RNA-directed DNA methylation

Last updated
Overview of several of the biological functions of RdDM.Top left: TE silencing by RdDM prevents TE activation and transposition. Without RdDM, active TEs are free to transpose into genes or promoters, which can disrupt gene expression or result in a mutant protein. Top right: RdDM is involved in several aspects of development; for example, RdDM affects flowering time by repressing FWA. In pollen, TEs become activated in a support cell, leading to the production of sRNAs for RdDM that move to the germ cell in order to reinforce TE silencing. Bottom left: sRNAs involved in RdDM are mobile, and can move between cells through plasmodesmata, or systemically via the vasculature, so RdDM-mediated silencing can spread from its point of origin to distal tissues. Bottom right: RdDM is involved in several abiotic stress responses including the heat shock response, and can silence TEs that would otherwise become active and transpose under heat stress. RdDM is also involved in pathogen defense, and can silence viral DNA (either as a viral minichromosome, shown, or as an integrated provirus) using sRNAs derived from viral mRNAs. RdDM biological functions.png
Overview of several of the biological functions of RdDM.Top left: TE silencing by RdDM prevents TE activation and transposition. Without RdDM, active TEs are free to transpose into genes or promoters, which can disrupt gene expression or result in a mutant protein. Top right: RdDM is involved in several aspects of development; for example, RdDM affects flowering time by repressing FWA. In pollen, TEs become activated in a support cell, leading to the production of sRNAs for RdDM that move to the germ cell in order to reinforce TE silencing. Bottom left: sRNAs involved in RdDM are mobile, and can move between cells through plasmodesmata, or systemically via the vasculature, so RdDM-mediated silencing can spread from its point of origin to distal tissues. Bottom right: RdDM is involved in several abiotic stress responses including the heat shock response, and can silence TEs that would otherwise become active and transpose under heat stress. RdDM is also involved in pathogen defense, and can silence viral DNA (either as a viral minichromosome, shown, or as an integrated provirus) using sRNAs derived from viral mRNAs.

RNA-directed DNA methylation (RdDM) is a biological process in which non-coding RNA molecules direct the addition of DNA methylation to specific DNA sequences. The RdDM pathway is unique to plants, although other mechanisms of RNA-directed chromatin modification have also been described in fungi and animals. To date, the RdDM pathway is best characterized within angiosperms (flowering plants), and particularly within the model plant Arabidopsis thaliana. However, conserved RdDM pathway components and associated small RNAs (sRNAs) have also been found in other groups of plants, such as gymnosperms and ferns. The RdDM pathway closely resembles other sRNA pathways, particularly the highly conserved RNAi pathway found in fungi, plants, and animals. Both the RdDM and RNAi pathways produce sRNAs and involve conserved Argonaute, Dicer and RNA-dependent RNA polymerase proteins.

Contents

RdDM has been implicated in a number of regulatory processes in plants. The DNA methylation added by RdDM is generally associated with transcriptional repression of the genetic sequences targeted by the pathway. Since DNA methylation patterns in plants are heritable, these changes can often be stably transmitted to progeny. As a result, one prominent role of RdDM is the stable, transgenerational suppression of transposable element (TE) activity. RdDM has also been linked to pathogen defense, abiotic stress responses, and the regulation of several key developmental transitions. Although the RdDM pathway has a number of important functions, RdDM-defective mutants in Arabidopsis thaliana are viable and can reproduce, which has enabled detailed genetic studies of the pathway. However, RdDM mutants can have a range of defects in different plant species, including lethality, altered reproductive phenotypes, TE upregulation and genome instability, and increased pathogen sensitivity. Overall, RdDM is an important pathway in plants that regulates a number of processes by establishing and reinforcing specific DNA methylation patterns, which can lead to transgenerational epigenetic effects on gene expression and phenotype.

Biological functions

RdDM is involved in a number of biological processes in the plant, including stress responses, cell-to-cell communication, and the maintenance of genome stability through TE silencing.

Transposable element silencing and genome stability

TEs are pieces of DNA that, when expressed, can move around the genome through a copy-and-paste or cut-and-paste mechanism. New TE insertions can disrupt protein coding or gene regulatory sequences, which can harm or kill the host cell or organism. [1] As a result, most organisms have mechanisms for preventing TE expression. This is particularly key in plant genomes, which are often TE-rich. Some plant species, including important crops like maize and wheat, have genomes consisting of upwards of 80% TEs. [1] [2] RdDM plays a key role in silencing these mobile DNA elements in plants by adding DNA methylation over new TE insertions and constantly reinforcing DNA methylation over existing TEs, inhibiting transposition and maintaining long-term genome stability. [3] Although the RdDM mechanism itself is unique to plants, using DNA methylation to silence TEs is a common strategy among eukaryotes. [4]

RdDM primarily targets small TEs and TE fragments near genes, which are usually in open, accessible euchromatic regions of the genome that are permissive for gene expression. [3] [5] In these regions, the ‘active’ chromatin state has a tendency to spread from expressed genes to nearby repressed regions, like TEs, which can cause these TEs to become activated and transpose. [3] Continuous activity by RdDM opposes the spread of active chromatin, maintaining a silent, repressive heterochromatic state over TEs in these otherwise euchromatic regions. In turn, RdDM activity recruits other pathways that help establish and propagate the silent, heterochromatic state (see 'Interactions between RdDM and other chromatin modifying pathways'). Because of the self-reinforcing nature of these silencing pathways, excessive RdDM activity can also cause the silent, heterochromatic chromatin state over TEs to spread to nearby genes and repress them, with potentially harmful consequences for the organism. [3] [5] Therefore, RdDM activity must be finely tuned to maintain a balance between repressing TEs and allowing expression of nearby genes. [3]

In addition to maintaining stable silencing of TEs, RdDM can also initiate transcriptional silencing of foreign DNA, including novel TE insertions, virus-derived sequences, and transgenes (also see 'Biotic stresses' and 'Transgene silencing' below). [6] [7] [8] [9] [10] When TEs integrate near genes, RdDM-mediated silencing of the TEs often affects gene expression. [3] [1] However, this is not always deleterious, and can sometimes be overcome by other processes, [11] or alter gene expression in ways beneficial to the plant. Over evolutionary time, beneficial TEs can become an important part of the mechanism by which a gene is regulated. [3] [1] In one example, the gene ROS1 lies adjacent to a small helitron TE that is normally methylated by RdDM. [12] [13] While DNA methylation is normally associated with transcriptional repression, this is not the case at the ROS1 locus. Instead, methylation of the helitron TE promotes ROS1 expression, so ROS1 expression is lost in mutants of the RdDM pathway that cannot methylate the TE. [12] [13] Interestingly, ROS1 encodes a DNA glycosylase that functions to remove DNA methylation from the genome. [14] The link between ROS1 expression and RdDM activity at this TE ensures that DNA methylation and demethylation activities remain in balance, helping to maintain DNA methylation homeostasis genome-wide. [12] [13] Thus, RdDM-mediated regulation of TEs can lead to beneficial regulatory outcomes.

Some TEs have evolved mechanisms to suppress or escape RdDM-based silencing in order to facilitate their own proliferation, leading to an evolutionary arms race between TEs and their host genomes. In one example, a TE-derived sequence was found to produce sRNAs that trigger post-transcriptional repression of a component of the RdDM pathway, inhibiting RdDM. [15] This sequence may have helped the original TE escape RdDM-based silencing and insert itself into the host genome.

Studying how RdDM targets and represses different types of TEs has led to many major insights into how the RdDM mechanism works. The retrotransposon EVADÉ (EVD) was one of the first TEs specifically shown to be repressed by RdDM-derived sRNAs. [16] Later work used EVD to trace the mechanism by which a novel TE insertion became silenced, revealing an important mechanistic link between post-transcriptional gene silencing and RdDM. [9] Studies of other retrotransposons, including ONSEN, which is regulated by both RdDM and heat stress, [17] [18] and Athila family TEs, [10] among many others, have also provided valuable insights into RdDM-mediated TE silencing.

Development and reproduction

A number of epigenetic changes required for normal development and reproduction in flowering plants involve RdDM. In a well-studied example, RdDM is required for repression of the FWA gene, which allows for proper timing of flowering in Arabidopsis. [19] The FWA promoter contains tandem repeats that are usually methylated by RdDM, leading to transcriptional repression. [20] Loss of this methylation re-activates FWA expression, causing a late-flowering phenotype. [19] [20] The loss of DNA methylation and associated late-flowering phenotype can be stably transmitted to progeny. Since the demethylated fwa allele leads to a stable, heritable change in the expression of FWA without any change to the DNA sequence, it is a classic example of an epiallele.

Mutations in the RdDM pathway can strongly affect gamete formation and seed viability, particularly in plant species with high TE content like maize and Brassica rapa, highlighting the importance of this pathway in plant reproduction. [21] [22] [23] During gamete formation, it has been hypothesized, and in some cases shown, that RdDM helps reinforce TE silencing in the germ cells. [24] [25] In both pollen and ovules, a support cell undergoes epigenetic reprogramming, losing DNA methylation and other epigenetic marks at a number of loci, including TEs. [26] [24] This causes TE re-activation and encourages the production of RdDM-derived sRNAs against these TEs in the support cells. The sRNAs are then thought to move from the support cell to the germ cell in order to reinforce TE silencing in the next generation. This phenomenon has been observed in pollen, but has yet to be shown definitively in the ovule. [27] [28] This role for sRNAs in plants resembles the role of piRNAs in germline development in Drosophila and some other animals. [29] [30] A similar phenomenon may also occur in roots to preserve TE silencing in important stem cell populations. [31]

The RdDM pathway is also involved in regulating imprinted expression at some genes. [32] This unusual parent-of-origin-specific expression pattern occurs at several loci in the endosperm during seed development in flowering plants. A few factors involved in the RdDM pathway are themselves imprinted (favoring expression from the paternal allele) in diverse species, including A. thaliana, A. lyrata, C. rubella, and maize. [33] [34] [35] [36] RdDM also plays a role in mediating the gene dosage effects seen in seeds derived from interploid crosses, [37] [38] though the mechanism for this remains largely unknown.

There is also evidence that RdDM plays a role in several other aspects of plant development, including seed dormancy, [39] fruit ripening, [40] and other pathways involved in flowering. [41] However, most of these data are correlative, and further study is necessary to understand the role of RdDM in these processes.

Stress response

Abiotic stresses

RdDM helps plants respond to a number of abiotic stresses, such as heat stress, drought, phosphate starvation, salt stress, and others. [42] Many TEs become upregulated under abiotic stress conditions, [43] [44] and thus one function of RdDM in stress response is to help counter this activation. In one example, the retrotransposon ONSEN is upregulated by heat stress, but normally remains suppressed by RdDM-associated sRNAs and can only transpose efficiently in heat-stressed plants that are also deficient in RdDM. [17] [18] More generally, in plants exposed to heat stress, several components of the RdDM pathway become upregulated, and mutations in some components of the RdDM machinery reduce heat tolerance, suggesting RdDM plays an important role during heat stress. [45] [46] In addition to regulating TEs under stress conditions, RdDM can also regulate genes in order to trigger appropriate stress responses. Under low humidity, leaves produce fewer stomata due to RdDM-mediated downregulation of two genes involved in stomatal development. [47] Similarly, RdDM becomes downregulated in response to salt stress, and this has been shown to trigger the expression of a transcription factor important in salt stress resistance. [48]

Biotic stresses

RdDM was initially discovered as a response to infection by viroids, [49] and along with RNAi plays an important role in defending the plant against viroids and viruses. The RdDM and RNAi machinery recognize viral RNAs and process them into sRNAs, which can then be used both pathways to degrade viral RNA (RNAi) and silence viral DNA (RdDM). [50] [51] [52] However, little is known about how the RdDM and RNAi machinery distinguish between viral RNAs and RNAs produced by the host plant. Mutants defective in RdDM and other methylation-deficient mutants are often hypersensitive to viral infection. [53] [54] Virus-host interactions are another example of an evolutionary arms race, and many plant viruses encode suppressors of both RdDM and RNAi in an attempt to evade the host plant's defenses. [55] [53] [56] [57]

RdDM is also involved in protecting the plant from other biotic stresses, [50] including bacterial infections, [58] fungal infections, [59] and predation. [60] Loss of RdDM can have opposing effects on resistance for different pathogens. For example, some RdDM mutants have increased susceptibility to the bacterium Agrobacterium tumefaciens, [61] but those same mutants have decreased susceptibility to the bacterium Pseudomonas syringae, [58] highlighting the complexity of the different pathogen defense pathways and their interactions with RdDM. [62]

Transgene silencing

In addition to naturally-occurring foreign nucleic acid stressors like TEs and viruses, artificially introduced DNA sequences, like transgenes, are also targeted for repression by RdDM. [63] [6] Transgenes are widely used in genetics research to study gene function and regulation, and in plant breeding to introduce novel and desirable properties into a plant. Transgene silencing by RdDM and other mechanisms has therefore proved problematic for plant researchers. Efforts to understand how transgenes become silenced have ultimately helped reveal much of what we now know about the RdDM pathway (see 'History and discovery of RdDM'). In one early example, researchers sequentially transformed plants with two different transgenes that shared some of their DNA sequence. [64] They found that transforming the second transgene into the plants led to the first transgene gaining DNA methylation and becoming inactivated. [64] This provided an early clue that there existed a trans-acting, sequence-based mechanism for transcriptional silencing of foreign DNA, later shown to be RdDM.

Stress and RdDM-mediated epigenetic ‘memory’

Due to the heritability of DNA methylation patterns in plants, and the self-reinforcing nature of RdDM and other DNA methylation pathways, any DNA methylation changes caused by environmental stressors have the potential to be maintained and transmitted to future generations. This can allow stress-induced DNA methylation changes to act as a ‘memory’ of the stressor and help prime the plant or its progeny to respond more efficiently to the stress if re-exposed. [50] [65] For example, RdDM-derived sRNAs against TEs or viruses that have already integrated into the genome and been silenced serve as a 'memory' of those prior infections, protecting against future invasions by similar sequences. There is also evidence that DNA methylation changes due to other stressors, such as salt or heat stress, can persist in the progeny of stressed plants even in the absence of the original stressor. [66] In this study, the persistence of the stress-induced DNA methylation changes required several RdDM-related proteins, suggesting that RdDM was involved in maintaining the stress-altered DNA methylation patterns. In another example, resistance to insect attack was transmitted to progeny via DNA methylation changes, and this inheritance was also dependent on functional sRNA biogenesis pathways. [60] [50] Thus, RdDM can potentially alter the plant epigenome in response to stress, and helps maintain these changes to modulate future stress responses in the affected plant and its descendants.

Short and long-range signaling

The sRNA molecules produced by RdDM and other pathways are able to move between cells via plasmodesmata, and can also move systemically through the plant via the vasculature. [67] [68] [69] They therefore have the potential to act as signaling molecules. This has been demonstrated in plants engineered to express green fluorescent protein (GFP). [70] The GFP protein produced by these plants caused them to glow green under certain light conditions. When tissue from a second plant expressing a sRNA construct complementary to GFP was grafted onto the GFP-expressing plant, the GFP fluorescence was lost: after grafting, the sRNAs being produced in the second plant's tissues were moving into the tissues of the first, GFP-expressing plant, and triggering silencing of GFP. [70] The same study showed that a subset of these mobile sRNAs were triggering the addition of DNA methylation to the GFP locus via RdDM. Therefore, sRNAs involved in RdDM can act as signaling molecules and trigger the addition of DNA methylation at complementary loci in cells far away from where the sRNAs were originally generated. Since then, studies have shown that sRNAs can move and direct RdDM both from shoot to root and root to shoot, though the silencing effect is more robust when sRNAs move from shoot to root. [69] [70] [71] [72]

Movement of sRNAs that drive RdDM activity plays an important role in plant development, including during reproduction [23] [24] [27] and root development. [31] In both cases, sRNA movement seems to function primarily as a way to reinforce DNA methylation and silencing of TEs in developmentally important cell types, like germ cells and stem cells. Silencing TEs and maintaining genome integrity in these cells is particularly important because they give rise to many other cells, all of which will inherit any defects or mutations in the original stem cell or germ cell. sRNA movement is also involved in plant-pathogen interactions: sRNAs can move from infected cells to distal uninfected tissues in order to prime a defense response, though to date this has only been shown for RNAi, not RdDM. [73]

Pathways and mechanisms

This section focuses on the pathways and mechanisms by which RdDM leads to sequence-specific DNA methylation. The pathways presented here were characterized primarily in the model plant Arabidopsis thaliana, but are likely similar in other angiosperms. Conservation of RdDM in other plant species is discussed in more detail in 'Evolutionary conservation' below.

