Next Article in Journal
A Weed Population Dynamics Model for Integrated Weed-Management Decision-Making Support: Euphorbia davidii Subils in Soybean Crops as a Simulation Study
Previous Article in Journal
An Improved Lightweight Network for Real-Time Detection of Apple Leaf Diseases in Natural Scenes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Maize Breeding: From Domestication to Genomic Tools

1
Department of Plant Breeding, Faculty of Agriculture, University of Agricultural Sciences and Veterinary Medicine Cluj-Napoca, Calea Mănăştur 3-5, 400372 Cluj-Napoca, Romania
2
Department of Genetics, Faculty of Agriculture, University of Agricultural Sciences and Veterinary Medicine Cluj-Napoca, Calea Mănăştur 3-5, 400372 Cluj-Napoca, Romania
3
Agricultural Research Development Station Turda, 27 Agriculturii Street, 401100 Turda, Romania
4
Department of Crop Plants, Faculty of Agriculture, University of Agricultural Sciences and Veterinary Medicine Cluj-Napoca, Calea Mănăştur 3-5, 400372 Cluj-Napoca, Romania
*
Authors to whom correspondence should be addressed.
Agronomy 2022, 12(10), 2365; https://doi.org/10.3390/agronomy12102365
Submission received: 11 September 2022 / Revised: 26 September 2022 / Accepted: 28 September 2022 / Published: 30 September 2022

Abstract

:
Maize will continue to expand and diversify as an industrial resource and a feed and fuel crop in the near future. The United Nations estimate that in 2050 the global population will reach 9.7 billion people. In this context, food security is increasingly being discussed. Additionally, another threat to food security is global warming. It is predicted that both the quantity and the quality of crops will be seriously affected by climate change in the near future. Scientists and breeders need to speed up the process of creating new maize cultivars that are resistant to climate stress without diminishing yield or quality. The present paper provides a brief overview of some of the most important genomics tools that can be used to develop high-performance and well-adapted hybrids of maize and also emphasizes the contribution of bioinformatics to an advanced maize breeding. Genomics tools are essential for a precise, fast, and efficient breeding of crops especially in the context of climate challenges. Maize breeders are able now to develop new cultivars with better traits more easily as a result of the new genomic approaches, either by aiding the selection process or by increasing the available diversity through precision breeding procedures. Furthermore, the use of genomic tools may in the future represent a way to accelerate the processes of de novo domestication of the species.

1. Introduction

Food crops are vital energy sources for the human body. Additionally, they are playing an extremely important role in the economy. The United Nations estimate that in 2050 the global population will reach 9.7 billion people [1]. In this context, food security is increasingly being discussed. In addition to the growing population of the globe, another threat to food security is global warming. It is predicted that both the quantity and the quality of crops will be seriously affected by climate change in the near future [2,3,4]. Under these conditions, scientists and breeders need to speed up the process of creating new varieties that are resistant to climate stress without diminishing yield or quality.
Maize or corn (Zea mays L.), a plant species particularly generous in its production potential and in its wide diversity of uses, is fundamental for the development of a modern and efficient agriculture. Globally, maize ranks third in area and first in production, ahead of wheat and rice [5]. Due to the large areas corn occupies, but especially to its yielding performance, maize is a major source of food for the world’s population. In addition to its nutritional importance for humans, corn is also basic food for animal feed and valuable raw material for industrial processing. The importance of corn for human health results from the fact that it is a food rich in nutrients [6,7,8], especially energy of 355 kcal per 100 g flour with 15% moisture, compared to 352 kcal for wheat flour, 348 kcal for rye flour, and 346 kcal for hulled barley [9]. As food, maize also has some shortcomings, of which we note the low amount of some essential amino acids, such as lysine and tryptophan [10,11]. Maize is also a valuable raw material for industry, extracting oil, starch, alcohol, glucose, and other products such as syrup, pectin, dextrin, plastics, lactic acid, acetic acid, acetone, dyes, and synthetic rubber from its grains. Paper, cardboard, and nitrocellulose can be made from corn stalks [9]. Every part of the maize plant has economic worth, including the grain, leaves, stalk, tassel, and cob, which can be used to make a variety of food and non-food goods [11]. As a result, maize will become increasingly important in grain production around the world. Plant breeding research has traditionally concentrated on enhancing productive potential in newly created maize varieties because this criterion ensures a crop’s economic efficiency [8].
Plant breeding is the art and the science of changing the traits of plants in order to produce desired characteristics in new improved varieties [12]. For most crop plants, classical breeding strategies for selecting superior genotypes depending on phenotypic traits have proven to be of narrow success due low heritability, polygenic effects, gene–gene interactions (epistasis), environmental–genotype interactions, and a long period of time to achieve the goals [12]. Maize traditional breeding uses the same strategies in improvement programs [13]. Yield and several other important traits are controlled by many genes with small effects whose value are conditioned by the environment [14,15]. Around the 2000s, plant breeding tools assured, at a global level, a constant yield gain rate/year of 1.6% for maize [14]. Projections for the year 2050 indicate that we need to increase the rate of yield gain of 1.67 to 3.06% per year due to the rapid growth of earth population [5]. This aspect also suggests the importance of developing new breeding methods to meet these goals [16,17]. In order to achieve this goal in the shortest time, it is necessary to use advanced breeding techniques based on plant genomics that help to identify the phenotypic and genotypic changes in plants faster than conventional methods [4,18,19]. Through plant genomics it is possible to examine the genetic material at the molecular level [20], which leads to a lot of new information. To be able to organize huge amounts of data, advanced methods and powerful tools are needed [4]. Bioinformatics provides us with these tools which scientists and breeders use to get closer to the genomic information.
Given the global importance of maize and the market demand for new maize crops with maximum yield and quality performance, mainly because of the rapid growth of the population as well as the accelerated change of climatic factors, the research and knowledge of modern techniques and tools for maize improvement is essential. Breeders now have advanced techniques, resources, and technologies to create superior maize cultivars that can withstand the vagaries of climate and insect pest attacks, thanks to new innovations and trends in genomics, bioinformatics, and phenomics. Therefore, this review paper provides a brief overview of some of the most important genomics tools used to obtain high-performance and well-adapted hybrids of maize and also emphasizes the contribution of bioinformatics to advanced maize breeding.

2. Evolution of Maize: From Domestication to Precise Breeding

Over time, there have been many hypotheses about the exact origin of maize and its domestication process [18]. Considered to be originating from the high mountains of Mexico for a long time, the newest molecular analyses suggest that the origin of species is the Central Balsas River Valley from southwestern Mexico. It also seems that maize domestication was not a long time process with multiple steps but rather a very rapid one [21] from its wild ancestor, a subspecies of teosinte known as Zea mays ssp. parviglumis [22,23]. Following the archeological researches carried out in Balsas River Valley, the archaeologists Piperno and Ranere highlight the fact that maize was present in the area about 8700 years ago, during the early Holocene [22,23,24], when the first agriculture practices were organized in small groups which were moving their shelters seasonally and carrying out a subsistence activity. From Mexico, maize spread over the Americas. Christopher Columbus brought the first seeds of maize to Europe, and in 1525 the first fields were planted in Spain [25].
The most widespread method of genetic improvement since the domestication of plants was mass selection, where the best ear of the healthiest plant is selected to be sown in the next year [26]. Even if the method perfected over the years from a simple mass selection to an ear-to-row selection, it was not very efficient for traits with low heritability like yield [26]. The selection method involves the choice of individuals for breeding based on how they differ in the desired characteristics, which are usually observable or measurable traits. In one of the world’s longest studies, the breeders from the Illinois Agricultural Experiment Station began selecting maize kernels for high oil content as early as 1896. Maize breeders have regularly selected to change the oil composition for nearly a century. The shift in oil concentration after 100 generations was significant: from a base of about 5%, the high oil-producing line currently has about 20% oil [25]. The evolution of maize breeding over time is synthetized in Figure 1.
The first attempts to direct pollen from one plant to another began in 1867, when H.S. Bidwell introduced maize detasseling as a pollination control measure in his breeding program. In 1878–1881, the professor W.J. Beal from Michigan State Agricultural College proposed crossing maize varieties to grow the production of commercial cultivars. W.A. Kellerman and W.T. Swingle conducted the first segregation counts on maize ears in 1891, for the starchy gene. In 1893, during the World’s Colombian Exposition in Chicago, R. and J. Reid’s “Yellow Dent Maize” won the golden prize as “the world’s most beautiful corn.” Reid’s maize became a very important parent of modern hybrid maize [25]. A revolutionary step in maize breeding began with the discovery of ‘hybrid vigor’ or heterosis in 1909, simultaneously by E.M. East at Harvard University in collaboration with the Connecticut Agricultural Experiment Station [27] and G.H. Schull at the Carnegie Institution Station at Cold Spring Harbor [27,28]. This was the first step toward the maize scientific breeding. It took until 1933 for the first hybrid maize created by crossing inbred lines to be introduced.
The modern breeding of corn started in the 1970s–1980s with the first regenerated plant from a cell culture and sequencing the first maize genes. In 1975, in vitro selection of maize found callus cultures resistant to fungal plant pathogen Helminthosporium maydis T-toxin [29]. Molecular markers were first used in 1980s–1990s in maize breeding and the first transgenic maize cultivar, Bacillus thuringiensis (Bt) hybrid, was marketed in 1996 [30,31]. Later, more transgenic hybrids with insect resistance, herbicide tolerance, high lysine content, and drought-resistance were commercialized [31]. Predictive breeding started with the decreasing costs of genome sequencing in the late 1990s to early 2000s [28], today being known as genomic prediction or genomic selection [32,33]. Genome editing by CRISPR technology represents the beginning of the precision plant breeding [34]. The first CRISPR-edited crop that was submitted to the US regulatory system in 2016 was Pioneer high amylopectin maize from DuPont [35].
New plant breeding techniques based on gene editing, cisgenesis, and epigenetic approaches [36] will hold an essential role in the developing of new plants cultivars, allowing for more precise, predictable and quick breeding works [37]. All of these new plant development strategies, however, would not be conceivable without the help of bioinformatics hardware and software.

