Next Article in Journal
Monocarbonyl Analogs of Curcumin Based on the Pseudopelletierine Scaffold: Synthesis and Anti-Inflammatory Activity
Next Article in Special Issue
Characterisation of LTR-Retrotransposons of Stevia rebaudiana and Their Use for the Analysis of Genetic Variability
Previous Article in Journal
Insight into Factors Influencing Wound Healing Using Phosphorylated Cellulose-Filled-Chitosan Nanocomposite Films
 
 
Correction published on 15 November 2022, see Int. J. Mol. Sci. 2022, 23(22), 14107.
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Dynamism of Transposon Methylation for Plant Development and Stress Adaptation

by
Muthusamy Ramakrishnan
1,2,
Lakkakula Satish
3,
Ruslan Kalendar
4,5,
Mathiyazhagan Narayanan
6,
Sabariswaran Kandasamy
7,
Anket Sharma
8,9,
Abolghassem Emamverdian
1,2,
Qiang Wei
1,2,*,† and
Mingbing Zhou
9,10,*,†
1
Co-Innovation Center for Sustainable Forestry in Southern China, Nanjing Forestry University, Nanjing 210037, China
2
Bamboo Research Institute, Nanjing Forestry University, Nanjing 210037, China
3
Department of Biotechnology Engineering, & The Jacob Blaustein Institutes for Desert Research, Ben-Gurion University of the Negev, Beer Sheva 84105, Israel
4
Helsinki Institute of Life Science HiLIFE, Biocenter 3, Viikinkaari 1, University of Helsinki, FI-00014 Helsinki, Finland
5
National Laboratory Astana, Nazarbayev University, Nur-Sultan 010000, Kazakhstan
6
PG and Research Centre in Biotechnology, MGR College, Adhiyamaan Educational Research Institute, Hosur 635 109, Tamil Nadu, India
7
Institute for Energy Research, Jiangsu University, Zhenjiang 212013, China
8
Department of Plant Science and Landscape Architecture, University of Maryland, College Park, MD 20742, USA
9
State Key Laboratory of Subtropical Silviculture, Zhejiang A&F University, Lin’an, Hangzhou 311300, China
10
Zhejiang Provincial Collaborative Innovation Center for Bamboo Resources and High-Efficiency Utilization, Zhejiang A&F University, Lin’an, Hangzhou 311300, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2021, 22(21), 11387; https://doi.org/10.3390/ijms222111387
Submission received: 30 July 2021 / Revised: 13 October 2021 / Accepted: 19 October 2021 / Published: 21 October 2021 / Corrected: 15 November 2022
(This article belongs to the Special Issue Transposable Elements and Phenotypic Variation in Plants)

Abstract

:
Plant development processes are regulated by epigenetic alterations that shape nuclear structure, gene expression, and phenotypic plasticity; these alterations can provide the plant with protection from environmental stresses. During plant growth and development, these processes play a significant role in regulating gene expression to remodel chromatin structure. These epigenetic alterations are mainly regulated by transposable elements (TEs) whose abundance in plant genomes results in their interaction with genomes. Thus, TEs are the main source of epigenetic changes and form a substantial part of the plant genome. Furthermore, TEs can be activated under stress conditions, and activated elements cause mutagenic effects and substantial genetic variability. This introduces novel gene functions and structural variation in the insertion sites and primarily contributes to epigenetic modifications. Altogether, these modifications indirectly or directly provide the ability to withstand environmental stresses. In recent years, many studies have shown that TE methylation plays a major role in the evolution of the plant genome through epigenetic process that regulate gene imprinting, thereby upholding genome stability. The induced genetic rearrangements and insertions of mobile genetic elements in regions of active euchromatin contribute to genome alteration, leading to genomic stress. These TE-mediated epigenetic modifications lead to phenotypic diversity, genetic variation, and environmental stress tolerance. Thus, TE methylation is essential for plant evolution and stress adaptation, and TEs hold a relevant military position in the plant genome. High-throughput techniques have greatly advanced the understanding of TE-mediated gene expression and its associations with genome methylation and suggest that controlled mobilization of TEs could be used for crop breeding. However, development application in this area has been limited, and an integrated view of TE function and subsequent processes is lacking. In this review, we explore the enormous diversity and likely functions of the TE repertoire in adaptive evolution and discuss some recent examples of how TEs impact gene expression in plant development and stress adaptation.

1. Introduction

Transposable elements (TEs), also known as jumping genes or mobile genetic elements, are key players in plant biological systems and genome evolution [1,2,3,4,5]. TEs were previously considered as genomic parasites since these self-replicating entities are ubiquitous [6,7] and abundant in nature [8]. In recent years, several evolutionary studies in eukaryote genomes emphasized the biological significance of TEs in animals and plant genomes [9,10,11]. For example, in mammals [12] and in the model organism Drosophila [13], TEs have a major role in disseminating cis-regulatory elements that help the host genome regulate its own genes both in the short-term (adaptation to environmental changes) and long-term (evolutionary changes). Furthermore, TEs act as key factors in diverse genetic mechanisms, such as chromosomal changes related to recombination processes of mobile genetic elements and other elements, regulation and expression of genes, genomic evolution, and genetic instability (Figure 1) [14,15,16]. TE transpositions may even cause mutations that lead to novel functional protein-coding sequences [17,18]. For example, Rag1 and Rag2 are TE-derived conserved genes that catalyse V(D)J somatic recombination in the vertebrate immune system [19,20]. As a consequence of the biological significance of TEs, TEs have recently been used as an integration tool in fundamental research [21] and in gene therapy [22]. TEs, or parts thereof, can also be implemented into common molecular biology tools, such as expression vectors [23]. In addition, TEs have been suggested as new markers (together with mitochondrial polymorphisms and Y-chromosome polymorphisms) to describe the evolutionary history of a species, or even of single individuals [24,25].
However, TEs are the most erratic components in plants and are species-dependent [26,27,28,29]. The host applies several strategies to control TE activities to avoid potential deleterious actions by other TEs, such as retrotransposon elements (RTEs). While most of the long terminal repeat (LTR) RTEs were recently inserted in most plant genomes, these insertions are unique in the genome. For example, some RTEs are transcriptionally inactive under normal conditions, but under different stress conditions, most of the RTEs are active [30]. The flexible genomic alterations in RTEs can be considered suitable for most plant adaptation mechanisms under various stresses, including biotic and abiotic stress [31,32,33]. However, plants possess a potent response that restrains TE activity, leading to epigenetic silencing of these elements, which results in alteration in plant gene function [15,34,35,36]. For instance, in the African oil palm (Elaeis guineensis), DNA hypomethylation of a LINE (non-LTR RTEs), related to rice Karma, is linked with alternative splicing and yield loss, whereas hypermethylation near the Karma splice enhanced the normal fruit set [37]. Typically, TE insertion did not impact the genome or related biomolecular products because of TE silencing [38]. For instance, in Arabidopsis and corn (Zea mays), methylation of mutated TEs is not harmful to the genome [26,33,39,40]. TE silencing is caused by miRNAs or epigenetic mechanisms, such as DNA methylation or chromatin remodelling [38,41]. The addition of a methyl group to the cytosine bases of DNA to generate 5-methylcytosine is called DNA methylation [42].
Figure 1. Primary regulatory roles of transposable elements (TEs). TEs are a rich source of host genome innovations. TE functions are either harmful or beneficial to the host genome, and their integration in the genome may induce deleterious mutations. Silenced TEs, mostly covered with DNA methylation, can affect the expression of nearby genes. In contrast, active TEs can act as regulatory elements by producing noncoding RNA (ncRNA) and alternative promoters. The illustration was adapted and redrawn from Jönsson et al. [43], with copyright permission from the Licensor Elsevier (Trends in Genetics: Cell Press publisher) and Copyright Clearance Center (https://www.copyright.com) (Supplementary File S1).
Figure 1. Primary regulatory roles of transposable elements (TEs). TEs are a rich source of host genome innovations. TE functions are either harmful or beneficial to the host genome, and their integration in the genome may induce deleterious mutations. Silenced TEs, mostly covered with DNA methylation, can affect the expression of nearby genes. In contrast, active TEs can act as regulatory elements by producing noncoding RNA (ncRNA) and alternative promoters. The illustration was adapted and redrawn from Jönsson et al. [43], with copyright permission from the Licensor Elsevier (Trends in Genetics: Cell Press publisher) and Copyright Clearance Center (https://www.copyright.com) (Supplementary File S1).
Ijms 22 11387 g001
Among several epigenetic mechanisms, DNA methylation and chromatin remodelling are more commonly implicated in the inactivation of TEs in plants and animals [40,44,45,46,47,48]. TEs are transcribed steadily in methylation-deficient plants and cause mutant phenotypes that are directly linked to TE insertion [14,42]. The other most significant epigenetic mechanism is chromatin remodelling. The altered chromatin structure results in constricted chromatin at the particular site of the genome where genes and transposons are inactivated, as the RNA polymerase is unable to access those sites. For example, in Arabidopsis, decondensed chromatin regulates the expression of small RNAs to help maintain TE methylome homeostasis during post-embryogenesis [49]. Hence, most elements are not transcribed [50]. Nevertheless, further investigations are required to understand the possible mechanisms of TEs involved in plant evolutionary processes and stress adaptation mechanisms. This review addresses TE methylation mechanisms and their significance in plant evolution and stress adaptation.

2. TE Classification and Copy Number in Plants

According to TE structure, the plant evolves and adapts as a consequence of dynamic changes in the TE. Based on the method of transposition (movement), TEs are classified into two major classes, class I and class II (Table 1) [51]. Class II (DNA transposons) are usually present in low copy numbers and are mobilized through a DNA intermediate by “cut-and-paste” mechanisms [52], as in the case of the Helitrons transposon, which is a “peel-and-paste” replicative mechanism via a circular DNA intermediate [53]. Class I transposons or RTEs are mobilized by copy-and-paste using RNA as an intermediate, whereby RNA is reverse-transcribed into cDNA then integrated into a target site of the genome [54,55,56].
Based on its structure and mechanism of integration, RTEs are further divided into different superfamilies, such as long terminal repeat (LTR) RTEs, non-LTR RTEs, and dictyostelium repetitive sequences (DIRS) [30]. LTR RTEs are the most common superfamily, contributing up to 80% of plant genome size [57], and have significantly higher copy numbers than other superfamilies and classes (Table 1).
According to Wicker et al. [51], class I (retrotransposons) do not require subclasses but superfamilies. However, class II transposons are classified into two subclasses distinguished by the number of DNA strands and do not move via an RNA intermediate. Each subclass is further classified into different superfamilies and families, with wide variations in the organization, but with shared common genetic structures and monophyletic origin. For example, the families of Ty3/gypsy and Ty1/copia are superfamilies of LTR RTEs found in virtually all major groups of eukaryotes [58]. Similarly, Tcl/mariner, hATs (hobo-Ac-Tam3), and MULEs (Mutator-like elements) are subclasses of DNA transposons that are widespread in eukaryotes [59]. Although conversion to the wild-type sequence at the insertion site can occur upon transposition, many types of transposons leave a detectable footprint upon mobilization. However, the net excision of the donor site of cut-and-paste transposons is generally challenging to detect since the donor site is converted to a normal site either by using a homolog as a template or a sister chromatid [59].
Both class I and class II TEs have autonomous (containing open reading frames, ORFs) and non-autonomous (absence of encoding potential while lacking transposition ability) TEs [12,23,60,61]. Class II autonomous TEs can encode transposase and helicase enzymes for cut-and-paste mechanisms [62]. Class I autonomous TEs can encode specialized Gag packaging proteins and reverse transcriptase for transposition [1]. The transposition-competent TEs have not only coding ability but also bear intact cis-acting elements that interact with the transposition complexes. LTRs (class I) and terminal-inverted repeats (TIRs) (class II) are examples of such cis-acting elements. Thus, autonomous elements are not dependent on any other factors for their movement [33], whereas non-autonomous TEs depend on autonomous TEs to migrate. However, non-autonomous elements can still express transposition-related proteins while lacking transposition ability [61]. For example, Ac (Activator) TEs can translocate their position as they are autonomous. In contrast, Ds (Dissociation) TEs are non-autonomous and can only be transposed by the availability of Ac or any other autonomous element [63]. The continuous transposition of TEs in the plant genome leads to significant evolutionary changes, constant divergences, and integrations that result in, as yet, uncharacterized variations in TE forms and shapes [3].
Table 1. Class- and family-wise examples of transposable elements (TEs) in different plant species. The table was adapted and recreated from Feschotte et al. [64], with copyright permission from the Licensor Springer Nature (Nature Reviews Genetics: Nature publisher) and Copyright Clearance Center (https://www.copyright.com) (Supplementary File S2).
Table 1. Class- and family-wise examples of transposable elements (TEs) in different plant species. The table was adapted and recreated from Feschotte et al. [64], with copyright permission from the Licensor Springer Nature (Nature Reviews Genetics: Nature publisher) and Copyright Clearance Center (https://www.copyright.com) (Supplementary File S2).
ClassSubclassSuperfamily/FamilyPlantsAutonomous MembersNon-Autonomous MembersCopy Number of the Entire FamilyReferences
Class I LTR Retrotransposonscopia-likeO. sativaTos17-(2–5) 30[65]
copia-likeHordeum sp.BARE-1-5000–22,000[66]
copia-likeN. tabacumTto1-30 (300)[67]
copia-likeN. tabacumTnt1A->100[68]
copia-likeZ. maysHopscotch-5–8[69]
copia-likeZ. mays-BS11–5[70]
copia-likeZ. maysOpie-2-100,000[71]
gypsy-likeO. sativaRIRE2Dasheng1200[72]
gypsy-likeZ. maysMagellan-4–8[73]
gypsy-likeZ. maysHuck-2-200,000[71]
gypsy-likeArabidopsisAthila 4-22[74]
gypsy-likeArabidopsisTa3-1[75]
gypsy-likeArabidopsisAthila 6-11[74]
gypsy-likeArabidopsisTar17-2[67]
Non-LTR RetrotransposonsLINEs; L1-cladeLilium speciosumDel2-250,000[76]
LINEs; L1-cladeZ. maysCin4-50–100[77]
LINEs; L1-cladeArabidopsisTal1-1–6[78]
SINEsN. tabacum-TS50,000[79]
SINEsB. napus-S1500[80]
Class IIDNA transposons MutatorZ. maysMuDRMu110–100[81]
MutatorArabidopsisAtMu1-1 (4)[82]
CACTAZ. maysSpmdSpm50–100[83]
CACTAArabidopsisCAC1CAC2(4) 20[84]
hATZ. maysAcDs50–100[85]
PIF/HarbingerZ. maysPIFamPIF6000[86]
PIF/HarbingerAngiospermsPIF-likeTourist-likeVariable[86,87]
Tc1/MarinerAngiospermsMLEsStowaway-likeVariable[88,89]
The copy numbers indicated are approximate and collected from various research articles. Those in parentheses result from transpositional activation (Tos17 and Tto1) or in mutant backgrounds (CAC and AtMu1). LINE: Long Interspersed Nuclear Element, SINE: Short Interspersed Nuclear Element, Ac: Activator; LTR: Long Terminal Repeat, MLE: Mariner-Like Element, Ds: Dissociation, mPIF: miniature P Instability Factor, Spm: SuppressorMutator, PIF: P Instability Factor.

3. Surprising Traits of TEs

In plants, TEs are located within or near the gene or promoters. The position of the TE determines plant gene expression and other regulatory mechanisms for growth and development and stress adaptation. TEs are aligned at a suitable location in the genome through transposition [90]. The aligned position should positively interact with the organelles of the cell [91]. This location-based, genomic-level adaptation through these various shapes of TEs surprised plant biologists by their outstanding genomic parasitism, optimistic competition, and cooperation with other cellular processes [15,92]. Another significant and surprising property of TEs is the spectrum of site selection for transposition in the plant genome [12]. However, the TE selection mechanism in the genome is still unclear as TEs insertion sites are not detrimental and not strongly counter-selected [93]. This indicates that natural selection and genetics are the most significant and forceful genome-shaping factors, acting through the adequate distribution and accumulation of various TEs in the plant genome [94]. Under certain circumstances, this insertion by transposition could cause positive effects that isolate the species from native populations.
In most cases, the insertion will have little or no effect on gene activity. In some cases, such insertions might alter gene expression such that the plant is better adapted to environmental and ecological conditions. The impact of such insertions might differ significantly among species [95]. The position of some TEs in the genome is more stable than that of other TEs. This genome stability is directly related to the forces of selection [96,97,98]. Such properties of various classes of TEs have shaped the genomes of plant species, thereby maintaining genome stability and function. A clear understanding of how natural forces of selection impact the transposition of TEs in the plant genome can provide valuable insights into evolutionary processes in plant biological systems.

4. Contribution of TEs in the Plant Genome

The average genomic fraction occupied by TEs in plant genomes is about 50% of the entire genome. This percentage can range from 15% in small to >85% in large plant genomes (Table 2) [99,100,101]. RTEs occupy a significant portion of the plant genome and are the most significant factor in the plant genome, thus contributing to plant growth [99,102] (Table 2). This variation was reported by researchers [103,104], who examined the possible relationship between LTR-RTEs and the total physical length of the plant genome. The total genomic content of plant species is a linear function of TE content. Thus, LTR-RTEs are significant components of the plant genome and contribute to the genome differences among plants [105].
The proportion of RTEs in the total genome of several plant species is directly correlated [100,106,107]. For example, the total proportion of RTEs in the total genome of Arabidopsis is 14% (total genome size: 125 Mb) [108]; it is 35% in Oryza sativa (total genome size 389 Mb) [109], and 85% in Zea mays (total genome size: 2.3 Gb) [110]. Among these plant species, Z. mays contains more RTEs than any other plant species investigated thus far (Table 2) [99]. Hence, the existence of an excess volume of RTEs in Z. mays has gradually increased (doubled) the total genome size in the past 3 million years due to the swift propagation of RTE families [99,111]. Similarly, the genome size of O. australiensis has doubled due to the rapid proliferation of three LTR-RTEs families (RIRE1, Kangourou, and Wallabi) [112].
Table 2. Proportion of class I and class II transposable elements (TEs) in the total genome of different plant species [99,100,101,102,104,110,113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129]. The table was adapted and recreated from Ragupathy et al. [99], with copyright permission from the Licensor Elsevier (Trends in Plant Science: Cell Press publisher) and Copyright Clearance Center (https://www.copyright.com) (Supplementary File S3).
Table 2. Proportion of class I and class II transposable elements (TEs) in the total genome of different plant species [99,100,101,102,104,110,113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129]. The table was adapted and recreated from Ragupathy et al. [99], with copyright permission from the Licensor Elsevier (Trends in Plant Science: Cell Press publisher) and Copyright Clearance Center (https://www.copyright.com) (Supplementary File S3).
Plant GenomeTotal Genome Size (Mb)Total TE Content (% of the Genome)Total Class I or RNA (Retroelements) (% of the Genome)Total Class II or DNA Transposons (% of the Genome)
Aegilops tauschii4.9868.2013.3053.50
Arabidopsis lyrata230.0029.7015.994.80
Arabidopsis thaliana125.0014.00–18.507.5011.00
Brachypodium distachyon355.0028.1023.334.77
Brassica oleracea600.0020.0014.006.00
Brassica rapa529.0039.5129.903.20
Cajanus cajan833.0051.6719.184.53
Carica papaya372.0051.9042.800.60
Cicer arietinum738.0049.4145.649.32
Citrus sinensis367.0020.5018.212.28
Cucumis melo450.0019.7014.705.00
Cucumis sativus367.0024.0112.161.24
Fragaria vesca240.0022.8116.376.44
Glycine max1115.0058.7442.2416.50
Gossypium herbaceum1660.0052.1052.000.10
Gossypium raimondii880.0056.9548.994.54
Gossypium raimondii880.0061.3054.901.50
Hordeum vulgare5100.0058.8952.835.25
Linum usitatissimum370.0024.2920.623.80
Lotus japonicus472.0030.8010.4–19.230.97–8.10
Malus domestica742.0042.4037.600.90
Medicago truncatula475.0038.009.60ND
Medicago truncatula550.0030.5026.503.40
Musa acuminata523.0032.6331.171.42
Oryza sativa389.0034.7919.3512.96
Phyllostachys edulis1908.0045.4538.207.25
Populus trichocarpa485.0042.0010.302.50
Populus trichocarpa550.0034.907.022.10
Pyrus bretschneideri527.0053.1045.9712.12
Ricinus communis320.0050.3318.160.91
Secale cereale8090.0069.3064.305.00
Setaria italica (Accession Zhang gu)510.0046.3031.609.40
Setaria italica (Inbred Yugu1)510.0040.0025.00ND
Solanum lycopersicum900.0063.2062.300.90
Solanum tuberosum844.0062.2032.293.94
Sorghum bicolor730.0062.0054.527.46
Theobroma cacao430.0025.7017.708.00
Vitis vinifera475.0041.4017.040.43
Zea mays2300.0084.2075.608.60

5. Distribution of TEs in the Plant Genome

Each TE is distributed in the plant genome with a specific insertion preference [130]. LTR- RTEs, such as the Ty3/gypsy and Ty1/copia superfamilies, are present in the centromere regions of the plant genome and play significant and perilous parts in the formation and function of centromeres [12,106,131]. In addition, Ty3/gypsy and Ty1/copia exhibit nested insertions, particularly in large genomes bearing a high number of elements and prefer older copies of the same family. This suggests that nesting of LTR-RTEs is not random and depends on chromatin modifications. Class II TEs can also lead to TE nesting, although nesting is common in LTR-RTEs [130].
Similarly, nonautonomous LTR-RTEs, such as Dasheng, are positioned in the pericentromeric regions of the genome of O. sativa [72]. The grapevine RTE 1 (Gret1) is a type of LTR retroelement. At the same time, the insertion and rearrangement of Gret1 in Vitis vinifera occurred close to the region of the VvmbyA1 gene, which led to development of colour variation in the fruit of Vitis vinifera [132]. Similarly, Rider is a type of LTR element. While Rider is inserted into another region, it acts as a novel regulatory element and enhances the expression of the Ruby gene, which leads to enhanced synthesis of anthocyanin production in the fruit of Citrus sinensis [133]. Consistently, the fruit shape of Solanum lycopersicum has been altered from round to oval due to the retroposition of the IQD12 gene [134]. In Arabidopsis (Landsberg erecta (Ler) accession early flowering), when mutator-like TEs are subjected to epigenetic modification, alteration in the first intron of Flowering Locus C (FLC) results in a delay in the flowering process [135]. Likewise, Ac/Ds are composed of autonomous and nonautonomous members of the maize hAT family, respectively. Ac/Ds can also stimulate structural rearrangements of other TEs in Z. mays [136,137,138] and can induce chromosomal rearrangements at the rice OsRLG5 locus [139].

