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:

The endocytic TPLATE complex internalizes ubiquitinated plasma membrane cargo

An Author Correction to this article was published on 12 January 2023

This article has been updated

Abstract

Endocytosis controls the perception of stimuli by modulating protein abundance at the plasma membrane. In plants, clathrin-mediated endocytosis is the most prominent internalization pathway and relies on two multimeric adaptor complexes, the AP-2 and the TPLATE complex (TPC). Ubiquitination is a well-established modification triggering endocytosis of cargo proteins, but how this modification is recognized to initiate the endocytic event remains elusive. Here we show that TASH3, one of the large subunits of TPC, recognizes ubiquitinated cargo at the plasma membrane via its SH3 domain-containing appendage. TASH3 lacking this evolutionary specific appendage modification allows TPC formation but the plants show severely reduced endocytic densities, which correlates with reduced endocytic flux. Moreover, comparative plasma membrane proteomics identified differential accumulation of multiple ubiquitinated cargo proteins for which we confirm altered trafficking. Our findings position TPC as a key player for ubiquitinated cargo internalization, allowing future identification of target proteins under specific stress conditions.

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: The viable nosh mutant in the TPC subunit TASH3 has reduced endocytic capacity.
Fig. 2: The mutant nosh is defective in endocytosis.
Fig. 3: The association between the hexameric core an the AtEH/PAn1 subunits is destabilized in nosh.
Fig. 4: The SH3 domain of TASH3 binds poly-ubiquitin chains.
Fig. 5: Ubiquitinated PM proteins accumulate in nosh seedlings.
Fig. 6: The SH3 domain of TASH3 recognizes ubiquitinated cargo.

Similar content being viewed by others

Data availability

All materials are available from the corresponding authors upon request. All data generated or analysed during this study are included in this article and its Extended Data) and/or in public repositories. The raw mass spectrometry data and MaxQuant result files have been deposited to the ProteomeXchange Consortium via PRIDE (PXD035444). The script for comparative quantification of fluorescent signal at PM versus cytoplasm is available for download (https://github.com/pegro-psb/Cyto-PM-signal-quantification). Araport11plus database consisting of the Araport11_genes.2016.06.pep.fasta downloaded from arabidopsis.org was used for MS analysis. Source data are provided with this paper.

Change history

References

  1. Bitsikas, V., Corrêa, I. R. & Nichols, B. J. Clathrin-independent pathways do not contribute significantly to endocytic flux. eLife 3, e03970 (2014).

  2. Kaksonen, M. & Roux, A. Mechanisms of clathrin-mediated endocytosis. Nat. Rev. Mol. Cell Biol. 19, 313–326 (2018).

    Article  CAS  Google Scholar 

  3. Reynolds, G. D., Wang, C., Pan, J. & Bednarek, S. Y. Inroads into internalization: five years of endocytic exploration. Plant Physiol. 176, 208–218 (2018).

    Article  CAS  Google Scholar 

  4. Gadeyne, A. et al. The TPLATE adaptor complex drives clathrin-mediated endocytosis in plants. Cell 156, 691–704 (2014).

    Article  CAS  Google Scholar 

  5. Hirst, J. et al. Characterization of TSET, an ancient and widespread membrane trafficking complex. eLife 3, e02866 (2014).

    Article  Google Scholar 

  6. Van Damme, D. et al. Somatic cytokinesis and pollen maturation in Arabidopsis depend on TPLATE, which has domains similar to coat proteins. Plant Cell 18, 3502–3518 (2006).

    Article  Google Scholar 

  7. Bashline, L., Li, S., Zhu, X. & Gu, Y. The TWD40-2 protein and the AP2 complex cooperate in the clathrin-mediated endocytosis of cellulose synthase to regulate cellulose biosynthesis. Proc. Natl Acad. Sci. USA 112, 12870–12875 (2015).

    Article  CAS  Google Scholar 

  8. Ohno, H. et al. The medium subunits of adaptor complexes recognize distinct but overlapping sets of tyrosine-based sorting signals. J. Biol. Chem. 273, 25915–25921 (1998).

    Article  CAS  Google Scholar 

  9. Mattera, R., Boehm, M., Chaudhuri, R., Prabhu, Y. & Bonifacino, J. S. Conservation and diversification of dileucine signal recognition by adaptor protein (AP) complex variants. J. Biol. Chem. 286, 2022 (2011).

    Article  CAS  Google Scholar 

  10. Arora, D. & van Damme, D. Motif-based endomembrane trafficking. Plant Physiol. 186, 221–238 (2021).

    Article  CAS  Google Scholar 

  11. Happel, N. et al. Arabidopsis µA-adaptin interacts with the tyrosine motif of the vacuolar sorting receptor VSR-PS1. Plant J. 37, 678–693 (2004).

    Article  CAS  Google Scholar 

  12. Takano, J. et al. Polar localization and degradation of Arabidopsis boron transporters through distinct trafficking pathways. Proc. Natl Acad. Sci. USA 107, 5220–5225 (2010).

    Article  CAS  Google Scholar 

  13. Mravec, J. et al. Subcellular homeostasis of phytohormone auxin is mediated by the ER-localized PIN5 transporter. Nature 459, 1136–1140 (2009).

    Article  CAS  Google Scholar 

  14. Yoshinari, A. et al. Polar localization of the borate exporter BOR1 requires AP2-dependent endocytosis. Plant Physiol. 179, 1569–1580 (2019).

