Next Article in Journal
Disinfection Efficacy of Tobamovirus-Contaminated Soil in Greenhouse-Grown Crops
Next Article in Special Issue
Cell Division Controls Final Fruit Size in Three Apple (Malus x domestica) Cultivars
Previous Article in Journal
Seaweed Extracts as Substitutes of Synthetic Hormones for Rooting Promotion in Rose Cuttings
Previous Article in Special Issue
Comprehensive Analysis of N6-Methyladenosine Regulatory Genes from Citrus grandis and Expression Profilings in the Fruits of “Huajuhong” (C. grandis “Tomentosa”) during Various Development Stages
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in DNA Methylation and Their Potential Breeding Applications in Plants

1
College of Life Sciences & Technology, Huazhong Agricultural University, Wuhan 430070, China
2
Department of Biotechnology, Faculty of Crop Production, Sindh Agriculture University, Tandojam 70060, Pakistan
3
Department of Horticulture, Faculty of Crop Production, Sindh Agriculture University, Tandojam 70060, Pakistan
4
Department of Crop Physiology, Faculty of Crop Production, Sindh Agriculture University, Tandojam 70060, Pakistan
5
State Key Laboratory of Crop Stress Biology for Arid Areas, College of Horticulture, Northwest A&F University, Xianyang 712100, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Horticulturae 2022, 8(7), 562; https://doi.org/10.3390/horticulturae8070562
Submission received: 13 May 2022 / Revised: 12 June 2022 / Accepted: 19 June 2022 / Published: 21 June 2022

Abstract

:
Traditional plant breeding encompasses repetitive crossing and selection based on morphological traits, while phenotypic selection has been complemented by molecular methods in recent decades. Genome editing with techniques like the CRISPR-Cas9 system is still a novel approach that is being used to make direct modifications to nucleotide sequences of crops. In addition to these genetic alterations, an improved understanding of epigenetic variations such as DNA methylation on the phenotype of plants has led to increased opportunities to accelerate crop improvement. DNA methylation is the most widely studied epigenetic mark in plants and other eukaryotes. These epigenetic marks are highly conserved and involved in altering the activities and functions of developmental signals by catalyzing changes in the chromatin structure through methylation and demethylation. Cytosine methylation (5mC) is the most prevalent modification found in DNA. However, recent identification of N6-methyladenosine (6mA) in plants starts to reveal their critical role in plant development. Epigenetic modifications are actively involved in creating the phenotype by controlling essential biological mechanisms. Epigenetic modifications could be heritable and metastable causing variation in epigenetic status between or within species. However, both genetic and heritable epigenetic variation has the potential to drive natural variation. Hence, epigenome editing might help overcome some of the shortcomings of genome editing (such as gene knockout), which can have significant off-target effects and only enables the loss of a gene’s function. In this review, we have discussed the mechanism underlying DNA methylation and demethylation in plants. Methyltransferases and demethylases are involved in catalyzing specific types of modification. We also discuss the potential role of DNA modifications in crop improvement for meeting the requirements of sustainable and green agriculture.

1. Introduction

Epigenetics is one of the most fascinating and topical fields of genetics, adding to classical knowledge about how genes interact with phenotypes, which has perplexed scientists for decades [1]. The concept of epigenetics was proposed in the mid-twentieth century by Waddington, by integrating epigenesis and genetics to elucidate the phenotypic traits of plants as a result of the causal interaction between genes and their products [2]. However, our present understanding of molecular biology has led us to a narrower definition that comprises the study of molecular processes in and around DNA that regulate genome-related activity and phenotype, regardless of DNA sequence. It can be inherited via mitosis or meiosis, and several studies have shown that stress-induced epimutations are successfully passed down to the next generation [3,4,5,6,7]. Epigenetics, as defined by Arthur Riggs and his colleagues, is the study of mitotically and meiotically inherited changes in gene function that cannot be explained by changes in DNA sequence [8]. Epigenetics has had considerable success due to its applications in plant breeding, which has been used to examine the transmission of epigenetic marks through generations to enhance desirable traits in crops [9]. Epigenetics has the potential to be employed as a crop improvement tool. Epigenetic modifications, including DNA methylation, histone modifications, chromatin remodeling, and activity of small RNAs (sRNAs), are thus inherited but do not follow the known inheritance patterns [1].
Because of its heritability and potential to influence plant phenotypes, DNA methylation seems to be a viable source of variation contributing to crop fitness and production. Variations in DNA methylation impact agronomically important traits including seed dormancy, flowering time, and yield, and the partial heritability of DNA methylation patterns reveals epigenetics had a role in plant domestication and evolution [10,11,12]. Understanding DNA methylation modifications would enable breeders to concurrently create favorable variations while limiting undesirable epigenetic changes that can be caused by breeding practices such as tissue culture [13,14,15,16,17]. Various approaches have previously been established for this purpose, and efficient induction of epigenetic modifications for plant breeding necessitates a technical grasp of the molecular mechanisms involved in both its introduction and its maintenance over successive generations.
DNA methylation being a conserved epigenetic mark regulates various cellular processes in eukaryotes including plants. Several mechanisms have been reported associated with this epigenetic mark such as genomic imprinting, X-chromosome inactivation, cell type differentiation during embryonic development, genome stability, chromatin architecture, transcription activation, embryogenesis, chromosome stability; it is associated with gene and transposon silencing [18,19,20,21,22,23]. The methylation patterns are heritable and have transgenerational effects [24]. However, the methylation marks are dynamic and show variations during reprogramming stages throughout the life cycle of an organism [25]. Moreover, methylation is significantly involved in stem cell pluripotency and cellular differentiation to distinct behaviour, memory and ageing [26,27,28]. At the population level, DNA methylation participates in the biodiversity of natural populations, disease susceptibility and response to environmental stimuli [29].
5mC has a significant role in regulating gene expression and TE silencing [30,31,32]. DNA methylation starts by stabilizing the DNA double helix, strongly affecting the binding of various proteins on DNA including regulatory proteins. Moreover, sometimes it prevents the binding of certain nuclear proteins which are vital for transcription and play significant roles in cellular pathways [33]. In contrast, some proteins arrange an entire ensemble on a methylated DNA sequence by specifically binding on it and regulate gene expression. Hence, DNA methylation can be defined as a factor of positive or negative control of transcription. The proliferation and accumulation of TE sequences are also influenced by epigenetic mechanisms. TEs are targeted for DNA methylation by small RNA (sRNA)–mediated mechanisms in several eukaryotic lineages [34]. DICER-LIKE RNase enzymes in plants generate 24-bp sRNA that directs ARGONAUTE and other downstream proteins to complementary DNA sequences, promoting and sustaining DNA and histone methylation [35,36]. In Arabidopsis, TE methylation has been found to silence transposition, as evidenced by drastically increased levels of TE transcription in met1 methylation mutants [37,38]. Silencing TEs near genes may also inhibit the generation of aberrant transcripts via read-through transcription beyond TE termini [39], in addition to preventing the proliferation of new TE sequences. However, methylated sequences can influence the expression of nearby genes, typically reducing expression [40]. Expression of flowering time gene FWA has been found to be correlated with the methylation status of nearby SINE-like TE. Hence, changes in gene expression caused by methylation of neighboring TEs may have a deleterious impact on gene and genome function.
Differences in cytosine DNA methylation patterns can contribute to phenotypic variability because they are inherited through both mitotic and meiotic cell divisions. High-throughput sequencing technologies have made it possible to generate a large amount of DNA sequence data. The underlying mechanisms and functions of DNA methylation have been discovered by integrated investigations of genome-wide gene expression patterns and DNA methylation patterns. There have also been associations discovered between DNA methylation and agronomic characteristics. The results could be relevant for future crop breeding applications of natural epigenomic variation. Artificial epigenome editing could also be a promising new plant breeding strategy for developing novel varieties with improved agronomic traits [41]. With its potential for disease resistance enhancement, DNA methylation presents a new direction to both scientists and breeders by providing a new source of variation [42]. Moreover, a recent study discovered that DNA methylation can cause epigenetic modifications in plants in response to abiotic stresses [43].
In this review, we explain the DNA 5mC methylation machinery, diversity and dynamics, molecular and biological functions in plants. We also discuss the newly identified N6-methyladenosine (6mA) dynamic distribution and responses to developmental signals in model plants such as Arabidopsis thaliana and rice (Oryza sativa). We consider DNA methyltransferases and demethylases that catalyze specific DNA modifications (5mC, 6mA), recent advances and understanding in DNA methylation and demethylation in genetic regulation and functions with emphasis on DNA methylation in plants. We also highlight knowledge gaps, and discuss challenges and opportunities for exploitation of DNA methylation in breeding applications.

2. DNA 5mC in Plants

Methylation marks in DNA show dynamic regulation of establishing, maintaining, and active removal of their activities. Among all the methylation modifications discovered to date, 5mC is the well-known and well-studied DNA mark and has been widely identified in plants and other eukaryotes. 5mC has been located in all sequence contexts such as CG, CHG and CHH (where H = A, T or C) [44]. De novo methylation on all sequence contexts is regulated by production of small interference RNAs (siRNA) comprising ~21–24 nucleotides [45]; it is an intricate process that is highly controlled by the RNA directed DNA methylation (RdDM) pathway. RdDM is a transcriptional gene silencing pathway that is specific to plants only, and is directed by siRNA molecules leading to DNA de novo methylation. The multistep pathway consists of polymerase IV-dependent siRNA biogenesis, polymerase V mediated de novo methylation and chromatin alteration [23,45,46]. It initiates with the production of siRNA, which consists of specific DNA dependent RNA polymerase Pol IV and V, the Pol IV interacting Sawadee homeodomain homolog 1 (SHH1), double-stranded RNA endonuclease DCL3, RDR2, and Argonaute proteins (AGO4 and AGO6) with the effector kow-domain containing transcription factor 1 (KTF1) [45,47,48,49].
The family of methyltransferases catalyze the establishment of 5mC by transferring a methyl group from S-adenosylmethionine (SAM) to the substrate (Figure 1) [28], while the maintenance of pre-existing methylation continues particularly on CG and CHG motifs, which are known as symmetrical sites as they are read in 5′ to 3′ direction. This symmetry after each round of replication provided a base for cytosine methylation on a mother strand and transfer to daughter strands. Following the DNA repair mechanism, methylation can take place, allowing newly synthesized strands to be methylated. Comparable to CHG and CHH methylation, CG methylation also takes place over gene bodies, while the CHG and CHH sequence context is fully involved in transcriptional gene silencing (TGS) and heterochromatin formation [50].

2.1. Distribution of 5mC in Plant Genome

So far, DNA methylation has been detected in a variety of plant species, including algae, cereal crops, vegetables and trees [51]. The importance of DNA methylation cannot only be described by its abundance, but is also dependent on its location on different parts of the gene and its positioning such as CH, CHG and CHH. A widespread diversity in the type, amount, and location of DNA methylation in a gene has been observed at different developmental stages in various plant species. In mammals, 5mC is mostly found on the CG sequence context, while in plants its presence in all sequence contexts has been observed. In Arabidopsis, gene bodies are mainly targeted by CG methylation; while CHG and CHH methylation mostly occur on TEs and repeats, CG methylation is also found on TEs. Genome-wide methylation in Arabidopsis shows that heterochromatin is a highly methylated region, which is enriched by transposons and other repetitive sequences [52].
Phylogenetic investigations of methylomes of more than 30 eukaryotes, a majority being plant species, have revealed the significant variation in CG and non-CG methylation among different species [51,53]. The lowest methylation levels, 5.4% CG, 2.6% CHG and 2.5% CHH were found in C. reinhardtii [51], while the highest levels were observed in Beta vulgaris, where 92.6% CG, 81.2% CHG and 18.9% CHH methylation were present. In Arabidopsis leaf, CG, CHG and CHH methylations were reported as 30.5%, 10.0% and 3.9%, respectively. Rice leaves displayed intermediate levels of CG, CHG and CHH methylation as 54.4%, 31.0 and 5.1% respectively. In the maize genome, 86% CG, 74% CHG and 5% CHH methylation have been reported [51,53]. Hence, methylation levels in all three sequence contexts vary significantly among different organisms.
Compared to CHG and CHH methylation, CG methylation has been found predominant. CG methylation in angiosperms contributes to more than 50% of total cytosine methylation [53]. Higher variation in CHG and CHH levels is found as compared to CG methylation, suggesting that DNA methylation patterns are diverse in various plant species. In plants, non-CG methylation plays vital role in exogenous DNA silencing via the RdDM pathway [21]. In Arabidopsis, during the early stage of plant development, higher invariable CG methylation was found on TEs and repeats while increasing levels of CHG and CHH methylation were observed. This non-CG methylation culminates in the mature embryo, where it reaches higher levels as compared to seedlings or adult plants [54].

2.2. Derivatives of 5mC

Apart from 5mC, the most common modification, several other types of changes on the same position have also been found to exist, for instance, 5-hydroxymethylcytosine (5hmC), 5-formylcytosine (5fC) and 5-carboxylcytosine (5caC); these modifications have also been considered the most critical marks [55]. Active demethylation through ten eleven translocation cytosine dioxygenases (TETs) oxidize 5mC to 5hmC and successive oxidation produces 5fC and 5caC, which are found in lower abundance in the genome and can be removed by BER proteins [56,57,58]. 5hmC is also regarded as the sixth nucleobase in DNA; its presence is reported in mammals and it is found to play key functions in gene regulation [59]. Recent studies revealed that 5hmC regulates various cellular and developmental pathways such as neuron development in mammals, tumorigenesis and embryonic stem cells pluripotency [55].
DME and ROS1 purified from E. coli have the ability to excise 35-mer oligonucleotides comprising of Cytosine, 5mC, 5hmC, 5fC and 5caC, suggesting that they can cleave both the 5mC and 5hmC in vitro. Previously, several attempts have been made for the detection of 5hmC in plants [59,60,61,62,63]. However, the Arabidopsis genome does not contain 5hmC [64]. Based on these studies, it can be concluded that either a lower abundance of 5hmC is present in plants or there is a distinct mechanism and independent evolution of DNA demethylation in plants as compared to animals. The presence of 5hmC in plants is still under debate. Recently, 5hmC has been detected in a plant genome. Wang et al. (2015) conducted a study using three rice cultivars and found low levels of 5hmC at 1.39 ± 0.16 and 2.17 ± 0.03 per million nucleotides in leaf and panicle, respectively. Levels of 5hmC varied significantly among different tissues. 5hmC sites were mainly enriched in heterochromatin regions and TEs, particularly around retrotransposons [59]. These findings will provide a base for studying, detecting and understanding 5hmC in plants and their functions in plant development and DNA demethylation.

2.3. Writers, Erasers and Readers of DNA Methylation

In plants, the DNA methylation mechanism is highly complicated and much richer as compared to animals [65]. These methylation modifications are enzymatically installed by: writers, that introduce various chemical modifications on DNA (i.e., methyltransferases); erasers, the unique set of enzymes capable of erasing these chemical tags (i.e., demethylases); and readers, the specialized domain comprising proteins that recognize and interpret such alterations (Table 1). Several DNA methyltransferases (DNMTs) have been reported in plants. However, some are limited to for specific organisms and possess a ubiquitin binding domain. Loss or knockdown of methyltransferases in plants is not as lethal as in animals [66]. In Arabidopsis, methyltransferase including Chromomethylase 1 (CMT1), CMT2 and CMT3, Domain rearrange DNA methylase 1 (DRM1) and DRM2, methyltransferase 1 (MET1) and MET2 have been identified as catalyzing DNA methylation [21]. Maintenance of cytosine methylation is carried out by MET1, a homolog of mammalian methyltransferase DNMT1, which recognizes hemimethylated CG sites and maintains symmetrical CG methylation along with different methylation proteins during DNA replication [67]. In contrast, CMT3, being a plant-specific methyltransferase, targets particular sequences and maintains methylation at CHG sites by interacting with KRYPTONINE (KYP), SUVH5 and SUVH6 [21,68]. Also, de novo DNA methylation levels in all sequence contexts are regulated by DRM2. DRM2 or CMT2 regulate the maintenance of CHH sequence contexts through different pathways [21,69,70]. These consist of nucleosomes remodelers DRD1 and DDM1, respectively [46,69]
Diverse and dynamic DNA methylation patterns in different plant species contain different sets of methyltransferases and demethylases [51,71]. For example, Arabidopsis possesses MET1; maize contains ZmMET1, peach has PsMET and rice OsMET1-1 and OSMET1-2 [72,73,74]. Also, Arabidopsis possesses CMT2 and CMT3 [75] and Brassica rapa BrCMT [76,77], while maize and rice contain ZMET2, ZMET5 and OsCMTL, OsMET2a respectively. Likewise, for the DRM family, Arabidopsis possesses DRM1 and DRM2 [78,79] and maize ZmDMT106 and Zmet3, while rice has OsDMT106 and OsZmet3 [80]. However, their biological functions are still unclear.
Demethylation complementary to DNA methylation causes the removal of methylation from CpG sites [81,82,83]. It is a complicated process that can be achieved actively by demethylases or passively during DNA replication [84,85]. Demethylation mediated by either deamination or oxidation of cytosine, results in the formation of primary sites, which is followed by the base excision repair (BER) pathway. TET proteins catalyze the oxidation pathway and thymine DNA glycosylase (TGD), while apolipoprotein B mRNA editing enzyme (APOBEC) and activation-induced deaminase (AID) catalyze the deamination pathway [83,86,87].
DNA glycosylase lyases can catalyze identification and active demethylation on specific loci (DNA-GL), a DNA demethylase, which removes 5mCs forming non-methylated cytosines [21,81]. The active removal of methylation using the BER pathway is associated with ROS1 and other DME family members such as DME, DML2 and DML3, apyrimidinic (AP) lyase and the bifunctional DNA glycosylase [88,89,90]. They are crucial for inhibiting hypermethylation at thousands of genomic regions and play a vital role in regulating transposable elements, transgenes and numerous endogenous gene expressions [91,92,93]. The removal of 5mC by DME forms a primary site which is subjected to the BER pathway [94]. In Arabidopsis, DME functions in gene imprinting in the endosperm [88], where DML2 and DML3 carried out removal of improperly placed methylation [71,95]. Hypermethylation on both endogenous genes and transgenes is prevented by ROS1 [96,97], which consequently affects the gene expression during developmental process such as, maternal imprinting [81,98], epidermal cell differentiation [93], development of male gametophytes [99] and response to pathogen attack [100]. ROS activity and function seems to be regulated by INCREASE IN DNA METHYLATION 1 (IDM1), IDM2, Histone 3 (H3) acetyltransferase, an alpha protein and methyl-binding protein7 (MBD7) [92,101,102]. Recent studies revealed that the total level of DNA methylation is determined by the combined action of DNA methyltransferases and demethylases in the regulatory loop where methylation level determines the expression of ROS1 gene [103,104].
Phylogenetic studies have revealed that the rice genome has six bifunctional DNA demethylases that are responsible for demethylating cytosine, two DML3 orthologs (DML3a and DML3b), and four ROS1 orthologs (ROS1a, 1b, 1c and 1d), while no DME ortholog has been reported in rice yet [105]. DNA GLYCOSYLASE 701 (DNG 701) also known as ROS1c, a functional DNA glycosylase, carried out demethylation and is involved in controlling the activity of retrotransposon Tos17 [84]. The knockout mutants of dng701 exhibit retrotransposon Tos17 transposition in calli compared to wildtype, while overexpression of DNG701 displays reduced methylation level and frequent transposition of Tos17 in calli, indicating that TEs repression is regulated by DNA methylation [84]. It has been revealed by a qRT-PCR investigation that DMAL3, ROS1a and ROS1d are expressed in diverse plant tissues [85]. By constructing the knock in null allele of ROS1a, it has been revealed that failure of early-stage endosperm and embryo development is caused by the maternal ros1a allele and neither paternal nor maternal allele is transmittable to progeny, showing that ROS1a demethylation plays a vital role in both female and male gametophytes [85].
The epigenetic information of DNA methylation can thus be further translated by various reader proteins to direct downstream functions. To a certain extent, reader proteins may ultimately affect the biological outcome. For readers of DNA 5mC, recent discoveries have shown that these proteins with a methyl-CpG (mCpG)-binding domain (MBD) can recognize and further bind to 5mC sites (Table 1). In mammals, the MBD family has five members, including methyl-CpG-binding protein 2 (MeCP2), MBD1, MBD2, MBD3 and MBD4 [106]. Among these MBD proteins, MBD3 does not bind to methylated DNA, while other MBD proteins bind to methylated DNA. In plants, these MBD proteins have also been identified in various pant species, such as Arabidopsis, soybean, watermelon, and Brassica napus [107]. Remarkably, recent reports have presented evidence that plant SU(VAR)3-9 homologs were among the methylated-DNA reader candidates. In Arabidopsis, the transcriptional anti-silencing factor SUVH1, and SUVH3, bound methylated DNA in vitro, were associated with euchromatic methylation in vivo, and formed a complex with two DNAJ domain-containing homologs, DNAJ1 and DNAJ2 [108]. Rice OsSUVH7 acts as a DNA 5mC reader forming a complex of BCL-2-ASSOCIATED ATHANOGENE4 (OsBAG4)-OsMYB106-OsSUVH7 expression in response to salt stress [109]. Thus, the SUVH proteins may also function as 5mC readers that bind to methylated DNA sites in plants.