DNA methylation context

DNA methylation sequence contexts and related DNA methyltransferases. DNA methylation at cytosines followed by guanines (CG methylation) is maintained by MET1, while CHG and CHH methylation are maintained by CMT3 and CMT2, respectively. The methyltransferase involved in RdDM, DRM2, can add DNA methylation regardless of sequence context. Methylation context.png
DNA methylation sequence contexts and related DNA methyltransferases. DNA methylation at cytosines followed by guanines (CG methylation) is maintained by MET1, while CHG and CHH methylation are maintained by CMT3 and CMT2, respectively. The methyltransferase involved in RdDM, DRM2, can add DNA methylation regardless of sequence context.

RdDM is the only mechanism in plants that can add DNA methylation to cytosines regardless of sequence context. [55] DNA methylation in plants is typically divided into three categories based on the sequence context of the methylated cytosine: CG, CHG, and CHH, where H is any nucleotide except G. These reflect the different sequence contexts targeted by several DNA methylation pathways in plants. These context-specific pathways are primarily involved in maintaining existing DNA methylation patterns. The highly conserved methyltransferase MET1 (homolog of mammalian DNMT1) maintains DNA methylation in the CG context, while two conserved plant-specific methyltransferases, Chromomethylase 3 (CMT3) and CMT2, help maintain CHG and CHH methylation, respectively. [74] [75] [76] [77] Unlike these pathways, RdDM leads to the addition of DNA methylation at all cytosines regardless of their sequence context. Like MET1, CMT2 and CMT3, RdDM is primarily involved in maintaining existing DNA methylation patterns. [55] However, RdDM is also the only pathway capable of adding DNA methylation de novo to previously unmethylated regions in plants.

Mechanism

The RdDM pathway can be split up into two main processes: the production of sRNAs, and the recruitment of DNA methylation machinery by those sRNAs to specific target loci in the DNA. [78] [55] [79] These two activities together constitute RdDM, and ultimately lead to DNA methylation being added to cytosines at specific target loci.

Schematic of the canonical RdDM pathway (top), and non-canonical RdDM and RNAi/PTGS (bottom). The canonical RdDM pathway can be broken into (1) sRNA production and (2) targeting DNA methylation to sites of sRNA production. The non-canonical RdDM pathway is closely related to RNAi and other PTGS pathways, and differs from canonical RdDM primarily in the source of sRNAs and sRNA processing. H3K9 = lysine 9 on histone H3; H3K4 = lysine 4 on histone H3; ssRNA = single-stranded RNA; dsRNA = double-stranded RNA, miRNA = microRNA RdDM pathway summary.png
Schematic of the canonical RdDM pathway (top), and non-canonical RdDM and RNAi/PTGS (bottom). The canonical RdDM pathway can be broken into (1) sRNA production and (2) targeting DNA methylation to sites of sRNA production. The non-canonical RdDM pathway is closely related to RNAi and other PTGS pathways, and differs from canonical RdDM primarily in the source of sRNAs and sRNA processing. H3K9 = lysine 9 on histone H3; H3K4 = lysine 4 on histone H3; ssRNA = single-stranded RNA; dsRNA = double-stranded RNA, miRNA = microRNA

Canonical RdDM

The canonical RdDM pathway is, as its name suggests, the most well-characterized RdDM pathway to date. Canonical RdDM is preferentially recruited to regions that are already DNA-methylated and heterochromatic, and acts to reinforce existing DNA methylation patterns at these loci, forming a positive feedback loop. [55] [79] Canonical RdDM makes up the majority of RdDM activity in a cell. [79]

sRNA production

The first part of the RdDM pathway revolves around the biogenesis of sRNAs. A plant-specific RNA polymerase complex, RNA Polymerase IV (Pol IV), is first recruited to silent heterochromatin via its interaction with CLASSY (CLSY) proteins and SAWADEE homeodomain homolog 1 (SHH1) (also see 'Interactions between RdDM and other chromatin modifying pathways' below). [80] [79] [81] Pol IV transcribes these regions to produce short single-stranded RNAs (ssRNAs) roughly 30 to 45 nucleotides in length, each of which is the precursor for a single sRNA. [82] [83] [84] These ssRNAs are converted into double-stranded RNAs (dsRNAs) co-transcriptionally by RNA-directed RNA polymerase 2 (RDR2), which physically associates with Pol IV. [83] The dsRNAs are then cleaved by the endoribonuclease Dicer-like 3 (DCL3) into 24 nucleotide (nt) sRNAs. Pol IV, RDR2, and DCL3 alone are sufficient for the production of 24 nt sRNAs in vitro, [84] suggesting that while other factors involved in this part of the pathway may help increase efficiency or specificity, they are not required for Pol IV-mediated sRNA production.

While nearly all 24 nt sRNAs involved in RdDM are produced through the Pol IV-RDR2-DCL3 pathway, a small proportion are produced through other pathways. For example, some RNA Polymerase II (Pol II) transcripts that contain an inverted repeat sequence form double-stranded hairpin structures that can be directly cleaved by DCL3 to form 24 nt sRNAs. [85] [79]

DNA methylation of target loci

In the second part of the pathway, the RdDM DNA methylation machinery is guided to DNA sequences complementary to the sRNAs generated in the first part of the pathway. One strand from each 24 nt double-stranded sRNA is loaded into Argonaute (AGO) proteins AGO4, AGO6, or AGO9. [55] AGO3 may also be able to function in this pathway. [86] Argonautes are a large, highly conserved family of proteins that can bind sRNAs, forming a protein-sRNA duplex that enables them to recognize and bind other RNA sequences complementary to their sRNA partner. [87] Once formed, the AGO-sRNA duplex finds and binds complementary sequences along an RNA ‘scaffold’ produced by the plant-specific RNA polymerase V (Pol V), with the help of interactions with Suppressor of Ty insertion 5-like (SPT5L), the Involved in de novo 2 - IDN2 Paralog (IDN2-IDP) complex, and the Pol V subunit NRPE1. [88] This leads to the recruitment of the DNA methyltransferase enzyme Domains Rearranged Methyltransferase 2 (DRM2), which methylates nearby DNA. [89] [55] [79] The mechanism by which the AGO-sRNA duplex recruits DRM2 is not yet well understood. [90]

Non-canonical RdDM

Recent work has revealed a number of variations of the RdDM pathway, collectively referred to as non-canonical RdDM. [79] Unlike canonical RdDM, the non-canonical pathways are generally involved in establishing initial DNA methylation at new target loci, like novel TE insertions, rather than maintaining existing heterochromatin. Actively expressing elements like new TE insertions are normally strongly targeted by post-transcriptional gene silencing (PTGS/RNAi) pathways. Non-canonical RdDM occurs primarily as a byproduct of these PTGS pathways, leading to the initial establishment of a silent, heterochromatic state over the new TE or other target locus. Once that initial silent state is established, Pol IV can be recruited to the locus by CLSY and SHH1, and the canonical RdDM pathway takes over the long-term maintenance of silencing. [79] Therefore, the non-canonical RdDM pathways often act as a temporary bridge between initial post-transcriptional silencing of novel elements by RNAi, and long-term transgenerational transcriptional silencing via canonical RdDM. [10] [9] [79] Consistent with this role in initiation of novel silencing, non-canonical RdDM targets relatively few loci in comparison to canonical RdDM. [79]

The primary difference between the canonical and non-canonical RdDM pathways lies in the origin and biogenesis of the sRNAs involved. The canonical RdDM pathway involves 24 nt sRNAs, which are specific to that pathway and come predominantly from a single source (the Pol IV-RDR2 complex). In contrast, the non-canonical RdDM pathways involve 21-22 nt sRNAs from a variety of sources, allowing de novo DNA methylation to be initiated at many different types of loci. These 21-22 nt sRNAs are not specific to non-canonical RdDM, and also function in other PTGS pathways. In fact, only a small fraction of 21-22nt sRNAs are involved in RdDM, with the majority instead driving a positive feedback loop amplifying the PTGS response. [91] The functional outcome of a specific 21-22 nt sRNA depends on the AGO protein it ultimately associates with: sRNAs that associate with AGO4, AGO6 or AGO9 result in RdDM and DNA methylation, while sRNAs that associate with other AGOs, like AGO1, primarily result in PTGS. [55] [79]

By using 21-22 nt sRNAs derived from a variety of sources, non-canonical RdDM can flexibly induce de novo DNA methylation and silencing at many different types of loci. One of the primary sources of 21-22 nt sRNAs is Pol II transcripts. Some of these transcripts, particularly those produced from TEs, viruses, or certain non-protein-coding transcripts, are targeted by PTGS pathways like miRNAs or RNAi, leading to cleavage of the transcript. The resulting fragments can be converted into dsRNA by RDR6 and then processed into 21-22 nt sRNAs by DCL2 or DCL4. [8] Most of these 21-22 nt sRNAs are loaded into AGO1 and feed back into PTGS, amplifying PTGS efficiency. [79] However, some will instead associate with AGO6, leading to RdDM. [10] dsRNAs resulting from RDR6 activity can also sometimes processed by DCL3 instead of DCL2/4 and trigger RdDM. [9] Additionally, some Pol II transcripts contain inverted repeat sequences, which can form double-stranded hairpin-like structures. These can be cleaved by DCL proteins independent of RDRs to produce either 21-22 nt or 24 nt sRNAs that can participate in RdDM. [79] Similarly, miRNA precursors, which also form hairpin structures and are normally cleaved by DCL1 to produce miRNAs, can instead be cleaved by other DCLs to form sRNAs for RdDM. [79] While most non-canonical RdDM occurs via AGO6 or AGO4, there is also a version of the pathway where sRNAs instead associate with AGO2, which together with the NERD complex (Needed for RDR2-independent DNA methylation) recruits DRM2 to target loci and triggers DNA methylation. [92] Since the non-canonical pathways are not yet as well characterized as the canonical RdDM pathway, [79] there likely remain additional sources of sRNAs used for RdDM that have not yet been uncovered.

Factors involved

A number of factors involved in RdDM are listed below, along with additional details about their function and corresponding references. Several factors primarily involved in PTGS that sometimes participate in RdDM are also listed.

