Next Article in Journal
Using Regulatory Flexibility to Address Market Informality in Seed Systems: A Global Study
Next Article in Special Issue
Genetic Diversity and Population Structure of Soybean Lines Adapted to Sub-Saharan Africa Using Single Nucleotide Polymorphism (SNP) Markers
Previous Article in Journal
Resistance of European Spring 2-Row Barley Cultivars to Pyrenophora graminea and Detection of Associated Loci
Previous Article in Special Issue
Genome-Wide Identification of the NHX Gene Family in Punica granatum L. and Their Expressional Patterns under Salt Stress
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Modern Approaches for the Genetic Improvement of Rice, Wheat and Maize for Abiotic Constraints-Related Traits: A Comparative Overview

Department of Biotechnology-Plant Biology, School of Agricultural, Food and Biosystems Engineering, Universidad Politécnica de Madrid, 28040 Madrid, Spain
*
Authors to whom correspondence should be addressed.
Submission received: 30 December 2020 / Revised: 12 February 2021 / Accepted: 13 February 2021 / Published: 20 February 2021

Abstract

:
After a basic description of the different sets of genetic tools and genomic approaches most relevant for modern crop breeding (e. g., QTL mapping, GWAS and genomic selection; transcriptomics, qPCR and RNA-seq; transgenesis and gene editing), this review paper describes their history and the main achievements in rice, wheat and maize research, with a further focus on crop traits related to the improvement of plant responses to face major abiotic constrains, including nutritional limitations, drought and heat tolerance, and nitrogen-use efficiency (NUE). Remarkable differences have been evidenced regarding the timing and degree of development of some genetic approaches among these major crops. The underlying reasons related to their distinct genome complexity, are also considered. Based on bibliographic records, drought tolerance and related topics (i.e., water-use efficiency) are by far the most abundantly addressed by molecular tools among the breeding objectives considered. Heat tolerance is usually more relevant than NUE in rice and wheat, while the opposite is true for maize.

1. Introduction

The genetic structure of plants has been manipulated by farmers since agriculture began 10,000 years ago. For centuries, the domestication process has progressed on the basis of selecting seeds of the grains or fruits from the individuals best adapted to local human needs and practices. This manipulation of natural diversity has outlined the genetic architecture of currently existing crops, which in some instances differ radically from their wild ancestors [1].
Plant improvement has been carried out throughout history using the scientific knowledge available. Therefore, soon after the discovery of plant sexual reproduction, early breeders implemented crossing designs to increase the variability available for selection. The rediscovery of Mendel’s laws of inheritance and the spread of Darwin’s ideas on the evolution of species in the early twentieth century provided a founding theoretical basis; thus, that crop improvement experienced a major transformation. The recognition of the value of genetic variability and the ability to predict the phenotypic outcome of designed crosses were a revolution that turned crop improvement into a scientific discipline in a few years [2].
Conventional genetic improvement, based on the application of classical genetic principles of trait transmission, has successfully introduced suitable characteristics into cultivars. During the breeding process, offspring must be tested in each generation to select superior individuals as parents of the next generation until the desired combination of crop traits is found.
The classical breeding approach has two main limitations: (i) only genetic variability existing in the crop species itself or in cross-compatible species can be used for crop improvement; (ii) the selection of superior genotypes is based on phenotypes. This may significantly affect the investment to develop new varieties; for instance, if trait evaluation requires time-consuming methods or expensive equipment, or when a large number of plants needs to be grown until maturity in field conditions. Moreover, for some complex polygenic traits, individual’s phenotypes can be unreliable, and the identification of valuable genotypes requires phenotyping of their progenies (i.e., progeny testing) (Chapters 5 and 8 in ref [3]).
Modern breeding has addressed these limitations by incorporating biotechnological tools that can overcome cross-sexual barriers or create novel variability in the crop species, or that are ultimately applied to improve the reliability of selection programs based on the genotyping of segregant progenies [2]. The development of advanced tools for the qualitative and quantitative analysis of gene expression in organs and tissues must also be highlighted, because these methodologies provide the molecular basis to establish functional links between genes and traits [4]. Some of the earliest methods, introduced by the 1980s, have been virtually replaced by new techniques or approaches that allow the addressing of similar goals either with increased efficiency, by saving time, or both. The most recent incorporations to the molecular breeding toolbox have emerged from the current capacity to sequence organisms at the genome level combined with bioinformatics developments. The genome contains all the individual’s hereditary information and Genomics has become essential to understand how genes control the traits exhibited by an organism and how the traits are transmitted to its offspring. Therefore, plant genomics has enormous potential, providing very valuable information that can be exploited for varietal improvement in breeding programs [5].
Rice, wheat, and maize are acknowledged as the most relevant crops worldwide, not only because of their leading position in agricultural and economic terms (Figure 1), but also because these three crops represent more than 50% of the human daily caloric intake. It can be added that maize has also a major indirect role in human nutrition by supplying most of the nutrients for livestock, along with other cereal grains (i.e., barley and sorghum) [6].
Some distinctive biological features of rice, wheat and maize must be taken into account to understand the remarkable differences regarding the timing and usage of some biotechnological developments in each of these crops (see below), in spite of a similar interest which can be assumed to exist on the breeders’ side.
Rice (Oryza sativa L.) is a diploid species (2n = 2x = 24) with the smallest-sized genome among all the crop plants of economic importance (430 Mbp). It explains the earliness of its full genome sequencing compared to any other main crops [7,8]. Other advantages, such as the availability of easier and efficient transformation protocols by the early 1990s, led to the adoption of rice as a model cereal for molecular studies [9]. Accordingly, all fine mapping and reverse-genetic tools, first developed in the model plant Arabidopsis thaliana, were soon applied to the structural and functional analysis of rice genes. Rice is likely the crop plant with the largest genomic toolbox [10,11,12,13]. Its panel of analytical resources also includes several wide collections of germplasm resources, most of which were developed and are being used by international consortia [14,15].
Bread wheat (Triticum aestivum L.) is a hexaploid species (2n = 6x = 42 chromosomes) with one of the largest genomes among all crop plants (around 16,000 Mbp). The wheat genome is composed of three homoeologous sub-genomes, with a basic number of x = 7 each (AABBDD genome composition). Its large and structurally complex genome has hindered the application of plant molecular tools to wheat and explains why the sequencing of the wheat genome has not been completed until very recently [16]. Notwithstanding, the work of Sears in the 1940–1950s [17,18] made a stock of more than 250 aneuploid lines of the wheat cultivar Chinese Spring available, which has been used for the mapping and expression studies of many wheat genes, even before the advent of DNA-based analytical tools [19]. For the same purposes, chromosome-engineered deletion, substitution, and recombinant lines derived from the initial Sears’ lines, or created from other wheat cultivars, have also been (and are still being) extensively employed [20]. Durum or pasta wheat (Triticum turgidum L.; 2n = 4x = 28, AABB genome composition) is not as relevant as bread wheat from an economical point of view, although has often been pioneering in the application of DNA-based methodologies to commercial wheat crops because of its less complex genetic/genomic architecture (tetraploid versus hexaploid). Although the genome of the reference cultivar Svevo was not published until 2019 [21], previous assemblies of the durum wheat genome were key to complete the sequencing of the bread wheat genome [22].
Maize (Zea mays L.; 2n = 2x = 20) has a medium-sized genome of about 2300 Mbp (in the range of the human genome) that was sequenced in 2009 [23]. In spite of its strictly diploid behavior, there is much molecular evidence of its ancient polyploid origin [24,25]. From a breeder’s perspective, maize is a crop radically different from rice and wheat. The distinct mode of reproduction (allogamy in maize and autogamy in rice and wheat) is on the basis of the different classical breeding methods that can be applied in each case [3]. Maize was the first crop where the superiority of hybrids over their parental lines, a biological phenomenon known as heterosis, was discovered and exploited for plant improvement; hence, most current elite maize varieties are F1 hybrids. Rice and wheat varieties are mostly inbred pure-lines that can be easily multiplied in the farm by the naturally occurring autogamous reproduction of these self-pollinated species, However, hybrid genotypes can only be multiplied by the parental lines’ owner, and hybrid seed must be bought every year. This makes maize varieties the most profitable for seed companies and cost-effective for expensive biotechnology investment.
This review describes the most common breeding-assistant tools that are based on the molecular knowledge of genes and genomes, grouped in three categories according to their primary purpose: marker-assisted breeding, gene expression analysis, and genetic modification. We have aimed to provide a comparative overview of their timing and usage in rice, wheat, and maize, with a special focus on milestones and specific achievements regarding the improvement of crop attributes involved in plant response to major climatic constraints and nutritional limitations—drought and heat tolerance and nitrogen-use efficiency. These three goals are the main crop breeding objectives to overcome the most challenging abiotic threats to global food security [26]. With that in mind, we have conducted: (i) a bibliographic search on the first documented evidence of use of the main molecular tools in these major crops; and (ii) a bibliometric analysis on the number of records retrieved from the Web of Science database for specific tools, globally and locally, for the breeding topics of interest. Details on the search queries and the methodology followed for these analyses are provided as Supplementary Materials.

2. Marker-Assisted Breeding

A molecular marker is any polymorphism at the DNA level. DNA-based markers are much more numerous (theoretically, up to one marker per base-pair of the crop genome) and more evenly distributed throughout the whole genome than classical genetic markers (i.e., morphological or biochemical). Other practical advantages in a crop breeding context are that their presence or absence is not affected by the environment and that can be detected at any stage of plant growth.
The term marker-assisted breeding (MAB) is used to describe several modern breeding strategies that are based on using the molecular marker profiling determined in segregant progenies or populations as the criterion to select the superior individuals for the trait of interest. The different MAB approaches can be classified into two main groups, marker-assisted selection (MAS) and genomic selection (GS), which mainly differ in the number of markers used for a genotype-based selection.

2.1. Marker-Assisted Selection

Marker-assisted selection, or MAS, is the most fruitful practical application of biotechnological tools for crop breeding to date. To be useful for selection, a marker must be reliably associated to the phenotype of interest. An ideal DNA marker is located within the sequence of the gene controlling the target trait, allowing the discrimination of favorable from unfavorable alleles. However, it requires previous structural and functional gene characterization, which is frequently lacking, especially regarding polygenic quantitative traits. Nevertheless, a molecular marker closely linked to the target gene or quantitative trait locus (QTL) may serve to conduct a preliminary selection of promising genotypes in large segregating populations or at early plant developmental stages, thus making MAS a cost-saving strategy.
Some pre-requisites must be fulfilled for the successful implementation of a marker-assisted breeding strategy [27]:
(a)
A suitable marker system. The relatively recent development of high-throughput Next Generation Sequencing (NGS) platforms has revolutionized the field, enabling and generalizing the use of single nucleotide polymorphisms (SNPs) as molecular markers. Earlier marker systems not based on NGS methods had several disadvantages related to expensiveness, time-consumption, or reproducibility.
(b)
The development of genetic or physical maps, where the marker–trait associations can be contextualized at a genome level and the most suited markers can be chosen for MAS. High-density genetic linkage maps, based on the segregation of markers and genes in experimental populations, have been built in most economically important plant species for MAS applications. However, because crop genome sequences are available, fine physical maps have become a popular alternative, mainly because of their faster development and almost unlimited resolution (i.e., at the base-pair level).
(c)
The identification of marker–trait associations. As mentioned, the genetic linkage between the trait of interest and the marker is a key aspect for marker-assisted breeding. The success in breeding programs can only be guaranteed if markers are tightly linked to the genes or QTLs, or closely associated with the target traits.
The most popular methodology to identify genes or genomic regions (QTLs) controlling target traits is referred to as “gene mapping” or “QTL mapping” (Figure 2). These mapping analyses are based on the identification of molecular markers that co-segregate with the trait of interest in an experimental population derived from a bi-parental cross. Different types of populations can be used, such as F2 populations, double-haploid (DH) populations, backcross, or recombinant inbred line (RIL) families. However, these populations present certain drawbacks; firstly, because the mapping resolution depends on the amount of recombination events happening during the population generation [28], which is usually limited; and secondly, because only those loci showing allelic variation between the parental lines can be analyzed. These limitations can be partially overcome by using multiparent RILs (as MAGIC populations: Multiparent Advanced Generation InterCross), which increases both the allelic variability considered for analysis and the number of recombination events occurring during population development [29].
At the beginning of the 21st century, Genome-Wide Association Studies (GWAS) emerged as a powerful complementary mapping tool [30,31]. The GWAS approaches test the statistical association between the markers, usually SNPs, and the phenotype of a target trait scored in large panels of distantly related individuals (i.e., natural populations) or lines (i.e., wide germplasm collections including landraces or cultivar representatives). Compared to the family-based traditional mapping methods, GWAS not only hugely increases the allelic diversity covered by the analysis, but also improves the map resolution by taking into account all genetic recombinations that historically occurred in the generation of the (unrelated) genotypes composing the mapping panel [32]. Nevertheless, the genetic structure of the mapping population must be taken into account to avoid false, inaccurate associations [33]. An additional limitation of GWAS is that most of the association methods only consider bi-allelic variants and the underlying function of multi-allelic sites, which if not properly handled, may be ignored [34,35]. Efforts are being made in order to address this drawback [36]. It can be noted that the GWAS approaches have benefited from the parallel implementation of high-throughput sequencing technologies, such as Genotyping by Sequencing (GBS), that allow for the rapid and cost-effective genotyping of thousands of molecular markers [37].
Once a statistically significant association between a marker (or a group of markers) and a crop trait is demonstrated, it is assumed that the marker is located within or close to a genic sequence causative of the phenotype of interest. This is usually the first step to clone and functionally characterize genes. However, as already mentioned, even with limited or no knowledge of the biological function of the marked gene, the marker can be successfully applied for MAS.
Some economically important crop characteristics are controlled by one single gene or a few genes having a major effect on the plant phenotype. That is the case, for instance, of many resistances to pests/diseases, male sterility, self-incompatibility, and certain characteristic related to plant morphology or quality. For these, namely qualitative traits, the mapping approaches usually yield reliable markers and MAS designs are currently common in breeding programs. But many relevant traits, of quantitative inheritance, have a complex genetic architecture as being controlled by multiple genes or QTLs with small individual effects on the phenotypic variation. This holds for yield components as well as for most morphological and physiological plant characteristics determining crop responses to abiotic limitations. Theoretically, all the QTLs contributing to the quantitative trait of interest could be taken into account in an MAS-based breeding scheme. However, the main obstacle remains to be the lack of efficient, reliable markers. For polygenic traits, non-genotypic factors usually have a great influence on the plant phenotype, with QTL × E and multiple epistatic interactions also being possible. This seriously hinders the validation of putative causative associations between markers or mapped QTLs and trait phenotypes [27].
When the association between genes/QTLs and traits has been established, marker-assisted backcrossing (MABC) is regarded as the simplest form of MAB. MABC facilitates the transfer of one or a few favorable genes or QTLs from one individual used as the genetic source (the donor parent, that may be agronomically unsuitable) into the elite breeding line or superior cultivar to be improved for the associated trait (the recurrent parent). Unlike traditional backcrossing, based on phenotypic presentation of the trait of interest, MABC is focused on the presence/absence of alleles of the markers linked with the target favorable gene(s)/QTL(s). MAS can also aid the concurrent selection of the elite parent alleles, representing a clearly time-saving strategy [38]. Marker-assisted gene pyramiding is another useful MAS application that allows the breeding of a novel cultivar by selecting gene/QTL-allele-linked markers of different donors mediating multiple-parent crossing or complex crossing, back-crossing, and recurrent selection. It may be a successful strategy to enhance a quantitatively inherited trait [27]. One of the most problematic issues in crossing breeding programs is linkage drag, which refers to the reduction in quality, yield, or other elite traits that can result from the introduction of deleterious genes or QTLs of a donor parent together with the favorable gene. It must be noted that MAS also offers a tool for monitoring undesirable genes in the offspring.

2.2. Genomic Selection

The sequencing of the whole genome and the development of powerful bioinformatic tools for associating characters and molecular markers has led to the design of new, more ambitious tools in order to assist in breeding programs.
Genomic selection (GS) or genome-wide selection (GWS) is a methodology based on MAS but characterized by the simultaneous selection for numerous markers (until hundreds of thousands) that cover the entire genome. Thanks to the high density of markers distributed along the genome, all genes/QTLs are expected to be in linkage disequilibrium with at least some of the markers [39]. GS programs consist of two distinct, consecutive, phases [40]. During the first phase, the effect on the phenotype of allelic variation at multiple loci is estimated and a “genomic estimated breeding value” (GEBV) formula is developed. For this, phenotypes and genome-wide genotypes of a reference population (called the training population) are analyzed, and significant associations between phenotypes and genotypes are prognosticated by means of statistical analyses. In the second phase, the selection of desirable individuals within a breeding population is carried out by firstly determining the genome-wide molecular marker profiles of candidate individuals, and then using the GEBV equation previously obtained. Crop breeding has already benefited from GS research (see Section 2.3. [40]), which is clearly more intense in wheat and maize than in rice, among the main crops (Figure 2). Some currently addressed issues (for example, the approximation of non-additive genetic effects and the combination of multiple traits or environments) will be key for improving the accuracy of GS predictions [41].