3. DNA Adenine Methylation 6mA in Plants

Although 6mA was discovered during the same time period as 5mC a few decades ago, its existence was thought to be limited to prokaryotes only. However, 6mA could not achieve the same level of research attention in eukaryotes as 5mC due to its low abundance and technological limitations. 6mA is a widely known prokaryotic DNA modification which involves critical regulatory functions in eukaryotes [110,111,112,113,114,115]. Recent studies have demonstrated its potential epigenetic role in cellular processes with biological consequences in diverse eukaryotes [116], which are dynamic and probably even contradictory to 5mC, a well-known epigenetic mark. The recent discovery and characterization of 6mA in different eukaryotes including plants such as Caenorhabditis elegans (nematode) [117], Chlamydomonas reinhardtii (algae) [118], Drosophila melanogaster (insect) [119], Tetrahymena (Protozoan) [120], fungi [121], yeast [122], Bombyx mori (silkworm) [123], vertebrates such as Xenopus laevis (Frog) [124], mouse [125,126], human [127], pig and zebrafish [128], plants such as Arabidopsis thaliana [129,130], and rice [131,132] have been reported, suggesting its potential regulatory functions in all organisms [133,134].
6mA is a vital element of the gene regulatory system in mammals and plays an essential role in gene silencing. Unlike animals, adenine DNA-methyltransferases have been found associated with mitochondria such as in wheat. However, cytosine DNA-methyltransferase is absent in mitochondria. An enzyme that was isolated from mitochondria of wheat methylates an internal adenine residue in the TGATCA sequence in DNA [135] and seems to regulate mitochondrial DNA replication. Surprisingly, it has also been reported that both adenine and cytosine methylations can coincide on the one and same gene. Cytosine DNA methylation can affect adenine methylation and vice versa. Hence, adenine methylation may regulate DNA replication and gene expression such as like cytokine methylation. Plants are known as masters of epigenetic regulation. All of the significant epigenetic controls that are known to occur in eukaryotes are used by plants [50]. Because of its involvement in diverse cellular functions, it has captivated the attention of researchers globally.
DNA 6mA has been revealed as an important and potential epigenetic mark in eukaryotes including plants, such as Arabidopsis and rice. Moreover, the conservation, distribution and functions of 6mA methylomes in plants have also been reported very recently [131,132,136]. Methylomes analysis suggest that 6mA distribution and function is rather conserved among Arabidopsis and rice. 6mA is found to be enriched around the transcriptional start site (TSS). In Arabidopsis [136], Chlamydomonas [137] and fungi [138], it is correlated with active transcription. The distribution and function of 6mA has been found divergent in different organisms. For example, 6mA is evenly distributed in the C. elegans genome [117], while in Drosophila it is found mainly on transposable elements [139]. In Tetrahymena it is not associated with active transcription, but is correlated with RNA polymerase II transcribed gene [120] and in the frog it is depleted in gene exons [140], while in mouse it is associated with gene expression and is depleted in exons [126,141].
Although 6mA was found in plants long ago [142,143,144], its genomic distribution and function remained elusive for a long time. Very recently, with advances in detection methods, finally the detection of 6mA in plant genomes has been made possible. In Arabidopsis, 6mA is widely distributed across the genome and more enriched on gene bodies [136] while its function and regulatory mechanism in Arabidopsis is yet unknown. Moreover, the distribution and genomic function of 6mA has also been reported in the rice genome, where 0.2% of all adenines were 6mA methylated and its presence was mostly found on GAGG motifs that was mapped to around 20% of genes and 14% of TEs. 6mA in gene bodies was found to be correlated with gene activity while in promoters it marked silent genes. In the rice genome, 6mA was also found to overlap with 5mC and in some regions it was also found complementary to 5mC. Additionally, OsALKBH1 was also proposed to be potentially involved in 6mA demethylation (Table 1) [131], suggesting distinct roles of 6mA. In another study among two rice cultivars Japonica and Indica, Zhang et al. (2018) revealed that 6mA was broadly distributed across genomes of both cultivars, while most abundantly found on exons and promoter regions. 6mA was positively correlated with key genes related to plant development and stress response. Additionally, DDM1 was found to be crucial for 6mA modification and decreased 6mA levels; defects in plant growth were observed due to DDM1 mutations, and expression of vital genes involved in plant growth was also affected by DDM1 [132]. This suggests that 6mA is conserved as a DNA methylation modification, is positively correlated with gene expression and significantly contributes to important agronomic traits in plants [132].

4. Molecular and Biological Function of DNA Methylation

4.1. Molecular Functions of DNA Methylation

DNA methylation has been found to be involved in the regulation of various cellular processes in plants and animals, playing a leading role in genome functioning, stability and development, TE silencing, genomic imprinting and X-chromosome inactivation [21,51,145,146]. It is also involved in both gene expression and repression. For example, methylation in promoter regions is associated with gene expression and ultimately regulates growth and development in plants and other eukaryotes. Methylation of cytosine residues strongly affects DNA binding with several proteins including regulatory proteins. Usually, it blocks the binding of certain nuclear proteins involved in transcription and multiple other pathways. In contrast, some proteins specifically bind to methylated DNA sequences and assemble an entire group of proteins on DNA regulating gene expression [147]. 5mC in plants is mainly found in transposons and repetitive sequences, preventing their transcription and transposition, and therefore is vital for gene silencing and genome stability. Thus, DNA methylation in plants and animals is a relatively stable but reversible epigenetic mark regulating gene expression.
In somatic cells of multicellular organisms, genome-wide epigenetic modifications are stable, while in early embryos and germ cells genome-wide epigenetic reprogramming takes place involving DNA demethylation and remodelling of histones. This epigenetic reprogramming drives various important cellular processes such as epigenetic inheritance across generations, TEs control and imprinting [148]. In plants and mammals, cytosine methylation is crucial for gene regulation. This epigenetic mark is extensively reprogrammed in the germline and early embryos of mammals, while in plants the extent by which it is reset between generations remains largely unknown [54]. Histone modifications and small RNAs may also play role in epigenetic inheritance and reprogramming [148].
DNA methylation in association with histone modification and non-histone proteins defines chromatin structure and accessibility. DNA methylation is therefore involved in gene expression and various other cellular processes. In plants and mammals, DNA methylation in promotor regions or within a transcribed gene body is associated with gene expression and plays vital role in growth and development. The accessibility of genomic regions to regulatory proteins and protein complexes is affected by cytosine methylation, subsequently affecting the chromatin structure and transcription rate of the gene [149]. Cytosine methylation in promotor and enhancer regions usually represses gene transcription; hence, in some cases it also promotes gene transcription, for instance, ROS1 in Arabidopsis and numerous genes in tomato that inhibit fruit ripening [103,104,150]. DNA methylation on promotor regions either directly suppresses gene transcription by inhibiting the binding of transcription promotors or by promoting the binding of transcription repressors or indirectly represses transcription by promoting repressive histone modifications like H3K9 and inhibiting permissive histone modifications such as histone acetylation [66,151,152].
Moreover, genome stability can also be threatened by transposons through their relocation or the insertion of new copies of retrotransposons. Cytosine methylation is crucial for transposon silencing and regulation of endogenous genes. In Arabidopsis, heavily methylated cytosines are found in all three contexts on pericentromeric heterochromatin regions and some transposon or repeat containing euchromatic regions and regions producing siRNA. More than one third of expressed genes are methylated within transcribed regions and promotor methylation was found on only 5% of the genes. Remarkably, genes methylated on transcribed regions are highly expressed and constitutively active, while the expression of promotor methylated genes is mostly tissue specific [52,153].

4.2. Dynamic DNA Methylation in Plant Development

DNA methylation is conserved in plants and animals, as its genome-wide precise patterns are crucial for growth and development. It has been revealed that along with silencing TEs and repeats, DNA methylation regulates transition to flowering, vernalization, leaf morphology, developmental reprogramming, fertility, floral organ identity and embryogenesis, seed development, response to environmental stimuli and genome protection [154,155,156,157,158,159,160,161,162]. Recent investigations have revealed that DNA methylation plays a crucial role in controlling agronomical relevance traits, such as heterosis [163], ripening in tomato and other fleshy fruits [164,165,166,167] and response to biotic and abiotic stress [168,169]. Recent studies revealed that DNA methylation is dynamic during embryogenesis and early vegetative development. Higher levels of CHH methylation have been found in embryos compared to seedlings or adult plants, suggesting that DNA methylation has a vital role to play in embryogenesis [54]. Mutations in MET1 and DDM1 have been shown to affect seed size and development [170], indicating that DNA methylation is crucial for seed development.
Recent studies have shown that DNA methylation is dynamic during plant development. In plants, experimentally induced hypomethylations in plant genomes have shown numerous developmental defects confirming that DNA methylation is vital for proper growth and development. CMT2, CMT3, DRM and MET1 mutants in plants are viable, even in combination with one or another, in contrast to mammals where DNMT1 and DNMT3 mutants show lethality during embryo development. Thus, DNA methylation in plants is involved in a variety of processes such as fertilization, gametogenesis, vegetative and reproductive development and crosstalk between DNA methylation and histone modification [171,172]. Seed development has also been found to be affected by DDM1, an ATP dependent SWI2/SNF2 chromatin remodelling factor and MET1. DNA methylation profiles in rice have shown strong fluctuations in DNA methylation levels in embryo and endosperm during seed development, indicating the association of controlled and complex DNA methylation with seed development [173]. In soybeans, DNA methylation levels can change in gene promotors during different stages of seed maturation [174]. Moreover, methylation levels vary in different cell types in gametophytes that have a similar origin and are separated by a few cell divisions. Nevertheless, the molecular mechanism regulating these dynamic methylation patterns remains to be elucidated.
Likewise, DNA methylation has been shown to regulate the growth progress of cell division and cell expansion within a growing organ of maize plants. Maintenance of methylation is transcriptionally regulated throughout the growth zone of maize leaf (division zone, transition zone, elongation zone and mature zone). Interestingly differentially methylated sequences mainly exist at or near gene bodies; numerous other genes involved in chromatin remodelling, cell cycle progression and growth regulation were also differentially methylated [162], suggesting an essential role of DNA methylation in leaf development in maize. In Arabidopsis, dysfunction of ROS1 has been found to cause hypermethylation of promoters and repression of the gene encoding EPIDERMAL PATTERNING FACTOR 2 (EPF2), which represses stomata formation; this leads to stomatal lineage cells over-proliferation, showing the important role of DNA methylation in pattern formation of some leaf epidermal cells [93]. DME DNA glycosylase expression is found significantly in vegetative cells of male gametophyte and central cells of male gametophyte, suggesting that DME mediated demethylation in Arabidopsis is vital for endosperm development. In many fruit ripening genes, active DNA methylation is found, as their promoter regions contain binding sites for RIPENING INHIBITOR (RIN), that is a major ripening transcription factor [150,165]. RIN binding to target promoters has been confirmed in most of the known ripening genes whose expression is negatively correlated with promoter DNA methylation levels [165].

4.3. DNA Methylation Heritability in Plants

In contrast to genomic variance, the potential of epigenetic markers to transmit acquired environmentally adapted traits to offspring is extremely advantageous. Several examples of DNA methylation pattern heritability in plant species such as Nicotiana tabacum, A. thaliana, Taraxacum officinale and Picea abies, as well as crop species such as rice and wheat, have been described [175,176,177,178,179,180]. DNA methylation modification transmission in A. thaliana has been witnessed for at least eight generations [11]. Stress-induced epigenetic modifications can be passed down to offspring, but their long-term stability and impact on a population are unknown [178]. A recent study on epigenetic changes that occurred during cotton domestication indicated that epialleles were passed between generations, resulting in phenotypic diversity and allowing the widespread geographic expansion of species [181,182]. Intergenerational transfer of phenotypic traits induced by DNA methylation modifications, such as delayed flowering time, drought stress adapted roots, and modified plant architecture, has also been reported [183,184]. Because DNA methylation is heritable and can result in phenotypic alterations, inducing/removing DNA methylation has become a focus of research with the goal of altering the expression of genes already present in crop genomes. The increased interest in epigenomics has resulted in the development of approaches to induce DNA methylation, first at the genome-wide level using non-specific approaches, and now at the base-specific level using new epigenetic editing tools [185].

4.4. DNA Methylation in Response to Stresses

Plant cells can sense environmental changes or stressors, which cause epigenetic modifications such as methylation pattern modification. At a late embryonic stage, germ cells are derived from somatic tissues. As a result, epigenetic modifications caused by environmental changes or stress, such as changes in methylation pattern, can be transferred down through generations [186,187]. This is crucial for understanding phenotypic variation in nature and biological stress adaptation [188], and it also indicates that methylation changes could be useful in breeding. This potential benefit is clearly expressed by the impact of DNA methylation variations on important agronomic traits such as flowering time and plant height [189,190], pathogen resistance [191,192], and yield [193]. Methylation polymorphisms have been reported in many varieties, biotypes, and intra-species of rice [194], Arabidopsis [195], Brassica oleracea [196], and cotton (Gossypium hirsutum L.) [196]. This methylation diversity may result in phenotypic changes [197,198], which provide the raw materials for plant breeding. Breeding selection, on the other hand, can be effective only if the methylation alleles are heritable. The amount of total and strand-specific methylation in DNA from different batches of seeds varies significantly. This difference is assumed to be caused during seed production, perhaps as a result of changing environmental conditions. Seeds grown in a more stressful environment produce seedlings with increased strand-specific methylation. Such patterns are likely to emerge early in development and are heritable during several days of seedling growth [199]. Recent studies on TMV-infected tobacco [200], blight pathogen-infected rice [201], and MSAP polymorphism in Brassica oleracea populations [202] have supported the inheritability of methylation modifications generated from environmental stresses. Crops like rice and corn are bred primarily to exploit heterosis, highlighting the relevance of inherited methylation diversity patterns in F1 hybrids. Rice hybrids had lower levels of methylation at some loci than their parents, but higher levels at others [203]. Although the majority of methylated loci in corn can be inherited by hybrid progeny, 6.59–11.92% of loci in hybrids differ from their parents, and these loci represent a variety of functional proteins or ESTs. These modifications can be directed or stochastic [204], implying that methylated locus inheritance is complicated and that methylation variations could be beneficial in plant hybrid breeding.

