Introduction

RNA modifications are critical regulators of mRNA processing and metabolism, which include splicing, 3′-end processing, nuclear export, translation, and decay. m6A is identified as the most abundant internal mRNA modification in mammals and plants1,2,3,4,5,6,7,8,9,10,11,12. In Arabidopsis, two different m6A writer complexes have been suggested to install m6A on mRNA. The first methyltransferase complex is composed of five respective orthologs of the components of the mammalian m6A methyltransferase complex that include mRNA adenosine methylase (MTA), MTB, VIRILIZER (VIR), FKBP12 INTERACTING PROTEIN 37KD (FIP37), and an E3 ubiquitin ligase HAKAI13,14,15,16. FIONA1 (FIO1), by contrast, is the Arabidopsis ortholog of the human methyltransferase METTL16, also depositing m6A modifications in U6 small nuclear RNA and a subset of mRNAs17,18,19. Defects in the m6A writer protein complex, such as MTA failure in plants, cause an embryo-lethal phenotype15,20,21,22 and stress responses23,24,25,26,27. Studies of other components of the writer complex in Arabidopsis revealed that FIP37 regulates shoot stem cell fate, FIO1 regulates floral transition and chlorophyll homeostasis, and VIRILIZER (VIR) is critical to vascular development13,14,18,19,28; whereas the defect of FIP37 in rice leads to early degeneration of microspores29,30. The m6A methylation can be reversed31,32. RNA m6A demethylases ALKBH10B and ALKBH9B, homologs of the human m6A demethylase ALKBH532, affect floral transition33 and viral infection34 in Arabidopsis. In our recent study, overexpression of the mammalian m6A demethylase FTO in rice dramatically increased the biomass and yield of rice and potato35, revealing that modulation of RNA m6A methylation could be a promising breeding or engineering strategy for crop improvement in the future. Although these observations indicate conserved mRNA m6A methylation regulators in plants as compared with animals, so far the high-resolution mRNA m6A maps in plants are still missing and the molecular level connection of m6A to plant development and other pathways are mostly unknown. We proceeded to fill this gap by using the newly developed high-resolution sequencing method36,37.

Methylated RNA m6A immunoprecipitation sequencing (MeRIP-seq or m6A-seq)38 has been widely used in identifying m6A-enriched transcripts in animals and plants. However, this method lacks single-base resolution and cannot quantify the extent of the modification. Variations of MeRIP-seq, including m6A individual-nucleotide-resolution crosslinking and immunoprecipitation (miCLIP)39, have been developed to detect m6A sites at single-base resolution. Unfortunately, these methods typically display low efficiency of UV crosslinking and cannot assess modification stoichiometry. Antibody-independent single-base m6A profiling methods, such as m6A-REF-seq40 or MAZTER-seq41, have been introduced. However, these methods can only identify RNA modifications in the ACA motif, and fail to identify other methylation motif. Nanopore direct RNA-seq (DRS) has been utilized to map m6A sites, but quantifying the differences in m6A modification levels with DRS is still challenging42,43,44. We have recently reported m6A-selective allyl chemical labeling and sequencing (m6A-SAC-seq)36,37 as a method capable of precisely mapping of individual m6A-modified sites in whole transcriptomes at single-nucleotide resolution. Although there are recently reported deamination-based methods that can also map m6A at base resolution, including one from us and collaborators45,46, m6A-SAC-seq is capable of reading m6A as a positive mutation signal without subtraction, and the optimized protocol works with 2–5 ng of input RNA36. We, therefore, decided to deploy m6A-SAC-seq to establish comprehensive maps of mRNA m6A at single-nucleotide resolution across various tissue types in two different plant species: Arabidopsis and rice.

A considerable number of high-confident m6A sites were identified spanning the entire life cycle of both rice and Arabidopsis. The evolutionarily conserved m6A mRNA modification sites across rice and Arabidopsis ortholog gene pairs play regulatory roles in tissue development, photosynthesis, and stress response. m6A levels are positively correlated with the length of the internal exon, but such correlation is missing in the last exon. Through comparative base-resolution m6A analysis across humans, rice and Arabidopsis, we unveil a distinct m6A distribution pattern that a suppression-activation dual model governs the m6A deposition in humans and plants. Using the robust method for comparison of m6A levels at single base resolution, we noticed that rice and Arabidopsis possess higher percentages of the overall m6A modifications in the 3’ UTR of their mRNAs than those of mammalian systems. These 3’ UTR m6A modifications generally stabilize mRNA and enhance translation, and these effects correlate well with the m6A fraction. These observations indicate that both fraction and position of m6A modification are critical for mRNA metabolism in plants. In addition, we noticed that the MTA-mediates m6A deposition in photosynthesis-related genes that transcribed from both nuclear and chloroplast genome in Arabidopsis. These m6A sites can either promote or reduce translation efficiency in a pathway-dependent manner. Altogether, our base-resolution and quantitative m6A sites across rice and Arabidopsis, provide a foundation for future studies to explore the regulatory roles of m6A in regulating plant development and evolution and for future plant engineering.

Results

m6A-SAC-seq identifies m6A modification sites in rice and Arabidopsis

m6A-SAC-seq utilizes the dimethyltransferase MjDim1 to introduce an allyl group to m6A, which upon chemical-induced cyclization could be read as mutation signals during reverse transcription (Supplementary Fig. 1a)36,37,47. We extracted total RNAs from nine Arabidopsis tissues (seedling, shoot, root, rosette leaf, cauline leaf, stem, flower, silique and seed) as well as eight rice tissues (plumule dark, plumule light, seedling at 8 days, seedling at 2 weeks, panicle, flag leaf at 10 days after anthesis, endosperm at 10 days after anthesis, and embryo at 10 days after anthesis) with two biological replicates for each sample (Fig. 1a, b). PolyA-tailed RNA of each biological replicate was purified and subjected to LC-MS/MS to measure the m6A/A ratio. The ratio of m6A/A in polyA-tailed RNA from these different tissues varied within the range of 0.36–0.75% in Arabidopsis (Fig. 1c) and in the range of 0.52–0.67% in rice (Fig. 1d). The remaining polyA-tailed RNAs were then processed following the m6A-SAC-seq library construction protocol36,37 to map m6A sites at the base resolution.