2.3. Marker-Assisted Breeding Oriented to Crop Improvement for Abiotic Limitations’ Response

The timing of application of the main MAB approaches to drought tolerance (sensu lato, also including water-use efficiency and related topics), heat tolerance, and NUE in rice, wheat, and maize is showed in Table 1. For temporal contextualization, the year of publication of the first article where QTL mapping, GWAS and GS methodologies have been realized (not merely mentioned) on any breeding topic are also showed. It supports that water stress has usually been the first research goal among abiotic limitations but that, despite their main relevance, the topics focused here have not always been among those initially addressed. Among crops, maize improvement received an earlier interest for the application of QTL mapping and GS approaches, despite the fact that the overall research efforts have been lesser than in wheat according to bibliometrics (Figure 2).
Some features of QTL mapping, GWAS and GS in each of these crops and technical achievements regarding sustainability-related traits are described in the following sub-sections.

2.3.1. Rice

Rice is the most drought-sensitive cereal. Thus, improvement of water-use efficiency and tolerance to water deficits are the main rice breeding goals. Two of the earliest reports on QTL mapping in rice did indeed deal with the identification of genome regions involved in traits related to plant responses to drought, i.e., root morphology and osmotic adjustment [76]. The long list of crop sustainability related traits for which QTLs or genes have been mapped in rice is headed, firstly, by traits improving plant response to drought, and secondly, by heat tolerance and traits related to nitrogen-use efficiency NUE (Figure 3) [63,77,78,79,80]. However, it also includes morphological or physiological characteristics related to salinity tolerance [81,82,83]; photosynthetic efficiency [84,85]; phosphorous and potassium use efficiency [86,87]; tolerance to deficiency in other essential nutrients [88]; or plant growth under aluminium stress [89], among others. The contemporary application of molecular markers to accelerate breeding programs is not a potential advantage of promising modern approaches, but a current selection strategy in rice, also for sustainability-related traits [90,91,92].
GWAS is relatively underused for the genetic dissection of rice traits compared to the predominance of this crop over wheat and maize, regarding the application of QTL mapping (Figure 2) and most of the modern analytical tools (see below). This is particularly remarkable considering that association studies are facilitated when dense genetic or physical maps are available in the species, and that, as noted above, the rice genome was well assembled years in advance of the maize and the wheat genomes. A possible explanation can be related to the easier construction of genetic maps in diploids, compared to polyploid species [93]. It may make GWAS comparatively less advantageous than other mapping tools in rice than in crops with more complex genomes such as maize (an ancient polyploid) and, specially, wheat. Nevertheless, several studies have taken advantage of its applicability for the easy identification of novel QTLs involved in plant response to environmental constraints by exploring wide rice germplasm collections of diverse geographical origins [73,94]. As noted for GWAS, genome selection will likely have greater impacts as breeding assisting tools in crops with more complex genomes (Figure 2 and Figure 3), but its predicting ability for developing rice varieties adapted to dryer environments has also been demonstrated [69,95].

2.3.2. Wheat

The first studies to determine the genome location of QTLs in wheat analyzed the association between the phenotype for the quantitative trait of interest and morphological or isozyme traits (i.e., polymorphisms for qualitatively-inherited genes) that had been already mapped by using the Sears’ cytogenetic stocks [43] (Table 2). These and other cytogenetic stocks derived of cv. Chinese Spring were also key materials to map the molecular markers (i.e., polymorphisms for DNA sequences) that were subsequently employed for QTL mapping by means of genetic linkage analyses [46,96]. In a comprehensive review, Gupta and coworkers have compiled the many QTLs related to agronomic, physiological, and root-related traits that have been identified under water, heat, and salinity stresses in wheat identified to date [97]. The mapping of wheat genome regions related to nutrient-use efficiency has received less attention (Figure 3). Nevertheless, QTLs with a significant effect, not only on biomass, yield or grain protein content, but also on N uptake and N utilization efficiency, have been detected under different nitrogen regimes [98,99].
Compared to rice and maize, the emergence of GWAS is relevant as the alternative to the classical biparental-population mapping (Figure 2), especially for the location of drought and heat tolerance-related QTLs in the complex wheat genome (Figure 3) [100]. The much greater variability detected at the genomic level in local landraces compared to commercial wheat cultivars makes this approach the most useful for the identification of novel alleles involved in abiotic stress tolerance and adaptation to low-input management practices [101]. Once an accurate and complete wheat genome sequence has been released, the availability of increasingly more reliable high-density SNP arrays [102] will prompt mass genotyping for GWAS approaches.
Wheat is likely the crop for which more studies confirming the applicability of molecular markers to assist selection programs are available. However, most of the documented examples refer to markers linked to breadmaking quality traits and pathogen resistances [103,104]. Breeding for complex polygenic traits such as tolerance to water stress or other abiotic constrains is still conducted by the direct selection of better performing plants in limiting conditions [105]. Nevertheless, the panel of new tools implemented after the wheat genome has been fully sequenced has boosted the design of predictive models and novel genome selection-based breeding strategies (Figure 2 and Figure 3), some of which are being tested in several international programs. A worldwide leading institution, CIMMYT (International Maize and Wheat Improvement Center), is following a genome selection strategy for wheat quality breeding [106,107]. Promising results have already been obtained in the scope of durum wheat adaptations to drought environments by ICARDA (International Center for Agricultural Research in the Dry Areas) [108] as well as for wheat N-use efficiency improvement [74].

2.3.3. Maize

Maize was the earliest main crop where QTL mapping studies were conducted [44] (Table 1). Since then, genetic loci associated to traits involved in plant response under environmental constraints such as drought, heat, salinity, and waterlogging, or under nitrogen deficit, have been identified by using either biparental populations or panels of maize lines encompassing wider genotypic variability [53,62,109,110,111,112]. The bibliometric analysis evidences a higher relevance of NUE and related traits in QTL mapping studies conducted in maize compared with rice and wheat (Figure 3).
Despite a lower abundance of research documents, maize has been pioneering in the breeding-oriented application of molecular markers, which is surely related to the higher profitability of biotech investments for hybrid cultivar improvement. So, as early as in 1992, the usefulness of marker-assisted selection of agronomic and yield-related quantitative traits was demonstrated in maize [113]. For context, the early uses of MAS in rice and wheat were not only later, but associated to monogenic traits or based on markers linked to alleles of known major effects on the traits of interest [114,115]. Maize was also the first, among these main crops, where experimental data were used to test the accuracy of breeding-by-design models [68] (Table 1). Since then, the ability of a number of genomic prediction indexes to select superior maize lines in varied stressed environments has been documented [71,72,75].

3. Gene Expression Analysis

3.1. Gene Expression Analysis Approaches

For decades, gene expression studies were almost exclusively based on the identification of individual proteins by electrophoretic methods or enzymatic activity assays [116,117,118]. However, over the past 20 years, the field of gene expression profiling has undergone a dramatic revolution with the incorporation of analytical tools that can provide a dynamic picture of the complete set of transcripts in specific tissues of an organism. The transcriptome includes the messenger RNA (mRNA) and non-coding RNA (ncRNA) molecules. Unlike the genome, which is nearly static for each organism, the transcriptome is active and dynamic, and can change in a specific cell, tissue, or organ according to the developmental stage or in response to external stimuli. Thus, transcriptomics is considered to be the major large-scale platform for studying the biological functioning of a living organism. By analyzing the transcriptome, researchers can determine which sets of genes are turned on or off in a particular condition and can quantify the changes in gene expression among different biological contexts [119].
Numerous molecular biology techniques have been used for transcription quantification and expression profiling [120]. The first of these techniques (Northern blotting, in situ hybridization and Reverse transcription polymerase chain reaction (RT-PCR)) served to determine the presence of single or few transcripts in a qualitative way [121,122]. A posterior technical improvement that allows the detection of mRNA present even at low levels (real-time RT-PCR or Quantitative PCR (qPCR)) has been the most widely used for absolute and relative gene expression quantification [123,124]. The expression of thousands of known genes can be studied simultaneously since the development of microarrays, which consist of a collection of specific-DNA spots attached to a solid surface that are hybridized with a cDNA or cRNA sample [125]. This methodology has provided a powerful tool for monitoring global changes of gene activity, needed to understand key molecular pathways involved in plant development and in physiological responses to external constraints [126,127]. More recently, NGS-based RNA sequencing (RNA-seq) has entailed another revolution in gene expression analysis [128]. RNA-seq uses NGS to determine the presence and quantity of RNA in a biological sample at a given moment, allowing a very sensitive monitoring of the changes in the cellular transcriptome during development processes or under biotic or abiotic stresses. In addition to mRNA transcripts, RNA-seq can analyze different populations of RNA including lncRNA (long ncRNA) and small RNA, such as micro-RNA, transfer RNA, and ribosomal RNA. The advent of high-throughput RNA sequencing technologies has also made it possible to map transcripts onto the genome for studying the structure of genes and allele variants, alternative splicing patterns, or post-transcriptional modifications [4,128].
All these molecular techniques have been extensively used for gene expression profiling in crop plants, resulting in the identification of key genes and pathways regulating traits of breeding interest. Particularly remarkable is their ability to identify genes coding for transcription factors (TFs), regulatory proteins that control the turning on/off of multiple downstream genes which are coordinately regulated by external or developmental signals [129]. TF genes are assumed to be the best breeding target for the successful improvement of polygenic crop traits [130].
Figure 4 shows the number of research documents referring to transcriptomics and the main modern methodologies for gene expression analysis (i.e., qPCR and RNA-seq) in rice, wheat, and maize. Like for most other molecular breeding tools, rice is the crop where these approaches have been more extensively used, but the distance with wheat and maize has narrowed considerably regarding the most recent technical incorporations.

3.2. Gene Expression Analysis Oriented to Crop Improvement for Abiotic Limitations’ Response

Table 2 shows the timing of application of molecular approaches based on mRNA expression to the functional analysis of genes or QTLs putatively involved in drought tolerance, heat tolerance and NUE in rice, wheat, and maize. For temporal contextualization, the year of publication of the first articles which documented gene expression analyses are showed, in addition to the year of publication of the first reports describing a real use of these novel methodologies to any breeding topic.
It is worth noting the lack of NUE-related gene expression studies in rice and wheat prior to the implementation of transcriptomic approaches, although no other temporal biases can be clearly established either for focused topics or crops. Some usage particularities and technical achievements regarding sustainability-related traits in rice, wheat and maize are described in the following.

3.2.1. Rice

As already noted, global expression studies using transcriptomic approaches, combined with bioinformatics, can serve to identify novel genes putatively involved in favorable responses. An early version of this strategy was applied to the genetic dissection of drought tolerance in rice as early as in 1988 [131], more than 10 years before the term “transcriptomics” was first used in plant studies [173] (Table 2). Those “pre-omic”-era studies were based on mRNA in vitro translation or protein extraction and sequencing for further comparison with the existing protein databases available. Interestingly, the conceptually equivalent evolved approach, RNA-seq, has demonstrated the significant effect of water-by-nitrogen interactions at the gene expression level [174]. It is also worth noting that several among the first rice studies where cDNA microarrays were employed to demonstrate differences in transcriptional profiles between genotypes or treatments allowed the identification of genes that were up- or down-regulated in abiotic stress conditions, different from those reflected in Table 2 such as salinity and nutritional starvation [175,176]. The use of qPCR for quantifying the differential expression of specific rice genes involved in plant response to major abiotic constrains is also documented (Figure 5 and Table 2). This is another example of a molecular analytical tool whose first report in rice was oriented to the study the functional genetic basis underlying the plant response to an abiotic limitation, i.e., phosphorous starvation [151].

3.2.2. Wheat

In wheat, the differential expression of genes depending on external conditions was initially studied only for Mendelian-inherited (monogenic) traits, and deduced from the associate phenotypic change [117]. By the early 1980s, the up-/down-regulation of specific wheat genes began to be determined by in vitro translation or cloning, as cDNA, of messenger RNAs extracted from treated cells [177]. Dozens of studies have used transcriptomic approaches to demonstrate the active role of TFs and genes involved in cellular defense mechanisms, mainly against oxidative or osmotic stress, in the wheat plant response to abiotic limiting conditions [178,179,180,181]. The transcriptomic studies dealing with the functional characterization of wheat genes presumably involved in the response to low N input are significantly scarce (Figure 5). However, the differential tissue-specific expression of several genes related to N assimilation, N and C metabolism, or coding transcription factors, has been demonstrated and quantified in durum wheat cultivar Svevo [182].

3.2.3. Maize

The first documented reports of changes in mRNA synthesis were conducted before the transcriptomic era and corresponded to plastid but not nuclear maize genes [183,184]. Since then, differential gene expression analyses at the transcriptome level have helped to identify many maize genes that are up- or down-regulated by environmental stresses [145,148,185], or by nitrogen levels [186,187]. Studies combining QTL mapping with the characterization of transcripts that are differentially expressed under abiotic stressed conditions have been especially successful in identifying gene targets for the breeding of maize drought-tolerant cultivars [188,189]. The role of genes involved in C and N metabolism in plant drought response has also been evidenced by transcriptomic approaches [190]. Among the reports addressing the genetic basis of NUE in maize, it is worth noting the study of Yang et al., who determined that around 7% of the maize transcriptome is nitrogen-responsive [186]. These authors have developed a set of gene biomarkers whose expression profiles can be a cost-effective tool to monitor the nitrogen status of maize plants growing under field conditions. On the other hand, and as noted for rice, transcriptomic analyses have demonstrated that simultaneous water and nitrogen stresses have a greater impact on gene function than those expected according to the individual effects of these input limitations [191]. As for QTL mapping, it is again remarkable the relative relevance of gene expression studies addressing NUE and related traits in maize compared to rice and wheat, crops where thermotolerance occupies a clear second position in the research interest after drought tolerance (Figure 5).

4. Genetic Modification

The third main group of breeding assistant tools includes a number of techniques allowing that crop improvement can be based on genes or alleles that do not naturally exist within the crop species or within its sexually compatible relatives. These tools are supplementary to plant breeding and always need to be coupled with classical breeding program for evolving varieties of commercial value. It must be noted that some of them (i.e., transgenesis and gene editing) are not only employed to increase genetic variability but for gene function analyses that characterize how a specific change in the coding or regulatory elements of a gene alters the plant phenotype.

4.1. Mutagenesis

Mutagenesis cannot be strictly considered as a modern crop breeding tool. Chemical or radiation treatment of plants has been used since the second half of the 20th century to generate mutations that could give rise to interesting new alleles [192,193]. In fact, these techniques have been highly successful, with more than 3200 varieties of around 200 species having been obtained by mutagenesis over the last 70 years (IAEA/FAO database, https://mvd.iaea.org/, (accessed on 30 December 2020)). The type of mutations introduced by these techniques does not essentially differ from spontaneous mutations. A high percentage of induced mutations are simple nucleotide changes (single nucleotide polymorphisms, SNPs), although more extensive gene or chromosomal rearrangements may also occur [193]. All of these modifications occur essentially randomly in the genome, and the task of genetic breeding is to identify those responsible for novel variability of interest and eventually transfer them to yielding varieties. With this in mind, numerous mutant collections have been built in different crop species. A genomic approach, termed TILLING (Targeting Induced Local Lesions IN Genomes), that combines chemical mutagenesis with a sensitive mutation detection PCR-based strategy, has been the most successful to develop and to screen mutant populations for allelic variants in target genes [193,194,195].

