Next Article in Journal
R-Spondin2, a Positive Canonical WNT Signaling Regulator, Controls the Expansion and Differentiation of Distal Lung Epithelial Stem/Progenitor Cells in Mice
Next Article in Special Issue
Combined Analysis of the Metabolome and Transcriptome to Explore Heat Stress Responses and Adaptation Mechanisms in Celery (Apium graveolens L.)
Previous Article in Journal
Neuroinflammation and COVID-19 Ischemic Stroke Recovery—Evolving Evidence for the Mediating Roles of the ACE2/Angiotensin-(1–7)/Mas Receptor Axis and NLRP3 Inflammasome
Previous Article in Special Issue
The Plastome Sequences of Triticum sphaerococcum (ABD) and Triticum turgidum subsp. durum (AB) Exhibit Evolutionary Changes, Structural Characterization, Comparative Analysis, Phylogenomics and Time Divergence
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Applications of Genomic Tools in Plant Breeding: Crop Biofortification

by
Inés Medina-Lozano
1,2 and
Aurora Díaz
1,2,*
1
Departamento de Ciencia Vegetal, Centro de Investigación y Tecnología Agroalimentaria de Aragón (CITA), Universidad de Zaragoza, Avda. Montañana 930, 50059 Zaragoza, Spain
2
Instituto Agroalimentario de Aragón—IA2, Centro de Investigación y Tecnología Agroalimentaria de Aragón (CITA), Universidad de Zaragoza, 50013 Zaragoza, Spain
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(6), 3086; https://doi.org/10.3390/ijms23063086
Submission received: 31 January 2022 / Revised: 4 March 2022 / Accepted: 10 March 2022 / Published: 13 March 2022
(This article belongs to the Special Issue Plant Genomics and Genome Editing)

Abstract

:
Crop breeding has mainly been focused on increasing productivity, either directly or by decreasing the losses caused by biotic and abiotic stresses (that is, incorporating resistance to diseases and enhancing tolerance to adverse conditions, respectively). Quite the opposite, little attention has been paid to improve the nutritional value of crops. It has not been until recently that crop biofortification has become an objective within breeding programs, through either conventional methods or genetic engineering. There are many steps along this long path, from the initial evaluation of germplasm for the content of nutrients and health-promoting compounds to the development of biofortified varieties, with the available and future genomic tools assisting scientists and breeders in reaching their objectives as well as speeding up the process. This review offers a compendium of the genomic technologies used to explore and create biodiversity, to associate the traits of interest to the genome, and to transfer the genomic regions responsible for the desirable characteristics into potential new varieties. Finally, a glimpse of future perspectives and challenges in this emerging area is offered by taking the present scenario and the slow progress of the regulatory framework as the starting point.

1. Introduction

Malnutrition is known to be a global public health problem and it has worsened with the COVID-19 pandemic. In 2020, about 768 million people in the world faced hunger, around 118 million more than in 2019 [1]. In addition, around 2.37 billion people (nearly one in three people in the world) suffered food insecurity (i.e., an inadequate access to safe, nutritious and sufficient food) in 2020, almost 320 million people more in just one year [1]. In fact, it is the first time that food insecurity has increased in North America and Europe since 2014 [1]. However, malnutrition is not only caused by the lack of food but also by a low dietary intake of essential nutrients (micronutrients included), known as hidden hunger [2]. This problem affects mainly developing countries in which the diet is usually based on more affordable major staple crops, characterized by a low micronutrient content. That being true, malnutrition is also present in developed countries, although in this case it is possibly due to unhealthy habits, such as extreme weight loss diets or substance abuse. It does not alleviate this situation given the fact that crop breeding has been mainly focused on increasing production, incorporating resistance to diseases, and enhancing tolerance to abiotic stresses, which has resulted in commercial varieties with low nutritional value [3].
Biofortification, i.e., the development of food crops with a high nutritional value per se through both conventional breeding and modern biotechnology techniques, could help in preventing hidden hunger. Micronutrients, minerals [4,5,6,7,8,9,10,11,12], vitamins [13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28], or both [29,], are the most common nutritional targets for biofortification strategies, though the improvement in fatty acid composition [30,31,32,33] and the increase in essential amino acids [34,35,36,37] and antioxidants [38,39,40,41] have also been recently included as aims of biofortification programs. This strategy carries multiple advantages. For example, it is a cost-effective approach, as shown by studies that report that for every dollar invested in the development of biofortified crops, as much as USD 17 of benefits may be obtained [42]. This is because, after a one-time investment to obtain the biofortified crops, they are able to synthesize larger amounts of the particular compounds without the need of adding any external micronutrients (fertilizers), which was the case in classical fortification. Therefore, as well as economic benefits, biofortification also brings environmental benefits. Moreover, it seems that breeding for a higher content in micronutrients does not entail a yield penalty [43,44]. This could be really helpful in developing countries, especially in areas with a limited access to marketed crops, as farmers could grow biofortified crops in the same way as conventional crops. Consequently, biofortification could be considered a sustainable and long-term solution to hidden hunger. In fact, the expected increase in population up to 9.7 billion by 2050 [45] makes it even more necessary.
Nevertheless, since the biofortification of a crop is tackled until the product is released to the market, a series of key steps have to be taken. The first would be to choose the species and the micronutrient to be enhanced. To maximize the positive impact on society, most consumed crops should be the target. This is what has been actually happening as, among the biofortified crops already developed, we can find staple crops, such as cereals (barley, maize, rice, and wheat) and beans, and some of the most consumed vegetables (tomato and potato) and fruits (apple and banana). One of the first steps consists of an evaluation of germplasm for their content in nutrients and health-promoting compounds; thus, outstanding alleles for those metabolic traits can be selected. Alternatively, the variability can be generated through induced mutagenesis (widely used in plant breeding since optimized during the second half of the 20th century) or by other more modern techniques of gene editing (i.e., clustered regularly interspaced short palindromic repeats (CRISPR)-associated system (CRISPR/Cas)). Secondly, genetic studies are usually conducted and molecular markers have to be developed to associate the trait of interest to the genomic regions. Finally, the allelic variants responsible for an increased content of the particular phytochemical have to be introduced to obtain the biofortified crop, either by conventional breeding or by modern biotechnology techniques. In this review, we will describe these steps in depth and, within the modern methods to introduce the allelic variants responsible for the increase in the specific compound, we will focus on transgenesis, cisgenesis, and intragenesis. Other simultaneous efforts will have to be made in order to ensure success both in the commercialization of the biofortified product and in the impact on consumers’ health. For the first goal, studies of market potential and consumers’ behavior and acceptability will have to be undertaken in advance, as was the case of selenium-biofortified apples [46] and iodine-biofortified fruits and vegetables [47], for example, both in Germany. This point is especially important in the case of controversial goods, such as transgenic biofortified food. That should be accompanied by promotion campaigns to make the product’s beneficial properties public, as the one carried out with the orange-flesh sweet potato biofortified in pro-vitamin A in Ghana and Nigeria [48]. For the second objective, analyses of micronutrient bioavailability and their efficacy of conversion in the human body will have to be performed, as reported in intervention studies which supply vitamin A-biofortified maize to Zambian children with promising results [49].
Taking all the above into account, the present review aims, firstly, to summarize the genomic tools available to explore the variability through single-nucleotide polymorphism (SNP) genotyping, and the analytical methods to determine the phytochemical profile and/or content of plant food. Secondly, a compendium of the researches carried out on the genomic association of metabolic data in crops is also presented here. Thirdly, different methods used to transfer the genomic regions responsible for a raise in the compound synthesis to the crops in order to create new biofortified varieties are shown, as well as some examples of their applications. These methods are either encompassed in conventional breeding strategies or modern biotechnology approaches, such as transgenesis, cisgenesis, and intragenesis. Finally, an overview of the current regulation and the future prospects of developing nutritionally enriched crops is also offered.
All the information needed to deal with the subjects mentioned above is obtained through searches in public databases and webpages, as described in Supplementary File S1.

2. Exploring Biodiversity: Searching for Outstanding Material

2.1. Genomic Diversity Enquired by SNP Genotyping

SNPs are not only the most frequent sequence variations among all practically genomes [50], but also the most amenable to automation. Even if a long list of molecular markers, and, more specifically, genetic markers, has been used in plant breeding since the 1980′s (restriction fragment length polymorphisms (RFLPs), random amplified polymorphic DNAs (RAPDs), amplified fragment length polymorphisms (AFLPs), simple sequence repeats (SSRs), inter-simple sequence repeats (ISSRs), cleaved amplified polymorphic sequences (CAPS), etc.) [51], all of them have been unarguably ousted by SNPs. Their predominance is also a consequence of the development of next-generation sequencing (NGS), including second- and third-generation sequencing (SGS and TGS), mainly SGS, which evolved from the sequencing of short DNA fragments (first-generation sequencing, FGS) to high-throughput technologies (SGS) and, finally, single-molecule sequencing (TGS). This soon made necessary high-throughput SNP genotyping platforms that could produce a massive volume of data more cost-effectively in a short period of time. Among the wide variety of techniques developed to genotype SNPs and the different detection methods coupled to them, we will highlight those more commonly used nowadays with crops and those that process a medium (normally, in the laboratory) to high number of markers and samples (commercial platforms). All of them are based on hybridization, amplification, sequencing, or a combination of them, and they have been grouped according to the type of platform employed (Figure 1).

2.1.1. SNP Genotyping Microarrays

Among the assays available, the Affymetrix (Axiom) is a hybridization-based microarray that uses probes for both alleles. Independent of the allele at the particular locus, both probes hybridize with the DNA sample though the signal become dimmer in the case of a mismatch. So, the genotype of each SNP marker is called by the probes, showing the highest intensity in their signal. SNP Affymetrix arrays (either Axiom or GeneChip) have been used in a number of food crops, including cereals (maize [52,53], rice [54,55], rye [56], and wheat [57,58]), horticultural crops (chickpea [59], lettuce [60], potato [61], soybean [62], and strawberry [63]), and woody crops (apple tree [64] and peanut tree [65]), among others.
In the Illumina BeadArray (Infinium), the silica beads are coated with probes targeting a specific SNP locus. They bind the region just upstream the polymorphic site. Then, by single-base extension (SBE), a labelled nucleotide will be incorporated, emitting a different signal depending on the base. Illumina developed other BeadArray (GoldenGate) that uses fluorescent universal primers that hybridize to the allele-specific oligos. These technologies have been extensively used to discover and genotype SNPs in food crops, including cereals (barley [66], maize [67,68], oat [69], rice [70,71], and wheat [72]), oil crops (oilseed rape [73] and sunflower [74,75]), horticultural crops (cowpea [76], potato [77], tomato [78], and soybean [79]), and woody crops (apple tree [80,81], cherry tee [82], peach tree [83,84,85], pear tree [86], and vine [87,88]), among others.
The immobilization of samples, probes, ddNTP, etc. on chips (depending on the technique) is what makes interrogating hundreds of thousands or even millions of markers simultaneously feasible (Figure 1). In both cases, there are predesigned chips for some crops, which is the most affordable choice, but there is also the possibility of designing custom chips with the SNP markers of interest.

2.1.2. Real-Time PCR for SNP Genotyping

One of the commercially available assays within this category is the TaqMan SNP genotyping. This technology is also based in DNA hybridization and amplification, the signal is generated by fluorescence resonance energy transfer (FRET), and it is amenable to automation by real-time PCR though it does not reach the same high-throughput format than microarrays (Figure 1). Briefly, two allele-specific probes are designed for each SNP locus with two different fluorescent dyes attached to them. When the probe is free, the fluorescence is suppressed by quenching. Only when the probe perfectly hybridizes with the DNA fragment containing the SNP allele and is extended by PCR, the fluorophore is released by the exonuclease activity of the DNA polymerase and its signal is captured by the appropriate detector. These techniques have been mainly used in plants to diagnose pathogens and, to a smaller extent, to identify transgenes and detect food frauds, though there are also some cases where they are used to study the genetics behind some traits of interest in food crops, such as the presence of anthocyanins in potato skin [89].
As the previous one, the Kompetitive allele-specific PCR (KASP) is also a FRET method that makes use of hybridization and amplification though, unlike the TaqMan assay, the reagents for the allele-specific amplification, on the one hand, and the dye and quenchers, on the other, act in two phases. During a first round of PCR, the allele-specific and the common reverse primer amplifies the region by harboring the target SNP. After this, one of the fluor-labelled oligo that was quenched until now binds as a tail to the corresponding amplified allele, generating a fluorescent signal. KASP assays have been extensively used in different crops, mainly cereals, becoming very helpful for MAS in wheat [90,91,92,93,94,95,96,97,98,99], barley [100], rice [101,102,103,104], sorghum [105], pea [106], watermelon [107,108], faba bean [109], tomato [110,111], and Brassica oleracea (cabbage, broccoli, kohlrabi, and Chinese kale [112]).
Another methodology included here is the high-resolution melting (HRM) analysis. After the amplification by PCR of the region containing the SNP of interest in the presence of a dye that binds to double-stranded DNA, the products are melted into a single strand. This then causes the release of the dye and a decrease in its fluorescence. The real-time PCR is able to detect those changes and generate a melt curve that is different for each of the genotypes at the SNP locus. Apart from cultivar identification, species authentication and pathogen diagnose, HRM has also been used for MAS to enhance the quality of soybean [113], rice [114], strawberry [115], and barley [116].
These methods normally do not reach the same high-throughput format than microarrays (Figure 1). However, nowadays, there are TaqMan and KASP arrays which help to process a high sample throughput for mid-density genotyping. In the case of TaqMan SNP genotyping, there are pre-designed and custom assays. Regarding HRM, as of recent, there are no commercial panels; however, it is the user who is in charge of designing and carrying out the assays. In the case of HRM and TaqMan (but not KASP) analyses, a low degree of multiplexing is possible (i.e., duplex).

2.1.3. Mass Spectrometry SNP Genotyping

Primers are designed immediately adjacent to the SNP locus and an SBE is carried out using mass-modified dideoxynucleotide terminators. The mass of the allele-specific product is determined by using matrix-assisted laser desorption–ionization time-of-flight (MALDI-TOF) mass spectrometry. Like all the other SNP genotyping technologies, this is used with identification purposes in crops. Besides, it is applied in MAS for quality traits in cereals, such as barley [117], rice [118], legumes (including pea) [119], and mung bean [120].
This is a high-throughput technology (Figure 1) which can process thousands of samples per day, which also allows the simultaneous amplification and detection of multiple markers per reaction (i.e., Agena iPLEX Gold, previously known as Sequenom iPLEX Gold). This method avoids the problems derived from a background signal typical from those based on hybridization. As the previous ones, this type of assay can be custom-designed.

2.1.4. SNP Analysis by NGS

With the increasing affordability of sequencing methods, these SNP genotyping platforms based on sequencing are becoming very popular. The main strategy nowadays consist of building reduced representation libraries (RRLs). By reducing the complexity of the targeted genome (normally digesting it with restriction enzymes), the depth of the sequencing can be increased. Among all the developed methods, including restriction site-associated DNA sequencing (RAD-Seq), diversity array technology sequencing (DArT-Seq), restriction fragment sequencing (REST-Seq), multiplex shotgun genotyping (MSG), sequence-based genotyping (SBG), specific-locus amplified fragment sequencing (SAF-Seq), etc., one of the most widely used in crops is genotyping by sequencing (GBS). Briefly, the whole genome is fragmented using restriction enzymes and short-read sequencing is performed on the ends (paired-end sequencing). Libraries for each sample are prepared using different barcodes; thus, a multiplex approach in which thousands of genotype SNPs across thousands of samples simultaneously was possible (Figure 1). GBS is used in studies on some traits that influence the nutritional value of food crops, such as the soluble solid content in plum [121]; sugar and acid content in apple [122]; sugar and carotenoid content in melon [123]; and certain mineral content in maize [124], pea [125], and spinach [126].
As in the previous technologies, pre-designed assays are available for some crops though custom panels of markers are also possible.
Thanks to this profusion of technologies that are becoming more and more affordable, a large number of SNP databases in crops is made available (Table 1). The data that have been made public in this way feed back into the agrigenomic field, as they can be used by other researchers to design their assays. Some of them only include marker information, but others also supplied the genotypes in different accessions (cultivars and wild crops relatives) as well as other useful tools, including genetic maps, genome sequences, etc. Table 1 clearly shows a higher representation of staple crops (i.e., cereals), given the very intense genetic breeding in recent decades, though other crops with a great economic importance, such as fruit trees (i.e., within Rosaceae family) or vegetables (i.e., tomato), are also present.
For the above, SNPs are the preferred markers to both, carry out genetic studies and undertake breeding programs in crops. Actually, genotyping assays have been developed for a large number of plants, including all major crops.

2.2. Nutritional and Phytochemical Profiles Assessed by Analytical Methods

The “omic” era has also reached the characterization of food plants in terms of their nutritional content, making use of metabolomic technologies. Thus, it is now possible (though still prohibitive, in many cases) to obtain the complete profiles of phytochemicals in complex extracts in a high number of samples. In this way, the compounds are identified by metabolic profiling and then quantified by target analysis. This has huge potential in plant breeding, especially in crop biofortification, which is still to be fully exploited. The different techniques normally used for metabolome analysis are enlisted here very briefly, as that is not the main scope of this review.

2.2.1. Mass Spectrometry (MS)

This is a very sensitive analytical technique, either used directly (non-hyphenated methods) or coupled with others (hyphenated methods), such as gas chromatography (GC), liquid chromatography (LC), or capillary electrophoresis (CE). In the first case, it is possible to process a high number of samples in a short period of time, though the identification capacity is limited. The hyphenated methods, on the other hand, are undoubtedly more powerful when it comes to identifying and quantifying metabolites, and there is also the possibility of reducing the running times by using more advanced techniques in chromatography (i.e., ultra-high-performance liquid chromatography (UPLC) instead of high-performance liquid chromatography (HPLC) [129]). In any case, the metabolite identification generally requires the availability of libraries in order to compare the spectra obtained.

2.2.2. Nuclear Magnetic Resonance (NMR)

This is a very reproducible spectroscopic technique used to quantify metabolite levels. It allows a high-throughput process of samples, though it is generally less sensitive and has less resolution power than MS. Moreover, it is a non-destructive method, which makes it the perfect choice for studying the metabolome evolution (for instance, in different plant stages), instead of simply obtaining a snapshot of the plants at a particular moment.
Both techniques can be actually combined, resulting in the detection of a higher number of metabolites.
Until recently, the most common nutritional studies in food crops have focused on the quantification of a discrete number of compounds with a high impact in their nutritional value (targeted metabolic studies), though some widely targeted metabolomics analyses are starting to be carried out even in minor crops [130]. The initial steps which deal with the germplasm evaluation for nutrients and health-promoting compounds are essential for harnessing the biodiversity harbored by cultivated varieties, but also by breeding material and crop wild relatives. Some examples of these characterization works can be found in all groups of food crops, cereals [131], fruits [132], legumes [133], and vegetables [134,135], among others. In this sense, a considerable number of researches has compared different plant material within the same crop (for instance, landraces vs. commercial varieties) in order to identify outstanding accessions for future breeding programs aimed at enhancing the content of nutritious and beneficial compounds (reviewed in [3]). Metabolomic offers the opportunity to study the huge range of metabolites present in a sample (untargeted metabolic studies) and not only some specific compounds.
Another metabolome approach, apart from profiling commented above, consists of performing metabolomic fingerprints, where compounds are not individually identified. However, the metabolite profiles are compared among samples, for instance, to study the plants at different developmental stages [136] or under several biotic [137] and abiotic [138] stresses. We will not go into depth in the latter, as it is not related to the subject of this review, though it is noteworthy to mention that some studies use a combination of both approaches, i.e., by carrying out metabolomic fingerprint experiments in which the compounds are actually identified [136].

