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:

ROS regulated reversible protein phase separation synchronizes plant flowering

Abstract

How aerobic organisms exploit inevitably generated but potentially dangerous reactive oxygen species (ROS) to benefit normal life is a fundamental biological question. Locally accumulated ROS have been reported to prime stem cell differentiation. However, the underlying molecular mechanism is unclear. Here, we reveal that developmentally produced H2O2 in plant shoot apical meristem (SAM) triggers reversible protein phase separation of TERMINATING FLOWER (TMF), a transcription factor that times flowering transition in the tomato by repressing pre-maturation of SAM. Cysteine residues within TMF sense cellular redox to form disulfide bonds that concatenate multiple TMF molecules and elevate the amount of intrinsically disordered regions to drive phase separation. Oxidation triggered phase separation enables TMF to bind and sequester the promoter of a floral identity gene ANANTHA to repress its expression. The reversible transcriptional condensation via redox-regulated phase separation endows aerobic organisms with the flexibility of gene control in dealing with developmental cues.

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: H2O2 represses flowering transition through TMF.
Fig. 2: TMF undergoes phase separation in vitro and in vivo.
Fig. 3: Redox-regulated disulfide bonding determines phase status of TMF.
Fig. 4: TMF directly targets the AN gene.
Fig. 5: H2O2 promoted formation of TMF transcriptional condensates regulate flowering.

Similar content being viewed by others

Data availability

Source data are provided with this paper. All other data supporting the findings of this study are available within the paper and its Supplementary information files, or are available from the corresponding author upon reasonable request.

References

  1. Irish, V. F. & Sussex, I. M. A fate map of the Arabidopsis embryonic shoot apical meristem. Development 115, 745–753 (1992).

    Article  Google Scholar 

  2. Park, S. J., Eshed, Y. & Lippman, Z. B. Meristem maturation and inflorescence architecture-lessons from the Solanaceae. Curr. Opin. Plant Biol. 17, 70–71 (2014).

    Article  PubMed  Google Scholar 

  3. MacAlister, C. A. et al. Synchronization of the flowering transition by the tomato TERMINATING FLOWER gene. Nat. Genet. 44, 1393–1398 (2012).

    Article  CAS  PubMed  Google Scholar 

  4. Xu, C., Park, S. J., Van Eck, J. & Lippman, Z. B. Control of inflorescence architecture in tomato by BTB/POZ transcriptional regulators. Genes Dev. 30, 2048–2061 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Iyer, L. M. & Aravind, L. ALOG domains: provenance of plant homeotic and developmental regulators from the DNA-binding domain of a novel class of DIRS1-type retroposons. Biol. Direct 7, 39 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Lippman, Z. B. et al. The making of a compound inflorescence in tomato and related nightshades. PLoS Biol. 6, 2424–2435 (2008).

    Article  CAS  Google Scholar 

  7. Lei, Y., Su, S., He, L., Hu, X. & Luo, D. A member of the ALOG gene family has a novel role in regulating nodulation in Lotus japonicus. J. Integr. Plant Biol. 61, 463–477 (2019).

    Article  CAS  PubMed  Google Scholar 

  8. Takeda, S. et al. CUP-SHAPED COTYLEDON1 transcription factor activates the expression of LSH4 and LSH3, two members of the ALOG gene family, in shoot organ boundary cells. Plant J. 66, 1066–1077 (2011).

    Article  CAS  PubMed  Google Scholar 

  9. Zhao, L. et al. Overexpression of LSH1, a member of an uncharacterised gene family, causes enhanced light regulation of seedling development. Plant J. 37, 694–706 (2004).

    Article  CAS  PubMed  Google Scholar 

  10. Yoshida, A., Suzaki, T., Tanaka, W. & Hirano, H.-Y. The homeotic gene long sterile lemma (G1) specifies sterile lemma identity in the rice spikelet. Proc. Natl Acad. Sci. USA 106, 20103–20108 (2009).

