Skip to main content

Methylome decoding of RdDM-mediated reprogramming effects in the Arabidopsis MSH1 system

Abstract

Background

Plants undergo programmed chromatin changes in response to environment, influencing heritable phenotypic plasticity. The RNA-directed DNA methylation (RdDM) pathway is an essential component of this reprogramming process. The relationship of epigenomic changes to gene networks on a genome-wide basis has been elusive, particularly for intragenic DNA methylation repatterning.

Results

Epigenomic reprogramming is tractable to detailed study and cross-species modeling in the MSH1 system, where perturbation of the plant-specific gene MSH1 triggers at least four distinct nongenetic states to impact plant stress response and growth vigor. Within this system, we have defined RdDM target loci toward decoding phenotype-relevant methylome data. We analyze intragenic methylome repatterning associated with phenotype transitions, identifying state-specific cytosine methylation changes in pivotal growth-versus-stress, chromatin remodeling, and RNA spliceosome gene networks that encompass 871 genes. Over 77% of these genes, and 81% of their central network hubs, are functionally confirmed as RdDM targets based on analysis of mutant datasets and sRNA cluster associations. These dcl2/dcl3/dcl4-sensitive gene methylation sites, many present as singular cytosines, reside within identifiable sequence motifs. These data reflect intragenic methylation repatterning that is targeted and amenable to prediction.

Conclusions

A prevailing assumption that biologically relevant DNA methylation variation occurs predominantly in density-defined differentially methylated regions overlooks behavioral features of intragenic, single-site cytosine methylation variation. RdDM-dependent methylation changes within identifiable sequence motifs reveal gene hubs within networks discriminating stress response and growth vigor epigenetic phenotypes. This study uncovers components of a methylome “code” for de novo intragenic methylation repatterning during plant phenotype transitions.

Background

Plants deploy a range of seed dispersal mechanisms that can span significant distance and environmental variation [1], necessitating robust environmental adaptation strategies. Successful germination and seedling establishment in a new location favors inducible and sustainable resilience mechanisms to facilitate survival under diverse environmental conditions. Modeling heritable phenotypic response assumes a plant’s ability to manage internal metabolic stochasticity in a manner that provides for phenotype stability (canalization) together with the requisite plasticity to respond to environmental change [2, 3]. It is thought that plants persist in new and changing environments through epigenetic processes [4, 5], but directional targeting of specific gene networks for enhanced phenotypic resilience in plants has not been previously defined.

Targeted epigenetic changes in plants take place, in part, through the RNA-directed DNA methylation (RdDM) pathway. This process effects de novo DNA methylation within transposable element and heterochromatic regions of the genome [6, 7], with less clear influence on genic methylation. Small RNAs (sRNA) associated with the RdDM pathway establish target sites and are generated by the activity of DCL3 in the canonical pathway [6], and DCL2 and DCL4 in associated pathways [7]. Site-specific DNA methylation is directed by the methyltransferase DRM2 [8] in a process that links to histone modifications influencing local chromatin architecture. These effects involve the histone deacetylase HDA6 [9]. HDA6 is a component of chromatin-mediated stress response [10, 11], presumably participating in re-establishment of chromatin homeostasis following environmental or developmental changes [12].

MSH1 is a plant-specific, nuclear-encoded gene for a DNA-binding protein [13] that maintains mitochondrial [14] and plastid [15] genome stability. Disruption of the plastid-targeted MSH1 protein causes plant developmental and epigenetic reprogramming [16, 17] and a variable, stress-responsive, daylength-sensitive growth phenotype (state 1) (Fig. 1). The msh1 reprogramming process is HDA6-dependent, such that an msh1,hda6 double mutant is lethal under standard (12-h daylength) growth conditions [18]. RNAi knockdown of MSH1 and subsequent transgene segregation leads to recurrently heritable, nongenetic msh1 “memory” (state 2). The msh1 memory state is a DRM2-dependent condition of reduced growth, delayed maturity, and chronic abiotic stress response [18].

Fig. 1
figure 1

The msh1 system is composed of four msh1-derived epigenetic states. In Arabidopsis, four distinct plant states originate from MSH1 knockdown or knockout. States 1 and 2 derive directly from msh1 disruption, resulting in highly stress-responsive phenotypes. State 1 at short daylength is variable, including a low-frequency “perennial-like” phenotype [16]. States 3 and 4 involve interaction of msh1-modified and naïve (wild type) genomes through grafting or crossing, resulting in growth vigor phenotypes. Genetic evidence of RdDM dependence is indicated at relevant transitions in gray shading

Grafting an isogenic wild type scion to msh1 rootstock produces progeny with measurably enhanced growth vigor, seed production, and resilience (state 3). This phenotype is markedly different from msh1(state 1) and derived memory (state 2). The graft progeny growth vigor, referred to as heritable enhancement through grafting (HEG), is sustained over multiple successive generations [17]. HEG effects depend on transgenerational sRNA transmission. Progeny growth and differential methylation outcomes are obviated by including dcl2/dcl3/dcl4 triple mutations with msh1 in the rootstock [19]. A fourth msh1-derived nongenetic state exists, detailed in this report. This state arises from crossing of msh1 to isogenic wild type, leading to a range of enhanced growth vigor in the epi-line progeny. In epi-lines, these effects are more variable and diminish after 5 to 6 generations, with evidence of sporadic reversion to a condition resembling state 2 memory as shown in this study.

Here, we exploit the MSH1 system as a model to address two fundamental questions of plant phenotypic plasticity: What are the criteria for identifying plant de novo RdDM gene targets that influence directional phenotype adjustment, and do these epigenomic effects lend themselves to prediction? We integrated nongenetic plant phenotype variation, DNA methylation, sRNA, mutant, and gene expression datasets from the four msh1-derived nongenetic states, to home in on putative target pathways that participate in msh1 developmental reprogramming. With a methylation analysis procedure enhanced for sensitivity, we discriminated gene networks undergoing directional changes in methylation and transcription. These networks appeared to underpin plant phenotype transitions. The DNA methylation analysis incorporated signal detection, machine learning, and sRNA and RdDM mutant datasets to classify candidate RdDM target loci that participate in msh1 effects without regard to cytosine context or numbers of proximal methylation changes at a site. High-resolution analysis of these putative targets pinpointed differential methylation positions within identifiable sequence motifs to discriminate the four msh1-derived states. These observations support a model of msh1-triggered epigenetic effects as targeted, programmed actions that may serve to promote phenotype plasticity in a manner far more precise than has been previously reported.

Results

msh1 epi-lines comprise a distinct nongenetic state based on phenotype

Manipulation of the msh1 mutant leads to four distinct phenotypes, with states 1 and 2 characterized by slowed growth, delayed flowering, and persistent stress response, and states 3 and 4 producing enhanced growth vigor and greater seed set over wild type (WT) (Fig. 1). State 4 results from direct or reciprocal crossing of Col-0 msh1 mutant (state 1) or memory line (state 2) x Col-0 WT and generation of epi-F2 and epi-F3 families in Arabidopsis. Similar results were obtained regardless of the direction of the cross, which indicates that these progeny phenotype changes do not derive from altered plastid genome effects. Regardless, we used Col-0 WT as female in crosses presented in this study to retain wild type cytoplasm. Progeny populations showed more variable distribution of growth-enhanced phenotypes within the F2 generation than occurs in state 3 graft progeny [19], and individual epi-line populations could be categorized with either enhanced vegetative growth, greater seed set, or both (Fig. 2; Additional file 1: Fig. S1).

Fig 2
figure 2

Characteristics of epi-lines derived by crossing msh1 T-DNA mutant with isogenic wild type. a The phenotypes of different epi-line F3 populations at 34 DAP. The lines derive from WT x msh1 crosses, with Epi 24 and Epi 8 from one parental cross, and Epi 10 and Epi 19 from a second parental cross. b Total leaf area (34 DAP), c, days to bolting, and d, seed weight (mg) are shown for the four populations along with WT control. b–d Bars represent means ± SE. The Mann–Whitney U-test with two-sided alternative hypothesis was used to test significance of the difference of mean between each Epi F3 population and WT control. e Root phenotype of the four different Epi F3 populations grown in sand (33 DAP). f Total leaf area (33 DAP), dry leaf weight (mg), and dry root weight (mg) are shown for the four populations grown on sand along with WT control. f–h Bars represent means ± SE. The Mann–Whitney U-test with two-sided alternative hypothesis was used to test significance of the difference of mean between each Epi F3 population and WT control. Significance codes: *p < 0.05, **p < 0.01, ***p < 0.001, ns – not significant

We followed 4 F3 populations of cross-derived epi-lines. Epi 8 and Epi 24 were sibling lines from one WT x msh1 cross event, and Epi 10 and Epi 19 were sibling lines from a second WT x msh1 cross. All four F3 epi-lines showed uniform phenotypes within each population, but significant variation between the four populations (Fig. 2; Additional file 1: Fig. S1). Epi-line populations were increased in aboveground vegetative growth and underground root development (Fig. 2a, e). Three of the populations, Epi 8, 10, and 19, had significantly higher total leaf area than the WT control (Fig. 2b, f), and all four populations had higher dry leaf weight than WT (Fig. 2g). The four populations also showed higher dry root weight than WT (Fig. 2h).

Enhanced reproductive growth occurred in the three epi-lines, Epi 8, 10, and 19, while Epi 8 showed early flowering (Fig. 2c). Importantly, Epi 8 and Epi 24, full-sib populations from the same original cross, showed distinct phenotypes. Epi 24 was highest in seed weight and lowest in leaf area among the epi-lines, while Epi 8 was highest in leaf area and lowest in seed weight. These observations suggest that nongenetic growth vigor can vary in vegetative versus reproductive allocations within closely related populations, and we focused on Epi 8 versus Epi 24 for more detailed studies.

The epi-line phenotype receded back to wild type by the fifth or sixth (S5, S6) generation. On occasion, we observed sporadic, low-frequency incidence of reversion to a condition resembling memory (state 2) phenotype (Fig. 3a). Putative revertants displayed reduced growth rate, altered leaf morphology, and enhanced stress response. Frequency of these putative reversion events ranged from 15 to 19% in Arabidopsis (Table 1). We have not observed this type of reversion phenomenon in graft progeny to date, with all graft progeny outcomes resulting in either enhanced growth or wild type growth phenotypes [17, 19].

