Open access peer-reviewed chapter

Maize (Zea mays L.) as a Model System for Plant Genetic, Genomic, and Applied Research

Written By

Fakhriddin N. Kushanov, Ozod S. Turaev, Oybek A. Muhammadiyev, Ramziddin F. Umarov, Nargiza M. Rakhimova and Noilabonu N. Mamadaliyeva

Submitted: 17 October 2021 Reviewed: 24 March 2022 Published: 23 June 2022

DOI: 10.5772/intechopen.104658

From the Edited Volume

Model Organisms in Plant Genetics

Edited by Ibrokhim Y. Abdurakhmonov

Chapter metrics overview

215 Chapter Downloads

View Full Metrics

Abstract

Maize leads the world’s cereals after wheat and rice in terms of cultivated area, because of its economic importance for the production of both food purposes and raw materials for industry. The maize genus Zea L. belonging to the family of cereals (Poaceae or Graminaceae) includes six species. However, all cultivated maize belongs specifically to Zea mays L. subsp. mays (2n = 2× = 20) is the only cultivated species of the genus Zea L., and the remaining species of this genus are mostly wild herbaceous plants. In addition to meeting the nutritional needs of the world’s population, Zea mays L. is one of the classic model objects of genetic and physiological research, as well as in the field of breeding not only cereals but also other important agricultural plants. Especially, this model object has been used in genetic mapping of loci of quantitative traits and genes associated with economically valuable traits, such as yield, resistance to diseases and pests, grain quality, etc. in cereal crops.

Keywords

  • Zea mays L.
  • hybridization
  • cytoplasmic male sterility
  • QTL
  • mapping
  • GWAS

1. Introduction

Due to the constant growth of the world’s population, the demand for high-calorie foods is increasing. Although maize was developed as an American crop, it is now grown all over the world and today it has become the third most important food crop after wheat and rice [1, 2]. Maize is the world’s leading cereal after wheat and rice in terms of sown area, as currently makes up about 21% of the human diet worldwide and more than 500 different staples and additives are produced from it (FAOSTAT data). According to the International Grains Council (IGC), the corn grain harvest in 2021 was about 1.12 billion tons.

A special role in the genetic analysis is played by model objects, by working with which the researcher can significantly speed up and facilitate the process of analysis. Maize (Zea mays L.) is one of the main classical models for fundamental research in the fields of plant genetics and breeding. Especially, this model object has been used in genetic mapping of loci of quantitative traits and genes associated with economically valuable traits, such as yield, resistance to diseases and pests, grain quality, etc. in cereal crops. Since its chromosomes are easily analyzed under an optical microscope, maize is also suitable for plant cytogenetic analysis. The simplicity of castration (removal of male inflorescences—panicle), the presence of mutations that cause male sterility, the possibility of setting seeds both during cross-pollination and during self-pollination, the presence of a huge number of various mutations facilitates hybridization.

The genus Zea L. from the grass family (Poaceae or Graminaceae) is represented by four diploid (2n = 2× = 20) and one tetraploid (2n = 4× = 40) species;

  1. Zea diploperennis—diploperennial teosinte,

  2. Zea luxurians—teosinte,

  3. Zea nicaraguensis,

  4. Zea mays L.—corn are diploids, and

  5. Zea perennis—perennial teosinte is a tetraploid species.

In turn, Zea mays L. is divided into three subspecies, such as,

  1. Zea mays subsp. huehuetenangensis—maize,

  2. Zea mays L. subsp. mays and

  3. Zea mays subsp. parviglumis.

As well as there are three subspecies including

  1. Zea mays subsp. mays—corn,

  2. Zea mays subsp. mexicana (Schrad.), and

  3. Zea mays subsp. parviglumis.

All species belonging to the genus Zea L. cross with other maize diploid species, except the perennial tetraploid Z. perennis. While, the diploid maize and its wild relative, teosinte Zea perennis (Hitchc.) Reeves & Mangelsd, readily intercrossing [3].

Maize (Zea mays L. subsp. mays) is the single cultivated species of the genus Zea L. The genome size of the diploid maize species ranges from 2.2 to 2.7 GB, including approximately 32.000–42.000 protein-coding genes [4, 5]. The first tetraploid species of Zea mays L. was obtained by Randolph (1932), using the heat shock method [6]. The tetraploid species is characterized by the strong development of all plant organs, resistance to abiotic and biotic environmental factors, and increased content of nutrients compared to diploid species [7].

Currently, genetic, chromosomal, genomic, and cytoplasmic modifications have been identified in maize, in particular, gene mutations have been best studied [489]. Especially, the genes that control the behavior of chromosomes in mitosis and meiosis, enzyme systems, the formation of chlorophyll and other pigments have been studied and described; structures and functions of vegetative organs, structure, and color of the endosperm, regulatory systems responsible for the mutability and expression of other genes, for the development of various elements of the reproductive system, which determine male and female sterility, selective fertilization, etc. [4, 8, 10].

Moreover, in maize were found and well-studied spontaneous and induced chromosome rearrangements such as deficiencies, translocations, duplications, and inversions [4]. In recent years, translocations have been widely used in maize to determine linkage groups [8].

Since the discovery of polyploidy forms of maize, many of them have been well studied. They are found in the aneuploids—trisomics, and monosomics in maize [8]. On maize, cytological evidence of crossing over was obtained for the first time in plants and mobile genetic elements were discovered [8, 11]. It studied the influence of long-term inbreeding and the effects of heterosis in plants and developed hybrid breeding techniques based on obtaining and crossing pure lines (interline and double interline hybrids); cytoplasmic mutations are well studied, especially mutations associated with cytoplasmic male sterility (CMS), the use of which is one of the achievements of maize genetics and plant genetics in general.

Advertisement

2. The origin and evolution of maize

Even though the origin of maize has been studied in-depth, it remains controversial. Prehistoric breeder practitioners who cannot live and reproduce in their current form without human help [12] domesticated maize (Zea mays L.). After the American continent was discovered, it became clear that corn was the staple food of the endemic Indians on the continent. According to several authors, maize was introduced into cultivation 7–12.000 years ago in the territory of southwestern Mexico. Cave excavations in arid regions of Mexico have unearthed small grains of corn grown for food 5.000 years ago [13]. The oldest finds of cultivated corn kernels were discovered in the caves of Gwila Nakitz and Tehuacan, located in the northwestern state of Oaxaca and the southeastern state of Puebla in central Mexico. In addition, according to archaeobotanists Ranere et al. (2009), the first straight there is evidence that maize was domesticated about 8.700 years ago in the Balsas region from the wild teosinte plant [14].

Taxonomic and evolutionary studies indicate that the teosinte is the closest wild relative of four annual and perennial maize species of the genus Zea L. [3]. However, according to the authors, some species of teosinte are genetically and taxonomically differing from Zea mays L. While, in the study of maize genetics, researchers consider teosinte to be an important resource.

Thus, according to archaeobotanical findings and the results of traditional analyzes [12, 13], there are various theories about the origin of Zea mays L. ssp. mays:

  1. As a result of the selection of the wild subspecies Zea mays ssp. parviglumis. In addition, due to the possible introgressive hybridization with the ancestral form Z. mays ssp. mexicana, the genetic material of up to 12% of cultivated form might be obtained from this subspecies.

  2. Caused by the hybridization of small cultivated wild form with another species of this genus—either Z. luxurians or Z. diploperennis.