3. Bioinformatics: The Fundamental Tool Underlying Modern Methods of Breeding

Knowledge that helps us to deeply understand the biological data through methods and software tools is bioinformatics (BI). A link between biology and computer science [38], BI plays an essential role in modern methods of plant breeding. The baggage of tools that bioinformatics provides facilitates the understanding of the biological and genetic processes that underlie the selection process. As Pauline Hogeweg, one of the promoters of BI states, the term was used for the first time by her and Ben Hesper, in the beginning of 1970s and defining “the study of informatics processes in biotic systems” [39,40]. Over the past years, advanced research in the field of biology, genetics, but also artificial intelligence, has led to the reconfiguration and development of the meaning of BI [38,40,41]. At present, evolutionary biology goes ‘hand-in-hand’ with bioinformatics, due to the fact that the first provides the scientific data and the second manages and analyzes them [42].
Major goals of BI can be stated in the analysis, reconstruction, discovery, assembly, comparison, classification, simulation, modeling, and prediction using biological data [38]. In order to be able to organize the multitude of existing data and to be able to process them, the data are collected and organized systematically in databases. The process of BI implication in plant breeding is shown in Figure 2. Most BI databases contain nucleic acid and protein sequences. Those are a key factor in understanding and developing the applied sciences [43], such as plant breeding, especially in the cases of the species that are rich in phenotypic and genetic diversity, such as maize. The most relevant BI databases for Zea mays species, which contain in addition to the interest sequences also the necessary genomic tools for maize breeding, are summarized in Table 1.
Having available the entire genome of a species opens the way to advanced research, because it can be used to create better crops, with a higher quality or more resistant to biotic and abiotic stress. Knowing exactly how closely related are the plants is the base for choosing the right genitors to produce crops with desirable traits. BI databases help researchers study and understand better functions and expression of genes especially under different stress conditions so that they can identify, analyze, and test the right plants for their purpose. Through its essence, BI is not only a provider of information but also leads to new experiments. By constantly updating databases, scientists will have new data in a timely manner that they can use for further and more advanced research.

4. Marker-Assisted Selection

Plant breeding for crop improvement that uses direct phenotypic selection is labor-intensive, expensive, and time-consuming. Target gene expression, their unique biological or environmental state, and the trait’s heredity all influence this selection. For quantitative traits that are regularly under selection process, phenotypic selection is less effective [72,73,74].
Marker-assisted selection (MAS), also known as marker-aided selection, is an indirect selection method in which a trait of interest is chosen based on a marker (morphological, biochemical, or DNA/RNA variation) [25]. In MAS, a marker is close to a gene that controls the trait, so the presence of the marker indicates the presence of a desired allele [75]. Molecular marker technology, automated equipment, and new statistical tools to analyze data are all aiding the breeders to be ready to face the growing food demand. MAS is now widely applied and proven to be an effective breeding technique [76]. Genotypic markers are used to help phenotypic selections. MAS is cheaper to use in crops for traits with simple inheritance [75] but also can be applied to quantitative traits controlled by many genes with small effects, such as yield, stress tolerance, and resistance to pests [75,77], even if using the method in this case is more complicated [78]. It is primarily utilized by plant breeders in their breeding programs to find desired dominant or recessive alleles throughout generations as well as to help identify the best genotypes from segregating generations [79]. Reliable markers, a high-quality DNA extraction procedure, genetic maps, knowledge of marker-trait associations, quick and efficient data processing, and the availability of high-throughput marker detection equipment are all prerequisites for an effective MAS program [72].
Marker-assisted backcross (MABC) is a particular case of MAS where backcrossing strategy is assisted by molecular markers to speed up recurrent parent selection and genome recovery [80]. Conventional backcrossing is a method that has been widely used for transferring oligogenic traits from donor parents to recurrent parents by recovering the whole genome of recurrent parents except for the trait of interest after 6–7 generations of backcrossing [81]. By transferring genes of interest or quantitative trait loci (QTLs) from donor parent, the MABC technique has been widely utilized to remove undesired features such as insect and disease susceptibility, anti-nutritional factors, and other undesirable traits from high yielding popular varieties [77]. The close connection of markers with genes or QTLs is the foundation of MABC [72].
Maize was one of the first important crop species for which a comprehensive molecular marker map was developed [82]. First molecular markers used were restriction fragment length polymorphism (RFLPs) [77]. Following that, random amplified polymorphic DNA markers (RAPDs) evolved as a result of the development of polymerase chain reaction (PCR) technology [83]. Because of their limitations, simple sequence repeats (SSRs) or microsatellite markers began to be increasingly used [84]. The amplified fragment length polymorphism (AFLP) uses PCR to preferentially amplify DNA fragments [84]. Single-nucleotide polymorphisms (SNPs) identify single base pair changes and are used more recently [85]. SSRs and SNPs appears to be most common markers used to identify different levels of genetic diversity [86,87]. The most significant molecular markers for maize breeding, widely used in the development of many important modern maize lines, are summarized in Table 2.

5. Genomic Selection

The new technology with perhaps the greatest potential to improve breeding efficiency for quantitative traits is genomic selection (GS). GS was first successfully used in animals [99], and then it also spread in the case of crops [100]. GS requires genotyping thousands of individuals with thousands of markers; it became more popular in breeding programs due to advancement in genotyping technology. By GS, breeders obtain the complete genetic profile of an individual, and the data are used to identify individuals that have superior breeding value [101]. Marker and phenotypic data are co-analyzed to develop a model that predicts which genomes produce the best breeding value meaning they produce superior progeny [102]. Those individuals are intermated to produce progeny that are genotyped and the best are selected using GS models, and intermated again to start a new cycle [103,104]. The goal is to assemble the best genome as quickly as possible [100,101]. GS can predict the yield value of individual plants quickly and cheaply, thereby reducing the duration and cost of a breeding cycle [105]. Apart from yield GS has the potential to improve complex traits with low heritability [106,107,108] as well as simple traits with higher heritability [109,110].
The application of GS in plant breeding could be limited by genotyping costs and circumstances where GS can be efficiently applied in a breeding program [111]. Limitations of GS arise from the size and diversity of the training population and the heritability of the traits to be predicted [112]. Statistical difficulties of GS are related to the high amplitude of marker data, where the number of markers is much larger than the number of observations [113].
There is a close relation between GS and phenotypic selection (PS). Both utilize cycles of selection and phenotypic data from multiple parent populations which are selected based on estimation of the value of their whole genome. Depending on the crop, a single cycle of GS can be completed in a few months instead of years with PS, which requires reliable phenotypes to estimate breeding values. Thus, gain-per-year can be greatly increased, making GS a promising technology to meet the pressing need to improve it [114,115].
In maize breeding the advantage of GS over PS consists of reducing costs and saving time. In terms of cost reduction, the breeder can testcross 50% of all available lines, evaluating them in first-stage multi-locational trials and can then use the phenotypic data to predict the remaining 50% by GS. Selecting lines directly for stage II instead of going through stage I (used in PS) significantly reduces the cost of testcross formation and evaluation at each stage of multi-location trials. The GS time efficacy in maize breeding could come from the second cycle of selection, which uses the training population from the previous cycle to predict the new doubled haploid lines, therefore eliminating testcross formation and first-stage multi-location evaluation trials. Based on this, the best lines could go directly to the second stage of multi-location evaluations [111].
For instance, Beyenne et al. (2021) used GS in maize to predict the breeding value for the new created double haploid lines for further testing. Using rAmpSeq markers, his team evaluated as test crosses 3068 inbred lines of tropical maize in well-watered and water-stress environments [116]. GS was also used in predicting the performance of maize hybrids under different genetic designs: Griffing’s methods, North Carolina Design II, partial diallel, and test crossing [33,117,118]. Using GS, inbreds with high combining ability, either general or specific, are revealed [109]. Genomic estimation of breeding value can help to identify inbred lines of maize with favorable disease resistance [119]. GS models can include multiple levels of the biological hierarchy of an agronomic trait based on the available omics data (genomics, transcriptomics, proteomics, epigenomics, and metabolomics) in maize [33,119].

6. Breeding by Design

Approaches made on genomics and breeding technologies facilitate the development of breeding by design (BD) where the allelic variation of the genes with agronomic importance can be controlled [120] so the breeder can design new in silico superior genotypes [121] (Figure 3). Due to the accuracy of this process, the breeder can ignore the selection by phenotyping and evaluate only the superior varieties for field performance [121]. The BD technique involves three stages: mapping loci that are influencing all agronomical relevant traits; assessing the allelic variation at those loci; and breeding by design using desirable alleles of different loci [121,122]. Thus, the BD approaches the possibility to predict the result of a set of crosses based on molecular marker information. This method has been mentioned in the development of higher-quality maize cultivars. The most effective use of the BD method, on the other hand, will come from incorporating the most powerful genomic tools into the breeding process, which will allow the improvement of predictions [123].

7. Genome Sequencing

Genome sequencing (GSq) implies the determination of the order of DNA nucleotides in the genome of an organism [123]. Using this technique, scientists try to localize the genes and understand how a certain gene develops and influences the entire organism. In addition to the genes, GSq can help to a better understanding of the parts of genome outside the genes that control how genes are turned on or off [123].
Due to its complex genetics, maize is mostly used as a model plant for genetics analyses [124,125] and has become one of the first species considered for GSq [126]. It is estimated that maize genome is almost 80% of the human genome size and contains approximately 59,000 genes [127]. First sequencing attempts on maize were started during the early 2000s with The Sequencing the Maize Genome Project (STMG) [128,129] and The Consortium for Maize Genomics (CMG) [129,130]. A full-scale program was launched in 2005 by the National Science Foundation [129], to sequence the cultivar B73, one of the most used maize inbred lines for the production of the high-yielding commercial hybrids [131,132]. Three years later, the first draft sequence for maize genome was published [133]. According to a multi-genome investigation carried out in 2021, new variation in gene content, genome organization, and methylation was discovered in maize [134].
GSq is an excellent technique for a wide range of studies, from single-gene markers to whole-genome profiling [101]. Plant breeders use GSq to genotype breeding populations quickly, allowing them to conduct genome-wide studies, genomic diversity studies, genetic linkage analysis, molecular marker identification, and genomic selection on a large scale [101]. Additionally, GSq is critical for identifying representative single-nucleotide polymorphism that can be used to improve crop genotypes [135]. In the near future, by the development of the next-generation DNA sequencing technology, it will be possible to sequence any maize genotypes [129,136]. After all, breeders’ ultimate goal is to connect phenotype and genetics together [104]. With a better understanding of the link between heritable genetic variables and the phenotypes they produce, the genotype can then be used to predict phenotypes and choose improved cultivars in plant breeding [104]. Genome prediction combined with crops modeling can lead to significant gains in maize breeding [32].