4.2. Transgenesis

Transgenic plants are plants into which one or more genes (transgenes) have been artificially inserted using genetic engineering techniques. Usually, the entire transgene or some of its components (i.e., promoter, coding region) come from an unrelated plant or from an organism belonging to a completely different kingdom, including bacteria and viruses. However, in the approach referred as “cisgenesis” a gene from the same or a close species is first isolated and engineered, and then introduced back into the recipient organism [196,197]. Many distinct protocols have been developed since the early experiments conducted in the 1980s, but most genetically modified plants are obtained by either of two transformation methods: the biolistic approach (particle gun method) or mediated by Agrobacterium tumefaciens [198].
The main breeding-applied purpose of inserting transgenes in a crop plant is to introduce a trait which does not occur naturally in the species, in order to improve its agronomic performance or product quality. Since 1996, the process has materialized in more than 500 authorized genetically modified (GM) events, some of which are currently cultivated in a global area of around 200 million hectares [199]. Most GM cultivars correspond to maize, soybean, cotton and canola, and express herbicide tolerance and/or insect resistance; however, other crops and traits successfully engineered have also reached the market (see updates in the database of the International Service for the Acquisition of Agri-biotech Applications, ISAAA, at www.isaaa.org (accessed on 22 December 2020)).
However, transgenesis may also be used to test the presumed function of a crop gene by genetically manipulating either its coding region or the non-coding sequences that control the level and pattern, either spatial or temporal, of its expression (i.e., promoters, enhancers). These kinds of experiments, frequently conducted in model plants such as Arabidospsis or tobacco, have been essential to confirm the real involvement of mapped genes on crop traits [200,201].
Gene silencing and gene overexpression are the main transgenic strategies used to determine plant gene function. Both approaches aim to deduce the role of a gene by examining the phenotypic changes that are produced when its expression is altered. Gene silencing constructs inactivate the target gene by a phenomenon called RNA interference, by which the expression of antisense RNA (i.e., complementary to the RNAm sequence) or double-stranded RNA stretches triggers the enzymatic degradation of the target mRNA, that cannot be further translated [202]. Gene overexpression, usually achieved by inserting strong promoters upstream of the gene coding region, may be specially interesting when the target gene is functionally redundant with another gene (i.e., when the role of one of them may cover up the function of the other, and vice versa, in gene silencing mutants), or when the silencing of the gene is particularly deleterious or even lethal [203]. In a breeding-applied context, gene silencing has been very successful to improve relevant crop traits, such as increased shelf-life of fruits and flowers, enhanced nutritional quality of oil, or decreased lignin content of cellulosic biomass, among others; meanwhile, overexpression has been used to confer to plant resistance to herbicides or to different biotic stresses.
Figure 6 depicts a remarkable difference between the research documents on transgenic approaches in wheat and maize, which does not hold for any other of the tools analyzed. The likely reasons for this peculiarity are considered below.

4.3. Gene Editing

Genome editing technologies address one of the most demanding challenges of current plant biotechnology: the custom design of new crop varieties. By using this set of tools, the repression or activation of gene expression, modifications of gene function, or the creation of gene knockouts can be performed, mediating the targeted manipulation of DNA sequences [204].
Gene editing requires that a nuclease enzyme introduces a double-stranded DNA break (DSB) at the site where a given gene wants to be edited. Afterwards, the end-joining or recombination pathways can repair the induced DSBs introducing simultaneously the desired modifications at the target locus [205]. However, off-target cleavage events and/or erroneous restoring of the physical integrity of the chromosomal DNA are far from negligible, and many target editing attempts provoke undesired mutations, even deleterious, or gross chromosomal rearrangements [206]. As in the other genetic modification techniques, screening procedures must be followed subsequently to identify plants that have been edited at the intended locus.
Until the generalization of editing systems based on the clustered, regularly interspaced, short palindromic repeats (CRISPR)-associated nuclease 9 (Cas9) [207], different classes of endonucleases have been used to design genomic editing tools, such as the mega-nucleases, the zinc finger nucleases (ZFNs), and the transcription activator-like effector nucleases (TALENs) [208,209,210]. Basically, these earlier systems consisted of chimeric nucleases with two functional domains: one of the domains, through binding to a specific DNA sequence, provides the specific action site to the nuclease domain, which catalyzes the cleavage of DNA [211]. TALEN-based tools have been the most used among these former gene editing systems. The main reason for their quick and full replacement by CRISPR/Cas-based tools is that TALEN site-specificity is based on the nuclease sequence, and a new protein must be designed for targeting each gene or gene location, while CRISPR/Cas specificity is based on the sequence of a guide RNA, which makes the system simpler and more versatile at a lower cost [204]. It is likely that the limited usage of any these approaches in wheat and maize (Figure 6) can be due to the easier molecular dissection of genes and their functions in rice, which is a pre-requisite for a suitable design of gene manipulation strategies.

4.4. Genetic Modification Oriented to Crop Improvement for Abiotic Limitations’ Response

The timing of application of mutagenesis, transgenesis and gene editing tools in rice, wheat, and maize, oriented to the improvement of drought tolerance, heat tolerance and NUE, is showed in Table 3. Again, for temporal contextualization, the year of publication of the first article where these genetic modification methodologies have been realized on any breeding topic are also showed.
According to the bibliographic and bibliometric analyses performed, and despite its relatively early introduction as breeding-assistant tool in the major crops (Table 3), the use of mutagenesis for the specific breeding goals of interest in the present study has been really limited, with a total of 36, 22, and 8 documents on rice, wheat, and maize, respectively, regarding drought, heat tolerance, or NUE. For that reason, this approach has not been further considered. The time gap between the first use of transgenesis in the major crops and its first application to abiotic limitations, of around 10 years, is the longest among all the molecular breeding approaches considered (Table 1, Table 2 and Table 3). This can be attributed to the mostly polygenic nature of plant responses to environmental conditions. It greatly hinders the structural and functional characterization of single genes with a significant effect on the target trait, always required for implementing genetic modification strategies. Some technical particularities and achievements regarding each of the crops under focus are described in the following sub-sections.

4.4.1. Rice

Most of the earliest studies that followed transgenic approaches for the functional characterization of crop genes actually engineered model plants with constructs carrying coding regions or control elements of the gene of interest. In the case of rice, one of these pioneering applications of transgenesis used transgenic tobacco plants to study the function of a glutamine synthetase rice promoter [200]. Since then, many rice genes putatively conferring tolerance to drought and other limiting plant stresses have been functionally characterized either in heterologous systems (i.e., tobacco, Arabidopsis) and in transgenic rice [240] (Figure 7). Rice, as a model monocot for transformation experiments, has also been engineered to demonstrate the protecting function of alien crop genes against abiotic stresses such as drought and salinity [241,242]. Gene editing has been the latest methodology incorporated to the toolbox for the functional characterization of genes. Its use for plant traits in the scope of this review, is comparatively novel (Table 3) and still scarce (Figure 7); however, a modified CRISPR protocol has already successfully inactivated a rice gene regulating the stomatal density, an important determinant of water use efficiency [243].
In addition to the genetic engineering experiments strictly oriented to basic-research purposes, transgenic rice plants have also been created with the applied goal of developing genetically modified (GM) commercial varieties. However, the current number of rice GM varieties is very limited according to the information in the most complete, cumulative, database of the International Service for the Acquisition of Agri-biotech Applications (ISAAA, www.isaaa.org, accessed on 22 December 2020). There are only eight rice events out of the 526 events ever approved in any country; and all, except the famous “Golden Rice”, carry transgenes related to biotic stresses, i.e., insect resistance and herbicide tolerance. A major reason for the limited representation of rice, compared to maize, among GM varieties will be considered below, but the null presence of abiotic stresses among the target traits is very likely due to the already noted complex genetic basis of the plant physiological response in limiting abiotic conditions. This makes it unlikely that single genes which may have demonstrated significant tolerant effects when checked under controlled experimental conditions can actually improve crop performance in a real field environment [244].

4.4.2. Wheat

Due to the technical problems to transform monocots, the first uses of transgenesis to analyze the function of wheat genes were made by transforming the model dicot tobacco with wheat genes and analyzing the expression of different constructs [245]. A 1989 report transformed rice instead of tobacco, for the first time, to identify functional elements of a wheat gene regulated by the phytohormone abscisic acid [246]. After several transforming protocol attempts [223], it was by 1992 when the first transgenic fertile wheat plants were reported, the last among the major crops [247]. In wheat, the transgenic approaches have evidenced a special interest on sustainability-related traits, drought tolerance and heat tolerance, representing 10% and 2% of the total documents, respectively; meanwhile, the corresponding values are 7.4% and 1% in rice, and 3.3% and 0,4% in maize (see data in Figure 6 and Figure 7). Most genes have been engineered with the final aim of developing abiotic stress-tolerant wheat cultivars code transcription factors, i.e., regulatory proteins that control and coordinate the expression of multiple genes in response to external stimuli [248,249]. This is the more promising strategy due to the complex genetic basis governing the different physiological pathways that underlie crop performance in real field conditions that, in the case of wheat, is additionally hindered by the existence of three homoeologous copies of many functional genes. Regarding the improvement of N use efficiency, the successful introduction into a winter wheat variety of a favorable allele of the wheat GS2 gene, coding the plastidic isoform of the glutamine synthetase enzyme has been reported [250]. Whether any of these developments will materialize in the next GM wheat cultivars cannot be predicted, but the history of GM wheat approvals is even more unsuccessful than that of rice, with only a single herbicide-tolerant event in the GM database of the ISAAA (https://www.isaaa.org/gmapprovaldatabase/; accessed on 22 December 2020). TALEN and CRISPR systems have been used for gene editing of a number of wheat genes since a pioneering work on cell suspensions conducted in 2013 [234]. However, the induction of targeted mutations on genes involved in the traits under focus is not yet documented in wheat (Table 3), despite a few reports having claimed the suitability of this approach for addressing the creation of drought-tolerant cultivars (Figure 7).

4.4.3. Maize

As described for rice and wheat, the first uses of transgenesis for functional analyses of maize genes were made by transforming dicots (petunia and tobacco) with maize genes and analyzing the expression of different genic constructs. Among these early reports, attempts were made to dissect the molecular basis of plant response to an abiotic constraint such as high temperature [251]. Currently, maize is by far the crop most represented among the GM cultivars, with 283 authorized events in the GM database of the ISAAA (https://www.isaaa.org/gmapprovaldatabase/; accessed on 22 December 2020), which may explain the particular abundance of transgenic research in maize (Figure 6). The greater interest of breeding companies and institutions to develop GM-maize but not GM-rice or GM-wheat varieties can be ultimately attributed to the already mentioned predominance of hybrid elite cultivars in maize. Herbicide tolerance and insect resistance are the commercial traits more frequent in maize GM cultivars, with almost 90% of authorized events carrying transgenes that protect plants against one or both of these biotic stresses. A number of engineered maize lines have been reported to perform with enhanced tolerance to drought, heat, salinity, or NUE [227,232,252,253,254,255], since the first GM maize patent related to abiotic limitations was issued in 2000 [254]. However, the interest in these topics seems relatively low (comparing data for maize in Figure 6 and Figure 7) and, to date, there is a single genetic engineered construction that has been successful to develop GM maize commercial cultivars with improved abiotic stress tolerance. That construction carries a bacterial gene coding for an RNA chaperone protein whose expression has been demonstrated to maintain cellular functions under water stress conditions [256]. This transgene is currently present in seven authorized GM events, most of which carry additional constructions for herbicide tolerance and/or insect resistance. Successful attempts based on targeted mutagenesis approaches (TALENS or CRISPR systems) have also been reported either for drought tolerance or NUE [129,257] (Table 3), but are still limited compared to the tool usage in rice (Figure 7).

5. Future Prospects

The improvement of crop plant response to the main environment constraints and nutrient limitations represents a major breeding goal to assure food security in the future, particularly under the threats of the changing climate and the expanding population [258]. The low success in creating abiotic stress-tolerant cultivars, despite the great number of genes for which a protective function has been demonstrated in controlled conditions, has been considered in earlier reviews [244,259]. Particularly, the actual reliability of phenotype data if real field conditions are replaced by high-throughput phenotyping platforms has been pointed out as the bottleneck for a confident application of the many localized markers [260]. Wide germplasm collections, including accessions adapted to unfavorable environments or low-input cultivation management (i.e., landraces, and local varieties), are likely the haystack where the needles will be found [15,101,261]. The current ability to cheaply sequence thousands of materials together with the growing development of bioinformatics tools will surely provide breeders with finer tuning of GS strategies [262], acknowledged as the most promising approach for crop improvement for traits that involve complex plant physiological responses [40]. As far as genetic modification is concerned, the current list of engineered sustainability-related traits in commercial varieties is almost limited to herbicide tolerance and insect resistance. It seems very unlike that gene editing tools can substantially change this bias, mostly due to the polygenic control of crop performance in abiotic stress conditions. However, it is worth mentioning that innovative biotechnological tools are being developed to create cereal plants capable of fixing atmospheric N. This goal is being addressed by the Gates Foundation with support to two distinct long-term appealing approaches: the transference of bacterial genes involved in N fixation [263]; and the promotion of symbiotic interactions with N-fixing microorganisms [264]. Genetic engineering of RuBisCo and other components of the carbon fixation metabolism are also being targeted to improve plant productivity [201]. Furthermore, a collaborative project aims to develop C4 rice by engineering the photosynthetic machinery of rice plants to include functional components of the C4-type plant pathway (https://c4rice.com/ (accessed on accessed on 22 December 2020)). It is thus feasible that GM cereal varieties with fewer fertilizer requirements and improved photosynthetic efficiency may be released in the ongoing decade [265].

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/2073-4395/11/2/376/s1, Table S1: Methodology followed for the bibliometric and bibliographic analyses.

Author Contributions

Conceptualization, E.B. and E.G.; methodology, E.B.; formal analysis, E.G.; writing—original draft preparation, E.B. and E.G.; writing—review and editing, E.B. and E.G. All authors have read and agreed to the published version of the manuscript.

Funding

The authors were funded by the Spanish Ministry of Science and Innovation (Grant No. PID2019-109089RB-C32), and by Comunidad de Madrid (Spain) and Structural EU Funds 2014-2020 (ERDF and ESF) (Grant No. AGRISOST-CM S2018/BAA-4330).