  3. One wild form has been introduced into the crop several times.

  4. From the hybridization of Zea diploperennis with some representatives of the closely related genus Tripsacum.

Advertisement

3. Maize polyploidy studies

Polyploidy or whole-genome duplication (WGD) is considered as a major process in plant evolution. A polyploidy event 160 million years ago is theorized to have created the ancestral line that led to all modern flowering plants [15]. Genome duplication is categorized into two events paleopolyploidy (ancient polyploidy) and neopolyploidy (recent polyploidy). Ancient genome duplications are widespread throughout eukaryotic lineages, particularly in plants. Paleopolyploidy has occurred at least several million years ago. Both phenomena, ancient and recent polyploids could occur through the doubling of the genome of single species (autopolyploidy) or combining genomes of two different species (allopolyploidy).

According to the maize DNA sequence data, the genome duplications event occurred after the divergence between sorghum and maize [10]. The duplications event that happened approximately 11.4 million years ago resulted from an ancient polyploid [16]. The maize WGD resulted in the subgenomes maize1 and maize2 [17]. Polyploids are found almost in all groups of eukaryotic organisms as a result of incorrect meiosis, fertilization, or cell division [18]. Polyploids can be obtained experimentally by treatment with chemicals such as colchicine, oryzalin, trifluralin and amiprophosmethyl or by combining diploid nuclei.

Niazi et al. (2014) have studied induced polyploidy in maize hybrids to increase heterosis and restore reproductive fertility [19]. The seeds of open-pollinated maize breeding lines and a maize × teosinte cross were germinated in colchicine solution (0.25, 0.5, or 1.0%) until they had a thick radical and protruded plumule. The highest number of tetraploids with the lowest number of chimeric plants induced at 0.5% colchicine. Scientists have reported that the leaf area, total soluble solids, leaf oil percentage, and leaf crude protein contents were significantly increased in leaves of the induced tetraploids of maize and maize × teosinte crosses relative to the diploid subspecies.

Iqbal et al. (2018) have conducted research aimed to clarify the mysterious meiotic behavior of autopolyploid and allopolyploid maize [20]. Scientists have explored the stability of the chromosomes during meiosis in both auto- and allopolyploid maize. Furthermore, they have identified an association of chromosomes between maize and Z. perennis by obtaining a numerous of auto- and allopolyploid maize hybrids. The results showed a higher level of chromosome stability in allopolyploid maize during meiosis than in autopolyploid maize. Additionally, the meiotic behavior of Z. perennis was relatively more stable than the allopolyploid maize. As well as, 10 chromosomes of maize “A” subgenomes were homologous to 20 chromosomes of Z. perennis genome with little evolutionary differentiation and a higher pairing frequency. However, “A” subgenome chromosomes have shown a little evolutionary differentiation, while “B” subgenome chromosomes had a lower pairing frequency and higher evolutionary differentiation in maize.

The diversity analysis of wild relatives of maize showed that various genes have different histories and domestication such as intensive breeding processes have had heterogeneous effects on genetic diversity across genes [10].

Advertisement

4. Maize is a model in genetic mapping studies for cereal crop improvement

4.1 QTL mapping and GWAS for dissecting genetic architecture of complex traits

A quantitative trait is a measurable phenotype and varies continuously among individuals in a population. Quantitative trait loci (QTL) are genomic regions in which genetic segregation within a population is statistically associated with variation in a quantitative trait. Genetic mapping of QTL is a process of locating genes with effects on quantitative traits using molecular markers. QTL mapping is a powerful method for improving agricultural crops, which allows using marker-assisted selection technology to introgression the genes of interest from one genotype to another [21].

QTL mapping and genome-wide association study (GWAS) both are similar and they typically measure the associations between genotype and phenotype [21, 22]. GWAS is a powerful tool for dissecting the genetic architecture of complex traits in many crop species [22, 23, 24]. The main goal of GWAS is to link genotypic variations with phenotypic differences. With the development of whole-genome sequencing technology and high-density single-nucleotide polymorphism (SNP) BeadCap, GWAS has also begun to be widely used to identify candidate genes that control quantitative traits in crops [22].

Maize is a suitable crop for the GWAS approach and considerable progress has been made over the last decade [25]. GWAS has been successfully used in maize to detect a great many candidate QTL/genes attending to control diverse morpho-biological and economically important traits, such as salt [26, 27, 28] and drought tolerance [8, 23, 29, 30, 31], kernel traits [32, 33, 34, 35, 36] and many other traits of interest. GWAS facilitates to achieve advances in current studies in quantitative genetics.

4.2 Genetic analysis and fine mapping of QTL for kernel traits

Grain traits are the most important in maize commercialization over the world. Kernel sizes and weight are major traits for grain yield in maize. Liu et al. (2014) was conducted the genetic analysis and identified major QTL for maize kernel size and weight [37]. They have identified a total of 55 and 28 QTL of maize kernel-size traits and kernel weight using composite interval mapping (CIM) for single-environment analysis along with mixed linear model-based CIM for joint analysis, respectively.

Wang et al. (2020) have conducted QTL analysis and fine mapping using a composite interval mapping (CIM) method aimed to map QTLs and predict candidate genes for kernel size in maize [38]. Five QTL were identified for kernel length and five QTL for kernel width out of 10 QTL.

Pan et al. (2017) reported the results of QTL mapping in six environments and consensus loci for grain weight detected by meta-analysis [39]. Subsequently, a meta-analysis was performed and 62 QTLs were determined for grain weight, ear weight, and kernel weight per plant in six environments.

Li et al. (2010) have carried out QTL mapping for grain yield and yield components under high and low phosphorus treatments in maize [40]. 69 QTL were identified for the six traits at two sites. Thirty-six distinct QTL were identified from Taian, in which 7 out of 36 for grain yield, 7 for 100 kernel weight, 5 for ear length, 5 for per ear, 6 for kernel number per row, and 6 for ear diameter, while 33 distinct QTLs were identified at Yantai, in which 6 out of 33 for grain yield, 5 for 100 kernel weight, 5 for ear length, 7 for row number per ear, 5 for kernel number per row and 5 for ear diameter.

Liu et al. (2020) have identification of QTL for kernel-related traits and the heterosis for these traits [34]. They developed and evaluated 301 RILs population for six kernel-related traits and the mid-parent heterosis (MPH) for these traits. A total of 100 QTLs were identified in both mapping populations. As well, 20 QTL clusters including 46 QTLs were identified across ten chromosomes. These results may provide additional insights into the genetic basis for the mid-parent heterosis for kernel-related traits.

Liu et al. (2015) conducted a genetic analysis of kernel traits in maize-teosinte introgression populations [33]. Scientists have analyzed kernel morphological traits in 10 maize-teosinte introgression populations using digital imaging software. QTLs were identified for kernel area and length with moderate allelic effects that colocalize with kernel weight QTL.

Another group of researchers has conducted linkage and association mapping aims to the analysis of the genetic architecture of maize kernel size [34]. Three kernel traits of maize, kernel length, kernel width, and kernel thickness, were studied in germplasm accessions and a biparental population. A total of 21 SNPs were identified under four environments. Besides, 50 QTL were determined in seven environments doubled haploid population. Combining the two mapping populations revealed that 56 SNPs fell within 18 of the QTL confidence intervals. A total of 73 candidate genes were detected, regulating seed development. As well, seven miRNAs were found to locate within the linkage disequilibrium regions of the colocalized SNPs.

