Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

DHH1/DDX6-like RNA helicases maintain ephemeral half-lives of stress-response mRNAs

Abstract

Gene transcription is counterbalanced by messenger RNA decay processes that regulate transcript quality and quantity. We show here that the evolutionarily conserved DHH1/DDX6-like RNA hellicases of Arabidopsis thaliana control the ephemerality of a subset of cellular mRNAs. These RNA helicases co-localize with key markers of processing bodies and stress granules and contribute to their subcellular dynamics. They function to limit the precocious accumulation and ribosome association of stress-responsive mRNAs involved in auto-immunity and growth inhibition under non-stress conditions. Given the conservation of this RNA helicase subfamily, they may control basal levels of conditionally regulated mRNAs in diverse eukaryotes, accelerating responses without penalty.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Arabidopsis RH6, RH8 and RH12 overlap in their contribution to growth and development.
Fig. 2: RH6, RH8 and RH12 form cytoplasmic mRNPs contributing to PB and SG formation.
Fig. 3: Attenuation of RH6, RH8 and RH12 function shifts the seedling steady-state transcriptome and translatome from a general growth to stress-responsive state.
Fig. 4: RHs facilitate decay of short-lived mRNAs.
Fig. 5: Stabilization of stress-response mRNAs in the rh6812 genotype is associated with an increase in abundance and translational status.
Fig. 6: The triple rh6812 mutant exhibits a constitutive immune response.

Similar content being viewed by others

Data availability

Sequence data are deposited in GEO accession no. GSE136713. All other data needed to evaluate the conclusions are in the Supplementary Information. Source Data for Figs. 1, 2 and 6, and Extended Data Figs. 2–5 and 7 are provided with the paper. Sequence Read Archive, http://www.ncbi.nlm.nih.gov/sra; NCBI Gene Expression Omnibus, https://www.ncbi.nlm.nih.gov/geo/.

References

  1. Chantarachot, T. & Bailey-Serres, J. Polysomes, stress granules, and processing bodies: a dynamic triumvirate controlling cytoplasmic mRNA fate and function. Plant Physiol. 176, 254–269 (2018).

    Article  CAS  PubMed  Google Scholar 

  2. Sheth, U. & Parker, R. Decapping and decay of messenger RNA occur in cytoplasmic processing bodies. Science 300, 805–808 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Hubstenberger, A. et al. P-body purification reveals the condensation of repressed mRNA regulons. Mol. Cell 68, 144–157 (2017).

    Article  CAS  PubMed  Google Scholar 

  4. Sorenson, R. & Bailey-Serres, J. Selective mRNA sequestration by OLIGOURIDYLATE-BINDING PROTEIN 1 contributes to translational control during hypoxia in Arabidopsis. Proc. Natl Acad. Sci. USA 111, 2373–2378 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Merret, R. et al. Heat-shock protein HSP101 affects the release of ribosomal protein mRNAs for recovery after heat shock. Plant Physiol. 174, 1216–1225 (2017).

  6. Zhang, W., Murphy, C. & Sieburth, L. E. Conserved RNase II domain protein functions in cytoplasmic mRNA decay and suppresses Arabidopsis decapping mutant phenotypes. Proc. Natl Acad. Sci. USA 107, 15981–15985 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Sorenson, R. S., Deshotel, M. J., Johnson, K., Adler, F. R. & Sieburth, L. E. Arabidopsis mRNA decay landscape arises from specialized RNA decay substrates, decapping-mediated feedback, and redundancy. Proc. Natl Acad. Sci. USA 115, E1485–E1494 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Siwaszek, A., Ukleja, M. & Dziembowski, A. Proteins involved in the degradation of cytoplasmic mRNA in the major eukaryotic model systems. RNA Biol. 11, 1122–1136 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  9. Nagarajan, V. K., Jones, C. I., Newbury, S. F. & Green, P. J. XRN 5′→3′ exoribonucleases: structure, mechanisms and functions. Biochimica Biophys. Acta Gene Regul. Mech. 1829, 590–603 (2013).