3. Association between the Traits of Interest and the Genomic Regions: Fishing for Genes

On one hand, one of the most useful and exploited genetic tools in crop breeding has been the linkage maps. Large SNP genotyping arrays have been used to build high and ultra-high-density genetic maps that allow the efficient marker-assisted selection (MAS) of beneficial alleles for the traits of interest. Nowadays, there are consensus and saturated genetic maps (mainly built with SSR and SNP markers) in virtually all the important crops and, in many cases, they are used to localize quantitatively trait loci (QTL). This fine mapping (often together with the QTL analysis) has led to the identification and cloning of the underlying gene(s), mainly in cereals (i.e., barley, maize, rice, and wheat), but also in some legumes (i.e., soybean) and vegetables (i.e., tomato) [139], though there are few cases for traits related to their nutritional value. An emerging application involves integrating metabolic/metabolomic and quantitative data to render metabolic QTL (mQTL). Until now, a number of these studies have been carried out, mainly in cereals (wheat, barley, rice, and maize) but also in oilseed rape and tomato [140]. As a result, numerous mQTL have been identified in those crops and some of them have eventually led to the identification of putative candidate genes controlling metabolic traits [140].
On the other hand, in genomics (the field that concerns us in this review), the whole genome of an organism is studied. As could be expected, the development of NGS technologies has led to a real boost for its applications, such as genome-wide association studies (GWAS). With the SNP genotyping by NGS, it is possible and affordable to rapidly scan markers across the complete genome of many individuals to find variations associated with a particular trait. In fact, the genotypes for thousands of SNPs are currently available for many crop species, as shown in Table 1. In order to make the most of all this already existing information, it can be combined with the results derived from the technology to analyze metabolites. In this line, researches which combine metabolic/metabolome and genome association results (metabolic/metabolomic GWAS, mGWAS) are starting to be carried out in crops (Table 2) and they are expected to become very helpful in genomic-assisted breeding programs by whole-genome selection and eventually in identifying some of the genes potentially influencing the nutritional value and the content of health-promoting compounds.
A potential drawback of this methodology, especially in the case of complex traits (as is the case of metabolism-related traits), is that the most significant variant obtained (i.e., allele of a SNP) is sometimes not responsible for metabolic differences. Actually, it is also common, as in any statistical analysis, to obtain spurious associations, for instance, when the trait heritability is low (high environmental effect). For this reason, it will still be necessary to carry out the validation of the candidate genes identified. In this sense, in many of those mGWAS involving compounds with a potential use to biofortify the respective crop (Table 2), other “omics” technologies, mainly transcriptomics, have assisted researchers in untangling the relationships between genotype and phenotype and in pinpointing the causal gene(s). Furthermore, it is also common to validate those findings by using mutants (knockout and/or overexpressing lines) and transgenic plants. Such an encompassing approach will undoubtedly speed up the process of obtaining healthier and nutritionally richer crops. Even if it is not the purpose of many of those studies, aimed at evaluating the metabolic changes that plants undergo during their development or to face environmental challenges (i.e., biotic and abiotic stresses), that knowledge about the genes responsible for the changes in metabolite contents is applicable in order to enhance the food in phytochemicals with beneficial properties.

4. Introducing Allelic Variants to Biofortify Crops

The last stages of the biofortification process in crop plants can be tackled through different approaches, including both conventional and modern biotechnology techniques, such as transgenesis, cisgenesis, intragenesis, or gene editing (i.e., CRISPR/Cas), in order to introduce genetic variation into the gene pool of the crop. Here, we will describe conventional breeding, transgenesis, cisgenesis, and intragenesis, as well as their applications in crop biofortification.

4.1. Conventional Breeding Assisted by Genomic Tools

Biofortification through conventional breeding is based on crosses within a sexually compatible group, specifically between donor plants with nutritional properties of interest and recipient ones with good agronomic characteristics. Many types of populations have been developed to perform genetic mapping, QTL identification, and association studies (i.e., both temporal (F2, backcrosses (BCs) and advance backcrosses (ABs)) and immortal (double haploid lines (DHLs), recombinant inbred lines (RILs), near isogenic lines (NILs), multi-parent advanced generation inter-cross (MAGIC), and nested association mapping (NAM)) ones). Among them, the most widely used in plant breeding to introgress DNA regions that harbor beneficial alleles for the trait of interest from the donor into the recipient parent are ABs, NILs, and RILs (Figure 2).
Several generations are needed, so it takes a substantial amount of time to obtain crops with the desired nutritional and agronomic characteristics by using this strategy. An emerging alternative approach to save time, effort, and money consists of carrying out a genomic selection (GS), as coined by Meuwissen et al. [183], based on a genomic prediction (GP) (Figure 1D). Instead of phenotyping at every stage of the population building (like in the MAS strategy), it is only carried out in what is known as training population (DHs, F2, marker-assisted recurrent selections, etc.). These data, together with genome-wide genotypic data from that same training population, are used to calculate the genomic estimated breeding value (GEBV) through processes of machine learning by means of different regression models. So, GEBV is a parameter used to quantify the genetic merit of a certain individual in order to improve the crop in the trait of interest. Finally, the GP is carried out with the data coming from genotyping the testing population (the breeding population) without the need to phenotype it. In this way, the individuals selected by the testing population are expected to show a genetic gain, i.e., an increase in performance thanks to the gene variant(s) responsible for the aforementioned trait. With this method, all markers are taken into account, not only those which show a significant association with the trait (as in MAS); thus, loci with little additive effects can also be detected. Until now, this approach has been scarcely used in crops for metabolite and nutritional content, such as in tomato [184] and wheat [185].
Furthermore, the chances of achieving biofortification by conventional breeding depends on the crop itself, since the strategy relies on the genetic variability available within its gene pool, which is usually limited in commercial varieties. This could be overcome by crossing plants with landraces or with more distant wild relatives that normally harbor higher genetic variability and, sometimes, can be richer in nutrients [3,186]. However, in some cases, it would be impracticable to obtain biofortified crops using conventional breeding. That would be the case when the genetic variability needed for a specific trait is insufficient within the gene pool, or when the investment of time and resources would be excessive, especially with non-diploid species, when the trait heritability is low or when linkage drag is unavoidable.
In spite of its limitations, conventional breeding is currently the most accepted method, as it is sustainable and it is not subject to regulatory obstacles. Nowadays, an important number of crops have been conventionally bred to enhance their nutritional content. In fact, several international organizations have initiated different programs to accomplish this objective. Harvest Plus, launched in 2003, is the most important one and is focused on enhancing the content of provitamin A, iron, and zinc in staple food crops across Asia and Africa [187]. It has managed to biofortify a large number of crops, many of which have been already released. Until 2019, there is a total of 242 across 30 developing countries [188]. Different studies have demonstrated the efficacy of biofortification through conventional methods, specifically increasing the content of micronutrients [189,190]. Furthermore, other smaller institutions are working on developing conventionally biofortified crops. For example, the International Potato Centre (CIP) has obtained, tested, and advertised an orange sweet potato enriched in provitamin A [191], and the International Maize and Wheat Improvement Centre (CIMMYT) has released different hybrid varieties with increased levels of the amino acids lysine and tryptophan through the incorporation of the naturally occurring mutation opaque-2 (o2) into different maize varieties [192].
The assistance of genomic tools has facilitated the development of many conventional biofortified crops, as they allow breeders to exploit the available genetic variability more efficiently; thus, time and costs can be significantly reduced. Plant breeding has existed since plant domestication started around 10,000 years ago, and the selection carried out at the beginning merely attends to the phenotype. However, with the application of genetic and genomic tools, genetic variants can be associated with differences in phenotypes, which then enables the selection at early stages of the plant. For that, the construction of genetic maps has been essential, as previously mentioned. Many studies have found markers linked to genes or QTL which can control the content of nutritional compounds, for example, those related with carotenoid variation in sorghum [193], mineral micronutrients in beans and wheat [194,195], vitamins levels in different cereal crops [196], etc. Thus, individuals with the best gene combination have been identified and used as potential donors in breeding programs to enhance the content in micronutrients (minerals and vitamins) and health-promoting compounds (polyphenols, carotenoids) in all kinds of crops, including cereals, fruits, legumes, and vegetables (Table 3).

4.2. Modern Biotechnology Techniques

4.2.1. Transgenesis

In biofortification, transgenic approaches consist of the transference of one or more alleles from genes responsible for the increase in the nutritional value from one or more organisms to the crop of interest. They are really helpful in overcoming the main handicap of conventional breeding, i.e., the limited genetic variation within the same or sexually compatible species [207]. Moreover, genetic transformation through transgenesis can achieve the expression of a gene independently of its origin, in terms of evolution, taxonomy, and even kingdom [19,208,209]. Hence, when a specific nutrient or a bioactive compound is not naturally synthesized in a crop, transgenesis is the only way to engineer the crop to produce it. Therefore, this strategy helps to exploit a much larger gene pool and transfers more than one gene and their regulatory regions simultaneously (Figure 3A). In this way, the crop can be enriched in more than one nutrient at the same time, as it has already been successfully engineered in rice [29]. However, it is important to take into account that some crops are recalcitrant to transformation and/or regeneration, for example, some cereals [210] or legumes [211].
Transgenic approaches require a lot of time and resources. The identification and characterization of the gene(s) are needed to eventually introduce them in the crop. Nevertheless, transgenesis is less time-consuming than the conventional alternative and more cost-effective than the agronomic fortification, which is ineffective in the long term because it requires regular applications of fertilizers [212]. This, together with the absence of taxonomic constrictions and the possibility of designing almost any synthetic gene, has resulted in a big number of biofortified crops developed through transgenic strategies (Table 3). One of the most remarkable examples is Golden Rice, obtained to alleviate the vitamin A deficiency [17]. It was the first application of transgenic biofortification, in which a carotenoid-free rice endosperm was genetically engineered to produce β-carotene (provitamin A) by expressing the genes codifying for the phytoene synthase and the carotene desaturase [17]. In addition, a clinical trial in humans has demonstrated that Golden Rice could be an alternative source of vitamin A for adults [213]. As in the case of conventional breeding, many different strategies have been applied to almost any kind of crop, including cereals, legumes, vegetables, fruits, and oilseeds, whereby the targets of biofortification are fatty acids, essential amino acids, and antioxidants, among others (Table 3).
The main disadvantage of these crops is the strict regulation to which they are subject to, at least, in Europe (more deeply described further on). However, some biofortified crops have gone beyond this limitation and they have been released. Some of these crops are cassava with improved levels of zinc, iron, β-carotene, or proteins, released by Biocassava Plus; canola with a higher availability of phosphate due to phytate degradation, released by BASF; and linseed enhanced in essential amino acids, released by the University Saskatchewan (Saskatoon, Canada).

4.2.2. Cisgenesis and Intragenesis

Cisgenesis and intragenesis are approaches that, to some extent, were developed to overcome the main limitation of transgenesis—its strict regulation [214]. The gene pool exploited here can only come from naturally crossable species; therefore, they might be a suitable alternative to obtain biofortified crops.
On the one hand, the terms “cisgenic plant” were first introduced in 2006 as “a crop plant that has been genetically modified with one or more genes (containing introns and flanking regions such as native promoter and terminator regions in a sense orientation) isolated from a crossable donor plant” [215] (Figure 3B). This donor plant has to belong to the same species than the modified crop or to a sexually compatible species; thus, the gene pool available for cisgenesis is identical to the gene pool exploited by conventional breeding. Nevertheless, unlike conventional breeding, only the gene(s) of interest, and no undesired sequences (linkage drag), are transferred to the final cisgenic crops.
On the other hand, the terms “intragenic plant” were introduced in 2004 and they refer to the isolation of specific genetic elements from a plant, the recombination of these elements in vitro, and the insertion of the resulting expression cassettes into a sexually compatible plant [216] (Figure 3C). Intragenesis can also be carried out using constructs with RNA interference (RNAi) [28,206] or genes edited, for instance, by CRISPR/Cas, as this technology has been successfully used to edit the genome of crops [217,218]. Therefore, intragenesis provides the possibility of creating novel combinations that render higher variability and novel expression patterns to develop new genetically modified organisms (GMOs) with new properties that will not happen spontaneously in nature or through conventional breeding.
The main difference between cisgenesis and intragenesis is related to the regulatory regions. In cisgenesis, the transgene is a complete DNA copy of the gene as it can be found in the donor plant (with promoter, introns, and terminator) in the normal-sense orientation (Figure 3B). In intragenesis, there is not any requisite about these regulatory elements, as long as all the genetic elements come from crossable donor plants, so that they can be engineered before being used in the transformation (Figure 3C). Consequently, intragenesis is not considered as close to conventional breeding as cisgenesis.
In both cases, when Agrobacterium-mediated transformation is used, T-DNA borders (flanking sequences of the DNA to be transferred) can be also inserted in the plant genome. This is a controversial topic as some authors are in favor of using T-DNA borders, claiming that they are safe because they are short non-coding sequences that can be found in plant genomes naturally too [219]. The evident argument against T-DNA borders is that all DNA sequences integrated into the recipient plant should come from a sexually compatible DNA pool, as established by both cisgenesis and intragenesis definitions [215,216]. Thus, both cisgenic and intragenic crops should be free of those T-DNA borders, and also of selection markers and vector backbones, as both of them are supposed to be genetically modified plants that do not contain foreign genes (only genes coming from cross compatible species). Two alternative solutions have been proposed. First, plants without T-DNA borders can be selected just by carrying out a PCR. In fact, the integration rate of the T-DNA borders in the plant genome is relatively low, as is the case of transgenic potatoes carrying R genes for late blight, in which only 45% of transformants possessed T-DNA borders [220]. Second, T-DNA border-like sequences found in the plant genomes, known as P-DNA borders, can be used upstream and downstream the gene to be transferred [216,221]. A rearrangement of the original gene is thus required, as it was in the donor plant, which is why this option should only be chosen in the case of intragenic plants. Furthermore, the presence of T-DNA borders in both types of plants could be a problem for the public acceptance and in terms of regulation [222]. Regarding the other non-plant sequences, the use of selection markers is not necessary when the transformation efficiency is high [223] or the product codified by the introduced gene can be visually detected, including a pigmented compound (i.e., carotenes, anthocyanins) [224]. There are also methods to eliminate markers based on site-specific recombination (marker genes are flanked by specific recombination sites) [225], or by carrying out a co-transformation, which allows the segregation of the transgene and the marker gene in the progeny, as they are integrated in different positions of the genome [226].
In comparison to transgenesis, cisgenesis and intragenesis have two clear limitations (Table 4). The first one is that the available variability only exists in plants from the same sexual compatibility group, as in conventional breeding. However, this disadvantage could be overcome, to some extent, by gene edition (in the case of intragenesis) or by making use of the higher biodiversity present in landraces [3] or wild relatives [186]. The second limitation is the need to remove the selection markers and the vector backbones, which could be both time- and labor-consuming. On the other hand, although the three technologies are subject to the same regulation, cisgenic and intragenic crops are more accepted by the general public than transgenic ones [227,228,229].
When compared to conventional breeding, cisgenesis and intragenesis are considered fast alternatives to transfer genes between plants from the same sexual compatibility group, especially for species with long lifetimes and high heterozygosity levels (Table 4). Additionally, these two approaches are able to avoid linkage drag issues associated with backcrosses in conventional breeding, as only the sequences of interest are transferred (Table 4). Changes in the gene expression levels can also be achieved with both techniques (Table 4). The introduction of the complete natural gene (cisgenesis) and changes in promoters and terminators (intragenesis) may increase the levels of expression, whereas the use of silencing constructs (intragenesis) could reduce them. Moreover, new genetic variability can be generated with different combinations of genetic elements with intragenic approaches.
Although most of the new traits incorporated to relevant crops through cisgenesis and intragenesis are related to disease resistance [216,225] and abiotic stress tolerance [230], these strategies have been also applied with biofortification purposes (Table 3). For example, Holme et al. [202] obtained a cisgenic barley by inserting copies of a barley phytase gene (HvPAPhy_a). Those barley plants with a single copy of the gene showed a 2.8-fold increase in the phytase activity and an enhanced bioavailability of phosphate. A cisgenic potato was developed by suppressing the starch phosphorylase L gene through dsRNAi (double-strand RNA interference) technology to decrease starch degradation [203]. Then, the accumulation of reducing (glucose, fructose) and non-reducing (sucrose) sugars was lower in tubers stored at 4 °C. Finally, cisgenic red-fleshed apples, rich in anthocyanins, were developed by expressing the MdMYB10 gene, a transcription factor involved in anthocyanin biosynthesis flanked by its native promoter and terminator [41]. In the case of intragenesis, potato is the most recurrently used crop for gene silencing strategies. In fact, the first intragenic application was the increase in amylopectin content in potato by silencing the granule-bound starch synthase gene (GBSS), responsible for the synthesis of amylose in potato [204]. The silencing construct contains an antisense GBSS gene composed of only potato sequences and is controlled by the potato GBSS promoter. However, the terminator is the one of the nopaline synthase gene (nos) from A. tumefaciens; thus, this crop could not be considered as completely intragenic. Nevertheless, this potato was released to the field in the EU in 2007 (B/NL/07/04) with the potato GBSS terminator, i.e., a fully intragenic potato plant. Another intragenic potato was engineered to reduce the acrylamide content in processed potatoes (without yield penalty or affecting the tuber shape) by silencing one asparagine synthase gene (StAs1) [205]. The development of other intragenic potatoes was achieved by overexpressing the lycopene b-ciclase (StLYCb) gene controlled by the potato GBSS promoter, which incited β-carotene accumulation in potato tubers [27]. In the case of tomato, carotenoid and flavonoid contents were enhanced simultaneously through the suppression of the DE-ETIOLATED1 (DET1) gene by using RNAi technology and fruit-specific promoters [28]. A gluten-free wheat has also been obtained using this technology by silencing a γ-gliadin gene [206]. The iron content in wheat flour has been increased by more than 2-fold following the expression of a vacuolar iron transporter gene (TaVIT2) under the control of a wheat endosperm-specific promoter [12]. Finally, Dupont-Pioneer and Monsanto have developed two high oleic soybean oils, Plenish® and Vistive® Gold, respectively, which are currently available in the USA market.

5. Regulation of Plant Breeding Methods

The current regulatory framework could present an obstacle when the above-described techniques are used in crop biofortification, except for conventional breeding, which is not subject to any specific law. However, this is not the case for modern biotechnology techniques. Genetically modified (GM) crops have been demonstrated to be safe countless times, as supported by more than 100 Nobel laureates [231]. In addition, thousands of risk assessments conducted by independent federal regulatory agencies on GM crops have found that there is not different risks between GM and non-GM crops [232]. Nevertheless, there is a widespread lack of acceptance associated with the artificial combination of foreign genetic elements and the use of antibiotic or herbicide resistance selectable markers. All this has triggered alerts about potential health and environmental risks in case gene flow from GM to other non-GM crops [233]. Furthermore, the legislation continues to be strict and differs largely in each country.
In 2019, genetically engineered crops were cultivated in 29 countries, covering a total of 190 million hectares worldwide [232]. North and South America are the biggest producers, followed by Asia, where the law is more flexible. In fact, out of these 190 million hectares of biotech crops cultivation, 174 (90% of the total area) are located in only five countries: USA, Brazil, Argentina, Canada, and India (sorted in descending order) [232]. In the case of the European Union (EU), GMO regulation is one of the most severe, since it assumes that GM crops are intrinsically different (potentially dangerous) [234]. Thus, most countries have used the opt-out clause in relation to the GM crop cultivation and only six countries allow it, having permitted only the cultivation of a GM crop, Bt maize. This led to a decline in research and development (RD) investment in Europa from one-third of the global expenses in agriculture in the mid-1990s to less than 10% by 2013 [235]. Nevertheless, it is worthy to remark that, in England, the rules have been recently relaxed as a consequence of Brexit. Field trials of gene-edited crops with research purposes will be allowed without the current impediments and “red tape”, being only necessary to notify it to the Department for Environment, Food, and Rural Affairs (DEFRA) (https://www.gov.uk/government/news, accessed on 3 March 2022). In addition, these measures are likely to be extended to the rest of UK and a redefinition of the law about genetic modification is also expected. However, until then, gene-edited plants will still be considered GMOs and their commercial cultivation will have to be authorized under the actual law. In many African countries, there is either not any regulatory framework, or it is very restrictive, in spite of being regarded as the part of the world with the largest potential to benefit from the adoption of GM crops due to the high rates of hunger and malnutrition. Notwithstanding, the number of countries embracing GM crops in this continent has been doubled from three in 2018 to six in 2019 [232].
Despite the huge number of developed crops with enhanced traits through genetic engineering, only four different biotech crops cover more than 95% of the cultivated area (soybean, maize, cotton, and canola) and, in most cases, the modified traits are related to herbicide tolerance and insect resistance [232]. Therefore, additional efforts are needed to approve GM crops with enhanced nutritional value in order to contribute to the end of world hunger. Nowadays, transgenesis, cisgenesis, and intragenesis are subject to the same regulation in the vast majority of countries. However, cisgenic and intragenic crops are generally more accepted by the general public and are expected to be regulated less severely in the coming years in some countries [236]. In fact, in Canada, the regulation system is based on the final product rather than on the process to obtain it, which has relaxed the control of these kinds of crops in comparison with the transgenic ones [237]. In Australia, cisgenic plants are not considered GMOs, as stated in Gene Technology Regulations, whereby organisms that are not GMO include “a mutant organism in which the mutational event did not involve the introduction of any foreign nucleic acid” [238]. Other countries are also evaluating cisgenic and intragenic crop regulation. For example, in 2012, the European Food Safety Authority (EFSA) proposed a less precautionary approach to regulate cisgenesis, as it is supposed to entail similar hazards to conventional breeding as introduced by unmodified genes [239]. In the case of intragenesis, the EFSA affirmed that hazards are less predictable due to the recombination of different genetic elements, despite belonging to the same gene pool [240]. However, crops developed by RNAi technology, considered an intragenic approach, have recently received a positive opinion from this organization after determination of risk assessments [240]. In USA, the Environmental Protection Agency (EPA) is also discussing a less strict regulatory framework for cisgenesis and intragenesis approaches, especially when enhanced traits are related to pest resistance [241]. Furthermore, a lot of studies have confirmed a higher consumer and farmer acceptance of cisgenic and intragenic crops than transgenic ones because they are considered to be more natural [227,228,229]. This, together with the favorable opinions about cisgenesis and intragenesis from public organizations, should pave the way to less stringent regulations for these types of crops. Furthermore, a recent worldwide study has shown that consumers are willing to pay up to 23.9% more for GM-biofortified crops [242].

6. Future Perspectives

The Sustainable Development Goal 2 of the United Nations (UN) consists of ending all forms of hunger, including hidden hunger, before 2030. Nevertheless, projections show that unless serious actions are taken to accelerate the process, hunger will not be eradicated by that year. In fact, current progress is stalled or worsening [1]. Biofortification could substantially help to achieve that objective, as there are cost-effective strategies available. The technologies to explore genomic (i.e., SNP genotyping) and metabolic diversity are evolving astonishingly fast and becoming more and more high-throughput and, at least in the first case, affordable. Similarly, the approaches to identify and introduce the genomic regions responsible for the crop biortification, in this case, are becoming more accurate. However, all of them present some limitations, as we have discussed before. In the case of conventional breeding, the lack of genetic variability and the investment of time, although alleviated by the use of genomic tools (Table 4), to some extent, make it an insufficient strategy to reach the expected food demands [243]. Modern biotechnological techniques allow us to overcome those hurdles, though they are hampered by regulatory barriers, either non-existing specific laws or especially strict ones, as described in the previous section. Technology is progressing faster than the regulations and this gap is holding us back, for instance, to achieve the UN Sustainable Development Goals.
In parallel, an effort to illuminate the safety of genetically engineered crops in a clear and understandable manner is essential in order to increase their acceptance among the general public and political organizations. It would be also interesting to improve research and development of biofortified crops in developing countries, where malnutrition is a real burden.

7. Conclusions

Considering the expected increase in population in the next years, the challenge is not only to produce enough quantity of food to feed the global population, but also to ensure that food is nutritionally rich to ensure balanced diets. It is well established that biofortification is a cost-effective strategy and a promising approach to fight against global hunger, especially in developing countries. Currently, a large number of biofortified crops have been developed and even released, mainly those obtained through conventional breeding, but also some of them through modern biotechnological techniques. Nevertheless, GMO rejection implies an obstacle and it is frequently based on political preferences in spite of scientific evidences that support the safety of GM-biofortified crops. Here, it is necessary to set aside political and populist views not built on scientific results in order to guarantee food security, a global priority matter. Thus, the likely approval of cisgeneic and intragenic crops, and the less likely but also possible approval of transgenic ones, combined with conventional breeding and genome editing technologies, would place us closer and faster to the zero-hunger goal.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/ijms23063086/s1, Supplementary File S1: Bibliographic search criteria to elaborate the present review.

Author Contributions

Conceptualization, A.D.; writing—original draft preparation, A.D. and I.M.-L.; writing—review and editing, A.D.; supervision and funding acquisition, A.D. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the National Institute for Agricultural and Food Research and Technology (INIA) through the project RTA2017-00093-00-00, and by the Government of Aragón through the project LMP164_18 and the funds granted to the consolidated research group A12_20R: “Grupo de investigación en fruticultura: caracterización, adaptación y mejora genética”. I.M.L. was supported by a predoctoral contract for training doctors from the Spanish Ministry of Science, Innovation and Universities (MCIU), and the Spanish State Research Agency (AEI).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We thank J.R. Bertolín from “Laboratorio de Valoración Nutritiva (Department of Animal Science, CITA)” for his expert advice on the analytical methods included in this review.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the writing of the manuscript, or in the decision to publish it.

References

  1. FAO; IFAD; UNICEF; WPF; WHO. The State of Food Security and Nutrition in the World 2021. Trasforming Food System for Food Security, Improved Nutrition and Affordable Healthy Diets for All, 1st ed.; FAO: Rome, Italy, 2021; ISBN 978-92-5-134325-8. [Google Scholar]
  2. UNICEF. First call for children. In World Declaration and 1990–2000 Plan of Action on the Survival, Protection and Development of Children; UNICEF: New York, NY, USA, 1990. [Google Scholar]
  3. Medina-Lozano, I.; Díaz, A. Nutritional Value and Phytochemical Content of Crop Landraces and Traditional Varieties. In Landraces—Traditional Variety and Natural Breed; Elkelish, A., Ed.; IntechOpen: Rijeka, Croatia, 2021; pp. 95–116. ISBN 978-1-83968-718-1. [Google Scholar]
  4. Paul, S.; Ali, N.; Datta, S.K.; Datta, K. Development of an Iron-enriched High-yieldings Indica Rice Cultivar by Introgression of A High-iron Trait from Transgenic Iron-biofortified Rice. Plant Foods Hum. Nutr. 2014, 69, 203–208. [Google Scholar] [CrossRef] [PubMed]
  5. Neelam, K.; Rawat, N.; Tiwari, V.K.; Kumar, S.; Chhuneja, P.; Singh, K.; Randhawa, G.S.; Dhaliwal, H.S. Introgression of group 4 and 7 chromosomes of Ae. peregrina in wheat enhances grain iron and zinc density. Mol. Breed. 2011, 28, 623–634. [Google Scholar] [CrossRef]
  6. Haynes, K.G.; Yencho, G.C.; Clough, M.E.; Henninger, M.R.; Sterrett, S.B. Genetic Variation for Potato Tuber Micronutrient Content and Implications for Biofortification of Potatoes to Reduce Micronutrient Malnutrition. Am. J. Potato Res. 2012, 89, 192–198. [Google Scholar] [CrossRef]
  7. Blair, M.W.; Astudillo, C.; Grusak, M.A.; Graham, R.; Beebe, S.E. Inheritance of seed iron and zinc concentrations in common bean (Phaseolus vulgaris L.). Mol. Breed. 2009, 23, 197–207. [Google Scholar] [CrossRef]
  8. Aluru, M.R.; Rodermel, S.R.; Reddy, M.B. Genetic modification of low phytic acid 1-1 maize to enhance iron content and bioavailability. J. Agric. Food Chem. 2011, 59, 12954–12962. [Google Scholar] [CrossRef] [PubMed]
  9. Beasley, J.T.; Bonneau, J.P.; Sánchez-Palacios, J.T.; Moreno-Moyano, L.T.; Callahan, D.L.; Tako, E.; Glahn, R.P.; Lombi, E.; Johnson, A.A.T. Metabolic engineering of bread wheat improves grain iron concentration and bioavailability. Plant Biotechnol. J. 2019, 17, 1514–1526. [Google Scholar] [CrossRef] [PubMed]
  10. Narayanan, N.; Beyene, G.; Chauhan, R.D.; Gaitán-Solís, E.; Gehan, J.; Butts, P.; Siritunga, D.; Okwuonu, I.; Woll, A.; Jiménez-Aguilar, D.M.; et al. Biofortification of field-grown cassava by engineering expression of an iron transporter and ferritin. Nat. Biotechnol. 2019, 37, 144–151. [Google Scholar] [CrossRef]
  11. Nozoye, T.; Otani, M.; Senoura, T.; Nakanishi, H.; Nishizawa, N.K. Overexpression of barley nicotianamine synthase 1 confers tolerance in the sweet potato to iron deficiency in calcareous soil. Plant Soil 2017, 418, 75–88. [Google Scholar] [CrossRef]
  12. Connorton, J.M.; Jones, E.R.; Rodríguez-Ramiro, I.; Fairweather-Tait, S.; Uauy, C.; Balk, J. Wheat Vacuolar Iron Transporter TaVIT2 Transports Fe and Mn and Is Effective for Biofortification. Plant Physiol. 2017, 174, 2434–2444. [Google Scholar] [CrossRef] [Green Version]
  13. Menkir, A.; Maziya-Dixon, B.; Mengesha, W.; Rocheford, T.; Alamu, E.O. Accruing genetic gain in pro-vitamin A enrichment from harnessing diverse maize germplasm. Euphytica 2017, 213, 105. [Google Scholar] [CrossRef]
  14. Hossain, F.; Muthusamy, V.; Zunjare, R.U. Molecular Breeding for Development of Biofortified Maize Hybrids in India. In Proceedings of the Extended Summaries: 13th Asian Maize Conference on and Expert Consultation on Maize for Food, Feed, Nutrition and Environmental Security, Ludhiana, India, 8–10 October 2018; pp. 220–230. [Google Scholar]
  15. Ceballos, H.; Morante, N.; Sanchez, T.; Ortiz, D.; Aragón, I.; Chávez, A.; Pizarro, M.; Calle, F.; Dominique, D. Rapid Cycling Recurrent Selection for Increased Carotenoids Content in Cassava Roots. Crop Sci. 2013, 53, 2342–2351. [Google Scholar] [CrossRef] [Green Version]
  16. Rezaei, M.; Deokar, A.; Arganosa, G.; Roorkiwal, M.; Pandey, S.; Warkentin, T.; Varshney, R.; Tar’an, B. Mapping Quantitative Trait Loci for Carotenoid Concentration in Three F Populations of Chickpea. Plant Genome 2019, 12, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Ye, X.; Al-Babili, S.; Klöti, A.; Zhang, J.; Lucca, P.; Beyer, P.; Potrykus, I. Engineering the provitamin A (β-carotene) biosynthetic pathway into (carotenoid-free) rice endosperm. Science 2000, 287, 303–305. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Broad, R.C.; Bonneau, J.P.; Beasley, J.T.; Roden, S.; Sadowski, P.; Jewell, N.; Brien, C.; Berger, B.; Tako, E.; Glahn, R.P.; et al. Effect of Rice GDP-L-Galactose Phosphorylase Constitutive Overexpression on Ascorbate Concentration, Stress Tolerance, and Iron Bioavailability in Rice. Front. Plant Sci. 2020, 11, 595439. [Google Scholar] [CrossRef] [PubMed]
  19. Aluru, M.; Xu, Y.; Guo, R.; Wang, Z.; Li, S.; White, W.; Wang, K.; Rodermel, S. Generation of transgenic maize with enhanced provitamin A content. J. Exp. Bot. 2008, 59, 3551–3562. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Liang, Q.; Wang, K.; Liu, X.; Riaz, B.; Jiang, L.; Wan, X.; Ye, X.; Zhang, C. Improved folate accumulation in genetically modified maize and wheat. J. Exp. Bot. 2019, 70, 1539–1551. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. De Lepeleire, J.; Strobbe, S.; Verstraete, J.; Blancquaert, D.; Ambach, L.; Visser, R.G.F.; Stove, C.; Van Der Straeten, D. Folate Biofortification of Potato by Tuber-Specific Expression of Four Folate Biosynthesis Genes. Mol. Plant 2018, 11, 175–188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Yazdani, M.; Sun, Z.; Yuan, H.; Zeng, S.; Thannhauser, T.W.; Vrebalov, J.; Ma, Q.; Xu, Y.; Fei, Z.; Van Eck, J.; et al. Ectopic expression of ORANGE promotes carotenoid accumulation and fruit development in tomato. Plant Biotechnol. J. 2019, 17, 33–49. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Zhang, Y.; Butelli, E.; Alseekh, S.; Tohge, T.; Rallapalli, G.; Luo, J.; Kawar, P.G.; Hill, L.; Santino, A.; Fernie, A.R.; et al. Multi-level engineering facilitates the production of phenylpropanoid compounds in tomato. Nat. Commun. 2015, 6, 8635. [Google Scholar] [CrossRef] [Green Version]
  24. Paul, J.-Y.; Khanna, H.; Kleidon, J.; Hoang, P.; Geijskes, J.; Daniells, J.; Zaplin, E.; Rosenberg, Y.; James, A.; Mlalazi, B.; et al. Golden bananas in the field: Elevated fruit pro-vitamin A from the expression of a single banana transgene. Plant Biotechnol. J. 2017, 15, 520–532. [Google Scholar] [CrossRef] [Green Version]
  25. Ramírez Rivera, N.G.; García-Salinas, C.; Aragão, F.J.L.; Díaz de la Garza, R.I. Metabolic engineering of folate and its precursors in Mexican common bean (Phaseolus vulgaris L.). Plant Biotechnol. J. 2016, 14, 2021–2032. [Google Scholar] [CrossRef] [PubMed]
  26. Yu, B.; Lydiate, D.J.; Young, L.W.; Schäfer, U.A.; Hannoufa, A. Enhancing the carotenoid content of Brassica napus seeds by downregulating lycopene epsilon cyclase. Transgenic Res. 2008, 17, 573–585. [Google Scholar] [CrossRef] [PubMed]
  27. Song, X.; Zhu, W.; Tang, R.; Cai, J.; Chen, M.; Yang, Q. Over-expression of StLCYb increases β-carotene accumulation in potato tubers. Plant Biotechnol. Rep. 2016, 10, 95–104. [Google Scholar] [CrossRef]
  28. Davuluri, G.R.; van Tuinen, A.; Fraser, P.D.; Manfredonia, A.; Newman, R.; Burgess, D.; Brummell, D.A.; King, S.R.; Palys, J.; Uhlig, J.; et al. Fruit-specific RNAi-mediated suppression of DET1 enhances carotenoid and flavonoid content in tomatoes. Nat. Biotechnol. 2005, 23, 890–895. [Google Scholar] [CrossRef] [PubMed]
  29. Singh, S.P.; Gruissem, W.; Bhullar, N.K. Single genetic locus improvement of iron, zinc and β-carotene content in rice grains. Sci. Rep. 2017, 7, 6883. [Google Scholar] [CrossRef] [PubMed]
  30. Ruiz-López, N.; Haslam, R.P.; Venegas-Calerón, M.; Larson, T.R.; Graham, I.A.; Napier, J.A.; Sayanova, O. The synthesis and accumulation of stearidonic acid in transgenic plants: A novel source of “heart-healthy” omega-3 fatty acids. Plant Biotechnol. J. 2009, 7, 704–716. [Google Scholar] [CrossRef] [PubMed]
  31. Cheng, B.; Wu, G.; Vrinten, P.; Falk, K.; Bauer, J.; Qiu, X. Towards the production of high levels of eicosapentaenoic acid in transgenic plants: The effects of different host species, genes and promoters. Transgenic Res. 2010, 19, 221–229. [Google Scholar] [CrossRef]
  32. Jing, G.; Tang, D.; Yao, Y.; Su, Y.; Shen, Y.; Bai, Y.; Jing, W.; Zhang, Q.; Lin, F.; Guo, D.; et al. Seed specifically over-expressing DGAT2A enhances oil and linoleic acid contents in soybean seeds. Biochem. Biophys. Res. Commun. 2021, 568, 143–150. [Google Scholar] [CrossRef]
  33. Liu, Q.; Guo, Q.; Akbar, S.; Zhi, Y.; El Tahchy, A.; Mitchell, M.; Li, Z.; Shrestha, P.; Vanhercke, T.; Ral, J.P.; et al. Genetic enhancement of oil content in potato tuber (Solanum tuberosum L.) through an integrated metabolic engineering strategy. Plant Biotechnol. J. 2017, 15, 56–67. [Google Scholar] [CrossRef] [Green Version]
  34. Kim, W.-S.; Sun-Hyung, J.; Oehrle, N.W.; Jez, J.M.; Krishnan, H.B. Overexpression of ATP sulfurylase improves the sulfur amino acid content, enhances the accumulation of Bowman-Birk protease inhibitor and suppresses the accumulation of the β-subunit of β-conglycinin in soybean seeds. Sci. Rep. 2020, 10, 14989. [Google Scholar] [CrossRef] [PubMed]
  35. Tang, M.; He, X.; Luo, Y.; Ma, L.; Tang, X.; Huang, K. Nutritional assessment of transgenic lysine-rich maize compared with conventional quality protein maize. J. Sci. Food Agric. 2013, 93, 1049–1054. [Google Scholar] [CrossRef] [PubMed]
  36. Yang, Q.Q.; Zhang, C.Q.; Chan, M.L.; Zhao, D.S.; Chen, J.Z.; Wang, Q.; Li, Q.F.; Yu, H.X.; Gu, M.H.; Sun, S.S.M.; et al. Biofortification of rice with the essential amino acid lysine: Molecular characterization, nutritional evaluation, and field performance. J. Exp. Bot. 2016, 67, 4285–4296. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Xu, M.; Zhao, S.; Zhang, Y.; Yin, H.; Peng, X.; Cheng, Z.; Yang, Z.; Zheng, J. Production of Marker-Free Transgenic Rice (Oryza sativa L.) with Improved Nutritive Quality Expressing AmA1. Iran. J. Biotechnol. 2017, 15, 102–110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Garg, M.; Chawla, M.; Chunduri, V.; Kumar, R.; Sharma, S.; Sharma, N.K.; Kaur, N.; Kumar, A.; Mundey, J.K.; Saini, M.K.; et al. Transfer of grain colors to elite wheat cultivars and their characterization. J. Cereal Sci. 2016, 71, 138–144. [Google Scholar] [CrossRef]
  39. Ingallina, C.; Maccelli, A.; Spano, M.; Di Matteo, G.; Di Sotto, A.; Giusti, A.M.; Vinci, G.; Di Giacomo, S.; Rapa, M.; Ciano, S.; et al. Chemico-biological characterization of torpedino di fondi® tomato fruits: A comparison with san marzano cultivar at two ripeness stages. Antioxidants 2020, 9, 1027. [Google Scholar] [CrossRef] [PubMed]
  40. Zhu, Q.; Zeng, D.; Yu, S.; Cui, C.; Li, J.; Li, H.; Chen, J.; Zhang, R.; Zhao, X.; Chen, L.; et al. From Golden Rice to aSTARice: Bioengineering Astaxanthin Biosynthesis in Rice Endosperm. Mol. Plant 2018, 11, 1440–1448. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Krens, F.A.; Schaart, J.G.; van der Burgh, A.M.; TinnenbroekCapel, I.E.M.; Groenwold, R.; Kodde, L.P.; Broggini, G.A.L.; Gessler, C.; Schouten, H.J. Cisgenic apple trees; development, characterization, and performance. Front. Plant Sci. 2015, 6, 286. [Google Scholar] [CrossRef] [Green Version]
  42. Hoddinott, J.F.; Rosegrant, M.W.; Torero, M. Investments to reduce hunger and undernutrition. In Global Problems, Smart Solutions; Lomborg, B., Ed.; Cambridge University Press: Cambridge, UK, 2013; pp. 332–367. [Google Scholar]
  43. Graham, R.D.; Welch, R.M. Breeding for Staple-Food Crops with High Micronutrient Density: Working Papers on Agricultural Strategies for Micronutrients, 3rd ed.; International Food Policy Institute: Washington, DC, USA, 1996. [Google Scholar]
  44. Graham, R.D.; Welch, R.M.; Bouis, H.E. Addressing Micronutrient Malnutrition Through Enhancing the Nutritional Quality of Staple Foods: Principles, Perspectives and Knowledge Gaps. Adv. Agron. 2001, 70, 77–142. [Google Scholar] [CrossRef]
  45. United Nations; Department of Economic and Social; Affairs, P.D. World Population Prospects, Medium Prognosis; The 2019 Revision: New York, NY, USA, 2019. [Google Scholar]
  46. Wortmann, L.; Enneking, U.; Daum, D. German Consumers’ Attitude Towards Selenium-Biofortified Apples and Acceptance of Related Nutrition and Health Claims. Nutrients 2018, 10, 190. [Google Scholar] [CrossRef] [Green Version]
  47. Welk, A.K.; Kleine-kalmer, R.; Daum, D.; Enneking, U. Consumer Acceptance and Market Potential of Iodine-Biofortified Fruit and Vegetables in Germany. Nutrients 2021, 13, 4198. [Google Scholar] [CrossRef] [PubMed]
  48. Adekambi, S.A.; Okello, J.J.; Rajendran, S.; Acheremu, K.; Carey, E.E.; Low, J.; Abidin, P.E. Effect of varietal attributes on the adoption of an orange-fleshed sweetpotato variety in Upper East and Northern Ghana. Outlook Agric. 2020, 49, 311–320. [Google Scholar] [CrossRef] [PubMed]
  49. Gannon, B.; Kaliwile, C.; Arscott, S.A.; Schmaelzle, S.; Chileshe, J.; Kalungwana, N.; Mosonda, M.; Pixley, K.; Masi, C.; Tanumihardjo, S.A. Biofortified orange maize is as efficacious as a vitamin A supplement in Zambian children even in the presence of high liver reserves of vitamin A: A community-based, randomized placebo-controlled trial. Am. J. Clin. Nutr. 2014, 100, 1541–1550. [Google Scholar] [CrossRef] [PubMed]
  50. Komar, A.A. Single Nucleotide Polymorphisms. Methods Mol. Biol. 2009, 578, 23–39. [Google Scholar]
  51. Collard, B.C.Y.; Jahufer, M.Z.Z.; Brouwer, J.B.; Pang, E.C.K. An introduction to markers, quantitative trait loci (QTL) mapping and marker-assisted selection for crop improvement: The basic concepts. Euphytica 2005, 142, 169–196. [Google Scholar] [CrossRef]
  52. Unterseer, S.; Bauer, E.; Haberer, G.; Seidel, M.; Knaak, C.; Ouzunova, M.; Meitinger, T.; Strom, T.M.; Fries, R.; Pausch, H.; et al. A powerful tool for genome analysis in maize: Development and evaluation of the high density 600 k SNP genotyping array. BMC Genomics 2014, 15, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Xu, C.; Ren, Y.; Jian, Y.; Guo, Z.; Zhang, Y.; Xie, C.; Fu, J.; Wang, H.; Wang, G.; Xu, Y.; et al. Development of a maize 55 K SNP array with improved genome coverage for molecular breeding. Mol. Breed. 2017, 37, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Tung, C.W.; Zhao, K.; Wright, M.H.; Ali, M.L.; Jung, J.; Kimball, J.; Tyagi, W.; Thomson, M.J.; McNally, K.; Leung, H.; et al. Development of a research platform for dissecting phenotype-genotype associations in rice (Oryza spp.). Rice 2010, 3, 205–217. [Google Scholar] [CrossRef] [Green Version]
  55. Singh, N.; Jayaswal, P.K.; Panda, K.; Mandal, P.; Kumar, V.; Singh, B.; Mishra, S.; Singh, Y.; Singh, R.; Rai, V.; et al. Single-copy gene based 50 K SNP chip for genetic studies and molecular breeding in rice. Sci. Rep. 2015, 5, 1–9. [Google Scholar] [CrossRef]
  56. Bauer, E.; Schmutzer, T.; Barilar, I.; Mascher, M.; Gundlach, H.; Martis, M.M.; Twardziok, S.O.; Hackauf, B.; Gordillo, A.; Wilde, P.; et al. Towards a whole-genome sequence for rye (Secale cereale L.). Plant J. 2017, 89, 853–869. [Google Scholar] [CrossRef] [Green Version]
  57. Winfield, M.O.; Allen, A.M.; Burridge, A.J.; Barker, G.L.A.; Benbow, H.R.; Wilkinson, P.A.; Coghill, J.; Waterfall, C.; Davassi, A.; Scopes, G.; et al. High-density SNP genotyping array for hexaploid wheat and its secondary and tertiary gene pool. Plant Biotechnol. J. 2016, 14, 1195–1206. [Google Scholar] [CrossRef] [PubMed]
  58. Allen, A.M.; Winfield, M.O.; Burridge, A.J.; Downie, R.C.; Benbow, H.R.; Barker, G.L.A.; Wilkinson, P.A.; Coghill, J.; Waterfall, C.; Davassi, A.; et al. Characterization of a Wheat Breeders’ Array suitable for high-throughput SNP genotyping of global accessions of hexaploid bread wheat (Triticum aestivum). Plant Biotechnol. J. 2017, 15, 390–401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Karadi, A.; Samineni, S.; Sajja, S.; Sharma, M.; Thudi, M.; Mallikarjuna, B.P.; Viswanatha, K.P.; Varshney, R.K.; Gaur, P.M. Molecular mapping of dry root rot resistance genes in chickpea (Cicer arietinum L.). Euphytica 2021, 217, 1–13. [Google Scholar] [CrossRef]
  60. Stoffel, K.; van Leeuwen, H.; Kozik, A.; Caldwell, D.; Ashrafi, H.; Cui, X.; Tan, X.; Hill, T.; Reyes-Chin-Wo, S.; Truco, M.J.; et al. Development and application of a 6.5 million feature Affymetrix Genechip® for massively parallel discovery of single position polymorphisms in lettuce (Lactuca spp.). BMC Genomics 2012, 13, 1–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Vos, P.G.; Uitdewilligen, J.G.A.M.L.; Voorrips, R.E.; Visser, R.G.F.; van Eck, H.J. Development and analysis of a 20K SNP array for potato (Solanum tuberosum): An insight into the breeding history. Theor. Appl. Genet. 2015, 128, 2387–2401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Lee, Y.G.; Jeong, N.; Kim, J.H.; Lee, K.; Kim, K.H.; Pirani, A.; Ha, B.K.; Kang, S.T.; Park, B.S.; Moon, J.K.; et al. Development, validation and genetic analysis of a large soybean SNP genotyping array. Plant J. 2015, 81, 625–636. [Google Scholar] [CrossRef]
  63. Bassil, N.V.; Davis, T.M.; Zhang, H.; Ficklin, S.; Mittmann, M.; Webster, T.; Mahoney, L.; Wood, D.; Alperin, E.S.; Rosyara, U.R.; et al. Development and preliminary evaluation of a 90 K Axiom® SNP array for the allo-octoploid cultivated strawberry Fragaria × ananassa. BMC Genomics 2015, 16, 1–30. [Google Scholar] [CrossRef] [Green Version]
  64. Bianco, L.; Cestaro, A.; Linsmith, G.; Muranty, H.; Denancé, C.; Théron, A.; Poncet, C.; Micheletti, D.; Kerschbamer, E.; Di Pierro, E.A.; et al. Development and validation of the Axiom®Apple480K SNP genotyping array. Plant J. 2016, 86, 62–74. [Google Scholar] [CrossRef] [Green Version]
  65. Pandey, M.K.; Agarwal, G.; Kale, S.M.; Clevenger, J.; Nayak, S.N.; Sriswathi, M.; Chitikineni, A.; Chavarro, C.; Chen, X.; Upadhyaya, H.D.; et al. Development and Evaluation of a High Density Genotyping “Axiom-Arachis” Array with 58 K SNPs for Accelerating Genetics and Breeding in Groundnut. Sci. Rep. 2017, 7, 1–10. [Google Scholar] [CrossRef] [PubMed]
  66. Comadran, J.; Kilian, B.; Russell, J.; Ramsay, L.; Stein, N.; Ganal, M.; Shaw, P.; Bayer, M.; Thomas, W.; Marshall, D.; et al. Natural variation in a homolog of Antirrhinum CENTRORADIALIS contributed to spring growth habit and environmental adaptation in cultivated barley. Nat. Genet. 2012, 44, 1388–1391. [Google Scholar] [CrossRef]
  67. Ganal, M.W.; Durstewitz, G.; Polley, A.; Bérard, A.; Buckler, E.S.; Charcosset, A.; Clarke, J.D.; Graner, E.M.; Hansen, M.; Joets, J.; et al. A Large Maize (Zea mays L.) SNP Genotyping Array: Development and Germplasm Genotyping, and Genetic Mapping to Compare with the B73 Reference Genome. PLoS ONE 2011, 6, e28334. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Rousselle, Y.; Jones, E.; Charcosset, A.; Moreau, P.; Robbins, K.; Stich, B.; Knaak, C.; Flament, P.; Karaman, Z.; Martinant, J.P.; et al. Study on Essential Derivation in Maize: III. Selection and Evaluation of a Panel of Single Nucleotide Polymorphism Loci for Use in European and North American Germplasm. Crop Sci. 2015, 55, 1170–1180. [Google Scholar] [CrossRef]
  69. Tinker, N.A.; Chao, S.; Lazo, G.R.; Oliver, R.E.; Huang, Y.; Poland, J.A.; Jellen, E.N.; Maughan, P.J.; Kilian, A.; Jackson, E.W. A SNP Genotyping Array for Hexaploid Oat. Plant Genome 2014, 7, 1–8. [Google Scholar] [CrossRef]
  70. Chen, W.; Gao, Y.; Xie, W.; Gong, L.; Lu, K.; Wang, W.; Li, Y.; Liu, X.; Zhang, H.; Dong, H.; et al. Genome-wide association analyses provide genetic and biochemical insights into natural variation in rice metabolism. Nat. Genet. 2014, 46, 714–721. [Google Scholar] [CrossRef] [PubMed]
  71. Yu, H.; Xie, W.; Li, J.; Zhou, F.; Zhang, Q. A whole-genome SNP array (RICE6K) for genomic breeding in rice. Plant Biotechnol. J. 2014, 12, 28–37. [Google Scholar] [CrossRef] [PubMed]
  72. Cavanagh, C.R.; Chao, S.; Wang, S.; Huang, B.E.; Stephen, S.; Kiani, S.; Forrest, K.; Saintenac, C.; Brown-Guedira, G.L.; Akhunova, A.; et al. Genome-wide comparative diversity uncovers multiple targets of selection for improvement in hexaploid wheat landraces and cultivars. Proc. Natl. Acad. Sci. USA 2013, 110, 8057–8062. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Clarke, W.E.; Higgins, E.E.; Plieske, J.; Wieseke, R.; Sidebottom, C.; Khedikar, Y.; Batley, J.; Edwards, D.; Meng, J.; Li, R.; et al. A high-density SNP genotyping array for Brassica napus and its ancestral diploid species based on optimised selection of single-locus markers in the allotetraploid genome. Theor. Appl. Genet. 2016, 129, 1887–1899. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Bachlava, E.; Taylor, C.A.; Tang, S.; Bowers, J.E.; Mandel, J.R.; Burke, J.M.; Knapp, S.J. SNP Discovery and Development of a High-Density Genotyping Array for Sunflower. PLoS ONE 2012, 7, e29814. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Livaja, M.; Unterseer, S.; Erath, W.; Lehermeier, C.; Wieseke, R.; Plieske, J.; Polley, A.; Luerßen, H.; Wieckhorst, S.; Mascher, M.; et al. Diversity analysis and genomic prediction of Sclerotinia resistance in sunflower using a new 25 K SNP genotyping array. Theor. Appl. Genet. 2016, 129, 317–329. [Google Scholar] [CrossRef] [PubMed]
  76. Close, T.J.; Lucas, M.R.; Muñoz-Amatriain, M.; Mirebrahim, H.; Wanamaker, S.; Barkley, N.A.; Clair, S.S.; Guo, Y.-N.; Lo, S.; Huynh, B.L. A new SNP-genotyping resource for cowpea and its deployment for breeding. In Proceedings of the Plant and Animal Genome Conference, San Diego, CA, USA, 10–14 January 2015; Volume 23, p. P0784. [Google Scholar]
  77. Hamilton, J.P.; Hansey, C.N.; Whitty, B.R.; Stoffel, K.; Massa, A.N.; Van Deynze, A.; De Jong, W.S.; Douches, D.S.; Buell, C.R. Single nucleotide polymorphism discovery in elite North American potato germplasm. BMC Genomics 2011, 12, 1–11. [Google Scholar] [CrossRef] [Green Version]
  78. Sim, S.C.; Durstewitz, G.; Plieske, J.; Wieseke, R.; Ganal, M.W.; van Deynze, A.; Hamilton, J.P.; Buell, C.R.; Causse, M.; Wijeratne, S.; et al. Development of a Large SNP Genotyping Array and Generation of High-Density Genetic Maps in Tomato. PLoS ONE 2012, 7, e40563. [Google Scholar] [CrossRef] [PubMed]
  79. Song, Q.; Hyten, D.L.; Jia, G.; Quigley, C.V.; Fickus, E.W.; Nelson, R.L.; Cregan, P.B. Development and Evaluation of SoySNP50K, a High-Density Genotyping Array for Soybean. PLoS ONE 2013, 8, e54985. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Bianco, L.; Cestaro, A.; Sargent, D.J.; Banchi, E.; Derdak, S.; Di Guardo, M.; Salvi, S.; Jansen, J.; Viola, R.; Gut, I.; et al. Development and validation of a 20K Single Nucleotide Polymorphism (SNP) whole genome genotyping array for apple (Malus × domestica Borkh). PLoS ONE 2014, 9, e110377. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Chagné, D.; Crowhurst, R.N.; Troggio, M.; Davey, M.W.; Gilmore, B.; Lawley, C.; Vanderzande, S.; Hellens, R.P.; Kumar, S.; Cestaro, A.; et al. Genome-Wide SNP Detection, Validation, and Development of an 8K SNP Array for Apple. PLoS ONE 2012, 7, e31745. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Peace, C.; Bassil, N.; Main, D.; Ficklin, S.; Rosyara, U.R.; Stegmeir, T.; Sebolt, A.; Gilmore, B.; Lawley, C.; Mockler, T.C.; et al. Development and Evaluation of a Genome-Wide 6K SNP Array for Diploid Sweet Cherry and Tetraploid Sour Cherry. PLoS ONE 2012, 7, e48305. [Google Scholar] [CrossRef] [PubMed]
  83. Verde, I.; Bassil, N.; Scalabrin, S.; Gilmore, B.; Lawley, C.T.; Gasic, K.; Micheletti, D.; Rosyara, U.R.; Cattonaro, F.; Vendramin, E.; et al. Development and Evaluation of a 9k SNP Array for Peach by Internationally Coordinated SNP Detection and Validation in Breeding Germplasm. PLoS ONE 2012, 7, e35668. [Google Scholar] [CrossRef]
  84. Micheletti, D.; Dettori, M.T.; Micali, S.; Aramini, V.; Pacheco, I.; Da Silva Linge, C.; Foschi, S.; Banchi, E.; Barreneche, T.; Quilot-Turion, B.; et al. Whole-Genome Analysis of Diversity and SNP-Major Gene Association in Peach Germplasm. PLoS ONE 2015, 10, e0136803. [Google Scholar] [CrossRef]
  85. Mas-Gómez, J.; Cantín, C.M.; Moreno, M.; Prudencio, Á.S.; Gómez-Abajo, M.; Bianco, L.; Troggio, M.; Martínez-Gómez, P.; Rubio, M.; Martínez-García, P.J. Exploring genome-wide diversity in the national peach (Prunus persica) germplasm collection at CITA (Zaragoza, Spain). Agronomy 2021, 11, 481. [Google Scholar] [CrossRef]
  86. Montanari, S.; Saeed, M.; Knäbel, M.; Kim, Y.K.; Troggio, M.; Malnoy, M.; Velasco, R.; Fontana, P.; Won, K.H.; Durel, C.E.; et al. Identification of Pyrus Single Nucleotide Polymorphisms (SNPs) and Evaluation for Genetic Mapping in European Pear and Interspecific Pyrus Hybrids. PLoS ONE 2013, 8, e77022. [Google Scholar] [CrossRef]
  87. Myles, S.; Chia, J.M.; Hurwitz, B.; Simon, C.; Zhong, G.Y.; Buckler, E.; Ware, D. Rapid genomic characterization of the genus Vitis. PLoS ONE 2010, 5, e8219. [Google Scholar] [CrossRef] [Green Version]
  88. Le Paslier, M.-C.; Choisne, N.; Scalabrin, S.; Bacilieri, R.; Berard, A.; Bounon, R.; Boursiquot, J.-M.; Bras, M.; Brunel, D.; Chauveau, A.; et al. The GrapeReSeq 18K Vitis genotyping chip. IX International Symposium on Grapevine Physiology Biotechnology, La Serena, Chile, 2–26 April 2013; p. 18. [Google Scholar]
  89. De Jong, W.S.; De Jong, D.M.; Bodis, M. A fluorogenic 5′ nuclease (TaqMan) assay to assess dosage of a marker tightly linked to red skin color in autotetraploid potato. Theor. Appl. Genet. 2003, 107, 1384–1390. [Google Scholar] [CrossRef]
  90. Wu, J.; Wang, Q.; Kang, Z.; Liu, S.; Li, H.; Mu, J.; Dai, M.; Han, D.; Zeng, Q.; Chen, X. Development and validation of KASP-SNP markers for QTL underlying resistance to stripe rust in common wheat cultivar P10057. Plant Dis. 2017, 101, 2079–2087. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Qureshi, N.; Kandiah, P.; Gessese, M.K.; Nsabiyera, V.; Wells, V.; Babu, P.; Wong, D.; Hayden, M.; Bariana, H.; Bansal, U. Development of co-dominant KASP markers co-segregating with Ug99 effective stem rust resistance gene Sr26 in wheat. Mol. Breed. 2018, 38, 1–9. [Google Scholar] [CrossRef]
  92. Collins, D.; Emebiri, L.; Tan, M.K.; El Bouhssini, M.; Wildman, O. Association of KASP markers with Hessian fly resistance in wheat of diverse origin. Euphytica 2018, 214, 1–8. [Google Scholar] [CrossRef]
  93. Wang, R.; Liu, Y.; Isham, K.; Zhao, W.; Wheeler, J.; Klassen, N.; Hu, Y.; Bonman, J.M.; Chen, J. QTL identification and KASP marker development for productive tiller and fertile spikelet numbers in two high-yielding hard white spring wheat cultivars. Mol. Breed. 2018, 38, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Singh, L.; Anderson, J.A.; Chen, J.; Gill, B.S.; Tiwari, V.K.; Rawat, N. Development and validation of a perfect KASP marker for fusarium head blight resistance gene Fhb1 in wheat. Plant Pathol. J. 2019, 35, 200–207. [Google Scholar] [CrossRef] [PubMed]
  95. Fang, T.; Lei, L.; Li, G.; Powers, C.; Hunger, R.M.; Carver, B.F.; Yan, L. Development and deployment of KASP markers for multiple alleles of Lr34 in wheat. Theor. Appl. Genet. 2020, 133, 2183–2195. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Grewal, S.; Othmeni, M.; Walker, J.; Hubbart-Edwards, S.; Yang, C.Y.; Scholefield, D.; Ashling, S.; Isaac, P.; King, I.P.; King, J. Development of Wheat-Aegilops caudata Introgression Lines and Their Characterization Using Genome-Specific KASP Markers. Front. Plant Sci. 2020, 11, 606. [Google Scholar] [CrossRef] [PubMed]
  97. Makhoul, M.; Rambla, C.; Voss-Fels, K.P.; Hickey, L.T.; Snowdon, R.J.; Obermeier, C. Overcoming polyploidy pitfalls: A user guide for effective SNP conversion into KASP markers in wheat. Theor. Appl. Genet. 2020, 133, 2413–2430. [Google Scholar] [CrossRef] [PubMed]
  98. Zhang, S.; Fan, C.; Luo, J.; Huang, L.; Xie, D.; Li, Y.; Chen, Z.; Jiang, B.; Ning, S.; Yuan, Z.; et al. KASP markers to detect sub-chromosomal arm translocations between 6VS of Haynaldia villosa and 6AS of wheat. Euphytica 2021, 217, 10. [Google Scholar] [CrossRef]
  99. Xu, X.; Li, G.; Bai, G.; Bernardo, A.; Carver, B.F.; Amand, P.S.; Armstrong, J.S. Development of KASP markers for wheat greenbug resistance gene Gb5. Crop Sci. 2021, 61, 490–499. [Google Scholar] [CrossRef]
  100. Sangha, J.; Tucker, J.R.; Legge, W.G.; Badea, A. Use of KASP assays for the analysis of rpg4/Rpg5 gene complex for marker-assisted selection for Ug99 stem rust resistance in barley. Can. J. Plant Pathol. 2017, 39, 578. [Google Scholar]
  101. Steele, K.A.; Quinton-Tulloch, M.J.; Amgai, R.B.; Dhakal, R.; Khatiwada, S.P.; Vyas, D.; Heine, M.; Witcombe, J.R. Accelerating public sector rice breeding with high-density KASP markers derived from whole genome sequencing of indica rice. Mol. Breed. 2018, 38, 1–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Yang, G.; Chen, S.; Chen, L.; Sun, K.; Huang, C.; Zhou, D.; Huang, Y.; Wang, J.; Liu, Y.; Wang, H.; et al. Development of a core SNP arrays based on the KASP method for molecular breeding of rice. Rice 2019, 12, 1–18. [Google Scholar] [CrossRef] [PubMed]
  103. Yang, Y.; Basnet, B.R.; Ibrahim, A.M.H.; Rudd, J.C.; Chen, X.; Bowden, R.L.; Xue, Q.; Wang, S.; Johnson, C.D.; Metz, R.; et al. Developing KASP Markers on a Major Stripe Rust Resistance QTL in a Popular Wheat TAM 111 Using 90K Array and Genotyping-by-Sequencing SNPs. Crop Sci. 2019, 59, 165–175. [Google Scholar] [CrossRef]
  104. Addison, C.K.; Angira, B.; Kongchum, M.; Harrell, D.L.; Baisakh, N.; Linscombe, S.D.; Famoso, A.N. Characterization of Haplotype Diversity in the BADH2 Aroma Gene and Development of a KASP SNP Assay for Predicting Aroma in U.S. Rice. Rice 2020, 13, 1–9. [Google Scholar] [CrossRef] [PubMed]
  105. Burow, G.; Chopra, R.; Hughes, H.; Xin, Z.; Burke, J. Marker Assisted Selection in Sorghum Using KASP Assay for the Detection of Single Nucleotide Polymorphism/Insertion Deletion. Methods Mol. Biol. 2019, 1931, 75–84. [Google Scholar] [CrossRef] [PubMed]
  106. Grimm, K.D.S.; Porter, L.D. Development and Validation of KASP Markers for the Identification of Pea seedborne mosaic virus Pathotype P1 Resistance in Pisum sativum. Plant Dis. 2020, 104, 1824–1830. [Google Scholar] [CrossRef]
  107. Legendre, R.; McGregor, C. KASP (TM) Markers for Selection for Fruit Shape in Watermelon. Hortscience 2019, 54, S399. [Google Scholar]
  108. Paudel, L.; Clevenger, J.; McGregor, C. Refining of the egusi locus in watermelon using KASP assays. Sci. Hortic. 2019, 257, 108665. [Google Scholar] [CrossRef]
  109. Zanotto, S.; Vandenberg, A.; Khazaei, H. Development and validation of a robust KASP marker for zt2 locus in faba bean (Vicia faba). Plant Breed. 2020, 139, 375–380. [Google Scholar] [CrossRef]
  110. Devran, Z.; Göknur, A.; Mesci, L. Development of molecular markers for the Mi-1 gene in tomato using the KASP genotyping assay. Hortic. Environ. Biotechnol. 2016, 57, 156–160. [Google Scholar] [CrossRef]
  111. Devran, Z.; Kahveci, E. Development and validation of a user-friendly KASP marker for the Sw-5 locus in tomato. Australas. Plant Pathol. 2019, 48, 503–507. [Google Scholar] [CrossRef]
  112. Han, F.; Zhang, X.; Yuan, K.; Fang, Z.; Yang, L.; Zhuang, M.; Zhang, Y.; Wang, Y.; Liu, Y.; Li, Z.; et al. A user-friendly KASP molecular marker developed for the DGMS-based breeding system in Brassica oleracea species. Mol. Breed. 2019, 39, 1–7. [Google Scholar] [CrossRef]
  113. da Cruz, M.F.A.; Bueno, R.D.; de Souza, F.B.; Moreira, M.A.; de Barros, E.G. Identification of SNPs for fatty acid content in soybean by the HRM technique. Pesqui. Agropecu. Bras. 2013, 48, 1596–1600. [Google Scholar] [CrossRef] [Green Version]
  114. Rai, V.P.; Singh, A.K.; Jaiswal, H.K.; Singh, S.P.; Singh, R.P.; Waza, S.A. Evaluation of molecular markers linked to fragrance and genetic diversity in Indian aromatic rice. Turk. J. Botany 2015, 39, 209–217. [Google Scholar] [CrossRef]
  115. Noh, Y.H.; Lee, S.; Whitaker, V.M.; Cearley, K.R.; Cha, J.S. A high-throughput marker-assisted selection system combining rapid DNA extraction and high-resolution melting and simple sequence repeat analysis: Strawberry as a model for fruit crops. J. Berry Res. 2017, 7, 23–31. [Google Scholar] [CrossRef]
  116. Geng, L.; Li, M.; Xie, S.; Wu, D.; Ye, L.; Zhang, G. Identification of genetic loci and candidate genes related to β-glucan content in barley grain by genome-wide association study in International Barley Core Selected Collection. Mol. Breed. 2021, 41, 1–12. [Google Scholar] [CrossRef]
  117. Paris, M.; Jones, M.G.K.; Eglinton, J.K. Genotyping Single Nucleotide Polymorphisms for Selection of Barley β-amylase Alleles. Plant Mol. Biol. Report. 2002, 20, 149–159. [Google Scholar] [CrossRef]
  118. Masouleh, A.K.; Waters, D.L.E.; Reinke, R.F.; Henry, R.J. A high-throughput assay for rapid and simultaneous analysis of perfect markers for important quality and agronomic traits in rice using multiplexed MALDI-TOF mass spectrometry. Plant Biotechnol. J. 2009, 7, 355–363. [Google Scholar] [CrossRef]
  119. Cheng, P.; Holdsworth, W.; Ma, Y.; Coyne, C.J.; Mazourek, M.; Grusak, M.A.; Fuchs, S.; McGee, R.J. Association mapping of agronomic and quality traits in USDA pea single-plant collection. Mol. Breed. 2015, 35, 1–13. [Google Scholar] [CrossRef]
  120. Lum, H.-K.; Lee, C.-H.; Butt, Y.K.-C.; Lo, S.C.-L. Sodium nitroprusside affects the level of photosynthetic enzymes and glucose metabolism in Phaseolus aureus (mung bean). Nitric Oxide 2005, 12, 220–230. [Google Scholar] [CrossRef] [PubMed]
  121. Salazar, J.A.; Pacheco, I.; Shinya, P.; Zapata, P.; Silva, C.; Aradhya, M.; Velasco, D.; Ruiz, D.; Martínez-Gómez, P.; Infante, R. Genotyping by Sequencing for SNP-Based Linkage Analysis and Identification of QTLs Linked to Fruit Quality Traits in Japanese Plum (Prunus salicina Lindl.). Front. Plant Sci. 2017, 8, 1–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Larsen, B.; Migicovsky, Z.; Jeppesen, A.A.; Gardner, K.M.; Toldam-Andersen, T.B.; Myles, S.; Ørgaard, M.; Petersen, M.A.; Pedersen, C. Genome-Wide Association Studies in Apple Reveal Loci for Aroma Volatiles, Sugar Composition, and Harvest Date. Plant Genome 2019, 12, 180104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Pereira, L.; Ruggieri, V.; Pérez, S.; Alexiou, K.G.; Fernández, M.; Jahrmann, T.; Pujol, M.; Garcia-Mas, J. QTL mapping of melon fruit quality traits using a high-density GBS-based genetic map. BMC Plant Biol. 2018, 18, 1–17. [Google Scholar] [CrossRef] [PubMed]
  124. Guo, R.; Dhliwayo, T.; Mageto, E.K.; Palacios-Rojas, N.; Lee, M.; Yu, D.; Ruan, Y.; Zhang, A.; San Vicente, F.; Olsen, M.; et al. Genomic Prediction of Kernel Zinc Concentration in Multiple Maize Populations Using Genotyping-by-Sequencing and Repeat Amplification Sequencing Markers. Front. Plant Sci. 2020, 11, 534. [Google Scholar] [CrossRef] [PubMed]
  125. Ma, Y.; Coyne, C.J.; Grusak, M.A.; Mazourek, M.; Cheng, P.; Main, D.; McGee, R.J. Genome-wide SNP identification, linkage map construction and QTL mapping for seed mineral concentrations and contents in pea (Pisum sativum L.). BMC Plant Biol. 2017, 17, 1–17. [Google Scholar] [CrossRef] [Green Version]
  126. Qin, J.; Shi, A.; Mou, B.; Grusak, M.A.; Weng, Y.; Ravelombola, W.; Bhattarai, G.; Dong, L.; Yang, W. Genetic diversity and association mapping of mineral element concentrations in spinach leaves. BMC Genomics 2017, 18, 1–14. [Google Scholar] [CrossRef] [Green Version]
  127. Tian, H.; Yang, Y.; Wang, R.; Fan, Y.; Yi, H.; Jiang, B.; Wang, L.; Ren, J.; Xu, L.; Zhang, Y.; et al. Screening of 200 core SNPs and the Construction of a Systematic SNP-DNA Standard Fingerprint Database with More Than 20,000 Maize Varieties. Agriculture 2021, 11, 597. [Google Scholar] [CrossRef]
  128. Kuhn, D.N.; Groh, A.; Rahaman, J.; Freeman, B.; Arpaia, M.L.; Van den Berg, N.; Abeysekara, N.; Manosalva, P.; Chambers, A.H. Creation of an avocado unambiguous genotype SNP database for germplasm curation and as an aid to breeders. Tree Genet. Genomes 2019, 15, 71. [Google Scholar] [CrossRef] [Green Version]
  129. Medina-lozano, I.; Bertolín, J.R.; Zufiaurre, R.; Diaz, A. Improved UPLC-UV Method for the Quantification of Vitamin C in Lettuce Varieties (Lactuca sativa L.) and Crop Wild Relatives (Lactuca spp.). J. Vis. Exp. 2020, 160, e61440. [Google Scholar] [CrossRef] [PubMed]
  130. Dossou, S.S.K.; Xu, F.; You, J.; Zhou, R.; Li, D.; Wang, L. Widely targeted metabolome profiling of different colored sesame (Sesamum indicum L.) seeds provides new insight into their antioxidant activities. Food Res. Int. 2022, 151, 110850. [Google Scholar] [CrossRef] [PubMed]
  131. Cheng, Z.Q.; Huang, X.Q.; Zhang, Y.Z.; Qian, J.; Yang, M.Z.; Wu, C.J.; Liu, J.F. Diversity in the content of some nutritional components in husked seeds of three wild rice species and rice varieties in Yunnan Province of China. J. Integr. Plant Biol. 2005, 47, 1260–1270. [Google Scholar] [CrossRef]
  132. Esteras, C.; Rambla, J.L.; Sánchez, G.; López-Gresa, M.P.; González-Mas, M.C.; Fernández-Trujillo, J.P.; Bellés, J.M.; Granell, A.; Picó, M.B. Fruit flesh volatile and carotenoid profile analysis within the Cucumis melo L. species reveals unexploited variability for future genetic breeding. J. Sci. Food Agric. 2018, 98, 3915–3925. [Google Scholar] [CrossRef] [PubMed]
  133. Burbano-Erazo, E.; León-Pacheco, R.I.; Cordero-Cordero, C.C.; López-Hernández, F.; Cortés, A.J.; Tofiño-Rivera, A.P. Multi-Environment Yield Components in Advanced Common Bean (Phaseolus vulgaris L.) × Tepary Bean (P. acutifolius A. Gray) Interspecific Lines for Heat and Drought tolerance. Agronomy 2021, 11, 1978. [Google Scholar] [CrossRef]
  134. Herraiz, F.J.; Raigón, M.D.; Vilanova, S.; García-Martínez, M.D.; Gramazio, P.; Plazas, M.; Rodríguez-Burruezo, A.; Prohens, J. Fruit composition diversity in land races and modern pepino (Solanum muricatum) varieties and wild related species. Food Chem. 2016, 203, 49–58. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Medina-Lozano, I.; Bertolín, J.R.; Díaz, A. Nutritional value of commercial and traditional lettuce (Lactuca sativa L.) and wild relatives: Vitamin C and anthocyanin content. Food Chem. 2021, 359. [Google Scholar] [CrossRef] [PubMed]
  136. Li, K.; Wang, D.; Gong, L.; Lyu, Y.; Guo, H.; Chen, W.; Jin, C.; Liu, X.; Fang, C.; Luo, J. Comparative analysis of metabolome of rice seeds at three developmental stages using a recombinant inbred line population. Plant J. 2019, 100, 908–922. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Jeon, J.E.; Kim, J.-G.; Fischer, C.R.; Mehta, N.; Dufour-Schroif, C.; Wemmer, K.; Mudgett, M.B.; Sattely, E. Pathogen-responsive gene cluster for highly modified fatty acids in tomato. Cell 2020, 180, 176–187. [Google Scholar] [CrossRef] [PubMed]
  138. Raza, A. Metabolomics: A systems biology approach for enhancing heat stress tolerance in plants. Plant Cell Rep. 2020. [Google Scholar] [CrossRef] [PubMed]
  139. Jaganathan, D.; Bohra, A.; Thudi, M.; Varshney, R.K. Fine mapping and gene cloning in the post-NGS era: Advances and prospects. Theor. Appl. Genet. 2020, 133, 1791–1810. [Google Scholar] [CrossRef] [Green Version]
  140. Sharma, V.; Gupta, P.; Priscilla, K.; Sharankumar; Hangargi, B.; Veershetty, A.; Ramrao, D.P.; Suresh, S.; Narasanna, R.; Naik, G.R.; et al. Metabolomics Intervention Towards Better Understanding of Plant Traits. Cells 2021, 10, 346. [Google Scholar] [CrossRef] [PubMed]
  141. Bilbrey, E.A.; Williamson, K.; Hatzakis, E.; Miller, D.D.; Fresnedo-Ramírez, J.; Cooperstone, J.L. Integrating genomics and multiplatform metabolomics enables metabolite quantitative trait loci detection in breeding-relevant apple germplasm. New Phytol. 2021, 232, 1944–1958. [Google Scholar] [CrossRef] [PubMed]
  142. Templer, S.E.; Ammon, A.; Pscheidt, D.; Ciobotea, O.; Schuy, C.; McCollum, C.; Sonnewald, U.; Hanemann, A.; Förster, J.; Ordon, F.; et al. Metabolite profiling of barley flag leaves under drought and combined heat and drought stress reveals metabolic QTLs for metabolites associated with antioxidant defense. J. Exp. Bot. 2017, 68, 1697–1713. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Matros, A.; Houston, K.; Tucker, M.R.; Schreiber, M.; Berger, B.; Aubert, M.K.; Wilkinson, L.G.; Witzel, K.; Waugh, R.; Seiffert, U.; et al. Genome-wide association study reveals the genetic complexity of fructan accumulation patterns in barley grain. J. Exp. Bot. 2021, 72, 2383–2402. [Google Scholar] [CrossRef] [PubMed]
  144. Mahalingam, R.; Sallam, A.H.; Steffenson, B.J.; Fiedler, J.D.; Walling, J.G. Genome-wide association analysis of natural variation in seed tocochromanols of barley. Plant Genome 2020, 13. [Google Scholar] [CrossRef] [PubMed]
  145. Peng, M.; Shahzad, R.; Gul, A.; Subthain, H.; Shen, S.; Lei, L.; Zheng, Z.; Zhou, J.; Lu, D.; Wang, S.; et al. Differentially evolved glucosyltransferases determine natural variation of rice flavone accumulation and UV-tolerance. Nat. Commun. 2017, 8, 1–12. [Google Scholar] [CrossRef] [PubMed]
  146. Ferrão, L.F.V.; Johnson, T.S.; Benevenuto, J.; Edger, P.P.; Colquhoun, T.A.; Munoz, P.R. Genome-wide association of volatiles reveals candidate loci for blueberry flavor. New Phytol. 2020, 226, 1725–1737. [Google Scholar] [CrossRef] [PubMed]
  147. Matros, A.; Liu, G.; Hartmann, A.; Jiang, Y.; Zhao, Y.; Wang, H.; Ebmeyer, E.; Korzun, V.; Schachschneider, R.; Kazman, E.; et al. Genome-metabolite associations revealed low heritability, high genetic complexity, and causal relations for leaf metabolites in winter wheat (Triticum aestivum). J. Exp. Bot. 2017, 68, 415–428. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Wei, W.; Li, S.; Wang, Y.; Wang, B.; Fan, G.; Zeng, Q.; Zhao, F.; Xu, C.; Zhang, X.; Tang, T.; et al. Metabolome-Based Genome-Wide Association Study Provides Genetic Insights Into the Natural Variation of Foxtail Millet. Front. Plant Sci. 2021, 12, 665530. [Google Scholar] [CrossRef] [PubMed]
  149. Zhang, W.; Alseekh, S.; Zhu, X.; Zhang, Q.; Fernie, A.R.; Kuang, H.; Wen, W. Dissection of the domestication-shaped genetic architecture of lettuce primary metabolism. Plant J. 2020, 104, 613–630. [Google Scholar] [CrossRef]
  150. Wang, Y. A draft genome, resequencing, and metabolomes reveal the genetic background and molecular basis of the nutritional and medicinal properties of loquat (Eriobotrya japonica (Thunb.) Lindl). Hortic. Res. 2021, 8, 231. [Google Scholar] [CrossRef] [PubMed]
  151. Li, H.; Peng, Z.; Yang, X.; Wang, W.; Fu, J.; Wang, J.; Han, Y.; Chai, Y.; Guo, T.; Yang, N.; et al. Genome-wide association study dissects the genetic architecture of oil biosynthesis in maize kernels. Nat. Genet. 2013, 45, 43–50. [Google Scholar] [CrossRef] [PubMed]
  152. Wen, W.; Li, D.; Li, X.; Gao, Y.; Li, W.; Li, H.; Liu, J.; Liu, H.; Chen, W.; Luo, J.; et al. Metabolome-based genome-wide association study of maize kernel leads to novel biochemical insights. Nat. Commun. 2014, 5, 3438. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Riedelsheimer, C.; Lisec, J.; Czedik-Eysenberg, A.; Sulpice, R.; Flis, A.; Grieder, C.; Altmann, T.; Stitt, M.; Willmitzer, L.; Melchinger, A.E. Genome-wide association mapping of leaf metabolic profiles for dissecting complex traits in maize. Proc. Natl. Acad. Sci. USA 2012, 109, 8872–8877. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Lipka, A.E.; Gore, M.A.; Magallanes-Lundback, M.; Mesberg, A.; Lin, H.; Tiede, T.; Chen, H.; Robin Buell, C.; Buckler, E.S.; Rocheford, T.; et al. Genome-Wide Association Study and Pathway-Level Analysis of Tocohromanol Levels in Maize Grain. G3 Genes Genomes Genet. 2013, 3, 1287–1299. [Google Scholar] [CrossRef] [Green Version]
  155. Owens, B.F.; Gore, M.A.; Magallanes-Lundback, M.; Tiede, T.; Diepenbrock, C.H.; Kandianis, C.B.; Kim, E.; Cepela, J.; Mateos-Hernandez, M.; Robin Buell, C.; et al. A Foundation for Provitamin A Biofortification of Maize: Genome-Wide Association and Genomic Prediction Models of Carotenoid Levels. Genetics 2014, 198, 1699–1716. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Liang, X.; Liu, S.; Wang, T.; Li, F.; Cheng, J.; Lai, J.; Qin, F.; Li, Z.; Wang, X.; Jiang, C. Metabolomics-driven gene mining and genetic improvement of tolerance to salt-induced osmotic stress in maize. New Phytol. 2021, 230, 2355–2370. [Google Scholar] [CrossRef]
  157. Suwarno, W.B.; Pixley, K.V.; Palacios-Rojas, N.; Kaeppler, S.M.; Babu, R. Genome-wide association analysis reveals new targets for carotenoid biofortification in maize. Theor. Appl. Genet. 2015, 128, 851–864. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Deng, M.; Li, D.; Luo, J.; Xiao, Y.; Liu, H.; Pan, Q.; Zhang, X.; Jin, M.; Zhao, M.; Yan, J. The genetic architecture of amino acids dissection by association and linkage analysis in maize. Plant Biotechnol. J. 2017, 15, 1250–1263. [Google Scholar] [CrossRef] [Green Version]
  159. Jin, M.; Zhang, X.; Zhao, M.; Deng, M.; Du, Y.; Zhou, Y.; Wang, S.; Tohge, T.; Fernie, A.R.; Willmitzer, L.; et al. Integrated genomics-based mapping reveals the genetics underlying maize flavonoid biosynthesis. BMC Plant Biol. 2017, 17, 17. [Google Scholar] [CrossRef] [Green Version]
  160. Azmach, G.; Menkir, A.; Spillane, C.; Gedil, M. Genetic Loci Controlling Carotenoid Biosynthesis in Diverse Tropical Maize Lines. G3 Genes Genomes Genet. 2018, 8, 1049–1065. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  161. Wang, H.; Xu, S.; Fan, Y.; Liu, N.; Zhan, W.; Liu, H.; Xiao, Y.; Li, K.; Pan, Q.; Li, W.; et al. Beyond pathways: Genetic dissection of tocopherol content in maize kernels by combining linkage and association analyses. Plant Biotechnol. J. 2018, 16, 1464–1475. [Google Scholar] [CrossRef] [PubMed]
  162. Wen, W.; Jin, M.; Li, K.; Liu, H.; Xiao, Y.; Zhao, M.; Alseekh, S.; Li, W.; de Abreu e Lima, F.; Brotman, Y.; et al. An integrated multi-layered analysis of the metabolic networks of different tissues uncovers key genetic components of primary metabolism in maize. Plant J. 2018, 93, 1116–1128. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Alves, M.L.; Bento-Silva, A.; Carbas, B.; Gaspar, D.; Paulo, M.; Brites, C.; Mendes-Moreira, P.; Brites, C.M.; Bronze, M.D.R.; Malosetti, M.; et al. Alleles to Enhance Antioxidant Content in Maize-A Genome-Wide Association Approach. J. Agric. Food Chem. 2020, 68, 4051–4061. [Google Scholar] [CrossRef]
  164. Baseggio, M.; Murray, M.; Magallanes-Lundback, M.; Kaczmar, N.; Chamness, J.; Buckler, E.S.; Smith, M.E.; DellaPenna, D.; Tracy, W.F.; Gore, M.A. Genome-Wide Association and Genomic Prediction Models of Tocochromanols in Fresh Sweet Corn Kernels. Plant Genome 2019, 12, 180038. [Google Scholar] [CrossRef] [Green Version]
  165. Diepenbrock, C.H.; Ilut, D.C.; Magallanes-Lundback, M.; Kandianis, C.B.; Lipka, A.E.; Bradbury, P.J.; Holland, J.B.; Hamilton, J.P.; Wooldridge, E.; Vaillancourt, B.; et al. Eleven biosynthetic genes explain the majority of natural variation in carotenoid levels in maize grain. Plant Cell 2021, 33, 882–900. [Google Scholar] [CrossRef]
  166. Chatham, L.A.; Juvik, J.A. Linking anthocyanin diversity, hue, and genetics in purple corn. G3 Genes Genomes Genet. 2021, 11, jkaa062. [Google Scholar] [CrossRef]
  167. Levina, A.V.; Hoekenga, O.; Gordin, M.; Broeckling, C.; De Jong, W.S. Genetic analysis of potato tuber metabolite composition: Genome-wide association studies applied to a nontargeted metabolome. Crop. Sci. 2021, 61, 591–603. [Google Scholar] [CrossRef]
  168. Dong, X.; Gao, Y.; Chen, W.; Wang, W.; Gong, L.; Liu, X.; Luo, J. Spatiotemporal distribution of phenolamides and the genetics of natural variation of hydroxycinnamoyl spermidine in rice. Mol. Plant 2015, 8, 111–121. [Google Scholar] [CrossRef] [Green Version]
  169. Brotman, Y.; Llorente-Wiegand, C.; Oyong, G.; Badoni, S.; Misra, G.; Anacleto, R.; Parween, S.; Pasion, E.; Tiozon, R.N.; Anonuevo, J.J.; et al. The genetics underlying metabolic signatures in a brown rice diversity panel and their vital role in human nutrition. Plant J. 2021, 106, 507–525. [Google Scholar] [CrossRef]
  170. Chen, W.; Wang, W.; Peng, M.; Gong, L.; Gao, Y.; Wan, J.; Wang, S.; Shi, L.; Zhou, B.; Li, Z.; et al. Comparative and parallel genome-wide association studies for metabolic and agronomic traits in cereals. Nat. Commun. 2016, 7, 12767. [Google Scholar] [CrossRef] [PubMed]
  171. Matsuda, F.; Nakabayashi, R.; Yang, Z.; Okazaki, Y.; Yonemaru, J.I.; Ebana, K.; Yano, M.; Saito, K. Metabolome-genome-wide association study dissects genetic architecture for generating natural variation in rice secondary metabolism. Plant J. 2015, 81, 13–23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Zhang, F.; Guo, H.; Huang, J.; Yang, C.; Li, Y.; Wang, X.; Qu, L.; Liu, X.; Luo, J. A UV-B-responsive glycosyltransferase, OsUGT706C2, modulates flavonoid metabolism in rice. Sci. China Life Sci. 2020, 63, 1037–1052. [Google Scholar] [CrossRef] [PubMed]
  173. Li, X.; Tian, R.; Shao, Z.; Zhang, H.; Chu, J.; Li, W.; Kong, Y.; Du, H.; Zhang, C. Genetic loci and causal genes for seed fatty acids accumulation across multiple environments and genetic backgrounds in soybean. Mol. Breed. 2021, 41, 31. [Google Scholar] [CrossRef]
  174. Wu, D.; Li, D.; Zhao, X.; Zhan, Y.; Teng, W.; Qiu, L.; Zheng, H.; Li, W.; Han, Y. Identification of a candidate gene associated with isoflavone content in soybean seeds using genome-wide association and linkage mapping. Plant J. 2020, 104, 950–963. [Google Scholar] [CrossRef] [PubMed]
  175. Liu, J.Y.; Li, P.; Zhang, Y.W.; Zuo, J.F.; Li, G.; Han, X.; Dunwell, J.M.; Zhang, Y.M. Three-dimensional genetic networks among seed oil-related traits, metabolites and genes reveal the genetic foundations of oil synthesis in soybean. Plant J. 2020, 103, 1103–1124. [Google Scholar] [CrossRef] [PubMed]
  176. Fang, K.; Xia, Z.; Li, H.; Jiang, X.; Qin, D.; Wang, Q.; Wang, Q.; Pan, C.; Li, B.; Wu, H. Genome-wide association analysis identified molecular markers associated with important tea flavor-related metabolites. Hortic. Res. 2021, 8, 42. [Google Scholar] [CrossRef]
  177. Yamashita, H.; Uchida, T.; Tanaka, Y.; Katai, H.; Nagano, A.J.; Morita, A.; Ikka, T. Genomic predictions and genome-wide association studies based on RAD-seq of quality-related metabolites for the genomics-assisted breeding of tea plants. Sci. Rep. 2020, 10, 17480. [Google Scholar] [CrossRef]
  178. Tieman, D.; Zhu, G.; Resende, M.F.R.; Lin, T.; Nguyen, C.; Bies, D.; Rambla, J.L.; Beltran, K.S.O.; Taylor, M.; Zhang, B.; et al. A chemical genetic roadmap to improved tomato flavor. Plant Sci. 2017, 355, 6323. [Google Scholar] [CrossRef]
  179. Sauvage, C.; Segura, V.; Bauchet, G.; Stevens, R.; Do, P.T.; Nikoloski, Z.; Fernie, A.R.; Causse, M. Genome-Wide Association in Tomato Reveals 44 Candidate Loci for Fruit Metabolic Traits. Plant Physiol. 2014, 165, 1120–1132. [Google Scholar] [CrossRef] [Green Version]
  180. Zhu, G.; Wang, S.; Huang, Z.; Zhang, S.; Liao, Q.; Zhang, C.; Lin, T.; Qin, M.; Peng, M.; Yang, C.; et al. Rewiring of the Fruit Metabolome in Tomato Breeding. Cell 2018, 172, 249–261.e12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  181. Burgos, E.; Belen De Luca, M.; Diouf, I.; de Haro, L.A.; Albert, E.; Sauvage, C.; Tao, Z.J.; Bermudez, L.; Asís, R.; Nesi, A.N.; et al. Validated MAGIC and GWAS population mapping reveals the link between vitamin E content and natural variation in chorismate metabolism in tomato. Plant J. 2021, 105, 907–923. [Google Scholar] [CrossRef] [PubMed]