Jiang et al. (2013) performed a meta-analysis of 584 QTLs related to grain yield components [41]. A total of 73 Meta-QTLs for grain yield components such as 22 QTLs for row number, 7 QTLs for kernel number per row, and 44 QTLs for kernel weight were estimated. Another group of Chinese scientists carried out combining meta-QTL with RNA-seq data to identify candidate genes of kernel row number traits [42]. A total of 373 QTL for grain yield and kernel row number was meta-analyzed. Fifty-four meta-QTL were determined, including 19 for grain yield and 35 for kernel row number. A total of 1.588 genes located in the kernel row number meta-QTL regions were identified by gene expression data.

DNA markers associated with kernel traits could be applied to marker-assisted selection (MAS) to facilitate yield architecture, QTL fine mapping, and gene cloning in the maize community [42].

4.3 The maize multiparental populations advance mapping resolution and power

4.3.1 Maize-NAM population as a template for other crops

Molecular mapping is typically carried out using genetically segregated F2, backcross (BC), recombinant inbred lines (RIL), doubled haploids (DH), and near-isogenic lines (NIL). These commonly used biparental populations have their weaknesses such as lower power, limited recombination, temporary nature, the impossibility of estimation of dominant effects, time requirement, and expense. To overcome some of the shortcomings in quantitative trait locus (QTL) mapping in biparental populations, schemes for creating mapping populations with multiple parental genotypes have been developed. The genetic diversity of these types of populations along with causing a wide range of phenotypes, it makes possible to identify QTLs with high accuracy.

The nested association mapping (NAM) population is also an example of the experimental design for multiparental populations (Figure 1). The NAM strategy was first developed in collaboration with researchers Buckler et al. [43] to study the genetic architecture of complex traits of maize (Zea mays L.). It should be noted that, unlike association mapping, NAM is a unique method that is performed only in a specially developed population [43].

Figure 1.

The nested association mapping (NAM) population scheme.

The theory of nested association mapping

The main goal of the NAM is to efficiently link phenotypic traits with genotypic data, as in a traditional QTL mapping strategy. The NAM method with low marker density, high allele richness, high mapping resolution, and high statistical power overcomes the disadvantages of Linkage analysis and Association mapping as well took advantages of both methods.

The NAM strategy allows researchers to effectively apply systematic methods of genetics and genomics and create sources such as general mapping populations as well to explore complex traits of plants at the fundamental level.

The NAM strategy involves the following stages [43]:

  1. Selection of diverse founders and the development of a mapping population (RILs with a stable set of phenotypic traits are preferred);

  2. Sequencing or high-density genotyping of parental genotypes;

  3. Genotyping of both the founders and the progenies with a smaller number of tagging markers to explain the inheritance of chromosome segments and to project the high-density marker information from the founders to the progenies;

  4. Phenotyping of hybrids/RILs for various complex traits;

  5. Conducting genome-wide association analysis using genotypic and phenotypic data.

The maize-NAM population was developed by crossing 25 diverse founders to a single common inbred line, B73, resulting in 5.000 RILs. Buckler et al. (2009) reported the results of the study on the genetic architecture of flowering time using the maize-NAM population [44, 45]. Subsequently, several NAM populations in maize were developed such as in Dent and Flint maize [46], Chinese inbred lines population-based NAM [47], and teosinte [48].

The maize-NAM design served as a model for other crops such as sorghum [149, 50, 51], peanut [52, 53], barley [11, 54, 55, 56, 57], oilseed rape [22, 58], wheat [59, 60, 61, 62, 63], rice [64, 65], soybean [66, 67] and cotton [68, 69] as well as NAM were developed in the model plant Arabidopsis thaliana [70, 71].

4.3.2 MAGIC population with greatly reduce mating design

Over the past decade, the use of multiparental populations was increased in plant genetic research. The two most popular multiparental population designs in crops are NAM and MAGIC (multi-parent advanced generation inter-cross) populations (Figure 2) [72]. MAGIC populations offer new opportunities in genetic mapping strategies and crop breeding approaches due to their complex pedigree structure [73]. MAGIC was first proposed and applied in mice [74], as well as Mackay and Powell (2007), and Cavanagh et al. (2008) first discussed in plants [75, 76]. The first plant-MAGIC population was developed in A. thaliana parents [77].

Figure 2.

Multi-parent advanced generation inter-cross (MAGIC) population scheme.

According to the MAGIC designs [75], several inbred lines are intercrossed many times over in aiming to assemble mosaic parental alleles in a single line (Figure 1). Two different MAGIC populations were developed in maize [78, 79, 80].

Dell’Acqua et al. (2015) produced 1.636 MAGIC maize RILs derived from eight genetically diverse inbred lines [79]. They show how MAGIC maize may find strong candidate genes by incorporating genome sequencing and transcriptomics data. They discuss several QTL for grain yield and flowering time, reporting candidate genes. Anderson et al. (2018) were developed four parent maize populations with five different mating designs used in MAGIC and bi-parental populations including 1.149 individuals [78]. The combined population here is comprised of 118.509 genetic markers. They conducted association mapping and identified 2, 5, 7, and 6 QTL for plant height, ear height, days to anthesis, and silking, respectively [78].

Advertisement

5. “Omics” tools to understand molecular mechanisms of major traits

5.1 RNAi and genome editing tools for control gene expression

The possibility of using an organism’s own gene and systematically inducing and triggering RNA interference (RNAi) for any desired sequence made RNAi an effective approach for functional genomics [81]. RNAi is a major biological process in plants that causes gene silencing both transcriptionally and post-transcriptionally. RNAi has been widely used in crops since its discovery. To date, this approach has been conventionally based on the use of transgenic plants expressing double-stranded RNAs (dsRNAs) against selected targets [82].

Segal et al. (2003) have conducted initial studies on RNAi mechanisms in maize [83]. They found that maize transformed RNAi constructs for 22-kD zein gene suppression could produce a dominant opaque phenotype. This phenotype suppresses 22-kD zeins without affecting the accumulation of other zein proteins.

Casati et al. (2006) have conducted GWAS of high-altitude maize and gene knockdown stocks implicate chromatin-remodeling proteins in response to UV-B [84]. They implemented comparative analysis by expression profiling of maize aim to determine new components in the mechanisms of maize responses to UV-B. Microarray analysis illustrated that among the UV-B responsive transcripts, various types of genes implicated in chromatin remodeling are differentially expressed before and after UV-B treatment in high-altitude lines. Transgenic RNAi plants with lower expression of four chromatin-associated genes showed hypersensitivity to UV-B, and altered UV-B regulation of selected genes. The results showed that genes attended in chromatin remodeling are crucial for UV-B acclimation and that some lines illustrate adaptations to this challenge.

Besides, Casati and Walbot (2008) have reported different transcriptome changes in RNAi lines. They used 44 K Agilent oligonucleotide array platform to compare RNAi lines to each other and to UV-B tolerant nontransgenic siblings both before and after 8 h of UV-B exposure [85]. Maize leaves express more than 20.000 different transcripts under greenhouse conditions; after UV-B exposure 267 transcripts exhibit expression changes in control genotypes of B73.

In recent years, RNAi research in maize has been developing rapidly. One of the agricultural economic problem is aflatoxins that are produced by fungus species such as Aspergillus. In spite of control efforts, aflatoxin contamination is causing the global loss of crops productions each year. Thakare et al. (2017) have obtained aflatoxin-free transgenic maize using host-induced gene silencing [86]. Scientists show that host-induced gene silencing is an effective method for eliminating this toxin in transgenic maize. They transformed RNAi-gene cassette targeting aflC gene, which encodes an enzyme in the Aspergillus aflatoxin biosynthetic pathway to the maize plants. The aflatoxin was not be detected in kernels of RNAi-maize plants after pathogen infection, while toxin loads reached thousands of parts per billion in nontransgenic control kernels. The results show that siRNA molecules can be used to silence aflatoxin biosynthesis in maize.