    Article  CAS  Google Scholar 

  10. Xu, J., Yang, J.-Y., Niu, Q.-W. & Chua, N.-H. Arabidopsis DCP2, DCP1, and VARICOSE form a decapping complex required for postembryonic development. Plant Cell 18, 3386–3398 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Goeres, D. C. et al. Components of the Arabidopsis mRNA decapping complex are required for early seedling development. Plant Cell 19, 1549–1564 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Xu, J. & Chua, N.-H. Arabidopsis decapping 5 is required for mRNA decapping, P-body formation, and translational repression during postembryonic development. Plant Cell 21, 3270–3279 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Perea-Resa, C. et al. LSM proteins provide accurate splicing and decay of selected transcripts to ensure normal Arabidopsis development. Plant Cell 24, 4930–4947 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Roux, M. E. et al. The mRNA decay factor PAT1 functions in a pathway including MAP kinase 4 and immune receptor SUMM2. EMBO J. 34, 593–608 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Martínez de Alba, A. E. et al. In plants, decapping prevents RDR6-dependent production of small interfering RNAs from endogenous mRNAs. Nucleic Acids Res. 43, 2902–2913 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  16. Xu, J. & Chua, N.-H. Dehydration stress activates Arabidopsis MPK6 to signal DCP1 phosphorylation. EMBO J. 31, 1975–1984 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Soma, F. et al. ABA-unresponsive SnRK2 protein kinases regulate mRNA decay under osmotic stress in plants. Nat. Plants 3, 16204 (2017).

    Article  CAS  PubMed  Google Scholar 

  18. Perea-Resa, C. et al. The LSM1-7 complex differentially regulates Arabidopsis tolerance to abiotic stress conditions by promoting selective mRNA decapping. Plant Cell 28, 505–520 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Gloggnitzer, J. et al. Nonsense-mediated mRNA decay modulates immune receptor levels to regulate plant antibacterial defense. Cell Host Microbe 16, 376–390 (2014).

    Article  CAS  PubMed  Google Scholar 

  20. Presnyak, V. & Coller, J. The DHH1/RCKp54 family of helicases: an ancient family of proteins that promote translational silencing. Biochimica Biophys. Acta Gene Regul. Mech. 1829, 817–823 (2013).

    Article  CAS  Google Scholar 

  21. Kami, D. et al. The DEAD-box RNA-binding protein DDX6 regulates parental RNA decay for cellular reprogramming to pluripotency. PLoS ONE 13, e0203708 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  22. Boag, P. R., Atalay, A., Robida, S., Reinke, V. & Blackwell, T. K. Protection of specific maternal messenger RNAs by the P body protein CGH-1 (Dhh1/RCK) during Caenorhabditis elegans oogenesis. J. Cell Biol. 182, 543–557 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Kamenska, A. et al. The DDX6-4E-T interaction mediates translational repression and P-body assembly. Nucleic Acids Res. 44, 6318–6334 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Wang, M. et al. ME31B globally represses maternal mRNAs by two distinct mechanisms during the Drosophila maternal-to-zygotic transition. eLife 6, e27891 (2017).

  25. Coller, J. M., Tucker, M., Sheth, U., Valencia-Sanchez, M. A. & Parker, R. The DEAD box helicase, Dhh1p, functions in mRNA decapping and interacts with both the decapping and deadenylase complexes. RNA 7, 1717–1727 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Coller, J. & Parker, R. General translational repression by activators of mRNA decapping. Cell 122, 875–886 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Radhakrishnan, A. et al. The DEAD-box protein Dhh1p couples mRNA decay and translation by monitoring codon optimality. Cell 167, 122–132 (2016).

  28. Zhang, X. et al. Suppression of endogenous gene silencing by bidirectional cytoplasmic RNA decay in Arabidopsis. Science 348, 120–123 (2015).

    Article  CAS  PubMed  Google Scholar 

  29. Hondele, M. et al. DEAD-box ATPases are global regulators of phase-separated organelles. Nature 573, 144–148 (2019).

  30. Mustroph, A. et al. Profiling translatomes of discrete cell populations resolves altered cellular priorities during hypoxia in Arabidopsis. Proc. Natl Acad. Sci. USA 106, 18843–18848 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Riehs-Kearnan, N., Gloggnitzer, J., Dekrout, B., Jonak, C. & Riha, K. Aberrant growth and lethality of Arabidopsis deficient in nonsense-mediated RNA decay factors is caused by autoimmune-like response. Nucleic Acids Res. 40, 5615–5624 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Feys, B. J., Moisan, L. J., Newman, M. A. & Parker, J. E. Direct interaction between the Arabidopsis disease resistance signaling proteins, EDS1 and PAD4. EMBO J. 20, 5400–5411 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Cui, H. et al. A core function of EDS1 with PAD4 is to protect the salicylic acid defense sector in Arabidopsis immunity. New Phytol. 213, 1802–1817 (2017).