Fig. 3
figure 3

Reversion phenotype in Arabidopsis. a Plant growth phenotype of three F3 epi-line populations, Epi 10 derived by crossing to msh1 T-DNA mutant, and Epi 6 and Epi 9 derived by crossing to msh1 memory line. Dashed circles indicate putative revertants. Col-0 wild type and msh1 memory are shown as controls. b Principal component analysis-linear discriminant analysis of methylome data from three epi-line (non-revertant) and three revertant full-sib F3 (state 4) progeny compared to three independently derived msh1 memory (state 2) plants

Table 1 Reversion frequencies within epi-lines

The reversion phenomenon implies that the epi-line state 4 and memory state 2 conditions are closely related. Reversion occurred in only epi-line populations deriving from crosses to an msh1 memory line (state 2) pollen parent, but not with msh1 null mutant (state 1) as pollen parent (Table 1). We interpret this observation to indicate that msh1 signal is weaker in the memory line than null mutant, leading to a less stable epi-line outcome. This interpretation would be consistent with the relatively weaker methylome repatterning effects reported in msh1 memory (state 2) relative to msh1 (state 1) [18].

The msh1 states 1 to 4 comprise discrete epigenetic phases by whole-genome methylome analysis

Significant changes in DNA methylation were detected in the four Arabidopsis epi-lines (F3), with gene-associated changes predominantly in CG context (Additional file 1: Fig. S2) and variable hyper and hypomethylation in CHH context within and between epi-lines (Additional file 1: Fig. S3). To estimate the relationship of intragenic methylation changes with plant phenotype, we used high-resolution methylome analysis. The procedure incorporates signal detection and machine learning to discriminate high-probability, treatment-associated methylation changes within gene regions. We consider three features of this analysis to be vital to this study: (1) The analysis focused on intragenic methylation repatterning without exclusion of single cytosine methylation changes. (2) DNA methylation variation within wild type samples was measured and subtracted to allow discrimination of treatment-associated (msh1-derived) methylation changes; and (3) individual cytosine methylation status, reflecting the outcome of pooled cell types, was analyzed as probability distributions using Hellinger divergence [20] rather than binary classification. The resulting datasets of treatment-associated differentially methylated positions (DMPs) could then be used for unbiased comparative analysis in all subsequent steps.

Hierarchical clustering of methylome data for individual plants from all four states in Arabidopsis used methylation level changes (computed as Hellinger divergence) at intragenic DMPs. The result produced clustering of individual plants (biological replicates) from the same population, with plants from different states separating to 6 branches distinct from WT control groups (Fig. 4). Despite originating from the same cross, Epi 8 and Epi 24 separated to two subgroups (Fig. 4a), in agreement with their distinct phenotypes (Fig. 2). The data suggested that Epi 24 genic DNA methylation repatterning was more closely related to that of graft progeny (HEG; state 3) than to its full-sib Epi 8 (Fig. 4a). Epi 10 and Epi 19 also formed two closely related clusters consistent with their phenotypic similarity. In contrast, hierarchical clustering of transposable element-associated DMPs in the four epi-lines produced outcomes in keeping with lineage, so that Epi 8 and Epi 24 now clustered together, as did Epi 10 and Epi 19 within a single sub-cluster, and both clusters separated from the graft progeny (HEG) state 3 (Fig. 4b). Thus, gene-associated methylation repatterning could be discerned to reflect plant phenotype among lines that are genetically uniform.

Fig. 4
figure 4

Discrimination of methylation repatterning among different msh1 states. a Hierarchical clustering results with genic methylome data from three different msh1 states in Arabidopsis: msh1 null mutant (state 1), Col-0/Col-0msh1 graft progeny (state 3), Epi F3 populations (state 4), and relevant Col-0 controls. b Hierarchical clustering results with TE DMP data from the same Arabidopsis datasets presented in panel a. Individuals were represented as vectors of the sum of Hellinger divergences (HD) at DMP positions within gene regions (a) or TE regions (b). The hierarchical clustering was built using Ward agglomeration method, and Hellinger divergence (HD) was computed by using the centroid of corresponding wild type samples. HD formula is reported elsewhere [21]

Methylome data support epi-line revertants as closely related to msh1 memory state 2

We included Epi-line (state 4) revertant samples in Arabidopsis in methylome analyses to test for evidence of methylation repatterning in revertants versus epi-line full-sib samples. Principal component with linear discriminant analysis (PCA-LDA) of these datasets using Hellinger divergence produced distinct clustering of epi-line (non-revertant) from revertant individuals deriving from the same progeny population. The epigenomic features that distinguished revertant from epi-line full-sib progeny presumably reflect inherent epigenomic stochasticity. Revertant methylome datasets showed a closer relationship with data from msh1 memory state 2 (Fig. 3b). Specifically, of 2192 differentially methylated genes (DMGs) identified in the memory state 2 dataset, 1536 (70%) were shared with the revertant dataset, where overlap with a non-revertant epi-line was 1227 (55%) DMGs.

The msh1 memory plants were derived from an independent population, so the similarity detected is expected to be meaningful rather than merely the consequence of lineage. A relationship of the revertant methylome profile with msh1 memory would be consistent with their demonstrated similarity in plant phenotype (Fig. 3). These results support previous indications of msh1 memory as distinctive in methylome features [18] and confirm low-frequency conversion of epigenetic state 4, enhanced in growth vigor, to something resembling the state 2 persistent stress response phenotype.

Epi-line methylome datasets are altered in growth-related gene networks

To investigate the relationship of intragenic differential methylation to emergent phenotypes, we focused on the identification of differentially methylated genes (DMGs). DMGs for each epi-line population were identified by applying generalized linear regression analysis (GLM) to test significance of the difference between group DMP counts (WT vs. epi-lines) in genes. This analysis identified 3204, 2860, 3208, and 2797 DMGs in Epi 8, Epi 10, Epi 19, and Epi 24, respectively (Additional file 2), with significant overlap between the epi-line datasets (Fig. S4). To investigate coincidence of DMG functional relationships in each population, we conducted a gene network-based enrichment analysis. DMGs from the different epi-lines shared enrichment for specific functional networks, particularly epigenetic-related processes (blue arrows) and plant developmental pathways (red arrows) (Fig. 5a; Additional file 2). For example, we detected enrichment for pathways involved in covalent chromatin modifications, DNA repair, maintenance of DNA methylation, RNA splicing and processing, and production of miRNAs involved in gene silencing. Toward plant development, we identified pathways for vegetative to reproductive phase transition, sugar metabolism and signaling, response to strigolactone, root development, auxin transport, and red/far-red light phototransduction. These outcomes reflect a non-random quality of DMP datasets, implying that methylation machinery-targeted gene loci and their respective networks can be identified based on methylome signal.

Fig. 5
figure 5

Significant enriched GO pathways by DMG and DEG analysis in epi-lines. a Enriched GO pathways shared by the four epi-lines in Arabidopsis as well as by F1 hybrid DMGs from the C24 x Ler cross by our analysis. Original methylation data for the F1 hybrid were previously reported [22]. Heat map was generated using the fold enrichment of GO terms (FDR < 0.05) common between all four epi-lines. Blue arrows indicate pathways likely contributing to epigenetic change and red arrows indicate pathways likely associated with plant developmental changes. Complete list of GO terms is available in  Extended Data 2 and 3. b Enriched GO pathways from DEGs (FDR<0.05) shared by Epi 8 and Epi 24 floral stem tissues. Heat map displays the fold enrichment of GO terms (FDR < 0.05) common between both the epi-lines. Complete list of GO terms is available in Extended Data 4. DAVID GO (version 6.8) [23] was used for GO enrichment analysis. GO terms with EASE score (a modified Fisher exact P-value) < 0.05 were used for FDR calculation. FDR was calculated using package stats (version 3.6.0; p.adjust method = FDR) in R. Package ggplot2 (version 3.3.3) in R was used to generate heatmap

We analyzed methylome data previously reported for the Arabidopsis F1 heterotic cross of ecotypes C24 and Ler [22] with the methylome analysis methods that were applied to the msh1 datasets (Additional file 3). The enriched networks emerging from F1 hybrid (C24 x Ler, and Ler x C24) data showed partial conformity to what was identified in msh1-derived epi-line data, emphasizing many of the same developmental pathways (Fig 5a). DMG overlap between hybrid data and epi-lines ranged from 47% overlap with Epi 10 to 54% overlap with Epi 8. This apparent alignment of methylome-identified gene pathways between the heterotic hybrids and msh1 epi-lines supports linkage of phenotype with methylome repatterning.

Examination of gene transcript changes in Arabidopsis epi-line populations involved sampling floral stem tissues, where MSH1-expression is enriched [13, 24]. Epi 8 and Epi 24, selected as full-sib lines differing in plant phenotype, were compared by RNAseq analysis of floral stem tissues. Our primary purpose in deriving gene transcript abundance data was to derive differentially expressed gene (DEG) datasets that could be integrated with methylome data for more robust gene network analysis. We have shown in previous studies that the direct overlap between DEG and DMG datasets is limited, with DMGs emphasized in upstream regulatory components and DEG data often identified in downstream components of a pathway [18, 19]. Epi 8 and Epi 24 analysis revealed 1375 DEGs from floral stem in Epi 8 and 1033 for Epi24 (p < 0.05; FDR < 0.05; Additional file 4).

To assess the impact of tissue type in gene expression changes, we also compared Epi 8 and Epi 24 root DEG datasets and included Epi 8 leaf DEG data to compare Epi 8 leaf, root, and floral datasets (Additional file 1: Fig. S5, Additional file 4). GO pathways identified in Epi 8 and Epi 24 root samples were consistent with those identified in floral stem samples and showed relatively few differences between the two lines except for notable enrichment in defense response-related pathways evident in Epi 8. Comparison of Epi 8 root and leaf samples showed enrichment for root morphogenesis and development pathways in the root DEG dataset, and differential representation of circadian and light response pathways. These tissue comparisons suggested that a feature distinguishing the Epi 8 and Epi 24 DEG datasets was Epi 24 enrichment in defense pathways in root tissues, whereas Epi 8 showed this enrichment in leaf tissues. Greatest DEG overlap occurred between root samples or between leaf and floral stem samples (Additional file 1: Fig. S5b), with DMG datasets overlapping to the greatest extent with floral stem DEG data (Additional file 1: Fig. S5c).

Network enrichment analysis of derived floral stem DEG datasets revealed that potential pathways that are altered in response to msh1 effects in floral stem emphasize environmental responsiveness (Fig. 5b). Several of the pathways identified in DEG data appear to reside downstream to those identified by methylome data for gene regulation and development. These DEG-associated pathways involved abiotic and biotic stress responses, circadian rhythm- and phytohormone response-related networks. Regulation of transcription was also prominent. Integration of DEG and DMG datasets served to amplify the msh1-associated network enrichments relative to either dataset alone.

Differential methylation and expression analysis identified putative central gene hubs for the epi-line state

To investigate the interaction of DMG and DEG datasets in epi-lines involved inputting DMG and DEG data to Cytoscape (3.8.2) to construct protein-protein interaction (PPI) maps, followed by k-means cluster analysis to identify putative core networks carrying central gene hubs. A k-means cluster machine learning algorithm uses betweenness centrality, closeness centrality, average shortest path length, clustering coefficient, degree, and eccentricity as parameters, allowing the identification of clusters that contain the most centralized nodes (proteins) in the PPI network.

In Arabidopsis Epi 8, a total of 3647 unique loci from DMGs and DEGs were used in the analysis to yield a PPI network formed by 430 genes. Functional enrichment analysis (FDR < 0.05) of these putative hub genes with the STRING [25] database functional enrichment tool revealed a PPI network of 153 hub genes and associated functional networks (Fig. 6a). These core networks included developmental process, response to hormone (particularly auxin), response to cold, chromosome organization, mRNA processing, spliceosome, and ribosome biogenesis. Comparison to Epi 24, with a total of 3523 unique loci from DMGs and DEGs forming a much larger PPI network of 346 genes, showed that several of these same processes were represented (Fig. 6b). However, Epi 24 analysis revealed more prominent changes in development-related gene expression and, most notably, a robust and overlapping effect in ribosome biogenesis-related gene expression (Fig. 6b). Ribosome biogenesis is integral to growth response [26]. Because Epi 8 and Epi 24 are full-sib populations from the same cross, these differing features are predicted to relate to their distinct phenotypes, with Epi 8 enhanced in vegetative growth and Epi 24 showing significantly increased reproductive vigor (Fig. 2).

Fig. 6
figure 6

PPI hubs derived from subsets of network-related DMGs and DEGs in Epi 8 and Epi 24. Epi 8 and Epi 24 represent progeny lines derived from the same cross, with Epi 8 displaying enhanced vegetative growth rate and Epi 24 significantly enhanced seed yield. The main subnetwork of hubs was obtained with the application of machine learning k-means clustering on the set of 3647 (Epi 8, panel a) and 3523 (Epi 24, panel b) network-related DMGs and DEGs (p < 0.1) identified in the Arabidopsis epi-line vs WT comparison. The analyses yielded 153 (a) and 346 (b) hub genes to form a closely related subnetwork. GO network enrichment analysis from the string application in Cytoscape [27] was used to identify enriched gene function pathways (FDR < 0.05) within the network. Blue represents DMGs, green represents DEGs, and magenta represents both DMGs and DEGs. Size of each node is proportional to its value of node degree and label font size is proportional to its betweenness centrality

Epi-state comparisons in Arabidopsis reveal conserved msh1 epigenome targets within biologically meaningful gene networks

We attempted to better define the relationship of genic methylation repatterning among the four distinct msh1-derived states by first assessing DMG overlap. Figure 7a shows that 871 DMGs were shared among the four msh1 states (Additional file 5), comprising 17.6% of msh1 (state 1), 39.7% of memory (state 2), 31.6% of graft progeny (HEG, state 3), and 31.1% of epi-line (state 4, Epi 24) DMGs. The overlap established a conserved msh1 “core” DMG dataset for further analysis (Fig. 7a). Each state was also classified by state-specific DMGs, accounting for 33.8% in state 1, 15.4% in state 2, 10.2% in state 3, and 15.2 % in state 4 (Fig. 7a). Of the 871 “core” msh1-associated DMGs, 531 (61%) could be placed within known networks (Fig. 7b).

Fig. 7
figure 7

The relationship of 871 core hub genes to msh1-derived states and biologically meaningful core networks. a Venn diagram of Arabidopsis DMGs identified from four different msh1-derived states (Col-0 genetic background): msh1 mutant (state 1), msh1 memory (state 2), graft progeny (HEG, state 3), and epi-line (Epi 24, state 4). b An overview of the PPI networks and individual 871 core hub genes. Blue represents DMGs, orange represents the DMGs that are dcl2/dcl3/dcl4-sensitive in graft progeny contrasting mutant rootstock analyses. c Hierarchical clustering of individual plant datasets from four different msh1-derived states based on the sum of Bayesian methylation level difference of DMPs over the 871 core genes from panel a. The hierarchical clustering was built using Ward agglomeration method. The Bayesian methylation level difference was computed as described previously [18]. d Main subnetwork of hubs obtained with the application of a machine learning k-means clustering algorithm on the set of 871 core genes from panel a. GO network enrichment analysis from the string application in Cytoscape [27] was used to identify enriched gene function pathways (FDR < 0.05) within the network. In total, 67 genes involved in enriched networks were identified. Blue nodes represent DMGs, with orange representing the DMGs that are dcl2/dcl3/dcl4-dependent in graft progeny contrasting mutant rootstock analysis. This 67-gene set supplied the RdDM candidate target genes for further study. Blue gene text represents DMGs proximal to only TE sequences, red text designates genes that are proximal to only sRNA clusters, and black text represents genes proximal to both TE and sRNA clusters. For both b and d, the size of each node is proportional to its value of node degree and the label font size is proportional to its betweenness centrality

We previously showed that components of the RdDM pathway were necessary for induction of msh1 state 1, transition from state 1 to state 2, and generation of state 3 following grafting [18, 19]. Therefore, we incorporated RdDM mutants to functionally identify RdDM-targeted DMGs within this core dataset. We contrasted methylome datasets for msh1 (state 1) versus dcl2/dcl3/dcl4/msh1 quadruple mutant and graft progeny from Col-0/Col-0msh1 (state 3) versus Col-0/Col-0dcl1/dcl3/dcl4/msh1 grafts to catalog DMGs that were RdDM (sRNA)-dependent [19]. These subtractive datasets confirmed that 674 (77%) of the 871 core DMG loci were predicted to be DCL2,DCL3,DCL4-dependent by obtaining the overlap between the 871 DMG core dataset and the msh1 vs dcl2/dcl3/dcl4/msh1 and graft datasets (Additional file 5). Figure 7b shows the PPI core hub network of DMGs with DCL2,DCL3,DCL4-dependent colored orange. These outcomes are evidence of a relationship between msh1-directed epigenetic effects and RdDM pathway influence on targeted methylome repatterning.

Whereas 871 core DMGs were shared among the four msh1 states, the methylation changes within these 871 loci also served to discriminate the four states. Figure 7c shows the results of hierarchical clustering for individual plants from the four epigenetic states with separation to four distinct clades by using only DMP methylation information over the 871 msh1 core genes. This clustering relationship helped validate our selection of a single memory (state 2), graft progeny (state 3), and epi-line (Epi 24, state 4) dataset for the identification of the overlap. The four epi-lines produced highly related datasets (Additional file 1: Fig. S4). Based on the PPI network derived for the 871 DMGs, we conducted k-means clustering to identify putative central hubs and conduct functional enrichment analysis. Figure 7d shows the resulting network of 67 candidate hubs (Additional file 7) involved in response to stress, developmental (growth) process, gene expression, spliceosome function, histone modification, and chromosome organization networks. Over 80% of the 67 hubs were predicted to be RdDM targets, shown in orange, with several central players in plant development and environmental stress response, including the cell growth regulator TARGET OF RAPAMYCIN (TOR), the SWI2/SNF2-like, meristem identity gene SPLAYED (SYD), its close homolog BRAHMA (BRM), and the chromatin remodelers PICKLE (PKL) and PICKLE-RELATED 1 (PKR1/CHR4).

Network-based analysis outcomes can be influenced by the available gene annotation and gene relationship information for any given species. To independently evaluate gene associations identified in these network-based studies, we performed weighted correlation network analysis [28, 29] with the 871 core msh1-associated DMGs. This alternative approach used pairwise comparisons within a multi-dimensional dataset to define clusters that could be assessed for network features in the absence of gene annotation data. At a correlation coefficient of 0.5, we identified 300 connected genes, with 44 (66%) overlapping with the 67 hub loci identified by k-means clustering (p-value = 0.01015 Fisher exact test). At a correlation coefficient of 0.6, we identified 166 genes, with 33 (49%) overlapping with the 67 hub loci (p-value = 0.000333 Fisher exact test) (Additional file 1: Fig. S5). These data reflect meaningful associations between the identified 67 hub loci for their methylation content and the discriminatory power of such information to distinguish phenotype among the four states.

Transposable elements and sRNAs associate with candidate RdDM target loci among the 871 msh1 DMGs

The RdDM pathway actively targets transposable element (TE) sequences [6, 7], prompting investigation of TE association with msh1-responsive loci. Looking at the 871 DMGs common among msh1 states, we detected association of these loci with TEs and sRNA (20-24 nt) clusters that was higher than genome-wide levels (61% DMGs within 2 kb of TE vs. 47% genome-wide; 77% DMGs within 2 kb of an sRNA cluster vs. 46% genome-wide). This enrichment for TE and sRNA cluster proximity increased further when we subset the dcl2,dcl3,dcl4-sensitive DMGs (65% DMGs within 2 kb of TE; 81% DMGs within 2 kb of an sRNA cluster). Yet, when we focused on the 67 hub DMGs derived by k-means clustering, only sRNA cluster enrichment was seen (49% of DMGs within 2 kb of TE; 72% DMGs within 2 kb of an sRNA cluster) (Additional file 6). Figure 7d shows only TE-associated genes (7) in blue text, only sRNA-associated genes (22) in red text, and genes associated with both TE and sRNA (26) in black text. Hence, although TEs may influence the methylation status of proximal genes and act as RdDM targets, sRNA cluster association regardless of TE proximity could define RdDM targets for the msh1 effect.

We found possible evidence of association between TE family and DMG proximity (Additional file 6). Comparing the number of members in TE families between observed and expected revealed significant overrepresentation in L1 and Gypsy families and underrepresentation in the Helitron family, both in the 871 DMGs common between msh1 states and 674 DMGs sensitive to dcl2,dcl3,dcl4 (Additional file 6). Further investigation is needed to reveal any biological significance of these associations. Based on the various analyses, four criteria served to classify RdDM target loci in our study (Fig. 7d): Positive loci were (1) confirmed DMGs in all four msh1 states that (2) comprised putative network hubs based on k-means cluster analysis, with genetic evidence for (3) dcl2,dcl3,dcl4-sensitivity, and/or (4) association with sRNA clusters (Additional file 6).

Detailed methylation analysis of selected RdDM target genes in the four different nongenetic states reveals sequence motifs encompassing dcl2/3/4-sensitive DMPs

Annotations of the 67 candidate hub loci supported their relevance to phenotype effects observed in the four msh1 states. In addition to gene networks for altered gene expression and chromatin behavior, major overlapping networks appeared to reflect the transition between stress response (Fig. 7d) that underpins msh1 and memory stress phenotypes versus growth vigor in HEG and epi-line states. Methylation features of the 67 candidate hub loci that discriminate the four msh1 states support a model of precisely targeted repatterning (Fig. 7c). Therefore, we investigated these loci for evidence of site-specific methylation changes.

Figure 8 and Table 2 show results of analysis for seven selected candidate RdDM target loci: TARGET OF RAPAMYCIN (TOR), a regulator of cell growth, SPLAYED (SYD), a chromatin remodeling component, UBIQUITIN-SPECIFIC PROTEASE 26 (UBP26), required for heterochromatin silencing, NUCLEAR RNA POLYMERASE D1A (NRPD1), the largest subunit of RNA polymerase IV, RNA POLYMERASE II LARGE SUBUNIT (NRPB1) involved in transcription, SU(VAR)3-9-RELATED PROTEIN 5 (SUVR5), a gene involved in H3K9me2 deposition and gene silencing, and UP-FRAMESHIFT1 (UPF1), a gene involved in RNA interference and nonsense-mediated RNA decay. These loci were selected to represent the various gene networks identified.

Fig. 8
figure 8

Investigation of methylation repatterning within candidate RdDM target genes that discriminate msh1-derived states. a Two of the putative cluster motifs identified based on differential gene methylation across four msh1-derived states. Hierarchical clustering on a set of 14-bp regions encompassing DMPs from seven (TOR, SYD, NRPB1, NRPD1, UBP26, SUVR5, UPF1) of the 67 core hub loci followed by DNA multiple sequence alignment of each cluster permitted the identification of methylation motifs (see “Methods” for more detail). b Difference of methylation levels on gene body DMPs within motif cluster 11 in the putative RdDM target gene UBP26. Variations on motif methylation repatterning at DMPs are shown with chromosome and position. Individual detected methylation changes are shown as color-coded dots for each plant assayed in each msh1 state, with positive (orange) indicating DMP and negative (blue) for no detected methylation change. Each line represents a single plant dataset. c Sample DMPs within motif cluster 11 in UBP26 and UPF1 that show dcl2/dcl3/dcl4-sensitivity in graft rootstock (state 1) and graft progeny (state 3) from contrasted mutant rootstock experiments. Note that dcl2/dcl3/dcl4 effects can only be assayed for msh1 and graft progeny data. All data associated with the seven genes in this figure are shown in Extended data 7. d Sample putative cluster motifs identified based on differential gene methylation across four msh1-derived states in analyses of the 67 core hub loci followed by DNA multiple sequence alignment of each cluster for methylation motifs. A complete data report for motifs identified based on all 67 genes is shown in Additional file 8

Table 2 Summary data for gene methylation within motif clusters 11 and 18*

Differential methylation analysis within the seven loci revealed evidence of state-specific repatterning (Table 2; Additional file 7). Changes in methylation at each locus were associated with identifiable sequence motifs that spanned approximately 14 nucleotides. Two sample composite motifs are shown in Fig. 8a, with examples of UBP26 and UPF1 state-specific methylation repatterning within cluster 11 shown in Fig. 8, panels b and c. Systematic analysis of all 67 loci revealed similar state-specific changes (Additional file 8). DMPs were assayed individually, although multiple DMPs could be detected within a single identified motif domain. DMPs were discriminated among replicated datasets, with several sites shown to be dcl2/dcl3/dcl4-sensitive in state 1 and state 3 (Fig. 8c; Table 2; Additional file 7). It was feasible to test dcl2/dcl3/dcl4 sensitivity in just the two states that encompass the msh1 rootstock [19], although these data confirm RdDM influence at single cytosine resolution. Table 2 summarizes scoring outcomes for cluster 11 and 18 motifs within the seven selected genes.

Identified DMPs did not show an obvious pattern of exon, intron or junction localization, and each gene contained multiple motif sites (Additional file 8). Evaluation of cluster motifs that encompass DMPs within the 67 msh1 core hub loci revealed, in many cases, evidence of high-order dependencies (Fig. 8d; Additional file 8). Multiple sequence alignment (MSA) of a given DNA motif can reveal a dependence relationship between two nucleotides located at different positions within the motif, reflected in their frequencies of simultaneous occurrence. First-order dependence refers to adjacent nucleotides, typically found in CG methylation context, second-order to nucleotides spaced two nucleotides apart, and high-order to nucleotides with intervening distance of more than two nucleotides. The relationships derive from the study of Markov dependence in DNA sequences, the basis for application of hidden Markov modeling of motif findings [30].

For the motifs identified, individual consensus nucleotides were evident at variable distance from the target cytosine, positioned at nucleotide 7 on plus or minus strand within each motif. For example, the motif from cluster 65 showed invariant T at position 14 and a consensus A at position 12, while the motif from cluster 66 showed invariant G at position 14 and an AG pair consensus at positions 2 and 3, respectively (Fig. 8d; Additional file 8). In motif 76, only G resides at positions 6 and 9 with consensus T at position 5, while motifs 82 and 86 showed consensus or invariant T at position 2. These observations support the hypothesis that target methylation sites within the genes are characterized by invariant or nearly invariant sequence features.

Discussion

MSH1 disruption or knockdown in plants leads to aberrant recombination in both mitochondrial [14] and plastid genomes [15], with the latter triggering nuclear epigenetic changes [13, 17] that are the focus of this study. Information from these investigations, integrated with earlier data, show that msh1, its memory lineage, graft-derived progeny and crossing-derived epi-lines comprise distinct de novo nongenetic states. These heritable states differ phenotypically from WT, consistently displaying nascent stress response (msh1/memory) or growth vigor (HEG/epi-line) phenotype changes across multiple plant species. Integrating carefully designed datasets confirmed RdDM influence on dynamic and potentially adaptive phenotype plasticity in plants.

In an earlier study, we speculated that msh1 epi-line variation in Arabidopsis was likely to be indistinguishable from graft progeny (HEG) outcomes [17]. This misinterpretation appears to be a consequence of insufficient sampling from only a single msh1 epi-line population. In fact, DNA methylation and gene expression analysis in this study provided unambiguous discrimination of the four msh1-derived states on a uniform Col-0 genetic background. Yet, it was possible to identify a subset of 871 DMGs shared among the four distinct states. We interpret this finding to indicate that reprogramming events in the original msh1 mutant are foundational to the other derived states.

The condition of msh1 memory, and its apparent relationship to epi-line reversion, raises important questions about stable transgenerational retention of a memory state. To date, we have not observed diminishment of msh1 memory as a whole-plant or methylome phenotype over seven sequential generations of self-pollination in Arabidopsis [18]. However, each memory generation shows evidence of DNA methylation repatterning at conserved sites [18]. This dynamic feature, together with dependence of memory induction on RdDM methyltransferase DRM2 and sRNA variation [18], suggest that transgenerational stability involves re-establishment of de novo methylation sites by inherited sRNAs each generation. Recurrent inheritance of epigenetic memory states in C. elegans points to small RNAs as the most likely epigenetic vector for stable transmission [31, 32] and to direct chromatin modification [33]. In C. elegans, two features of sRNAs facilitate their inheritance: the ability to circumvent sRNA degradation processes and to amplify by templating their own synthesis [34,35,36]. The extent to which sRNA transgenerational stability is sustained similarly in plants is unclear, but parental sRNA activity occurs in gametes and developing embryo [37, 38]. Epi-line (state 4) reversion back to a condition resembling msh1 memory (state 2) implies that interaction of the msh1 memory genome with naïve (WT) through reciprocal crossing does not impede reformation, at low frequency, of intact memory in subsequent generations. These sustained memory epigenomic features are presumed to be sRNA-directed.

Incorporation of analysis methods that discriminate treatment-associated intragenic DMP variation [21] provided sufficient resolution to conduct gene network-based analysis. A similar application of these methods to contrasting datasets from msh1, dcl2/dcl3/dcl4/msh1, and the corresponding graft-derived progeny functionally identified candidate RdDM target genes. Target loci were derived based on their classification as network hubs, differential methylation in all four msh1 states, DCL2/DCL3/DCL4-dependent effects, and/or association with sRNA clusters. Data from this analysis, which could not be accomplished using traditional DMR-based methods [18, 21, 28], opens the door to investigation of plant environmental or heterotic response as a precise and predictable epigenomic phenomenon.

Candidate RdDM targets identified in the study represent prominent members within pathways relevant to epigenetic regulation and plant adjustment to environmental change. The known function of these loci aligned with observed msh1-derived phenotypes. msh1 and its memory lineage display complex stress phenotypes, altered expression of abiotic and biotic stress pathways and circadian clock components, and suppressed growth rate and delayed flowering [16, 18, 39]. Graft [19] and epi-line progeny effects include changes in phytohormone and environmental response pathways and heterosis-like gene expression.

As anticipated, we found only modest overlap between the 67 core hub loci associated with msh1 effects and DEG data; this outcome is common to such studies. RNAseq analysis was conducted with samples from pooled cell types, which confounds detection of spatio-temporal expression, and does not reveal changes in transcription rate or splicing activity. Of the 67 core hub loci, at least 43% (29) show cell-specific differential expression in the absence of MSH 1 [24], and 66% (44) exhibit cell-type specific expression in the inflorescence [40]. Instead, our analysis focused on integration of gene pathways identified in common by DMG and DEG data, permitting greater resolution of these pathways than was achieved by either dataset alone. We have shown previously with this type of integrated pathway analysis that DMGs tend to represent upstream regulators with DEGs more abundant as downstream pathway components [18, 19]. In this study, our analysis permitted detection of the more robust ribosome biogenesis pathway component within datasets from epi-lines with greater growth vigor. A majority of this enhanced pathway was comprised of DEGs, several essential to plant growth and development (e.g., ROOT INITIATION DEFECTIVE 3 (RID3)), organ size control and cell proliferation (e.g. OLIGOCELLULA2 (OLI2)), mitotic division cycles and gametogenesis (e.g., SLOW WALKER 1 (SWA1)), and active cell division (e.g., BLOCK OF CELL PROLIFERATION 1 (BOP1)). We suspect that these DEGS operate downstream to central growth regulators.