Fig. 1: m6A-SAC-seq effectively identifies m6A sites across plant tissues.
figure 1

a, b, Tissues from eight different rice organs (a) and nine different Arabidopsis organs (b), collected throughout their respective life cycles, were subjected to m6A-SAC-seq. c, d, mRNA m6A levels in the harvested samples were quantified using LC-MS/MS for both rice (c) and Arabidopsis (d). The m6A-to-A ratio was determined using calibration standards. Data are means ± SD, n = 6. e, Venn diagram showing m6A sites detected by SAC-seq overlapping with those identified by miCLIP and DRS, with the ±5 nt sliding window around each m6A site. f, A comparison of the percentages of overlapping m6A sites identified by different methods. g, h, The number of m6A sites and their motif distribution in rice (g) and Arabidopsis (h) tissues were shown. Motif sequences were separated into three groups, RAC, GAT and others. Base “R” denotes either A or G. 8D represents 8 days and 2 W represents 2 weeks. Source data are provided as a Source Data file.

To reduce background noise and eliminate any potential batch effect, we added 2% spike-in calibration probes into each sample. These calibration probes contained varying fractions of m6A within the NNm6ANN motif. By determining the mutation rates of both A and m6A within different motifs in these probes, we could best determine the m6A modification fraction at individual sites in each sample. Our initial results showed that the labeling efficiency of the calibration probes in plant cells was very high, comparable to that observed in mammalian cells (Supplementary Fig. 1b). Additionally, the average background noise of unmodified A sites was as low as 0.49%, which is 10 times lower than the cutoff (<5%) used for m6A site detection, indicating high m6A-SAC-seq data quality obtained from these plant samples. Furthermore, we assessed the relative conversion ratio and background noise in each sample. The relative conversion ratio ranged from 0.92 to 1.11 in Arabidopsis samples (Supplementary Fig. 1c) and from 0.96 to 1.07 in rice samples (Supplementary Fig. 1d), showing consistent efficiency of m6A labeling across samples. By carefully calibrating the differences observed in each sample, we ensured a fair and accurate comparison of m6A levels among different samples. This calibration step is crucial for obtaining reliable and meaningful results, enabling us to effectively compare the m6A modification among various plant samples.

Before proceeding to downstream analysis, we compared our m6A-SAC-seq data with previously published m6A sites profiled by DRS48 and miCLIP39 in Arabidopsis. Only sites with a sufficient number of sequence coverage among all tissues (depth> 10) were selected for further analysis. A total of 188,282 m6A sites were obtained when we combined the m6A sites detected by SAC-seq in this study, with ~42% and ~17% of the m6A sites measured by DRS and miCLIP overlapping with SAC-seq m6A sites, respectively (Supplementary Fig. 1e). With a sliding window of ±5 nt around the m6A sites, many more previously detected m6A sites overlap with the m6A-SAC-seq sites (Fig. 1e, f). As expected, a much higher percentage of m6A sites overlap between m6A-SAC-seq and DRS than that between m6A-SAC-seq and miCLIP (Fig. 1f). This might suggest the low accuracy of miCLIP in identifying m6A site, which could be caused by low crosslinking efficiency in plant tissues.

Principal component analysis (PCA) revealed a distinct clustering based on the m6A fractions of the different tissues (Supplementary Fig. 2a–d). An average of 49,791 m6A sites from 12,652 genes with at least 20 reads were identified in Arabidopsis libraries. In rice, an average of 67,173 m6A sites were detected from 15,138 genes. The number of high-confidence m6A sites from eight different rice tissues varied within the range of 52,646 in seedling 2 W to 82,157 in flag leaf (Fig. 1g). While the number of m6A sites among nine Arabidopsis tissues ranged from 25,990 in seed to 74,259 in silique (Fig. 1h). In comparison with previously published Arabidopsis MeRIP data of seedling (7,489 m6A peaks), we identified approximately 59,212 m6A sites within 14,180 genes in the seedling datasets, demonstrating high sensitivity of the SAC-seq method. Therefore, Arabidopsis seedling transcriptome contains ~4.2 m6A sites per gene, which is four-fold higher than that observed in the MeRIP data49. Interestingly, ~4.4 m6A sites per gene were observed among rice tissues, suggesting a likely conserved distribution density across different plant species. Consistent with the findings in mammals36,37 and plants9,50,38,51, the RAC (R = A or G) motif displayed the highest frequency among methylated motifs in both rice and Arabidopsis (Fig. 1g, h).

Base-resolution mRNA m6A maps from different tissues of rice and Arabidopsis

We next analyzed the distribution of m6A in the whole transcriptome for both Arabidopsis and rice. All the identified m6A sites in Arabidopsis and rice are accessible under the GEO numbers GSE245738 and GSE243722, respectively. As observed in the metagene profile, most m6A sites are highly enriched within 3′-untranslated region (3′ UTR), followed by coding DNA sequence (CDS) and 5′-untranslated region (5′ UTR) in both Arabidopsis (Fig. 2a) and rice (Supplementary Fig. 3a). Although both mammalian and plant mRNAs highly enrich m6A in the 3’ UTR (Fig. 2b), rice and Arabidopsis mRNAs harbor noticeably higher percentages of the overall m6A modifications in the 3’ UTR compared with that of human HeLa cells36 (Fig. 2b). In addition to m6A sites in the regions of 3’ UTRs and CDSs, we also observed a considerable number of m6A sites in the intronic, and 5’ UTRs regions in both rice and Arabidopsis, which are consist with past results observed in mammals37 (Supplementary Fig. 3b, c). The average m6A fraction is notably higher in intronic regions than in the 5’ UTR and CDS, but lower than that in the 3’ UTR (Supplementary Fig. 3b, c).