    Article  CAS  Google Scholar 

  15. Liu, D. et al. Endocytosis of BRASSINOSTEROID INSENSITIVE1 is partly driven by a canonical Tyr-based motif. Plant Cell 32, 3598–3612 (2020).

    Article  CAS  Google Scholar 

  16. Yperman, K. et al. Distinct EH domains of the endocytic TPLATE complex confer lipid and protein binding. Nat. Commun. 12, 3050 (2021).

  17. Robatzek, S., Chinchilla, D. & Boller, T. Ligand-induced endocytosis of the pattern recognition receptor FLS2 in Arabidopsis. Genes Dev. 20, 537 (2006).

    Article  CAS  Google Scholar 

  18. Wang, S. et al. Auxin-related gene families in abiotic stress response in Sorghum bicolor. Funct. Integr. Genomics 10, 533–546 (2010).

    Article  CAS  Google Scholar 

  19. Erwig, J. et al. Chitin-induced and CHITIN ELICITOR RECEPTOR KINASE1 (CERK1) phosphorylation-dependent endocytosis of Arabidopsis thaliana LYSIN MOTIF-CONTAINING RECEPTOR-LIKE KINASE5 (LYK5). New Phytol. 215, 382–396 (2017).

    Article  CAS  Google Scholar 

  20. Traub, L. M. Tickets to ride: selecting cargo for clathrin-regulated internalization. Nat. Rev. Mol. Cell Biol. 10, 583–596 (2009).

    Article  CAS  Google Scholar 

  21. Dubeaux, G. & Vert, G. Zooming into plant ubiquitin-mediated endocytosis. Curr. Opin. Plant Biol. 40, 56–62 (2017).

    Article  CAS  Google Scholar 

  22. Fan, L., Li, R., Pan, J., Ding, Z. & Lin, J. Endocytosis and its regulation in plants. Trends Plant Sci. 20, 388–397 (2015).

    Article  CAS  Google Scholar 

  23. Nagel, M. K. et al. Arabidopsis SH3P2 is an ubiquitin-binding protein that functions together with ESCRT-I and the deubiquitylating enzyme AMSH3. Proc. Natl Acad. Sci. USA 114, E7197–E7204 (2017).

    Article  CAS  Google Scholar 

  24. Weinberg, J. S. & Drubin, D. G. Regulation of clathrin-mediated endocytosis by dynamic ubiquitination and deubiquitination. Curr. Biol. 24, 951–959 (2014).

    Article  CAS  Google Scholar 

  25. Yoshinari, A. et al. DRP1-dependent endocytosis is essential for polar localization and boron-induced degradation of the borate transporter BOR1 in Arabidopsis thaliana. Plant Cell Physiol. 57, 1985–2000 (2016).

    Article  CAS  Google Scholar 

  26. Leitner, J. et al. Lysine63-linked ubiquitylation of PIN2 auxin carrier protein governs hormonally controlled adaptation of Arabidopsis root growth. Proc. Natl Acad. Sci. USA 109, 8322–8327 (2012).

    Article  CAS  Google Scholar 

  27. Barberon, M. et al. Monoubiquitin-dependent endocytosis of the iron-regulated transporter 1 (IRT1) transporter controls iron uptake in plants. Proc. Natl Acad. Sci. USA 108, E450–E458 (2011).

    Article  CAS  Google Scholar 

  28. Vert, G. et al. IRT1, an Arabidopsis transporter essential for iron uptake from the soil and for plant growth. Plant Cell 14, 1223–1233 (2002).

    Article  CAS  Google Scholar 

  29. Liao, D. et al. Arabidopsis E3 ubiquitin ligase PLANT U-BOX13 (PUB13) regulates chitin receptor LYSIN MOTIF RECEPTOR KINASE5 (LYK5) protein abundance. New Phytol. 214, 1646–1656 (2017).

    Article  CAS  Google Scholar 

  30. Lu, D. et al. Direct ubiquitination of pattern recognition receptor FLS2 attenuates plant innate immunity. Science 332, 1439–1442 (2011).

    Article  CAS  Google Scholar 

  31. Zhou, J. et al. Regulation of Arabidopsis brassinosteroid receptor BRI1 endocytosis and degradation by plant U-box PUB12/PUB13-mediated ubiquitination. Proc. Natl Acad. Sci. USA 115, E1906–E1915 (2018).

    Article  CAS  Google Scholar 

  32. Korbei, B. et al. Arabidopsis TOL proteins act as gatekeepers for vacuolar sorting of PIN2 plasma membrane protein. Curr. Biol. 23, 2500–2505 (2013).

    Article  CAS  Google Scholar 

  33. Moulinier-Anzola, J. et al. TOLs function as ubiquitin receptors in the early steps of the ESCRT pathway in higher plants. Mol. Plant 13, 717–731 (2020).

    Article  CAS  Google Scholar 

  34. Arora, D. et al. Establishment of proximity-dependent biotinylation approaches in different plant model systems. Plant Cell 32, 3388–3407 (2020).

    Article  CAS  Google Scholar 

  35. Wang, J. et al. Conditional destabilization of the TPLATE complex impairs endocytic internalization. Proc. Natl Acad. Sci. USA 118, e202345611 (2021).

    Google Scholar 

  36. Yperman, K. et al. Molecular architecture of the endocytic TPLATE complex. Sci. Adv. 7, eabe7999 (2021).

    Article  CAS  Google Scholar 

  37. Wang, P. et al. Plant AtEH/Pan1 proteins drive autophagosome formation at ER-PM contact sites with actin and endocytic machinery. Nat. Commun. 10, 5132 (2019).