    Article  CAS  PubMed  Google Scholar 

  11. Cho, E. & Zambryski, P. C. ORGAN BOUNDARY1 defines a gene expressed at the junction between the shoot apical meristem and lateral organs. Proc. Natl Acad. Sci. USA 108, 2154–2159 (2011).

    Article  CAS  PubMed  Google Scholar 

  12. Sato, D.-S., Ohmori, Y., Nagashima, H., Toriba, T. & Hirano, H.-Y. A role for TRIANGULAR HULL1 in fine-tuning spikelet morphogenesis in rice. Genes Genet. Syst. 89, 61–69 (2014).

    Article  CAS  PubMed  Google Scholar 

  13. Naramoto, S. et al. A conserved regulatory mechanism mediates the convergent evolution of plant shoot lateral organs. PLoS Biol. 17, e3000560 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  14. Yoshida, A. et al. TAWAWA1, a regulator of rice inflorescence architecture, functions through the suppression of meristem phase transition. Proc. Natl Acad Sci. USA 110, 767–772 (2013).

    Article  CAS  PubMed  Google Scholar 

  15. Chen, F. et al. Genome-wide identification and characterization of the ALOG gene family in Petunia. BMC Plant Biol. 19, 600 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Owusu-Ansah, E. & Banerjee, U. Reactive oxygen species prime Drosophila haematopoietic progenitors for differentiation. Nature 461, 537–541 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Morimoto, H. et al. ROS are required for mouse spermatogonial stem cell self-renewal. Cell Stem Cell. 12, 774–786 (2013).

    Article  CAS  PubMed  Google Scholar 

  18. Zeng, J., Dong, Z., Wu, H., Tian, Z. & Zhao, Z. Redox regulation of plant stem cell fate. EMBO J. 36, 2844–2855 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Tsukagoshi, H., Busch, W. & Benfey, P. N. Transcriptional regulation of ROS controls transition from proliferation to differentiation in the root. Cell 143, 606–616 (2010).

    Article  CAS  PubMed  Google Scholar 

  20. Yang, S. et al. ROS: the fine-tuner of plant stem cell fate. Trends Plant Sci. 23, 850–853 (2018).

    Article  CAS  PubMed  Google Scholar 

  21. Rampon, C., Volovitch, M., Joliot, A. & Vriz, S. Hydrogen peroxide and redox regulation of developments. Antioxidants 7, 159 (2018).

    Article  PubMed Central  Google Scholar 

  22. Sies, H. & Jones, D. P. Reactive oxygen species (ROS) as pleiotropic physiological signalling agents. Nat. Rev. Mol. Cell Biol. 21, 363–383 (2020).

    Article  CAS  Google Scholar 

  23. Foyer, C. H. & Noctor, G. Redox homeostasis and signaling in a higher-CO2 world. Annu. Rev. Plant Biol. 71, 157–182 (2020).

    Article  CAS  PubMed  Google Scholar 

  24. Paulsen, C. E. & Carroll, K. S. Orchestrating redox signaling networks through regulatory cysteine switches. ACS Chem. Biol. 5, 47–62 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Poole, L. B. The basics of thiols and cysteines in redox biology and chemistry. Free Radic. Biol. Med. 80, 148–157 (2015).

    Article  CAS  PubMed  Google Scholar 

  26. Zeida, A. et al. Catalysis of peroxide reduction by fast reacting protein thiols. Chem. Rev. 119, 10829–10855 (2019).

    Article  CAS  PubMed  Google Scholar 

  27. Banani, S. F., Lee, H. O., Hyman, A. A. & Rosen, M. K. Biomolecular condensates: organizers of cellular biochemistry. Nat. Rev. Mol. Cell Biol. 18, 285–298 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Shin, Y. & Brangwynne, C. P. Liquid phase condensation in cell physiology and disease. Science 357, eaaf4382 (2017).