    Article  CAS  PubMed  Google Scholar 

  34. Baek, W., Lim, C. W. & Lee, S. C. A DEAD-box RNA helicase, RH8, is critical for regulation of ABA signalling and the drought stress response via inhibition of PP2CA activity. Plant Cell Environ. 41, 1593–1604 (2018).

    Article  CAS  PubMed  Google Scholar 

  35. Chicois, C. et al. The UPF1 interactome reveals interaction networks between RNA degradation and translation repression factors in Arabidopsis. Plant J. 96, 119–132 (2018).

  36. Sulkowska, A. et al. RNA helicases From the DEA(D/H)-box family contribute to plant NMD efficiency. Plant Cell Physiol. 61, 144–157 (2019).

  37. Protter, D. S. W. & Parker, R. Principles and properties of stress granules. Trends Cell Biol. 26, 668–679 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Luo, Y., Na, Z. & Slavoff, S. A. P-bodies: composition, properties, and functions. Biochemistry 57, 2424–2431 (2018).

    Article  CAS  PubMed  Google Scholar 

  39. Serman, A. et al. GW body disassembly triggered by siRNAs independently of their silencing activity. Nucleic Acids Res. 35, 4715–4727 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Minshall, N., Kress, M., Weil, D. & Standart, N. Role of p54 RNA helicase activity and its C-terminal domain in translational repression, P-body localization and assembly. Mol. Biol. Cell 20, 2464–2472 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  41. Ayache, J. et al. P-body assembly requires DDX6 repression complexes rather than decay or Ataxin2/2L complexes. Mol. Biol. Cell 26, 2579–2595 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Jang, G.-J., Yang, J.-Y., Hsieh, H.-L. & Wu, S.-H. Processing bodies control the selective translation for optimal development of Arabidopsis young seedlings. Proc. Natl Acad. Sci. USA 116, 6451–6456 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Borrelli, G. M. et al. Regulation and evolution of NLR genes: a close interconnection for plant immunity. Int. J. Mol. Sci. 19, 1662 (2018).

  44. Lai, Y. & Eulgem, T. Transcript-level expression control of plant NLR genes. Mol. Plant Pathol. 19, 1267–1281 (2018).

    Article  PubMed  Google Scholar 

  45. Zhang, X.-N. et al. Transcriptome analyses reveal SR45 to be a neutral splicing regulator and a suppressor of innate immunity in Arabidopsis thaliana. BMC Genomics 18, 772 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  46. Tsuchiya, T. & Eulgem, T. An alternative polyadenylation mechanism coopted to the Arabidopsis RPP7 gene through intronic retrotransposon domestication. Proc. Natl Acad. Sci. USA 110, E3535–E3543 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Boccara, M. et al. The Arabidopsis miR472-RDR6 silencing pathway modulates PAMP- and effector-triggered immunity through the post-transcriptional control of disease resistance genes. PLoS Pathog. 10, e1003883 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  48. Wu, Z. et al. Regulation of plant immune receptor accumulation through translational repression by a glycine-tyrosine-phenylalanine (GYF) domain protein. eLife 6, e23684 (2017).

  49. Yu, X. et al. Orchestration of processing body dynamics and mRNA decay in Arabidopsis Immunity. Cell Rep. 28, 2194–2205 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Sharif, H. et al. Structural analysis of the yeast Dhh1-Pat1 complex reveals how Dhh1 engages Pat1, Edc3 and RNA in mutually exclusive interactions. Nucleic Acids Res. 41, 8377–8390 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Lumb, J. H. et al. DDX6 represses aberrant activation of interferon-stimulated genes. Cell Rep. 20, 819–831 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. van Wersch, R., Li, X. & Zhang, Y. Mighty dwarfs: Arabidopsis autoimmune mutants and their usages in genetic dissection of plant immunity. Front. Plant Sci. 7, 1717 (2016).