5. Epigenetics Prospective for Plant Breeding

The applications of epigenome editing tools have facilitated the use of methylation modification in plant breeding. Although TALE-based systems are simpler to construct, the first-generation tool, ZFs, remains important due to its tiny size and binding properties. TALEs have a problematic sensitivity to DNA methylation, which hampers their usage in epigenome editing. The CRISPR-Cas9 approach has comparable efficiency, but it is easier and less expensive to implement, and it can target several sites (multiplex). CRISPR-Cas9 has already been widely used in genome editing and is rapidly becoming the most popular technique for inducing methylation in plants [184]. CRISPR-Cas9 has already been efficiently employed to edit a variety of crop genomes, including maize, rice, cotton, potato, tomato, soybean and sweet orange [205,206,207,208,209,210,211]. Endonucleases, on the other hand, run the danger of producing off-target double-strand breaks (DSBs), which can result in unanticipated changes elsewhere in the genome. Because the effects of CRISPR-Cas9 epigenome editing are gradual and proportionate to target binding, it produces less dramatic off-target activity than genome editing. Furthermore, most off-target impacts will occur in non-regulatory regions, rendering them silent [212]. Furthermore, unlike most genetic variations, which tend to produce a loss of function, epigenetic mutagenesis might result in a genetic gain of function [213].
Plant breeding could benefit from epigenome editing on multiple levels. Firstly, as with genetic alterations, it has the potential to speed the process of domestication of wild plants by modifying traits involved in growth habit, flowering time, yield, nutrition, and seed and fruit size and number in a single generation. In contrast to gene knockout, however, induced (de)methylation has a variable impact and can affect both gene expression and repression. Induced DNA methylation may also enhance hybrid breeding and plant propagation via tissue culture, which can create new gene expression patterns in their offspring [214,215,216,217], and thus help manage offspring phenotypic. It could be used to remove undesirable traits or to introduce desirable traits to crops, such as biotic and abiotic stress resistance (Table 2). Finally, it has the potential to generate plant resistance to viruses or even modify a pathogen population, making it susceptible to the plant’s defenses [218]. The extensive use of induced DNA methylation in plant breeding is hindered by a number of limitations. On a biological level, the selection of genes to target, the durable maintenance of induced epialleles, and the occurrence of off-target effects. On the other hand, improper legislation involving new breeding techniques has been introduced, which has restricted research and the adoption of sophisticated breeding approaches [219]. Numerous GMO regulations are based on out-of-date definitions that do not take into account new plant breeding techniques and are incompatible with modern genome editing technology. The European Court of Justice ruled in July 2018 that plants resulting from genome editing methods, including ones that do not employ recombinant DNA, are subject to current GMO regulations [220]. Some nations, such as Canada, define the safety of novel crops based on the end product’s attributes (product-based) rather than the technology used to create it (process-based). The United States has implemented a hybrid system in which only plants featuring novel traits are subject to stringent regulations [221]. The Australian government deregulated “DNA free” gene editing methods in 2019. It is critical to amend regulations governing novel plant breeding techniques, since it has both scientific and commercial implications [222].

5.1. Applications of DNA Methylation in Plant Breeding

Plant species, unlike animals, have certain methylation loci that can be inherited for numerous generations [245,246]. Treatment with 5-azacytidine, for example, reduces methylation in flax (Linum usitatissimum) and produces dwarfism and early maturity traits [189]. The methylation level is likewise lower in F2 and F3 progenies resulting from crossings between wild type and mutant plants selected for the phenotype of early maturity [188]. This shows that 5-azacytidine-induced hypomethylation can be passed down for at least three generations. Resistance to the blight pathogen, which is caused by variations in rice methylation, can be passed on for at least nine generations [191]. Because the market vitality of new rice varieties is often fewer than 5 years, this is a good option for commercial development. Some methylation differences can be passed down for hundreds of years, having an impact on evolution. Linnaeus 250 years ago discovered a toadflax (Linaria vulgaris) mutant in which the flower’s fundamental symmetry has been altered. This mutant’s Lcyc gene, which is crucial in floral development, has been discovered to be extensively methylated and silenced [197]. These stably transmitted DNA methylation alterations are identical to DNA sequence mutations, implying that they have genetic behavior similar to that specified by Mendelian genetics [247]. For example, traditional hybridization and segregation approaches revealed that three methylation alleles govern the phenotypic of early maturity in flax, two of which are recessive and one dominant or codominant [248]. Because of the similarities to classical genetics, traditional breeding processes such as hybridization, selection, and purification can be used to create new plant types containing novel methylation alleles.
Methylation patterns are dynamic and reversible during eukaryotic life cycles [249] and various stages of cancer development [250], it also occurs in plants sometimes [251]. The distribution patterns of methylated loci in individual clover and flax plants are largely identical before and after metal stress [252], implying that methylation pattern alterations may occur with stress and then dissipate once the stress is eliminated. Similar situation has also been observed in Bryonia dioica and Antirrhinum majus. When Bryonia dioica is exposed to rubbing stress, the level of methylation drops, but it returns to normal after the stress is removed [253]. Low temperature-induced methylation changes in Antirrhinum majus are reversed when the temperature is returned to normal even within the same generation [254]. Lower methylation levels are generally associated with higher levels of gene expression. Gene overexpression and energy loss are inescapable if a plant’s methylation level does not rebound when the stress is removed. We deduce that alterations in methylation that are reversed once the stress is removed are more common in nature, since they allow for more rational use of biological energy, just as the resistance response to pathogens occurs only when the disease is attacking. This emphasizes the possible significance of methylated alleles in breeding. This type of methylation may be studied in the same way that pathogen resistance is. Stress is first induced, either naturally or artificially, and then the resulting methylation variations in the plant genome are discovered. The methylation pattern must be assessed at multiple time points, and the amount of methylation must also be considered. A numerical approach can be used to represent the level of methylation variation in disease resistance responses. With the help of molecular markers, the methylation variations that are shown to be induced reversibly in response to stress by several rounds of multidrop experiments can be identified. Classic hybrid breeding strategies can be used to produce new varieties if the loci are closely related to stress resistance [255].

5.2. Potential Applications of RdDM in Crop Improvement

RdDM is a biological mechanism in which non-coding RNA molecules direct the addition of DNA methylation to specific sequence [256]. The mechanism underlying the sequence-specificity of RdDM is well understood. RdDM can be ‘tricked’ into targeting and silencing endogenous genes in a highly specific, which offers a variety of potential biotechnological and breeding applications. Numerous methods can be employed to induce RdDM based DNA methylation and silencing of specific genes. One approach, known as virus-induced gene silencing (VIGS), involves introducing a portion of the target gene’s promoter region into a virus [257]. As the virus replicates its own RNA, it will generate the chunk of promoter sequence as part of its own RNA, which is foreign to the plant. The viral RNA being foreign will be targeted for PTGS and processed into sRNAs, some of which will be complementary to the promoter of the original target gene. Certain sRNAs add DNA methylation to the target gene by recruiting RdDM machinery. Various studies have used this approach such as fruit ripening gene in tomato and a gene affecting flower pigmentation in Petunia have been silenced using Cucumber Mosaic Virus to recruit RdDM to silence the gene [258]. Arabidopsis FWA locus has also been silenced using VIGS, which resulted in delayed flowering [257]. RdDM-induced changes can occasionally be retained and inherited through multiple generations without outside intervention or manipulation, implying that RdDM can be a useful tool for targeted epigenome editing.

5.3. DNA Methylation and Heterosis

Heterosis, often known as hybrid vigor, refers to hybrid offspring outperforming their parents. It is a significant biological phenomenon that is often used to enhance agricultural productivity. Heterosis can be seen in plants in traits including environmental adaption yield, and growth [259]. Heterosis’ molecular foundation is mainly unknown, which limits its application in plant development. Recent omics research has demonstrated that substantial changes in gene activity between hybrids and their parents produce gene expression complementation and consequently hybrid vigor [259]. A variety of biological processes are affected by these changes, including, stress responses, light and hormone signaling, ageing and energy utilization and metabolism [260]. Epigenetic variation is also implicated in heterosis, as it regulates gene expression and stabilizes the genome [261]. As an epigenetic alteration, DNA methylation at cytosine occurs in all sequence contexts, including CG, CHG, and CHH (where H = A, T, or C). Numerous studies have demonstrated that hybrids of many crops, such as rice [262], maize [263], and pigeonpea [264], undergo DNA methylation rearrangements, leading their methylome to differ from their parental lines. Additionally, loss genome methylation of TE in hybrids resulted in loss of phenotypic superiority [265,266]. According to research in rice, paternal CHG methylation divergence may play a role in allelic-specific expression and hybrid vigor [267]. In pigeon pea, there was a substantial association between expression levels and methylation for certain heterosis-related genes [264]. Dissecting the significance of DNA methylation in heterosis, on the other hand, is difficult since genome-wide DNA methylation levels differ depending on developmental stages and environmental factors. Increased DNA methylation levels in reciprocal hybrids were found in studies employing Arabidopsis and pigeon pea seedlings [264,265]. In contrast to their parents, rice developing seeds revealed that overall DNA methylation of reciprocal hybrids decreased [268]. Changing DNA methylation in hybrids did not generate significant changes in target gene expression in 6-week-old rice seedling leaf or root [269]. In maize, there was no association between expanded leaves at the 7–8 leaf stage and grain yield heterosis identified for parental variations in CG or total (CG + CHG) methylation levels [270].
The accumulation of various phenotypic advantages at distinct developmental stages can lead to heterosis of a trait, which can start as early as double fertilization, as seen in maize hybrid embryos with phenotypic dominance at 6 days after fertilization [271]. When reciprocal hybrids and their parents are examined, embryos and endosperm experience synchronized significant alterations in the DNA methylome, which is thought to be the cause of heterosis of immature seed-related characteristics [268]. During soybean seed development, DNA methylation is altered differently in different cytosine contexts: CHH increases significantly throughout the seed and drops precipitously within the germinating seedling, whereas only minor changes in the global level of CG/CHG occur during the same developmental period [173,272]. Furthermore, in Arabidopsis, loss of CHH and CHG methylation has no effect on seed development, and the key genes for soybean seed development are found in non-methylated genomic regions [272]. According to the information above, methylation profiles produced during seed development may have other functions. Hybridization, on the other hand, can cause DNA methylation changes in hybrids, which are maintained in the hybrid progenies [273]. Additionally, epigenetic modifications that arise during seed development can be transmitted via cell division to exert their effect once the seed reaches a certain stage of development [274]. A recent study was carried out on reciprocal hybrids of soyabean. They established three soybean reciprocal hybrid combinations and then utilized MethylRAD-seq to identify CCGG and CCWGG (W = A or T) methylation in the whole genome of these hybrids and their parents during the middle development stage of contemporary seed. The study demonstrated that alterations in DNA methylation patterns occurred in immature hybrid seeds, and that parental variation had a role in differential expression between reciprocal hybrids. In hybrids, non-additive differential methylation sites (DMSs) were also found in considerable numbers. Surprisingly, the majority of these DMSs were hyper-methylated and localized in gene regions, contrary to the natural distribution of methylation sites. Further investigation of the non-additive DMSs found in gene regions revealed their involvement in a variety of biological processes, including those linked to transcriptional control and hormone function. These findings demonstrated a pattern of DNA methylation reprogramming in hybrid soybeans, which is linked to phenotypic variance and heterosis initiation [275].

5.4. Challenges and Opportunities for DNA Methylation in Breeding Applications

Plant genomic methylation modification can respond quickly to stress. For example, rubbing stress can reduce methylation levels to zero in 1 h [253]. However, some methylation alterations can only be noticed after comparatively prolonged periods of stress, such as 12 h for wounding and chilling stresses [276]. Methylation, on the other hand, varies dynamically [251]. Because of the dynamic and flexible nature of DNA methylation changes, we can precisely specify the time points at which observations are made. Errors of a few hours can lead to vastly divergent conclusions. To assess if the alterations in methylation are reversed after the stress is removed, multiple time points of observation and repeated analysis under several rounds of stress are required. For example, treatment of flax with 5-azacytidine causes the reduced level of methylation and early maturity phenotype to be reversed partly in the first generation and completely in the fourth generation of one line [248]. Under stress, quantitative changes in methylation variation can also be detected. For example, the degree of decrease in DNA methylation stressed clover and hemp, which ranges from 20% to 40%, is clearly dependent on the concentration of metal ions to which they are exposed [252]. In Stellaria longipes subjected to red/far red light, a similar phenomenon was found [277]. Quantitative analysis is far more difficult than qualitative analysis. Bisulfite sequencing and Southern blotting, the two most used approaches, are now time-consuming and expensive procedures, making their application to breeding challenging.
In the field of breeding, the utilization of methylation alleles is currently uncommon. Treatment with 5-azacytidine has been used in all previous studies to generate a decline in plant methylation levels and produce unique phenotypes that will be studied in future breeding studies. Early maturity and dwarfism in flax [188,189] and resistance to the blight pathogen in rice [191] are two examples. We believe that 5-azacytidine-induced methylation variation in breeding has numerous drawbacks. First, while the changes in methylation may not be stochastic [188], the approach does not allow for directed variation, akin to breeding with radiation-induced mutations. Second, the entire genome is influenced. As a result, favorable phenotypes are inextricably associated to a significant number of unfavorable phenotypes. The majority of plants do not survive the treatment [191], reducing its usefulness in breeding. For example, 5-azacytidine-induced early maturity has numerous impacts [278] and is associated with dwarfism [189]. To allow breeding selection with a clear target for such differences in methylation pattern, an approach comparable to molecular marker-assisted breeding in classical genetics can be developed. As a result, in the quest for strategies that can be applied to these aims, conventional genetics and breeding procedures should be examined.
Classical genetics has advanced from DNA sequencing to the post-genomic period, which is focused on the biological roles of genes. For major crops including rice, corn, and wheat, molecular linkage maps have been created, providing a platform for molecular-assisted selection and gene function analysis. These linkage maps may be employed in gene mapping and cloning, allowing the DNA sequence, gene function, and phenotype to be linked together. The approach of constructing genomic methylation maps is similar to that of DNA sequencing in classical genetics. This effort, however, is still in its early phases. The Arabidopsis high-resolution methylation map created by Zhang et al. [52] and Cokus et al. [152] is now the only example accessible. There hasn’t been a map like this for other plant species, including crops. One factor for this is that we place far too much emphasis on the DNA sequence, resulting in a lack of understanding of the critical role of DNA methylation in genetics and evolution. The absence of appropriate technologies for genome methylation mapping is another concern. The bisulfite sequencing was the initial step. It is, however, far too time and labor intensive to be used in a large-scale genomic study. Zhang et al. used chromatin immunoprecipitation and a tiling microarray approach to discover the distribution of methylated cytosines in the Arabidopsis genome for the first time in 2006. They were able to attain a resolution of as high as 35 bp [52]. However, most plant species, including crops, do not have commercial tiling microarrays, which severely restricts the technique’s potential uses. Peng et al. [278] proposed to replace chip hybridization with Solexa sequencing to generate high-resolution genome methylation maps in order to solve this issue and the limitations of microarrays. This high-throughput sequencing-based design potentially enhances the accuracy and detail of the generated map [279]. The key benefits of Solexa sequencing over microarrays are the digital signal, the lack of hybridization interference, and the removal of the need for PCR amplification [280].
Molecular marker-assisted selection was widely used in breeding practice even before the advent of genome sequencing. It is theoretically plausible that it can also be used to select methylated alleles. However, challenges exist in breeding practice. The MSAP marker, which is derived from the amplified fragment length polymorphism (AFLP) marker, is currently the most commonly used type of molecular marker for methylated alleles. The primary distinction is that MSAP employs restriction enzymes that are variably sensitive to methylation. The use of MSAP in markers assisted selection has numerous drawbacks. First, in a given sequence context, a single enzyme can only detect one type of methylation. Distinct stresses, on the other hand, may result in different methylation variations. Salt stress, for example, causes hypermethylation in CCWGG sequences in Mesembryanthemum crystallinum but leaving CCGG sequences untouched [281]. Second, the AFLP technique is time-consuming and difficult, making it only ideal for theoretical research, implying a bleak future for MSAP in breeding practices. Finally, the MSAP approach makes it difficult to quantify methylation variation.
Methyl cytosine immunoprecipitation separates methylated and unmethylated DNA into two pools (mCIP). Primers are designed to allow for target gene amplification and are used to amplify the target gene in both pools. The methylation level of the target gene is represented by the ratio of the quantity of PCR product produced from each pool. This design’s detection process is comparable to that of microsatellite (SSR: simple sequence repeat) markers in classical genetics for qualitative determination. It is comparable to RT-PCR in terms of quantitative analysis. The advantages of this design over MSAP include the creation of a single band, the capacity to investigate any locus, and the ease of detection by agarose gel electrophoresis. To identify several genes in a single sample, just one round of mCIP is required. However, numerous rounds of mCIP are required to detect a single gene in several samples, resulting in a high cost at present.

6. Conclusions and Future Perspectives

Currently sustainable agriculture has been regarded as a concept for feeding an increasing population on our planet to face the undergoing climactic changes. The continuous increases in crop yields with traditional breeding strategies meets on a bottleneck in agricultural production, constantly raising questions about the capacity of traditional genetic improvement to face emerging needs. Thus, novel breeding strategies and new technologies will be required for filling the predicted yield gaps. Benefiting from the progresses in biological and agricultural research, we are better known that DNA methylation beyond genome, have vital functions during plant growth, development and response to environmental factors. In addition to genetic factors, DNA methylation has been found to excitingly regulate a variety of genes that are vital for many important processes ranging from plant growth and development to biotic and abiotic stress response in plants having more complex genomes than Arabidopsis. These and numerous other discoveries expand our knowledge of understanding the epigenetic regulations especially DNA methylation dynamics not only in plants but other prokaryotes and eukaryotes including mammals. Further studies in future are required for undoubtedly revealing new DNA methylation modifications, their derivatives and functions in plants, novel mechanisms for targeting DNA methyltransferases and demethylases and the mechanisms by which DNA methylation patterns are generated, maintained, converted and erased. It will also be interesting to understand whether there is any interaction between the two different methylation modifications such as 5mC and 6mA, if yes, how they coordinate and affect each other. Very recently, an expansion in our understanding about how these epigenetic mechanisms control plant development and stress responses has unveiled to likely develop improved crops, heralding the promise of “synthetic epigenetics” [282]. This strategy will enable to breed engineered smart crops that can actively monitor and rapidly respond to environmental cues for meeting the requirements of sustainable and green agriculture.

Author Contributions

A.A.S. and S.C. wrote the draft review. M.C. and N.A. helped for manuscript modification. P.Z., C.G. and S.C. modified the paper and approved the final version. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Key Research and Development Program of China (2019YFD1000600), the National Undergraduate Training Program for Innovation and Entrepreneurship (201310504024), the China Postdoctoral Science Foundation (2022M711268), and the Natural Science Foundation of Hubei Province of China (2022CFB638).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Correction Statement

This article has been republished with a minor correction to the correspondence contact information and the Funding statement. This change does not affect the scientific content of the article.