Fig. 2: Comprehensive base-resolution maps of m6A sites in rice and Arabidopsis.
figure 2

a Metagene profiles showing the m6A site distribution across transcripts in nine Arabidopsis tissues. Each transcript is segmented into three regions: 5′ UTR, CDS and 3′ UTR. The black dashed line represents the average m6A fraction across the nine tissues. The m6A site number (N) is indicated in the figure. The percentage (%) of overall m6A modifications distributed in 5′ UTR, CDS, and 3′ UTR regions within different tissues were shown. b Percentages of total m6A modifications distributed in 5′ UTR, CDS, and 3′ UTR regions in Hela cells, rice and Arabidopsis. We combined m6A sites from all the rice tissues and Arabidopsis tissues, respectively, to calculate percentages of total m6A fractions within 5′ UTR, CDS, and 3′ UTR regions. c, d Bar plot showing the number of tissue-unique and tissue-common m6A sites in eight rice tissues (c) and nine Arabidopsis tissues (d). The m6A site number is the overlapped m6A sites between two biological replicates for each tissue. e Metagene profiles showing common m6A sites among nine Arabidopsis tissues distributed across transcripts. f Metagene profiles of tissue-unique m6A sites among nine Arabidopsis tissues. g Metagene profiles showing the rice reproductive unique-, vegetative unique- and common m6A sites distributed across transcripts. Tissues of flower, seed and silique were grouped as reproductive tissues, while the remaining tissues were considered vegetative. Each transcript is divided into three regions: 5′ UTR, CDS and 3′ UTR. For e.g., the numbers of m6A sites are indicated in the figures.

Both the m6A/A ratio and m6A site number vary among different tissues, suggesting the presence of both tissue-shared and tissue-specific deposition of m6A methylation in plants. We, therefore, analyzed tissue-specific and commonly shared m6A modification sites among all the tissues in rice and Arabidopsis, and identified 14,953 and 6,815 shared m6A sites among all rice and Arabidopsis tissues, respectively (Fig. 2c, d). These shared m6A modification sites were predominately enriched in the 3′ UTR region relative to this tissue-specific m6A both in rice (Supplementary Fig. 3d, e) and Arabidopsis (Fig. 2e, f), which may suggest that 3′ UTR m6A sites, rather than the CDS or 5′ UTR, play general roles in maintaining plant transcriptome metabolism across the entire life cycle. To find the biological difference between reproductive- and vegetative unique m6A sites, we further divided the different tissues into reproductive tissue and vegetative tissue (see Method). The reproductive unique and vegetative unique m6A sites were identified in both rice and Arabidopsis, respectively (Supplementary Data 1). The metagene profile revealed that reproductive unique m6A sites show increased distribution in the 3′ UTR region compared to the vegetative unique one, implying the significance of m6A regulation through 3′ UTR in the reproductive phase (Fig. 2g and Supplementary Fig. 3f). GO enrichment analysis showed that genes containing reproductive unique m6A sites in both rice and Arabidopsis are enriched in similar pathways, such as reproductive structure development, embryo development, immune response, photosynthesis, and chloroplast organization (Supplementary Fig. 3g). Meanwhile genes containing vegetative unique m6A modifications are enriched in stimulus response, such as genes of ARF1, ARF7 and ARF9 involved in response to hormone stimulus in rice (Supplementary Fig. 3h), although these pathways are not significant enriched in Arabidopsis. This likely suggests tissue-specific m6A methylations play regulatory roles in plant growth regulation.

Evolutionary conservation and variability of m6A regulation across rice and Arabidopsis

Next, we investigated the evolutionary conservation of mRNA m6A modification in Arabidopsis and rice orthologous gene pairs. We found a total of 12,359 pairs of conserved m6A sites (Supplementary Data 2) in orthologous genes, while 108,856 and 226,673 m6A sites were only identified in Arabidopsis and rice, respectively (Fig. 3a). Interestingly, the average m6A fractions of unique m6A sites are much higher than the rice-Arabidopsis conserved sites in rice (Supplementary Fig. 4a) and Arabidopsis (Supplementary Fig. 4b). Amongst the conserved m6A site pairs, 7,734 pairs of conserved m6A sites exist within the same motif sequence (Supplementary Data 3, Supplementary Fig. 4c and Fig. 3b). The fractions of the conserved m6A sites within homologous genes are weakly correlated (Fig. 3c) and vary among different tissues (Fig. 3d and Supplementary Fig. 4d), suggesting that the modification levels of conserved m6A sites are tissue dependent even though they are universal among all the tissues. Since the presence of m6A is critical for normal plant development19,20,21,52, we explored functional insights about the rice-Arabidopsis conserved m6A sites. GO analysis showed that genes with conserved m6A sites are significantly enriched in stimulus response and plant development-related pathways, such as the chloroplast, photosynthesis, photomorphogenesis, embryo development, shoot morphogenesis, flower development, leaf development, root development, and ovule development (Fig. 3e). For example, the light harvest related genes, CAB3, Lhca5, LHCA3, and LHB1B2; root epidermal cell differentiation and root hair cell differentiation-related genes, POM1, GN, UBC36, GCS1, UBC35, GEM, SCN1, and MRH1; flower development genes PFT1, PS1, RDR6, DCL4, ARF8, and MET1 are all conserved in their m6A methylation sites between rice and Arabidopsis. Overall, these results provide a foundation for future studies to explore the potential roles of m6A under evolutionary pressure in plants.

Fig. 3: Evolutionary conservation and divergence in m6A regulation between rice and Arabidopsis.
figure 3

a Number of conserved and species-unique m6A sites between rice and Arabidopsis. b Motif of RAC is the most prevalent one among the conserved m6A sites. Motif sequences were divided into three groups, including RAC, GAT and others. Base “R” denotes either A or G. c Correlation analysis of the m6A fraction in the ortholog gene pair between rice and Arabidopsis. The R2 value is labeled in the figure. d Principal component analysis (PCA) of the m6A fraction in those conserved m6A sites across all tissues in rice and Arabidopsis. The modification levels of these m6A sites were normalized based on its tissue rankings within the respective species. e Gene ontology (GO) analysis of genes containing conserved m6A sites. One-sided Fisher’s exact test. Adjusted P values using the linear step-up method. f, g The ratios of standard deviation to mean for m6A levels across different tissues in rice (f) and Arabidopsis (g) are presented for various transcript regions: 5′ UTR, CDS and 3′ UTR. Generally, both species exhibit reduced variance from the 5′ UTR to 3′ UTR. Notably, rice showed a gradually reduced variance across the gene structure, while Arabidopsis showed rather stable variance in the CDS region followed by a sharp reduction in the 3′ UTR. Source data are provided as a Source Data file.

Despite the evolutionarily conserved m6A modification sites across rice and Arabidopsis, the metagene profile showed some differences, especially regarding the distribution of the m6A sites in the 3′UTR region. To illustrate these differences, we combined all the m6A sites within different tissues in rice and Arabidopsis, respectively. We calculated the ratio of standard deviation to the mean of m6A site counts. We observed that in general, the reduced variance from 5′ UTR to 3′ UTR in rice (Fig. 3f) and Arabidopsis (Fig. 3g), while the variance in the 5′ UTR region is fluctuant (Fig. 3f, g). Despite the similarity, rice showed gradually reduced variance across the gene structure, while Arabidopsis showed rather stable variance in the CDS region followed by a sharp reduction in variance in the 3′ UTR (Fig. 3f, g). Together these results may suggest differential m6A deposition regulations, especially in the CDS and 3′ UTR regions between rice and Arabidopsis.

Divergent paradigms governing m6A deposition in plant genomes

Earlier research has elucidated the distribution patterns of m6A in both mammalian and plant genomes38,49,53,54, suggesting that m6A is predominantly enriched in the last exon and long internal exons53,54. This distribution is additionally shaped by the underlying exon architecture and is regulated by the Exon Junction Complex (EJC)55,56,57. We categorize these unique distribution patterns into three basic rules: the “long exon,” “last exon,” and “exon structure” rules. Using these quantitative m6A sites at single-base resolution, we investigated whether these rules are conserved across the plant kingdom. It’s worth noting that exons in the human genome are generally longer than those in rice and Arabidopsis (Supplementary Fig. 5a-c), which could lead to differences in m6A distribution both per exon and per sliding window. To improve the accuracy of our measurements, we introduced two metrics: “m6A density,” which normalizes the total m6A level within each exon by its length, and “m6A likelihood,” which normalizes the m6A level within each sliding window by the pileup coverage of the exons in that window (Supplementary Fig. 5d).

We re-analyzed single-base mRNA m6A data from mammalian samples, specifically from the HeLa cell line36, and compared them with data from rice and Arabidopsis. We found that the m6A level per exon generally increases with internal exon length when that length is under 1000 nucleotides (nt) in both human and plant genomes (Fig. 4a-c and Supplementary Fig. 5e-g). This pattern does not hold for internal exons longer than 1000 nt in humans. Rice and Arabidopsis do not display this trend, largely because they have extremely rare internal exons exceeding 1000 nt (Fig. 4a-c). Examining “m6A density,” we observed an inverse correlation with internal exon length in both humans and plants (Fig. 4d-f and Supplementary Fig 5h-j). This suggests that m6A modifications accumulate more slowly than the exon length increases. Interestingly, overall m6A density is higher in plants than in humans, with a peak at around 100 nt in exon length. This indicates that excessively long or short exons could reduce m6A modification more effectively in plants.

Fig. 4: Divergent rules governing m6A deposition in plant and mammalian genomes.
figure 4

a-c Internal exons of human (a), rice (b), and Arabidopsis (c) transcripts were grouped into 100 bins of equal size based on their length, and the average m6A level for each bin was plotted against exon length, represented by blue dots. d-f ‘m6A density’ of each bin of the internal exons of human (d), rice (e), and Arabidopsis (f) transcripts were shown against exon length, represented by yellow dots. m6A density was calculated as the m6A level within each exon, normalized by its length and multiplied by 1,000. g-i All internal exons were aligned at their internal exon junction sites in human (g), rice (h), and Arabidopsis (i) genomes, and the overall m6A level per sliding window in the flanking regions was shown against the distance to exon junction sites, represented by brown dots. j-l Distribution of ‘m6A likelihood’ near the internal exon junction in human (j), rice (k), and Arabidopsis (l) transcripts were shown in dark green line, with 95% confidential intervals shadowed. m-o Similar to panel (a-c) but the average m6A level per exon in the last exons of human (m), rice (n), and Arabidopsis (o) genomes were shown. p-r Similar to panel (d-f) ‘m6A density’ in the last exons of human (p), rice (q), and Arabidopsis (r) genomes. s-u Similar to panel (g-i) m6A level per sliding window flanking last exon junction site in human (s), rice (t), and Arabidopsis (u) genomes. v-x Similar to panel (j-l) Distribution of ‘m6A likelihood’ near the last exon junction in human (v), rice (w), and Arabidopsis (x) genomes. y, Diagram showing the inhibition mode in humans and activation mode in plants contributes to distinct m6A distribution pattern. For j-l and v-x data are presented as median values.

