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:

Conoid extrusion regulates glideosome assembly to control motility and invasion in Apicomplexa

Abstract

Members of Apicomplexa are defined by apical cytoskeletal structures and secretory organelles, tailored for motility, invasion and egress. Gliding is powered by actomyosin-dependent rearward translocation of apically secreted transmembrane adhesins. In the human parasite Toxoplasma gondii, the conoid, composed of tubulin fibres and preconoidal rings (PCRs), is a dynamic organelle of undefined function. Here, using ultrastructure expansion microscopy, we established that PCRs serve as a hub for glideosome components including Formin1. We also identified components of the PCRs conserved in Apicomplexa, Pcr4 and Pcr5, that contain B-box zinc-finger domains, assemble in heterodimer and are essential for the formation of the structure. The fitness conferring Pcr6 tethers the PCRs to the cone of tubulin fibres. F-actin produced by Formin1 is used by Myosin H to generate the force for conoid extrusion which directs the flux of F-actin to the pellicular space, serving as gatekeeper to control parasite motility.

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: Conoid extrusion is an actomyosin-dependent process powered by MyoH and fuelled by FRM1.
Fig. 2: Five new preconoidal ring proteins.
Fig. 3: Depletion of Pcr4 or Pcr5 causes loss of PCRs.
Fig. 4: Pcr4 and Pcr5 form a complex in vitro.
Fig. 5: Pcr4 and Pcr5 localization and phenotypes are recapitulated in P. berghei.
Fig. 6: PCRs serve as a gatekeeper for glideosome entry in the pellicular space.

Similar content being viewed by others

Data availability

All data are available within the paper and its supplementary information. Sequences used in this study have been obtained from EuPathDB. HDX–MS data are available via ProteomeXchange with identifier PXD031816. All biological materials and data are available from the author upon request. Source data are provided with this paper.

References

  1. Adl, S. M. et al. Diversity, nomenclature, and taxonomy of protists. Syst. Biol. 56, 684–689 (2007).

    Article  PubMed  Google Scholar 

  2. Montoya, J. G. & Liesenfeld, O. Toxoplasmosis. Lancet 363, 1965–1976 (2004).

    Article  CAS  PubMed  Google Scholar 

  3. Phillips, M. A. et al. Malaria. Nat. Rev. Dis. Prim. 3, 17050 (2017).

    Article  PubMed  Google Scholar 

  4. Rashid, M. et al. A systematic review on modelling approaches for economic losses studies caused by parasites and their associated diseases in cattle. Parasitology 146, 129–141 (2019).

    Article  PubMed  Google Scholar 

  5. Guerin, A. & Striepen, B. The biology of the intestinal intracellular parasite Cryptosporidium. Cell Host Microbe 28, 509–515 (2020).

    Article  CAS  PubMed  Google Scholar 

  6. Frenal, K., Dubremetz, J. F., Lebrun, M. & Soldati-Favre, D. Gliding motility powers invasion and egress in Apicomplexa. Nat. Rev. Microbiol. 15, 645–660 (2017).

    Article  CAS  PubMed  Google Scholar 

  7. Gubbels, M. J. & Duraisingh, M. T. Evolution of apicomplexan secretory organelles. Int J. Parasitol. 42, 1071–1081 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Dos Santos Pacheco, N., Tosetti, N., Koreny, L., Waller, R. F. & Soldati-Favre, D. Evolution, composition, assembly, and function of the conoid in Apicomplexa. Trends Parasitol. 36, 688–704 (2020).

    Article  PubMed  Google Scholar 

  9. D’Haese, J., Mehlhorn, H. & Peters, W. Comparative electron microscope study of pellicular structures in coccidia (Sarcocystis, Besnoitia and Eimeria). Int. J. Parasitol. 7, 505–518 (1977).