8. New Plant-Breeding Techniques

The start of the precision breeding was represented by the first actions of editing the plant genome [34]. Genome editing (GE) is the newest approach of modifying certain genes to improve the expression of interest traits. This technique can increase the genetic gain of the plants and also accelerate the breeding process at the same time [137]. GE modifications are precise, and the resulting changes are minor [138]. The GE experiments with high chances of success are the ones in which the target traits are determined by a single gene. The needed tools for GE are provided by the Novel Genomics Techniques (NGTs)/New Plant-Breeding Techniques (NBTs) [139]. GE instruments are based on DNA modification with the purpose of changing the traits of the plant [140,141]. Genome editing through NTBs implies:
  • cutting and modifying the genome using one of the following tools: CRISPR/Cas, Zinc finger nuclease or TALENs [142,143,144,145,146,147];
  • introducing a few base pairs by oligonucleotide-directed mutagenesis [148,149,150];
  • deploying processes that alter the genes, but without the alteration of DNA by epigenetic methods [151,152,153];
  • transferring beneficial natural genes from compatible species for crossings (cisgenesis) [154,155,156];
  • transferring recreated genes in vitro from genetic elements of compatible species (intragenesis) [155,157,158];
  • grafting an unmodified plant with a modified one (only one part is transgenic, scion or rootstock) [159,160,161].
In the 1980s, CRISPR (clustered regularly interspaced short palindromic repeats) was identified for the first time in Escherichia coli. It was only in 2007 that the CRISPR/Cas9 system was discovered to be a bacterial defense mechanism against viruses and foreign DNA [162]. E.M. Charpentier and J.A. Doudna published works on the resistance mechanism of bacteria to phages. Their discoveries have led to the development of new technologies for manipulating DNA, which are today known as GE [163]. The procedure is comparable to induced mutations caused by physical or chemical mutagens, but it is more accurate due to the protein called Cas9 that can easily cut desired gene or DNA sequence [163 A CRISPR/Cas9 toolkit was developed by Xing and his team for GE in maize and other plant species [164]. Kouranov and his collaborators used the CRISPR Cas12a system to increase the chromosomal crossovers produced in the somatic cells of the F1 hybrid generation and therefore a way to expand the genetic diversity in maize [165].
Zinc finger nuclease (ZFN) and transcription activator-like effector proteins (TALENs) are other technologies in addition to CRISPR that can be used to alter the genome [166]. ZFNs are made up of zinc finger protein domains that can bind DNA with specificity to certain sequences and a nuclease domain that can cleave DNA [167]. ZFNs can be designed for the majority of genes and interesting areas of most genomes because they have few target sequence limitations. In contrast to the CRISPR/Cas system, ZFNs can attach and cleave with high fidelity at the majority of loci [168]. As shown in various crops including maize, canola, and wheat, high-quality ZFN designs appear to minimize off-target effects [169]. With the ability to alter a specific gene sequences, ZFN technique has become an important tool for creating targeted mutations. For example, low phytated maize lines were created by inducing maize mutations in IPK1 [164,169,170]. As with ZFNs, TALENs are modular proteins, made up of two parts: a TALE that directs the protein to a particular DNA sequence and a nuclease that breaks DNA; the Fok1 nuclease is frequently used [171]. TALEN has been applied like an efficient genome editing tool to produce beneficial characteristics in crops; over 50 genes were targeted for mutation in plants using this technology, including in maize (genes GL2, IPK1A, IPK, and MRP4) [172].
Oligonucleotide-directed mutagenesis (ODM) provides a quick, accurate, non-transgenic alternative for breeding traits in plants [173]. Using chemically synthetized oligonucleotides, this technology uses the cell’s natural DNA repair machinery to modify a set of targeted bases inside the genome. Random mutations are avoided due to the precise control. After the successful application of ODM in tobacco plants, maize followed closely [173,174].
Processes that modify gene activity without affecting the DNA sequence are referred to as epigenetic and can result in alterations that can be passed on to daughter cells [151,175]. DNA methylation is probably the most well-known epigenetic process [176]. Some of the possible applications of this technology in maize breeding are the development of climate-smart hybrids by improvement adaptation and resilience of plants to environmental stresses [177], creation of new molecular markers associated with plants fertility [178] or to predict key phenotypes [179].
Cisgenesis is based on transferring a natural gene from a certain plant on a new compatible plant, which can be crossed also in a traditional way, being sexually compatible. The most relevant advantages of this method are the elimination of the linkage process [156] and time saving compared to the classic methods of plant improvement [180]. For example, using this technique, maize breeders are able to transfer the cisgene T-urf13 implied in the nucleocytoplasmic male sterility of plants [181].
Another method of genetic modification that not implies the presence of a transgene is intragenesis that allows combination of genes or intergenic sequences from related plants [182,183]. Intragenic transformations are recommended for studying plant adaptation and induced stress response but also to increase nutrient accumulation in maize [182].
A long-standing agricultural practice called grafting now has ramifications for the creation of transgenic plants. Only the scion or rootstock of transgrafted plants is transgenic, and the other component is left unaltered [159]. Numerous vegetable and fruit species have been propagated and produced using this technology [184,185].
Using these types of tools combined with the classical breeding methods could remodel the maize breeding science [186] by speeding the entire process with shortening the generation time [73].

9. Conclusions

Maize will continue to expand and diversify as an industrial resource and a feed and fuel crop in the near future. Its enormous diversity and complexity of genetic study ensure its role further as a model plant for biological studies. New plant breeding techniques used together with classic breeding approaches have the potential to revolutionize maize breeding science by speeding up the entire process and shortening the generation time. Maize breeders are able to obtain new cultivars with improved features more easily as a result of new genomic developments, either by aiding the selection process or by increasing the available diversity through precision breeding procedures. Incorporating the most powerful genomics tools into the breeding process allows the improvement of predictions; therefore, maize breeders are able to develop new maize crops with the highest yield and quality performance in the shortest time possible. Furthermore, the use of genomic tools may in the future represent a way to accelerate the processes of de novo domestication of the species, an idea that is gaining interest among researchers in the field.