    PubMed  PubMed Central  Google Scholar 

  53. Arribere, J. A., Doudna, J. A. & Gilbert, W. V. Reconsidering movement of eukaryotic mRNAs between polysomes and P bodies. Mol. Cell 44, 745–758 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank J. Parker for kindly providing eds1-2 and pad4-1 seeds, M. Crespi for rdr6sgs2-1 and sgs3-1 and Y. Watanabe for proDCP2:DCP2-GFP in dcp2-1. We thank C. Bousquet-Antonelli for sharing RH6/8/12 antisera. We thank the members of the Bailey-Serres and Sieburth laboratories for discussion. This research was supported by the United States National Science Foundation (grant nos. MCB-1021969 to J.B.-S. and L.E.S. and MCB-1716913 to J.B.-S.) and a UC MacArthur Foundation Chair award to J.B.-S. T.C. was supported by a Royal Thai Government Development and Promotion of Science and Technology Talents Project scholarship.

Author information

Authors and Affiliations

Authors

Contributions

T.C., R.S.S., L.S. and J.B.-S. conceived and designed experiments. T.C., R.S., M.H., H.K., A.T.K. and K.A. performed experiments. D.C. contributed to the preparation of genetic material. K.D., T.E., L.S. and J.B.-S. provided reagents. T.C., R.S.S., M.H., L.S. and J.B.-S. analysed data. T.C., R.S.S., L.S. and J.B.-S. wrote the manuscript.

Corresponding author

Correspondence to Julia Bailey-Serres.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature Plants thanks Yuichiro Watanabe and the other, anonymous, reviewers for their contribution to the peer review of this work.

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

Extended data

Extended Data Fig. 1 Phylogenetic relationship and schematic representation of the primary sequence structure of eukaryotic DHH1/DDX6-like DEAD-box RNA helicases.

The evolutionary history was inferred from 58 representative DHH1/DDX6-like helicases across different eukaryotic lineages. The sequences were aligned using the Muscle algorithm, and the phylogenetic tree was generated with the Maximum Likelihood method with 1000 bootstrap replicates. Evolutionary rate differences among sites were modeled with a discrete Gamma distribution. The tree is shown to scale, with branch lengths measured as the number of substitutions per site. Black rectangular boxes on the tree branches depict bootstrap values greater than 50%, with the size of the boxes proportional to the bootstrap values. Grey boxes represent the N- and C-terminal extensions; blue boxes, RecA-like domains; yellow boxes, linker regions.

Extended Data Fig. 2 Characterization of Arabidopsis rh6-1, rh8-1 and rh12-2 mutants and their higher-order mutant combinations.

a, Schematic representation of the RH6, RH8 and RH12 loci and their associated T-DNA insertion alleles. Untranslated and coding regions are depicted as white and black boxes, respectively. Introns are depicted as lines. ‘ATG’ and ‘TAA’ indicate the translational initiation and stop codon, respectively. Inverted white triangles refer to the positions of T-DNA insertions in the rh6-1, rh8-1 and rh12-2 alleles. Arrows indicate the orientation and approximate position of primers used for molecular genotyping of the genes or the T-DNAs: blue, forward primers; red, reverse primers. b, PCR-based genotyping of the homozygous rh6-1, rh8-1 and rh12-2 alleles. Gene and T-DNA-specific primers represented in a were used for PCR amplification of genomic DNA from Col-0 and homozygous rh6-1, rh8-1 and rh12-2 insertion alleles. Results validated with other primers. c, Reverse transcriptase (RT)-PCR analysis of RH6, RH8 and RH12 transcripts. Primers flanking the T-DNAs as indicated in a were used for PCR-based specific detection of RH6, RH8 and RH12 transcripts from the homozygous rh6-1, rh8-1 and rh12-2 alleles in comparison to Col-0. UBQ10 was used as an internal control. d, RT-qPCR analysis of RH6, RH8 and RH12 transcript levels in 7-day-old seedlings of the homozygous rh6-1, rh8-1 and rh12-2 mutants. Gene-specific primers located downstream of the T-DNA as indicated in a were used. Relative transcript fold-change was calculated using PP2AA2 as a reference. Error bars, s.d. (n = 3). Statistical significance was determined by ANOVA, followed by Tukey’s HSD test. Means that are significantly different from each other (P < 0.05) are denoted by different letters. See Source Data for fold change and P values. e, Western blot analysis of RH6, RH8 and RH12 protein levels in 5-day-old seedlings of the homozygous rh6-1, rh8-1 and rh12-2 mutants. f, Western blot analysis of RH6, RH8 and RH12 protein levels in rosette leaves of 32-day-old plants in the double (rh68, rh612 and rh812), double homozygous hemizygous (rh6+/−812, rh68+/−12 and rh6812+/−) and triple (rh6812) mutants. For e,f, each lane was loaded with an equal quantity of protein from a crude homogenate, RH specific antisera were used20. The molecular weight marker (M) ladder and Ponceau S staining as loading control were imaged in visible light. Data are representative of two experiments.