    Article  PubMed  Google Scholar 

  10. Mann, T. & Beckers, C. J. Characterization of the subpellicular network, a filamentous membrane skeletal component in the parasite Toxoplasma gondii. Mol. Biochem Parasitol. 115, 257–268 (2001).

    Article  CAS  PubMed  Google Scholar 

  11. Morrissette, N. S. & Sibley, L. D. Cytoskeleton of apicomplexan parasites. Microbiol Mol. Biol. Rev. 66, 21–38 (2002).

    Article  PubMed  PubMed Central  Google Scholar 

  12. Tran, J. Q. et al. RNG1 is a late marker of the apical polar ring in Toxoplasma gondii. Cytoskeleton (Hoboken) 67, 586–598 (2010).

    Article  CAS  Google Scholar 

  13. Leung, J. M. et al. Stability and function of a putative microtubule-organizing center in the human parasite Toxoplasma gondii. Mol. Biol. Cell 28, 1361–1378 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Dubois, D. J. & Soldati-Favre, D. Biogenesis and secretion of micronemes in Toxoplasma gondii. Cell Microbiol. 21, e13018 (2019).

    Article  PubMed  Google Scholar 

  15. Ben Chaabene, R., Lentini, G. & Soldati-Favre, D. Biogenesis and discharge of the rhoptries: key organelles for entry and hijack of host cells by the Apicomplexa. Mol. Microbiol. 115, 453–465 (2021).

    Article  PubMed  Google Scholar 

  16. Besteiro, S., Michelin, A., Poncet, J., Dubremetz, J. F. & Lebrun, M. Export of a Toxoplasma gondii rhoptry neck protein complex at the host cell membrane to form the moving junction during invasion. PLoS Pathog. 5, e1000309 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  17. Jacot, D. et al. An Apicomplexan actin-binding protein serves as a connector and lipid sensor to coordinate motility and invasion. Cell Host Microbe 20, 731–743 (2016).

    Article  CAS  PubMed  Google Scholar 

  18. Tosetti, N., Dos Santos Pacheco, N., Soldati-Favre, D. & Jacot, D. Three F-actin assembly centers regulate organelle inheritance, cell-cell communication and motility in Toxoplasma gondii. eLife 8, e42669 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  19. Daher, W., Plattner, F., Carlier, M. F. & Soldati-Favre, D. Concerted action of two formins in gliding motility and host cell invasion by Toxoplasma gondii. PLoS Pathog. 6, e1001132 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  20. Baum, J. et al. A malaria parasite formin regulates actin polymerization and localizes to the parasite-erythrocyte moving junction during invasion. Cell Host Microbe 3, 188–198 (2008).

    Article  CAS  PubMed  Google Scholar 

  21. Dubremetz, J. F. Rhoptries are major players in Toxoplasma gondii invasion and host cell interaction. Cell Microbiol. 9, 841–848 (2007).

    Article  CAS  PubMed  Google Scholar 

  22. Frenal, K., Marq, J. B., Jacot, D., Polonais, V. & Soldati-Favre, D. Plasticity between MyoC- and MyoA-glideosomes: an example of functional compensation in Toxoplasma gondii invasion. PLoS Pathog. 10, e1004504 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  23. Graindorge, A. et al. The conoid associated motor MyoH is indispensable for Toxoplasma gondii entry and exit from host cells. PLoS Pathog. 12, e1005388 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  24. Heaslip, A. T., Nishi, M., Stein, B. & Hu, K. The motility of a human parasite, Toxoplasma gondii, is regulated by a novel lysine methyltransferase. PLoS Pathog. 7, e1002201 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Bertiaux, E. et al. Expansion microscopy provides new insights into the cytoskeleton of malaria parasites including the conservation of a conoid. PLoS Biol. 19, e3001020 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Koreny, L. et al. Molecular characterization of the conoid complex in Toxoplasma reveals its conservation in all apicomplexans, including Plasmodium species. PLoS Biol. 19, e3001081 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Wall, R. J. et al. SAS6-like protein in Plasmodium indicates that conoid-associated apical complex proteins persist in invasive stages within the mosquito vector. Sci. Rep. 6, 28604 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Ferreira, J. L. et al. Form follows function: variable microtubule architecture in the malaria parasite. Preprint at bioRxiv https://doi.org/10.1101/2022.04.13.488170 (2022).