Acknowledgments

The authors at grateful to P. Giraldo and L. Pascual for their helpful comments to a former version of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gross, B.L.; Olsen, K.M. Genetic perspectives on crop domestication. Trends Plant Sci. 2010, 15, 529–537. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Langridge, P. Innovation in breeding and biotechnology. In Agriculture and Food Systems to 2050. Global Trends, Challenges and Opportunities; Serraj, R., Pingali, P.L., Eds.; World Scientific Publishing Co. Pte. Ltd.: Singapore, 2019; pp. 245–284. [Google Scholar] [CrossRef]
  3. Acquaah, G. Principles of Plant Genetics and Breeding; John Wiley & Sons: Malden, MA, USA, 2009. [Google Scholar]
  4. Nejat, N.; Ramalingam, A.; Mantri, N. Advances in transcriptomics of plants. In Plant Genetics and Molecular Biology. Advances in Biochemical Engineering/Biotechnology; Varshney, R., Pandey, M., Chitikineni, A., Eds.; Springer: Berlin/Heidelberg, Germany, 2018; Volume 164, pp. 161–185. [Google Scholar]
  5. Kumpatla, S.P.; Buyyarapu, R.; Abdurakhmonov, I.Y.; Mammadov, J.A. Genomics-assisted plant breeding in the 21st century: Technological advances and progress. In Plant Breeding; Abdurakhmonov, I., Ed.; Intechopen: London, UK, 2012; pp. 131–184. Available online: https://www.intechopen.com/books/plant-breeding/genomics-assisted-plant-breeding-in-the-21st-century-technological-advances-and-progresspp (accessed on 8 October 2020).
  6. Awika, J.M. Major Cereal Grains Production and Use around the World. In ACS Symposium Series; American Chemical Society (ACS): Washington, DC, USA, 2011; pp. 1–13. [Google Scholar] [CrossRef]
  7. Yu, J.; Hu, S.; Wang, J.; Wong, G.K.-S.; Li, S.; Liu, B.; Deng, Y.; Dai, L.; Zhou, Y.; Zhang, X.; et al. A Draft Sequence of the Rice Genome (Oryza sativa L. ssp. indica). Science 2002, 296, 79–92. [Google Scholar] [CrossRef]
  8. Jackson, S.A. Rice: The First Crop Genome. Rice 2016, 9, 1–3. [Google Scholar] [CrossRef] [Green Version]
  9. Shimamoto, K. Genetic manipulation of rice: From protoplasts to transgenic plants. Jpn. J. Genet. 1992, 67, 273–290. [Google Scholar] [CrossRef] [Green Version]
  10. Helmy, M.; Tomita, M.; Ishihama, Y. OryzaPG-DB: Rice proteome database based on shotgun proteogenomics. BMC Plant Biol. 2011, 11, 1–9. [Google Scholar] [CrossRef] [Green Version]
  11. McCouch, S.R.; Zhao, K.; Wright, M.; Tung, C.-W.; Ebana, K.; Thomson, M.; Reynolds, A.; Wang, D.; Declerck, G.; Ali, L.; et al. Development of genome-wide SNP assays for rice. Breed. Sci. 2010, 60, 524–535. [Google Scholar] [CrossRef] [Green Version]
  12. Zhao, K.; Tung, C.-W.; Eizenga, G.C.; Wright, M.H.; Ali, M.L.; Price, A.H.; Norton, G.J.; Islam, M.R.; Reynolds, A.R.; Mezey, J.G.; et al. Genome-wide association mapping reveals a rich genetic architecture of complex traits in Oryza sativa. Nat. Commun. 2011, 2, 467. [Google Scholar] [CrossRef] [PubMed]
  13. Viana, V.E.; Pegoraro, C.; Busanello, C.; De Oliveira, A.C. Mutagenesis in Rice: The Basis for Breeding a New Super Plant. Front. Plant Sci. 2019, 10, 1326. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Mohapatra, T.; Robin, S.; Sarla, N.; Sheshasayee, M.; Singh, A.K.; Singh, K.; Mithra, S.V.A.; Sharma, R.P. EMS Induced Mutants of Upland Rice Variety Nagina22: Generation and Characterization. Proc. Indian Natl. Sci. Acad. 2014, 80, 163. [Google Scholar] [CrossRef]
  15. Li, J.-Y.; Wang, J.; Zeigler, R.S. The 3,000 rice genomes project: New opportunities and challenges for future rice research. Gigascience 2014, 3, 8. [Google Scholar] [CrossRef] [Green Version]
  16. Appels, R.; Eversole, K.; Stein, N.; Feuillet, C.; Keller, B.; Rogers, J.; Pozniak, C.J.; Choulet, F.; Distelfeld, A.; Poland, J.; et al. Shifting the limits in wheat research and breeding using a fully annotated reference genome. Science 2018, 361, 7191. [Google Scholar] [CrossRef] [Green Version]
  17. Sears, E. Cytogenetic studies with polyploid species of wheat. I. Chromosomal aberrations in the progeny of a haploid of Triticum vulgare. Genetics 1939, 24, 509–523. [Google Scholar] [CrossRef] [PubMed]
  18. Sears, E.; Miller, T. The history of Chinese Spring wheat. Cereal Res. Commun. 1985, 13, 261–263. [Google Scholar]
  19. De Caleya, R.F.; Hernandez-Lucas, C.; Carbonero, P.; Garcia-Olmedo, F. Gene Expression in Alloploids: Genetic Control of Lipopurothionins in Wheat. Genetics 1976, 83, 687–699. [Google Scholar]
  20. Gupta, P.K.; Vasistha, N.K. Wheat cytogenetics and cytogenomics: The present status. Nucleus 2018, 61, 195–212. [Google Scholar] [CrossRef]
  21. Maccaferri, M.; Harris, N.S.; Twardziok, S.O.; Pasam, R.K.; Gundlach, H.; Spannagl, M.; Ormanbekova, D.; Lux, T.; Prade, V.M.; Milner, S.G.; et al. Durum wheat genome highlights past domestication signatures and future improvement targets. Nat. Genet. 2019, 51, 885–895. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Guan, J.; Garcia, D.F.; Zhou, Y.; Appels, R.; Li, A.; Mao, L. The Battle to Sequence the Bread Wheat Genome: A Tale of the Three Kingdoms. Genom. Proteom. Bioinform. 2020, 18, 221–229. [Google Scholar] [CrossRef]
  23. Schnable, P.S.; Ware, D.; Fulton, R.S.; Stein, J.C.; Wei, F.; Pasternak, S.; Liang, C.; Zhang, J.; Fulton, L.; Graves, T.A.; et al. The B73 Maize Genome: Complexity, Diversity, and Dynamics. Science 2009, 326, 1112–1115. [Google Scholar] [CrossRef] [Green Version]
  24. Helentjaris, T.; Weber, D.; Wright, S. Identification of the genomic locations of duplicate nucleotide sequences in maize by analysis of restriction fragment length polymorphisms. Genetics 1988, 118, 353–363. [Google Scholar] [CrossRef]
  25. Gale, M.D. Plant Comparative Genetics after 10 Years. Science 1998, 282, 656–659. [Google Scholar] [CrossRef] [Green Version]
  26. Fischer, R.; Byerlee, D.; Edmeades, G. Crop Yields and Global Food Security; Australian Centre for International Agricultural Research: Canberra, Australia, 2014. [Google Scholar]
  27. Jiang, G.-L. Molecular markers and marker-assisted breeding in plants. In Plant Breeding from Laboratories to Fields; Andersen, S.B., Ed.; IntechOpen: London, UK, 2013; pp. 45–83. [Google Scholar] [CrossRef] [Green Version]
  28. Mitchell-Olds, T. Complex-trait analysis in plants. Genome Biol. 2010, 11, 1–3. [Google Scholar] [CrossRef]
  29. Liller, C.B.; Walla, A.; Boer, M.P.; Hedley, P.; Macaulay, M.; Effgen, S.; Von Korff, M.; Van Esse, G.W.; Koornneef, M. Fine mapping of a major QTL for awn length in barley using a multiparent mapping population. Theor. Appl. Genet. 2016, 130, 269–281. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Flint-Garcia, S.A.; Thornsberry, J.M.; Buckler, E.S. Structure of Linkage Disequilibrium in Plants. Annu. Rev. Plant Biol. 2003, 54, 357–374. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Korte, A.; Farlow, A. The advantages and limitations of trait analysis with GWAS: A review. Plant Methods 2013, 9, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Alqudah, A.M.; Sallam, A.; Baenziger, P.S.; Börner, A. GWAS: Fast-forwarding gene identification and characterization in temperate Cereals: Lessons from Barley—A review. J. Adv. Res. 2020, 22, 119–135. [Google Scholar] [CrossRef]
  33. Zhao, K.; Aranzana, M.J.; Kim, S.; Lister, C.; Shindo, C.; Tang, C.; Toomajian, C.; Zheng, H.; Dean, C.; Marjoram, P. An Arabidopsis example of association mapping in structured samples. PLoS Genet. 2007, 3, e4. [Google Scholar] [CrossRef] [Green Version]
  34. Purcell, S.; Neale, B.; Todd-Brown, K.; Thomas, L.; Ferreira, M.A.R.; Bender, D.; Maller, J.; Sklar, P.; De Bakker, P.I.W.; Daly, M.J.; et al. PLINK: A Tool Set for Whole-Genome Association and Population-Based Linkage Analyses. Am. J. Hum. Genet. 2007, 81, 559–575. [Google Scholar] [CrossRef] [Green Version]
  35. Chang, C.C.; Chow, C.C.; Tellier, L.C.A.M.; Vattikuti, S.; Purcell, S.M.; Lee, J.J. Second-generation PLINK: Rising to the challenge of larger and richer datasets. GigaScience 2015, 4, 7. [Google Scholar] [CrossRef]
  36. Xiao, Y.; Liu, H.; Wu, L.; Warburton, M.; Yan, J. Genome-wide Association Studies in Maize: Praise and Stargaze. Mol. Plant 2017, 10, 359–374. [Google Scholar] [CrossRef] [Green Version]
  37. He, J.; Zhao, X.; Laroche, A.; Lu, Z.-X.; Liu, H.; Li, Z. Genotyping-by-sequencing (GBS), an ultimate marker-assisted selection (MAS) tool to accelerate plant breeding. Front. Plant Sci. 2014, 5, 484. [Google Scholar] [CrossRef] [Green Version]
  38. Frisch, M. Breeding strategies: Optimum design of marker-assisted backcross programs. In Molecular Marker Systems in Plant Bredding and Crop Improvement; Lorz, H., Wenzel, M., Eds.; Springer: Berlin/Heidelberg, Germany, 2005; Volume 55, pp. 319–334. [Google Scholar]
  39. Meuwissen, T. Genomic selection: Marker assisted selection on a genome wide scale. J. Anim. Breed. Genet. 2007, 124, 321–322. [Google Scholar] [CrossRef] [PubMed]
  40. Voss-Fels, K.P.; Cooper, M.; Hayes, B.J. Accelerating crop genetic gains with genomic selection. Theor. Appl. Genet. 2019, 132, 669–686. [Google Scholar] [CrossRef] [PubMed]
  41. Wang, X.; Xu, Y.; Hu, Z.; Xu, C. Genomic selection methods for crop improvement: Current status and prospects. Crop. J. 2018, 6, 330–340. [Google Scholar] [CrossRef]
  42. Pham, J.-L. Identification of genetic markers for quantitative traits in rice (Oryza sativa L.). Comptes Rendus Acad. Sci. Ser. 3 Sci. Vie 1990, 310, 477–483. [Google Scholar]
  43. Miura, H.; Parker, B.B.; Snape, J.W. The location of major genes and associated quantitative trait loci on chromosome arm 5BL of wheat. Theor. Appl. Genet. 1992, 85, 197–204. [Google Scholar] [CrossRef]
  44. Edwards, M.D.; Stuber, C.W.; Wendel, J.F. Molecular-Marker-Facilitated Investigations of Quantitative-Trait Loci in Maize. I.; Numbers, Genomic Distribution and Types of Gene Action. Genetics 1987, 116, 113–125. [Google Scholar] [CrossRef]
  45. Champoux, M.C.; Wang, G.; Sarkarung, S.; Mackill, D.J.; O’Toole, J.C.; Huang, N.; McCouch, S.R. Locating genes associated with root morphology and drought avoidance in rice via linkage to molecular markers. Theor. Appl. Genet. 1995, 90, 969–981. [Google Scholar] [CrossRef]
  46. Quarrie, S.A.; Gullì, M.; Calestani, C.; Steed, A.; Marmiroli, N. Location of a gene regulating drought-induced abscisic acid production on the long arm of chromosome 5A of wheat. Theor. Appl. Genet. 1994, 89, 794–800. [Google Scholar] [CrossRef] [PubMed]
  47. Lebreton, C.; Lazić-Jančić, V.; Steed, A.; Pekić, S.; Quarrie, S. Identification of QTL for drought responses in maize and their use in testing causal relationships between traits. J. Exp. Bot. 1995, 46, 853–865. [Google Scholar] [CrossRef]
  48. Tripathy, J.N.; Zhang, J.; Robin, S.; Nguyen, T.T.; Nguyen, H.T. QTLs for cell-membrane stability mapped in rice (Oryza sativa L.) under drought stress. Theor. Appl. Genet. 2000, 100, 1197–1202. [Google Scholar] [CrossRef]
  49. Yang, J.; Sears, R.; Gill, B.; Paulsen, G. Quantitative and molecular characterization of heat tolerance in hexaploid wheat. Euphytica 2002, 126, 275–282. [Google Scholar] [CrossRef]
  50. Ottaviano, E.; Gorla, M.S.; Pè, E.; Frova, C. Molecular markers (RFLPs and HSPs) for the genetic dissection of thermotolerance in maize. Theor. Appl. Genet. 1991, 81, 713–719. [Google Scholar] [CrossRef] [PubMed]
  51. Ishimaru, K.; Kobayashi, N.; Ono, K.; Yano, M.; Ohsugi, R. Are contents of Rubisco, soluble protein and nitrogen in flag leaves of rice controlled by the same genetics? J. Exp. Bot. 2001, 52, 1827–1833. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Verma, V.; Foulkes, M.J.; Worland, A.J.; Sylvester-Bradley, R.; Caligari, P.D.S.; Snape, J.W. Mapping quantitative trait loci for flag leaf senescence as a yield determinant in winter wheat under optimal and drought-stressed environments. Euphytica 2004, 135, 255–263. [Google Scholar] [CrossRef]
  53. Agrama, H.; Zakaria, A.; Said, F.; Tuinstra, M. Identification of quantitative trait loci for nitrogen use efficiency in maize. Mol. Breed. 1999, 5, 187–195. [Google Scholar] [CrossRef]
  54. Yang, X.; Jawdy, S.; Tschaplinski, T.J.; Tuskan, G.A. Genome-wide identification of lineage-specific genes in Arabidopsis, Oryza and Populus. Genomics 2009, 93, 473–480. [Google Scholar] [CrossRef] [Green Version]
  55. Yao, J.; Wang, L.; Liu, L.; Zhao, C.; Zheng, Y. Association mapping of agronomic traits on chromosome 2A of wheat. Genetica 2009, 137, 67–75. [Google Scholar] [CrossRef]
  56. Tian, F.; Bradbury, P.J.; Brown, P.J.; Hung, H.; Sun, Q.; Flintgarcia, S.A.; Rocheford, T.R.; McMullen, M.D.; Holland, J.B.; Buckler, E.S. Genome-wide association study of leaf architecture in the maize nested association mapping population. Nat. Genet. 2011, 43, 159–162. [Google Scholar] [CrossRef] [PubMed]
  57. Nuruzzaman, M.; Manimekalai, R.; Sharoni, A.M.; Satoh, K.; Kondoh, H.; Ooka, H.; Kikuchi, S. Genome-wide analysis of NAC transcription factor family in rice. Gene 2010, 465, 30–44. [Google Scholar] [CrossRef]
  58. Edae, E.A.; Byrne, P.F.; Haley, S.D.; Lopes, M.S.; Reynolds, M.P. Genome-wide association mapping of yield and yield com-ponents of spring wheat under contrasting moisture regimes. Theor. Appl. Genet. 2014, 127, 791–807. [Google Scholar] [CrossRef]
  59. Li, L.; Hao, Z.; Li, X.; Xie, C.; Li, M.; Zhang, D.; Weng, J.; Su, Z.; Liang, X.; Zhang, S. An analysis of the polymorphisms in a gene for being involved in drought tolerance in maize. Genetica 2011, 139, 479–487. [Google Scholar] [CrossRef] [PubMed]
  60. Dingkuhn, M.; Pasco, R.; Pasuquin, J.M.; Damo, J.; Soulié, J.-C.; Raboin, L.-M.; Dusserre, J.; Sow, A.; Manneh, B.; Shrestha, S.; et al. Crop-model assisted phenomics and genome-wide association study for climate adaptation of indica rice. 2. Thermal stress and spikelet sterility. J. Exp. Bot. 2017, 68, 4389–4406. [Google Scholar] [CrossRef] [PubMed]
  61. Elbasyoni, I.; Saadalla, M.; Baenziger, S.; Bockelman, H.; Morsy, S. Cell Membrane Stability and Association Mapping for Drought and Heat Tolerance in a Worldwide Wheat Collection. Sustainability 2017, 9, 1606. [Google Scholar] [CrossRef] [Green Version]
  62. Gao, J.; Wang, S.; Zhou, Z.; Wang, S.; Dong, C.; Mu, C.; Song, Y.; Ma, P.; Li, C.; Wang, Z.; et al. Linkage mapping and genome-wide association reveal candidate genes conferring thermotolerance of seed-set in maize. J. Exp. Bot. 2019, 70, 4849–4864. [Google Scholar] [CrossRef] [PubMed]