References

  1. 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]
  2. Waddington, C.H. The Epigenotype. Int. J. Epidemiol. 2012, 41, 10–13. [Google Scholar] [CrossRef]
  3. Kumar, S.; Singh, A. Epigenetic Regulation of Abiotic Stress Tolerance in Plants. Adv. Plants Agric. Res. 2016, 5, 00179. [Google Scholar] [CrossRef]
  4. Saze, H. Epigenetic Memory Transmission through Mitosis and Meiosis in Plants. Semin. Cell Dev. Biol. 2008, 19, 527–536. [Google Scholar] [CrossRef]
  5. Thiebaut, F.; Hemerly, A.S.; Ferreira, P.C.G. A Role for Epigenetic Regulation in the Adaptation and Stress Responses of Non-Model Plants. Front. Plant. Sci. 2019, 10, 246. [Google Scholar] [CrossRef]
  6. Henderson, I.R.; Jacobsen, S.E. Epigenetic Inheritance in Plants. Nature 2007, 447, 418–424. [Google Scholar] [CrossRef]
  7. Stassen, J.H.M.; López, A.; Jain, R.; Pascual-Pardo, D.; Luna, E.; Smith, L.M.; Ton, J. The Relationship between Transgenerational Acquired Resistance and Global DNA Methylation in Arabidopsis. Sci. Rep. 2018, 8, 1–13. [Google Scholar] [CrossRef]
  8. Russo, V.E.A.; Martienssen, R.A.; Riggs, A.D. Epigenetic Mechanisms of Gene Regulation; Cold Spring Harbor Laboratory Press: New York, NY, USA, 1996; ISBN 0879694904. [Google Scholar]
  9. Slotkin, R.K. Plant epigenetics: From genotype to phenotype and back again. Genome Biol. 2016, 17, 57. [Google Scholar] [CrossRef]
  10. Liu, B.; Wendel, J.F. Epigenetic Phenomena and the Evolution of Plant Allopolyploids. Mol. Phylogenet Evol. 2003, 29, 365–379. [Google Scholar] [CrossRef]
  11. Song, Q.; Zhang, T.; Stelly, D.M.; Chen, Z.J. Epigenomic and Functional Analyses Reveal Roles of Epialleles in the Loss of Photoperiod Sensitivity during Domestication of Allotetraploid Cottons. Genome Biol. 2017, 18, 1–14. [Google Scholar] [CrossRef]
  12. Zhang, Y.; Fischer, M.; Colot, V.; Bossdorf, O. Epigenetic Variation Creates Potential for Evolution of Plant Phenotypic Plasticity. New Phytol. 2013, 197, 314–322. [Google Scholar] [CrossRef]
  13. Stroud, H.; Ding, B.; Simon, S.A.; Feng, S.; Bellizzi, M.; Pellegrini, M.; Wang, G.-L.; Meyers, B.C.; Jacobsen, S.E. Plants Regenerated from Tissue Culture Contain Stable Epigenome Changes in Rice. elife 2013, 2, e00354. [Google Scholar] [CrossRef] [PubMed]
  14. Lin, W.; Xiao, X.; Zhang, H.; Li, Y.; Liu, S.; Sun, W.; Zhang, X.; Wu, Q. Whole-Genome Bisulfite Sequencing Reveals a Role for Dna Methylation in Variants from Callus Culture of Pineapple (Ananas comosus L.). Genes 2019, 10, 877. [Google Scholar] [CrossRef] [PubMed]
  15. Stelpflug, S.C.; Eichten, S.R.; Hermanson, P.J.; Springer, N.M.; Kaeppler, S.M. Consistent and Heritable Alterations of DNA Methylation Are Induced by Tissue Culture in Maize. Genetics 2014, 198, 209–218. [Google Scholar] [CrossRef] [PubMed]
  16. Han, Z.; Crisp, P.A.; Stelpflug, S.; Kaeppler, S.M.; Li, Q.; Springer, N.M. Heritable Epigenomic Changes to the Maize Methylome Resulting from Tissue Culture. Genetics 2018, 209, 983–995. [Google Scholar] [CrossRef] [PubMed]
  17. Zhang, D.; Wang, Z.; Wang, N.; Gao, Y.; Liu, Y.; Wu, Y.; Bai, Y.; Zhang, Z.; Lin, X.; Dong, Y. Tissue Culture-Induced Heritable Genomic Variation in Rice, and Their Phenotypic Implications. PLoS ONE 2014, 9, e96879. [Google Scholar] [CrossRef] [PubMed]
  18. Suzuki, M.M.; Bird, A. DNA Methylation Landscapes: Provocative Insights from Epigenomics. Nat. Rev. Genet. 2008, 9, 465–476. [Google Scholar] [CrossRef]
  19. Mazzio, E.A.; Soliman, K.F.A. Basic Concepts of Epigenetics: Impact of Environmental Signals on Gene Expression. Epigenetics 2012, 7, 119–130. [Google Scholar] [CrossRef]
  20. García-Aguilar, M.; Gillmor, C.S. Zygotic Genome Activation and Imprinting: Parent-of-Origin Gene Regulation in Plant Embryogenesis. Curr. Opin. Plant. Biol. 2015, 27, 29–35. [Google Scholar] [CrossRef]
  21. Law, J.A.; Jacobsen, S.E. Establishing, Maintaining and Modifying DNA Methylation Patterns in Plants and Animals. Nat. Rev. Genet. 2010, 11, 204–220. [Google Scholar] [CrossRef]
  22. He, G.; Elling, A.A.; Deng, X.W. The Epigenome and Plant Development. Annu. Rev. Plant Biol. 2011, 62, 411–435. [Google Scholar] [CrossRef]
  23. He, X.-J.; Chen, T.; Zhu, J.-K. Regulation and Function of DNA Methylation in Plants and Animals. Cell Res. 2011, 21, 442–465. [Google Scholar] [CrossRef]
  24. Li, Q.; Xu, W.; Cui, Y.; Ma, L.; Richards, J.; Li, W.; Ma, Y.; Fu, G.; Bythwood, T.; Wang, Y. A Preliminary Exploration on DNA Methylation of Transgene across Generations in Transgenic Rats. Sci. Rep. 2015, 5, 8292. [Google Scholar] [CrossRef] [PubMed]
  25. Hackett, J.A.; Surani, M.A. DNA Methylation Dynamics during the Mammalian Life Cycle. Phil. Trans. R. Soc. B 2013, 368, 20110328. [Google Scholar] [CrossRef] [PubMed]
  26. Delcuve, G.P.; Rastegar, M.; Davie, J.R. Epigenetic Control. J. Cell. Physiol. 2009, 219, 243–250. [Google Scholar] [CrossRef] [PubMed]
  27. Altun, G.; Loring, J.F.; Laurent, L.C. DNA Methylation in Embryonic Stem Cells. J. Cell. Biochem. 2010, 109, 1–6. [Google Scholar] [CrossRef]
  28. Smith, Z.D.; Meissner, A. DNA Methylation: Roles in Mammalian Development. Nat. Rev. Genet. 2013, 14, 204–220. [Google Scholar] [CrossRef]
  29. Elhamamsy, A.R. DNA Methylation Dynamics in Plants and Mammals: Overview of Regulation and Dysregulation. Cell Biochem. Funct. 2016, 34, 289–298. [Google Scholar] [CrossRef]
  30. Goettel, W.; Messing, J. Epiallele Biogenesis in Maize. Gene 2013, 516, 8–23. [Google Scholar] [CrossRef]
  31. Ito, H. Small RNAs and Regulation of Transposons in Plants. Genes Genet. Syst. 2013, 88, 3–7. [Google Scholar] [CrossRef]
  32. Shi, J.; Shi, W.; Ni, L.; Xu, X.; Su, X.; Xia, L.; Xu, F.; Chen, J.; Zhu, J. OCT4 Is Epigenetically Regulated by DNA Hypomethylation of Promoter and Exon in Primary Gliomas. Oncol. Rep. 2013, 30, 201–206. [Google Scholar] [CrossRef] [PubMed]
  33. Zou, X.; Ma, W.; Solov’Yov, I.A.; Chipot, C.; Schulten, K. Recognition of Methylated DNA through Methyl-CpG Binding Domain Proteins. Nucleic Acids Res. 2012, 40, 2747–2758. [Google Scholar] [CrossRef] [PubMed]
  34. Almeida, R.; Allshire, R.C. RNA Silencing and Genome Regulation. Trends Cell Biol. 2005, 15, 251–258. [Google Scholar] [CrossRef] [PubMed]
  35. Teixeira, F.K.; Heredia, F.; Sarazin, A.; Roudier, F.; Boccara, M.; Ciaudo, C.; Cruaud, C.; Poulain, J.; Berdasco, M.; Fraga, M.F. A Role for RNAi in the Selective Correction of DNA Methylation Defects. Science 2009, 323, 1600–1604. [Google Scholar] [CrossRef] [PubMed]
  36. Zhang, X. The Epigenetic Landscape of Plants. Science 2008, 320, 489–492. [Google Scholar] [CrossRef] [PubMed]
  37. Zilberman, D.; Gehring, M.; Tran, R.K.; Ballinger, T.; Henikoff, S. Genome-Wide Analysis of Arabidopsis thaliana DNA Methylation Uncovers an Interdependence between Methylation and Transcription. Nat. Genet. 2007, 39, 61–69. [Google Scholar] [CrossRef] [PubMed]
  38. Lister, R.; O’Malley, R.C.; Tonti-Filippini, J.; Gregory, B.D.; Berry, C.C.; Millar, A.H.; Ecker, J.R. Highly Integrated Single-Base Resolution Maps of the Epigenome in Arabidopsis. Cell 2008, 133, 523–536. [Google Scholar] [CrossRef]
  39. Barkan, A.; Martienssen, R.A. Inactivation of Maize Transposon Mu Suppresses a Mutant Phenotype by Activating an Outward-Reading Promoter near the End of Mu1. Proc. Natl. Acad. Sci. USA 1991, 88, 3502–3506. [Google Scholar] [CrossRef]
  40. Zhang, X.; Shiu, S.; Cal, A.; Borevitz, J.O. Global Analysis of Genetic, Epigenetic and Transcriptional Polymorphisms in Arabidopsis thaliana Using Whole Genome Tiling Arrays. PLoS Genet. 2008, 4, e1000032. [Google Scholar] [CrossRef]
  41. Kawakatsu, T.; Ecker, J.R. Diversity and Dynamics of DNA Methylation: Epigenomic Resources and Tools for Crop Breeding. Breed. Sci. 2019, 69, 191–204. [Google Scholar] [CrossRef]
  42. Tirnaz, S.; Batley, J. DNA Methylation: Toward Crop Disease Resistance Improvement. Trends Plant. Sci. 2019, 24, 1137–1150. [Google Scholar] [CrossRef] [PubMed]
  43. Yang, X.; Sanchez, R.; Kundariya, H.; Maher, T.; Dopp, I.; Schwegel, R.; Virdi, K.; Axtell, M.J.; Mackenzie, S.A. Segregation of an MSH1 RNAi Transgene Produces Heritable Non-Genetic Memory in Association with Methylome Reprogramming. Nat. Commun. 2020, 11, 1–17. [Google Scholar] [CrossRef] [PubMed]
  44. Do Kim, K.; El Baidouri, M.; Jackson, S.A. Accessing Epigenetic Variation in the Plant Methylome. Brief. Funct. Genom. 2014, 13, 318–327. [Google Scholar] [CrossRef] [PubMed]
  45. Liu, W.; Duttke, S.H.; Hetzel, J.; Groth, M.; Feng, S.; Gallego-Bartolome, J.; Zhong, Z.; Kuo, H.Y.; Wang, Z.; Zhai, J.; et al. RNA-Directed DNA Methylation Involves Co-Transcriptional Small-RNA-Guided Slicing of Polymerase V Transcripts in Arabidopsis. Nat. Plants 2018, 4, 181–188. [Google Scholar] [CrossRef]
  46. Matzke, M.A.; Mosher, R.A. RNA-Directed DNA Methylation: An Epigenetic Pathway of Increasing Complexity. Nat. Rev. Genet. 2014, 15, 394–408. [Google Scholar] [CrossRef]
  47. Zhong, X.; Du, J.; Hale, C.J.; Gallego-Bartolome, J.; Feng, S.; Vashisht, A.A.; Chory, J.; Wohlschlegel, J.A.; Patel, D.J.; Jacobsen, S.E. Molecular Mechanism of Action of Plant DRM de Novo DNA Methyltransferases. Cell 2014, 157, 1050–1060. [Google Scholar] [CrossRef]
  48. Law, J.A.; Du, J.; Hale, C.J.; Feng, S.; Krajewski, K.; Palanca, A.M.S.; Strahl, B.D.; Patel, D.J.; Jacobsen, S.E. Polymerase IV Occupancy at RNA-Directed DNA Methylation Sites Requires SHH1. Nature 2013, 498, 385–389. [Google Scholar] [CrossRef]
  49. Zhang, H.; Ma, Z.-Y.; Zeng, L.; Tanaka, K.; Zhang, C.-J.; Ma, J.; Bai, G.; Wang, P.; Zhang, S.-W.; Liu, Z.-W. DTF1 Is a Core Component of RNA-Directed DNA Methylation and May Assist in the Recruitment of Pol IV. Proc. Natl. Acad. Sci. USA 2013, 110, 8290–8295. [Google Scholar] [CrossRef]
  50. Pikaard, C.S.; Mittelsten Scheid, O. Epigenetic Regulation in Plants. Cold Spring Harb. Perspect. Biol. 2014, 6, a019315. [Google Scholar] [CrossRef]
  51. Bartels, A.; Han, Q.; Nair, P.; Stacey, L.; Gaynier, H.; Mosley, M.; Huang, Q.Q.; Pearson, J.K.; Hsieh, T.-F.; An, Y.-Q.C.; et al. Dynamic DNA Methylation in Plant Growth and Development. Int. J. Mol. Sci. 2018, 19, 2144. [Google Scholar] [CrossRef]
  52. Zhang, X.; Yazaki, J.; Sundaresan, A.; Cokus, S.; Chan, S.W.-L.; Chen, H.; Henderson, I.R.; Shinn, P.; Pellegrini, M.; Jacobsen, S.E. Genome-Wide High-Resolution Mapping and Functional Analysis of DNA Methylation in Arabidopsis. Cell 2006, 126, 1189–1201. [Google Scholar] [CrossRef] [PubMed]
  53. Niederhuth, C.E.; Bewick, A.J.; Ji, L.; Alabady, M.S.; Kim, K.D.; Li, Q.; Rohr, N.A.; Rambani, A.; Burke, J.M.; Udall, J.A.; et al. Widespread Natural Variation of DNA Methylation within Angiosperms. Genome Biol. 2016, 17, 194. [Google Scholar] [CrossRef] [PubMed]
  54. Bouyer, D.; Kramdi, A.; Kassam, M.; Heese, M.; Schnittger, A.; Roudier, F.; Colot, V. DNA Methylation Dynamics during Early Plant Life. Genome Biol. 2017, 18, 1–12. [Google Scholar] [CrossRef] [PubMed]
  55. Shi, D.Q.; Ali, I.; Tang, J.; Yang, W.C. New Insights into 5hmC DNA Modification: Generation, Distribution and Function. Front. Genet. 2017, 8, 100. [Google Scholar] [CrossRef]
  56. Plongthongkum, N.; Diep, D.H.; Zhang, K. Advances in the Profiling of DNA Modifications: Cytosine Methylation and Beyond. Nat. Rev. Genet. 2014, 15, 647–661. [Google Scholar] [CrossRef]
  57. Rajakumara, E.; Nakarakanti, N.K.; Nivya, M.A.; Satish, M. Mechanistic Insights into the Recognition of 5-Methylcytosine Oxidation Derivatives by the SUVH5 SRA Domain. Sci. Rep. 2016, 6, 20161. [Google Scholar] [CrossRef]
  58. He, Y.F.; Li, B.Z.; Li, Z.; Liu, P.; Wang, Y.; Tang, Q.; Ding, J.; Jia, Y.; Chen, Z.; Li, N.; et al. Tet-Mediated Formation of 5-Carboxylcytosine and Its Excision by TDG in Mammalian DNA. Science 2011, 333, 1303–1307. [Google Scholar] [CrossRef]
  59. Wang, X.L.; Song, S.H.; Wu, Y.S.; Li, Y.L.; Chen, T.T.; Huang, Z.Y.; Liu, S.; Dunwell, T.L.; Pfeifer, G.P.; Dunwell, J.M.; et al. Genome-Wide Mapping of 5-Hydroxymethylcytosine in Three Rice Cultivars Reveals Its Preferential Localization in Transcriptionally Silent Transposable Element Genes. J. Exp. Bot. 2015, 66, 6651–6663. [Google Scholar] [CrossRef]
  60. Terragni, J.; Bitinaite, J.; Zheng, Y.; Pradhan, S. Biochemical Characterization of Recombinant β-Glucosyltransferase and Analysis of Global 5-Hydroxymethylcytosine in Unique Genomes. Biochemistry 2012, 51, 1009–1019. [Google Scholar] [CrossRef]
  61. Yao, Q.; Song, C.X.; He, C.; Kumaran, D.; Dunn, J.J. Heterologous Expression and Purification of Arabidopsis thaliana VIM1 Protein: In Vitro Evidence for Its Inability to Recognize Hydroxymethylcytosine, a Rare Base in Arabidopsis DNA. Protein Expr. Purif. 2012, 83, 104–111. [Google Scholar] [CrossRef]
  62. Liu, S.; Dunwell, T.L.; Pfeifer, G.P.; Dunwell, J.M.; Ullah, I.; Wang, Y. Detection of Oxidation Products of 5-Methyl-2′-Deoxycytidine in Arabidopsis DNA. PLoS ONE 2013, 8, e84620. [Google Scholar] [CrossRef] [PubMed]
  63. Moricová, P.; Ondřej, V.; Navrátilová, B.; Luhová, L. Changes of DNA Methylation and Hydroxymethylation in Plant Protoplast Cultures. Acta Biochim. Pol. 2013, 60, 33–36. [Google Scholar] [CrossRef] [PubMed]
  64. Jang, H.; Shin, H.; Eichman, B.F.; Huh, J.H. Excision of 5-Hydroxymethylcytosine by DEMETER Family DNA Glycosylases. Biochem. Biophys. Res. Commun. 2014, 446, 1067–1072. [Google Scholar] [CrossRef] [PubMed]
  65. Vanyushin, B.F.; Ashapkin, V.V. DNA Methylation in Higher Plants: Past, Present and Future. Biochim. Biophys. Acta 2011, 1809, 360–368. [Google Scholar] [CrossRef]
  66. Zhang, M.; Kimatu, J.N.; Xu, K.; Liu, B. DNA Cytosine Methylation in Plant Development. J. Genet. Genom. 2010, 37, 1–12. [Google Scholar] [CrossRef]
  67. Hirsch, S.; Baumberger, R.; Grossniklaus, U. Epigenetic Variation, Inheritance, and Selection in Plant Populations. In Cold Spring Harbor Symposia on Quantitative Biology; Cold Spring Harbor Laboratory Press: New York, NY, USA, 2012; Volume 77, pp. 97–104. [Google Scholar]
  68. Du, J.; Johnson, L.M.; Groth, M.; Feng, S.; Hale, C.J.; Li, S.; Vashisht, A.A.; Gallego-Bartolome, J.; Wohlschlegel, J.A.; Patel, D.J. Mechanism of DNA Methylation-Directed Histone Methylation by KRYPTONITE. Mol. Cell 2014, 55, 495–504. [Google Scholar] [CrossRef]
  69. Zemach, A.; Kim, M.Y.; Hsieh, P.-H.; Coleman-Derr, D.; Eshed-Williams, L.; Thao, K.; Harmer, S.L.; Zilberman, D. The Arabidopsis Nucleosome Remodeler DDM1 Allows DNA Methyltransferases to Access H1-Containing Heterochromatin. Cell 2013, 153, 193–205. [Google Scholar] [CrossRef]
  70. Stroud, H.; Do, T.; Du, J.; Zhong, X.; Feng, S.; Johnson, L.; Patel, D.J.; Jacobsen, S.E. Non-CG Methylation Patterns Shape the Epigenetic Landscape in Arabidopsis. Nat. Struct. Mol. Biol. 2014, 21, 64–72. [Google Scholar] [CrossRef]
  71. Ortega-Galisteo, A.P.; Morales-Ruiz, T.; Ariza, R.R.; Roldán-Arjona, T. Arabidopsis DEMETER-LIKE Proteins DML2 and DML3 Are Required for Appropriate Distribution of DNA Methylation Marks. Plant. Mol. Biol. 2008, 67, 671–681. [Google Scholar] [CrossRef]
  72. Teerawanichpan, P.; Krittanai, P.; Chauvatcharin, N.; Narangajavana, J. Purification and Characterization of Rice DNA Methyltransferase. Plant. Physiol. Biochem. 2009, 47, 671–680. [Google Scholar] [CrossRef]
  73. Pradhan, S.; Cummings, M.; Roberts, R.J.; Adams, R.L.P. Isolation, Characterization and Baculovirus-Mediated Expression of the CDNA Encoding Cytosine DNA Methyltransferase from Pisum Sativum. Nucleic Acids Res. 1998, 26, 1214–1222. [Google Scholar] [CrossRef] [PubMed]
  74. Steward, N.; Kusano, T.; Sano, H. Expression of ZmMET1, a Gene Encoding a DNA Methyltransferase from Maize, Is Associated Not Only with DNA Replication in Actively Proliferating Cells, but Also with Altered DNA Methylation Status in Cold-Stressed Quiescent Cells. Nucleic Acids Res. 2000, 28, 3250–3259. [Google Scholar] [CrossRef] [PubMed]
  75. Lindroth, A.M.; Cao, X.; Jackson, J.P.; Zilberman, D.; McCallum, C.M.; Henikoff, S.; Jacobsen, S.E. Requirement of CHROMOMETHYLASE3 for Maintenance of CpXpG Methylation. Science 2001, 292, 2077–2080. [Google Scholar] [CrossRef] [PubMed]
  76. Bartee, L.; Malagnac, F.; Bender, J. Arabidopsis Cmt3 Chromomethylase Mutations Block Non-CG Methylation and Silencing of an Endogenous Gene. Genes Dev. 2001, 15, 1753–1758. [Google Scholar] [CrossRef]
  77. Fujimoto, R.; Sasaki, T.; Nishio, T. Characterization of DNA Methyltransferase Genes in Brassica Rapa. Genes Genet. Syst. 2006, 81, 235–242. [Google Scholar] [CrossRef]
  78. Cao, X.; Jacobsen, S.E. Locus-Specific Control of Asymmetric and CpNpG Methylation by the DRM and CMT3 Methyltransferase Genes. Proc. Natl. Acad. Sci. USA 2002, 99, 16491–16498. [Google Scholar] [CrossRef]
  79. Zhang, X.; Jacobsen, S.E. Genetic Analyses of DNA Methyltransferases in Arabidopsis thaliana. In Cold Spring Harbor Symposia on Quantitative Biology; Cold Spring Harbor Laboratory Press: New York, NY, USA, 2006. [Google Scholar]
  80. Pavlopoulou, A.; Kossida, S. Plant Cytosine-5 DNA Methyltransferases: Structure, Function, and Molecular Evolution. Genomics 2007, 90, 530–541. [Google Scholar] [CrossRef]
  81. Zhu, J.-K. Active DNA Demethylation Mediated by DNA Glycosylases. Annu. Rev. Genet. 2009, 43, 143. [Google Scholar] [CrossRef]
  82. Wu, S.C.; Zhang, Y. Active DNA Demethylation: Many Roads Lead to Rome. Nat. Rev. Mol. Cell Biol. 2010, 11, 607–620. [Google Scholar] [CrossRef]
  83. Bhutani, N.; Burns, D.M.; Blau, H.M. DNA Demethylation Dynamics. Cell 2011, 146, 866–872. [Google Scholar] [CrossRef]
  84. La, H.; Ding, B.; Mishra, G.P.; Zhou, B.; Yang, H.; del Rosario Bellizzi, M.; Chen, S.; Meyers, B.C.; Peng, Z.; Zhu, J.-K. A 5-Methylcytosine DNA Glycosylase/Lyase Demethylates the Retrotransposon Tos17 and Promotes Its Transposition in Rice. Proc. Natl. Acad. Sci. USA 2011, 108, 15498–15503. [Google Scholar] [CrossRef] [PubMed]
  85. Ono, A.; Yamaguchi, K.; Fukada-Tanaka, S.; Terada, R.; Mitsui, T.; Iida, S. A Null Mutation of ROS1a for DNA Demethylation in Rice Is Not Transmittable to Progeny. Plant. J. 2012, 71, 564–574. [Google Scholar] [CrossRef] [PubMed]
  86. Nabel, C.S.; Manning, S.A.; Kohli, R.M. The Curious Chemical Biology of Cytosine: Deamination, Methylation, and Oxidation as Modulators of Genomic Potential. ACS Chem. Biol. 2011, 7, 20–30. [Google Scholar] [CrossRef] [PubMed]
  87. Kohli, R.M.; Zhang, Y. TET Enzymes, TDG and the Dynamics of DNA Demethylation. Nature 2013, 502, 472–479. [Google Scholar] [CrossRef]
  88. Gehring, M.; Huh, J.H.; Hsieh, T.-F.; Penterman, J.; Choi, Y.; Harada, J.J.; Goldberg, R.B.; Fischer, R.L. DEMETER DNA Glycosylase Establishes MEDEA Polycomb Gene Self-Imprinting by Allele-Specific Demethylation. Cell 2006, 124, 495–506. [Google Scholar] [CrossRef]
  89. Morales-Ruiz, T.; Ortega-Galisteo, A.P.; Ponferrada-Marín, M.I.; Martínez-Macías, M.I.; Ariza, R.R.; Roldán-Arjona, T. Demeter and Repressor of Silencing 1 Encode 5-Methylcytosine DNA Glycosylases. Proc. Natl. Acad. Sci. USA 2006, 103, 6853–6858. [Google Scholar] [CrossRef]
  90. Zhang, H.; Zhu, J.-K. Active DNA Demethylation in Plants and Animals. In Cold Spring Harbor Symposia on Quantitative Biology; Cold Spring Harbor Laboratory Press: New York, NY, USA, 2012; Volume 77, pp. 161–173. [Google Scholar]
  91. Zhang, P.; Li, X.; Wang, Y.; Guo, W.; Chachar, S.; Riaz, A.; Geng, Y.; Gu, X.; Yang, L. PRMT6 physically associates with nuclear factor Y to regulate photoperiodic flowering in Arabidopsis. aBiotech 2021, 2, 403–414. [Google Scholar] [CrossRef]
  92. Qian, W.; Miki, D.; Zhang, H.; Liu, Y.; Zhang, X.; Tang, K.; Kan, Y.; La, H.; Li, X.; Li, S. A Histone Acetyltransferase Regulates Active DNA Demethylation in Arabidopsis. Science 2012, 336, 1445–1448. [Google Scholar] [CrossRef]
  93. Yamamuro, C.; Miki, D.; Zheng, Z.; Ma, J.; Wang, J.; Yang, Z.; Dong, J.; Zhu, J.-K. Overproduction of Stomatal Lineage Cells in Arabidopsis Mutants Defective in Active DNA Demethylation. Nat. Commun. 2014, 5, 1–7. [Google Scholar] [CrossRef]
  94. Gong, Z.; Zhu, J.-K. Active DNA Demethylation by Oxidation and Repair. Cell Res. 2011, 21, 1649–1651. [Google Scholar] [CrossRef]
  95. Penterman, J.; Zilberman, D.; Huh, J.H.; Ballinger, T.; Henikoff, S.; Fischer, R.L. DNA Demethylation in the Arabidopsis Genome. Proc. Natl. Acad. Sci. USA 2007, 104, 6752–6757. [Google Scholar] [CrossRef] [PubMed]
  96. Gong, Z.; Morales-Ruiz, T.; Ariza, R.R.; Roldán-Arjona, T.; David, L.; Zhu, J.-K. ROS1, a Repressor of Transcriptional Gene Silencing in Arabidopsis, Encodes a DNA Glycosylase/Lyase. Cell 2002, 111, 803–814. [Google Scholar] [CrossRef]
  97. Agius, F.; Kapoor, A.; Zhu, J.-K. Role of the Arabidopsis DNA Glycosylase/Lyase ROS1 in Active DNA Demethylation. Proc. Natl. Acad. Sci. USA 2006, 103, 11796–11801. [Google Scholar] [CrossRef]
  98. Gehring, M.; Reik, W.; Henikoff, S. DNA Demethylation by DNA Repair. Trends Genet. 2009, 25, 82–90. [Google Scholar] [CrossRef] [PubMed]
  99. Schoft, V.K.; Chumak, N.; Choi, Y.; Hannon, M.; Garcia-Aguilar, M.; Machlicova, A.; Slusarz, L.; Mosiolek, M.; Park, J.-S.; Park, G.T. Function of the DEMETER DNA Glycosylase in the Arabidopsis thaliana Male Gametophyte. Proc. Natl. Acad. Sci. USA 2011, 108, 8042–8047. [Google Scholar] [CrossRef] [PubMed]
  100. Yu, A.; Lepère, G.; Jay, F.; Wang, J.; Bapaume, L.; Wang, Y.; Abraham, A.-L.; Penterman, J.; Fischer, R.L.; Voinnet, O. Dynamics and Biological Relevance of DNA Demethylation in Arabidopsis Antibacterial Defense. Proc. Natl. Acad. Sci. USA 2013, 110, 2389–2394. [Google Scholar] [CrossRef]
  101. Qian, W.; Miki, D.; Lei, M.; Zhu, X.; Zhang, H.; Liu, Y.; Li, Y.; Lang, Z.; Wang, J.; Tang, K. Regulation of Active DNA Demethylation by an α-Crystallin Domain Protein in Arabidopsis. Mol. Cell 2014, 55, 361–371. [Google Scholar] [CrossRef]
  102. Wang, C.; Dong, X.; Jin, D.; Zhao, Y.; Xie, S.; Li, X.; He, X.; Lang, Z.; Lai, J.; Zhu, J.-K. Methyl-CpG-Binding Domain Protein MBD7 Is Required for Active DNA Demethylation in Arabidopsis. Plant. Physiol. 2015, 167, 905–914. [Google Scholar] [CrossRef]
  103. Lei, M.; Zhang, H.; Julian, R.; Tang, K.; Xie, S.; Zhu, J.-K. Regulatory Link between DNA Methylation and Active Demethylation in Arabidopsis. Proc. Natl. Acad. Sci. USA 2015, 112, 3553–3557. [Google Scholar] [CrossRef]
  104. Williams, B.P.; Pignatta, D.; Henikoff, S.; Gehring, M. Methylation-Sensitive Expression of a DNA Demethylase Gene Serves as an Epigenetic Rheostat. PLoS Genet. 2015, 11, e1005142. [Google Scholar] [CrossRef]
  105. Zemach, A.; Kim, M.Y.; Silva, P.; Rodrigues, J.A.; Dotson, B.; Brooks, M.D.; Zilberman, D. Local DNA Hypomethylation Activates Genes in Rice Endosperm. Proc. Natl. Acad. Sci. USA 2010, 107, 18729–18734. [Google Scholar] [CrossRef]
  106. Zhu, H.; Wang, G.; Qian, J. Transcription Factors as Readers and Effectors of DNA Methylation. Nat. Rev. Genet. 2016, 17, 551–565. [Google Scholar] [CrossRef] [PubMed]
  107. Coelho, F.S.; Sangi, S.; Moraes, J.L.; da Silva Santos, W.; Gamosa, E.A.; Fernandes, K.V.S.; Grativol, C. Methyl-CpG Binding Proteins (MBD) Family Evolution and Conservation in Plants. Gene 2022, 824, 146404. [Google Scholar] [CrossRef] [PubMed]
  108. Harris, C.J.; Scheibe, M.; Wongpalee, S.P.; Liu, W.; Cornett, E.M.; Vaughan, R.M.; Li, X.; Chen, W.; Xue, Y.; Zhong, Z. A DNA Methylation Reader Complex That Enhances Gene Transcription. Science 2018, 362, 1182–1186. [Google Scholar] [CrossRef]
  109. Wang, J.; Nan, N.; Li, N.; Liu, Y.; Wang, T.-J.; Hwang, I.; Liu, B.; Xu, Z.-Y. A DNA Methylation Reader–Chaperone Regulator–Transcription Factor Complex Activates OsHKT1; 5 Expression during Salinity Stress. Plant. Cell 2020, 32, 3535–3558. [Google Scholar] [CrossRef] [PubMed]
  110. Li, X.; Qian, W.; Zhao, Y.; Wang, C.; Shen, J.; Zhu, J.-K.; Gong, Z. Antisilencing Role of the RNA-Directed DNA Methylation Pathway and a Histone Acetyltransferase in Arabidopsis. Proc. Natl. Acad. Sci. USA 2012, 109, 11425–11430. [Google Scholar] [CrossRef]
  111. Tang, K.; Lang, Z.; Zhang, H.; Zhu, J.K. The DNA Demethylase ROS1 Targets Genomic Regions with Distinct Chromatin Modifications. Nat. Plants 2016, 2, 1–10. [Google Scholar] [CrossRef]
  112. Mathieu, O.; Reinders, J.; Čaikovski, M.; Smathajitt, C.; Paszkowski, J. Transgenerational Stability of the Arabidopsis Epigenome Is Coordinated by CG Methylation. Cell 2007, 130, 851–862. [Google Scholar] [CrossRef]
  113. Hu, L.; Li, N.; Xu, C.; Zhong, S.; Lin, X.; Yang, J.; Zhou, T.; Yuliang, A.; Wu, Y.; Chen, Y.-R. Mutation of a Major CG Methylase in Rice Causes Genome-Wide Hypomethylation, Dysregulated Genome Expression, and Seedling Lethality. Proc. Natl. Acad. Sci. USA 2014, 111, 10642–10647. [Google Scholar] [CrossRef]
  114. Erhard, K.F.; Talbot, J.E.R.B.; Deans, N.C.; McClish, A.E.; Hollick, J.B. Nascent Transcription Affected by RNA Polymerase IV in Zea Mays. Genetics 2015, 199, 1107–1125. [Google Scholar] [CrossRef]
  115. Liu, J.; Zhu, Y.; Luo, G.Z.; Wang, X.; Yue, Y.; Wang, X.; Zong, X.; Chen, K.; Yin, H.; Fu, Y.; et al. Abundant DNA 6mA Methylation during Early Embryogenesis of Zebrafish and Pig. Nat. Commun. 2016, 7, 13052. [Google Scholar] [CrossRef] [PubMed]
  116. Luo, G.Z.; Wang, F.; Weng, X.; Chen, K.; Hao, Z.; Yu, M.; Deng, X.; Liu, J.; He, C. Characterization of Eukaryotic DNA N6-Methyladenine by a Highly Sensitive Restriction Enzyme-Assisted Sequencing. Nat. Commun. 2016, 7, 11301. [Google Scholar] [CrossRef] [PubMed]
  117. Greer, E.L.; Blanco, M.A.; Gu, L.; Sendinc, E.; Liu, J.; Aristizábal-Corrales, D.; Hsu, C.H.; Aravind, L.; He, C.; Shi, Y. DNA Methylation on N6-Adenine in C. Elegans. Cell 2015, 161, 868–878. [Google Scholar] [CrossRef] [PubMed]
  118. Fu, Y.; Luo, G.Z.; Chen, K.; Deng, X.; Yu, M.; Han, D.; Hao, Z.; Liu, J.; Lu, X.; Doré, L.C.; et al. N6-Methyldeoxyadenosine Marks Active Transcription Start Sites in Chlamydomonas. Cell 2015, 161, 879–892. [Google Scholar] [CrossRef]
  119. Zhang, G.; Huang, H.; Liu, D.; Cheng, Y.; Liu, X.; Zhang, W.; Yin, R.; Zhang, D.; Zhang, P.; Liu, J.; et al. N6-Methyladenine DNA Modification in Drosophila. Cell 2015, 161, 893–906. [Google Scholar] [CrossRef]
  120. Wang, Y.; Chen, X.; Sheng, Y.; Liu, Y.; Gao, S. N6-Adenine DNA Methylation Is Associated with the Linker DNA of H2A.Z-Containing Well-Positioned Nucleosomes in Pol II-Transcribed Genes in Tetrahymena. Nucleic Acids Res. 2017, 45, 11594–11606. [Google Scholar] [CrossRef]
  121. Mondo, S.J.; Dannebaum, R.O.; Kuo, R.C.; Louie, K.B.; Bewick, A.J.; LaButti, K.; Haridas, S.; Kuo, A.; Salamov, A.; Ahrendt, S.R.; et al. Widespread Adenine N6-Methylation of Active Genes in Fungi. Nat. Genet. 2017, 49, 964–968. [Google Scholar] [CrossRef]
  122. Liang, Z.; Yu, G.; Liu, J.; Geng, Y.; Mao, J.; Wang, D.; Zhou, J.; Gu, X. The N6-Adenine Methylation in Yeast Genome Profiled by Single-Molecule Technology. J. Genet. Genom. 2018, 45, 223–225. [Google Scholar] [CrossRef]
  123. Wang, X.; Li, Z.; Zhang, Q.; Li, B.; Lu, C.; Li, W.; Cheng, T.; Xia, Q.; Zhao, P. DNA Methylation on N6-Adenine in Lepidopteran Bombyx Mori. Biochim. Biophys. Acta Gene Regul. Mech. 2018, 1861, 815–825. [Google Scholar] [CrossRef]
  124. Wu, T.P.; Wang, T.; Seetin, M.G.; Lai, Y.; Zhu, S.; Lin, K.; Liu, Y.; Byrum, S.D.; Mackintosh, S.G.; Zhong, M.; et al. DNA Methylation on N-Adenine in Mammalian Embryonic Stem Cells. Nature 2016, 532, 329–333. [Google Scholar] [CrossRef]
  125. Yao, B.; Cheng, Y.; Wang, Z.; Li, Y.; Chen, L.; Huang, L.; Zhang, W.; Chen, D.; Wu, H.; Tang, B.; et al. DNA N6-Methyladenine Is Dynamically Regulated in the Mouse Brain Following Environmental Stress. Nat. Commun. 2017, 8, 1–10. [Google Scholar] [CrossRef] [PubMed]
  126. Xiao, C.-L.; Zhu, S.; He, M.-H.; Chen, Y.; Yu, G.-L.; Chen, D.; Xie, S.-Q.; Luo, F.; Liang, Z.; Gu, X.-F.; et al. N6-Methyladenine DNA Modification in Human Genome. Mol. Cell 2018, 71, 306–318. [Google Scholar] [CrossRef] [PubMed]
  127. Liang, Z.; Shen, L.; Cui, X.; Bao, S.; Geng, Y.; Yu, G.; Liang, F.; Xie, S.; Lu, T.; Gu, X.; et al. DNA N-Adenine Methylation in Arabidopsis thaliana. Dev. Cell 2018, 45, 406–416. [Google Scholar] [CrossRef] [PubMed]
  128. Liang, Z.; Shen, L.; Cui, X.; Bao, S.; Geng, Y.; Yu, G.; Liang, F.; Lu, T.; Gu, X.; Yu, H. DNA N6-Adenine Methylation in Arabidopsis thaliana. Dev. Cell 2017, 45, 406–416. [Google Scholar] [CrossRef]
  129. Zhou, C.; Wang, C.; Liu, H.; Zhou, Q.; Liu, Q.; Guo, Y.; Peng, T.; Song, J.; Zhang, J.; Chen, L.; et al. Identification and Analysis of Adenine N6-Methylation Sites in the Rice Genome. Nat. Plants 2018, 4, 554–563. [Google Scholar] [CrossRef]
  130. Zhang, Q.; Liang, Z.; Cui, X.; Ji, C.; Li, Y.; Zhang, P.; Liu, J.; Riaz, A.; Yao, P.; Liu, M.; et al. N(6)-Methyladenine DNA Methylation in Japonica and Indica Rice Genomes and Its Association with Gene Expression, Plant Development and Stress Responses. Mol. Plant. 2018, 11, 1492–1508. [Google Scholar] [CrossRef]
  131. Heyn, H.; Esteller, M. An Adenine Code for DNA: A Second Life for N6-Methyladenine. Cell 2015, 161, 710–713. [Google Scholar] [CrossRef]
  132. Luo, G.Z.; Blanco, M.A.; Greer, E.L.; He, C.; Shi, Y. DNA N6-Methyladenine: A New Epigenetic Mark in Eukaryotes? Nat. Rev. Mol. Cell Biol. 2015, 16, 705–710. [Google Scholar] [CrossRef]
  133. Fedoreyeva, L.I.; Vanyushin, B.F. N6-Adenine DNA-Methyltransferase in Wheat Seedlings. FEBS Lett. 2002, 514, 305–308. [Google Scholar] [CrossRef]
  134. Koziol, M.J.; Bradshaw, C.R.; Allen, G.E.; Costa, A.S.H.; Frezza, C.; Gurdon, J.B. Identification of Methylated Deoxyadenosines in Vertebrates Reveals Diversity in DNA Modifications. Nat. Struct. Mol. Biol. 2016, 23, 24–30. [Google Scholar] [CrossRef]
  135. Dhar, M.S.; Pethe, V.V.; Gupta, V.S.; Ranjekar, P.K. Predominance and Tissue Specificity of Adenine Methylation in Rice. Theor. Appl. Genet. 1990, 80, 402–408. [Google Scholar] [CrossRef] [PubMed]
  136. Ashapkin, V.V.; Kutueva, L.I.; Vanyushin, B.F. The Gene for Domains Rearranged Methyltransferase (DRM2) in Arabidopsis thaliana Plants Is Methylated at Both Cytosine and Adenine Residues. FEBS Lett. 2002, 532, 367–372. [Google Scholar] [CrossRef]
  137. Schübeler, D. Function and Information Content of DNA Methylation. Nature 2015, 517, 321–326. [Google Scholar] [CrossRef] [PubMed]
  138. Rodrigues, J.A.; Zilberman, D. Evolution and Function of Genomic Imprinting in Plants. Genes Dev. 2015, 29, 2517–2531. [Google Scholar] [CrossRef] [PubMed]
  139. Vanyushin, B.F. JSM Genetics & Genomics Enzymatic DNA Methylation: What It Is Needed for in the Cell. JSM Genet. Genom. 2016, 3, 1010–1014. [Google Scholar]
  140. Feng, S.; Jacobsen, S.E.; Reik, W. Epigenetic Reprogramming in Plant and Animal Development. Science 2010, 330, 622–627. [Google Scholar] [CrossRef]
  141. Charlet, J.; Duymich, C.E.; Lay, F.D.; Mundbjerg, K.; Dalsgaard Sørensen, K.; Liang, G.; Jones, P.A. Bivalent Regions of Cytosine Methylation and H3K27 Acetylation Suggest an Active Role for DNA Methylation at Enhancers. Mol. Cell 2016, 62, 422–431. [Google Scholar] [CrossRef]
  142. Lang, Z.; Wang, Y.; Tang, K.; Tang, D.; Datsenka, T.; Cheng, J.; Zhang, Y.; Handa, A.K.; Zhu, J.-K. Critical Roles of DNA Demethylation in the Activation of Ripening-Induced Genes and Inhibition of Ripening-Repressed Genes in Tomato Fruit. Proc. Natl. Acad. Sci. USA 2017, 114, E4511–E4519. [Google Scholar] [CrossRef]
  143. Domcke, S.; Bardet, A.F.; Adrian Ginno, P.; Hartl, D.; Burger, L.; Schübeler, D. Competition between DNA Methylation and Transcription Factors Determines Binding of NRF1. Nature 2015, 528, 575–579. [Google Scholar] [CrossRef]
  144. Liang, Z.; Riaz, A.; Chachar, S.; Ding, Y.; Du, H.; Gu, X. Epigenetic modifications of mRNA and DNA methylation in plants. Mol. Plant 2020, 13, 14–30. [Google Scholar] [CrossRef]
  145. Cokus, S.J.; Feng, S.; Zhang, X.; Chen, Z.; Merriman, B.; Haudenschild, C.D.; Pradhan, S.; Nelson, S.F.; Pellegrini, M.; Jacobsen, S.E. Shotgun Bisulphite Sequencing of the Arabidopsis Genome Reveals DNA Methylation Patterning. Nature 2008, 452, 215–219. [Google Scholar] [CrossRef] [PubMed]
  146. Finnegan, E.J.; Peacock, W.J.; Dennis, E.S. Reduced DNA Methylation in Arabidopsis thaliana Results in Abnormal Plant Development. Proc. Natl. Acad. Sci. USA 1996, 93, 8449–8454. [Google Scholar] [CrossRef]
  147. Xiao, W.; Gehring, M.; Choi, Y.; Margossian, L.; Pu, H.; Harada, J.J.; Goldberg, R.B.; Pennell, R.I.; Fischer, R.L. Imprinting of the MEA Polycomb Gene Is Controlled by Antagonism between MET1 Methyltransferase and DME Glycosylase. Dev. Cell 2003, 5, 891–901. [Google Scholar] [CrossRef]
  148. Xiao, W.; Custard, K.D.; Brown, R.C.; Lemmon, B.E.; Harada, J.J.; Goldberg, R.B.; Fischer, R.L. DNA Methylation Is Critical for Arabidopsis Embryogenesis and Seed Viability. Plant. Cell 2006, 18, 805–814. [Google Scholar] [CrossRef] [PubMed]
  149. Rea, M.; Zheng, W.; Chen, M.; Braud, C.; Bhangu, D.; Rognan, T.N.; Xiao, W. Histone H1 Affects Gene Imprinting and DNA Methylation in Arabidopsis. Plant. J. 2012, 71, 776–786. [Google Scholar] [CrossRef] [PubMed]
  150. Jacobsen, S.E.; Sakai, H.; Finnegan, E.J.; Cao, X.; Meyerowitz, E.M. Ectopic Hypermethylation of Flower-Specific Genes in Arabidopsis. Curr. Biol. 2000, 10, 179–186. [Google Scholar] [CrossRef]
  151. Kim, M.; Ohr, H.; Lee, J.W.; Hyun, Y.; Fischer, R.L.; Choi, Y. Temporal and Spatial Downregulation of Arabidopsis MET1 Activity Results in Global DNA Hypomethylation and Developmental Defects. Mol. Cells 2014, 17, 280. [Google Scholar] [CrossRef]
  152. Choi, Y.; Gehring, M.; Johnson, L.; Hannon, M.; Harada, J.J.; Goldberg, R.B.; Jacobsen, S.E.; Fischer, R.L. DEMETER, a DNA Glycosylase Domain Protein, Is Required for Endosperm Gene Imprinting and Seed Viability in Arabidopsis. Cell 2002, 110, 33–42. [Google Scholar] [CrossRef]
  153. Kawakatsu, T.; Nery, J.R.; Castanon, R.; Ecker, J.R. Dynamic DNA Methylation Reconfiguration during Seed Development and Germination. Genome Biol. 2017, 18, 1–12. [Google Scholar] [CrossRef]
  154. Candaele, J.; Demuynck, K.; Mosoti, D.; Beemster, G.T.S.; Inze, D.; Nelissen, H. Differential Methylation during Maize Leaf Growth Targets Developmentally Regulated Genes. Plant Physiol. 2014, 164, 1350–1364. [Google Scholar] [CrossRef]
  155. Shen, H.; He, H.; Li, J.; Chen, W.; Wang, X.; Guo, L.; Peng, Z.; He, G.; Zhong, S.; Qi, Y. Genome-Wide Analysis of DNA Methylation and Gene Expression Changes in Two Arabidopsis Ecotypes and Their Reciprocal Hybrids. Plant Cell 2012, 24, 875–892. [Google Scholar] [CrossRef]
  156. Msogoya, T.J.; Grout, B.W.; Roberts, A. Reduction in Genome Size and DNA Methylation Alters Plant and Fruit Development in Tissue Culture Induced Off-Type Banana (Musa spp.). J. Anim. Plant. Sci. 2011, 3, 1450–1456. [Google Scholar]
  157. Zhong, S.; Fei, Z.; Chen, Y.-R.; Zheng, Y.; Huang, M.; Vrebalov, J.; McQuinn, R.; Gapper, N.; Liu, B.; Xiang, J. Single-Base Resolution Methylomes of Tomato Fruit Development Reveal Epigenome Modifications Associated with Ripening. Nat. Biotechnol. 2013, 31, 154–159. [Google Scholar] [CrossRef]
  158. Liu, R.; How-Kit, A.; Stammitti, L.; Teyssier, E.; Rolin, D.; Mortain-Bertrand, A.; Halle, S.; Liu, M.; Kong, J.; Wu, C. A DEMETER-like DNA Demethylase Governs Tomato Fruit Ripening. Proc. Natl. Acad. Sci. USA 2015, 112, 10804–10809. [Google Scholar] [CrossRef]
  159. Xu, J.; Xu, H.; Xu, Q.; Deng, X. Characterization of DNA Methylation Variations During Fruit Development and Ripening of Sweet Orange. Plant. Mol. Biol. Rep. 2015, 33, 1–11. [Google Scholar] [CrossRef]
  160. Baulcombe, D.C.; Dean, C. Epigenetic Regulation in Plant Responses to the Environment. Cold Spring Harb. Perspect. Biol. 2014, 6, a019471. [Google Scholar] [CrossRef] [PubMed]
  161. Probst, A.V.; Scheid, O.M. Stress-Induced Structural Changes in Plant Chromatin. Curr. Opin. Plant Biol. 2015, 27, 8–16. [Google Scholar] [CrossRef] [PubMed]
  162. Xiao, W.; Brown, R.C.; Lemmon, B.E.; Harada, J.J.; Goldberg, R.B.; Fischer, R.L. Regulation of Seed Size by Hypomethylation of Maternal and Paternal Genomes. Plant Physiol. 2006, 142, 1160–1168. [Google Scholar] [CrossRef]
  163. Furner, I.J.; Matzke, M. Methylation and Demethylation of the Arabidopsis Genome. Curr. Opin. Plant Biol. 2011, 14, 137–141. [Google Scholar] [CrossRef] [PubMed]
  164. Meyer, P. DNA Methylation Systems and Targets in Plants. FEBS Lett. 2011, 585, 2008–2015. [Google Scholar] [CrossRef] [PubMed]
  165. Xing, M.-Q.; Zhang, Y.; Zhou, S.-R.; Hu, W.-Y.; Wu, X.-T.; Ye, Y.-J.; Xiao, Y.-P.; Li, X.; Xue, H.-W. Global Analysis Reveals the Crucial Roles of DNA Methylation during Rice Seed Development. Plant Physiol. 2015, 168, 1417–1432. [Google Scholar] [CrossRef] [PubMed]
  166. An, Y.C.; Goettel, W.; Han, Q.; Bartels, A.; Liu, Z.; Xiao, W. Dynamic Changes of Genome-Wide DNA Methylation during Soybean Seed Development. Sci. Rep. 2017, 7, 12263. [Google Scholar] [CrossRef] [PubMed]
  167. Kou, H.P.; Li, Y.; Song, X.X.; Ou, X.F.; Xing, S.C.; Ma, J.; von Wettstein, D.; Liu, B. Heritable Alteration in DNA Methylation Induced by Nitrogen-Deficiency Stress Accompanies Enhanced Tolerance by Progenies to the Stress in Rice (Oryza sativa L.). J. Plant. Physiol. 2011, 168, 1685–1693. [Google Scholar] [CrossRef] [PubMed]
  168. Wang, X.; Xin, C.; Cai, J.; Zhou, Q.; Dai, T.; Cao, W.; Jiang, D. Heat Priming Induces Trans-Generational Tolerance to High Temperature Stress in Wheat. Front. Plant Sci. 2016, 7, 501. [Google Scholar] [CrossRef]
  169. Kathiria, P.; Sidler, C.; Golubov, A.; Kalischuk, M.; Kawchuk, L.M.; Kovalchuk, I. Tobacco Mosaic Virus Infection Results in an Increase in Recombination Frequency and Resistance to Viral, Bacterial, and Fungal Pathogens in the Progeny of Infected Tobacco Plants. Plant Physiol. 2010, 153, 1859–1870. [Google Scholar] [CrossRef] [PubMed]
  170. Verhoeven, K.J.F.; Jansen, J.J.; van Dijk, P.J.; Biere, A. Stress-induced DNA Methylation Changes and Their Heritability in Asexual Dandelions. New Phytol. 2010, 185, 1108–1118. [Google Scholar] [CrossRef]
  171. Johnsen, Ø.; Dæhlen, O.G.; Østreng, G.; Skrøppa, T. Daylength and Temperature during Seed Production Interactively Affect Adaptive Performance of Picea Abies Progenies. New Phytol. 2005, 168, 589–596. [Google Scholar] [CrossRef]
  172. Johannes, F.; Porcher, E.; Teixeira, F.K.; Saliba-Colombani, V.; Simon, M.; Agier, N.; Bulski, A.; Albuisson, J.; Heredia, F.; Audigier, P. Assessing the Impact of Transgenerational Epigenetic Variation on Complex Traits. PLoS Genet. 2009, 5, e1000530. [Google Scholar] [CrossRef]
  173. Schmid, M.W.; Heichinger, C.; Coman Schmid, D.; Guthörl, D.; Gagliardini, V.; Bruggmann, R.; Aluri, S.; Aquino, C.; Schmid, B.; Turnbull, L.A. Contribution of Epigenetic Variation to Adaptation in Arabidopsis. Nat. Commun. 2018, 9, 1–12. [Google Scholar] [CrossRef]
  174. Herman, J.J.; Sultan, S.E. DNA Methylation Mediates Genetic Variation for Adaptive Transgenerational Plasticity. Proc. R. Soc. B Biol. Sci. 2016, 283, 20160988. [Google Scholar] [CrossRef]
  175. Mercé, C.; Bayer, P.E.; Tay Fernandez, C.; Batley, J.; Edwards, D. Induced Methylation in Plants as a Crop Improvement Tool: Progress and Perspectives. Agronomy 2020, 10, 1484. [Google Scholar] [CrossRef]
  176. Molinier, J.; Ries, G.; Zipfel, C.; Hohn, B. Transgeneration Memory of Stress in Plants. Nature 2006, 442, 1046–1049. [Google Scholar] [CrossRef] [PubMed]
  177. Lukens, L.N.; Zhan, S. The Plant Genome’s Methylation Status and Response to Stress: Implications for Plant Improvement. Curr. Opin. Plant Biol. 2007, 10, 317–322. [Google Scholar] [CrossRef] [PubMed]
  178. Bossdorf, O.; Richards, C.L.; Pigliucci, M. Epigenetics for Ecologists. Ecol. Lett. 2008, 11, 106–115. [Google Scholar] [CrossRef] [PubMed]
  179. Fieldes, M.A.; Schaeffer, S.M.; Krech, M.J.; Brown, J.C.L. DNA Hypomethylation in 5-Azacytidine-Induced Early-Flowering Lines of Flax. Theor. Appl. Genet. 2005, 111, 136–149. [Google Scholar] [CrossRef]
  180. Fieldes, M.A. Heritable Effects of 5-Azacytidine Treatments on the Growth and Development of Flax (Linum usitatissimum) Genotrophs and Genotypes. Genome 1994, 37, 1–11. [Google Scholar] [CrossRef]
  181. Boyko, A.; Kathiria, P.; Zemp, F.J.; Yao, Y.; Pogribny, I.; Kovalchuk, I. Transgenerational Changes in the Genome Stability and Methylation in Pathogen-Infected Plants: (Virus-Induced Plant Genome Instability). Nucleic Acids Res. 2007, 35, 1714–1725. [Google Scholar] [CrossRef]
  182. Akimoto, K.; Katakami, H.; Kim, H.-J.; Ogawa, E.; Sano, C.M.; Wada, Y.; Sano, H. Epigenetic Inheritance in Rice Plants. Ann. Bot. 2007, 100, 205–217. [Google Scholar] [CrossRef]