    Article  PubMed  Google Scholar 

  29. Kroschwald, S. et al. Different material states of Pub1 condensates define distinct modes of stress adaptation and recovery. Cell Rep. 23, 3327–3339 (2018).

    Article  CAS  PubMed  Google Scholar 

  30. Franzmann, T. M. et al. Phase separation of a yeast prion protein promotes cellular fitness. Science 359, eaao5654 (2018).

    Article  PubMed  Google Scholar 

  31. Riback, J. A. et al. Stress-triggered phase separation is an adaptive, evolutionarily tuned response. Cell 168, 1028–1040.e19 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Kato, M. et al. Redox state controls phase separation of the yeast Ataxin-2 protein via reversible oxidation of its methionine-rich low-complexity domain. Cell 177, 711–721.e8 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Yang, Y. S. et al. Yeast Ataxin-2 forms an intracellular condensate required for the inhibition of TORC1 signaling during respiratory growth. Cell 177, 697–710.e17 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Zhang, G., Wang, Z., Du, Z. & Zhang, H. mTOR regulates phase separation of PGL granules to modulate their autophagic degradation. Cell 174, 1492–1506.e22 (2018).

    Article  CAS  PubMed  Google Scholar 

  35. Wu, X., Cai, Q., Feng, Z. & Zhang, M. Liquid–liquid phase separation in neuronal development and synaptic signaling. Dev. Cell 55, 18–29 (2020).

    Article  CAS  PubMed  Google Scholar 

  36. Dunand, C., Crèvecoeur, M. & Penel, C. Distribution of superoxide and hydrogen peroxide in Arabidopsis root and their influence on root development: possible interaction with peroxidases. N. Phytol. 174, 332–341 (2007).

    Article  CAS  Google Scholar 

  37. Park, S. J., Jiang, K., Schatz, M. C. & Lippman, Z. B. Rate of meristem maturation determines inflorescence architecture in tomato. Proc. Natl Acad. Sci. USA 109, 639–644 (2012).

    Article  CAS  PubMed  Google Scholar 

  38. Doussiere, J. & Vignais, P. V. Diphenylene iodonium as an inhibitor of the NADPH oxidase complex of bovine neutrophils. Factors controlling the inhibitory potency of diphenylene iodonium in a cell-free system of oxidase activation. Eur. J. Biochem. 208, 61–71 (1992).

    Article  CAS  PubMed  Google Scholar 

  39. Chen, X. et al. Apoplastic H2O2 plays a critical role in axillary bud outgrowth by altering auxin and cytokinin homeostasis in tomato plants. N. Phytol. 211, 1266–1278 (2016).

    Article  CAS  Google Scholar 

  40. Mei, Y., Chen, H., Shen, W., Shen, W. & Huang, L. Hydrogen peroxide is involved in hydrogen sulfide-induced lateral root formation in tomato seedlings. BMC Plant Biol. 17, 162 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  41. Li, P. et al. Phase transitions in the assembly of multivalent signalling proteins. Nature 483, 336–340 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Kato, M. et al. Cell-free formation of RNA granules: low complexity sequence domains form dynamic fibers within hydrogels. Cell 149, 753–767 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Lin, Y., Protter, D. S. W., Rosen, M. K. & Parker, R. Formation and maturation of phase-separated liquid droplets by RNA-binding proteins. Mol. Cell 60, 208–219 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Jiao, C.-J. et al. β-ODAP accumulation could be related to low levels of superoxide anion and hydrogen peroxide in Lathyrus sativus L. Food Chem. Toxicol. 49, 556–562 (2011).

    Article  CAS  PubMed  Google Scholar 

  45. Wang, L., Wang, X. & Wang, C. Protein disulfide–isomerase, a folding catalyst and a redox-regulated chaperone. Free Radic. Biol. Med. 83, 305–313 (2015).