Source data

Extended Data Fig. 3 Reduced expression of RH6, RH8 and RH12 affects plant growth.

a, Schematic diagram and sequences of a 21-nucleotide artificial miRNA and its target site on RH6 expressed in the Arabidopsis MIR319a backbone under the 35S promoter. b, Levels of the artificial amiRH6 in 7-day-old seedlings of three independent amiRH6 lines generated in rh812 background (amiRH6 rh812) determined by pulsed stem-loop RT-qPCR. miRNA fold-change was calculated relative to line #21 using U6 RNA as a reference. Error bars, s.d. (n = 3); n.d., not detectable. c, RT-qPCR analysis of RH6, RH8 and RH12 transcript levels in 7-day-old seedlings of Col-0, rh812 and three amiRH6 rh812 lines. Relative transcript fold-change was calculated using PP2AA2 as a reference. Error bars, s.d. (n = 3). d, Representative rosette growth phenotype of 28-day-old plants of Col-0, rh812 and the amiRH6 rh812 lines. e,, f, Rosette diameters (n = 24) and fresh weights (n = 18-24) of 28-day-old plants of the genotypes presented in d. g, h, Phenotype (g) and primary root lengths (h, n = 27) of 7-day-old seedlings of the wild-type Col-0, the double rh812 mutant and the amiRH6 rh812 lines. Statistical significance was determined by ANOVA, followed by Tukey’s HSD test. Means that are significantly different from each other (P < 0.05) are denoted by different letters. For e, f and h, boxplot boundaries represent the first and third quartiles; a horizontal line divides the interquartile range, median; red diamond, mean. See Source Data for exact sample sizes, fold change or P values for b, c, e, f and h.

Source data

Extended Data Fig. 4 Complementation of the rh6812 phenotype by a genomic fragment of RH6 with a C-terminal FLAG epitope tag (gRH6-FLAG).

a, Rosette growth phenotype of three independent rh6812 gRH6-FLAG lines in the T2 population. 28-day-old plants are shown in comparison to the triple rh6812 mutant and the wild-type Col-0. The rh6812 genotype does not flower (0/36 plants), whereas rh6812 gRH6-FLAG plants are fertile (36/36 plants). Lower panels are magnified images of framed areas in the upper panels. b, Western blot analysis of the RH6-FLAG protein in three independent rh6812 gRH6-FLAG lines presented in a using anti-FLAG M2 antibody. Numbers represent individual plants in the T2 population. c-e, RH6, RH8 and RH12 protein abundance in the rh6812 and rh6812 gRH6-FLAG-22 genotypes. Western blot analysis of RH6 (c, upper panel), RH8 (d, upper panel), RH12 (c, upper panel) and RPS6 (a, b and c, lower panel). Protein levels were determined in triplicate (n = 3) in 5-day-old seedling tissues of Col-0 pro35S:HF-GFP-RPL18 and rh6812 pro35S:HF-GFP-RPL18 and 7-day-old seedling tissues of rh6812 gRH6-FLAG-22 homozygotes. RH specific antisera were used. Data are representative of two experiments.

Source data

Extended Data Fig. 5 The triple rh6812 mutant phenotype is RNA-DEPENDENT RNA POLYMERASE 6 and SUPPRESSOR OF GENE SILENCING 3 independent.

a, Representative rosette growth phenotype of 28-day-old plants of Col-0, single rdr6sgs2-1 and sgs3-1 mutants, triple rh6812 mutant, and quadruple rh6812 rdr6sgs2-1 and rh6812 sgs3-1 mutants. b, Rosette diameter and fresh weight of 28-day-old plants of the genotypes presented in a. Boxplot boundaries represent the first and third quartiles; a horizontal line divides the interquartile range, median; red diamond, mean. Statistical significance was determined by ANOVA (n = 30), followed by Tukey’s HSD test. Data were log-transformed. Means that are significantly different from one another (P < 0.05) are denoted by different letters. Data are representative of two experiments. See Source Data for P values.