Factors involved in RdDM
Factor(s)Factor typePathwayRole in RdDMKnown direct interactorsDescriptionReferences
NRPD1 and the Pol IV complexRNA polymeraseCanonical RdDMsRNA productionCLSY proteins, RDR2Pol IV is a plant-specific RNA polymerase complex and NRPD1, its largest subunit, is specific to the complex. Through its interaction with the CLSY proteins and SHH1, Pol IV is recruited to heterochromatic regions (specifically to H3K9me2- and H3K4me0-containing chromatin), and transcribes single-stranded RNAs precursors of the sRNAs used in the canonical RdDM pathway. [93] [80] [94] [81]
NRPE1 and the Pol V complexRNA polymeraseAll RdDMDNA methylation of target lociPol V is a plant-specific RNA polymerase complex and NRPE1, its largest subunit, is specific to the complex. Pol V transcribes non-coding RNAs that serve as scaffolds for several other RdDM components, most importantly the AGO-sRNA duplex, but also SPT5L, and the IDN2-IDP complex. Both NRPE1 and SPT5L contain an AGO hook motif that helps recruit AGO4 to Pol V transcripts. Mutating the AGO hook motifs on both proteins results in reduced DNA methylation at RdDM target loci, resembling nrpe1 null mutant phenotypes. Binding of the AGO-sRNA duplex to complementary sites along the Pol V transcript leads to recruitment of DRM2 and addition of DNA methylation to target loci. [93] [80] [94] [95] [81]
RDR2RNA-dependent RNA polymeraseCanonical RdDMsRNA productionPol IVExists in a complex with Pol IV and converts the nascent Pol IV transcript to double-stranded RNA, which can then be processed by DCL3 to generate sRNAs for canonical RdDM. [83] [80]
RDR6RNA-dependent RNA polymerasePTGS, non-canonical RdDMsRNA productionConverts single-stranded RNAs to double-stranded RNAs for processing into 21-22 nt sRNAs by DCL2 and DCL4. Most of these sRNAs lead to PTGS, but some are loaded into AGO6 and participate in non-canonical RdDM. [9] [80]
DCL1 EndoribonucleasePTGS, non-canonical RdDMmiRNA production, sRNA productionAn endoribonuclease that cleaves double-stranded RNA, primarily involved in the production of microRNAs that lead to PTGS via AGO1. Can also catalyze the production of 21 nt sRNAs from mRNAs containing inverted repeats, which can be used in either PTGS or non-canonical RdDM depending on the AGO protein they associate with. The four DCL proteins in A. thaliana (DCL1,2,3,4) compete for access to dsRNA substrates. [96] [97] [81] [98]
DCL2 EndoribonucleasePTGS, Non-canonical RdDMsRNA productionAn endoribonuclease that cleaves double-stranded RNA, resulting in 22 nt sRNAs that can be used in both PTGS and non-canonical RdDM. The four DCL proteins in A. thaliana (DCL1,2,3,4) compete for access to dsRNA substrates, and DCL2,4 can substitute for loss of DCL3 for most RdDM targets. [96] [99] [97] [80]
DCL3 EndoribonucleaseCanonical RdDMsRNA productionAn endoribonuclease that cleaves double-stranded RNA, resulting in 24 nt sRNAs used in canonical RdDM. Preferentially targets the short dsRNAs produced by Pol IV-RDR2, but can also slice other dsRNA substrates, including mRNAs containing inverted repeats or miRNA precursors. The four DCL proteins in A. thaliana (DCL1,2,3,4) compete for access to dsRNA substrates, and DCL2,4 can substitute for loss of DCL3 for most RdDM targets. When PTGS pathways via DCL2,4 become saturated, DCL3 can step in and process the DCL2,4 dsRNA substrates, triggering a switch from PTGS to RdDM-mediated TGS. [96] [9] [99] [97] [80]
DCL4 EndoribonucleasePTGS, Non-canonical RdDMsRNA productionAn endoribonuclease that cleaves double-stranded RNA, resulting in 21 nt sRNAs that can be used for both PTGS and non-canonical RdDM. The four DCL proteins in A. thaliana (DCL1,2,3,4) compete for access to dsRNA substrates, and DCL2,4 can substitute for loss of DCL3 for most RdDM targets. [96] [99] [97]
AGO4Argonaute proteinCanonical RdDMDNA methylation of target lociNRPE1, SPT5LThe main Argonaute protein involved in canonical RdDM. AGO4 is partially redundant with AGO6, which can also function in this pathway, as well as with AGO9 in reproductive tissues. It binds the 24 nt sRNAs produced by the pathway to form an AGO4-sRNA duplex, which can recognize sequences complementary to the sRNA. Assisted by interactions with SPT5L, NRPE1, and the IDN2-IDP complex, the AGO4-sRNA duplex binds a single-stranded, noncoding RNA produced by Pol V, and helps recruit DRM2 to the DNA. [93] [80] [100]
AGO6Argonaute proteinAll RdDMDNA methylation of target lociAn argonaute protein that can function in either canonical or non-canonical RdDM pathways. Partially redundant with AGO4 (the main canonical RdDM AGO). Can associate with either 24 nt or 21-22 nt sRNAs to trigger RdDM at complementary loci. By interacting with both 21-22 nt and 24 nt sRNAs, AGO6 helps in the transition from PTGS (normally mediated by 21-22 nt sRNAs) to stable silencing by RdDM (normally mediated by 24 nt sRNAs). Expressed particularly in the root and shoot meristems, which are the two main stem cell populations in plants. This may indicate that plants increase surveillance for novel TEs in order to ensure genome integrity in the key cells that will give rise to most of the other cells in the plant. [93] [101] [10] [80] [100]
AGO9Argonaute proteinCanonical RdDMDNA methylation of target lociA highly specialized AGO expressed primarily in the germline, where it is required for proper female gamete formation. Interacts with 24 nt sRNAs to silence TEs in the germline, similar to the role of PIWI Argonaute proteins in animals. [102] [25] [100]
AGO1Argonaute proteinPTGS, non-canonical RdDMsRNA productionBinds microRNAs or 21-22 nt sRNAs, which it uses to recognize complementary sequences on other RNAs. When an AGO1-sRNA duplex (often called the RISC) finds a complementary single-stranded mRNA, the RNA is cleaved by AGO1, destroying the mRNA and causing PTGS. The resulting RNA fragments can then be converted to dsRNAs by RDR6 and processed by DCL2,4 to form secondary 21-22 nt sRNAs. These are predominantly loaded back into AGO1, forming a self-reinforcing ‘RNAi loop’. However, some of the 21-22 nt sRNAs are loaded into AGO6 instead, leading to RdDM. [91] [97] [80] [100]
DRM2DNA methyltransferaseAll RdDMDNA methylation of target lociThe main DNA methyltransferase involved in RdDM. Catalyzes the addition of a methyl group to cytosines in DNA. Recruited by the AGO4-sRNA duplex after it binds to a complementary sequence in a Pol V transcript, but the mechanism by which this happens is not well understood. [103] [80]
SHH1/DTF1DNA and chromatin binding proteinCanonicalsRNA productionCLSY1Required for Pol IV-derived sRNA production at a subset of RdDM loci. Via its SAWADEE domain, SHH1 binds histone H3 with specific modifications associated with heterochromatin and DNA methylation: methylation of the 9th lysine (H3K9me2) and unmethylated K4 (H3K4me0). By interacting with SHH1 via the CLSY proteins, Pol IV is recruited to heterochromatic/silent chromatin. To date, SHH1 has only been shown to directly interact with CLSY1. The ability of SHH1 to associate with Pol IV/NRPD1 is mostly abolished in clsy1,2 double mutants, so recruitment of Pol IV by SHH1 likely requires CLSY proteins. [104] [105] [106] [107]
CLSY1, CLSY2putative chromatin remodelersCanonicalsRNA productionPol IV, SHH1Required for SHH1 interaction with and recruitment of Pol IV to a subset of target loci. Mutually exclusive with loci regulated by CLSY3 and CLSY4. Together, the four CLSY proteins regulate nearly all Pol IV-derived sRNAs, and loss of all four results in a near total loss of 24-nucleotide sRNA production. Requires H3K9me2, likely through interaction with SHH1. sRNAs regulated by CLSY1,2 are enriched in the chromosome arms, while those regulated by CLSY3,4 are enriched in the pericentromere. [107] [108]
CLSY3, CLSY4putative chromatin remodelersCanonicalsRNA production, Pol IV targetingPol IVInvolved in recruitment of Pol IV to a subset of target loci. Mutually exclusive with loci regulated by CLSY1 and CLSY2. Together, the four CLSY proteins regulate nearly all Pol IV-sRNAs, and loss of all four results in a near total loss of 24-nucleotide sRNA production. sRNAs regulated by CLSY3,4 are enriched in the pericentromere, while sRNAs regulated by CLSY1,2 are enriched in the chromosome arms. [107] [108]
HEN1RNA methylaseBothsRNA productionnoneStabilizes sRNAs by adding methylation to the 3'-OH groups. [109]
SUVH2, SUVH9methyl-DNA binding proteinsBothDNA methylation of target lociDDR complex, MORC1, MORC6A pair of closely related methyl-DNA binding proteins that interact with the DDR complex and are required for proper localization of the DDR complex and Pol V. By recruiting Pol V to regions with DNA methylation, which tend to be silent, heterochromatic regions, SU(VAR)3-9 homolog (SUVH) 2 and 9 help form a positive feedback loop that reinforces RdDM-mediated silencing. May also associate with MORCs. [110]
DDR complex (RDM1, DMS3, DRD1)putative chromatin remodeling complexBothDNA methylation of target lociSUVH2, SUVH9The DDR complex, composed of DRD1, DMS3, and RDM1, is thought to facilitate access of Pol V to its target sites, possibly by unwinding DNA downstream of Pol V. Interacts with SUVH2,9, which bind methylated DNA, and this interaction may help recruit Pol V to regions of existing heterochromatin. RDM1 also binds single-stranded DNA, which may help unwind the DNA to facilitate recruitment of DRM2. [88] [111] [112] [113] [110]
SPT5L/RDM3/KTF1transcription factorBothDNA methylation of target lociAGO4, Pol V transcriptsInteracts with AGO4 and helps recruit it to the RNA scaffold produced by Pol V. Like the Pol V subunit NRPE1, SPT5L contains an AGO hook motif in its C-terminal domain. The motifs on both NRPE1 and SPT5L redundantly help recruit AGO4 to loci being transcribed by Pol V. Mutating the AGO hook motifs on both proteins results in reduced DNA methylation at RdDM target loci, resembling nrpe1 null mutant phenotypes. Also required for co-transcriptional slicing of Pol V transcripts. [114] [95] [115]
SWI/SNF complexchromatin remodeling complexBothDNA methylation of target lociIDN2The Switch/Sucrose non-fermentable (SWI/SNF) complex is a chromatin remodeling complex that is recruited to Pol V scaffolds by the IDN2-IDP complex, where it affects nucleosome positioning. SWI/SNF may promote RdDM by making the chromatin more accessible, which may facilitate access of DRM2 to DNA. [116]
IDN2-IDP complexdsRNA-binding proteinBothDNA methylation of target lociSWI/SNF complexA complex composed of IDN2 and IDP1 (also called IDNL1) or IDP2 (IDNL2). IDN2, and possibly IDP1, can bind the dsRNA duplex formed when AGO-associated sRNAs hybridize with the Pol V scaffold. This complex is thought to help stabilize base pairing between the AGO-sRNA and Pol V scaffold RNA. IDN2-IDP may also facilitate recruitment of the SWI/SNF complex to Pol V scaffolds. Additionally, IDP1 can bind unmethylated DNA, which may help recruit DRM2 to regions lacking DNA methylation. [117] [116] [118]
NERDGW repeat- and PHD finger-containing proteinNon-canonical RdDMsRNA production, DNA methylation of target lociAGO2Forms a non-canonical RdDM pathway that includes a number of genes involved in PTGS, including AGO2. Binds histone H3 and AGO2. Required for 21 nt sRNA accumulation at some non-canonical RdDM targets, including novel TE insertions. Leads to histone tail modifications associated with transcriptional repression; because these modifications can recruit other DNA methylation machinery, including canonical RdDM, it is unclear if the effect of NERD on DNA methylation is direct or indirect. [92] [79]
MORC1, MORC6GHKL ATPasesBothDNA methylation of target loci (?)SUVH2, SUVH9, IDN2, DMS3Microrchidia 1 (MORC1) and MORC6 form a heterodimer and may interact with the DDR complex to recruit Pol V. However, they are thought to mainly act downstream of DNA methylation to promote silencing. Their precise role in RdDM is still unclear. [110] [80] [90]
DRM1DNA methyltransferaseAll RdDMDNA methylation of target lociA homolog of DRM2 that is only expressed during sexual reproduction, specifically in the egg cell and potentially the early embryo. DRM2 is likely the main RdDM methyltransferase in all other tissues. [119]
HDA6Histone deacetylaseCanonical RdDMsRNA productionMay facilitate Pol IV recruitment by creating a permissive chromatin state for SHH1 binding by removing histone acetylation, promoting H3K9 methylation. In histone deacetylase 6 (hda6) mutant plants, HDA6 target loci lose Pol IV targeting and sRNA biogenesis, suggesting HDA6 is involved in Pol IV recruitment at a subset of RdDM target loci. Further, normal Pol IV targeting cannot be restored after re-introduction of functional HDA6, suggesting that HDA6 is also required to propagate the trans-generational 'memory' of where Pol IV should be targeted. HDA6 physically associates with MET1 and facilitates CG methylation maintenance by MET1, which may also be important for sRNA production at HDA6-dependent loci. [120] [80]

Interactions with other chromatin modifying pathways

Different chromatin states, like active euchromatin or silent heterochromatin, are defined by a combination of specific histone modification and DNA methylation patterns. Repressive chromatin modifications, like DNA methylation, help promote DNA compaction and reduce DNA accessibility, while other modifications help open chromatin and increase accessibility. Methylation of the 9th lysine of histone H3 (H3K9), primarily in the form of H3K9 trimethylation (H3K9me3) in animals and H3K9 dimethylation (H3K9me2) in plants, is a highly conserved repressive modification. [121] [122] Lack of H3K4 methylation (H3K4me0) is also associated with repression, along with several other histone modifications and variants. The combination of DNA methylation, H3K9me2, and H3K4me0 is strongly associated with heterochromatin in plants.

Since DNA methylation and repressive histone modifications together define heterochromatin, most DNA methylation pathways in plants recognize and interact with repressive histone marks and vice versa, forming positive feedback loops that help maintain the repressive chromatin state. [123] The RdDM-associated protein SHH1 recognizes H3K4me0 and H3K9me2 at heterochromatic loci and recruits Pol IV to these loci to trigger additional DNA methylation at these regions. [106] Similarly, SUVH2 and SUVH9 help recruit Pol V to loci with DNA methylation. [110] Thus, both major parts of the canonical RdDM pathway are preferentially recruited to regions that are already in the silent, heterochromatic state marked by DNA methylation, H3K9me2, and H3K4me0. DNA methylation at these same heterochromatic loci is also recognized by the histone methyltransferases SUVH4/KYP, SUVH5, and SUVH6, which bind to non-CG methylation and add H3K9me2 to nearby histones, [123] [124] closing the positive feedback loop. Similarly, CMT3 and CMT2, the two DNA methyltransferases involved in the maintenance of CHG and CHH methylation respectively, [75] both bind and add DNA methylation to H3K9me2-marked heterochromatin, forming their own feedback loop with SUVH4/5/6. [125] [123] These interactions help strongly reinforce silencing at TEs and other heterochromatic regions.

A similar feedback loop occurs in animals. HP1 plays a vital role in maintaining heterochromatin by propagating H3K9 methylation through a positive feedback loop with the H3K9 methyltransferase SUV39H. [126] H3K9 methylation recruits HP1, which recruits SUV39H to deposit more H3K9 methylation. [126] Though HP1 is conserved in plants, its function in this feedback loop is not. [127] Instead, the positive feedback loops between H3K9me2 and the RdDM and CMT2/3 DNA methylation pathways fulfill a similar function in propagating H3K9me2. More recently, a plant-specific protein, Agenet Domain Containing Protein 1 (ADCP1), was also identified that may function analogously to HP1 in maintaining H3K9me2 levels in heterochromatin, facilitating heterochromatin formation. [128]

Ultimately, the constant reinforcement of silencing chromatin modifications at heterochromatic loci creates a repressive chromatin state wherein the DNA and histones (nucleosomes) become tightly packed together. This helps silence gene expression by physically inhibiting access to the DNA, preventing RNA Polymerase II, transcription factors and other proteins from initiating transcription. [129] However, this same compaction also prevents factors involved in heterochromatin maintenance from accessing the DNA, which could lead to the silent, compact state being lost. This is particularly true in the dense constitutive heterochromatin surrounding the centromere. In these regions, the chromatin remodeler DDM1 plays a crucial role in DNA methylation maintenance by displacing nucleosomes temporarily to allow methyltransferases and other factors access the DNA. [130] [131] [5] However, since most RdDM targets are small TEs in open, accessible and gene-rich regions (see “TE silencing and genome stability”), few RdDM sites require DDM1. [5] [99] In fact, dense heterochromatin inhibits RdDM. [5] By contrast, CMT2 and CMT3 preferentially function in constitutive heterochromatin and depend strongly on DDM1 to maintain silencing over these regions. [131] [5] [3] Similarly, MET1, which maintains DNA methylation at CG sites after replication, requires DDM1 to access heterochromatin and maintain CG methylation in those regions. [132] Thus, DDM1 is a key regulator of DNA methylation in dense heterochromatin, but regulates sites mostly independently from RdDM. [5] [99]

Interactions between RdDM and the other three maintenance DNA methylation pathways are limited and predominantly indirect. The DNA methyltransferase MET1 robustly maintains CG methylation genome-wide, including at RdDM target sites. In RdDM mutants, non-CG methylation at RdDM target sites is lost, but CG methylation is still maintained, suggesting that MET1 activity is independent of RdDM. [99] However, although met1 mutants lose CG methylation as expected, they also lose much of their non-CG methylation, including at RdDM target loci. [99] At these sites, silencing can still be initiated by RdDM in met1 mutants, but it is not maintained or transmitted to progeny, suggesting that MET1 is important for the maintenance, but not initiation, of silencing at a subset of RdDM target loci. [133] [120] This effect is likely indirect: loss of MET1 leads to loss of H3K9me2 at some sites, which inhibits the recruitment of Pol IV and therefore prevents maintenance of DNA methylation via canonical RdDM, although the non-canonical pathways (which do not involve Pol IV) are not affected. [99] [120] Loss of the histone deacetylase HDA6, which facilitates maintenance methylation by MET1 at some loci, has a similar effect, suggesting that multiple different factors involved in maintaining heterochromatin likely facilitate RdDM-mediated DNA methylation maintenance. [120]

Loss of RdDM leads to strong loss of non-CG methylation at TEs in gene-rich regions in the chromosome arms, but has little effect on DNA methylation levels in the constitutive heterochromatin around the centromere. [99] [5] [3] This suggests that CMT2 and CMT3, which function primarily to maintain CHG and CHH methylation in dense constitutive heterochromatin, do not depend on RdDM activity. [99] [5] [3] Similarly, in cmt2,cmt3 double mutants, many TEs in the chromosome arms remain methylated, presumably due to the persistent activity of RdDM, indicating that loss of CMT2/3 has little effect on RdDM activity. [5] [3] This suggests that RdDM and CMT2/3 function mostly independently and at distinct loci: RdDM is the main pathway responsible for maintaining non-CG DNA methylation in euchromatic, gene rich regions, while CMT2 and CMT3 maintain non-CG DNA methylation in constitutive heterochromatin. In mutants defective in both RdDM and CMT2/CMT3, all non-CG methylation in the genome is eliminated, [74] demonstrating that together RdDM and CMT2/CMT3 account for all non-CG methylation in the genome.

Balance between DNA methylation and demethylation

Most DNA methylation mechanisms in plants are self-reinforcing (see above), including RdDM: Pol IV and Pol V are both recruited to heterochromatic regions that already have DNA methylation, encouraging additional DNA methylation via canonical RdDM. [55] Positive feedback loops like these can cause DNA methylation activity to spread out from the intended methylated target sites into genes or other regulatory elements, which can negatively affect gene expression. To prevent this spreading, DNA methylation pathways are opposed by passive and active DNA demethylation. DNA methylation can be lost passively with each cell division, because newly synthesized strands of DNA lack DNA methylation until it is re-added by one of the maintenance DNA methylation pathways. [134] DNA methylation can also be actively removed in plants by DNA glycosylases, which remove methylated cytosines via the base excision repair pathway. In Arabidopsis, there are four proteins responsible for removing DNA methylation: Repressor of silencing 1 (ROS1), Demeter (DME), Demeter-like 2 (DML2), and Demeter-like 3 (DML3). [135] [136] These DNA glycosylases help prevent the spread of DNA methylation from RdDM targets to active genes. [137] [14] Loss of active DNA demethylation in ros1;dml2;dml3 triple mutants leads to a widespread increase in DNA methylation levels, whereas ectopic expression of ROS1 leads to progressive loss of DNA methylation at many loci, [138] highlighting the importance of balancing DNA methylation and demethylation activity.

Interestingly, expression of the DNA demethylase ROS1 is directly tied to RdDM activity: DNA methylation over a TE targeted by RdDM in the ROS1 promoter is required for ROS1 expression, [12] [13] though other factors are also involved in regulating ROS1. [139] [140] Since ROS1 expression is tied to DNA methylation at a specific TE, ROS1 expression is also strongly reduced in plants with defective RdDM that lose the ability to methylate that TE and others. [12] This general mechanism helps maintain DNA methylation homeostasis by tuning DNA demethylation activity to DNA methylation activity, helping to ensure that DNA methylation patterns can be stably maintained over time.

Evolutionary conservation

Origins of RdDM pathway members

NRPDE phylogenetic.png
A schematic depicting the evolutionary conservation of selected Pol IV and V subunit orthologs within the plant kingdom. Subunits beginning with NRPD are Pol IV subunits, subunits beginning with NRPE are Pol V subunits, and subunits labeled as NRPD/E are found in both Pol IV and V. [141] A filled circle for a subunit indicates that an ortholog for that subunit has been identified within the associated lineage.
RdDM phylogenetic.png
A schematic depicting the evolutionary conservation of selected RdDM pathway component orthologs within the plant kingdom. A filled circle for a subunit indicates that an ortholog for that subunit has been identified within the associated lineage.

While all eukaryotes share three RNA polymerases (RNA Pol I, II and III), plants have two additional polymerases, Pol IV and Pol V. Both Pol IV and V share an evolutionary origin, deriving from Pol II. [141] [94] In other eukaryotic kingdoms that lack these two specialized RNA polymerases, Pol II transcribes the precursors of small RNAs used in silencing pathways – in fact, Pol II transcripts are also sometimes processed into sRNAs in plants. It has been hypothesized that the origin of both Pol IV and Pol V is rooted in “escape from adaptive conflict”. [142] The idea is that potential tensions between the “traditional” function of Pol II and the small RNA biogenesis function could be relieved by duplication of Pol II and subfunctionalization of the resulting multiple RNA polymerases.

Analyses of evolutionary lineage for Pol IV and Pol V are complicated to some extent by the fact that each enzyme is actually composed of at least 12 subunits. [141] In Arabidopsis thaliana, some subunits are shared between Pol IV and Pol V, some are unique to each polymerase, and some are shared between Pol II, IV, and V. [143] Orthologs of certain Pol IV and V subunits have been found in all lineages of land plants, including ferns, liverworts, and mosses. [144] [142] These findings argue for a shared origin of Pol IV and V dating back to early land / vascular plants.

Much of the work done to elucidate the genes and proteins involved in the RdDM pathway has been performed in Arabidopsis thaliana, a model angiosperm. However, studies of Pol IV and V conducted in maize show some key differences with Arabidopsis. Maize Pol IV and V differ from each other in terms of only one subunit (the largest one). In Arabidopsis, Pol IV and V differ from each other in terms of three subunits. [145] However, maize utilizes a set of interchangeable catalytic subunits – two in the case of Pol IV and three in the case of Pol V – that provide additional specialization of polymerase functionality. [145] While differences exist, overall there is a broad overlap in RdDM functions and components between the different angiosperm species studied to date.

Outside of Pol IV and Pol V, a large proportion of key RdDM component proteins (for example, DCL3 and AGO4) have orthologs found within each class of land plants, which provides support for the hypothesis that some form of the RdDM pathway evolved early within the plant lineage. [142] However, RdDM pathway functionality does appear to change to an appreciable extent between different plant species and lineages. For example, while gymnosperms have functional Pol IV and produce 24 nt small RNAs, the biogenesis of sRNAs within gymnosperms is much more heavily skewed towards 21 nt than 24 nt sRNAs. [146] This suggests that canonical RdDM may be rarer or less pronounced in gymnosperms than in angiosperms. Similarly, while orthologs of DRM2 are found in various angiosperms, there are no known DRM2 orthologs in other plant lineages. [147] One possibility is that angiosperms have the “most complete” version of the RdDM pathway, with all other plant lineages possessing robust and functional subsets of the pathway. However, since nearly all of the work on RdDM has been done in angiosperms, it is also possible that alternative versions of RdDM in other lineages have simply not yet been uncovered, particularly if these alternative versions include different proteins or proteins without clear homologs in angiosperms.

Relationships with sRNA silencing pathways in other kingdoms

All eukaryotic kingdoms host some form of small RNAs. One such class of sRNAs is the Piwi-interacting RNAs (piRNAs). Much like in RdDM, piRNAs primarily function to target and silence transposons, particularly in the germline. [29] [30] However, piRNAs are only found in animals, are longer than the small RNAs functioning in RdDM (24-32 nucleotides), and mediate their functions through interactions with a different subclass of AGO proteins, the PIWI subfamily, which are absent from plants. [29] [30] MicroRNAs (miRNAs) are another class of small RNA with silencing properties. [148] While miRNAs are in a similar size range as RdDM sRNAs (~21 nt), miRNAs associate with a distinct set of Argonaute proteins that silence target RNAs by initiating their degradation or blocking their downstream translation into proteins, rather than recruiting DRM2 to add DNA methylation to nearby DNA. Both RdDM and the miRNA pathways involve related proteins from the Argonaute and Dicer families. [148]

Perhaps the most analogous pathways to RdDM in another eukaryotic kingdom are the sRNA directed transcriptional gene silencing (TGS) and co-transcriptional gene silencing (CTGS) pathways in Schizosaccharomyces pombe. [149] In S. pombe, TGS directs methylation of H3K9, leading to heterochromatin formation, and is directed by sRNAs produced from the targeted regions. [150] Similar to canonical RdDM, this pathway is a positive feedback loop: sRNAs are generated preferentially from heterochromatin-rich areas of the genome, and these sRNAs direct the addition of K3K9 methylation to maintain/spread heterochromatin. Meanwhile, CTGS is directed by AGO1-bound sRNAs, similar to PTGS within plants, and results in the inhibition of transcription by Pol II, as well as to Pol II release. [151] [152] Unlike RdDM, TGS and CTGS in S. pombe do not rely on transcription from non-Pol II sources or lead to the addition of DNA methylation. However, the S. pombe pathways and RdDM share many of the same components, like RNA-directed RNA polymerases and sRNAs, and have similar functions in maintaining heterochromatin.

History

Introducing transgenes into organisms has been a widely used tool in plant genetics research for decades. However, researchers often find that their introduced transgenes are not expressed as strongly as expected, or sometimes even at all, a phenomenon called transgene silencing. [153] The discovery of transgene silencing in the 1990s spurred a great deal of interest in understanding the mechanisms behind this silencing. [154] [155] [156] Researchers found that transgene silencing was ubiquitous, occurring in multiple species (including Arabidopsis, Tobacco, and Petunia), and was associated with increased DNA methylation over and around the silenced transgene. [157] [158] [159]

Around the same time in 1994, work in tobacco plants had revealed a new pathway involving RNAs that resulted in DNA methylation. Researchers found that when viroids were introduced into the plant and integrated into the plant genome, the viroid sequences, but not the host genome, gained DNA methylation. [49] The deposition of methylation over these foreign viroid sequences helped inhibit viroid replication, and was therefore thought to represent a plant pathogen defense mechanism. The evidence suggested that the viroid RNAs produced during viroid replication were being used by the plant as a template to help target DNA methylation to the viroid sequences. This mechanism was therefore named RNA-directed DNA methylation, or RdDM. [49]

RdDM turned out to be the solution to the transgene mystery: like viroids and viruses, transgenes are foreign sequences, and as a result they are often recognized as foreign invaders and targeted for silencing by RdDM and PTGS. Since transgene silencing was a reliable marker of RdDM activity, researchers were able to design genetic screens to identify mutants that failed to trigger silencing at transgenes, reasoning that these genes were likely to be involved in the RdDM pathway. These experiments revealed many parts of the pathway, including RNA Pol IV and V, Dicer-like proteins, Argonautes, and others. [6] [160] [161]

The involvement of sRNAs in RdDM was initially suspected due to the similarity between RdDM and RNAi, the latter of which had recently been shown to involve small RNAs. [49] [162] To test whether sRNAs were involved in RdDM, RNA hairpin structures complementary to a specific gene promoter were introduced into Arabidopsis and Tobacco. [163] The hairpin RNAs were processed into sRNAs, which were able to trigger the addition of DNA methylation to the targeted promoter and silence the gene. [163] This demonstrated that sRNAs could direct DNA methylation to specific loci. Later efforts showed that the sRNAs involved in RdDM were approximately 24-26 nt long, while the sRNAs associated with RNAi were only about 21-22 nt in length. [164] [165] Soon after, the identification of AGO4 and characterization of its role in RdDM led to predictions, later confirmed, that 24 nt sRNAs were associating with AGO4 and directing DNA methylation to complementary loci. [166] [165]

Early work on transgene silencing and RdDM also identified SDE4 as required for the production of most sRNAs involved in RdDM. [167] SDE4 would later be identified as the largest subunit of Pol IV, and renamed NRPD1. A number of studies published in quick succession from multiple research groups, utilizing both forward and reverse genetic approaches, went on to identify and characterize Pol IV and Pol V as highly specialized plant RNA polymerases involved in RdDM. [168] [169] [170] [171] The Pol IV / Pol V naming convention was adopted shortly thereafter. [88] [141]

Potential biotechnology applications

Since the mechanism underlying the sequence-specificity of RdDM is well known, RdDM can be ‘tricked’ into targeting and silencing endogenous genes in a highly specific manner, which has a number of potential biotechnological and bioengineering applications. Several different methods can be used to trigger RdDM-based DNA methylation and silencing of specific genes. One method, called virus-induced gene silencing (VIGS), involves inserting part of the promoter sequence of the desired target gene into a virus. [172] The virus will reproduce the chunk of promoter sequence as part of its own RNA, which is otherwise foreign to the plant. Because the viral RNA is foreign, it will be targeted for PTGS and processed into sRNAs, some of which will be complementary to the original target gene's promoter. A subset of these sRNAs will recruit the RdDM machinery to the target gene to add DNA methylation. In one study, researchers used this method with an engineered Cucumber Mosaic Virus to recruit RdDM to silence a gene that affected flower pigmentation in petunia, and another that affected fruit ripening in tomato. [173] In both cases, they showed that DNA methylation was added to the locus as expected. In petunia, both the gain of DNA methylation and changes in flower coloration were heritable, while only partial silencing and heritability were observed in tomato. VIGS has also been used to silence the FLOWERING WAGENINGEN (FWA) locus in Arabidopsis, which resulted in plants that flowered later than normal. [172] The same study also showed that the inhibitory effect of VIGS on FWA and flowering can become stronger over the course of successful generations. [172]

Another method to target RdDM to a desired target gene involves introducing a hairpin RNA construct that is complementary to the target locus. Hairpin RNAs contain an inverted repeat, which causes the RNA molecule to form a double-stranded RNA (dsRNA) structure called an RNA hairpin. The dsRNA hairpin can be processed by DCL proteins into sRNAs which are complementary to the target locus, triggering RdDM at that locus. This method has been used in several studies. [12] [174] [175]

Changes induced by RdDM can sometimes be maintained and inherited over multiple generations without outside intervention or manipulation, suggesting that RdDM can be a valuable tool for targeted epigenome editing. Recent work has even bypassed RdDM altogether by artificially tethering DRM2 (or other components of the RdDM pathway) directly to specific target loci, using either zinc finger nucleases or CRISPR. [90] [176] In these experiments, tethering the RdDM machinery to a specific locus led to gain of DNA methylation at the target site that was often heritable for multiple generations, even once the artificial construct was removed through crossing. For all of these methods, however, more work on minimizing off-target effects and increasing DNA methylation efficiency is needed.

Genetically Modified Organisms (GMOs) have played a large role in recent agricultural research and practice, but have proven controversial, and face regulatory barriers to implementation in some jurisdictions. GMOs are defined by the inclusion of “foreign” genetic material into the genome. The treatment of plants with engineered RNAs or viruses intended to trigger RdDM does not change the underlying DNA sequence of the treated plant's genome; only the epigenetic state of portions of the DNA sequence already present are altered. As a result, these plants are not considered GMOs. This has led to efforts to utilize RdDM and other RNA-mediated effects to induce agriculturally-beneficial traits, like altering pathogen or herbicide susceptibility, or speeding up plant breeding by quickly inducing favorable traits. [177] [178] [179] However, while this is an area of active interest, there are few broadly implemented applications as of now.

Related Research Articles

<span class="mw-page-title-main">Histone</span> Family proteins package and order the DNA into structural units called nucleosomes.

In biology, histones are highly basic proteins abundant in lysine and arginine residues that are found in eukaryotic cell nuclei and in most Archaeal phyla. They act as spools around which DNA winds to create structural units called nucleosomes. Nucleosomes in turn are wrapped into 30-nanometer fibers that form tightly packed chromatin. Histones prevent DNA from becoming tangled and protect it from DNA damage. In addition, histones play important roles in gene regulation and DNA replication. Without histones, unwound DNA in chromosomes would be very long. For example, each human cell has about 1.8 meters of DNA if completely stretched out; however, when wound about histones, this length is reduced to about 90 micrometers (0.09 mm) of 30 nm diameter chromatin fibers.

<span class="mw-page-title-main">Epigenetics</span> Study of DNA modifications that do not change its sequence

In biology, epigenetics is the study of heritable traits, or a stable change of cell function, that happen without changes to the DNA sequence. The Greek prefix epi- in epigenetics implies features that are "on top of" or "in addition to" the traditional genetic mechanism of inheritance. Epigenetics usually involves a change that is not erased by cell division, and affects the regulation of gene expression. Such effects on cellular and physiological phenotypic traits may result from environmental factors, or be part of normal development. They can lead to cancer.

<span class="mw-page-title-main">Euchromatin</span> Lightly packed form of chromatin that is enriched in genes

Euchromatin is a lightly packed form of chromatin that is enriched in genes, and is often under active transcription. Euchromatin stands in contrast to heterochromatin, which is tightly packed and less accessible for transcription. 92% of the human genome is euchromatic.

Heterochromatin is a tightly packed form of DNA or condensed DNA, which comes in multiple varieties. These varieties lie on a continuum between the two extremes of constitutive heterochromatin and facultative heterochromatin. Both play a role in the expression of genes. Because it is tightly packed, it was thought to be inaccessible to polymerases and therefore not transcribed; however, according to Volpe et al. (2002), and many other papers since, much of this DNA is in fact transcribed, but it is continuously turned over via RNA-induced transcriptional silencing (RITS). Recent studies with electron microscopy and OsO4 staining reveal that the dense packing is not due to the chromatin.

In molecular biology and genetics, transcriptional regulation is the means by which a cell regulates the conversion of DNA to RNA (transcription), thereby orchestrating gene activity. A single gene can be regulated in a range of ways, from altering the number of copies of RNA that are transcribed, to the temporal control of when the gene is transcribed. This control allows the cell or organism to respond to a variety of intra- and extracellular signals and thus mount a response. Some examples of this include producing the mRNA that encode enzymes to adapt to a change in a food source, producing the gene products involved in cell cycle specific activities, and producing the gene products responsible for cellular differentiation in multicellular eukaryotes, as studied in evolutionary developmental biology.

<span class="mw-page-title-main">Regulation of gene expression</span> Modifying mechanisms used by cells to increase or decrease the production of specific gene products

Regulation of gene expression, or gene regulation, includes a wide range of mechanisms that are used by cells to increase or decrease the production of specific gene products. Sophisticated programs of gene expression are widely observed in biology, for example to trigger developmental pathways, respond to environmental stimuli, or adapt to new food sources. Virtually any step of gene expression can be modulated, from transcriptional initiation, to RNA processing, and to the post-translational modification of a protein. Often, one gene regulator controls another, and so on, in a gene regulatory network.

<span class="mw-page-title-main">DNA repair</span> Cellular mechanism

DNA repair is a collection of processes by which a cell identifies and corrects damage to the DNA molecules that encodes its genome. In human cells, both normal metabolic activities and environmental factors such as radiation can cause DNA damage, resulting in tens of thousands of individual molecular lesions per cell per day. Many of these lesions cause structural damage to the DNA molecule and can alter or eliminate the cell's ability to transcribe the gene that the affected DNA encodes. Other lesions induce potentially harmful mutations in the cell's genome, which affect the survival of its daughter cells after it undergoes mitosis. As a consequence, the DNA repair process is constantly active as it responds to damage in the DNA structure. When normal repair processes fail, and when cellular apoptosis does not occur, irreparable DNA damage may occur. This can eventually lead to malignant tumors, or cancer as per the two-hit hypothesis.

<span class="mw-page-title-main">DNA methylation</span> Biological process

DNA methylation is a biological process by which methyl groups are added to the DNA molecule. Methylation can change the activity of a DNA segment without changing the sequence. When located in a gene promoter, DNA methylation typically acts to repress gene transcription. In mammals, DNA methylation is essential for normal development and is associated with a number of key processes including genomic imprinting, X-chromosome inactivation, repression of transposable elements, aging, and carcinogenesis.

<span class="mw-page-title-main">Paramutation</span>

In epigenetics, a paramutation is an interaction between two alleles at a single locus, whereby one allele induces a heritable change in the other allele. The change may be in the pattern of DNA methylation or histone modifications. The allele inducing the change is said to be paramutagenic, while the allele that has been epigenetically altered is termed paramutable. A paramutable allele may have altered levels of gene expression, which may continue in offspring which inherit that allele, even though the paramutagenic allele may no longer be present. Through proper breeding, paramutation can result in siblings that have the same genetic sequence, but with drastically different phenotypes.

The family of heterochromatin protein 1 (HP1) consists of highly conserved proteins, which have important functions in the cell nucleus. These functions include gene repression by heterochromatin formation, transcriptional activation, regulation of binding of cohesion complexes to centromeres, sequestration of genes to the nuclear periphery, transcriptional arrest, maintenance of heterochromatin integrity, gene repression at the single nucleosome level, gene repression by heterochromatization of euchromatin, and DNA repair. HP1 proteins are fundamental units of heterochromatin packaging that are enriched at the centromeres and telomeres of nearly all eukaryotic chromosomes with the notable exception of budding yeast, in which a yeast-specific silencing complex of SIR proteins serve a similar function. Members of the HP1 family are characterized by an N-terminal chromodomain and a C-terminal chromoshadow domain, separated by a hinge region. HP1 is also found at some euchromatic sites, where its binding can correlate with either gene repression or gene activation. HP1 was originally discovered by Tharappel C James and Sarah Elgin in 1986 as a factor in the phenomenon known as position effect variegation in Drosophila melanogaster.

RNA-induced transcriptional silencing (RITS) is a form of RNA interference by which short RNA molecules – such as small interfering RNA (siRNA) – trigger the downregulation of transcription of a particular gene or genomic region. This is usually accomplished by posttranslational modification of histone tails which target the genomic region for heterochromatin formation. The protein complex that binds to siRNAs and interacts with the methylated lysine 9 residue of histones H3 (H3K9me2) is the RITS complex.

<span class="mw-page-title-main">Eukaryotic transcription</span> Transcription is heterocatalytic function of DNA

Eukaryotic transcription is the elaborate process that eukaryotic cells use to copy genetic information stored in DNA into units of transportable complementary RNA replica. Gene transcription occurs in both eukaryotic and prokaryotic cells. Unlike prokaryotic RNA polymerase that initiates the transcription of all different types of RNA, RNA polymerase in eukaryotes comes in three variations, each translating a different type of gene. A eukaryotic cell has a nucleus that separates the processes of transcription and translation. Eukaryotic transcription occurs within the nucleus where DNA is packaged into nucleosomes and higher order chromatin structures. The complexity of the eukaryotic genome necessitates a great variety and complexity of gene expression control.

RNA polymerase IV is an enzyme that synthesizes small interfering RNA (siRNA) in plants, which silence gene expression. RNAP IV belongs to a family of enzymes that catalyze the process of transcription known as RNA Polymerases, which synthesize RNA from DNA templates. Discovered via phylogenetic studies of land plants, genes of RNAP IV are thought to have resulted from multistep evolution processes that occurred in RNA Polymerase II phylogenies. Such an evolutionary pathway is supported by the fact that RNAP IV is composed of 12 protein subunits that are either similar or identical to RNA polymerase II, and is specific to plant genomes. Via its synthesis of siRNA, RNAP IV is involved in regulation of heterochromatin formation in a process known as RNA directed DNA Methylation (RdDM).

RNA polymerase V, previously known as RNA polymerase IVb, is a multisubunit plant specific RNA polymerase. It is required for normal function and biogenesis of small interfering RNA (siRNA). Together with RNA polymerase IV, Pol V is involved in an siRNA-dependent epigenetic pathway known as RNA-directed DNA methylation (RdDM), which establishes and maintains heterochromatic silencing in plants.

<span class="mw-page-title-main">Thomas Jenuwein</span> German scientist

Thomas Jenuwein is a German scientist working in the fields of epigenetics, chromatin biology, gene regulation and genome function.

HSV epigenetics is the epigenetic modification of herpes simplex virus (HSV) genetic code.

Decreased DNA Methylation I (DDM1), is a plant gene that encodes a nucleosome remodeler which facilitates DNA methylation. The DDM1 gene has been described extensively in Arabidopsis thaliana and also in maize. The protein has been described to be similar to the SWI2/SNF2 chromatin remodeling proteins.

<span class="mw-page-title-main">Julie Law</span> American molecular and cellular biologist

Julie Law is an American molecular and cellular biologist. Law's pioneering work on DNA methylation patterns led to the discovery of the role of the CLASSY protein family in DNA methylation. Law is currently an associate professor at the Salk Institute for Biological Studies.

Jian-Kang Zhu is a plant scientist, researcher and academic. He is a Senior Principal Investigator in the Shanghai Center for Plant Stress Biology, Chinese Academy of Sciences (CAS). He is also the Academic Director of CAS Center of Excellence in Plant Sciences.

Transgenerational epigenetic inheritance in plants involves mechanisms for the passing of epigenetic marks from parent to offspring that differ from those reported in animals. There are several kinds of epigenetic markers, but they all provide a mechanism to facilitate greater phenotypic plasticity by influencing the expression of genes without altering the DNA code. These modifications represent responses to environmental input and are reversible changes to gene expression patterns that can be passed down through generations. In plants, transgenerational epigenetic inheritance could potentially represent an evolutionary adaptation for sessile organisms to quickly adapt to their changing environment.

References

Open Access logo PLoS transparent.svg This article was adapted from the following source under a CC BY 4.0 license (2020) (reviewer reports): Robert M. Erdmann; Colette Lafontaine Picard (8 October 2020). "RNA-directed DNA Methylation". PLOS Genetics . 16 (10): e1009034. doi:10.1371/JOURNAL.PGEN.1009034. ISSN   1553-7390. PMID   33031395. Wikidata   Q100233435.{{cite journal}}: CS1 maint: unflagged free DOI (link)

  1. 1 2 3 4 Dubin MJ, Mittelsten Scheid O, Becker C (April 2018). "Transposons: a blessing curse". Current Opinion in Plant Biology. 42: 23–29. Bibcode:2018COPB...42...23D. doi: 10.1016/j.pbi.2018.01.003 . PMID   29453028.
  2. Wicker T, Gundlach H, Spannagl M, Uauy C, Borrill P, Ramírez-González RH, et al. (August 2018). "Impact of transposable elements on genome structure and evolution in bread wheat". Genome Biology. 19 (1): 103. doi: 10.1186/s13059-018-1479-0 . PMC   6097303 . PMID   30115100.
  3. 1 2 3 4 5 6 7 8 9 10 11 Sigman MJ, Slotkin RK (February 2016). "The First Rule of Plant Transposable Element Silencing: Location, Location, Location". The Plant Cell. 28 (2): 304–13. doi:10.1105/tpc.15.00869. PMC   4790875 . PMID   26869697.
  4. Deniz Ö, Frost JM, Branco MR (July 2019). "Regulation of transposable elements by DNA modifications". Nature Reviews. Genetics. 20 (7): 417–431. doi:10.1038/s41576-019-0106-6. PMID   30867571. S2CID   76662244.
  5. 1 2 3 4 5 6 7 8 9 10 Zemach A, Kim MY, Hsieh PH, Coleman-Derr D, Eshed-Williams L, Thao K, et al. (March 2013). "The Arabidopsis nucleosome remodeler DDM1 allows DNA methyltransferases to access H1-containing heterochromatin". Cell. 153 (1): 193–205. doi:10.1016/j.cell.2013.02.033. PMC   4035305 . PMID   23540698.
  6. 1 2 3 Chan SW, Zilberman D, Xie Z, Johansen LK, Carrington JC, Jacobsen SE (February 2004). "RNA silencing genes control de novo DNA methylation". Science. 303 (5662): 1336. doi:10.1126/science.1095989. PMID   14988555. S2CID   44659873.
  7. Pérez-Hormaeche J, Potet F, Beauclair L, Le Masson I, Courtial B, Bouché N, Lucas H (July 2008). "Invasion of the Arabidopsis genome by the tobacco retrotransposon Tnt1 is controlled by reversible transcriptional gene silencing". Plant Physiology. 147 (3): 1264–78. doi:10.1104/pp.108.117846. PMC   2442547 . PMID   18467467.
  8. 1 2 Nuthikattu S, McCue AD, Panda K, Fultz D, DeFraia C, Thomas EN, Slotkin RK (May 2013). "The initiation of epigenetic silencing of active transposable elements is triggered by RDR6 and 21-22 nucleotide small interfering RNAs". Plant Physiology. 162 (1): 116–31. doi:10.1104/pp.113.216481. PMC   3641197 . PMID   23542151.
  9. 1 2 3 4 5 6 Marí-Ordóñez A, Marchais A, Etcheverry M, Martin A, Colot V, Voinnet O (September 2013). "Reconstructing de novo silencing of an active plant retrotransposon". Nature Genetics. 45 (9): 1029–39. doi:10.1038/ng.2703. PMID   23852169. S2CID   13122409.
  10. 1 2 3 4 5 McCue AD, Panda K, Nuthikattu S, Choudury SG, Thomas EN, Slotkin RK (January 2015). "ARGONAUTE 6 bridges transposable element mRNA-derived siRNAs to the establishment of DNA methylation". The EMBO Journal. 34 (1): 20–35. doi:10.15252/embj.201489499. PMC   4291478 . PMID   25388951.
  11. Harris CJ, Scheibe M, Wongpalee SP, Liu W, Cornett EM, Vaughan RM, et al. (December 2018). "A DNA methylation reader complex that enhances gene transcription". Science. 362 (6419): 1182–1186. Bibcode:2018Sci...362.1182H. doi:10.1126/science.aar7854. PMC   6353633 . PMID   30523112.
  12. 1 2 3 4 5 6 Williams BP, Pignatta D, Henikoff S, Gehring M (March 2015). "Methylation-sensitive expression of a DNA demethylase gene serves as an epigenetic rheostat". PLOS Genetics. 11 (3): e1005142. doi: 10.1371/journal.pgen.1005142 . PMC   4380477 . PMID   25826366.
  13. 1 2 3 4 Lei M, Zhang H, Julian R, Tang K, Xie S, Zhu JK (March 2015). "Regulatory link between DNA methylation and active demethylation in Arabidopsis". Proceedings of the National Academy of Sciences of the United States of America. 112 (11): 3553–7. Bibcode:2015PNAS..112.3553L. doi: 10.1073/pnas.1502279112 . PMC   4371987 . PMID   25733903.
  14. 1 2 Penterman J, Zilberman D, Huh JH, Ballinger T, Henikoff S, Fischer RL (April 2007). "DNA demethylation in the Arabidopsis genome". Proceedings of the National Academy of Sciences of the United States of America. 104 (16): 6752–7. Bibcode:2007PNAS..104.6752P. doi: 10.1073/pnas.0701861104 . PMC   1847597 . PMID   17409185.
  15. Cho J (2018). "Transposon-Derived Non-coding RNAs and Their Function in Plants". Frontiers in Plant Science. 9: 600. doi: 10.3389/fpls.2018.00600 . PMC   5943564 . PMID   29774045.
  16. Mirouze M, Reinders J, Bucher E, Nishimura T, Schneeberger K, Ossowski S, et al. (September 2009). "Selective epigenetic control of retrotransposition in Arabidopsis". Nature. 461 (7262): 427–30. Bibcode:2009Natur.461..427M. doi:10.1038/nature08328. PMID   19734882. S2CID   205218044.
  17. 1 2 Ito H, Gaubert H, Bucher E, Mirouze M, Vaillant I, Paszkowski J (April 2011). "An siRNA pathway prevents transgenerational retrotransposition in plants subjected to stress". Nature. 472 (7341): 115–9. Bibcode:2011Natur.472..115I. doi:10.1038/nature09861. PMID   21399627. S2CID   4426724.
  18. 1 2 Cavrak VV, Lettner N, Jamge S, Kosarewicz A, Bayer LM, Mittelsten Scheid O (January 2014). "How a retrotransposon exploits the plant's heat stress response for its activation". PLOS Genetics. 10 (1): e1004115. doi: 10.1371/journal.pgen.1004115 . PMC   3907296 . PMID   24497839.
  19. 1 2 Soppe WJ, Jacobsen SE, Alonso-Blanco C, Jackson JP, Kakutani T, Koornneef M, Peeters AJ (October 2000). "The late flowering phenotype of fwa mutants is caused by gain-of-function epigenetic alleles of a homeodomain gene". Molecular Cell. 6 (4): 791–802. doi: 10.1016/s1097-2765(05)00090-0 . PMID   11090618.
  20. 1 2 Kinoshita Y, Saze H, Kinoshita T, Miura A, Soppe WJ, Koornneef M, Kakutani T (January 2007). "Control of FWA gene silencing in Arabidopsis thaliana by SINE-related direct repeats". The Plant Journal. 49 (1): 38–45. doi:10.1111/j.1365-313X.2006.02936.x. hdl: 11858/00-001M-0000-0012-38D2-5 . PMID   17144899.
  21. Gouil Q, Baulcombe DC (December 2016). "DNA Methylation Signatures of the Plant Chromomethyltransferases". PLOS Genetics. 12 (12): e1006526. doi: 10.1371/journal.pgen.1006526 . PMC   5221884 . PMID   27997534.
  22. Grover JW, Kendall T, Baten A, Burgess D, Freeling M, King GJ, Mosher RA (May 2018). "Maternal components of RNA-directed DNA methylation are required for seed development in Brassica rapa". The Plant Journal. 94 (4): 575–582. doi: 10.1111/tpj.13910 . hdl: 10150/628261 . PMID   29569777. S2CID   4212729.
  23. 1 2 Wang G, Köhler C (February 2017). "Epigenetic processes in flowering plant reproduction". Journal of Experimental Botany. 68 (4): 797–807. doi: 10.1093/jxb/erw486 . PMID   28062591. S2CID   23237961.
  24. 1 2 3 Martinez G, Köhler C (April 2017). "Role of small RNAs in epigenetic reprogramming during plant sexual reproduction". Current Opinion in Plant Biology. 36: 22–28. Bibcode:2017COPB...36...22M. doi:10.1016/j.pbi.2016.12.006. PMID   28088028.
  25. 1 2 Olmedo-Monfil V, Durán-Figueroa N, Arteaga-Vázquez M, Demesa-Arévalo E, Autran D, Grimanelli D, et al. (March 2010). "Control of female gamete formation by a small RNA pathway in Arabidopsis". Nature. 464 (7288): 628–32. Bibcode:2010Natur.464..628O. doi:10.1038/nature08828. PMC   4613780 . PMID   20208518.
  26. Slotkin RK, Vaughn M, Borges F, Tanurdzić M, Becker JD, Feijó JA, Martienssen RA (February 2009). "Epigenetic reprogramming and small RNA silencing of transposable elements in pollen". Cell. 136 (3): 461–72. doi:10.1016/j.cell.2008.12.038. PMC   2661848 . PMID   19203581.
  27. 1 2 Martínez G, Panda K, Köhler C, Slotkin RK (March 2016). "Silencing in sperm cells is directed by RNA movement from the surrounding nurse cell". Nature Plants. 2 (4): 16030. doi:10.1038/nplants.2016.30. PMID   27249563. S2CID   24746649.
  28. Erdmann RM, Hoffmann A, Walter HK, Wagenknecht HA, Groß-Hardt R, Gehring M (September 2017). "Molecular movement in the Arabidopsis thaliana female gametophyte". Plant Reproduction. 30 (3): 141–146. doi:10.1007/s00497-017-0304-3. PMC   5599461 . PMID   28695277.
  29. 1 2 3 Siomi MC, Sato K, Pezic D, Aravin AA (April 2011). "PIWI-interacting small RNAs: the vanguard of genome defence". Nature Reviews. Molecular Cell Biology. 12 (4): 246–58. doi:10.1038/nrm3089. PMID   21427766. S2CID   5710813.
  30. 1 2 3 Ernst C, Odom DT, Kutter C (November 2017). "The emergence of piRNAs against transposon invasion to preserve mammalian genome integrity". Nature Communications. 8 (1): 1411. Bibcode:2017NatCo...8.1411E. doi:10.1038/s41467-017-01049-7. PMC   5681665 . PMID   29127279.
  31. 1 2 Kawakatsu T, Stuart T, Valdes M, Breakfield N, Schmitz RJ, Nery JR, et al. (April 2016). "Unique cell-type-specific patterns of DNA methylation in the root meristem". Nature Plants. 2 (5): 16058. doi:10.1038/nplants.2016.58. PMC   4855458 . PMID   27243651.
  32. Vu TM, Nakamura M, Calarco JP, Susaki D, Lim PQ, Kinoshita T, et al. (July 2013). "RNA-directed DNA methylation regulates parental genomic imprinting at several loci in Arabidopsis". Development. 140 (14): 2953–60. doi:10.1242/dev.092981. PMC   3879202 . PMID   23760956.
  33. Waters AJ, Bilinski P, Eichten SR, Vaughn MW, Ross-Ibarra J, Gehring M, Springer NM (November 2013). "Comprehensive analysis of imprinted genes in maize reveals allelic variation for imprinting and limited conservation with other species". Proceedings of the National Academy of Sciences of the United States of America. 110 (48): 19639–44. Bibcode:2013PNAS..11019639W. doi: 10.1073/pnas.1309182110 . PMC   3845156 . PMID   24218619.
  34. Pignatta D, Erdmann RM, Scheer E, Picard CL, Bell GW, Gehring M (July 2014). "Natural epigenetic polymorphisms lead to intraspecific variation in Arabidopsis gene imprinting". eLife. 3: e03198. doi: 10.7554/eLife.03198 . PMC   4115658 . PMID   24994762.
  35. Klosinska M, Picard CL, Gehring M (September 2016). "Conserved imprinting associated with unique epigenetic signatures in the Arabidopsis genus". Nature Plants. 2 (10): 16145. doi:10.1038/nplants.2016.145. PMC   5367468 . PMID   27643534.
  36. Hatorangan MR, Laenen B, Steige KA, Slotte T, Köhler C (August 2016). "Rapid Evolution of Genomic Imprinting in Two Species of the Brassicaceae". The Plant Cell. 28 (8): 1815–27. doi:10.1105/tpc.16.00304. PMC   5006707 . PMID   27465027.
  37. Erdmann RM, Satyaki PR, Klosinska M, Gehring M (December 2017). "A Small RNA Pathway Mediates Allelic Dosage in Endosperm". Cell Reports. 21 (12): 3364–3372. doi: 10.1016/j.celrep.2017.11.078 . PMID   29262317.
  38. Satyaki PR, Gehring M (July 2019). "Paternally Acting Canonical RNA-Directed DNA Methylation Pathway Genes Sensitize Arabidopsis Endosperm to Paternal Genome Dosage". The Plant Cell. 31 (7): 1563–1578. doi:10.1105/tpc.19.00047. PMC   6635864 . PMID   31064867.
  39. Iwasaki M, Hyvärinen L, Piskurewicz U, Lopez-Molina L (March 2019). "Non-canonical RNA-directed DNA methylation participates in maternal and environmental control of seed dormancy". eLife. 8. doi: 10.7554/eLife.37434 . PMC   6435323 . PMID   30910007.
  40. Cheng J, Niu Q, Zhang B, Chen K, Yang R, Zhu JK, et al. (December 2018). "Downregulation of RdDM during strawberry fruit ripening". Genome Biology. 19 (1): 212. doi: 10.1186/s13059-018-1587-x . PMC   6280534 . PMID   30514401.
  41. Guo X, Ma Z, Zhang Z, Cheng L, Zhang X, Li T (2017). "Small RNA-Sequencing Links Physiological Changes and RdDM Process to Vegetative-to-Floral Transition in Apple". Frontiers in Plant Science. 8: 873. doi: 10.3389/fpls.2017.00873 . PMC   5447065 . PMID   28611800.
  42. Fortes AM, Gallusci P (2017-02-06). "Plant Stress Responses and Phenotypic Plasticity in the Epigenomics Era: Perspectives on the Grapevine Scenario, a Model for Perennial Crop Plants". Frontiers in Plant Science. 8: 82. doi: 10.3389/fpls.2017.00082 . PMC   5292615 . PMID   28220131.
  43. Kumar A, Bennetzen JL (1999). "Plant retrotransposons". Annual Review of Genetics. 33: 479–532. doi:10.1146/annurev.genet.33.1.479. PMID   10690416.
  44. Ito H, Kim JM, Matsunaga W, Saze H, Matsui A, Endo TA, et al. (March 2016). "A Stress-Activated Transposon in Arabidopsis Induces Transgenerational Abscisic Acid Insensitivity". Scientific Reports. 6 (1): 23181. Bibcode:2016NatSR...623181I. doi:10.1038/srep23181. PMC   4791638 . PMID   26976262.
  45. Liu J, Feng L, Li J, He Z (2015-04-24). "Genetic and epigenetic control of plant heat responses". Frontiers in Plant Science. 6: 267. doi: 10.3389/fpls.2015.00267 . PMC   4408840 . PMID   25964789.
  46. Popova OV, Dinh HQ, Aufsatz W, Jonak C (March 2013). "The RdDM pathway is required for basal heat tolerance in Arabidopsis". Molecular Plant. 6 (2): 396–410. doi:10.1093/mp/sst023. PMC   3603006 . PMID   23376771.
  47. Tricker PJ, Gibbings JG, Rodríguez López CM, Hadley P, Wilkinson MJ (June 2012). "Low relative humidity triggers RNA-directed de novo DNA methylation and suppression of genes controlling stomatal development". Journal of Experimental Botany. 63 (10): 3799–813. doi:10.1093/jxb/ers076. PMC   3733579 . PMID   22442411.
  48. Xu R, Wang Y, Zheng H, Lu W, Wu C, Huang J, et al. (September 2015). "Salt-induced transcription factor MYB74 is regulated by the RNA-directed DNA methylation pathway in Arabidopsis". Journal of Experimental Botany. 66 (19): 5997–6008. doi:10.1093/jxb/erv312. PMC   4566987 . PMID   26139822.
  49. 1 2 3 4 Wassenegger M, Heimes S, Riedel L, Sänger HL (February 1994). "RNA-directed de novo methylation of genomic sequences in plants". Cell. 76 (3): 567–76. doi:10.1016/0092-8674(94)90119-8. PMID   8313476. S2CID   35858018.
  50. 1 2 3 4 Huang J, Yang M, Zhang X (April 2016). "The function of small RNAs in plant biotic stress response". Journal of Integrative Plant Biology. 58 (4): 312–27. doi: 10.1111/jipb.12463 . PMID   26748943.
  51. Raja P, Jackel JN, Li S, Heard IM, Bisaro DM (March 2014). "Arabidopsis double-stranded RNA binding protein DRB3 participates in methylation-mediated defense against geminiviruses". Journal of Virology. 88 (5): 2611–22. doi:10.1128/JVI.02305-13. PMC   3958096 . PMID   24352449.
  52. Jackel JN, Storer JM, Coursey T, Bisaro DM (August 2016). Simon A (ed.). "Arabidopsis RNA Polymerases IV and V Are Required To Establish H3K9 Methylation, but Not Cytosine Methylation, on Geminivirus Chromatin". Journal of Virology. 90 (16): 7529–7540. doi:10.1128/JVI.00656-16. PMC   4984644 . PMID   27279611.
  53. 1 2 Diezma-Navas L, Pérez-González A, Artaza H, Alonso L, Caro E, Llave C, Ruiz-Ferrer V (October 2019). "Crosstalk between epigenetic silencing and infection bytobacco rattle virus in Arabidopsis". Molecular Plant Pathology. 20 (10): 1439–1452. doi:10.1111/mpp.12850. PMC   6792132 . PMID   31274236.
  54. Calil IP, Fontes EP (March 2017). "Plant immunity against viruses: antiviral immune receptors in focus". Annals of Botany. 119 (5): 711–723. doi:10.1093/aob/mcw200. PMC   5604577 . PMID   27780814.
  55. 1 2 3 4 5 6 7 8 9 Matzke MA, Mosher RA (June 2014). "RNA-directed DNA methylation: an epigenetic pathway of increasing complexity". Nature Reviews. Genetics. 15 (6): 394–408. doi:10.1038/nrg3683. PMID   24805120. S2CID   54489227.
  56. Wang MB, Masuta C, Smith NA, Shimura H (October 2012). "RNA silencing and plant viral diseases". Molecular Plant-Microbe Interactions. 25 (10): 1275–85. doi: 10.1094/MPMI-04-12-0093-CR . PMID   22670757.
  57. Wang Y, Wu Y, Gong Q, Ismayil A, Yuan Y, Lian B, et al. (March 2019). Simon AE (ed.). "Geminiviral V2 Protein Suppresses Transcriptional Gene Silencing through Interaction with AGO4". Journal of Virology. 93 (6): e01675–18, /jvi/93/6/JVI.01675–18.atom. doi:10.1128/JVI.01675-18. PMC   6401443 . PMID   30626668.
  58. 1 2 Dowen RH, Pelizzola M, Schmitz RJ, Lister R, Dowen JM, Nery JR, et al. (August 2012). "Widespread dynamic DNA methylation in response to biotic stress". Proceedings of the National Academy of Sciences of the United States of America. 109 (32): E2183-91. doi: 10.1073/pnas.1209329109 . PMC   3420206 . PMID   22733782.
  59. López A, Ramírez V, García-Andrade J, Flors V, Vera P (December 2011). Pikaard CS (ed.). "The RNA silencing enzyme RNA polymerase v is required for plant immunity". PLOS Genetics. 7 (12): e1002434. doi: 10.1371/journal.pgen.1002434 . PMC   3248562 . PMID   22242006.
  60. 1 2 Rasmann S, De Vos M, Casteel CL, Tian D, Halitschke R, Sun JY, et al. (February 2012). "Herbivory in the previous generation primes plants for enhanced insect resistance". Plant Physiology. 158 (2): 854–63. doi:10.1104/pp.111.187831. PMC   3271773 . PMID   22209873.
  61. Gohlke J, Scholz CJ, Kneitz S, Weber D, Fuchs J, Hedrich R, Deeken R (2013-02-07). McDowell JM (ed.). "DNA methylation mediated control of gene expression is critical for development of crown gall tumors". PLOS Genetics. 9 (2): e1003267. doi: 10.1371/journal.pgen.1003267 . PMC   3567176 . PMID   23408907.
  62. Espinas NA, Saze H, Saijo Y (2016-08-11). "Epigenetic Control of Defense Signaling and Priming in Plants". Frontiers in Plant Science. 7: 1201. doi: 10.3389/fpls.2016.01201 . PMC   4980392 . PMID   27563304.
  63. Aufsatz W, Mette MF, van der Winden J, Matzke AJ, Matzke M (December 2002). "RNA-directed DNA methylation in Arabidopsis". Proceedings of the National Academy of Sciences of the United States of America. 99 Suppl 4 (Supplement 4): 16499–506. Bibcode:2002PNAS...9916499A. doi: 10.1073/pnas.162371499 . PMC   139914 . PMID   12169664.
  64. 1 2 Matzke MA, Primig M, Trnovsky J, Matzke AJ (March 1989). "Reversible methylation and inactivation of marker genes in sequentially transformed tobacco plants". The EMBO Journal. 8 (3): 643–9. doi:10.1002/j.1460-2075.1989.tb03421.x. PMC   400855 . PMID   16453872.
  65. Gutzat R, Mittelsten Scheid O (November 2012). "Epigenetic responses to stress: triple defense?". Current Opinion in Plant Biology. 15 (5): 568–73. Bibcode:2012COPB...15..568G. doi:10.1016/j.pbi.2012.08.007. PMC   3508409 . PMID   22960026.
  66. Boyko A, Kovalchuk I (August 2010). Shiu SH (ed.). "Transgenerational response to stress in Arabidopsis thaliana". Plant Signaling & Behavior. 5 (8): 995–8. Bibcode:2010PlSiB...5..995B. doi:10.4161/psb.5.8.12227. PMC   3115178 . PMID   20724818.
  67. Mermigka G, Verret F, Kalantidis K (April 2016). "RNA silencing movement in plants". Journal of Integrative Plant Biology. 58 (4): 328–42. doi: 10.1111/jipb.12423 . PMID   26297506.
  68. Lewsey MG, Hardcastle TJ, Melnyk CW, Molnar A, Valli A, Urich MA, et al. (February 2016). "Mobile small RNAs regulate genome-wide DNA methylation". Proceedings of the National Academy of Sciences of the United States of America. 113 (6): E801-10. Bibcode:2016PNAS..113E.801L. doi: 10.1073/pnas.1515072113 . PMC   4760824 . PMID   26787884.
  69. 1 2 Tamiru M, Hardcastle TJ, Lewsey MG (January 2018). "Regulation of genome-wide DNA methylation by mobile small RNAs". The New Phytologist. 217 (2): 540–546. doi: 10.1111/nph.14874 . PMID   29105762.
  70. 1 2 3 Molnar A, Melnyk CW, Bassett A, Hardcastle TJ, Dunn R, Baulcombe DC (May 2010). "Small silencing RNAs in plants are mobile and direct epigenetic modification in recipient cells". Science. 328 (5980): 872–5. Bibcode:2010Sci...328..872M. doi: 10.1126/science.1187959 . PMID   20413459. S2CID   206525853.
  71. Bai S, Kasai A, Yamada K, Li T, Harada T (August 2011). "A mobile signal transported over a long distance induces systemic transcriptional gene silencing in a grafted partner". Journal of Experimental Botany. 62 (13): 4561–70. doi:10.1093/jxb/err163. PMC   3170550 . PMID   21652532.
  72. Zhang W, Kollwig G, Stecyk E, Apelt F, Dirks R, Kragler F (October 2014). "Graft-transmissible movement of inverted-repeat-induced siRNA signals into flowers". The Plant Journal. 80 (1): 106–21. doi: 10.1111/tpj.12622 . PMID   25039964.
  73. Parent JS, Martínez de Alba AE, Vaucheret H (2012). "The origin and effect of small RNA signaling in plants". Frontiers in Plant Science. 3: 179. doi: 10.3389/fpls.2012.00179 . PMC   3414853 . PMID   22908024.
  74. 1 2 Stroud H, Do T, Du J, Zhong X, Feng S, Johnson L, et al. (January 2014). "Non-CG methylation patterns shape the epigenetic landscape in Arabidopsis". Nature Structural & Molecular Biology. 21 (1): 64–72. doi:10.1038/nsmb.2735. PMC   4103798 . PMID   24336224.
  75. 1 2 Bewick AJ, Niederhuth CE, Ji L, Rohr NA, Griffin PT, Leebens-Mack J, Schmitz RJ (May 2017). "The evolution of CHROMOMETHYLASES and gene body DNA methylation in plants". Genome Biology. 18 (1): 65. doi: 10.1186/s13059-017-1195-1 . PMC   5410703 . PMID   28457232.
  76. Bartels A, Han Q, Nair P, Stacey L, Gaynier H, Mosley M, et al. (July 2018). "Dynamic DNA Methylation in Plant Growth and Development". International Journal of Molecular Sciences. 19 (7): 2144. doi: 10.3390/ijms19072144 . PMC   6073778 . PMID   30041459.
  77. Wendte JM, Schmitz RJ (March 2018). "Specifications of Targeting Heterochromatin Modifications in Plants". Molecular Plant. 11 (3): 381–387. doi: 10.1016/j.molp.2017.10.002 . PMID   29032247.
  78. Law JA, Jacobsen SE (March 2010). "Establishing, maintaining and modifying DNA methylation patterns in plants and animals". Nature Reviews. Genetics. 11 (3): 204–20. doi:10.1038/nrg2719. PMC   3034103 . PMID   20142834.
  79. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 Cuerda-Gil D, Slotkin RK (November 2016). "Non-canonical RNA-directed DNA methylation". Nature Plants. 2 (11): 16163. doi:10.1038/nplants.2016.163. PMID   27808230. S2CID   4248951.
  80. 1 2 3 4 5 6 7 8 9 10 11 12 13 Matzke MA, Kanno T, Matzke AJ (2015). "RNA-Directed DNA Methylation: The Evolution of a Complex Epigenetic Pathway in Flowering Plants". Annual Review of Plant Biology. 66: 243–67. doi: 10.1146/annurev-arplant-043014-114633 . PMID   25494460.
  81. 1 2 3 4 Wendte JM, Pikaard CS (January 2017). "The RNAs of RNA-directed DNA methylation". Biochimica et Biophysica Acta (BBA) - Gene Regulatory Mechanisms. 1860 (1): 140–148. doi:10.1016/j.bbagrm.2016.08.004. PMC   5203809 . PMID   27521981.
  82. Zhai J, Bischof S, Wang H, Feng S, Lee TF, Teng C, et al. (October 2015). "A One Precursor One siRNA Model for Pol IV-Dependent siRNA Biogenesis". Cell. 163 (2): 445–55. doi:10.1016/j.cell.2015.09.032. PMC   5023148 . PMID   26451488.
  83. 1 2 3 Blevins T, Podicheti R, Mishra V, Marasco M, Wang J, Rusch D, et al. (October 2015). "Identification of Pol IV and RDR2-dependent precursors of 24 nt siRNAs guiding de novo DNA methylation in Arabidopsis". eLife. 4: e09591. doi: 10.7554/eLife.09591 . PMC   4716838 . PMID   26430765.
  84. 1 2 Singh J, Mishra V, Wang F, Huang HY, Pikaard CS (August 2019). "Reaction Mechanisms of Pol IV, RDR2, and DCL3 Drive RNA Channeling in the siRNA-Directed DNA Methylation Pathway". Molecular Cell. 75 (3): 576–589.e5. doi:10.1016/j.molcel.2019.07.008. PMC   6698059 . PMID   31398324.
  85. Panda K, Ji L, Neumann DA, Daron J, Schmitz RJ, Slotkin RK (August 2016). "Full-length autonomous transposable elements are preferentially targeted by expression-dependent forms of RNA-directed DNA methylation". Genome Biology. 17 (1): 170. doi: 10.1186/s13059-016-1032-y . PMC   4977677 . PMID   27506905.
  86. Zhang Z, Liu X, Guo X, Wang XJ, Zhang X (April 2016). "Arabidopsis AGO3 predominantly recruits 24-nt small RNAs to regulate epigenetic silencing". Nature Plants. 2 (5): 16049. doi:10.1038/nplants.2016.49. PMID   27243648. S2CID   8933827.
  87. Meister G (July 2013). "Argonaute proteins: functional insights and emerging roles". Nature Reviews. Genetics. 14 (7): 447–59. doi:10.1038/nrg3462. PMID   23732335. S2CID   5210500.
  88. 1 2 3 Wierzbicki AT, Haag JR, Pikaard CS (November 2008). "Noncoding transcription by RNA polymerase Pol IVb/Pol V mediates transcriptional silencing of overlapping and adjacent genes". Cell. 135 (4): 635–48. doi:10.1016/j.cell.2008.09.035. PMC   2602798 . PMID   19013275.
  89. Cao X, Jacobsen SE (July 2002). "Role of the arabidopsis DRM methyltransferases in de novo DNA methylation and gene silencing". Current Biology. 12 (13): 1138–44. Bibcode:2002CBio...12.1138C. doi: 10.1016/s0960-9822(02)00925-9 . PMID   12121623. S2CID   15695949.
  90. 1 2 3 Gallego-Bartolomé J, Liu W, Kuo PH, Feng S, Ghoshal B, Gardiner J, et al. (February 2019). "Co-targeting RNA Polymerases IV and V Promotes Efficient De Novo DNA Methylation in Arabidopsis". Cell. 176 (5): 1068–1082.e19. doi:10.1016/j.cell.2019.01.029. PMC   6386582 . PMID   30739798.
  91. 1 2 Voinnet O (July 2008). "Use, tolerance and avoidance of amplified RNA silencing by plants". Trends in Plant Science. 13 (7): 317–28. doi:10.1016/j.tplants.2008.05.004. PMID   18565786.
  92. 1 2 Pontier D, Picart C, Roudier F, Garcia D, Lahmy S, Azevedo J, et al. (October 2012). "NERD, a plant-specific GW protein, defines an additional RNAi-dependent chromatin-based pathway in Arabidopsis". Molecular Cell. 48 (1): 121–32. doi: 10.1016/j.molcel.2012.07.027 . PMID   22940247.
  93. 1 2 3 4 Haag JR, Pikaard CS (July 2011). "Multisubunit RNA polymerases IV and V: purveyors of non-coding RNA for plant gene silencing". Nature Reviews. Molecular Cell Biology. 12 (8): 483–92. doi:10.1038/nrm3152. PMID   21779025. S2CID   9970159.
  94. 1 2 3 Zhou M, Law JA (October 2015). "RNA Pol IV and V in gene silencing: Rebel polymerases evolving away from Pol II's rules". Current Opinion in Plant Biology. 27: 154–64. Bibcode:2015COPB...27..154Z. doi:10.1016/j.pbi.2015.07.005. PMC   4618083 . PMID   26344361.
  95. 1 2 Lahmy S, Pontier D, Bies-Etheve N, Laudié M, Feng S, Jobet E, et al. (December 2016). "Evidence for ARGONAUTE4-DNA interactions in RNA-directed DNA methylation in plants". Genes & Development. 30 (23): 2565–2570. doi:10.1101/gad.289553.116. PMC   5204349 . PMID   27986858.
  96. 1 2 3 4 Henderson IR, Zhang X, Lu C, Johnson L, Meyers BC, Green PJ, Jacobsen SE (June 2006). "Dissecting Arabidopsis thaliana DICER function in small RNA processing, gene silencing and DNA methylation patterning". Nature Genetics. 38 (6): 721–5. doi:10.1038/ng1804. PMID   16699516. S2CID   10261689.
  97. 1 2 3 4 5 Bologna NG, Voinnet O (2014). "The diversity, biogenesis, and activities of endogenous silencing small RNAs in Arabidopsis". Annual Review of Plant Biology. 65: 473–503. doi:10.1146/annurev-arplant-050213-035728. PMID   24579988.
  98. Wang J, Mei J, Ren G (2019). "Plant microRNAs: Biogenesis, Homeostasis, and Degradation". Frontiers in Plant Science. 10: 360. doi: 10.3389/fpls.2019.00360 . PMC   6445950 . PMID   30972093.
  99. 1 2 3 4 5 6 7 8 9 10 Stroud H, Greenberg MV, Feng S, Bernatavichute YV, Jacobsen SE (January 2013). "Comprehensive analysis of silencing mutants reveals complex regulation of the Arabidopsis methylome". Cell. 152 (1–2): 352–64. doi:10.1016/j.cell.2012.10.054. PMC   3597350 . PMID   23313553.
  100. 1 2 3 4 Fang X, Qi Y (February 2016). "RNAi in Plants: An Argonaute-Centered View". The Plant Cell. 28 (2): 272–85. doi:10.1105/tpc.15.00920. PMC   4790879 . PMID   26869699.
  101. Eun C, Lorkovic ZJ, Naumann U, Long Q, Havecker ER, Simon SA, et al. (2011). "AGO6 functions in RNA-mediated transcriptional gene silencing in shoot and root meristems in Arabidopsis thaliana". PLOS ONE. 6 (10): e25730. Bibcode:2011PLoSO...625730E. doi: 10.1371/journal.pone.0025730 . PMC   3187791 . PMID   21998686.
  102. Durán-Figueroa N, Vielle-Calzada JP (November 2010). "ARGONAUTE9-dependent silencing of transposable elements in pericentromeric regions of Arabidopsis". Plant Signaling & Behavior. 5 (11): 1476–9. Bibcode:2010PlSiB...5.1476D. doi:10.4161/psb.5.11.13548. PMC   3115260 . PMID   21057207.
  103. Cao X, Aufsatz W, Zilberman D, Mette MF, Huang MS, Matzke M, Jacobsen SE (December 2003). "Role of the DRM and CMT3 methyltransferases in RNA-directed DNA methylation". Current Biology. 13 (24): 2212–7. Bibcode:2003CBio...13.2212C. doi: 10.1016/j.cub.2003.11.052 . PMID   14680640. S2CID   8232599.
  104. Law JA, Vashisht AA, Wohlschlegel JA, Jacobsen SE (July 2011). "SHH1, a homeodomain protein required for DNA methylation, as well as RDR2, RDM4, and chromatin remodeling factors, associate with RNA polymerase IV". PLOS Genetics. 7 (7): e1002195. doi: 10.1371/journal.pgen.1002195 . PMC   3141008 . PMID   21811420.
  105. Zhang H, Ma ZY, Zeng L, Tanaka K, Zhang CJ, Ma J, et al. (May 2013). "DTF1 is a core component of RNA-directed DNA methylation and may assist in the recruitment of Pol IV". Proceedings of the National Academy of Sciences of the United States of America. 110 (20): 8290–5. Bibcode:2013PNAS..110.8290Z. doi: 10.1073/pnas.1300585110 . PMC   3657815 . PMID   23637343.
  106. 1 2 Law JA, Du J, Hale CJ, Feng S, Krajewski K, Palanca AM, et al. (June 2013). "Polymerase IV occupancy at RNA-directed DNA methylation sites requires SHH1". Nature. 498 (7454): 385–9. Bibcode:2013Natur.498..385L. doi:10.1038/nature12178. PMC   4119789 . PMID   23636332.
  107. 1 2 3 Zhou M, Palanca AM, Law JA (June 2018). "Locus-specific control of the de novo DNA methylation pathway in Arabidopsis by the CLASSY family". Nature Genetics. 50 (6): 865–873. doi:10.1038/s41588-018-0115-y. PMC   6317521 . PMID   29736015.
  108. 1 2 Yang DL, Zhang G, Wang L, Li J, Xu D, Di C, et al. (2018). "Four putative SWI2/SNF2 chromatin remodelers have dual roles in regulating DNA methylation in Arabidopsis". Cell Discovery. 4: 55. doi:10.1038/s41421-018-0056-8. PMC   6189096 . PMID   30345072.
  109. Ji L, Chen X (April 2012). "Regulation of small RNA stability: methylation and beyond". Cell Research. 22 (4): 624–36. doi:10.1038/cr.2012.36. PMC   3317568 . PMID   22410795.
  110. 1 2 3 4 Liu ZW, Shao CR, Zhang CJ, Zhou JX, Zhang SW, Li L, et al. (January 2014). "The SET domain proteins SUVH2 and SUVH9 are required for Pol V occupancy at RNA-directed DNA methylation loci". PLOS Genetics. 10 (1): e1003948. doi: 10.1371/journal.pgen.1003948 . PMC   3898904 . PMID   24465213.
  111. Wierzbicki AT, Ream TS, Haag JR, Pikaard CS (May 2009). "RNA polymerase V transcription guides ARGONAUTE4 to chromatin". Nature Genetics. 41 (5): 630–4. doi:10.1038/ng.365. PMC   2674513 . PMID   19377477.
  112. Zhong X, Hale CJ, Law JA, Johnson LM, Feng S, Tu A, Jacobsen SE (September 2012). "DDR complex facilitates global association of RNA polymerase V to promoters and evolutionarily young transposons". Nature Structural & Molecular Biology. 19 (9): 870–5. doi:10.1038/nsmb.2354. PMC   3443314 . PMID   22864289.
  113. Pikaard CS, Haag JR, Pontes OM, Blevins T, Cocklin R (2012). "A transcription fork model for Pol IV and Pol V-dependent RNA-directed DNA methylation". Cold Spring Harbor Symposia on Quantitative Biology. 77: 205–12. doi: 10.1101/sqb.2013.77.014803 . PMID   23567894.
  114. He XJ, Hsu YF, Zhu S, Wierzbicki AT, Pontes O, Pikaard CS, et al. (May 2009). "An effector of RNA-directed DNA methylation in arabidopsis is an ARGONAUTE 4- and RNA-binding protein". Cell. 137 (3): 498–508. doi:10.1016/j.cell.2009.04.028. PMC   2700824 . PMID   19410546.
  115. Liu W, Duttke SH, Hetzel J, Groth M, Feng S, Gallego-Bartolome J, et al. (March 2018). "RNA-directed DNA methylation involves co-transcriptional small-RNA-guided slicing of polymerase V transcripts in Arabidopsis". Nature Plants. 4 (3): 181–188. doi:10.1038/s41477-017-0100-y. PMC   5832601 . PMID   29379150.
  116. 1 2 Zhu Y, Rowley MJ, Böhmdorfer G, Wierzbicki AT (January 2013). "A SWI/SNF chromatin-remodeling complex acts in noncoding RNA-mediated transcriptional silencing". Molecular Cell. 49 (2): 298–309. doi:10.1016/j.molcel.2012.11.011. PMC   3560041 . PMID   23246435.
  117. Ausin I, Mockler TC, Chory J, Jacobsen SE (December 2009). "IDN1 and IDN2 are required for de novo DNA methylation in Arabidopsis thaliana". Nature Structural & Molecular Biology. 16 (12): 1325–7. doi:10.1038/nsmb.1690. PMC   2842998 . PMID   19915591.
  118. Xie M, Ren G, Zhang C, Yu B (November 2012). "The DNA- and RNA-binding protein FACTOR of DNA METHYLATION 1 requires XH domain-mediated complex formation for its function in RNA-directed DNA methylation". The Plant Journal. 72 (3): 491–500. doi: 10.1111/j.1365-313X.2012.05092.x . PMID   22757778.
  119. Jullien PE, Susaki D, Yelagandula R, Higashiyama T, Berger F (October 2012). "DNA methylation dynamics during sexual reproduction in Arabidopsis thaliana". Current Biology. 22 (19): 1825–30. Bibcode:2012CBio...22.1825J. doi: 10.1016/j.cub.2012.07.061 . PMID   22940470. S2CID   18586419.
  120. 1 2 3 4 Blevins T, Pontvianne F, Cocklin R, Podicheti R, Chandrasekhara C, Yerneni S, et al. (April 2014). "A two-step process for epigenetic inheritance in Arabidopsis". Molecular Cell. 54 (1): 30–42. doi:10.1016/j.molcel.2014.02.019. PMC   3988221 . PMID   24657166.
  121. Peters AH, Kubicek S, Mechtler K, O'Sullivan RJ, Derijck AA, Perez-Burgos L, et al. (December 2003). "Partitioning and plasticity of repressive histone methylation states in mammalian chromatin". Molecular Cell. 12 (6): 1577–89. doi: 10.1016/s1097-2765(03)00477-5 . PMID   14690609.
  122. Jackson JP, Johnson L, Jasencakova Z, Zhang X, PerezBurgos L, Singh PB, et al. (March 2004). "Dimethylation of histone H3 lysine 9 is a critical mark for DNA methylation and gene silencing in Arabidopsis thaliana". Chromosoma. 112 (6): 308–15. doi:10.1007/s00412-004-0275-7. PMID   15014946. S2CID   17798608.
  123. 1 2 3 Du J, Johnson LM, Jacobsen SE, Patel DJ (September 2015). "DNA methylation pathways and their crosstalk with histone methylation". Nature Reviews. Molecular Cell Biology. 16 (9): 519–32. doi:10.1038/nrm4043. PMC   4672940 . PMID   26296162.
  124. Li X, Harris CJ, Zhong Z, Chen W, Liu R, Jia B, et al. (September 2018). "Mechanistic insights into plant SUVH family H3K9 methyltransferases and their binding to context-biased non-CG DNA methylation". Proceedings of the National Academy of Sciences of the United States of America. 115 (37): E8793–E8802. Bibcode:2018PNAS..115E8793L. doi: 10.1073/pnas.1809841115 . PMC   6140468 . PMID   30150382.
  125. Du J, Zhong X, Bernatavichute YV, Stroud H, Feng S, Caro E, et al. (September 2012). "Dual binding of chromomethylase domains to H3K9me2-containing nucleosomes directs DNA methylation in plants". Cell. 151 (1): 167–80. doi:10.1016/j.cell.2012.07.034. PMC   3471781 . PMID   23021223.
  126. 1 2 Lachner M, O'Carroll D, Rea S, Mechtler K, Jenuwein T (March 2001). "Methylation of histone H3 lysine 9 creates a binding site for HP1 proteins". Nature. 410 (6824): 116–20. Bibcode:2001Natur.410..116L. doi:10.1038/35065132. PMID   11242053. S2CID   4331863.
  127. Mylne JS, Barrett L, Tessadori F, Mesnage S, Johnson L, Bernatavichute YV, et al. (March 2006). "LHP1, the Arabidopsis homologue of HETEROCHROMATIN PROTEIN1, is required for epigenetic silencing of FLC". Proceedings of the National Academy of Sciences of the United States of America. 103 (13): 5012–7. Bibcode:2006PNAS..103.5012M. doi: 10.1073/pnas.0507427103 . PMC   1458786 . PMID   16549797.
  128. Zhao S, Cheng L, Gao Y, Zhang B, Zheng X, Wang L, et al. (January 2019). "Plant HP1 protein ADCP1 links multivalent H3K9 methylation readout to heterochromatin formation". Cell Research. 29 (1): 54–66. doi:10.1038/s41422-018-0104-9. PMC   6318295 . PMID   30425322.
  129. Klemm SL, Shipony Z, Greenleaf WJ (April 2019). "Chromatin accessibility and the regulatory epigenome". Nature Reviews. Genetics. 20 (4): 207–220. doi:10.1038/s41576-018-0089-8. PMID   30675018. S2CID   59159906.
  130. Vongs A, Kakutani T, Martienssen RA, Richards EJ (June 1993). "Arabidopsis thaliana DNA methylation mutants". Science. 260 (5116): 1926–8. Bibcode:1993Sci...260.1926V. doi:10.1126/science.8316832. PMID   8316832.
  131. 1 2 Jeddeloh JA, Stokes TL, Richards EJ (May 1999). "Maintenance of genomic methylation requires a SWI2/SNF2-like protein". Nature Genetics. 22 (1): 94–7. doi:10.1038/8803. PMID   10319870. S2CID   20199014.
  132. Kankel MW, Ramsey DE, Stokes TL, Flowers SK, Haag JR, Jeddeloh JA, et al. (March 2003). "Arabidopsis MET1 cytosine methyltransferase mutants". Genetics. 163 (3): 1109–22. doi:10.1093/genetics/163.3.1109. PMC   1462485 . PMID   12663548.
  133. Jones L, Ratcliff F, Baulcombe DC (May 2001). "RNA-directed transcriptional gene silencing in plants can be inherited independently of the RNA trigger and requires Met1 for maintenance". Current Biology. 11 (10): 747–57. Bibcode:2001CBio...11..747J. doi: 10.1016/s0960-9822(01)00226-3 . PMID   11378384. S2CID   16789197.
  134. Chan SW, Henderson IR, Jacobsen SE (May 2005). "Gardening the genome: DNA methylation in Arabidopsis thaliana". Nature Reviews. Genetics. 6 (5): 351–60. doi:10.1038/nrg1601. PMID   15861207. S2CID   20083628.
  135. Li Y, Kumar S, Qian W (January 2018). "Active DNA demethylation: mechanism and role in plant development". Plant Cell Reports. 37 (1): 77–85. doi:10.1007/s00299-017-2215-z. PMC   5758694 . PMID   29026973.
  136. Choi Y, Gehring M, Johnson L, Hannon M, Harada JJ, Goldberg RB, et al. (July 2002). "DEMETER, a DNA glycosylase domain protein, is required for endosperm gene imprinting and seed viability in arabidopsis". Cell. 110 (1): 33–42. doi: 10.1016/s0092-8674(02)00807-3 . PMID   12150995. S2CID   14828646.
  137. Zhu J, Kapoor A, Sridhar VV, Agius F, Zhu JK (January 2007). "The DNA glycosylase/lyase ROS1 functions in pruning DNA methylation patterns in Arabidopsis". Current Biology. 17 (1): 54–9. Bibcode:2007CBio...17...54Z. doi: 10.1016/j.cub.2006.10.059 . PMID   17208187. S2CID   3955783.
  138. Williams BP, Gehring M (December 2017). "Stable transgenerational epigenetic inheritance requires a DNA methylation-sensing circuit". Nature Communications. 8 (1): 2124. Bibcode:2017NatCo...8.2124W. doi:10.1038/s41467-017-02219-3. PMC   5730562 . PMID   29242626.
  139. Wang J, Blevins T, Podicheti R, Haag JR, Tan EH, Wang F, Pikaard CS (August 2017). "Arabidopsis SMC4 identifies condensin as a corepressor of pericentromeric transposons and conditionally expressed genes". Genes & Development. 31 (15): 1601–1614. doi:10.1101/gad.301499.117. PMC   5630024 . PMID   28882854.
  140. Córdoba-Cañero D, Cognat V, Ariza RR, Roldán Arjona T, Molinier J (December 2017). "Dual control of ROS1-mediated active DNA demethylation by DNA damage-binding protein 2 (DDB2)". The Plant Journal. 92 (6): 1170–1181. doi: 10.1111/tpj.13753 . PMID   29078035. S2CID   37919309.
  141. 1 2 3 4 Ream TS, Haag JR, Wierzbicki AT, Nicora CD, Norbeck AD, Zhu JK, et al. (January 2009). "Subunit compositions of the RNA-silencing enzymes Pol IV and Pol V reveal their origins as specialized forms of RNA polymerase II". Molecular Cell. 33 (2): 192–203. doi:10.1016/j.molcel.2008.12.015. PMC   2946823 . PMID   19110459.
  142. 1 2 3 Huang Y, Kendall T, Forsythe ES, Dorantes-Acosta A, Li S, Caballero-Pérez J, et al. (July 2015). "Ancient Origin and Recent Innovations of RNA Polymerase IV and V". Molecular Biology and Evolution. 32 (7): 1788–99. doi:10.1093/molbev/msv060. PMC   4476159 . PMID   25767205.
  143. Tucker SL, Reece J, Ream TS, Pikaard CS (2010). "Evolutionary history of plant multisubunit RNA polymerases IV and V: subunit origins via genome-wide and segmental gene duplications, retrotransposition, and lineage-specific subfunctionalization". Cold Spring Harbor Symposia on Quantitative Biology. 75: 285–97. doi: 10.1101/sqb.2010.75.037 . PMID   21447813.
  144. Luo J, Hall BD (January 2007). "A multistep process gave rise to RNA polymerase IV of land plants". Journal of Molecular Evolution. 64 (1): 101–12. Bibcode:2007JMolE..64..101L. doi:10.1007/s00239-006-0093-z. PMID   17160640. S2CID   37590716.
  145. 1 2 Haag JR, Brower-Toland B, Krieger EK, Sidorenko L, Nicora CD, Norbeck AD, et al. (October 2014). "Functional diversification of maize RNA polymerase IV and V subtypes via alternative catalytic subunits". Cell Reports. 9 (1): 378–390. doi:10.1016/j.celrep.2014.08.067. PMC   4196699 . PMID   25284785.
  146. Ma L, Hatlen A, Kelly LJ, Becher H, Wang W, Kovarik A, et al. (September 2015). "Angiosperms Are Unique among Land Plant Lineages in the Occurrence of Key Genes in the RNA-Directed DNA Methylation (RdDM) Pathway". Genome Biology and Evolution. 7 (9): 2648–62. doi:10.1093/gbe/evv171. PMC   4607528 . PMID   26338185.
  147. Yaari R, Katz A, Domb K, Harris KD, Zemach A, Ohad N (April 2019). "RdDM-independent de novo and heterochromatin DNA methylation by plant CMT and DNMT3 orthologs". Nature Communications. 10 (1): 1613. Bibcode:2019NatCo..10.1613Y. doi:10.1038/s41467-019-09496-0. PMC   6453930 . PMID   30962443.
  148. 1 2 Moran Y, Agron M, Praher D, Technau U (February 2017). "The evolutionary origin of plant and animal microRNAs". Nature Ecology & Evolution. 1 (3): 27. Bibcode:2017NatEE...1...27M. doi:10.1038/s41559-016-0027. PMC   5435108 . PMID   28529980.
  149. Castel SE, Martienssen RA (February 2013). "RNA interference in the nucleus: roles for small RNAs in transcription, epigenetics and beyond". Nature Reviews. Genetics. 14 (2): 100–12. doi:10.1038/nrg3355. PMC   4205957 . PMID   23329111.
  150. Volpe TA, Kidner C, Hall IM, Teng G, Grewal SI, Martienssen RA (September 2002). "Regulation of heterochromatic silencing and histone H3 lysine-9 methylation by RNAi". Science. 297 (5588): 1833–7. Bibcode:2002Sci...297.1833V. doi: 10.1126/science.1074973 . PMID   12193640. S2CID   2613813.
  151. Bühler M, Verdel A, Moazed D (June 2006). "Tethering RITS to a nascent transcript initiates RNAi- and heterochromatin-dependent gene silencing". Cell. 125 (5): 873–86. doi: 10.1016/j.cell.2006.04.025 . PMID   16751098. S2CID   2938057.
  152. Zaratiegui M, Castel SE, Irvine DV, Kloc A, Ren J, Li F, et al. (October 2011). "RNAi promotes heterochromatic silencing through replication-coupled release of RNA Pol II". Nature. 479 (7371): 135–8. Bibcode:2011Natur.479..135Z. doi:10.1038/nature10501. PMC   3391703 . PMID   22002604.
  153. Fagard M, Vaucheret H (June 2000). "(TRANS)GENE SILENCING IN PLANTS: How Many Mechanisms?". Annual Review of Plant Physiology and Plant Molecular Biology. 51 (1): 167–194. doi:10.1146/annurev.arplant.51.1.167. PMID   15012190.
  154. Napoli C, Lemieux C, Jorgensen R (April 1990). "Introduction of a Chimeric Chalcone Synthase Gene into Petunia Results in Reversible Co-Suppression of Homologous Genes in trans". The Plant Cell. 2 (4): 279–289. doi:10.1105/tpc.2.4.279. PMC   159885 . PMID   12354959.
  155. van der Krol AR, Mur LA, Beld M, Mol JN, Stuitje AR (April 1990). "Flavonoid genes in petunia: addition of a limited number of gene copies may lead to a suppression of gene expression". The Plant Cell. 2 (4): 291–9. doi:10.1105/tpc.2.4.291. PMC   159886 . PMID   2152117.
  156. Depicker A, Montagu MV (June 1997). "Post-transcriptional gene silencing in plants". Current Opinion in Cell Biology. 9 (3): 373–82. doi:10.1016/s0955-0674(97)80010-5. PMID   9159078.
  157. Assaad FF, Tucker KL, Signer ER (September 1993). "Epigenetic repeat-induced gene silencing (RIGS) in Arabidopsis". Plant Molecular Biology. 22 (6): 1067–85. doi:10.1007/BF00028978. PMID   8400126. S2CID   26576784.
  158. Ingelbrecht I, Van Houdt H, Van Montagu M, Depicker A (October 1994). "Posttranscriptional silencing of reporter transgenes in tobacco correlates with DNA methylation". Proceedings of the National Academy of Sciences of the United States of America. 91 (22): 10502–6. Bibcode:1994PNAS...9110502I. doi: 10.1073/pnas.91.22.10502 . PMC   45049 . PMID   7937983.
  159. Meyer P, Heidmann I (May 1994). "Epigenetic variants of a transgenic petunia line show hypermethylation in transgene DNA: an indication for specific recognition of foreign DNA in transgenic plants". Molecular & General Genetics. 243 (4): 390–9. doi:10.1007/BF00280469. PMID   8202084. S2CID   10429039.
  160. Greenberg MV, Ausin I, Chan SW, Cokus SJ, Cuperus JT, Feng S, et al. (March 2011). "Identification of genes required for de novo DNA methylation in Arabidopsis". Epigenetics. 6 (3): 344–54. doi:10.4161/epi.6.3.14242. PMC   3092683 . PMID   21150311.
  161. Meyer P (2013). "Transgenes and their contributions to epigenetic research". The International Journal of Developmental Biology. 57 (6–8): 509–15. doi: 10.1387/ijdb.120254pm . PMID   24166433.
  162. Hamilton AJ, Baulcombe DC (October 1999). "A species of small antisense RNA in posttranscriptional gene silencing in plants". Science. 286 (5441): 950–2. doi:10.1126/science.286.5441.950. PMID   10542148.
  163. 1 2 Mette MF, Aufsatz W, van der Winden J, Matzke MA, Matzke AJ (October 2000). "Transcriptional silencing and promoter methylation triggered by double-stranded RNA". The EMBO Journal. 19 (19): 5194–201. doi:10.1093/emboj/19.19.5194. PMC   302106 . PMID   11013221.
  164. Hamilton A, Voinnet O, Chappell L, Baulcombe D (September 2002). "Two classes of short interfering RNA in RNA silencing". The EMBO Journal. 21 (17): 4671–9. doi:10.1093/emboj/cdf464. PMC   125409 . PMID   12198169.
  165. 1 2 Xie Z, Johansen LK, Gustafson AM, Kasschau KD, Lellis AD, Zilberman D, et al. (May 2004). "Genetic and functional diversification of small RNA pathways in plants". PLOS Biology. 2 (5): E104. doi: 10.1371/journal.pbio.0020104 . PMC   350667 . PMID   15024409.
  166. Zilberman D, Cao X, Jacobsen SE (January 2003). "ARGONAUTE4 control of locus-specific siRNA accumulation and DNA and histone methylation". Science. 299 (5607): 716–9. Bibcode:2003Sci...299..716Z. doi:10.1126/science.1079695. PMID   12522258. S2CID   8498615.
  167. Dalmay T, Hamilton A, Rudd S, Angell S, Baulcombe DC (May 2000). "An RNA-dependent RNA polymerase gene in Arabidopsis is required for posttranscriptional gene silencing mediated by a transgene but not by a virus". Cell. 101 (5): 543–53. doi: 10.1016/s0092-8674(00)80864-8 . PMID   10850496. S2CID   2103803.
  168. Herr AJ, Jensen MB, Dalmay T, Baulcombe DC (April 2005). "RNA polymerase IV directs silencing of endogenous DNA". Science. 308 (5718): 118–20. Bibcode:2005Sci...308..118H. doi: 10.1126/science.1106910 . PMID   15692015. S2CID   206507767.
  169. Onodera Y, Haag JR, Ream T, Costa Nunes P, Pontes O, Pikaard CS (March 2005). "Plant nuclear RNA polymerase IV mediates siRNA and DNA methylation-dependent heterochromatin formation". Cell. 120 (5): 613–22. doi: 10.1016/j.cell.2005.02.007 . PMID   15766525. S2CID   1695604.
  170. Kanno T, Huettel B, Mette MF, Aufsatz W, Jaligot E, Daxinger L, et al. (July 2005). "Atypical RNA polymerase subunits required for RNA-directed DNA methylation". Nature Genetics. 37 (7): 761–5. doi:10.1038/ng1580. PMID   15924141. S2CID   20032369.
  171. Pontier D, Yahubyan G, Vega D, Bulski A, Saez-Vasquez J, Hakimi MA, et al. (September 2005). "Reinforcement of silencing at transposons and highly repeated sequences requires the concerted action of two distinct RNA polymerases IV in Arabidopsis". Genes & Development. 19 (17): 2030–40. doi:10.1101/gad.348405. PMC   1199573 . PMID   16140984.
  172. 1 2 3 Bond DM, Baulcombe DC (January 2015). "Epigenetic transitions leading to heritable, RNA-mediated de novo silencing in Arabidopsis thaliana". Proceedings of the National Academy of Sciences of the United States of America. 112 (3): 917–22. Bibcode:2015PNAS..112..917B. doi: 10.1073/pnas.1413053112 . PMC   4311854 . PMID   25561534.
  173. Kanazawa A, Inaba JI, Shimura H, Otagaki S, Tsukahara S, Matsuzawa A, et al. (January 2011). "Virus-mediated efficient induction of epigenetic modifications of endogenous genes with phenotypic changes in plants". The Plant Journal. 65 (1): 156–168. doi: 10.1111/j.1365-313X.2010.04401.x . hdl: 2115/49399 . PMID   21175898.
  174. Dalakouras A, Moser M, Zwiebel M, Krczal G, Hell R, Wassenegger M (December 2009). "A hairpin RNA construct residing in an intron efficiently triggered RNA-directed DNA methylation in tobacco". The Plant Journal. 60 (5): 840–51. doi: 10.1111/j.1365-313X.2009.04003.x . PMID   19702668.
  175. Pignatta D, Novitzky K, Satyaki PR, Gehring M (November 2018). "A variably imprinted epiallele impacts seed development". PLOS Genetics. 14 (11): e1007469. doi: 10.1371/journal.pgen.1007469 . PMC   6237401 . PMID   30395602.
  176. Papikian A, Liu W, Gallego-Bartolomé J, Jacobsen SE (February 2019). "Site-specific manipulation of Arabidopsis loci using CRISPR-Cas9 SunTag systems". Nature Communications. 10 (1): 729. Bibcode:2019NatCo..10..729P. doi:10.1038/s41467-019-08736-7. PMC   6374409 . PMID   30760722.
  177. Dalakouras A, Wassenegger M, Dadami E, Ganopoulos I, Pappas ML, Papadopoulou K (January 2020). "Genetically Modified Organism-Free RNA Interference: Exogenous Application of RNA Molecules in Plants". Plant Physiology. 182 (1): 38–50. doi:10.1104/pp.19.00570. PMC   6945881 . PMID   31285292.
  178. Regalado A (11 August 2015). "The Next Great GMO Debate". MIT Technology Review.
  179. Gohlke J, Mosher RA (September 2015). "Exploiting mobile RNA silencing for crop improvement". American Journal of Botany. 102 (9): 1399–400. doi: 10.3732/ajb.1500173 . PMID   26391704.