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:

Stomatal regulators are co-opted for seta development in the astomatous liverwort Marchantia polymorpha

Abstract

The evolution of special types of cells requires the acquisition of new gene regulatory networks controlled by transcription factors (TFs). In stomatous plants, a TF module formed by subfamilies Ia and IIIb basic helix–loop–helix TFs (Ia-IIIb bHLH) regulates stomatal formation; however, how this module evolved during land plant diversification remains unclear. Here we show that, in the astomatous liverwort Marchantia polymorpha, a Ia-IIIb bHLH module regulates the development of a unique sporophyte tissue, the seta, which is found in mosses and liverworts. The sole Ia bHLH gene, MpSETA, and a IIIb bHLH gene, MpICE2, regulate the cell division and/or differentiation of seta lineage cells. MpSETA can partially replace the stomatal function of Ia bHLH TFs in Arabidopsis thaliana, suggesting that a common regulatory mechanism underlies setal and stomatal formation. Our findings reveal the co-option of a Ia-IIIb bHLH TF module for regulating cell fate determination and/or cell division of distinct types of cells during land plant evolution.

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: MpSETA is the only bHLH TF that belongs to subfamily Ia in M. polymorpha.
Fig. 2: MpSETA is preferentially expressed in developing sporophyte.
Fig. 3: Mpsetako mutants show developmental defects in the seta.
Fig. 4: MpICE2 regulates setal development in cooperation with MpSETA.
Fig. 5: Function and co-option of the Ia-IIIb bHLH module during land plant evolution.

Similar content being viewed by others

Data availability

Public RNA-seq data can be downloaded from SRA repository (https://www.ncbi.nlm.nih.gov/sra) under accession PRJNA350270, PRJDB6579, PRJDB4420, PRJDB9329 and PRJNA265205. All data supporting the conclusions in the paper are available within the paper or the supplementary materials. Sequence data and gene ID can be found in the database, as follows: MarpolBase (https://marchantia.info), Phytozome v.13 (https://phytozome-next.jgi.doe.gov/), OneKP (https://db.cngb.org/onekp/), TAIR (http://www.arabidopsis.org/) and NCBI (https://www.ncbi.nlm.nih.gov/genome/?term=PRJNA701193). Source data are provided with this paper.

References

  1. Ligrone, R., Duckett, J. G. & Renzaglia, K. S. Major transitions in the evolution of early land plants: a bryological perspective. Ann. Bot. 109, 851–871 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  2. Romani, F. & Moreno, J. E. Molecular mechanisms involved in functional macroevolution of plant transcription factors. New Phytol. 230, 1345–1353 (2021).

    Article  CAS  PubMed  Google Scholar 

  3. Catarino, B., Hetherington, A. J., Emms, D. M., Kelly, S. & Dolan, L. The stepwise increase in the number of transcription factor families in the Precambrian predated the diversification of plants on land. Mol. Biol. Evol. 33, 2815–2819 (2016).

    Article  CAS  PubMed  Google Scholar 

  4. Bowman, J. L. et al. Insights into land plant evolution garnered from the Marchantia polymorpha genome. Cell 171, 287–304.e15 (2017).

    Article  CAS  PubMed  Google Scholar 

  5. Pires, N. & Dolan, L. Origin and diversification of basic-helix-loop-helix proteins in plants. Mol. Biol. Evol. 27, 862–874 (2010).

    Article  CAS  PubMed  Google Scholar 

  6. MacAlister, C. A., Ohashi-Ito, K. & Bergmann, D. C. Transcription factor control of asymmetric cell divisions that establish the stomatal lineage. Nature 445, 537–540 (2007).

    Article  CAS  PubMed  Google Scholar 

  7. Pillitteri, L. J., Sloan, D. B., Bogenschutz, N. L. & Torii, K. U. Termination of asymmetric cell division and differentiation of stomata. Nature 445, 501–505 (2007).