    Article  Google Scholar 

  38. Van Damme, D. et al. Adaptin-like protein TPLATE and clathrin recruitment during plant somatic cytokinesis occurs via two distinct pathways. Proc. Natl Acad. Sci. USA 108, 615–620 (2011).

    Article  Google Scholar 

  39. Boutté, Y. et al. Endocytosis restricts Arabidopsis KNOLLE syntaxin to the cell division plane during late cytokinesis. EMBO J. 29, 546–558 (2010).

    Article  Google Scholar 

  40. Mayer, B. J., Hamaguchi, M. & Hanafusa, H. A novel viral oncogene with structural similarity to phospholipase C. Nature 332, 272–275 (1988).

    Article  CAS  Google Scholar 

  41. Kurochkina, N. & Guha, U. SH3 domains: modules of protein–protein interactions. Biophys. Rev. 5, 29–39 (2013).

    Article  CAS  Google Scholar 

  42. Stamenova, S. D. et al. Ubiquitin binds to and regulates a subset of SH3 domains. Mol. Cell 25, 273 (2007).

    Article  CAS  Google Scholar 

  43. Kang, J., Kang, S., Hyuk, N. K., He, W. & Park, S. Distinct interactions between ubiquitin and the SH3 domains involved in immune signaling. Biochim. Biophys. Acta 1784, 1335–1341 (2008).

    Article  CAS  Google Scholar 

  44. Bu, F., Yang, M., Guo, X., Huang, W. & Chen, L. Multiple functions of ATG8 family proteins in plant autophagy. Front. Cell Dev. Biol. 8, 466 (2020).

    Article  Google Scholar 

  45. Schwihla, M. & Korbei, B. The beginning of the end: initial steps in the degradation of plasma membrane proteins. Front. Plant Sci. 11, 680 (2020).

    Article  Google Scholar 

  46. Bashline, L., Li, S., Anderson, C. T., Lei, L. & Gu, Y. The endocytosis of cellulose synthase in Arabidopsis is dependent on μ2, a clathrin-mediated endocytosis adaptin. Plant Physiol. 163, 150–160 (2013).

    Article  CAS  Google Scholar 

  47. Lian, N. et al. COP1 mediates dark-specific degradation of microtubule-associated protein WDL3 in regulating Arabidopsis hypocotyl elongation. Proc. Natl Acad. Sci. USA 114, 12321–12326 (2017).

    Article  CAS  Google Scholar 

  48. Kim, J. H. & Kim, W. T. The Arabidopsis RING E3 ubiquitin ligase AtAIRP3/LOG2 participates in positive regulation of high-salt and drought stress responses. Plant Physiol. 162, 1733–1749 (2013).

    Article  CAS  Google Scholar 

  49. Svozil, J., Hirsch-Hoffmann, M., Dudler, R., Gruissem, W. & Baerenfaller, K. Protein abundance changes and ubiquitylation targets identified after inhibition of the proteasome with syringolin A. Mol. Cell. Proteom. 13, 1523–1536 (2014).

    Article  CAS  Google Scholar 

  50. Walton, A. et al. It’s time for some ‘site’-seeing: novel tools to monitor the ubiquitin landscape in Arabidopsis thaliana. Plant Cell 28, 6–16 (2016).

    Article  CAS  Google Scholar 

  51. Johnson, A. & Vert, G. Unraveling K63 polyubiquitination networks by sensor-based proteomics. Plant Physiol. 171, 1808–1820 (2016).

    Article  CAS  Google Scholar 

  52. Aguilar-Hernández, V. et al. Mass spectrometric analyses reveal a central role for ubiquitylation in remodeling the Arabidopsis proteome during photomorphogenesis. Mol. Plant 10, 846–865 (2017).

    Article  Google Scholar 

  53. Romero-Barrios, N. et al. Advanced cataloging of lysine-63 polyubiquitin networks by genomic, interactome, and sensor-based proteomic analyses. Plant Cell 32, 123–138 (2020).

    Article  CAS  Google Scholar 

  54. Grubb, L. E. et al. Large-scale identification of ubiquitination sites on membrane-associated proteins in Arabidopsis thaliana seedlings. Plant Physiol. 185, 1483–1488 (2021).

    Article  CAS  Google Scholar 

  55. Martins, S. et al. Internalization and vacuolar targeting of the brassinosteroid hormone receptor BRI1 are regulated by ubiquitination. Nat. Commun. 6, 6151 (2015).

    Article  CAS  Google Scholar 

  56. Winkler, J. et al. Nanobody-dependent delocalization of endocytic machinery in Arabidopsis root cells dampens their internalization capacity. Front. Plant Sci. 12, 538580 (2021).

    Article  Google Scholar 

  57. Fan, L. et al. Dynamic analysis of Arabidopsis AP2 σ subunit reveals a key role in clathrin-mediated endocytosis and plant development. Development 140, 3826–3837 (2013).

    Article  CAS  Google Scholar 

  58. Wendrich, J. R. et al. Vascular transcription factors guide plant epidermal responses to limiting phosphate conditions. Science 370, eaay4970 (2020).

    Article  CAS  Google Scholar 

  59. Berrío, R. T. et al. Single-cell transcriptomics sheds light on the identity and metabolism of developing leaf cells. Plant Physiol. 188, 898–918 (2022).

    Article  Google Scholar 

  60. Song, Q., Ando, A., Jiang, N., Ikeda, Y. & Chen, Z. J. Single-cell RNA-seq analysis reveals ploidy-dependent and cell-specific transcriptome changes in Arabidopsis female gametophytes. Genome Biol. 21, 178 (2020).

    Article  CAS  Google Scholar 

  61. Lin, D. et al. Rho GTPase signaling activates microtubule severing to promote microtubule ordering in Arabidopsis. Curr. Biol. 23, 290–297 (2013).

    Article  CAS  Google Scholar 

  62. Abas, L. et al. Intracellular trafficking and proteolysis of the Arabidopsis auxin-efflux facilitator PIN2 are involved in root gravitropism. Nat. Cell Biol. 8, 249–256 (2006).

    Article  CAS  Google Scholar 

  63. Leitner, J., Retzer, K., Korbei, B. & Luschnig, C. Dynamics in PIN2 auxin carrier ubiquitylation in gravity-responding Arabidopsis roots. Plant Signal Behav. 7, 1271–1273 (2012).

    Article  CAS  Google Scholar 

  64. Yoshinari, A. et al. Transport-coupled ubiquitination of the borate transporter BOR1 for its boron-dependent degradation. Plant Cell 33, 420–438 (2021).

    Article  Google Scholar 

  65. Luo, Y. et al. Deubiquitinating enzymes UBP12 and UBP13 stabilize the brassinosteroid receptor BRI1. EMBO Rep. 23, e53354 (2022).

    Article  CAS  Google Scholar 

  66. Hicke, L., Schubert, H. L. & Hill, C. P. Ubiquitin-binding domains. Nat. Rev. Mol. Cell Biol. 6, 610–621 (2005).

    Article  CAS  Google Scholar 

  67. Hawryluk, M. J. et al. Epsin 1 is a polyubiquitin-selective clathrin-associated sorting protein. Traffic 7, 262–281 (2006).

    Article  CAS  Google Scholar 

  68. Sorkina, T. et al. RNA interference screen reveals an essential role of Nedd4-2 in dopamine transporter ubiquitination and endocytosis. J. Neurosci. 26, 8195–8205 (2006).

    Article  CAS  Google Scholar 

  69. Haglund, K. & Dikic, I. The role of ubiquitylation in receptor endocytosis and endosomal sorting. J. Cell Sci. 125, 265–275 (2012).

    Article  CAS  Google Scholar 

  70. Mayers, J. R. et al. Regulation of ubiquitin-dependent cargo sorting by multiple endocytic adaptors at the plasma membrane. Proc. Natl Acad. Sci. USA 110, 11857–11862 (2013).

    Article  CAS  Google Scholar 

  71. Schuh, A. L. & Audhya, A. The ESCRT machinery: from the plasma membrane to endosomes and back again. Crit. Rev. Biochem. Mol. Biol. 49, 242–261 (2014).

    Article  CAS  Google Scholar 

  72. Adamowski, M., Matijević, I., Narasimhan, M. & Friml, J. SH3Ps recruit auxilin-like vesicle uncoating factors into clathrin-mediated endocytosis. Preprint at bioRxiv https://doi.org/10.1101/2022.01.07.475403 (2022).

  73. Pashkova, N. et al. Article WD40 repeat propellers define a ubiquitin-binding domain that regulates turnover of F box proteins. Mol. Cell 40, 433–443 (2010).

    Article  CAS  Google Scholar 

  74. Isono, E. et al. The deubiquitinating enzyme AMSH3 is required for intracellular trafficking and vacuole biogenesis in Arabidopsis thaliana. Plant Cell 22, 1826–1837 (2010).

    Article  CAS  Google Scholar 

  75. Katsiarimpa, A. et al. The Arabidopsis deubiquitinating enzyme AMSH3 interacts with ESCRT-III subunits and regulates their localization. Plant Cell 23, 3026–3040 (2011).

    Article  CAS  Google Scholar 

  76. Katsiarimpa, A. et al. The deubiquitinating enzyme AMSH1 and the ESCRT-III subunit VPS2.1 are required for autophagic degradation in Arabidopsis. Plant Cell 25, 2236 (2013).

    Article  CAS  Google Scholar 

  77. Katsiarimpa, A. et al. The ESCRT-III-interacting deubiquitinating enzyme AMSH3 is essential for degradation of ubiquitinated membrane proteins in Arabidopsis thaliana. Plant Cell Physiol. 55, 727–736 (2014).

    Article  CAS  Google Scholar 

  78. Ingouff, M. et al. Live-cell analysis of DNA methylation during sexual reproduction in Arabidopsis reveals context and sex-specific dynamics controlled by noncanonical RdDM. Genes Dev. 31, 72–83 (2017).

    Article  CAS  Google Scholar 

  79. Karimi, M., Depicker, A. & Hilson, P. Recombinational cloning with plant gateway vectors. Plant Physiol. 145, 1144 (2007).

    Article  CAS  Google Scholar 

  80. Lampropoulos, A. et al. GreenGate—a novel, versatile, and efficient cloning system for plant transgenesis. PLoS ONE 8, e83043 (2013).

    Article  Google Scholar 

  81. Waadt, R. et al. Dual-reporting transcriptionally linked genetically encoded fluorescent indicators resolve the spatiotemporal coordination of cytosolic abscisic acid and second messenger dynamics in Arabidopsis. Plant Cell 32, 2582–2601 (2020).

    Article  CAS  Google Scholar 

  82. Waadt, R., Krebs, M., Kudla, J. & Schumacher, K. Multiparameter imaging of calcium and abscisic acid and high-resolution quantitative calcium measurements using R-GECO1-mTurquoise in Arabidopsis. New Phytol. 216, 303–320 (2017).

    Article  CAS  Google Scholar 

  83. Decaestecker, W. et al. CRISPR-TSKO: a technique for efficient mutagenesis in specific cell types, tissues, or organs in Arabidopsis. Plant Cell 31, 2868–2887 (2019).

    Article  CAS  Google Scholar 

  84. Mertens, N., Remaut, E. & Fiers, W. Versatile, multi-featured plasmids for high-level expression of heterologous genes in Escherichia coli: overproduction of human and murine cytokines. Gene 164, 9–15 (1995).

    Article  CAS  Google Scholar 

  85. Kim, S. Y. et al. Adaptor protein complex 2-mediated endocytosis is crucial for male reproductive organ development in Arabidopsis. Plant Cell 25, 2970–2985 (2013).

    Article  CAS  Google Scholar 

  86. Mravec, J. et al. Report cell plate restricted association of DRP1A and PIN proteins is required for cell polarity establishment in Arabidopsis. Curr. Biol. 21, 1055–1060 (2011).

    Article  CAS  Google Scholar 

  87. Reichardt, I. et al. Plant cytokinesis requires de novo secretory trafficking but not endocytosis. Curr. Biol. 17, 2047–2053 (2007).

    Article  CAS  Google Scholar 

  88. Dejonghe, W. et al. Disruption of endocytosis through chemical inhibition of clathrin heavy chain function. Nat. Chem. Biol. 15, 641–649 (2019).

    Article  CAS  Google Scholar 

  89. Xu, T. et al. Cell surface ABP1-TMK auxin-sensing complex activates ROP GTPase signaling. Science 343, 1025–1028 (2014).

    Article  CAS  Google Scholar 

  90. Di Rubbo, S. et al. The clathrin adaptor complex AP-2 mediates endocytosis of brassinosteroid insensitive1 in Arabidopsis. Plant Cell 25, 2986–2997 (2013).

    Article  Google Scholar 

  91. Belkhadir, Y. et al. Brassinosteroids modulate the efficiency of plant immune responses to microbe-associated molecular patterns. Proc. Natl Acad. Sci. USA 109, 297–302 (2012).

    Article  CAS  Google Scholar 

  92. Sparkes, I. A., Runions, J., Kearns, A. & Hawes, C. Rapid, transient expression of fluorescent fusion proteins in tobacco plants and generation of stably transformed plants. Nat. Protoc. 1, 2019–2025 (2006).

    Article  CAS  Google Scholar 

  93. Schindelin, J. et al. Fiji: an open-source platform for biological-image analysis. Nat. Methods 9, 676–682 (2012).

    Article  CAS  Google Scholar 

  94. Altschul, S. F. et al. Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res. 25, 3389–3402 (1997).

    Article  CAS  Google Scholar 

  95. Letunic, I. & Bork, P. 20 years of the SMART protein domain annotation resource. Nucleic Acids Res. 46, D493–D496 (2018).

    Article  CAS  Google Scholar 

  96. Katoh, K., Rozewicki, J. & Yamada, K. D. MAFFT online service: multiple sequence alignment, interactive sequence choice and visualization. Brief Bioinform. 20, 1160–1166 (2019).

    Article  CAS  Google Scholar 

  97. Guindon, S. et al. New algorithms and methods to estimate maximum-likelihood phylogenies: assessing the performance of PhyML 3.0. Syst. Biol. 59, 307–321 (2010).

    Article  CAS  Google Scholar 

  98. Lefort, V., Longueville, J. E. & Gascuel, O. SMS: smart model selection in PhyML. Mol. Biol. Evol. 34, 2422–2424 (2017).

    Article  CAS  Google Scholar 

  99. Letunic, I. & Bork, P. Interactive Tree Of Life (iTOL) v5: an online tool for phylogenetic tree display and annotation. Nucleic Acids Res. 49, W293–W296 (2021).

    Article  CAS  Google Scholar 

  100. Varadi, M. et al. AlphaFold protein structure database: massively expanding the structural coverage of protein-sequence space with high-accuracy models. Nucleic Acids Res. 50, D439–D444 (2022).

    Article  CAS  Google Scholar 

  101. Mirdita, M. et al. ColabFold: making protein folding accessible to all. Nat. Methods 19, 679–682 (2022).

    Article  CAS  Google Scholar 

  102. Desta, I. T., Porter, K. A., Xia, B., Kozakov, D. & Vajda, S. Performance and its limits in rigid body protein–protein docking. Structure 28, 1071–1081 (2020).

    Article  CAS  Google Scholar 

  103. Yan, Y., Tao, H., He, J. & Huang, S. Y. The HDOCK server for integrated protein–protein docking. Nat. Protoc. 15, 1829–1852 (2020).

    Article  CAS  Google Scholar 

  104. Pierce, B. G. et al. ZDOCK server: interactive docking prediction of protein–protein complexes and symmetric multimers. Bioinformatics 30, 1771–1773 (2014).

    Article  CAS  Google Scholar 

  105. Pettersen, E. F. et al. UCSF ChimeraX: structure visualization for researchers, educators, and developers. Protein Sci. 30, 70–82 (2021).

    Article  CAS  Google Scholar 

  106. Ashkenazy, H. et al. ConSurf 2016: an improved methodology to estimate and visualize evolutionary conservation in macromolecules. Nucleic Acids Res. 44, W344–W350 (2016).