Methylome-derived gene network information revealed a surprising number of genes likely to be integral to the observed msh1-triggered “pivot” from stress to growth vigor phenotypes. For example, TARGET OF RAPAMYCIN (TOR), SPLAYED (SYD), and BRAHMA (BRM) emerged as putative RdDM target loci that undergo differential methylation in msh1-derived states. TOR kinase serves as a key developmental regulator in the plant that directs growth by integrating nutrient and environmental signals [41,42,43]. TOR functions antagonistically to stress response pathways and is postulated to serve as a decision point in the plant’s energy allocation for defense versus growth [43]. TOR influences plant growth and yield and acts to regulate the circadian clock in plants [44, 45]. In animal systems, TOR is known to also function within the nucleus, where it triggers chromatin and epigenomic responses [46]. These features suggest that TOR may be intrinsic to msh1 developmental reprogramming.

SPLAYED is a SWI2/SNF2-like protein in the SNF2 subclass that acts to regulate shoot apical meristem identity. SYD functions in meristem maintenance and regulates several early developmental processes [47,48,49]. BRAHMA is also a SWI2/SNF2 ATP-dependent chromatin remodeler and a SYD homolog; both mutants display pleiotropic phenotypes. BRM and SYD appear, based on mutant phenotypes and genome-wide occupancy studies, to carry out both redundant and distinct functions [48]. Further, SYD and BRM can act antagonistically with Polycomb-group (PcG) proteins [49]. For example, BRM promotes vegetative growth by suppressing PcG-associated H3K27me to upregulate the SHORT VEGETATIVE PHASE flowering repressor [50]. Likewise, BRM is required to establish epigenetic heat stress memory in association with FORGETTER 1 [51]. In genome-wide occupancy studies, SYD directly targets nearly 6000 mainly developmental and stress-responsive genes in Arabidopsis, influencing H3K27me3 levels at several target loci [49]. Consequently, these genes serve as innate regulators of plant vegetative and reproductive growth.

In the context of the msh1 system, modulation of these msh1 RdDM targets could influence the plant growth program for heightened stress response versus growth vigor, as well as epi-line vegetative growth vigor versus reproductive growth variation. A fourth chromatin remodeling component, PICKLE (PKL), similarly identified as an RdDM target in msh1, functions to maintain RdDM-targeted methylation [52]. PKL acts synergistically with RdDM pathways to influence plant development [53].

Transition from msh1 “stress” to “growth” state entails interaction of the msh1-modified nuclear genome with naïve (WT) achieved through crossing or grafting. Similar epigenomic interaction between epigenetically modified and native states is implicated in heterosis [54, 55]. Manipulation of the msh1 system leads to a condition that resembles classic F1 heterosis, but differs in its subsequent heritability.

What is striking about a majority of msh1 RdDM target genes is their broad regulatory capacity in stress response and development, interpreted to reflect the vital role of integrated signaling and metabolic networks to calibrate environmental response in plants [56]. In this study, TE methylation data aligned according to lineage relationships, whereas gene methylation agreed with phenotype changes in epi-line comparisons. This observation further links RdDM effects on intragenic methylation with possible plant phenotype modulation. Identification of loci based on site-specific methylation suggests that adjustments in chromatin poise, or amenability to expression, can involve precise methylation changes. These gene methylation effects can facilitate alternative transcript splicing behavior in response to cellular cues [57] that would not likely be detectable in DEG datasets. The 871 DMGs shared across msh1 states, and their 67 core hub loci, include significant RNA spliceosome pathway components (Fig. 7d).

The ability to predict sites of differential gene methylation within major regulatory and chromatin remodeling genes serves to mark these loci during environmental response or growth changes in the plant. Results suggest the existence of methylation motifs with variable order interdependency of specific motif nucleotide positions. Overall, motif consensus sequences showed statistically significant departure from randomness, supporting the participation of DNA features in local epigenomic behavior [58]. We expect that the high-order dependencies observed within methylation motifs reported here can be captured by future deep neural network machine learning models [59, 60]. Moreover, the MSH1 system, as a biological model, can generate sufficiently robust methylation motif libraries to feed such machine learning predictive models. We consider the system of intragenic DMP motifs described in this study to reflect a previously uncharacterized fine-tuning mechanism of epigenomic regulation, in contrast to the high-density DMR changes that characterize RdDM-mediated TE and promoter silencing [61]. Epigenomic effects reported here should prove valuable to deciphering environmentally influenced complex trait expression.

Conclusions

The dynamic nature of DNA methylation during developmental and environmental changes renders elusive the essential features of methylome variation. Analyses based on changes in methylation density and magnitude are less informative for phenotype-associated methylome interpretation than are studies of low-density, intragenic methylation repatterning. The MSH1 system provides unique materials for these initial efforts toward methylome decoding, identifying, and confirming RdDM gene targets within networks that undergo epigenomic reprogramming during growth-versus-stress response transitions. The extent that information from this study extrapolates across plant species will have important implications for understanding genotype x environment influences on plant fitness.

Methods

Plant materials and growth conditions

For Arabidopsis thaliana material used in the study, seeds (Col-0 genetic background) were sown on peat mix in square pots, with stratification at 4 °C for 2 days before transfer to growth chambers (22 °C, 12 h DL, 120−150 μmol m−2 s−1 light). To generate epi-lines, two third-generation msh1 T-DNA mutant (SAIL_877_F01) plants were used to pollinate Col-0 to generate two independent F1 populations. Derived F1 progenies were self-pollinated to generate F2 populations that were genotyped for the msh1 T-DNA mutation. MSH1 F2 plants were self-pollinated to produce F3 families that were used in the study.

Phenotypic data

For leaf area, photos of trays containing randomly placed treatment and control plants were taken at 33–34 DAP using a Canon EOS DIGITAL REBEL XSi camera. Total leaf area was measured using ImageJ (v1.51j8; https://imagej.nih.gov/ij/index.html) software. For root-related analysis, seeds were sown in square pots containing sand media topped with standard peat mix to allow seed germination. At 33 DAP, roots were cleaned under running water to remove sand particles and allowed to dry at room temperature for at least 3 weeks before measuring dry weight. At the same time, aboveground rosettes were also preserved, cleaned, and dried for dry leaf weight measurements.

Whole-genome bisulfite sequencing and DNA methylation analysis

In Arabidopsis, Col-0 WT, Epi 8, Epi 10, Epi 19, and Epi 24 plants were grown together as one experiment, and WT, epi-line revertant, and epi-line non-revertant plants were grown together as a second experiment, with three plants from each line selected for sequencing. For the reversion experiment, an F3 population obtained through Col-0 WT x msh1 memory was selected based on evidence of reversion activity. From this population, revertant and non-revertant full-sibs and a 4th generation memory sample in triplicate were used for bisulfite sequencing. Whole rosettes at the bolting stage from three biological replicates were flash frozen and ground in liquid nitrogen.

Tissues were ground and processed with the DNeasy Plant Kit (Qiagen, Germany) for genomic DNA with RNA removal according to the manufacturer’s protocol for subsequent bisulfite sequencing. Whole-genome bisulfite sequencing was conducted on the Hiseq 4000 (Illumina, USA) at BGI-Tech (Shenzhen, China) and Novogene (Beijing, China), according to the manufacturer’s instructions. Then, 150 bp read length and at least 4 Gb data per sample were derived for Arabidopsis samples and at least 20 Gb per sample was collected for tomato samples. Raw sequencing reads were quality-controlled with FastQC (version 0.11.5), trimmed with TrimGalore (version 0.4.1) and Cutadapt (version 1.15), then aligned to the reference genome using Bismark (version 0.19.0) with bowtie2 (version 2.3.3.1). DMPs were identified using Methyl-IT (version 0.3.2) R package as described previously [18, 19, 21]. Briefly, Hellinger Divergence (HD) was calculated with a pool of control (wild type) samples as reference as described previously [21]. Cytosines with methylation level difference >20% in the treatment vs. reference comparison were selected and further filtered by estimating the optimal cutoff for HD based on a machine learning approach to obtain DMPs (https://genomaths.github.io/methylit/). To identify the DMGs, we selected loci with at least seven DMPs and minimum DMP density of 3 per 10 kbp, followed by group comparison using likelihood ratio test (LRT) to select loci with log2fold change >1 and adjusted p-value < 0.05 (Benjamini and Hochberg method). TAIR10 version 38 GTF was used for Arabidopsis annotation. DAVID Bioinformatics Resources 6.8 was used for GO function enrichment analysis. GO terms with EASE score (a modified Fisher exact P-value) < 0.05 were selected and further used for FDR calculation. FDR was calculated using package stats (version 3.6.0; p.adjust method = FDR) in R. Package ggplot2 (version 3.3.3) in R was used to generate heatmap.

For hierarchical clustering, we used the hclust from the R package stats (version 3.6.0). Either sum of Hellinger divergence (HD) or sum of methylation level over a given locus was used for analysis. A complete documentation and package (with Methyl-IT) is available at https://genomaths.github.io/.

RNA sequencing and analysis

For the Arabidopsis WT and epi-lines, floral stem tissues from 3 individual 6-week-old plants of epi24, epi8, and WT; root tissue from 3 individual 10-day-old plants of epi24, epi8, and WT; leaf tissues from 3 individual 4-week-old plants of epi8 and WT were flash frozen and ground in liquid nitrogen. Tissues were subjected to RNA extraction, including DNA removal, using the NucleoSpin RNA Plant Kit (Macherey-Nagel, Germany). sRNA extraction used the NucleoSpin miRNA Plant Kit (Macherey-Nagel, Germany) following the manufacturer’s protocol. RNA libraries were constructed as described in the TruSeq RNA Sample Preparation v2 Guide and were sequenced with the 150 bp read, paired-end option, in the Novaseq analyzer (Illumina, USA) at Novogene (Beijing, China). Raw sequencing reads were quality-controlled, trimmed with TrimGalore (version 0.4.1), and then aligned to the TAIR10 reference genome using STAR (version 2.7.3a) with –twopassMode = Basic and –outFilterMultimapNmax = 1 parameters, retaining only uniquely mapped reads. QoRTs software package (version v1.3.0) with –minMAPQ = 25 option were used to generate BAM files. edgeR package (version 3.26.8) was used for gene count normalization and to identify DEGs (p < 0.05, FDR < 0.05, |log2FC| ≥ 0.5). DAVID Bioinformatics Resources 6.8 was used for GO function enrichment analysis. GO terms with EASE score (a modified Fisher exact P-value) < 0.05 were selected and further used for FDR calculation. FDR was calculated using package stats (version 3.6.0; p.adjust method = FDR) in R. Package ggplot2 (version 3.3.3) in R was used to generate heatmap.

Core hub PPI networks, functional enrichment, and weighted correlation networks

Cytoscape (ver. 3.8.2), and the built-in application STRING, clusterMaker was used to construct the core hub PPI network. The input DMGs and DEGs were placed into the STRING data network by Analyze Network function, and then classified into 3 clusters by using k-means cluster function, with Euclidean distance as the distance metric, betweenness centrality, closeness centrality, average shortest path length, clustering coefficient, degree and eccentricity as node attributes, and 300 iterations were carried out. Genes in the cluster highest overall centrality were selected as core hubs and subject to functional enrichment analysis using the string enrichment function (FDR < 0.05).

To confirm network associations, we conducted weighted correlation network analysis. Individuals were represented as vector of genes of 600 (67)-dimensional space of DMGs, where each gene was represented by the sum of Hellinger divergence at each DMP on the gene body. Next, principal component analysis (PCA) was applied to identify the genes with the highest contributions to the discrimination between the four msh1 states and controls. PCA is a standard statistical procedure to reduce data dimensionality, to represent the set of DMGs by new orthogonal (uncorrelated) variables, the principal components (PCs) [62], and to identify the variables with the main contribution to the PCs carrying most of the whole sample variance. In the current case, components carrying 1% of the sample variance (Guttman-Kaiser criterion [63]) and accumulating at least 95% of the whole sample variance were selected. Next, genes were represented as vectors of principal components, where vector (gene) coordinates are PC loadings. A PC score (genes-core) was computed as the Euclidean norm of the vector of principal components. Since the sum of the squared of variable loadings over a principal component is equal to 1, the squared loadings tell us the proportion of variance of the given principal components explained by the gene. Thus, the greater the PC score value, the greater the fraction of cumulative variance from the whole sample variance is carried by the gene and, consequently, the greater the discriminatory power carried by the gene.

The sum of HD on the gene body was computed with Methyl-IT function getGRegionsStat and the principal component with function pcaLDA, which applies the PCA calling function prcomp from the R package “stats.”

Motif clustering

Small regions (14 bp) encompassing DMPs in at least three samples were identified and considered as DNA methylation motif candidates in the 67 identified msh1 core hub genes. A distance matrix was estimated on the set of selected regions using function dist.dna from ape R package (version 5.5). Hierarchical cluster analysis on the set of selected regions (using the previous estimated distance matrix) was accomplished with function hclust from stats R package (version 4.1.1) and grouped to 100 clusters. UPGMA approach was applied as agglomeration algorithm. Clusters with fewer than 10 regions where discarded. A DNA multiple sequence alignment on each cluster of sequences was accomplished with MUSCLE algorithm implemented on Bioconductor R package muscle (version 3.14). The sample motifs presented in Fig. 8 resulted from the specific analysis of genes TOR, SYD, NRPB1, NRPD1, UBP26, SUVR5, and UPF1.

Motif score

We defined the motif score sjkof the aligned sequences j and k of length N as the logarithm base 2 of the number of matched bases found in the alignment. Formally:

$${s}_{jk}={\mathit{\log}}_2\sum_{i=1}^N{I}_{jk}^i$$
(1)

where \({I}_{jk}=\left\{\begin{array}{c}1\ if\ {b}_j^i={b}_k^i\\ {}0\ \mathrm{otherwise}\end{array}\right.\) for every base position i on sequences j and k. The maximum motif score is then: Max{sjk} = log2N. The motif score in a MSA is, thus, defined as:

$$S=\frac{1}{m}\sum_{j=1}^{M-1}\sum_{k=j+1}^M{s}_{jk}$$
(2)

For a MSA with M sequences of length N each, the number of pairwise comparisons is: \(m=\frac{M\times \left(M-1\right)}{2}\). For a fixed value of motif length N, the perfect MSA of DNA sequence motifs would have the maximum score: Max{S} = log2N. In other words, the maximum amount of information carried on an MSA in this modeling is Imax = log2N, and the amount of information (the uncertainty change) carried by a MSA is given by the expression:

$$I=-\left(S-{\mathit{\log}}_2N\right)$$
(3)

The same result is obtained if the letter frequencies in the MSA and Shannon entropies, before (perfect alignment) and after, are estimated instead of the number of matches. Then, the alignment information is computed as: I =  − (Hafter − Hbefore) [64].

Monte Carlo testing of MSA randomness

A Monte Carlo test (MCT) of how a given DNA multiple sequence alignment differs from randomly generated MSAs was accomplished assuming:

  1. (1)

    The count vectors nk summarizing the observed DNA base frequencies in N column (k = A, C, G, T) are distributed according to the multinomial distribution with parameters pk.

  2. (2)

    In Bayesian framework, since Dirichlet distribution is the conjugate prior distribution of the multinomial distribution, the parameter vector P = (pA, pC, pG , pT) is drawn from a Dirichlet distribution \({\mathcal{D}}_{\alpha P}\) with parameter vectors 𝛼 = (𝛼1, 𝛼2, 𝛼3, 𝛼4) and Q = (qA, qC, qG , qT).

A N × 4 matrix of counts was computed from the entire set of identified DNA motif candidates as indicated by Sjolander et al. [65], and the parameter vector 𝛼 = (𝛼1, 𝛼2, 𝛼3, 𝛼4) from a Dirichlet distribution was estimated applying function estimateDirichDist from the R package usefr available at GitHub: https://github.com/genomaths/usefr (version 0.1). As shown by Sjolander et al. [65], the posterior probabilities \({\hat{p}}_i\) are given by the expression:

$${\hat{p}}_i=\frac{n_i+{\alpha}_i}{\left|\overrightarrow{n}\right|+\left|\overrightarrow{\alpha}\right|}$$
(4)

where \(\left|\overrightarrow{n}\right|\) and \(\left|\overrightarrow{\alpha}\right|\) are the total number of counts and pseudo-counts (sum of αi), respectively. Random DNA MSA sequences were generated according to the estimated Dirichlet distribution with function rdirichlet from the R package usefr.

For MSAs of fixed length N, the log2N is a constant. Thus, for the purposes of MCT, we could consider the score statistic given in Eq. 2 and evaluate how much an observed aligned motif differs statistically from Monte Carlo simulated aligned sequences. The Monte Carlo p-value was estimated as:

$${p}_{\mathrm{value}}=\frac{1}{N_S+1}\sum_{i=0}^{N_s}{\delta}_i\left({S}_i,{S}_0\right)$$
(5)

where S0 is the motif score from the observed MSA to be tested, Si is the motif score for the ith Monte Carlo simulated MSA (including S0) and \({\delta}_i\left({S}_i,{S}_0\right)=\left\{\begin{array}{c}1\ if\ {S}_i\ge {S}_0\\ {}0\ \mathrm{otherwise}\end{array}\right.\) (i = 0, 1, 2, …).

Preserved methylation motifs were identified (visualized) with function seqLogo from Bioconductor R package seqLogo (version 4.1). The final visualization of DNA methylation sequence motifs was accomplished with function ggseqlogo from the CRAN R package ggseqlogo (version 0.1) [66]. DNA sequence motifs (Additional file 8) were selected based on their motif scores, all of which differed statistically from random MSA.

Availability of data and materials

The datasets generated and analyzed during the current study are available from the corresponding author upon request. All next-generation sequencing data generated by this study, related to msh1 state 4, were deposited to Gene Expression Omnibus database under the following accessions: GSE192428 (Arabidopsis epi BSseq) [67], GSE192432 (Arabidopsis msh1 memory reversion BSseq) [68], and GSE192429 (Arabidopsis epi RNAseq) [69]. Sequencing data for msh1 states 1 and 3 can be accessed at GSE152570 [19] and data for msh1 state 2 can be accessed under GSE129303 [18]. The customized code used in this study, including methylation analysis with Methyl-IT, DMP classification based on signal detection, and identification of DMGs, is deposited to Zenodo [70]. Codes for principal component and linear discriminant analyses are deposited to Zenodo [71].

References

  1. Gunn CR, Dennis JV. World Guide to Tropical Drift Seeds and Fruits. NY: Quadrangle/New York Times Book Co; 1976.

    Google Scholar 

  2. Joseph B, Corwin JA, Kliebenstein DJ. Genetic variation in the nuclear and organellar genomes modulates stochastic variation in the metabolome, growth, and defense. PLoS Genet. 2015;11:e1004779.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  3. Mackenzie SA, Kundariya H. Organellar protein multifunctionality and phenotypic plasticity in plants. Philos Trans R Soc Lond Ser B Biol Sci. 2020;375:20190182.

    Article  CAS  Google Scholar 

  4. Balao F, Paun O, Alonso C. Uncovering the contribution of epigenetics to plant phenotypic variation in Mediterranean ecosystems. Plant Biol (Stuttg). 2018;20 Suppl 1:38–49.

    Article  CAS  Google Scholar 

  5. Lämke J, Bäurle I. Epigenetic and chromatin-based mechanisms in environmental stress adaptation and stress memory in plants. Genome Biol. 2017;18:124.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  6. Castel S, Martienssen R. RNA interference in the nucleus: roles for small RNAs in transcription, epigenetics and beyond. Nat Rev Genet. 2013;14:100–12.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  7. Cuerda-Gil D, Slotkin R. Non-canonical RNA-directed DNA methylation. Nat Plants. 2016;2:16163.

    Article  CAS  PubMed  Google Scholar 

  8. Cao X, Jacobsen SE. Role of the Arabidopsis DRM methyltransferases in de novo DNA methylation and gene silencing. Curr Biol. 2002;12:1138–44.

    Article  CAS  PubMed  Google Scholar 

  9. To TK, Kim J-M, Matsui A, Kurihara Y, Morosawa T, Ishida J, et al. Arabidopsis HDA6 regulates locus-directed heterochromatin silencing in cooperation with MET1. PLoS Genet. 2011;7:e1002055.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  10. Zhou X, He J, Velanis CN, Zhu Y, He Y, Tang K, et al. A domesticated Harbinger transposase forms a complex with HDA6 and promotes histone H3 deacetylation at genes but not TEs in Arabidopsis. J Integr Plant Biol. 2021;63:1462–74.

    Article  CAS  PubMed  Google Scholar 

  11. Chen LT, Wu K. Role of histone deacetylases HDA6 and HDA19 in ABA and abiotic stress response. Plant Signal Behav. 2010;5:1318–20.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  12. Saha RN, Pahan K. HATs and HDACs in neurodegeneration: a tale of disconcerted acetylation homeostasis. Cell Death Differ. 2006;13:539–50.

    Article  CAS  PubMed  Google Scholar 

  13. Virdi KS, Wamboldt Y, Kundariya H, Laurie JD, Keren I, Kumar SKR, et al. MSH1 is a plant organellar DNA binding and thylakoid protein under precise spatial regulation to alter development. Mol Plant. 2016;9:245–60.

    Article  CAS  PubMed  Google Scholar 

  14. Davila JI, Arrieta-Montiel MP, Wamboldt Y, Cao J, Hagmann J, Shedge V, et al. Double-strand break repair processes drive evolution of the mitochondrial genome in Arabidopsis. BMC Biol. 2011;9:64.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  15. Xu YZ, Arrieta-Montiel MP, Virdi KS, de Paula WBM, Widhalm JR, Basset GJ, et al. MutS HOMOLOG1 is a nucleoid protein that alters mitochondrial and plastid properties and plant response to high light. Plant Cell. 2011;23:3428–41.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  16. Xu YZ, de la Rosa SR, Virdi KS, Arrieta-Montiel MP, Razvi F, Li S, et al. The chloroplast triggers developmental reprogramming when MutS HOMOLOG1 is suppressed in plants. Plant Physiol. 2012;159:710–20.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  17. Virdi KS, Laurie JD, Xu Y-Z, Yu J, Shao M-R, Sanchez R, et al. Arabidopsis MSH1 mutation alters the epigenome and produces heritable changes in plant growth. Nat Commun. 2015;6:6386.

    Article  CAS  PubMed  Google Scholar 

  18. Yang X, Sanchez R, Kundariya H, Maher T, Dopp I, Schwegel R, et al. Segregation of an MSH1 RNAi transgene produces heritable non-genetic memory in association with methylome reprogramming. Nat Commun. 2020;11:2214.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  19. Kundariya H, Yang X, Morton K, Sanchez R, Axtell MJ, Hutton SF, et al. MSH1-induced heritable enhanced growth vigor through grafting is associated with the RdDM pathway in plants. Nat Commun. 2020;11:5343.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  20. Basu A, Mandal A, Pardo L. Hypothesis testing for two discrete populations based on the Hellinger distance. Stat Probab Lett. 2010;80:206–14.

    Article  Google Scholar 

  21. Sanchez R, Yang X, Maher T, Mackenzie SA. Discrimination of DNA methylation signal from background variation for clinical diagnostics. Int J Mol Sci. 2019;20:5343.

    Article  CAS  PubMed Central  Google Scholar 

  22. Shen H, He H, Li J, Chen W, Wang X, Guo L, et al. Genome-wide analysis of DNA methylation and gene expression changes in two Arabidopsis ecotypes and their reciprocal hybrids. Plant Cell. 2012;24:875–92.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  23. Huang D, Sherman B, Lempicki R. Systematic and integrative analysis of large gene lists using DAVID bioinformatics resources. Nat Protoc. 2009;4:44–57.

    Article  CAS  Google Scholar 

  24. Beltrán J, Wamboldt Y, Sanchez R, LaBrant EW, Kundariya H, Virdi KS, et al. Specialized plastids trigger tissue-specific signaling for systemic stress response in plants. Plant Physiol. 2018;178:672–83.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  25. Szklarczyk D, Gable AL, Nastou KC, Lyon D, Kirsch R, Pyysalo S, et al. The STRING database in 2021: customizable protein-protein networks, and functional characterization of user-uploaded gene/measurement sets. Nucleic Acids Res. 2021;49:D605–12.

    Article  CAS  PubMed  Google Scholar 

  26. Maekawa S, Yanagisawa S. Nucleolar stress and sugar response in plants. Plant Signal Behav. 2018;13:e1442975.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  27. Zhang Q, Burdette JE, Wang JP. Integrative network analysis of TCGA data for ovarian cancer. BMC Syst Biol. 2014;8:1338.

    Article  PubMed Central  PubMed  Google Scholar 

  28. Sanchez R, Mackenzie SA. Integrative network analysis of differentially methylated and expressed genes for biomarker identification in Leukemia. Sci Rep. 2020;10:2123.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  29. Toubiana D, Sade N, Liu L, Wilhelmi MDMR, Brotman Y, Luzarowska U, et al. Correlation-based network analysis combined with machine learning techniques highlight the role of the GABA shunt in Brachypodium sylvaticum freezing tolerance. Sci Rep. 2020;10:4489.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  30. Wu J, Xie J. Hidden Markov model and its applications in motif findings. Methods Mol Biol. 2010;620:405–16.

    Article  CAS  PubMed  Google Scholar 

  31. Burton NO, Burkhart KB, Kennedy S. Nuclear RNAi maintains heritable gene silencing in Caenorhabditis elegans. Proc Natl Acad Sci. 2011;108:19683–8.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  32. Rechavi O, Minevich G, Hobert O. Transgenerational inheritance of an acquired small RNA-based antiviral response in C. elegans. Cell. 2011;147:1248–56.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  33. Gu SG, Pak J, Guang S, Maniar JM, Kennedy S, Fire A. Amplification of siRNA in Caenorhabditis elegans generates a transgenerational sequence-targeted histone H3 lysine 9 methylation footprint. Nat Genet. 2012;44:157–64.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  34. Ru¨egger S, Großhans H. MicroRNA turnover: when, how, and why. Trends Biochem Sci. 2012;37:436–46.

    Article  CAS  Google Scholar 

  35. Pisacane P, Halic M. Tailing and degradation of Argonaute-bound small RNAs protect the genome from uncontrolled RNAi. Nat Commun. 2017;8:15332.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  36. Houri-Zeevi L, Rechavi O. A matter of time: small RNAs regulate the duration of epigenetic inheritance. Trends Genet. 2017;33:46–57.

    Article  CAS  PubMed  Google Scholar 

  37. Long J. Nurse cell-derived small RNAs define paternal epigenetic inheritance in Arabidopsis. Science. 2021;373:eabh0556.

    Article  CAS  PubMed  Google Scholar 

  38. Parent JS, Cahn J, Herridge RP, Grimanelli D, Martienssen RA. Small RNAs guide histone methylation in Arabidopsis embryos. Genes Dev. 2021;35:841–6.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  39. Shao MR, Raju SKK, Laurie JD, Sanchez R, Mackenzie SA. Stress-responsive pathways and small RNA changes distinguish variable developmental phenotypes caused by MSH1 loss. BMC Plant Biol. 2017;17:47.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  40. Shi D, Jouannet V, Agusti J, Kaul, Levitsky V, Sanchez P. Tissue-specific transcriptome profiling of the Arabidopsis inflorescence stem reveals local cellular signatures. Plant Cell. 2021;33:200–23.

    Article  PubMed  Google Scholar 

  41. Xiong Y, McCormack M, Li L, Hall Q, Xiang C, Sheen J. Glucose-TOR signaling reprograms the transcriptome and activates meristems. Nature. 2013;496:181–6.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  42. Rosenberger CL, Chen J. To grow or not to grow: TOR and SnRK2 coordinate growth and stress response in Arabidopsis. Mol Cell. 2018;69:3–4.

    Article  CAS  PubMed  Google Scholar 

  43. Margalha L, Confraria A, Baena-González E. SnRK1 and TOR: modulating growth–defense trade-offs in plant stress responses. J Exp Bot. 2019;70:2261–74.

    Article  CAS  PubMed  Google Scholar 

  44. Deprost D, Yao L, Sormani R, Moreau M, Leterreux G, Nicolai M, et al. The Arabidopsis TOR kinase links plant growth, yield, stress resistance and mRNA translation. EMBO Rep. 2007;8:864–70.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  45. Zhang N, Meng Y, Li X, Zhou Y, Ma L, Fu L. Metabolite-mediated TOR signaling regulates the circadian clock in Arabidopsis. Proc Natl Acad Sci. 2019;116:25395–7.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  46. Laribee RN, Weisman R. Nuclear functions of TOR: impact on transcription and the epigenome. Genes. 2020;11:641.

    Article  CAS  PubMed Central  Google Scholar 

  47. Kwon CS, Chen C, Wagner D. WUSCHEL is a primary target for transcriptional regulation by SPLAYED in dynamic control of stem cell fate in Arabidopsis. Genes Dev. 2005;19:992–1003.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  48. Kwon CS, Hibara K, Pfluger J, Bezhani S, Metha H, Aida M, et al. A role for chromatin remodeling in regulation of CUC gene expression in the Arabidopsis cotyledon boundary. Development. 2006;133:3223–30.

    Article  CAS  PubMed  Google Scholar 

  49. Shu J, Chen C, Li C, Thapa RK, Song J, Xie X. Genome-wide occupancy of Arabidopsis SWI/SNF chromatin remodeler SPLAYED provides insights into its interplay with its close homolog BRAHMA and Polycomb proteins. Plant J. 2021;106:200–13.

    Article  CAS  PubMed  Google Scholar 

  50. Li C, Chen C, Gao L, Yang S, Nguyen V, Shi X, et al. The Arabidopsis SWI2/SNF2 chromatin Remodeler BRAHMA regulates polycomb function during vegetative development and directly activates the flowering repressor gene SVP. PLoS Genet. 2015;11:e1004944.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  51. Brzezinka K, Altmann S, Czesnick H, Nicolas P, Gorka M, Benke E, et al. Arabidopsis FORGETTER1 mediates stress-induced chromatin memory through nucleosome remodeling. eLife. 2016;5:e17061.

    Article  PubMed Central  PubMed  Google Scholar 

  52. Yang R, Zheng Z, Chen Q, Yang L, Huang H, Miki D, et al. The developmental regulator PKL is required to maintain correct DNA methylation patterns at RNA-directed DNA methylation loci. Genome Biol. 2017;18:103.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  53. Yang R, He L, Huang H, Zhu J-K, Lozano-Duran R, Zhang H. RNA-directed DNA methylation has an important developmental function in Arabidopsis that is masked by the chromatin remodeler PICKLE. J Integr Plant Biol. 2020;62:1647–52.

    Article  CAS  PubMed  Google Scholar 

  54. Groszmann M, Greaves IK, Albertyn ZI, Scofield GN, Peacock WJ, Dennis ES. Changes in 24-nt siRNA levels in Arabidopsis hybrids suggest an epigenetic contribution to hybrid vigor. Proc Natl Acad Sci. 2011;108:2617–22.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  55. Lauss K, Wardenaar R, Oka R, van Hulten MHA, Guryev V, Keurentjes JJB, et al. Parental DNA methylation states are associated with heterosis in epigenetic hybrids. Plant Physiol. 2018;176:1627–45.

    Article  CAS  PubMed  Google Scholar 

  56. Li L, Sheen J. Dynamic and diverse sugar signaling. Curr Opin Plant Biol. 2016;33:116–25.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  57. Jabre I, Reddy ASN, Kalyna M, Chaudhary S, Khokhar W, Byrne KJ, et al. Does co-transcriptional regulation of alternative splicing mediate plant stress responses? Nucleic Acids Res. 2019;47:2716–26.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  58. Chua EYD, Vasudevan D, Davey GE, Wu B, Davey CA. The mechanics behind DNA sequence-dependent properties of the nucleosome. Nucleic Acids Res. 2012;40:6338–52.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  59. Ramachandran A, Lumetta SS, Klee E, Chen D. A recurrent Markov state-space generative model for sequences. Proceedings of the Twenty-Second International Conference on Artificial Intelligence and Statistics, PMLR 2019;89:3070–79.

  60. Wainberg M, Merico D, Delong A, Frey BJ. Deep learning in biomedicine. Nat Biotechnol. 2018;36:829–38.

    Article  CAS  PubMed  Google Scholar 

  61. Erdmann RM, Picard CL. RNA-directed DNA Methylation. PLoS Genet. 2020;16:e1009034.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  62. Stevens JP. Applied multivariate statistics for the social sciences: Routledge Academic; 2009.

    Google Scholar 

  63. Jackson DA. Stopping rules in principal components analysis: a comparison of heuristical and statistical approaches. Ecology. 1993;74:2204–14.

    Article  Google Scholar 

  64. Schneider TD. Theory of molecular machines. II. Energy dissipation from molecular machines. J Theor Biol. 1991;148:125–37.

    Article  CAS  PubMed  Google Scholar 

  65. Sjolander K, Karplus K, Brown M, Hughey R, Krogh A, Mian IS, et al. Dirichlet mixtures: a method for improved detection of weak but significant protein sequence homology. Bioinformatics. 1996;12:327–45.

    Article  CAS  Google Scholar 

  66. Wagih O. ggseqlogo: a versatile R package for drawing sequence logos. Bioinformatics. 2017;33:3645–7.

    Article  CAS  PubMed  Google Scholar 

  67. Kundariya H, Sanchez R, Yang X, Hafner A, Mackenzie SA. Bisulfite sequencing of msh1-derived epi-lines (msh1 state 4). Datasets Gene Expression Omnibus. 2022. https://0-www-ncbi-nlm-nih-gov.brum.beds.ac.uk/geo/query/acc.cgi?acc=GSE192428.

  68. Kundariya H, Sanchez R, Yang X, Hafner A, Mackenzie SA. Bisulfite sequencing of msh1 memory revertant plants. Datasets Gene Expression Omnibus. 2022; https://0-www-ncbi-nlm-nih-gov.brum.beds.ac.uk/geo/query/acc.cgi?acc=GSE192432.

  69. Kundariya H, Sanchez R, Yang X, Hafner A, Mackenzie SA. RNA sequencing of msh1-derived epi-lines (msh1 state 4). Datasets Gene Expression Omnibus. 2022; https://0-www-ncbi-nlm-nih-gov.brum.beds.ac.uk/geo/query/acc.cgi?acc=GSE192429.

  70. Kundariya H. Methyl-IT pipeline for the analysis of DNA methylation of Arabidopsis. Zenodo. 2022. https://0-doi-org.brum.beds.ac.uk/10.5281/zenodo.6812417.

  71. Sanchez R. Methyl-IT: R package based on signal detection and machine learning (v0.3.2.2). Zenodo. 2020. https://0-doi-org.brum.beds.ac.uk/10.5281/zenodo.4041756.

Download references

Acknowledgements

We thank Kelsey Castellanos for technical assistance, and Drs. Sally Assmann and Michael Axtell for thoughtful input.

Review history

The review history is available as Additional file 9.

Peer review information

Wenjing She was the primary editor of this article and managed its editorial process and peer review in collaboration with the rest of the editorial team.

Funding

This work was funded by grant NIH (R01 GM134056-01) to S.A.M. and support to the Mackenzie lab from the Huck Institutes of the Life Sciences.

Author information

Authors and Affiliations

Authors

Contributions

H.K. and X.Y. designed and conducted the laboratory experiments. H.K., X.Y., A.H., and R.S. performed methylome and network analysis, H.K. and X.Y. conducted RNAseq and network analysis, H.K. conducted RdDM mutant studies, H.K. and R.S. conducted sequence motif studies, and S.A.M. derived funding, designed experiments, participated in data analysis, and wrote the manuscript, edited by H.K., R.S., and A.K. All author(s) read and approved the final manuscript.

Corresponding author

Correspondence to Sally A. Mackenzie.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Competing interests

S. Mackenzie is a co-founder for EpiCrop Technologies Inc, a small start-up company that investigates the MSH1 system in an agricultural context.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

13059_2022_2731_MOESM1_ESM.pptx

Additional file 1: Figure S1. Phenotype of four selected F3 epi-lines and wild type used for methylome sequencing. Photo was taken at 27 days after planting. Epi24 and Epi8 are sister lines derived from the same crossing event; Epi10 and Epi19 are sister lines derived from a second crossing event. Figure S2. Relative DMP frequencies among different Arabidopsis epi-lines. DMPs were assigned to genic (blue), TE-related (red) and other genomic regions (green). The centroid of the wild type samples was used as reference. Relative DMP frequencies in each genomic feature were estimated as number of DMPs divided by number of total genomic cytosine positions in each genomic feature. A DMP was defined as ‘hyper’ if cytosine methylation level difference was greater than 0 and defined as ‘hypo’ if methylation level difference was less than 0. Figure S3. Total hyper- and hypo-DMP counts in epi-lines. Arabidopsis epi-lines vs WT. Each bar graph represents a single plant in each population. Cytosine context (CG, CHG, CHH) is shown separately for each plant. A DMP was defined as ‘hyper’ if cytosine methylation level difference was greater than 0 and defined as ‘hypo’ if methylation level difference was less than 0. Control centroid was used a reference. Figure S4. Upset plot showing the intersection of DMGs between different msh1 states. The horizontal bar graph on the right (set size) shows the total number of DMGs in each dataset. The upper bar graph (intersection size) shows the number of DMGs shared by different datasets represented by black connected dots. Package ComplexHeatmap (version 2.0.0) in R was used to make graphs. Red arrows indicate relationship among the four epi-lines (state 4) and between selected Epi 24 (state 4) and other individual datasets representing states l-3 in our study. Figure S5. Significant enriched GO pathways by DEG analysis in Epi 8 and Epi 24. a, Enriched GO pathways shared by epi-lines Epi 8 and Epi 24 root RNA (DEG) datasets and Epi 8 leaf sample DEG dataset in Arabidopsis. Heat map was generated using the fold enrichment of GO terms (FDR < 0.05) common between the datasets. GO terms with EASE score (a modified Fisher Exact P-value) < 0.05 were used for FDR calculation. FDR was calculated using package stats (version 3.6.0; p.adjust method = FDR) in R. Package ggplot2 (version 3.3.3) in R was used to generate heatmap. b, Upset plot showing the intersection of DEGs between different Epi 8 and Epi 24 tissues. The horizontal bar graph on the right (set size) shows the total number of DEGs in each dataset. The upper bar graph (intersection size) shows the number of DEGs shared by Epi 8 and Epi 24 datasets represented by black connected dots. c, Upset plots showing the intersection of DMG and DEG datasets in Epi 8 or Epi 24. The horizontal bar graphs on the right (set size) shows the total number of DMGs or DEGs in each dataset. The upper bar graphs (intersection size) shows the number of DMGs or DEGs shared by Epi 8 datasets or Epi 24 datasets represented by black connected dots. Package ComplexHeatmap (version 2.0.0) in R was used to make graphs. Figure S6. Results of weighted correlation network analysis. Network derived from the 871 core msh1-associated DMGs resulted in 166 DMGs with Spearman’s correlation coefficient of 0.6. Loci shown in red are 33 DMGs that overlap (49%) with the 67 core hub loci from k-means cluster analysis (p-value = 0.000333 Fisher’s Exact Test). The node and label sizes are in correspondence with the node degree (from 40 to 105). That is, a large node indicates a significant correlation with a large number of other genes for its/their epigenetic effect on phenotype. Node color represents the gene score (PC-score), from light-green (small value/contribution) to dark maroon (large value/contribution). The greater the PC-score value, the greater the fraction of cumulative variance from the whole sample variance is carried by the gene and, consequently, the greater the discriminatory power carried by the gene. The pairwise correlation between genes was computed as described in Methods.

Additional file 2. Epi-line state 4 DMGs and GO enrichment.

Additional file 3. Arabidopsis C24 and Ler F1 DMGs and network enrichments.

Additional file 4. Arabidopsis msh1 epi-line state 4 DEG and GO enrichment.

Additional file 5. 871 DMGs shared between four msh1 states.

Additional file 6. TE and sRNA associations with 871 DMGs shared between four msh1 states.

Additional file 7. Cluster 11 and 18 with dcl2/3/4-sensitive DMPs.

Additional file 8. Cluster motifs within 67 core hub loci.

Additional file 9. Review history.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kundariya, H., Sanchez, R., Yang, X. et al. Methylome decoding of RdDM-mediated reprogramming effects in the Arabidopsis MSH1 system. Genome Biol 23, 167 (2022). https://0-doi-org.brum.beds.ac.uk/10.1186/s13059-022-02731-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://0-doi-org.brum.beds.ac.uk/10.1186/s13059-022-02731-w

Keywords