    Article  CAS  PubMed  Google Scholar 

  46. Lyles, M. M. & Gilbert, H. F. Catalysis of the oxidative folding of ribonuclease A by protein disulfide isomerase: dependence of the rate on the composition of the redox buffer. Biochemistry 30, 613–619 (1991).

    Article  CAS  PubMed  Google Scholar 

  47. Levy, Y. Y. & Dean, C. The transition to flowering. Plant Cell 10, 1973–1989 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Jones, D. P. Radical-free biology of oxidative stress. Am. J. Physiol.-Cell Physiol. 295, C849–C868 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Eck, J. V., Kirk, D. D. & Walmsley, A. M. In Agrobacterium Protocols. Methods in Molecular Biology Vol. 343 (ed. Wang, K.) (Humana Press, 2006); https://doi.org/10.1385/1-59745-130-4:459

  50. Jackson, D. Molecular Plant Pathology: A Practical Approach (Oxford Univ. Press, 1992).

  51. Xu, C. et al. A cascade of arabinosyltransferases controls shoot meristem size in tomato. Nat. Genet. 47, 784–792 (2015).

    Article  CAS  PubMed  Google Scholar 

  52. Lin, R. et al. Transposase-derived transcription factors regulate light signaling in arabidopsis. Science 318, 1302–1305 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Gendrel, A.-V., Lippman, Z., Martienssen, R. & Colot, V. Profiling histone modification patterns in plants using genomic tiling microarrays. Nat. Methods 2, 213–218 (2005).

    Article  CAS  PubMed  Google Scholar 

  54. Shen, G. et al. Warfarin traps human vitamin K epoxide reductase in an intermediate state during electron transfer. Nat. Struct. Mol. Biol. 24, 69–76 (2017).

    Article  CAS  PubMed  Google Scholar 

  55. Li, H. et al. Crystal and solution structures of human protein-disulfide isomerase-like protein of the testis (PDILT) provide insight into its chaperone activity. J. Biol. Chem. 293, 1192–1202 (2018).

    Article  CAS  PubMed  Google Scholar 

  56. Shi, Y. et al. Ethylene signaling negatively regulates freezing tolerance by repressing expression of CBF and Type-A ARR genes in Arabidopsis. Plant Cell 24, 2578–2595 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank C.-C. Wang and L. Wang from the Institute of Biophysics, CAS for sharing PDI enzyme and valuable suggestions for related experiments, Z. B. Lippman (Cold Spring Harbor Laboratory), S. J. Park (Wonkwang University) and J. V. Eck (Boyce Thompson Institute) for sharing seeds, in situ images and the tomato transformation protocol, respectively. We thank J. Y. Li, Z. B. Lippman, J. Yang, K. Jiang and Q. Y. Wu for valuable comments and discussions; S. H. Yang and J. G. Li for sharing transformation vectors; Y. B. Tian, L. Fu, Z. Lu and S. Q. Jia for technical support and Y. F. Chen, W. X. Dong, Q. L. Guo, Y. Yu, Y. Xie, T. H. Zhang and X. Tong in the Xu laboratory for their help in this study. We are grateful to the staff from facilities of State Key Laboratory of Plant Genomics, Institute of Genetics and Developmental Biology, CAS; Institute of Botany, CAS and Tsinghua University for technical assistance. This study was financially supported by the Key Research Program of Frontier Sciences of the Chinese Academy of Science (grant no. ZDBS-LY-SM021) and the Strategic Priority Research Program of Chinese Academy of Sciences (grant no. XDA24030503) to C.X., a National Key R&D grant (no. 2019YFA0508403) and a NSFC grant (no. 31871443) to P.L. and start-up funding from State Key Laboratory of Plant Genomics and Institute of Genetics and Developmental Biology to C.X. and a NSFC grant (no. 31900174) to X.H.