  63. Tang, W.; Ye, J.; Yao, X.; Zhao, P.; Xuan, W.; Tian, Y.; Zhang, Y.; Xu, S.; An, H.; Chen, G.; et al. Genome-wide associated study identifies NAC42-activated nitrate transporter conferring high nitrogen use efficiency in rice. Nat. Commun. 2019, 10, 5279. [Google Scholar] [CrossRef] [Green Version]
  64. Cormier, F.; Le Gouis, J.; Dubreuil, P.; Lafarge, S.; Praud, S. A genome-wide identification of chromosomal regions determining nitrogen use efficiency components in wheat (Triticum aestivum L.). Theor. Appl. Genet. 2014, 127, 2679–2693. [Google Scholar] [CrossRef] [PubMed]
  65. Morosini, J.S.; Mendonça, L.D.F.; Lyra, D.H.; Galli, G.; Vidotti, M.S.; Fritsche-Neto, R. Association mapping for traits related to nitrogen use efficiency in tropical maize lines under field conditions. Plant Soil 2017, 421, 453–463. [Google Scholar] [CrossRef]
  66. Xu, S.; Zhu, D.; Zhang, Q. Predicting hybrid performance in rice using genomic best linear unbiased prediction. Proc. Natl. Acad. Sci. USA 2014, 111, 12456–12461. [Google Scholar] [CrossRef] [Green Version]
  67. Heffner, E.L.; Jannink, J.L.; Sorrells, M.E. Genomic selection accuracy using multifamily prediction models in a wheat breeding program. Plant Genome 2011, 4, 65–75. [Google Scholar] [CrossRef] [Green Version]
  68. Bernardo, R.; Yu, J. Prospects for Genomewide Selection for Quantitative Traits in Maize. Crop. Sci. 2007, 47, 1082–1090. [Google Scholar] [CrossRef] [Green Version]
  69. Ben Hassen, M.; Bartholomé, J.; Valè, G.; Cao, T.-V.; Ahmadi, N. Genomic Prediction Accounting for Genotype by Environment Interaction Offers an Effective Framework for Breeding Simultaneously for Adaptation to an Abiotic Stress and Performance Under Normal Cropping Conditions in Rice. G3 Genes Genomes Genet. 2018, 8, 2319–2332. [Google Scholar] [CrossRef] [PubMed]
  70. Ly, D.; Huet, S.; Gauffreteau, A.; Rincent, R.; Touzy, G.; Mini, A.; Jannink, J.-L.; Cormier, F.; Paux, E.; Lafarge, S.; et al. Whole-genome prediction of reaction norms to environmental stress in bread wheat (Triticum aestivum L.) by genomic random regression. Field Crop. Res. 2018, 216, 32–41. [Google Scholar] [CrossRef]
  71. Ziyomo, C.; Bernardo, R. Drought Tolerance in Maize: Indirect Selection through Secondary Traits versus Genomewide Selection. Crop. Sci. 2013, 53, 1269–1275. [Google Scholar] [CrossRef] [Green Version]
  72. Van Inghelandt, D.; Frey, F.P.; Ries, D.; Stich, B. QTL mapping and genome-wide prediction of heat tolerance in multiple connected populations of temperate maize. Sci. Rep. 2019, 9, 1–16. [Google Scholar] [CrossRef]
  73. Liu, Z.; Zhu, C.; Jiang, Y.; Tian, Y.; Yu, J.; An, H.; Tang, W.; Sun, J.; Tang, J.; Chen, G. Association mapping and genetic dis-section of nitrogen use efficiency-related traits in rice (Oryza sativa L.). Funct. Integr. Genom. 2016, 16, 323–333. [Google Scholar] [CrossRef]
  74. Michel, S.; Löschenberger, F.; Ametz, C.; Pachler, B.; Sparry, E.; Bürstmayr, H. Simultaneous selection for grain yield and protein content in genomics-assisted wheat breeding. Theor. Appl. Genet. 2019, 132, 1745–1760. [Google Scholar] [CrossRef]
  75. Semagn, K.; Beyene, Y.; Babu, R.; Nair, S.; Gowda, M.; Das, B.; Tarekegne, A.; Mugo, S.; Mahuku, G.; Worku, M.; et al. Quantitative Trait Loci Mapping and Molecular Breeding for Developing Stress Resilient Maize for Sub-Saharan Africa. Crop. Sci. 2015, 55, 1449–1459. [Google Scholar] [CrossRef]
  76. McCough, S.R.; Doerge, R.W. QTL mapping in rice. Trends Genet. 1995, 11, 482–487. [Google Scholar] [CrossRef]
  77. Jagadish, S.V.K.; Cairns, J.; Lafitte, R.; Wheeler, T.R.; Price, A.H.; Craufurd, P.Q. Genetic Analysis of Heat Tolerance at Anthesis in Rice. Crop. Sci. 2010, 50, 1633–1641. [Google Scholar] [CrossRef]
  78. Wei, D.; Cui, K.; Ye, G.; Pan, J.; Xiang, J.; Huang, J.; Nie, L. QTL mapping for nitrogen-use efficiency and nitrogen-deficiency tolerance traits in rice. Plant Soil 2012, 359, 281–295. [Google Scholar] [CrossRef]
  79. Raju, B.R.; Mohankumar, M.V.; Sumanth, K.K.; Rajanna, M.P.; Udayakumar, M.; Prasad, T.G.; Sheshshayee, M.S. Discovery of QTLs for water mining and water use efficiency traits in rice under water-limited condition through association mapping. Mol. Breed. 2016, 36, 35. [Google Scholar] [CrossRef]
  80. Roja, V.; Patil, S.; Deborah, D.A.; Srividhya, A.; Ranjitkumar, N.; Kadambari, G.; Ramanarao, P.V.; Siddiq, E.A.; Vemireddy, L.R. Finding genomic regions and candidate genes governing water use efficiency in rice. Biol. Plant. 2016, 60, 757–766. [Google Scholar] [CrossRef]
  81. An, H.; Liu, K.; Wang, B.; Tian, Y.; Ge, Y.; Zhang, Y.; Tang, W.; Chen, G.; Yu, J.; Wu, W.; et al. Genome-wide association study identifies QTLs conferring salt tolerance in rice. Plant Breed. 2020, 139, 73–82. [Google Scholar] [CrossRef]
  82. Ghomi, K.; Rabiei, B.; Sabouri, H.; Sabouri, A. Mapping QTLs for Traits Related to Salinity Tolerance at Seedling Stage of Rice (Oryza sativa L.): An Agrigenomics Study of an Iranian Rice Population. OMICS J. Integr. Biol. 2013, 17, 242–251. [Google Scholar] [CrossRef] [PubMed]
  83. Bonilla, P.; Dvorak, J.; Mackell, D.; Deal, K.; Gregorio, G. RFLP and SSLP mapping of salinity tolerance genes in chromosome 1 of rice (Oryza sativa L.) using recombinant inbred lines. Philipp. Agric. Sci. 2002, 85, 68–76. [Google Scholar]
  84. Fukuda, A.; Kondo, K.; Ikka, T.; Takai, T.; Tanabata, T.; Yamamoto, T. A novel QTL associated with rice canopy temperature difference affects stomatal conductance and leaf photosynthesis. Breed. Sci. 2018, 68, 305–315. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Takai, T.; Ohsumi, A.; Arai, Y.; Iwasawa, N.; Yano, M.; Yamamoto, T.; Yoshinaga, S.; Kondo, M. QTL Analysis of Leaf Photosynthesis in Rice. Jpn. Agric. Res. Q. JARQ 2013, 47, 227–235. [Google Scholar] [CrossRef] [Green Version]
  86. Vinod, K.K.; Heuer, S. Approaches towards nitrogen- and phosphorus-efficient rice. AoB Plants 2012, 2012, pls028. [Google Scholar] [CrossRef] [Green Version]
  87. Hartley, T.N.; Thomas, A.S.; Maathuis, F.J.M. A role for the OsHKT 2;1 sodium transporter in potassium use efficiency in rice. J. Exp. Bot. 2020, 71, 699–706. [Google Scholar] [CrossRef] [Green Version]
  88. Pariasca-Tanaka, J.; Baertschi, C.; Wissuwa, M. Identification of Loci Through Genome-Wide Association Studies to Improve Tolerance to Sulfur Deficiency in Rice. Front. Plant Sci. 2020, 10, 1668. [Google Scholar] [CrossRef]
  89. Nguyen, V.T.; Burow, M.D.; Le, B.T.; Le, T.D.; Paterson, A.H. Molecular mapping of genes conferring aluminum tolerance in rice (Oryza sativa L.). Theor. Appl. Genet. 2001, 102, 1002–1010. [Google Scholar] [CrossRef]
  90. Singh, R.; Singh, Y.; Xalaxo, S.; Verulkar, S.; Yadav, N.; Singh, S.; Singh, N.; Prasad, K.; Kondayya, K.; Rao, P.R. From QTL to variety-harnessing the benefits of QTLs for drought, flood and salt tolerance in mega rice varieties of India through a multi-institutional network. Plant Sci. 2016, 242, 278–287. [Google Scholar] [CrossRef] [PubMed]
  91. Feng, B.; Chen, K.; Cui, Y.; Wu, Z.; Zheng, T.; Zhu, Y.; Ali, J.; Wang, B.; Xu, J.; Zhang, W.; et al. Genetic Dissection and Simultaneous Improvement of Drought and Low Nitrogen Tolerances by Designed QTL Pyramiding in Rice. Front. Plant Sci. 2018, 9, 306. [Google Scholar] [CrossRef] [Green Version]
  92. Dharmappa, P.M.; Doddaraju, P.; Malagondanahalli, M.V.; Rangappa, R.B.; Mallikarjuna, N.M.; Rajendrareddy, S.H.; Ramanjinappa, R.; Mavinahalli, R.P.; Prasad, T.G.; Udayakumar, M.; et al. Introgression of Root and Water Use Efficiency Traits Enhances Water Productivity: An Evidence for Physiological Breeding in Rice (Oryza sativa L.). Rice 2019, 12, 14. [Google Scholar] [CrossRef] [PubMed]
  93. Aversano, R.; Ercolano, M.R.; Caruso, I.; Fasano, C.; Rosellini, D.; Carputo, D. Molecular Tools for Exploring Polyploid Genomes in Plants. Int. J. Mol. Sci. 2012, 13, 10316–10335. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Muthukumar, C.; Subathra, T.; Aiswarya, J.; Gayathri, V.; Babu, R.C. Comparative genome-wide association studies for plant production traits under drought in diverse rice (Oryza sativa L.) lines using SNP and SSR markers. Curr. Sci. 2015, 109, 139–147. [Google Scholar]
  95. Bhandari, A.; Bartholomé, J.; Cao-Hamadoun, T.-V.; Kumari, N.; Frouin, J.; Kumar, A.; Ahmadi, N. Selection of trait-specific markers and multi-environment models improve genomic predictive ability in rice. PLoS ONE 2019, 14, e0208871. [Google Scholar] [CrossRef] [Green Version]
  96. Anderson, J.A.; Sorrells, M.E.; Tanksley, S.D. RFLP Analysis of Genomic Regions Associated with Resistance to Preharvest Sprouting in Wheat. Crop. Sci. 1993, 33, 453–459. [Google Scholar] [CrossRef]
  97. Gupta, P.K.; Balyan, H.S.; Sharma, S.; Kumar, R. Genetics of yield, abiotic stress tolerance and biofortification in wheat (Triticum aestivum L.). Theor. Appl. Genet. 2020, 133, 1569–1602. [Google Scholar] [CrossRef]
  98. Cormier, F.; Foulkes, J.; Hirel, B.; Gouache, D.; Moënne-Loccoz, Y.; Le Gouis, J. Breeding for increased nitrogen-use efficiency: A review for wheat (T. aestivum L.). Plant Breed. 2016, 135, 255–278. [Google Scholar] [CrossRef] [Green Version]
  99. Zhang, M.; Gao, M.; Zheng, H.; Yuan, Y.; Zhou, X.; Guo, Y.; Zhang, G.; Zhao, Y.; Kong, F.; An, Y.; et al. QTL mapping for nitrogen use efficiency and agronomic traits at the seedling and maturity stages in wheat. Mol. Breed. 2019, 39, 71. [Google Scholar] [CrossRef]
  100. Mérida-García, R.; Bentley, A.R.; Gálvez, S.; Dorado, G.; Solís, I.; Ammar, K.; Hernandez, P. Mapping Agronomic and Quality Traits in Elite Durum Wheat Lines under Differing Water Regimes. Agronomy 2020, 10, 144. [Google Scholar] [CrossRef] [Green Version]
  101. Pascual, L.; Ruiz, M.; López-Fernández, M.; Pérez-Peña, H.; Benavente, E.; Vázquez, J.F.; Sansaloni, C.; Giraldo, P. Genomic analysis of Spanish wheat landraces reveals their variability and potential for breeding. BMC Genom. 2020, 21, 1–17. [Google Scholar] [CrossRef] [Green Version]
  102. Sun, C.; Dong, Z.; Zhao, L.; Ren, Y.; Zhang, N.; Chen, F. The Wheat 660K SNP array demonstrates great potential for marker-assisted selection in polyploid wheat. Plant Biotechnol. J. 2020, 18, 1354–1360. [Google Scholar] [CrossRef] [PubMed]
  103. Ahmad, M. Molecular marker-assisted selection of HMW glutenin alleles related to wheat bread quality by PCR-generated DNA markers. Theor. Appl. Genet. 2000, 101, 892–896. [Google Scholar] [CrossRef]
  104. Vida, G.; Gál, M.; Uhrin, A.; Veisz, O.; Syed, N.H.; Flavell, A.J.; Wang, Z.; Bedö, Z. Molecular markers for the identification of resistance genes and marker-assisted selection in breeding wheat for leaf rust resistance. Euphytica 2009, 170, 67–76. [Google Scholar] [CrossRef]
  105. Mohammadi, R. Breeding for increased drought tolerance in wheat: A review. Crop. Pasture Sci. 2018, 69, 223–241. [Google Scholar] [CrossRef]
  106. Juliana, P.; Poland, J.; Huerta-Espino, J.; Shrestha, S.; Crossa, J.; Crespo-Herrera, L.; Toledo, F.H.; Govindan, V.; Mondal, S.; Kumar, U.; et al. Improving grain yield, stress resilience and quality of bread wheat using large-scale genomics. Nat. Genet. 2019, 51, 1530–1539. [Google Scholar] [CrossRef] [PubMed]
  107. Guzman, C.; Peña, R.J.; Singh, R.; Autrique, E.; Dreisigacker, S.; Crossa, J.; Rutkoski, J.; Poland, J.; Battenfield, S. Wheat quality improvement at CIMMYT and the use of genomic selection on it. Appl. Transl. Genom. 2016, 11, 3–8. [Google Scholar] [CrossRef] [Green Version]
  108. Zaïm, M.; Kabbaj, H.; Kehel, Z.; Gorjanc, G.; Filali-Maltouf, A.; Belkadi, B.; Nachit, M.M.; Bassi, F.M. Combining QTL Analysis and Genomic Predictions for Four Durum Wheat Populations Under Drought Conditions. Front. Genet. 2020, 11, 316. [Google Scholar] [CrossRef]
  109. Hao, Z.; Li, X.; Liu, X.; Xie, C.; Li, M.; Zhang, D.; Zhang, S. Meta-analysis of constitutive and adaptive QTL for drought tol-erance in maize. Euphytica 2010, 174, 165–177. [Google Scholar] [CrossRef]
  110. Zhang, X.; Tang, B.; Yu, F.; Li, L.; Wang, M.; Xue, Y.; Zhang, Z.; Yan, J.; Yue, B.; Zheng, Y.; et al. Identification of Major QTL for Waterlogging Tolerance Using Genome-Wide Association and Linkage Mapping of Maize Seedlings. Plant Mol. Biol. Rep. 2012, 31, 594–606. [Google Scholar] [CrossRef]
  111. Pestsova, E.; Lichtblau, D.; Wever, C.; Presterl, T.; Bolduan, T.; Ouzunova, M.; Westhoff, P. QTL mapping of seedling root traits associated with nitrogen and water use efficiency in maize. Euphytica 2015, 209, 585–602. [Google Scholar] [CrossRef]
  112. Luo, M.; Zhao, Y.; Zhang, R.; Xing, J.; Duan, M.; Li, J.; Wang, N.; Wang, W.; Zhang, S.; Chen, Z.; et al. Mapping of a major QTL for salt tolerance of mature field-grown maize plants based on SNP markers. BMC Plant Biol. 2017, 17, 1–10. [Google Scholar] [CrossRef] [Green Version]
  113. Zehr, B.E.; Dudley, J.W.; Chojecki, J.; Maroof, M.A.S.; Mowers, R.P. Use of RFLP markers to search for alleles in a maize population for improvement of an elite hybrid. Theor. Appl. Genet. 1992, 83, 903–911. [Google Scholar] [CrossRef]
  114. Silvela, L.; Ayuso, M.; Gil-Delgado, L.; Salaices, L. Genetic and environmental contributions to bread-wheat flour quality using the SDS sedimentation test as an index. Theor. Appl. Genet. 1993, 86, 894–899. [Google Scholar] [CrossRef] [PubMed]
  115. Cho, Y.G.; Eun, M.Y.; McCouch, S.R.; Chae, Y.A. The semidwarf gene, sd-1, of rice (Oryza sativa L.). II. Molecular mapping and marker-assisted selection. Theor. Appl. Genet. 1994, 89, 54–59. [Google Scholar] [CrossRef]
  116. Scandalios, J.G.; Felder, M.R. Developmental expression of alcohol dehydrogenases in maize. Dev. Biol. 1971, 25, 641–654. [Google Scholar] [CrossRef]
  117. Luig, N.; Rajaram, S. The effect of temperature and genetic background on host gene expression and interaction to Puccinia graminis tritici. Phytopathology 1972, 62, 1171–1174. [Google Scholar] [CrossRef]
  118. Sano, Y. Differential regulation of waxy gene expression in rice endosperm. Theor. Appl. Genet. 1984, 68, 467–473. [Google Scholar] [CrossRef]
  119. Rodrigues, C.M.; Mafra, V.S.; Machado, M.A. Transcriptomics. In Omics in Plant Breeding; Borem, A., Fritsche-Neto, R., Eds.; Wiley-Blackwell: Hoboken, NJ, USA, 2014; pp. 33–57. [Google Scholar]
  120. Morozova, O.; Hirst, M.; Marra, M.A. Applications of New Sequencing Technologies for Transcriptome Analysis. Annu. Rev. Genom. Hum. Genet. 2009, 10, 135–151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Weis, J.H.; Tan, S.S.; Martin, B.K.; Wittwer, C.T. Detection of rare mRNAs via quantitative RT-PCR. Trends Genet. 1992, 8, 263–264. [Google Scholar] [CrossRef]
  122. Parker, R.M.; Barnes, N.M. mRNA: Detection by in Situ and northern hybridization. Methods Mol. Biol. 1999, 106, 247–283. [Google Scholar] [PubMed]
  123. Bustin, S.A. Absolute quantification of mRNA using real-time reverse transcription polymerase chain reaction assays. J. Mol. Endocrinol. 2000, 25, 169–193. [Google Scholar] [CrossRef] [Green Version]
  124. Deepak, S.; Kottapalli, K.; Rakwal, R.; Oros, G.; Rangappa, K.; Iwahashi, H.; Masuo, Y.; Agrawal, G. Real-time PCR: Revolutionizing detection and expression analysis of genes. Curr. Genom. 2007, 8, 234–251. [Google Scholar] [CrossRef]
  125. Schena, M.; Shalon, D.; Davis, R.W.; Brown, P.O. Quantitative Monitoring of Gene Expression Patterns with a Complementary DNA Microarray. Science 1995, 270, 467–470. [Google Scholar] [CrossRef] [Green Version]
  126. Zik, M.; Irish, V.F. Global Identification of Target Genes Regulated by APETALA3 and PISTILLATA Floral Homeotic Gene Action. Plant Cell 2002, 15, 207–222. [Google Scholar] [CrossRef] [Green Version]
  127. Harmer, S.L.; HogenEsch, J.B.; Straume, M.; Chang, H.-S.; Han, B.; Zhu, T.; Wang, X.; Kreps, J.A.; Kay, S.A. Orchestrated Transcription of Key Pathways in Arabidopsis by the Circadian Clock. Science 2000, 290, 2110–2113. [Google Scholar] [CrossRef]
  128. Wang, Z.; Gerstein, M.; Snyder, M. RNA-Seq: A revolutionary tool for transcriptomics. Nat. Rev. Genet. 2009, 10, 57–63. [Google Scholar] [CrossRef]
  129. Wang, Y.; Yao, Q.; Zhang, Y.; Zhang, Y.; Xing, J.; Yang, B.; Mi, G.; Li, Z.; Zhang, M. The Role of Gibberellins in Regulation of Nitrogen Uptake and Physiological Traits in Maize Responding to Nitrogen Availability. Int. J. Mol. Sci. 2020, 21, 1824. [Google Scholar] [CrossRef] [Green Version]
  130. Wang, H.; Wang, H.; Shao, H.; Tang, X. Recent Advances in Utilizing Transcription Factors to Improve Plant Abiotic Stress Tolerance by Transgenic Technology. Front. Plant Sci. 2016, 7, 67. [Google Scholar] [CrossRef] [Green Version]
  131. Mundy, J.; Chua, N.H. Abscisic acid and water-stress induce the expression of a novel rice gene. EMBO J. 1988, 7, 2279–2286. [Google Scholar] [CrossRef]
  132. Curry, J.; Morris, C.F.; Walker-Simmons, M.K. Sequence analysis of a cDNA encoding a Group 3 LEA mRNA inducible by ABA or dehydration stress in wheat. Plant Mol. Biol. 1991, 16, 1073–1076. [Google Scholar] [CrossRef]
  133. Bochicchio, A.; Vazzana, C.; Velasco, R.; Singh, M.; Bartels, D. Exogenous ABA induces desiccation tolerance and leads to the synthesis of specific gene transcription in immature embryos of maize. Maydica (Italy) 1991, 36, 11–16. [Google Scholar]
  134. Borkird, C.; Claes, B.; Caplan, A.; Simoens, C.; Van Montagu, M. Differential Expression of Water-Stress Associated Genes in Tissues of Rice Plants. J. Plant Physiol. 1991, 138, 591–595. [Google Scholar] [CrossRef]
  135. Vierling, R.A.; Nguyen, H.T. Heat-Shock Protein Gene Expression in Diploid Wheat Genotypes Differing in Thermal Tolerance. Crop. Sci. 1992, 32, 370–377. [Google Scholar] [CrossRef]
  136. Frova, C.; Gorla, M.S. Quantitative expression of maize HSPs: Genetic dissection and association with thermotolerance. Theor. Appl. Genet. 1993, 86, 213–220. [Google Scholar] [CrossRef] [PubMed]
  137. Lian, X.; Wang, S.; Zhang, J.; Feng, Q.; Zhang, L.; Fan, D.; Li, X.; Yuan, D.; Han, B.; Zhang, Q. Expression Profiles of 10,422 Genes at Early Stage of Low Nitrogen Stress in Rice Assayed using a cDNA Microarray. Plant Mol. Biol. 2006, 60, 617–631. [Google Scholar] [CrossRef]
  138. Bernard, S.M.; Møller, A.L.B.; Dionisio, G.; Kichey, T.; Jahn, T.P.; Dubois, F.; Baudo, M.; Lopes, M.S.; Tercé-Laforgue, T.; Foyer, C.H.; et al. Gene expression, cellular localisation and function of glutamine synthetase isozymes in wheat (Triticum aestivum L.). Plant Mol. Biol. 2008, 67, 89–105. [Google Scholar] [CrossRef]
  139. Martin, A.; Lee, J.; Kichey, T.; Gerentes, D.; Zivy, M.; Tatout, C.; Dubois, F.; Balliau, T.; Valot, B.; Davanture, M.; et al. Two Cytosolic Glutamine Synthetase Isoforms of Maize Are Specifically Involved in the Control of Grain Production. Plant Cell 2006, 18, 3252–3274. [Google Scholar] [CrossRef] [Green Version]
  140. Kohli, A.; Xiong, J.; Greco, R.; Christou, P.; Pereira, A. Tagged Transcriptome Display (TTD) in indica rice using Ac transposition. Mol. Genet. Genom. 2001, 266, 1–11. [Google Scholar] [CrossRef] [PubMed]
  141. Kashkush, K.; Feldman, M.; Levy, A.A. Gene loss, silencing and activation in a newly synthesized wheat allotetraploid. Genetics 2002, 160, 1651–1659. [Google Scholar]
  142. Wang, H.; Miyazaki, S.; Kawai, K.; Deyholos, M.; Galbraith, D.W.; Bohnert, H.J. Temporal progression of gene expression responses to salt shock in maize roots. Plant Mol. Biol. 2003, 52, 873–891. [Google Scholar] [CrossRef]
  143. Wu, C.-Q.; Hu, H.-H.; Zeng, Y.; Liang, D.-C.; Xie, K.-B.; Zhang, J.-W.; Chu, Z.-H.; Xiong, L.-Z. Identification of Novel Stress-responsive Transcription Factor Genes in Rice by cDNA Array Analysis. J. Integr. Plant Biol. 2006, 48, 1216–1224. [Google Scholar] [CrossRef]
  144. Ergen, N.Z.; Thimmapuram, J.; Bohnert, H.J.; Budak, H. Transcriptome pathways unique to dehydration tolerant relatives of modern wheat. Funct. Integr. Genom. 2009, 9, 377–396. [Google Scholar] [CrossRef] [Green Version]
  145. Zhuang, Y.; Ren, G.; Yue, G.; Li, Z.; Qu, X.; Hou, G.; Zhu, Y.; Zhang, J. Effects of water-deficit stress on the transcriptomes of developing immature ear and tassel in maize. Plant Cell Rep. 2007, 26, 2137–2147. [Google Scholar] [CrossRef]
  146. Antoine, W.; Stewart, J.M.; Reyes, B.G.D.L. The rice homolog of the sodium/lithium tolerance gene functions as molecular chaperon in vitro. Physiol. Plant. 2005, 125, 299–310. [Google Scholar] [CrossRef]
  147. Hays, D.; Mason, E.; Do, J.H.; Menz, M.; Reynolds, M. Expression quantitative trait loci mapping heat tolerance during reproductive development in wheat (Triticum aestivum). In Wheat Production in Stressed Environments; Buck, H.T., Nisi, J.E., Salomon, N., Eds.; Springer: Dordrecht, The Netherlands, 2007; Volume 12, p. 373. [Google Scholar]
  148. Dutra, S.; Von Pinho, E.; Santos, H.; Lima, A.; Von Pinho, R.; Carvalho, M. Genes related to high temperature tolerance during maize seed germination. Genet. Mol. Res. 2015, 14, 18047–18058. [Google Scholar] [CrossRef]
  149. Gregersen, P.L.; Holm, P.B.; Krupinska, K. Leaf senescence and nutrient remobilisation in barley and wheat. Plant Biol. 2008, 10, 37–49. [Google Scholar] [CrossRef]
  150. Clay, S.A.; Clay, D.E.; Horvath, D.P.; Pullis, J.; Carlson, C.G.; Hansen, S.; Reicks, G. Corn Response to Competition: Growth Alteration vs. Yield Limiting Factors. Agron. J. 2009, 101, 1522–1529. [Google Scholar] [CrossRef] [Green Version]
  151. Wasaki, J.; Yonetani, R.; Shinano, T.; Kai, M.; Osaki, M. Expression of the OsPI1 gene, cloned from rice roots using cDNA microarray, rapidly responds to phosphorus status. New Phytol. 2003, 158, 239–248. [Google Scholar] [CrossRef]
  152. Terzi, V.; Ferrari, B.; Finocchiaro, F.; Di Fonzo, N.; Stanca, A.M.; Lamacchia, C.; Napier, J.; Shewry, P.R.; Faccioli, P. TaqMan PCR for detection of genetically modified durum wheat. J. Cereal Sci. 2003, 37, 157–163. [Google Scholar] [CrossRef]
  153. Vaïtilingom, M.; Pijnenburg, H.; Gendre, F.; Brignon, P. Real-Time Quantitative PCR Detection of Genetically Modified Maximizer Maize and Roundup Ready Soybean in Some Representative Foods. J. Agric. Food Chem. 1999, 47, 5261–5266. [Google Scholar] [CrossRef]
  154. Kumar, K.; Rao, K.P.; Sharma, P.; Sinha, A.K. Differential regulation of rice mitogen activated protein kinase kinase (MKK) by abiotic stress. Plant Physiol. Biochem. 2008, 46, 891–897. [Google Scholar] [CrossRef] [PubMed]
  155. Gallé, Á.; Csiszár, J.; Secenji, M.; Guóth, A.; Cseuz, L.; Tari, I.; Györgyey, J.; Erdei, L. Glutathione transferase activity and expression patterns during grain filling in flag leaves of wheat genotypes differing in drought tolerance: Response to water deficit. J. Plant Physiol. 2009, 166, 1878–1891. [Google Scholar] [CrossRef] [Green Version]
  156. Qin, F.; Kakimoto, M.; Sakuma, Y.; Maruyama, K.; Osakabe, Y.; Tran, L.-S.P.; Shinozaki, K.; Yamaguchi-Shinozaki, K. Regulation and functional analysis of ZmDREB2A in response to drought and heat stresses in Zea mays L. Plant J. 2007, 50, 54–69. [Google Scholar] [CrossRef]
  157. Khurana, P.; Chauhan, H. Characterization and expression of high temperature stress responsive genes in bread wheat (Triticum aestivum L.). Czech J. Genet. Plant Breed. 2011, 47, S94–S97. [Google Scholar] [CrossRef] [Green Version]
  158. Duan, Y.H.; Zhang, Y.L.; Ye, L.T.; Fan, X.R.; Xu, G.H.; Shen, Q.R. Responses of Rice Cultivars with Different Nitrogen Use Efficiency to Partial Nitrate Nutrition. Ann. Bot. 2007, 99, 1153–1160. [Google Scholar] [CrossRef] [Green Version]
  159. Nigro, D.; Gu, Y.Q.; Huo, N.; Marcotuli, I.; Blanco, A.; Gadaleta, A.; Anderson, O.D. Structural Analysis of the Wheat Genes Encoding NADH-Dependent Glutamine-2-oxoglutarate Amidotransferases and Correlation with Grain Protein Content. PLoS ONE 2013, 8, e73751. [Google Scholar] [CrossRef]
  160. Chen, R.; Tian, M.; Wu, X.; Huang, Y. Differential global gene expression changes in response to low nitrogen stress in two maize inbred lines with contrasting low nitrogen tolerance. Genes Genom. 2011, 33, 491–497. [Google Scholar] [CrossRef]
  161. Zhang, G.; Guo, G.; Hu, X.; Zhang, Y.; Li, Q.; Li, R.; Zhuang, R.; Lu, Z.; He, Z.; Fang, X.; et al. Deep RNA sequencing at single base-pair resolution reveals high complexity of the rice transcriptome. Genome Res. 2010, 20, 646–654. [Google Scholar] [CrossRef] [Green Version]
  162. Pont, C.; Murat, F.; Confolent, C.; Balzergue, S.; Salse, J. RNA-seq in grain unveils fate of neo- and paleopolyploidization events in bread wheat (Triticum aestivum L.). Genome Biol. 2011, 12, 119. [Google Scholar] [CrossRef] [Green Version]
  163. Davidson, R.M.; Hansey, C.N.; Gowda, M.; Childs, K.L.; Lin, H.; Vaillancourt, B.; Sekhon, R.S.; De Leon, N.; Kaeppler, S.M.; Jiang, N.; et al. Utility of RNA Sequencing for Analysis of Maize Reproductive Transcriptomes. Plant Genome 2011, 4, 191–203. [Google Scholar] [CrossRef]
  164. Silveira, R.; Abreu, F.; Mamidi, S.; McClean, P.; Vianello, R.; Lanna, A.; Carneiro, N.; Brondani, C. Expression of drought tolerance genes in tropical upland rice cultivars (Oryza sativa). Genet. Mol. Res. 2015, 14, 8181–8200. [Google Scholar] [CrossRef]
  165. Okay, S.; Derelli, E.; Unver, T. Transcriptome-wide identification of bread wheat WRKY transcription factors in response to drought stress. Mol. Genet. Genom. 2014, 289, 765–781. [Google Scholar] [CrossRef]
  166. Kakumanu, A.; Ambavaram, M.M.; Klumas, C.; Krishnan, A.; Batlang, U.; Myers, E.; Grene, R.; Pereira, A. Effects of Drought on Gene Expression in Maize Reproductive and Leaf Meristem Tissue Revealed by RNA-Seq. Plant Physiol. 2012, 160, 846–867. [Google Scholar] [CrossRef] [Green Version]
  167. Liao, J.-L.; Zhou, H.-W.; Peng, Q.; Zhong, P.-A.; Zhang, H.-Y.; He, C.; Huang, Y.-J. Transcriptome changes in rice (Oryza sativa L.) in response to high night temperature stress at the early milky stage. BMC Genom. 2015, 16, 18. [Google Scholar] [CrossRef] [Green Version]
  168. Kumar, R.R.; Goswami, S.; Sharma, S.K.; Kala, Y.K.; Rai, G.K.; Mishra, D.C.; Grover, M.; Singh, G.P.; Pathak, H.; Rai, A.; et al. Harnessing Next Generation Sequencing in Climate Change: RNA-Seq Analysis of Heat Stress-Responsive Genes in Wheat (Triticum aestivum L.). OMICS J. Integr. Biol. 2015, 19, 632–647. [Google Scholar] [CrossRef] [Green Version]
  169. Frey, F.P.; Urbany, C.; Hüttel, B.; Reinhardt, R.; Stich, B. Genome-wide expression profiling and phenotypic evaluation of European maize inbreds at seedling stage in response to heat stress. BMC Genom. 2015, 16, 123. [Google Scholar] [CrossRef] [Green Version]
  170. Sinha, S.K.; Sevanthi, V.A.M.; Chaudhary, S.; Tyagi, P.; Venkadesan, S.; Rani, M.; Mandal, P.K. Transcriptome analysis of two rice varieties contrasting for nitrogen use efficiency under chronic N starvation reveals differences in chloroplast and starch metabolism-related genes. Genes 2018, 9, 206. [Google Scholar] [CrossRef] [Green Version]
  171. Camilios-Neto, D.; Bonato, P.; Wassem, R.; Tadra-Sfeir, M.Z.; Brusamarello-Santos, L.C.C.; Valdameri, G.; Donatti, L.; Faoro, H.; Weiss, V.A.; Chubatsu, L.S.; et al. Dual RNA-seq transcriptional analysis of wheat roots colonized by Azospirillum brasilense reveals up-regulation of nutrient acquisition and cell cycle genes. BMC Genom. 2014, 15, 1–13. [Google Scholar] [CrossRef] [Green Version]
  172. Trevisan, S.; Manoli, A.; Ravazzolo, L.; Botton, A.; Pivato, M.; Masi, A.; Quaggiotti, S. Nitrate sensing by the maize root apex transition zone: A merged transcriptomic and proteomic survey. J. Exp. Bot. 2015, 66, 3699–3715. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Maleck, K.; Levine, A.; Eulgem, T.; Morgan, A.; Schmid, J.; Lawton, K.A.; Dangl, J.L.; Dietrich, R.A. The transcriptome of Arabidopsis thaliana during systemic acquired resistance. Nat. Genet. 2000, 26, 403–410. [Google Scholar] [CrossRef]
  174. Swift, J.; Adame, M.; Tranchina, D.; Henry, A.; Coruzzi, G.M. Water impacts nutrient dose responses genome-wide to affect crop production. Nat. Commun. 2019, 10, 1374. [Google Scholar] [CrossRef] [PubMed]
  175. Wasaki, J.; Yonetani, R.; Kuroda, S.; Shinano, T.; Yazaki, J.; Fujii, F.; Shimbo, K.; Yamamoto, K.; Sakata, K.; Sasaki, T. Tran-scriptomic analysis of metabolic changes by phosphorus stress in rice plant roots. Plant Cell Environ. 2003, 26, 1515–1523. [Google Scholar] [CrossRef]
  176. Pillai, M.A.; Yanagihara, S.; Akiyama, T. Molecular cloning and characterization of salt responsive genes in rice (Oryza sativa). J. Plant Physiol. 2001, 158, 1189–1194. [Google Scholar] [CrossRef]
  177. Baulcombe, D.C.; Buffard, D. Gibberellic-acid-regulated expression of α-amylase and six other genes in wheat aleurone layers. Planta 1983, 157, 493–501. [Google Scholar] [CrossRef] [PubMed]
  178. Winfield, M.O.; Lu, C.; Wilson, I.D.; Coghill, J.A.; Edwards, K.J. Plant responses to cold: Transcriptome analysis of wheat. Plant Biotechnol. J. 2010, 8, 749–771. [Google Scholar] [CrossRef] [Green Version]
  179. Henry, R.J.; Furtado, A.; Rangan, P. Wheat seed transcriptome reveals genes controlling key traits for human preference and crop adaptation. Curr. Opin. Plant Biol. 2018, 45, 231–236. [Google Scholar] [CrossRef] [Green Version]
  180. Hu, L.; Xie, Y.; Fan, S.; Wang, Z.; Wang, F.; Zhang, B.; Li, H.; Song, J.; Kong, L. Comparative analysis of root transcriptome profiles between drought-tolerant and susceptible wheat genotypes in response to water stress. Plant Sci. 2018, 272, 276–293. [Google Scholar] [CrossRef]
  181. Amirbakhtiar, N.; Ismaili, A.; Ghaffari, M.R.; Firouzabadi, F.N.; Shobbar, Z.-S. Transcriptome response of roots to salt stress in a salinity-tolerant bread wheat cultivar. PLoS ONE 2019, 14, e0213305. [Google Scholar] [CrossRef]
  182. Curci, P.L.; Cigliano, R.A.; Zuluaga, D.L.; Janni, M.; Sanseverino, W.; Sonnante, G. Transcriptomic response of durum wheat to nitrogen starvation. Sci. Rep. 2017, 7, 1–14. [Google Scholar] [CrossRef]
  183. Link, G.; Coen, N.M.; Bogorad, L. Differential expression of the gene for the large subunit of ribulose bisphosphate carboxylase in maize leaf cell types. Cell 1978, 15, 725–731. [Google Scholar] [CrossRef]
  184. Bedbrook, J.R.; Link, G.; Coen, D.M.; Bogorad, L. Maize plastid gene expressed during photoregulated development. Proc. Natl. Acad. Sci. USA 1978, 75, 3060–3064. [Google Scholar] [CrossRef] [Green Version]
  185. Wang, M.; Wang, Y.; Zhang, Y.; Li, C.; Gong, S.; Yan, S.; Li, G.; Hu, G.; Ren, H.; Yang, J.; et al. Comparative transcriptome analysis of salt-sensitive and salt-tolerant maize reveals potential mechanisms to enhance salt resistance. Genes Genom. 2019, 41, 781–801. [Google Scholar] [CrossRef] [PubMed]
  186. Yang, X.S.; Wu, J.; Ziegler, T.E.; Yang, X.; Zayed, A.; Rajani, M.; Zhou, D.; Basra, A.S.; Schachtman, D.P.; Peng, M.; et al. Gene Expression Biomarkers Provide Sensitive Indicators of in Planta Nitrogen Status in Maize. Plant Physiol. 2011, 157, 1841–1852. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Zamboni, A.; Astolfi, S.; Zuchi, S.; Pii, Y.; Guardini, K.; Tononi, P.; Varanini, Z. Nitrate induction triggers different transcrip-tional changes in a high and a low nitrogen use efficiency maize inbred line. J. Integr. Plant Biol. 2014, 56, 1080–1094. [Google Scholar] [CrossRef] [PubMed]
  188. Marino, R.; Ponnaiah, M.; Krajewski, P.; Frova, C.; Gianfranceschi, L.; Pè, M.E.; Sari-Gorla, M. Addressing drought tolerance in maize by transcriptional profiling and mapping. Mol. Genet. Genom. 2008, 281, 163–179. [Google Scholar] [CrossRef]
  189. Guo, J.; Li, C.; Zhang, X.; Li, Y.; Zhang, D.; Shi, Y.; Song, Y.; Li, Y.; Yang, D.; Wang, T. Transcriptome and GWAS analyses reveal candidate gene for seminal root length of maize seedlings under drought stress. Plant Sci. 2020, 292, 110380. [Google Scholar] [CrossRef]
  190. Ren, J.; Xie, T.; Wang, Y.; Li, H.; Liu, T.; Zhang, S.; Yin, L.; Wang, S.; Deng, X.; Ke, Q. Coordinated regulation of carbon and nitrogen assimilation confers drought tolerance in maize (Zea mays L.). Environ. Exp. Bot. 2020, 176, 104086. [Google Scholar] [CrossRef]
  191. Humbert, S.; Subedi, S.; Cohn, J.; Zeng, B.; Bi, Y.-M.; Chen, X.; Zhu, T.; McNicholas, P.D.; Rothstein, S.J. Genome-wide ex-pression profiling of maize in response to individual and combined water and nitrogen stresses. BMC Genomics 2013, 14, 3. [Google Scholar] [CrossRef] [Green Version]
  192. Koornneeff, M.; Dellaert, L.; Van Der Veen, J. EMS- and relation-induced mutation frequencies at individual loci in Arabidopsis thaliana (L.) Heynh. Mutat. Res. Mol. Mech. Mutagen. 1982, 93, 109–123. [Google Scholar] [CrossRef]
  193. Pathirana, R. Plant mutation breeding in agriculture. CAB Rev. Perspect. Agric. Veter. Sci. Nutr. Nat. Resour. 2011, 6, 107–126. [Google Scholar] [CrossRef]
  194. McCallum, C.M.; Comai, L.; Greene, E.A.; Henikoff, S. Targeting induced locallesions in genomes (TILLING) for plant functional genomics. Plant Physiol. 2000, 123, 439–442. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Taheri, S.; Abdullah, T.L.; Jain, S.M.; Sahebi, M.; Azizi, P. TILLING, high-resolution melting (HRM), and next-generation sequencing (NGS) techniques in plant mutation breeding. Mol. Breed. 2017, 37, 40. [Google Scholar] [CrossRef]
  196. 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]
  197. Cardi, T. Cisgenesis and genome editing: Combining concepts and efforts for a smarter use of genetic resources in crop breeding. Plant Breed. 2016, 135, 139–147. [Google Scholar] [CrossRef]
  198. Shewry, P.R.; Jones, H.D.; Halford, N.G. Plant Biotechnology: Transgenic Crops. In Food Biotechnology; Stahl, U., Donalies, U.E.B., Nevoigt, E., Eds.; Springer: Berlin/Heidelberg, Germany, 2008; Volume 111, pp. 149–186. [Google Scholar]
  199. International Service for Acquisition of Agri-biotech Applications. Brief 55: Global Status of Commercialized Biotech/GM Crops: 2019; ISAAA: Ithaca, NY, USA, 2019. [Google Scholar]
  200. Kozaki, A.; Sakamoto, A.; Tanaka, K.; Takeba, G. The promoter of the gene for glutamine synthetase from rice shows or-gan-specific and substrate-induced expression in transgenic tobacco plants. Plant Cell Physiol. 1991, 32, 353–358. [Google Scholar] [CrossRef]
  201. López-Calcagno, P.E.; Fisk, S.; Brown, K.L.; Bull, S.E.; South, P.F.; Raines, C.A. Overexpressing the H-protein of the glycine cleavage system increases biomass yield in glasshouse and field-grown transgenic tobacco plants. Plant Biotechnol. J. 2019, 17, 141–151. [Google Scholar] [CrossRef]
  202. Bosher, J.M.; Labouesse, M. RNA interference: Genetic wand and genetic watchdog. Nat. Cell Biol. 2000, 2, E31–E36. [Google Scholar] [CrossRef]
  203. Abe, K.; Ichikawa, H. Gene Overexpression Resources in Cereals for Functional Genomics and Discovery of Useful Genes. Front. Plant Sci. 2016, 7, 1359. [Google Scholar] [CrossRef] [Green Version]
  204. Kamburova, V.S.; Nikitina, E.V.; Shermatov, S.E.; Buriev, Z.T.; Kumpatla, S.P.; Emani, C.; Abdurakhmonov, I.Y. Genome Editing in Plants: An Overview of Tools and Applications. Int. J. Agron. 2017, 2017, 1–15. [Google Scholar] [CrossRef] [Green Version]
  205. Shrivastav, M.; De Haro, L.P.; Nickoloff, J.A. Regulation of DNA double-strand break repair pathway choice. Cell Res. 2008, 18, 134–147. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Cheng, J.K.; Alper, H.S. The genome editing toolbox: A spectrum of approaches for targeted modification. Curr. Opin. Biotechnol. 2014, 30, 87–94. [Google Scholar] [CrossRef] [PubMed]
  207. Noman, A.; Aqeel, M.; He, S. CRISPR-Cas9: Tool for Qualitative and Quantitative Plant Genome Editing. Front. Plant Sci. 2016, 7, 1740. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Gong, W.J.; Golic, K.G. Ends-out, or replacement, gene targeting in Drosophila. Proc. Natl. Acad. Sci. USA 2003, 100, 2556–2561. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Kim, Y.G.; Cha, J.; Chandrasegaran, S. Hybrid restriction enzymes: Zinc finger fusions to Fok I cleavage domain. Proc. Natl. Acad. Sci. USA 1996, 93, 1156. [Google Scholar] [CrossRef] [Green Version]
  210. Cermak, T.; Doyle, E.L.; Christian, M.; Wang, L.; Zhang, Y.; Schmidt, C.; Baller, J.A.; Somia, N.V.; Bogdanove, A.J.; Voytas, D.F. Efficient design and assembly of custom TALEN and other TAL effector-based constructs for DNA targeting. Nucleic Acids Res. 2011, 39, e82. [Google Scholar] [CrossRef] [Green Version]
  211. Puchta, H.; Fauser, F. Gene targeting in plants: 25 years later. Int. J. Dev. Biol. 2013, 57, 629–637. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Ratho, S.; Jachuck, P. A method of inducing awnless condition in rice by chemical mutagenesis. Curr. Sci. 1971, 40, 274–276. [Google Scholar]
  213. Rao, H.S.; Sears, E. Chemical mutagenesis in Triticum aestivum. Mutat. Res. Mol. Mech. Mutagen. 1964, 1, 387–399. [Google Scholar] [CrossRef]
  214. Bellini, G.; Bianchi, A.; Ottaviano, E. The use of interchanges involving B-type chromosomes in studying artificial mutagenesis in maize. Mol. Genet. Genom. 1961, 92, 85–99. [Google Scholar] [CrossRef]
  215. Jiang, S.-Y.; Bachmann, D.; La, H.; Ma, Z.; Venkatesh, P.N.; Ramamoorthy, R.; Ramachandran, S. Ds insertion mutagenesis as an efficient tool to produce diverse variations for rice breeding. Plant Mol. Biol. 2007, 65, 385–402. [Google Scholar] [CrossRef] [PubMed]
  216. Khan, A.J.; Hassan, S.; Tariq, M.; Khan, T. Haploidy breeding and mutagenesis for drought tolerance in wheat. In Mutations, In Vitro and Molecular Techniques for Environmentally Sustainable Crop Improvement; Springer International Publishing: Berlin/Heidelberg, Germany, 2002; pp. 75–82. [Google Scholar]
  217. Gao, Z.; Liu, H.; Wang, H.; Li, N.; Wang, D.; Song, Y.; Miao, Y.; Song, C. Generation of the genetic mutant population for the screening and characterization of the mutants in response to drought in maize. Chin. Sci. Bull. 2014, 59, 766–775. [Google Scholar] [CrossRef]
  218. Agarwal, M.; Sahi, C.; Katiyar-Agarwal, S.; Agarwal, S.; Young, T.; Gallie, D.R.; Sharma, V.M.; Ganesan, K.; Grover, A. Molecular characterization of rice hsp101: Complementation of yeast hsp104 mutation by disaggregation of protein granules and differential expression in indica and japonica rice types. Plant Mol. Biol. 2003, 51, 543–553. [Google Scholar] [CrossRef]
  219. Xue, G.-P.; Sadat, S.; Drenth, J.; McIntyre, C.L. The heat shock factor family from Triticum aestivumin response to heat and other major abiotic stresses and their role in regulation of heat shock protein genes. J. Exp. Bot. 2014, 65, 539–557. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  220. Zhao, Y.; Du, H.; Wang, Y.; Wang, H.; Yang, S.; Li, C.; Chen, N.; Yang, H.; Zhang, Y.; Zhu, Y.; et al. The calcium-dependent protein kinase ZmCDPK7 functions in heat-stress tolerance in maize (Zea mays L.). J. Integr. Plant Biol. 2020. [Google Scholar] [CrossRef]
  221. Brauer, E.K.; Rochon, A.; Bi, Y.M.; Bozzo, G.G.; Rothstein, S.J.; Shelp, B.J. Reappraisal of nitrogen use efficiency in rice over-expressing glutamine synthetase1. Physiol. Plant. 2011, 141, 361–372. [Google Scholar] [CrossRef]
  222. Toriyama, K.; Arimoto, Y.; Uchimiya, H.; Hinata, K. Transgenic Rice Plants After Direct Gene Transfer into Protoplasts. Nat. Biotechnol. 1988, 6, 1072–1074. [Google Scholar] [CrossRef]
  223. Hess, D.; Dressler, K.; Nimmrichter, R. Transformation experiments by pipetting Agrobacterium into the spikelets of wheat (Triticum aestivum L.). Plant Sci. 1990, 72, 233–244. [Google Scholar] [CrossRef]
  224. Klein, T.M.; Fromm, M.; Weissinger, A.; Tomes, D.; Schaaf, S.; Sletten, M.; Sanford, J.C. Transfer of foreign genes into intact maize cells with high-velocity microprojectiles. Proc. Natl. Acad. Sci. USA 1988, 85, 4305–4309. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Su, J.; Shen, Q.; Ho, T.-H.D.; Wu, R. Dehydration-Stress-Regulated Transgene Expression in Stably Transformed Rice Plants. Plant Physiol. 1998, 117, 913–922. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Sivamani, E.; Bahieldin, A.; Wraith, J.M.; Alniemi, T.S.; 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]
  227. Jeanneau, M.; Gerentes, D.; Foueillassar, X.; Zivy, M.; Vidal, J.; Toppan, A.; Perez, P. 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 2002, 84, 1127–1135. [Google Scholar] [CrossRef]
  228. Kishitani, S.; Takanami, T.; Suzuki, M.; Oikawa, M.; Yokoi, S.; Ishitani, M.; Alvarez-Nakase, A.M.; Takabe, T. Compatibility of glycinebetaine in rice plants: Evaluation using transgenic rice plants with a gene for peroxisomal betaine aldehyde dehydrogenase from barley. Plant Cell Environ. 2000, 23, 107–114. [Google Scholar] [CrossRef]
  229. Fu, J.; Momcilović, I.; Clemente, T.E.; Nersesian, N.; Trick, H.N.; Ristic, Z. Heterologous expression of a plastid EF-Tu reduces protein thermal aggregation and enhances CO2 fixation in wheat (Triticum aestivum) following heat stress. Plant Mol. Biol. 2008, 68, 277–288. [Google Scholar] [CrossRef] [Green Version]
  230. Makino, A.; Shimada, T.; Takumi, S.; Kaneko, K.; Matsuoka, M.; Shimamoto, K.; Nakano, H.; Miyao-Tokutomi, M.; Mae, T.; Yamamoto, N. Does Decrease in Ribulose-1,5-Bisphosphate Carboxylase by Antisense RbcS Lead to a Higher N-Use Efficiency of Photosynthesis under Conditions of Saturating CO2 and Light in Rice Plants? Plant Physiol. 1997, 114, 483–491. [Google Scholar] [CrossRef] [Green Version]
  231. Habash, D.Z.; Massiah, A.J.; Rong, H.L.; Wallsgrove, R.M.; Leigh, R.A. The role of cytosolic glutamine synthetase in wheat. Ann. Appl. Biol. 2001, 138, 83–89. [Google Scholar] [CrossRef]
  232. Liu, G.-W.; Sun, A.-L.; Li, D.-Q.; Athman, A.; Gilliham, M.; Liu, L.-H. Molecular identification and functional analysis of a maize (Zea mays) DUR3 homolog that transports urea with high affinity. Planta 2014, 241, 861–874. [Google Scholar] [CrossRef]
  233. Li, T.; Liu, B.; Spalding, M.H.; Weeks, D.P.; Yang, B. High-efficiency TALEN-based gene editing produces disease-resistant rice. Nat. Biotechnol. 2012, 30, 390–392. [Google Scholar] [CrossRef]
  234. Upadhyay, S.K.; Kumar, J.; Alok, A.; Tuli, R. RNA-guided genome editing for target gene mutations in wheat. G3 Genes Genomes Genet. 2013, 3, 2233–2238. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Liang, Z.; Zhang, K.; Chen, K.; Gao, C. Targeted Mutagenesis in Zea mays Using TALENs and the CRISPR/Cas System. J. Genet. Genom. 2014, 41, 63–68. [Google Scholar] [CrossRef]
  236. Lou, D.; Wang, H.; Liang, G.; Yu, D. OsSAPK2 Confers Abscisic Acid Sensitivity and Tolerance to Drought Stress in Rice. Front. Plant Sci. 2017, 8, 993. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  237. Chilcoat, D.; Liu, Z.-B.; Sander, J. Use of CRISPR/Cas9 for Crop Improvement in Maize and Soybean. Prog. Mol. Biol. Transl. Sci. 2017, 149, 27–46. [Google Scholar] [CrossRef] [PubMed]