  182. Chen, J.; Hu, X.; Shi, T.; Yin, H.; Sun, D.; Hao, Y.; Xia, X.; Luo, J.; Fernie, A.R.; He, Z.; et al. Metabolite-based genome-wide association study enables dissection of the flavonoid decoration pathway of wheat kernels. Plant Biotechnol. J. 2020, 18, 1722–1735. [Google Scholar] [CrossRef] [PubMed]
  183. Meuwissen, T.H.E.; Hayes, B.J.; Goddard, M.E. Prediction of total genetic value using genome-wide dense marker maps. Genetics 2001, 157, 1819–1829. [Google Scholar] [CrossRef] [PubMed]
  184. Duangjit, J.; Causse, M.; Sauvage, C. Efficiency of genomic selection for tomato fruit quality. Mol. Breed. 2016, 36, 29. [Google Scholar] [CrossRef]
  185. Battenfield, S.D.; Guzmán, C.; Gaynor, R.C.; Singh, R.P.; Peña, R.J.; Dreisigacker, S.; Fritz, A.K.; Poland, J.A. Genomic Selection for Processing and End-Use Quality Traits in the CIMMYT Spring Bread Wheat Breeding Program. Plant Genome 2016, 9, plantgenome2016-01. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Dempewolf, H.; Baute, G.; Anderson, J.; Kilian, B.; Smith, C.; Guarino, L. Past and Future Use of Wild Relatives in Crop Breeding. Crop. Sci. 2017, 57, 1070–1082. [Google Scholar] [CrossRef]
  187. Bouis, H.E.; Saltzman, A. Improving nutrition through biofortification: A review of evidence from HarvestPlus, 2003 through 2016. Glob. Food Sec. 2017, 12, 49–58. [Google Scholar] [CrossRef] [PubMed]
  188. HarvestPlus. Getting Biofortified Food on Everyone’s Plate—2019 Annual Report; HarvestPlus: Washington, DC, USA, 2019. [Google Scholar]
  189. Sazawal, S.; Dhingra, U.; Dhingra, P.; Dutta, A.; Deb, S.; Kumar, J.; Devi, P.; Prakash, A. Efficacy of high zinc biofortified wheat in improvement of micronutrient status, and prevention of morbidity among preschool children and women—A double masked, randomized, controlled trial. Nutr. J. 2018, 17, 86. [Google Scholar] [CrossRef] [Green Version]
  190. Palmer, A.C.; Healy, K.; Barffour, M.A.; Siamusantu, W.; Chileshe, J.; Schulze, K.J.; West, K.P.J.; Labrique, A.B. Provitamin A Carotenoid-Biofortified Maize Consumption Increases Pupillary Responsiveness among Zambian Children in a Randomized Controlled Trial. J. Nutr. 2016, 146, 2551–2558. [Google Scholar] [CrossRef] [Green Version]
  191. Low, J.W.; Mwanga, R.O.M.; Andrade, M.; Carey, E.; Ball, A.-M. Tackling vitamin A deficiency with biofortified sweetpotato in sub-Saharan Africa. Glob. Food Sec. 2017, 14, 23–30. [Google Scholar] [CrossRef] [PubMed]
  192. Prasanna, B.M.; Palacios-Rojas, N.; Hossain, F.; Muthusamy, V.; Menkir, A.; Dhliwayo, T.; Ndhlela, T.; San Vicente, F.; Nair, S.K.; Vivek, B.S.; et al. Molecular Breeding for Nutritionally Enriched Maize: Status and Prospects. Front. Genet. 2020, 10, 1392. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Cruet-Burgos, C.; Cox, S.; Ioerger, B.P.; Perumal, R.; Hu, Z.; Herald, T.J.; Bean, S.R.; Rhodes, D.H. Advancing provitamin A biofortification in sorghum: Genome-wide association studies of grain carotenoids in global germplasm. Plant Genome 2020, 13, e20013. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Izquierdo, P.; Astudillo, C.; Blair, M.W.; Iqbal, A.M.; Raatz, B.; Cichy, K.A. Meta-QTL analysis of seed iron and zinc concentration and content in common bean (Phaseolus vulgaris L.). Theor. Appl. Genet. 2018, 131, 1645–1658. [Google Scholar] [CrossRef] [PubMed]
  195. Krishnappa, G.; Rathan, N.D.; Sehgal, D.; Ahlawat, A.K.; Singh, S.K.; Singh, S.K.; Shukla, R.B.; Jaiswal, J.P.; Solanki, I.S.; Singh, G.P.; et al. Identification of Novel Genomic Regions for Biofortification Traits Using an SNP Marker-Enriched Linkage Map in Wheat (Triticum aestivum L.). Front. Nutr. 2021, 8, 669444. [Google Scholar] [CrossRef] [PubMed]
  196. Garg, M.; Sharma, A.; Vats, S.; Tiwari, V.; Kumari, A.; Mishra, V.; Krishania, M. Vitamins in Cereals: A Critical Review of Content, Health Effects, Processing Losses, Bioaccessibility, Fortification, and Biofortification Strategies for Their Improvement. Front. Nutr. 2021, 8, 586815. [Google Scholar] [CrossRef] [PubMed]
  197. Bhatt, V.; Muthysamy, V.; Jha, S.; Zunjare, R.U.; Baveja, A.; Sosad, S. Development of low phytic acid maize through marker assisted introgression of lpa1-1 and lpa2-1 genes. In Proceedings of the 13th Asian Maize Conference on and Expert Consultation on Maize for Food, Feed, Nutrition and Environmental Security, Ludhiana, India, 8–10 October 2018; pp. 143–144. [Google Scholar]
  198. Storozhenko, S.; De Brouwer, V.; Volckaert, M.; Navarrete, O.; Blancquaert, D.; Zhang, G.F.; Lambert, W.; Van Der Straeten, D. Folate fortification of rice by metabolic engineering. Nat. Biotechnol. 2007, 25, 1277–1279. [Google Scholar] [CrossRef] [PubMed]
  199. Díaz de La Garza, R.I.; Gregory, J.F.; Hanson, A.D. Folate biofortification of tomato fruit. Proc. Natl. Acad. Sci. USA 2007, 104, 4218–4222. [Google Scholar] [CrossRef] [Green Version]
  200. Bulley, S.; Wright, M.; Rommens, C.; Yan, H.; Rassam, M.; Lin-Wang, K.; Andre, C.; Brewster, D.; Karunairetnam, S.; Allan, A.C.; et al. Enhancing ascorbate in fruits and tubers through over-expression of the L-galactose pathway gene GDP-L-galactose phosphorylase. Plant Biotechnol. J. 2012, 10, 390–397. [Google Scholar] [CrossRef]
  201. Pierce, E.C.; LaFayette, P.R.; Ortega, M.A.; Joyce, B.L.; Kopsell, D.A.; Parrott, W.A. Ketocarotenoid production in soybean seeds through metabolic engineering. PLoS ONE 2015, 10, e0138196. [Google Scholar] [CrossRef] [Green Version]
  202. Holme, I.B.; Dionisio, G.; Brinch-Pedersen, H.; Wendt, T.; Madsen, C.K.; Vincze, E.; Holm, P.B. Cisgenic barley with improved phytase activity. Plant Biotechnol. J. 2012, 10, 237–247. [Google Scholar] [CrossRef] [PubMed]
  203. Kamrani, M.; Kohnehrouz, B.B.; Gholizadeh, A. Cisgenic inhibition of the potato cold induced phosphorylase L gene expression and decrease in sugar contents. Afr. J. Biotechnol. 2011, 10, 10076–10082. [Google Scholar] [CrossRef]
  204. De Vetten, N.; Wolters, A.M.; Raemakers, K.; Van der Meer, I.; Ter Stege, R.; Heeres, E.; Heeres, P.; Visser, R. A transformation method for obtaining marker-free plants of a cross-pollinating and vegetatively propagated crop. Nat. Biotechnol. 2003, 21, 439–442. [Google Scholar] [CrossRef] [PubMed]
  205. Chawla, R.; Shakya, R.; Rommens, C.M. Tuber-specific silencing of asparagine synthetase-1 reduces the acrylamide-forming potential of potatoes grown in the field without affecting tuber shape and yield. Plant Biotechnol. J. 2012, 10, 913–924. [Google Scholar] [CrossRef] [PubMed]
  206. Gil-Humanes, J.; Pistón, F.; Hernando, A.; Alvarez, J.B.; Shewry, P.R.; Barro, F. Silencing of γ-gliadins by RNA interference (RNAi) in bread wheat. J. Cereal Sci. 2008, 48, 565–568. [Google Scholar] [CrossRef]
  207. Brinch-Pedersen, H.; Borg, S.; Tauris, B.; Holm, P.B. Molecular genetic approaches to increasing mineral availability and vitamin content of cereals. J. Cereal Sci. 2007, 46, 308–326. [Google Scholar] [CrossRef]
  208. Coca, M.; Peñas, G.; Gómez, J.; Campo, S.; Bortolotti, C.; Messeguer, J.; Segundo, B.S. Enhanced resistance to the rice blast fungus Magnaporthe grisea conferred by expression of a cecropin A gene in transgenic rice. Planta 2006, 223, 392–406. [Google Scholar] [CrossRef] [Green Version]
  209. Girgi, M.; Breese, W.A.; Lörz, H.; Oldach, K.H. Rust and Downy Mildew Resistance in Pearl Millet (Pennisetum glaucum) Mediated by Heterologous Expression of the afp Gene from Aspergillus giganteus. Transgenic Res. 2006, 15, 313–324. [Google Scholar] [CrossRef]
  210. Yadav, H.; Malik, K.; Kumar, S.; Jaiwal, P.K. Comparative regeneration in six bread wheat (Triticum aestivum L.) varieties from immature and mature scutella for developing efficient and genotype-independent protocol prerequisite for genetic improvement of wheat. Vitr. Cell. Dev. Biol.-Plant 2020, 56, 610–617. [Google Scholar] [CrossRef]
  211. Sainger, M.; Chaudhary, D.; Dahiya, S.; Jaiwal, R.; Jaiwal, P.K. Development of an efficient in vitro plant regeneration system amenable to Agrobacterium-mediated transformation of a recalcitrant grain legume blackgram (Vigna mungo L. Hepper). Physiol. Mol. Biol. Plants 2015, 21, 505–517. [Google Scholar] [CrossRef] [Green Version]
  212. Garg, M.; Sharma, N.; Sharma, S.; Kapoor, P.; Kumar, A.; Chunduri, V.; Arora, P. Biofortified crops generated by breeding, agronomy, and transgenic approaches are improving lives of millions of people around the world. Front. Nutr. 2018, 5, 12. [Google Scholar] [CrossRef] [PubMed]
  213. Tang, G.; Qin, J.; Dolnikowski, G.G.; Russell, R.M.; Grusak, M.A. Golden Rice is an effective source of vitamin A. Am. J. Clin. Nutr. 2009, 89, 1776–1783. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Smyth, S.J. Genetically modified crops, regulatory delays, and international trade. Food Energy Secur. 2017, 6, 78–86. [Google Scholar] [CrossRef] [Green Version]
  215. Schouten, H.J.; Krens, F.A.; Jacobsen, E. Do cisgenic plants warrant less stringent oversight? Nat. Biotechnol. 2006, 24, 753. [Google Scholar] [CrossRef] [PubMed]
  216. Rommens, C.M. All-native DNA transformation: A new approach to plant genetic engineering. Trends Plant Sci. 2004, 9, 457–464. [Google Scholar] [CrossRef] [PubMed]
  217. Jiang, M.; Liu, Y.; Liu, Y.; Tan, Y.; Huang, J.; Shu, Q. Mutation of Inositol 1,3,4-trisphosphate 5/6-kinase6 Impairs Plant Growth and Phytic Acid Synthesis in Rice. Plants 2019, 8, 114. [Google Scholar] [CrossRef] [Green Version]
  218. Sun, Y.; Jiao, G.; Liu, Z.; Zhang, X.; Li, J.; Guo, X.; Du, W.; Du, J.; Francis, F.; Zhao, Y.; et al. Generation of high-amylose rice through CRISPR/Cas9-mediated targeted mutagenesis of starch branching enzymes. Front. Plant Sci. 2017, 8, 298. [Google Scholar] [CrossRef]
  219. Schouten, H.J.; Krens, F.A.; Jacobsen, E. Cisgenic plants are similar to traditionally bred plants: International regulations for genetically modified organisms should be altered to exempt cisgenesis. EMBO Rep. 2006, 7, 750–753. [Google Scholar] [CrossRef] [Green Version]
  220. Zhu, S.; Duwal, A.; Su, Q.; Vossen, J.H.; Visser, R.G.F.; Jacobsen, E. Vector integration in triple R gene transformants and the clustered inheritance of resistance against potato late blight. Transgenic Res. 2013, 22, 315–325. [Google Scholar] [CrossRef]
  221. Conner, A.J.; Barrell, P.J.; Baldwin, S.J.; Lokerse, A.S.; Cooper, P.A.; Erasmuson, A.K.; Nap, J.P.; Jacobs, J.M.E. Intragenic vectors for gene transfer without foreign DNA. Euphytica 2007, 154, 341–353. [Google Scholar] [CrossRef]
  222. Schouten, H.J.; Jacobsen, E. Cisgenesis and intragenesis, sisters in innovative plant breeding. Trends Plant Sci. 2008, 13, 260–261. [Google Scholar] [CrossRef] [PubMed]
  223. Bhatnagar, M.; Prasad, K.; Bhatnagar-Mathur, P.; Narasu, M.L.; Waliyar, F.; Sharma, K.K. An efficient method for the production of marker-free transgenic plants of peanut (Arachis hypogaea L.). Plant Cell Rep. 2010, 29, 495–502. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Doshi, K.M.; Eudes, F.; Laroche, A.; Gaudet, D. Anthocyanin expression in marker free transgenic wheat and triticale embryos. Vitr. Cell. Dev. Biol.-Plant 2007, 43, 429–435. [Google Scholar] [CrossRef]
  225. Vanblaere, T.; Szankowski, I.; Schaart, J.; Schouten, H.; Flachowsky, H.; Broggini, G.A.L.; Gessler, C. The development of a cisgenic apple plant. J. Biotechnol. 2011, 154, 304–311. [Google Scholar] [CrossRef]
  226. Ling, F.; Zhou, F.; Chen, H.; Lin, Y. Development of marker-free insect-resistant indica rice by Agrobacterium tumefaciens-mediated co-transformation. Front. Plant Sci. 2016, 7, 1608. [Google Scholar] [CrossRef] [Green Version]
  227. Delwaide, A.C.; Nalley, L.L.; Dixon, B.L.; Danforth, D.M.; Nayga, R.M.; Van Loo, E.J.; Verbeke, W. Revisiting GMOs: Are There Differences in European Consumers’ Acceptance and Valuation for Cisgenically vs Transgenically Bred Rice? PLoS ONE 2015, 10, e0126060. [Google Scholar] [CrossRef] [Green Version]
  228. Shew, A.M.; Nalley, L.L.; Danforth, D.M.; Dixon, B.L.; Nayga, R.M.; Delwaide, A.C.; Valent, B. Are all GMOs the same? Consumer acceptance of cisgenic rice in India. Plant Biotechnol. J. 2016, 14, 4–7. [Google Scholar] [CrossRef] [Green Version]
  229. Edenbrandt, A.K.; Gamborg, C.; Thorsen, B.J. Consumers’ Preferences for Bread: Transgenic, Cisgenic, Organic or Pesticide-free? J. Agric. Econ. 2018, 69, 121–141. [Google Scholar] [CrossRef]
  230. Raj, R.S.; Singh, C.; Modi, A.; Subhash, N. Genetic transformation of lowland rice variety GR11 for drought tolerance and its ratification for upland paddy cultivation. Indian J. Genet. Plant Breed. 2015, 75, 30–40. [Google Scholar] [CrossRef]
  231. Roberts, R.J. The Nobel Laureates’ Campaign Supporting GMOs. J. Innov. Knowl. 2018, 3, 61–65. [Google Scholar] [CrossRef]
  232. June, M.K. International Service for the Acquisition of Agri-Biotech (ISAAA). Global Status of Commercialized Biotech/GM Crops: 2019; ISAAA: Ithaca, NY, USA, 2020; ISBN 978-1-892456-69-9. [Google Scholar]
  233. Purchase, I.F.H. What determines the acceptability of genetically modified food that can improve human nutrition? Toxicol. Appl. Pharmacol. 2005, 207, 19–27. [Google Scholar] [CrossRef] [PubMed]
  234. Davison, J.; Ammann, K. New GMO regulations for old: Determining a new future for EU crop biotechnology. GM Crop. Food 2017, 8, 13–34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. McDougall, P. R&D Trends for Chemical Crop Protection Products and the Position of the European Market; Phillips McDougall Ltd.: Pathhead, UK, 2013. [Google Scholar]
  236. 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] [PubMed]
  237. Hou, H.; Atlihan, N.; Lu, Z.X. New biotechnology enhances the application of cisgenesis in plant breeding. Front. Plant Sci. 2014, 5, 389. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  238. Russell, A.W.; Sparrow, R. The case for regulating intragenic GMOs. J. Agric. Environ. Ethics 2008, 21, 153–181. [Google Scholar] [CrossRef]
  239. EFSA Panel on Genetically Modified Organisms (GMO). Scientific opinion addressing the safety assessment of plants developed through cisgenesis and intragenesis. EFSA J. 2012, 10, 2561. [Google Scholar] [CrossRef] [Green Version]
  240. European Food Safety Authority (EFSA). International Scientific Workshop ‘Risk Assessment Considerations for RNAi-Based GM Plants’; EFSA Supporting Publication: Brussels, Belgium, 2014; Volume EN-705. [Google Scholar]
  241. Environmental Protection Agency (EPA). Pesticides; Data Requirements for Plant.-Incorporated Protectants (PIPs) and Certain Exemptions for PIPs; EPA: Whasington, DC, USA, 2011; Volume 76.
  242. De Steur, H.; Wesana, J.; Blancquaert, D.; Van Der Straeten, D.; Gellynck, X. The socioeconomics of genetically modified biofortified crops: A systematic review and meta-analysis. Ann. N. Y. Acad. Sci. 2017, 1390, 14–33. [Google Scholar] [CrossRef] [Green Version]
  243. Xu, Y.; Li, P.; Zou, C.; Lu, Y.; Xie, C.; Zhang, X.; Prasanna, B.M.; Olsen, M.S. Enhancing genetic gain in the era of molecular breeding. J. Exp. Bot. 2017, 68, 2641–2666. [Google Scholar] [CrossRef]
Figure 1. Comparison of the SNP genotyping techniques most commonly used in crops grouped by the platforms in the throughput level.
Figure 1. Comparison of the SNP genotyping techniques most commonly used in crops grouped by the platforms in the throughput level.
Ijms 23 03086 g001
Figure 2. Construction of breeding populations: (A) recombinant inbred lines (RILs); (B) near isogenic lines (NILs); (C) advanced backcross; and (D) their use for Genomic Selection (GS). Only some of the possible crossing designs are shown.
Figure 2. Construction of breeding populations: (A) recombinant inbred lines (RILs); (B) near isogenic lines (NILs); (C) advanced backcross; and (D) their use for Genomic Selection (GS). Only some of the possible crossing designs are shown.
Ijms 23 03086 g002
Figure 3. Schematic representation of three modern biotechnology techniques to introduce allelic variants of interest in a recipient organism: (A) transgeneis; (B) cisgenesis; and (C) intragenesis.
Figure 3. Schematic representation of three modern biotechnology techniques to introduce allelic variants of interest in a recipient organism: (A) transgeneis; (B) cisgenesis; and (C) intragenesis.
Ijms 23 03086 g003
Table 1. List of the main public SNP databases in food crops. The type of information available ranges from the marker description to the genotype and map and/or genome location.
Table 1. List of the main public SNP databases in food crops. The type of information available ranges from the marker description to the genotype and map and/or genome location.
Database NameUrlCrop
CerealsDBhttps://www.cerealsdb.uk.net/cerealgenomics/CerealsDB/indexNEW.php, accessed on 17 January 2022Bread wheat (Triticum eastivum L.)
Chickpea SNP-InDel Database (CicArVarDB)https://cegresources.icrisat.org/cicarvardb, accessed on the 17 of January 2022Chickpea (Cicer arietinum L.)
CropSNPdbhttp://snpdb.appliedbioinformatics.com.au/, accessed on 17 January 2022Bread wheat (T. eastivum L.)
Cabbage (Brassica rapa L.)
Cauliflower (Brassica oleracea L.)
Indian mustard (Brassica juncea L.)
Oilseed rape (Brassica napus L.)
Cucurbit Genomics Database (CuGeDG)http://cucurbitgenomics.org/, accessed on 17 January 2022Cucumber (Cucumis sativus L.)
Melon (Cucumis melo L.)
Pumpkin (Cucurbita spp.)
Watermelon (Citrullus lanatus Thumb.)