Author information

Authors and Affiliations

Authors

Contributions

C.X. and P.L. designed the research. X.H. performed in vivo condensate assays, chemical treatments, gene expression, immunoblots, ChIP–qPCR and transcriptional activity assays with help from N.X., S.C. and W. Li performed in vitro phase separation assays. S.C. performed EMSA, protein sedimentation assays and PDI treatments. X.H. and N.X. produced CRISPR mutants and transgenic lines with the help from X.Z. L.T. performed the yeast one-hybrid assay. Y. Zhang and W. Liu performed in situ hybridization and H2O2 staining. N.Y. performed mass spectrometric analysis. Y. Zou performed bioinformatic analysis. C.X. and P.L. wrote the paper with input from all other authors.

Corresponding authors

Correspondence to Pilong Li or Cao Xu.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

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

Extended data

Extended Data Fig. 1 Images and quantification data for hydrogen peroxide staining and redox chemical treatments.

a, DAB staining and stereoscope imaging showing the accumulation of H2O2 in tomato young leaves treated with or without H2O2 (10 mM) for 36 h. b,c, HPF staining (b) and quantitative data (c) showing the accumulation of H2O2 in the meristem treated with or without H2O2 (10 mM) for 48 h. (n=3). d,e, Stereoscope images (d) and quantitative data (e) comparing flowering transition indicated by leaf production until floral meristem stage transition from mock and H2O2 (10 mM) treated WT (upper) and tmf-2 (bottom). Leaf production is indicated by leaf number. L, Leaf. Scale bar, 100 μm. Data are presented as means (± s.d.). Sample size used for statistics of mock and H2O2 treatment for WT and tmf-2 is 19, 16, 43, 44, respectively. f, Phylogeny tree showing RBOH gene family in tomato. g, Expression of SlRBOH genes during meristem maturation of tomato. h, CRISPR/Cas9 gRNAs for targeting SlRBOH genes. i, DAB staining showing decreased H2O2 level in CRISPR mutant of slrboh1 slrboh2. Scale bar, 0.9 cm. j, Stereoscope images comparing flowering transition indicated by leaf production until floral meristem stage transition from WT and slrboh1 slrboh2 mutant, L, leaf. Scale bar, 100 μm. Three independent assays with similar results were carried out. In c and e, data are presented as means (± s.d.)(two-tailed t-test).

Source data

Extended Data Fig. 2 Droplet property and FRAP analysis of TMF-GFP condensates.

a, Amino acid sequence of TMF indicating IDRs, putative DNA binding domain and cysteine residues. b, Aspect ratios (maximal diameter/ minimal diameter) of droplets formed GFP-TMF. Gray ellipses show a guide to the eye of different aspect ratios. Totally, 202 droplets were measured for aspect ratio calculation. c, Quantification of TMF-GFP transfected tomato protoplast cells with or without condensates in nuclei. Three independent experiments were performed for quantification. Data are presented as (± s.d.) (n = 48). d,e, Image (d) and quantitative data (e) showing the recovery of TMF-GFP condensates after photobleaching in tomato protoplasts. The bleached (green line) event occurs at time = 0 s. The unbleached (blue line) was used as control. Quantitative data are representative of three independent photobleaching events. Data are presented as (± s.d.) (n = 3).

Source data

Extended Data Fig. 3 Quantification data and representative images for protein behavior of TMF and its mutated variants after redox chemicals treatment.

a, Quantification of integrated fluorescence density of the liquid-like droplets formed by GFP-TMF protein under various concentration combinations of H2O2 and DTT with constant protein concentration (25 μM). b,c, Representative confocal images (b) and quantification data (c) showing effects of TCEP treatment on droplet formation of GFP-TMF proteins. Protein concentration, 20 μM; Salt concentration, 25 mM. Scale bar, 20 μm. d, Schematic sedimentation assay for redox regulated phase separation. e,f, Immunoblotting (e) and quantification data (f) showing the distribution of TMF proteins between aqueous-solution/supernatant (S) and condensed liquid phase/pellet (P) fractions after H2O2 or DTT treatments. g, Schematics showing TMF variants with IDR or cysteine mutations. h, Quantification of integrated fluorescence density of the droplets formed by TMF variants with IDR or cysteine mutations. i, Immunoblot analysis showing the expression for TMF and variants with IDR or cysteine mutations in tomato protoplast. Actin serves as a loading control. In a,c,f,h, three technical replicates data are presented as means (± s.d.) (n = 3, two-tailed t-test).