To estimate the probability of m6A modifications, we aligned all exons at their junction sites and calculated the m6A levels per sliding window for both humans and plants. Regions closer to these junction sites typically have higher coverage of exons, resulting in increased m6A levels (Fig. 4g-i). However, when normalized by the coverage (Supplementary Fig. 5q-s), “m6A likelihood” showed suppression behaviors at these junction sites in the HeLa cell line, consistent with prior mammalian studies. In contrast, this pattern was not observed in rice and Arabidopsis; instead, we found a slight inverse trend. This divergence suggests that the EJC complex may passively suppress m6A deposition in mammalian cells but not in plants. The opposite trend raises the intriguing question of whether the “exon structure” rule governing m6A distribution is universally conserved in plants. Active m6A deposition pathways may also shape mRNA m6A distribution in certain plants.

Regarding the last exons, a clear correlation exists between increasing m6A levels per exon and exon length in the HeLa cell line, even when the exon length exceeds 1000 nt (Fig. 4m and Supplementary Fig. 5k). This correlation is notably absent in rice and Arabidopsis (Fig. 4n, o and Supplementary Fig. 5l, m). Furthermore, an inverse correlation between ‘m6A density’ and exon length was observed for the last exon in both humans and plants (Fig. 4p-r and Supplementary Fig. 5n-p). In contrast to internal exons, m6A is more condensed in exons around 100 nt in both humans and plants. Remarkably, the ‘m6A density’ in the last exons of rice and Arabidopsis is higher than in humans, even though the overall m6A level is lower. These findings suggest that while the general pattern of m6A enrichment in the last exon is evolutionarily conserved, significant differences exist in how this modification correlates with last exon length across species.

In terms of ‘m6A likelihood’ near the last exon junction sites, we observed an asymmetric peak downstream from these sites in both mammals and plants (Fig. 4s-u and Supplementary Fig 5t-v). Similar to internal exons, ‘m6A likelihood’ decreases as it approaches the last exon junctions in mammals (Fig. 4v). Notably, specific breakpoints in this distribution pattern occur right at the exon junctions (Fig. 4j, v). This suggests that in mammals, the EJC complex plays a significant role in suppressing m6A deposition and that overall m6A distribution is regulated by passive processes40. In contrast, rice and Arabidopsis display a pronounced peak about 300 nt downstream from last exon junction sites, followed by a decrease (Fig. 4w, x). This may indicate an active process driving m6A deposition in plants, suggesting enhanced recruitment of m6A methyltransferases to the peak region. Given that most stop codons are situated within the last exon, we aligned the exons based on their distance to the stop codon for a more nuanced analysis. We observed a pronounced breakpoint immediately adjacent to the stop codons of rice and Arabidopsis, and found that the peak of m6A likelihood is closer to the stop codon than to the starting point of the last exon (Supplementary Fig. 5w-y). These findings suggest that the active deposition of m6A in plants may be influenced by specific genomic features near the stop codon (Fig. 4y). Collectively, these insights point to potentially significant differences in m6A modification mechanisms between plant and mammalian kingdoms, indicating that there may be novel mechanisms for plant mRNA m6A deposition yet to be explored.

m6A modification enhances mRNA stability and translation mainly through 3′ UTR in Arabidopsis seedling

m6A modification promotes mRNA turnover37, mainly through 3′ UTR sites in mammals58,59,60. In contrast to mammals, the effects of m6A are less clear in plants, while several reports demonstrated that m6A modification stabilizes modified mRNAs in Arabidopsis21,27,61,62,63,64. It seems that the role of m6A modification in plants differs from that of mammals. To obtain a more accurate correlation between m6A level and mRNA turnover transcriptome-wide, we retrieved the public RNA lifetime data in Arabidopsis seedlings65 and found that m6A-modified transcripts tend to have a longer lifetime than the unmodified transcripts63 (Supplementary Fig. 6a). Plants have redundant ECT proteins that bind preferentially to m6A-modified mRNAs. We next studied the effects on mRNA degradation regulated by the m6A reader of ECT2 with the publicly available ECT2 CLIP-seq data61 and mRNA lifetime data65 of Arabidopsis seedlings. We observed considerable overlap between the m6A sites and ECT2 targets (Fig. 5a), confirming m6A binding by ECT2. ECT2 target genes containing m6A sites display higher lifetime than those not bound by ECT2 (Supplementary Fig. 6b), indicating that m6A stabilizes mRNA and that m6A readers could enhance mRNA stability in Arabidopsis seedlings.

Fig. 5: Impact of m6A modifications on mRNA stability and translation.
figure 5

a Density plot displaying the distance between the peak centers of ECT2 targets identified by CLIP-seq and m6A sites identified in Arabidopsis seedlings by m6A-SAC-seq. b Lifetime difference of transcripts with 3′ UTR-only m6A modification and non-3′ UTR m6A modification compared to transcripts without (w/o) m6A modifications. 3′ UTR, n = 1,813; non-3′ UTR, n = 2,645; w/o m6A, n = 3,319. c Cumulative curves and box plots showing the mRNA lifetime distribution for transcripts with 3′ UTR-only m6A modification. Transcripts were grouped into five categories (0,0.2); (0.2, 0.4); (0.4, 0.6); (0.6, 0.8); and (0.8, 1)) based on the sum of their m6A fractions. (0,0.2), n = 62; (0.2, 0.4), n = 228; (0.4, 0.6), n = 353; (0.6, 0.8), n = 483; (0.8, 1), n = 612. d Cumulative curves and box plots showing the mRNA lifetime distribution for transcripts with m6A modification outside the 3′ UTR. Transcripts were grouped into five categories (0,0.2); (0.2, 0.4); (0.4, 0.6); (0.6, 0.8); and (0.8, 1)) based on the sum of their m6A fractions. (0,0.2), n = 466; (0.2, 0.4), n = 981; (0.4, 0.6), n = 566; (0.6, 0.8), n = 280; (0.8, 1), n = 294. e GO enrichment analysis for genes associated with 3′ UTR-only m6A sites. f GO enrichment analysis of non-3′ UTR m6A-associated mRNAs. For e, f one-sided Fisher’s exact test. Adjusted P values using the linear step-up method. g Transcripts with 3′ UTR-only m6A sites exhibit strong positive correlations with translation efficiency. Transcripts were grouped into five categories (0,0.2); (0.2, 0.4); (0.4, 0.6); (0.6, 0.8); and (0.8, 1)) based on the sum of their m6A fractions. (0,0.2), n = 88; (0.2, 0.4), n = 293; (0.4, 0.6), n = 489; (0.6, 0.8), n = 620; (0.8, 1), n = 749. For c, d and g, the Arabidopsis seedling lifetime data set GSE86361 was used for mRNA decay analysis and Arabidopsis seedling translation efficiency data set GSE206292 was used for translation efficiency analysis. For b-d and g the p-value was determined by a one-tailed Wilcoxon rank-sum test. In box plots, the center line represents the median, and the red dot represents the mean. Upper and lower quartiles were the box limits. Source data are provided as a Source Data file.