Author Contributions

Conceptualization, A.O. and L.M.; methodology, I.B.; software, A.O.; validation, L.M., A.O. and I.R.; formal analysis, I.R.; investigation, I.B. and S.M.; resources, I.R.; data curation, A.O.; writing–original draft preparation, L.M.; writing–review and editing, S.M.; visualization, L.M. and I.B.; supervision, S.M.; project administration, L.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Roser, M. Future Population Growth. Our World in Data, 9 May 2013. [Google Scholar]
  2. Hu, H.; Scheben, A.; Edwards, D. Advances in Integrating Genomics and Bioinformatics in the Plant Breeding Pipeline. Agriculture 2018, 8, 75. [Google Scholar] [CrossRef]
  3. Merca, N.C.; Rusu, T.; Merca, I.; Ona, A.D. Agroecology: A Real Opportunity to Fight against the Climate Challenges. Sci. Pap. Ser. Manag. Econ. Eng. Agric. Rural. Dev. 2021, 21, 6. [Google Scholar]
  4. Shafi, A.; Zahoor, I.; Haq, E.; Fazili, K.M. Impact of Bioinformatics on Plant Science Research and Crop Improvement. In Essentials of Bioinformatics, Volume III: In Silico Life Sciences: Agriculture; Hakeem, K.R., Shaik, N.A., Banaganapalli, B., Elango, R., Eds.; Springer International Publishing: Cham, Switzerland, 2019; pp. 29–46. ISBN 978-3-030-19318-8. [Google Scholar]
  5. FAOSTAT. Available online: https://www.fao.org/faostat/en/#home (accessed on 29 March 2022).
  6. Ghețe, A.B.; Duda, M.M.; Vârban, D.I.; Vârban, R.; Moldovan, C.; Muntean, S. Maize (Zea Mays), a Prospective Medicinal Plant in Romania. Hop Med. Plants 2018, 26, 44–51. [Google Scholar]
  7. Vârban, D.I.; Gheţe, A.; Chirițoiu, I.A.; Odagiu, A. Study of the Behavior of Some Maize Hybrids, in the Pedo- Climatic Conditions from Stănilești Commune Vaslui County. ProEnvironment 2020, 13, 43. [Google Scholar]
  8. Merca, N.C.; Rusu, T.; Merca, I.; Ona, A.D. Maize under the Climate Change Impact. Agricultura 2021, 118, 1–2. [Google Scholar] [CrossRef]
  9. Cristea, M. Importanţa economică, răspândirea geografică, producţia şi comerţul cu porumb. In Monografia Porumbului; Editura Academiei Române: București, Romania, 2004; Volume 1, pp. 17–27. ISBN 973-27-1055-1. [Google Scholar]
  10. Ufaz, S.; Galili, G. Improving the Content of Essential Amino Acids in Crop Plants: Goals and Opportunities. Plant Physiol. 2008, 147, 954–961. [Google Scholar] [CrossRef]
  11. Lone, A.A.; Dar, Z.A.; Gull, A.; Gazal, A.; Naseer, S.; Khan, M.H.; Ahangar, A.; Iqbal, A.M. Breeding Maize for Food and Nutritional Security; IntechOpen: London, UK, 2021; ISBN 978-1-83969-164-5. [Google Scholar]
  12. Poehlman, J.M.; Sleper, D.A. Breeding Field Crops; Springer: Berlin/Heidelberg, Germany, 2013. [Google Scholar]
  13. Cushman, J.C.; Bohnert, H.J. Genomic Approaches to Plant Stress Tolerance. Curr. Opin. Plant Biol. 2000, 3, 117–124. [Google Scholar] [CrossRef]
  14. Hafner, S. Trends in Maize, Rice, and Wheat Yields for 188 Nations over the Past 40 Years: A Prevalence of Linear Growth. Agric. Ecosyst. Environ. 2003, 1–3, 275–283. [Google Scholar] [CrossRef]
  15. Crouch, D.J.M.; Bodmer, W.F. Polygenic Inheritance, GWAS, Polygenic Risk Scores, and the Search for Functional Variants. Proc. Natl. Acad. Sci. USA 2020, 117, 18924–18933. [Google Scholar] [CrossRef]
  16. Tester, M.; Langridge, P. Breeding Technologies to Increase Crop Production in a Changing World. Science 2010, 327, 818–822. [Google Scholar] [CrossRef]
  17. Lloyd, J.R.; Kossmann, J. Improving Crops for a Changing World. Front. Plant Sci. 2021, 12, 728328. [Google Scholar] [CrossRef] [PubMed]
  18. Ahmad, P.; Ashraf, M.; Younis, M.; Hu, X.; Kumar, A.; Akram, N.A.; Al-Qurainy, F. Role of Transgenic Plants in Agriculture and Biopharming. Biotechnol. Adv. 2012, 30, 524–540. [Google Scholar] [CrossRef] [PubMed]
  19. Maghuly, F.; Myśków, B.; Till, B.J. Functional Genomics for Plant Breeding. Int. J. Mol. Sci. 2021, 22, 11854. [Google Scholar] [CrossRef] [PubMed]
  20. Govindaraj, M.; Vetriventhan, M.; Srinivasan, M. Importance of Genetic Diversity Assessment in Crop Plants and Its Recent Advances: An Overview of Its Analytical Perspectives. Genet. Res. Int. 2015, 2015, e431487. [Google Scholar] [CrossRef]
  21. Liu, C.; Li, X.; Meng, D.; Zhong, Y.; Chen, C.; Dong, X.; Xu, X.; Chen, B.; Li, W.; Li, L.; et al. A 4-Bp Insertion at ZmPLA1 Encoding a Putative Phospholipase a Generates Haploid Induction in Maize. Mol. Plant. 2017, 10, 520–522. [Google Scholar] [CrossRef]
  22. Piperno, D.R.; Ranere, A.J.; Holst, I.; Iriarte, J.; Dickau, R. Starch Grain and Phytolith Evidence for Early Ninth Millennium B.P. Maize from the Central Balsas River Valley, Mexico. Proc. Natl. Acad. Sci. USA 2009, 106, 5019–5024. [Google Scholar] [CrossRef]
  23. Ranere, A.J.; Piperno, D.R.; Holst, I.; Dickau, R.; Iriarte, J. The Cultural and Chronological Context of Early Holocene Maize and Squash Domestication in the Central Balsas River Valley, Mexico. Proc. Natl. Acad. Sci. USA 2009, 106, 5014–5018. [Google Scholar] [CrossRef]
  24. Tenaillon, M.I.; Charcosset, A. A European Perspective on Maize History. Comptes Rendus. Biol. 2011, 334, 221–228. [Google Scholar] [CrossRef]
  25. Schlegel, R.H.J. History of Plant Breeding; CRC Press: Boca Raton, FL, USA, 2017. [Google Scholar]
  26. Lee, E.A.; Tracy, W.F. Modern Maize Breeding. In Handbook of Maize: Genetics and Genomics; Bennetzen, J.L., Hake, S., Eds.; Springer: New York, NY, USA, 2009; pp. 141–160. ISBN 978-0-387-77863-1. [Google Scholar]
  27. Byerlee, D. The Globalization of Hybrid Maize, 1921–1970. J. Glob. Hist. 2020, 15, 101–122. [Google Scholar] [CrossRef]
  28. Varshney, R.K.; Tuberosa, R. Genomics-Assisted Crop Improvement: An Overview. In Genomics-Assisted Crop Improvement: Vol. 1: Genomics Approaches and Platforms; Varshney, R.K., Tuberosa, R., Eds.; Springer Netherlands: Dordrecht, The Netherlands, 2007; pp. 1–12. ISBN 978-1-4020-6295-7. [Google Scholar]
  29. Gengenbach, B.G.; Green, C.E. Selection of T-Cytoplasm Maize Callus Cultures Resistant to Helminthosporium Maydis Race T Pathotoxin1. Crop Sci. 1975, 15, 645–649. [Google Scholar] [CrossRef]
  30. Troyer, A.F. Development of Hybrid Corn and the Seed Corn Industry. In Handbook of Maize: Genetics and Genomics; Bennetzen, J.L., Hake, S., Eds.; Springer: New York, NY, USA, 2009; pp. 87–114. ISBN 978-0-387-77863-1. [Google Scholar]
  31. Lundmark, C. Genetically Modified Maize. BioScience 2007, 57, 996. [Google Scholar] [CrossRef]
  32. Washburn, J.D.; Burch, M.B.; Franco, J.A.V. Predictive Breeding for Maize: Making Use of Molecular Phenotypes, Machine Learning, and Physiological Crop Models. Crop Sci. 2020, 60, 622–638. [Google Scholar] [CrossRef]
  33. Fritsche-Neto, R.; Galli, G.; Borges, K.L.R.; Costa-Neto, G.; Alves, F.C.; Sabadin, F.; Lyra, D.H.; Morais, P.P.P.; Braatz de Andrade, L.R.; Granato, I.; et al. Optimizing Genomic-Enabled Prediction in Small-Scale Maize Hybrid Breeding Programs: A Roadmap Review. Front. Plant Sci. 2021, 12, 658267. [Google Scholar] [CrossRef] [PubMed]
  34. Kim, J.-I.; Kim, J.-Y. New Era of Precision Plant Breeding Using Genome Editing. Plant Biotechnol. Rep. 2019, 13, 419–421. [Google Scholar] [CrossRef]
  35. Waltz, E. CRISPR-Edited Crops Free to Enter Market, Skip Regulation. Nat. Biotechnol. 2016, 34, 582–583. [Google Scholar] [CrossRef] [PubMed]
  36. Salgotra, R.K.; Stewart, C.N. Functional Markers for Precision Plant Breeding. Int. J. Mol. Sci. 2020, 21, 4792. [Google Scholar] [CrossRef]
  37. Veillet, F.; Durand, M.; Kroj, T.; Cesari, S.; Gallois, J.-L. Precision Breeding Made Real with CRISPR: Illustration through Genetic Resistance to Pathogens. Plant Commun. 2020, 1, 100102. [Google Scholar] [CrossRef]
  38. Kushwaha, U.K.S.; Deo, I.; Jaiswal, J.P.; Prasad, B. Role of Bioinformatics in Crop Improvement. Glob. J. Sci. Front Res. D Agric. Vet. 2017, 17, 13–24. [Google Scholar]
  39. Hesper, B.; Hogeweg, P. Bio-Informatics: A Working Concept. A Translation of “Bio-Informatica: Een Werkconcept” by B. Hesper and P. Hogeweg. arXiv 2021, arXiv:2111.11832. [Google Scholar]
  40. Hogeweg, P. The Roots of Bioinformatics in Theoretical Biology. PLoS Comput. Biol. 2011, 7, e1002021. [Google Scholar] [CrossRef]
  41. Attwood, T.K.; Pettifer, S.R.; Thorne, D. Bioinformatics Challenges at the Interface of Biology and Computer Science: Mind the Gap; John Wiley & Sons: Hoboken, NJ, USA, 2016; ISBN 978-1-119-24344-1. [Google Scholar]
  42. Choudhuri, S. Bioinformatics for Beginners: Genes, Genomes, Molecular Evolution, Databases and Analytical Tools; Elsevier Science: Amsterdam, The Netherlands, 2014; ISBN 978-0-12-410471-6. [Google Scholar]
  43. Jimenez-Gutierrez, L.R.; Barrios-Hernández, C.J.; Pedraza-Ferreira, G.R.; Vera-Cala, L.; Martinez-Perez, F. Importance of Databases of Nucleic Acids for Bioinformatic Analysis Focused to Genomics. J. Phys. Conf. Ser. 2016, 743, 012009. [Google Scholar] [CrossRef]
  44. Lawrence, C.J.; Dong, Q.; Polacco, M.L.; Seigfried, T.E.; Brendel, V. MaizeGDB, the Community Database for Maize Genetics and Genomics. Nucleic Acids Res. 2004, 32, D393–D397. [Google Scholar] [CrossRef] [PubMed]
  45. Lawrence, C.J.; Seigfried, T.E.; Brendel, V. The Maize Genetics and Genomics Database. The Community Resource for Access to Diverse Maize Data. Plant Physiol. 2005, 138, 55–58. [Google Scholar] [CrossRef] [PubMed]
  46. Harper, L.; Gardiner, J.; Andorf, C.; Lawrence, C.J. MaizeGDB: The Maize Genetics and Genomics Database. Methods Mol. Biol. 2016, 1374, 187–202. [Google Scholar] [CrossRef]
  47. Andorf, C.M.; Cannon, E.K.; Portwood, J.L.; Gardiner, J.M.; Harper, L.C.; Schaeffer, M.L.; Braun, B.L.; Campbell, D.A.; Vinnakota, A.G.; Sribalusu, V.V.; et al. MaizeGDB Update: New Tools, Data and Interface for the Maize Model Organism Database. Nucleic Acids Res. 2016, 44, D1195–D1201. [Google Scholar] [CrossRef] [Green Version]
  48. Dong, Q.; Roy, L.; Freeling, M.; Walbot, V.; Brendel, V. ZmDB, an Integrated Database for Maize Genome Research. Nucleic Acids Res. 2003, 31, 244–247. [Google Scholar] [CrossRef]
  49. Gai, X.; Lal, S.; Xing, L.; Brendel, V.; Walbot, V. Gene Discovery Using the Maize Genome Database ZmDB. Nucleic Acids Res. 2000, 28, 94–96. [Google Scholar] [CrossRef]
  50. Cho, K.T.; Portwood, J.L.; Gardiner, J.M.; Harper, L.C.; Lawrence-Dill, C.J.; Friedberg, I.; Andorf, C.M. MaizeDIG: Maize Database of Images and Genomes. Front. Plant Sci. 2019, 10, 1050. [Google Scholar] [CrossRef]
  51. Garnett, T.; Lawrence-Dill, C.J.; Pridmore, T.; Watt, M.; Pommier, C.; Pieruschka, R.; Ghamkhar, K. Phenotyping: From Plant, to Data, to Impact and Highlights of the International Plant Phenotyping Symposium—IPPS 2018; Frontiers Media SA: Lausanne, Switzerland, 2021; ISBN 978-2-88966-397-2. [Google Scholar]
  52. Zhou, W.; Wang, L.; Zheng, W.; Yao, W. MaizeSNPDB: A Comprehensive Database for Efficient Retrieve and Analysis of SNPs among 1210 Maize Lines. Comput. Struct. Biotechnol. J. 2019, 17, 1377–1383. [Google Scholar] [CrossRef]
  53. Gui, S.; Yang, L.; Li, J.; Luo, J.; Xu, X.; Yuan, J.; Chen, L.; Li, W.; Yang, X.; Wu, S.; et al. ZEAMAP, a Comprehensive Database Adapted to the Maize Multi-Omics Era. iScience 2020, 23, 101241. [Google Scholar] [CrossRef]
  54. Peng, Z.; Li, H.; Sun, G.; Dai, P.; Geng, X.; Wang, X.; Zhang, X.; Wang, Z.; Jia, Y.; Pan, Z.; et al. CottonGVD: A Comprehensive Genomic Variation Database for Cultivated Cottons. Front. Plant Sci. 2021, 12, 2896. [Google Scholar] [CrossRef] [PubMed]
  55. Xiao, Y.; Jiang, S.; Cheng, Q.; Wang, X.; Yan, J.; Zhang, R.; Qiao, F.; Ma, C.; Luo, J.; Li, W.; et al. The Genetic Mechanism of Heterosis Utilization in Maize Improvement. Genome Biol. 2021, 22, 148. [Google Scholar] [CrossRef] [PubMed]
  56. Zhao, W.; Canaran, P.; Jurkuta, R.; Fulton, T.; Glaubitz, J.; Buckler, E.; Doebley, J.; Gaut, B.; Goodman, M.; Holland, J.; et al. Panzea: A Database and Resource for Molecular and Functional Diversity in the Maize Genome. Nucleic Acids Res. 2006, 34, D752–D757. [Google Scholar] [CrossRef] [PubMed]
  57. Du, C.; Buckler, E.; Muse, S. Development of a Maize Molecular Evolutionary Genomic Database. Comp. Funct. Genom. 2003, 4, 246–249. [Google Scholar] [CrossRef]
  58. Canaran, P.; Buckler, E.S.; Glaubitz, J.C.; Stein, L.; Sun, Q.; Zhao, W.; Ware, D. Panzea: An Update on New Content and Features. Nucleic Acids Res. 2008, 36, D1041–D1043. [Google Scholar] [CrossRef]
  59. Gupta, P.; Naithani, S.; Tello-Ruiz, M.K.; Chougule, K.; D’Eustachio, P.; Fabregat, A.; Jiao, Y.; Keays, M.; Lee, Y.K.; Kumari, S.; et al. Gramene Database: Navigating Plant Comparative Genomics Resources. Curr. Plant Biol. 2016, 7–8, 10–15. [Google Scholar] [CrossRef]
  60. Jaiswal, P. Gramene Database: A Hub for Comparative Plant Genomics. Methods Mol. Biol. 2011, 678, 247–275. [Google Scholar] [CrossRef]
  61. Monaco, M.K.; Stein, J.; Naithani, S.; Wei, S.; Dharmawardhana, P.; Kumari, S.; Amarasinghe, V.; Youens-Clark, K.; Thomason, J.; Preece, J.; et al. Gramene 2013: Comparative Plant Genomics Resources. Nucleic Acids Res. 2014, 42, D1193–D1199. [Google Scholar] [CrossRef]
  62. Tello-Ruiz, M.K.; Naithani, S.; Gupta, P.; Olson, A.; Wei, S.; Preece, J.; Jiao, Y.; Wang, B.; Chougule, K.; Garg, P.; et al. Gramene 2021: Harnessing the Power of Comparative Genomics and Pathways for Plant Research. Nucleic Acids Res. 2021, 49, D1452–D1463. [Google Scholar] [CrossRef]
  63. Chen, F.; Dong, W.; Zhang, J.; Guo, X.; Chen, J.; Wang, Z.; Lin, Z.; Tang, H.; Zhang, L. The Sequenced Angiosperm Genomes and Genome Databases. Front. Plant Sci. 2018, 9, 418. [Google Scholar] [CrossRef]
  64. The UniProt Consortium. UniProt: The Universal Protein Knowledgebase in 2021. Nucleic Acids Res. 2021, 49, D480–D489. [Google Scholar] [CrossRef]
  65. Bairoch, A.; Apweiler, R.; Wu, C.H.; Barker, W.C.; Boeckmann, B.; Ferro, S.; Gasteiger, E.; Huang, H.; Lopez, R.; Magrane, M.; et al. The Universal Protein Resource (UniProt). Nucleic Acids Res. 2005, 33, D154–D159. [Google Scholar] [CrossRef] [PubMed]
  66. Benson, D.A.; Karsch-Mizrachi, I.; Lipman, D.J.; Ostell, J.; Wheeler, D.L. GenBank. Nucleic Acids Res. 2005, 33, D34–D38. [Google Scholar] [CrossRef] [PubMed]
  67. O’Sullivan, C.; Busby, B.; Mizrachi, I.K. Managing Sequence Data. In Bioinformatics; Keith, J.M., Ed.; Methods in Molecular Biology; Springer New York: New York, NY, USA, 2017; Volume 1525, pp. 79–106. ISBN 978-1-4939-6620-2. [Google Scholar]
  68. Leray, M.; Knowlton, N.; Ho, S.-L.; Nguyen, B.N.; Machida, R.J. GenBank Is a Reliable Resource for 21st Century Biodiversity Research. Proc. Natl. Acad. Sci. USA 2019, 116, 22651–22656. [Google Scholar] [CrossRef] [PubMed]
  69. Baker, W.; van den Broek, A.; Camon, E.; Hingamp, P.; Sterk, P.; Stoesser, G.; Tuli, M.A. The EMBL Nucleotide Sequence Database. Nucleic Acids Res. 2000, 28, 19–23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Goodstein, D.M.; Shu, S.; Howson, R.; Neupane, R.; Hayes, R.D.; Fazo, J.; Mitros, T.; Dirks, W.; Hellsten, U.; Putnam, N.; et al. Phytozome: A Comparative Platform for Green Plant Genomics. Nucleic Acids Res. 2012, 40, D1178–D1186. [Google Scholar] [CrossRef]
  71. Bolser, D.; Staines, D.M.; Pritchard, E.; Kersey, P. Ensembl Plants: Integrating Tools for Visualizing, Mining, and Analyzing Plant Genomics Data. In Plant Bioinformatics: Methods and Protocols; Edwards, D., Ed.; Methods in Molecular Biology; Springer: New York, NY, USA, 2016; pp. 115–140. ISBN 978-1-4939-3167-5. [Google Scholar]
  72. Kumawat, G.; Kumawat, C.K.; Chandra, K.; Pandey, S.; Chand, S.; Mishra, U.N.; Lenka, D.; Sharma, R. Insights into Marker Assisted Selection and Its Applications in Plant Breeding; IntechOpen: London, UK, 2020; ISBN 978-1-83968-310-7. [Google Scholar]
  73. Watson, A.; Ghosh, S.; Williams, M.J.; Cuddy, W.S.; Simmonds, J.; Rey, M.-D.; Asyraf Md Hatta, M.; Hinchliffe, A.; Steed, A.; Reynolds, D.; et al. Speed Breeding Is a Powerful Tool to Accelerate Crop Research and Breeding. Nat. Plants 2018, 4, 23–29. [Google Scholar] [CrossRef]
  74. Watson, A.; Hickey, L.T.; Christopher, J.; Rutkoski, J.; Poland, J.; Hayes, B.J. Multivariate Genomic Selection and Potential of Rapid Indirect Selection with Speed Breeding in Spring Wheat. Crop Sci. 2019, 59, 1945–1959. [Google Scholar] [CrossRef]
  75. Stevens, R. Prospects for Using Marker-Assisted Breeding to Improve Maize Production in Africa. J. Sci. Food Agric. 2008, 88, 745–755. [Google Scholar] [CrossRef]
  76. Ramesh, P.; Mallikarjuna, G.; Sameena, S.; Kumar, A.; Gurulakshmi, K.; Reddy, B.V.; Reddy, P.C.O.; Sekhar, A.C. Advancements in Molecular Marker Technologies and Their Applications in Diversity Studies. J. Biosci. 2020, 45, 123. [Google Scholar] [CrossRef]
  77. Ribaut, J.; Hoisington, D. Marker-Assisted Selection: New Tools and Strategies. Trends Plant Sci. 1998, 3, 236–239. [Google Scholar] [CrossRef]
  78. Ashraf, M.; Akram, N.A.; Rahman, M.U.; Foolad, M.R. Marker-Assisted Selection in Plant Breeding for Salinity Tolerance. Methods Mol. Biol. 2012, 913, 305–333. [Google Scholar] [CrossRef] [PubMed]
  79. Francia, E.; Tacconi, G.; Crosatti, C.; Barabaschi, D.; Bulgarelli, D.; Dall’Aglio, E.; Valè, G. Marker Assisted Selection in Crop Plants. Plant Cell Tissue Organ Cult. 2005, 82, 317–342. [Google Scholar] [CrossRef]
  80. Hasan, M.M.; Rafii, M.Y.; Ismail, M.R.; Mahmood, M.; Rahim, H.A.; Alam, M.A.; Ashkani, S.; Malek, M.A.; Latif, M.A. Marker-Assisted Backcrossing: A Useful Method for Rice Improvement. Biotechnol. Biotechnol. Equip. 2015, 29, 237–254. [Google Scholar] [CrossRef] [PubMed]
  81. Miah, G.; Rafii, M.Y.; Ismail, M.R.; Puteh, A.B.; Rahim, H.A.; Latif, M.A. Recurrent Parent Genome Recovery Analysis in a Marker-Assisted Backcrossing Program of Rice (Oryza Sativa L.). Comptes Rendus. Biol. 2015, 338, 83–94. [Google Scholar] [CrossRef]
  82. Helentjaris, T.; Slocum, M.; Wright, S.; Schaefer, A.; Nienhuis, J. Construction of Genetic Linkage Maps in Maize and Tomato Using Restriction Fragment Length Polymorphisms. Theoret. Appl. Genet. 1986, 72, 761–769. [Google Scholar] [CrossRef]