Source data

Extended Data Fig. 4 Inter- and intramolecular disulfide bonds identified by LC-MS/MS.

a, LC-MS/MS spectrum of intramolecular and intermolecular disulfide bonds from normal TMF. b, Schematic diagrams showing the working model of PDI in different redox status. c, LC-MS/MS spectrum of intramolecular and intermolecular disulfide bonds from PDI treated TMF protein.

Extended Data Fig. 5 Representative colonies and quantification data of yeast one-hybrid assay.

a, Promoter regions upstream of the AN gene selected for yeast one hybrid assays in (b-d). b-d, Colony growth assessment (b) and quantification of β-galactosidase activity (c,d) in yeast one-hybrid assay. e, Promoter regions upstream of the AN gene selected for evaluating the ChIP enrichments. In c and d, three biological replicates data are presented as means (± s.d.) (n = 3, two-tailed t-test).

Source data

Extended Data Fig. 6 Quantification data for droplets, phase sedimentation assay and leaf production for flower transition.

a, Phase diagram showing droplets formed by GFP-TMF protein and Cy-3 labeled DNA fragments under various concentration combinations of H2O2 and DTT with constant protein concentration. Scale bar, 5 μm. b,c, Immunoblotting (b) and quantification data (c) showing the distribution of TMF-DNA complex between aqueous-solution/supernatant (S) and condensed liquid phase/pellet (P) fractions after H2O2 or DTT treatments. Three technical replicates data are presented. d, Western blot analysis showing expression of proteins in transgenic plants. The Wild-type (WT) plant sample served as a negative control, actin served as a loading control. e, Quantification of leaf number to flower transition on primary shoots. In c and e, data are means(± s.d.) (n = 3 for c, n = 8 for e, two-tailed t-test).

Source data

Supplementary information

Supplementary Information

Supplementary Fig. 1 and Table 1.

Reporting Summary

Supplementary Video 1

Fusion process of two separated GFP–TMF droplets for Fig. 2f.

Supplementary Video 2

FRAP analysis of recombinantly expressed GFP–TMF droplets for Fig. 2g.

Supplementary Video 3

FRAP analysis of TMF–GFP condensates in tomato protoplasts for Extended Data Fig. 2d.

Supplementary Video 4

FRAP analysis of TMF–GFP condensates in the young leaf of 35S:TMFGFP transgenic tomato plants for Fig. 2j.

Supplementary Video 5

FRAP analysis of droplets formed by GFP–TMF–DNA complex for Fig. 4e.

Source data

Source Data Fig. 1

Statistical Source Data.

Source Data Fig. 1

Unprocessed western blots.

Source Data Fig. 2

Statistical Source Data.

Source Data Fig. 3

Statistical Source Data.

Source Data Fig. 4

Statistical Source Data.

Source Data Fig. 5

Statistical Source Data.

Source Data Extended Data Fig. 1

Statistical Source Data.

Source Data Extended Data Fig. 2

Statistical Source Data.

Source Data Extended Data Fig. 3

Statistical Source Data.

Source Data Extended Data Fig. 3

Unprocessed western blots.

Source Data Extended Data Fig. 5

Statistical Source Data.

Source Data Extended Data Fig. 6

Statistical Source Data.

Source Data Extended Data Fig. 6

Unprocessed western blots.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Huang, X., Chen, S., Li, W. et al. ROS regulated reversible protein phase separation synchronizes plant flowering. Nat Chem Biol 17, 549–557 (2021). https://doi.org/10.1038/s41589-021-00739-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41589-021-00739-0

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