  183. Tani, E.; Polidoros, A.N.; Nianiou-Obeidat, I.; Tsaftaris, A.S. DNA Methylation Patterns Are Differently Affected by Planting Density in Maize Inbreds and Their Hybrids. Maydica 2005, 50, 19. [Google Scholar]
  184. Ashikawa, I. Surveying CpG Methylation at 5′-CCGG in the Genomes of Rice Cultivars. Plant. Mol. Biol. 2001, 45, 31–39. [Google Scholar] [CrossRef]
  185. Cervera, M.-T.; Ruiz-García, L.; Martinez-Zapater, J. Analysis of DNA Methylation in Arabidopsis thaliana Based on Methylation-Sensitive AFLP Markers. Mol. Genet. Genom. 2002, 268, 543–552. [Google Scholar] [CrossRef] [PubMed]
  186. Salmon, A.; Clotault, J.; Jenczewski, E.; Chable, V.; Manzanares-Dauleux, M.J. Brassica Oleracea Displays a High Level of DNA Methylation Polymorphism. Plant Sci. 2008, 174, 61–70. [Google Scholar] [CrossRef]
  187. Keyte, A.L.; Percifield, R.; Liu, B.; Wendel, J.F. Infraspecific DNA Methylation Polymorphism in Cotton (Gossypium hirsutum L.). J. Hered. 2006, 97, 444–450. [Google Scholar] [CrossRef]
  188. Cubas, P.; Vincent, C.; Coen, E. An Epigenetic Mutation Responsible for Natural Variation in Floral Symmetry. Nature 1999, 401, 157–161. [Google Scholar] [CrossRef] [PubMed]
  189. Kalisz, S.; Purugganan, M.D. Epialleles via DNA Methylation: Consequences for Plant Evolution. Trends Ecol. Evol. 2004, 19, 309–314. [Google Scholar] [CrossRef] [PubMed]
  190. Luo, S.; Preuss, D. Strand-Biased DNA Methylation Associated with Centromeric Regions in Arabidopsis. Proc. Natl. Acad. Sci. USA 2003, 100, 11133–11138. [Google Scholar] [CrossRef] [PubMed]
  191. Xiong, L.Z.; Xu, C.G.; Maroof, S.; Zhang, Q. Patterns of Cytosine Methylation in an Elite Rice Hybrid and Its Parental Lines, Detected by a Methylation-Sensitive Amplification Polymorphism Technique. Mol. Gen. Genet. 1999, 261, 439–446. [Google Scholar] [CrossRef] [PubMed]
  192. Zhao, X.; Chai, Y.; Liu, B. Epigenetic Inheritance and Variation of DNA Methylation Level and Pattern in Maize Intra-Specific Hybrids. Plant. Sci. 2007, 172, 930–938. [Google Scholar] [CrossRef]
  193. Xu, R.; Li, H.; Qin, R.; Wang, L.; Li, L.; Wei, P.; Yang, J. Gene Targeting Using the Agrobacterium Tumefaciens-Mediated CRISPR-Cas System in Rice. Rice 2014, 7, 1–4. [Google Scholar] [CrossRef]
  194. Wang, S.; Zhang, S.; Wang, W.; Xiong, X.; Meng, F.; Cui, X. Efficient Targeted Mutagenesis in Potato by the CRISPR/Cas9 System. Plant Cell Rep. 2015, 34, 1473–1476. [Google Scholar] [CrossRef]
  195. Jia, H.; Wang, N. Targeted Genome Editing of Sweet Orange Using Cas9/SgRNA. PLoS ONE 2014, 9, e93806. [Google Scholar] [CrossRef] [PubMed]
  196. Cai, Y.; Chen, L.; Liu, X.; Sun, S.; Wu, C.; Jiang, B.; Han, T.; Hou, W. CRISPR/Cas9-Mediated Genome Editing in Soybean Hairy Roots. PLoS ONE 2015, 10, e0136064. [Google Scholar] [CrossRef] [PubMed]
  197. Brooks, C.; Nekrasov, V.; Lippman, Z.B.; van Eck, J. Efficient Gene Editing in Tomato in the First Generation Using the Clustered Regularly Interspaced Short Palindromic Repeats/CRISPR-Associated9 System. Plant Physiol. 2014, 166, 1292–1297. [Google Scholar] [CrossRef] [PubMed]
  198. Liang, Z.; Zhang, K.; Chen, K.; Gao, C. Targeted Mutagenesis in Zea Mays Using TALENs and the CRISPR/Cas System. J. Genet. Genom. 2014, 41, 63–68. [Google Scholar] [CrossRef] [PubMed]
  199. Chen, X.; Lu, X.; Shu, N.; Wang, S.; Wang, J.; Wang, D.; Guo, L.; Ye, W. Targeted Mutagenesis in Cotton (Gossypium hirsutum L.) Using the CRISPR/Cas9 System. Sci. Rep. 2017, 7, 1–7. [Google Scholar] [CrossRef]
  200. Mlambo, T.; Nitsch, S.; Hildenbeutel, M.; Romito, M.; Müller, M.; Bossen, C.; Diederichs, S.; Cornu, T.I.; Cathomen, T.; Mussolino, C. Designer Epigenome Modifiers Enable Robust and Sustained Gene Silencing in Clinically Relevant Human Cells. Nucleic Acids Res. 2018, 46, 4456–4468. [Google Scholar] [CrossRef] [PubMed]
  201. Springer, N.M. Epigenetics and Crop Improvement. Trends Genet. 2013, 29, 241–247. [Google Scholar] [CrossRef]
  202. Hegarty, M.J.; Barker, G.L.; Wilson, I.D.; Abbott, R.J.; Edwards, K.J.; Hiscock, S.J. Transcriptome Shock after Interspecific Hybridization in Senecio Is Ameliorated by Genome Duplication. Curr. Biol. 2006, 16, 1652–1659. [Google Scholar] [CrossRef]
  203. Buggs, R.J.A.; Zhang, L.; Miles, N.; Tate, J.A.; Gao, L.; Wei, W.; Schnable, P.S.; Barbazuk, W.B.; Soltis, P.S.; Soltis, D.E. Transcriptomic Shock Generates Evolutionary Novelty in a Newly Formed, Natural Allopolyploid Plant. Curr. Biol. 2011, 21, 551–556. [Google Scholar] [CrossRef]
  204. Chen, K.; Wang, Y.; Zhang, R.; Zhang, H.; Gao, C. CRISPR/Cas Genome Editing and Precision Plant Breeding in Agriculture. Annu. Rev. Plant. Biol. 2019, 70, 667–697. [Google Scholar] [CrossRef]
  205. Metje-Sprink, J.; Menz, J.; Modrzejewski, D.; Sprink, T. DNA-Free Genome Editing: Past, Present and Future. Front. Plant Sci. 2019, 9, 1957. [Google Scholar] [CrossRef]
  206. Agapito-Tenfen, S.Z.; Okoli, A.S.; Bernstein, M.J.; Wikmark, O.-G.; Myhr, A.I. Revisiting Risk Governance of GM Plants: The Need to Consider New and Emerging Gene-Editing Techniques. Front. Plant Sci. 2018, 9, 1874. [Google Scholar] [CrossRef] [PubMed]
  207. Davison, J.; Ammann, K. New GMO Regulations for Old: Determining a New Future for EU Crop Biotechnology. GM Crop. Food 2017, 8, 13–34. [Google Scholar] [CrossRef] [PubMed]
  208. Gayacharan; Joel, A.J. Epigenetic Responses to Drought Stress in Rice (Oryza sativa L.). Physiol. Mol. Biol. Plants 2013, 19, 379–387. [Google Scholar] [CrossRef] [PubMed]
  209. Mutum, R.D.; Balyan, S.C.; Kansal, S.; Agarwal, P.; Kumar, S.; Kumar, M.; Raghuvanshi, S. Evolution of Variety-specific Regulatory Schema for Expression of Osa-miR408 in Indica Rice Varieties under Drought Stress. FEBS J. 2013, 280, 1717–1730. [Google Scholar] [CrossRef] [PubMed]
  210. Wang, W.-S.; Pan, Y.-J.; Zhao, X.-Q.; Dwivedi, D.; Zhu, L.-H.; Ali, J.; Fu, B.-Y.; Li, Z.-K. Drought-Induced Site-Specific DNA Methylation and Its Association with Drought Tolerance in Rice (Oryza sativa L.). J. Exp. Bot. 2011, 62, 1951–1960. [Google Scholar] [CrossRef] [PubMed]
  211. Wang, W.; Zhao, X.; Pan, Y.; Zhu, L.; Fu, B.; Li, Z. DNA Methylation Changes Detected by Methylation-Sensitive Amplified Polymorphism in Two Contrasting Rice Genotypes under Salt Stress. J. Genet. Genom. 2011, 38, 419–424. [Google Scholar] [CrossRef]
  212. Forestan, C.; Farinati, S.; Rouster, J.; Lassagne, H.; Lauria, M.; Dal Ferro, N.; Varotto, S. Control of Maize Vegetative and Reproductive Development, Fertility, and RRNAs Silencing by HISTONE DEACETYLASE 108. Genetics 2018, 208, 1443–1466. [Google Scholar] [CrossRef]
  213. Xu, J.; Wang, Q.; Freeling, M.; Zhang, X.; Xu, Y.; Mao, Y.; Tang, X.; Wu, F.; Lan, H.; Cao, M. Natural Antisense Transcripts Are Significantly Involved in Regulation of Drought Stress in Maize. Nucleic Acids Res. 2017, 45, 5126–5141. [Google Scholar] [CrossRef]
  214. Sosa-Valencia, G.; Palomar, M.; Covarrubias, A.A.; Reyes, J.L. The Legume MiR1514a Modulates a NAC Transcription Factor Transcript to Trigger PhasiRNA Formation in Response to Drought. J. Exp. Bot. 2017, 68, 2013–2026. [Google Scholar] [CrossRef]
  215. Kulcheski, F.R.; Oliveira, L.F.; Molina, L.G.; Almerão, M.P.; Rodrigues, F.A.; Marcolino, J.; Abdelnoor, R.V. Identification of Novel Soybean MicroRNAs Involved in Abiotic and Biotic Stresses. BMC Genom. 2011, 12, 307. [Google Scholar] [CrossRef] [PubMed]
  216. Surdonja, K.; Eggert, K.; Hajirezaei, M.-R.; Harshavardhan, V.T.; Seiler, C.; von Wirén, N.; Sreenivasulu, N.; Kuhlmann, M. Increase of DNA Methylation at the HvCKX2. 1 Promoter by Terminal Drought Stress in Barley. Epigenomes 2017, 1, 9. [Google Scholar] [CrossRef]
  217. Temel, A.; Janack, B.; Humbeck, K. Drought Stress-Related Physiological Changes and Histone Modifications in Barley Primary Leaves at HSP17 Gene. Agronomy 2017, 7, 43. [Google Scholar] [CrossRef]
  218. Kantar, M.; Unver, T.; Budak, H. Regulation of Barley MiRNAs upon Dehydration Stress Correlated with Target Gene Expression. Funct. Integr. Genom. 2010, 10, 493–507. [Google Scholar] [CrossRef]
  219. Benoit, M.; Drost, H.-G.; Catoni, M.; Gouil, Q.; Lopez-Gomollon, S.; Baulcombe, D.; Paszkowski, J. Environmental and Epigenetic Regulation of Rider Retrotransposons in Tomato. PLoS Genet. 2019, 15, e1008370. [Google Scholar] [CrossRef]
  220. González, R.M.; Ricardi, M.M.; Iusem, N.D. Epigenetic Marks in an Adaptive Water Stress-Responsive Gene in Tomato Roots under Normal and Drought Conditions. Epigenetics 2013, 8, 864–872. [Google Scholar] [CrossRef]
  221. González, R.M.; Ricardi, M.M.; Iusem, N.D. Atypical Epigenetic Mark in an Atypical Location: Cytosine Methylation at Asymmetric (CNN) Sites within the Body of a Non-Repetitive Tomato Gene. BMC Plant Biol. 2011, 11, 1–11. [Google Scholar] [CrossRef]
  222. Labra, M.; Ghiani, A.; Citterio, S.; Sgorbati, S.; Sala, F.; Vannini, C.; Ruffini-Castiglione, M.; Bracale, M. Analysis of Cytosine Methylation Pattern in Response to Water Deficit in Pea Root Tips. Plant Biol. 2002, 4, 694–699. [Google Scholar] [CrossRef]
  223. Shui, X.-R.; Chen, Z.-W.; Li, J.-X. MicroRNA Prediction and Its Function in Regulating Drought-Related Genes in Cowpea. Plant Sci. 2013, 210, 25–35. [Google Scholar] [CrossRef]
  224. De la Rosa, C.; Covarrubias, A.A.; Reyes, J.L. A Dicistronic Precursor Encoding MiR398 and the Legume-specific MiR2119 Coregulates CSD1 and ADH1 MRNAs in Response to Water Deficit. Plant Cell Environ. 2019, 42, 133–144. [Google Scholar] [CrossRef]
  225. Abid, G.; Mingeot, D.; Muhovski, Y.; Mergeai, G.; Aouida, M.; Abdelkarim, S.; Aroua, I.; el Ayed, M.; M’hamdi, M.; Sassi, K. Analysis of DNA Methylation Patterns Associated with Drought Stress Response in Faba Bean (Vicia faba L.) Using Methylation-Sensitive Amplification Polymorphism (MSAP). Environ. Exp. Bot. 2017, 142, 34–44. [Google Scholar] [CrossRef]
  226. Arshad, M.; Gruber, M.; Hannoufa, A. Transcriptome Analysis of MicroRNA156 Overexpression Alfalfa Roots under Drought Stress. Sci. Rep. 2018, 8, 9363. [Google Scholar] [CrossRef] [PubMed]
  227. Hajyzadeh, M.; Turktas, M.; Khawar, K.M.; Unver, T. MiR408 Overexpression Causes Increased Drought Tolerance in Chickpea. Gene 2015, 555, 186–193. [Google Scholar] [CrossRef] [PubMed]
  228. Khandal, H.; Parween, S.; Roy, R.; Meena, M.K.; Chattopadhyay, D. MicroRNA Profiling Provides Insights into Post-Transcriptional Regulation of Gene Expression in Chickpea Root Apex under Salinity and Water Deficiency. Sci. Rep. 2017, 7, 1–14. [Google Scholar] [CrossRef]
  229. Kumar, S.; Beena, A.S.; Awana, M.; Singh, A. Physiological, Biochemical, Epigenetic and Molecular Analyses of Wheat (Triticum aestivum) Genotypes with Contrasting Salt Tolerance. Front. Plant Sci. 2017, 8, 1151. [Google Scholar] [CrossRef]
  230. Zhong, L.; Xu, Y.; Wang, J. DNA-Methylation Changes Induced by Salt Stress in Wheat Triticum Aestivum. Afr. J. Biotechnol. 2009, 8, 6201–6207. [Google Scholar]
  231. Zhu, N.; Cheng, S.; Liu, X.; Du, H.; Dai, M.; Zhou, D.-X.; Yang, W.; Zhao, Y. The R2R3-Type MYB Gene OsMYB91 Has a Function in Coordinating Plant Growth and Salt Stress Tolerance in Rice. Plant. Sci. 2015, 236, 146–156. [Google Scholar] [CrossRef]
  232. Karan, R.; DeLeon, T.; Biradar, H.; Subudhi, P.K. Salt Stress Induced Variation in DNA Methylation Pattern and Its Influence on Gene Expression in Contrasting Rice Genotypes. PLoS ONE 2012, 7, e40203. [Google Scholar] [CrossRef]
  233. Ferreira, L.J.; Donoghue, M.T.A.; Barros, P.; Saibo, N.J.; Santos, A.P.; Oliveira, M.M. Uncovering Differentially Methylated Regions (DMRs) in a Salt-Tolerant Rice Variety under Stress: One Step towards New Regulatory Regions for Enhanced Salt Tolerance. Epigenomes 2019, 3, 4. [Google Scholar] [CrossRef]
  234. Marconi, G.; Pace, R.; Traini, A.; Raggi, L.; Lutts, S.; Chiusano, M.; Guiducci, M.; Falcinelli, M.; Benincasa, P.; Albertini, E. Use of MSAP Markers to Analyse the Effects of Salt Stress on DNA Methylation in Rapeseed (Brassica napus Var. Oleifera). PLoS ONE 2013, 8, e75597. [Google Scholar] [CrossRef]
  235. Sako, K.; Kim, J.-M.; Matsui, A.; Nakamura, K.; Tanaka, M.; Kobayashi, M.; Saito, K.; Nishino, N.; Kusano, M.; Taji, T.; et al. Ky-2, a Histone Deacetylase Inhibitor, Enhances High-Salinity Stress Tolerance in Arabidopsis thaliana. Plant Cell Physiol. 2016, 57, 776–783. [Google Scholar] [CrossRef] [PubMed]
  236. Hou, H.; Zhao, L.; Zheng, X.; Gautam, M.; Yue, M.; Hou, J.; Chen, Z.; Wang, P.; Li, L. Dynamic Changes in Histone Modification Are Associated with Upregulation of Hsf and RRNA Genes during Heat Stress in Maize Seedlings. Protoplasma 2019, 256, 1245–1256. [Google Scholar] [CrossRef] [PubMed]
  237. Wang, N.; Ku, L.; Chen, Y.; Wang, W. Comparative Proteomic Analysis of Leaves between Photoperiod-Sensitive and Photoperiod-Insensitive Maize Inbred Seedlings under Long Day Treatments. Acta Physiol. Plant. 2015, 37, 1–7. [Google Scholar] [CrossRef]
  238. Wang, Y.; Li, H.; Sun, Q.; Yao, Y. Characterization of Small RNAs Derived from TRNAs, RRNAs and SnoRNAs and Their Response to Heat Stress in Wheat Seedlings. PLoS ONE 2016, 11, e0150933. [Google Scholar] [CrossRef]
  239. Hossain, M.S.; Kawakatsu, T.; Kim, K.D.; Zhang, N.; Nguyen, C.T.; Khan, S.M.; Batek, J.M.; Joshi, T.; Schmutz, J.; Grimwood, J. Divergent Cytosine DNA Methylation Patterns in Single-cell, Soybean Root Hairs. New Phytol. 2017, 214, 808–819. [Google Scholar] [CrossRef]
  240. Gao, G.; Li, J.; Li, H.; Li, F.; Xu, K.; Yan, G.; Chen, B.; Qiao, J.; Wu, X. Comparison of the Heat Stress Induced Variations in DNA Methylation between Heat-Tolerant and Heat-Sensitive Rapeseed Seedlings. Breed. Sci. 2014, 64, 125–133. [Google Scholar] [CrossRef]