Velez et al. (2020) have studied the lethal and sublethal effects of Sec23 dsRNA in maize RNAi lines to control western corn rootworm (WCR) [87]. They determined Sec23 as a highly lethal RNAi target using WCR adult feeding assays. Scientists explain Sec23 dsRNA as an RNAi target for planta rootworm control.

In the last few years, new genome editing methods have emerged that use four types of engineered nucleases: Meganucleases, ZFN (Zinc-Finger Nucleases), TALENs (Transcription Activator Like Effector Nucleases), and the CRISPR (Clustered Regularly Interspaced Short Palindromic Repeats) system. Among model plants, including maize, all kinds of nucleases mentioned above have been used to create targeted genome modifications [88, 89]. For example, D’Halluin et al. (2008) reported the first use of targeted genome modification using customizable endonucleases in maize. They succeeded in inserting a 35S promoter upstream of a promoterless herbicide resistance transgene using a meganuclease [90]. Later, Shukla et al. (2009) reported the first use of ZFNs for site-directed mutagenesis at the maize IPK1 gene as well as site-directed DNA insertion of a PAT gene [91]. Especially, among the DNA-Free Genome Editing technologies TALEN and CRISPR/cas9 technologies, have become powerful tools for genomic research. For the first time in maize, a group of scientists [92] using both TALEN and CRISPR systems reported the results of site-directed mutagenesis in maize. They designed 5 TALEN and two CRISPR constructions that target three genes involved in phytic acid (PA) biosynthesis. The results of this study served to reduce the content of PA in the seed and this led the authors to conclude that both technologies can be used to modify the maize genome. The following year, using this technology has been obtained the generation of stable, heritable mutations at the maize glossy2 locus [93]. As a result of this study, transgenic plants containing mono- or diallelic mutations were obtained with a frequency of about 10%. However, the TALEN method is more labor-intensive, requiring more time for construction than CRISPR/Cas9.

Thus, the development of the TALEN and CRISPR/Cas9 systems is an important step in the development of modern genomic engineering. The emergence of these systems, due to their low cost and ease of design, has become a powerful impetus for the development of both fundamental and applied science. More precisely, directional editing of plant genomes can be used to solve both; the study of gene functions and obtain plants with new properties, such as resistance to pathogens, herbicides, metabolism changes, yield indicators, etc. [94].

5.2 Maize proteomics opens the way for an insight into the biology of cereal crop

In the natural conditions of growth or cultivation of a species, plants in the process of their growth and development are often affected by adverse environmental factors [95]. Under the influence of unfavorable conditions, the decrease in physiological processes and functions can reach critical levels that do not ensure the implementation of the genetic program of ontogenesis, energy metabolism, regulatory systems, protein metabolism, and other vital functions of the plant organism are disrupted [96].

The main feature of protein research at the end of the twentieth century is proteomics. Since proteomics complements the research of genomics, transcriptomics, and metabolomics, it plays a central role in systems biology. Over the past three decades, significant progress has been made in the proteomic studies of maize as a model object [97]. Maize proteomic studies can be divided into two categories [98]:

  1. Profiling (or mapping) of the identified proteins of biological material, with the aim of separating, identifying, and cataloging as many proteins as possible, and, thus, the most complete scanning of the expressed genome sequences in individual representatives, at certain phases of development.

  2. Functional (cell-mapped) proteomics—studies polymorphism between different protein populations. With the help of two-dimensional electrophoresis, a comparative analysis of protein extracts of control and experimental plants is carried out. Both types of analysis became more real and informative after the sequencing of the reference genotype of maize B73 was completed [10, 98].

To date, published maize proteomic studies have used major proteomic technologies such as SDS-PAGE and two-dimensional electrophoresis (2-DE), laser capture microdissection, a combination of 2-DE with time-of-flight mass spectrometry (MALDI TOF), gas chromatography-mass spectrometry technologies [99, 100, 101, 102, 103, 104]. The main proteomic studies served to study changes in the composition of proteins under the influence of biotic and abiotic factors. For example, the effect of salicylic acid under high-temperature stress on the growth of seedlings and the antioxidant defense system of corn was studied [105]. In addition, the effect of some phytohormones, such as salicylic acid (SA), abscisic acid (ABA), jasmonic acid (JA), and methyl jasmonate (MeJA), on the protein composition of corn roots and leaves has been studied, and their important role in plant defenses has been proven [7, 98, 101, 105, 106].

However, despite the potential role of proteomics in advancing the study of stress tolerance in plants (also in the model), little useful information has been obtained so far for crop improvement and breeding [99].

5.3 Maize is a paragon for investigation of epigenetic studies

Epigenetic gene regulation is essential for the proper development of organisms. Epigenetic changes such as DNA methylation, histone modification, and RNA processing influence gene expression without changing the DNA sequence. Epigenetic studies have been the focus of many questions in plant research over the last decade [107]. The maize genome is relatively large and complex that includes abundant repetitive sequences, which are regularly silenced by epigenetic changes, making it an ideal organism to study epigenetic gene regulation. The application of new technologies to characterize maize epigenomes allows an understanding of the relationship between epigenetic mechanisms and genome organization [108].

Initial examples of epigenetic regulation were related to the transposable elements, starting with McClintock’s early work in the 1950s [109]. Implementation of advanced technologies to describe maize epigenomes allows a more clear understanding of the association between epigenetic mechanisms and genome organization. In maize, the genome-wide analysis of cytosine methylation was carried out using the combination of high-throughput DNA sequencing with the enzymatic characterization of methylated bases through bisulfite conversion. In recent years, numerous genome-wide studies of cytosine methylation have been published in maize [30, 80, 110, 111, 112, 113, 114, 115].

Eichten et al. (2011) have studied heritable epigenetic variation among maize inbred lines [111]. The comparison analysis of the DNA methylation degree of B73 and Mo17 maize lines permitted determining of about 700 differentially methylated regions (DMRs). Some DMRs occur in genomic regions that are apparently identical by descent in B73 and Mo17 suggesting that they may be examples of pure epigenetic variation. The results of this study showed the naturally occurring epigenetic variation in maize, including a pure epigenetic variation that is not conditioned by genetic differences. The identified epigenetic variation may provide complex trait variation.

Regulski et al. (2013) present the genome-wide map of cytosine methylation for two maize inbred lines, B73 and Mo17 [116]. Results showed that CpG (65%) and CpHpG (50%) islands (where H = A, C, or T) are highest methylated in transposons while CpHpH methylated is likely guided by 24-nucleotide (nt), but not 21-nt, small interfering RNAs (siRNAs). Scientists concluded that CpG methylation in exons (8%) may deter insertion of Mutator transposon insertion, while CpHpG methylation at splice acceptor sites may inhibit RNA splicing. The methylation map developed in this study will be an invaluable resource for maize epigenetic studies.

West et al. (2014) have studied the genomic distribution of H3K9me2 and DNA methylation in a maize genome [117]. They have investigated H3K9me2 in seedling tissue for the maize inbred B73 and compared to patterns of these modifications observed in Arabidopsis thaliana. This study gives a clear view of the relationship between DNA methylation and H3K9me2 in the maize genome and how the distribution of these modifications is shaped by the interplay of genes and transposons.