    Article  CAS  PubMed  Google Scholar 

  8. Ohashi-Ito, K. & Bergmann, D. C. Arabidopsis FAMA controls the final proliferation/differentiation switch during stomatal development. Plant Cell 18, 2493–2505 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Kanaoka, M. M. et al. SCREAM/ICE1 and SCREAM2 specify three cell-state transitional steps leading to Arabidopsis stomatal differentiation. Plant Cell 20, 1775–1785 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Chater, C. C. et al. Origin and function of stomata in the moss Physcomitrella patens. Nat. Plants 2, 16179 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Caine, R. S., Chater, C. C. C., Fleming, A. J. & Gray, J. E. Stomata and sporophytes of the model moss Physcomitrium patens. Front. Plant Sci. 11, 643 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  12. Chater, C. C. C., Caine, R. S., Fleming, A. J. & Gray, J. E. Origins and evolution of stomatal development. Plant Physiol. 174, 624–638 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Harris, B. J., Harrison, C. J., Hetherington, A. M. & Williams, T. A. Phylogenomic evidence for the monophyly of bryophytes and the reductive evolution of stomata. Curr. Biol. 30, 2001–2012.e2 (2020).

    Article  CAS  PubMed  Google Scholar 

  14. Flores-Sandoval, E., Romani, F. & Bowman, J. L. Co-expression and transcriptome analysis of Marchantia polymorpha transcription factors supports class C ARFs as independent actors of an ancient auxin regulatory module. Front. Plant Sci. 9, 1345 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  15. Feller, A., Hernandez, J. M. & Grotewold, E. An ACT-like domain participates in the dimerization of several plant basic-helix-loop-helix transcription factors. J. Biol. Chem. 281, 28964–28974 (2006).

    Article  CAS  PubMed  Google Scholar 

  16. Seo, H. et al. Intragenic suppressors unravel the role of the SCREAM ACT-like domain for bHLH partner selectivity in stomatal development. Proc. Natl Acad. Sci. USA 119, e2117774119 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Linde, A. M., Eklund, D. M., Cronberg, N., Bowman, J. L. & Lagercrantz, U. Rates and patterns of molecular evolution in bryophyte genomes, with focus on complex thalloid liverworts, Marchantiopsida. Mol. Phylogenet. Evol. 165, 107295 (2021).

    Article  CAS  PubMed  Google Scholar 

  18. Pillitteri, L. J., Bogenschutz, N. L. & Torii, K. U. The bHLH protein, MUTE, controls differentiation of stomata and the hydathode pore in Arabidopsis. Plant Cell Physiol. 49, 934–943 (2008).

    Article  CAS  PubMed  Google Scholar 

  19. Frank, M. H. & Scanlon, M. J. Transcriptomic evidence for the evolution of shoot meristem function in sporophyte-dominant land plants through concerted selection of ancestral gametophytic and sporophytic genetic programs. Mol. Biol. Evol. 32, 355–367 (2015).

    Article  CAS  PubMed  Google Scholar 

  20. Yamaoka, S. et al. Generative cell specification requires transcription factors evolutionarily conserved in land plants. Curr. Biol. 28, 479–486.e5 (2018).

    Article  CAS  PubMed  Google Scholar 

  21. Higo, A. et al. Transcriptional framework of male gametogenesis in the liverwort Marchantia polymorpha L. Plant Cell Physiol. 57, 325–338 (2016).

    Article  CAS  PubMed  Google Scholar 

  22. Hisanaga, T. et al. Deep evolutionary origin of gamete-directed zygote activation by KNOX/BELL transcription factors in green plants. eLife 10, e57090 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Shimamura, M. Marchantia polymorpha: taxonomy, phylogeny and morphology of a model system. Plant Cell Physiol. 57, 230–256 (2016).

    Article  CAS  PubMed  Google Scholar 

  24. Haig, D. Filial mistletoes: the functional morphology of moss sporophytes. Ann. Bot. 111, 337–345 (2013).

    Article  PubMed  Google Scholar 

  25. Ligrone, R., Duckett, J. G. & Renzaglia, K. S. The origin of the sporophyte shoot in land plants: a bryological perspective. Ann. Bot. 110, 935–941 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  26. Durand, E. J. The development of the sexual organs and sporogonium of Marchantia polymorpha. Bull. Torrey Bot. Club 35, 321–335 (1908).

    Article  Google Scholar 

  27. Hernández-García, J. et al. Coordination between growth and stress responses by DELLA in the liverwort Marchantia polymorpha. Curr. Biol. 31, 3678–3686.e11 (2021).

    Article  PubMed  Google Scholar 

  28. Ishizaki, K., Johzuka-Hisatomi, Y., Ishida, S., Iida, S. & Kohchi, T. Homologous recombination-mediated gene targeting in the liverwort Marchantia polymorpha L. Sci. Rep. 3, 1532 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  29. Shaw, J. & Renzaglia, K. Phylogeny and diversification of bryophytes. Am. J. Bot. 91, 1557–1581 (2004).

    Article  PubMed  Google Scholar 

  30. Wu, Z. et al. Multiple transcriptional factors control stomata development in rice. New Phytol. 223, 220–232 (2019).

    Article  CAS  PubMed  Google Scholar 

  31. Sugano, S. S. et al. Efficient CRISPR/Cas9-based genome editing and its application to conditional genetic analysis in Marchantia polymorpha. PLoS ONE 13, e0205117 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  32. Sugano, S. S. et al. CRISPR/Cas9-mediated targeted mutagenesis in the liverwort Marchantia polymorpha L. Plant Cell Physiol. 55, 475–481 (2014).

    Article  CAS  PubMed  Google Scholar 

  33. Chinnusamy, V. et al. ICE1: a regulator of cold-induced transcriptome and freezing tolerance in Arabidopsis. Genes Dev. 17, 1043–1054 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Davies, K. A. & Bergmann, D. C. Functional specialization of stomatal bHLHs through modification of DNA-binding and phosphoregulation potential. Proc. Natl Acad. Sci. USA 111, 15585–15590 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. MacAlister, C. A. & Bergmann, D. C. Sequence and function of basic helix-loop-helix proteins required for stomatal development in Arabidopsis are deeply conserved in land plants. Evol. Dev. 13, 182–192 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Merced, A. & Renzaglia, K. S. Structure, function and evolution of stomata from a bryological perspective. Bryophyt. Divers. Evol. 39, 7–20 (2017).

    Article  Google Scholar 

  37. Wei, D. et al. INDUCER OF CBF EXPRESSION 1 is a male fertility regulator impacting anther dehydration in Arabidopsis. PLoS Genet. 14, e1007695 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  38. Higo, A. et al. Transcription factor DUO1 generated by neo-functionalization is associated with evolution of sperm differentiation in plants. Nat. Commun. 9, 5283 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Menand, B. et al. An ancient mechanism controls the development of cells with a rooting function in land plants. Science 316, 1477–1480 (2007).

    Article  CAS  PubMed  Google Scholar 

  40. Koshimizu, S. et al. Physcomitrella MADS-box genes regulate water supply and sperm movement for fertilization. Nat. Plants 4, 36–45 (2018).

    Article  CAS  PubMed  Google Scholar 

  41. Shirakawa, M. et al. FAMA is an essential component for the differentiation of two distinct cell types, myrosin cells and guard cells, in Arabidopsis. Plant Cell 26, 4039–4052 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Yasui, Y. et al. GEMMA CUP-ASSOCIATED MYB1, an ortholog of axillary meristem regulators, is essential in vegetative reproduction in Marchantia polymorpha. Curr. Biol. 29, 3987–3995.e5 (2019).

    Article  CAS  PubMed  Google Scholar 

  43. Proust, H. et al. RSL class I genes controlled the development of epidermal structures in the common ancestor of land plants. Curr. Biol. 26, 93–99 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Puttick, M. N. et al. The interrelationships of land plants and the nature of the ancestral embryophyte. Curr. Biol. 28, 733–745.e2 (2018).

    Article  CAS  PubMed  Google Scholar 

  45. Renzaglia, K. S. & Garbary, D. J. Motile gametes of land plants: diversity, development, and evolution. Crit. Rev. Plant Sci. 20, 107–213 (2001).

    Article  Google Scholar 

  46. Sakakibara, K., Nishiyama, T., Deguchi, H. & Hasebe, M. Class 1 KNOX genes are not involved in shoot development in the moss Physcomitrella patens but do function in sporophyte development. Evol. Dev. 10, 555–566 (2008).

    Article  CAS  PubMed  Google Scholar 

  47. Coudert, Y., Novák, O. & Harrison, C. J. A KNOX-cytokinin regulatory module predates the origin of indeterminate vascular plants. Curr. Biol. 29, 2743–2750.e5 (2019).

    Article  CAS  PubMed  Google Scholar 

  48. Li, M. & Sack, F. D. Myrosin idioblast cell fate and development are regulated by the Arabidopsis transcription factor FAMA, the auxin pathway, and vesicular trafficking. Plant Cell 26, 4053–4066 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Shirakawa, M., Tanida, M. & Ito, T. The cell differentiation of idioblast myrosin cells: similarities with vascular and guard cells. Front. Plant Sci. 12, 829541 (2021).

    Article  PubMed  Google Scholar 

  50. Li, F. W. et al. Anthoceros genomes illuminate the origin of land plants and the unique biology of hornworts. Nat. Plants 6, 259–272 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Han, S. et al. MUTE directly orchestrates cell-state switch and the single symmetric division to create stomata. Dev. Cell 45, 303–315.e5 (2018).

    Article  CAS  PubMed  Google Scholar 

  52. Lau, O. S. et al. Direct roles of SPEECHLESS in the specification of stomatal self-renewing cells. Science 345, 1605–1609 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. One Thousand Plant Transcriptomes Initiative. One thousand plant transcriptomes and the phylogenomics of green plants. Nature 574, 679–685 (2019).

    Article  CAS  Google Scholar 

  54. Katoh, K. & Toh, H. Recent developments in the MAFFT multiple sequence alignment program. Brief. Bioinform. 9, 286–298 (2008).

    Article  CAS  PubMed  Google Scholar 

  55. Waterhouse, A. M., Procter, J. B., Martin, D. M. A., Clamp, M. & Barton, G. J. Jalview version 2-A multiple sequence alignment editor and analysis workbench. Bioinformatics 25, 1189–1191 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Kumar, S., Stecher, G. & Tamura, K. MEGA7: Molecular Evolutionary Genetics Analysis version 7.0 for bigger datasets. Mol. Biol. Evol. 33, 1870–1874 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Nakagawa, T. et al. Improved gateway binary vectors: high-performance vectors for creation of fusion constructs in transgenic analysis of plants. Biosci. Biotechnol. Biochem. 71, 2095–2100 (2007).

    Article  CAS  PubMed  Google Scholar 

  58. Nakamura, S. et al. Gateway binary vectors with the bialaphos resistance gene, bar, as a selection marker for plant transformation. Biosci. Biotechnol. Biochem. 74, 1315–1319 (2010).

    Article  CAS  PubMed  Google Scholar 

  59. Clough, S. J. & Bent, A. F. Floral dip: a simplified method for Agrobacterium-mediated transformation of Arabidopsis thaliana. Plant J. 16, 735–743 (1998).

    Article  CAS  PubMed  Google Scholar 

  60. Shimada, T. L., Shimada, T. & Hara-Nishimura, I. A rapid and non-destructive screenable marker, FAST, for identifying transformed seeds of Arabidopsis thaliana. Plant J. 61, 519–528 (2010).

    Article  CAS  PubMed  Google Scholar 

  61. Chen, S., Zhou, Y., Chen, Y. & Gu, J. Fastp: an ultra-fast all-in-one FASTQ preprocessor. Bioinformatics 34, i884–i890 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  62. Dobin, A. et al. STAR: ultrafast universal RNA-seq aligner. Bioinformatics 29, 15–21 (2013).

    Article  CAS  PubMed  Google Scholar 

  63. Li, H. et al. The Sequence Alignment/Map format and SAMtools. Bioinformatics 25, 2078–2079 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  64. Li, B. & Dewey, C. N. RSEM: accurate transcript quantification from RNA-Seq data with or without a reference genome. BMC Bioinformatics 12, 323 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Ishizaki, K. et al. Development of gateway binary vector series with four different selection markers for the liverwort Marchantia polymorpha. PLoS ONE 10, e0138876 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  66. Matsushita, T., Mochizuki, N. & Nagatani, A. Dimers of the N-terminal domain of phytochrome B are functional in the nucleus. Nature 424, 571–574 (2003).

    Article  CAS  PubMed  Google Scholar 

  67. Tsuboyama, S., Nonaka, S., Ezura, H. & Kodama, Y. Improved G-AgarTrap: a highly efficient transformation method for intact gemmalings of the liverwort Marchantia polymorpha. Sci. Rep. 8, 10800 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  68. Ishizaki, K., Chiyoda, S., Yamato, K. T. & Kohchi, T. Agrobacterium-mediated transformation of the haploid liverwort Marchantia polymorpha L., an emerging model for plant biology. Plant Cell Physiol. 49, 1084–1091 (2008).

    Article  CAS  PubMed  Google Scholar 

  69. Rövekamp, M., Bowman, J. L. & Grossniklaus, U. Marchantia MpRKD regulates the gametophyte–sporophyte transition by keeping egg cells quiescent in the absence of fertilization. Curr. Biol. 26, 1782–1789 (2016).

    Article  PubMed  Google Scholar 

  70. Rossignol, P., Collier, S., Bush, M., Shaw, P. & Doonan, J. H. Arabidopsis POT1A interacts with TERT-V(18), an N-terminal splicing variant of telomerase. J. Cell Sci. 120, 3678–3687 (2007).

    Article  CAS  PubMed  Google Scholar 

  71. Hino, T. et al. Two Sec13p homologs, AtSec13A and AtSec13B, redundantly contribute to the formation of COPII transport vesicles in Arabidopsis thaliana. Biosci. Biotechnol. Biochem. 75, 1848–1852 (2011).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank T. Nakagawa (Shimane University, Japan), S. Mano (National Institute for Basic Biology, Japan), S. S. Sugano (National Institute of Advanced Industrial Science and Technology, Japan) and K. U. Torii (The University of Texas at Austin, USA) for sharing the materials. We also thank K. Nakajima (Nara Institute of Science and Technology, Japan) for sharing the figures of plants in Fig. 5b. We are grateful to J. Raymond for his critical readings of this manuscript. This work was supported by Ministry of Education, Culture, Sports, Science and Technology/Japan Society for the Promotion of Science (MEXT/JSPS) KAKENHI grants to M.S. (JP19K06722 and JP20H05416), K.T. (JP26711017 and JP18K06283), Y.O. (JP18K19964), T.M. (JP20H05905 and JP20H05906), I.H.-N. (JP15H05776), R.N. (JP20H04884) and T.S. (JP18K06284); Grants-in-Aid JSPS Fellows to K.C.M. (JP21J14990) and M.S. (JP12J05453); and the Takeda Science Foundation, the Kato Memorial Bioscience Foundation and the Ohsumi Frontier Science Foundation to M.S. J.L.-M. and Y.-T.L. were supported by PhD studentships from the Darwin Trust of Edinburgh.

Author information

Authors and Affiliations

Authors

Contributions

K.C.M. and T.S. conceived and designed the research in general. K.C.M. performed most of the experiments and analysed the data. M.S. and Y.M. performed the experiments on MpSETA. J.L.-M., Y.-T. L., G.I. and J.G. performed the experiments on MpICE2. K.T., Y.O., T.M., I.H.-N., R.N., J.G. and T.K. supervised the experiments. K.C.M. and T.S. wrote the manuscript. All authors read, edited and approved the manuscript.

Corresponding author

Correspondence to Tomoo Shimada.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Plants thanks John Bowman and Caspar Chater for their contribution to the peer review of this work.

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 Comparison of the domain architecture of Ia bHLHs in land plants.

a, A diagram of the domain architecture of MpSETA (M. polymorpha), PpSMF1, PpSMF2 (P. patens), AtSPCH, AtMUTE, and AtFAMA (A. thaliana). While no PEST domain was identified, MpSETA has a bHLH domain and SMF domain conserved at the C-terminus like other Ia bHLH proteins. SMF domain is structurally considered to be the ACT-like domain, which is a putative domain for protein-protein dimerization. b, Sequence alignment of the bHLH domain of Ia bHLH proteins. Ia bHLHs are surrounded by a black box, and others are Ib(1) bHLHs. Asterisks indicate amino acids that are assumed to be important for binding to the E-box (CANNTG), and the triangles indicate amino acids that are assumed to be important for the dimerization of the bHLH domain. The yellow box indicates the LxCxE motif, which is a binding motif with Retinoblastoma-related (RBR). c, Sequence alignment of the C-terminal SMF domain of Ia bHLH proteins.

Extended Data Fig. 2 Function of MpSETA in A. thaliana Ia bHLH mutants.

a, Confocal images of A. thaliana abaxial cotyledons of wild type (Col-0), spch-3, and proAtSPCH:MpSETA spch-3 at 9 days after stratification (DAS). b, Confocal images of A. thaliana abaxial cotyledons of wild type (Col-0), fama-1, and proAtFAMA:MpSETA fama-1 at 9 DAS. Brackets and arrows indicate fama tumors and stomatal-lineage cells, respectively. c, Quantitative data of the distribution of the number of cell divisions that occurred in the stomatal lineage in each genotype. (n > 320 cells per genotype, 9 DAS cotyledons). d, Y2H assays in which the MpSETA fused with the GAL4 DNA-binding domain (DBD) was used as bait, and the AtICE1 and AtSCRM2 fused with the GAL4 activation domain (AD) were used as prey. DBD alone and AD alone were used as negative controls. e, BiFC assays showing the interaction between MpSETA and AtICE1 or AtSCRM2 in N. benthamiana leaf epidermal cells. MpSETA was fused to the N-terminal fragment of EYFP (nYFP), whereas AtICE1 or AtSCRM2 was fused to the C-terminal fragment of EYFP (cYFP). nYFP alone and cYFP alone were used as the negative controls. Nuclei were stained with DAPI. The experiments in this figure were repeated at least three times with similar results. Bars, 10 µm (e), and 100 µm (a,b).

Extended Data Fig. 3 Expression analysis of MpSETA in the gametophytic tissues.

Histochemical detection of β-glucuronidase (GUS) activity driven by the MpSETA promoter in the developing antheridia. The experiments in this figure were repeated at three times with similar results. Bars, 1 mm.

Extended Data Fig. 4 Generation and phenotypes of MpSETA knock-out lines.

a, Structure of the MpSETA locus disrupted by homologous recombination. Knock-out lines have a deletion in the bHLH domain coding region. White boxes indicate the exons of the MpSETA coding sequence. DT-A, diphtheria toxin A fragment gene; HgrR, hygromycin- resistance gene. b, Genotyping of the Mpsetako lines used in this study to distinguish sex. rbm27, a male-specific marker; rhf73, a female-specific marker. c, Genotyping of the Mpsetako lines. The position of the primers used for PCR is shown in (a). M, Male; F, Female. d, RT-PCR to confirm the loss of the full-length MpSETA transcript in Mpsetako lines in 21 DPF sporophytes. MpEF1α was used as an internal control. e, Spermatogenesis process in the wild type (WT) and Mpsetako lines. All the images are at the same scale. The experiments in this figure were repeated at least three times with similar results. Bars, 10 μm (e).

Extended Data Fig. 5 Phylogenetic tree of IIIb bHLH TFs.

A maximum-likelihood bHLH phylogenetic tree of subfamilies IIIb, III (a + c) (light blue), and III(d + e) (grey, outgroup) is shown. Numbers at branches indicate bootstrap values calculated from 1,000 replicates. The scale bar indicates the substitution rate per residue. IIIb bHLHs are divided into 2 groups: ICE/SCRM clade (orange) and NFL clade (magenta). Species are abbreviated as follows: Mp, M. polymorpha (liverwort); Lc, L. cruciata (liverwort); Pp, P. patens (moss); Cepur, Ceratodon purpureus (moss); Aagr, Anthoceros agrestis (hornwort); Sm, Selaginella moellendorffii (lycophyte); AmTr, Amborella trichopoda (basal angiosperm); Os, Oryza sativa (monocot); At, A. thaliana (dicot). Arrows indicate MpICE1 (Mp4g04910) and MpICE2 (Mp4g04920). For the phylogenetic construction of subfamilies III(a + c) and III(d + e), we used amino acid sequences from only A. thaliana and M. polymorpha.

Extended Data Fig. 6 Comparison of the domain architecture of IIIb bHLHs in land plants.

a, A diagram of the domain architecture of MpICE1, MpICE2 (M. polymorpha), PpSCRM1 (P. patens), AtICE1, and AtSCRM2 (A. thaliana). MpICE1 and MpICE2 have a bHLH domain and ACT-like domain conserved at the C-terminus like other IIIb bHLH proteins. b, Sequence alignment of the bHLH domain of the IIIb bHLH proteins. IIIb bHLHs are surrounded by a black box, and others are an outgroup. Asterisks indicate amino acids that are assumed to be important for binding to the E-box (CANNTG). c, Sequence alignment of the C-terminal ACT-like domain of the IIIb bHLH proteins.

Extended Data Fig. 7 The expression analysis of MpICE2.

a, Histochemical detection of β-glucuronidase (GUS) activity driven by the MpICE2 promoter in the vegetative thallus. b, Confocal images of the dorsal epidermis of proMpICE2:Citrine-GUS-NLS line. The upper and lower panels indicate the epidermis around the apical notch and the epidermis around the midrib, respectively. Arrows indicate the air pores. c,d, Histochemical detection of GUS activity driven by the MpICE2 promoter in the gametophytic reproductive organs. An antheridiophore (c) and an archegoniophore (d) are shown. e, Expression pattern of MpICE2 in developing sporophytes. f, foot; s, seta; at, archesporial tissue; sp, sporangium; ca, calyptra; p, pseudoperianth (n). Arrowheads indicate the cell wall of the first cell division. The experiments in this figure were repeated at least three times with similar results. Bars, 5 mm (a,c,d), 100 μm (b and e).

Extended Data Fig. 8 Generation of Mpice2 mutants by CRISPR/Cas9.

a, Schematic representation of the MpICE2 gene and the resulting mutations in the obtained CRISPR/Cas9-generated alleles. Gray, white, and blue boxes indicate the coding sequences (CDS), the untranslated regions (UTR), and the bHLH domain coding region, respectively. b, Sequence alignment of putative translational products of wild type (WT) and Mpice2ge mutants. Asterisks indicate the amino acids that are assumed to be important for binding to the E-box.

Extended Data Fig. 9 Functional analysis of MpICE1 and MpICE2 in A. thaliana mutants.

a, Confocal images of A. thaliana abaxial cotyledons of wild type (Ws-4), ice1-2 scrm2-1, and proAtMUTE:MpSETA mute-2 expressing MpICE2 at 9 DAS. Arrowheads and asterisks indicate stomata and hydathode pores, respectively. The experiments were performed once. b, Confocal images of A. thaliana abaxial leaves of wild type (Col-0), ice1-2 scrm2-1, proAtICE1:MpICE1 ice1-2 scrm2-1, and proAtICE1:MpICE2 ice1-2 scrm2-1 at 13 DAS. The experiments were repeated three times with similar results. Bars, 100 µm.

Supplementary information

Supplementary Information

Supplementary Tables 1 and 2.

Reporting Summary

Source data

Source Data Fig. 1

Unprocessed gels.

Source Data Fig. 2

Statistical source data.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Moriya, K.C., Shirakawa, M., Loue-Manifel, J. et al. Stomatal regulators are co-opted for seta development in the astomatous liverwort Marchantia polymorpha. Nat. Plants 9, 302–314 (2023). https://doi.org/10.1038/s41477-022-01325-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41477-022-01325-5

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