We then asked whether the position of m6A modification could underlie mRNA stability differences as that in mammals59,60. To answer this, we first clustered the m6A sites into 3′ UTR-only m6A and non- 3′ UTR m6A, and found that 3′ UTR-only m6A significantly stabilizes mRNA, and only a slight increased lifetime was observed with genes carrying non- 3′ UTR m6A compared to genes without (w/o) m6A modification (Fig. 5b). This likely suggests a more predominant role of 3′ UTR-only m6A in regulating mRNA stability, especially stabilizing the modified transcripts. To further explore the correlation between m6A fraction levels and mRNA stability, we divided mRNA carrying the 3′ UTR-only and non- 3′ UTR m6A sites into five groups based on m6A levels, and found that higher m6A fractions are associated with higher mRNA stability in those genes bearing the 3′ UTR-only m6A sites (Fig. 5c), while no significant correlations were observed in genes containing non-3′ UTR m6A sites (Fig. 5d and Supplementary Fig. 6c). The above results were further confirmed with the mRNA metabolic data from Sorenson, R. S et al. 66 (Supplementary Fig. 6d, e). Although different reader proteins can recognize m6A in different regions to exert either stabilization or decay function, given that a majority of mRNA m6A modification enriches in the 3′ UTR (Fig. 2a, b), this observation may suggest an overall mRNA stabilization effect by m6A in plants.

The distinct effects of m6A position in controlling mRNA stability prompted us to investigate whether m6A position underlies biological function differences. The GO enrichment analysis showed that genes associated with the 3′ UTR-only m6A are significantly enriched in general biological pathways like gene expression, RNA processing, and ribosome biogenesis (Fig. 5e), while non-3′ UTR m6A-associated mRNAs were enriched in more specific pathways, such as response to hormone, response to fungus, response to salt stress, leaf development, and plant organ senescence (Fig. 5f). In addition, we also observed a positive correlation between m6A levels and translation efficiency in transcripts modified with 3′ UTR-only m6A (Fig. 5g), but no significant correlations were noticed in transcripts bearing non-3′ UTR m6A sites (Supplementary Fig. 6f). These observations suggest more diverse effects of m6A in plant, affected by its location and downstream binding proteins.

m6A installed by MTA in the chloroplast transcriptome in Arabidopsis seedlings

We next profiled the m6A sites in Arabidopsis mta mutant seedlings (Supplementary Fig. 6g, h) using m6A-SAC-seq and compared the m6A sites with those of WT (col) seedlings. A total of 14,125 MTA-dependent m6A sites within 2,894 RNAs were identified in WT. The methylation levels of these m6A sites were noticeably reduced (10,505 m6A sites) or completely abolished (3,621 m6A sites) in mta compared with WT; these m6A sites were hereafter defined as MTA-dependent m6A sites. We noticed that the MTA-dependent m6A sites tend to be more preferentially located in the 3’ UTR region than CDS and 5’ UTR regions (Fig. 6a). GO enrichment analysis showed that mRNAs containing these m6A sites are mostly enriched in stimulus response, chloroplast, photosynthesis membrane and post-embryonic development (Fig. 6b). This is consistent with previous studies in Arabidopsis that MTA is involved in response to salt stress26, blue light response21 and embryonic development15. Notably, m6A levels of chloroplast (538 genes) and photosynthesis membrane (59 genes) related transcripts were significantly reduced in mta, among which twenty m6A sites showed dramatically reduced m6A levels in the chloroplast transcriptome (Fig. 6c). The methylation levels of the chloroplast encoded transcripts vary in different tissues (Fig. 6c), suggesting dynamic MTA-dependent m6A modifications in the chloroplast transcriptome across Arabidopsis life cycle.

Fig. 6: The MTA-mediated m6A modification in chloroplast transcriptome of Arabidopsis seedlings.
figure 6

a, Distribution of MTA-dependent m6A sites along transcripts in Arabidopsis seedlings. b, GO enrichment analysis of mRNAs containing MTA-dependent m6A sites. One-sided Fisher’s exact test. Adjusted P values using the linear step-up method. c, Methylation level heatmap of the chloroplast-encoded genes across different tissues. The position of m6A site on chloroplast-encoded genes was shown. m6A levels and fold change were also shown. Source data are provided as a Source Data file.