Kravets and Sokolova (2020) have studied the relationship between epigenetic variability with different individual radiosensitivity and adaptive capacity [118]. The researchers found significant differences in chromosomal aberration yield and DNA methylation profile under control and UV-C exposure for seedlings of subpopulations that differed in germination time. These significant differences in the control seedlings of different germination terms show the effect of the DNA methylation profile on DNA damage by regular metabolic factors including reactive oxygen species or thermal vibrations. The results showed the importance of epigenetic factors in identifying the radio-resistance and adaptive capacity of organisms.

Han et al. (2021) have reported epigenetic links to inbreeding depression in maize [119]. Throughout the subsequent inbreeding between inbred lines, thousands of genomic regions across TPC (teosinte branched1/cycloidea/proliferating cell factor)-binding sites (TBS) are hypermethylated across the H3K9me2-mediated pathway. Thus, several hundred TCP-target genes attended in mitochondrion, chloroplast, and ribosome functions are down-regulated, causing decreased growth vigor. On the contrary, random mating can reverse corresponding hypermethylation sites and TCP-target gene expression, restoring growth vigor.

A sufficiently large and highly repetitive maize genome provides an excellent model for other crop genomes to study gene regulation.

Advertisement

6. Conclusion and future prospect

More recently, plants have not played an important role in various genetic research because of their large and complex plant genomes. The role of model objects in understanding the patterns of historical and individual development of organisms is exceptionally great. The choice of an object for experimental scientific research, as a rule, becomes a separate task that requires special attention. It is necessary to clearly understand the criteria that the object of study must meet in order not only to solve the scientific problem, but also in its own way to facilitate the direct setting of the experiment. Plant genetics, physiology, and biochemistry have developed along this path, and the formation of modern sections of the biology of individual development is proceeding along this path.

Gradual studies carried out on model objects such as Arabidopsis, tobacco, and rice, including corn, proved that plants could also play a key role in molecular genetic experiments. Currently, genetic, chromosomal, genomic, and cytoplasmic modifications have been identified in maize; in particular, gene mutations have been best studied. To date, Zea mays L. is widely used in scientific research of the plant world. Day after day, it becomes a real classical model in plant biology; it has unconditional advantages in solving many current issues of genetics and individual development of plants, including cereals. Nevertheless, world science is moving forward and posing tasks that are ever more complex for researchers, for which corn alone is not enough.

References

  1. 1. Higgins R, Thurber C, Assaranurak I, Brown P. Multiparental mapping of plant height and flowering time QTL in partially isogenic sorghum families. G3: Genes, Genomes. Genetics. 2014;4:1593-1602
  2. 2. Horsfall JG. The fire brigade stops a raging corn epidemic. In: Hayes J, editor. The 1975 Yearbook of a Agriculture: That We May Eat. Washington DC: US Gov; 1975. pp. 105-114
  3. 3. Fukunaga K, Hill J, Vigouroux Y, Matsuoka Y, Sanchez GJ, Liu K, et al. Genetic diversity and population structure of teosinte. Genetics. 2005;169(4):2241-2254
  4. 4. Bennetzen JL. Maize genome structure and evolution. In: Bennetzen JL, Hake S, editors. Handbook of Maize: Genetics and Genomics. Berlin, Germany: Springer; 2009. pp. 179-199
  5. 5. Liu Y, Wang L, Sun C, Zhang Z, Zheng Y, Qiu F. Genetic analysis and major QTL detection for maize kernel size and weight in multi-environments. Theoretical and Applied Genetics. 2014;127(5):1019-1037
  6. 6. Randolph LF. Some effects of high temperature on polyploidy and other variations in maize. Proceedings of the National Academy of Science USA. 1932;18:222-229
  7. 7. Shahzad AN, Pitann B, Ali H, Qayyum MF, Fatima A, Bakhat HF. Maize genotypes differing in salt resistance vary in jasmonic acid accumulation during the first phase of salt stress. Journal of Agronomy and Crop Science. 2015;201:443-451
  8. 8. Liu J, Fernie AR, Yan J. The past, present, and future of maize improvement: Domestication, genomics, and functional genomic routes toward crop enhancement. Plant Communication. 2019;1(1):100010
  9. 9. McClintock B. The production of homozygous deficient tissues with mutant characteristics by means of the aberrant mitotic behavior of ring-shaped chromosomes. Genetics. 1938;23(4):315-376
  10. 10. Gaut BS, Doebley JF. DNA sequence evidence for the segmental allotetraploid origin of maize. Proceedings of the National Academy of Sciences. 1997;94(13):6809-6814. DOI: 10.1073/pnas.94.13.6809
  11. 11. Liller CB, Walla A, Boer MP, et al. Fine mapping of a major QTL for awn length in barley using a multiparent mapping population. Theoretical and Applied Genetics. 2017;130:269-281
  12. 12. Brandolini A. Razze europee di mais. Maydica. 1970;15:5-27
  13. 13. Harpstead DD. Man-molded cereal: Hybrid corn’s story. In: Hayes J, editor. The 1975 Yearbook of Agriculture: That We May Eat. Washington DC: US Gov; 1975. pp. 213-224
  14. 14. Ranere, A. J.; Piperno, D. R.; Holst, I.; Dickau, R.; Iriarte, J. (2009). The cultural and chronological context of early Holocene maize and squash domestication in the Central Balsas River Valley, Mexico. Proceedings of the National Academy of Sciences 106(13):5014-5018; Anthony Ranere, Dolores Piperno et al. The cultural and chronological context of early Holocene maize and squash domestication in the Central Balsas River Valley, Mexcio. PNAS, March 24, 2009
  15. 15. Callaway E. Shrub genome reveals secrets of flower power. Nature. 2013
  16. 16. Gangurde SS, Kumar R, Pandey AK, Burow M, Laza HE, Nayak SN, et al. Climate-smart groundnuts for achieving high productivity and improved quality: current status, challenges, and opportunities. In: Kole C, editor. Genomic Designing of Climate-Smart Oilseed Crops. Cham: Springer Nature Switzerland AG; 2019. pp. 133-172
  17. 17. Li L, Briskine R, Schaefer R, Schnable P, Myers C, Flagel L, et al. Co-expression network analysis of duplicate genes in maize (Zea mays L.) reveals no subgenome bias. BMC Genomics. 2016;17:875
  18. 18. Heslop-Harrison JS. Polyploidy. In: Maloy S, Hughes K, editors. Brenner’s Encyclopedia of Genetics (Second Edition). San Diego, US: Academic Press; 2013. pp. 402-403. DOI: 10.1016/B978-0-12-374984-0.01192-X
  19. 19. Niazi IAK, Rauf S, Teixeira da Silva JA, Iqbal Z, Munir H. Induced polyploidy in inter-subspecific maize hybrids to reduce heterosis breakdown and restore reproductive fertility. Grass and Forage Science. 2015;70(4):682-694
  20. 20. Iqbal MZ, Cheng M, Zhao Y, Wen X, Zhang P, Zhang L, et al. Mysterious meiotic behavior of autopolyploid and allopolyploid maize. Comparative Cytogenetics. 2018;12(2):247-265
  21. 21. Kushanov FN, Turaev OS, Ernazarova DK, Gapparov BM, Oripova BB, Kudratova MK, et al. Genetic diversity, QTL mapping, and marker-assisted selection technology in cotton (Gossypium spp.). Frontiers in Plant Science. 2021;12:779386
  22. 22. Wu X, Chen F, Zhao X, Pang C, Shi R, Liu C, et al. QTL mapping and GWAS reveal the genetic mechanism controlling soluble solids content in Brassica napus shoots. Food. 2021;10:2400
  23. 23. Guo J, Li C, Zhang X, Li Y, Zhang D, Shi Y, et al. Transcriptome and GWAS analyses reveal candidate gene for seminal root length of maize seedlings under drought stress. Plant Science. 2020;292:110380
  24. 24. Edwards D, Batley J, Snowdon RJ. Accessing complex crop genomes with next-generation sequencing. Theoretical and Applied Genetics. 2013;126:1-11. DOI: 10.1007/s00122-012-1964-x
  25. 25. Xiao Y, Liu H, Wu L, Warburton M, Yan J. Genome-wide association studies in maize: Praise and stargaze. Molecular Plant. 2017;10(3):359-374
  26. 26. Luo M, Zhang Y, Li J, et al. Molecular dissection of maize seedling salt tolerance using a genome-wide association analysis method. Plant Biotechnology Journal. 2021;19(10):1937-1951. DOI: 10.1111/pbi.13607
  27. 27. Xie Y, Feng Y, Chen Q, Zhao F, Zhou S, Ding Y, et al. Genome-wide association analysis of salt tolerance QTLs with SNP markers in maize (Zea mays L.). Genes Genomics. 2019;41(10):1135-1145
  28. 28. Yuan J, Wang X, Zhao Y, et al. Genetic basis and identification of candidate genes for salt tolerance in rice by GWAS. Scientific Reports. 2020;10:9958
  29. 29. Liu M, Tan X, Yang Y, Liu P, Zhang X, Zhang Y, et al. Analysis of the genetic architecture of maize kernel size traits by combined linkage and association mapping. Plant Biotechnology Journal. 2020;18:207-221
  30. 30. Wang X, Wang H, Liu S, et al. Genetic variation in ZmVPP1 contributes to drought tolerance in maize seedlings. Nature Genetics. 2016;48:1233-1241
  31. 31. Yuan Y, Cairns JE, Babu R, Gowda M, Makumbi D, Magorokosho C, et al. Genome-wide association mapping and genomic prediction analyses reveal the genetic architecture of grain yield and flowering time under drought and heat stress conditions in maize. Frontiers in Plant Science. 2019;9:1919
  32. 32. Dai LQ, Wu L, Dong QS, Yan G, Qu J, Wang PW. Genome-wide association analysis of maize kernel length. Journal of Northwest A&F University (Natural Science). 2018;46:20-28
  33. 33. Liu S, Wang X, Wang H, Xin H, Yang X, Yan J, et al. Genome-wide analysis of ZmDREB genes and their association with natural variation in drought tolerance at seedling stage of Zea mays L. PLoS Genetics. 2013;9:e1003790
  34. 34. Liu L, Du Y, Shen X, Li M, Sun W, Huang J, et al. KRN4 controls quantitative variation in maize kernel row number. PLoS Genetics. 2015;11:e1005670
  35. 35. Zhang S, Thakare D, Yadegari R. Laser-capture microdissection of maize kernel compartments for RNA-Seq-based expression analysis. Methods in Molecular Biology. 2018;1676:153-163
  36. 36. Zheng Y, Yuan F, Huang Y, et al. Genome-wide association studies of grain quality traits in maize. Scientific Reports. 2021;11:9797
  37. 37. Liu Y, Yi Q, Hou X, et al. Identification of quantitative trait loci for kernel-related traits and the heterosis for these traits in maize (Zea mays L.). Molecular Genetics and Genomics. 2020;295:121-133
  38. 38. Wang G, Zhao Y, Mao W, Ma X, Su C. QTL analysis and fine mapping of a major QTL conferring kernel size in maize (Zea mays). Frontiers in Genetics. 2020;11:603920
  39. 39. Pan L, Wang N, Wu Z, Guo R, Yu X, Zheng Y, et al. A high density genetic map derived from RAD sequencing and its application in QTL analysis of yield-related traits in Vigna unguiculata. Frontiers in Plant Science. 2017;8:1544
  40. 40. Li M, Guo X, Zhang M, Wang X, Zhang G, Tian Y, et al. Mapping QTLs for grain yield and yield components under high and low phosphorus treatments in maize (Zea mays L.). 2010;178(5):462