6. TE-Induced Mutations

Active TEs induce heritable mutations in the genome that have been fully characterized at both the genetic and molecular levels. Several reports also state that TEs are mutagens and may be responsible for mutation through various means, such as by inserting themselves into active genes or near genes that contain promoter and enhancer elements. Although all active genes contain at least a promoter and many are influenced by enhancers, TE insertion still causes heritable mutations or alters gene activity [1]. Therefore, TEs are considered as the most potent natural evolutionary and adaptation mediators within the genome of plant species. TEs play a critical role in adaptation and new species formation by evolution, as TE insertions generate gene (DNA) rearrangements and can act as new coding and regulatory sequences (Figure 1) [140]. The high copy number (3000 to 10,000 per genome) of both classes (I and II) of TEs have site-specific (e.g., TAA or TA) insertions or transitions in plants. Tourist and stowaway elements belonging to MITEs in maize and sorghum, respectively, are preferably located at the 5′ and 3′ noncoding sections in the genes of these plants [141]. Furthermore, these elements are interconnected with the regulatory portion of genes in different flowering plants [142]. In cut-and-paste transposition, a faulty repair process may seal the gap formed during transposition. Moreover, identical repetitive sequences create a problem in the pairing process, especially during meiosis [106]. In some cases, TEs may insert the stopping codon that results in the production of truncated proteins [143].
Arabidopsis is a genetic model plant used for evolutionary biology and mutation-related studies and has significantly contributed to our TE research. However, an in-depth analysis of the active TEs of Arabidopsis mutation accumulation lines showed an absence of TE-induced direct mutation [144,145]. Surprisingly, study of Arabidopsis mutation accumulation lines revealed the limited scale of TE-induced mutations, which were approximately 1/haploid genome/generation. TEs involved in the insertion process could be analysed through purifying selection and population genomic analyses of polymorphic TEs, which provide a partial view of TE migration or transition [93,146].

7. Association of RTEs with Genomes

Approximately 7.5% to 75% of the genomes of many plant species consist of RTEs (Table 2), which play a vital role in the evolutionary process. According to recent studies on genome analyses, approximately 67% of the hexaploid Triticum aestivum (wheat) genome is made up of RTEs, which are primarily TEs of the class I Ty3/gypsy and Ty1/copia. The chromosome content of hexaploid wheat has been improved with highly repetitive RTE elements [147]. The latest assemblage of hexaploid wheat (bread wheat) enhanced the extremely recurring RTE elements positioned in the A, B, and D sub-genomes of the species. Wheat is an important crop where repetitive RTEs occupy approximately 67% of the genome, as RTEs undergo a large amplification process [147]. Moreover, the TE proportion is very similar in the A, B, and D sub-genomes, which evolved approximately two to three million years ago (Mya) (based on molecular dating of chloroplast DNA) [148]. This two to three-million-year evolution by rapid amplification of various RTEs led to the development of an intricate, surplus, and allohexaploid genetic material. These lengthy evolutionary processes by RTEs made the genetic material of wheat by far the most prevalent and most intricate in form in the plant kingdom.
RTEs associated with plant genomes may further show both positive and negative impacts on genomic and phenotypic activities, such as alterations in gene activity and genome organization. This occurs through amending gene expression, disrupting protein-coding regions, and stimulating chromosomal rearrangements at a large scale [149]. Such RTE activities may create a mutation that expels the particular plant from its population. For example, RTEs are the predominant source of cis-regulatory elements and cause rapid alteration in the transcriptional unit of various genes under biotic and abiotic stresses [17,150]. Moreover, large RTEs and related repetitive elements may be involved in DNA double-strand break repair mechanisms and enhance chromosomal rearrangements through translocations, inversions, duplications, and deletions [1,17,149].

8. Balance between TE Expression and Repression

To ensure survival, plants and other organisms must evolve and adapt to the surrounding biotic and abiotic stresses [151]. Large portions of the genomes of many organisms are composed of RTEs that balance the expression and repression of essential gene sequences [152]. TEs are usually assumed to insert anywhere in the genome, but some TEs are biased in their insertion locations to balance both expression and repression. For example, Athila RTEs and other RTEs are inserted in the pericentromeric regions and less proximal regions of the chromosome arms, respectively. This suggests that these regions could help balance TE expression and repression through epigenetic modification [5,153]. Furthermore, for successful evolution, regulatory elements with TE insertions should balance gene expression, as overexpression may be a disadvantage and increased copy numbers may be unusable [154,155]. Insufficiency of enzymes encoded by TEs may explain the insufficient quantity for the transposition process. For instance, transposition of Ppmar1 and Ppmar2 (Mariner-like elements (MLEs) isolated from Moso bamboo) is determined by the quantity of transposases present inside the nucleus [156,157]. This suggests that MLEs generally have the potential to develop a self-regulatory strategy that can control their amplification and copy numbers by minimization of transposases. This is a well-known regulation mechanism known as overproduction (or overexpression) inhibition [158]. TE expression or its transposition may also be influenced by some default factors, such as chromatin, DNA alteration pathways, small interfering RNAs (siRNA), specific gene repressors under abiotic stress [5]. For example, Wang et al. [159] performed an experiment on three strains of Arabidopsis to demonstrate the significance of siRNAs and epigenetic processes (such as DNA methylation) to identify the balance between the expression and repression of genes. They found an optimistic correlation and interspecific alteration in gene expression of TE sequence polymorphisms and the existence of associated TEs. Small gene (<2 kb) sequences that possess conserved TEs are more stable than larger TEs inserted into adjacent gene polymorphisms. siRNAs serve to repress TEs (stopping proliferation) situated near coding genes, which leads to strong suppression of adjacent gene expression [48].
In some cases, such as the pollen of flowering plant species, the host cell could employ a cohort cell (that does not pass hereditary information to subsequent generations) produced simultaneously during the meiosis process, which ensures TE repression [160]. Moreover, the balance of expression and repression of TEs is also determined, and their degrees vary among tissue types and with the age of the organism (i.e., stage of life cycle). Furthermore, TEs are expressed only in germline cells and not in the somatic cells in many plant species. Hence, TEs are retained in the germline (also called micronucleus) and are actively deleted from the somatic macronucleus [160].

9. TE Transposition and Genome Stability

TEs associated with genes are transposed into other sites of the same genome with transposase enzymes and TE transposition machinery. Moreover, TEs involved in this transposition process can exist as replicates or conservative in form. In replicative transposition (copy-and-paste process), TEs are copied and relocated in the same genome, leading to duplicate TEs in the genome [1,17,59,149]. The cut-and-paste process is involved in conservative transposition, in which TEs are excised from their original position and transposed to the new position in the same genome. In this cut-and-paste process, the adjacent sequence of a neighbour gene sequence can be cut and reinserted into a new site in the same genome; this phenomenon can also be called exon shuffling. This transposition can cause damage to the genome by disrupting the expression of critical genes [161,162].
The plant can silence transposition through various mechanisms, such as via mutations in TEs, epigenetic silencing (e.g., DNA methylation), and siRNA silencing [163]. In certain situations, the transposition properties of TEs may assist the plant species to rapidly adapt to biotic and abiotic stresses and expand genome size [150,164]. For example, a heat-activated RTE in Arabidopsis, ONSEN, increases abiotic stress tolerance through a mutation in an abscisic acid (ABA) responsive gene and epigenetic mechanisms [165]. Initially in the transposition process, RTE generates its transcription by reverse transcriptase and reintegration into the genome, a process termed retrotransposition. In both cases, the transposase enzyme is involved in the insertion of TEs at another site. In retrotransposition, RTEs inhabit approximately 74% of the 240-kb maize genomes (Adh region). These elements comprise 11 different families from 23 members of RTEs [4,166]. In the transposition process, insertion age correlates with the retrotransposition process, as the ends of RTEs are probably identical during the element insertion mechanisms [26].
Although Barbara McClintock discovered TEs approximately 70 years ago, several studies have revealed new information about TEs in both prokaryotes and eukaryotes. It is now recognized that the excision and insertion traits of TEs can cause genetic instability in both prokaryotes and eukaryotes, which can lead to genomic innovations and facilitate the emergence of new species [167]. The effects of TEs on genetic stability remains poorly understood. Available data suggest that the genomic instability of TEs has both positive and negative impacts on the host. For example, genomic instability can increase genetic diversity, give an optimistic outcome, facilitate evolution, and involve gene regulation [18,168]. In contrast, genomic instability in plants may also lead to unusable phenotypic changes, such as flowering, yield reduction, and reduction in stress tolerance [97,169,170,171].

10. TE Is the Source of Non-Coding RNAs (ncRNAs)

Non-coding RNAs (ncRNAs) are a group of various RNA complexes that act as key factors in regulating gene expression. Based on the source and mode of action, ncRNAs are classified into housekeeping ncRNAs (tRNAs, rRNAs, and snoRNAs) and regulatory ncRNAs. Moreover, regulatory ncRNAs are sub-classified into small ncRNAs (siRNAs and miRNAs) and long ncRNAs (intronic ncRNAs (incRNAs) and long intergenic ncRNAs (lincRNAs)) [172]. Several theories, such as duplication, pseudogenization of protein-coding sequences, double-stranded RNAs (dsRNAs) from heterochromatin regions, evolution (genomic) from existing transposons, replication of RNA viruses, and random hairpin structures have been proposed to explain the source of different ncRNAs, especially regulatory ncRNAs [172]. However, a significant amount of ncRNAs is transcribed from TEs [173]. These ncRNAs, especially regulatory ncRNAs, can modify RNA stability, prevent RNA translation, and, most importantly, play a key role in the modulation of gene expression at transcriptional and post-transcriptional levels [172]. Interestingly, recently published literature suggests that ncRNAs may be involved in various stress responses in plants [174,175]. For instance, siRNAs are involved in transcriptional and post-transcriptional processes [176].

11. Role of ncRNAs in Plant Response to Abiotic Stress

TEs influence phenotype through the production of ncRNAs, which play a significant role in responding to and balancing abiotic stress. Several recent research findings have revealed that the active expression of ncRNAs, either directly or indirectly, is involved in plant responses to abiotic stress [177]. miRNA expression might be enhanced or suppressed in response to different abiotic stresses [178]. For example, salt stress in Arabidopsis induces miR393 expression, and miR393 is involved in repression of lateral root initiation, emergence, and elongation and increases levels of reactive oxygen species (ROS) in the lateral root [179].
Similarly, siRNAs contribute significantly to abiotic stress responses. For example, in Arabidopsis, nat-siRNA, along with SRO5, regulate proline metabolism through pyrroline-5-carboxylate dehydrogenase (P5CDH), which reduces the increased production of ROS under high salt stress [180]. Similarly, lncRNAs from plants exhibit a significant mimicry response to different abiotic stresses [181]. lncRNAs serve as competitive endogenous RNAs (ceRNAs) that have been overwhelmed by miRNAs. Thus, lncRNAs inhibit the interaction of the original miRNA at the target site [182]. For example, in Arabidopsis grown under phosphate deficiency stress conditions, lncRNA IPS1 is activated to mimic miRNA399, which inhibits binding of native miRNA399s to their target site, such as in the case of PHO2 [183]. Similarly, various types of siRNAs and lncRNAs from various plants mediate responses to various abiotic stresses (Table 3) [172].
Table 3. Abiotic stress response mechanisms of non-coding RNAs (siRNAs and lncRNAs) from various plant species.
Table 3. Abiotic stress response mechanisms of non-coding RNAs (siRNAs and lncRNAs) from various plant species.
Plant SpeciessiRNA MechanismsAbiotic Stresses Induced/SuppressedReferences
ArabidopsisSRO5-P5CDH nat-siRNARegulation of proline metabolismSalt stress ↓[180]
ArabidopsisTAS1, TAS2, TAS3 ta-siRNAElevated expressionHypoxia stress ↑[184,185,186]
ArabidopsisHTT1, HTT2-TAS1NYEHeat stress ↑[187,188]
ArabidopsisTAS4 ta-siRNAsBiosynthesis of anthocyaninsPhosphate deficiency ↑[189,190]
ArabidopsisTAS4-siR81(-)Accumulation of anthocyaninNitrogen deficiency ↑[190]
ArabidopsishcsiRNAs (ONSEN)DNA methylationHeat stress ↑[191,192,193]
ArabidopsishcsiRNAs (HD2C, HDA6)DNA methylationDrought and ABA stresses ↑ and ↓[194,195,196,197,198,199]
ArabidopsisIPS1 *miR399 target mimicryPhosphate deficiency ↑ [183,200,201]
ArabidopsislncRNAs *Antisense transcriptionLight stress ↑[202]
ArabidopsisasHSFB2a *Antisense transcriptionHeat stress ↑[203]
ArabidopsisCOOLAIR *Chromatin remodellingCold stress ↑[204]
ArabidopsislncRNAs *Histone modificationLight stress ↑[202]
ArabidopsisCOLDAIR *Histone modificationCold stress ↑[205]
ArabidopsislncRNAs *RdDM pathwayHeat stress ↑[206]
ArabidopsislncRNAs *RdDM pathwaySalt stress ↓[207]
Brassica oleraceanat-siRNAsDNA methylationHeat stress ↑[208,209]
Brassica rapanat-siRNAsDNA methylationHeat stress ↑ and ↓[209]
Brassica rapalincRNAs *miRNA precursorsCold and heat stresses ↑ and ↓[210]
Craterostigma plantagineumCDT1-siRNANYEDehydration stress ↑[211]
Manihot esculenta2 nat-siRNA, 3 ta-siRNAsNYECold stress ↑ and ↓[212]
Oryza sativalncRNAs *target mimicryPhosphate deficiency ↑ and ↓[213]
Phaeodactylum tricornutumpti-MIR5472 *miR5472 precursorsPhosphate deficiency ↑[214]
Phaeodactylum tricornutumpti-MIR5471 *miR5471 precursorsPhosphate deficiency ↑ [214]
Populus tomentosalincRNAs *miRNA precursorsNitrogen deficiency ↑ and ↓[215]
Populus tomentosalincRNAs *Antisense transcriptionNitrogen deficiency ↑ and ↓[215]
Populus trichocarpalincRNA1128 *ptc-miR482a.1 target mimicryDrought stress ↓ [216]
Populus trichocarpalincRNA1393 *ptc-miR6459b target mimicryDrought stress ↓[216]
Populus trichocarpalincRNA3018 *ptc-miR399i target mimicryDrought stress ↓[216]
Populus trichocarpalincRNA2752 *ptc-miR169o target mimicryDrought stress ↑[216]
Populus trichocarpalincRNA1795 *ptc-miR476a target mimicryDrought stress ↓[216]
Populus trichocarpalincRNA20 *ptc-miR476a target mimicryDrought stress ↑[216]
Populus trichocarpalincRNA2623 *ptc-miR156k target mimicryDrought stress ↓[216]
Populus trichocarpalincRNA2623 *ptc-miR156c target mimicryDrought stress ↓[216]
Populus trichocarpalincRNA967 *ptc-miR6462e target mimicryNo response to drought stress [216]
Populus trichocarpalincRNA2762 *ptc-miR156k target mimicryDrought stress ↓[216]
Populus trichocarpalincRNA1449 *ptc-miR156k target mimicryNo response to drought stress [216]
Populus trichocarpalincRNA179 *ptc-miR156a target mimicryNo response to drought stress [216]
Populus trichocarpalincRNA2198 *, lincRNA2131 *, lincRNA2085 *, lincRNA2962 *
lincRNA1534 *, lincRNA1039 *
lincRNA2962 *
NYEDrought stress ↑[216]
Solanum lycopersicumlncRNAs *RdDM pathwaySalt and drought stresses ↓[217]
Triticum aestivum002061_0636_3054.1 siRNANYEHeat, NaCl, and dehydration ↓[218]
Triticum aestivum005047_0654_1904.1 siRNANYEHeat, NaCl, and dehydration ↓[218]
Triticum aestivum005047_0654_1904.1 siRNANYECold stress ↑[218]
Triticum aestivum080621_1340_ 0098.1 siRNANYECold stress ↑ and heat stress ↓[218]
Triticum aestivum007927_0100_2975.1 siRNANYECold, NaCl, and dehydration ↓[218]
Triticum aestivumta-siRNA TAS3a-50D6 (+)Auxin signalling pathwayCold stress ↑[219]
Triticum aestivumTalnRNA5 *ta-miR2004 precursorsHeat stress ↑[218,220]
Triticum aestivumTahlnRNA27 *ta-miR2010 precursorsHeat stress ↑[218,220]
Triticum aestivumTalnRNA21 *, TahlnRNA3 *, TahlnRNA14 *, TahlnRNA19 *
TahlnRNA36 *, TahlnRNA41 *
TahlnRNA42 *, TahlnRNA47 *
TahlnRNA52 *
siRNA precursorsHeat stress ↑[218,220]
Zea mayslncRNAs *siRNA precursors and antisense transcriptionDrought stress ↑[221]
Star symbol “*” indicates lncRNA; no symbol indicates siRNA. Up arrow “↑” indicates that siRANs/lncRNAs are enhanced in response to the corresponding abiotic stress while the down arrow “↓” indicates that siRANs/lncRNAs are suppressed in response to the corresponding abiotic stress. NYE indicates that the mechanism/process of that particular siRNA or lncRNA has not yet been established. RdDM, small RNA-directed DNA methylation.

12. Epigenetic Effects of TEs

As previously mentioned, all types of TEs from both classes have a unique role in genome instability and evolution and organism adaptation to abnormal conditions [222]. Nevertheless, insertion or transposition of TEs in normal conditions may cause harmful effects to organisms, including plants. Hence, under normal conditions (i.e., absence of mutations or biotic or abiotic stress), TEs are silenced or inactivated by epigenetic silencing mechanisms, such as DNA methylation or suppressive chromatin alterations (Figure 2) [223]. The epigenetic silencing process is more active in plants than in any other organisms. In this process, TEs can be in an inactive form, when the epigenetic silencing process is turned off, or in alleviated conditions, such as under mutant backgrounds and biotic or abiotic stress [104,224]. Recently, several studies have revealed that the promoter sequence of TEs enhances expression of genes situated nearby in plants and how this expression is controlled by epigenetic regulation, which mediates phenotypic diversity and adaptation (Figure 3) [150,225].
In some eukaryotic organisms, epigenetic effects can also participate in the proliferation and accumulation of TEs, leading to an enlargement in genome size, in which siRNA-mediated pathways can occur and end with DNA methylation in TEs [1].
In eukaryotes, biochemical modifications of DNA that lead to chromatin remodelling via histone binding are known as epigenetic modifications. These modifications provide information on gene regulation. In general, histone lysine and arginine residues are subjected to epigenetic modification. Several types of lysine residues (H3K4, H3K9, and H3K27 with mono/di/tri-methylation) have been extensively studied in animals and plants. Among these types, H3K9me2 is associated with TE methylation [33]. These suppressive epigenetic effects promote packaging of chromatin into compacted nuclear partitions of the cell [226]. In eukaryotes, especially in plants, the epigenetic silencing mechanisms directly act on TEs via the small RNA-directed DNA methylation (RdDM) pathway. Briefly, the siRNA matching regions of TEs are targeted by either AGO4 or AGO6 directed by siRNA. These targeted regions (scaffolding RNA) are transcribed by polymerase V [227,228]. These scaffolded dsRNA elements react with methyltransferases DRM1 and DRM2, leading to the methylation of TEs [226].

13. TE Methylation

As TEs possess self-replication potential and exist as genomic parasites, they can cause detrimental effects on essential active genes and generate ectopic recombination of DNA. These damaging effects can be avoided and controlled by epigenetic silencing, such as through DNA methylation [193]. siRNAs are interconnected with various TEs and act as mediators and stimulate DNA methylation [252]. This DNA methylation may lead to suppression of transposition through transcription reduction, along with the formation of loops among DNA and histone methylations (Figure 4) through siRNAs [253]. For example, siRNA-mediated epigenetic modification of TEs results in a delay of the flowering process in Arabidopsis [135]. This suggests that TE epigenetic modification regulates FLC expression. Hence, these siRNAs act as a strong substitute for DNA methylation in TEs, and siRNA-targeted TEs have strong effects on nearby gene transcription than those without. In some plant species, the cytosine methylation process occurs at CG, CHG, and CHH (H represents A, T, or C) sites of TEs. Most of these sites are unmethylated, and some sites (approximately 15%) are similar to DNA methylation patterns. Interestingly, siRNA-mediated DNA methylation can spread about 500 bp into unmethylated neighbouring TEs. In the case of DNA methylation in euchromatin TEs, it can spread approximately 200 bp beyond the siRNA target positions. This depends on the effect of siRNAs on the expression of proximal genes that are 400 bp in size [104,254,255].
In most cases, siRNA-mediated methylated TEs are probably situated fewer base pairs away from active genes than the location of unmethylated or partially methylated TEs. A possible reason for the partial methylation of TEs is the nucleotide composition of siRNAs. This phenomenon suggests that under unfavourable conditions, such as biotic or abiotic stress, active TEs are involved in the evolutionary process. In normal circumstances, TEs have been targeted by siRNAs for DNA methylation of cytosine to maintain genomic stability of the plant under usual conditions [42]. Moreover, to maintain TE methylome homeostasis in Arabidopsis, altered chromatin structure also increases siRNA production from heterochromatic TEs during post-embryogenesis [49].

14. TE Methylation in Plant Evolution

Since DNA methylation is positively correlated with repetitive sequences, such as RTEs and centromeric repeats non-randomly distributed across the entire plant genome, it is also enriched in centromeres in replicated regions [256]. Active TEs are mutagenic and disrupt genes, regulatory regions, and genome integrity. In contrast, the remaining new RTEs are silent and permanently or partially disabled [257]. One of the earliest known functions of the DNA methylation pathway is the inhibition of RTEs (Figure 3). In plant genomes, RTEs have significantly higher DNA methylation levels than non-coding regions (specifically CHG and CG) across all contexts [258], but some RTEs can easily escape host silencing by activating anti-silencing factors [30]. Maintenance of LTR-RTE silencing in Arabidopsis is based on a combination of RdDM and RNA-independent mechanisms. TE silencing accepts a distinct chromatin state. For instance, silent or increased histone H3K9 and DNA methylation in conjunction with H3 lysine results in TE suppression in Arabidopsis, thus protecting the genomes from TE transposition and genome instability [259]. This distinctive three-layered state of silent heterochromatin is distinguishable from the polycomb gene cluster transcribed and active heterochromatin gene expression and is linked to the Arabidopsis genome [260]. There can be several different reasons for the collapse of structures and reactivation of previously silenced TEs [261].
Changes in environmental conditions may lead to RTE reactivation. Alternatively, polyploidy and hybridization may cause another kind of systematic shock for RTE activation [262]. Polyploidy frequently occurs in plant genomes, making the periodic expansion of RTEs possible. For example, autopolyploidy promotes retention of TEs instead of eliminating them. Eukaryotic species seem to be linked to large population sizes, and small genomes are unusual for the few organisms known to have lost cytosine methylation. Active transpositions of DNA methylation may be less effectively eliminated in such populations [263]. Moreover, DNA methylation and gene expression patterns must be understood to understand gene expression. Although DNA methylation patterns are conserved across organisms, promoter DNA methylation is widely divergent. DNA methylation in genes and promoters are perhaps the most well-known DNA methylation pattern in plants [264].

15. TE Methylation in Plant Stress Response

Epigenetic modifications, including DNA and histone methylation, play a significant role in managing stress responses in plants through memory of abiotic and biotic stress factors. DNA methylation is a primary mediator of plant stress responses.

15.1. Abiotic Stress

Under both abiotic stress (such as extremes of temperature, salinity, low nutrient levels) and under normal conditions, recent studies have shown variable expression of epigenetic gene regulators depending on the local environment, thus demonstrating the need for epigenetic regulation (Table 4) [265]. Epigenomic reprogramming research on histone-associated chromatin and DNA modification has shown that plants exhibit a genome-wide reorganization response to stress [266]. A recent study on drought response in Arabidopsis revealed that trimethylation at lysine 4 on histone H3 (H3K4me3/H3K9me2) is complex and directly correlates with gene expression in stressed cells (Figure 5C) [267]. Increasing histone H3 phosphorylation at alkaline pH also helps maintain heterochromatin structure. H3 threonine 3 (H3T3ph) also tends to interact with H3K4me3 during osmotic stress [268], and this could potentially impact gene expression; this has previously been proposed for histone deacetylase HDA9. The epigenomic environment also contains the repressive H3K27me3 as a partial result of priming in Arabidopsis [269]. DNA methylation requires a specific histone H1 variant, and two DEAD-box helicases are needed for the epigenetic silencing of gene expression in plants, leading to stress [270]. Arabidopsis mutants defective in all stages of the RdDM pathway or CHG maintenance have an altered stomatal index or aversion to moisture starvation [271]. This supports the hypothesis that DNA methylation regulates abiotic gene expression. Drought in several plant species leads to substantial remodelling of DNA methylation, which allows plants to respond more effectively to recurring stress and prepares offspring for future stress responses [272]. However, in this case, modifying DNA methylation still seems to be essential to regulate neighbouring gene expression [273]. Phosphate starvation induced high-level TE methylation in rice but had a very limited effect in Arabidopsis, suggesting species-specific TE methylation in response to stress [274].
Table 4. Various roles of DNA methylation in plant responses to abiotic stresses [275].
Table 4. Various roles of DNA methylation in plant responses to abiotic stresses [275].
Abiotic StressPlantsChanges in DNA Methylation LevelsMajor EffectsReferences
Cold stressArabidopsisEnhanced methylation in the ALN promoterPromotes seed dormancy[276]
Cold stressArabidopsisVariation in ICE1 methylationCold tolerance divergence in different accessions[277,278]
Cold stressB. rapaDecreased DNA methylation levels in the BramMDH1 promoterIncreased heat tolerance and growth rate[279]
Cold stressB. rapaDemethylation of BrCKA2 and BrCKB4Regulation of floral transition. Regulation of temperature-dependent sex determination[280]
Cold stressCucumis sativusDemethylation of CHH sitesRegulation of temperature-dependent sex determination[281]
Cold stressRosa hybridaEnhanced CHH methylation of the RhAG promoterRegulation of floral organ development[282]
Drought stressArabidopsisIncreased 5mC methylation partly depending on H1.3Adaptive response to water deficiency[283]
Drought stressBrachypodium distachyonDecreased global 5mC while Bacillus subtilis strain B26 inoculation increasesIncreased drought stress resilience[284]
Drought stressG. hirsutumGlobal hypermethylation in all three contexts Acclimation to drought stress[285]
Drought stressO. sativaDifferential 5mC methylation alterationsConstitutive drought tolerance[286]
Drought stressPopulus trichocarpaIncreased methylation of upstream and downstream 2 kb and TEsRegulation of drought responses[287]
Drought stressZ. maysSuppression of ZmNAC111 by MITE through RdDMNatural variation in maize drought tolerance[288]
Heat stressArabidopsisAltered methylation of transposon remnantsRegulation of basal thermotolerance[206]
Heat stressArabidopsisChanges in genome-wide CHH-methylation patternsNatural adaptation to different temperatures[289]
Heat stressB. napusDNA hypomethylationRegulation of heat stress responses in cultured microspores[290]
Heat stressBrassica napusIncreased DNA methylation in heat-sensitive genotypesAdaptation to heat stress[291]
Heat stressGlycine maxHypomethylation in all contextsAffects the expression of genes or TEs under heat stress[292]
Heat stressGossypium hirsutumReduced DNA methylation level in a heat-sensitive lineMicrospore sterility[293,294]
Heat stressO. sativaDecreased DNA methylation levels of OsFIE1Regulation of seed size under heat stress[295]
Heat, salt, cold stressesO. sativaIncreased 6mA levels in heat and salt stress, decreased 6mA levels in cold stressRegulation of plant responses to environmental stresses[296]
Salt and drought stressesS. melongenaExpression changes of C5-MTases and demethylasesResponse to salt and drought stresses[297]
Salt and drought stressesSolanum lycopersicumActivation of Rider retrotransposon Modulation of salt and drought stress responses[298]
Salt stressB. napusDecreased methylation in the salinity-tolerant cultivar but increased methylation in the salinity-sensitive cultivarAcclimation to salt stress[299]
Salt stressO. sativaDecreased 5mC levels in the OsMYB91promoterEnhanced salt tolerance[207]
Salt stressO. sativaIncreased methylation level of the osa-miR393a promoterImproved salt tolerance[300]
Salt stressT. aestivumIncreased 5mC levels in TaHKT2;1 and TaHKT2;3Improved salt tolerance [301]
Salt stressTriticum aestivumReduced methylation levels in the promoter of salinity-responsive genesContributes to superior salinity tolerance[302]
Salt stressZea maysIncreased methylation of root ZmPP2C and demethylation of leaf ZmGSTAcclimation to salt stress[303]
Salt, heat and drought stressesO. sativaActivation of an LTR retrotransposon, HUOModulation of stress responses[304]

15.2. Biotic Stress

When compared with abiotic stress, less information is available on DNA methylation and histone post-translational modifications in response to biotic stress. Recent literature indicates that both necrotrophic and biotrophic pathogens are involved in changes to chromatin structure [305]. Chromatin modification is another layer of regulation for plant disease resistance. E3 ubiquitin ligase genes and histone monoubiquitination 1 (HUB1) and HUB2 regulate the expression of R genes, which induce constitutive immune responses in an Arabidopsis mutant. Histone ubiquitination is directly induced at the R gene locus [306]. Loss of histone deacetylase HDA19 mediates Arabidopsis immune responses to the pathogen Pseudomonas syringae pathovar tomato (Pst) strain DC3000 [307]. Silent or suppressed genes in stress regulation are characterized by the dimethylation and trimethylation of histone H3 Lys 27 (H3K27me2/3).
The rice gene, Jumonji C (jmjC) histone lysine protein gene (JMJ705) encoding histone lysine demethylase is involved in reversing Lys DNA methylation. In transgenic plants, increased JMJ705 expression removes H3K27me3 from defence-related genes, induces their expression with the aid of jasmonic acid, and improves resistance to the bacterial blight disease pathogen Xanthomonas oryzae pathovar oryzae [308]. In contrast, impaired JMJ703 activity raised levels of H3K4me3 and reactivated two families of non-LTR-RTE, and loss of JMJ703 did not change silencing of TE silencing [309]. This suggests that histone modifications are involved in TE silencing to regulate the plant immune response. It is also fascinating to note that the role of TEs is also important in plant pathogens to facilitate infection. For example, the ascomycete fungal pathogen Leptosphaeria maculans secretes an arsenal of small, secreted proteins (SSPs) that act as effectors to modulate host immunity to facilitate infection in B. napus. Chromatin-based transcriptional regulation of SSP-encoding genes associated with TEs in fungi impacts disease development during infection [310].
Many differentially methylated stress-response genes were discovered in plants exposed to different pathogens. Differentially methylated regions in the genome are also linked to gene expression. Mutations in the non-CG methyltransferases (DRM1, DMR2, and CMT3) and the CG methyltransferase (MET1) lead to genome-wide hypomethylation and pleiotropic developmental defects [311]. However, the met1 and the drm1, drm2, and cmt3 (ddc) mutants showed more disease resistance to the bacterial pathogen P. syringae pv. tomato DC3000 (Pst). These dynamic changes in DNA methylation and the functional consequences of differential methylation in regulating defence-related genes following pathogen attack in Arabidopsis are facilitated by TEs. In the Arabidopsis triple mutant rdd (ros1 dml2 dml3), defence-related genes are typically downregulated and therefore susceptibility to the fungal pathogen Fusarium oxysporum is increased. These genes in the mutant contain hypermethylated TE in their promoters. In contrast, these promotors are actively demethylated in the wild-type strain. Furthermore, ROS1, DML2, and DML3 demethylase activities are linked to fungal disease resistance, and DNA demethylation of TE sequences is largely regulated by ROS1 [312]. In addition, DNA methylation can prime TEs to cause activation of epigenetic transducers and can also directly induce gene silencing. Repeat components of DNA regions are known as DNA methylation interferes with expression of some biotic stress response genes. Loss of TE methylation also makes it easier to start the transcription process [313]. DNA methylation regulates stress-related genes by selective suppression of active TEs in their regulatory regions [314]. However, in addition to these mechanisms, a full understanding of epigenetic changes is also essential to better understand new key factors underlying plant stress responses. For example, TE methylation changes may lead to the activation of the SA signalling pathway to trigger widespread cell death during biotic stress. However, no direct evidence linking cell death to differential methylation has been observed.

16. Detection of TE Modifications and Measurement of TE Expression

Detecting TE modifications and measuring TE expression can facilitate understanding how TEs alter gene expression. A wide range of molecular techniques and analytical approaches are available to assess TE expression and modifications. However, these approaches should be carefully considered before implementation [317]. Analysis of TE sequencing results or TE-derived reads is challenging, as TEs are usually present in multiple copies in the plant genome, and ncRNAs and several mRNA genes are derived from TEs. However, there are several methods to detect TE methylation. These include methylation-sensitive amplified polymorphisms (MSAPs), methylation-specific PCR (MSP), sequencing of specific genes, and high-performance liquid chromatography (HPLC). However, these techniques are not suitable for broad identification of TE-modified sites. Whole genome bisulfite sequencing (WGBS) and reduced representation bisulfite sequencing (RRBS) are widely used methods to study TE modifications. Standard methods used for next-generation sequencing (NGS) are becoming routine. Several low-cost NGS platforms, including 454 sequencings, Illumina Genome analyser, Illumina, HeliScope Single Molecular Sequencer, Helicos BioSciences, and Nanopore sequencing are available to systematically study TE methylation [318]. Similar to DNA, RNA also undergoes various modifications (known as epitranscriptomics) and plays a significant role in biological processes [319]. This will lead to new discoveries in TE epitranscriptomics. As the present techniques cannot accurately detect TE modifications, focused research is necessary to generate new NGS platforms that can advance the understanding of all types of TE modifications in plants.
Recent molecular biology approaches such as ALE-seq, mobilome-seq, and VLP DNA-seq are more applicable in detecting active TEs in plants [320,321]. However, multi-mapped reads are typically discarded or not considered for analysis because of short-read sequencing. Thus, long-read sequencing technologies have recently been used as promising alternative methods that can easily separate different copies of the same family of TEs. For instance, unique transcripts containing various TEs were identified in maize using PacBio single-molecule RNA sequencing [322]. In addition, Oxford Nanopore Technology (ONT) can generate complete gene-like transcript annotation for TEs [323], suggesting that long-read sequencing allows the mapping of TE reads to a unique position of the plant genome.
However, conventional molecular biology techniques are still commonly used to study TEs. Although some approaches provide unique information, these are not applicable with genome-wide approaches. Although TE-derived transcripts are commonly quantified using qRT-PCR, this method has several major limitations. First, the main portion of the raw material starts with high-quality RNA, which contains pre-mRNA. Accordingly, the process begins with autonomous and passive transcription. Second, it is challenging to develop probes and primers that are truly different for a specific TE family. Third, the order of the amplified fragment cannot be predicted and is more likely to be a shortened transcript [317,324]. Unlike Southern blotting, Northern blotting assesses the size distribution of TE transcripts and whether full-length transcripts are present. Finally, programming individual TE loci with a reporter gene knock-in can be used to measure and parallelize gene expression levels accurately and rapidly [317,325]. This methodology has been employed in measuring each individual Ty1 RTE present in S. cerevisiae [326]; however, the results cannot be easily applied or generalized. Detection of TE proteins is also important. Internal TE mutations often inhibit translation of TE proteins, and post-translational modifications limit RTEs downstream. Western blotting and immunofluorescence experiments can address this issue. However, all conventional molecular biology techniques have several major limitations and advantages [317]. Thus, new approaches are needed to study a genome-wide view of TE expression.

17. Recent Machine Learning and Computational Tools for Analysing

Genome-wide analyses of TE methylation are limited due to the complex structures and high diversity of TEs. Several TE-dedicated computational tools (Table 5) are available for genome-wide analysis of TE expression and TE classification. These tools use various approaches, such as structure-based, homology-based, comparative genomics, and de novo. However, using these tools can still be challenging due to the polymorphic structures of TEs; thus, there are still debates on TE classification and annotation. No single bioinformatics tool can give reliable results on different types of TEs, and all tools have a high rate of false positives [30,327]. In general, RNA-seq data is mostly used for genome-wide approaches but mapping strategies of TEs with reference genomes mainly differ. Consequently, in addition to computational tools, the use of machine learning algorithms in bioinformatics has rapidly increased in recent years due to their demonstrable achievements in handling the difficult task of managing large datasets. Examples include genome annotation, classification of various plant genotypes with morphological and molecular markers, modularity and prediction of important quantitative properties in plants, analysis of complex, non-linear plant characteristics, and prediction and optimization of in vitro breeding methods. Various types of machine learning have been developed, each with its own methods, strengths, and disadvantages, thus making certain approaches more suited to specific tasks. Machine learning is divided into two categories (supervised and unsupervised), both of which improve the accuracy of TE detection by using results obtained by conventional software [30]. Machine learning can classify autonomous and non-autonomous TEs derived from LTR-RTEs using different features, such as LTR and ORF lengths. This can also distinguish between retroviral LTRs and other RTEs. Using machine learning, it is possible to discover new information on TEs, such as arrays of TEs, new transposition, TE methylation, new ncRNAs, and new DNA motifs. Using machine learning applications, detection of single nucleotide polymorphisms (SNPs) associated with TEs are useful for creating TE population models. Variation in allele frequencies may be used to reveal TE positive selection. However, very few tools, such as Red and TEClass, apply machine learning for TEs and their application in TEs is still limited [30].
Some online TE libraries also use machine learning approaches. For instance, InpactorDB (a semi-curated dataset composed of 130 439 LTR- RTEs from 195 plant genomes of 108 plant species) is an RTE library (e.g., RepeatMasker) for identifying and annotating LTR-RTEs using a machine learning approach [30]. Deep learning is a sub-discipline of machine learning and has shown successful results in genomics; hence, the use of deep learning in machine learning is also rapidly increasing. Deep learning and machine learning are more efficient approaches that use selected histograms or expected histograms to define TE genomic windows and hierarchical classification. However, machine learning has limited potential because of the repetitive nature and diverse polymorphisms of TEs and the species specificity of TEs. Furthermore, although deep learning is useful for genomic research, thus far no software has been developed to use deep learning for the identification and classification of TEs. Despite these challenges, a well-developed machine learning tool for TE classification would advance TE research [327]. Using data mining along with several key features, such as LTR length, TDS, ORFs, TATA boxes, AATAAA, and poly-A tails, developing machine learning for TE classification is possible. Thus, researchers should consider using computational tools and machine learning with deep learning and integrating different TE analyses, which can facilitate development of new applications for TE measurement, transposition, methylation levels, classification, and annotation.
Table 5. Analysis of transposable element (TE) unit expression from RNA-seq results using statistical methods and approaches. The table was adapted and recreated from Lanciano et al. [317], with copyright permission from the Licensor Springer Nature (Nature Reviews Genetics: Nature publisher) and Copyright Clearance Center (https://www.copyright.com) (Supplementary File S5).
Table 5. Analysis of transposable element (TE) unit expression from RNA-seq results using statistical methods and approaches. The table was adapted and recreated from Lanciano et al. [317], with copyright permission from the Licensor Springer Nature (Nature Reviews Genetics: Nature publisher) and Copyright Clearance Center (https://www.copyright.com) (Supplementary File S5).
Approaches or ToolsMapping or Pseudo-MappingFate of MultimappersType of QuantificationDistinguishes Unit-Length Transcripts from other TE-Derived TranscriptsIncludes Polymorphic TE ExpressionNotesReferences
Endogenous retrovirus (ERV) mapReference genomeDiscardedLocus specific--Uses a curated full-length human ERV database[328]
L1EMModel transcriptomeEM algorithmLocus specific+-Proof-of-principle on human long interspersed element 1 (L1) could be generalized[329]
Manual curationReference genome DiscardedLocus specific+-Difficult to generalize[324]
Multi-omics 1Reference genomeNALocus specific++Combines targeted DNA sequencing, RNA-seq, and ChIP-seq (chromatin immunoprecipitation followed by sequencing)[330]
Multi-omics 2Reference genome NALocus specific++Combines whole-genome sequencing and RNA-seq[331]
Random assignment of multimappersReference genomeRandomly assignedLocus specific--Locus-specific transcription not reliable on youngest TEs[332]
RE discover TEModel transcriptomeEM algorithm FFamily specific+-Uses Salmon TE algorithm[333]
Rep EnrichReference genomeRemapped on TE pseudogenomeFamily specific---[334]
Salmon TEConsensus transcriptomeExpectation-maximization (EM) algorithmFamily specific--Rapid pseudo mapping[335]
SQuIREReference genomeEM algorithmLocus specific-+/−Polymorphic insertion can be added as extra chromosome if internal sequence known[336]
TE toolsTE pseudo genomeRandomly assignedFamily specific--Applicable to unassembled genomes[337]
TEcandidatesReference genomeRemapped on partially masked reference genomeLocus specific---[338]
TelescopeReference genome EM algorithmLocus specific+--[339]
TEtranscriptsReference genomeEM algorithmFamily specific--Commonly used tool, tested on a wide variety of organisms[340]
TeXPReference genomeRandomly assignedFamily specific+/--Subtracts signal from pervasive transcription but not from other forms of chimeric transcripts[341]

18. Future Perspectives and Biotechnological Opportunities

Plant research has addressed important questions on whether TE-associated DNA variants contribute to evolutionary transition without affecting the genome. To better understand the impact on evolution, extensive molecular studies on the forms, origins, and impacts of TE activation in Arabidopsis have been performed. The results are also applicable to other organisms, especially maize [5]. In particular, the epigenetic and genetic influence of TEs on both hosts and TEs remains relatively understudied. The impact of TEs is attributed to the influence on the genome by suppressing genome recombination in the locality of TEs [226]. In the long term, peripheral transmission effects could theoretically influence overall evolution and have significant implications for genetic and molecular experiments that employ epigenomics [342]. Long-read technologies may elucidate the function of TEs from diverse plants [343]. Similar to DNA methylation, epitranscriptomic modification of RNAs (posttranscriptional RNA modifications) found in eukaryotes is a new layer of gene regulation and may function against TE transcripts [344]. Furthermore, single-cell genomics technologies, for example, appear to be a promising alternative for investigating DNA context in individual cells. Digital droplet PCR (ddPCR) is cost-effective and easy to use [345]. Since ddPCR performs a PCR on many thousands of tiny droplets, the digital presence or absence of TE in each droplet is easily identified by counting the number of droplets. Overall, the latest advances in DNA sequencing have radically changed the direction of transposon research. Relying on new types of epigenomics would open up knowledge and allow engineering of non-genetically modified crops [320].

19. Conclusions

It is generally agreed that TEs facilitate genetic and evolutionary diversification. Although some circumstantial evidence supports the above hypothesis, none of it is substantial and there is no direct proof that TEs facilitate ripening inhibitors. TEs are most often thought to create new genetic and phenotypic diversity via the introduction of new regulatory elements and gene and chromosomal disruptions. TEs also often play a crucial role in lineage-specific regulatory and coding sequence evolutions, contributing to new gene functions. Thus, TEs play a key role in the emergence of new phenotypes. For example, TEs are the primary source of novel regulatory sequence variations in primates. Adaptive novelty is mainly due to TE behaviour, which results in a large variety of genetic alterations, such as gene replication, enhanced expression, and newly created genes. Until now, most analyses of TEs only addressed occurrences of TEs and gene activity or transcript and phenotype relationships. A better understanding of the 3D chromatin structure organization within the nucleus may increase our understanding on the function of chromatin structure and its relation to mechanistic genome variations. This review highlighted the need to assess the regulation of TEs and their influence on the adaptive genome. This may facilitate development of improved traits for climate resilience and stress tolerance in the future.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms222111387/s1. Supplementary File S1 contains the copyright transfer documents of Figure 1 and Figure 4. Supplementary File S2 contains the copyright transfer documents of Table 1. Supplementary File S3 contains the copyright transfer documents of Table 2. Supplementary File S4 contains the copyright transfer documents of Figure 2, Figure 3 and Figure 5. Supplementary File S5 contains the copyright transfer documents of Table 5.

Author Contributions

M.R. planned, designed, and wrote the review. M.R., Q.W., R.K. and M.Z. outlined and edited the review. M.R., S.K., M.N. and L.S. drew the images. M.R., R.K., Q.W., M.N., S.K., L.S., A.S., A.E. and M.Z. edited and revised the review. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by grants from the National Key Point Research and Invention Program of the 13th Five-Year (Grant No. 2018YFD0600101), the National Natural Science Foundation of China (32071848, 31670602, 31960336), Jiangxi “Shuangqian” Program (S2019DQKJ2030), the Qing Lan Project of Jiangsu Higher Education Institutions, the Natural Science Foundation for Distinguished Young Scholars of Nanjing Forestry University (JC2019004), and a project funded by the Priority Academic Program Development of Jiangsu Higher Education Institutions. This work was also funded by Metasequoia Faculty Research Start-up Funding (grant number 163100028) at the Bamboo Research Institute, Nanjing Forestry University for first author MR. The research was supported by grants from the National Natural Science Foundation of China (grant numbers 31870656 and 31470615) and the Zhejiang Provincial Natural Science Foundation of China (grant number LZ19C160001). This work was supported by the Science Committee of the Ministry of Education and Science of the Republic of Kazakhstan in the framework of science and technology funding for the research program OR11465424 for 2021-2022.

Acknowledgments

The authors wish to thank Derek Ho (The University of Helsinki Language Centre, Finland) for outstanding editing and proofreading of the manuscript. We sincerely thank Kunnummal K. Vinod, Division of Genetics, ICAR-Indian Agricultural Research Institute, New Delhi, 110012, India for technical support.

Conflicts of Interest

All the authors have declared no conflict of interest.

References

  1. Bourque, G.; Burns, K.H.; Gehring, M.; Gorbunova, V.; Seluanov, A.; Hammell, M.; Imbeault, M.; Izsvák, Z.; Levin, H.L.; Macfarlan, T.S.; et al. Ten things you should know about transposable elements. Genome Biol. 2018, 19, 199. [Google Scholar] [CrossRef] [PubMed]
  2. Jangam, D.; Feschotte, C.; Betrán, E. Transposable Element Domestication as an Adaptation to Evolutionary Conflicts. Trends Genet. 2017, 33, 817–831. [Google Scholar] [CrossRef] [PubMed]
  3. Bennetzen, J.L.; Wang, H. The contributions of transposable elements to the structure, function, and evolution of plant genomes. Annu. Rev. Plant Biol. 2014, 65, 505–530. [Google Scholar] [CrossRef] [PubMed]
  4. Fedoroff, N. Transposons and genome evolution in plants. Proc. Natl. Acad. Sci. USA 2000, 97, 7002–7007. [Google Scholar] [CrossRef]
  5. Quesneville, H. Twenty years of transposable element analysis in the Arabidopsis thaliana genome. Mob. DNA 2020, 11, 28. [Google Scholar] [CrossRef]
  6. Kalendar, R.; Muterko, A.; Boronnikova, S. Retrotransposable Elements: DNA Fingerprinting and the Assessment of Genetic Diversity. Methods Mol. Biol. 2021, 2222, 263–286. [Google Scholar]
  7. Ariel, F.D.; Manavella, P.A. When junk DNA turns functional: Transposon-derived non-coding RNAs in plants. J. Exp. Bot. 2021, 72, 4132–4143. [Google Scholar] [CrossRef]
  8. Moelling, K.; Broecker, F.; Russo, G.; Sunagawa, S. RNase H As Gene Modifier, Driver of Evolution and Antiviral Defense. Front. Microbiol. 2017, 8, 1745. [Google Scholar] [CrossRef]
  9. Kumar, S. Epigenomics of Plant Responses to Environmental Stress. Epigenomes 2018, 2, 6. [Google Scholar] [CrossRef]
  10. Ayarpadikannan, S.; Kim, H.-S. The impact of transposable elements in genome evolution and genetic instability and their implications in various diseases. Genom. Inform. 2014, 12, 98–104. [Google Scholar] [CrossRef]
  11. Fattash, I.; Rooke, R.; Wong, A.; Hui, C.; Luu, T.; Bhardwaj, P.; Yang, G. Miniature inverted-repeat transposable elements: Discovery, distribution, and activity. Genome/Natl. Res. Counc. Can. = Genome/Cons. Natl. Rech. Can. 2013, 56, 475–486. [Google Scholar] [CrossRef]
  12. Sundaram, V.; Wysocka, J. Transposable elements as a potent source of diverse cis-regulatory sequences in mammalian genomes. Philos.Trans. R. Soc. Lond. B Biol. Sci. 2020, 375, 20190347. [Google Scholar] [CrossRef]
  13. Moschetti, R.; Palazzo, A.; Lorusso, P.; Viggiano, L.; Massimiliano Marsano, R. “What You Need, Baby, I Got It”: Transposable Elements as Suppliers of Cis-Operating Sequences in Drosophila. Biology 2020, 9, 25. [Google Scholar] [CrossRef]
  14. Sahebi, M.; Hanafi, M.M.; van Wijnen, A.J.; Rice, D.; Rafii, M.Y.; Azizi, P.; Osman, M.; Taheri, S.; Bakar, M.F.A.; Isa, M.N.M.; et al. Contribution of transposable elements in the plant’s genome. Gene 2018, 665, 155–166. [Google Scholar] [CrossRef]
  15. Payer, L.M.; Burns, K.H. Transposable elements in human genetic disease. Nat. Rev. Genet. 2019, 20, 760–772. [Google Scholar] [CrossRef]
  16. Hirsch, C.D.; Springer, N.M. Transposable element influences on gene expression in plants. Biochim. Biophys. Acta Gene Regul. Mech. 2017, 1860, 157–165. [Google Scholar] [CrossRef]
  17. Chuong, E.B.; Elde, N.C.; Feschotte, C. Regulatory activities of transposable elements: From conflicts to benefits. Nat. Rev. Genet. 2017, 18, 71–86. [Google Scholar] [CrossRef]
  18. Pisoschi, A.M.; Pop, A.; Iordache, F.; Stanca, L.; Predoi, G.; Serban, A.I. Oxidative stress mitigation by antioxidants—An overview on their chemistry and influences on health status. Eur. J. Med. Chem. 2021, 209, 112891. [Google Scholar] [CrossRef]
  19. Huang, S.; Tao, X.; Yuan, S.; Zhang, Y.; Li, P.; Beilinson, H.A.; Zhang, Y.; Yu, W.; Pontarotti, P.; Escriva, H.; et al. Discovery of an Active RAG Transposon Illuminates the Origins of V(D)J Recombination. Cell 2016, 166, 102–114. [Google Scholar] [CrossRef]
  20. Kapitonov, V.V.; Koonin, E.V. Evolution of the RAG1-RAG2 locus: Both proteins came from the same transposon. Biol. Direct 2015, 10, 20. [Google Scholar] [CrossRef]
  21. Sandoval-Villegas, N.; Nurieva, W.; Amberger, M.; Ivics, Z. Contemporary Transposon Tools: A Review and Guide through Mechanisms and Applications of Sleeping Beauty, piggyBac and Tol2 for Genome Engineering. Int. J. Mol. Sci. 2021, 22, 5084. [Google Scholar] [CrossRef]
  22. Tipanee, J.; VandenDriessche, T.; Chuah, M.K. Transposons: Moving Forward from Preclinical Studies to Clinical Trials. Hum. Gene Ther. 2017, 28, 1087–1104. [Google Scholar] [CrossRef]
  23. Palazzo, A.; Marsano, R.M. Transposable elements: A jump toward the future of expression vectors. Crit. Rev. Biotechnol. 2021, 41, 792–808. [Google Scholar] [CrossRef]
  24. Walker, J.A.; Jordan, V.E.; Steely, C.J.; Beckstrom, T.O.; McDaniel, C.L.; St. Romain, C.P.; Bennett, E.C.; Robichaux, A.; Clement, B.N.; Konkel, M.K.; et al. Papio Baboon Species Indicative Alu Elements. Genome Biol. Evol. 2017, 9, 1788–1796. [Google Scholar] [CrossRef]
  25. Steely, C.J.; Walker, J.A.; Jordan, V.E.; Beckstrom, T.O.; McDaniel, C.L.; St. Romain, C.P.; Bennett, E.C.; Robichaux, A.; Clement, B.N.; Raveendran, M.; et al. Alu Insertion Polymorphisms as Evidence for Population Structure in Baboons. Genome Biol. Evol. 2017, 9, 2418–2427. [Google Scholar] [CrossRef]
  26. Kalendar, R.; Amenov, A.; Daniyarov, A. Use of retrotransposon-derived genetic markers to analyse genomic variability in plants. Funct. Plant Biol. 2018, 46, 15–29. [Google Scholar] [CrossRef]
  27. Kalendar, R.; Raskina, O.; Belyayev, A.; Schulman, A.H. Long Tandem Arrays of Cassandra Retroelements and Their Role in Genome Dynamics in Plants. Int. J. Mol. Sci. 2020, 21, 2931. [Google Scholar] [CrossRef]
  28. Kalendar, R.; Tanskanen, J.; Chang, W.; Antonius, K.; Sela, H.; Peleg, O.; Schulman, A.H. Cassandra retrotransposons carry independently transcribed 5S RNA. Proc. Natl. Acad. Sci. USA 2008, 105, 5833–5838. [Google Scholar] [CrossRef]
  29. Kalendar, R.; Vicient, C.M.; Peleg, O.; Anamthawat-Jonsson, K.; Bolshoy, A.; Schulman, A.H. Large Retrotransposon Derivatives: Abundant, Conserved but Nonautonomous Retroelements of Barley and Related Genomes. Genetics 2004, 166, 1437–1450. [Google Scholar] [CrossRef]
  30. Orozco-Arias, S.; Isaza, G.; Guyot, R. Retrotransposons in Plant Genomes: Structure, Identification, and Classification through Bioinformatics and Machine Learning. Int. J. Mol. Sci. 2019, 20, 3837. [Google Scholar] [CrossRef]
  31. Hickman, A.B.; Dyda, F. DNA Transposition at Work. Chem. Rev. 2016, 116, 12758–12784. [Google Scholar] [CrossRef] [PubMed]
  32. Khan, N.; Bano, A.; Ali, S.; Babar, M.A. Crosstalk amongst phytohormones from planta and PGPR under biotic and abiotic stresses. Plant Growth Regul. 2020, 90, 189–203. [Google Scholar] [CrossRef]
  33. Ashapkin, V.V.; Kutueva, L.I.; Aleksandrushkina, N.I.; Vanyushin, B.F. Epigenetic Mechanisms of Plant Adaptation to Biotic and Abiotic Stresses. Int. J. Mol. Sci. 2020, 21, 7457. [Google Scholar] [CrossRef] [PubMed]
  34. Deniz, Ö.; Frost, J.M.; Branco, M.R. Regulation of transposable elements by DNA modifications. Nat. Rev. Genet. 2019, 20, 417–431. [Google Scholar] [CrossRef]
  35. 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]
  36. Lindermayr, C.; Rudolf, E.E.; Durner, J.; Groth, M. Interactions between metabolism and chromatin in plant models. Mol. Metab. 2020, 38, 100951. [Google Scholar] [CrossRef]
  37. Ong-Abdullah, M.; Ordway, J.M.; Jiang, N.; Ooi, S.E.; Kok, S.Y.; Sarpan, N.; Azimi, N.; Hashim, A.T.; Ishak, Z.; Rosli, S.K.; et al. Loss of Karma transposon methylation underlies the mantled somaclonal variant of oil palm. Nature 2015, 525, 533–537. [Google Scholar] [CrossRef]
  38. Khan, T.; Relitti, N.; Brindisi, M.; Magnano, S.; Zisterer, D.; Gemma, S.; Butini, S.; Campiani, G. Autophagy modulators for the treatment of oral and esophageal squamous cell carcinomas. Med. Res. Rev. 2020, 40, 1002–1060. [Google Scholar] [CrossRef]
  39. Belyayev, A.; Kalendar, R.; Brodsky, L.; Nevo, E.; Schulman, A.H.; Raskina, O. Transposable elements in a marginal plant population: Temporal fluctuations provide new insights into genome evolution of wild diploid wheat. Mob. DNA 2010, 1, 6. [Google Scholar] [CrossRef]
  40. 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]
  41. Zhou, W.; Liang, G.; Molloy, P.L.; Jones, P.A. DNA methylation enables transposable element-driven genome expansion. Proc. Natl. Acad. Sci. USA 2020, 117, 19359–19366. [Google Scholar] [CrossRef]
  42. Zhang, H.; Lang, Z.; Zhu, J.K. Dynamics and function of DNA methylation in plants. Nat. Rev. Mol. Cell Biol. 2018, 19, 489–506. [Google Scholar] [CrossRef]
  43. Jönsson, M.E.; Garza, R.; Johansson, P.A.; Jakobsson, J. Transposable Elements: A Common Feature of Neurodevelopmental and Neurodegenerative Disorders. Trends Genet. 2020, 36, 610–623. [Google Scholar] [CrossRef]
  44. Wambui Mbichi, R.; Wang, Q.-F.; Wan, T. RNA directed DNA methylation and seed plant genome evolution. Plant Cell Rep. 2020, 39, 983–996. [Google Scholar] [CrossRef]
  45. Slotkin, R.K.; Martienssen, R. Transposable elements and the epigenetic regulation of the genome. Nat. Rev. Genet. 2007, 8, 272–285. [Google Scholar] [CrossRef]
  46. Kabelitz, T.; Brzezinka, K.; Friedrich, T.; Górka, M.; Graf, A.; Kappel, C.; Bäurle, I. A JUMONJI Protein with E3 Ligase and Histone H3 Binding Activities Affects Transposon Silencing in Arabidopsis. Plant Physiol. 2016, 171, 344–358. [Google Scholar] [CrossRef]
  47. Kalavacharla, V.; Subramani, M.; Ayyappan, V.; Dworkin, M.C.; Hayford, R.K. Plant Epigenomics. In Handbook of Epigenetics, 2nd ed.; Tollefsbol, T.O., Ed.; Academic Press: Cambridge, MA, USA, 2017; Chapter 16; pp. 245–258. [Google Scholar]
  48. Wang, X.; Weigel, D.; Smith, L.M. Transposon variants and their effects on gene expression in Arabidopsis. PLoS Genet. 2013, 9, e1003255. [Google Scholar] [CrossRef]
  49. Papareddy, R.K.; Páldi, K.; Paulraj, S.; Kao, P.; Lutzmayer, S.; Nodine, M.D. Chromatin regulates expression of small RNAs to help maintain transposon methylome homeostasis in Arabidopsis. Genome Biol. 2020, 21, 251. [Google Scholar] [CrossRef]
  50. Feng, J.X.; Riddle, N.C. Epigenetics and genome stability. Mamm. Genome 2020, 31, 181–195. [Google Scholar] [CrossRef]
  51. Wicker, T.; Sabot, F.; Hua-Van, A.; Bennetzen, J.L.; Capy, P.; Chalhoub, B.; Flavell, A.; Leroy, P.; Morgante, M.; Panaud, O.; et al. A unified classification system for eukaryotic transposable elements. Nat. Rev. Genet. 2007, 8, 973–982. [Google Scholar] [CrossRef]
  52. Levin, H.L.; Moran, J.V. Dynamic interactions between transposable elements and their hosts. Nat. Rev. Genet. 2011, 12, 615–627. [Google Scholar] [CrossRef]
  53. Grabundzija, I.; Hickman, A.B.; Dyda, F. Helraiser intermediates provide insight into the mechanism of eukaryotic replicative transposition. Nat. Commun. 2018, 9, 1278. [Google Scholar] [CrossRef] [PubMed]
  54. Boeke, J.D.; Garfinkel, D.J.; Styles, C.A.; Fink, G.R. Ty elements transpose through an RNA intermediate. Cell 1985, 40, 491–500. [Google Scholar] [CrossRef]
  55. Griffiths, J.; Catoni, M.; Iwasaki, M.; Paszkowski, J. Sequence-Independent Identification of Active LTR Retrotransposons in Arabidopsis. Mol. Plant 2018, 11, 508–511. [Google Scholar] [CrossRef]
  56. Pastuzyn, E.D.; Day, C.E.; Kearns, R.B.; Kyrke-Smith, M.; Taibi, A.V.; McCormick, J.; Yoder, N.; Belnap, D.M.; Erlendsson, S.; Morado, D.R.; et al. The Neuronal Gene Arc Encodes a Repurposed Retrotransposon Gag Protein that Mediates Intercellular RNA Transfer. Cell 2018, 172, 275–288.e18. [Google Scholar] [CrossRef]
  57. Gao, D.; Jimenez-Lopez, J.C.; Iwata, A.; Gill, N.; Jackson, S.A. Functional and Structural Divergence of an Unusual LTR Retrotransposon Family in Plants. PLoS ONE 2012, 7, e48595. [Google Scholar] [CrossRef]
  58. Malik, H.S.; Eickbush, T.H. Phylogenetic analysis of ribonuclease H domains suggests a late, chimeric origin of LTR retrotransposable elements and retroviruses. Genome Res. 2001, 11, 1187–1197. [Google Scholar] [CrossRef]
  59. Feschotte, C.; Pritham, E.J. DNA transposons and the evolution of eukaryotic genomes. Annu. Rev. Genet. 2007, 41, 331–368. [Google Scholar] [CrossRef]
  60. Naorem, S.S.; Han, J.; Wang, S.; Lee, W.R.; Heng, X.; Miller, J.F.; Guo, H. DGR mutagenic transposition occurs via hypermutagenic reverse transcription primed by nicked template RNA. Proc. Natl. Acad. Sci. USA 2017, 114, E10187–E10195. [Google Scholar] [CrossRef]
  61. Piégu, B.; Bire, S.; Arensburger, P.; Bigot, Y. A survey of transposable element classification systems--a call for a fundamental update to meet the challenge of their diversity and complexity. Mol. Phylogenet. Evol. 2015, 86, 90–109. [Google Scholar] [CrossRef]
  62. Zhao, D.; Ferguson, A.A.; Jiang, N. What makes up plant genomes: The vanishing line between transposable elements and genes. Biochim. Biophys. Acta 2016, 1859, 366–380. [Google Scholar] [CrossRef] [PubMed]
  63. Berthelier, J.; Casse, N.; Daccord, N.; Jamilloux, V.; Saint-Jean, B.; Carrier, G. A transposable element annotation pipeline and expression analysis reveal potentially active elements in the microalga Tisochrysis lutea. BMC Genom. 2018, 19, 378. [Google Scholar] [CrossRef] [PubMed]
  64. Feschotte, C.; Jiang, N.; Wessler, S.R. Plant transposable elements: Where genetics meets genomics. Nat. Rev. Genet. 2002, 3, 329–341. [Google Scholar] [CrossRef] [PubMed]
  65. Hirochika, H.; Sugimoto, K.; Otsuki, Y.; Tsugawa, H.; Kanda, M. Retrotransposons of rice involved in mutations induced by tissue culture. Proc. Natl. Acad. Sci. USA 1996, 93, 7783–7788. [Google Scholar] [CrossRef]
  66. Vicient, C.M.; Suoniemi, A.; Anamthawat-Jónsson, K.; Tanskanen, J.; Beharav, A.; Nevo, E.; Schulman, A.H. Retrotransposon BARE-1 and Its Role in Genome Evolution in the Genus Hordeum. Plant Cell 1999, 11, 1769–1784. [Google Scholar] [CrossRef]
  67. Hirochika, H. Activation of tobacco retrotransposons during tissue culture. EMBO J. 1993, 12, 2521–2528. [Google Scholar] [CrossRef] [PubMed]
  68. Grandbastien, M.A.; Spielmann, A.; Caboche, M. Tnt1, a mobile retroviral-like transposable element of tobacco isolated by plant cell genetics. Nature 1989, 337, 376–380. [Google Scholar] [CrossRef]
  69. White, S.E.; Habera, L.F.; Wessler, S.R. Retrotransposons in the flanking regions of normal plant genes: A role for copia-like elements in the evolution of gene structure and expression. Proc. Natl. Acad. Sci. USA 1994, 91, 11792–11796. [Google Scholar] [CrossRef]
  70. Jin, Y.K.; Bennetzen, J.L. Structure and coding properties of Bs1, a maize retrovirus-like transposon. Proc. Natl. Acad. Sci. USA 1989, 86, 6235–6239. [Google Scholar] [CrossRef]
  71. Meyers, B.C.; Tingey, S.V.; Morgante, M. Abundance, distribution, and transcriptional activity of repetitive elements in the maize genome. Genome Res. 2001, 11, 1660–1676. [Google Scholar] [CrossRef]
  72. Jiang, N.; Bao, Z.; Temnykh, S.; Cheng, Z.; Jiang, J.; Wing, R.A.; McCouch, S.R.; Wessler, S.R. Dasheng: A recently amplified nonautonomous long terminal repeat element that is a major component of pericentromeric regions in rice. Genetics 2002, 161, 1293–1305. [Google Scholar] [CrossRef] [PubMed]
  73. Purugganan, M.D.; Wessler, S.R. Molecular evolution of magellan, a maize Ty3/gypsy-like retrotransposon. Proc. Natl. Acad. Sci. USA 1994, 91, 11674–11678. [Google Scholar] [CrossRef] [PubMed]
  74. Wright, D.A.; Voytas, D.F. Athila4 of Arabidopsis and Calypso of soybean define a lineage of endogenous plant retroviruses. Genome Res. 2002, 12, 122–131. [Google Scholar] [CrossRef] [PubMed]
  75. Konieczny, A.; Voytas, D.F.; Cummings, M.P.; Ausubel, F.M. A superfamily of Arabidopsis thaliana retrotransposons. Genetics 1991, 127, 801–809. [Google Scholar] [CrossRef]
  76. Leeton, P.R.; Smyth, D.R. An abundant LINE-like element amplified in the genome of Lilium speciosum. Mol. Gen. Genet. 1993, 237, 97–104. [Google Scholar] [CrossRef]
  77. Schwarz-Sommer, Z.; Leclercq, L.; Göbel, E.; Saedler, H. Cin4, an insert altering the structure of the A1 gene in Zea mays, exhibits properties of nonviral retrotransposons. EMBO J. 1987, 6, 3873–3880. [Google Scholar] [CrossRef]
  78. Wright, D.A.; Ke, N.; Smalle, J.; Hauge, B.M.; Goodman, H.M.; Voytas, D.F. Multiple non-LTR retrotransposons in the genome of Arabidopsis thaliana. Genetics 1996, 142, 569–578. [Google Scholar] [CrossRef]
  79. Yoshioka, Y.; Matsumoto, S.; Kojima, S.; Ohshima, K.; Okada, N.; Machida, Y. Molecular characterization of a short interspersed repetitive element from tobacco that exhibits sequence homology to specific tRNAs. Proc. Natl. Acad. Sci. USA 1993, 90, 6562–6566. [Google Scholar] [CrossRef]
  80. Deragon, J.M.; Landry, B.S.; Pélissier, T.; Tutois, S.; Tourmente, S.; Picard, G. An analysis of retroposition in plants based on a family of SINEs from Brassica napus. J. Mol. Evol. 1994, 39, 378–386. [Google Scholar] [CrossRef]
  81. Chandler, V.; Rivin, C.; Walbot, V. Stable non-mutator stocks of maize have sequences homologous to the Mu1 transposable element. Genetics 1986, 114, 1007–1021. [Google Scholar] [CrossRef]
  82. Singer, T.; Yordan, C.; Martienssen, R.A. Robertson’s Mutator transposons in A. thaliana are regulated by the chromatin-remodeling gene Decrease in DNA Methylation (DDM1). Genes Dev. 2001, 15, 591–602. [Google Scholar] [CrossRef]
  83. Gierl, A. The En/Spm transposable element of maize. Curr. Top. Microbiol. Immunol. 1996, 204, 145–159. [Google Scholar]
  84. Miura, A.; Yonebayashi, S.; Watanabe, K.; Toyama, T.; Shimada, H.; Kakutani, T. Mobilization of transposons by a mutation abolishing full DNA methylation in Arabidopsis. Nature 2001, 411, 212–214. [Google Scholar] [CrossRef]
  85. Wessler, S.R. Phenotypic diversity mediated by the maize transposable elements Ac and Spm. Science 1988, 242, 399–405. [Google Scholar] [CrossRef]
  86. Zhang, X.; Feschotte, C.; Zhang, Q.; Jiang, N.; Eggleston, W.B.; Wessler, S.R. P instability factor: An active maize transposon system associated with the amplification of Tourist-like MITEs and a new superfamily of transposases. Proc. Natl. Acad. Sci. USA 2001, 98, 12572–12577. [Google Scholar] [CrossRef]
  87. Kapitonov, V.V.; Jurka, J. Molecular paleontology of transposable elements from Arabidopsis thaliana. Genetica 1999, 107, 27–37. [Google Scholar] [CrossRef]
  88. Turcotte, K.; Srinivasan, S.; Bureau, T. Survey of transposable elements from rice genomic sequences. Plant J. 2001, 25, 169–179. [Google Scholar] [CrossRef]
  89. Feschotte, C.; Wessler, S.R. Mariner-like transposases are widespread and diverse in flowering plants. Proc. Natl. Acad. Sci. USA 2002, 99, 280–285. [Google Scholar] [CrossRef]
  90. Svedberg, J.; Hosseini, S.; Chen, J.; Vogan, A.A.; Mozgova, I.; Hennig, L.; Manitchotpisit, P.; Abusharekh, A.; Hammond, T.M.; Lascoux, M.; et al. Convergent evolution of complex genomic rearrangements in two fungal meiotic drive elements. Nat. Commun. 2018, 9, 4242. [Google Scholar] [CrossRef]
  91. Cornejo, E.; Abreu, N.; Komeili, A. Compartmentalization and organelle formation in bacteria. Curr. Opin. Cell Biol. 2014, 26, 132–138. [Google Scholar] [CrossRef]
  92. Neumann, P.; Novák, P.; Hoštáková, N.; Macas, J. Systematic survey of plant LTR-retrotransposons elucidates phylogenetic relationships of their polyprotein domains and provides a reference for element classification. Mob. DNA 2019, 10, 1. [Google Scholar] [CrossRef]
  93. Bourgeois, Y.; Ruggiero, R.P.; Hariyani, I.; Boissinot, S. Disentangling the determinants of transposable elements dynamics in vertebrate genomes using empirical evidences and simulations. PLoS Genet. 2020, 16, e1009082. [Google Scholar] [CrossRef] [PubMed]
  94. Maxwell, P.H. Diverse transposable element landscapes in pathogenic and nonpathogenic yeast models: The value of a comparative perspective. Mob. DNA 2020, 11, 16. [Google Scholar] [CrossRef]
  95. Yang, N.; Yan, J. New genomic approaches for enhancing maize genetic improvement. Curr. Opin. Plant Biol. 2021, 60, 101977. [Google Scholar] [CrossRef] [PubMed]
  96. Schubert, I.; Vu, G.T.H. Genome Stability and Evolution: Attempting a Holistic View. Trends Plant Sci. 2016, 21, 749–757. [Google Scholar] [CrossRef] [PubMed]
  97. Bennetzen, J.L.; Wang, X. Relationships between Gene Structure and Genome Instability in Flowering Plants. Mol. Plant 2018, 11, 407–413. [Google Scholar] [CrossRef]
  98. Soltis, P.S.; Marchant, D.B.; Van de Peer, Y.; Soltis, D.E. Polyploidy and genome evolution in plants. Curr. Opin. Genet. Dev. 2015, 35, 119–125. [Google Scholar] [CrossRef] [PubMed]
  99. Ragupathy, R.; You, F.M.; Cloutier, S. Arguments for standardizing transposable element annotation in plant genomes. Trends Plant Sci. 2013, 18, 367–376. [Google Scholar] [CrossRef]
  100. Hu, T.T.; Pattyn, P.; Bakker, E.G.; Cao, J.; Cheng, J.-F.; Clark, R.M.; Fahlgren, N.; Fawcett, J.A.; Grimwood, J.; Gundlach, H.; et al. The Arabidopsis lyrata genome sequence and the basis of rapid genome size change. Nat. Genet. 2011, 43, 476–481. [Google Scholar] [CrossRef]
  101. Civáň, P.; Švec, M.; Hauptvogel, P. On the Coevolution of Transposable Elements and Plant Genomes. J. Bot. 2011, 2011, 893546. [Google Scholar] [CrossRef]
  102. Dodsworth, S.; Leitch, A.R.; Leitch, I.J. Genome size diversity in angiosperms and its influence on gene space. Curr. Opin. Genet. Dev. 2015, 35, 73–78. [Google Scholar] [CrossRef]
  103. Galindo-Gonzalez, L.; Mhiri, C.; Deyholos, M.K.; Grandbastien, M.A. LTR-retrotransposons in plants: Engines of evolution. Gene 2017, 626, 14–25. [Google Scholar] [CrossRef]
  104. Hollister, J.D.; Gaut, B.S. Epigenetic silencing of transposable elements: A trade-off between reduced transposition and deleterious effects on neighboring gene expression. Genome Res. 2009, 19, 1419–1428. [Google Scholar] [CrossRef]
  105. Haberer, G.; Kamal, N.; Bauer, E.; Gundlach, H.; Fischer, I.; Seidel, M.A.; Spannagl, M.; Marcon, C.; Ruban, A.; Urbany, C.; et al. European maize genomes highlight intraspecies variation in repeat and gene content. Nat. Genet. 2020, 52, 950–957. [Google Scholar] [CrossRef]
  106. Ma, B.; Xin, Y.; Kuang, L.; He, N. Distribution and Characteristics of Transposable Elements in the Mulberry Genome. Plant Genome 2019, 12, 180094. [Google Scholar] [CrossRef]
  107. Tenaillon, M.I.; Hollister, J.D.; Gaut, B.S. A triptych of the evolution of plant transposable elements. Trends Plant Sci. 2010, 15, 471–478. [Google Scholar] [CrossRef]
  108. The Arabidopsis Genome, I. Analysis of the genome sequence of the flowering plant Arabidopsis thaliana. Nature 2000, 408, 796–815. [Google Scholar] [CrossRef]
  109. Sasaki, T.; International Rice Genome Sequencing, P. The map-based sequence of the rice genome. Nature 2005, 436, 793–800. [Google Scholar] [CrossRef]
  110. Schnable, P.S.; Ware, D.; Fulton, R.S.; Stein, J.C.; Wei, F.; Pasternak, S.; Liang, C.; Zhang, J.; Fulton, L.; Graves, T.A.; et al. The B73 maize genome: Complexity, diversity, and dynamics. Science 2009, 326, 1112–1115. [Google Scholar] [CrossRef]
  111. Estep, M.C.; DeBarry, J.D.; Bennetzen, J.L. The dynamics of LTR retrotransposon accumulation across 25 million years of panicoid grass evolution. Heredity 2013, 110, 194–204. [Google Scholar] [CrossRef]
  112. Piegu, B.; Guyot, R.; Picault, N.; Roulin, A.; Sanyal, A.; Kim, H.; Collura, K.; Brar, D.S.; Jackson, S.; Wing, R.A.; et al. Doubling genome size without polyploidization: Dynamics of retrotransposition-driven genomic expansions in Oryza australiensis, a wild relative of rice. Genome Res. 2006, 16, 1262–1269. [Google Scholar] [CrossRef]
  113. Bennett, M.D.; Leitch, I.J. “Plant DNA C-Values Database,” Release 5.0, December 2010. 2010. Available online: http://data.kew.org/cvalues/ (accessed on 29 June 2021).
  114. Bartos, J.; Paux, E.; Kofler, R.; Havránková, M.; Kopecký, D.; Suchánková, P.; Safár, J.; Simková, H.; Town, C.D.; Lelley, T.; et al. A first survey of the rye (Secale cereale) genome composition through BAC end sequencing of the short arm of chromosome 1R. BMC Plant Biol. 2008, 8, 95. [Google Scholar] [CrossRef]
  115. Wang, H.; Liu, J.S. LTR retrotransposon landscape in Medicago truncatula: More rapid removal than in rice. BMC Genom. 2008, 9, 382. [Google Scholar] [CrossRef]
  116. Velasco, R.; Zharkikh, A.; Troggio, M.; Cartwright, D.A.; Cestaro, A.; Pruss, D.; Pindo, M.; Fitzgerald, L.M.; Vezzulli, S.; Reid, J.; et al. A high quality draft consensus sequence of the genome of a heterozygous grapevine variety. PLoS ONE 2007, 2, e1326. [Google Scholar] [CrossRef]
  117. Jaillon, O.; Aury, J.M.; Noel, B.; Policriti, A.; Clepet, C.; Casagrande, A.; Choisne, N.; Aubourg, S.; Vitulo, N.; Jubin, C.; et al. The grapevine genome sequence suggests ancestral hexaploidization in major angiosperm phyla. Nature 2007, 449, 463–467. [Google Scholar]
  118. Sato, S.; Nakamura, Y.; Kaneko, T.; Asamizu, E.; Kato, T.; Nakao, M.; Sasamoto, S.; Watanabe, A.; Ono, A.; Kawashima, K.; et al. Genome Structure of the Legume, Lotus japonicus. DNA Res. 2008, 15, 227–239. [Google Scholar] [CrossRef]
  119. Holligan, D.; Zhang, X.; Jiang, N.; Pritham, E.J.; Wessler, S.R. The transposable element landscape of the model legume Lotus japonicus. Genetics 2006, 174, 2215–2228. [Google Scholar] [CrossRef]
  120. Tuskan, G.A.; Difazio, S.; Jansson, S.; Bohlmann, J.; Grigoriev, I.; Hellsten, U.; Putnam, N.; Ralph, S.; Rombauts, S.; Salamov, A.; et al. The genome of black cottonwood, Populus trichocarpa (Torr. & Gray). Science 2006, 313, 1596–1604. [Google Scholar]
  121. Wicker, T.; Zimmermann, W.; Perovic, D.; Paterson, A.H.; Ganal, M.; Graner, A.; Stein, N. A detailed look at 7 million years of genome evolution in a 439 kb contiguous sequence at the barley Hv-eIF4E locus: Recombination, rearrangements and repeats. Plant J. 2005, 41, 184–194. [Google Scholar] [CrossRef]
  122. Ming, R.; Hou, S.; Feng, Y.; Yu, Q.; Dionne-Laporte, A.; Saw, J.H.; Senin, P.; Wang, W.; Ly, B.V.; Lewis, K.L.; et al. The draft genome of the transgenic tropical fruit tree papaya (Carica papaya Linnaeus). Nature 2008, 452, 991–996. [Google Scholar] [CrossRef]
  123. Du, J.; Grant, D.; Tian, Z.; Nelson, R.T.; Zhu, L.; Shoemaker, R.C.; Ma, J. SoyTEdb: A comprehensive database of transposable elements in the soybean genome. BMC Genom. 2010, 11, 113. [Google Scholar] [CrossRef] [PubMed]
  124. Shen, Y.; Liu, J.; Geng, H.; Zhang, J.; Liu, Y.; Zhang, H.; Xing, S.; Du, J.; Ma, S.; Tian, Z. De novo assembly of a Chinese soybean genome. Sci. China Life Sci. 2018, 61, 871–884. [Google Scholar] [CrossRef] [PubMed]
  125. González, V.M.; Benjak, A.; Hénaff, E.M.; Mir, G.; Casacuberta, J.M.; Garcia-Mas, J.; Puigdomènech, P. Sequencing of 6.7 Mb of the melon genome using a BAC pooling strategy. BMC Plant Biol. 2010, 10, 246. [Google Scholar] [CrossRef] [PubMed]
  126. Ruggieri, V.; Alexiou, K.G.; Morata, J.; Argyris, J.; Pujol, M.; Yano, R.; Nonaka, S.; Ezura, H.; Latrasse, D.; Boualem, A.; et al. An improved assembly and annotation of the melon (Cucumis melo L.) reference genome. Sci. Rep. 2018, 8, 8088. [Google Scholar] [CrossRef]
  127. Hawkins, J.S.; Kim, H.; Nason, J.D.; Wing, R.A.; Wendel, J.F. Differential lineage-specific amplification of transposable elements is responsible for genome size variation in Gossypium. Genome Res. 2006, 16, 1252–1261. [Google Scholar] [CrossRef]
  128. Zhang, X.; Wessler, S.R. Genome-wide comparative analysis of the transposable elements in the related species Arabidopsis thaliana and Brassica oleracea. Proc. Natl. Acad. Sci. USA 2004, 101, 5589–5594. [Google Scholar] [CrossRef]
  129. Guo, Z.-H.; Ma, P.-F.; Yang, G.-Q.; Hu, J.-Y.; Liu, Y.-L.; Xia, E.-H.; Zhong, M.-C.; Zhao, L.; Sun, G.-L.; Xu, Y.-X.; et al. Genome Sequences Provide Insights into the Reticulate Origin and Unique Traits of Woody Bamboos. Mol. Plant 2019, 12, 1353–1365. [Google Scholar] [CrossRef]
  130. Jedlicka, P.; Lexa, M.; Vanat, I.; Hobza, R.; Kejnovsky, E. Nested plant LTR retrotransposons target specific regions of other elements, while all LTR retrotransposons often target palindromes and nucleosome-occupied regions: In silico study. Mob. DNA 2019, 10, 50. [Google Scholar] [CrossRef]
  131. Wei, L.; Xiao, M.; An, Z.; Ma, B.; Mason, A.S.; Qian, W.; Li, J.; Fu, D. New Insights into Nested Long Terminal Repeat Retrotransposons in Brassica Species. Mol. Plant 2013, 6, 470–482. [Google Scholar] [CrossRef]
  132. Pereira, H.S.; Barão, A.; Delgado, M.; Morais-Cecílio, L.; Viegas, W. Genomic analysis of Grapevine Retrotransposon 1 (Gret1) in Vitis vinifera. Theor. Appl. Genet. 2005, 111, 871–878. [Google Scholar] [CrossRef]
  133. Butelli, E.; Licciardello, C.; Zhang, Y.; Liu, J.; Mackay, S.; Bailey, P.; Reforgiato-Recupero, G.; Martin, C. Retrotransposons control fruit-specific, cold-dependent accumulation of anthocyanins in blood oranges. Plant Cell 2012, 24, 1242–1255. [Google Scholar] [CrossRef]
  134. Xiao, H.; Jiang, N.; Schaffner, E.; Stockinger, E.J.; van der Knaap, E. A Retrotransposon-Mediated Gene Duplication Underlies Morphological Variation of Tomato Fruit. Science 2008, 319, 1527–1530. [Google Scholar] [CrossRef]
  135. Michaels, S.D.; Ditta, G.; Gustafson-Brown, C.; Pelaz, S.; Yanofsky, M.; Amasino, R.M. AGL24 acts as a promoter of flowering in Arabidopsis and is positively regulated by vernalization. Plant J. 2003, 33, 867–874. [Google Scholar] [CrossRef]
  136. Zhang, J.; Zhang, F.; Peterson, T. Transposition of Reversed Ac Element Ends Generates Novel Chimeric Genes in Maize. PLoS Genet. 2006, 2, e164. [Google Scholar] [CrossRef]
  137. Yu, C.; Zhang, J.; Peterson, T. Genome rearrangements in maize induced by alternative transposition of reversed ac/ds termini. Genetics 2011, 188, 59–67. [Google Scholar] [CrossRef]
  138. Du, C.; Hoffman, A.; He, L.; Caronna, J.; Dooner, H.K. The complete Ac/Ds transposon family of maize. BMC Genom. 2011, 12, 588. [Google Scholar] [CrossRef]
  139. Xuan, Y.H.; Piao, H.L.; Je, B.I.; Park, S.J.; Park, S.H.; Huang, J.; Zhang, J.B.; Peterson, T.; Han, C.-D. Transposon Ac/Ds-induced chromosomal rearrangements at the rice OsRLG5 locus. Nucleic Acids Res. 2011, 39, e149. [Google Scholar] [CrossRef]
  140. Shen, D.; Song, C.; Miskey, C.; Chan, S.; Guan, Z.; Sang, Y.; Wang, Y.; Chen, C.; Wang, X.; Müller, F.; et al. A native, highly active Tc1/mariner transposon from zebrafish (ZB) offers an efficient genetic manipulation tool for vertebrates. Nucleic Acids Res. 2021, 49, 2126–2140. [Google Scholar] [CrossRef]
  141. Jiang, N.; Wessler, S.R. Insertion preference of maize and rice miniature inverted repeat transposable elements as revealed by the analysis of nested elements. Plant Cell 2001, 13, 2553–2564. [Google Scholar]
  142. Bureau, T.E.; Wessler, S.R. Stowaway: A new family of inverted repeat elements associated with the genes of both monocotyledonous and dicotyledonous plants. Plant Cell 1994, 6, 907–916. [Google Scholar]
  143. Panini, M.; Chiesa, O.; Troczka, B.J.; Mallott, M.; Manicardi, G.C.; Cassanelli, S.; Cominelli, F.; Hayward, A.; Mazzoni, E.; Bass, C. Transposon-mediated insertional mutagenesis unmasks recessive insecticide resistance in the aphid Myzus persicae. Proc. Natl. Acad. Sci. USA 2021, 118, e2100559118. [Google Scholar] [CrossRef]
  144. Quadrana, L.; Bortolini Silveira, A.; Mayhew, G.F.; LeBlanc, C.; Martienssen, R.A.; Jeddeloh, J.A.; Colot, V. The Arabidopsis thaliana mobilome and its impact at the species level. eLife 2016, 5, e15716. [Google Scholar] [CrossRef]
  145. Weng, M.L.; Becker, C.; Hildebrandt, J.; Neumann, M.; Rutter, M.T.; Shaw, R.G.; Weigel, D.; Fenster, C.B. Fine-Grained Analysis of Spontaneous Mutation Spectrum and Frequency in Arabidopsis thaliana. Genetics 2019, 211, 703–714. [Google Scholar] [CrossRef]
  146. Bourgeois, Y.; Boissinot, S. On the Population Dynamics of Junk: A Review on the Population Genomics of Transposable Elements. Genes 2019, 10, 419. [Google Scholar] [CrossRef]
  147. Wicker, T.; Gundlach, H.; Spannagl, M.; Uauy, C.; Borrill, P.; Ramírez-González, R.H.; De Oliveira, R.; Mayer, K.F.X.; Paux, E.; Choulet, F.; et al. Impact of transposable elements on genome structure and evolution in bread wheat. Genome Biol. 2018, 19, 103. [Google Scholar] [CrossRef]
  148. Middleton, C.P.; Senerchia, N.; Stein, N.; Akhunov, E.D.; Keller, B.; Wicker, T.; Kilian, B. Sequencing of chloroplast genomes from wheat, barley, rye and their relatives provides a detailed insight into the evolution of the Triticeae tribe. PLoS ONE 2014, 9, e85761. [Google Scholar] [CrossRef]
  149. Lisch, D. How important are transposons for plant evolution? Nat. Rev. Genet. 2013, 14, 49–61. [Google Scholar] [CrossRef]
  150. Negi, P.; Rai, A.N.; Suprasanna, P. Moving through the Stressed Genome: Emerging Regulatory Roles for Transposons in Plant Stress Response. Front. Plant Sci. 2016, 7, 1448. [Google Scholar] [CrossRef]
  151. Rejeb, I.B.; Pastor, V.; Mauch-Mani, B. Plant Responses to Simultaneous Biotic and Abiotic Stress: Molecular Mechanisms. Plants 2014, 3, 458–475. [Google Scholar] [CrossRef]
  152. Etchegaray, E.; Naville, M.; Volff, J.-N.; Haftek-Terreau, Z. Transposable element-derived sequences in vertebrate development. Mob. DNA 2021, 12, 1. [Google Scholar] [CrossRef] [PubMed]
  153. Pereira, V. Insertion bias and purifying selection of retrotransposons in the Arabidopsis thaliana genome. Genome Biol. 2004, 5, R79. [Google Scholar] [CrossRef] [PubMed]
  154. Lee, S.-I.; Kim, N.-S. Transposable elements and genome size variations in plants. Genom. Inform. 2014, 12, 87–97. [Google Scholar] [CrossRef] [PubMed]
  155. Galindo-González, L.; Sarmiento, F.; Quimbaya, M.A. Shaping Plant Adaptability, Genome Structure and Gene Expression through Transposable Element Epigenetic Control: Focus on Methylation. Agronomy 2018, 8, 180. [Google Scholar] [CrossRef]
  156. Ramakrishnan, M.; Zhou, M.-B.; Pan, C.-F.; Hänninen, H.; Tang, D.-Q.; Vinod, K.K. Nuclear export signal (NES) of transposases affects the transposition activity of mariner-like elements Ppmar1 and Ppmar2 of moso bamboo. Mob. DNA 2019, 10, 35. [Google Scholar] [CrossRef]
  157. Ramakrishnan, M.; Zhou, M.; Pan, C.; Hanninen, H.; Yrjala, K.; Vinod, K.K.; Tang, D. Affinities of Terminal Inverted Repeats to DNA Binding Domain of Transposase Affect the Transposition Activity of Bamboo Ppmar2 Mariner-Like Element. Int. J. Mol. Sci. 2019, 20, 3692. [Google Scholar] [CrossRef]
  158. Lohe, A.R.; Hartl, D.L. Autoregulation of mariner transposase activity by overproduction and dominant-negative complementation. Mol. Biol. Evol. 1996, 13, 549–555. [Google Scholar] [CrossRef]
  159. Wang, X.; Duan, C.G.; Tang, K.; Wang, B.; Zhang, H.; Lei, M.; Lu, K.; Mangrauthia, S.K.; Wang, P.; Zhu, G.; et al. RNA-binding protein regulates plant DNA methylation by controlling mRNA processing at the intronic heterochromatin-containing gene IBM1. Proc. Natl. Acad. Sci. USA 2013, 110, 15467–15472. [Google Scholar] [CrossRef]
  160. Slotkin, R.K.; Vaughn, M.; Borges, F.; Tanurdzić, M.; Becker, J.D.; Feijó, J.A.; Martienssen, R.A. Epigenetic reprogramming and small RNA silencing of transposable elements in pollen. Cell 2009, 136, 461–472. [Google Scholar] [CrossRef]
  161. Klein, S.J.; O’Neill, R.J. Transposable elements: Genome innovation, chromosome diversity, and centromere conflict. Chromosome Res. 2018, 26, 5–23. [Google Scholar] [CrossRef]
  162. Martin, M.-L.; Jose, L.G.-P. DNA Transposons: Nature and Applications in Genomics. Curr. Genom. 2010, 11, 115–128. [Google Scholar]
  163. Sigman, M.J.; Slotkin, R.K. The First Rule of Plant Transposable Element Silencing: Location, Location, Location. Plant Cell 2016, 28, 304–313. [Google Scholar] [CrossRef]
  164. Joly-Lopez, Z.; Forczek, E.; Vello, E.; Hoen, D.R.; Tomita, A.; Bureau, T.E. Abiotic Stress Phenotypes Are Associated with Conserved Genes Derived from Transposable Elements. Front. Plant Sci. 2017, 8, 2027. [Google Scholar] [CrossRef]
  165. Ito, H.; Kim, J.-M.; Matsunaga, W.; Saze, H.; Matsui, A.; Endo, T.A.; Harukawa, Y.; Takagi, H.; Yaegashi, H.; Masuta, Y.; et al. A Stress-Activated Transposon in Arabidopsis Induces Transgenerational Abscisic Acid Insensitivity. Sci. Rep. 2016, 6, 23181. [Google Scholar] [CrossRef]
  166. Tikhonov, A.P.; SanMiguel, P.J.; Nakajima, Y.; Gorenstein, N.M.; Bennetzen, J.L.; Avramova, Z. Colinearity and its exceptions in orthologous adh regions of maize and sorghum. Proc. Natl. Acad. Sci. USA 1999, 96, 7409–7414. [Google Scholar] [CrossRef]
  167. Serrato-Capuchina, A.; Matute, D.R. The role of transposable elements in speciation. Genes 2018, 9, 254. [Google Scholar] [CrossRef]
  168. Li, S.; Ramakrishnan, M.; Vinod, K.K.; Kalendar, R.; Yrjälä, K.; Zhou, M. Development and Deployment of High-Throughput Retrotransposon-Based Markers Reveal Genetic Diversity and Population Structure of Asian Bamboo. Forests 2020, 11, 31. [Google Scholar] [CrossRef]
  169. Roy, S. Maintenance of genome stability in plants: Repairing DNA double strand breaks and chromatin structure stability. Front. Plant Sci. 2014, 5, 487. [Google Scholar] [CrossRef]
  170. Kovalchuk, I. Transgenerational Genome Instability in Plants. In Genome Stability; Kovalchuk, I., Kovalchuk, O., Eds.; Academic Press: Boston, MA, USA, 2016; Chpater 36; pp. 615–633. [Google Scholar]
  171. Manova, V.; Gruszka, D. DNA damage and repair in plants—From models to crops. Front. Plant Sci. 2015, 6, 885. [Google Scholar] [CrossRef]
  172. Waititu, J.K.; Zhang, C.; Liu, J.; Wang, H. Plant Non-Coding RNAs: Origin, Biogenesis, Mode of Action and Their Roles in Abiotic Stress. Int. J. Mol. Sci. 2020, 21, 8401. [Google Scholar] [CrossRef]
  173. Nosaka, M.; Itoh, J.; Nagato, Y.; Ono, A.; Ishiwata, A.; Sato, Y. Role of transposon-derived small RNAs in the interplay between genomes and parasitic DNA in rice. PLoS Genet. 2012, 8, e1002953. [Google Scholar] [CrossRef]
  174. Ahmed, W.; Xia, Y.; Li, R.; Bai, G.; Siddique, K.H.M.; Guo, P. Non-coding RNAs: Functional roles in the regulation of stress response in Brassica crops. Genomics 2020, 112, 1419–1424. [Google Scholar] [CrossRef] [PubMed]
  175. Zhao, J.; He, Q.; Chen, G.; Wang, L.; Jin, B. Regulation of Non-coding RNAs in Heat Stress Responses of Plants. Front. Plant Sci. 2016, 7, 1213. [Google Scholar] [CrossRef] [PubMed]
  176. Guo, Q.; Liu, Q.; Smith, N.A.; Liang, G.; Wang, M.-B. RNA Silencing in Plants: Mechanisms, Technologies and Applications in Horticultural Crops. Curr. Genom. 2016, 17, 476–489. [Google Scholar] [CrossRef] [PubMed]
  177. Asefpour Vakilian, K. Machine learning improves our knowledge about miRNA functions towards plant abiotic stresses. Sci. Rep. 2020, 10, 3041. [Google Scholar] [CrossRef]
  178. Huang, S.; Zhou, J.; Gao, L.; Tang, Y. Plant miR397 and its functions. Funct. Plant Biol. 2021, 48, 361–370. [Google Scholar] [CrossRef]
  179. Iglesias, M.J.; Terrile, M.C.; Windels, D.; Lombardo, M.C.; Bartoli, C.G.; Vazquez, F.; Estelle, M.; Casalongué, C.A. MiR393 Regulation of Auxin Signaling and Redox-Related Components during Acclimation to Salinity in Arabidopsis. PLoS ONE 2014, 9, e107678. [Google Scholar] [CrossRef]
  180. Borsani, O.; Zhu, J.; Verslues, P.E.; Sunkar, R.; Zhu, J.K. Endogenous siRNAs derived from a pair of natural cis-antisense transcripts regulate salt tolerance in Arabidopsis. Cell 2005, 123, 1279–1291. [Google Scholar] [CrossRef]
  181. Jha, U.C.; Nayyar, H.; Jha, R.; Khurshid, M.; Zhou, M.; Mantri, N.; Siddique, K.H.M. Long non-coding RNAs: Emerging players regulating plant abiotic stress response and adaptation. BMC Plant Biol. 2020, 20, 466. [Google Scholar] [CrossRef]
  182. Wang, L.; Cho, K.B.; Li, Y.; Tao, G.; Xie, Z.; Guo, B. Long Noncoding RNA (lncRNA)-Mediated Competing Endogenous RNA Networks Provide Novel Potential Biomarkers and Therapeutic Targets for Colorectal Cancer. Int. J. Mol. Sci. 2019, 20, 5758. [Google Scholar] [CrossRef]
  183. Wu, H.-J.; Wang, Z.-M.; Wang, M.; Wang, X.-J. Widespread Long Noncoding RNAs as Endogenous Target Mimics for MicroRNAs in Plants. Plant Physiol. 2013, 161, 1875–1884. [Google Scholar] [CrossRef]
  184. Allen, E.; Xie, Z.; Gustafson, A.M.; Carrington, J.C. microRNA-directed phasing during trans-acting siRNA biogenesis in plants. Cell 2005, 121, 207–221. [Google Scholar] [CrossRef]
  185. Rajagopalan, R.; Vaucheret, H.; Trejo, J.; Bartel, D.P. A diverse and evolutionarily fluid set of microRNAs in Arabidopsis thaliana. Genes Dev. 2006, 20, 3407–3425. [Google Scholar] [CrossRef]
  186. Lurin, C.; Andrés, C.; Aubourg, S.; Bellaoui, M.; Bitton, F.; Bruyère, C.; Caboche, M.; Debast, C.; Gualberto, J.; Hoffmann, B.; et al. Genome-wide analysis of Arabidopsis pentatricopeptide repeat proteins reveals their essential role in organelle biogenesis. Plant Cell 2004, 16, 2089–2103. [Google Scholar] [CrossRef]
  187. Kume, K.; Tsutsumi, K.; Saitoh, Y. TAS1 trans-acting siRNA targets are differentially regulated at low temperature, and TAS1 trans-acting siRNA mediates temperature-controlled At1g51670 expression. Biosci. Biotechnol. Biochem. 2010, 74, 1435–1440. [Google Scholar] [CrossRef]
  188. Li, S.; Liu, J.; Liu, Z.; Li, X.; Wu, F.; He, Y. HEAT-INDUCED TAS1 TARGET1 Mediates Thermotolerance via HEAT STRESS TRANSCRIPTION FACTOR A1a-Directed Pathways in Arabidopsis. Plant Cell 2014, 26, 1764–1780. [Google Scholar] [CrossRef]
  189. Hsieh, L.C.; Lin, S.I.; Shih, A.C.; Chen, J.W.; Lin, W.Y.; Tseng, C.Y.; Li, W.H.; Chiou, T.J. Uncovering small RNA-mediated responses to phosphate deficiency in Arabidopsis by deep sequencing. Plant Physiol. 2009, 151, 2120–2132. [Google Scholar] [CrossRef]
  190. Luo, Q.J.; Mittal, A.; Jia, F.; Rock, C.D. An autoregulatory feedback loop involving PAP1 and TAS4 in response to sugars in Arabidopsis. Plant Mol. Biol. 2012, 80, 117–129. [Google Scholar] [CrossRef]
  191. Ito, H.; Gaubert, H.; Bucher, E.; Mirouze, M.; Vaillant, I.; Paszkowski, J. An siRNA pathway prevents transgenerational retrotransposition in plants subjected to stress. Nature 2011, 472, 115–119. [Google Scholar] [CrossRef]
  192. Matsunaga, W.; Kobayashi, A.; Kato, A.; Ito, H. The effects of heat induction and the siRNA biogenesis pathway on the transgenerational transposition of ONSEN, a copia-like retrotransposon in Arabidopsis thaliana. Plant Cell Physiol. 2012, 53, 824–833. [Google Scholar] [CrossRef]
  193. Stief, A.; Brzezinka, K.; Lämke, J.; Bäurle, I. Epigenetic responses to heat stress at different time scales and the involvement of small RNAs. Plant Signal. Behav. 2014, 9, e970430. [Google Scholar] [CrossRef]
  194. Boyko, A.; Kovalchuk, I. Transgenerational response to stress in Arabidopsis thaliana. Plant Signal. Behav. 2010, 5, 995–998. [Google Scholar] [CrossRef]
  195. Chen, M.; Lv, S.; Meng, Y. Epigenetic performers in plants. Dev. Growth Differ. 2010, 52, 555–566. [Google Scholar] [CrossRef]
  196. Sridha, S.; Wu, K. Identification of AtHD2C as a novel regulator of abscisic acid responses in Arabidopsis. Plant J. 2006, 46, 124–133. [Google Scholar] [CrossRef]
  197. Aufsatz, W.; Mette, M.F.; Van der Winden, J.; Matzke, M.; Matzke, A.J. HDA6, a putative histone deacetylase needed to enhance DNA methylation induced by double-stranded RNA. EMBO J. 2002, 21, 6832–6841. [Google Scholar] [CrossRef]
  198. To, T.K.; Kim, J.-M.; Matsui, A.; Kurihara, Y.; Morosawa, T.; Ishida, J.; Tanaka, M.; Endo, T.; Kakutani, T.; Toyoda, T. Arabidopsis HDA6 regulates locus-directed heterochromatin silencing in cooperation with MET1. PLoS Genet. 2011, 7, e1002055. [Google Scholar] [CrossRef]
  199. Song, Y.; Wu, K.; Dhaubhadel, S.; An, L.; Tian, L. Arabidopsis DNA methyltransferase AtDNMT2 associates with histone deacetylase AtHD2s activity. Biochem. Biophys. Res. Commun. 2010, 396, 187–192. [Google Scholar] [CrossRef]
  200. Aung, K.; Lin, S.-I.; Wu, C.-C.; Huang, Y.-T.; Su, C.-l.; Chiou, T.J. pho2, a phosphate overaccumulator, is caused by a nonsense mutation in a microRNA399 target gene. Plant Physiol. 2006, 141, 1000–1011. [Google Scholar] [CrossRef]
  201. Franco-Zorrilla, J.M.; Valli, A.; Todesco, M.; Mateos, I.; Puga, M.I.; Rubio-Somoza, I.; Leyva, A.; Weigel, D.; García, J.A.; Paz-Ares, J. Target mimicry provides a new mechanism for regulation of microRNA activity. Nat. Genet. 2007, 39, 1033–1037. [Google Scholar] [CrossRef]
  202. Wang, H.; Chung, P.J.; Liu, J.; Jang, I.C.; Kean, M.J.; Xu, J.; Chua, N.H. Genome-wide identification of long noncoding natural antisense transcripts and their responses to light in Arabidopsis. Genome Res. 2014, 24, 444–453. [Google Scholar] [CrossRef]
  203. Wunderlich, M.; Gross-Hardt, R.; Schöffl, F. Heat shock factor HSFB2a involved in gametophyte development of Arabidopsis thaliana and its expression is controlled by a heat-inducible long non-coding antisense RNA. Plant Mol. Biol. 2014, 85, 541–550. [Google Scholar] [CrossRef]
  204. Csorba, T.; Questa, J.I.; Sun, Q.; Dean, C. Antisense COOLAIR mediates the coordinated switching of chromatin states at FLC during vernalization. Proc. Natl. Acad. Sci. USA 2014, 111, 16160–16165. [Google Scholar] [CrossRef] [PubMed]
  205. Heo, J.B.; Sung, S. Vernalization-mediated epigenetic silencing by a long intronic noncoding RNA. Science 2011, 331, 76–79. [Google Scholar] [CrossRef] [PubMed]
  206. Popova, O.V.; Dinh, H.Q.; Aufsatz, W.; Jonak, C. The RdDM pathway is required for basal heat tolerance in Arabidopsis. Mol. Plant. 2013, 6, 396–410. [Google Scholar] [CrossRef] [PubMed]
  207. Xu, R.; Wang, Y.; Zheng, H.; Lu, W.; Wu, C.; Huang, J.; Yan, K.; Yang, G.; Zheng, C. Salt-induced transcription factor MYB74 is regulated by the RNA-directed DNA methylation pathway in Arabidopsis. J. Exp. Bot. 2015, 66, 5997–6008. [Google Scholar] [CrossRef]
  208. Wang, L.; Yu, X.; Wang, H.; Lu, Y.Z.; de Ruiter, M.; Prins, M.; He, Y.K. A novel class of heat-responsive small RNAs derived from the chloroplast genome of Chinese cabbage (Brassica rapa). BMC Genom. 2011, 12, 289. [Google Scholar] [CrossRef]
  209. Yu, X.; Yang, J.; Li, X.; Liu, X.; Sun, C.; Wu, F.; He, Y. Global analysis of cis-natural antisense transcripts and their heat-responsive nat-siRNAs in Brassica rapa. BMC Plant Biol. 2013, 13, 208. [Google Scholar] [CrossRef]
  210. Song, X.; Liu, G.; Huang, Z.; Duan, W.; Tan, H.; Li, Y.; Hou, X. Temperature expression patterns of genes and their coexpression with LncRNAs revealed by RNA-Seq in non-heading Chinese cabbage. BMC Genom. 2016, 17, 297. [Google Scholar] [CrossRef]
  211. Furini, A.; Koncz, C.; Salamini, F.; Bartels, D.J.T.E.J. High level transcription of a member of a repeated gene family confers dehydration tolerance to callus tissue of Craterostigma plantagineum. EMBO J. 1997, 16, 3599–3608. [Google Scholar] [CrossRef]
  212. Xia, J.; Zeng, C.; Chen, Z.; Zhang, K.; Chen, X.; Zhou, Y.; Song, S.; Lu, C.; Yang, R.; Yang, Z. Endogenous small-noncoding RNAs and their roles in chilling response and stress acclimation in Cassava. BMC Genom. 2014, 15, 634. [Google Scholar] [CrossRef]
  213. Xu, X.W.; Zhou, X.H.; Wang, R.R.; Peng, W.L.; An, Y.; Chen, L.L. Functional analysis of long intergenic non-coding RNAs in phosphate-starved rice using competing endogenous RNA network. Sci. Rep. 2016, 6, 20715. [Google Scholar] [CrossRef]
  214. Cruz de Carvalho, M.H.; Sun, H.X.; Bowler, C.; Chua, N.H. Noncoding and coding transcriptome responses of a marine diatom to phosphate fluctuations. New Phytol. 2016, 210, 497–510. [Google Scholar] [CrossRef]
  215. Chen, M.; Wang, C.; Bao, H.; Chen, H.; Wang, Y. Genome-wide identification and characterization of novel lncRNAs in Populus under nitrogen deficiency. Mol. Genet Genom. 2016, 291, 1663–1680. [Google Scholar] [CrossRef]
  216. Shuai, P.; Liang, D.; Tang, S.; Zhang, Z.; Ye, C.Y.; Su, Y.; Xia, X.; Yin, W. Genome-wide identification and functional prediction of novel and drought-responsive lincRNAs in Populus trichocarpa. J. Exp. Bot. 2014, 65, 4975–4983. [Google Scholar] [CrossRef]
  217. Huang, W.; Xian, Z.; Hu, G.; Li, Z. SlAGO4A, a core factor of RNA-directed DNA methylation (RdDM) pathway, plays an important role under salt and drought stress in tomato. Mol. Breed. 2016, 36, 28. [Google Scholar] [CrossRef]
  218. Yao, Y.; Ni, Z.; Peng, H.; Sun, F.; Xin, M.; Sunkar, R.; Zhu, J.K.; Sun, Q. Non-coding small RNAs responsive to abiotic stress in wheat (Triticum aestivum L.). Funct. Integr. Genom. 2010, 10, 187–190. [Google Scholar] [CrossRef]
  219. Tang, Z.; Zhang, L.; Xu, C.; Yuan, S.; Zhang, F.; Zheng, Y.; Zhao, C. Uncovering small RNA-mediated responses to cold stress in a wheat thermosensitive genic male-sterile line by deep sequencing. Plant Physiol. 2012, 159, 721–738. [Google Scholar] [CrossRef]
  220. Xin, M.; Wang, Y.; Yao, Y.; Song, N.; Hu, Z.; Qin, D.; Xie, C.; Peng, H.; Ni, Z.; Sun, Q. Identification and characterization of wheat long non-protein coding RNAs responsive to powdery mildew infection and heat stress by using microarray analysis and SBS sequencing. BMC Plant Biol. 2011, 11, 61. [Google Scholar] [CrossRef]
  221. Zhang, W.; Han, Z.; Guo, Q.; Liu, Y.; Zheng, Y.; Wu, F.; Jin, W. Identification of maize long non-coding RNAs responsive to drought stress. PLoS ONE 2014, 9, e98958. [Google Scholar] [CrossRef]
  222. Kalendar, R.; Sabot, F.; Rodriguez, F.; Karlov, G.I.; Natali, L.; Alix, K. Editorial: Mobile Elements and Plant Genome Evolution, Comparative Analyzes and Computational Tools. Front. Plant Sci. 2021, 12, 735134. [Google Scholar] [CrossRef]
  223. Lerat, E.; Casacuberta, J.; Chaparro, C.; Vieira, C. On the Importance to Acknowledge Transposable Elements in Epigenomic Analyses. Genes 2019, 10, 258. [Google Scholar] [CrossRef]
  224. Sanchez, D.H.; Paszkowski, J. Heat-Induced Release of Epigenetic Silencing Reveals the Concealed Role of an Imprinted Plant Gene. PLoS Genet. 2014, 10, e1004806. [Google Scholar] [CrossRef]
  225. Cavrak, V.V.; Lettner, N.; Jamge, S.; Kosarewicz, A.; Bayer, L.M.; Mittelsten Scheid, O. How a Retrotransposon Exploits the Plant’s Heat Stress Response for Its Activation. PLoS Genet. 2014, 10, e1004115. [Google Scholar] [CrossRef]
  226. Choi, J.Y.; Lee, Y.C.G. Double-edged sword: The evolutionary consequences of the epigenetic silencing of transposable elements. PLoS Genet. 2020, 16, e1008872. [Google Scholar] [CrossRef]
  227. Law, J.A.; Ausin, I.; Johnson, L.M.; Vashisht, A.A.; Zhu, J.K.; Wohlschlegel, J.A.; Jacobsen, S.E. A protein complex required for polymerase V transcripts and RNA- directed DNA methylation in Arabidopsis. Curr. Biol. 2010, 20, 951–956. [Google Scholar] [CrossRef]
  228. McCue, A.D.; Panda, K.; Nuthikattu, S.; Choudury, S.G.; Thomas, E.N.; Slotkin, R.K. ARGONAUTE 6 bridges transposable element mRNA-derived siRNAs to the establishment of DNA methylation. EMBO J. 2015, 34, 20–35. [Google Scholar] [CrossRef]
  229. 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]
  230. Zhang, H.; Zhu, J.K. Active DNA demethylation in plants and animals. Cold Spring Harb. Symp. Quant. Biol. 2012, 77, 161–173. [Google Scholar] [CrossRef]
  231. Zhang, X.; Yazaki, J.; Sundaresan, A.; Cokus, S.; Chan, S.W.; Chen, H.; Henderson, I.R.; Shinn, P.; Pellegrini, M.; Jacobsen, S.E.; et al. Genome-wide high-resolution mapping and functional analysis of DNA methylation in arabidopsis. Cell 2006, 126, 1189–1201. [Google Scholar] [CrossRef]
  232. 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]
  233. 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]
  234. 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] [PubMed]
  235. 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]
  236. Takuno, S.; Gaut, B.S. Gene body methylation is conserved between plant orthologs and is of evolutionary consequence. Proc. Natl. Acad. Sci. USA 2013, 110, 1797–1802. [Google Scholar] [CrossRef] [PubMed]
  237. Saze, H.; Kitayama, J.; Takashima, K.; Miura, S.; Harukawa, Y.; Ito, T.; Kakutani, T. Mechanism for full-length RNA processing of Arabidopsis genes containing intragenic heterochromatin. Nat. Commun. 2013, 4, 2301. [Google Scholar] [CrossRef]
  238. Lei, M.; La, H.; Lu, K.; Wang, P.; Miki, D.; Ren, Z.; Duan, C.G.; Wang, X.; Tang, K.; Zeng, L.; et al. Arabidopsis EDM2 promotes IBM1 distal polyadenylation and regulates genome DNA methylation patterns. Proc. Natl. Acad. Sci. USA 2014, 111, 527–532. [Google Scholar] [CrossRef]
  239. Duan, C.G.; Wang, X.; Zhang, L.; Xiong, X.; Zhang, Z.; Tang, K.; Pan, L.; Hsu, C.C.; Xu, H.; Tao, W.A.; et al. A protein complex regulates RNA processing of intronic heterochromatin-containing genes in Arabidopsis. Proc. Natl. Acad. Sci. USA 2017, 114, E7377–E7384. [Google Scholar] [CrossRef]
  240. Feng, S.; Cokus, S.J.; Schubert, V.; Zhai, J.; Pellegrini, M.; Jacobsen, S.E. Genome-wide Hi-C analyses in wild-type and mutants reveal high-resolution chromatin interactions in Arabidopsis. Mol. Cell 2014, 55, 694–707. [Google Scholar] [CrossRef]
  241. Grob, S.; Schmid, M.W.; Grossniklaus, U. Hi-C analysis in Arabidopsis identifies the KNOT, a structure with similarities to the flamenco locus of Drosophila. Mol. Cell 2014, 55, 678–693. [Google Scholar] [CrossRef]
  242. Ibarra, C.A.; Feng, X.; Schoft, V.K.; Hsieh, T.F.; Uzawa, R.; Rodrigues, J.A.; Zemach, A.; Chumak, N.; Machlicova, A.; Nishimura, T.; et al. Active DNA demethylation in plant companion cells reinforces transposon methylation in gametes. Science 2012, 337, 1360–1364. [Google Scholar] [CrossRef]
  243. Martínez, G.; Panda, K.; Köhler, C.; Slotkin, R.K. Silencing in sperm cells is directed by RNA movement from the surrounding nurse cell. Nat. Plants 2016, 2, 16030. [Google Scholar] [CrossRef]
  244. Gehring, M.; Bubb, K.L.; Henikoff, S. Extensive demethylation of repetitive elements during seed development underlies gene imprinting. Science 2009, 324, 1447–1451. [Google Scholar] [CrossRef]
  245. Ingouff, M.; Selles, B.; Michaud, C.; Vu, T.M.; Berger, F.; Schorn, A.J.; Autran, D.; Van Durme, M.; Nowack, M.K.; Martienssen, R.A.; et al. Live-cell analysis of DNA methylation during sexual reproduction in Arabidopsis reveals context and sex-specific dynamics controlled by noncanonical RdDM. Genes Dev. 2017, 31, 72–83. [Google Scholar] [CrossRef]
  246. 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]
  247. Jullien, P.E.; Katz, A.; Oliva, M.; Ohad, N.; Berger, F. Polycomb group complexes self-regulate imprinting of the Polycomb group gene MEDEA in Arabidopsis. Curr. Biol. 2006, 16, 486–492. [Google Scholar] [CrossRef]
  248. Dong, X.; Zhang, M.; Chen, J.; Peng, L.; Zhang, N.; Wang, X.; Lai, J. Dynamic and Antagonistic Allele-Specific Epigenetic Modifications Controlling the Expression of Imprinted Genes in Maize Endosperm. Mol. Plant 2017, 10, 442–455. [Google Scholar] [CrossRef]
  249. 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, 4062. [Google Scholar] [CrossRef]
  250. Wang, Y.; Xue, X.; Zhu, J.K.; Dong, J. Demethylation of ERECTA receptor genes by IBM1 histone demethylase affects stomatal development. Development 2016, 143, 4452–4461. [Google Scholar] [CrossRef]
  251. Liu, R.; How-Kit, A.; Stammitti, L.; Teyssier, E.; Rolin, D.; Mortain-Bertrand, A.; Halle, S.; Liu, M.; Kong, J.; Wu, C.; et al. A DEMETER-like DNA demethylase governs tomato fruit ripening. Proc. Natl. Acad. Sci. USA 2015, 112, 10804–10809. [Google Scholar] [CrossRef]
  252. Torres, D.E.; Thomma, B.P.H.J.; Seidl, M.F. Transposable Elements Contribute to Genome Dynamics and Gene Expression Variation in the Fungal Plant Pathogen Verticillium dahliae. Genome Biol. Evol. 2021, 13, evab135. [Google Scholar] [CrossRef]
  253. Lisch, D. Epigenetic regulation of transposable elements in plants. Annu. Rev. Plant Biol. 2009, 60, 43–66. [Google Scholar] [CrossRef]
  254. Ahmed, I.; Sarazin, A.; Bowler, C.; Colot, V.; Quesneville, H. Genome-wide evidence for local DNA methylation spreading from small RNA-targeted sequences in Arabidopsis. Nucleic Acids Res. 2011, 39, 6919–6931. [Google Scholar] [CrossRef]
  255. Hollister, J.D.; Smith, L.M.; Guo, Y.L.; Ott, F.; Weigel, D.; Gaut, B.S. Transposable elements and small RNAs contribute to gene expression divergence between Arabidopsis thaliana and Arabidopsis lyrata. Proc. Natl. Acad. Sci. USA 2011, 108, 2322–2327. [Google Scholar] [CrossRef]
  256. Sotelo-Silveira, M.; Chávez Montes, R.A.; Sotelo-Silveira, J.R.; Marsch-Martínez, N.; de Folter, S. Entering the Next Dimension: Plant Genomes in 3D. Trends Plant Sci. 2018, 23, 598–612. [Google Scholar] [CrossRef]
  257. Kumar, A. Jump around: Transposons in and out of the laboratory. F1000Research 2020, 9, 135. [Google Scholar] [CrossRef]
  258. Rymen, B.; Ferrafiat, L.; Blevins, T. Non-coding RNA polymerases that silence transposable elements and reprogram gene expression in plants. Transcription 2020, 11, 172–191. [Google Scholar] [CrossRef]
  259. Xu, L.; Jiang, H. Writing and Reading Histone H3 Lysine 9 Methylation in Arabidopsis. Front. Plant Sci. 2020, 11, 452. [Google Scholar] [CrossRef]
  260. Scheid, R.; Chen, J.; Zhong, X. Biological role and mechanism of chromatin readers in plants. Curr. Opin. Plant Biol. 2021, 61, 102008. [Google Scholar] [CrossRef]
  261. Pease, N.A.; Nguyen, P.H.B.; Woodworth, M.A.; Ng, K.K.H.; Irwin, B.; Vaughan, J.C.; Kueh, H.Y. Tunable, division-independent control of gene activation timing by a polycomb switch. Cell Rep. 2021, 34, 108888. [Google Scholar] [CrossRef]
  262. Alix, K.; Gérard, P.R.; Schwarzacher, T.; Heslop-Harrison, J.S. Polyploidy and interspecific hybridization: Partners for adaptation, speciation and evolution in plants. Ann. Bot. 2017, 120, 183–194. [Google Scholar] [CrossRef]
  263. Kyriakidou, M.; Tai, H.H.; Anglin, N.L.; Ellis, D.; Strömvik, M.V. Current Strategies of Polyploid Plant Genome Sequence Assembly. Front. Plant Sci. 2018, 9, 1660. [Google Scholar] [CrossRef]
  264. Ritter, E.J.; Niederhuth, C.E. Intertwined evolution of plant epigenomes and genomes. Curr. Opin. Plant Biol. 2021, 61, 101990. [Google Scholar] [CrossRef] [PubMed]
  265. Dwivedi, S.L.; Scheben, A.; Edwards, D.; Spillane, C.; Ortiz, R. Assessing and Exploiting Functional Diversity in Germplasm Pools to Enhance Abiotic Stress Adaptation and Yield in Cereals and Food Legumes. Front. Plant Sci. 2017, 8, 1461. [Google Scholar] [CrossRef] [PubMed]
  266. Kim, J.-H. Chromatin Remodeling and Epigenetic Regulation in Plant DNA Damage Repair. Int. J. Mol. Sci. 2019, 20, 4093. [Google Scholar] [CrossRef] [PubMed]
  267. Liu, N.; Fromm, M.; Avramova, Z. H3K27me3 and H3K4me3 chromatin environment at super-induced dehydration stress memory genes of Arabidopsis thaliana. Mol. Plant 2014, 7, 502–513. [Google Scholar] [CrossRef]
  268. Mehraj, H.; Takahashi, S.; Miyaji, N.; Akter, A.; Suzuki, Y.; Seki, M.; Dennis, E.S.; Fujimoto, R. Characterization of Histone H3 Lysine 4 and 36 Tri-methylation in Brassica rapa L. Front. Plant Sci. 2021, 12, 785. [Google Scholar] [CrossRef]
  269. Gan, E.-S.; Xu, Y.; Ito, T. Dynamics of H3K27me3 methylation and demethylation in plant development. Plant Signal. Behav. 2015, 10, e1027851. [Google Scholar] [CrossRef]
  270. Bhadouriya, S.L.; Mehrotra, S.; Basantani, M.K.; Loake, G.J.; Mehrotra, R. Role of Chromatin Architecture in Plant Stress Responses: An Update. Front. Plant Sci. 2021, 11, 603380. [Google Scholar] [CrossRef]
  271. Tricker, P.J.; Gibbings, J.G.; Rodríguez López, C.M.; Hadley, P.; Wilkinson, M.J. Low relative humidity triggers RNA-directed de novo DNA methylation and suppression of genes controlling stomatal development. J. Exp. Bot. 2012, 63, 3799–3813. [Google Scholar] [CrossRef]
  272. Miryeganeh, M. Plants’ Epigenetic Mechanisms and Abiotic Stress. Genes 2021, 12, 1106. [Google Scholar] [CrossRef]
  273. Kong, L.; Liu, Y.; Wang, X.; Chang, C. Insight into the Role of Epigenetic Processes in Abiotic and Biotic Stress Response in Wheat and Barley. Int. J. Mol. Sci. 2020, 21, 1480. [Google Scholar] [CrossRef]
  274. Secco, D.; Wang, C.; Shou, H.; Schultz, M.D.; Chiarenza, S.; Nussaume, L.; Ecker, J.R.; Whelan, J.; Lister, R. Stress induced gene expression drives transient DNA methylation changes at adjacent repetitive elements. eLife 2015, 4, e09343. [Google Scholar] [CrossRef]
  275. Liu, J.; He, Z. Small DNA Methylation, Big Player in Plant Abiotic Stress Responses and Memory. Front. Plant Sci. 2020, 11, 1977. [Google Scholar] [CrossRef]
  276. Iwasaki, M.; Hyvärinen, L.; Piskurewicz, U.; Lopez-Molina, L. Non-canonical RNA-directed DNA methylation participates in maternal and environmental control of seed dormancy. eLife 2019, 8, e37434. [Google Scholar] [CrossRef]
  277. Xie, H.J.; Li, H.; Liu, D.; Dai, W.M.; He, J.Y.; Lin, S.; Duan, H.; Liu, L.L.; Chen, S.G.; Song, X.L.; et al. ICE1 demethylation drives the range expansion of a plant invader through cold tolerance divergence. Mol. Ecol. 2015, 24, 835–850. [Google Scholar] [CrossRef]
  278. Xie, H.; Sun, Y.; Cheng, B.; Xue, S.; Cheng, D.; Liu, L.; Meng, L.; Qiang, S. Variation in ICE1 Methylation Primarily Determines Phenotypic Variation in Freezing Tolerance in Arabidopsis thaliana. Plant Cell Physiol. 2019, 60, 152–165. [Google Scholar] [CrossRef]
  279. Liu, T.; Li, Y.; Duan, W.; Huang, F.; Hou, X. Cold acclimation alters DNA methylation patterns and confers tolerance to heat and increases growth rate in Brassica rapa. J. Exp. Bot. 2017, 68, 1213–1224. [Google Scholar] [CrossRef]
  280. Duan, W.; Zhang, H.; Zhang, B.; Wu, X.; Shao, S.; Li, Y.; Hou, X.; Liu, T. Role of vernalization-mediated demethylation in the floral transition of Brassica rapa. Planta 2017, 245, 227–233. [Google Scholar] [CrossRef]
  281. Lai, Y.S.; Zhang, X.; Zhang, W.; Shen, D.; Wang, H.; Xia, Y.; Qiu, Y.; Song, J.; Wang, C.; Li, X. The association of changes in DNA methylation with temperature-dependent sex determination in cucumber. J. Exp. Bot. 2017, 68, 2899–2912. [Google Scholar] [CrossRef]
  282. Ma, N.; Chen, W.; Fan, T.; Tian, Y.; Zhang, S.; Zeng, D.; Li, Y. Low temperature-induced DNA hypermethylation attenuates expression of RhAG, an AGAMOUS homolog, and increases petal number in rose (Rosa hybrida). BMC Plant Biol. 2015, 15, 237. [Google Scholar] [CrossRef]
  283. Rutowicz, K.; Puzio, M.; Halibart-Puzio, J.; Lirski, M.; Kotliński, M.; Kroteń, M.A.; Knizewski, L.; Lange, B.; Muszewska, A.; Śniegowska-Świerk, K.; et al. A Specialized Histone H1 Variant Is Required for Adaptive Responses to Complex Abiotic Stress and Related DNA Methylation in Arabidopsis. Plant Physiol. 2015, 169, 2080–2101. [Google Scholar] [CrossRef]
  284. Gagné-Bourque, F.; Mayer, B.F.; Charron, J.B.; Vali, H.; Bertrand, A.; Jabaji, S. Accelerated Growth Rate and Increased Drought Stress Resilience of the Model Grass Brachypodium distachyon Colonized by Bacillus subtilis B26. PLoS ONE 2015, 10, e0130456. [Google Scholar] [CrossRef]
  285. Lu, X.; Wang, X.; Chen, X.; Shu, N.; Wang, J.; Wang, D.; Wang, S.; Fan, W.; Guo, L.; Guo, X.; et al. Single-base resolution methylomes of upland cotton (Gossypium hirsutum L.) reveal epigenome modifications in response to drought stress. BMC Genom. 2017, 18, 297. [Google Scholar] [CrossRef]
  286. Wang, W.; Qin, Q.; Sun, F.; Wang, Y.; Xu, D.; Li, Z.; Fu, B. Genome-Wide Differences in DNA Methylation Changes in Two Contrasting Rice Genotypes in Response to Drought Conditions. Front. Plant Sci. 2016, 7, 1675. [Google Scholar] [CrossRef]
  287. Liang, D.; Zhang, Z.; Wu, H.; Huang, C.; Shuai, P.; Ye, C.Y.; Tang, S.; Wang, Y.; Yang, L.; Wang, J.; et al. Single-base-resolution methylomes of Populus trichocarpa reveal the association between DNA methylation and drought stress. BMC Genet. 2014, 15 (Suppl. 1), S9. [Google Scholar] [CrossRef]
  288. Mao, H.; Wang, H.; Liu, S.; Li, Z.; Yang, X.; Yan, J.; Li, J.; Tran, L.S.; Qin, F. A transposable element in a NAC gene is associated with drought tolerance in maize seedlings. Nat. Commun. 2015, 6, 8326. [Google Scholar] [CrossRef]
  289. Shen, X.; De Jonge, J.; Forsberg, S.K.; Pettersson, M.E.; Sheng, Z.; Hennig, L.; Carlborg, Ö. Natural CMT2 variation is associated with genome-wide methylation changes and temperature seasonality. PLoS Genet. 2014, 10, e1004842. [Google Scholar] [CrossRef]
  290. Li, J.; Huang, Q.; Sun, M.; Zhang, T.; Li, H.; Chen, B.; Xu, K.; Gao, G.; Li, F.; Yan, G.; et al. Global DNA methylation variations after short-term heat shock treatment in cultured microspores of Brassica napus cv. Topas. Sci. Rep. 2016, 6, 38401. [Google Scholar] [CrossRef]
  291. 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]
  292. 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.; et al. Divergent cytosine DNA methylation patterns in single-cell, soybean root hairs. New Phytol. 2017, 214, 808–819. [Google Scholar] [CrossRef]
  293. Ma, Y.; Min, L.; Wang, M.; Wang, C.; Zhao, Y.; Li, Y.; Fang, Q.; Wu, Y.; Xie, S.; Ding, Y.; et al. Disrupted Genome Methylation in Response to High Temperature Has Distinct Affects on Microspore Abortion and Anther Indehiscence. Plant Cell 2018, 30, 1387–1403. [Google Scholar] [CrossRef]
  294. Min, L.; Li, Y.; Hu, Q.; Zhu, L.; Gao, W.; Wu, Y.; Ding, Y.; Liu, S.; Yang, X.; Zhang, X. Sugar and auxin signaling pathways respond to high-temperature stress during anther development as revealed by transcript profiling analysis in cotton. Plant Physiol. 2014, 164, 1293–1308. [Google Scholar] [CrossRef] [PubMed]
  295. Folsom, J.J.; Begcy, K.; Hao, X.; Wang, D.; Walia, H. Rice fertilization-Independent Endosperm1 regulates seed size under heat stress by controlling early endosperm development. Plant Physiol. 2014, 165, 238–248. [Google Scholar] [CrossRef] [PubMed]
  296. 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] [PubMed]
  297. Moglia, A.; Gianoglio, S.; Acquadro, A.; Valentino, D.; Milani, A.M.; Lanteri, S.; Comino, C. Identification of DNA methyltransferases and demethylases in Solanum melongena L. and their transcription dynamics during fruit development and after salt and drought stresses. PLoS ONE 2019, 14, e0223581. [Google Scholar] [CrossRef]
  298. 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]
  299. 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]
  300. Ganie, S.A.; Dey, N.; Mondal, T.K. Promoter methylation regulates the abundance of osa-miR393a in contrasting rice genotypes under salinity stress. Funct. Integr. Genom. 2016, 16, 1–11. [Google Scholar] [CrossRef]
  301. Kumar, S.; Beena, A.S.; Awana, M.; Singh, A. Salt-Induced Tissue-Specific Cytosine Methylation Downregulates Expression of HKT Genes in Contrasting Wheat (Triticum aestivum L.) Genotypes. DNA Cell Biol. 2017, 36, 283–294. [Google Scholar] [CrossRef]
  302. Wang, M.; Qin, L.; Xie, C.; Li, W.; Yuan, J.; Kong, L.; Yu, W.; Xia, G.; Liu, S. Induced and constitutive DNA methylation in a salinity-tolerant wheat introgression line. Plant Cell Physiol. 2014, 55, 1354–1365. [Google Scholar] [CrossRef]
  303. Tan, M.P. Analysis of DNA methylation of maize in response to osmotic and salt stress based on methylation-sensitive amplified polymorphism. Plant Physiol. Biochem. 2010, 48, 21–26. [Google Scholar] [CrossRef]
  304. Peng, Y.; Zhang, Y.; Gui, Y.; An, D.; Liu, J.; Xu, X.; Li, Q.; Wang, J.; Wang, W.; Shi, C.; et al. Elimination of a Retrotransposon for Quenching Genome Instability in Modern Rice. Mol. Plant 2019, 12, 1395–1407. [Google Scholar] [CrossRef]
  305. Wang, X.; Jiang, N.; Liu, J.; Liu, W.; Wang, G.-L. The role of effectors and host immunity in plant-necrotrophic fungal interactions. Virulence 2014, 5, 722–732. [Google Scholar] [CrossRef]