  29. Mondragon, R. & Frixione, E. Ca2+-dependence of conoid extrusion in Toxoplasma gondii tachyzoites. J. Eukaryot. Microbiol 43, 120–127 (1996).

    Article  CAS  PubMed  Google Scholar 

  30. Monteiro, V. G., de Melo, E. J., Attias, M. & de Souza, W. Morphological changes during conoid extrusion in Toxoplasma gondii tachyzoites treated with calcium ionophore. J. Struct. Biol. 136, 181–189 (2001).

    Article  CAS  PubMed  Google Scholar 

  31. Del Carmen, M. G., Mondragon, M., Gonzalez, S. & Mondragon, R. Induction and regulation of conoid extrusion in Toxoplasma gondii. Cell Microbiol. 11, 967–982 (2009).

    Article  PubMed  Google Scholar 

  32. Hortua Triana, M. A., Marquez-Nogueras, K. M., Vella, S. A. & Moreno, S. N. J. Calcium signaling and the lytic cycle of the Apicomplexan parasite Toxoplasma gondii. Biochim. Biophys. Acta Mol. Cell Res. 1865, 1846–1856 (2018).

    Article  CAS  PubMed  Google Scholar 

  33. Heaslip, A. T., Ems-McClung, S. C. & Hu, K. TgICMAP1 is a novel microtubule binding protein in Toxoplasma gondii. PLoS ONE 4, e7406 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  34. Mageswaran, S. K. et al. In situ ultrastructures of two evolutionarily distant apicomplexan rhoptry secretion systems. Nat. Commun. 12, 4983 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Leung, J. M., Liu, J., Wetzel, L. A. & Hu, K. Centrin2 from the human parasite Toxoplasma gondii is required for its invasion and intracellular replication. J. Cell Sci. 132, jcs228791 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Dos Santos Pacheco, N. & Soldati-Favre, D. Coupling auxin-inducible degron system with ultrastructure expansion microscopy to accelerate the discovery of gene function in Toxoplasma gondii. Methods Mol. Biol. 2369, 121–137 (2021).

    Article  PubMed  Google Scholar 

  37. de Leon, J. C. et al. A SAS-6-like protein suggests that the Toxoplasma conoid complex evolved from flagellar components. Eukaryot. Cell 12, 1009–1019 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  38. Long, S., Anthony, B., Drewry, L. L. & Sibley, L. D. A conserved ankyrin repeat-containing protein regulates conoid stability, motility and cell invasion in Toxoplasma gondii. Nat. Commun. 8, 2236 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  39. Nagayasu, E., Hwang, Y. C., Liu, J., Murray, J. M. & Hu, K. Loss of a doublecortin (DCX)-domain protein causes structural defects in a tubulin-based organelle of Toxoplasma gondii and impairs host-cell invasion. Mol. Biol. Cell 28, 411–428 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Howard, B. L. et al. Identification of potent phosphodiesterase inhibitors that demonstrate cyclic nucleotide-dependent functions in apicomplexan parasites. ACS Chem. Biol. 10, 1145–1154 (2015).

    Article  CAS  PubMed  Google Scholar 

  41. Hu, K. et al. Cytoskeletal components of an invasion machine–the apical complex of Toxoplasma gondii. PLoS Pathog. 2, e13 (2006).

    Article  PubMed  PubMed Central  Google Scholar 

  42. Lentini, G., Dubois, D. J., Maco, B., Soldati-Favre, D. & Frenal, K. The roles of Centrin 2 and Dynein Light Chain 8a in apical secretory organelles discharge of Toxoplasma gondii. Traffic 20, 583–600 (2019).