  83. Ristić, D.; Babić, V.; Anđelković, V.; Vančetović, J.; Mladenović-Drinić, S.; Ignjatović-Micić, D. Genetic Diversity in Maize Dent Landraces Assessed by Morphological and Molecular Markers. Genetika 2013, 45, 811–824. [Google Scholar] [CrossRef]
  84. Zhang, F.; Wan, X.-Q.; Pan, G.-T. QTL Mapping of Fusarium Moniliforme Ear Rot Resistance in Maize. 1. Map Construction with Microsatellite and AFLP Markers. J. Appl. Genet. 2006, 47, 9–15. [Google Scholar] [CrossRef]
  85. Xia, X.C.; Reif, J.C.; Melchinger, A.E.; Frisch, M.; Hoisington, D.A.; Beck, D.; Pixley, K.; Warburton, M.L. Genetic Diversity among CIMMYT Maize Inbred Lines Investigated with SSR Markers: II. Subtropical, Tropical Midaltitude, and Highland Maize Inbred Lines and Their Relationships with Elite U.S. and European Maize. Crop Sci. 2005, 45, 2573–2582. [Google Scholar] [CrossRef]
  86. Owens, B.F.; Mathew, D.; Diepenbrock, C.H.; Tiede, T.; Wu, D.; Mateos-Hernandez, M.; Gore, M.A.; Rocheford, T. Genome-Wide Association Study and Pathway-Level Analysis of Kernel Color in Maize. Genetics 2019, 9, 1945–1955. [Google Scholar] [CrossRef]
  87. Kumar, A.; Longmei, N.; Kumar, P.; Kaushik, P. Molecular Marker Analysis of Genetic Diversity in Maize: A Review. OBM Genet. 2022, 6, 1. [Google Scholar] [CrossRef]
  88. Bentolila, S.; Guitton, C.; Bouvet, N.; Sailland, A.; Nykaza, S.; Freyssinet, G. Identification of an RFLP Marker Tightly Linked to TheHt1 Gene in Maize. Theoret. Appl. Genet. 1991, 82, 393–398. [Google Scholar] [CrossRef]
  89. Chandran, S.; Pukalenthy, B.; Adhimoolam, K.; Manickam, D.; Sampathrajan, V.; Chocklingam, V.; Eswaran, K.; Arunachalam, K.; Joikumarmeetei, L.; Rajasekaran, R.; et al. Marker-Assisted Selection to Pyramid the Opaque-2 (O2) and β-Carotene (CrtRB1) Genes in Maize. Front. Genet. 2019, 10, 859. [Google Scholar] [CrossRef] [PubMed]
  90. Singh, J.; Sharma, S.; Kaur, A.; Vikal, Y.; Cheema, A.K.; Bains, B.K.; Kaur, N.; Gill, G.K.; Malhotra, P.K.; Kumar, A.; et al. Marker-Assisted Pyramiding of Lycopene-ε-Cyclase, β-Carotene Hydroxylase1 and Opaque2 Genes for Development of Biofortified Maize Hybrids. Sci. Rep. 2021, 11, 12642. [Google Scholar] [CrossRef] [PubMed]
  91. Xu, Z.; Hua, J.; Wang, F.; Cheng, Z.; Meng, Q.; Chen, Y.; Han, X.; Tie, S.; Liu, C.; Li, X.; et al. Marker-Assisted Selection of QMrdd8 to Improve Maize Resistance to Rough Dwarf Disease. Breed Sci. 2020, 70, 183–192. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Abdel-Rahman, M.M.; Bayoumi, S.R.; Barakat, M.N. Identification of Molecular Markers Linked to Fusarium Ear Rot Genes in Maize Plants Zea Mays L. Biotechnol. Biotechnol. Equip. 2016, 30, 692–699. [Google Scholar] [CrossRef]
  93. Chen, Q.; Song, J.; Du, W.-P.; Xu, L.-Y.; Jiang, Y.; Zhang, J.; Xiang, X.-L.; Yu, G.-R. Identification, Mapping, and Molecular Marker Development for Rgsr8.1: A New Quantitative Trait Locus Conferring Resistance to Gibberella Stalk Rot in Maize (Zea Mays L.). Front. Plant Sci. 2017, 8, 1355. [Google Scholar] [CrossRef]
  94. Anderson, D.L. Functional Markers in Zea Mays L.; Iowa State University: Ames, IO, USA, 2021; p. 31. [Google Scholar]
  95. Duan, C.; Song, F.; Sun, S.; Guo, C.; Zhu, Z.; Wang, X. Characterization and Molecular Mapping of Two Novel Genes Resistant to Pythium Stalk Rot in Maize. Phytopathology 2019, 109, 804–809. [Google Scholar] [CrossRef]
  96. Zhang, N.; Huang, X. Mapping Quantitative Trait Loci and Predicting Candidate Genes for Leaf Angle in Maize. PLoS ONE 2021, 16, e0245129. [Google Scholar] [CrossRef]
  97. Zuo, Z.; Lu, Y.; Zhu, M.; Chen, R.; Zhang, E.; Hao, D.; Huang, Q.; Wang, H.; Su, Y.; Wang, Z.; et al. Nucleotide Diversity of the Maize ZmCNR13 Gene and Association with Ear Traits. Front. Genet. 2021, 12, 773597. [Google Scholar] [CrossRef]
  98. Makarevitch, I.; Thompson, A.; Muehlbauer, G.J.; Springer, N.M. Brd1 Gene in Maize Encodes a Brassinosteroid C-6 Oxidase. PLoS ONE 2012, 7, e30798. [Google Scholar] [CrossRef] [PubMed]
  99. Hayes, B.J.; Bowman, P.J.; Chamberlain, A.J.; Goddard, M.E. Invited Review: Genomic Selection in Dairy Cattle: Progress and Challenges. J. Dairy Sci. 2009, 92, 433–443. [Google Scholar] [CrossRef] [PubMed]
  100. Genomic Selection for Crop Improvement—Heffner—2009—Crop Science—Wiley Online Library. Available online: https://acsess.onlinelibrary.wiley.com/doi/abs/10.2135/cropsci2008.08.0512 (accessed on 30 March 2022).
  101. Poland, J.A.; Rife, T.W. Genotyping-by-Sequencing for Plant Breeding and Genetics. Plant Genome 2012, 5, 92–102. [Google Scholar] [CrossRef]
  102. Heslot, N.; Yang, H.-P.; Sorrells, M.E.; Jannink, J.-L. Genomic Selection in Plant Breeding: A Comparison of Models. Crop Sci. 2012, 52, 146–160. [Google Scholar] [CrossRef]
  103. R2D2 Consortium; Fugeray-Scarbel, A.; Bastien, C.; Dupont-Nivet, M.; Lemarié, S. Why and How to Switch to Genomic Selection: Lessons from Plant and Animal Breeding Experience. Front. Genet. 2021, 12, 629737. [Google Scholar] [CrossRef]
  104. Atanda, S.A.; Olsen, M.; Burgueño, J.; Crossa, J.; Dzidzienyo, D.; Beyene, Y.; Gowda, M.; Dreher, K.; Zhang, X.; Prasanna, B.M.; et al. Maximizing Efficiency of Genomic Selection in CIMMYT’s Tropical Maize Breeding Program. Theor. Appl. Genet. 2021, 134, 279–294. [Google Scholar] [CrossRef]
  105. Robertsen, C.D.; Hjortshøj, R.L.; Janss, L.L. Genomic Selection in Cereal Breeding. Agronomy 2019, 9, 95. [Google Scholar] [CrossRef]
  106. Pérez-Rodríguez, P.; Gianola, D.; González-Camacho, J.M.; Crossa, J.; Manès, Y.; Dreisigacker, S. Comparison Between Linear and Non-Parametric Regression Models for Genome-Enabled Prediction in Wheat. Genetics 2012, 2, 1595–1605. [Google Scholar] [CrossRef]
  107. Windhausen, V.S.; Atlin, G.N.; Hickey, J.M.; Crossa, J.; Jannink, J.-L.; Sorrells, M.E.; Raman, B.; Cairns, J.E.; Tarekegne, A.; Semagn, K.; et al. Effectiveness of Genomic Prediction of Maize Hybrid Performance in Different Breeding Populations and Environments. Genetics 2012, 2, 1427–1436. [Google Scholar] [CrossRef]
  108. Technow, F.; Bürger, A.; Melchinger, A.E. Genomic Prediction of Northern Corn Leaf Blight Resistance in Maize with Combined or Separated Training Sets for Heterotic Groups. Genetics 2013, 3, 197–203. [Google Scholar] [CrossRef]
  109. Massman, J.M.; Jung, H.-J.G.; Bernardo, R. Genomewide Selection versus Marker-Assisted Recurrent Selection to Improve Grain Yield and Stover-Quality Traits for Cellulosic Ethanol in Maize. Crop Sci. 2013, 53, 58–66. [Google Scholar] [CrossRef]
  110. Combs, E.; Bernardo, R. Genomewide Selection to Introgress Semidwarf Maize Germplasm into U.S. Corn Belt Inbreds. Crop Sci. 2013, 53, 1427–1436. [Google Scholar] [CrossRef]
  111. Crossa, J.; Pérez-Rodríguez, P.; Cuevas, J.; Montesinos-López, O.; Jarquín, D.; de los Campos, G.; Burgueño, J.; González-Camacho, J.M.; Pérez-Elizalde, S.; Beyene, Y.; et al. Genomic Selection in Plant Breeding: Methods, Models, and Perspectives. Trends Plant Sci. 2017, 22, 961–975. [Google Scholar] [CrossRef]
  112. Zhang, A.; Wang, H.; Beyene, Y.; Semagn, K.; Liu, Y.; Cao, S.; Cui, Z.; Ruan, Y.; Burgueño, J.; San Vicente, F.; et al. Effect of Trait Heritability, Training Population Size and Marker Density on Genomic Prediction Accuracy Estimation in 22 Bi-Parental Tropical Maize Populations. Front. Plant Sci. 2017, 8, 1916. [Google Scholar] [CrossRef]
  113. Shepherd, R.K.; Meuwissen, T.H.; Woolliams, J.A. Genomic Selection and Complex Trait Prediction Using a Fast EM Algorithm Applied to Genome-Wide Markers. BMC Bioinform. 2010, 11, 529. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Plant Breeding with Genomic Selection: Gain per Unit Time and Cost—Heffner—2010—Crop Science—Wiley Online Library. Available online: https://acsess.onlinelibrary.wiley.com/doi/abs/10.2135/cropsci2009.11.0662 (accessed on 30 March 2022).
  115. Jannink, J.-L.; Lorenz, A.J.; Iwata, H. Genomic Selection in Plant Breeding: From Theory to Practice. Brief. Funct. Genom. 2010, 9, 166–177. [Google Scholar] [CrossRef] [PubMed]
  116. Beyene, Y.; Gowda, M.; Pérez-Rodríguez, P.; Olsen, M.; Robbins, K.R.; Burgueño, J.; Prasanna, B.M.; Crossa, J. Application of Genomic Selection at the Early Stage of Breeding Pipeline in Tropical Maize. Front. Plant Sci. 2021, 12, 685488. [Google Scholar] [CrossRef]
  117. Hallauer, A.R.; Filho, J.B.M.; Carena, M.J. Breeding Plans. In Quantitative Genetics in Maize Breeding; Carena, M.J., Hallauer, A.R., Miranda Filho, J.B., Eds.; Handbook of Plant Breeding; Springer: New York, NY, USA, 2010; pp. 577–653. ISBN 978-1-4419-0766-0. [Google Scholar]
  118. Fritsche-Neto, R.; Akdemir, D.; Jannink, J.-L. Accuracy of Genomic Selection to Predict Maize Single-Crosses Obtained through Different Mating Designs. Theor. Appl. Genet. 2018, 131, 1153–1162. [Google Scholar] [CrossRef] [PubMed]
  119. Rice, B.R.; Lipka, A.E. Diversifying Maize Genomic Selection Models. Mol. Breed. 2021, 41, 33. [Google Scholar] [CrossRef]
  120. Xu, J. Breeding by Design for Future Rice: Genes and Genome Technologies. Crop J. 2021, 6, 491–496. [Google Scholar] [CrossRef]
  121. Peleman, J.D.; van der Voort, J.R. Breeding by Design. Trends Plant Sci. 2003, 8, 330–334. [Google Scholar] [CrossRef]
  122. Zhang, G. Target Chromosome-Segment Substitution: A Way to Breeding by Design in Rice. Crop J. 2021, 9, 658–668. [Google Scholar] [CrossRef]
  123. Pérez-de-Castro, A.M.; Vilanova, S.; Cañizares, J.; Pascual, L.; Blanca, J.M.; Díez, M.J.; Prohens, J.; Picó, B. Application of Genomic Tools in Plant Breeding. Curr. Genom. 2012, 13, 179–195. [Google Scholar] [CrossRef]
  124. Hake, S.; Ross-Ibarra, J. Genetic, Evolutionary and Plant Breeding Insights from the Domestication of Maize. eLife 2015, 4, e05861. [Google Scholar] [CrossRef] [PubMed]
  125. Scott, M.P.; Emery, M. Maize: Overview. In Reference Module in Food Science; Elsevier: Amsterdam, The Netherlands, 2016; ISBN 978-0-08-100596-5. [Google Scholar]
  126. Haberer, G.; Young, S.; Bharti, A.K.; Gundlach, H.; Raymond, C.; Fuks, G.; Butler, E.; Wing, R.A.; Rounsley, S.; Birren, B.; et al. Structure and Architecture of the Maize Genome. Plant Physiol. 2005, 139, 1612–1624. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Messing, J.; Bharti, A.K.; Karlowski, W.M.; Gundlach, H.; Kim, H.R.; Yu, Y.; Wei, F.; Fuks, G.; Soderlund, C.A.; Mayer, K.F.X.; et al. Sequence Composition and Genome Organization of Maize. Proc. Natl. Acad. Sci. USA 2004, 101, 14349–14354. [Google Scholar] [CrossRef] [PubMed]
  128. Chandler, V.L.; Brendel, V. The Maize Genome Sequencing Project. Plant Physiol. 2002, 130, 1594–1597. [Google Scholar] [CrossRef]
  129. Xu, Y.; Skinner, D.J.; Wu, H.; Palacios-Rojas, N.; Araus, J.L.; Yan, J.; Gao, S.; Warburton, M.L.; Crouch, J.H. Advances in Maize Genomics and Their Value for Enhancing Genetic Gains from Breeding. Int. J. Plant Genom. 2009, 2009, e957602. [Google Scholar] [CrossRef]
  130. Palmer, L.E.; Rabinowicz, P.D.; O’Shaughnessy, A.L.; Balija, V.S.; Nascimento, L.U.; Dike, S.; de la Bastide, M.; Martienssen, R.A.; McCombie, W.R. Maize Genome Sequencing by Methylation Filtration. Science 2003, 302, 2115–2117. [Google Scholar] [CrossRef]
  131. Mascher, M.; Gerlach, N.; Gahrtz, M.; Bucher, M.; Scholz, U.; Dresselhaus, T. Sequence and Ionomic Analysis of Divergent Strains of Maize Inbred Line B73 with an Altered Growth Phenotype. PLoS ONE 2014, 9, e96782. [Google Scholar] [CrossRef]
  132. Stojaković, M.; Ivanović, M.; Bekavac, G.; Stojaković, Ž. Grain Yield of B73 x Mo17-Type Maize Hybrids from Different Periods of Breeding. Cereal Res. Commun. 2010, 38, 440–448. [Google Scholar] [CrossRef]
  133. Stojaković, M.; Bekavac, G.; Vasić, N. B73 and Related Inbred Lines in Maize Breeding. Genetika 2005, 37, 245–252. [Google Scholar] [CrossRef]
  134. Hufford, M.B.; Seetharam, A.S.; Woodhouse, M.R.; Chougule, K.M.; Ou, S.; Liu, J.; Ricci, W.A.; Guo, T.; Olson, A.; Qiu, Y.; et al. De Novo Assembly, Annotation, and Comparative Analysis of 26 Diverse Maize Genomes. Science 2021, 373, 105655–105662. [Google Scholar] [CrossRef] [PubMed]
  135. He, J.; Zhao, X.; Laroche, A.; Lu, Z.-X.; Liu, H.; Li, Z. Genotyping-by-Sequencing (GBS), an Ultimate Marker-Assisted Selection (MAS) Tool to Accelerate Plant Breeding. Front. Plant Sci. 2014, 5, 484. [Google Scholar] [CrossRef]
  136. Shendure, J.; Ji, H. Next-Generation DNA Sequencing. Nat. Biotechnol. 2008, 26, 1135–1145. [Google Scholar] [CrossRef] [PubMed]
  137. Nepolean, T.; Kaul, J.; Mukri, G.; Mittal, S. Genomics-Enabled Next-Generation Breeding Approaches for Developing System-Specific Drought Tolerant Hybrids in Maize. Front. Plant Sci. 2018, 9, 361. [Google Scholar] [CrossRef] [PubMed]
  138. Savadi, S.; Mangalassery, S.; Sandesh, M.S. Advances in Genomics and Genome Editing for Breeding next Generation of Fruit and Nut Crops. Genomics 2021, 113, 3718–3734. [Google Scholar] [CrossRef] [PubMed]
  139. EU Regulation of New Plant Breeding Technologies and Their Possible Economic Implications for the EU and Beyond—Purnhagen—2021—Applied Economic Perspectives and Policy—Wiley Online Library. Available online: https://onlinelibrary.wiley.com/doi/full/10.1002/aepp.13084 (accessed on 30 March 2022).
  140. Eckerstorfer, M.F.; Engelhard, M.; Heissenberger, A.; Simon, S.; Teichmann, H. Plants Developed by New Genetic Modification Techniques—Comparison of Existing Regulatory Frameworks in the EU and Non-EU Countries. Front. Bioeng. Biotechnol. 2019, 7, 26. [Google Scholar] [CrossRef]
  141. Enfissi, E.M.A.; Drapal, M.; Perez-Fons, L.; Nogueira, M.; Berry, H.M.; Almeida, J.; Fraser, P.D. New Plant Breeding Techniques and Their Regulatory Implications: An Opportunity to Advance Metabolomics Approaches. J. Plant Physiol. 2021, 258–259, 153378. [Google Scholar] [CrossRef]
  142. Aglawe, S.B.; Barbadikar, K.M.; Mangrauthia, S.K.; Madhav, M.S. New Breeding Technique “Genome Editing” for Crop Improvement: Applications, Potentials and Challenges. 3 Biotech 2018, 8, 336. [Google Scholar] [CrossRef]
  143. El-Mounadi, K.; Morales-Floriano, M.L.; Garcia-Ruiz, H. Principles, Applications, and Biosafety of Plant Genome Editing Using CRISPR-Cas9. Front. Plant Sci. 2020, 11, 56. [Google Scholar] [CrossRef] [PubMed]
  144. Jaganathan, D.; Ramasamy, K.; Sellamuthu, G.; Jayabalan, S.; Venkataraman, G. CRISPR for Crop Improvement: An Update Review. Front. Plant Sci. 2018, 9, 985. [Google Scholar] [CrossRef] [PubMed]
  145. Petolino, J.F.; Kumar, S. Transgenic Trait Deployment Using Designed Nucleases. Plant Biotechnol. J. 2016, 14, 503–509. [Google Scholar] [CrossRef] [PubMed]
  146. Smith, V.; Wesseler, J.H.H.; Zilberman, D. New Plant Breeding Technologies: An Assessment of the Political Economy of the Regulatory Environment and Implications for Sustainability. Sustainability 2021, 13, 3687. [Google Scholar] [CrossRef]
  147. Wada, N.; Ueta, R.; Osakabe, Y.; Osakabe, K. Precision Genome Editing in Plants: State-of-the-Art in CRISPR/Cas9-Based Genome Engineering. BMC Plant Biol. 2020, 20, 234. [Google Scholar] [CrossRef]
  148. Lusser, M.; Parisi, C.; Plan, D.; Rodríguez-Cerezo, E. Deployment of New Biotechnologies in Plant Breeding. Nat. Biotechnol. 2012, 30, 231–239. [Google Scholar] [CrossRef]
  149. Modrzejewski, D.; Hartung, F.; Sprink, T.; Krause, D.; Kohl, C.; Schiemann, J.; Wilhelm, R. What Is the Available Evidence for the Application of Genome Editing as a New Tool for Plant Trait Modification and the Potential Occurrence of Associated Off-Target Effects: A Systematic Map Protocol. Environ. Evid. 2018, 7, 18. [Google Scholar] [CrossRef]
  150. Sauer, N.J.; Narváez-Vásquez, J.; Mozoruk, J.; Miller, R.B.; Warburg, Z.J.; Woodward, M.J.; Mihiret, Y.A.; Lincoln, T.A.; Segami, R.E.; Sanders, S.L.; et al. Oligonucleotide-Mediated Genome Editing Provides Precision and Function to Engineered Nucleases and Antibiotics in Plants. Plant Physiol. 2016, 170, 1917–1928. [Google Scholar] [CrossRef]
  151. Dalakouras, A.; Vlachostergios, D. Epigenetic Approaches to Crop Breeding: Current Status and Perspectives. J. Exp. Bot. 2021, 72, 5356–5371. [Google Scholar] [CrossRef]
  152. Matzke, A.J.; Matzke, M.A. Position Effects and Epigenetic Silencing of Plant Transgenes. Curr. Opin. Plant Biol. 1998, 1, 142–148. [Google Scholar] [CrossRef]
  153. Songstad, D.D.; Petolino, J.F.; Voytas, D.F.; Reichert, N.A. Genome Editing of Plants. Crit. Rev. Plant Sci. 2017, 36, 1–23. [Google Scholar] [CrossRef]
  154. Cisgenesis and Genome Editing: Combining Concepts and Efforts for a Smarter Use of Genetic Resources in Crop Breeding—Cardi—2016—Plant Breeding—Wiley Online Library. Available online: https://onlinelibrary.wiley.com/doi/full/10.1111/pbr.12345 (accessed on 30 March 2022).
  155. Espinoza, C.; Schlechter, R.; Herrera, D.; Torres, E.; Serrano, A.; Medina, C.; Arce-Johnson, P. Cisgenesis and Intragenesis: New Tools for Improving Crops. Biol. Res. 2013, 46, 323–331. [Google Scholar] [CrossRef] [PubMed]
  156. 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]
  157. 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]
  158. Súnico, V.; Higuera, J.J.; Molina-Hidalgo, F.J.; Blanco-Portales, R.; Moyano, E.; Rodríguez-Franco, A.; Muñoz-Blanco, J.; Caballero, J.L. The Intragenesis and Synthetic Biology Approach towards Accelerating Genetic Gains on Strawberry: Development of New Tools to Improve Fruit Quality and Resistance to Pathogens. Plants 2022, 11, 57. [Google Scholar] [CrossRef]
  159. Song, G.; Walworth, A.E.; Loescher, W.H. Grafting of Genetically Engineered Plants. J. Am. Soc. Hortic. Sci. 2015, 140, 203–213. [Google Scholar] [CrossRef]
  160. Kodama, H.; Miyahara, T.; Oguchi, T.; Tsujimoto, T.; Ozeki, Y.; Ogawa, T.; Yamaguchi, Y.; Ohta, D. Effect of Transgenic Rootstock Grafting on the Omics Profiles in Tomato. Food Saf. 2021, 9, 32–47. [Google Scholar] [CrossRef]
  161. Tsaballa, A.; Xanthopoulou, A.; Madesis, P.; Tsaftaris, A.; Nianiou-Obeidat, I. Vegetable Grafting from a Molecular Point of View: The Involvement of Epigenetics in Rootstock-Scion Interactions. Front. Plant Sci. 2021, 11, 621999. [Google Scholar] [CrossRef]
  162. Ishino, Y.; Krupovic, M.; Forterre, P. History of CRISPR-Cas from Encounter with a Mysterious Repeated Sequence to Genome Editing Technology. J. Bacteriol. 2018, 200, e00580-17. [Google Scholar] [CrossRef]