  241. Hu, X.; Wu, X.; Li, C.; Lu, M.; Liu, T.; Wang, Y.; Wang, W. Abscisic Acid Refines the Synthesis of Chloroplast Proteins in Maize (Zea Mays) in Response to Drought and Light. PLoS ONE 2012, 7, e49500. [Google Scholar] [CrossRef]
  242. Hu, X.; Lu, M.; Li, C.; Liu, T.; Wang, W.; Wu, J.; Tai, F.; Li, X.; Zhang, J. Differential Expression of Proteins in Maize Roots in Response to Abscisic Acid and Drought. Acta Physiol. Plant. 2011, 33, 2437–2446. [Google Scholar] [CrossRef]
  243. Steward, N.; Ito, M.; Yamaguchi, Y.; Koizumi, N.; Sano, H. Periodic DNA Methylation in Maize Nucleosomes and Demethylation by Environmental Stress. J. Biol. Chem. 2002, 277, 37741–37746. [Google Scholar] [CrossRef]
  244. Zhang, B.; Tieman, D.M.; Jiao, C.; Xu, Y.; Chen, K.; Fei, Z.; Giovannoni, J.J.; Klee, H.J. Chilling-Induced Tomato Flavor Loss Is Associated with Altered Volatile Synthesis and Transient Changes in DNA Methylation. Proc. Natl. Acad. Sci. USA 2016, 113, 12580–12585. [Google Scholar] [CrossRef]
  245. Habu, Y.; Kakutani, T.; Paszkowski, J. Epigenetic Developmental Mechanisms in Plants: Molecules and Targets of Plant Epigenetic Regulation. Curr. Opin. Genet. Dev. 2001, 11, 215–220. [Google Scholar] [CrossRef]
  246. Kakutani, T. Epi-Alleles in Plants: Inheritance of Epigenetic Information over Generations. Plant Cell Physiol. 2002, 43, 1106–1111. [Google Scholar] [CrossRef] [PubMed]
  247. Messeguer, R.; Ganal, M.W.; Steffens, J.C.; Tanksley, S.D. Characterization of the Level, Target Sites and Inheritance of Cytosine Methylation in Tomato Nuclear DNA. Plant Mol. Biol. 1991, 16, 753–770. [Google Scholar] [CrossRef] [PubMed]
  248. Fieldes, M.A.; Amyot, L.M. Epigenetic Control of Early Flowering in Flax Lines Induced by 5-Azacytidine Applied to Germinating Seed. J. Hered. 1999, 90, 199–206. [Google Scholar] [CrossRef]
  249. Bird, A. DNA Methylation Patterns and Epigenetic Memory. Genes Dev. 2002, 16, 6–21. [Google Scholar] [CrossRef]
  250. Ehrlich, M.; Lacey, M. DNA Hypomethylation and Hemimethylation in Cancer. Epigenetic Alter. Oncog. 2013, 754, 31–56. [Google Scholar]
  251. Gehring, M.; Henikoff, S. DNA Methylation Dynamics in Plant Genomes. Biochim. Biophys. Acta 2007, 1769, 276–286. [Google Scholar] [CrossRef]
  252. Aina, R.; Sgorbati, S.; Santagostino, A.; Labra, M.; Ghiani, A.; Citterio, S. Specific Hypomethylation of DNA Is Induced by Heavy Metals in White Clover and Industrial Hemp. Physiol. Plant. 2004, 121, 472–480. [Google Scholar] [CrossRef]
  253. Galaud, J.; Gaspar, T.; Boyer, N. Inhibition of Internode Growth Due to Mechanical Stress in Bryonia Dioica: Relationship between Changes in DNA Methylation and Ethylene Metabolism. Physiol. Plant. 1993, 87, 25–30. [Google Scholar] [CrossRef]
  254. Hashida, S.; Kitamura, K.; Mikami, T.; Kishima, Y. Temperature Shift Coordinately Changes the Activity and the Methylation State of Transposon Tam3 in Antirrhinum Majus. Plant Physiol. 2003, 132, 1207–1216. [Google Scholar] [CrossRef]
  255. Peng, H.; Zhang, J. Plant Genomic DNA Methylation in Response to Stresses: Potential Applications and Challenges in Plant Breeding. Prog. Nat. Sci. 2009, 19, 1037–1045. [Google Scholar] [CrossRef]
  256. Erdmann, R.M.; Picard, C.L. RNA-Directed DNA Methylation. PLoS Genet. 2020, 16, e1009034. [Google Scholar] [CrossRef] [PubMed]
  257. Bond, D.M.; Baulcombe, D.C. Epigenetic Transitions Leading to Heritable, RNA-Mediated de Novo Silencing in Arabidopsis thaliana. Proc. Natl. Acad. Sci. USA 2015, 112, 917–922. [Google Scholar] [CrossRef] [PubMed]
  258. Kanazawa, A.; Inaba, J.; Shimura, H.; Otagaki, S.; Tsukahara, S.; Matsuzawa, A.; Kim, B.M.; Goto, K.; Masuta, C. Virus-mediated Efficient Induction of Epigenetic Modifications of Endogenous Genes with Phenotypic Changes in Plants. Plant J. 2011, 65, 156–168. [Google Scholar] [CrossRef]
  259. Liu, W.; Zhang, Y.; He, H.; He, G.; Deng, X.W. From Hybrid Genomes to Heterotic Trait Output: Challenges and Opportunities. Curr. Opin. Plant Biol. 2022, 66, 102193. [Google Scholar] [CrossRef]
  260. Chen, Z.J. Genomic and Epigenetic Insights into the Molecular Bases of Heterosis. Nat. Rev. Genet. 2013, 14, 471–482. [Google Scholar] [CrossRef]
  261. He, G.; He, H.; Deng, X.W. Epigenetic Variations in Plant Hybrids and Their Potential Roles in Heterosis. J. Genet. Genom. 2013, 40, 205–210. [Google Scholar] [CrossRef]
  262. Chodavarapu, R.K.; Feng, S.; Ding, B.; Simon, S.A.; Lopez, D.; Jia, Y.; Wang, G.-L.; Meyers, B.C.; Jacobsen, S.E.; Pellegrini, M. Transcriptome and Methylome Interactions in Rice Hybrids. Proc. Natl. Acad. Sci. USA 2012, 109, 12040–12045. [Google Scholar] [CrossRef]
  263. Wang, X.; Elling, A.A.; Li, X.; Li, N.; Peng, Z.; He, G.; Sun, H.; Qi, Y.; Liu, X.S.; Deng, X.W. Genome-Wide and Organ-Specific Landscapes of Epigenetic Modifications and Their Relationships to MRNA and Small RNA Transcriptomes in Maize. Plant Cell 2009, 21, 1053–1069. [Google Scholar] [CrossRef]
  264. Sinha, P.; Singh, V.K.; Saxena, R.K.; Kale, S.M.; Li, Y.; Garg, V.; Meifang, T.; Khan, A.W.; Kim, K.D.; Chitikineni, A. Genome-wide Analysis of Epigenetic and Transcriptional Changes Associated with Heterosis in Pigeonpea. Plant Biotechnol. J. 2020, 18, 1697–1710. [Google Scholar] [CrossRef]
  265. Borges, F.; Donoghue, M.T.A.; LeBlanc, C.; Wear, E.E.; Tanurdžić, M.; Berube, B.; Brooks, A.; Thompson, W.F.; Hanley-Bowdoin, L.; Martienssen, R.A. Loss of Small-RNA-Directed DNA Methylation in the Plant Cell Cycle Promotes Germline Reprogramming and Somaclonal Variation. Curr. Biol. 2021, 31, 591–600. [Google Scholar] [CrossRef] [PubMed]
  266. Ma, X.; Xing, F.; Jia, Q.; Zhang, Q.; Hu, T.; Wu, B.; Shao, L.; Zhao, Y.; Zhang, Q.; Zhou, D.-X. Parental Variation in CHG Methylation Is Associated with Allelic-Specific Expression in Elite Hybrid Rice. Plant Physiol. 2021, 186, 1025–1041. [Google Scholar] [CrossRef] [PubMed]
  267. Zhou, S.; Xing, M.; Zhao, Z.; Gu, Y.; Xiao, Y.; Liu, Q.; Xue, H. DNA Methylation Modification in Heterosis Initiation through Analyzing Rice Hybrid Contemporary Seeds. Crop. J. 2021, 9, 1179–1190. [Google Scholar] [CrossRef]
  268. Lv, Z.; Zhang, W.; Wu, Y.; Huang, S.; Zhou, Y.; Zhang, A.; Deng, X.; Xu, C.; Xu, Z.; Gong, L. Extensive Allele-level Remodeling of Histone Methylation Modification in Reciprocal F1 Hybrids of Rice Subspecies. Plant J. 2019, 97, 571–586. [Google Scholar] [CrossRef] [PubMed]
  269. Qi, X.; Li, Z.H.; Jiang, L.L.; Yu, X.M.; Ngezahayo, F.; Liu, B. Grain-Yield Heterosis in Zea mays L. Shows Positive Correlation with Parental Difference in CHG Methylation. Crop Sci. 2010, 50, 2338–2346. [Google Scholar] [CrossRef]
  270. Jahnke, S.; Sarholz, B.; Thiemann, A.; Kühr, V.; Gutiérrez-Marcos, J.F.; Geiger, H.H.; Piepho, H.-P.; Scholten, S. Heterosis in Early Seed Development: A Comparative Study of F1 Embryo and Endosperm Tissues 6 Days after Fertilization. Theor. Appl. Genet. 2010, 120, 389–400. [Google Scholar] [CrossRef]
  271. Lin, J.-Y.; Le, B.H.; Chen, M.; Henry, K.F.; Hur, J.; Hsieh, T.-F.; Chen, P.-Y.; Pelletier, J.M.; Pellegrini, M.; Fischer, R.L. Similarity between Soybean and Arabidopsis Seed Methylomes and Loss of Non-CG Methylation Does Not Affect Seed Development. Proc. Natl. Acad. Sci. USA 2017, 114, E9730–E9739. [Google Scholar] [CrossRef]
  272. Cao, S.; Wang, L.; Han, T.; Ye, W.; Liu, Y.; Sun, Y.; Moose, S.P.; Song, Q.; Chen, Z.J. Small RNAs Mediate Transgenerational Inheritance of Genome-Wide Trans-Acting Epialleles in Maize. Genome Biol. 2022, 23, 1–25. [Google Scholar] [CrossRef]
  273. Tao, Z.; Shen, L.; Gu, X.; Wang, Y.; Yu, H.; He, Y. Embryonic Epigenetic Reprogramming by a Pioneer Transcription Factor in Plants. Nature 2017, 551, 124–128. [Google Scholar] [CrossRef]
  274. Chen, L.; Zhu, Y.; Ren, X.; Yao, D.; Song, Y.; Fan, S.; Li, X.; Zhang, Z.; Yang, S.; Zhang, J. Heterosis and Differential DNA Methylation in Soybean Hybrids and Their Parental Lines. Plants 2022, 11, 1136. [Google Scholar] [CrossRef]
  275. Wada, Y.; Miyamoto, K.; Kusano, T.; Sano, H. Association between Up-Regulation of Stress-Responsive Genes and Hypomethylation of Genomic DNA in Tobacco Plants. Mol. Genet. Genom. 2004, 271, 658–666. [Google Scholar] [CrossRef] [PubMed]
  276. Tatra, G.S.; Miranda, J.; Chinnappa, C.C.; Reid, D.M. Effect of Light Quality and 5-azacytidine on Genomic Methylation and Stem Elongation in Two Ecotypes of Stellaria Longipes. Physiol. Plant. 2000, 109, 313–321. [Google Scholar] [CrossRef]
  277. Fieldes, M.A.; Harvey, C.G. Differences in Developmental Programming and Node Number at Flowering in the 5-Azacytidine-Induced Early Flowering Flax Lines and Their Controls. Int. J. Plant. Sci. 2004, 165, 695–706. [Google Scholar] [CrossRef]
  278. Peng, H.; Zhang, J.; Wu, X. The Ploidy Effects in Plant Gene Expression: Progress, Problems and Prospects. Sci. China Ser. C Life Sci. 2008, 51, 295–301. [Google Scholar] [CrossRef] [PubMed]
  279. Johnson, D.S.; Mortazavi, A.; Myers, R.M.; Wold, B. Genome-Wide Mapping of in Vivo Protein-DNA Interactions. Science 2007, 316, 1497–1502. [Google Scholar] [CrossRef]
  280. Zilberman, D.; Henikoff, S. Genome-Wide Analysis of DNA Methylation Patterns. Development 2007, 134, 3959–3965. [Google Scholar] [CrossRef]
  281. Dyachenko, O.V.; Zakharchenko, N.S.; Shevchuk, T.V.; Bohnert, H.J.; Cushman, J.C.; Buryanov, Y.I. Effect of Hypermethylation of CCWGG Sequences in DNA of Mesembryanthemum Crystallinum Plants on Their Adaptation to Salt Stress. Biochemistry 2006, 71, 461–465. [Google Scholar] [CrossRef]
  282. Yang, L.; Zhang, P.; Wang, Y.; Hu, G.; Guo, W.; Gu, X.; Pu, L. Plant synthetic epigenomic engineering for crop improvement. Sci. China Life Sci. 2022. Available online: https://engine.scichina.com/doi/10.1007/s11427-021-2131-6 (accessed on 10 June 2022).
Figure 1. Maintenance of methylation and demethylation in plants.
Figure 1. Maintenance of methylation and demethylation in plants.
Horticulturae 08 00562 g001
Table 1. Summary of writers, erasers and readers of 5mC and 6mA.
Table 1. Summary of writers, erasers and readers of 5mC and 6mA.
ModificationSubstrate SpecificityPutative FunctionsExample Proteins/Domains
HumanArabidopsisRice
5mCCytosineRepression
Writer DNMT1MET1
DNMT3CMT3
CMT2
DRM2
Eraser TET1ROS1
TET2DME
TDGDML2
DML3
Reader MECP2SUVH2SUVH7
MBDSUVH9
6mADeoxyadenosineActivation
Writer N6AMT1AtN6AMT1OsN6AMT1
Eraser ALKBH1AtALKBH1OsALKBH1
Table 2. Known epigenetic mechanisms involved in different abiotic stresses in plants.
Table 2. Known epigenetic mechanisms involved in different abiotic stresses in plants.
Abiotic StressCropEpigenetic Mechanism(s)Reference
DroughtRiceHypomethylation[208]
Up-regulation of miR408 expression[209]
Site-specific DNA methylation[210,211]
MaizeModifications of H3K4me3 and H3K9ac dynamics[212]
Enrichment in H3K36me3, H3K9ac, and H3K4me3[213]
SoybeanmiR1514a modulation of a NAC transcription factor transcript[214]
Up-regulation of isomiRNAs[215]
BarleyHc-siRNA-mediated hyper-methylation at CYTOKININ-OXIDASE 2.1 promoter[216]
Increase in H3 and loss in H3K9me2[217]
Accumulation of miR408 transcripts[218]
TomatoRNA-dependent DNA methylation[219]
Increased Asr1 and Asr2 expression due to demethylation of putative regulatory and transcribed regions[220,221]
PeaHypermethylation of cytosine residues[221]
CowpeaIncrease of P5CS transcripts and very low expression of vun-miR5021 and vun-miR156b-3p[222]
BeanDicistronic arrangement of miR398a and miR2119[223]
Faba beanIncreased DNA demethylation[224]
AlfalfaOverexpression of miR156[225]
ChickpeaAccumulation of miR408 transcripts[226]
Accumulation of miRNAs at root apex[227]
SalinityWheatHypermethylation of cytosines at HKT genes[228]
5-mC depletion[229]
RiceDemethylation at promoter region of OsMYB91 gene and rapid histone modifications at OsMYB9 locus[230]
DNA methylation[231,232]
RapeseedIncreased DNA demethylation[233]
ChickpeaAccumulation of miRNAs at root apex[227]
ArabidopsisIncreased acetylation of histone H4 at AtSOS1 due to inhibition of de-acetylase[234]
HeatMaizeH3K4me2 and H3K9ac alterations[235]
Increased histone acetylation and decreased H3K9me3[236]
WheatIncreased histone demethylation of the various genes[237]
SoyabeanHypomethylation of cytosine[238]
RapeseedIncreased DNA demethylation[239]
ColdMaizeEnrichment in H3K9ac and decrease in DNA methylation and H3K9me2[240]
Reduction in histone acetylation in euchromatin-associated gene regions[241]
DNA demethylation[242]
TomatoIncreased DNA methylation[243]
ArabidopsisNon-CG hypermethylation under cold and low light stress[244]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shaikh, A.A.; Chachar, S.; Chachar, M.; Ahmed, N.; Guan, C.; Zhang, P. Recent Advances in DNA Methylation and Their Potential Breeding Applications in Plants. Horticulturae 2022, 8, 562. https://doi.org/10.3390/horticulturae8070562

AMA Style

Shaikh AA, Chachar S, Chachar M, Ahmed N, Guan C, Zhang P. Recent Advances in DNA Methylation and Their Potential Breeding Applications in Plants. Horticulturae. 2022; 8(7):562. https://doi.org/10.3390/horticulturae8070562

Chicago/Turabian Style

Shaikh, Aamir Ali, Sadaruddin Chachar, Muzafaruddin Chachar, Nazir Ahmed, Changfei Guan, and Pingxian Zhang. 2022. "Recent Advances in DNA Methylation and Their Potential Breeding Applications in Plants" Horticulturae 8, no. 7: 562. https://doi.org/10.3390/horticulturae8070562

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