  306. Zou, B.; Sun, Q.; Zhang, W.; Ding, Y.; Yang, D.L.; Shi, Z.; Hua, J. The Arabidopsis Chromatin-Remodeling Factor CHR5 Regulates Plant Immune Responses and Nucleosome Occupancy. Plant Cell Physiol. 2017, 58, 2202–2216. [Google Scholar] [CrossRef]
  307. Choi, S.-M.; Song, H.-R.; Han, S.-K.; Han, M.; Kim, C.-Y.; Park, J.; Lee, Y.-H.; Jeon, J.-S.; Noh, Y.-S.; Noh, B. HDA19 is required for the repression of salicylic acid biosynthesis and salicylic acid-mediated defense responses in Arabidopsis. Plant J. 2012, 71, 135–146. [Google Scholar] [CrossRef]
  308. Li, T.; Chen, X.; Zhong, X.; Zhao, Y.; Liu, X.; Zhou, S.; Cheng, S.; Zhou, D.-X. Jumonji C domain protein JMJ705-mediated removal of histone H3 lysine 27 trimethylation is involved in defense-related gene activation in rice. Plant Cell 2013, 25, 4725–4736. [Google Scholar] [CrossRef]
  309. Cui, X.; Jin, P.; Cui, X.; Gu, L.; Lu, Z.; Xue, Y.; Wei, L.; Qi, J.; Song, X.; Luo, M.; et al. Control of transposon activity by a histone H3K4 demethylase in rice. Proc. Natl. Acad. Sci. USA 2013, 110, 1953–1958. [Google Scholar] [CrossRef]
  310. Soyer, J.L.; El Ghalid, M.; Glaser, N.; Ollivier, B.; Linglin, J.; Grandaubert, J.; Balesdent, M.-H.; Connolly, L.R.; Freitag, M.; Rouxel, T.; et al. Epigenetic control of effector gene expression in the plant pathogenic fungus Leptosphaeria maculans. PLoS Genet. 2014, 10, e1004227. [Google Scholar] [CrossRef]
  311. Zhang, X. Dynamic differential methylation facilitates pathogen stress response in Arabidopsis. Proc. Natl. Acad. Sci. USA 2012, 109, 12842–12843. [Google Scholar] [CrossRef]
  312. Schumann, U.; Lee, J.; Kazan, K.; Ayliffe, M.; Wang, M.-B. DNA-Demethylase Regulated Genes Show Methylation-Independent Spatiotemporal Expression Patterns. Front. Plant Sci. 2017, 8, 1449. [Google Scholar] [CrossRef]
  313. Annacondia, M.L.; Magerøy, M.H.; Martinez, G. Stress response regulation by epigenetic mechanisms: Changing of the guards. Physiol. Plant. 2018, 162, 239–250. [Google Scholar] [CrossRef]
  314. Liu, S.; de Jonge, J.; Trejo-Arellano, M.S.; Santos-González, J.; Köhler, C.; Hennig, L. Role of H1 and DNA methylation in selective regulation of transposable elements during heat stress. New Phytol. 2021, 229, 2238–2250. [Google Scholar] [CrossRef] [PubMed]
  315. Le, T.-N.; Schumann, U.; Smith, N.A.; Tiwari, S.; Au, P.C.K.; Zhu, Q.-H.; Taylor, J.M.; Kazan, K.; Llewellyn, D.J.; Zhang, R.; et al. DNA demethylases target promoter transposable elements to positively regulate stress responsive genes in Arabidopsis. Genome Biol. 2014, 15, 458. [Google Scholar] [CrossRef] [PubMed]
  316. Iwasaki, M.; Paszkowski, J. Identification of genes preventing transgenerational transmission of stress-induced epigenetic states. Proc. Natl. Acad. Sci. USA 2014, 111, 8547–8552. [Google Scholar] [CrossRef] [PubMed]
  317. Lanciano, S.; Cristofari, G. Measuring and interpreting transposable element expression. Nat. Rev. Genet. 2020, 21, 721–736. [Google Scholar] [CrossRef]
  318. Kulski, J.K. Next-generation sequencing—an overview of the history, tools, and “Omic” applications. In Next Generation Sequencing—Advances, Applications and Challenges; Kulski, J.K., Ed.; IntechOpen: London, UK, 2016; Volume 3, p. 60. [Google Scholar]
  319. Boo, S.H.; Kim, Y.K. The emerging role of RNA modifications in the regulation of mRNA stability. Exp. Mol. Med. 2020, 52, 400–408. [Google Scholar] [CrossRef]
  320. Satheesh, V.; Fan, W.; Chu, J.; Cho, J. Recent advancement of NGS technologies to detect active transposable elements in plants. Genes Genom. 2021, 43, 289–294. [Google Scholar] [CrossRef]
  321. Cho, J.; Benoit, M.; Catoni, M.; Drost, H.G.; Brestovitsky, A.; Oosterbeek, M.; Paszkowski, J. Sensitive detection of pre-integration intermediates of long terminal repeat retrotransposons in crop plants. Nat. Plants 2019, 5, 26–33. [Google Scholar] [CrossRef]
  322. Wang, B.; Tseng, E.; Regulski, M.; Clark, T.A.; Hon, T.; Jiao, Y.; Lu, Z.; Olson, A.; Stein, J.C.; Ware, D. Unveiling the complexity of the maize transcriptome by single-molecule long-read sequencing. Nat. Commun. 2016, 7, 11708. [Google Scholar] [CrossRef]
  323. Panda, K.; Slotkin, R.K. Long-Read cDNA Sequencing Enables a “Gene-Like” Transcript Annotation of Transposable Elements. Plant Cell 2020, 32, 2687–2698. [Google Scholar] [CrossRef]
  324. Deininger, P.; Morales, M.E.; White, T.B.; Baddoo, M.; Hedges, D.J.; Servant, G.; Srivastav, S.; Smither, M.E.; Concha, M.; DeHaro, D.L.; et al. A comprehensive approach to expression of L1 loci. Nucleic Acids Res. 2017, 45, e31. [Google Scholar] [CrossRef]
  325. Belancio, V.P.; Roy-Engel, A.M.; Pochampally, R.R.; Deininger, P. Somatic expression of LINE-1 elements in human tissues. Nucleic Acids Res. 2010, 38, 3909–3922. [Google Scholar] [CrossRef]
  326. Morillon, A.; Bénard, L.; Springer, M.; Lesage, P. Differential effects of chromatin and Gcn4 on the 50-fold range of expression among individual yeast Ty1 retrotransposons. Mol. Cell Biol. 2002, 22, 2078–2088. [Google Scholar] [CrossRef]
  327. Orozco-Arias, S.; Isaza, G.; Guyot, R.; Tabares-Soto, R. A systematic review of the application of machine learning in the detection and classification of transposable elements. PeerJ 2019, 7, e8311. [Google Scholar] [CrossRef]
  328. Tokuyama, M.; Kong, Y.; Song, E.; Jayewickreme, T.; Kang, I.; Iwasaki, A. ERVmap analysis reveals genome-wide transcription of human endogenous retroviruses. Proc. Natl. Acad. Sci. USA 2018, 115, 12565–12572. [Google Scholar] [CrossRef]
  329. McKerrow, W.; Fenyö, D. L1EM: A tool for accurate locus specific LINE-1 RNA quantification. Bioinformatics 2020, 36, 1167–1173. [Google Scholar] [CrossRef]
  330. Philippe, C.; Vargas-Landin, D.B.; Doucet, A.J.; van Essen, D.; Vera-Otarola, J.; Kuciak, M.; Corbin, A.; Nigumann, P.; Cristofari, G. Activation of individual L1 retrotransposon instances is restricted to cell-type dependent permissive loci. eLife 2016, 5, e13926. [Google Scholar] [CrossRef]
  331. Scott, E.C.; Gardner, E.J.; Masood, A.; Chuang, N.T.; Vertino, P.M.; Devine, S.E. A hot L1 retrotransposon evades somatic repression and initiates human colorectal cancer. Genome Res. 2016, 26, 745–755. [Google Scholar] [CrossRef]
  332. Teissandier, A.; Servant, N.; Barillot, E.; Bourc’his, D. Tools and best practices for retrotransposon analysis using high-throughput sequencing data. Mob. DNA 2019, 10, 52. [Google Scholar] [CrossRef]
  333. Kong, Y.; Rose, C.M.; Cass, A.A.; Williams, A.G.; Darwish, M.; Lianoglou, S.; Haverty, P.M.; Tong, A.J.; Blanchette, C.; Albert, M.L.; et al. Transposable element expression in tumors is associated with immune infiltration and increased antigenicity. Nat. Commun. 2019, 10, 5228. [Google Scholar] [CrossRef]
  334. Criscione, S.W.; Zhang, Y.; Thompson, W.; Sedivy, J.M.; Neretti, N. Transcriptional landscape of repetitive elements in normal and cancer human cells. BMC Genom. 2014, 15, 583. [Google Scholar] [CrossRef]
  335. Jeong, H.H.; Yalamanchili, H.K.; Guo, C.; Shulman, J.M.; Liu, Z. An ultra-fast and scalable quantification pipeline for transposable elements from next generation sequencing data. Pac. Symp. Biocomput. 2018, 23, 168–179. [Google Scholar] [PubMed]
  336. Yang, W.R.; Ardeljan, D.; Pacyna, C.N.; Payer, L.M.; Burns, K.H. SQuIRE reveals locus-specific regulation of interspersed repeat expression. Nucleic Acids Res. 2019, 47, e27. [Google Scholar] [CrossRef] [PubMed]
  337. Lerat, E.; Fablet, M.; Modolo, L.; Lopez-Maestre, H.; Vieira, C. TEtools facilitates big data expression analysis of transposable elements and reveals an antagonism between their activity and that of piRNA genes. Nucleic Acids Res. 2017, 45, e17. [Google Scholar] [CrossRef] [PubMed]
  338. Valdebenito-Maturana, B.; Riadi, G. TEcandidates: Prediction of genomic origin of expressed transposable elements using RNA-seq data. Bioinformatics 2018, 34, 3915–3916. [Google Scholar] [CrossRef]
  339. Bendall, M.L.; de Mulder, M.; Iñiguez, L.P.; Lecanda-Sánchez, A.; Pérez-Losada, M.; Ostrowski, M.A.; Jones, R.B.; Mulder, L.C.F.; Reyes-Terán, G.; Crandall, K.A.; et al. Telescope: Characterization of the retrotranscriptome by accurate estimation of transposable element expression. PLoS Comput. Biol. 2019, 15, e1006453. [Google Scholar] [CrossRef]
  340. Jin, Y.; Tam, O.H.; Paniagua, E.; Hammell, M. TEtranscripts: A package for including transposable elements in differential expression analysis of RNA-seq datasets. Bioinformatics 2015, 31, 3593–3599. [Google Scholar] [CrossRef]
  341. Navarro, F.C.; Hoops, J.; Bellfy, L.; Cerveira, E.; Zhu, Q.; Zhang, C.; Lee, C.; Gerstein, M.B. TeXP: Deconvolving the effects of pervasive and autonomous transcription of transposable elements. PLoS Comput. Biol. 2019, 15, e1007293. [Google Scholar] [CrossRef]
  342. Lacal, I.; Ventura, R. Epigenetic Inheritance: Concepts, Mechanisms and Perspectives. Front. Mol. Neurosci. 2018, 11, 292. [Google Scholar] [CrossRef]
  343. Shahid, S.; Slotkin, R.K. The current revolution in transposable element biology enabled by long reads. Curr. Opin. Plant Biol. 2020, 54, 49–56. [Google Scholar] [CrossRef]
  344. Hu, J.; Manduzio, S.; Kang, H. Epitranscriptomic RNA Methylation in Plant Development and Abiotic Stress Responses. Front. Plant Sci. 2019, 10, 500. [Google Scholar] [CrossRef]
  345. Valihrach, L.; Androvic, P.; Kubista, M. Platforms for Single-Cell Collection and Analysis. Int. J. Mol. Sci. 2018, 19, 807. [Google Scholar] [CrossRef]
Figure 2. Cellular functions of DNA methylation (m) in the plant genome. DNA methylation regulates transposon activation, gene regulation, and chromosome interactions. (A) Methylation in the gene promoter either represses or activates transcription [229,230,231,232,233]. (B) Gene body methylations mainly occur in the CG context, although its function remains unknown [42,231,234,235,236]. (C) DNA methylation in heterochromatin regions causes the ASI1-AIPP1-EDM2 complex to enhance polyadenylation sites (red stars). ASI1 binds RNA and associates with chromatin, and EDM2 catches demethylated histone H3 lysine in the heterochromatin region [159,237,238,239]. (D) The methylation of transposons and other DNA repeats mainly occurs in pericentromeric heterochromatin regions [231,235]. (E) Chromosome interactions among pericentromeric and heterochromatin islands are regulated by DNA methylation, and repressive chromatin regions are characterized by abundant transposons and small RNAs [240,241]. ASI1, anti-silencing 1; AIPP1, immunoprecipitated protein 1; EDM2, enhanced downy mildew 2; POL II, RNA polymerase II. The illustration was adapted and redrawn from Zhang et al. [42], with copyright permission from the Licensor Springer Nature (Nature Reviews Molecular Cell Biology: Nature publisher) and Copyright Clearance Center (https://www.copyright.com) (Supplementary File S4).
Figure 2. Cellular functions of DNA methylation (m) in the plant genome. DNA methylation regulates transposon activation, gene regulation, and chromosome interactions. (A) Methylation in the gene promoter either represses or activates transcription [229,230,231,232,233]. (B) Gene body methylations mainly occur in the CG context, although its function remains unknown [42,231,234,235,236]. (C) DNA methylation in heterochromatin regions causes the ASI1-AIPP1-EDM2 complex to enhance polyadenylation sites (red stars). ASI1 binds RNA and associates with chromatin, and EDM2 catches demethylated histone H3 lysine in the heterochromatin region [159,237,238,239]. (D) The methylation of transposons and other DNA repeats mainly occurs in pericentromeric heterochromatin regions [231,235]. (E) Chromosome interactions among pericentromeric and heterochromatin islands are regulated by DNA methylation, and repressive chromatin regions are characterized by abundant transposons and small RNAs [240,241]. ASI1, anti-silencing 1; AIPP1, immunoprecipitated protein 1; EDM2, enhanced downy mildew 2; POL II, RNA polymerase II. The illustration was adapted and redrawn from Zhang et al. [42], with copyright permission from the Licensor Springer Nature (Nature Reviews Molecular Cell Biology: Nature publisher) and Copyright Clearance Center (https://www.copyright.com) (Supplementary File S4).
Ijms 22 11387 g002
Figure 3. Functions of transposable element (TE) methylation in plant growth, stomata formation, and fruit ripening. (A) In the vegetative cell (male gamete) of Arabidopsis, the TE is silenced by DME-mediated DNA methylation by downregulating the chromatin remodeller DDM1. Small interfering RNAs (siRNAs) derived from TE transcripts travel from the vegetative cell to the sperm cells to reinforce global demethylation (m) in the endosperm with reinforced CHH methylation (H represents A, T, or C) [160,242,243,244,245]. (B) Gene imprinting in the endosperm occurs either at MEGs or PEGs through DNA and histone H3 lysine methylations [246,247,248]. (C) Methylation at the promoter of the gene encoding epidermal patterning factor 2 (EPF2) that suppresses stomata formation is pruned by ROS1, whose mutation silences the EPF2 or the ERECTA genes, thus resulting in stomata formation in Arabidopsis [249,250]. (D) Gradual expression of DML2 during tomato fruit ripening reduces 5-methylcytosine (mC) DNA methylation at several genes (such as CNR, involved in fruit ripening) and epimutation of those genes inhibits fruit ripening [42,229,251]. DME, transcriptional activator demeter; DDM1, decreased DNA methylation 1; MEGs, maternally expressed genes; PEGs, paternally expressed genes; ROS1, repressor of silencing 1; DML2, DNA demethylase DME-LIKE 2; MET1, methyltransferase 1. The illustration was adapted and redrawn from Zhang et al. [42], with copyright permission from the Licensor Springer Nature (Nature Reviews Molecular Cell Biology: Nature publisher) and Copyright Clearance Center (https://www.copyright.com) (Supplementary File S4).
Figure 3. Functions of transposable element (TE) methylation in plant growth, stomata formation, and fruit ripening. (A) In the vegetative cell (male gamete) of Arabidopsis, the TE is silenced by DME-mediated DNA methylation by downregulating the chromatin remodeller DDM1. Small interfering RNAs (siRNAs) derived from TE transcripts travel from the vegetative cell to the sperm cells to reinforce global demethylation (m) in the endosperm with reinforced CHH methylation (H represents A, T, or C) [160,242,243,244,245]. (B) Gene imprinting in the endosperm occurs either at MEGs or PEGs through DNA and histone H3 lysine methylations [246,247,248]. (C) Methylation at the promoter of the gene encoding epidermal patterning factor 2 (EPF2) that suppresses stomata formation is pruned by ROS1, whose mutation silences the EPF2 or the ERECTA genes, thus resulting in stomata formation in Arabidopsis [249,250]. (D) Gradual expression of DML2 during tomato fruit ripening reduces 5-methylcytosine (mC) DNA methylation at several genes (such as CNR, involved in fruit ripening) and epimutation of those genes inhibits fruit ripening [42,229,251]. DME, transcriptional activator demeter; DDM1, decreased DNA methylation 1; MEGs, maternally expressed genes; PEGs, paternally expressed genes; ROS1, repressor of silencing 1; DML2, DNA demethylase DME-LIKE 2; MET1, methyltransferase 1. The illustration was adapted and redrawn from Zhang et al. [42], with copyright permission from the Licensor Springer Nature (Nature Reviews Molecular Cell Biology: Nature publisher) and Copyright Clearance Center (https://www.copyright.com) (Supplementary File S4).
Ijms 22 11387 g003
Figure 4. Transposable elements (TEs) are suppressed by DNA and histone methylations. (A) TE methylation is most commonly found in the CG context. The de novo DNA methylation is performed by DNA methyltransferases DNMT3A and 3B; the pattern of DNA methylation is maintained by DNMT1 by adding a methyl group to the newly synthesized DNA strand (a complementary strand of the hemi-methylated DNA strand), thus ensuring that the epigenetic modifications are inherited by the daughter cell. (B) Nucleosomes are made up of DNA and eight histone proteins. These proteins can be modified in several ways for chromatin accessibility, thereby either activating or inactivating gene expression (gene imprinting). TRIM28, a silencing complex, recognizes KRAB-ZNFs (Kruppel-associated box zinc-finger proteins), which contain a TE-binding domain and deposits H3K9me3 on TE (euchromatin region), thus causing TE repression and heterochromatin formation. The illustration was adapted and redrawn from Jönsson et al. [43], with copyright permission from the Licensor Elsevier (Trends in Genetics: Cell Press publisher) and Copyright Clearance Center (https://www.copyright.com) (Supplementary File S1).
Figure 4. Transposable elements (TEs) are suppressed by DNA and histone methylations. (A) TE methylation is most commonly found in the CG context. The de novo DNA methylation is performed by DNA methyltransferases DNMT3A and 3B; the pattern of DNA methylation is maintained by DNMT1 by adding a methyl group to the newly synthesized DNA strand (a complementary strand of the hemi-methylated DNA strand), thus ensuring that the epigenetic modifications are inherited by the daughter cell. (B) Nucleosomes are made up of DNA and eight histone proteins. These proteins can be modified in several ways for chromatin accessibility, thereby either activating or inactivating gene expression (gene imprinting). TRIM28, a silencing complex, recognizes KRAB-ZNFs (Kruppel-associated box zinc-finger proteins), which contain a TE-binding domain and deposits H3K9me3 on TE (euchromatin region), thus causing TE repression and heterochromatin formation. The illustration was adapted and redrawn from Jönsson et al. [43], with copyright permission from the Licensor Elsevier (Trends in Genetics: Cell Press publisher) and Copyright Clearance Center (https://www.copyright.com) (Supplementary File S1).
Ijms 22 11387 g004
Figure 5. Epigenetic modifications under stress conditions and possible stress memory. (A) Both biotic and abiotic stresses can induce or change DNA methylation (5-methylcytosine, mC) and induce other epigenetic changes in the genome; such modifications are associated with the expression of stress-response genes, which conversely may lead to epigenetic processes. Reprogrammed epigenetic modifications (stress memory) are inherited by the offspring. (B) In Arabidopsis, ROS1, DML2, and DML3 remove DNA methylation, thus collectively regulating stress responsive genes in their vicinity. Defects in demethylases, such as ROS1, DML2 and DML3, exhibit increased susceptibility to the fungal pathogen Fusarium oxysporum [315]. (C) During Arabidopsis recovery from heat stress, DDM1 and MOM1 regulate the deletion of stress-induced epigenetic memory. Mutations in DDM1, a chromatin remodeller, assuages transcriptional silence with a significant loss of DNA methylation. MOM1 intermediates facilitate transcriptional silence via an unknown mechanism without loss of DNA methylation. Dysfunction of DDM1 and MOM1 in heat stress-induced gene de-silencing can be inherited in plants exposed to repeated stress [316]. ROS1, repressor of silencing 1; DMEL2 and DML3, transcriptional activator demeter (DME)-Like 2 and 3, respectively; DDM1, decreased DNA methylation 1; MOM1, morpheus molecule 1; H3K9me2, demethylated histone H3 lysine 9. The illustration was adapted and redrawn from Zhang et al. [42], with copyright permission from the Licensor Springer Nature (Nature Reviews Molecular Cell Biology: Nature publisher) and Copyright Clearance Center (https://www.copyright.com) (Supplementary File S4).
Figure 5. Epigenetic modifications under stress conditions and possible stress memory. (A) Both biotic and abiotic stresses can induce or change DNA methylation (5-methylcytosine, mC) and induce other epigenetic changes in the genome; such modifications are associated with the expression of stress-response genes, which conversely may lead to epigenetic processes. Reprogrammed epigenetic modifications (stress memory) are inherited by the offspring. (B) In Arabidopsis, ROS1, DML2, and DML3 remove DNA methylation, thus collectively regulating stress responsive genes in their vicinity. Defects in demethylases, such as ROS1, DML2 and DML3, exhibit increased susceptibility to the fungal pathogen Fusarium oxysporum [315]. (C) During Arabidopsis recovery from heat stress, DDM1 and MOM1 regulate the deletion of stress-induced epigenetic memory. Mutations in DDM1, a chromatin remodeller, assuages transcriptional silence with a significant loss of DNA methylation. MOM1 intermediates facilitate transcriptional silence via an unknown mechanism without loss of DNA methylation. Dysfunction of DDM1 and MOM1 in heat stress-induced gene de-silencing can be inherited in plants exposed to repeated stress [316]. ROS1, repressor of silencing 1; DMEL2 and DML3, transcriptional activator demeter (DME)-Like 2 and 3, respectively; DDM1, decreased DNA methylation 1; MOM1, morpheus molecule 1; H3K9me2, demethylated histone H3 lysine 9. The illustration was adapted and redrawn from Zhang et al. [42], with copyright permission from the Licensor Springer Nature (Nature Reviews Molecular Cell Biology: Nature publisher) and Copyright Clearance Center (https://www.copyright.com) (Supplementary File S4).
Ijms 22 11387 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ramakrishnan, M.; Satish, L.; Kalendar, R.; Narayanan, M.; Kandasamy, S.; Sharma, A.; Emamverdian, A.; Wei, Q.; Zhou, M. The Dynamism of Transposon Methylation for Plant Development and Stress Adaptation. Int. J. Mol. Sci. 2021, 22, 11387. https://doi.org/10.3390/ijms222111387

AMA Style

Ramakrishnan M, Satish L, Kalendar R, Narayanan M, Kandasamy S, Sharma A, Emamverdian A, Wei Q, Zhou M. The Dynamism of Transposon Methylation for Plant Development and Stress Adaptation. International Journal of Molecular Sciences. 2021; 22(21):11387. https://doi.org/10.3390/ijms222111387

Chicago/Turabian Style

Ramakrishnan, Muthusamy, Lakkakula Satish, Ruslan Kalendar, Mathiyazhagan Narayanan, Sabariswaran Kandasamy, Anket Sharma, Abolghassem Emamverdian, Qiang Wei, and Mingbing Zhou. 2021. "The Dynamism of Transposon Methylation for Plant Development and Stress Adaptation" International Journal of Molecular Sciences 22, no. 21: 11387. https://doi.org/10.3390/ijms222111387

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