  163. Charpentier, E.; Doudna, J.A. Rewriting a Genome. Nature 2013, 495, 50–51. [Google Scholar] [CrossRef]
  164. Xing, H.-L.; Dong, L.; Wang, Z.-P.; Zhang, H.-Y.; Han, C.-Y.; Liu, B.; Wang, X.-C.; Chen, Q.-J. A CRISPR/Cas9 Toolkit for Multiplex Genome Editing in Plants. BMC Plant Biol. 2014, 14, 327. [Google Scholar] [CrossRef] [PubMed]
  165. Kouranov, A.; Armstrong, C.; Shrawat, A.; Sidorov, V.; Huesgen, S.; Lemke, B.; Boyle, T.; Gasper, M.; Lawrence, R.; Yang, S. Demonstration of Targeted Crossovers in Hybrid Maize Using CRISPR Technology. Commun. Biol. 2022, 5, 1–11. [Google Scholar] [CrossRef] [PubMed]
  166. Shankar, S.; Sreekumar, A.; Prasad, D.; Das, A.V.; Pillai, M.R. Genome Editing of Oncogenes with ZFNs and TALENs: Caveats in Nuclease Design. Cancer Cell Int. 2018, 18, 169. [Google Scholar] [CrossRef]
  167. Petolino, J.F. Genome Editing in Plants via Designed Zinc Finger Nucleases. Vitr. Cell. Dev. Biol.-Plant 2015, 51, 1–8. [Google Scholar] [CrossRef]
  168. Davies, J.P.; Kumar, S.; Sastry-Dent, L. Use of Zinc-Finger Nucleases for Crop Improvement. Prog. Mol. Biol. Transl. Sci. 2017, 149, 47–63. [Google Scholar] [CrossRef]
  169. Shukla, V.K.; Doyon, Y.; Miller, J.C.; DeKelver, R.C.; Moehle, E.A.; Worden, S.E.; Mitchell, J.C.; Arnold, N.L.; Gopalan, S.; Meng, X.; et al. Precise Genome Modification in the Crop Species Zea Mays Using Zinc-Finger Nucleases. Nature 2009, 459, 437–441. [Google Scholar] [CrossRef]
  170. Weinthal, D.; Tovkach, A.; Zeevi, V.; Tzfira, T. Genome Editing in Plant Cells by Zinc Finger Nucleases. Trends Plant Sci. 2010, 15, 308–321. [Google Scholar] [CrossRef] [PubMed]
  171. ISAAA Inc. Pocket K59 (English). Available online: https://www.isaaa.org/resources/publications/ (accessed on 25 September 2022).
  172. Malzahn, A.; Lowder, L.; Qi, Y. Plant Genome Editing with TALEN and CRISPR. Cell Biosci. 2017, 7, 21. [Google Scholar] [CrossRef]
  173. Sauer, N.J.; Mozoruk, J.; Miller, R.B.; Warburg, Z.J.; Walker, K.A.; Beetham, P.R.; Schöpke, C.R.; Gocal, G.F.W. Oligonucleotide-Directed Mutagenesis for Precision Gene Editing. Plant Biotechnol. J. 2016, 14, 496–502. [Google Scholar] [CrossRef]
  174. Zhu, T.; Mettenburg, K.; Peterson, D.J.; Tagliani, L.; Baszczynski, C.L. Engineering Herbicide-Resistant Maize Using Chimeric RNA/DNA Oligonucleotides. Nat. Biotechnol. 2000, 18, 555–558. [Google Scholar] [CrossRef]
  175. Weinhold, B. Epigenetics: The Science of Change. Environ. Health Perspect. 2006, 114, A160–A167. [Google Scholar] [CrossRef] [PubMed]
  176. Kakoulidou, I.; Avramidou, E.V.; Baránek, M.; Brunel-Muguet, S.; Farrona, S.; Johannes, F.; Kaiserli, E.; Lieberman-Lazarovich, M.; Martinelli, F.; Mladenov, V.; et al. Epigenetics for Crop Improvement in Times of Global Change. Biology 2021, 10, 766. [Google Scholar] [CrossRef] [PubMed]
  177. Varotto, S.; Tani, E.; Abraham, E.; Krugman, T.; Kapazoglou, A.; Melzer, R.; Radanović, A.; Miladinović, D. Epigenetics: Possible Applications in Climate-Smart Crop Breeding. J. Exp. Bot. 2020, 71, 5223–5236. [Google Scholar] [CrossRef]
  178. Chen, Y.; Müller, F.; Rieu, I.; Winter, P. Epigenetic Events in Plant Male Germ Cell Heat Stress Responses. Plant Reprod. 2016, 29, 21–29. [Google Scholar] [CrossRef]
  179. Regulski, M.; Lu, Z.; Kendall, J.; Donoghue, M.T.A.; Reinders, J.; Llaca, V.; Deschamps, S.; Smith, A.; Levy, D.; McCombie, W.R.; et al. The Maize Methylome Influences MRNA Splice Sites and Reveals Widespread Paramutation-like Switches Guided by Small RNA. Genome Res. 2013, 23, 1651–1662. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Telem, R.S.; Wani, S.H.; Singh, N.B.; Nandini, R.; Sadhukhan, R.; Bhattacharya, S.; Mandal, N. Cisgenics—A Sustainable Approach for Crop Improvement. Curr. Genom. 2013, 14, 468–476. [Google Scholar] [CrossRef]
  181. Jacobsen, E.; Schouten, H.J. Cisgenesis. In Molecular Techniques in Crop Improvement, 2nd ed.; Jain, S.M., Brar, D.S., Eds.; Springer Netherlands: Dordrecht, The Netherlands, 2009; pp. 591–611. ISBN 978-90-481-2967-6. [Google Scholar]
  182. Almeraya, E.V.; Sánchez-de-Jiménez, E. Intragenic Modification of Maize. J. Biotechnol. 2016, 238, 35–41. [Google Scholar] [CrossRef]
  183. Yadava, P.; Abhishek, A.; Singh, R.; Singh, I.; Kaul, T.; Pattanayak, A.; Agrawal, P.K. Advances in Maize Transformation Technologies and Development of Transgenic Maize. Front. Plant Sci. 2017, 7, 1949. [Google Scholar] [CrossRef]
  184. Zhao, D.; Song, G. Rootstock-to-Scion Transfer of Transgene-Derived Small Interfering RNAs and Their Effect on Virus Resistance in Nontransgenic Sweet Cherry. Plant Biotechnol. J. 2014, 12, 1319–1328. [Google Scholar] [CrossRef]
  185. Yang, Y.; Mao, L.; Jittayasothorn, Y.; Kang, Y.; Jiao, C.; Fei, Z.; Zhong, G.-Y. Messenger RNA Exchange between Scions and Rootstocks in Grafted Grapevines. BMC Plant Biol. 2015, 15, 251. [Google Scholar] [CrossRef]
  186. Gedil, M.; Menkir, A. An Integrated Molecular and Conventional Breeding Scheme for Enhancing Genetic Gain in Maize in Africa. Front. Plant Sci. 2019, 10, 1430. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Maize breeding stages: from plant domestication to precise improvement.
Figure 1. Maize breeding stages: from plant domestication to precise improvement.
Agronomy 12 02365 g001
Figure 2. Operation process of bioinformatics in plant breeding.
Figure 2. Operation process of bioinformatics in plant breeding.
Agronomy 12 02365 g002
Figure 3. A simplified approach of design in silico.
Figure 3. A simplified approach of design in silico.
Agronomy 12 02365 g003
Table 1. Zea mays breeding-related biological datasets and bioinformatics databases.
Table 1. Zea mays breeding-related biological datasets and bioinformatics databases.
DatabaseDescriptionReferenceLink
Maize Genetic and Genomic Database (MaizeGBD)Genomic sequence data
Gene expression analysis
Phenotypes of plants
[44,45,46,47,48,49]https://www.maizegdb.org/, accessed on 26 September 2022
Maize Database of Images
and Genomes (MaizeDIG)
Genotype/phenotype database for maize[50,51]http://maizedig.maizegdb.org/, accessed on 26 September 2022
MaizeSNPDBDatabase of maize lines and SNP sites[52]https://venyao.xyz/MaizeSNPDB/, accessed on 26 September 2022
ZeaMapMaize and teosinte comparative genomics
Multi-omics data
Genetic maps
[53,54,55,56]http://www.zeamap.com/, accessed on 26 September 2022
PanzeaMolecular and functional diversity of the maize genome [56,57,58]https://www.panzea.org/data, accessed on 26 September 2022
GrameneComparative genomics of plants[59,60,61,62]https://www.gramene.org/, accessed on 26 September 2022
Plant Genome Database
(GBD)
Plant genome information,
Molecular sequence data
[48,63]https://www.plantgdb.org/, accessed on 26 September 2022
UniProtDatabase of protein sequences[64,65]https://www.uniprot.org/, accessed on 26 September 2022
GenbankDatabase of nucleic acids sequences[66,67,68]https://www.ncbi.nlm.nih.gov/genbank/, accessed on 26 September 2022
European Molecular Biology Laboratory (EMBL) DatabaseNucleotide sequence database[69]http://www.ebi.ac.uk, accessed on 26 September 2022
PhytozomePlant comparative genomics[63,70]https://phytozome-next.jgi.doe.gov/, accessed on 26 September 2022
EsemblPlantsGenome-scale information[63,71]http://plants.ensembl.org/species.html, accessed on 26 September 2022
Table 2. Summary of significant molecular markers for Zea mays.
Table 2. Summary of significant molecular markers for Zea mays.
Gene/LocusMarkerTarget TraitReferences
Ht1umc150Bresistance to the fungal pathogen Helminthosporium turcicum[88]
o2umc1066quality, lysine, and
tryptophan content
[89,90]
crtRB1crtRB1 3′TE
lcyElcyE 5′TE, SNP216, SNP2238, lcyE 3′InDel[90]
qMrdd8IDP25Kresistance to rough dwarf disease[91]
QTLbnlg1063, umc2082, bnlg1621, umc2013, bnlg1740, umc2059, SSR85, STS06resistance to Fusarium verticillioides
ear root
[92]
Zm953, Zm972/Rgsr8.1SSR78resistance to Gibberella stalk rot[93]
su1SuDel6FR, SuDel36FR, SNP2703, SNP2703-CG85/89encodes a starch debranching enzyme that influences the sweetness
of kernels
[94]
RpiX178-1umc2047, IDP2347, SSRZ8resistance to Pythium stalk rot[95]
RpiX178-2bnlg1444, umc2041
Zm00001d006675bnlg2077DNA repair[96]
ZmCNR13InDel1413, SNP2305, SNP2337ear traits[97]
brd1umc2240, SNP M1-8, IDP7806brassinosteroid synthesis[98]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Muntean, L.; Ona, A.; Berindean, I.; Racz, I.; Muntean, S. Maize Breeding: From Domestication to Genomic Tools. Agronomy 2022, 12, 2365. https://doi.org/10.3390/agronomy12102365

AMA Style

Muntean L, Ona A, Berindean I, Racz I, Muntean S. Maize Breeding: From Domestication to Genomic Tools. Agronomy. 2022; 12(10):2365. https://doi.org/10.3390/agronomy12102365

Chicago/Turabian Style

Muntean, Leon, Andreea Ona, Ioana Berindean, Ionuț Racz, and Sorin Muntean. 2022. "Maize Breeding: From Domestication to Genomic Tools" Agronomy 12, no. 10: 2365. https://doi.org/10.3390/agronomy12102365

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