Source data

Extended Data Fig. 6 The triple rh6812 mutant transcriptome and translatome are enriched with stress/defense-responsive mRNAs but depleted of those required for growth and development.

a, Volcano plots of change in Total and TRAP mRNA abundance of rh6812 relative to Col-0 based on differential abundance analysis by edgeR. The log2 fold-change (FC) is shown on the x-axis, and the negative log10 of the false-discovery rate (FDR) is shown on the y-axis. Genes with |log2 FC| ≥ 1 and FDR < 0.01 were considered differentially expressed. The total number of genes in the analysis is given in parentheses. b, GO functional categories (biological process) of gene transcripts enriched or depleted in rh6812 relative to Col-0 as identified in a. Twenty non-redundant terms determined by a hypergeometric test with the lowest Bonferroni-adjusted P values presented. Fold enrichment (fold enrich.) represents the number of genes observed relative to the expected number in each category. c, Volcano plots of change in translational status calculated by comparison of steady-state Total and TRAP mRNA abundance in Col-0 and rh6812 based on differential abundance analysis by edgeR. The log2 FC is shown on the x-axis, and the negative log10 of the FDR is shown on the y-axis. Genes with |log2 FC| ≥ 1, FDR < 0.01 were considered enhanced or repressed ribosome-association. The total number of genes used in the analysis is presented in parentheses.

Extended Data Fig. 7 The rh6812 mutant exhibits autoimmunity in a PAD4-partially dependent but EDS1-independent manner.

a, Heatmap representing relative fold change in Total and TRAP mRNA abundance for 104 ‘core’ EDS1/PAD-induced genes33 in rh6812 relative to Col-0. Individual genes are presented as a column with the upper two rows showing their log2 FC following 12 and 24 hours of EDS1 and PAD4 overexpression33, whereas the lower two rows representing their log2 FC in Total and TRAP populations of rh6812 relative to Col-0. b, Representative rosette growth phenotype of 28-day-old plants of Col-0, eds1-2, pad4-1, rh6812, rh6812 eds1-2, rh6812 pad4-1, rh6(+/−)812, rh6(+/−)812 eds1-2 and rh6(+/−)812 pad4-1 genotypes. c, Rosette diameter and fresh weight of 28-day-old plants (n = 12 or 30) of the genotypes presented in b. Boxplot boundaries represent the first and third quartiles; a horizontal line divides the interquartile range, median; red diamond, mean. d, RT-qPCR analysis of PR1 transcripts in 7-day-old seedlings of the rh6812 eds1-2 and rh6812 pad4-1 genotypes relative to rh6812 and the control genotypes Col-0, eds1-2, and pad4-1. Relative transcript fold-change was calculated using ACT1 as a reference. Error bars, s.d. (n = 3). e, Comparison of SA levels in 7-day-old seedlings of the genotypes as in d. Error bars, s.d. (n = 3). Statistical significance in c-e was determined by ANOVA, followed by Tukey’s HSD test. Data in c-e were log-transformed. Significantly differences of means (P < 0.05) are represented by different letters. See Source Data for P values for c, d and e.

Source data

Supplementary information

Source data

Source Data Fig. 1

Statistical source data for Fig. 1.

Source Data Fig. 2

Statistical source data for Fig. 2.

Source Data Fig. 6

Statistical source data for Fig. 6.

Source Data Extended Data Fig. 2

Uncropped immunoblots for Extended Data Fig. 2.

Source Data Extended Data Fig. 2

Statistical source data for Extended Data Fig. 2.

Source Data Extended Data Fig. 3

Statistical source data for Extended Data Fig. 3

Source Data Extended Data Fig. 4

Uncropped immunoblots for Extended Data Fig. 4c–e.

Source Data Extended Data Fig. 5

Statistical source data for Extended Data Fig. 5.

Source Data Extended Data Fig. 7

Statistical source data for Extended Data Fig. 7.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chantarachot, T., Sorenson, R.S., Hummel, M. et al. DHH1/DDX6-like RNA helicases maintain ephemeral half-lives of stress-response mRNAs. Nat. Plants 6, 675–685 (2020). https://doi.org/10.1038/s41477-020-0681-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41477-020-0681-8

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing