stressed environments: Technical options and research resource allocation. Part 1 of CIMMYT 1997/98 world maize facts and trends

Google Scholar

Ho, J. C., McCouch, S. R., and Smith, M. E. (2002). Improvement of hybrid yield by advanced backcross QTL analysis in elite maize. Theor. Appl. Genet. 105 (2-3), 440–448. doi:10.1007/s00122-002-0945-x

PubMed Abstract | CrossRef Full Text | Google Scholar

Jansen, R. C., Jannink, J-L., and Beavis, W. D. (2003). Mapping quantitative trait loci in plant breeding populations: Use of parental haplotype sharing. Crop Sci. 44 (3), 829–834. doi:10.2135/cropsci2003.8290

CrossRef Full Text | Google Scholar

Jeanneau, M., Gerentes, D., Foueillassar, X., Zivy, M., Vidal, J., Toppan, A., et al. (2002). Improvement of drought tolerance in maize: Towards the functional validation of the zm-asr1 gene and increase of water use efficiency by over-expressing C4–PEPC. Biochimie 84 (11), 1127–1135. doi:10.1016/S0300-9084(02)00024-X

PubMed Abstract | CrossRef Full Text | Google Scholar

Kamali, B., Abbaspou, K. C., Lehmann, A., Wehrli, B., and Yang, H. (2018). Spatial assessment of maize physical drought vulnerability in sub-Saharan Africa: Linking drought exposure with crop failure. Environ. Res. Lett. 12, 074010. doi:10.1088/1748-9326/aacb37

CrossRef Full Text | Google Scholar

Ke, F., and Ma, X. (2021). Responses of maize hybrids with contrasting maturity to planting date in Northeast China. Sci. Rep. 11 (1), 15776. doi:10.1038/s41598-021-95328-5

PubMed Abstract | CrossRef Full Text | Google Scholar

Khan, S. U., Zheng, Y., Chachar, Z., Zhang, X., Zhou, G., Zong, N., et al. (2022). Dissection of maize drought tolerance at the flowering stage using genome-wide association studies. Genes 13, 564. doi:10.3390/genes13040564

PubMed Abstract | CrossRef Full Text | Google Scholar

Korte, A., and Farlow, A. (2013). The advantages and limitations of trait analysis with GWAS: A review. Plant Methods 9 (1), 29. doi:10.1186/1746-4811-9-29

PubMed Abstract | CrossRef Full Text | Google Scholar

Levitt, J. (1972). Responses of plants to environmental stresses. New York: Academic Press.

Google Scholar

Li, C., Sun, B., Li, Y., Liu, C., Wu, X., Zhang, D., et al. (2016). Numerous genetic loci identified for drought tolerance in the maize nested association mapping populations. BMC genomics 17 (1), 894. doi:10.1186/s12864-016-3170-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu, S., Li, C., Wang, H., Wang, S., Yang, S., Liu, X., et al. (2020). Mapping regulatory variants controlling gene expression in drought response and tolerance in maize. Genome Biol. 21 (1), 163. doi:10.1186/s13059-020-02069-1

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu, S., and Qin, F. (2021). Genetic dissection of maize drought tolerance for trait improvement. Mol. Breed. 41, 8. doi:10.1007/s11032-020-01194-w

CrossRef Full Text | Google Scholar

Liu, S., Yang, J., Yang, X., Drury, C. F., Jiang, R., and Reynolds, W. D. (2021). Simulating maize yield at county scale in southern Ontario using the decision support system for agrotechnology transfer model. Can. J. Soil Sci. 101 (4), 734–748. doi:10.1139/cjss-2020-0116

CrossRef Full Text | Google Scholar

Lobell, D. B., Schlenker, W., and Costa-Roberts, J. (2011). Climate trends and global crop production since 1980. Science 333 (6042), 616–620. doi:10.1126/science.1204531

PubMed Abstract | CrossRef Full Text | Google Scholar

Lu, Y., Zhang, S., Shah, T., Xie, C., Hao, Z., Li, X., et al. (2010). Joint linkage–linkage disequilibrium mapping is a powerful approach to detecting quantitative trait loci underlying drought tolerance in maize. Proc. Natl. Acad. Sci. U. S. A. 107 (45), 19585–19590. doi:10.1073/pnas.1006105107

PubMed Abstract | CrossRef Full Text | Google Scholar

Lunduka, R. W., Mateva, K. I., Magorokosho, C., and Manjeru, P. (2017). Impact of adoption of drought-tolerant maize varieties on total maize production in south Eastern Zimbabwe. Clim. Dev. 11, 35–46. doi:10.1080/17565529.2017.1372269

PubMed Abstract | CrossRef Full Text | Google Scholar

Mao, H., Wang, H., Liu, S., Li, Z., Yang, X., Yan, J., et al. (2015). A transposable element in a NAC gene is associated with drought tolerance in maize seedlings. Nat. Commun. 6, 8326. doi:10.1038/ncomms9326

PubMed Abstract | CrossRef Full Text | Google Scholar

Marino, R., Ponnaiah, M., Krajewski, P., Frova, C., Gianfranceschi, L., Pè, M. E., et al. (2008). Addressing drought tolerance in maize by transcriptional profiling and mapping. Mol. Genet. Genomics 281 (2), 163–179. doi:10.1007/s00438-008-0401-y

PubMed Abstract | CrossRef Full Text | Google Scholar

Masuka, B., Atlin, G. N., Olsen, M., Magorokosho, C., Labuschagne, M., Crossa, J., et al. (2017). Gains in maize genetic improvement in eastern and southern Africa: I. CIMMYT hybrid breeding pipeline. Crop Sci. 57 (1), 168–179. doi:10.2135/cropsci2016.05.0343

CrossRef Full Text | Google Scholar

Meseka, S., Fakorede, M., Ajala, S., Badu-Apraku, B., and Menkir, A. (2013). Introgression of alleles from maize landraces to improve drought tolerance in an adapted germplasm. J. Crop Improv. 27 (1), 96–112. doi:10.1080/15427528.2012.729259

CrossRef Full Text | Google Scholar

Meseka, S., Menkir, A., Bossey, B., and Mengesha, W. (2018). Performance assessment of drought tolerant maize hybrids under combined drought and heat stress. Agronomy 8, 274. doi:10.3390/agronomy8120274

PubMed Abstract | CrossRef Full Text | Google Scholar

Messina, C. D., Podlich, D., Zhanshan, D., Samples, M., and Cooper, M. (2011). Yield–trait performance landscapes: From theory to application in breeding maize for drought tolerance. J. Exp. Bot. 62 (3), 855–868. doi:10.1093/jxb/erq329

PubMed Abstract | CrossRef Full Text | Google Scholar

Messina, C., McDonald, D., Poffenberger, H., Clark, R., Salinas, A., Fang, Y., et al. (2021). Reproductive resilience but not root architecture underpins yield improvement under drought in maize. J. Exp. Bot. 72, 5235–5245. doi:10.1093/jxb/erab231

PubMed Abstract | CrossRef Full Text | Google Scholar

Meuwissen, T. H. E., Hayes, B. J., and Goddard, M. E. (2001). Prediction of total genetic value using genome-wide dense marker maps. Genetics 157 (4), 1819–1829. doi:10.1093/genetics/157.4.1819

PubMed Abstract | CrossRef Full Text | Google Scholar

Mir, R. R., Zaman-Allah, M., Sreenivasulu, N., Trethowan, R., and Varshney, R. K. (2012). Integrated genomics, physiology and breeding approaches for improving drought tolerance in crops. Theor. Appl. Genet. 125 (4), 625–645. doi:10.1007/s00122-012-1904-9

PubMed Abstract | CrossRef Full Text | Google Scholar

Monneveux, P., and Ribaut, J.-M. (2006). “Secondary traits for drought tolerance improvement in cereals,” in Drought adaptation in cereals. Editor J.-M. Ribaut (Binghamton, NY: The Haworth Press, Inc;), 97–143.

Google Scholar

Myles, S., Peiffer, J., Brown, P. J., Ersoz, E. S., Zhang, Z., Costich, D. E., et al. (2009). Association mapping: Critical considerations shift from genotyping to experimental design. Plant Cell 21 (8), 2194–2202. doi:10.1105/tpc.109.068437

PubMed Abstract | CrossRef Full Text | Google Scholar

Nasser, L. M., Badu-Apraku, B., Gracen, V. E., and Mafouasson, N. A. (2020). Combining ability of early-maturing Yellow maize inbreds under combined drought and heat stress and well-watered environments. Agronomy 10 (10), 1585. doi:10.3390/agronomy10101585

CrossRef Full Text | Google Scholar

Nelson, D. E., Repetti, P. P., Adams, T. R., Creelman, R. A., Wu, J., Warner, D. C., et al. (2007). Plant nuclear factor Y (NF-Y) B subunits confer drought tolerance and lead to improved corn yields on water-limited acres. Proc. Natl. Acad. Sci. U. S. A. 104 (42), 16450–16455. doi:10.1073/pnas.0707193104

PubMed Abstract | CrossRef Full Text | Google Scholar

Nuccio, M. L., Wu, J., Mowers, R., Zhou, H.-P., Meghji, M., Primavesi, L. F., et al. (2015). Expression of trehalose-6-phosphate phosphatase in maize ears improves yield in well-watered and drought conditions. Nat. Biotechnol. 33 (8), 862–869. doi:10.1038/nbt.3277

PubMed Abstract | CrossRef Full Text | Google Scholar

Oikeh, S., Ngonyamo-Majee, D., Mugo, S. I. N., Mashingaidze, K., Cook, V., and Stephens, M. (2014). “The water efficient maize for Africa project as an example of a public–private partnership,” in Convergence of food security, Energy security and sustainable agriculture. Biotechnology in agriculture and Forestry. Editors D. Songstad, J. Hatfield, and D. Tomes (Berlin, Heidelberg: Springer). doi:10.1007/978-3-642-55262-5_13

CrossRef Full Text | Google Scholar

Otegui, M. E., and Andrade, F. H. (2000). “New relationships between light interception, ear growth, and kernel set in maize,” in Physiology and modeling kernel set in maize. Editors M. Westgate, and K. Boote (Baltimore, Maryland: CSSA Special Publication Number 29), 89–102.

Google Scholar

Podlich, D. W., and Cooper, M. (1998). QU-GENE: A simulation platform for quantitative analysis of genetic models. Bioinformatics 14 (7), 632–653. doi:10.1093/bioinformatics/14.7.632

PubMed Abstract | CrossRef Full Text | Google Scholar

Podlich, D. W., Winkler, C. R., and Cooper, M. (2004). Mapping as you go: An effective approach for marker-assisted selection of complex traits. Crop Sci. 44 (5), 1560–1571. doi:10.2135/cropsci2004.1560

CrossRef Full Text | Google Scholar

Ramirez-Villegas, J., Milan, A. M., Alexandrov, N., Asseng, S., Challinor, A. J., Crossa, J., et al. (2020). CGIAR modeling approaches for resource-constrained scenarios: I. Accelerating crop breeding for a changing climate. Crop Sci. 60, 547–567. doi:10.1002/csc2.20048

CrossRef Full Text | Google Scholar

Resende, M. D. V. d., Silva, F. F. e., Resende, M. F. R., and Azevedo, C. F. (2014). “Chapter 5 - genome-wide selection (GWS),” in Biotechnology and plant breeding. Editors A. Borem, and R. Fritsche-Neto (San Diego: Academic Press), 105–133.

CrossRef Full Text | Google Scholar

Reymond, M., Muller, B., Leonardi, A., Charcosset, A., and Tardieu, F. (2003). Combining quantitative trait loci analysis and an ecophysiological model to analyze the genetic variability of the responses of maize leaf growth to temperature and water deficit. Plant Physiol. 131 (2), 664–675. doi:10.1104/pp.013839

PubMed Abstract | CrossRef Full Text | Google Scholar

Ribaut, J.-M., Betran, J., Monneveux, P., and Setter, T. (2009). “Drought tolerance in maize,” in Handbook of maize: Its biology. Editors J. L. Bennetzen, and S. C. Hake (New York: Springer), 311–344.

CrossRef Full Text | Google Scholar

Ribaut, J.-M. (2006). Drought adaptation in cereals. Binghamton, NY: The Haworth Press.

Google Scholar

Ribaut, J.-M., and Ragot, M. (2006). Marker-assisted selection to improve drought adaptation in maize: The backcross approach, perspectives, limitations, and alternatives. J. Exp. Bot. 58 (2), 351–360. doi:10.1093/jxb/erl214

PubMed Abstract | CrossRef Full Text | Google Scholar

Riedelsheimer, C., Lisec, J., Czedik-Eysenberg, A., Sulpice, R., Flis, A., Grieder, C., et al. (2012). Genome-wide association mapping of leaf metabolic profiles for dissecting complex traits in maize. Proc. Natl. Acad. Sci. U. S. A. 109 (23), 8872–8877. doi:10.1073/pnas.1120813109

PubMed Abstract | CrossRef Full Text | Google Scholar

Rosegrant, M. R., Ringler, C., Sulser, T. B., Ewing, M., Palazzo, A., Zhu, T., et al. (2009). Agriculture and food security under global change: Prospects for 2025/2050. Washington, DC: International Food Policy Research Institute.

Google Scholar

Ruswandi, D., AgustianAnggia, E. P., Canama, A., Marta, H., Ruswandi, S., et al. (2014). Mutation breeding of maize for anticipating global climate change in Indonesia. Asian J. Agric. Res. 8, 234–247. doi:10.3923/ajar.2014.234.247

CrossRef Full Text | Google Scholar

Sah, R. P., Chakraborty, M., Prasad, K., Pandit, M., Tudu, V. K., Chakravarty, M. K., et al. (2020). Impact of water deficit stress in maize: Phenology and yield components. Sci. Rep. 10, 2944–2945. doi:10.1038/s41598-020-59689-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Salvi, S., Sponza, G., Morgante, M., Tomes, D., Niu, X., Fengler, K. A., et al. (2007). Conserved noncoding genomic sequences associated with a flowering-time quantitative trait locus in maize. Proc. Natl. Acad. Sci. U. S. A. 104 (27), 11376–11381. doi:10.1073/pnas.0704145104

PubMed Abstract | CrossRef Full Text | Google Scholar

Sammons, D. J. (1987). Origin and early history of maize in East Africa with special reference to Kenya. Njoro, Kenya: Egerton University Research Paper Series, 2–11.

Google Scholar

Sawkins, M. C., Farmer, A. D., Hoisington, D., Sullivan, J., Tolopko, A., Jiang, Z., et al. (2004). Comparative map and trait viewer (CMTV): An integrated bioinformatic tool to construct consensus maps and compare QTL and functional genomics data across genomes and experiments. Plant Mol. Biol. 56 (3), 465–480. doi:10.1007/s11103-004-4950-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Schlenker, W., and Lobell, D. B. (2010). Robust negative impacts of climate change on African agriculture. Environ. Res. Lett. 5 (1), 014010. doi:10.1088/1748-9326/5/1/014010

CrossRef Full Text | Google Scholar

Semagn, K., Beyene, Y., Babu, R., Nair, S., Gowda, M., Das, B., et al. (2015). Quantitative trait loci mapping and molecular breeding for developing stress resilient maize for sub-saharan Africa. Crop Sci. 55 (4), 1449–1459. doi:10.2135/cropsci2014.09.0646

CrossRef Full Text | Google Scholar

Semagn, K., Beyene, Y., Warburton, M. L., Tarekegne, A., Mugo, S., Meisel, B., et al. (2013). Meta-analyses of QTL for grain yield and anthesis silking interval in 18 maize populations evaluated under water-stressed and well-watered environments. BMC genomics 14 (1), 313. doi:10.1186/1471-2164-14-313

PubMed Abstract | CrossRef Full Text | Google Scholar

Setter, T. L., Yan, J., Warburton, M., Ribaut, J.-M., Xu, Y., Sawkins, M., et al. (2011). Genetic association mapping identifies single nucleotide polymorphisms in genes that affect abscisic acid levels in maize floral tissues during drought. J. Exp. Bot. 62 (2), 701–716. doi:10.1093/jxb/erq308

PubMed Abstract | CrossRef Full Text | Google Scholar

Seyoum, S., Chauhan, Y., Rachaputi, R., Fekybelu, S., and Boddupalli, P. (2017). Characterising production environments for maize in eastern and southern Africa using the APSIM Model. Agric. For. Meteorology 247, 445–453. doi:10.1016/j.agrformet.2017.08.023

CrossRef Full Text | Google Scholar

Shi, J., Gao, H., Wang, H., Lafitte, H. R., Archibald, R. L., Yang, M., et al. (2017). ARGOS8 variants generated by CRISPR‐Cas9 improve maize grain yield under field drought stress conditions. Plant Biotechnol. J. 15 (2), 207–216. doi:10.1111/pbi.12603

PubMed Abstract | CrossRef Full Text | Google Scholar

Shiferaw, B., Prasanna, B. M., Hellin, J., and Bänziger, M. (2011). Crops that feed the world 6. Past successes and future challenges to the role played by maize in global food security. Food Secur. 3 (3), 307–327. doi:10.1007/s12571-011-0140-5

CrossRef Full Text | Google Scholar

Sinclair, T. R., and Purcell, L. C. (2005). Is a physiological perspective relevant in a ‘genocentric’ age? J. Exp. Bot. 56 (421), 2777–2782. doi:10.1093/jxb/eri297

PubMed Abstract | CrossRef Full Text | Google Scholar

Singh, B. D. (2010). Plant breeding: Principles and methods. New Delhi - Ludhiana: Kalyani Publishers.

Google Scholar

Suza, W. P., Gibson, P., Edema, R., Akromah, R., Sibiya, J., Madakadze, R., et al. (2016). Plant breeding capacity building in Africa. Nat. Clim. Chang. 6, 976. doi:10.1038/nclimate3139

CrossRef Full Text | Google Scholar

Tanksley, S. D., and Nelson, J. C. (1996). Advanced backcross QTL analysis: A method for the simultaneous discovery and transfer of valuable QTLs from unadapted germplasm into elite breeding lines. Theor. Appl. Genet. 92 (2), 191–203. doi:10.1007/BF00223376

PubMed Abstract | CrossRef Full Text | Google Scholar

Tardieu, F. (2012). Any trait or trait-related allele can confer drought tolerance: Just design the right drought scenario. J. Exp. Bot. 63 (1), 25–31. doi:10.1093/jxb/err269

PubMed Abstract | CrossRef Full Text | Google Scholar

Tardieu, F. (2006). “Leaf growth under water-limited conditions,” in Drought adaptation in cereals. Editor J.-M. Ribaut (Binghamton, NY: The Haworth Press).

Google Scholar

Tardieu, F., and Tuberosa, R. (2010). Dissection and modelling of abiotic stress tolerance in plants. Curr. Opin. Plant Biol. 13 (2), 206–212. doi:10.1016/j.pbi.2009.12.012

PubMed Abstract | CrossRef Full Text | Google Scholar

Tesfaye, K., Kruseman, G., Cairns, J. E., Zaman-Allah, M., Wegary, D., Zaidi, P. H., et al. (2018). Potential benefits of drought and heat tolerance for adapting maize to climate change in tropical environments. Clim. Risk Manag. 19, 106–119. doi:10.1016/j.crm.2017.10.001

CrossRef Full Text | Google Scholar

Tollenaar, M., and Lee, E. (2006). Physiological dissection of grain yield in maize by examining genetic improvement and heterosis. Maydica 51 (2), 399. doi:10.2135/cropsci2007.04.0010IPBS

CrossRef Full Text | Google Scholar

Tsonev, S., Todorovska, E., Avramova, V., Kolev, S. E., Abu-Mhadi, N., and Christov, N. K. (2009). Genomics assisted improvement of drought tolerance in maize: QTL approaches. Biotechnol. Biotechnol. Equip. 23, 1410–1413. doi:10.2478/V10133-009-0004-8

CrossRef Full Text | Google Scholar

Tuberosa, R., and Salvi, S. (2006). Genomics-based approaches to improve drought tolerance of crops. Trends Plant Sci. 11 (8), 405–412. doi:10.1016/j.tplants.2006.06.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Tuberosa, R., Salvi, S., Giuliani, S., Sanguineti, M. C., Bellotti, M., Conti, S., et al. (2007). Genome‐wide approaches to investigate and improve maize response to drought. Crop Sci. 47, S-120–S-141. doi:10.2135/cropsci2007.04.0001IPBS

CrossRef Full Text | Google Scholar

Tuberosa, R., Salvi, S., Sanguineti, M. C., Landi, P., Maccaferri, M., and Conti, S. (2002). Mapping QTLs regulating morpho-physiological traits and yield: Case studies, shortcomings and perspectives in drought-stressed maize. Ann. Bot. 89 (7), 941–963. doi:10.1093/aob/mcf134

PubMed Abstract | CrossRef Full Text | Google Scholar

Vargas, M., van Eeuwijk, F. A., Crossa, J., and Ribaut, J.-M. (2006). Mapping QTLs and QTL x environment interaction for CIMMYT maize drought stress program using factorial regression and partial least squares methods. Theor. Appl. Genet. 112 (6), 1009–1023. doi:10.1007/s00122-005-0204-z

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, N., Cheng, M., Chen, Y., Liu, B., Wang, X., Li, G., et al. (2021). Natural variations in the non-coding region of ZmNAC080308 contributes maintaining grain yield under drought stress in maize. BMC Plant Biol. 21, 305. doi:10.1186/s12870-021-03072-9

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, X., Wang, H., Liu, S., Ferjani, A., Li, J., Yan, J., et al. (2016). Genetic variation in ZmVPP1 contributes to drought tolerance in maize seedlings. Nat. Genet. 48, 1233–1241. doi:10.1038/ng.3636

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, Z., Zhao, X., Ren, Z., Abou-Elwafa, S. F., Pu, X., Zhu, Y., et al. (2022). ZmERF21 directly regulates hormone signaling and stress-responsive gene expression to influence drought tolerance in maize seedlings. Plant Cell Environ. 45, 312–328. doi:10.1111/pce.14243

PubMed Abstract | CrossRef Full Text | Google Scholar

Welcker, C., Boussuge, B., Bencivenni, C., Ribaut, J. M., and Tardieu, F. (2006). Are source and sink strengths genetically linked in maize plants subjected to water deficit? A QTL study of the responses of leaf growth and of anthesis-silking interval to water deficit. J. Exp. Bot. 58 (2), 339–349. doi:10.1093/jxb/erl227

PubMed Abstract | CrossRef Full Text | Google Scholar

Westgate, M. E., and Boyer, J. S. (1986). Reproduction at low and pollen water potentials in maize. Crop Sci. 26 (5), 951–956. doi:10.2135/cropsci1986.0011183X002600050023x