Genome Database for Rosaceae (GDR)https://www.rosaceae.org, accessed on 17 January 2022Apple tree (Malus spp.)
Blackberry (Rubus spp.)
Peach tree (Prunus spp.)
Pear tree (Pyrus spp.)
Strawberry (Fragaria spp.)
Gramenehttps://www.gramene.org, accessed on 17 January 2022African rice (Oryza galberrina Steud)
Asian rice (Oryza sativa L.)
Barley (Hordeum vulgare L.)
Foxtail millet (Setaria italica (L.) Beauv.)
Maize (Zea mays L.)
Sorghum (Sorghum bicolor (L.) Moench)
Wheat (Triticum spp.)
Kazusa Tomato Genomics Database (KaTomicsDB)https://www.kazusa.or.jp/tomato/, accessed on 17 January 2022Tomato (Solanum lycopersicum L.)
Lettuce Genome Database (LettuceGDB)https://www.lettucegdb.com, accessed on 17 January 2022Lettuce (Lactuca sativa L.)
Maize Genetics and Genomics Database (MaizeGDB)https://www.maizegdb.org/, accessed on 17 January 2022Maize (Z. mays L.)
Maize SNP-DNA Fingerprint Databasehttp://doi.org/10.3390/agriculture11070597 (Tables S1 and S2; [127]), accessed on 18 January 2022Maize (Z. mays L.)
Q-TARO (QTL Annotation Rice Online) databasehttp://qtaro.abr.affrc.go.jp/index.html, accessed on 18 January 2022Asian rice (O. sativa L.)
SNP genotype database for avocadohttps://doi.org/10.1007/s11295-019-1374-1 (Table S2; [128]), accessed on 18 January 2022Avocado (Persea americana Mill.)
Sol Genomics Networkhttps://solgenomics.net, accessed on 18 January 2022Tomato (S. lycopersicum L.)
SorGSDhttps://ngdc.cncb.ac.cn/sorgsd, accessed on 18 January 2022Sorghum (S. bicolor (L.) Moench)
SpinachBasehttp://www.spinachbase.org, accessed on 19 January 2022Spinach (Spinacia oleracea L.)
Rice SNP-Seek Databasehttps://snp-seek.irri.org, accessed on 19 January 2022Asian rice (O. sativa L.)
The IPK Crop EST Database (CR-EST)http://pgrc.ipk-gatersleben.de/cr-est, accessed on 19 January 2022Barley (H. vulgare L.)
Bread wheat (T. eastivum L.)
Pea (Pisum sativum L.)
Potato (Solanum tuberosum L.)
The Tomato Integrated Database (Tomatonics)http://plantomics.mind.meiji.ac.jp/tomatomics, accessed on 19 January 2022Tomato (S. lycopersicum L.)
TropGENE-DBhttp://tropgenedb.cirad.fr/tropgene/JSP/index.jsp, accessed on 19 January 2022Asian rice (O. sativa L.)
Banana (Musa acuminata Juss.)
Bread fruit (Artocarpus altilis (Parkinson) Fosberg)
Cassava (Manihot esculenta Crantz)
Clemetine (Citrus clementina L.)
Cocoa (Theobroma cacao L.)
Coconut (Cocos nucifera L.)
Coffee (Coffea canephora L.)
Cupuassu (Theobroma grandiflorum Schum.)
Oil palm (Elaeis guineensis Jacq.)
Pummelo (Citrus grandis (L.) Osbeck)
Sorghum (S. bicolor L. Moench)
Sugarcane (Saccharum officinarus L.)
Sweet orange (Citrus sinensis Osbeck)
Vitis International Variety Catalogue (VIVC)https://www.vivc.de/index.php?r=site%2Findex, accessed on 19 January 2022Grapevine (Vitis spp.)
Even if there are more species in some databases, they were not included if there is no SNP information available or they are not food crops.
Table 2. Metabolomic genome-wide association studies (mGWAS). Only groups of compounds that play an important role in human nutrition and/or health status are shown.
Table 2. Metabolomic genome-wide association studies (mGWAS). Only groups of compounds that play an important role in human nutrition and/or health status are shown.
CropSpeciesAnalytical Technique MetaboliteReference
Apple treeMalus × domestica Borkh.UHPLC–ESI-QTOF-MS, NMRFlavonoids, polyphenols, sugars, terpenoids[141]
BarleyH. vulgareHPLC-FL, HPLC-MS, IC-MS/MSAmino acids, glutathione, organic acids, starch, sugars, vitamin E (tocopherol)[142]
HPAEC-PAD, HPLC-ELSD, HPLC-MALDITOF-MSSugars[143]
HPLC-Fluorescence detectionCarotenoids (i.e., tocopherols and tocotrienols: vitamin E)[144]
Barley
Bread wheat
Maize
Potato
Rice
Sweet orange tree
H. vulgare
T. aestivum
Z.mays
S. tuberosum
O.sativa
Citrus xsinensis (L.) Osbeck
GC-TOF-MSFlavonoids[145]
BlueberryVaccinium spp.GC-MSFatty acids, phenylpropanoids, terpenoids[146]
Bread wheatT.aestivumGC-MSAmino acids, organic acid ‡‡ sugars[147]
Foxtail milletS.italicaHPLC-ESI-QTRAP-MS/MSAlkaloids, amino acids, fatty acids, organic acids, phenolamides, polyphenols (i.e., flavonoids, anthocyanins...), sugars, vitamins[148]
LettuceL.sativaGC-TOF-MSAlkaloids, amino acids, organic acids, polyamines, polyphenols, sugars, vitamins, etc.[149]
LoquatEriobotryajaponica (Thunb.) Lindl.UPLC-ESI-MS/MSAlkaloids, flavonoids, phenolic acids, polysaccharides, terpenoids[150]
MaizeZ.maysLC-MS/MSFatty acids[151]
LC-ESI-(QTRAP or QqTOF)-MS/MSAmino acids, fatty acids, flavonoids[152]
GC-MSAmino acids, organic acids, phenylpropanoids[153]
HPLC-Fluorescence detectionTocochromanols (tocopherols and tocotrienols)[154]
HPLC-PDACarotenoids[155]
UPLC-HRMSAmino acids, fatty acids, flavonoids, benzoxazinoids, terpenoids[156]
HPLC, UPLCCarotenoids[157]
CECAmino acids[158]
LC-ESI-QqTOF-MS/MSFlavonoids[159]
HPLCCarotenoids[160]
UPLC-PDATocopherol (part of vitamin E)[161]
GC-TOF-MSAmino acids, (poly)amines, organic acids, sugars, vitamin E (tocopherol)[162]
HPLC-PDA, HPLC-fluorescence detectionCarotenoids, phenolics, tocopherol (a form of vitamin E)[163]
HPLC-fluorescence detectionCarotenoids (i.e., tocopherols and tocotrienols: vitamin E)[164]
HPLC-PDACarotenoids (α-carotene, β-carotene, β-cryptoxanthin, lutein, phytofluene, zeaxanthin, zeinoxanthin)[165]
HPLC-UV/VisAnthocyanins[166]
PotatoS.tuberosumUPLC-Q-TOF-MSAlkaloids, amino acids[167]
RiceO.sativaLC-ESI-Q TRAP-MS/MSPhenolamides[168]
GC-TOF-MSAmino acids, flavonoids, organic acids[169]
LC-ESI-MS/MSAmino acids, fatty acids, flavonoids[170]
HPLC-ESI-QTOF/MSAmino acids, flavonoids, phenolamines, terpenoids[171]
HPLC-ESI-(QTRAP or QqTOF)-MSAmino acids, flavonoids, phenolamines, terpenoids[70]
LC-ESI-Q TRAP-MS/MSFlavonoids[172]
SoybeanGlycinemax L.GCFatty acids[173]
HPLC-DADIsoflavones[174]
HPLC-MSAminoacids, isoflavones, lipids, organic acids[175]
TeaCamelliasinensis L.HPLCTheanine, caffeine, catechins[176]
HPLC-PDAAmino acids, caffeine, catechins[177]
TomatoS.lycopersicumGC-MSOrganic acids, sugars[178]
GC-MSAmino acids, organic acid ‡‡, sugars[179]
HPLC-MS/MSAlkaloids ‡‡‡[180]
GC-MSFatty acids, lipids, carotenoids (i.e., tocopherols and tocotrienols: vitamin E)[181]
WheatT.aestivumHPLC-ESI-QTRAP-MS/MSAmino acids, (poly)amines, flavonoids, organic acids, sugars, vitamins, etc.[182]
CEC: cation exchange chromatography; ELSD: evaporative light scattering detection; GC: gas chromatography; GC-MS: GC mass spectrometry; GC-TOF-MS: GC time-of-flight mass spectrometry; HPAEC-PAD: high-pH anion-exchange chromatography with pulsed amperometric detection; HPLC: high-performance liquid chromatography; HPLC-ESI-(QTRAP or QqTOF)-MS: HPLC-ESI-quadrupole TRAP or TOF tandem mass spectrometry; HPLC-MALDITOF-MS: HPLC matrix-assisted laser desorption–ionization time-of-flight mass spectrometry; IC-MS/MS: ion chromatography tandem mass spectrometry; LC-ESI-MS/MS: liquid chromatography–electrospray ionization tandem mass spectrometry; LC-Q-TOF-MS: liquid chromatography quadrupole TOF mass spectrometry; NMR: nuclear magnetic resonance; UPLC-ESI-MS/MS: ultra-high-performance liquid chromatography ESI tandem mass spectrometry; UPLC-HRMS: UPLC high-resolution mass spectrometry. ‡‡ Oxalic acid (anti-nutrient). ‡‡‡ Steroidal glycoalkaloids (SGAs): most of them are considered anti-nutrients.
Table 3. Biofortified crops through different techniques.
Table 3. Biofortified crops through different techniques.
TechniqueCropMethodBiofortified TraitReference
Conventional breedingRiceBackcrosses between a high-yielding cultivar and the IR68144 lineA 2.54-fold increase in iron and 1.54-fold increase in zinc[4]
MaizeBackcrosses involving diverse exotic donor linesLines with high provitamin A content by accumulating mainly high β-carotene and lines with high provitamin A by promoting accumulation of high levels of both carotenes and xanthophylls[13]
Marker-assisted introgression of lpa1-1 and lpa2-1 alleles in elite lines of provitamin A-enriched quality protein maize (QPM)A reduction in phytic acid content and improvement in the mineral bioavailability in lines of QPM rich in provitamin A[197]
Introgression of VTE4 (γ-tocopherol
methyl transferase) allele into four provitamin-A rich QPM elite inbreds using marker-assisted backcross breeding
An increase in α-tocopherol to 15.2 ppm over 8.0 ppm in the original inbreds[14]
WheatMarker-assisted introgression of group 4 and 7 chromosomes of the wild ancestor Aegilops peregrina in a commercial variety of wheatHigher content in iron and zinc in wheat grains[5]
Backcrosses between low-yielding exotic donor lines and commercial varietiesBlack, purple, and blue lines with high content in anthocyanins[38]
CassavaRapid cycling recurrent selectionSignificant gains for total carotenoid content and total β-carotene[15]
Potato‘Atlantic’ and 17 4x-2x hybrids between S. tuberosum and diploid hybrids of Solanum phureja-Solanum stenotomumHigher contents of copper, iron, manganese, and zinc[6]
TomatoBackcrosses between landraces of tomatoHybrid with increased concentration of polyphenols and high antioxidant activity in pink ripeness stage[39]
BeanBackcrosses between low and high mineral genotypes using a QTL mapping approachIncreased iron and zinc content[7]
ChickpeaCrosses between different cultivarsHigher content of carotenoids[16]
TransgenesisRiceEndosperm-specific overexpression of Arabidopsis thaliana GTP cyclohydrolase I (GTPCHI) and aminodeoxychorismate synthase (ADCS) genesAn enhancement of 100 times in folate[198]
Overexpression of phytoene synthaseHigher content in β-carotene[17]
Expression of four synthetic genes: sZmPSY1, sPaCrtI, sCrBKT, and sHpBHY (for phytoene synthase, phytoene desaturase, β-carotene ketolase, and β-carotene hydroxylase, respectively)Synthesis de novo of the carotenoid astaxanthin[40]
Coexpression of an Arabidopsis nicotianamine synthase (AtNAS1), bean ferritin (PvFerritin), bacterial carotene desaturase (CRTI), and maize phytoene synthase (ZmPSY)Simultaneous increase in iron, zinc, and β-carotene content in the rice endosperm[29]
Constitutive overexpression of the rice GDP-L-galactose phosphorylase (35S-OsGGP) geneIncrease in ascorbate concentrations in germinated brown rice[18]
Expression bacterial aspartate kinase (AK) and dihydrodipicolinate synthase (DHPS), downregulation of rice lysine ketoglutarate reductase/saccharopine dehydrogenase (LKR/SD) and selection of marker-free transgenic linesUp to 25-fold increase in free lysine levels[36]
Expression of an AmA1 gene from Amaranthus hypochondriacusA significant increase in the content of several EAAs, including lysine, threonine, and valine, as well as a 1.06~12.87% increase in the total protein content[37]
MaizeOverexpression of the bacterial genes crtB (for phytoene synthase) and crtI (for the four desaturation steps of the carotenoid pathway) under the control of a endosperm-specific promoterAn increase in total carotenoids of up to 34-fold with a preferential accumulation of β-carotene in the maize endosperm[19]
Endosperm-specific overexpression of soybean ferritinA 2-fold improvement in seed iron bioavailability[8]
Coexpression of Gm8gGCHI and GmADCS genes driven by endosperm-specific promotersA 4.2-fold increase in folate (vitamin B9) level in transgenic maize grains[20]
Insertion of the lysine-rich sb401 geneSignificantly higher levels of lysine total protein in maize seeds[35]
WheatConstitutive expression of the rice nicotianamine synthase 2 (OsNAS2) geneHigher concentrations of grain iron and zinc, and enhanced localization of iron and zinc in endosperm and crease tissues, respectively[9]
CassavaCoexpression of ferritin (FER1) and mutated Iron transporter (IRT1) from A. thalianaAccumulation of iron levels 7–18 times higher and zinc levels 3–10 times higher[10]
PotatoOverexpression of AtGTPCHI, AtADCS, OsHPPK/DHPS and AtFPGS genesA 2-fold increase in folate content in mature tubers and stable accumulation of folates for up to 9 months of storage[21]
Simultaneous expression of Wrinkled 1 (WRI1), Diacylglycerol acyltransferase 1 (DGAT1) and Oleosin under the transcriptional control of tuber-specific (patatin) and constitutive (CaMV-35S) promoters.Over a 100-fold increase in triacylglycerol accumulation to levels up to 3.3% of tuber dry weigh[33]
Sweet PotatoExpression of a barley NA synthase 1 (HvNAS1) geneA 3- and 2.9-fold increase in the concentrations of iron and zinc, respectively[11]
TomatoCross between GTPCHI and ADCS overexpressing plantsA 25-fold more in folate (Vitamin B9) level in fruits[199]
Overexpression of an A. thaliana Orange (AtOR) geneAn increase in total carotenoids in fruits[22]
Overexpression of GDP-l-galactose phosphorylase (GGP) gene from Actinidia chinensis under the control of the 35S promoterA 3- to 6-fold higher content in ascorbic acid in fruits[200]
Fruit-specific expression of the transcription factor AtMYB12Increased content of different phenylpropanoids[23]
StrawberryOverexpression of a GDP-l-galactose phosphorylase (GGP) gene from Actinidia chinensis under the control of the 35S promoterA 2-fold higher content in ascorbic acid in fruits[200]
BananaExpression of a Fe’i banana-derived phytoene synthase (MtPsy2a) gene under the maize polyubiquitin promoterEnhanced β-carotene content in fruit[24]
SoybeanOverexpression of the bacterial genes crtB (for phytoene synthase) and crtW and bkt1 (ketolase genes) under the control of seed-specific promotersEnhanced accumulation of ketocarotenoids in seeds[201]
Overexpression of adenosine 5’-phosphosulfate sulfurylase 1Higher amounts of sulfate, cysteine, and some sulfur-containing secondary metabolites in seeds[34]
Overexpression of a GmDGAT2A gene driven by a seed-specific promoter of Gmole1Significantly increased linoleic acid content specifically and total oil content[32]
BeanSeed-specific overexpression of a GTP cyclohydrolase I gene from Arabidopsis (AtGchI)Increased folate levels in raw desiccated seeds by up to 3-fold[25]
CanolaDownregulation of lycopene ε-cyclase (ε-CYC)Increased levels of β-carotene, zeaxanthin, violaxanthin, and lutein[26]
BrassicacarinataExpression of an 18-carbon ω3 desaturase (CpDesX) gene from Claviceps purpurea and a 20-carbon ω3 desaturase (Pir-ω3) gene from Pythium irregulareUp to 25% increase in eicosapentaenoic acid[31]
LinseedExpression of a Δ6-desaturase from Primula vialiiTransgenic lines that accumulate the omega-3 fatty acid stearidonic acid[30]
CisgenesisBarleyExpression of a barley phytase gene (HvPAPhy_a)Decrease in phytate concentration, which then increases phosphate bioavailability[202]
PotatoSuppression of a starch phosphorylase L. gene through dsRNAi technologyDecrease in starch degradation what reduces the accumulation of reducing (glucose, fructose) and non-reducing (sucrose) sugars in tubers stored at 4 °C[203]
AppleExpression of MdMYB10 transcription factorRed-fleshed ‘Gala’ apples rich in anthocyanins[41]
IntragenesisPotatoSilencing of a granule-bound starch synthase (GBSS) geneAn increase in amylopectin content[204]
Silencing of an asparagine synthase gene (StAs1)Reduced free asparagine concentration by up to 80% and consequent decrease in acrylamide content in processed potato[205]
Overexpression of a lycopene b-cyclase (StLYCb) gene under the GBSS promoterAn increase in β-carotene accumulation in potato tubers[27]
TomatoSuppression of a DE-ETIOLATED1 (DET1) gene through RNAi technologyEnhanced carotenoid and flavonoid content[28]
WheatSuppression of a γ-gliadin gene by using RNAi technologyGluten-free wheat[206]
Overexpression of a vacuolar Iron transporter (TaVIT2) under the control of a wheat endosperm-specific promoterAn increase in more than 2-fold of iron in white flour fractions[12]
SoybeanRNAi technologyPlenish® high oleicDupont-Pioneer (Johnston, IA, USA)
Vistive® Gold low saturated high oleicMonsanto (St. Louis, MO, USA)
Table 4. Comparison of the main characteristics of conventional breeding, transgenesis, cisgenesis, and intragenesis.
Table 4. Comparison of the main characteristics of conventional breeding, transgenesis, cisgenesis, and intragenesis.
CharacteristicConventional Breeding TransgenesisCisgenesisIntragenesis
Variability sourceSexually compatible groupAny organismSexually compatible groupSexually compatible group
MethodCrosses and selectionRecombinant DNABy AgrobacteriumBy Agrobacterium (recombinant DNA)
Introducing DNANaturalNatural and/or artificialNaturalNatural and/or artificial
Gene poolUnalteredAlteredUnalteredAltered
Borders-T-DNAT-DNA (to be eliminated)T-DNA or P-DNA
Linkage dragYesNoNoNo
Expression modulationNoYesYesYes
TimeHighMediumMediumMedium
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Medina-Lozano, I.; Díaz, A. Applications of Genomic Tools in Plant Breeding: Crop Biofortification. Int. J. Mol. Sci. 2022, 23, 3086. https://doi.org/10.3390/ijms23063086

AMA Style

Medina-Lozano I, Díaz A. Applications of Genomic Tools in Plant Breeding: Crop Biofortification. International Journal of Molecular Sciences. 2022; 23(6):3086. https://doi.org/10.3390/ijms23063086

Chicago/Turabian Style

Medina-Lozano, Inés, and Aurora Díaz. 2022. "Applications of Genomic Tools in Plant Breeding: Crop Biofortification" International Journal of Molecular Sciences 23, no. 6: 3086. https://doi.org/10.3390/ijms23063086

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