    Article  CAS  PubMed  Google Scholar 

  43. Hammoudi, P. M., Maco, B., Dogga, S. K., Frenal, K. & Soldati-Favre, D. Toxoplasma gondii TFP1 is an essential transporter family protein critical for microneme maturation and exocytosis. Mol Microbiol. 109, 225–244 (2018).

  44. Barylyuk, K. et al. A comprehensive subcellular atlas of the Toxoplasma proteome via hyperLOPIT provides spatial context for protein functions. Cell Host Microbe 28, 752–766 e759 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Sidik, S. M. et al. A genome-wide CRISPR screen in Toxoplasma identifies essential apicomplexan genes. Cell 166, 1423–1435.e1412 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Kluska, K., Adamczyk, J. & Krężel, A. Metal binding properties, stability and reactivity of zinc fingers. Coord. Chem. Rev. 367, 18–64 (2018).

    Article  CAS  Google Scholar 

  47. Munera Lopez, J. et al. An apical protein Pcr2 is required for persistent movement by the human parasite Toxoplasma gondii. PLOS Pathog. 18, e1010776 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Tosetti, N. et al. Essential function of the alveolin network in the subpellicular microtubules and conoid assembly in Toxoplasma gondii. eLife 9, e56635 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Beck, J. R. et al. A novel family of Toxoplasma IMC proteins displays a hierarchical organization and functions in coordinating parasite division. PLoS Pathog. 6, e1001094 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  50. Carruthers, V., Giddings, O. K. & Sibley, L. D. Secretion of micronemal proteins is associated with toxoplasma invasion of host cells. Cell. Microbiol. 1, 225–235 (1999).

    Article  CAS  PubMed  Google Scholar 

  51. Bisio, H., Lunghi, M., Brochet, M. & Soldati-Favre, D. Phosphatidic acid governs natural egress in Toxoplasma gondii via a guanylate cyclase receptor platform. Nat. Microbiol 4, 420–428 (2019).

    Article  CAS  PubMed  Google Scholar 

  52. Dos Santos Pacheco, N. et al. Revisiting the role of Toxoplasma gondii ERK7 in the maintenance and stability of the apical complex. mBio 12, e0205721 (2021).

    Article  PubMed  Google Scholar 

  53. Cassandri, M. et al. Zinc-finger proteins in health and disease. Cell Death Disco. 3, 17071 (2017).

    Article  Google Scholar 

  54. Jumper, J. et al. Highly accurate protein structure prediction with AlphaFold. Nature 596, 583–589 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. James, E. I., Murphree, T. A., Vorauer, C., Engen, J. R. & Guttman, M. Advances in hydrogen/deuterium exchange mass spectrometry and the pursuit of challenging biological systems. Chem. Rev. 122, 7562–7623 (2021).

  56. Bannister, L. H., Hopkins, J. M., Fowler, R. E., Krishna, S. & Mitchell, G. H. Ultrastructure of rhoptry development in Plasmodium falciparum erythrocytic schizonts. Parasitology 121, 273–287 (2000).

    Article  PubMed  Google Scholar 

  57. Hanssen, E. et al. Electron tomography of Plasmodium falciparum merozoites reveals core cellular events that underpin erythrocyte invasion. Cell Microbiol. 15, 1457–1472 (2013).

    Article  CAS  PubMed  Google Scholar 

  58. Bushell, E. et al. Functional profiling of a plasmodium genome reveals an abundance of essential genes. Cell 170, 260–272 e268 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Zhang, M. et al. Uncovering the essential genes of the human malaria parasite Plasmodium falciparum by saturation mutagenesis. Science 360, eaap7847 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  60. Laurentino, E. C. et al. Experimentally controlled downregulation of the histone chaperone FACT in Plasmodium berghei reveals that it is critical to male gamete fertility. Cell Microbiol. 13, 1956–1974 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Nichols, B. A. & Chiappino, M. L. Cytoskeleton of Toxoplasma gondii. J. Protozool. 34, 217–226 (1987).

    Article  CAS  PubMed  Google Scholar 

  62. Sivagurunathan, S., Heaslip, A., Liu, J. & Hu, K. Identification of functional modules of AKMT, a novel lysine methyltransferase regulating the motility of Toxoplasma gondii. Mol. Biochem. Parasitol. 189, 43–53 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Pavlou, G. et al. Coupling polar adhesion with traction, spring, and torque forces allows high-speed helical migration of the protozoan parasite toxoplasma. ACS Nano. 14, 7121–7139 (2020).

    Article  CAS  PubMed  Google Scholar 

  64. Pavlou, G. et al. Toxoplasma parasite twisting motion mechanically induces host cell membrane fission to complete invasion within a protective vacuole. Cell Host Microbe 24, 81–96 e85 (2018).

    Article  CAS  PubMed  Google Scholar 

  65. Soldati, D. & Boothroyd, J. Transient transfection and expression in the obligate intracellular parasite Toxoplasma gondii. Science 260, 349–352 (1993).

    Article  CAS  PubMed  Google Scholar 

  66. Brown, K. M., Long, S. & Sibley, L. D. Plasma membrane association by N-acylation governs PKG function in Toxoplasma gondii. mBio 8, e00375-17 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  67. Meissner, M., Brecht, S., Bujard, H. & Soldati, D. Modulation of myosin A expression by a newly established tetracycline repressor-based inducible system in Toxoplasma gondii. Nucleic Acid Res. 29, E115 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Billker, O. et al. Calcium and a calcium-dependent protein kinase regulate gamete formation and mosquito transmission in a malaria parasite. Cell 117, 503–514 (2004).

    Article  CAS  PubMed  Google Scholar 

  69. Fang, H. et al. Epistasis studies reveal redundancy among calcium-dependent protein kinases in motility and invasion of malaria parasites. Nat. Commun. 9, 4248 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  70. Brusini, L., Dos Santos Pacheco, N., Tromer, E. C., Soldati-Favre, D. & Brochet, M. Composition and organization of kinetochores show plasticity in apicomplexan chromosome segregation. J. Cell Biol. 221, e202111084 (2022).

    Article  CAS  PubMed  Google Scholar 

  71. Sebastian, S. et al. A Plasmodium calcium-dependent protein kinase controls zygote development and transmission by translationally activating repressed mRNAs. Cell Host Microbe 12, 9–19 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Gurnett, A. M. et al. Purification and molecular characterization of cGMP-dependent protein kinase from Apicomplexan parasites. A novel chemotherapeutic target. J. Biol. Chem. 277, 15913–15922 (2002).

    Article  CAS  PubMed  Google Scholar 

  73. Plattner, F. et al. Toxoplasma profilin is essential for host cell invasion and TLR11-dependent induction of an interleukin-12 response. Cell Host Microbe 3, 77–87 (2008).

    Article  CAS  PubMed  Google Scholar 

  74. Sheiner, L. et al. A systematic screen to discover and analyze apicoplast proteins identifies a conserved and essential protein import factor. PLoS Pathog. 7, e1002392 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Lentini, G. et al. Structural insights into an atypical secretory pathway kinase crucial for Toxoplasma gondii invasion. Nat. Commun. 12, 3788 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Moon, R. W. et al. A cyclic GMP signalling module that regulates gliding motility in a malaria parasite. PLoS Pathog. 5, e1000599 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  77. Mathur, V. et al. Multiple independent origins of apicomplexan-like parasites. Curr. Biol. 29, 2936–2941.e5 (2019).

    Article  CAS  PubMed  Google Scholar 

  78. Mathur, V. et al. Phylogenomics identifies a new major subgroup of apicomplexans, marosporida class nov., with extreme apicoplast genome reduction. Genome Biol. Evol. 13, evaa244 (2021).

    Article  PubMed  Google Scholar 

  79. Janouškovec, J. et al. Apicomplexan-like parasites are polyphyletic and widely but selectively dependent on cryptic plastid organelles. eLife 8, e49662 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  80. Emms, D. M. & Kelly, S. OrthoFinder: phylogenetic orthology inference for comparative genomics. Genome Biol. 20, 238 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  81. Katoh, K., Kuma, K., Toh, H. & Miyata, T. MAFFT version 5: improvement in accuracy of multiple sequence alignment. Nucleic Acids Res. 33, 511–518 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Eddy, S. R. A new generation of homology search tools based on probabilistic inference. Genome Informatics. International Conference on Genome Informatics 23, 205–211 (2009).

    PubMed  Google Scholar 

  83. Waterhouse, A. M., Procter, J. B., Martin, D. M., 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 

  84. Soding, J. Protein homology detection by HMM-HMM comparison. Bioinformatics 21, 951–960 (2005).

    Article  PubMed  Google Scholar 

  85. Minh, B. Q. et al. IQ-TREE 2: New models and efficient methods for phylogenetic inference in the genomic era. Mol. Biol. Evol. 37, 1530–1534 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Revell, L. J. phytools: an R package for phylogenetic comparative biology (and other things). Methods Ecol. Evolution 3, 217–223 (2012).

    Article  Google Scholar 

Download references

Acknowledgements

We thank for their technical assistance, the team at the Bioimaging Core Facility, R. Visentin at the Protein platform and A. Hainard at the proteomics platform of the faculty of Medicine University of Geneva as well as F. Pojer, L. Abriata and K. Lau from the ‘Protein Production and Structure Core Facility’ at the EPFL. We thank D. Baker for supply of C1. We thank Varsha Mathur and Jan Janouškovec for kindly providing critical sequence data and assemblies used for phylogenetic analysis. This work was supported by the Swiss National Foundation (grant Nos. 310030_185325 to D.S.-F., and BSSGI0_155852 and 310030_208151 to M.B.), the Novartis Foundation (grant No. 19C189 to L.B.) and the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (grant agreement No. 695596 to B.M.).

Author information

Authors and Affiliations

Authors

Contributions

D.S-F., O.V., N.D.S.P., L.B. and M.B. conceptualized the project. N.D.S.P., O.V., R.H., L.B., B.M. and N.T. designed the methodology, and N.D.S.P., O.V., R.H., L.B. and N.T. performed the investigations. N.D.S.P., O.V., R.H., L.B., N.T. and B.M. did the formal analysis. N.D.S.P., D.S-F. and O.V. wrote the original draft and N.D.S.P., D.S-F., O.V., R.H., L.B., N.T. and M.B reviewed and edited the manuscript. D.S-F. and M.B. did the funding acquisition, D.S-F., O.V. and M.B. obtained resources and D.S-F., O.V. and M.B. supervised. All authors contributed to the article and approved the submitted version.

Corresponding authors

Correspondence to Oscar Vadas or Dominique Soldati-Favre.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Microbiology thanks the anonymous reviewers for their contribution to the peer review of this work. Peer reviewer reports are available.

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 U-ExM allows high-resolution localization of apical proteins and conoid extrusion assessment.

a, U-ExM localization of conoidal proteins fused with a mAiD-HA cassette. Scale bar = 2 µm. b, Localization of AKMT and GAC in intracellular vacuoles as seen by U-ExM. Scale bar = 2 µm. c, Representative pictures of some U-ExM samples used for quantification of Fig. 1c. Dash boxes represents crop parasites displayed in Fig. 1b. Scale bar = 5 µm.

Extended Data Fig. 2 Additional phenotyping of Pcr1,4,5,6,7.

a, IFA of Pcr1,4,5,6,7 in non-dividing and dividing parasite showing that Pcr1,4,5,6 are also found in forming daughter cells. White arrowhead = apical pole of mature cells. Cyan arrowhead = apical pole of forming daughter cells. Scale bar = 2 µm. b, Additional U-ExM pictures of Pcr1,4,5,6,7. Scale bar = 2 µm. c, Quantification of the plaque size. The different replicates are displayed in shades of grey. Ten plaques were measures for each condition and replicate. All data are presented as mean ± SD (n = 3). Unpaired two-tailed Student’s t-tests where non-significant (n.s.) if P > 5E-02.

Source data

Extended Data Fig. 3 Conservation of Pcr4 and Pcr5 across Alveolata.

a, Conservation of Pcr4 and Pcr5 in the Alveolate superphylum. Presence (coloured) or absence (white) of homologues identified across alveolate predicted proteomes following. b, Results of a maximum likelihood inference based on an alignment of sequences retrieved for homologues of Pcr4 and Pcr5. Numbers beside nodes indicate bootstrap support (1000 ultrafast bootstrap replicates).

Extended Data Fig. 4 Protein downregulation and microneme secretion assays.

a, Regulation of Pcr4 and Pcr5 on intracellular parasites. Time of auxin incubations are indicated in hour. Actin is used as loading control. Molecular weight in kDa. A small scheme of the experimental timeline is presented. b, Regulation of Pcr4 and Pcr5 on extracellular parasites. Time of auxin incubations are indicated in hour. Molecular weight in kDa. Actin is used as loading control. A small scheme of the experimental timeline is presented. c, Stability of Pcr4,5,6 and FRM1 fused with a mAiD-HA cassette, after short IAA treatment. ISP1 is used as a reference marker appearing early during daughter cells biogenesis. White arrowhead = apical signal at the apical pole of mature cells. Cyan arrowhead = apical signal at the apical pole of forming daughter cells. Black filled arrowhead = absence of signal at the apical pole. d, Representative pictures of microneme secretion quantified in Fig. 2d. Molecular weight in kDa. White arrowhead = secreted MIC2. Black arrowhead = secreted MIC2 trimmed by SUB1 protease. GRA1 is used as loading control while catalase is used as a lysis control.

Source data

Extended Data Fig. 5 Integrity of apical markers and PCR structure.

a, IFA showing that the depletion of Pcr6 did not induce the loss of Pcr5. Scale bars = 2 µm. b, IFA showing that the depletion of FRM1 did not induce the loss of Pcr4 and Pcr5. Scale bars = 2 µm. c, IFA showing that the depletion of Pcr5 does not affect conoid (Cam2) and APR (RNG1/KinA) markers. d, Depletion of Pcr5 leads to the loss of the PCRs signal of AKMT in intracellular parasite by U-ExM. Scale bars = 2 µm. e, Depletion of Pcr5 leads to the loss of the PCRs signal of GAC in intracellular parasite by U-ExM. Scale bars = 2 µm. f, Gallery of additional EM images of Pcr4 and Pcr5 knockdown strains. Scale bar = 200 nm. g, Gallery of additional EM images of Pcr6 knockdown strains. Scale bar = 200 nm. h, Gallery of EM images of FRM1 knockdown strains. Scale bar = 200 nm.

Extended Data Fig. 6 Additional data on Pcr4,5 biochemical properties and interactions.

a, Schematic of Pcr5 two B-Box domains. The amino acid sequences of the four versions of Pcr5 used for complementation are presented. Cysteine and histidine residues (violet), known to chelate zinc ions, were mutated to alanine to create mutated versions of the B-Box domains. b, Representative Coomassie-stained gel of the solubility assay for the individual proteins quantified in Fig. 4a. L = Whole cell lysate. S = Soluble fraction. c, Representative Coomassie-stained gel of a solubility assay when Pcr4 is co-expressed with the four versions of Pcr5, quantified in Fig. 4b. d, Representative Coomassie-stained gels used for solubility quantification and corresponding quantified Western Blot by band densitometry. e, Mass photometry of the Pcr4–Pcr5 complex allowed the measure of the complex mass. f, SEC–MALS coupled with size-exclusion chromatography for the Pcr4–Pcr5 complex. g, Representative Coomassie-stained gel of pull-down assays performed with Pcr4 co-expressed with the four versions of Pcr5. Pull-down was performed using anti-Strep column (Pcr5 pull-down). h, HDX–MS analysis highlight a different dynamic of Pcr5 when mutated in the first B-Box domain (Pcr5MUT1) i, Immunofluorescence of the four complemented strains. The endogenous copy is presented in green (HA) and the second copy is presented in magenta (Ty). White arrowhead = presence of apical signal. Black arrowhead = no apical signal. Scale bars = 2 µm.

Source data

Extended Data Fig. 7 Generation of Plasmodium berghei strains and additional phenotyping.

a, Schematics of the promoter-swap strategy to obtain the Pclag9Pcr4 and Pclag9Pcr5 strains. On the right, integration PCR are presented. b, Quantification of microgametocyte exflagellation for the Pclag9Pcr4 and Pclag9Pcr5 strains. Data are presented as mean ± SD (individual biological replicates are also presented). c, Quantification of ookinete conversion for the Pclag9Pcr4 and Pclag9Pcr5 strains. All three states of the conversion (round, retort and ookinete) are presented. Data are presented as mean ± SD (individual biological replicates are also presented). For all statistical analysis, an unpaired two-tailed Student’s t-tests where non-significant (n.s.) if P>5E-02 was used. All data are presented as mean ± SD (n=3).

Source data

Supplementary information

Supplementary Information

Supplementary Text and Discussion, and HDX–MS peptide map.

Reporting Summary

Peer Review File

Supplementary Table 1

HDX–MS raw data.

Supplementary Table 2

List of strains, primers, plasmids, antibodies and supplies used in this study.

Supplementary Video 1

Representative live-imaging video of wild-type ookinete gliding in Matrigel. Several videos were used to calculate gliding speed and track individual parasites. Acquisition time, 10 min. Acquisition rate, three frames a min. Scale bar, 25 µm.

Supplementary Video 2

Representative live-imaging video of Pclag9Pcr4 ookinete gliding in Matrigel. Several videos were used to calculate gliding speed and track individual parasites. Acquisition time, 10 min. Acquisition rate, three frames a min. Scale bar, 25 µm

Supplementary Video 3

Representative live-imaging video of Pclag9Pcr5 ookinete gliding in Matrigel. Several videos were used to calculate gliding speed and track individual parasites. Acquisition time, 10 min. Acquisition rate, three frames a min. Scale bar, 25 µm

Source data

Source Data Fig. 1

Raw data and statistical analysis.

Source Data Fig. 2

Raw data and statistical analysis.

Source Data Fig. 2

Unprocessed gels.

Source Data Fig. 3

Unprocessed gels.

Source Data Fig. 4

Raw data and statistical analysis.

Source Data Fig. 4

Unprocessed gels.

Source Data Fig. 5

Raw data and statistical analysis.

Source Data Extended Data Fig. 2

Raw data and statistical analysis.

Source Data Extended Data Fig. 4

Unprocessed gels.

Source Data Extended Data Fig. 6

Unprocessed gels.

Source Data Extended Data Fig. 7

Raw data and statistical analysis.

Source Data Extended Data Fig. 7

Unprocessed gels.

Rights and permissions

Springer Nature or its licensor 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

Dos Santos Pacheco, N., Brusini, L., Haase, R. et al. Conoid extrusion regulates glideosome assembly to control motility and invasion in Apicomplexa. Nat Microbiol 7, 1777–1790 (2022). https://doi.org/10.1038/s41564-022-01212-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41564-022-01212-x

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