Next, we observed overall reduced translation efficiency in the mta mutant compared to that of WT using previously published datasets26 (Supplementary Fig. 6i). However, similar numbers of genes with upregulated translation efficiency (332 genes, fold change >2, p < 0.05) and downregulated translation efficiency (257 genes, fold change <0.5, p < 0.05) were observed in mta mutant relative to WT control. The effect of m6A installed by MTA on translation efficiency appear to be heterogeneous in Arabidopsis seedlings, resembling to that observed in mammals67. Perhaps consistently, the above differentially translated genes are also enriched in distinct GO terms. Genes with upregulated translation efficiency are mainly enriched in general pathways like the ribosome and nucleolus (Supplementary Fig. 6j), while genes with downregulated translation efficiency are specifically enriched in the chloroplast and photosynthesis membrane (Supplementary Fig. 6k), such as genes of FIBRILLIN 4 (FIB4) and SMO268. Collectively, MTA deposits m6A modifications both in the nuclear and chloroplast transcriptomes, which regulate photosynthesis.

Light-induced feedback regulation of the circadian clock through m6A

m6A methylation of mRNAs regulates the circadian clock in both plants21 and mammals69. To further probe the light effect on rice m6A methylation, we germinated rice seeds under dark (24 h of dark per 24 h) and light conditions (16 h of light per 24 h), respectively. Plumules under dark and light conditions at 3 days after germination were studied, respectively, by using m6A-SAC-seq (Fig. 7a). Our data revealed a pervasive increase of m6A methylation levels (23,253 hypermethylated and 2,607 hypomethylated m6A sites) under light compared to dark conditions (Supplementary Fig. 7a-c). The light-induced m6A sites are largely outside of 3′ UTR (Fig. 7b). GO enrichment analysis revealed that genes containing hypermethylated m6A sites are highly clustered in stimulus response, including light and hormone stimulus (Supplementary Fig. 7d). Noticeably, light significantly increased the m6A modification levels of photoreceptor transcripts, for example, PHYA (Chr3: 29172686; CDS; AGATA), PHYB (Chr3: 11021272; CDS; AGATA), PHYC, (Chr3:31007707;CDS; GGACA), CRY1a (Chr2: 21976854;5’ UTR; AGAGC), and CRY2 (Chr2:24921916; 3’ UTR; AAACT). Thus, m6A methylation levels of circadian clock genes within their transcripts were notably updated by light in rice, as the case observed in mammals70. Moreover, in the aforementioned mta mutant of Arabidopsis, a decreased m6A fraction of CRY1 in the 3′ UTR (Chr4:5727183; AAACA; 3′ UTR) was observed, potentially resulting in reduced translation efficiency, suggesting that MTA regulates CRY1 translation through m6A deposition (Fig. 7c, d). Given that light-induced phase separation of CRYs modulates MTA activity in plants69, and MTA controls m6A modification on CRY transcripts to regulate CRY translation, there appears to be a feedback loop of epitranscriptome-translation regulation of the circadian clock in plants.

Fig. 7: Light responsive regulation through m6A in rice and Arabidopsis.
figure 7

a Light increased the m6A fractions in the plumule of rice. Data are means ± SD, n = 2. The p-value was determined by a one-tailed Wilcoxon rank-sum test. b Relative distribution of m6A sites across 5′ UTR, CDS and 3′ UTR regions. c Reduced m6A fraction in CRY1 was observed in mta mutant. Data are means, n = 2. d Decreased translation efficiency of CRY1 in the mta mutant. Translation efficiency data set GSE206292 was used for the analysis. Data are means, n = 2. Student t-test was used to determine the statistic difference. Source data are provided as a Source Data file.

Discussion

m6A methylation of mRNA plays critical roles in both plant and mammal development as well as signaling and stimulation responses15,20,38,71,72. Previous RNA m6A studies in plants lack base-resolution, precision and modification stoichiometry information5,23,48,63. Benefit from the development of m6A sequencing at single-base resolution in mammalian transcriptome using the m6A-SAC-seq method for the first time36,37, we report here comprehensive maps of m6A at single-base precision with stoichiometry information in eight rice tissues and nine Arabidopsis tissues spanning their life cycle. We uncovered high-confidence, single-base resolution m6A sites across rice and Arabidopsis tissues, providing in-depth resources for future investigations of m6A functions in rice and Arabidopsis.

Evolutionarily conserved m6A modification sites in orthologous gene pairs of Arabidopsis and rice were identified. Orthologous genes containing conserved m6A sites are significant for tissue development, photosynthesis and stimulus response, and might be selected under evolutionary pressure. Despite these conserved m6A sites in orthologous genes, we observed different m6A distribution patterns in 3′ UTR between rice and Arabidopsis. Furthermore, although a positive correlation between the total level of m6A and the internal exon length was observed, different from that in mammals, no such correlation was seen regarding to the last exon in plants. While the mRNA m6A distribution in mammals appears to be shaped by repressive pathways mediated through the exon junction complexes or other protein factors55,56,57, plants, by contrast, also rely on active installation to control m6A deposition. In particular, our results suggest an active m6A deposition process occurring near the stop codon in plant mRNAs. This suppression-activation dual deposition regulation potentially depicts m6A distribution patterns across species. We and others have shown that m6A suppression is mediated by EJC in mammals55,56,57, the suppression-activation model predicts RNA-binding proteins that may recruit m6A writers and direct m6A deposition near the stop codon in plant mRNAs. These observations suggest an as-yet-unknown mechanism that regulates m6A distribution in plant transcriptomes, which requires future in-depth investigations.

A majority of plant m6A modification resides in 3′ UTR from our analyses. Arising evidence in plants suggested that m6A modification stabilizes mRNA21 although mammalian m6A modification tends to destabilize the modified mRNA37. Based on our single-base resolution results in Arabidopsis seedlings, we observed a positive correlation between m6A methylation level and mRNA stability for 3’ UTR m6A sites, confirming an overall mRNA stabilization effect by m6A in certain plant tissues. Consistently, the m6A sites bound by its reader protein ECT2 exhibit significantly elevated half lifetime as compared with transcripts not bound by ECT2, indicating the presence of m6A reader protein to enhance mRNA stability in plants61,62. In addition, m6A methylation installed by MTA can either promote or reduce translation efficiency in a pathway-dependent manner, resembling observations made in mammals60. Therefore, m6A modification stabilizes modified transcripts, with reader proteins stabilizing the bound mRNA in Arabidopsis seedlings. Translation effects can be complex and context-dependent but an overall translation promotion effect in Arabidopsis seedling was suggested from our data.