CrossRef Full Text | Google Scholar

White, G. C., Chapman, L. M., Slate, W. L., and Brown, B. A. (1922). A comparison of early, medium and late maturing varieties of silage corn for milk production: A report of progress. J. Dairy Sci. 5 (4), 333–347. doi:10.3168/jds.s0022-0302(22)94161-x

CrossRef Full Text | Google Scholar

Wossen, T., Abdoulaye, T., Alene, A., Feleke, S., Menkir, A., and Manyong, V. (2017). Measuring the impacts of adaptation strategies to drought stress: The case of drought tolerant maize varieties. J. Environ. Manage. 203 (1), 106–113. doi:10.1016/j.jenvman.2017.06.058

PubMed Abstract | CrossRef Full Text | Google Scholar

Yamada, T., and Dwiyanti, M. (2013). “Molecular mapping and breeding for genes/QTLS related to climate change,” in Genomics and breeding for climate -resilient crops. Editor C. Kole (Berlin, Heidelberg: Springer Berlin Heidelberg), 179–212.

CrossRef Full Text | Google Scholar

Yang, S., Vanderbeld, B., Wan, J., and Huang, Y. (2010). Narrowing down the targets: Towards successful genetic engineering of drought-tolerant crops. Mol. Plant 3 (3), 469–490. doi:10.1093/mp/ssq016

PubMed Abstract | CrossRef Full Text | Google Scholar

Yu, J., Holland, J. B., McMullen, M. D., and Buckler, E. S. (2008). Genetic design and statistical power of nested association mapping in maize. Genetics 178 (1), 539–551. doi:10.1534/genetics.107.074245

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, X., Warburton, M. L., Setter, T., Liu, H., Xue, Y., Yang, N., et al. (2016). Genome-wide association studies of drought-related metabolic changes in maize using an enlarged SNP panel. Theor. Appl. Genet. 129 (8), 1449–1463. doi:10.1007/s00122-016-2716-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhao, X., Peng, Y., Zhang, J., Fang, P., Wu, B., and Wang, Y. (2018). Shenmai injection protects mitochondria from oxidative injury in myocardial cells and its mechanism. Crop Sci. 58, 507–513. doi:10.3785/j.issn.1008-9292.2018.10.10

PubMed Abstract | CrossRef Full Text | Google Scholar

Zinselmeier, C., Sun, Y., Helentjaris, T., Beatty, M., Yang, S., Smith, H., et al. (2002). The use of gene expression profiling to dissect the stress sensitivity of reproductive development in maize. Field Crops Res. 75 (2), 111–121. doi:10.1016/S0378-4290(02)00021-7

CrossRef Full Text | Google Scholar

Ziyomo, C., and Bernardo, R. (2013). Drought tolerance in maize: Indirect selection through secondary traits versus genome wide selection. Crop Sci. 53 (4), 1269–1275. doi:10.2135/cropsci2012.11.0651

CrossRef Full Text | Google Scholar

Keywords: drought tolerance, food security, maize breeding, genomics assisted selection, genome mapping, model-assisted approaches, plant breeding education

Citation: McMillen MS, Mahama AA, Sibiya J, Lübberstedt T and Suza WP (2022) Improving drought tolerance in maize: Tools and techniques. Front. Genet. 13:1001001. doi: 10.3389/fgene.2022.1001001

Received: 22 July 2022; Accepted: 14 October 2022;
Published: 28 October 2022.

Edited by:

Reyazul Rouf Mir, Sher-e-Kashmir University of Agricultural Sciences and Technology, India

Reviewed by:

Jiuran Zhao, Beijing Academy of Agricultural and Forestry Sciences, China
Tim L. Setter, Cornell University, United States

Copyright © 2022 McMillen, Mahama, Sibiya, Lübberstedt and Suza. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Walter P. Suza, wpsuza@iastate.edu

Present address: Michael S. McMillen, Syngenta, Greensboro, NC, United States

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