  41. 41. Jiang GL. Molecular markers and marker-assisted breeding in plants. In: Andersen SB, editor. Plant Breeding from Laboratories to Fields. Rijeka, Croatia: IntechOpen; 2013. pp. 45-83
  42. 42. Jiang Q, Tang D, Hu C, Qu J, Liu J. Combining meta-QTL with RNA-seq data to identify candidate genes of kernel row number trait in maize. Maydica. 2016;61:1-9
  43. 43. Yu J, Holland JB, McMullen MD, Buckler ES. Genetic design and statistical power of nested association mapping in maize. Genetics. 2008;78(1):539-551
  44. 44. Buckler ES, Holland JB, Bradbury PJ, Acharya CB, Brown PJ, Browne C, et al. The genetic architecture of maize flowering time. Science. 2009;325(5941):714-718
  45. 45. McMullen MD, Kresovich S, Villeda HS, Bradbury P, Li H, Sun Q, et al. Genetic properties of the maize nested association mapping population. Science. 2009;325(737):737-740
  46. 46. Bauer E et al. Intraspecific variation of recombination rate in maize. Genome Biology. 2013;14:R103
  47. 47. Li H, Bowling AJ, Gandra P, et al. Systemic RNAi in western corn rootworm, Diabrotica virgifera virgifera, does not involve transitive pathways. Insect Science. 2018;25(1):45-56. DOI: 10.1111/1744-7917.12382
  48. 48. Chen Q et al. TeoNAM: A nested association mapping population for domestication and agronomic trait analysis in maize. Genetics. 2019;213:1065-1078
  49. 49. Bouchet S, Olatoye MO, Marla SR, Perumal R, Tesso T, Yu J, et al. Increased power to dissect adaptive traits in global sorghum diversity using a nested association mapping population. Genetics. 2017;206:573-585
  50. 50. Jordan DR, Mace ES, Cruickshank AW, Hunt CH, Henzell RG. Exploring and exploiting genetic variation from unadapted sorghum germplasm in a breeding program. Crop Science. 2011;51:1444-1457
  51. 51. Mace ES, Hunt CH, Jordan DR. Supermodels: Sorghum and maize provide mutual insight into the genetics of flowering time. Theoretical and Applied Genetics. 2013 May;126(5):1377-1395
  52. 52. Gangurde SS, Wang H, Yaduru S, Pandey MK, Fountain JC, Chu Y, et al. Nested-association mapping (NAM)-based genetic dissection uncovers candidate genes for seed and pod weights in peanut (Arachis hypogaea). Plant Biotechnology Journal. 2020 Jun;18(6):1457-1471. DOI: 10.1111/pbi.13311. [Epub Dec 25, 2019]. PMID: 31808273; PMCID: PMC7206994
  53. 53. Gaut BS, Le Thierry d’Ennequin M, Peek AS, Sawkins MC. Maize as a model for the evolution of plant nuclear genomes. Proceedings of the National Academy of Sciences. 2000;97(13):7008-7015. DOI: 10.1073/pnas.97.13.7008
  54. 54. Hemshrot A, Poets AM, Tyagi P, Lei L, Carter CK, Hirsch CN, et al. Development of a multiparent population for genetic mapping and allele discovery in six-row barley. Genetics. 2019;213:595-613
  55. 55. Maurer A, Draba V, Jiang Y, et al. Modelling the genetic architecture of flowering time control in barley through nested association mapping. BMC Genomics. 2015;16:290
  56. 56. Nice LM, Steffenson BJ, Brown-Guedira GL, Akhunov ED, Liu C, Kono TJY, et al. Development and genetic characterization of an advanced backcross-nested association mapping (AB-NAM) population of wild × cultivated barley. Genetics. 2016;203:1453-1467
  57. 57. Schnaithmann F, Kopahnke D, Pillen K. A first step toward the development of a barley NAM population and its utilization to detect QTLs conferring leaf rust seedling resistance. Theoretical and Applied Genetics. 2014;127:1513-1525
  58. 58. Schmutzer T, Samans B, Dyrszka E, et al. Species-wide genome sequence and nucleotide polymorphisms from the model allopolyploid plant Brassica napus. Sci Data. 2015;2:150072
  59. 59. Bajgain P, Rouse MN, Tsilo TJ, Macharia GK, Bhavani S, Jin Y, et al. Nested association mapping of stem rust resistance in wheat using genotyping by sequencing. PLoS One. 2016;11:1-22
  60. 60. Chidzanga C, Fleury D, Baumann U, Mullan D, Watanabe S, Kalambettu P, et al. Development of an Australian Bread Wheat Nested Association Mapping Population, a new genetic diversity resource for breeding under dry and hot climates. International Journal of Molecular Sciences. 2021;22:4348
  61. 61. Jordan KW, Wang S, He F, Chao S, Lun Y, Paux E, et al. The genetic architecture of genome-wide recombination rate variation in allopolyploid wheat revealed by nested association mapping. The Plant Journal. 2018 Sep;95(6):1039-1054
  62. 62. Kidane YG, Gesesse CA, Hailemariam BN, Desta EA, Mengistu DK, Fadda C, et al. A large nested association mapping population for breeding and quantitative trait locus mapping in Ethiopian durum wheat. Plant Biotechnology Journal. 2019;17:1380-1393
  63. 63. Wingen LU, West C, Leverington-Waite M, Collier S, Orford S, Goram R, et al. Wheat landrace genome diversity. Genetics. 2017;205(4):1657-1676. DOI: 10.1534/genetics.116.194688
  64. 64. Christopher AF, Moreno M, Wang Z, Heffelfinger C, Arbelaez LJ, Aguirre JA, et al. Genetic architecture of a rice nested association mapping population. G3 Genes|Genomes|Genetics. 2017;7(6):1913-1926
  65. 65. Kitony JK, Sunohara H, Tasaki M, Mori J-I, Shimazu A, Reyes VP, et al. Development of an Aus-derived nested association mapping (Aus-NAM) population in rice. Plants. 2021;10:1255
  66. 66. Song J, Lu D, Niu Y, Sun H, Zhang P, Dong W, et al. Label-free quantitative proteomics of maize roots from different root zones provides insight into proteins associated with enhance water uptake. BMC Genomics. 2022;23(1):184
  67. 67. Xavier A, Xu S, Muir WM, Rainey KM. NAM: Association studies in multiple populations. Bioinformatics. 2015;31:3862-3864
  68. 68. Abdurakhmonov I, Abdullaev A, Buriev Z, Shermatov S, Kushanov F, Makamov A, et al. Cotton germplasm collection of Uzbekistan. In: Abdurakhmonov I, editor. World Cotton Germplasm Resources. London: IntechOpen; 2014
  69. 69. Turaev O, Kushanov F, Makamov A, Darmonov M, Husenov N, Rakhmanov B. Statistical analysis for stability of fiber quality traits of cotton NAM founders. In: Proceedings of the III Tashkent International Innovation Forum. 2017. pp. 176-182
  70. 70. Li H, Bradbury P, Ersoz E, Buckler ES, Wang J. Joint QTL linkage mapping for multiple-cross mating design sharing one common parent. PLoS One. 2011;6:e0017573
  71. 71. Brock MT, Rubin MJ, DellaPenna D, Weinig C. A nested association mapping panel in Arabidopsis thaliana for mapping and characterizing genetic architecture. G3 Genes|Genomes|Genetics. 2020;10(10):3701-3708
  72. 72. Scott MF, Ladejobi O, Amer S, et al. Multi-parent populations in crops: A toolbox integrating genomics and genetic mapping with breeding. Heredity. 2020;125:396-416
  73. 73. Bevan H, Klara V, Arunas V, Chitra R, Vikas S, Pooran G, et al. MAGIC populations in crops: Current status and future prospects. Theoretical and Applied Genetics. 2015;128(6):999-1017. DOI: 10.1007/s00122-015-2506-0
  74. 74. Mott R, Talbot CJ, Turri MG, Collins AC, Flint J. A method for fine mapping quantitative trait loci in outbred animal stocks. Proceedings of the National Academy of Sciences. 2000;97(23):12649-12654
  75. 75. Cavanagh C, Morell M, Mackay I, Powell W. From mutations to MAGIC: Resources for gene discovery, validation and delivery in crop plants. Current Opinion in Plant Biology. 2008;11:215-221
  76. 76. Mackay IJ, Powell W. The significance and relevance of linkage disequilibrium and association mapping in crops. Trends in Plant Science. 2007;12:53
  77. 77. Kover PX, Valdar W, Trakalo J, Scarcelli N, Ehrenreich IM, et al. A multiparent advanced generation inter-cross to fine-map quantitative traits in Arabidopsis thaliana. PLoS Genetics. 2009;5(7):e1000551
  78. 78. Anderson SL, Mahan AL, Murray SC, Klein PE. Four Parent Maize (FPM) population: Effects of mating designs on linkage disequilibrium and mapping quantitative traits. The Plant Genome. 2018;11(2). DOI: 10.3835/plantgenome2017.11.0102
  79. 79. Dell’Acqua M, Gatti DM, Pea G, et al. Genetic properties of the MAGIC maize population: A new platform for high definition QTL mapping in Zea mays. Genome Biology. 2015;16:167