    Article  CAS  Google Scholar 

  107. Kapust, R. B. et al. Tobacco etch virus protease: mechanism of autolysis and rational design of stable mutants with wild-type catalytic proficiency. Protein Eng. Des. Sel. 14, 993–1000 (2001).

    Article  CAS  Google Scholar 

  108. Dahhan, D. A. et al. Proteomic characterization of isolated Arabidopsis clathrin-coated vesicles reveals evolutionarily conserved and plant-specific components. Plant Cell 34, 2150–2173 (2022).

    Article  Google Scholar 

  109. Wendrich, J. R., Boeren, S., Möller, B. K., Weijers, D. & De Rybel, B. In vivo identification of plant protein complexes using IP-MS/MS. Methods Mol. Biol. 1497, 147–158 (2017).

    Article  CAS  Google Scholar 

  110. Sauer, M., Paciorek, T., Benkovó, E. & Friml, J. Immunocytochemical techniques for whole-mount in situ protein localization in plants. Nat. Protoc. 1, 98–103 (2006).

    Article  CAS  Google Scholar 

  111. von Wangenheim, D. et al. Live tracking of moving samples in confocal microscopy for vertically grown roots. eLife 6, e26792 (2017).

    Article  Google Scholar 

  112. Johnson, A. & Vert, G. Single event resolution of plant plasma membrane protein endocytosis by TIRF microscopy. Front. Plant Sci. 8, 612 (2017).

    Article  Google Scholar 

  113. Narasimhan, M. et al. Evolutionarily unique mechanistic framework of clathrin-mediated endocytosis in plants. eLife 9, e52067 (2020).

    Article  CAS  Google Scholar 

  114. Winkler, J. et al. Visualizing protein–protein interactions in plants by rapamycin-dependent delocalization. Plant Cell 33, 1101–1117 (2021).

    Article  Google Scholar 

Download references

Acknowledgements

We would like to thank T. Xu (FAFU-UCR, Fuzhou) and J. Friml (IST Austria) for sharing TMK1 seeds, G. Vert (CNRS, Toulouse) for sharing BRI1 and BRI1-25KR seeds, E. Russinova for providing 35S::eGFP seeds and R. Owens (OPPF, Research Complex, Harwell) for sharing an aliquot of the HIS-GFP vector. We express our gratitude to the VIB proteomics core facility for the help and expertise with running all MS experiments and the VIB protein core facility for the help and expertise with protein purifications. This work was supported by the European Research Council Grant T-REX 682436 (D.V.D.); the Research Foundation–Flanders (FWO) 1226420N (P.G.), 12S7222N (J.M.D.), 1124621N (A.D.M.) and G017919N (M.K.); the Czech Science Foundation 22-35680 M (R.P.); and the China Scholarship Council Grant 201906760018 (Q.J.).

Author information

Authors and Affiliations

Authors

Contributions

P.G. and D.V.D. designed the research and wrote the manuscript. P.G. performed most of the experiments. A.D.M. and K.Y. cloned and purified the SH3 domain. A.D.M. performed the SH3 ubiquitin-binding assay, partitioning assay and PM fraction analysis. J.M.D. and M.V. helped with cloning. R.P. performed in silico docking and phylogenetic analysis. D.E. performed MS analysis. M.K. raised and characterized the AtEH1/Pan1 antibody. J.N. prepared samples for AP–MS analysis. Q.J. and B.P. created the script for measuring fluorescent signal on confocal images. E.M. performed tip-tracking experiments. B.D.R., G.D.J., D.V.D. and P.G. were responsible for experimental design and research supervision. All authors contributed to finalizing the text.

Corresponding authors

Correspondence to Peter Grones or Daniël Van Damme.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Plants thanks Takashi Ueda, Jian Feng Ma and the other, anonymous, reviewer(s) 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 Phenotypical assessment of the nosh truncation.

(a) Quantification of genotyping PCR reactions on the F1 progeny of tash3-1 (), tash3-2 () and nosh () backcrossed into Col-0 () to evaluate the transfer of the T-DNA allele via the pollen. The male sterility of tash3-1 and tash3-2 mutants prevents transfer of the T-DNA insertion to the next generation (only WT band amplified). nosh mutants produce viable pollen and can transfer the T-DNA insertion to the next generation (WT band and T-DNA band amplified). (b) Amino acid sequence of the TASH3 C terminus and predicted amino acid sequence of the nosh C terminus, based on the sequencing of the T-DNA insertion site. The body part sequence of TASH3 is marked in blue and the SH3 domain sequence is depicted in green. The sequence that is altered in nosh is underlined. (c) Representative examples of 5- week-old Col-0 and nosh plants grown under long-day conditions (16 h light/8 h dark). Under these conditions, nosh exhibits predominantly delayed flowering. (d) Representative examples of 8-week-old Col-0 and nosh plants grown at 12 h light/12 h dark conditions. Under these conditions, nosh exhibits reduced rosette growth and early senescence.

Extended Data Fig. 2 Complementation of nosh with full-length TASH3-GFP restores its endocytic defects.

(a, b) Representative single-slice spinning-disk images and box plot graphs of endocytic foci densities in epidermal dark-grown hypocotyl cells of TPLATE-GFP (tplate), TASH3-GFP (tash3-1) and two independent TASH3-GFP expressing lines in the nosh background. The densities of endocytic foci in both complemented nosh mutants are similar to the values in TPLATE-GFP (tplate) and to those in the complemented tash3-1 mutant allele (TASH3-GFP in tash3-1). Numbers of quantified cells (2 cells per seedling) are indicated. The top and bottom lines of box plots represent 25th and 75th percentiles, the centre line is the median and whiskers are the full data range. The statistical test used was a two-sided Wilcoxon-signed rank test by comparing mutants to wild type. No adjustment for multiple comparisons was performed. Scale bar = 5 µm. (c, d) Representative kymographs and violin plot graphs of the life-time measurements from the spinning-disk time lapses from panel a. Analogous life-time distributions of endocytic events were observed for TASH3-GFP in nosh as for TPLATE-GFP (tplate) and TASH3-GFP (tash3-1). The number of events analysed for each independent line is indicated at the bottom of each graph. At least 12 movies from 6 seedlings were imaged and analysed for each independent transgenic line. The widest part of the violin plot represents the highest point density, whereas the top and bottom are the maximum and minimum data respectively. Red circles represent the mean and the red line represents the standard deviation. The statistical test used was a two-sided Wilcoxon-signed rank test by comparing mutants to wild type. No adjustment for multiple comparisons was performed. Scale bar = 50 µm. n.s. = not significant.

Source data

Extended Data Fig. 3 Comparative interactomics and AtEH1/Pan1 antibody specification.

(a) Graph depicting the normalized peptide intensities of TPC subunits obtained from MS analysis. For each TPC subunit, the intensities of only those peptides that were present in all experiments (for both baits and in all replicas) were averaged and normalized to the values of the corresponding bait protein. Error bars correspond to ± SD and are based on three technical repeats. The results show that nosh does not affect the hexameric TPC formation, but that the association with the AtEH/Pan1 proteins is weakened. (b) Coomassie-stained gel of the obtained SEC fractions and a quality control HPLC analysis performed using a Superdex 200 increase 10/300 for the batch of recombinant AtEH1 C-term fragment used for rabbit immunization. Fractions 18–23 were pooled to immunize (marked in green). (c) Coomassie-stained gel of the obtained SEC fractions and a quality control HPLC analysis performed using a Superdex 200 increase 10/300 for the batch used for antibody purification. Fractions 25–29 were pooled for purification (marked in green). (d) Stain-free gel and blot for testing AtEH1/Pan1 antibody specificity. In the Col-0 sample only the native AtEH1/Pan1 band is prominently observed (marked with a black arrowhead), while in the pH3.3::AtEH1/Pan1-mRuby3 (Col-0) sample, both the native and the transgene fusion protein (marked with a white arrowhead) can be observed. Two homozygous pH3.3::AtEH1/Pan1-mRuby3 (ateh1/pan1 1-2 −/−) lines show only the transgene fusion protein (marked with white arrowhead).

Source data

Extended Data Fig. 4 A TSAUCER-like variant is produced in nosh.

(a) TASH3 protein structure with annotated peptides identified by MS analysis. Peptides marked in green were present in both samples, while peptides in magenta were missing in the TPLATE-GFP (nosh/tplate) sample. (b) Amino acid sequence alignment of TASH3 (Arabidopsis) and TSAUCER (Dictyostelium) showing that the linker and SH3 domain are missing at the C terminus of the TSAUCER protein (outlined in red). (c) Phylogenetic tree representing the maximum likelihood phylogeny of TASH3. Numbers at nodes correspond to the approximate likelihood ratio test with Shimodaira–Hasegawa-like support. Red circles mark the presence of SH3 domain(s).

Source data

Extended Data Fig. 5 TASH3 interacts with AtEH/Pan1 subunits through its body and not via its SH3 domain.

(a) Schematic representation of AtEH1/Pan1 and AtEH2/Pan1 proteins with PxxP motifs indicated by orange lines. (b–d) Maximal intensity projection of representative Z-stacks of transiently expressed TASH3-GFP, TASH3_body-GFP and mCherry-TASH3_linker_SH3 in epidermal N. benthamiana cells, respectively. (e, f) Maximal intensity projection of representative Z-stacks of TASH3-GFP recruitment to AtEH1/Pan1-mCherry and AtEH2/Pan1-mCherry positive foci upon transient co-expression in epidermal N. benthamiana cells. (g, h) Maximal intensity projection of representative Z-stacks of TASH3_body-GFP recruitment to AtEH1/Pan1-mCherry and AtEH2/Pan1-mCherry positive foci upon transient co-expression in epidermal N. benthamiana cells. (i, j) Representative Z-stacks showing that mCherry-TASH3_linker_SH3 is not recruited to AtEH1/Pan1-GFP or AtEH2/Pan1-GFP positive foci upon transient co-expression in epidermal N. benthamiana cells. White arrowheads indicate colocalization. Scale bar = 20 µm. (k, l) Box plot graphs of the particle/cytoplasm intensity of TASH3, TASH3_body and TASH3_linker_SH3 upon transient co-expression of AtEH1/Pan1 and AtEH2/Pan1 in N. benthamiana. The number of Z-stacks analysed for each combination is indicated at the bottom of each graph. The top and bottom lines of box plots represent 25th and 75th percentiles, the centre line is the median and whiskers are the full data range. Letters represent a two-sided mixed linear model statistic used to determine the difference between samples. No adjustment for multiple comparisons was performed. One outlier is marked with a red asterix. TASH3 and TASH3_body are recruited to AtEH/Pan1, whereas TASH3_linker_SH3 is not.

Source data

Extended Data Fig. 6 Modelling of protein–protein interaction between TASH3_SH3 and Ubiquitin10 or ATG8.

Three best-scoring models from two different modelling algorithms (HDOCK and AlphaFold2) show almost identical binding interfaces between TASH3_SH3 and Ubiquitin10. In contrast, modelling ATG8a-TASH3_SH3 did not result in a single orientation of ATG8a towards the TASH3_SH3 domain as shown for three best-scoring models calculated by AlphaFold2 or HDOCK.

Extended Data Fig. 7 Recombinant TASH3_SH3 domain purification.

(a) Protein sequence of the construct used for recombinant TASH3_SH3 production. (b) Predicted structural model of amino acids 1136–1198 of TASH3 by Alphafold2. (cf) Purification of recombinant HIS-TEV-SH3 domain (TASH3). The collected fractions are indicated by a dotted line. (c) Immobilized metal affinity chromatogram of the first step in the purification. (d) Size exclusion chromatogram of the second step in the purification. The fractions collected in the first step were loaded on a HiLoad® 16/600 superdex®75 pg column. (e) Reverse Immobilized metal affinity chromatogram of the third step in the purification. HIS-TEV-SH3 collected in the second step was subjected to overnight TEV cleavage with HIS-TEV protease and loaded on a HisprepTM Fast Flow 1 ml column (HIS-TEV was removed). (f) Size exclusion chromatogram of the final step in the purification. The fractions collected in the third step were loaded on a HiLoad® 16/600 superdex®75 pg column. (g) Coomassie-stained SDS–PAGE gel of the different steps of the TASH3_SH3 domain purification process. IMAC – collected fractions from c, SEC-I – collected fractions from d, TEV cleavage - the cleavage of the HIS-TEV from the recombinant protein using a HIS-TEV protease, R.IMAC - the collected fractions from e, SEC-II – the collected fractions from f. HIS-TEV-SH3: 8.8 kDa, SH3: 7.0 kDa, HIS-TEV protease: 28 kDa. (h) Western blot detection using an anti-5xHIS antibody on the collected fractions of the second step of purification (SEC-II), showing that the collected fractions contain HIS-TEV-SH3. (i) Coomassie-stained SDS–PAGE gel showing the coupling efficiency of HIS-HRV3C-GFP and TASH3_SH3 on PierceTM NHS-Activated agarose beads. Covalent coupling of the recombinant protein to the beads can be observed as a reduction in the intensity in the flow through. (j) Stain-freeTM SDS–PAGE gel showing Col-0 extracts which were incubated with the HIS-HRV3C-GFP or TASH3_SH3 coupled beads. (k) Western blot detection using a general anti-Ubiquitin antibody (P4D1) showing Col-0 extracts which were incubated with the HIS-HRV3C-GFP or TASH3_SH3 coupled beads. The smear in the bound fraction indicates that the SH3 domain from TASH3 can bind ubiquitinated proteins as opposed to the HIS-HRV3C-GFP control. I – Input, FT - flow through, B – bound fractions.

Extended Data Fig. 8 Analysis of PM fraction purity.

(a) Immunoblot analysis of different fractions acquired using the Minute™ Plant Plasma Membrane Protein Isolation Kit. Collected fractions: total protein fraction (TP), nuclei and debris fraction (ND), cytosolic fraction (C), total membrane fraction (TM), organelle membrane fraction (OM) and plasma membrane fraction (PM). The obtained fractions were separated by SDS–PAGE, visualized using stain-free gel, blotted and probed with antibodies against the Aquaporin PIP2;7 (PIP2;7, a PM marker), cytosolic Ascorbate Peroxidase (cAPX, a cytosol marker) and Cytochrome C (CytC, a mitochondrial marker). Full-length bands of the proteins are marked with arrowheads (PIP2;7 – black arrowhead; cAPX – grey arrowhead; CytC – white arrowhead).The antibodies used are listed at the right bottom corner beneath the blots.

Source data

Supplementary information

Reporting Summary

Supplementary Data

Dataset 1: Quantitative analysis of GFP pulldown TPLATE-GFP (tplate) versus TPLATE-GFP (noshtplate) based on the MaxQuant proteingroups file of GFP pulldown samples analysed by LC–MS/MS on Q Exactive (Thermo Fisher Scientific). Dataset 2: Differential analysis of PM purified samples nosh versus Col-O, based on the MaxQuant proteingroups file of the PM samples analysed by LC–MS/MS on Q Exactive HF (Thermo Fisher Scientific). Dataset 3: Primers used for cloning and genotyping.

Supplementary Alignment File

Fasta file containing the sequences used to generate the phylogenetic tree in Extended Data Fig. 4c.

Source data

Source Data Fig. 1

Statistical source data.

Source Data Fig. 2

Statistical source data.

Source Data Fig. 3

Statistical source data.

Source Data Fig. 4

Unprocessed western blots.

Source Data Fig. 5

Unprocessed western blots and statistical source data.

Source Data Fig. 6

Unprocessed western blots.

Source Data Extended Data Fig. 2

Statistical source data.

Source Data Extended Data Fig. 3

Unprocessed western blots and statistical source data.

Source Data Extended Data Fig. 4

Statistical source data.

Source Data Extended Data Fig. 5

Statistical source data.

Source Data Extended Data Fig. 8

Unprocessed western blots.

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

Grones, P., De Meyer, A., Pleskot, R. et al. The endocytic TPLATE complex internalizes ubiquitinated plasma membrane cargo. Nat. Plants 8, 1467–1483 (2022). https://doi.org/10.1038/s41477-022-01280-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41477-022-01280-1

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