In conclusion, these base-resolution and quantitative m6A modification maps across rice and Arabidopsis’ life cycle have filled a pronounced gap in plant research. The comparative analysis of single-base m6A maps between humans and plants reveals a suppression-activation dual regulation model in shaping m6A distribution patterns in different species.

Methods

Plant material

Col-0 accession of Arabidopsis thaliana and japonica rice (Oryza sativa) cultivar Nipponbare were used in this study. Arabidopsis, plants were grown at 22 °C with 16 h of light per 24 h. Arabidopsis seedlings were harvested after growing on 1/2 Murashige and Skoog medium (MS) plates for 7 d. Arabidopsis shoots and roots were harvested after growing on 1/2 Murashige and Skoog medium (MS) plates for 14 d. Arabidopsis Rosetta leaves were harvested after growing in soil for 30 d. Arabidopsis cauline leaf, flower, stem and silique were harvested after flowering. Arabidopsis seeds were collected after the seeds is totally dry. Seeds of Arabidopsis mta mutant (ABI3::MTA/mta)20,21 were sowed in the MS plates, and the mta seedlings were harvested at 8-day-old at 22 °C with 16 h of light per 24 h. For rice, plants were grown at 28 °C with 14 h of light per 24 h and 8-day-old, 2-week-old seedlings were harvested. The heading panicles, flag leaf at 10 days after anthesis, endosperms and embryos at 10 days after anthesis were harvested. The plumules under dark and light conditions at 3 days after germination were sampled. The tissue was flash-frozen in liquid nitrogen, ground using a mortar and pestle and stored at −80 °C. Total RNA were extracted using TRIzol™ Reagent (Catalog number: 15596026) according to the manufacturer′s instructions. All the plants were planted in the greenhouse of The Chinese University of Hong Kong.

mRNA capture from the extracted total RNA

A total of 50 μg total RNA for each of the two biological replicates was used for mRNA capture (Dynabeads mRNA DIRECT Purification Kit (Invitrogen)) following the manufacturer′s instructions with modification. Briefly, 50 ug total RNA was diluted with H2O in 100 ul volume and then denatured under 65 °C for 2 min, and immediately put on the ice for exactly 2 min. Next, a total of 100 ul Dynabeads were washed twice with 200 ul lysis/binding buffer provided in the above kit. The washed beads were then eluted in 100 ul lysis/binding buffer, which was further mixed with the denatured total RNA. Then, the sample was put on the rotor to bind for 15 min at room temperature. After binding, wash buffer B was used to wash the beads for twice. 30 ul H2O was used to elute the beads and immediately put on the 75 °C for 2 min. The eluted mRNA was collected after magnetic separation. The mRNA capture process was repeated as described above to obtain the more purified mRNA.

Quantification of m6A in RNA by LC–MS/MS

50 ng mRNAs were digested into nucleosides, and the amount of m6A was measured by using Agilent 6460 Triple Quad MS–MS with a 1290 UHPLC supplied with a ZORBAX Eclipse XDB-C18 column (UHPLC–QQQ–MS/MS) and calculated based on the standard curve generated by pure standards. For each sample, RNA was digested by using nuclease P1 (NEB) at 37 °C for 2 h. Then, 1 μl of Shrimp Alkaline Phosphatase (rSAP) and 3 μl of 10× rCutsmart buffer (NEB) were added, and the reaction was incubated at 37 °C for 2 h. Samples were then filtered using a 0.22-μm filter (Millipore) and injected into LC–MS/MS. The nucleosides were quantified by using the nucleoside-to-base ion mass transitions of 282 to 150 (m6A), and 268 to 136 (A). Quantification was performed in comparison to the standard curve obtained from pure nucleoside standards run on the same batch of samples. The ratio of m6A to A was calculated based on the calibrated concentrations.

m6A-SAC-seq library construction

50 ng mRNAs of each replicate were used for the library construction. All these libraries were constructed exactly following the previously published protocols36,37. The constructed libraries were sequenced on the Illumina HiSeq sequencing platform in pair-end mode with 150 bp per read.

m6A-SAC-seq data processing

After sequencing, the m6A sites were detected using the method of (https://github.com/y9c/m6A-SACseq)36,37. The analysis utilized the reference genome downloaded from the Ensemble database, with assembly versions TAIR10 and IRGSP-1.0 employed for Arabidopsis and rice respectively.

RNA lifetime profiling and translation data analysis

RNA lifetime and translation efficiency data of Arabidopsis Col seedling was downloaded from data set GSE206292 and GSE118462. The translation efficiency data of mta mutant in Arabidopsis was also retrieved from GSE206292.

Conserved m6A sites in ortholog genes between rice and Arabidopsis

The one-to-one ortholog genes between rice and Arabidopsis were first identified. The ortholog genes were then pairwise aligned to obtain the sites on the homolog positions which showed consistent flanking sequence (±1 nt) centered on A sites. The above A sites were defined as the conserved A sites. While the m6A modification on conserved A sites for both species are defined as conserved m6A sites.

Gene Ontology (GO) analysis

Functional GO enrichment analysis was performed by web-based toolkit for the agricultural community agriGO v2.038 (http://systemsbiology.cau.edu.cn/agriGOv2/). GO terms with a false discovery rate (FDR) < 0.05 were considered significantly enriched.

Statistics and reproducibility

All experiments were repeated independently at least twice and showed similar results. GraphPad Prism v.9 and R studio were deployed for the figure plotting.

Reporting summary

Further information on research design is available in the Nature Portfolio Reporting Summary linked to this article.