  80. 80. Mahan AL, Murray SC, Klein PE. Four-parent maize (FPM) population: Development and phenotypic characterization. Crop Science. 2018;58:1106-1117
  81. 81. Abdurakhmonov IY, Ayubov MS, Ubaydullaeva KA, Buriev ZT, Shermatov SE, Ruziboev HS, et al. RNA interference for functional genomics and improvement of cotton (Gossypium sp.). Frontiers in Plant Science. 2016;22(7):202
  82. 82. Dalakouras A, Wassenegger M, Dadami E, Ganopoulos I, Pappas ML, Papadopoulou K. Genetically modified organism-free RNA interference: Exogenous application of RNA molecules in plants. Plant Physiology. 2020;182(1):38-50
  83. 83. Segal G, Song R, Messing J. A new opaque variant of maize by a single dominant RNA-interference-inducing transgene. Genetics. 2003;165:387-397
  84. 84. Casati P, Stapleton AE, Blum JE, Walbot V. Genome-wide analysis of high-altitude maize and gene knockdown stocks implicates chromatin remodeling proteins in response to UV-B. The Plant Journal. 2006 May;46(4):613-627
  85. 85. Casati P, Walbot V. Maize lines expressing RNAi to chromatin remodeling factors are similarly hypersensitive to UV-B radiation but exhibit distinct transcriptome responses. Epigenetics. 2008;3(4):216-229
  86. 86. Thakare D, Zhang J, Wing RA, Cotty PJ, Schmidt MA. Aflatoxin-free transgenic maize using host-induced gene silencing. Science Advances. 2017;3(3):e1602382
  87. 87. Vélez AM, Fishilevich E, Rangasamy M, Khajuria C, McCaskill DG, Pereira AE, et al. Control of western corn rootworm via RNAi traits in maize: Lethal and sublethal effects of Sec23 dsRNA. Pest Management Science. Apr 2020;76(4):1500-1512. DOI: 10.1002/ps.5666
  88. 88. Metje-Sprink J, Menz J, Modrzejewski D, Sprink T. DNA-free genome editing: Past, present and future. Frontiers in Plant Science. 2019;9:1957
  89. 89. Nayak SN, Aravind B, Malavalli SS, Sukanth BS, Poornima R, Bharati P, et al. Omics technologies to enhance plant based functional foods: An overview. Frontiers in Genetics. 2021;8(12):742095
  90. 90. D’Halluin K, Ruiter R. Directed genome engineering for genome optimization. The International Journal of Developmental Biology. 2013;57:621-627
  91. 91. Shukla V, Doyon Y, Miller J, et al. Precise genome modification in the crop species Zea mays using zinc-finger nucleases. Nature. 2009;459:437-441
  92. 92. Liang Z, Zhang K, Chen KL, Gao CX. Targeted mutagenesis in Zea mays using TALENs and the CRISPR/Cas system. Journal of Genetics and Genomics. 2014;41(2):63-68
  93. 93. Char SN, Unger-Wallace E, Frame B, Briggs SA, Main M, Spalding MH, et al. Heritable site-specific mutagenesis using TALENs in maize. Plant Biotechnology Journal. 2015;13(7):1002-1010
  94. 94. Malzahn A, Lowder L, Qi Y. Plant genome editing with TALEN and CRISPR. Cell & Bioscience. 2017;7:21
  95. 95. Kubota C. Growth, Development, Transpiration and Translocation as Affected by Abiotic Environmental Factors. In: Kozai T, Niu G, Takagaki M, editors. Plant Factory: An Indoor Vertical Farming System for Efficient Quality Food Production. London, UK: Elsevier Inc.; 2015. pp. 151-164. DOI: 10.1016/B978-0-12-801775-3.00010-X
  96. 96. Goff SA. A unifying theory for general multigenic heterosis: Energy efficiency, protein metabolism, and implications for molecular breeding. The New Phytologist. 2011;189(4):923-937
  97. 97. Song Q et al. Genetic characterization of the soybean nested association mapping population. Plant Genome. 2017;10:1-14
  98. 98. Pechanova O, Takáč T, Samaj J, Pechan T. Maize proteomics: An insight into the biology of an important cereal crop. Proteomics. 2013;13(3-4):637-662
  99. 99. Eldakak M, Milad SI, Nawar AI, Rohila JS. Proteomics: A biotechnology tool for crop improvement. Frontiers in Plant Science. 2013;4:35
  100. 100. Flores I, Cabra V, Quirasco MC, Farres A, Galvez A. Emulsifying properties of chemically deamidated corn (Zea mays) gluten meal. Food Science and Technology International. 2010;16(3):241-250
  101. 101. Hochholdinger F, Marcon C, Baldauf JA, Yu P, Frey FP. Proteomics of maize root development. Frontiers in Plant Science. 2018;9:143
  102. 102. Usuda H, Shimogawara K. Phosphate deficiency in maize. VI. Changes in the two-dimensional electrophoretic patterns of soluble proteins from second leaf blades associated with induced senescence. Plant & Cell Physiology. 1995;36:1149-1155
  103. 103. Venkatesh TV, Chassy AW, Fiehn O, Flint-Garcia S, Zeng Q, Skogerson K, et al. Metabolomic assessment of key maize resources: GC-MS and NMR profiling of grain from B73 hybrids of the Nested Association Mapping (NAM) Founders and of geographically diverse landraces. Journal of Agricultural and Food Chemistry. 2016;64(10):2162-2172
  104. 104. Zhang X, Zhang R, Li L, Yang Y, Ding Y, Guan H, et al. Negligible transcriptome and metabolome alterations in RNAi insecticidal maize against Monolepta hieroglyphica. Plant Cell Reports. 2020;39(11):1539-1547
  105. 105. Khanna P, Kaur K, Gupta AK. Salicylic acid induces differential antioxidant response in spring maize under high temperature stress. Indian Journal of Experimental Biology. 2016;54(6):386-393
  106. 106. Wu LJ, Zu XF, Wang XT, Sun AG, Zhang J, Wang SX, et al. Comparative proteomic analysis of the effects of salicylic acid and abscisic acid on maize (Zea mays L.) leaves. Plant Molecular Biology Reporter. 2013;31:507-516
  107. 107. Mladenov V, Fotopoulos V, Kaiserli E, Karalija E, Maury S, Baranek M, et al. Deciphering the epigenetic alphabet involved in transgenerational stress memory in crops. International Journal of Molecular Sciences. 2021;22(13):7118
  108. 108. Huang J, Lynn JS, Schulte L, Vendramin S, McGinnis K. Epigenetic control of gene expression in maize. International Review of Cell and Molecular Biology. 2017;328:25-48
  109. 109. McClintock B. The origin and behavior of mutable loci in maize. Proceedings of the National Academy of Sciences of the United States of America. 1950;36(6):344-355
  110. 110. Ding H, Gao J, Qin C, et al. The dynamics of DNA methylation in maize roots under Pb stress. International Journal of Molecular Sciences. 2014;15(12):23537-23554
  111. 111. Eichten SR, Swanson-Wagner RA, Schnable JC, et al. Heritable epigenetic variation among maize inbreds. PLoS Genetics. 2011;7(11):e1002372
  112. 112. Forestan C, Farinati S, Aiese Cigliano R, et al. Maize RNA PolIV affects the expression of genes with nearby TE insertions and has a genome-wide repressive impact on transcription. BMC Plant Biology. 2017;17:161
  113. 113. Gent JI, Madzima TF, Bader R, Kent MR, Zhang X, Stam M, et al. Accessible DNA and relative depletion of H3K9me2 at maize loci undergoing RNA-directed DNA methylation. The Plant Cell. 2014;26(12):4903-4917
  114. 114. Li Q, Suzuki M, Wendt J, Patterson N, Eichten S, Hermanson P, et al. Post-conversion targeted capture of modified cytosines in mammalian and plant genomes. Nucleic Acids Research. 2015;43:e81
  115. 115. Wang QX, Xie WB, Xing KJ, Yan J, Meng XJ, Li XL, et al. Genetic architecture of natural variation in rice chlorophyll content revealed by a genome-wide association study. Molecular Plant. 2015;8:946-957
  116. 116. Regulski M, Lu Z, Kendall J, et al. The maize methylome influences mRNA splice sites and reveals widespread paramutation-like switches guided by small RNA. Genome Research. 2013;23(10):1651-1662
  117. 117. West PT, Li Q, Ji L, Eichten SR, Song J, Vaughn MW, et al. Genomic distribution of H3K9me2 and DNA methylation in a maize genome. PLoS One. 2014;9(8):e105267
  118. 118. Kravets O, Sokolova DO. Epigenetic factors of individual radiosensitivity and adaptive capacity. International Journal of Radiation Biology. 2020;96:1-29
  119. 119. Han T, Wang F, Song Q, Ye W, Liu T, Wang L, et al. An epigenetic basis of inbreeding depression in maize. Science Advances. 2021;7(35):eabg5442

Written By

Fakhriddin N. Kushanov, Ozod S. Turaev, Oybek A. Muhammadiyev, Ramziddin F. Umarov, Nargiza M. Rakhimova and Noilabonu N. Mamadaliyeva

Submitted: 17 October 2021 Reviewed: 24 March 2022 Published: 23 June 2022