  238. Wang, B.; Zhong, Z.; Wang, X.; Han, X.; Yu, D.; Wang, C.; Song, W.; Zheng, X.; Chen, C.; Zhang, Y. Knockout of the OsNAC006 Transcription Factor Causes Drought and Heat Sensitivity in Rice. Int. J. Mol. Sci. 2020, 21, 2288. [Google Scholar] [CrossRef] [Green Version]
  239. Li, J.; Zhang, X.; Sun, Y.; Zhang, J.; Du, W.; Guo, X.; Li, S.; Zhao, Y.; Xia, L. Efficient allelic replacement in rice by gene editing: A case study of the NRT1.1B gene. J. Integr. Plant Biol. 2018, 60, 536–540. [Google Scholar] [CrossRef] [Green Version]
  240. Todaka, D.; Shinozaki, K.; Yamaguchi-Shinozaki, K. Recent advances in the dissection of drought-stress regulatory networks and strategies for development of drought-tolerant transgenic rice plants. Front. Plant Sci. 2015, 6, 84. [Google Scholar] [CrossRef] [Green Version]
  241. Xu, D.; Duan, X.; Wang, B.; Hong, B.; Ho, T.; Wu, R. Expression of a Late Embryogenesis Abundant Protein Gene, HVA1, from Barley Confers Tolerance to Water Deficit and Salt Stress in Transgenic Rice. Plant Physiol. 1996, 110, 249–257. [Google Scholar] [CrossRef] [Green Version]
  242. Cheng, Z.; Targolli, J.; Huang, X.; Wu, R. Wheat LEA genes, PMA80 and PMA1959, enhance dehydration tolerance of trans-genic rice (Oryza sativa L.). Mol. Breed. 2002, 10, 71–82. [Google Scholar] [CrossRef]
  243. Yin, X.; Anand, A.; Quick, P.; Bandyopadhyay, A. Editing a stomatal developmental gene in rice with CRISPR/Cpf1. In Methods in Molecular Biology; Springer International Publishing: Berlin/Heidelberg, Germany, 2019; Volume 1917, pp. 257–268. [Google Scholar]
  244. Gaudin, A.C.M.; Henry, A.; Sparks, A.H.; Slamet-Loedin, I.H. Taking transgenic rice drought screening to the field. J. Exp. Bot. 2013, 64, 109–117. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Lamppa, G.; Nagy, F.; Chua, N.-H. Light-regulated and organ-specific expression of a wheat Cab gene in transgenic tobacco. Nat. Cell Biol. 1985, 316, 750–752. [Google Scholar] [CrossRef] [PubMed]
  246. Marcotte, W.R.; Russell, S.H.; Quatrano, R.S. Abscisic acid-responsive sequences from the em gene of wheat. Plant Cell 1989, 1, 969–976. [Google Scholar] [PubMed] [Green Version]
  247. Shewry, P. Transgenic wheat: Where do we stand after the first 12 years? Ann. Appl. Biol. 2005, 147, 1–14. [Google Scholar] [CrossRef]
  248. Khan, S.; Anwar, S.; Yu, S.; Sun, M.; Yang, Z.; Gao, Z.-Q. Development of Drought-Tolerant Transgenic Wheat: Achievements and Limitations. Int. J. Mol. Sci. 2019, 20, 3350. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  249. Gaponenko, A.K.; Shulga, O.A.; Mishutkina, Y.B.; Tsarkova, E.A.; Timoshenko, A.A.; Spechenkova, N.A. Perspectives of Use of Transcription Factors for Improving Resistance of Wheat Productive Varieties to Abiotic Stresses by Transgenic Technologies. Russ. J. Genet. 2018, 54, 27–35. [Google Scholar] [CrossRef]
  250. Hu, M.; Zhao, X.; Liu, Q.; Hong, X.; Zhang, W.; Zhang, Y.; Sun, L.; Li, H.; Tong, Y. Transgenic expression of plastidic glutamine synthetase increases nitrogen uptake and yield in wheat. Plant Biotechnol. J. 2018, 16, 1858–1867. [Google Scholar] [CrossRef] [PubMed]
  251. Rochester, D.E.; Winer, J.A.; Shah, D.M. The structure and expression of maize genes encoding the major heat shock protein, hsp70. EMBO J. 1986, 5, 451–458. [Google Scholar] [CrossRef] [PubMed]
  252. Streatfield, S.J.; Magallanes-Lundback, M.E.; Beifuss, K.K.; Brooks, C.A.; Harkey, R.L.; Love, R.T.; Bray, J.; Howard, J.A.; Jilka, J.M.; Hood, E.E. Analysis of the maize polyubiquitin-1 promoter heat shock elements and generation of promoter variants with modified expression characteristics. Transgenic Res. 2004, 13, 299–312. [Google Scholar] [CrossRef]
  253. Di, H.; Tian, Y.; Zu, H.; Meng, X.; Zeng, X.; Wang, Z. Enhanced salinity tolerance in transgenic maize plants expressing a BADH gene from Atriplex micrantha. Euphytica 2015, 206, 775–783. [Google Scholar] [CrossRef]
  254. Puskaric, V. Maize Inbred Line PH1MD Useful for Producing F1. Hybrid. Patent US6127610-A, 24 October 2000. [Google Scholar]
  255. Wu, J.; Lawit, S.J.; Weers, B.; Sun, J.; Mongar, N.; Van Hemert, J.; Melo, R.; Meng, X.; Rupe, M.; Clapp, J. Overexpression of zmm28 increases maize grain yield in the field. Proc. Natl. Acad. Sci. USA 2019, 116, 23850–23858. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Castiglioni, P.; Warner, D.; Bensen, R.J.; Anstrom, D.C.; Harrison, J.; Stoecker, M.; Abad, M.; Kumar, G.; Salvador, S.; D’Ordine, R.; et al. Bacterial RNA Chaperones Confer Abiotic Stress Tolerance in Plants and Improved Grain Yield in Maize under Water-Limited Conditions. Plant Physiol. 2008, 147, 446–455. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Shi, J.; Gao, H.; Wang, H.; Lafitte, H.R.; Archibald, R.L.; Yang, M.; Hakimi, S.M.; Mo, H.; Habben, J.E. ARGOS8 variants generated by CRISPR-Cas9 improve maize grain yield under field drought stress conditions. Plant Biotechnol. J. 2016, 15, 207–216. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  258. Voss-Fels, K.P.; Stahl, A.; Hickey, L.T. Q&A: Modern crop breeding for future food security. BMC Biol. 2019, 17, 1–7. [Google Scholar] [CrossRef] [Green Version]
  259. Blum, A. Genomics for drought resistance–getting down to earth. Funct. Plant Biol. 2014, 41, 1191–1198. [Google Scholar] [CrossRef]
  260. Araus, J.L.; Cairns, J.E. Field high-throughput phenotyping: The new crop breeding frontier. Trends Plant Sci. 2014, 19, 52–61. [Google Scholar] [CrossRef]
  261. Mayer, M.; Hölker, A.C.; González-Segovia, E.; Bauer, E.; Presterl, T.; Ouzunova, M.; Melchinger, A.E.; Schön, C.-C. Discovery of beneficial haplotypes for complex traits in maize landraces. Nat. Commun. 2020, 11, 1–10. [Google Scholar] [CrossRef]
  262. Sansaloni, C.; Franco, J.; Santos, B.; Percival-Alwyn, L.; Singh, S.; Petroli, C.; Campos, J.; Dreher, K.; Payne, T.; Marshall, D.; et al. Diversity analysis of 80,000 wheat accessions reveals consequences and opportunities of selection footprints. Nat. Commun. 2020, 11, 1–12. [Google Scholar] [CrossRef]
  263. Curatti, L.; Rubio, L.M. Challenges to develop nitrogen-fixing cereals by direct nif-gene transfer. Plant Sci. 2014, 225, 130–137. [Google Scholar] [CrossRef]
  264. Rogers, C.; Oldroyd, G.E.D. Synthetic biology approaches to engineering the nitrogen symbiosis in cereals. J. Exp. Bot. 2014, 65, 1939–1946. [Google Scholar] [CrossRef] [Green Version]
  265. National Academies of Sciences, Engineering and Medicine. Future genetically engineered crops. In Genetically Engineered Crops: Experiences and Prospects; National Academies Press: Washington, DC, USA, 2016. [Google Scholar] [CrossRef]
Figure 1. Global data illustrating the agricultural and economic relevance of rice, wheat, and maize. Area harvested and production correspond to 2019, while global gross production value corresponds to 2018. (Source: FAOSTAT database, in www.fao.org/faostat, accessed on 27 January 2021).
Figure 1. Global data illustrating the agricultural and economic relevance of rice, wheat, and maize. Area harvested and production correspond to 2019, while global gross production value corresponds to 2018. (Source: FAOSTAT database, in www.fao.org/faostat, accessed on 27 January 2021).
Agronomy 11 00376 g001
Figure 2. Number of bibliographic records referring to the main approaches for marker-assisted breeding in rice, wheat, and maize (source: Web of Science database, all years available; accessed 15 December 2020). QTL: Quantitative Trait Locus; GWAS: Genome-Wide Association Studies.
Figure 2. Number of bibliographic records referring to the main approaches for marker-assisted breeding in rice, wheat, and maize (source: Web of Science database, all years available; accessed 15 December 2020). QTL: Quantitative Trait Locus; GWAS: Genome-Wide Association Studies.
Agronomy 11 00376 g002
Figure 3. Relative relevance of breeding topics related to the main abiotic stresses and nutrient limitations in bibliographic records referring to QTL mapping, GWAS, and genomic selection, in rice, wheat, and maize (source: Web of Science database, all years available; accessed on 15 December 2020).
Figure 3. Relative relevance of breeding topics related to the main abiotic stresses and nutrient limitations in bibliographic records referring to QTL mapping, GWAS, and genomic selection, in rice, wheat, and maize (source: Web of Science database, all years available; accessed on 15 December 2020).
Agronomy 11 00376 g003
Figure 4. Number of bibliographic records referring to main modern approaches for gene expression analysis in rice, wheat, and maize (source: Web of Science database, all years available; accessed on 15 December 2020).
Figure 4. Number of bibliographic records referring to main modern approaches for gene expression analysis in rice, wheat, and maize (source: Web of Science database, all years available; accessed on 15 December 2020).
Agronomy 11 00376 g004
Figure 5. Relative relevance of breeding topics related to major abiotic stresses and nutrient-limiting topics in bibliographic records referring to gene expression and related tools in rice, wheat, and maize (source: Web of Science database, all years available; accessed on 15 December 2020).
Figure 5. Relative relevance of breeding topics related to major abiotic stresses and nutrient-limiting topics in bibliographic records referring to gene expression and related tools in rice, wheat, and maize (source: Web of Science database, all years available; accessed on 15 December 2020).
Agronomy 11 00376 g005
Figure 6. Number of bibliographic records referring to main approaches for genetic modification in rice, wheat and maize (Source: Web of Science database, all years available; accessed on 15 December 2020).
Figure 6. Number of bibliographic records referring to main approaches for genetic modification in rice, wheat and maize (Source: Web of Science database, all years available; accessed on 15 December 2020).
Agronomy 11 00376 g006
Figure 7. Relative relevance of breeding topics related to main abiotic stress and nutrient-limiting topics in bibliographic records referring to transgenesis and gene editing tools in rice, wheat, and maize (source: Web of Science database, all years available; accessed on 15 December 2020).
Figure 7. Relative relevance of breeding topics related to main abiotic stress and nutrient-limiting topics in bibliographic records referring to transgenesis and gene editing tools in rice, wheat, and maize (source: Web of Science database, all years available; accessed on 15 December 2020).
Agronomy 11 00376 g007
Table 1. Year and bibliographic reference of the first documented application of different marker-assisted breeding approaches in rice, wheat, and maize.
Table 1. Year and bibliographic reference of the first documented application of different marker-assisted breeding approaches in rice, wheat, and maize.
ApproachTopicRiceWheatMaize
QTL mapping(Any)1990 [42]1992 [43]1987 [44]
Drought tolerance1995 [45]1994 [46]1995 [47]
Heat tolerance2000 [48]2002 [49]1991 [50]
NUE2001 [51]2004 [52]1999 [53]
GWAS(Any)2009 [54]2009 [55]2011 [56]
Drought tolerance2010 [57]2014 [58]2011 [59]
Heat tolerance2017 [60]2017 [61]2019 [62]
NUE2019 [63]2014 [64]2017 [65]
Genome selection(Any)2014 [66]2011 [67]2007 [68]
Drought tolerance2018 [69]2018 [70]2013 [71]
Heat tolerance-2018 [70]2019 [72]
NUE2016 [73]2019 [74]2015 [75]
Table 2. Year and bibliographic reference of the first documented application of RNA-based approaches in rice, wheat, and maize. The timing of gene expression analyses by crop and topic is also indicated.
Table 2. Year and bibliographic reference of the first documented application of RNA-based approaches in rice, wheat, and maize. The timing of gene expression analyses by crop and topic is also indicated.
ApproachTopicRiceWheatMaize
Gene expression(Any)1984 [118]1972 [117]1971 [116]
Drought tolerance1988 [131]1991 [132]1991 [133]
Heat tolerance1991 [134]1992 [135]1993 [136]
NUE2006 [137]2008 [138]2006 [139]
Transcriptomics(Any)2001 [140]2002 [141]2003 [142]
Drought tolerance2006 [143]2009 [144]2007 [145]
Heat tolerance2005 [146]2007 [147]2015 [148]
NUE2006 [137]2008 [149]2009 [150]
Quantitative PCR(Any)2003 [151]2003 [152]1999 [153]
Drought tolerance2008 [154]2009 [155]2007 [156]
Heat tolerance2008 [154]2011 [157]2007 [156]
NUE2007 [158]2013 [159]2011 [160]
RNA-seq(Any)2010 [161]2011 [162]2011 [163]
Drought tolerance2015 [164]2014 [165]2012 [166]
Heat tolerance2015 [167]2015 [168]2015 [169]
NUE2018 [170]2014 [171]2015 [172]
Table 3. Year and bibliographic reference of the first documented use of mutagenesis, transgenesis and gene editing tools in rice, wheat, and maize.
Table 3. Year and bibliographic reference of the first documented use of mutagenesis, transgenesis and gene editing tools in rice, wheat, and maize.
ApproachTopicRiceWheatMaize
Mutagenesis(Any)1971 [212]1964 [213]1961 [214]
Drought tolerance, WUE2007 [215]2001 [216]2014 [217]
Heat tolerance2003 [218]2014 [219]2020 [220]
NUE2011 [221]-2006 [139]
Transgenesis(Any)1988 [222]1990 [223]1988 [224]
Drought tolerance, WUE1998 [225]2000 [226]2002 [227]
Heat tolerance2000 [228]2008 [229]2007 [156]
NUE1997 [230]2001 [231]2015 [232]
Gene editing(Any)2012 [233]2013 [234]2014 [235]
Drought tolerance, WUE2017 [236]-2017 [237]
Heat tolerance2020 [238]--
NUE2018 [239]-2020 [129]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Benavente, E.; Giménez, E. Modern Approaches for the Genetic Improvement of Rice, Wheat and Maize for Abiotic Constraints-Related Traits: A Comparative Overview. Agronomy 2021, 11, 376. https://0-doi-org.brum.beds.ac.uk/10.3390/agronomy11020376

AMA Style

Benavente E, Giménez E. Modern Approaches for the Genetic Improvement of Rice, Wheat and Maize for Abiotic Constraints-Related Traits: A Comparative Overview. Agronomy. 2021; 11(2):376. https://0-doi-org.brum.beds.ac.uk/10.3390/agronomy11020376

Chicago/Turabian Style

Benavente, Elena, and Estela Giménez. 2021. "Modern Approaches for the Genetic Improvement of Rice, Wheat and Maize for Abiotic Constraints-Related Traits: A Comparative Overview" Agronomy 11, no. 2: 376. https://0-doi-org.brum.beds.ac.uk/10.3390/agronomy11020376

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